29.10.2015 Views

October 17-21 2015 Baltimore Maryland

XwI2O0

XwI2O0

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

- - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - -<br />

ASRM <strong>2015</strong><br />

SCIENTIFIC ABSTRACTS to be presented at the 71st Annual Meeting<br />

of the American Society for Reproductive Medicine, <strong>October</strong> <strong>17</strong>-<strong>21</strong>,<br />

<strong>2015</strong>, <strong>Baltimore</strong>, <strong>Maryland</strong>.<br />

(e2) ORAL SESSION<br />

(e106) POSTER SESSION<br />

(e362) AUTHOR INDEX<br />

(e378) TOPIC INDEX<br />

(e380) AUTHOR AND SPOUSE/PARTNER DISCLOSURES INDEX<br />

<strong>October</strong> <strong>17</strong>-<strong>21</strong>, <strong>2015</strong><br />

<strong>Baltimore</strong>, <strong>Maryland</strong><br />

These abstracts of research studies, printed as submitted by the authors, are presented<br />

in the ASRM <strong>2015</strong> meeting sessions and are published in the order of their<br />

presentation. Abstracts of plenary lectures, symposia and interactive sessions are<br />

not included.<br />

Copyright ª<strong>2015</strong> American Society for Reproductive Medicine,<br />

1209 Montgomery Highway, Birmingham, Alabama 35<strong>21</strong>6-2809


The first six papers are candidates for the ASRM Scientific Program<br />

Prize Paper Awards. Six additional candidates will be presented during<br />

the Prize Paper Candidates’ session on Tuesday.<br />

SCIENTIFIC PROGRAM PRIZE PAPER SESSION 1<br />

O-1 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:15 AM<br />

DIETARY PROTEIN INTAKE AND REPRODUCTIVE HORMONES<br />

AND OVULATION: THE BIOCYCLE STUDY. S. L. Mumford, a<br />

A. Alohali, b J. Wactawski-Wende. c a NICHD, NIH, Rockville, MD; b George<br />

Mason University, Arlington, VA; c University at Buffalo, Buffalo, NY.<br />

OBJECTIVE: Protein intake has been associated with changes in steroidogenesis<br />

in women with polycystic ovary syndrome, likely through reducing<br />

hyperinsulinemia. However, the associations among premenopausal women<br />

without a history of infertility or polycystic ovary syndrome are not well<br />

characterized. The objective of this study was to evaluate the association between<br />

protein intake and reproductive hormone concentrations and ovulation<br />

in a healthy population.<br />

DESIGN: Prospective cohort study of 259 healthy premenopausal women<br />

followed for up to two menstrual cycles.<br />

MATERIALS AND METHODS: Estradiol, progesterone, luteinizing hormone,<br />

follicle-stimulating hormone, and testosterone were measured up to<br />

eight times per cycle for up to two cycles, with visits scheduled using fertility<br />

monitors to coincide with menstrual cycle phases. Percent energy from total<br />

protein, animal protein, and vegetable protein were assessed by 24-hour<br />

recall up to four times per cycle. Linear mixed models and generalized linear<br />

mixed models were used to evaluate the association between protein intake<br />

and reproductive hormone levels, and with ovulatory status (peak progesterone<br />

%5ng/mL with no LH peak on the mid or late luteal phase visit), respectively.<br />

All models were adjusted for total energy intake, age, body mass<br />

index, race, fat intake, and physical activity.<br />

RESULTS: Dietary protein consumption, specifically animal protein, was<br />

found to be inversely associated with testosterone concentrations. In particular,<br />

the highest tertile of percent energy from total protein intake was associated<br />

with lower testosterone concentrations (beta -0.05, 95% CI -0.01,<br />

-0.005) compared to the lowest tertile of intake. The highest tertile of percent<br />

energy from animal protein intake was also associated with lower testosterone<br />

(beta -0.02, 95% CI -0.04, -0.0001). No associations were observed<br />

between protein intake and other hormones, including estradiol, progesterone,<br />

luteinizing hormone, or follicle stimulating hormone levels, nor was<br />

there an association between protein intake and ovulation.<br />

CONCLUSIONS: These findings suggest that a diet high in protein, particularly<br />

animal protein, is significantly associated with reduced testosterone<br />

levels among healthy women. These results highlight the importance of<br />

diet on reproductive function and the potential role of protein intake in<br />

androgen synthesis.<br />

Supported by: Intramural Research Program, DIPHR, NICHD, NIH.<br />

O-2 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:30 AM<br />

ROCKING THE DOGMA OF SEMINAL ROUND CELLS. Q. V. Neri,<br />

T. Cozzubbo, Z. Rosenwaks, G. D. Palermo. Reproductive Medicine, Weill<br />

Cornell Medical College, New York, NY.<br />

OBJECTIVE: To revisit the origin and significance of the sporadic presence<br />

of round cells in the ejaculates of men screened for male infertility.<br />

DESIGN: In a prospective fashion, a total of 4,051 men undergoing male<br />

infertility screening in a period of 24 months were included in the study. RC<br />

cells were characterized for WBC components versus exfoliated germ cells<br />

by testing for multiple markers of ploidy as well as protamine assays. Cases<br />

displaying R2 million RC were screened for bacteria. The effect of RC on<br />

clinical outcome was assessed in specimens used for ART.<br />

MATERIALS AND METHODS: Raw specimens containing RC were<br />

processed by peroxidase and other leukocyte assays, specific stains for protamines<br />

were used to identify spermiogenic stage, aneuploidy (9 chromosome<br />

FISH) assessment was carried out, and the presence of various Sertoli-cell<br />

cytoplasmic remnants was analyzed to identify and characterize immature<br />

germ cells (IGC).<br />

RESULTS: A total of 4,810 ejaculated samples were processed for semen<br />

analysis. The average age of the men involved was 39.2 7yrs. Semen samples<br />

had a mean concentration of 40.7 31 million, motility of 42.6 35%,<br />

and morphology of 2.3 2%. Round cells were identified in 261 of the specimens<br />

evaluated, representing a proportion of 5.4%. Men presenting with<br />

round cell had comparable age but lower sperm concentration and<br />

morphology than the control (P < 0.0001). Aneuploidy was detected in 91<br />

specimens, of which 30.8% (28/91) presented with round cells. The aneuploidy<br />

rate of 4.3%, remarkably higher than the control (2.3%; P < 0.001).<br />

Sperm aneuploidy rate positively correlate with the amount of RC (P ¼<br />

0.0001).In 44 men, <strong>17</strong> of them in 18 cycles had up to 1.9 million RC without<br />

affecting fertilization and clinical pregnancy rates when compared to control<br />

(n¼365 cycles). In 27 men with 33 ICSI cycles with over R2 million round<br />

cells, the fertilization rate trended lower and the miscarriage rate was significantly<br />

increased (P ¼ 0.05).The absence of any correlation between RC and<br />

bacteriological growth as well as the results of marker testing indicates that<br />

seminal RC are mostly immature germ cells that stain for vimentin and<br />

inhibin B. Moreover, their modest protamine content and their haploid status<br />

confirm that they are post-meiotic. Sequential observation in the same man<br />

showed that the RC episode was followed by an amelioration of the semen<br />

parameters and that the presence of RC corresponds to flu season peaks.<br />

CONCLUSIONS: Seminal round cell presence is not a marker of infectiousness<br />

but rather a transient indicator of spermatogenic insult that possibly<br />

occurs in most men following a mild and transient ailment such as the flu.<br />

Supported by: WCMC.<br />

O-3 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:45 AM<br />

THE SUPEROVULATED ENVIRONMENT, INDEPENDENT OF<br />

EMBRYO VITRIFICATION, RESULTS IN LOW BIRTHWEIGHT<br />

FOLLOWING EMBRYO TRANSFER IN A MOUSE<br />

MODEL. R. S. Weinerman, T. S. Ord, C. Coutifaris, M. A. Mainigi. Division<br />

of Reproductive Endocrinology and Infertility, University of Pennsylvania,<br />

Philadelphia, PA.<br />

OBJECTIVE: Babies born following fresh embryo transfer are of lower<br />

birthweight than babies born following frozen embryo transfer. The objective<br />

of this study was to determine, in a mouse model, if this difference is due to<br />

the superovulated (SO) peri-implantation environment or embryo vitrification<br />

(VIT).<br />

DESIGN: Laboratory research.<br />

MATERIALS AND METHODS: Female CF1 mice were superovulated<br />

with gonadotropins and mated to male mice heterozygous for green-fluorescent<br />

protein (GFP). 2-pronuclear embryos were collected and cultured to the<br />

blastocyst stage and a subset of embryos were vitrified. For each transfer<br />

experiment, 10 blastocysts, 5 fresh and 5 vitrified/thawed, were transferred<br />

into a host mouse, using GFP to tag the embryos as fresh or frozen. Transfers<br />

were performed into pseudopregnant females created through either natural<br />

mating to vasectomized males (n¼30) or superovulation followed by mating<br />

to vasectomized males (n¼45). This resulted in 4 experimental groups: 1)<br />

Natural environment, fresh embryos (Nat-Fresh) 2) Natural environment,<br />

frozen embryos (Nat-VIT) 3) Superovulated environment, fresh embryos<br />

(SO-Fresh) 4) Superovulated environment, frozen embryos (SO-VIT). Pregnant<br />

mice were sacrificed near term (E18.5) for assessment of fetal and<br />

placental weights and GFP status. An a-priori power calculation determined<br />

that <strong>17</strong> fetuses per group would provide 80% power to detect a 15% difference<br />

in fetal weight with an alpha of 0.05.<br />

RESULTS: There was no difference in litter size between natural (n¼15<br />

litters) and SO (n¼13 litters) hosts. Although there was no difference in<br />

placental weight, there was a highly significant difference (p


O-4 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:00 PM<br />

THE SYNCHRONIZATION OF THE LEIOMYOMA<br />

EXTRACELLULAR MATRIX SIGNALING PATHWAYS OF<br />

SURGICAL SPECIMENS IN RESPONSE TO ULIPRISTAL<br />

ACETATE. M. Malik, a J. Cox, b J. Britten, a A. Patel, c L. K. Nieman, d<br />

W. H. Catherino. a a Obstetrics and Gynecology, Uniformed Services University<br />

of the Health Sciences, Bethesda, MD; b Program in Reproductive and<br />

Adult Endocrinology, Eunice Kennedy Shriver National Institute of Child<br />

Health and Human Development, Bethesda, MD; c Henry Jackson Foundation,<br />

Bethesda, MD; d Reproductive Biology and Medicine Branch, Eunice<br />

Kennedy Shriver National Institute of Child Health and Human Development,<br />

Bethesda, MD.<br />

OBJECTIVE: To elucidate the pathways that may be involved in the action<br />

of Ulipristal acetate, a selective progesterone receptor modulator, on leiomyoma<br />

extracellular matrix (ECM) production and breakdown.<br />

DESIGN: Uterine leiomyomas are characterized by increased stiffness as a<br />

result of excessive and disordered ECM which contributes to the total bulk of<br />

the tumor. Ulipristal acetate reduced leiomyoma size in several randomized<br />

placebo-controlled studies. We have previously demonstrated that Ulipristal<br />

acetate reduces the total ECM that forms the bulk of the tumor in over 60% of<br />

patients studied. Clinical tissue was sent for RNASeq analysis to identify<br />

pathways that may be involved in leiomyoma ECM breakdown. In-vitro analysis<br />

was carried out to analyze the effect of the medroxyprogesterone acetate<br />

(MPA) and Ulipristal acetate.<br />

MATERIALS AND METHODS: RNA isolated from clinically collected<br />

fibroid and myometrial tissue from patients treated with Ulipristal acetate<br />

in a placebo-controlled, randomized trial, underwent RNASeq analysis<br />

(Beckman Coulter Genomics Inc). To further evaluate the impact of Ulipristal<br />

acetate, 3-dimensional leiomyoma cultures were treated with MPA, Ulipristal<br />

acetate, and combinations. The expressions of ECM genes as well<br />

as Wnt/b-catenin pathway genes were analyzed. Proteins were analyzed using<br />

Western Blot.<br />

RESULTS: Both TGFb3 and TGFb1 transcripts were reduced in tumors of<br />

women treated with Ulipristal acetate. The reduction was also observed at<br />

protein levels. Multiple components of the TGFb signaling pathway demonstrated<br />

an overall reduction. Reduced transcript and protein expressions were<br />

observed in matrix metalloproteinases (MMPs), such as MMP-9 (2.38-fold)<br />

and MMP-13 (1.97-fold). Decreased WNT5a protein indicated an involvement<br />

of the WNT/b-catenin pathway that was Supported by changes in<br />

expression of Frizzled gene transcripts such as FZD6. In addition, there<br />

was increased expression of aquaporins and other osmotic stress regulators<br />

such as NFAT5, which are known to be regulated by WNT pathway.<br />

Leiomyoma cells in 3D culture, in response to MPA, demonstrated a<br />

3.23+/-0.15-fold increase in b-catenin expression. Ulipristal acetate alone<br />

and in combination with MPA decreased b-catenin expression below that<br />

of untreated leiomyoma cells.<br />

CONCLUSIONS: TGFb signaling pathway in association with both<br />

canonical and non-canonical WNT pathway may participate in Ulipristal<br />

acetate clinical activity by reducing the total amount of ECM proteins produced,<br />

and by decreasing osmotic stress. Ulipristal acetate may also reduce<br />

the progesterone-mediated fibrosis.<br />

Supported by: This research was Supported by Intramural grant from Uniformed<br />

Services University of the Health Sciences, QP85GF13 and<br />

RO85298815. The research was also Supported, in part, by the intramural<br />

research Program in Reproductive and Adult Endocrinology, NIH; NIH<br />

R<strong>21</strong> and EMD Serono.<br />

O-5 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:15 PM<br />

ELECTIVE SINGLE EMBRYO TRANSFERS (ESET) AS COMPARED<br />

TO COMPREHENSIVE CHROMOSOME SCREENING<br />

(CCS). A. R. Anderson, D. Taylor, K. C. Chiles, U. Balthazar,<br />

A. S. Browne. Reproductive Medicine Associates of Texas, San Antonio, TX.<br />

OBJECTIVE: To investigate the association with eSET with optimal embryo<br />

development and CCS testing for embryo selection.<br />

DESIGN: A prospective cohort study.<br />

MATERIALS AND METHODS: CCS aneuploidy testing was performed<br />

on 297 embryo transfers. Alternately, <strong>17</strong>2 eSET blastocyst transfers were<br />

completed based on at least two high quality day 5 embryos available for<br />

transfer where one is transferred and at least one selected for cryopreservation<br />

on day 5. In the CCS subgroup, 5 to 7 cells were biopsied from day 5 and day 6<br />

blastocysts in a HEPES buffered Medium. Biopsied cells were placed in<br />

separate PCR tubes and analyzed for 24 chromosomes at an outside reference<br />

laboratory. All embryos in the CCS subgroup were subjected to vitrification<br />

for a delayed embryo transfer in a subsequent frozen embryo cycle.<br />

RESULTS: There was no significant difference in ongoing pregnancy,<br />

average age, embryos transferred, or miscarriage rates when eSET criterion<br />

or CCS was utilized for embryo selection. In the eSET group there were <strong>17</strong>2<br />

transfers with 126 (73%) positive pregnancies, 25 (20%) losses, and 101<br />

(59%) ongoing pregnancies. From 297 CCS embryo transfers there were<br />

209 (70%) positive pregnancies, 47 (22%) losses, and 162 (54%) ongoing<br />

pregnancies respectively with no significant difference between these two<br />

subgroups. However, there was a significant (P


expression of AMH, FSHR, Inhibina and Inhibinb in growing follicles in<br />

group 3 versus group 2.Tracking studies demonstrated the human MSCs<br />

evenly infiltrating and repopulating growing follicles in treated ovaries.<br />

Finally, breeding data showed significant increase in both the number of<br />

pregnancies and total number of pups per litter in group 3 compared to group<br />

2(P¼ 0.02).<br />

CONCLUSIONS: Our study shows that intra-ovarian injected human<br />

BMSCs were able to restore ovarian hormone production, reactivate folliculogenesis<br />

in chemotherapy-damaged ovaries and reverse infertility in this<br />

preclinical model. This approach carries high promise to women with chemotherapy-induced,<br />

and potentially other types of, premature ovarian failure.<br />

MENOPAUSE<br />

O-7 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:15 AM<br />

DRUG METABOLIZING ENZME POLYMORPHISMS ARE<br />

ASSOCIATEDWITH CHEMOTHERAPY RELATEDAMENORRHEA<br />

IN YOUNG BREAST CANCER SURVIVORS. L. M. Charo, a<br />

M. V. Homer, a L. Natarajan, a C. Haunschild, b A. DeMichele, c I. Su. a<br />

a UC San Diego, San Diego, CA;<br />

b Stanford University, Stanford, CA;<br />

c University of Pennsylvania, Philadelphia, PA.<br />

OBJECTIVE: To test if candidate single nucleotide polymorphisms<br />

(SNPs) in enzymes involved in cyclophosphamide activation or detoxification<br />

are associated with time to ovarian failure after chemotherapy in young<br />

breast cancer survivors.<br />

DESIGN: Prospective cohort study.<br />

MATERIALS AND METHODS: Premenopausal breast cancer survivors<br />

(n ¼ 116) with Stages 0 to III disease and planned cyclophosphamide-based<br />

chemotherapy were enrolled at diagnosis from three academic breast programs<br />

and followed longitudinally for menstrual pattern. Participants were<br />

genotyped for SNPs in genes involved in cyclophosphamide activation<br />

(CYP3A4 [rs1067910] and CYP2C19 [rs42244285]) and detoxification<br />

(GSTP1 [rs1695] and GSTA1 [rs4715332]). The primary endpoint was<br />

time to chemotherapy-related amenorrhea (CRA), defined as >12 months<br />

of amenorrhea after the end of chemotherapy. Using the time-to-event<br />

method, the association between SNPs and CRA were assessed using Cox<br />

proportional hazard regression models. A priori sample size calculations estimated<br />

80% power to detect relative risks of 1.7-2.6.<br />

RESULTS: The cohort had a median age of 39.7 years (range 20.8-46.1) at<br />

end of chemotherapy and median follow up of 594.5 days (range 23-<strong>21</strong>19).<br />

28% experienced CRA. Survivors with at least one major allele of GSTA1<br />

had significantly lower hazards of developing CRA compared to survivors<br />

who were homozygous for the minor allele (HR 0.22 [95% CI 0.61-0.91],<br />

p¼0.04). Survivors with at least one major allele of CYP2C19 had significantly<br />

higher hazards of developing CRA compared to survivors who were<br />

homozygous for the minor allele (HR 4.56 [95% CI 1.54-13.57], p¼0.01).<br />

CYP3A4 and GSTP1 SNPs were not related to CRA. Increased age was<br />

also associated with CRA. In separate multivariable models adjusting for<br />

age and BMI, GSTA1 remained significantly associated with CRA (HR<br />

0.23 [95% CI 0.06-0.96], p¼0.04) while CYP2C19 was attenuated (HR<br />

2.75 [0.89-8.49], p¼0.08).<br />

CONCLUSIONS: In younger breast cancer patients undergoing cyclophosphamide-based<br />

chemotherapy, the presence of one or more major alleles<br />

of GSTA1 was found to have lower risk of developing CRA. Inter-individual<br />

variation in enzymes involved in chemotherapy metabolism is related to posttreatment<br />

ovarian function.<br />

Supported by: MRSG-08-110-01-CCE, HD058799, T32 HD007203.<br />

O-8 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:30 AM<br />

PRESCRIBING OF COMPOUNDED AND COMMERCIALLY<br />

AVAILABLE MENOPAUSAL HORMONE THERAPY BY<br />

OBSTETRICIAN-GYNECOLOGISTS AND FAMILY MEDICINE<br />

PHYSICIANS. J. P. Dubaut, F. Dong, B. L. Tjaden, D. A. Grainger,<br />

J. Duong, L. L. Tatpati. University of Kansas School of Medicine -<br />

Wichita, Wichita, KS.<br />

OBJECTIVE: Compounded bioidentical hormone use has risen in the<br />

United States (1). The American College of Obstetricians and Gynecologists<br />

(ACOG) released a committee opinion on this topic, including guidelines for<br />

its members (2). We explored factors that may influence prescribing practices<br />

of compounded and bioidentical menopausal hormonal therapy (MHT) and<br />

level of agreement with relevant ACOG statements among obstetrician-gynecologists<br />

(OB/GYNs) and family medicine physicians (FPs).<br />

DESIGN: Cross-sectional physician survey.<br />

MATERIALS AND METHODS: After Institutional Review Board<br />

approval, investigators created an online anonymous survey. The survey<br />

link was emailed to Kansas OB/GYN and FP physicians. The initial and<br />

reminder emails were sent 4 weeks apart; the survey was available for 3<br />

months. Survey constructs included: demographics, MHT knowledge,<br />

MHT prescribing practices, and opinions on statements published in the<br />

ACOG Committee Opinion 532. Chi-square analyses were conducted to<br />

identify associations between specialties, practices and opinions on ACOG<br />

statements.<br />

RESULTS: Overall response rate was 11.1% (150 of 1349). The response<br />

rate of OB/GYNs (27%) exceeded that of FPs (7.2%). Of 150 respondents,<br />

53.3% were FPs. The majority of respondents identified as female (64%),<br />

were under 50 years old (57.4%), and worked in cities with populations<br />

over 100,000 (70.5%). In the past year, 84.5% prescribed conventional<br />

MHT, 83% prescribed commercial bioidentical MHT, and 58.9% prescribed<br />

compounded bioidentical MHT. OB/GYNs prescribed more than<br />

FPs in each category; the difference in prescribing commercial bioidentical<br />

MHT was statistically significant (p¼0.03). Hormone levels were monitored<br />

in at least some patients by 40% of physicians. When asked whether<br />

compounded MHT was regulated by the Food and Drug Administration<br />

(FDA), 76.7% answered ‘‘no’’, 4% answered ‘‘yes’’, and 19.3% were<br />

‘‘not sure’’ or declined to answer. Most respondents stated efficacy, risks,<br />

tolerability, cost, patient preference, and experience of previous patients<br />

were important factors influencing their MHT prescribing practices. FDA<br />

regulation was not important to 15.3%, while customization was important<br />

to 62% of physicians. The majority of respondents agreed with 10 ACOG<br />

statements regarding MHT (range 53-97%), but at least 15% showed<br />

disagreement with 7 of 10 statements. More OB/GYNs than FPs agreed<br />

‘‘saliva levels are not biologically meaningful’’ (p


RESULTS: Ethnic distribution did not differ in each reproductive category:<br />

39 premenopausal South Asians and 34 Europeans; 10 perimenopausal<br />

South Asians and 14 Europeans; and 42 postmenoapusal South Asians and 39<br />

Europeans. Body fatness variables increased with reproductive ageing to an<br />

almost similar degree in both ethnic groups (p for trend greater than 0.05 for<br />

all associations) without ethnicity modifying the gradient of the associations<br />

between menopausal state and body fatness (p for interaction with ethnicity<br />

greater than 0.05 for all associations). Moderate to vigorous physical activity<br />

and VO2max decreased in a similar fashion across the reproductive stages in<br />

both ethnic groups whereas energy intake remained unchanged. Body<br />

fatness, physical activity and fitness did not differ among the ethnic groups<br />

for each reproductive stage either. Metabolic biomarkers (insulin, total<br />

cholesterol, triglycerides and blood pressure) deteriorated with reproductive<br />

ageing in both ethnic groups in a similar degree. Notably, HbA1c levels<br />

increased to a much greater degree with reproductive ageing in the South<br />

Asians than in the Europeans (p¼0.02 for interaction with ethnicity).<br />

CONCLUSIONS: The increase in HbA1c levels in healthy women<br />

without overt T2DM during menopausal transition was much greater in the<br />

South Asians than in their White counterparts and this discrepancy was not<br />

explained by a greater deterioration in body composition or physical activity<br />

variables along with reproductive ageing.<br />

O-10 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:00 PM<br />

ASSOCIATION BETWEEN POLYCYSTIC OVARY SYNDROME<br />

AND HOT FLASH PRESENTATION DURING THE<br />

MENOPAUSE TRANSITION. O. Yin, a H. A. Zacur, b J. A. Flaws, c<br />

M. S. Christianson. a a Johns Hopkins University School of Medicine,<br />

Lutherville, MD;<br />

b Reproductive Endocrinologist, Lutherville, MD;<br />

c University of Illinois, Urbana, IL.<br />

OBJECTIVE: While polycystic ovary syndrome (PCOS) is the most common<br />

endocrinopathy in reproductive-age women, most research has focused<br />

on young women and the impact of PCOS on the menopause transition remains<br />

poorly understood. This study aims to determine the influence of<br />

PCOS on hot flash presentation in midlife women.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Subjects were recruited from an ongoing<br />

cohort study involving 748 midlife women aged 45-54 from an urban<br />

metropolitan area. Subjects completed detailed questionnaires that included<br />

hot flash symptom and demographic information. Between June 2014 and<br />

March <strong>2015</strong>, 656 patients were contacted by telephone. Those who agreed to<br />

participate were screened for history of PCOS using the Rotterdam criteria.<br />

Chi square analysis and Wilcoxon rank sum test were used as needed to compare<br />

subjects with a history of PCOS with other midlife women. Multivariate logistic<br />

regression was performed to identify factors associated with hot flashes at<br />

midlife; odds ratios (OR) with 95% confidence intervals (CI) were calculated.<br />

RESULTS: A total of 453 women (69%) responded to the telephone interview<br />

and 9.3% (n¼42) met diagnostic criteria for PCOS. The remaining 411<br />

were included as controls. Mean PCOS subject age was 48.0 and body mass<br />

index (BMI) was 27.3. The majority of subjects were Caucasian (73%) with a<br />

smaller proportion African American (<strong>21</strong>%) and other ethnicities (2%).<br />

PCOS and control groups were not statistically different with respect to<br />

age, BMI, race, income, smoking, drinking, physical activity, number of periods<br />

in the last year, and testosterone, progesterone, or estradiol levels.<br />

Multivariate logistic regression analysis demonstrated that PCOS was not<br />

associated with increased odds of hot flash prevalence, frequency, duration,<br />

or severity. Smoking was the only variable that demonstrated an increased association<br />

with experiencing hot flashes (OR 2.0, 95% CI 1.05-3.98).<br />

CONCLUSIONS: A history of PCOS was not associated with increased<br />

hot flashes during the menopause transition in this study. These data suggest<br />

that women with PCOS have similar hot flash presentations in midlife as<br />

compared to the general population. Additional research should continue<br />

to investigate the health and quality of life implications associated with a<br />

history of PCOS in the aging population.<br />

Supported by: 2R01AG018400 - 05A2.<br />

O-11 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:15 PM<br />

ORAL TIBOLONE (2.5 MG) VERSUS TRANSDERMAL ESTRA-<br />

DIOL GEL (0.06%, 2.5 GM) - EFFECTS ON SERUM CALCIUM<br />

AND 25-HYDROXY VITAMIN D3 LEVELS AFTER SURGICAL<br />

MENOPAUSE. S. M. Bhattacharya a A. Jha. b a Obstetrics and Gynecology,<br />

KPC Medical College, Kolkata, India; b Research Associate, West Virginia,<br />

WV.<br />

OBJECTIVE: To compare the effects of oral Tibolone (2.5 mg) and Transdermal<br />

Estradiol gel (0.06%; 2.5g) on serum Calcium and 25-hydroxy<br />

Vitamin D3 levels in surgically postmenopausal women after 6 months of<br />

treatment.<br />

DESIGN: Open label randomized controlled study.<br />

MATERIALS AND METHODS: 144 women (40-52 years of age) with<br />

surgical menopause (duration 3-18 months and done for benign gynecological<br />

causes and having distressing menopausal symptoms) with preset inclusion<br />

and exclusion criteria were randomized in 1:1 ratio, between<br />

01.03 2013 and 30.06.2014 into 2 groups. Prior ethical approval and<br />

informed written consents (from all participants) were obtained. Group A<br />

received Tibolone (2.5mg, daily) orally and Group B received Transdermal<br />

Estradiol gel (0.06%; 2.5 g, daily). The primary outcomes were comparison<br />

of the absolute changes in serum Calcium and 25- hydroxy Vitamin D3<br />

levels following 6 months of treatment between the two groups. Body<br />

mass index (BMI) and blood pressure (systolic, SBP: diastolic, DBP)<br />

were recorded. Serum calcium and 25- hydroxyl vitamin D3 levels were<br />

measured. All parameters were studied at baseline and after 6 months of<br />

treatment. Sample size was calculated based on a pilot study where 6<br />

months treatment with Tibolone increased vitamin D3 level by 4.91<br />

ng/ml (SD ¼ 6.54) and Transdermal estradiol gel increased vitamin D3<br />

level by 2.08 ng/ml (SD ¼ 3.99). It was calculated that a sample size of<br />

58 patients per group would have 80% power and 95% confidence level<br />

with 2-sided test of significance to detect this mean difference between<br />

the two groups. 6 cases in the Group A and 9 cases in the Group B were<br />

lost to follow up after 6 months of treatment.<br />

RESULTS: Intent-to-treat analysis showed that after 6 months of treatment,<br />

effects of the two treatment interventions were identical in the studied<br />

population. In both intervention arms, patients recorded increase in serum<br />

Calcium and Vitamin D3 levels, but the mean differences were not statistically<br />

significant.<br />

CONCLUSIONS: There were no differences in therapeutic effects of oral<br />

Tibolone and Transdermal Estradiol gel on BMI, blood pressure, serum Calcium<br />

and Vitamin D3 levels in surgically postmenopausal women after 6<br />

months of treatment. Effects of the changes in serum Calcium and Vitamin<br />

D3 levels, by either mode of treatment, on bone remodelling in menopausal<br />

women would need more studies.<br />

Support: None.CTRI registration no. - CTRI/2013/02/003341.<br />

O-12 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:30 PM<br />

LONG TERM HORMONE REPLACEMENT THERAPY (HT)<br />

DOES NOT AFFECT POST-MENOPAUSAL TOTAL BODY<br />

COMPOSITION. A. H. Bayer, a K. N. Goldman, b R. Mauricio, a<br />

M. J. Nachtigall, b F. Naftolin, b L. E. Nachtigall. b a NYU School of Medicine,<br />

New York, NY; b Department of Obstetrics & Gynecology, NYU School of<br />

Medicine, New York, NY.<br />

OBJECTIVE: The impact of HT on menopause-related changes in body<br />

composition is not resolved. We sought to evaluate differences in total<br />

body composition in post-menopausal women who had taken HT for an<br />

average of 14 years.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Post-menopausal women (40-100 years)<br />

on HT for a minimum of 6 years in a university-affiliated menopause clinic<br />

underwent annual Dual Energy X-ray Absorptiometry (DXA) from August<br />

2004 to <strong>October</strong> 2014. Annual DXA scans from untreated post-menopausal<br />

women seen in the same clinic during the same time period were evaluated as<br />

controls. Exclusion criteria included the diagnosis of primary ovarian insufficiency,<br />

DXA data unavailable or incomplete, and DXA scan not performed<br />

while on HT (for HT group). Primary outcomes were percent (%) total body<br />

fat and % total lean body mass. Secondary outcomes included body mass index<br />

(BMI) and relevant co-morbidities. Data were analyzed using Student’s<br />

t-test and Fisher’s exact test where appropriate (p


differences in medical co-morbidities in women on HT compared to controls,<br />

including but not limited to osteoporosis, diabetes, hypertension, hyperlipidemia,<br />

coronary artery disease, fibrocystic breast disease, endometrial polyps,<br />

colonic polyps, endometrial cancer, ovarian cancer, mammogram BIRADS<br />

(breast imaging-reporting data system) 3, 4, or 5, or breast cancer.<br />

CONCLUSIONS: Evaluation by DXA of post-menopausal women<br />

receiving HT for up to 25 years demonstrates that long-term HT has no significant<br />

impact on body composition. It is notable that there was no increased<br />

prevalence of medical co-morbidities between the treated and control groups.<br />

These findings may inform the risk-benefit ratio when considering long-term<br />

HT for post-menopausal women.<br />

Supported by: Pfizer Corporation.<br />

REPRODUCTIVE ENDOCRINOLOGY: RESEARCH 1<br />

O-13 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:15 AM<br />

ELEVATED SERUM ANTI-MULLERIAN HORMONE (AMH)<br />

STALLS OVARIAN FOLLICLE DEVELOPMENT BY<br />

DOWNREGULATING FSH- AND LH-RECEPTORS AND<br />

INHIBIN-B PRODUCTION. L. Detti, a L. J. Williams, a<br />

S. E. Osborne, a N. M. Fletcher, b G. M. Saed. b a Obstetrics and Gynecology,<br />

University of Tennessee, Memphis, Memphis, TN; b Wayne State University,Detroit,MI.<br />

OBJECTIVE: In granulosa cell cultures AMH inhibits the FSH-dependent<br />

follicular growth and the cyclic selection for dominance [1-3]. Women with<br />

polycystic ovary syndrome (PCOS) have high serum AMH levels from<br />

increased production [4] which are correlated to the follicle number and<br />

AMH levels [5]. We tested the hypothesis that administration of recombinant<br />

AMH to ovarian cortex fragments would inhibit follicular development by<br />

downregulating hormone receptors’ expression.<br />

DESIGN: Pilot experimental study with ovarian cortex obtained from 3<br />

patients.<br />

MATERIALS AND METHODS: Immediately after explant the ovarian<br />

cortex specimens were divided into 5 equal fragments. One fragment was<br />

flash-frozen (untreated) and four were incubated for 48 hours 37 Cina<br />

pH-adjusted gamete buffer media with increasing AMH concentrations of<br />

0-5-25-50 ng/ml. After incubation, all specimens were rinsed and flashfrozen<br />

for PCR analyses, which were executed in triplicates. We utilized<br />

real time RT-PCR to determine mRNA levels for FSH-R, LH-R and<br />

Inhibin-B in ovarian cortex tissue. We performed ANOVA with Tukey post<br />

hoc tests to evaluate changes in mRNA levels among the five different fragments.<br />

A p


were collected. We used microanalytical assays to measure the levels of ATP<br />

and citrate in single oocytes. To evaluate spindle and chromosome alignment,<br />

mice were injected with PMSG and hCG, sacrificed and ovulated meiosis II<br />

(MII) oocytes were collected. Mature oocytes were stained with tubulin and<br />

DAPI. Imaging was performed on a Leica confocal microscope and analysis<br />

performed blindly with ImageJ software. ANOVA, students t-tests, and chisquare<br />

analysis were used in the statistical analysis as appropriate.<br />

RESULTS: HF mice weighed significantly more than mice on the control<br />

diet (28g vs <strong>21</strong>.7g, p


Supported by: Expanding the Boundaries Research Grant, Brigham &<br />

Women’s Hospital, Harvard Medical School NIH Grant RO1 HD053112,<br />

R<strong>21</strong> HD061259.<br />

O-18 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:30 PM<br />

SELECTIVE PROGESTERONE RECEPTOR MODULATOR<br />

(SPRM), CDB-4142 INHIBITS DECIDUALIZATION OF HUMAN<br />

AND MOUSE ENDOMETRIAL STROMAL CELLS AND PREVENTS<br />

EMBRYO IMPLANTATION IN THE MOUSE<br />

UTERUS. S. Kuokkanen, a L. Zhu, b B. McAvey, c J. Pollard. d a Obstetrics<br />

& Gynecology, Albert Einstein College of Medicine, Bronx, NY; b Department<br />

of Molecular Biology, Albert Einstein College of Medicine, Bronx,<br />

NY; c Obstetrics & Gynecology, Icahn School of Medicine, Mt. Sinai and<br />

RMA of NY, New York, NY; d College of Medicine and Veterinary Medicine,<br />

University of Edinburgh, Edinburg, United Kingdom.<br />

OBJECTIVE: SPRMs with selective effects on hormone responsive tissues<br />

hold promise for long-term medical management of fibroids and endometriosis.<br />

Telapristone acetate (CDB-4124) is a derivative form of ulipristal<br />

acetate, a reliable emergency contraceptive (EllaÒ). Here, we examined the<br />

impact of CDB-4124 on in vivo decidualization and embryo implantation in<br />

the mouse uterus and in vitro decidualization of human endometrial stromal<br />

cells (hESC).<br />

DESIGN: Controlled laboratory study.<br />

MATERIALS AND METHODS: Endometrial tissue was collected by biopsy<br />

from healthy volunteers. After isolation, stromal cells were seeded in 6-<br />

well plates in DMEM-F12 with 2% charcoal/dextran-treated FBS. The decidualization<br />

media was supplemented with 1uM progesterone (P), 30 nM<br />

estradiol (E) and CDB or 0.1% EtoH vehicle control. hESC decidualization<br />

was monitored by morphology and prolactin (PRL) and IGFBP1 mRNA by<br />

qrtPCR. Decidualization of the mouse uterus was induced in ovariectomized,<br />

E/P or E/P/CDB treated mice by intraluminal injection of peanut oil or PBS<br />

(control) and decidual response was assessed by uterine horn weight and<br />

morphology. For implantation study, copulation of wild type CD1 mice<br />

was monitored with daily vaginal plugs and CDB was administered on pregnancy<br />

d 3-6. The mice were euthanized on pregnancy d7 and implantation<br />

sites were quantified.<br />

RESULTS: After 9 days of incubation, hESC in the decid. media transformed<br />

from fibroblast-like cells to round decidual cells and produced<br />

increased levels of the decidualization markers compared to the control cells<br />

(PRL FC¼41, p¼0.0003; IGFBP1 FC¼76, p¼0.01). hESC decidualized in<br />

the decid. media with CDB 0.1uM (PRL FC¼8.5, p¼0.0006; IGFBP1<br />

FC¼10, p¼0.03). In contrast, hESC cultured in the decid. media with<br />

CDB at 1, 3 or 9 uM remained spindle-shaped and PRL and IGFBP1<br />

mRNAwere at the level of the control cells. CDB inhibited decidual response<br />

in the mouse uterus and the weight ratio of the oil stimulated to the control<br />

mouse horn was approximately 3 in E/P treated mice (0.1095 mg 0.034<br />

mg vs. 0.034 mg 0.0072, p¼0.01), but only 1.03 in E/P/CDB treated<br />

mice (0.042 mg 0.011 mg vs. 0.041 mg 0.011, p¼0.25). Implantation<br />

sites were absent in the CDB treated uteri after copulation compare to an<br />

average of 12 sites in the control mice.<br />

CONCLUSIONS: CDB-4124 acts as progesterone receptor antagonist in<br />

human and mouse endometrial stromal cells, completely inhibiting stromal<br />

decidualization and inhibiting embryo implantation in the mouse uterus.<br />

These findings describe a novel mechanism as to how SPRM can be efficacious<br />

in treating sex-steroid dependent conditions and, as post-coital contraceptive,<br />

delay endometrial maturation and extend contraceptive efficacy<br />

beyond the time of ovulation.<br />

Supported by: U54 HD058155, ABOG/AAOGF Bridge Funding (S.K).<br />

MALE REPRODUCTION AND UROLOGY:<br />

TRAVELING SCHOLARS<br />

O-19 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:15 AM<br />

COMPARISON OF THREE METHODS OF PENILE VIBRATORY<br />

STIMULATION (PVS) IN MEN WITH SPINAL CORD INJURY<br />

(SCI). W. Chong, a E. Ibrahim, b T. Aballa, b C. Lynne, a N. L. Brackett. b<br />

a University of Miami/Jackson Memorial Hospital, Miami, FL; b The Miami<br />

Project to Cure Paralysis, Miami, FL.<br />

OBJECTIVE: PVS is recommended as the first line of therapy for semen<br />

retrieval in anejaculatory men with SCI. This study compared ejaculatory<br />

success rates and patient preference for three methods of PVS within the<br />

same group of men with SCI.<br />

DESIGN: Prospective, three-way crossover design.<br />

MATERIALS AND METHODS: Subjects were 15 men with SCI whose<br />

level of injury was T10 or rostral. Each subject received the following three<br />

methods of PVS, with an interval of 2-4 weeks between each method.<br />

Method 1 (M1): applying one FertiCare PersonalÔ (Multicept, Denmark)<br />

device to the dorsum or frenulum of the glans penis; Method 2 (M2): ‘‘sandwiching’’<br />

the glans penis between two FertiCare devices; Method 3 (M3):<br />

sandwiching the glans penis between the two vibrating surfaces of the Viberect<br />

X3Ô (Reflexonic, Frederick, MD) device. To control for sequencing effects,<br />

5 subjects received PVS in the following sequence: M1, M2, M3; 5<br />

received the sequence M2, M3, M1; and 5 received the sequence M3, M2,<br />

M1. For trials with M1 and M2, FertiCare device(s) were set at 2.5 mm<br />

amplitude and 100 Hz. For trials with M3, stimulation parameters of the Viberect<br />

X3 were preset by the manufacturer and were not adjustable. For all<br />

methods, PVS was delivered in 2 minute increments with inspection of the<br />

penile skin between increments. PVS was stopped if ejaculation occurred<br />

or after 10 minutes of PVS with no ejaculation. Following each PVS trial,<br />

subjects were asked to rate their experience on a questionnaire with scaled<br />

responses.<br />

RESULTS: Please see Table 1. Ejaculation success rates were high for<br />

each method, however, ejaculation latency was significantly longer with<br />

M3 compared with M1 or M2. When analyzing subject responses to survey<br />

questions, there were no significant differences in ratings of M1 compared to<br />

M2. In contrast, M3 was rated lower than M1 and M2 for all survey questions.<br />

These differences were significant for survey questions 1, 2 and 4. Semen<br />

collection was more problematic with M3 versus M1 or M2 due to the configuration<br />

of the Viberect X3 device, which hampered proximity of the specimen<br />

cup to the urethral meatus.<br />

CONCLUSIONS: Based on these findings, our recommended algorithm is<br />

to attempt PVS with one FertiCare device. If that fails, use two FertiCare devices.<br />

Although the Viberect X3 was preferred less by patients, it is a lower<br />

cost alternative that may be suitable for home use by some patients.<br />

Table 1<br />

Success rate (% of men<br />

ejaculating)<br />

Ejaculation latency in<br />

seconds (mean SEM)<br />

Survey Questions:<br />

(Values represent<br />

means SEM)<br />

1. How much did this<br />

method meet your<br />

expectations?<br />

0 ¼ Did not meet<br />

expectations<br />

100 ¼ Met expectations<br />

2. How comfortable<br />

did you feel during<br />

stimulation?<br />

0 ¼ Not comfortable<br />

100 ¼ Very comfortable<br />

3. How comfortable do<br />

you feel about using<br />

this method at home<br />

either by yourself or<br />

with a partner?<br />

0 ¼ Not comfortable<br />

100 ¼ Very comfortable<br />

4. Would you recommend<br />

this method to other<br />

men with spinal<br />

cord injury?<br />

0 ¼ Would not recommend<br />

100 ¼ Would recommend<br />

One FertiCare<br />

Device (M1)<br />

Two Ferticare<br />

Devices (M2)<br />

87 100 87<br />

Viberect X3<br />

Device (M3)<br />

29.6 5.0 32.2 4.4 56.8 5.0 a<br />

74.9 6.1 77.2 6.1 52.8 5.9 b<br />

82.2 5.7 82.2 5.7 64.3 5.5 c<br />

79.1 6.5 71.5 6.5 67.5 6.2<br />

91.7 7.0 83.6 7.0 68.4 6.7 d<br />

SEM ¼ standard error of the mean<br />

Unless indicated by a superscripted letter, comparisons between groups<br />

were not significant.<br />

a Significantly different from M1 (p ¼ 0.0006) and M2 (p ¼ 0.001)<br />

b Significantly different from M1 (p ¼ 0.01) and M2 (p ¼ 0.03)<br />

c Significantly different from M1 (p ¼ 0.03) and M2 (p ¼ 0.03)<br />

d Significantly different from M1 (p ¼ 0.02)<br />

e8 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


O-20 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:30 AM<br />

REDEFINING AND CLARIFYING THE RELATIONSHIP<br />

BETWEEN TOTAL MOTILE SPERM COUNTS (TMSC) AND<br />

INTRAUTERINE INSEMINATION (IUI) PREGNANCY<br />

RATES. R. S. Rubin, a K. S. Richter, b F. Naeemi, b S. Shipley, b<br />

P. R. Shin, c J. E. O’Brien. b a Urology, Medstar Georgetown University<br />

Hospital, Washington, DC; b Shady Grove Fertility Reproductive Science<br />

Center, Rockville, MD; c Shady Grove Fertility Center, Washington, DC.<br />

OBJECTIVE: The value of post-wash TMSC for predicting IUI outcomes<br />

is not well defined. Limitations of previous reports include small sample<br />

sizes and attempts to identify clinically meaningful thresholds as opposed<br />

to gradual trends across a continuum of TMSC. To clarify the relationship between<br />

post-wash TMSC and IUI outcomes we evaluated a large single institution<br />

sample.<br />

DESIGN: Retrospective review<br />

MATERIALS AND METHODS: All stimulated clomiphene citrate, letrozole,<br />

and/or injectable gonadotropin IUI cycles performed at a single institution<br />

from 2004 to 2014 were reviewed, excluding double insemination.<br />

Generalized estimating equations (GEE) analysis was used to account for<br />

multiple cycles by individual patients and to adjust for age, BMI, infertility<br />

diagnoses, stimulation protocol, and pre-insemination endometrial thickness,<br />

serum estradiol, and numbers of follicles R14 mm.<br />

RESULTS: 47,553 insemination cycles were available to evaluate the relationship<br />

between TMSC and clinical pregnancy (defined as ultrasound<br />

confirmation of an intrauterine gestational sac). Pregnancy rates were highest<br />

with a clear threshold noted at R9 million TMSC. Pregnancy rates declined<br />

gradually as TMSC decreased.<br />

Complete data for the adjusted GEE analysis were available for 40,655 cycles.<br />

Adjusted GEE analysis among cycles with R9 million TMSC<br />

confirmed that TMSC in this range was unrelated to pregnancy (p¼0.47).<br />

Conversely, TMSC was highly predictive of pregnancy (Wald c 2 ¼120) in<br />

adjusted GEE analysis among cycles with


(not gender specific). Among all cancer centers, only 60% include information<br />

on fertility preservation specifically directed toward men, such as sperm<br />

cryopreservation. Survivorship information on family building after cancer<br />

was available on 32% of cancer center web sites. State population density<br />

had no significant effect on whether a web site included risks of treatment<br />

on fertility (p¼0.90) or information on fertility preservation (p¼0.29).<br />

CONCLUSIONS: Forty percent of NCI designated cancer center web sites<br />

do not discuss options for male fertility preservation, and over one-third<br />

make no mention of the ramifications of cancer treatment on male fertility.<br />

Given the increasing recognition of the importance of oncofertility in cancer<br />

survivorship, more education should be available about options for fertility<br />

preservation, particularly among men.<br />

References: [1] There are 68 NCI designated cancer centers, of which 61<br />

engage in clinical activity. Cleveland Clinic Taussig Cancer Institute was<br />

included as a separate data point although it is also a member of the Case<br />

Comprehensive Cancer Center, for final n¼62.<br />

O-23 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:15 PM<br />

BLOCKADE OF PHOSPHATIDYLSERINE ON MURINE SPERM<br />

INHIBITS IN VITRO FERTILIZATION. L. Smith-Harrison, a<br />

K. Wheeler, b C. Barberry, a W. Xu, b R. Smith, c J. Lysiak. b a Urology,<br />

University of Virginia, Charlottesville, VA; b University of Virginia, Charlottesville,<br />

VA; c UVA Urology, Charlottesville, VA.<br />

OBJECTIVE: Phosphatidylserine (PtdSer) expression is not only a marker<br />

for apoptotic cells but also a ligand for engulfment receptors essential for the<br />

recognition and efficient removal of these dead cells. The engulfment of the<br />

dead/apoptotic cells by phagocytes or neighboring cells is actually the final<br />

step of apoptotic cell death. Our lab has recently shown that efficient clearance<br />

of apoptotic germ cells is essential for normal spermatogenesis. Studies<br />

in the boar, ram, mouse, rat, and human have found mature sperm positive for<br />

the expression of PtdSer. PtdSer positive sperm were found to be functional<br />

and may even positively correlate with embryo formation in humans. In this<br />

current study we test the novel hypothesis that egg-sperm interactions during<br />

fertilization may hijack the machinery normally used for engulfing apoptotic<br />

cells, with the sperm mimicking an apoptotic cell and the oocyte a phagocyte.<br />

DESIGN: In vitro fertilization (IVF) experiments were performed using<br />

oocytes from super-ovulated mice and sperm isolated from the murine cauda<br />

epididymis.<br />

MATERIALS AND METHODS: Expression of PtdSer on sperm was performed<br />

with fluorescenctly tagged Annexin V (AnV). Blockade of PtdSer on<br />

sperm was accomplished with unlabeled AnV.<br />

RESULTS: The percentage of fertilized embryos after PtdSer blockade<br />

with AnV was 57.6% versus 87.2% in controls (p < 0.001). Sperm motility<br />

was not affected by AnV. Additionally, immunofluorescence identified<br />

PtdSer on the midpiece and acrosome regions of sperm from the cauda but<br />

not caput epididymis.<br />

CONCLUSIONS: The exact mechanisms involved in sperm entry into the<br />

oocyte remain elusive. These results suggest that PtdSer on sperm play an<br />

important role in fertilization and supports our hypothesis that ligands and<br />

receptors involved in apoptotic cell engulfment may be used during sperm<br />

oocyte interactions. Whether oocytes have receptors for PtdSer, or whether<br />

PtdSer interacts with cumulus cells or the zona pellucida remains to be elucidated.<br />

The results of this study will have potential impacts on in vitro fertilization<br />

and contraceptive technologies.<br />

O-24 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:30 PM<br />

PROTEOMIC PATHWAY IN SEMINAL PLASMA OF MEN WITH<br />

SPINAL CORD INJURY (SCI) BEFORE AND AFTER ORAL<br />

ADMINISTRATION OF PROBENECID. M. Camargo, a E. Ibrahim, b<br />

T. C. Aballa, b V. Carvalho, c K. Cardozo, c C. M. Lynne, b,d R. Bertolla, a<br />

N. Brackett. b,d a Department of Surgery, Division of Urology, Human Reproduction<br />

Section, Sao Paulo Federal University, Sao Paulo, Brazil; b Miami<br />

Project To Cure Paralysis, University of Miami, Miami, FL; c Fleury Group,<br />

Sao Paulo, Brazil; d Urology, University of Miami, Miller School of Medicine,<br />

Miami, FL.<br />

OBJECTIVE: Previous results demonstrate that oral administration of<br />

probenecid increases sperm motility in men with SCI (J Urol 193(4S)<br />

e344-e345). Our objective was to evaluate the proteome of seminal plasma<br />

in SCI patients before and after treatment with oral probenecid in order to<br />

demonstrate changes in associated pathways.<br />

DESIGN: Prospective Study.<br />

MATERIALS AND METHODS: This study included 10 men with SCI<br />

who ejaculated regularly by penile vibratory stimulation or electroejaculation.<br />

Probenecid tablets (Watson Pharma Inc., Corona, CA) were administered<br />

in two phases. In Phase 1, subjects received 250 mg orally twice a<br />

day for 1 week. In Phase 2, subjects who completed Phase 1 with no complications<br />

or side effects were administered 500 mg orally twice a day for 3<br />

weeks. Semen analyses were performed at two time points: Pre-treatment<br />

(Pre-Rx group, 1-2 days before Phase 1) and Post-treatment (Post-Rx group,<br />

within 4 weeks after completion of Phase 2). Seminal plasma proteomics was<br />

performed by a label-free quantitative approach, in which 50 mg of total proteins<br />

were pooled, digested into tryptic peptides and analyzed by liquid chromatography<br />

followed by tandem mass spectrometry (LC-MS/MS). Each<br />

sample was run in triplicate. Significant proteins (Student’s t-test) were<br />

used for functional enrichment analysis performed using the Cytoscape platform.<br />

RESULTS: In total, 806 proteins were identified, of which 695 were not<br />

significantly changed in both groups. Thirteen proteins were down-regulated<br />

and 65 other proteins were exclusive to the Pre-Rx group. Five proteins were<br />

up-regulated and 28 other proteins were exclusive to the Post-Rx group. The<br />

Pre-Rx group expressed increased immunologic functions, such as antigen<br />

processing and preservation of peptide antigen via MHC class I. Catabolic<br />

activities such as amino acid degradation, lysosome organization, as well<br />

as the pentose-phosphate shunt oxidative functions were also observed in<br />

the Pre-Rx group. The Post-Rx group expressed enriched energy production<br />

pathways (glycine-tRNA ligase activity) as well as 4-hydroxyproline catabolic<br />

processes.<br />

CONCLUSIONS: Oxidative and immune functions, as well as catabolic<br />

processes were enhanced in patients with SCI. Treatment with oral probenecid<br />

enhanced the energy-production pathways that play an important role<br />

in the biological process of improving sperm motility in men with SCI.<br />

Supported by: Craig H. Neilsen Foundation #224598.<br />

MALE REPRODUCTION AND UROLOGY: CLINICAL 1<br />

O-25 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:15 AM<br />

SEMINAL VESICLE SPERM ASPIRATION FROM WOUNDED<br />

WARRIORS: A CASE SERIES. M. W. Healy, a B. Yauger, a<br />

A. N. James, b R. Dean. c a Department of Obstetrics and Gynecology, Division<br />

of Reproductive Endocrinology and Infertility, Walter Reed National<br />

Military Medical Center, Bethesda, MD; b ART Institute of Washington,<br />

Inc, Bethesda, MD; c Department of Urology, Division of Andrology, Walter<br />

Reed National Military Medical Center, Bethesda, MD.<br />

OBJECTIVE: There has been an increase in blast injuries from dismounted<br />

Improvised Explosive Devices (IEDs) encountered on the battlefield<br />

(i.e. ambulatory soldiers not within the protection of a vehicle).<br />

Associated injuries involve the lower extremities, pelvis, and perineum.<br />

Thus, options to retrieve sperm may be limited due to type and extent of<br />

injury. An alternative technique to the standard Testicular Sperm Extraction<br />

or Microsurgical Epididymal Sperm Aspiration is seminal vesicle sperm<br />

aspiration (SVSA), described in cases of ejaculatory duct obstructions or primary<br />

anorgasmy. Given the type of pelvic injuries seen in these wounded<br />

warriors, we assessed SVSA as an option to retrieve sperm with the goal<br />

of cryopreservation for future use in In Vitro Fertilization (IVF) with intracytoplasmic<br />

sperm injection (ICSI).<br />

DESIGN: Retrospective case series.<br />

MATERIALS AND METHODS: Wounded warriors who underwent<br />

SVSA at Walter Reed National Military Medical Center between 2012-<br />

2014 were included. Patient age, type and date of injury, date of SVSA, specimen<br />

fluid analysis, post-thaw analysis, fertilization rates during IVF/ICSI,<br />

pregnancy rates, and live birth outcomes were evaluated.<br />

RESULTS: Six patients who presented with lower extremity, pelvic, and<br />

perineal injuries resulting from dismounted IEDs underwent SVSA within<br />

5-12 days of the initial injury. Sperm retrieved were analyzed (volume:<br />

0.4mL to 1.8mL, concentration: 40K to 2,200K, motility: 0% to 5%). Sperm<br />

was washed and cryopreserved. In two cases, IVF/ICSI cycles were performed<br />

using the frozen samples. Sperm retrieval for these cases occurred<br />

5 and 9 days after the initial injuries. In one couple, 13 mature oocytes underwent<br />

ICSI with morphologically normal sperm that responded to the touch<br />

test with a fertilization rate of 38%. One grade V embryo was transferred<br />

on day 4 with a negative pregnancy test. The second couple underwent<br />

two IVF/ICSI cycles. In the first cycle, 9 mature oocytes underwent ICSI<br />

with 4 sperm that responded to Pentoxifylline and 5 sperm that responded<br />

e10 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


to touch test notable for a 44% fertilization rate. On day 5, one expanded blastocyst,<br />

B/B, was transferred with a negative pregnancy test. In the second cycle,<br />

<strong>17</strong> mature oocytes underwent ICSI with sperm that responded to<br />

Pentoxifylline, with a 47% fertilization rate. Two B/B expanded blastocysts<br />

were cryopreserved due to concern for ovarian hyperstimulation in the female<br />

partner. Among all three cycles, fertilization rate was 43.5%.<br />

CONCLUSIONS: SVSA is a reasonable option to retrieve sperm in<br />

wounded warriors or trauma patients with pelvic or perineal injuries.<br />

Supported by: This research was Supported, in part, by the Program in<br />

Reproductive and Adult Endocrinology, NICHD, NIH.<br />

O-26 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:30 AM<br />

EVALUATION OF REPRODUCTIVE SYSTEM ANATOMY AND<br />

GONADAL FUNCTION IN PATIENTS WITH PRUNE-BELLY<br />

SYNDROME. M. Cocuzza, a B. C. Tiseo, a R. Park, a G. P. Padovani, a<br />

R. H. Baroni, b A. Tavares, a F. T. Denes, a M. Srougi. a a Division of Urology,<br />

Hospital das Clinicas, University of Sao Paulo Medical School, Sao Paulo,<br />

Brazil; b Radiology Institute, Hospital das Clinicas, University of Sao Paulo<br />

Medical School, Sao Paulo, Brazil.<br />

OBJECTIVE: To report the first series of Prune-belly syndrome (PBS) patients<br />

that were submitted to surgical treatment during early childhood<br />

focusing in the evaluation of reproductive system anatomy and gonadal function.<br />

To date, PBS patients were diagnosed as infertile and fertility was often<br />

given lowest priority. Their infertility is multifactorial and is probably related<br />

to the undescended testes, prostatic hypoplasia and retrograde ejaculation.<br />

There are no documented cases of unassisted paternity; few successful pregnancies<br />

have been reported through assisted reproduction, but all using<br />

epididymal or testicular sperm extraction.<br />

DESIGN: Series of case analysis.<br />

MATERIALS AND METHODS: We accessed 30 patients with PBS from<br />

our service that were submitted to any surgical procedure during their<br />

childhood and now are at least 14 years old. Patient records were accessed<br />

to identify age at orchidopexy. All patients were submitted to pelvic MRI in<br />

order to evaluate anatomical findings of reproductive system, including<br />

prostate size, characteristics of seminal vesicles and vas deferentia. Serum<br />

levels of LH, FSH, testosterone and also creatinine were evaluated. Sperm<br />

analysis was conducted and analysis of the urine after masturbation when<br />

needed.<br />

RESULTS: We contacted 18 patients. Of those, 15 had reliable data from<br />

patient records including physical examination and hormone profile. The<br />

average of age of this group at evaluation is 18.2 years. The average age at<br />

orchidopexy was <strong>17</strong> months and an average follow-up was <strong>17</strong>.4 years. All<br />

had normal physical development and stable anatomy and function of the reconstructed<br />

urinary tract with an average creatinine of 1.64 mg/dL. Fourteen<br />

patients had both testes in scrotum, and the testicular volume varied from<br />

2.1cc to 9.4cc, averaging 6.9cc. Eight patients collected semen with a sperm<br />

count of 5.07 million/mL, whereas motile sperm was found in 62.5%<br />

including three in the ejaculate and two in urine after masturbation. Average<br />

hormones levels were LH: 5.3 mg/dL, FSH: 6.9 mg/dL, total testosterone:<br />

531.2 mg/dL and free testosterone: 450.6 mg/dL. MRI revealed prostates hypoplastic<br />

in 55.6% and absent in 22.3%, while 55.6% had absence of at least<br />

one of the seminal vesicles. There was no vasal abnormality.<br />

CONCLUSIONS: Our data enlightens findings in patients with PBS that<br />

were not described yet. A high prevalence of hypoplastic or absent prostate<br />

and seminal vesicle abnormalities was observed in our patients; those findings<br />

may represent their main cause of infertility. Probably, early orchidopexy<br />

increases testicular function salvage preserving their fertility<br />

potential patients leading to the finding of motile sperm in the ejaculate or<br />

in urine after masturbation. The next step is to provide a better understanding<br />

of their fertility potential improving quality of life.<br />

O-27 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:45 AM<br />

THE EFFECTS OF TRANSVERSE AND LONGITUDINAL<br />

INCISION OF TUNICA ALBUGINEA IN MICRODISSECTION<br />

TESTICULAR SPERM EXTRACTION. T. Ishikawa, K. Yamaguchi,<br />

R. Nishiyama, Y. Takaya, K. Kitaya, H. Matsubayashi. Reproduction Clinic<br />

Osaka, Osaka, Japan.<br />

OBJECTIVE: The optimal technique of sperm extraction would be<br />

minimally invasive and avoid destruction of testicular function without<br />

compromising the chance of retrieval adequate numbers of spermatozoa to<br />

perform ICSI. In general, the tunica albuginea should be opened on an equatorial<br />

plane because antimesenteric incision increases the risk of testicular<br />

devascularization and may adversely affect access to the seminiferous tubules.<br />

To demonstrate the differences of sperm retrieval rate (SRR) and postoperative<br />

complication between transverse or longitudinal incision of tunica<br />

albuginea in microdissection testicular sperm extraction (micro TESE).<br />

DESIGN: A retrospective study.<br />

MATERIALS AND METHODS: A total of 1080 patients (including 970<br />

46XY males with NOA and 110 Klinefelter syndrome (KS) patients) underwent<br />

micro TESE were subjected to sperm retrieval procedures. All operations<br />

were performed by one surgeon (TI). Karyotyping test was<br />

performed on a sample of blood to all patients. SRR and postoperative<br />

complication were analyzed in the 960 patients with transverse incision<br />

(group T) and the 120 patients with longitudinal incision (group L). Ninety<br />

and 20 non-mosaic KS cases each were included in group T and group L,<br />

respectively.<br />

RESULTS: Testicular sperm were successfully retrieved by micro-TESE<br />

in 415 of 960 (43.2%) NOA (including 47 of 90 KS: 52.2%) and 56 of 120<br />

(46.7%) NOA (including 10 of 20 KS: 50%) patients with group T and group<br />

L incision in micro TESE, respectively. There was no significant difference<br />

of sperm retrieval rate with either approach. No patients had postoperative<br />

complications such as major hematoma or leakage of seminiferous tubules<br />

through this series. The hospital stay was all the same in both groups<br />

(a-day-surgery). For the 46XY males with NOA patients, after micro-<br />

TESE, the testosterone level did not drop significantly in both group T and<br />

group L. In addition, there were no significant differences in the levels of<br />

testosterone between group T and group L for the 46XY males with NOA<br />

patients. Of the KS patients who underwent micro TESE, the mean serum<br />

testosterone level showed an average decline of 30-35% from baseline<br />

when assessed at 1, 3, and 6 months after micro-TESE, but no significant differences<br />

were also shown in the levels of testosterone between group T and<br />

group L.<br />

CONCLUSIONS: There were no significant differences between transverse<br />

or longitudinal incision for SRR and postoperative complication in<br />

micro TESE. We should take care of the hypogonadism in KS patients after<br />

even microdissection procedure by either transverse or longitudinal incision.<br />

O-28 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:00 PM<br />

COMPARISON OF SEMEN QUALITY BETWEEN UNIVERSITY<br />

AND PRIVATE CLINIC LABORATORIES. O. Khan, a C. F. Jensen, b<br />

J. Sonksen, b M. Fode, b T. Shah, c D. A. Ohl. a a Urology, University of Michigan,<br />

Ann Arbor, MI;<br />

b Urology, Herlev University Hospital, Herlev,<br />

Denmark; c Obstetrics and Gynecology, University of Michigan, Ann Arbor,<br />

MI.<br />

OBJECTIVE: Obtaining an ejaculate and performing a semen analysis<br />

(SA) is an essential first step in evaluating the infertile man. Based on the<br />

SA the physician determines the nature of further evaluation and treatment.<br />

Multiple treatment options are available and range from simple counseling to<br />

the more complicated and expensive assisted reproduction techniques<br />

(ART). The objective of this study was to investigate inter-laboratory variation<br />

in semen quality between private ART laboratories and university-based<br />

ART laboratories respectively.<br />

DESIGN: Clinical retrospective study.<br />

MATERIALS AND METHODS: IRB approval was obtained to retrospectively<br />

evaluate patients who had undergone a SA at both the University<br />

of Michigan Center For Reproductive Medicine and private practice IVF<br />

clinics. When more than one SA was available from both clinics the SA<br />

with the highest total motile sperm was selected for analysis. Major semen<br />

parameters from both SA’s were compared using Wilcoxon signed-rank<br />

test.<br />

RESULTS: A total of 20 men aged 35 6 (mean SD) years were<br />

included in the study. Table 1 shows comparisons of the major semen parameters<br />

obtained at the two laboratories. Morphology was reported significantly<br />

lower at the private clinics.<br />

CONCLUSIONS: In this small series, morphology was significantly lower<br />

in semen analyses performed at private IVF clinics. Since sperm morphology<br />

under 5% is commonly used to recommend IVF with ISCI, underestimation<br />

of sperm morphology at private clinics may lead to over-utilization of high<br />

level assisted reproductive techniques at these sites.<br />

FERTILITY & STERILITY Ò<br />

e11


TABLE 1.<br />

University Based ART Laboratory<br />

Private ART Laboratory<br />

25th percentile Median 75th percentile 25th percentile Median 75th percentile<br />

p-value<br />

Volume (mL)(n¼19) 1.7 3.3 4.1 1.75 3.2 4.5 0.43<br />

Motility (%)(n¼19) 40 40 50 16.9 56 69.2 0.07<br />

Concentration (million/mL)(n¼20) 0.45 10 44 0.89 7 26 0.30<br />

Total Motile Sperm (N/ejaculate)(n¼19) 0.3 8 54 0.35 9 31 0.44<br />

Morphology (%)(n¼9) 6.5 7 10 2 3 6 0.01<br />

O-29 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:15 PM<br />

CLOMID HAS GOTA BRIGHT SIDE AND A DARK SIDE. WHAT DO<br />

WE REALLY KNOWAFTER ALLTHESE YEARS? EVIDENCE FOR<br />

TOXICITY. R. Wiehle G. K. Fontenot. Research and Development,<br />

Repros Therapeutics Inc, The Woodlands, TX.<br />

OBJECTIVE: Clomid has been approved in women with PCOS/ovulation<br />

induction. In men it is used off-label to increase testosterone or to relieve<br />

symptoms of androgen abuse. Clomid is a mixture of two isomers: zuclomiphene<br />

and enclomiphene. The former is a weak estrogen agonist and the latter<br />

is a strong estrogen antagonist. Repros Therapeutics is developing enclomiphene<br />

for elevating endogenous testosterone in men with secondary hypogonadism.<br />

DESIGN: We have administered each isomer of clomid to mice, both<br />

chronically and acutely, in an attempt to determine the relative differences.<br />

MATERIALS AND METHODS: Enclomiphene and zuclomiphene were<br />

separated from clomiphene as pure mono-isomers. Mice were administered<br />

each isomer chronically (90 days) then tissues were examined histologic.<br />

In a mouse ADME, pigmented mice were acutely administered each of the<br />

14C-labelled isomers. Tissue and fluid distribution were followed up to 45<br />

days.<br />

RESULTS: In animals chronically administered each isomer, zuclomiphene<br />

had pernicious effects on the male reproductive tract (testes, epididymis,<br />

and seminal vesicles) and the kidney. Significant reduction in size of<br />

testes with testicular degeneration was seen, including Leydig cell loss<br />

with absence of sperm in the seminiferous tubules and reduction in size of<br />

the epididymis, seminal vesicles and kidneys. In the ADME study, for 47 tissues<br />

assessed, 97.5% of the 14C-enclomiphene seen at 4 hours was lost by 24<br />

hours. In contrast, only 40.8% of the 14C-zuclomiphene seen at 4 hours was<br />

lost. Zuclomiphene was retained selectively. Among those that accumulated<br />

zuclomiphene that seen in the blood were the pigmented portions of the eye,<br />

the brain, adrenal gland, the kidneys and the testes.<br />

CONCLUSIONS: We infer that zuclomiphene accumulates in excess over<br />

enclomiphene. Human studies have suggested this as well. We concede that<br />

the antagonist effects of enclomiphene can overwhelm effects of zuclomiphene<br />

when used chronically, however the extreme high build-up of the<br />

agonist isomer may have lasting effects. These results justify the case for a<br />

monoisomeric preparation and the development of Enclomiphene citrate,<br />

for clinical use in men with secondary hypogonadism to increase testosterone.<br />

It is interesting to speculate whether clomiphene citrate would have<br />

been granted FDA approval given the differences between its constitutive<br />

isomers.<br />

Supported by: Repros Therapeutics Inc.<br />

O-30 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:30 PM<br />

THE EFFECT OF CLOMIPHENE CITRATE IN THE TREATMENT<br />

OF SUBFERTILE MALES WITH BODY MASS INDEX (BMI) ‡ 25<br />

KG/M 2 . B. Patel, a T. Shah, a D. Shin. b a Urology, Rutgers New Jersey<br />

Medical School, Newark, NJ; b Urology, Hackensack University Medical<br />

Center, Hackensack, NJ.<br />

OBJECTIVE: Elevated BMI has been shown to have a negative impact on<br />

male fertility. Clomiphene citrate (CC), a selective estrogen receptor modulator,<br />

is often used in the empiric treatment of subfertile males to increase<br />

testosterone levels and improve spermatogenesis. The objective of this study<br />

was to assess the effect of CC in the treatment of subfertile males with<br />

elevated BMI R 25 kg/m2.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Fifty-six subfertile males with BMI<br />

R 25 kg/m2 were treated with CC between 2009 and 2014. Semen analysis<br />

was conducted on all 56 patients at baseline and at minimum, 3 months<br />

follow-up. Fifty of the 56 patients had baseline and 5 month follow-up measurements<br />

of follicle-stimulating hormone (FSH), luteinizing hormone (LH),<br />

total testosterone (TT), bioavailable testosterone (BT) and estradiol (E)<br />

levels. Pregnancy status was obtained when possible. Paired t-test was<br />

used to compare baseline and follow-up hormonal profiles and semen analyses.<br />

RESULTS: Significant increase in sperm concentration, from 19.0 2.7<br />

M to 28.2 4.0 M (p


MATERIALS AND METHODS: Patients with normal ovarian reserve<br />

% 42 years of age were recruited. Following routine IVF care, trophectoderm<br />

biopsy was performed. Embryos were subsequently selected for transfer per<br />

routine. No NGS CCS analysis was done prior to transfer. A novel targeted<br />

amplification method of NGS based CCS which does not require whole<br />

genome amplification was then performed. The outcome for each transferred<br />

embryo was compared to the NGS screening result to determine the<br />

predictive value of that result. In the case of a two embryo transfer, DNA<br />

fingerprinting was utilized to ensure that the embryo responsible for the pregnancy<br />

was identified correctly. Implantation rates were compared between<br />

embryos designated euploid, aneuploid, and for the population as a whole.<br />

RESULTS: 1<strong>17</strong> patients had 187 blastocysts transferred. Of the 41 embryos<br />

assigned an aneuploid karyotype, none sustained implantation yielding<br />

a predictive value of an aneuploid result of 100%. 97 of 146 embryos designated<br />

as euploid implanted and progressed to delivery yielding a predictive<br />

value for a euploid result of 66.0%. No embryo designated as euploid subsequently<br />

developed into an aneuploid gestation. Euploid embryos had a higher<br />

sustained implantation rate than the population as a whole (50.8%,<br />

p


OBJECTIVE: The objective of this study was twofold; first we wanted to<br />

determine the number of blastocysts needed at a certain age to produce at<br />

least one euploid blastocyst (>95%) through PGS using whole comprehensive<br />

chromosome screening. Secondly, we evaluated the chance of obtaining<br />

a euploid embryo in the next cycle after obtaining all aneuploid embryos.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: 10,852 cycles of PGS were performed<br />

using blastocyst biopsy and analyzed by array Comparative Genomic<br />

Hybridization (aCGH). A total of 58,798 embryos were analyzed. IVF<br />

cycles were performed at multiple fertility centers and biopsies sent to a<br />

reference laboratory for analysis. Some poor prognosis patients accumulated<br />

embryos from several cycles and then performed PGS (‘‘embryo<br />

banking’’).<br />

RESULTS: There was no correlation between cohort size of blastocyst<br />

analyzed and euploidy rates, but there were significant differences between<br />

euploidy rates and maternal age (p< 0.001). Euploidy rates 65%, 56%,<br />

46%, 33%, 19% and 13% for egg donors were observed for ages 42 years<br />

old, respectively.The cumulative number of blastocyst needed to produce<br />

at least one euploid blastocyst with 95% or higher chance was calculated<br />

from single cycles or embryo banking cycles from the same patient and is<br />

shown in the below table.<br />

% of patients with normal blastocysts<br />

#of<br />

embryos<br />

egg<br />

donors<br />

42 years old, respectively, produced<br />

no euploid embryos in the first cycle. Of those with no euploid embryos<br />

in the first cycle, 38% (41-42 years old) and 25% (> 42 yeas old)<br />

of those that produced <strong>17</strong> embryos produced euploid embryos in successive<br />

cycles.<br />

CONCLUSIONS: This is the largest study so far on the relationship between<br />

aneuploidy and maternal age and cohort size. In women, 35 and<br />

older more than 50% of embryos are chromosomally abnormal, with those<br />

41 and older needing 18 or more embryos to secure one euploid one. However,<br />

a cycle with no euploid embryos does not preclude finding euploid<br />

ones in the next cycle(s) provided those cycles are not far apart and<br />

enough embryos are generated. This data will help couples assess their<br />

odds at producing a viable pregnancy and how many cycles and embryos<br />

it might involve.<br />

O-35 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:15 PM<br />

41-42<br />

years<br />

>42 years<br />

1-3 83% 80% 71% 57% 36% 22%<br />

4-6 97% 95% 92% 82% 59% 43%<br />

7-10 99% 98% 96% 89% 74% 50%<br />

10-<strong>17</strong> 100% 99% 99% 97% 88% banked 64% banked<br />

><strong>17</strong> 100% 100% 100% 99% 97% banked 87% banked<br />

WHY DO ARRAY-CGH (ACGH) EUPLOID EMBRYOS MISCARRY?<br />

REANALYSIS BY NGS REVEALS UNDETECTED ABNORMALITIES<br />

WHICH WOULD HAVE PREVENTED 56% OF THE<br />

MISCARRIAGES. J. Grifo, a P. Colls, b L. Ribustello, b T. Escudero, b<br />

E. Liu, b S. Munne. b a NYU Langone Medical Center, NY, NY; b Reprogenetics,<br />

Livingston, NJ.<br />

OBJECTIVE: aCGH, qPCR and SNP arrays have been extensively used<br />

for preimplantation genetic screening (PGS) but a more robust and sensitive<br />

technique, Next Generation Sequencing (NGS) can detect mosaicism and<br />

polyploidy more effectively. The objective of this study was to determine<br />

if miscarriages occurring after PGS were due to abnormalities not detected<br />

by aCGH.<br />

DESIGN: Retrospective Analysis with prospective sample re-analysis by<br />

aCGH and NGS<br />

MATERIALS AND METHODS: A total of 43 miscarriages were reported<br />

upon follow up of spontaneous pregnancy loss resulting from<br />

2442 cycles with euploid blastocysts obtained from PGS by blastocyst biopsy<br />

and array CGH. Karyotype analysis of products of conception (POC)<br />

in 18 samples revealed aneuploidy (full aneuploidy, mosaic aneuploidy<br />

and partial aneuploidy) which was inconsistent with the aCGH analysis<br />

from the blastocyst biopsy and went undetected prior to embryo transfer.<br />

In the remaining 25 samples no POC data was available. Saved amplified<br />

DNA samples from these 43 blastocyst biopsies previously diagnosed as<br />

‘‘euploid’’ by aCGH were re-analyzed by aCGH yielding the same result.<br />

A third aliquot of the same amplified DNA was then analyzed by NGS to<br />

determine if abnormalities not detectable by aCGH were present.<br />

RESULTS: The DNA from the embryos replaced that resulted in pregnancy<br />

and miscarriage were reanalyzed by NGS and the results are shown<br />

in the table below:<br />

NGS Reanalysis of TE Biopsy Specimen<br />

Normal<br />

Triploid<br />

Mosaic<br />

Whole<br />

Aneuploidy<br />

Mosaic<br />

Partial<br />

Aneuploidy<br />

Full<br />

Trisomy<br />

No<br />

Diagnosis<br />

POC Not 11 (44%) 2 (8%) 3 (12%) 7 (28%) 0 2 (8%)<br />

Analyzed<br />

N¼25<br />

POC 6 (33%) 0 9 (50%) 1 (6%) 0 2 (11%)<br />

Aneuploid<br />

N¼18<br />

Total<br />

N¼43<br />

<strong>17</strong> (40%) 2 (5%) 12 (28%) 8 (19%) 0 4 (9%)<br />

CONCLUSIONS: NGS is known to be able to detect mosaicism as well as<br />

triploidy, which is also well-known cause of spontaneous pregnancy loss.<br />

Whole and partial mosaicism could also play a role in spontaneous pregnancy<br />

loss depending on the chromosomes involved as well as the percentage of<br />

abnormal cells in the embryo. Of 16 embryos diagnosed euploid by aCGH<br />

that resulted in an aneuploid loss 10 (62%) were diagnosed as abnormal by<br />

NGS (2 not analyzable). Similarly, of the 23 embryos diagnosed euploid<br />

by aCGH that resulted in loss but not diagnosed by POC analysis, 12<br />

(52%) were abnormal by NGS (2 not analyzable). The use of NGS would<br />

have avoided 56%(22/39) of the pregnancy losses resulting from aCGH<br />

tested embryos. Likely, mosaicism and triploidy accounted for most of these<br />

losses and NGS is more sensitive at detecting them. These data support the<br />

notion that NGS is the most powerful technique for PGS analysis in predicting<br />

euploid outcome but will not predict all euploid losses. It reduces the<br />

miscarriage rate by more than 50% over aCGH and will significantly improve<br />

outcomes.<br />

O-36 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:30 PM<br />

GENETIC CARRIER SCREENING IN AN EGG DONOR<br />

PROGRAM. A. Quinteiro Retamar, a C. M. Borghi, a G. Fiszbajn, a<br />

S. Papier, a S. Munne, b J. Hamer, a C. Alvarez Sedo. a a CEGYR (Reproductive<br />

Medicine and Genetics), Buenos Aires, Argentina; b Reprogenetics, Livingston,<br />

NJ.<br />

OBJECTIVE: With the emergence of genomics platforms and the possibility<br />

of studying multiple diseases in only one study and the lack of local updated<br />

data, raised the necessity to establish the prevalence of genetic carriers<br />

(250 diseases) in our egg donor population, in order to improve our egg donation<br />

program.<br />

DESIGN: Retrospective prevalence study.<br />

MATERIALS AND METHODS: Three hundred two oocyte donors were<br />

included (<strong>21</strong>-33 y/o), from December 2013-March <strong>2015</strong>. All donor signed an<br />

informed consent to participate in this study. Donors were included in this<br />

study after fulfilling the egg donation program inclusion criteria (antral follicle<br />

count >16, negative serology, psychological counseling, genetic counseling,<br />

and normal karyotype). All patients were evaluated by a clinical<br />

geneticist, in order to detect relevant family history, before and after the<br />

study.A blood sample was taken, and DNA was extracted in our molecular<br />

biology laboratory. Dried DNA was sent to Recombine Laboratory (Livingston,<br />

NJ, USA). A genomics platform was used to detect more than <strong>17</strong>00 mutations<br />

that correspond to 250 autosomal recessive diseases by microarray<br />

SNPs array technique.<br />

RESULTS: Considering the 250 tested autosomal recessive diseases, our<br />

donors have mutations for 35 (14%), 74.3% (26) of the latter have a high<br />

impact over life expectancy and quality of life, and 25.7% (9) have a moderate<br />

impact.Among the high impact diseases, 61.54% (16) may be subject to<br />

medical treatment, but the 38.46% (10) do not. Cystic Fibrosis (1:19), the<br />

nonsyndromic hearing loss and deafness: GJB2-Related (1:50) and Biotinidase<br />

Deficiency (1:50) were the most prevalent high impact diseases<br />

with medical treatment. Spinal Muscular Atrophy: SMN1 Linked (1:23)<br />

was the most prevalent disease in the group without treatment.Moderate<br />

e14 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


impact diseases were the least present. Within this group, 44.4%(4) have not<br />

treatment and 55.6%(5) have a possible treatment. The <strong>21</strong>-Hydroxylase-<br />

Deficient Congenital Nonclassical Adrenal Hyperplasia (1:9), Familial Mediterranean<br />

Fever (1:100) and Pseudocholinesterase Deficiency (1:50) were<br />

the most prevalent detected diseases.<br />

CONCLUSIONS: It is clear that the screening of genetic diseases<br />

for oocyte donors is part of the necessary studies to reduce the risk of children<br />

born from this kind of treatment. We can infer that the genetic screening of<br />

recessive mutations in people who donate their gametes will further reduce<br />

the risk of certain genetic diseases transmission.<br />

OUTCOME PREDICTORS - CLINICAL: ART 1<br />

O-37 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:15 AM<br />

MATERNAL PREGNANCY AND BIRTH COMPLICATIONS BY<br />

FERTILITY STATUS: THE MASSACHUSETTS OUTCOMES STUDY<br />

OF ASSISTED REPRODUCTIVE TECHNOLOGIES. B. Luke, a<br />

D. Gopal, b J. E. Stern, c E. Declercq, b L. Hoang, b M. Kotelchuck, d<br />

H. Diop. e a Obstetrics, Gynecology, and Reproductive Biology, Michigan<br />

State University, East Lansing, MI; b Community Health Sciences, Boston<br />

University, Boston, MA; c Geisel School of Medicine at Dartmouth, Lebanon,<br />

NH; d MGH Center for Child & Adolscent Health Research and Policy, Mass-<br />

General Hospital for Children, Boston, MA; e Bureau of Family Health and<br />

Nutrition, Massachusetts Department of Public Health, Boston, MA.<br />

OBJECTIVE: To evaluate the effect of maternal fertility status on the risk<br />

of pregnancy and birth complications.<br />

DESIGN: Longitudinal cohort study, linking cycles from the SART<br />

CORS, hospital discharge, and vital records from 2004-2010 in Massachusetts.<br />

MATERIALS AND METHODS: The study included three fertility<br />

groups: women without ART or other infertility treatment (fertile); women<br />

with indicators of subfertility but no ART treatment (subfertile), and women<br />

with ART treatment. The risks of seven adverse outcomes were modeled using<br />

logistic regression, adjusted for parental ages, race/ethnicity, education,<br />

payor status, maternal pre-existing conditions (diabetes and chronic hypertension),<br />

and plurality, and reported as adjusted odds ratios (AORs) and<br />

95% confidence intervals.<br />

Pregnancy Outcomes and Multiple Pregnancy Rate based on Age and Number<br />

of Embryos Transferred<br />

Compliant<br />

(1 ET)<br />


O-39 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:45 AM<br />

ELECTIVE SINGLE EMBRYO TRANSFER (ESET) IS ASSOCIATED<br />

WITH NON-LOW BIRTHWEIGHT TERM SINGLETON<br />

OUTCOMES: AN ANALYSIS OF 263,375 CYCLES. A. K. Styer, a<br />

W. Vitek, b M. S. Christianson, c V. Baker, d A. Armstrong, e N. Santoro, f<br />

B. Luke, g A. J. Polotsky. f a Massachusetts General Hospital/Harvard Medical<br />

School, Boston, MA; b University of Rochester School of Medicine, Rochester,<br />

NY; c Johns Hopkins University School of Medicine, Lutherville, MD;<br />

d Stanford University, Palo Alto, CA; e NICHD, Bethesda, MD; f University<br />

of Colorado, Aurora, CO; g Research, Ann Arbor, MI.<br />

OBJECTIVE: Multiple gestation and its concomitant risk of prematurity is<br />

the major complication of ART. The adoption of ‘‘good perinatal outcome’’<br />

as a more relevant measure of term live birth with a normal birth weight<br />

neonate has been recently advocated. The objective of this study is to evaluate<br />

the likelihood of non-low birth weight (NLBW) term singleton live birth<br />

outcomes with single (SET) and double embryo transfer (DET) in the United<br />

States.<br />

DESIGN: Historical cohort study from the Society for Assisted Reproductive<br />

Technology Clinic Online Reporting System between 2004 and 2012.<br />

MATERIALS AND METHODS: Fresh autologous IVF cycles among<br />

women ages 18-37 years using partner’s semen with either SET or DET<br />

were assessed and categorized into groups with or without cryopreserved<br />

(cryo) supernumerary embryos. SET (cryo) was designated as eSET. SET<br />

(no cryo) were non elective SET due to limited embryo cohort. Gestational<br />

carrier, preimplantation genetic screening or diagnosis, and research cycles<br />

were excluded. Logistic regression (adjusted for age, race/ethnicity, infertility<br />

diagnosis, year of cycle, number of oocytes retrieved, day of transfer)<br />

was used to determine the likelihood of NLBW term singleton live birth<br />

outcome (birthweight R2,500 g and gestation R259 days). Results are reported<br />

as adjusted odds ratios (AORs) and 95% confidence intervals.<br />

RESULTS: 263,375 cycles were analyzed. Compared to eSET, the likelihood<br />

of clinical pregnancy and live birth was significantly increased for DET<br />

(cryo). The odds of NLBW term singleton live birth was significantly lower<br />

in all groups compared to eSET [Table].<br />

Pregnancy Outcomes Following SET and DET<br />

Group<br />

SET(cryo)<br />

[eSET]<br />

n¼20,996<br />

DET(cryo)<br />

n¼1<strong>17</strong>,091<br />

SET(no cryo)<br />

n¼<strong>21</strong>,9<strong>17</strong><br />

DET(no cryo)<br />

n¼103,371<br />

Clinical Pregnancy Live Birth NLBW Term Singleton<br />

% AOR (95% CI) % AOR (95% CI) % AOR (95% CI)<br />

57.6 Reference 49.3 Reference 39.1 Reference<br />

62.5 1.38 (1.34, 1.42) 55.0 1.40 (1.36, 1.44) 25.8 0.55 (0.53, 0.57)<br />

26.6 0.39 (0.37, 0.41) <strong>21</strong>.7 0.41 (0.39, 0.43) <strong>17</strong>.3 0.36 (0.34, 0.38)<br />

47.8 0.86 (0.83, 0.89) 40.9 0.89 (0.86, 0.91) 23.0 0.48 (0.46, 0.49)<br />

CONCLUSIONS: Elective SET increases the likelihood of NLBW by 45-<br />

52% as compared to DET. The use of NLBW term singleton live birth as a<br />

measure of ART success may be useful for the development of clinical strategies<br />

which optimize good perinatal birth outcomes.<br />

Supported by: Clinical Research Scientist Training Program/NICHD,<br />

R25HD075737.<br />

O-40 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:00 PM<br />

THE IMPACT OF INSURANCE MANDATES ON MULTIPLE BIRTH<br />

RATES FOLLOWING IN VITRO FERTILIZATION. M. P. Provost, a<br />

J. S. Yeh, a S. M. Thomas, b W. W. Hurd, a J. L. Eaton. a a Division of Reproductive<br />

Endocrinology & Infertility, Duke University Medical Center,<br />

Durham, NC; b Department of Biostatistics & Bioinformatics, Duke University<br />

Medical Center, Durham, NC.<br />

OBJECTIVE: To examine the relationship between state-mandated insurance<br />

coverage for in vitro fertilization (IVF) and multiple birth rate.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: We utilized the Society for Assisted<br />

Reproductive Technologies-Clinical Outcomes Reporting System (SART-<br />

CORS) database to identify fresh, autologous IVF cycles performed between<br />

2007 and 2011 in women aged 20-42 years. Only first IVF cycles performed<br />

in each woman’s state of residence were included in the analysis. Cycles were<br />

excluded if the indication for IVF was ‘‘non-infertile’’ or ’’preimplantation<br />

genetic diagnosis,’’ if they were performed in a state with only one reporting<br />

clinic, or if embryo transfer occurred on days other than 3 or 5. Among the 40<br />

states with more than one clinic, 6 have legislation requiring insurance<br />

coverage for at least one IVF cycle and were designated ‘‘mandated:’’ CT,<br />

HI, IL, MA, MD, and NJ. The remaining 34 states were designated ‘‘nonmandated.’’<br />

Student’s t-test and the X 2 test were used to analyze continuous<br />

and categorical variables, respectively. P < 0.00019 was considered statistically<br />

significant after Bonferroni adjustment for multiple comparisons. Logistic<br />

regression was performed to model IVF outcomes while controlling<br />

for potential confounders.<br />

RESULTS: Of the <strong>17</strong>3,968 cycles included in the analysis, 45,011 (26%)<br />

were performed in mandated states and 128,957 (74%) were performed in<br />

non-mandated states. The multiple birth rate per live birth was significantly<br />

lower in mandated vs. non-mandated states (29% vs. 33%, P < 0.00001).<br />

This association remained significant after adjusting for potential confounders<br />

(OR 0.87, 95% CI 0.83-0.91). After stratification by SARTage category<br />

and day of transfer, the relationship between mandate status and<br />

multiple birth rate remained statistically significant only in women


OBJECTIVE: ASRM and SART regularly update guidelines on the<br />

maximum number of embryos to transfer per cycle. Providers/patients, however,<br />

often elect to transfer more than the recommended number of embryos<br />

which can lead to an unnecessary increase in the number of multiple pregnancies.<br />

The objective of this study was to determine the prevalence of cleavage<br />

transfer cycles that do not conform to current recommended embryo transfer<br />

(ET) limits and assess the impact of noncompliance on multiple pregnancy<br />

rate (MPR) in first IVF cycles with a favorable prognosis.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: 30,567 first fresh autologous IVF cycles<br />

in women under 43 undergoing cleavage stage embryo transfer (ET) from the<br />

2011-2012 SART registry were stratified into cohorts based on ASRM<br />

defined age bins. Cycles were classified as compliant (C) versus noncompliant<br />

(NC) based on their adherence to the published 2009/2013 guidelines for<br />

first IVF cycles with a favorable prognosis. Main outcomes measured were<br />

percentage of C and NC cycles in each age group as well as the MPR (R2<br />

fetal heart beats on ultrasound) in each of these groups. Secondary outcomes<br />

included clinical pregnancy rate (CPR), live birth rate (LBR), and singletons<br />

(1 fetal heart beat on ultrasound). Data were analyzed using two-sided<br />

Welch’s t-test for each age category and the Benjamini-Hochberg procedure<br />

was used to control the false discovery rate.<br />

RESULTS: The percentage of NC cycles ranged from 2 to 27.4% in<br />

different age groups. Compared to C cycles, noncompliance resulted in<br />

higher MPR in every age group, but statistical significance was reached<br />

only in the two youngest age bins (p


CONCLUSIONS: There is no significant difference of epigenetic profiling<br />

between PBs and corresponding female nuclear genome in the ooplasm at<br />

specific stage from mouse to human, suggesting PB1 and PB2 could be the<br />

optimum candidate of oocyte genome replacement for preventing the transmission<br />

of inherited mitochondrial diseases.<br />

References:<br />

1. Wakayama T, Hayashi Y, Ogura A. Participation of the female pronucleus<br />

derived from the second polar body in full embryonic development<br />

of mice. J Reprod Fertil. 1997, 110(2): 263-266.<br />

2. Tachibana M, Sparman M, Sritanaudomchai H, Ma H, Clepper L,<br />

Woodward J, Li Y, Ramsey C, Kolotushkina O and Mitalipov S. Mitochondrial<br />

gene replacement in primate offspring and embryonic stem<br />

cells. Nature. 2009, 461(7262): 367-372.<br />

3. Wang T, Sha H, Ji D, Zhang HL, Chen D, Cao Y, Zhu J. Polar body<br />

genome transfer for preventing the transmission of inherited mitochondrial<br />

diseases. Cell. 2014, 157(7):1591-1604.<br />

4. Bartholomeusz R. Review of the longevity of the second polar body in<br />

the mouse. Zygote. 2003, 11(1): 23-34.<br />

5. Hino T, Kusakabe, H and Tateno H. Chromosomal stability of second<br />

polar bodies in mouse embryos. J Assist Reprod Genet. 2013, 30(1):<br />

91-98.<br />

6. Fabian D, Cikos S, Rehak P, Koppel J. Do embryonic polar bodies<br />

commit suicide? Zygote. 2012, 22(1): 10-<strong>17</strong>.<br />

7. Montag M, K€oster M, Strowitzki T, Toth B. Polar body biopsy. Fertil<br />

sertil; 2013, 100(3):603-607.<br />

8. Hou Y, Fan W, Yan L, Li R, Lian Y, Huang J, Li J, Xu L, Tang F, Xie XS,<br />

Qiao J. Genome analyses of single human oocytes. Cell. 2013, 155(7):<br />

1492-1506.<br />

Supported by: This study was Supported by grants from National Basic<br />

Research Program of China (2014CB 81471512 to H.S.,2014CB 81370691<br />

to C.Y.).<br />

O-44 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:30 AM<br />

ANEUPLOIDY AND RECOMBINATION IN IN VITRO FERTILIZED<br />

EMBYROS (BLASTOCYSTS) UNDERGOING PREIMPLANTATION<br />

GENETIC DIAGNOSIS (PGD). K. Ravichandran, a Z. K. Gunes, b<br />

B. Bankowski, c A. M. Rosen, d S. H. Chen, e A. Hershlag, f B. Sandler, g<br />

J. Grifo, h D. Wells, i M. Konstantinidis. a a Reprogenetics, Livingston, NJ;<br />

b Reprogenetics UK, Oxford, United Kingdom; c Oregon Reproductive Medicine,<br />

Portland, OR; d Mercy Hospital and Medical Center, Chicago, IL;<br />

e IRMS at Saint Barnabas, Livingston, NJ; f North Shore LIJ Health System,<br />

Cold Spring Harbor, NY; g Obstetrics, Gynecology and Reproductive Science,<br />

Icahn School of Medicine at Mount Sinai, New York City, NY;<br />

h NYU Langone Medical Center, NY, NY; i University of Oxford.<br />

OBJECTIVE: To investigate aneuploidy and recombination in human preimplantation<br />

embryos in an effort to advance knowledge regarding the occurrence<br />

of these events during reproduction.<br />

DESIGN: Single nucleotide polymorphism (SNP) arrays and array<br />

comparative genomic hybridization (aCGH) were used in parallel to screen<br />

blastocysts undergoing PGD.<br />

MATERIALS AND METHODS: The Infinium Karyomapping assay (Illumina,<br />

USA) was used to process 480 blastocysts derived from 133 couples.<br />

Recombination data were obtained from 460 embryos. Aneuploidy data were<br />

collected from <strong>21</strong>4 embryos determined to be either euploid or aneuploid<br />

affected with trisomies of meiotic origin.<br />

RESULTS: 49 embryos were found to carry meiotic trisomies. Almost<br />

all of the meiotic trisomies were determined to be of maternal origin<br />

(57/58). 66.1% of trisomies were determined to have occurred during<br />

meiosis I and 33.9% during meiosis II. Chromosomes 5, 16, <strong>21</strong>, and 22<br />

were more likely to be affected by meiosis I errors while meiosis II errors<br />

were more frequent in chromosomes 6 and 18. Recombination analysis<br />

was carried out in 12,545 chromosomes and 19,769 recombination events<br />

were recorded. For autosomes, the average recombination rate was<br />

24.10.28 for male meiosis and 40.010.47 for female meiosis. Sites<br />

of recombination tended to cluster towards the distal ends of the chromosome,<br />

a pattern obvious in male meiosis. Female meiosis showed no patterns<br />

in locations of crossover events but spikes in recombination activity<br />

were noticed in the centromeric regions of chromosomes 16, <strong>17</strong>, 18, <strong>21</strong>,<br />

and 22. Comparative analysis conducted between 165 euploid embryos<br />

and 29 aneuploid embryos presenting meiotic trisomies showed no differences<br />

(P>0.05) between the average recombination rates for male meiosis<br />

(24.831.15 in aneuploid vs. 24.310.42 in euploid) or female meiosis<br />

(37.<strong>17</strong>1.64 in aneuploid vs. 40.890.69 in euploid). No differences in<br />

the locations of crossover sites between euploid and aneuploid embryos<br />

were observed.<br />

CONCLUSIONS: This study, yielded unique data concerning recombination<br />

and the origin of aneuploidy in human embryos. The dominance of female<br />

meiotic errors, especially those occurring in meiosis I, was<br />

confirmed. Certain chromosomes were most often affected during meiosis<br />

I and others during meiosis II. This indicates that relative contribution of<br />

mechanisms causing aneuploidy may differ for individual chromosomes.<br />

Recombination rates as well as locations of crossover sites differed between<br />

male and female meiosis but not between euploid and aneuploid embryos.<br />

O-45 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:45 AM<br />

MORPHOKINETICS AND NUCLEATION STATUS OF EMBRYOS<br />

AS MARKERS OF IMPLANTATION POTENTIAL. J. A. Aguilar, a<br />

E. Munoz, b E. Taboas, b M. Perez, b A. Delgado, c T. Viloria, c<br />

M. Meseguer. d a IVF Laboratory, IVI Vigo, Vigo, Spain; b IVI Vigo, Vigo,<br />

Spain; c IVI Valencia, Valencia, Spain; d Clinical Embryology, Valencia,<br />

Spain.<br />

OBJECTIVE: to study the nucleation status of in vitro human embryos in<br />

the second cell cycle (2cells and 4 cells embryos), the effect of multinucleation<br />

in 2cells in the implantation rate, the reversibility of multinucleation status<br />

in the 4 cells embryos and the embryo morphokinetics their relationship<br />

with embryo implantation.<br />

DESIGN: Observational retrospective study conducted in IVI Vigo and<br />

IVI Valencia, between 2011 and 2013. We analyzed 1679 embryos,<br />

cultured in an Automated Time-lapse Incubator incubator in a 37 C,<br />

6% CO2 and 20% de O2 atmosphere from 940 oocyte donation cycles.<br />

The exact timing of the events cited below was evaluated in hours post<br />

insemination by ICSI.<br />

MATERIALS AND METHODS: Nucleation status was annotated according<br />

to the number of nuclei present in each cell in the 2-cells and<br />

4-cells embryos as follows: nMONO (mononucleated) where ‘n’ represents<br />

the number of blastomeres in which a unique nucleus is seen. nBI<br />

(binucleated) number of blastomeres in which two nuclei per cell are<br />

visible. nMULTI (multinucleated) number of blastomeres in which<br />

more than two nuclei are visible. This definition includes micronuclei.<br />

The time for each embryo division was defined as t2(time to 2cells), t3<br />

(3cells),t4 (4cells), t5(5cells), t8(8cells);cc2¼t3-t2; cc3¼t5-t3; s2¼t4-t3;<br />

s3¼t8-t5 ANOVA test and c2-test were performed when applicable to<br />

assess the influence of the length of S-phase in the implantation rate.<br />

RESULTS: From 1602 embryos analyzed. 40.08% showed multinucleation<br />

at 2cells, while it was present in 14,4% of 4 cells embryos. There<br />

were no significant differences in implantation rate between mononucleated<br />

embryos and multinucleated (one or both blastomeres) after analyzing multinucleation<br />

at 2cells embryos(p¼0,07), on the contrary there were when<br />

considering it at 4 cells embryos (p


potential of monosomy embryos is negligible compared to their trisomy<br />

counterparts, yet the cause for this is unknown. Our previous investigation<br />

of epigenetic modifiers uncovered a decreased expression of DNA methyltransferase<br />

enzymes in monosomy blastocysts, as well as a hypomethylated<br />

state limited to the monosomic chromosome. This study was thus extended to<br />

determine the DNA methylation profiles of imprinting control regions (ICRs)<br />

present on the affected chromosomes.<br />

DESIGN: Research study.<br />

MATERIALS AND METHODS: Surplus cryopreserved human blastocysts<br />

were donated with IRB approval. Trophectoderm biopsies followed<br />

by comprehensive chromosomal screening using qPCR (RMA-NJ) determined<br />

chromosome constitution. Individual monosomy 11 (n¼10) or 15<br />

(n¼10), trisomy 11 (n¼10) or 15 (n¼10), and euploid oocyte donor blastocysts<br />

(controls; n¼20) underwent bisulfite conversion and PCR amplification<br />

to determine the DNA methylation profiles of two maternally<br />

methylated ICRs, KCNQ1OT1 (11p15.5) and SNRPN (15q11.2). Statistical<br />

analysis was calculated using the Student’s t-Test, with significance at<br />

p


with chromosome aneuploidy and embryo demise will assist in improving<br />

clinical IVF protocols.<br />

POLYCYSTIC OVARY SYNDROME<br />

O-49 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:15 AM<br />

ABNORMAL EXPRESSION OF GENES GOVERNING<br />

ADIPOGENESIS AND EXTRACELLULAR MATRIX (ECM)<br />

FORMATION IN SUBCUTANEOUS (SC) ABDOMINAL ADIPOSE<br />

STEM CELLS (ASCS) OF LEAN POLYCYSTIC OVARY SYNDROME<br />

(PCOS) WOMEN. G. D. Chazenbalk, a J. D. Phan, a V. Madrigal, a<br />

X. Ding, b X. Li, b D. A. Dumesic. a a OB/GYN, UCLA, Los Angeles, CA;<br />

b Pathology and Laboratory Medicine, UCLA, Los Angeles, CA.<br />

OBJECTIVE: To identify differential expression of genes involved in adipogenesis<br />

and ECM formation in subcutaneous (SC) abdominal adipose stem<br />

cells (ASCs) of lean PCOS and age- and BMI-matched normoandrogenic<br />

(NL) women.<br />

DESIGN: Prospective Cohort Study.<br />

MATERIALS AND METHODS: SC abdominal fat biopsy was obtained<br />

from four lean (body mass index: 19-25 kg/m 2 )PCOSwomen,ages<strong>21</strong>-32<br />

years, and four BMI- and age-matched NL women. ASCs were isolated<br />

after digestion of SC abdominal adipose with collagenase and differential<br />

centrifugation. RNA from ASCs was extracted using Qiagen miRNeasy<br />

Mini Kit. Microarray and small non-coding RNA (miRNA) analysis<br />

were performed using Affymetrix Human Genome U1.33 2.0 and Human<br />

miRNA arrays, respectively. Data were analyzed using Ingenuity Pathway<br />

Analysis.<br />

RESULTS: Differential expression of >1.5 fold change between SC<br />

abdominal ASCs of PCOS and NL women indicated 64 up-regulated<br />

and 103 down-regulated genes (P


progesterone level>5 ng/ml. Blood samples were obtained during the<br />

early follicular phase. TNF- a, IL-27, IL-35, IL-37, a-1 acid glycoprotein<br />

were measured by a high sensitivity ELISA, while CRP, testosterone (T),<br />

androstenedione (AE), SHBG, DHEA-S levels were measured by RIA.<br />

Free androgen index (FAI) was calculated as percentage ratio of total<br />

testosterone to SHBG values.<br />

RESULTS: Levels of inflammatory mediators (CRP, TNF- a, a-1 acid<br />

glycoprotein) were significantly increased in obese PCOS patients<br />

compared to obese controls and in lean PCOS patients compared to<br />

lean controls (all p


and tended to positively correlate with SC abdominal fat volume<br />

(R 2 ¼0.29, P¼0.09) and waist size (R 2 ¼0.28, P¼0.08).<br />

CONCLUSIONS: SC abdominal adipocyte size in lean PCOS and NL<br />

women negatively correlates with Si as a predictor of abdominal fat content<br />

rather than androgen excess per se, and is accompanied by a population of<br />

small SC abdominal adipocytes in PCOS.<br />

Reference:<br />

1. Chazenbalk G, Singh P, Irge D, et al. Androgens inhibit adipogenesis<br />

during human adipose stem cell commitment to predipocyte formation.<br />

Steroids 2013;78:920-6.<br />

References:<br />

1. Chang WY, Clements D, Johnson SR. Effect of doxycycline on<br />

proliferation, MMP production, and adhesion in LAM-related<br />

cells. Am J Physiol Lung Cell Mol Physiol. 2010<br />

Sep;299(3):L393-400.<br />

2. Lewandowski, KC et al Increased circulating levels of matrix Metalloproteinase-2<br />

and -9 in Women with the Polycystic Ovary Syndrome<br />

JCEM 2006; 91:1<strong>17</strong>3-77.<br />

Supported by: NIH grant P50 HD071836.<br />

O-54 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:30 PM<br />

IMPACT ON OVULATORY PARAMETERS WITH THE USE OF<br />

MMP INHIBITORS IN PCOS, MI-PCOS: A PILOT<br />

STUDY. A. Light, a S. Alur, b S. Hammes, c K. Hoeger. b a Pathology and<br />

Laboratory Medicine, University of Rochester Medical Center, Rochester,<br />

NY; b Obstetrics and Gynecology, University of Rochester Medical Center,<br />

Rochester, NY; c Endocrinology, University of Rochester Medical Center,<br />

Rochester, NY.<br />

OBJECTIVE: PolycysticOvary Syndrome (PCOS) is characterized in<br />

part by anovulation and excess ovarian androgen mediated by excess<br />

LH. Wepreviously reported LH induced steroid production in freshly<br />

isolated murine ovarian follicles requires transactivation of the EGF<br />

receptor following release of EGF-like ligands by matrix metalloproteinases<br />

(MMPs). This steroidogenesis is suppressed by doxycycline<br />

(DOXY), anagent known to inhibit MMP2/9(1). Furthermore, we<br />

have shown that the addition of the MMP inhibitor Galardin to prenatally<br />

androgenized mice improved cycling. Given the association of<br />

LH induced steroidogenesis and evidence MMP2/9 activities are<br />

elevated in PCOS(2), MMP-I may treat the abnormal endocrine state<br />

in PCOS. In this pilot study, we examined the impact of MMP 2/9<br />

inhibition with DOXY on ovulation and androgens in women with<br />

PCOS.<br />

DESIGN: Randomized placebo controlled pilot trial<br />

MATERIALS AND METHODS: 13 women diagnosed with PCOS by<br />

Rotterdam criteria were recruited and 10 met criteria and were randomized<br />

to DOXY 100mg BID for 12 weeks or an identical placebo tablet<br />

(PL) BID for 12 weeks and then no treatment for 12 weeks. At baseline,<br />

total and free testosterone, and SHBG were obtained. Women collected<br />

urine for urinary pregnanediol glucuronide (UDPG) assessment weekly<br />

and were seen biweekly for serum progesterone (P4) and menstrual diary<br />

collection for 24 weeks. Androgens were assessed at 12 and 24<br />

weeks.<br />

RESULTS: 5 subjects received DOXY and 5 PL. One DOXY subject<br />

completed 4 weeks of UDPG and 9 P4. One PL subject completed 12<br />

weeks. Missing data were assumed to be not ovulatory. Over 24 weeks,<br />

DOXY had a mean of 2.4 ovulations versus 1.4 in PL (p¼0.12). 37%<br />

of the possible cycles were ovulatory in DOXY versus 23% of the PL<br />

(p¼0.13). 3/5 DOXY subjects had at least 3 ovulations versus 1/5 PL.<br />

DOXY did not impact total or free testosterone measures or SHBG<br />

compared to PL at 12 or 24 weeks.<br />

CONCLUSIONS: DOXY did not reduce serum androgens in this pilot<br />

study; however, there appeared to be a trend toward more ovulatory cycles<br />

in the DOXY group, though this did not reach significance. If ovulation<br />

is improved it may be through subtle changes in androgens that are<br />

not detectable using our current techniques or may be via a different<br />

mechanism than change in androgen production globally. A follow up<br />

larger trial is needed.<br />

Baseline Characteristics<br />

PL<br />

DOXY<br />

Age (years, range) 27.4 (24-31) 28.2 (<strong>21</strong>-39)<br />

BMI (kg/m2, range) 29.3 (24-37.6) 33.3 (24.6-47.7)<br />

Total T (ng/dL) (SD) 49.6 (14.6) 50.6 (8.2)<br />

Free T (ng/dL) (SD) 1.36 (0.39) 1.14 (0.57)<br />

SHBG (nmol/L) (SD) 44.0 (13.5) 55 (13.2)<br />

PROCEDURES AND TECHNIQUES<br />

O-55 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:15 AM<br />

EMBRYO SELECTION USING THE EEVA TEST, AN AUTOMATED<br />

TIME-LAPSE QUANTITATIVE ANALYSIS, IN CONJUNCTION<br />

WITH STANDARD MORPHOLOGY: A PROSPECTIVE<br />

MULTI-CENTER PILOT STUDY. D. C. Kieslinger, a S. De Gheselle, b<br />

P. De Sutter, b E. H. Kostelijk, a J. van Rijswijk, a C. B. Lambalk, a E. van<br />

den Abbeel, b C. G. Vergouw. a a Department of Obstetrics and Gynecology,<br />

VU University Medical Center, Amsterdam, Netherlands; b Centre for Reproductive<br />

Medicine, University Hospital Gent, Gent, Belgium.<br />

OBJECTIVE: To evaluate the impact of selecting embryos for transfer<br />

using Eeva Test prediction scores combined with traditional morphology<br />

on the pregnancy rate of IVF and ICSI patients undergoing Day 3 embryo<br />

transfer.<br />

DESIGN: Prospective, observational, multi-center pilot study. The analysis<br />

involved 280 of 302 enrolled patients; 560 control patients were identified<br />

from a cohort of <strong>17</strong>77 patients (year 2011-2013) using propensity<br />

matching based on age, cycle number, oocyte number and number of<br />

fertilized oocytes. The majority of transfers (98%) were single embryo<br />

transfers.<br />

MATERIALS AND METHODS: Two academic hospitals (VUmc Amsterdam<br />

and UZ Gent) enrolled patients


O-56 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:30 AM<br />

COMPARISON OF A BENCHTOP INCUBATOR AND A TIME<br />

LAPSE INCUBATOR FOR CULTURE OF HUMAN EMBRYOS:<br />

IMPACT OF CULTURE DISH. R. L. Krisher, a C. Pospisil, b<br />

A. F. Greene, a J. M. Stevens, b W. B. Schoolcraft, c J. E. Swain. d a National<br />

Foundation for Fertility Research, Lone Tree, CO; b Fertility Labs of Colorado,<br />

Lone Tree, CO; c Colorado Center for Reproductive Medicine, Lone<br />

Tree, CO; d Fertility Lab Sciences, Lone Tree, CO.<br />

OBJECTIVE: Much debate exists as to the optimal embryo culture incubator.<br />

However, several variables exist that must be controlled to properly<br />

examine the impact of the incubator itself, including the culture dish utilized<br />

and the gas atmosphere. Our objective was to compare human embryo development<br />

between two modern embryo culture incubators; the EmbryoScopeÔ<br />

and the benchtop G185.<br />

DESIGN: Prospective sibling embryo split<br />

MATERIALS AND METHODS: Zygotes from 26 patients (average age<br />

36.70.7) were placed individually into EmbryoSlideÔ dishes (25 mL under<br />

oil). One EmbryoSlideÔ was placed into the EmbryoScopeÔ (Vitrolife,<br />

n¼<strong>17</strong>2 2PNs), while the other was placed into a G185 incubator (K-Systems,<br />

n¼158 2PNs). Embryos were cultured in sequential media for 6 days. Both<br />

incubators utilized 7% CO2, 5% O2. To examine the impact of the dish,<br />

zygotes from 8 patients (average age 33.10.9) were randomly assigned to<br />

culture in either an EmbryoSlideÔ (n¼58 2PNs) or to standard group microdrop<br />

culture (up to 5 embryos per 30mL; n¼73 2PNs) and cultured in a G185<br />

incubator.<br />

RESULTS: Significantly more embryos developed to good quality day 5<br />

blastocysts (>3BB) in the EmbryoSlideÔ when it was incubated in the<br />

G185 benchtop incubator compared to the EmbryoScopeÔ (39.2% versus<br />

28.5%, respectively; p¼0.039), although there was no difference in the number<br />

of good quality blastocysts developing overall by day 6 (54.4% vs.<br />

52.3%, respectively). When cultured in the same incubator (G185), the EmbryoSlideÔ<br />

Supported good quality blastocyst development equal to that of<br />

microdrop group culture (EmbryoSlideÔ- 25.9% day 5 and 58.6% overall;<br />

group microdrop- 26.0% day 5 and 58.9% overall).<br />

CONCLUSIONS: In patient embryo splits with a controlled gas atmosphere<br />

and culture dish, the benchtop G185 incubator yielded more high<br />

quality day 5 blastocysts compared to the EmbryoScopeÔ. The EmbryoSlideÔ<br />

dish supports blastocyst development of individually cultured human<br />

embryos equally well compared to group culture in standard<br />

microdrops within the benchtop G185. Using sequential media and low oxygen,<br />

modern benchtop incubators function as well as time-lapse incubators<br />

and permit individual embryo culture, which is critical for the application of<br />

biomarker analysis.<br />

O-57 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:45 AM<br />

PATIENT TREATMENT JOURNEYS, THEN AND NOW: 10 YEARS<br />

OF CLINICAL PROGRESS?. A. Zgodic, a H. Karvir, a D. Parfitt, a<br />

S. T. Globus, a A. B. Copperman, b P. Yurttas Beim. a a Celmatix Inc, New<br />

York, NY; b Reproductive Medicine Associates of New York, New York, NY.<br />

OBJECTIVE: As assisted reproductive technology (ART)-based therapies<br />

are refined and clinical expertise is strengthened over time, improvements<br />

are expected in live-birth outcomes for patients beginning<br />

infertility treatment. To investigate this hypothesis, we analyzed trends in<br />

clinical outcomes over a ten-year period at a large academic reproductive<br />

medical center.<br />

DESIGN: Retrospective analysis using de-identified ART cycles (10801<br />

in-vitro fertilization (IVF) cycles, 1462 Egg Recipient (ER) cycles, and<br />

33947 Non-IVF (NIVF) cycles) from 2003-2012.<br />

MATERIALS AND METHODS: We performed trend analyses within:<br />

1) All; 2) IVF; and 3) NIVF cycles. We used two-sided chi-square test to<br />

investigate significant trends in live birth and multiples rates. The direction<br />

and magnitude of these trends were assessed using a Binomial model with<br />

treatment year as a predictor. For mean time-to-live birth (months), mean<br />

cost of a singleton live birth, mean time between cycles (months), and<br />

mean number of embryos transferred (IVF only), we used a linear model<br />

to detect significant trends.<br />

RESULTS: In all cycles, we observed no significant trends in odds of<br />

achieving live birth or multiples (p


Augmented PR activity by AKAP13 in vivo was dependent upon the presence<br />

of P4. These results suggest that in cells the interaction between<br />

AKAP13 and PR might require the presence of P4. Further studies are needed<br />

to characterize specific molecular mechanism(s) by which AKAP13 influences<br />

PR function.<br />

Supported by: HD 008737 to JHS.<br />

O-59 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:15 PM<br />

ART DOES NOT INCREASE THE MOLECULAR KARYOTYPE<br />

ABNORMALITY RATE OF MISCARRIAGE. G. Li, W. Niu, Y. Su,<br />

Y. Guo, Y. Sun. Reproductive Medical Center, The First Affiliated Hospital<br />

of Zhengzhou University, Zhengzhou, China.<br />

OBJECTIVE: To explore whether assisted reproductive technology (ART)<br />

increase the molecular karyotype abnormality rate of miscarriage.<br />

DESIGN: Retrospective study.<br />

MATERIALS AND METHODS: IRB approval was obtained. Miscarriage<br />

tissues underwent DNA extraction and 23-chromosome SNP microarray<br />

analysis using humanCytoSNP-12 DNA beadchips and GenomeStudio software.<br />

RESULTS: 618 coupes were enrolled. The mean age was 31.95.1<br />

years. 144 experienced natural conception (NC) and 474 were pregnant<br />

by ART. Of the 474 cases, 36 underwent artificial insemination (AI),<br />

244 fresh in vitro fertilization-embryo transfer (IVF-ET), 66 fresh intracytoplasmic<br />

sperm injection (ICSI)-embryo transfer and 128 thawed embryo<br />

transfer. The total abnormal molecular karyotype rate was 70.9%<br />

(438/618). There was no significant difference in the abnormal molecular<br />

karyotype rate between ART (70.5%) and NC (72.%). The abnormal molecular<br />

karyotype rate in the subgroup of ICSI was 65.6%,which was lower<br />

than that of AI+NC( 71.7%) and IVF( 72.2%) ,However, there was no significant<br />

difference between them. In the cases aged over 35 years, the<br />

abnormal molecular karyotype rate was 83.3%,which was higher than<br />

66.4% in the cases aged 35 years and under .<br />

CONCLUSIONS: ART does not increase abnormal molecular<br />

karyotype rate of abortion tissue. Moreover, in terms of fertilization<br />

mode, compared with the IVF and AI+NC, ICSI does not increase<br />

abnormal molecular karyotype rate of miscarriage. Abnormal molecular<br />

karyotype rate is increased in the cases aged over 35 years in both NC<br />

and ART. To our knowledge, this is the largest and most comprehesive<br />

trial conducted evaluating SNP molecular karyotyping of spontaneous<br />

miscarriage after ART.<br />

O-60 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:30 PM<br />

CYST ASPIRATION VERSUS GNRH ANTAGONIST<br />

ADMINITRATION FOR OVARIAN CYSTS DETECTED AT<br />

THE START OF FRESH IN VITRO FERTILIZATION<br />

CYCLES. N. Pereira, a S. Amrane, b E. Hobeika, c J. Lekovich, a<br />

P. Chung, d Z. Rosenwaks. d a The Ronald O. Perelman and Claudia Cohen<br />

Center for Reproductive Medicine, New York, NY; b New York Presbyterian<br />

Hospital, New York, NY; c North Shore LIJ at Staten Island, Staten<br />

Island, NY; d Weill Cornell Medical College, New York, NY.<br />

OBJECTIVE: To investigate effect of cyst aspiration versus GnRH antagonist<br />

(GnRH-ant) administration for ovarian cysts detected at the start of<br />

fresh in vitro fertilization (IVF)-embryo transfer (ET) on the outcomes of<br />

the same IVF-ET cycles.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: All patients undergoing fresh IVF-ET<br />

between January 2008 and July 2013 who had transvaginal ultrasonogram-guided<br />

cyst aspiration or treatment with GnRH-ant for ovarian<br />

cysts detected at the start of IVF-ET prior to controlled ovarian simulation<br />

(COS) were analyzed for potential inclusion. Exclusion criteria<br />

included donor oocyte cycles, non-GnRH antagonist based cycles, or patients<br />

having > 1 ovarian cysts at IVF start. Patients with known simple<br />

ovarian cysts or endometriomas were excluded. Demographic and baseline<br />

IVF characteristics were recorded from patient charts and included<br />

age, gravidity, parity, body mass index (kg/m 2 ), estradiol (E 2 ) level at<br />

cycle start (pg/mL), total days of ovarian stimulation, total dosage of<br />

gonadotropins administered (IU), peak E 2 level (pg/mL), number of<br />

mature oocytes retrieved, number of embryos transferred, and day<br />

of ET. Pregnancy outcomes (clinical pregnancy rate, spontaneous miscarriage<br />

rate, live birth rate) following ET were also noted. Student’s t-tests<br />

and Chi-squared tests were used as indicated. Statistical analyses were<br />

performed using STAT version 13 (College Station, TX: Stata Corp<br />

LP). Statistical significance was set at P < 0.05.<br />

RESULTS: 403 patients met inclusion criteria: 41 underwent cyst<br />

aspiration and 362 were treated with GnRH-ant. Patients who underwent<br />

cyst aspiration had a larger cyst size compared to GnRH-ant group i.e.,<br />

31.1 5.71 mm versus 22.3 6.33 (P < 0.01). There was no difference<br />

in the demographics or baseline IVF cycle characteristics of the two<br />

groups. Patients treated with GnRH-ant had a longer duration of COS<br />

(10.8 3.45 days vs. 9.05 4.06, P¼0.01) and required higher doses of<br />

gonadotropins (3887.7 1097.8 IU vs. 3293.7 990.5; P¼0.003)<br />

compared to the cyst aspiration group. The median number of mature<br />

oocytes and embryos transferred were comparable between the groups.<br />

There was no difference in the overall clinical pregnancy (43.9% vs.<br />

41.4%), spontaneous miscarriage (9.76% vs. 8.01%) and live birth (34.1<br />

vs. 33.4%) rates between the groups.<br />

CONCLUSIONS: Our findings suggest that cyst aspiration or GnRH-ant<br />

administration are equivalent modalities for management of solitary ovarian<br />

cysts detected at the start of IVF-ET. However, in our study cohort, patients<br />

treated with GnRH-ant had longer duration of COS and required higher<br />

gonadotropin doses without adversely impacting the yield of mature oocytes<br />

or pregnancy rates.<br />

IN VITRO FERTILIZATION 1<br />

O-61 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:15 AM<br />

EFFECT OF ELEVATED PROGESTERONE ON THE DAY OF<br />

TRIGGER AND PREGNANCY OUTCOMES IN GNRH AGONIST<br />

TRIGGER CYCLES. M. T. Connell, a G. Patounakis, a M. W. Healy, a<br />

A. DeCherney, a K. Devine, b E. A. Widra, c M. J. Hill. a a National Institutes<br />

of Health, Bethesda, MD; b Shady Grove Fertility Center, Washington, DC;<br />

c Shady Grove Fertility, Washington, DC.<br />

OBJECTIVE: The current published data on progesterone elevation and<br />

IVF outcomes have predominantly come from hCG trigger cycles. Trigger<br />

medication has been shown to affect luteal phase hormone profiles. The<br />

objective of this study was to determine whether elevated progesterone level<br />

was similarly associated with clinical pregnancy outcome when GnRH<br />

agonist trigger was used for IVF.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Fresh autologous IVF cycles from<br />

2011-2013 were included. The primary outcome was clinical pregnancy.<br />

GEE modelling with interaction testing was used to control for confounding<br />

variables and multiple patient cycles.<br />

RESULTS: 3226 cycles were evaluated of which 647 cycles utilized an<br />

agonist trigger. The agonist trigger cohort were younger and had higher<br />

oocyte yield, better quality embryos, and fewer embryos transferred<br />

compared to the hCG trigger cohort (P


O-62 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:30 AM<br />

SUCCESSFUL SURVIVAL IN CULTURE OF SPINAL CORD (SC)<br />

DORSAL ROOT GANGLIA (DRG) NEURONS AND GLIA<br />

CELLS AFTER AUTOMATIC VITRIFICATION. A. Arav, a<br />

A. Shahar, b O. Ziv-Polat, b Y. Natan, c P. Patrizio. d a IVF Research lab, FertileSafe<br />

Ltd., Nes Ziona, Israel; b NVR, Nes Ziona, Israel; c FertileSafe Ltd.,<br />

Nes-Ziona, Israel; d Yale Fertility Center & Fertility Preservati, New Haven,<br />

CT.<br />

OBJECTIVE: To assess the survival and re-growth of neuronal cells after<br />

cryopreservation with a new automatic vitrification system.<br />

DESIGN: Experimental research with SC and DRG isolated from rat fetuses.<br />

MATERIALS AND METHODS: Stationary organotypic DRG and SC<br />

cultures were prepared from rat fetuses (gestational day 15, Lewis inbred,<br />

Harlan, Israel). mmediately after dissection, the DRG and SC were cut<br />

into small slices (400 micron) and seeded in 12 well culture plates containing<br />

1 mL of NVR-Gel. Cultures were enriched once in gel with<br />

GDNF conjugated iron oxide nanoparticles (10 ng/mL) and subsequently<br />

with nutrient medium at each consecutive feeding. Monitoring of the<br />

DRG-SC growth pattern was done by daily phase contrast microscopic<br />

observations 24 hours after setting the cultures onward and by IF staining.<br />

The various neuronal samples (SC, DRG, glia cells) were cryopreserved<br />

with a new device for automatic vitrification (SarahÒ, Fertilesafe, IL),<br />

consisting of a special capsule containing EM gold grids attached to<br />

0.25ml straws (IMV,France). The straws loaded with neuronal tissue<br />

were placed into the mixing chamber of the device attached to a syringe<br />

pump dispensing timed and gradual increasing concentrations of equilibrium<br />

solution for 12-25 min. and vitrification solutions for 3-5 min.,<br />

before plunginginto liquid nitrogen slush (VitMaster, IMT IL). After 1<br />

week the samples were rewarmed according to the Origio’s kit instructions.<br />

To assess survival, the cells recovered were fixed (4% paraformaldehyde),<br />

permeabilized with 0.1% Triton X-100 in PBS and then<br />

immunoblocked with a 1% BSA in PBS for 1 h at RT prior to double<br />

staining with mouse anti S-100 antibodies (1:80, Acris Antibodies, glial<br />

cell marker) and rabbit anti neurofilament antibodies (NF, Novus Biologicals,<br />

1:500, neuronal cell marker). The primary antibodies were diluted<br />

in 0.1% BSA and 0.05% Tween-20 in PBS and incubated with the specimens<br />

overnight at 4 C. After rinsing, the DRG specimens were incubated<br />

for 1 h at RT with fluorescent secondary antibodies.<br />

RESULTS: Organotypic cultures of SC, DRG neurons and glia cells<br />

demonstrated a remarkable survival as well as nerve fiber regeneration after<br />

cryopreservation/rewarming. The SC neurons maintained their multipolar<br />

shape with re-growth of dendrites and axons. The round shaped DRG neurons<br />

exhibited euchromatic nuclei with prominent nucleoli and an active regeneration<br />

of nerve processes.<br />

CONCLUSIONS: This is the first report on successful vitrification of SC<br />

cells, DRG neurons and glia cells using an automatic freezing device. The<br />

remarkable resumption of growth for neuronal cells will greatly expand<br />

regenerative research opportunities.<br />

O-63 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:45 AM<br />

THE IN VITRO DEVELOPMENT OF HUMAN ZYGOTE<br />

RECONSTRUCTED BY PRONUCLEAR TRANSFER. H. Liu, a<br />

Z. Lu, a A. Chavez-Badiola, b P. Colls, c S. Munne, c J. Zhang. a a New<br />

Hope Fertility Center, New York, NY; b New Hope Fertility Center Mexico,<br />

Guadalajara, Jalisco, Mexico; c Reprogenetics, Livingston, NJ.<br />

OBJECTIVE: Nuclear transfer techniques, including metaphase II<br />

spindle transfer and pronuclear transfer (PNT), have been proposed as<br />

potential therapeutic measures to prevent maternal mitochondrial (mt)<br />

DNA disease transmission and to treat recurrent failure of implantation<br />

due to ooplasmic dysfunction. Although PNT has been widely tested on<br />

several animal models resulting in normal offsprings, the use of this<br />

technique in humans is still limited due to safety and efficacy concerns.<br />

This study aimed to examine the feasibility of PNT and its outcome<br />

on human chromosome complements and pre-implantation blastocyst<br />

development.<br />

DESIGN: Two pronuclei (2PN) were exchanged between two zygotes<br />

via PNT; the reconstructed zygotes were cultured to blastocysts followed<br />

by aneuploidy testing using array comparative genomic hybridization<br />

(aCGH).<br />

MATERIALS AND METHODS: Oocytes and sperm from gamete donors<br />

aged between <strong>21</strong> and 28 were collected and frozen for these experiments.<br />

Donor oocytes and sperm were then thawed and cultured in vitro<br />

for 3 hours before intracytoplasmic sperm injection (ICSI). Only zygotes<br />

with 2PN at 15 hours after ICSI were used in this study and were subjected<br />

to PNT via electric pulse to initiate membrane fusion between<br />

the isolated karyoplasm and ooplasm. Reconstructed zygotes were<br />

cultured in vitro until the blastocyst stage. The blastocysts were biopsied<br />

by extracting 3-5 cells of the trophectoderm followed by aCGH testing<br />

for aneuploidy. Blastocysts formed from fertilized donor oocytes and<br />

sperm that were not subjected to PNT were used as controls (non-<br />

PNT). Data were analyzed using chi-square test.<br />

RESULTS: Thirty-two donor oocytes were thawed with 93.8% survival<br />

rate, and 80% of the survived oocytes fertilized normally after ICSI with<br />

thawed donor sperm forming a total of 24 zygotes (2PN). The zygotes<br />

were then subjected to PNT, 37.5% of which developed to the blastocyst<br />

stage compared to 53.7% blastocyst formation rate in the non-PNT group<br />

(p>0.05). In addition, 7 out of 9 (77.8%) PNT blastocysts showed euploid<br />

chromosomes following aCGH compared to 55% (n¼158) euploidy rate in<br />

the non-PNT group (p¼0.6).<br />

CONCLUSIONS: PNT is applicable in human zygotes without imposing<br />

adverse impact onto the pre-implantation embryonic development and its<br />

chromosomal content.<br />

O-64 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:00 PM<br />

IDENTIFICATION OF EMBRYO MARKERS PREDICTING<br />

BLASTOCYST FORMATION BEFORE 1ST CLEAVAGE. Y. Kida,<br />

N. Fukunaga, H. Kitasaka, T. Yoshimura, K. Nakayama, H. Ohno,<br />

M. Takeuchi, M. Shimomura, S. Kounogi, Y. Asada. Asada Ladies Clinic<br />

Medical Corporation, Nagoya, Aichi, Japan.<br />

OBJECTIVE: To identify potential embryo markers predicting blastocyst<br />

formation before 1st cleavage using EmbryoScopeÒ.<br />

DESIGN: Prospective cohort study.<br />

MATERIALS AND METHODS: We examined 288 embryos resulting<br />

from normal fertilization in 25 recipient patients undergoing ICSI between<br />

March and July, 2013. The embryos were placed and cultured in the Embryo-<br />

ScopeÒ in droplets of 30ml of Continuous Single Culture Medium (Irvine<br />

Scientic:USA) till day 7. We analyzed the time of PB2 extrusion (tPB2),<br />

appearance (tPNa) and fade (tPNf) of pronuclei. For tPNa, the time of<br />

maternal pronuclei appearance was used. The following intervals were calculated:<br />

PB2-PNa, PNa-PNf. We used Gardner’s classification to evaluate blastocysts.<br />

Blastocysts evaluated more than 3BB were considered as goodquality<br />

ones. T-test was used for comparison of mean timings and Chisquared<br />

test for comparison of blastocyst rates. P-values


O-65 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:15 PM<br />

A LONGITUDINAL STUDY OF FAMILIES CREATED<br />

BY REPRODUCTIVE DONATION: FOLLOW-UP AT<br />

ADOLESCENCE. E. C. Ilioi, V. Jadva, S. Golombok. Centre for Family<br />

Research, University of Cambridge, Cambridge, United Kingdom.<br />

OBJECTIVE: This study examines parental psychological well-being,<br />

adolescent psychological adjustment and family functioning in families<br />

that have used donor insemination (DI), egg donation (ED), surrogacy,<br />

or natural conception (NC) to have a child. It is the first longitudinal<br />

study to compare children conceived by different forms of reproductive<br />

donation at adolescence, the time at which identity issues become<br />

prominent and difficulties in parent-child relationships are likely to<br />

increase.<br />

DESIGN: The study represents the sixth phase of a longitudinal investigation.<br />

Previous phases took place when the children were aged 1, 2, 3, 7, and<br />

10 years. The study uses a multi-informant (mother and adolescent), and<br />

multi-method (standardized interviews and questionnaires) approach to<br />

data collection.<br />

MATERIALS AND METHODS: Data were obtained from 31 DI families,<br />

28 ED families, 29 surrogacy families, and 57 NC families when<br />

the children were 13 to 14 years old. Questionnaires assessed parent psychological<br />

well-being (the State-Trait Anxiety Inventory and the Edinburgh<br />

Depression Scale), parenting and family functioning (the FAM<br />

III Communication Scale, the Index of Family Relationships, the Rhoner<br />

Short Parental Acceptance/Rejection Scale and the Rhoner Parental Control<br />

Scale) and adolescent psychological adjustment (Strengths and Difficulties<br />

Questionnaire).<br />

RESULTS: Preliminary analyses of questionnaire data found few differences<br />

between families who conceived through reproductive donation and<br />

those who conceived naturally. In terms of disclosure rates since age 10, 3<br />

DI families, 5 ED families, and 1 surrogacy family had told their child about<br />

their conception resulting in 39% of DI families, 64% of ED families, and<br />

86% of surrogacy families having disclosed this information to their child.<br />

Overall, adolescents who were aware of their conception showed a clear understanding<br />

of it, with the majority expressing a neutral or indifferent attitude.<br />

CONCLUSIONS: Despite concerns about the psychological impact<br />

of reproductive donation at adolescence, the findings indicate that the<br />

families in this study were highly functioning in relation to parent psychological<br />

well-being, and the quality of adolescents’ relationships with<br />

their parents, irrespective of the specific method used in the child’s<br />

conception. Initial findings indicate that parents who used reproductive<br />

donation are not more likely to reject their child or have increased<br />

strain in parent-child relationships at adolescence. Furthermore, adolescents<br />

did not differ in their psychological adjustment, despite some of<br />

them knowing they are not genetically related to one parent. Whilst<br />

disclosure rates differed between groups, most of the adolescents who<br />

were aware of their conception showed a clear understanding of it<br />

regardless of whether they were told earlier in childhood or at adolescence.<br />

Supported by: This research was Supported by the Wellcome Trust<br />

[097857/Z/11/Z]. ECI is additionally funded by Pembroke College,<br />

Cambridge.<br />

O-66 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:30 PM<br />

INTRALIPID FOR REPEATED IMPLANTATION FAILURE (RIF): A<br />

RANDOMIZED CONTROLLED TRIAL. W. El-khayat a,b M. El<br />

Sadek. a,b a Obstetrics and Gynecology, Cairo University, Cairo, Egypt;<br />

b Middle East Fertility Centre, Giza, Egypt.<br />

OBJECTIVE: To evaluate the effect of intravenous infusion of intralipid<br />

20% on the outcome of intracytoplasmic sperm injection (ICSI) cycles in<br />

women with history of repeated implantation failure (RIF).<br />

DESIGN: An open label prospective randomized controlled trial<br />

MATERIALS AND METHODS: This was a prospective randomized<br />

controlled trial which included two hundred and three infertile couples<br />

with a history of repeated implantation failure. Repeated implantation<br />

failure defined as failure to achieve pregnancy after two to six ICSI cycles<br />

with the transfer of more than ten high grade embryos. Sample size was<br />

calculated as prior data (1) indicated that the pregnancy rate among controls<br />

is 18%. If the pregnancy rate for experimental subjects will be<br />

doubled to be 36%, we will need to study 94 subjects in each arm. We<br />

will use an uncorrected chi-squared statistic to evaluate the null hypothesis.<br />

We excluded patients with uterine fibroid, endometrial polyp, endometriosis<br />

and hydrosalpinx, also disturbances of normal fat metabolism,<br />

allergy to eggs, soybean oil, and finally liver disease, kidney disease,<br />

lung disease and blood clotting disorder. The subjects were randomly<br />

allocated to either the intervention group (group A) or the control group<br />

(group B) with the use of computer-generated random number tables and<br />

opaque sealed envelopes containing the participants’ group allocation.<br />

Group A (n¼101) subjects were given Intralipid 20% ( Intralipid, Fresenius<br />

Kabi AB, Uppsala, Sweden) by intravenous infusion between day<br />

four and nine of ovarian stimulation during ICSI cycles & another dose<br />

when got pregnant within the 1st week of positive pregnancy test, while<br />

group B (n¼102) subjects were control group and not given Intralipid.<br />

The primary outcome measure was the clinical pregnancy rate. Clinical<br />

pregnancy was considered as the presence of a gestational sac with a fetal<br />

heart activity. Comparison of quantitative variables between the study<br />

groups was done with the use of Student t-test for independent samples.<br />

For comparing categorical data, chi-square test was performed (Clinical<br />

Trial Registration Number: NCT01540591).<br />

RESULTS: There were statistically significant differences between both<br />

groups regarding the clinical pregnancy rate, the implantation rate and<br />

live birth rate. The clinical pregnancy rate was 35% in group A while it<br />

was only 15% in group B (p ¼ 0.001) with RR 1.6 (1.25 - 2.09), the<br />

implantation rate in group A was 13% while in group B it was 5%<br />

(p ¼ 0.048) with RR 1.5 (1.09 - 2.08), and the live birth rate was 33%<br />

in group A while it was 13% in group B (p ¼ 0.001) with RR 1.6<br />

(1.28 - 2.13).<br />

CONCLUSIONS: Intravenous Intralipid 20% infusion improved<br />

significantly the clinical pregnancy rate, the implantation rate and the<br />

live birth rate in ICSI cycles in women with history of repeated implantation<br />

failure.<br />

Reference:<br />

1. Shohayeb A, El-Khayat W. Does a single endometrial biopsy regimen<br />

(S-EBR)improve ICSI outcome in patients with repeated implantation<br />

failure? A randomised controlled trial. Eur J Obstet Gynecol Reprod<br />

Biol. 2012 Oct;164(2):<strong>17</strong>6-9.<br />

e26 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


REPRODUCTIVE SURGERY<br />

O-67 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:15 AM<br />

THE IMPACT OF UNILATERAL SALPINGECTOMY ON ANTRAL<br />

FOLLICLE COUNT AND OVARIAN RESPONSE IN ICSI CYCLES:<br />

COMPARISON OF CONTRALATERAL SIDE. B. Demir, a<br />

G. Bozdag, b O. Sengul, c I. Kahyaoglu, d S. Mumusoglu, e D. Zengin. f<br />

a Obstetrics and Gynaecology, Etlik Zubeyde Hanim Women’s Health Teaching<br />

and Research Hospital, Ankara, Turkey; b Hacettepe University, School of Medicine,<br />

Ankara, Turkey; c Yildirim Beyazid University, School of Medicine, Ankara,<br />

Turkey; d Etlik Zubeyde Hanim Women’s Health Teaching and Research<br />

Hospital, Ankara, Turkey; e Obstetrics and Gynaecology, Hacettepe University,<br />

School of Medicine, Ankara, Turkey; f Hacettepe University, Ankara, Turkey.<br />

OBJECTIVE: Salpingectomy is a treatment option in cases with hydrosalpinx<br />

and ruptured ectopic pregnancy. Because of the close relationship<br />

between mesosalpinx and ovarian blood supply, salpingectomy may compromise<br />

ovarian reserve. However, salpingectomy related detrimental effect on<br />

the ovarian reserve is still controversial. The purpose of this study is to determine<br />

the impact of salpingectomy on the ovarian reserve. For this aim, antral<br />

follicle count and ovarian response were compared with the contralateral side<br />

in patients with unilateral salpingectomy undergoing ICSI cycles.<br />

DESIGN: Retrospective multicenter study<br />

TABLE 1. Antral follicle count and ovarian response of the patients<br />

Operated<br />

side<br />

MATERIALS AND METHODS: Between January-2009 and April-<strong>2015</strong>,<br />

all patients who had unilateral salpingectomy undergoing ICSI treatment in<br />

ART Center of Etlik Zubeyde Hanim Women’s Health Teaching and<br />

Research Hospital and Department of Obstetrics and Gynecology, Hacettepe<br />

University were retrospectively reviewed from patient files and computer<br />

based data. Inclusion criteria were history of unilateral salpingectomy, woman’s<br />

age %40 years, no history of ovarian surgery. Women with bilateral salpingectomy<br />

and a history of endometriosis were excluded from the study.<br />

Ovarian reserve was evaluated with the antral follicle count (AFC). AFC<br />

was measured at the beginning of the menstrual cycle. Controlled ovarian hyperstimulation<br />

(COH) parameters and number of collected oocytes were used<br />

for the evaluation of the ovarian response.<br />

RESULTS: A total of 56 patients included in this study. The mean age of<br />

the women is 31.64.7 years. The reason of the salpingectomy as follows:<br />

hydrosalpinx 39.3% (n¼22), ruptured ectopic pregnancy 60.7% (n¼34).<br />

The clinical pregnancy rate per embryo transfer is 30%. The AFC and<br />

COH parameters were presented in Table.<br />

CONCLUSIONS: The present study suggests that salpingectomy is not<br />

associated with detrimental effects on the antral follicle count and ovarian<br />

response.<br />

O-68 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:30 AM<br />

Non-operated<br />

side<br />

P value<br />

Antral follicle count 5.914 6.113.7 NS<br />

No. of follicle ><strong>17</strong> mm in size 1.781.7 2.092.2 NS<br />

on the hCG day<br />

No. follicle >10-<strong>17</strong> mm in size 6.234.4 5.923.8 NS<br />

on the hCG day<br />

No. of total collected oocytes 6.665.5 6.<strong>17</strong>4.2 NS<br />

THE COMPARISON OF ICSI-ET OUTCOMES IN SURGICALY<br />

TREATED MINIMAL ENDOMETRIOSIS, ENDOMETRIOMA AND<br />

TUBAL FACTOR WITHOUT UNDERLYING ENDOMETRIOSIS.<br />

I. Guler, a A. Erdem, a M. Erdem, a N. Bozkurt, a M. Oktem, a M. F. Mutlu, b<br />

Y. Oguz, a F. Cevher. a a Obstetrics and Gynecology, Gazi University Faculty<br />

of Medicine, Ankara, Turkey; b Koru Hospital, Ankara, Turkey.<br />

OBJECTIVE: To assess the effect of laparoscopic surgery in patients with<br />

minimal endometriosis and endometrioma on the outcome of IVF-ET by<br />

comparing with the results of patients with tubal factor without underlying<br />

endometriosis that is confirmed by laparoscopy.<br />

DESIGN: Retrospective case-control study<br />

MATERIALS AND METHODS: From 2003 to 2013 retrospective data of<br />

patients with minimal endometriosis or endometrioma who were diagnosed<br />

and treated by laparoscopic surgery before ICSI-ET were reviewed for the<br />

study. In all patients with endometrioma, laparoscopic unilateral or bilateral<br />

cystectomy was performed. The control group was chosen from the patients<br />

with the diagnosis of tubal factor who were underwent laparoscopy and not<br />

diagnosed with underlying endometriosis. Patients with additional male factor,<br />

thyroid disease, PCO in each group were also excluded. Statistical significance<br />

was defined as p < 0.05.<br />

RESULTS: A total number of 365 cycles of 206 patients, including 92<br />

cycles of 47 patients with minimal endometriosis, 65 cycles of 30 patients<br />

with endometrioma and 208 cycles of 129 patients with tubal factor were<br />

analyzed. Mean ages of patients among three groups were comparable. Basal<br />

antral follicle count, total dose of gonadotropins, duration of stimulation,<br />

total number of retrieved oocytes, fertilization, embryo clivage and clinical<br />

pregnancy rates were found to be significantly different between groups.<br />

All results were summarized in the Table 1.<br />

CONCLUSIONS: Even if minimal endometriosis or endometrioma was<br />

treated by laparoscopy prior to IVF, clinical pregnancy rates tended to be<br />

lower in endometriosis than tubal factor infertility that was not associated<br />

with endometriosis. Laparoscopic removal of endometrioma was also related<br />

with lower ovarian response and less number of oocyte retrieval.<br />

TABLE 1. Comparison of ICSI outcomes among study and control groups<br />

Variable<br />

Minimal<br />

Endometriosis<br />

(n¼ 47, 92<br />

cycles)<br />

Supported by: Gazi University.<br />

Endometrioma<br />

(n¼30, 65<br />

cycles)<br />

Tubal<br />

Factor<br />

(n¼129,<br />

208 cycles)<br />

p<br />

value<br />

Age (years) 31.9 4.6 32.4 4.3 32.5 4.9 0.586<br />

Total basal antral 8.9 5.4 6.1 3.9 6.8 5.4 0.003<br />

follicle count<br />

(n)<br />

Day 3 Basal 6.7 2.8 8.8 4.4 8.4 8.2 0.125<br />

FSH level<br />

(IU/L)<br />

Duration of 10.6 2.8 11.1 2.9 10.1 2.6 0.044<br />

stimulation<br />

(day)<br />

Total dose of<br />

gonadotropins<br />

(IU)<br />

3400.31 <br />

1903.98<br />

4255.71 <br />

1601.<strong>21</strong><br />

3113.63 <br />

1424.09<br />

0.000<br />

Serum E2 on the<br />

day of hCG<br />

injection<br />

(pg/mL)<br />

Fertilization<br />

rate (%)<br />

Embryo cleavage<br />

rate (%)<br />

Total number of<br />

retrieved<br />

oocytes (n)<br />

Total number of<br />

retrieved M2<br />

oocytes (n)<br />

Clinical pregnancy<br />

rate (%)<br />

1635.52 <br />

1402.38<br />

1287.13 <br />

1002.42<br />

<strong>17</strong>41.19 <br />

1508.87<br />

0.186<br />

75.9 23.4 66.3 28.3 75.7 23.2 0.053<br />

91.2 <strong>21</strong>.8 85.7 30.0 94.6 14.0 0.026<br />

10.9 8.2 7.1 4.5 10.6 8.7 0.018<br />

8.2 6.6 5.8 3.8 8.1 6.6 0.062<br />

25.0 14.3 58.3 0.000<br />

O-69 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:45 AM<br />

MISDIAGNOSED SEPTATE UTERUS: DO WE REALLY NEED<br />

THREE DIMENSIONAL TRANSVAGINAL ULTRASOUND (3D-US)<br />

TO REEVALUATE ARCUATE UTERI DIAGNOSED BY<br />

HYSTROSALPINGOGRAPHY (HSG)? A. M. Abdelmagied, a,b<br />

M. A. Kamel, b A. M. Abuelhasan, b T. A. Farghaly, b A. A. Nassr, a,b<br />

S. A. Shazly, a,b M. H. Makarem. b a Obstetrics and Gynecology, Mayo Clinic,<br />

Rochester, MN; b Obstetrics and Gynecology, Women Health Hospital,<br />

Assiut University, Assiut, Egypt.<br />

OBJECTIVE: It is well known that septate uterus is a common cause of<br />

reproductive failure and its correction has been linked to better prognosis.<br />

Also, there is evidence that infertile women both with small and large septae<br />

report decreased implantation and pregnancy rates. Our objective was to<br />

assess the utility of 3D-US after HSG for the differentiation between arcuate<br />

and septate uterus in infertile women.<br />

FERTILITY & STERILITY Ò<br />

e27


DESIGN: Prospective cohort study<br />

MATERIALS AND METHODS: Infertile women with diagnosis of<br />

arcuate and septate uterus based on HSG were recruited. All women<br />

were examined by 3D-US on day 22 cycle to allow for better delineation<br />

of the uterine contour. The outer and inner fundal contours and the length<br />

of the fundal notch were examined by 3D-US in the mid-coronal view of<br />

the multi-planar and multi-slice display of the uterus. The final diagnosis<br />

of the anomalies was based on combined hysteroscopy/laparoscopy examination<br />

(HL), the gold standard,. In cases of septate uterus, HL was accompanied<br />

by hysteroscopic resection of the septum. ASRM description of the<br />

congenital uterine anomalies was followed to categorize the findings. For<br />

imaging and endoscopic analysis of the uterine morphology, standards in<br />

the literature were utilized to appropriately define each anomaly. To calculate<br />

the level of agreement between the diagnostic modalities, kappa statistics<br />

were utilized.<br />

RESULTS: Thirty eight women were included. Based on HL diagnosis,<br />

3 groups of congenital uterine anolamlies were diagnosed: arcuate uterus<br />

(n¼12,31.5%), septate uterus (n¼22,58%) and bicornuate uterus (n¼4,<br />

10.5%). 3D-US showed perfect diagnostic accuracy (100%) and perfect<br />

concordance (k ¼ 1.00, 95% CI ¼ 1.00-1.00) with HL in identifying all<br />

anomalies in the 3 groups while HSG had 68.4% accuracy and fair concordance<br />

(k ¼ 0.37, 95% CI ¼ 0.09-0.66) with HL. For septate uterus, 8 cases<br />

were misdiagnosed by HSG as arcuate uterus (40%) while for bicornuate<br />

uterus; HSG misdiagnosed all of them as septate. Septae missed by HSG<br />

were less than 2 cm in length and shorter (Median, <strong>17</strong> vs. 29 mm) than those<br />

correctly diagnosed.<br />

CONCLUSIONS: In infertile women, diagnosis of arcuate uterus based on<br />

HSG needs further evaluation by 3D-US in order to diagnose possible missed<br />

small septae. Being in perfect concordance with HL, 3D-US averts the need<br />

to endoscopy to differentiate between common uterine anomalies and offers<br />

the preoperative confirmation of the anomaly that should be surgically corrected.<br />

References:<br />

1. Salim R, Woelfer B, Backos M, Regan L, Jurkovic D. Reproducibility<br />

of threedimensional ultrasound diagnosis of congenital uterine anomalies.<br />

Ultrasound Obstet Gynecol 2003; <strong>21</strong>:578-82.<br />

2. Ludwin A, Pity_nski K, Ludwin I, Banas T, Knafel A. Two- and three--<br />

dimensional ultrasonography and sonohysterography versus hysteroscopy<br />

with laparoscopy in the differential diagnosis of septate,<br />

bicornuate, and arcuate uteri. J Minim Invasive Gynecol 2013; 20:90-9.<br />

3. Bermejo C, Martınez Ten P, Cantarero R, Diaz D, P_erez Pedregosa J,<br />

Barr_on E, et al. Three-dimensional ultrasound in the diagnosis of<br />

mullerian duct anomalies and concordance with magnetic resonance<br />

imaging. Ultrasound Obstet Gynecol 2010; 35:593-601.<br />

OBJECTIVE: To determine if there is discrepancy between the measurement<br />

of the mid fundal length of subtle incomplete uterine septum or arcuate<br />

uterine anomaly on hysteroscopy and on trans-vaginal 3D ultrasound scan<br />

(TV 3D US) with or without saline infusion sonohysterogram (SIH).<br />

DESIGN: Retrospective study<br />

MATERIALS AND METHODS: Two hundred sixty three patients who<br />

had a subtle incomplete uterine septum (30.0%) or arcuate uterine anomaly<br />

(70.0%) on hysteroscopy between 2010 and <strong>2015</strong> were studied. All patients<br />

had TV 3D US with and without SIH as part of their work up for infertility<br />

(65.0%), history of miscarriage (6.6%) or both (28.3%) and subsequently had<br />

a hysteroscopy with or without laparoscopy. The length of the mid fundal<br />

protrusion in the endometrial cavity was calculated on 3D TVUS with and<br />

without SIH. If the 2 values were available, a mean value was obtained; otherwise<br />

the one available value was used. At time of hysteroscopy the type of<br />

uterine anomalies according to ASRM classification (Class IV b or<br />

Class V) was documented and its mid fundal length was measured using<br />

the tip of a straight resectoscope loop. If the incomplete uterine septum or<br />

arcuate uterine anomaly was deemed clinically significant (length > 10<br />

mm) it was divided at the same session. We compared the findings on hysteroscopy<br />

and on 3D TVUS with and without SIH. Paired t-test and correlation<br />

analysis were used.<br />

RESULTS: Mean age was 32.6 + 5.2 years, mean body mass index was<br />

27.6 + 6.6 kg/m 2, and 59.7% of patients had primary infertility. Overall<br />

mean fundal length on TV 3D US (6.4 + 2.9 cm) was significantly lower<br />

than mean actual length on hysteroscopy (13.5 + 3.3 cm), p


HEALTH DISPARITIES 1<br />

O-73 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:15 AM<br />

O-72 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:30 PM<br />

COMPARISON OF OPERATIVE AND PERI-OPERATIVE<br />

RESULTS FOR ROBOTIC AND LAPAROSCOPIC<br />

MYOMECTOMIES. C. E. Miller, a K. Sasaki, b C. Steller, c S. Sulo, d<br />

A. Cholkeri-Singh. e a The Advanced IVF Institute, Naperville, IL; b Advanced<br />

Gynecologic Surgery Institute, Naperville, IL; c Advocate Lutheran General<br />

Hospital, Park Ridge, IL; d Center for Advanced Care, Russell Institute of<br />

Research and Innovation, Park Ridge, IL; e The Advanced Gynecologic Surgery<br />

Institute, Naperville, IL.<br />

OBJECTIVE: To compare operative and peri-operative outcomes between<br />

laparoscopic and robotic myomectomies.<br />

DESIGN: A review of prospectively maintained data in 2:1 matched cases<br />

of 96 laparoscopic and 48 robotic myomectomies. The cases were matched<br />

for number and size of fibroids removed, and timing of post-operative ultrasound.<br />

Medical charts were abstracted for demographics, operative and perioperative<br />

outcomes.<br />

MATERIALS AND METHODS: Women over the age of 18 who underwent<br />

a laparoscopic or robotic myomectomy from July 2009 through<br />

December 2014.<br />

RESULTS: The data was analyzed via Student t-test for continuous variables<br />

and Chi-square or Fischer’s exact for categorical variables. A<br />

Spearman correlation was performed for conversion rate with estimated<br />

blood loss and operative time. A linear regression was performed for predictive<br />

variables on number of post-operative fibroids found. No differences<br />

were identified between the two cohorts with respect to age,<br />

pregnancy history, prior pelvic surgery, trocar number, or number or<br />

size of fibroids removed (p>.05). There were no differences in operative<br />

time, estimated blood loss, complication or overnight admission rates. An<br />

increased conversion rate in the robotic group to laparoscopy correlated<br />

with an increased estimated blood loss (r¼0.237, p¼ .004) and operative<br />

time (r¼.143, p¼ .038), due to case complexity. No differences in postoperative<br />

ultrasound findings for fibroid number or size were found. Number<br />

of fibroids removed intra-opeatively was a significant predictor of<br />

increased post-operative fibroid number (b¼ .262, 95% CI .027-.113,<br />

p¼ .002) for either cohort.<br />

CONCLUSIONS: Laparoscopic and robotic myomectomies have similar<br />

outcomes in skilled surgical hands. The only difference was an increased<br />

conversion rate for complex cases in the robotic cohort. The number of<br />

post-operative fibroids did not differ, thus demonstrating that although robotics<br />

lacks haptic feedback, it does not limit a surgeon’s ability to identify<br />

and remove fibroids.<br />

Comparison of Operative and Peri-operative Outcomes<br />

Laparoscopic<br />

Myomectomy<br />

N¼96<br />

Robotic<br />

Myomectomy<br />

N¼48 p-value<br />

Operative Time (min), 162.2 (80.2) 169.3 (87.2) .627<br />

mean (SD)<br />

Estimated Blood Loss 126.4 (189) 111.9 (<strong>17</strong>1.9) .674<br />

(EBL) (mL), mean (SD)<br />

Conversion of Surgical 0 3 (6.3%) .035<br />

Approach, N (%)<br />

Complication Rate, N (%) 2 (2.1%) 3 (6.3%) .333<br />

Admission Rate, N (%) 15 (15.6%) 11 (23%) .284<br />

Number of fibroids removed 4.54 (4.0) 4.3 (4.2) .739<br />

intra-operatively, mean (SD)<br />

Number of fibroids<br />

on post-operative<br />

ultrasound, mean (SD)<br />

0.4 (1.2) 0.3 (0.78) .591<br />

EARLY LIFE EXPOSURE TO ESTROGEN-MIMICS INCREASE<br />

THE OCCURRENCE OF UTERINE FIBROIDS VIA EXPANSION<br />

OF MYOMETRIAL STEM CELL POPULATION. A. Mas, a<br />

L. Elam, a C. Walker, b C. Simon, c M. P. Diamond, a W. Thompson, d<br />

A. Al-Hendy. a a Department of Obstetrics and Gynecology, Georgia Regents<br />

University, Augusta, GA; b Department of Obstetrics and Gynecology, Georgia<br />

Regents University, Augusta, TX; c Fundacion Instituto Valenciano de Infertilidad<br />

IVI-Universidad de Valencia-INCLIVA, Paterna (Valencia), Spain;<br />

d Department of Obstetrics and Gynecology, Morehouse School of Medicine,<br />

Atlanta, GA.<br />

OBJECTIVE: Experimental animal studies have shown that exposure of<br />

target tissues to xenoestrogens (XEs) during sensitive periods of development<br />

can increase risk of disease in adult life, a process known as developmental<br />

reprogramming. Normal uterine myometrium is sensitive to ovarian<br />

hormones, making it a potential target for XEs. Developmental exposures<br />

to XEs have been shown to cause alterations the epigenetic programming<br />

of exposed myometrial cells, increasing risk for development of uterine<br />

fibroids. Our goal was to determine if exposure to XEs during critical periods<br />

of development altered the number and/or function of myometrial stem cells<br />

in exposed tissues, contributing to fibroid tumorigenesis.<br />

DESIGN: Laboratory research studies using a murine model of a human<br />

disease.<br />

MATERIALS AND METHODS: Eker rats (animal model for uterine fibroids)<br />

were exposed to XEs diethylstilbestrol (DES) or vehicle control<br />

(VEH) during postnatal days (PND) 10-12, and maintained until adulthood<br />

at 5 months of age. Myometrial cells were isolated from the uterine horns<br />

and cervix of adult Eker rats, and sorted by flow cytometry (FACS) using<br />

Stro1/CD44 myometrial stem cell markers. Additionally, human myometrium<br />

isolated from women without fibroids (MyoN) or women with fibroids<br />

(MyoF) was examined by immunohistochemistry (IHC) for Stro1/CD44-<br />

positive (stem) cells.<br />

RESULTS: At 5 months of age, morphological and functional differences<br />

were observed in the uterus of DES versus VEH treated rats. Quantification<br />

of Stro1/CD44 myometrial stem cells by FACS revealed that the percentage<br />

of myometrial stem cells in cervix (44.8%) and horns (23.4%) from DES<br />

exposed animal was significantly higher than cervix (35.3%) and horns<br />

(11.69%) from VEH rats (p%0.05). Quantification of Stro1/CD44 human<br />

myometrial stem cells by IHC demonstrated that the percentage of Stro1/<br />

CD44 positive cells was significantly higher in MyoF versus MyoN<br />

(p


with sufficient certainty and accuracy. The imaging characteristics are shared<br />

by the more common benign leiomyoma and potential serum biomarkers are<br />

not reliable. Developing a diagnostic tool for LMS is a high priority and<br />

would improve health care of women with suspicious myometrial mass.<br />

we wanted to identify a LMS specific promoter that can potentially drive a<br />

reporter gene to be expressed only in LMS cells (LMS-ON/Leiomyoma-<br />

OFF)<br />

DESIGN: Laboratory studies using human cells<br />

MATERIALS AND METHODS: We screened several modified adenoviruses<br />

encompassing various neoplasm-related promoters driving the<br />

expression of luciferase reporter gene The Adenovirus construct that<br />

showed highest expression potential in LMS, Ad-ANS-886, was selected<br />

for further evaluation. In-vitro, we transfected 3 cell types, which are<br />

SK-UT-1 (LMS), primary fibroid (1ry F) and myometrium (Myo F) cells<br />

with Ad-ANS-886 at multiplicity of infection (MOI) 1, 5. Luciferase transactivation<br />

was evaluated by both luciferase assay as well as xenogen camera<br />

imaging of cells 24 and 48 hours after transduction. Further in vivo<br />

evaluation of Ad-ANS-886 was performed by injecting adenovirus transfected<br />

LMS cells (5^6 cells/mouse) both subcutaneous (SC) and intrauterine<br />

(IU) in 6 weeks old female nude mice, followed by imaging the<br />

animals with Xenogen camera. Control animals were implanted with<br />

1ryF cells.<br />

RESULTS: Ad-ANS-886 transfects LMS cells much more readily than<br />

benign or normal myometrial cells as evidenced by luciferase assay that<br />

showed 6-fold higher bioluminescence in LMS vs 1ryF, MyoF at MOI 5<br />

with p < 0.0001. IVIS of LMS gave 100 fold higher total photon emission<br />

per second (TPE/S) compared to 1ryF , MyoF (p< 0.0001). In-vivo studies<br />

showed highly significant rise in TPE/S emission in LMS-based lesions vs<br />

1ryF-related lesions at 48 hours post cell implantation both in SC as well<br />

as IU locations (pA, p.G44D)] plasmid constructs containing a flag-tag<br />

and a green fluorescence protein (GFP). Transfected cells were selected<br />

with puromycin to generate stable cell populations which were characterized<br />

to determine the effect of MED12 somatic mutations on expression levels of<br />

Wnt4 and b-catenin.<br />

RESULTS: In western blot analyses using lysates from the above stable<br />

cells, we demonstrated that cells expressing mutant MED12 protein<br />

exhibited higher levels of Wnt4 and b-catenin as compared with cells expressing<br />

WT-MED12 protein. We also observed that the exogenous<br />

mutant-MED12 protein can induce Wnt4 and b-catenin in both cytosolic<br />

and nuclear compartments as compared with WT-MED12 protein. Cell-cycle<br />

analyses by FACScan assay showed that the expression of mutant-<br />

MED12 protein induced UtSMC cells into S-phase (22%, p


O-77 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:15 PM<br />

THE IMPACT OF ALCOHOL CONSUMPTION ON OVARIAN<br />

RESERVE IN REPRODUCTIVE-AGE AFRICAN AMERICAN<br />

WOMEN. L. Hawkins Bressler, a L. A. Bernardi, b P. D. De Chavez, b<br />

D. Baird, c M. Carnethon, d E. E. Marsh. e a Department of Obstetrics and<br />

Gynecology, Northwestern University, Chicago, IL; b Northwestern University,<br />

Chicago, IL; c Epidemiology, National Institute of Environmental Health<br />

Sciences, Research Triangle Park, NC;<br />

d Department of Epidemiology,<br />

Northwestern University, Chicago, IL;<br />

e Feinberg SOM - Northwestern<br />

University, Chicago, IL.<br />

OBJECTIVE: To evaluate the relationship between alcohol intake and<br />

anti-m€ullerian hormone (AMH) levels as a marker of ovarian reserve in<br />

reproductive-age African American women. A racial dichotomy exists in<br />

the literature implying a differential impact of alcohol on AMH whereby consumption<br />

does not affect AMH levels in white women (1-2) but is associated<br />

with lower AMH levels in black women (3).<br />

DESIGN: Cross-sectional analysis from an ongoing prospective study.<br />

MATERIALS AND METHODS: A total of 1,654 African American<br />

women (AAW) ages 23-35 completed questionnaires on the duration, frequency<br />

and amount of their alcohol consumption and provided serum for<br />

AMH measurement. Multivariable linear and logistic regression accounting<br />

for age, body mass index (BMI), hormonal contraception (HC), irregular<br />

menses and thyroid conditions were used to assess the relationship between<br />

alcohol intake and AMH. Binge drinking was defined as four or more drinks<br />

consumed on one occasion.<br />

RESULTS: Most reproductive-age AAW (73.5%) drink alcohol. Of these,<br />

33.9% report weekly or more frequent drinking in the last 12 months. The<br />

majority (74.1%) report binge drinking, of which more than a fourth<br />

(26.5%) report binge drinking multiple times per month in the last year.<br />

Among AAW who drink alcohol, 83.7% report their heaviest consumption<br />

occurred over the age of 19. Frequency of binge drinking appeared to demonstrate<br />

an inverse association with AMH and approached significance<br />

among AAW who binge drink twice weekly or more compared to never bingers<br />

(b¼-0.27, P¼0.07). Ever drinking, years of regular drinking, frequency of<br />

alcohol consumption, number of drinks per day, type of alcohol consumed<br />

and age at heaviest consumption were not associated with AMH (all P>0.2).<br />

CONCLUSIONS: Findings represent the most comprehensive assessment<br />

to date of patterns of alcohol consumption on AMH levels. Among reproductive-age<br />

AAW, alcohol intake does not appear to meaningfully impact AMH.<br />

Additional investigation in the infertile population including functional<br />

ovarian reserve testing would assess reproducibility and permit further characterization<br />

of the impact of binge drinking on ovarian reserve.<br />

References:<br />

1. Dolleman, M., Verschuren, W. M., Eijkemans, M. J., Dolle, M. E., Jansen,<br />

E. H., Broekmans, F. J., & van der Schouw, Y. T. Reproductive and<br />

lifestyle determinants of anti-Mullerian hormone in a large populationbased<br />

study. J Clin Endocrinol Metab, 2013;98(5):<strong>21</strong>06-<strong>21</strong>15.<br />

2. Nardo, L. G., Christodoulou, D., Gould, D., Roberts, S. A., Fitzgerald,<br />

C. T., & Laing, I. Anti-Mullerian hormone levels and antral follicle<br />

count in women enrolled in in vitro fertilization cycles: relationship<br />

to lifestyle factors, chronological age and reproductive history. Gynecol<br />

Endocrinol, 2007;23(8):486-493.<br />

3. Whitworth, K. W., Baird, D. D., Steiner, A. Z., Bornman, R. M., Travlos,<br />

G. S., Wilson, R. E., & Longnecker, M. P. Anti-Mullerian Hormone<br />

and Lifestyle, Reproductive, and Environmental Factors Among<br />

Women in Rural South Africa. Epidemiology <strong>2015</strong>;26(3):429-35.<br />

Supported by: National Institutes of Health Grant R<strong>21</strong>HD077479-01; National<br />

Institutes of Health Grant K12HD0501<strong>21</strong>; Northwestern University<br />

Women’s Reproductive Health (WRHR) Scholar Award; Harold Amos Medical<br />

Faculty Development Award; Robert Wood Johnson Foundation.<br />

O-78 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:30 PM<br />

PREGNANCY RATES AMONG AFRICAN-AMERICAN WOMEN<br />

UNDERGOING SINGLE EMBRYO TRANSFER (SET). T. Plowden, a<br />

A. Christy, a A. DeCherney, a J. Csokmay. b a NIH, Bethesda, MD;<br />

b WRNMMC, Bethesda, MD.<br />

OBJECTIVE: One of the greatest risks of Assisted Reproductive Technologies<br />

is iatrogenic multi-fetal gestation, which is associated with significant<br />

fetal and maternal morbidity and mortality. In recent years, there is a national<br />

impetus for IVF practices to move towards SET to decrease the rate of multiple<br />

pregnancies. Previous IVF literature has shown that African-American<br />

women undergoing IVF are less likely to become pregnant than their Caucasian<br />

counterparts. To date, no studies have specifically explored the success<br />

of SET in African-American women. Our objective was to explore clinical<br />

pregnancy (IUP with cardiac activity) and spontaneous abortion (SAB) rates<br />

in African-American women undergoing SET.<br />

DESIGN: This study was a retrospective analysis of patients undergoing<br />

treatment from Jan 2012-Dec 2014.<br />

MATERIALS AND METHODS: Our institution routinely recommends<br />

transfer of a single embryo for good prognosis women ( 0 was equal to or greater than .95 or equal to or less than<br />

.05. Separate estimates for effects were calculated for men and women within<br />

the same model and the correlation coefficient within a couple was calculated<br />

for all sub-scale scores. In addition to the effect of time, age, infertility and<br />

mental health treatment history (MHT) were added to the models. The impact<br />

of miscarriage was measured using three sub-scales from the Revised Impact<br />

of Miscarriage Scale (RIMS)(Huffman, Swanson, & Lynn, 2014): The RIMS<br />

is comprised of three sub-scales: 1) Isolation and guilt (I/G), 2) Loss of baby<br />

(LB), and 3) Devastating event (DE). Higher scores indicate greater impact.<br />

RESULTS: 302 couples data were eligible for analysis.The mean age of<br />

women was 32.62(SD ¼ 6.01) and 34.08( SD ¼ 6.77) for men, mean gestational<br />

age at miscarriage was 9.9 weeks (SD ¼ 3.08). Seventy-four couples<br />

had a history of infertility, 48% of women and 24% of men were positive<br />

for MHT. When controlling for baseline scores, infertility had no effect on<br />

I/G, DE, or LB scores over the course of a year. Women were no more<br />

FERTILITY & STERILITY Ò<br />

e31


likely to feel isolated or guilty than their fertile counter parts (pp of an effect<br />

for infertility was .48 in women and .63 in men). Men with MHT were<br />

more isolated/guilty during the first year than men without such history<br />

(mean parameter estimate for MHT ¼ .62, pp ¼ .99). Older men and<br />

women were less likely to view their miscarriage as the loss of a baby<br />

than younger men and women, though this did not reach significance.<br />

Time had a significant effect in reducing I/G, DE, and LB in women.<br />

Men’s initial impact remained constant over the course of the year except<br />

for devastation, which showed a small decrease. The couples scores were<br />

significantly correlated, albeit, the correlation coefficients were low ( r ¼<br />

.07 - .16, pp ¼ .96 -.99)<br />

CONCLUSIONS: During the first year of loss, the impact of miscarriage is<br />

not influenced by infertility. Men with a history of grief, depression, or anxiety<br />

may experience more isolation and guilt. Women see a decrease in impact<br />

over time, where as men only see a decrease in the devastation experienced.<br />

References:<br />

1. Swanson KM, Chen HT, Graham, JC, Wojnar DM, Petras A. Resolution<br />

of depression and grief during the first year after miscarriage: a randomized<br />

controlled clinical trial of couples-focused interventions. J<br />

Womens Health (Larchmt), 2009; 18:1245-57.<br />

2. Huffman CS, Swanson KM, Lynn M. Measuring the meaning of<br />

miscarriage: revision of the Impact of Miscarriage Scale. J of Nurs<br />

Measurement 2014; 22: 29-45.<br />

Supported by: Funding for the original study was provided to K.M.S. by<br />

the NIH, National Institute of Nursing Research, 5 R01 NR005343. Trial<br />

registration number: NCT00194844.<br />

O-80 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:30 AM<br />

THE PREVALENCE OF ABNORMAL MAMMOGRAMS IN OVUM<br />

RECIPIENTS DOES NOT CORRELATE WITH RECIPIENT<br />

AGE. S. E. Yerkes, a J. Rodriguez-Purata, b J. A. Lee, a<br />

M. C. Whitehouse, b M. Daneyko, a B. Sandler, b,c A. B. Copperman. a,c a Reproductive<br />

Medicine Associates of New York, New York, NY; b Reproductive<br />

Medicine Associates of New York, New York City, NY; c Obstetrics, Gynecology<br />

and Reproductive Science, Mount Sinai School of Medicine, New<br />

York, NY.<br />

OBJECTIVE: Ovum recipients undergo endometrial preparation with estrogen<br />

to mimic the natural menstrual cycle. Recipients R40 years are<br />

required to complete a mammogram, but younger patients are not. The study<br />

sought to compare if the prevalence of an abnormal mammogram in ovum<br />

recipients increases with ovum recipient age.<br />

DESIGN: Retrospective analysis.<br />

MATERIALS AND METHODS: Patients who had a mammogram prior to<br />

starting an ovum donation cycle from 2010 - 2014 were included. The following<br />

Breast Imaging Reporting and Data System (BIRADS) categories were used for<br />

reporting mammographic results: 0: incomplete (needs additional image evaluation<br />

and/or prior mammograms for comparison); 1: negative; 2: benign; 3:<br />

probably benign; 4: suspicious; 5: highly suggestive of malignancy; and 6:<br />

known biopsy (proven malignancy). Patients were segregated into two groups<br />

according to age (Group A:


for statistical significance. A logistical regression model was also utilized to<br />

identify significance of multi-variables.<br />

RESULTS: Responses (N¼1<strong>17</strong>) were returned. The female rate was 66%<br />

(N¼88). The majority,98%, were Christian (N¼112). The majority, 88%,<br />

were college educated (N¼102) with 39% (N¼40) with some post-graduate<br />

education. Sixty-eight percent (N¼79) had no knowledge of PGS prior to<br />

their IVF cycle; however, after provider education, ninety-two percent<br />

(N¼108) correctly identified that PGS was elective and 93% (N¼109) reported<br />

their knowledge was sufficient to make an informed decision to accept<br />

or decline PGS. The additional cost of screening (P¼0.004), the provider information<br />

and influence (P¼0.0001), and social support or acceptance from<br />

partner, family, friends (P¼0.03), were the three variables affecting the decision.<br />

In a logistical regression model, additional cost (P¼0.003) and explanation<br />

from the provider (P¼0.0003) were the only significant<br />

determinants influencing the decision. Religious and ethical beliefs were<br />

important (P¼0.20), but not statistically significant nor was information<br />

regarding the disposition of abnormal embryos (P¼0.18) or the concerns<br />

regarding cryopreservation and transfer at a later date (P¼0.79).<br />

CONCLUSIONS: PGS use as an adjuvant therapy to aid in selection of the<br />

embryo with the greatest potential for success, has become readily available<br />

in many IVF clinics. Researchers have spent time, resources, and effort on the<br />

clinical benefits and limitations of PGS, yet little research on the patient<br />

perspective, knowledge, education, and awareness regarding the options<br />

and ultimate decision making capabilities of the patient. This is the first study<br />

to the authors knowledge to identify and assess the determinants of the patient<br />

decision making process when presented with the choice of PGS in a<br />

given IVF treatment cycle. Several factors contribute to the patient perceived<br />

determinants when choosing to accept or decline PGS, including cost, religious/ethical<br />

values, social influence, and the past experience of the patient.<br />

O-83 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:15 PM<br />

LESSENING THE BURDEN OF CARE: STEPPED VS. FIXED<br />

ESTRADIOL PROTOCOL FOR FROZEN EMBRYO<br />

TRANSFER. A. B. Tepper, a C. M. Bergh, a T. Molinaro, a,b<br />

P. A. Bergh. a,b a Reproductive Medicine Associates of New Jersey, Basking<br />

Ridge, NJ; b Obstetrics, Gynecology & Reproductive Sciences, Rutgers-<br />

Robert Wood Johnson Medical School, New Brunswick, NJ.<br />

OBJECTIVE: Simplification of cycle medication protocols without a<br />

reduction in efficacy is a key priority in lessening the patient’s burden of<br />

care. This study sought to compare frozen embryo transfer (FET) cycle outcomes<br />

between estradiol therapies using a 1mg stepped protocol or 2mg fixed<br />

protocol at cycle start.<br />

DESIGN: Retrospective database study.<br />

MATERIALS AND METHODS: Analysis was performed on 3,984 patients’<br />

electronic medical records at Reproductive Medicine Associates of<br />

New Jersey (RMANJ) from <strong>October</strong> 9, 1999 to April 30, <strong>2015</strong>. FET cycles<br />

using autologous oocytes and blastocyst transfers with a starting dose of<br />

either 1 or 2mg were included in this study. Gestational Carrier cycles<br />

were excluded. Chi-square and t-test were performed for our normally<br />

distributed population. The study was powered to detect a 5% difference between<br />

groups. Logistic regression was performed to control for confounders.<br />

RESULTS: There were small statistical significant differences in endometrial<br />

thickness, length of cycle and use of CCS. Implantation rates were<br />

similar between both groups. When controlling for age, CCS use, number<br />

of embryos transferred, and method of cryopreservation, there were no significant<br />

differences in clinical pregnancy (OR 0.96; 95%CI 0.79 - 1.15)<br />

nor clinical loss rate (OR 0.98; 95%CI 0.79 - 1.22).<br />

Cycle Characteristics Stratified By Starting Estradiol Dose.<br />

1mg. estradiol 2mg. estradiol<br />

p value<br />

Sample Size (n) 3102 882<br />

Oocyte Age (years) 34.2 34.2 p ¼ 0.887<br />

Endometrial Thickness (mm) 9.9 9.6 p ¼ 0.000<br />

Length of Cycle (days) 12.5 11.8 p ¼ 0.000<br />

Use of Comprehensive 53.5 57.4 p ¼ 0.041<br />

Chromosome Screening<br />

(CCS) (%)<br />

Clinical Pregnancy Rate (%) 76.7 77.0 p ¼ 0.835<br />

Implantation Rate (%) 68.7 69.5 p ¼ 0.618<br />

CONCLUSIONS: Clinical pregnancy and loss rates are not impacted by<br />

using a 2 mg fixed protocol for a FET cycle. Using a standardized fixed protocol<br />

may reduce provider ordering and patient administration errors, number<br />

of monitoring visits, and improve patient compliance. As described, a fixed<br />

FET protocol may reduce the patient’s burden of care without compromising<br />

outcomes.<br />

O-84 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:30 PM<br />

GAY SURROGACY-THE QUANDRY OF ACCESSING VERIFIABLE<br />

FACTS. D. Smotrich, a A. Botes, a X. Wang, a M. Gaona, b D. Batzofin. a<br />

a IVF, LaJolla IVF, La Jolla, CA; b Embryology/Andrology, LaJolla IVF, La<br />

Jolla, CA.<br />

OBJECTIVE: To bring attention to the lack of verifiable data pinpointing<br />

clinics that are accommodating to LGBT patients as well as being able to<br />

easily obtain LGBT friendly legitimate and objective statistical data<br />

regarding the use and success of gestational surrogate (GS) egg donor<br />

(ED) cycles to create their families.<br />

DESIGN: Retrospective analysis.<br />

MATERIALS AND METHODS: Retrospective review of IVF laboratory<br />

database and patients’ charts was conducted on 529 consecutive fresh GS/ED<br />

cycles performed for gay couples from January 2004 - December 2013 (a ten<br />

year period) at a private clinic in California. After May 2005 all gamete providers<br />

were subject to FDA regulations. Standard stimulation protocols,<br />

monitoring, egg retrievals and embryo transfers were performed. Embryos<br />

were created with ICSI (100%), PGD/S (75%) and blastocyst transfers<br />

(80%). A review of a subset of 200 randomly selected gay intended parents’<br />

charts from the 529 cycles was undertaken to analyze documented comments<br />

by patients regarding their country of origin, from where they learned about<br />

GS/ED as a treatment option and what additional information would have<br />

been useful in their decision making.<br />

RESULTS: From January 2004 - December 2013, 529 fresh GS/ED cycles<br />

were performed for gay couples. A clinical pregnancy was confirmed in 454<br />

GS for an 86% clinical pregnancy rate. 80% of the gay couples achieved a<br />

live birth after one GS/ED cycle. Data obtained from the clinical chart review:<br />

patients traveled from 54 countries and 47 US states; 80% learned<br />

about GS/ED from the internet and media outlets, 7% from GS/ED agencies,<br />

4% from peers and 9% from other patients. 96% stated some official statistics<br />

in a report (along the lines of the CDC Clinic Success Rates Report) as being<br />

the most authoritative guidance missing from their data search.<br />

CONCLUSIONS: In experienced hands, gestational surrogacy/egg donation<br />

is a highly effective treatment for gay couples in terms of family building.<br />

These patients deserve and should be able to rely on official statistical<br />

data in order to make a logical and informed decision in regards to their<br />

choice of the most appropriate treatment facility for their needs, just as other<br />

patients researching ART treatments are afforded.<br />

EARLY PREGNANCY 1<br />

O-85 Monday, <strong>October</strong> 19, <strong>2015</strong> 11:15 AM<br />

THE EFFECT OF A PUBLICLY FUNDEDED NORTH AMERICAN<br />

IVF PROGRAM WITH MANDATED SINGLE EMBRYO TRANSFER<br />

ON MATERNAL ANTENATAL ADMISSION RATES. M. Dahan, a<br />

T. Shaulov, b S. Belisle. c a McGill University, Montreal, QC, Canada; b McGill<br />

University Health Centre, Montreal, QC, Canada; c University of Montreal,<br />

Montreal, QC, Canada.<br />

OBJECTIVE: To determine the effect of a drastic drop in the multiple<br />

pregnancy rate on antenatal admission of women who underwent IVF.<br />

DESIGN: A retrospective study was performed by analyzing data concerning<br />

IVF cycle outcomes and hospital admission in Quebec. Single payer government<br />

health care with IVF coverage for all women enabled to track<br />

outcome details which were gathered from the Ministry of Health and Social<br />

Services ‘‘Pro-Assis’’ database and the hospital admission database (MED-<br />

ECHO). The government program has mandated single embryo transfer in<br />

most cases.<br />

MATERIALS AND METHODS: Data is presented by financial year, and<br />

involves more than 250 000 pregnancies. The government began covering<br />

IVF in august 2010. Statistical analysis was performed using Chi squared<br />

tests, odds ratios or correlation coefficients. Data is available from<br />

FERTILITY & STERILITY Ò<br />

e33


2009-2010, pre-program, to 2012-2013, the first year to reflect conceptions<br />

achieved exclusively under government mandate.<br />

RESULTS: The proportion of live births that are a result of IVF has<br />

increased by 64% since 2009-2010, from 1.23% to 2.02% (p


(OR¼1.36 [1.09-1.69]), while it had no effect on EP in antagonist cycles<br />

(OR¼0.99 [0.84-1.16], interaction test p¼0.01).<br />

CONCLUSIONS: The GnRH antagonist protocol was associated with<br />

elevated odds of EP as compared with the GnRH agonist protocol in fresh<br />

autologous cycles. These findings highlight a possible role for extrapituitary<br />

GnRH on the uterine environment and the need to account for multiple<br />

maternal and treatment factors when evaluating the risk of EP after IVF.<br />

O-88 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:00 PM<br />

DON’T DISCRIMINATE: EVEN A DISCRIMINATORY ZONE AS<br />

HIGH AS 4000 CAN RESULT IN INTERRUPTION OF AN<br />

IUP. K. T. Barnhart, a M. D. Sammel, b A. Singer, a A. Reid, a<br />

L. Taylor, b S. Senapati. a a Obstetrics and Gynecology, University of Pennsylvania<br />

Perelman School of Medicine, Philadelphia, PA; b Biostatistics and<br />

Epidemiology, University of Pennsylvania Perelman School of Medicine,<br />

Philadelphia, PA.<br />

OBJECTIVE: The discriminatory zone (DZ) is the level of hCG above<br />

which one can be assured the pregnancy outcome will be ectopic pregnancy<br />

(EP) or miscarriage (SAB) if a normal intrauterine pregnancy (IUP) is not<br />

visualized by ultrasound (US). In this study, we assess how often the use<br />

of a DZ may misclassify an IUP as a non-viable pregnancy.<br />

DESIGN: Retrospective population-based cohort study.<br />

MATERIALS AND METHODS: 4008 women with symptomatic first<br />

trimester pregnancies of unknown location (PUL, defined as no intrauterine<br />

or extrauterine gestational sac, yolk sac, or embryo) were evaluated between<br />

January 1990 and August 2009. Subjects were stratified based on initial hCG<br />

values at proposed DZ levels (R2000, R3000, R4000 mIU/mL) and final<br />

outcomes of IUP, SAB, and EP. Subjects were further stratified based on<br />

ultrasound findings that were suspicious for IUP (intrauterine echogenic<br />

sac-like structure), suspicious for EP (adnexal mass or extrauterine sac-like<br />

structure), or no ultrasound findings suspicious for IUP or EP. The proportion<br />

of PUL that went on to be diagnosed as definitive IUP was computed for each<br />

DZ threshold.<br />

RESULTS: The median hCG level at presentation for care was 751, IQR:<br />

198 - 2690 mIU/mL. The average age was 26.4 (sd¼6.9), 70.9% were<br />

black and 15.6% were white. Final outcome was EP in 22%(N¼860),<br />

IUP in 26%(N¼1054), and SAB in 52%(N¼2094) of the population.<br />

30% had a previous pregnancy, 24% had a history of SAB, and 1% had<br />

2 or more prior EP. 19-31% of the PUL population presented with an<br />

hCG level above a DZ, depending on threshold (Table 1). Within this population,<br />

most subjects were eventually diagnosed with a SAB. For those<br />

subjects not diagnosed with a SAB, an ultrasound finding of suspected<br />

IUP was more likely to result in a definitive IUP than an ultrasound finding<br />

of suspected EP was to result in a definitive EP (74% vs. 50%, p or ¼ 3000 mIU/mL N¼923 (23%) IUP: 255 (28%)<br />

EP: 187 (20%)SAB: 481 (52%)<br />

hCG > or ¼ 4000 mIU/mL N¼763 (19%) IUP: 198 (26%)<br />

EP: 154 (20%)<br />

SAB: 411(54%)<br />

IUP 253 (74%) 73 (<strong>21</strong>%) 15 (5%) 88 (7.2%)<br />

EP 9 (3%) 125 (47%) 132 (50%)<br />

IUP 198 (78%) 44 (<strong>17</strong>%) 13 (5%) 57 (6.2%)<br />

EP 7 (4%) 82 (44%) 98 (52%)<br />

IUP 155 (78%) 31 (16%) 12 (6%) 43 (5.6%)<br />

EP 7 (5%) 66 (43%) 81 (52%)<br />

FERTILITY & STERILITY Ò<br />

e35


O-90 Monday, <strong>October</strong> 19, <strong>2015</strong> 12:30 PM<br />

FIRST TRIMESTER PREGNANCY LOSS FOLLOWING FRESH<br />

AND FROZEN IN VITRO FERTILIZATION CYCLES. H. Hipp, a,b<br />

S. Crawford, b J. F. Kawwass, a,b J. Chang, b D. M. Kissin, b<br />

D. J. Jamieson. b a Gynecology and Obstetrics, Emory University School of<br />

Medicine, Atlanta, GA;<br />

b Centers for Disease Control and Prevention,<br />

Atlanta, GA.<br />

OBJECTIVE: To determine if there is an increased risk of first trimester<br />

pregnancy loss in frozen embryo transfer cycles as compared to fresh cycles<br />

following in vitro fertilization (IVF).<br />

DESIGN: Retrospective cohort study using data from Centers for Disease<br />

Control and Prevention National ART Surveillance System for 2007-2012.<br />

MATERIALS AND METHODS: Multivariable log binomial regression,<br />

stratified by maternal age (< 30, 30-34, 35-37, 38-40, >40 years), was performed<br />

to compare age- specific risk of first trimester pregnancy loss between<br />

fresh and frozen embryo transfers. The main analysis included fresh cycles<br />

(n¼203,970) and frozen cycles for which maternal age at oocyte retrieval<br />

could be ascertained (n¼45,660) that resulted in a pregnancy. A subgroup<br />

analysis was performed to compare miscarriage risks of similar quality<br />

embryos between fresh (n¼203,970) and frozen cycles (n¼7,885). Frozen<br />

cycles for the subgroup analysis were restricted to those for which their originating<br />

fresh cycle had not had an embryo transfer (i.e. the embryo(s) transferred<br />

during the frozen cycle were the first embryos transferred from the<br />

originating retrieval).<br />

RESULTS: There was an increased risk of first trimester loss in pregnancies<br />

achieved after frozen embryo transfers versus fresh cycles and the<br />

adjusted relative risk remained significant among women less than 37 years<br />

of age:


Supported by: NIH/NICHD R25 HD075737, U10 HD27049, U10<br />

HD38992, U10 HD055925, U10 HD39005, U10 HD33<strong>17</strong>2, U10<br />

HD38998, U10 HD055936, U10 HD055942, and U10 HD055944, U54-<br />

HD29834, U10HD03005-0851, U10HD055925-02W1.<br />

O-93 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 11:45 AM<br />

OVARIAN TUMOR RISK IN WOMEN AFTER ASSISTED REPRO-<br />

DUCTIVE THERAPY (ART); 2.2 MILLION PERSON YEARS OF<br />

OBSERVATION IN GREAT BRITAIN. A. G. Sutcliffe, a<br />

C. L. Williams, a M. E. Jones, b A. J. Swerdlow, b M. C. Davies, c<br />

I. Jacobs, d B. J. Botting. a a Institute of Child Health, University College London,<br />

London, United Kingdom; b Institute of Cancer Research, Sutton, Surrey,<br />

United Kingdom; c Reproductive Medicine Unit, University College<br />

Hospital, London, United Kingdom; d University of New South Wales, Sydney,<br />

Australia.<br />

OBJECTIVE: To determine the risk of ovarian cancer, including malignant<br />

and borderline ovarian tumors, in women who have been exposed to<br />

ART.<br />

DESIGN: Records from the Human Fertilisation & Embryology Authority<br />

(HFEA) of all women who had ART in Britain between 1991-2010, were<br />

linked to the National Health Service Central Registers (NHSCR) for England,<br />

Wales and Scotland to obtain follow up for cancer outcomes, deaths<br />

and emigrations. Reporting to the HFEA is mandatory.<br />

MATERIALS AND METHODS: Cancer incidence in the cohort was stratified<br />

by age and calendar period and compared with expectations derived<br />

from annual age-specific national rates over the same period. Data were<br />

also stratified for potential mediating/moderating factors such as repeated exposures,<br />

age at first exposure, parity and subfertility diagnoses. Trends across<br />

categories were evaluated using Poisson regression.<br />

RESULTS: With 8.8 years average follow-up, 386 ovarian cancers<br />

occurred in 255,786 women. An increased risk of developing an ovarian cancer<br />

was observed in the cohort (standardized incidence ratio (SIR) 1.37; 95%<br />

CI 1.24-1.51). No increased risk was found with increasing number of cycles<br />

of ART (P trend ¼0.80). Increasing risk was found with decreasing parity<br />

(P trend ¼0.002), with women who had no live births by the end of treatment<br />

being at greatest risk (SIR 1.54; 95%CI 1.34-1.76). Risk was increased for<br />

women with a ‘female factor’ cause of infertility (SIR 1.62; 95%CI 1.42-<br />

1.84), especially endometriosis (SIR 2.35; 95%CI 1.80-3.07), but ‘male factor’<br />

only infertility was not associated with risk (SIR 1.05; 95%CI 0.86-1.28).<br />

Younger age at starting ART carried greater cancer risk (P trend


action. While additional studies to improve the efficacy of the approach to<br />

100% are needed, this study supports further investigations of the approach<br />

as a method of nonsurgical permanent female contraception.<br />

Supported by: Bill and Melinda Gates Foundation OPP1025233,<br />

P51OD011092, P51 OD011133, U54-055744-07.<br />

O-96 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:30 PM<br />

URINARY PHTHALATE METABOLITE CONCENTRATIONS<br />

WERE ASSOCIATED WITH PREGNANCY LOSS AMONG WOMEN<br />

CONCEIVING WITH MEDICALLY ASSISTED<br />

REPRODUCTION. C. Messerlian, a B. Wylie, b P. Williams, c<br />

J. B. Ford, a M. Keller, a A. M. Calafat, d R. Hauser. a,b a Department of Environmental<br />

Health, Harvard T. H. Chan School of Public Health, Boston, MA;<br />

b Department of Obstetrics and Gynecology, Massachusetts General Hospital,<br />

Boston, MA; c Department of Biostatistics, Harvard T. H. Chan School of<br />

Public Health, Boston, MA; d Centers for Disease Control and Prevention, Atlanta,<br />

GA.<br />

OBJECTIVE: Several phthalates are reproductive toxicants in animals and<br />

human occupational exposure has been associated with decreased pregnancy<br />

rates. We examined the association of pregnancy loss in relation to urinary<br />

phthalate metabolites among women undergoing in-vitro fertilization and<br />

ovarian stimulation with and without intrauterine insemination.<br />

DESIGN: A prospective cohort of 249 women conceiving 294 pregnancies<br />

from the Environment and Reproductive Health Study (EARTH) at the<br />

Massachusetts General Hospital Fertility Center.<br />

MATERIALS AND METHODS: Four di(2-ethylhexyl) phthalate (DEHP)<br />

metabolites in urine were measured by the Centers for Disease Control and<br />

Prevention at two time points per woman during each fertility treatment cycle<br />

and the geometric mean urinary DEHP metabolite concentration was calculated.<br />

We estimated adjusted risk ratios (RRs) and 95% confidence intervals<br />

(CIs) for both biochemical and total pregnancy loss (


CONCLUSIONS: A nanocaged formulation of highly insoluble ORG<br />

9935 improved bioavailability but did not maintain drug concentrations in<br />

the range needed to prevent oocyte maturation. While the extended-release<br />

preparation of NC35 prevented oocyte maturation, this improved PD performance<br />

was associated with serum levels that varied widely and exceeded the<br />

desired peak by 2-fold at the time of oocyte recovery. While promising,<br />

further enhancements of the PK profile will be needed to improve feasibility<br />

of this approach as a contraceptive in women.<br />

Supported by: NIH U54 HD055744, P51OD011092.<br />

O-99 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 11:45 AM<br />

CANIDATE GENES AND PATHWAYS MEDIATING HORMONAL<br />

REGULATION OF THE MACAQUE FALLOPIAN<br />

TUBE. K. R. Bond a O. D. Slayden. b,c a Division of Reproductive & Developmental<br />

Sciences, Oregon National Primate Research Center, Beaverton,<br />

OR; b Oregon National Primate Research Center, Beaverton, OR; c Department<br />

of Obstetrics and Gynecology, Oregon Health & Science University,<br />

Portland, OR.<br />

OBJECTIVE: The fallopian tube plays an important role in gamete transport<br />

and fertilization, and may provide a target for reversible contraception.<br />

The aim of this study was to identify differentially expressed, hormonally<br />

regulated, genes and their related biological pathways in the rhesus macaque<br />

fallopian tube during the menstrual cycle.<br />

DESIGN: Affymetrix array analysis of macaque fallopian tube.<br />

MATERIALS AND METHODS: Fifteen adult cycling rhesus macaques<br />

were ovariectomized and treated sequentially with estradiol (E2) and then<br />

E2 plus progesterone (P) in order to create artificial menstrual cycles. Fallopian<br />

tubes were collected during the artificial proliferative phase (PP; E2<br />

treatment alone), secretory phase (SP; E2 + P treatment), and the menstrual<br />

phase (MP) cycle. RNA was isolated from the samples and analyzed using<br />

Affymetrix GeneChip Ò Rhesus Macaque Genome Arrays. Transcript expression<br />

for select targets were validated by real-time PCR with macaque specific<br />

TaqMan Ò primer/probe sets and presented relative to expression of ribosomal<br />

(S10) RNA expression.<br />

RESULTS: Affymetrix analysis revealed 6297 genes that were significantly<br />

regulated by steroid hormones (P


are four different PTGERs, our genomic data revealed that PTGER3 and<br />

PTGER2 are the predominant PGE2 receptors expressed in the macaque<br />

periovulatory follicle. Moreover, PTGER3 and PTGER2 mRNA levels<br />

both increase significantly 12 hr after an ovulatory stimulus. Thus, the objective<br />

of this study was to determine if PTGER3 is also directly involved in<br />

C-OE or follicle rupture in rhesus monkeys.<br />

DESIGN: Prospective non-human primate experimental study.<br />

MATERIALS AND METHODS: Cumulus oocyte complexes (COCs)<br />

aspirated from rhesus macaques undergoing an ovarian stimulation protocol<br />

were cultured (n¼13-18/group; 3 experiments) in the presence or absence of<br />

the PTGER3 agonist sulprostone (10 ng/mL) or PGE2 (500 ng/mL; positive<br />

control). Bright field images were obtained at 0 hr and 30 hr and immunofluorescent<br />

staining for hyaluronic acid (HA) expression (a marker for C-OE)<br />

was performed following COC collection at 30 hr. In rhesus monkeys undergoing<br />

a controlled ovulation protocol (n¼4/group), the dominant follicle was<br />

injected with either the PTGER3 antagonist (L798106; 250 ng/follicle) or<br />

vehicle (0.005% DMSO in saline) at the time the animals received an ovulatory<br />

stimulus. Laparoscopy was performed 72 hr later to check for the presence<br />

or absence of an ovulatory stigmata. Luteal phase length and circulating<br />

progesterone levels were also assessed.<br />

RESULTS: While PGE2 directly induced C-OE in vitro, sulprostone treatment<br />

of macaque COCs did not result in cumulus expansion or HA synthesis.<br />

Despite the absence of an effect on C-OE, intrafollicular injection of the<br />

PTGER3-selective antagonist L798106 in the preovulatory follicle at the time<br />

the animal received an ovulatory stimulus prevented (4 out of 4) the appearance<br />

of a typical rupture site relative to controls. Additionally, antagonist treated animals<br />

had an average luteal phase length of 18 days and an average peak progesterone<br />

level of 3.47ng/mL, which was not significantly different from controls.<br />

CONCLUSIONS: These findings demonstrate that PTGER3 plays an<br />

important role in ovulation, but not C-OE, in primates.<br />

Supported by: NIH T31HD007133 (JB), P51OD011092, and<br />

U54HD55744 (JDH, RLS).<br />

O-102 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:30 PM<br />

RATES OF INTRAUTERINE DEVICES (IUD) UTILIZATION AND<br />

STERILIZATION IN WOMEN OF REPRODUCTIVE<br />

AGE. B. Tang, K. McCarthy, S. Eid, G. Strachan, M. Tjoa,<br />

A. Pohlmeier, B. Howard. Teva Pharmaceutical, Frazer, PA.<br />

OBJECTIVE: While there are a myriad of contraceptive choices available<br />

to women, intrauterine devices (IUD) and female sterilization are two of the<br />

most effective methods. The objective was to evaluate trends of IUD insertion<br />

and sterilization rates in real world practice over the recent period of 2010 to<br />

2013.<br />

DESIGN: A retrospective study using US insurance administrative database<br />

was conducted.<br />

MATERIALS AND METHODS: Truven’s MarketScan Treatment Pathways<br />

3.0 tool was used to analyze the rate of IUD insertion and sterilization<br />

procedures among women from 2010 through 2013. Descriptive data on contraceptive<br />

use and age group distribution was evaluated using R 2 to determine<br />

the probability of change in rates predicted against time.<br />

RESULTS: A total of 530,6<strong>21</strong> IUD insertions and 95,650 women who<br />

chose sterilization were included in this study. While the overall trend<br />

increased for IUD insertions over the analysis period, we observed the utilizations<br />

of both IUD and sterilization decreased slightly in 2013 (R 2 ¼0.707).<br />

The greatest proportions of IUD and sterilization users were between the ages<br />

of 18-34 followed by the 35-44 age group. From 2010 to 2013, the proportion<br />

of IUD insertions in women aged 45-54 increased from 0.9% to 5.0%.<br />

CONCLUSIONS: Overall, IUD utilization rates of increased over time,<br />

while sterilization rates remained stagnant.<br />

Supported by: This study was Supported by Teva Pharmaceutical.<br />

MENTAL HEALTH<br />

O-103 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 11:15 AM<br />

QUALITY OF PARENTING, MOTHER AND CHILD WELLBEING<br />

AND ‘DADDY TALK’ IN SINGLE PARENT FAMILIES FORMED<br />

THROUGH THE USE OF DONOR INSEMINATION. S. Zadeh, a<br />

T. Freeman, b S. Golombok. c a Centre for Family Research, University of<br />

Cambridge, Cambridge, United Kingdom; b University of Cambridge, Cambridge,<br />

United Kingdom; c Centre for Family Research, Cambridge, United<br />

Kingdom.<br />

OBJECTIVE: The study examined parenting quality, mother and child<br />

wellbeing and communication in single parent families formed through the<br />

use of donor insemination (DI).<br />

DESIGN: Semi-structured interviews and questionnaires were administered<br />

to mothers at home.<br />

MATERIALS AND METHODS: Data were obtained from 51 single<br />

mothers and a comparison group of 52 partnered mothers with at least one<br />

child aged 4-9 years. Mothers were interviewed using an adaptation of a<br />

semi-structured interview designed to assess quality of parenting (Quinton<br />

& Rutter, 1988) and to establish thoughts and feelings about motherhood.<br />

Participants completed The Trait Anxiety Inventory (Spielberger, 1983), Edinburgh<br />

Depression Scale (Thorpe, 1993) and the short form of the Parenting<br />

Stress Index (Abidin, 1990). The Strengths and Difficulties Questionnaire<br />

(SDQ) (Goodman, 1997) was administered to assess the presence of children’s<br />

emotional and behavioural difficulties. Comparisons of parenting<br />

quality and mothers’ wellbeing between the two family types were conducted<br />

using multivariate analyses of covariance (MANCOVAs). Children’s psychological<br />

adjustment was compared between family types using analyses of<br />

variance (ANCOVAs). Hierarchical regression analysis was used to examine<br />

factors associated with child adjustment problems within the solo mother<br />

families. Qualitative analysis was used to supplement statistical findings.<br />

RESULTS: All of the single mothers described their use of DI in ambivalent<br />

terms. Most (n¼34, 67%) attributed their decision to use fertility treatment<br />

to the lack of a suitable partner. Several (n¼28, 55%) expressed<br />

concern about their child’s lack of a father. Different strategies were<br />

described by mothers in discussing this issue with their child. There were<br />

no significant differences in parenting quality or maternal wellbeing between<br />

the solo mother and two-parent families. In terms of children’s adjustment,<br />

there was no significant difference between family types for the total score<br />

of the SDQ. Within the solo mother families, higher levels of financial difficulties<br />

(F (2, 41) ¼ 5.00, p ¼ .01) and higher levels of parenting stress (F (7,<br />

36) ¼ 2.55, p ¼ .03) were each associated with higher levels of children’s<br />

emotional and behavioural problems.<br />

CONCLUSIONS: Findings suggest that despite mothers’ reported preferences<br />

and concerns, children in single parent DI families are no more likely<br />

to experience emotional or behavioural difficulties than their two-parent<br />

counterparts. Results indicate that single women seeking DI should not raise<br />

specific concerns for clinicians with regards to psychological wellbeing,<br />

quality of parenting, or child adjustment. As found in non-assisted single<br />

parent families, the present study suggests that family process variables are<br />

more influential in children’s adjustment than is family structure. ’Daddy<br />

talk’ nevertheless remains central.<br />

Frequency and Rates of Utilization of IUDs and Sterilization Over Time (Truven Marketscan Databases).<br />

Year 2010 2011 2012 2013 Total<br />

IUD (n, rate %) 109,504 (0.56) 129,457 (0.67) 151,699 (0.78) 139,961 (0.72) 530,6<strong>21</strong><br />

Age %<strong>17</strong> (%) 1.2 1.0 0.5 0<br />

Age 18-34 (%) 66.7 66.2 66.3 63.6<br />

Age 35-44 (%) 31.2 30.4 29.6 31.4<br />

Age 45-54 (%) 0.9 2.4 3.5 5.0<br />

Sterilization<br />

24,305 (0.12) 25,827 (0.13) 25,708 (0.13) 19,810 (0.10) 95,650<br />

(n, rate%)<br />

Age %<strong>17</strong> (%) 0 0 0 0<br />

Age 18-34 (%) 58.4 58.4 58.8 57.8<br />

Age 35-44 (%) 41.3 41.1 40.6 41.4<br />

Age 45-54 (%) 0.3 0.4 0.06 0.8<br />

e40 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


References:<br />

1. Quinton, D., & Rutter, M. Parenting Breakdown: The making and<br />

breaking of intergenerational links. Aldershot, UK: Avebury Gower<br />

Publishing, 1988.<br />

2. Spielberger, C. The Handbook of the State-Trait Anxiety Inventory.<br />

Palo Alto, CA: Consulting Psychologists Press, 1983.<br />

3. Thorpe, K. A study of the use of the Edinburgh Postnatal Depression<br />

Scale with parent groups outside the postpartum period. Journal of<br />

Reproductive and Infant Psychology, 1993;11:119-125.<br />

4. Abidin, R. Parenting Stress Index Test Manual. Charlottesville, VA: Pediatric<br />

Psychology Press, 1990.<br />

5. Goodman, R. The Strengths and Difficulties Questionnaire: A research<br />

note. Journal of Child Psychology and Psychiatry, 1997;38:581-586.<br />

Supported by: This study was Supported by The Wellcome Trust.<br />

O-104 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 11:30 AM<br />

NOW THAT THEY ARE FROZEN-WHAT NEXT? STEPS TAKEN<br />

AFTER OOCYTE CRYOPRESERVATION (OC) FOR DEFERRED<br />

REPRODUCTION (DR). B. Hodes-Wertz, M. E. Fino,<br />

K. N. Goldman, D. H. McCulloh, N. Noyes. NYU Langone Medical Center,<br />

New York, NY.<br />

OBJECTIVE: To assess post-treatment trends in patients completing OC<br />

cycles for DR.<br />

DESIGN: Retrospective cross-sectional study.<br />

MATERIALS AND METHODS: Post-treatment electronic medical records<br />

of all patients that underwent OC for DR from 2005-2014 were reviewed.<br />

Patients completing OC for medical indications or due to lack of<br />

sperm on day of egg retrieval (ER) and study patients were excluded. Any contact<br />

with the patient regarding her stored oocytes or her fertility after the last<br />

OC cycle was recorded (routine GYN matters were excluded). Post-ER treatments<br />

included ovulation induction (OI), IVF and oocyte thaw (OT).<br />

RESULTS: 1,394 patients underwent 1,759 cycles of OC for DR from<br />

2005-2014 with 60% of cycles completed from 2012-2014. The average<br />

age of patients that froze prior to 2012 was 383 y compared to 373 y<br />

for those that froze from 2012-2014 (p


DESIGN: Randomized, controlled, prospective study.<br />

MATERIALS AND METHODS: 166 women about to begin their first IVF<br />

cycle were randomized to the self-administered CCRI or to a routine care<br />

control group and then followed for 12 months<br />

RESULTS: The 12-month pregnancy rate was similar for the RCC and<br />

CCRI groups (OR¼1.02, 95% CI [.53 - 1.98]). Of the patients who were<br />

not pregnant on the first cycle, 15.2% (n¼15/46) of patients assigned to<br />

RC discontinued compared to 5.5% (n¼5/55) of patients assigned to the<br />

CCRI (OR¼3.11, 95% CI [.756 - 12.80]. The CCRI group engaged in significantly<br />

more positive reappraisal coping (OR¼.275, 95% CI [.16, .39] than<br />

the Routine Care control group (OR¼.097, 95% CI [-.03, .23]). The CCRI<br />

group had an improved Fertility Quality of Life (FertiQoL CORE,<br />

OR¼4.07, 95% CI [2.07, 6.06]; FertiQoL Emotional, OR¼5.95 , 95% CI<br />

[2.89, 9.00]) compared to the control group (Core OR¼ .67, 95% CI<br />

[-1.55, 2.89], Emotional OR¼-.02, 95% CI [-3.36, 3.32]). The CCRI group<br />

reported less global anxiety (OR¼ .275, 95% CI [.16, .39]) than the control<br />

group (OR¼ .471, , 95% CI [-2.40, 3.34]). The CCRI reported positive evaluations<br />

for the intervention (e.g., ease of use, helpfulness, perceived stress<br />

reduction).<br />

CONCLUSIONS: Use of the CCRI tool led to improved psychological status<br />

but not significantly more treatment cycles or a higher pregnancy rate.<br />

Supported by: This study as Supported by an unrestricted educational<br />

grant from Merck & Co, Inc., Kenilworth NJ..<br />

O-107 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:15 PM<br />

STRESS MANAGEMENT AND RESILIENCE TRAINING (SMART)<br />

THERAPY FOR COUPLES UNDERGOING IN VITRO FERTILIZA-<br />

TION (IVF): A RANDOMIZED CLINICAL TRIAL (RCT). Z. Khan,<br />

D. Fuehrer, C. Coddington, J. Bleess, G. Daftary, E. Stewart, J. Jensen,<br />

A. Sood. Mayo Clinic, Rochester, MN.<br />

OBJECTIVE: Infertility treatment is associated with high levels of stress<br />

and anxiety. Some countries mandate routine counseling to couples during<br />

infertility treatment. We conducted an RCT comparing the efficacy of Stress<br />

Management And Resiliency Training (SMART) to stress reduction CD.<br />

DESIGN: RCT<br />

MATERIALS AND METHODS: Heterosexual couples undergoing a first<br />

IVF cycle were eligible. Exclusion criteria were female age >40 years, recurrent<br />

miscarriage and current or recent (6 months) psychotic episode. Couples<br />

were randomized 1:1 to the SMART arm or CD. Those in the SMART arm<br />

attended a single 90-minutes group session that taught skills in self-awareness,<br />

and ways to strengthen attention, and incorporate greater gratitude,<br />

compassion, acceptance and purpose. Breathing-based relaxation was also<br />

incorporated. CD arm received a stress reduction CD. All participants<br />

completed validated instruments: Perceived Stress Scale (PSS-10), Fertility<br />

problem inventory (FPI), General Anxiety Disorder (GAD-7), Gratitude<br />

Questionnaire (GQ-6), Satisfaction with Life Scale (SWL) and Subjective<br />

Happiness Scale (SHS) at baseline, 6- and 12-weeks after enrollment. The<br />

study had 80% power to detect a significant difference with 28 people per<br />

group. Paired t-tests were used to compare the change in scores within a<br />

group and 2-sample t-tests were used for between group comparisons.<br />

RESULTS: Of 40 couples (n¼80) consented, 5 withdrew, 5 were lost to<br />

follow-up and 30 (n¼60) had complete data. Baseline demographics were<br />

comparable between the two groups. At 12 weeks, compared to baseline,<br />

Comparison of Score Differences within and Between Groups.<br />

Comparison within group<br />

SMART Arm<br />

Score<br />

Difference<br />

p-value<br />

CD Arm<br />

Score<br />

Difference p-value<br />

PSS-10 -6.57


Characteristics of Men Who Have Undergone Vasectomy without and with<br />

Children.<br />

No Children<br />

National Estimate<br />

(95% CI)<br />

completed their family size and do not plan on having more children. While<br />

uncommon, a childless man requesting a vasectomy can be ethically challenging<br />

scenario for urologists. We hypothesized that men who had undergone<br />

vasectomy prior to having children would have higher rates of<br />

vasectomy reversal and family planning attitudes inconsistent with being<br />

sterile.<br />

DESIGN: Retrospective analysis of the National Survey for Family<br />

Growth (NSFG).<br />

MATERIALS AND METHODS: The NSFG is a survey periodically<br />

administered in the United States by the National Center for Health Statistics.<br />

The survey uses complex survey design with oversampling of to ensure a nationally<br />

representative sample is queried on topics of sexuality and family<br />

planning. The three cycles since 2002 have included 20,146 men aged 15-<br />

44. We compared demographic information and family planning attitudes<br />

between men who had undergone vasectomy with and without children.<br />

We used the survey data analysis package in STATA to account for survey<br />

design.<br />

RESULTS: Of the 20,146 men surveyed, 696 (3.5%) reported undergoing<br />

a vasectomy. Using the complex survey design to estimate the national<br />

population, this is approximately 3,511,541 men aged 15-44 who have undergone<br />

a vasectomy. Of the men reporting vasectomy, 3.5% (95% CI 2.4-5.1)<br />

underwent the procedure without having children. Men without children<br />

were older at the time of vasectomy (Table) and have a higher household income.<br />

These men were less likely to have ever been married, and were more<br />

likely to be non-religious. There were no differences in race, education,<br />

employment status, or insurance status between the two groups (data not<br />

shown). Whereas 1.3% (0.7-2.5%) of men with children underwent vasectomy<br />

reversal during the follow-up, 0% of men without children underwent<br />

reversal, p¼0.441. When asked how many children they intended to have,<br />

men without children expected 0.00 children, whereas vasectomized men<br />

with children expected 0.01 (0.0-0.2), p¼0.007. The two groups of vasectomized<br />

men were of similar age at the time of the study, but the men without<br />

children had shorter follow up compared to men with children [47.1 months<br />

(30.8-63.4) vs. 80.8 (74.3-87.2), p


Number of veins ligated during varicocelectomy and improvement in pain post-surgery.<br />

Varicocelectomy<br />

(Number of Spermatic Veins Ligated)<br />

Variable<br />

All Subjects<br />

1-4 veins 5-10 veins R11 veins<br />

P-value<br />

Mean Age (years, y) 35.34 9.22<br />

N¼675 (100%)<br />

Unilateral (Left) Varicocele 394 (73.2%)<br />

n¼538<br />

Bilateral Varicocele 144 (26.8%)<br />

n¼538<br />

Presenting Indication:<br />

207 (30.8%)<br />

With Pain<br />

n¼673<br />

Presenting Indication:<br />

466 (69.2%)<br />

Without Pain<br />

n¼673<br />

Outcome Indication:<br />

94 (45.4%)<br />

Pain Improvement<br />

n¼207<br />

Outcome Indication:<br />

19 (9.2%)<br />

No Pain<br />

n¼207<br />

Improvement<br />

Outcome Indication:<br />

94 (45.4%)<br />

Unknown<br />

n¼207<br />

Outcome Indication (Only If Known): 94 (83.2%)<br />

Pain Improvement<br />

Outcome Indication (Only If Known):<br />

No Pain Improvement<br />

n¼113<br />

19 (16.8%)<br />

n¼113<br />

36.47 10.91<br />

n¼116 (20.0%)<br />

74 (81.3%)<br />

n¼91<br />

<strong>17</strong> (18.7%)<br />

n¼91<br />

29 (25.2%)<br />

n¼115<br />

86 (74.8%)<br />

n¼115<br />

18 (62.1%)<br />

n¼29<br />

2 (6.9%)<br />

n¼29<br />

9 (31.0%)<br />

n¼29<br />

18 (90.0%)<br />

n¼20<br />

2 (10.0%)<br />

n¼20<br />

32.98 8.56<br />

n¼3<strong>21</strong> (55.4%)<br />

<strong>21</strong>9 (82.6%)<br />

n¼265<br />

46 (<strong>17</strong>.4%)<br />

n¼265<br />

118 (36.9%)<br />

n¼320<br />

202 (63.1%)<br />

n¼320<br />

53 (44.9%)<br />

n¼118<br />

12 (10.2%)<br />

n¼118<br />

53 (44.9%)<br />

n¼118<br />

53 (81.5%)<br />

n¼65<br />

12 (18.5%)<br />

n¼65<br />

38.32 8.08<br />

n¼142 (24.5%)<br />

47 (53.4%)<br />

n¼88<br />

41 (46.6%)<br />

n¼88<br />

40 (28.2%)<br />

n¼142<br />

102 (71.8%)<br />

n¼142<br />

18 (45.0%)<br />

n¼40<br />

3 (7.5%)<br />

n¼40<br />

19 (47.5%)<br />

n¼40<br />

18 (85.7%)<br />

n¼<strong>21</strong><br />

3 (14.3%)<br />

n¼<strong>21</strong><br />

P¼0.033 (C)<br />

P¼0.033 (C)<br />

P¼0.57 (C)<br />

P¼0.57 (C)<br />

P¼0.57 (C)<br />

P¼0.81 (C)<br />

P¼0.81 (C)<br />

O-112 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:00 PM<br />

PREGNANCY RATES OF FRESH VERSUS CRYOPRESERVED<br />

SPERM OBTAINED BY PERCUTANEOUS TESTICULAR BIOPSY<br />

IN MEN WITH OBSTRUCTIVE AZOOSPERMIA IS DEPENDANT<br />

ON FEMALE AGE. S. Parsons, a J. A. Lee, a A. B. Copperman, a,b<br />

N. Bar-Chama. a,c a Reproductive Medicine Associates of New York, New<br />

York, NY; b Obstetrics, Gynecology and Reproductive Science, Icahn School<br />

of Medicine at Mount Sinai, New York, NY; c Urology, Icahn School of Medicine<br />

at Mount Sinai, New York, NY.<br />

OBJECTIVE: A percutaneous testicular biopsy (Perc BX) is a noninvasive<br />

technique of retrieving sperm for ART in men with obstructive azoospermia.<br />

This minimally invasive technique allows for excess testicular tissue to be<br />

cryopreserved for future IVF cycle(s). This study aims to determine factors<br />

related to success in Perc BX cycles of IVF.<br />

DESIGN: Retrospective cohort study<br />

MATERIALS AND METHODS: Couples undergoing fresh IVF cycles<br />

requiring a Perc BX for obstructive azoospermia by the same urologist<br />

from May 2003 to March <strong>2015</strong> were reviewed. Testicular tissue was obtained<br />

utilizing a 14 gauge core needle biopsy gun and typically under intravenous<br />

sedation. Female cohorts were segregated by age (


(micro TESE) and intracytoplasmic sperm injection (ICSI) outcome of<br />

AZFa, b, and c) in Japanese azoospermic patients.<br />

DESIGN: Retrospective, multicenter AZF deletion study in infertile Japanese<br />

men.<br />

MATERIALS AND METHODS: A total of 1032 azoospermic patinets<br />

men were examined genetic testing for AZF deletions by Promega Y Chromosome<br />

AZF Analysis System (version 2.0) in 12 Japanese medical centers<br />

from Octobar 2008 to March <strong>2015</strong>. We analyzed sperm retrieval rate (SRR)<br />

of the patiens with AZF microdeletion. In addition, we analyzed fertilization<br />

and clinical pregnancy rates of those in whom the sperm retrieval was successful.<br />

RESULTS: AZF microdeletions were found in 103 cases (9.8% of men<br />

with azoospermia): 9 AZFa, 6 AZFb, 43 AZFc, 30 AZFb+c, and 15 AZ-<br />

Fa+b+c. No spermatozoa were found with TESE or biopsy of the patients<br />

with AZFa, AZFb, AZFb+c and AZFa+b+c deletions. Spermatozoa was<br />

only obtained from patients with AZFc microdeletions. Of men with isolated<br />

AZFc deletion, spermatozoa were found in 60.0% (24/40) by micro<br />

TESE. Mean patient age was 34.8 4.7 years old. Testicular atrophy<br />

was seen in almost all men (mean testicular volume: 11.9 4.5ml). Follicular<br />

stimulating hormone (FSH), luteinizing hormone (LH), and testosterone<br />

(T) levels before surgery was 19.5 10.5 mIU/ml, 6.9 4.0<br />

mIU/ml, and 4.4 1.8 ng/ml, respectively. No correlation was found between<br />

serum FSH, LH, and T level with the success of sperm retrieval. Patients’<br />

age and testicular volume and also did not affect the SRR for micro<br />

TESE. Fertilization rate was 60.3% and clinical pregnancy rate per embryo<br />

transfer (ET) was 23.7%. Six children were safely delivered and 3 pregnancies<br />

were ongoing.<br />

CONCLUSIONS: The frequency of AZF deletions in Japan is similar to in<br />

other countries. AZFc deletions are associated with severe dysregulations of<br />

spermatogenesis, however, sperm retrieval is possible in more than half of the<br />

patients with AZFc microdeletions by micro TESE. It is useful to obtain reliable<br />

genetic information from azoospermic patients and to avoid unnecessary<br />

micro TESE.<br />

O-114 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:30 PM<br />

MALE UNDERWEAR AND SEMEN QUALITY IN A POPULATION-<br />

BASED PRECONCEPTION COHORT. K. J. Sapra, a S. Kim, a<br />

M. Eisenberg, b Z. Chen, a G. M. Buck Louis. c a Division of Intramural Population<br />

Health Research, Eunice Kennedy Shriver National Institute of Child,<br />

Rockville, MD; b Stanford University, Stanford, CA; c NICHD, Rockville,<br />

MD.<br />

OBJECTIVE: To evaluate the relationship between type of male underwear<br />

worn and semen quality in couples attempting pregnancy.<br />

DESIGN: Population-based cohort with preconception enrollment.<br />

MATERIALS AND METHODS: 501 couples were followed prospectively<br />

from cessation of contraception until positive pregnancy test or 12<br />

months of trying. At enrollment, male partners reported type of underwear<br />

worn (briefs, boxer-briefs, boxers, none) during the day and to bed, separately.<br />

Men also provided semen samples at baseline and one month later using<br />

in-home collection protocols for 24-hour semen analysis of 35 semen<br />

quality endpoints: 5 general, 8 motility, 6 sperm head, 14 morphology, and<br />

2 SCSA endpoints. Linear mixed models were used to estimate beta (b) coefficients<br />

and 95% CI for each Box-Cox transformed semen quality endpoint.<br />

Models were adjusted for statistically significant confounders: race/ethnicity<br />

and season of enrollment.<br />

RESULTS: Semen analysis was available for 62 men wearing briefs<br />

daytime/bed, 116 boxers-briefs daytime/bed, 129 boxers daytime/bed, 56<br />

briefs daytime/boxers or none bed, 71 boxer-briefs daytime/boxers or<br />

none bed, 29 boxers daytime/none bed. Compared with men wearing<br />

briefs daytime/bed (i.e. presumably highest scrotal heat exposure), changing<br />

to boxers or none for bed (i.e. lowering exposure) was associated with<br />

decreased DNA fragmentation index (DFI) in men wearing boxer-briefs in<br />

daytime (b -0.20, 95%CI -0.39, -0.01) though high DNA stainability<br />

(HDS) was increased (b 0.24, 95%CI 0.05, 0.44). Men wearing boxers<br />

in daytime and none to bed had decreased DFI (b -0.25, 95%CI -0.49,<br />

-0.01).<br />

CONCLUSIONS: Among men in the general population attempting pregnancy,<br />

type of underwear worn during the day and to bed is associated with<br />

semen quality. Reducing exposure for bed decreases DFI; better semen quality<br />

parameters are observed in men wearing boxers during the day and none<br />

to bed.<br />

Supported by: Intramural Research Program Eunice Kennedy Shriver National<br />

Institute of Child Health and Human Development.<br />

OUTCOME PREDICTORS-CLINICAL: ART 2<br />

O-115 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 11:15 AM<br />

SIBLINGS CONCEIVED WITH ASSISTED REPRODUCTIVE<br />

TECHNOLOGY: BIRTHWEIGHT AND GESTATION DIFFER-<br />

ENCES IN FRESH VS FROZEN CYCLES. B. Luke, a M. B. Brown, b<br />

E. Wantman, c J. E. Stern. d a Obstetrics, Gynecology, and Reproductive<br />

Biology, Michigan State University, East Lansing, MI; b University of Michigan,<br />

Ann Arbor, MI; c Redshift Technologies, New York, NY; d Geisel School<br />

of Medicine at Dartmouth, Lebanon, NH.<br />

OBJECTIVE: To evaluate birth outcomes of siblings conceived with ART<br />

in fresh versus frozen cycles<br />

DESIGN: Longitudinal cohort study<br />

MATERIALS AND METHODS: Cycles reported to the Society for Assisted<br />

Reproductive Technology Clinic Online Reporting System (SART<br />

CORS) during 2004-12 were linked to individual women by the woman’s<br />

birth date, first and last names, and social security number (when present);<br />

linkages across clinics also included partner’s name and sequence of ART<br />

outcomes. Included were cycles with at least one embryo transferred;<br />

excluded were research cycles and those using gestational carriers. The<br />

study population was limited to women with singleton live births conceived<br />

with ART during the study period, both using the same oocyte source<br />

(autologous or donor) but having live births from both fresh and frozen embryos<br />

(the most recent live birth was chosen if there was more than one<br />

eligible singleton birth). Mean differences in birthweight, birthweight z-<br />

score (weight-for-gestation), and length of gestation between pairs of births<br />

within each group were compared using Student’s t-test, after adjusting for<br />

mother’s age, gravidity, infant gender and, when appropriate, length of<br />

gestation.<br />

RESULTS: The study population included 10,222 autologous pairs of siblings<br />

and 2,240 donor pairs of siblings. In both groups, the frozen cycle was<br />

approximately 2 years later than the fresh cycle. Siblings born from frozen<br />

embryo transfers had shorter gestations (0.6 days for autologous and 1.1<br />

days for donor, both p¼0.0003), heavier birthweights (160 grams for autologous<br />

and 46 grams for donor, both p


ANOVA, Welch’s t-test and multivariate analysis. Significance was defined<br />

as p


Embryo Transfer Practices and Outcomes in Mandated States.<br />

CT HI IL MD MA NJ P Value<br />

Day 5 transfer (%) 30.0 34.4 52.5 44.8 <strong>17</strong>.8 42.7


Science, Icahn School of Medicine at Mount Sinai, New York, NY; b Reproductive<br />

Medicine Associates of New York, New York, NY; c RMANY-Mount<br />

Sinai, New York, NY.<br />

OBJECTIVE: The modern treatment of the infertile patient often includes<br />

ovarian stimulation, oocyte retrieval, fertilization, biopsy and chromosomal<br />

analysis (CCS), and vitrification. Embryos found to be normal can be transferred<br />

in a subsequent FET cycle. We were able to control for embryo quality<br />

by using only screened embryos and for synchronization by initiating progesterone<br />

treatment five days prior to embryo transfer and were able to ask the<br />

question of whether the number of days of estradiol preparation impacts the<br />

likelihood of implantation.<br />

DESIGN: Retrospective<br />

MATERIALS AND METHODS: Patients who underwent an IVF cycle<br />

with qPCR-based CCS that had R1 euploid embryo available for FET<br />

were included. Fresh ETs were excluded from this analysis. Patients started<br />

oral estradiol 2 mg twice daily for four days, then 2 mg three times daily.<br />

Endometrial thickness was assessed weekly until a thickness of R7mm<br />

was observed. Progesterone supplementation was then initiated, and five<br />

days later, re-warming and ultrasound guided transfer was performed. A<br />

Poisson regression model was used to predict if the total number of days<br />

of estradiol impacted implantation rates. Statistical difference of p


TABLE 1. TEBX-PGS cycle outcomes using FR vs. frozen oocytes in women 39y at retrieval.<br />

Age(y)<br />

39y<br />

groups<br />

FR: Total No. Retrieved OC: Total No. 674 5798 292 4762<br />

MI+MII Thawed<br />

OC: Overall Survival 559 (83%) 238 (82%)<br />

OC: Survived and Assessed as MII at Thaw 534 (79% ) 225 (77%)<br />

No. Cycles Achieving BL Suitable for TEBX 39/47 (83%) 352 /407 (86%) 0.5 15/22 (68%) 339/407 (83%) 0.08<br />

No. Biopsied BL/ 2PN Fertilization 145/373 (39%) 1911/3418 (56%) 0.0001 54/167 (32%) 1316/2628 (50%) 0.0001<br />

No. Euploid BL 58/145 (40% ) 9<strong>21</strong>/1911 (48%) 0.06 8/54 (15%) 254/1316 (19%) 0.5<br />

No. Aneuploid BL 86/145 (60%) 955/1911 (50%) 0.03 44/54 (85%) 1031/1316 (78%) 0.7<br />

No. TEBX Cycles with At Least One Euploid 28/39 (72%) 295/352 (84%) 0.07 7/15 (47%) 160/339 (47%) 1.0<br />

BL<br />

No. TEBX Cycles with All Aneuploid BL 11/39 (28%) 57/352 (16%) 0.07 8/15 (53%) <strong>17</strong>9/339 (53%) 1.0<br />

Embryo Implantation Rate 14/22 (64%) 188/287 (66%) 0.8 2/3 (67%) 89/129 (69%) 1.0<br />

Ongoing+Delivered Rate/Transfer 12/<strong>21</strong> (57%) 154/276 (56%) 1.0 2/3 (67%) 72/123 (59%) 1.0<br />

semen quality. All resultant zygotes were cultured for 5-7 days at which point<br />

any BL reaching Stage R2BC was subjected to TEBX, then vitrified until<br />

euploid ET. Thus, OC oocytes were subjected to freeze-thaw twice (oocyte+BL)<br />

while in FR cycles, only biopsied BL were vitrified. In the OC<br />

arm, average oocyte freezing duration was 3.42y. Data were compared<br />

for OC vs. FR cycles in women ages < vs. R 40y. Fisher’s was used for statistical<br />

comparisons.<br />

RESULTS: See Table. In 54 (78%) OC and 691 (85%) FR cycles, at<br />

least one 2PN-fertilized oocyte developed into a BL suitable for<br />

TEBX (p¼ 0.2) with a mean of 3/cycle for OC & 4/cycle for FR. A total<br />

of 199 BL (37% of 2PN fert) were suitable for biopsy in OC cycles vs.<br />

3227 BL (53% of 2PN fert) in FR attempts (p¼ 0.0001). The latter difference<br />

held across both age groups. Regardless of age, all but 12/423 (3%)<br />

women had single- ET. EIR and pregnancy rates were not different between<br />

the groups.<br />

CONCLUSIONS: Adding OC to TEBX-PGS resulted in significantly<br />

fewer 2PN-fertilized oocytes becoming BL suitable for TEBX, although<br />

the total no. of cycles with at least 1 biopsiable BL was not different. Notably,<br />

the euploidy rate for biopsied BL and implantation and ongoing/delivered<br />

rates for transferred euploid BL were similar whether derived from OC or<br />

FR oocytes, regardless of female age at retrieval. Thus, adding TEBX to<br />

OC appears reasonable and potentially beneficial as long as patients are<br />

counseled that OC results in a lower BFR, thus fewer biopsiable BL will<br />

be available for chromosomal evaluation.<br />

O-124 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:00 PM<br />

PREGNANCY OUTCOMES AFTER FRESH VERSUS FROZEN-<br />

THAWED BLASTOCYST TRANSFERS. N. Pereira, a A. C. Petrini, b<br />

J. Lekovich, a G. L. Schattman, c Z. Rosenwaks. b a The Ronald O. Perelman<br />

and Claudia Cohen Center for Reproductive Medicine, New York, NY;<br />

b Weill Cornell Medical College, New York, NY; c Weill Medical College/<br />

Cornell University, New York, NY.<br />

OBJECTIVE: There has been a shift towards frozen-thawed embryo transfers<br />

in the name of improved endometrial receptivity and pregnancy outcomes.<br />

To investigate the effect of ovarian stimulation on endometrial<br />

receptivity, we aim to compare pregnancy outcomes after fresh versus<br />

frozen-thawed blastocyst transfer cycles at our center.<br />

DESIGN: Retrospective single-center cohort study.<br />

MATERIALS AND METHODS: All patients undergoing transfer of fresh<br />

or frozen (vitrified)-thawed blastocysts after in vitro fertilization (IVF) between<br />

January 2010 and September 2013 were analyzed for potential inclusion.<br />

Cycles utilizing slow freezing protocols or pre-implantation genetic<br />

screening were excluded from the analysis. Demographic characteristics<br />

Comparision of Pregnancy Outcomes in Patients undergoing Fresh vs. Frozen-<br />

Thawed ET.<br />

Parameter<br />

Fresh<br />

Blastocyst<br />

Transfer<br />

(n¼918)<br />

Frozen-Thaw<br />

Blastocyst<br />

Transfer<br />

(n¼1273)<br />

Age (years) 35.8 (5.09) 36.1 (5.28) 0.18<br />

BMI (kg/m2) 23.2 (4.72) 23.4 (4.22) 0.60<br />

Peak endometrial 10.7 (3.73) 10.5 (3.81) 0.19<br />

stripe (mm)<br />

Blastocele grading<br />

1,2<br />

3<br />

151 (16.4%)<br />

767 (83.6%)<br />

220 (<strong>17</strong>.3%)<br />

1053 (82.7%)<br />

0.61<br />

ICM grading<br />

A<br />

B<br />

C<br />

Trophectoderm<br />

grading<br />

A<br />

B<br />

C<br />

61 (6.64%)<br />

709 (77.2%)<br />

148 (16.1%)<br />

73 (7.95%)<br />

701 (76.4%)<br />

144 (15.7%)<br />

60 (4.71%)<br />

1098 (86.2%)<br />

115 (9.03%)<br />

62 (4.87%)<br />

1111 (87.2%)<br />

120 (9.43%)<br />

compared were age and body mass index (kg/m 2 ), while baseline characteristics<br />

analyzed included endometrial stripe thickness (mm), blastocele<br />

grading (1-3), inner cell mass grading (A-C), and trophectoderm grading<br />

(A-C). Implantation, clinical pregnancy, spontaneous miscarriage and live<br />

birth rates were calculated. Student’s t-tests and Chi-square (c2) tests were<br />

used as indicated. P


(5.28) years in the frozen-thawed group (P ¼0.18). There was no difference<br />

in the mean BMI or peak endometrial stripe when comparing the groups.<br />

There was also no difference in the grading of the blastocele expansion, inner<br />

cell mass and trophectoderm grading of the blastocysts transferred in the<br />

two groups. No differences in the rates of implantation (37.3% vs. 37.7%),<br />

clinical pregnancy (50.2% vs. 49.4%), spontaneous miscarriage (7.30%<br />

vs. 9.27%), or live birth (42.9% vs. 40.6%) were found between the two<br />

groups.<br />

CONCLUSIONS: Our study highlights that fresh blastocyst transfers<br />

yield equivalent pregnancy outcomes as frozen-thawed blastocyst transfers.<br />

While there continues to be a shift towards frozen-thawed ET cycles in the<br />

name of improved pregnancy and perinatal outcomes, it is possible that<br />

conservative stimulation protocols with fresh transfer offer equivalent benefits.<br />

O-125 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:15 PM<br />

SINGLE EUPLOID BLASTOCYST TRANSFER IS A VIABLE CLIN-<br />

ICAL OPTION FOR AMA PATIENTS UP TO 42 YEARS: ANALYSIS<br />

OF 1,000 CONSECUTIVE SINGLE FROZEN EMBRYO<br />

TRANSFERS. M. Katz-Jaffe, a E. Surrey, a D. A. Minjarez, b<br />

R. L. Gustofson, c L. A. Kondapalli, a J. M. Stevens, d W. B. Schoolcraft. a<br />

a Colorado Center for Reproductive Medicine, Lone Tree, CO; b Colorado<br />

Center for Reproductive Medicine, Denver, CO; c CCRM, Lone Tree, CO;<br />

d Fertility Labs of Colorado, Lone Tree, CO.<br />

OBJECTIVE: Women of advanced maternal age (AMA) presenting for<br />

infertility treatment are considered to have poor prognosis. In order to increase<br />

the likelihood of a clinical pregnancy, AMA women typically have<br />

multiple embryos transferred. Often, this leads to undesirable, multiple gestations<br />

that result in poorer obstetrical and neonatal outcomes. The objective<br />

of this study was to evaluate the clinical efficacy of a single euploid frozen<br />

blastocyst transfer for AMA women.<br />

DESIGN: Research study<br />

MATERIALS AND METHODS: A consecutive cohort of 1,000<br />

infertility patients who received a single euploid, frozen blastocyst<br />

transfer. All embryos were routinely biopsied at the blastocyst stage<br />

for comprehensive chromosome screening prior to vitrification.<br />

Standard protocols for a hormone replacement frozen embryo transfer<br />

were employed. Analysis was performed to determine the probability<br />

of achieving a live birth following the transfer of a single euploid blastocyst<br />

in association with maternal age. Statistical analysis included<br />

Fisher’s exact and Chi-square test for independence, with significance<br />

at P


OVARIAN STIMULATION<br />

O-127 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 11:15 AM<br />

RANDOMIZED PROSPECTIVE TRIAL COMPARING THREE<br />

DOSES OF GNRH ANTGONIST PROTOCOL IN PATIENTS WITH<br />

POOR OVARIAN RESPONSE. E. Bastu, O. Dural, C. Yasa,<br />

M. Ozsurmeli, I. Demiral, B. Baysal, F. Buyru. Department of Obstetrics<br />

and Gynecology, Istanbul University School of Medicine, Istanbul, Turkey.<br />

OBJECTIVE: To examine whether patients with poor ovarian response<br />

(POR) according to the Bologna criteria, during conventional IVF/intracytoplasmic<br />

sperm injection (ICSI) cycle, may benefit from different doses of gonadotropins<br />

in ovulation induction protocols.<br />

DESIGN: Randomized prospective trial carried out in an infertility clinic<br />

of an university hospital.<br />

MATERIALS AND METHODS: Only POR patients strictly defined according<br />

to the Bologna criteria were included. 95 patients with POR were randomized<br />

to receive three different doses of gonadotropins during ovulation<br />

stimulation. Group 1 (31 patients) received 450 IU gonadotropins (225 IU<br />

HP-hMG+225 IU rFSH); Group 2 (31 patients) received 300 IU gonadotropins<br />

(150 IU HP-hMG+150 IU rFSH); Group 3 (33 patients) received 150 IU gonadotropins<br />

(75 IU HP-hMG+75 IU HP-hMG+5mg letrozole). All transferred<br />

embryos were Grade-1 according to the morphologic classification.<br />

RESULTS: There were no statistically significant differences between the<br />

age, body mass index (BMI), infertility period, day-3 E2, FSH, prolactin and<br />

testosterone serum levels between the three groups (all p>0.05). Between the<br />

three groups, there were no statistically significant difference in length of<br />

ovulation induction, number of retrieved oocytes, number of MII oocytes,<br />

number of fertilized oocytes, fertilization rate, number of transferred embryos,<br />

cycle cancellation, chemical and clinical pregnancy rates (all<br />

p>0.05). Although the difference was not statistically significant, rate of<br />

chemical pregnancy was higher in the third group (75 IU HP-hMG+75 IU<br />

rFSH+letrozole) compared to the first and second groups (0.160.37 vs.<br />

0.160.38 vs. 0.<strong>21</strong>0.42; p¼0.83).<br />

CONCLUSIONS: Increasing the daily dosage of administered gonadotropins<br />

is an intuitively appealing approach to the stimulation of the patients<br />

with POR. However, increasing the dosage does not improve the number<br />

of MII oocytes, fertilization rate, number of transferred embryos, chemical<br />

and clinical pregnancy rates. Utilizing a mild stimulation with letrozole is<br />

as effective as stimulation with higher doses.<br />

Comparison of Outcome parameters<br />

1st Group<br />

(225 IU<br />

HP-hMG<br />

+ 225 IU<br />

rFSH)<br />

n¼31<br />

References: The ClinicalTrials.gov identifier is NCT02293668.<br />

O-128 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 11:30 AM<br />

2nd Group<br />

(150 IU<br />

HP-hMG<br />

+ 150 IU<br />

rFSH)<br />

n¼31<br />

3rd Group<br />

(75 IU<br />

HP-hMG<br />

+75IU<br />

rFSH)<br />

n¼33<br />

Length of ovulation 8.771.65 9.652.74 9.881.82 0.097<br />

induction, months<br />

Gonadotropin dosage, IU 3929.03 2670.09 2309.09


presenting moderate/severe OHSS had significantly higher number of intermediate<br />

but similar number of dominant follicles than women without OHSS<br />

(respectively 124.9 and 7.25.3, p¼0.008; 4.44.1 and 4.33.7, p¼0.89).<br />

A multivariable logistic regression revealed that the number of intermediate<br />

follicles was the only statistically significant predictor of OHSS occurrence.<br />

CONCLUSIONS: Shoview study shows that moderate/severe OHSS incidence<br />

after COS with HP-hMG was 2.4% in this cohort where 41% of women<br />

were considered at high risk of developing OHSS. Although some cases still<br />

occurred, risk assessment and COS adaptation allowed satisfactory OHSS<br />

risk management.<br />

Supported by: Shoview cohort study was sponsored and conducted by Ferring<br />

SAS, Gentilly, France.<br />

O-130 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:00 PM<br />

NEUTRALIZATION OF VASCULAR ENDOTHELIAL GROWTH<br />

FACTOR (VEGF) BEFORE OR AFTER CG DURING COS CYCLES<br />

RESTORES OVARIAN VASCULAR PARAMETERS SIMILAR TO<br />

NATURAL MENSTRUAL CYCLES: A PILOT STUDY IN<br />

NONHUMAN PRIMATES. C. V. Bishop, a X. Li, b D. M. Lee. c a Division<br />

of Reproductive & Developmental Sciences, Oregon National Primate<br />

Research Center, Beaverton, OR; b Advanced Imaging Research Center, Oregon<br />

Health & Science University, Portland, OR; c Obstetrics & Gynecology,<br />

Oregon Health & Science University, Portland, OR.<br />

OBJECTIVE: To determine if VEGF neutralization either before or after<br />

hCG administration would result in restoration of vascular parameters<br />

similar to those observed in natural cycles in a nonhuman primate model<br />

of ovarian hyperstimulation syndrome (OHSS).<br />

DESIGN: Repeated measures in vivo pilot study.<br />

MATERIALS AND METHODS: Rhesus macaque females (n¼8) underwent<br />

baseline evaluation of ovarian vascular flow and blood volume (BV)<br />

during normal menstrual cycles at mid luteal phase by contrast enhanced ultrasound<br />

[1]. Ovarian vascular permeability to serum albumin was analyzed<br />

on the same day by Dynamic Contrast Enhanced-MRI [2]. Females were then<br />

treated with a Controlled Ovarian Stimulation [3] (COS) protocol with hCG<br />

trigger, and then randomly assigned to treatment groups: Control (n¼3),<br />

VEGF neutralizing agent Avastin 15-30hrs before hCG (single IV bolus;<br />

10mg/kg; n¼3), and Avastin 3-4 days post-hCG (single IV bolus; n¼2).<br />

Vascular flow, BV, and permeability of ovaries were analyzed 6-8 days<br />

post-hCG (similar to mid-luteal phase in normal menstrual cycles).<br />

RESULTS: OHSS symptoms developed in 2/3 Control, 0/3 Avastin prehCG,<br />

and 1/2 Avastin post-hCG females: abdominal fluid was present at<br />

time of oocyte retrieval and/or by MRI evaluation. Avastin administration<br />

prior to hCG reduced serum progesterone levels compared to other groups<br />

(P¼0.03); peak levels did not rise above 1.5 ng/ml. There was no significant<br />

difference in ovarian vascular flow; however, increased BV occurred in COS<br />

cycles (multiple CL) compared to natural cycles (single CL; P


DESIGN: Prospective randomized controlled trial at a university-affiliated<br />

private IVF clinic between <strong>October</strong> 2009 and September 2013.<br />

MATERIALS AND METHODS: Participants were IVF/ICSI candidates<br />

considered at risk of POR. Since the trial had started prior to the publication<br />

of the Bologna criteria, we had defined the risk of POR as the following: basal<br />

FSH > 10 IU/L or AMH


for the evaluation of oocyte quality secondary to bidirectional communication<br />

via gap junctions. The intent of this study was to investigate if differences<br />

in the CC transcriptome, as determined by RNA-Seq, are associated<br />

with embryo implantation.<br />

DESIGN: Paired analysis.<br />

MATERIALS AND METHODS: Patients with normal ovarian reserve<br />

% 42 years of age were recruited for participation. COCs were dissected,<br />

and CC RNA was isolated and then stored at -80 C. Embryo identity was<br />

maintained, and blastocysts underwent CCS. Patients who had a double<br />

embryo transfer resulting in a singleton live birth were potentially informative<br />

and were selected for further analysis. DNA fingerprinting was utilized<br />

to determine which embryo implanted. A cDNA library was<br />

constructed from CC RNA and sequenced using a Life Technologies<br />

(LT) Ion Torrent Sequencing platform. Filtered reads were aligned to the<br />

hg19 AmpliSeq transcriptome reference (LT). Both negative-binomial<br />

based and nonparametric methods were used to identify differentially expressed<br />

genes with regards to live birth.<br />

RESULTS: CC RNA sequencing of 38 samples (from 19 patients) generated<br />

an average of 10.5 4.3 million reads/sample. 23 differentially expressed<br />

genes between sibling embryos that resulted in a live birth and<br />

those that did not were identified (P < 0.05). However, after correcting for<br />

multiple testing none of the genes remained significantly differentially expressed<br />

(false discovery rate < 0.05). <strong>17</strong>3 genes identified in previous studies<br />

as predicting implantation, including RGS2, ANG, PTX3, and ZNF132, were<br />

not differentially expressed in this analysis.<br />

CONCLUSIONS: This study demonstrates that there were no candidate<br />

biomarkers identified in the CC transcriptome that were differentially<br />

expressed between euploid embryos resulting in a live birth and<br />

those that failed. This is the first trial of this question to utilize paired<br />

euploid sibling embryos which is the optimal study design for identifying<br />

effective biomarkers within an individual patient’s cohort of<br />

euploid embryos.<br />

O-136 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:00 PM<br />

MIRNAS ISOLATED FROM EXTRACELLULAR VESICLES IN<br />

FOLLICULAR FLUID AND OOCYTE DEVELOPMENT<br />

POTENTIAL. R. Machtinger, a R. Rodosthenous, b A. Mansour, c<br />

M. Adir, c C. Racowsky, d R. Hauser, b A. A. Baccarelli. b a Obstetrics and Gynecology,<br />

Sheba Medical Center and Tel-Aviv University, Tel Aviv, Israel;<br />

b Harvard Chan School of Public Health, Boston, MA; c Obstetrics and Gynecology,<br />

Sheba Medical Center and Tel-Aviv University, Ramat gan, Israel;<br />

d Brigham and Women’s Hospital ART Center.<br />

OBJECTIVE: Bidirectional communication between the oocyte and granulosa<br />

compartment is critical for oocyte maturation, fertilization and embryonic<br />

development. Recent studies have shown that extracellular vesicles<br />

(EVs) within the ovarian follicle serve as vehicles for the transfer of mediators<br />

such as miRNAs between the oocyte and the somatic cells. However, the<br />

profile of these miRNAs remains largely unexplored. The aim of this study<br />

was to characterize EV-encapsulated miRNAs from human follicular fluid<br />

(FF) and to test their association with fertilization and day 3 embryo quality.<br />

DESIGN: Prospective study.<br />

MATERIALS AND METHODS: Follicular fluid (FF) samples were<br />

collected from a single follicle of ICSI patients stimulated with GnRH antagonist<br />

(n¼46). From this cohort, samples from non-severe male factor patients<br />

and that contained a mature (MII) oocyte were analyzed (n¼40). Each<br />

oocyte, its corresponding embryo and FF sample were tracked. Expression<br />

profiles of EV-encapsulated miRNAs in FF were assessed using TaqMan<br />

OpenArrayÒ. Only miRNAs that were detected in >50% of the samples<br />

and exerted a Ct value %35 were included in the statistical analysis. Statistical<br />

significance was adjusted for multiple comparisons using the Benjamini-Hochberg<br />

method.<br />

RESULTS: Of the 754 screened miRNAs, <strong>21</strong>5 were detected in all samples,<br />

with several (e.g. miR-30d, miR-320b, miR-10b and miR-1291), being<br />

detected in >85% of the samples. After excluding cases of abnormal fertilization<br />

(n¼5), we compared 42 miRNAs from FF that contained normal<br />

fertilized (n¼30) or failed to fertilize (n¼5) oocytes, and found 2.7-fold<br />

and 1.5-fold higher expression of miR-202-5p (p¼0.001) and miR-7-1-3p<br />

(p¼0.001) in EVs from follicles containing oocytes that fertilized normally<br />

as compared those that failed to fertilize. Twenty nine embryos were evaluated<br />

on day 3. No significant difference in expression of any EV-encapsulated<br />

miRNA was observed in the miRNA profile (n¼35) for FF samples linked<br />

with top quality embryos (n¼19) versus those linked to embryos of lesser<br />

quality (n¼10).<br />

CONCLUSIONS: Our preliminary results in this novel area of investigation<br />

identified specific miRNAs carried within EVs in FF that may be associated<br />

with fertilization potential. If our findings are Supported by further<br />

studies, differential expression profiles of these miRNAs may prove to be<br />

useful biomarkers for oocyte developmental potential.<br />

Supported by: This study was funded by Grant Award no. RPGA1301 from<br />

the Environment and Health Fund Israel and by grants P30ES00002 and<br />

R<strong>21</strong>ES024236 from the National Institute of Environmental Health<br />

Research.<br />

O-137 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:15 PM<br />

MICROBIOME AT THE TIME OF EMBRYO TRANSFER: NEXT<br />

GENERATION SEQUENCING OF THE 16S RIBOSOMAL<br />

SUBUNIT. J. M. Franasiak, M. D. Werner, C. R. Juneau, X. Tao,<br />

J. N. Landis, Y. Zhan, N. R. Treff, R. T. Scott. RMA, NJ, NJ.<br />

OBJECTIVE: Characterization of the human microbiome has benefited<br />

from the application of powerful molecular tools utilizing the unique 16S ribosomal<br />

subunit’s hypervariable regions to greatly increase sensitivity. The<br />

microbiome of the lower genital tract can prognosticate obstetrical outcome<br />

while the upper reproductive tract remains poorly characterized. Issues<br />

outside of obstetrics are not well studied. Indeed, the microbiome at the<br />

time of embryo transfer has only been evaluated utilizing culture based technology<br />

which is known to capture only a fraction of dominant and major organisms.<br />

This study analyzes the microbiome at the time of single embryo<br />

transfer and characterizes the microbiome by reproductive outcome.<br />

DESIGN: Prospective cohort.<br />

MATERIALS AND METHODS: Consecutive patients undergoing<br />

euploid, single embryo transfer were included in the analysis. Patients underwent<br />

routine IVF care, oocyte retrieval, and embryo culture. The embryo<br />

transfer was performed per routine. After transfer, the most distal 5mm<br />

portion of the transfer catheter was sterilely placed in a DNA free PCR tubes.<br />

Cell lysis and DNA purification were performed and DNA sequencing was<br />

performed using Next Generation Sequencing of the bacteria specific 16S<br />

ribosome gene. The sequences were compared to reference databases to procure<br />

genus and species calls for microorganisms. This allowed for identification<br />

and relative quantification of bacteria present in the sample.<br />

RESULTS: Taxonomy assignments were made on 35 samples from 33 patients<br />

and 2 E. coli controls. Of the 35 patients, 18 had ongoing pregnancies<br />

and 15 did not. There were a total of 276 different genus calls present across<br />

patient samples. The microbiome at time of transfer for those patients with<br />

ongoing pregnancy vs. those without ongoing pregnancy is characterized<br />

by top genera by sum fraction. Results are summarized in Table 1. Several<br />

dominant species were present in both as would be expected; however, there<br />

were varied major species by outcome.<br />

CONCLUSIONS: The data presented here show the microbiome at the<br />

time of embryo transfer may differ by pregnancy outcome. This novel<br />

approach, both in specimen collection and analysis, is the first step towards<br />

goal of determining physiologic from pathophysiologic microbiota. Further<br />

studies will help delineate if differences in the microbiome at the time of embryo<br />

transfer have a reliable impact on pregnancy outcome.<br />

Microbiome at time of transfer which resulted in ongoing and non-ongoing<br />

gestations after SET.<br />

Ongoing<br />

Pregnancy<br />

(Genus)<br />

Genus<br />

Sum<br />

fraction<br />

Non-ongoing<br />

Pregnancy<br />

(Genus)<br />

Genus<br />

Sum<br />

fraction<br />

Flavobacterium 4.58 Flavobacterium 3.42<br />

Lactobacillus 3.35 Lactobacillus 3.29<br />

Limnohabitans 0.85 Limnohabitans 0.73<br />

Polynucleobacter 0.78 Polynucleobacter 0.71<br />

Bdellovibrio 0.68 Bdellovibrio 0.58<br />

Chryseobacterium 0.64 Blvii28 0.53<br />

Spirochaeta 0.63 Clostridium 0.52<br />

Clostridium 0.55 Chryseobacterium 0.52<br />

Blvii28 0.50 Spirochaeta 0.51<br />

Paludibacter 0.39 Pseudomonas 0.31<br />

e54 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


O-138 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:30 PM<br />

DO THERMAL PROPERTIES OF A CULTURE DISH INFLUENCE<br />

EMBRYO DEVELOPMENT?. L. B. Ramirez, a H. L. Shi, a E. Lvov, a<br />

C. Racowsky, a,b C. L. Bormann. a,b a Department of Obstetrics, Gynecology<br />

and Reproductive Biology, Brigham and Women’s Hospital, Boston, MA;<br />

b Harvard Medical School, Boston, MA.<br />

OBJECTIVE: Embryos display sensitivity to temperature fluctuations. As<br />

rate of heat gain and loss may vary by type of culture dish, the question arises<br />

whether dish-specific rate of temperature exchange (RTE) influences embryo<br />

development during the short excursions of dishes from incubators to room<br />

temperature (RT). If so, dish RTE consideration would be important when<br />

choosing the type of dish used in clinical IVF. This study tested the hypothesis<br />

that high dish RTE negates any adverse effect of temperature fluctuation<br />

on embryo development.<br />

DESIGN: Experimental study.<br />

MATERIALS AND METHODS: Thirteen commercial dishes containing<br />

medium and pre-equilibrated to 37 C were assessed for their RTE’s by moving<br />

the dishes from 37 C to RT and then back to 37 C. The RTE’s were determined<br />

from times required for a 5 C temperature change after these<br />

excursions. The two dishes selected for embryo experiments were at the extremes<br />

for RTE values (Falcon 1007, low RTE¼1.360.04 C/100s; Vitro-<br />

Life micro-droplet dish, high RTE¼1.710.16 C/100s). A total of 464<br />

fresh 2-cell mouse embryos were randomly allocated to 4 groups (Falcon<br />

control [F-C], Falcon treatment [F-T], Vitrolife control [V-C], and Vitrolife<br />

treatment [V-T]) in triplicate experiments. Embryos were cultured in 25 ml<br />

drops of Global TotalÒ, overlaid with 9 ml OvoilÔ at 37 C in 5% O2,<br />

6.2% CO2 and balanced N2. All dishes were removed twice a day from<br />

the incubator and placed on a RT surface (treatments) or 37 C surface (controls)<br />

for 5 min, and then returned to the incubator. Development was assessed<br />

on day 4.5, blastocyst number and quality were recorded, and<br />

embryos were stained with Hoescht for total cell counts. Groups were<br />

compared using Fisher’s Exact and t-test where appropriate, with p


O-141 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 11:45 AM<br />

DISTRIBUTION OF THE FMR1 GENE IN FEMALES BY RACE-<br />

ETHNICITY: WOMEN WITH DIMINISHED OVARIAN RESERVE<br />

VERSUS WOMEN WITH NORMAL FERTILITY (SWAN<br />

STUDY). L. Pastore, a S. L. Young, b A. Manichaikul, c V. Baker. d a Stony<br />

Brook Medicine, Stony Brook, NY; b UNC School of Medicine, Chapel Hill,<br />

NC; c University of Virginia, Charlottesville, VA; d Stanford University, Stanford,<br />

CA.<br />

OBJECTIVE: Premutation level trinucleotide repeat lengths in the<br />

FMR1 gene (CGG 55-199) are associated with primary ovarian insufficiency<br />

before age 40. However, an association between the repeat length<br />

with other ovarian aging phenotypes is not established. We examined<br />

whether reported associations between the FMR1 CGG repeat lengths<br />

in the intermediate, high normal, or low normal range differentiate<br />

women diagnosed with diminished ovarian reserve (DOR) from those<br />

with normal reproductive histories, and whether associations vary by<br />

race-ethnic group.<br />

DESIGN: DOR cases (n¼129) enrolled from 5 US fertility clinics vs.<br />

normal fertility female controls (n¼803) from the US-based Study of<br />

Women’s Health Across the Nation (SWAN).<br />

MATERIALS AND METHODS: Cases’ (95 Caucasians, 22 Asian, 12<br />

other) and controls’ (386 Caucasians, <strong>21</strong>9 African-Americans, 102 Japanese,<br />

96 Chinese) banked DNA was analyzed for FMR1 CGG repeat<br />

lengths. Cases were clinically diagnosed with DOR, with regular menses<br />

and no fragile X syndrome family history. Controls had >¼1 menstrual<br />

period in the 3 months pre-enrollment, >¼1 pregnancy, no history of infertility<br />

or hormonal therapy, and menopause R46 years. In a previous analysis<br />

the SWAN Chinese and Japanese groups had similar FMR1 CGG<br />

repeat lengths, so those two groups were combined. We used Fisher’s exact<br />

tests to analyze data.<br />

RESULTS: We found fewer CGG trinucleotide repeats in the lower of the<br />

two FMR1 alleles (i.e., allele with fewer repeats) in DOR cases relative to<br />

controls among Caucasians (p&lt0.0001), but not Asians (p¼0.24). Caucasian<br />

DOR cases were more likely to have fewer CGG repeats in the lower<br />

allele compared with Asian DOR cases (p¼0.027). No significant differences<br />

were found in the CGG repeat length distribution of the higher FMR1 allele<br />

among DOR cases compared with controls (p&gt0.20), or by race/ethnic<br />

group (p¼0.41).<br />

CONCLUSIONS: This study did not find an association between DOR and<br />

high normal/intermediate repeats, but there was an association between DOR<br />

and low normal repeats (


membrane permeability and compromises oocyte quality. These findings<br />

contribute to understanding of the premature ovarian insufficiency and infertility<br />

seen commonly in women with classic galactosemia.<br />

O-144 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:30 PM<br />

SELECTING EMBRYOS WITH OPTIMAL CLEAVAGE KINETICS<br />

IMPROVES ONGOING PREGNANCY RATE FOLLOWING BLAS-<br />

TOCYST TRANSFER IN A MOUSE MODEL. R. S. Weinerman, a<br />

T. S. Ord, a R. Feng, b C. Coutifaris, a M. A. Mainigi. a a Division of Reproductive<br />

Endocrinology and Infertility, University of Pennsylvania, Philadelphia,<br />

PA; b Department of Biostatistics and Epidemiology, University of Pennsylvania,<br />

Philadelphia, PA.<br />

OBJECTIVE: To develop a model predicting blastocyst (blast) formation<br />

based on early cleavage parameters in the mouse and to determine if the<br />

model predicts pregnancy outcome following blast transfer. Microarray analysis<br />

was performed to identify genes associated with early embryo development.<br />

DESIGN: Laboratory research.<br />

MATERIALS AND METHODS: 313 2-pronuclear (PN) embryos were<br />

collected from superovulated (SO) female mice and were cultured to blast<br />

in 5% (n¼180) and 20% (n¼133) oxygen (O2). Time-lapse videos were<br />

collected in the EevaÔ System (Progyny). Classification and regression<br />

tree analysis (CART) was used to predict blast formation based on 4 cleavage<br />

parameters: 1st cytokinesis duration, and time from 2-3 cell, 3-4 cell, and 4-5<br />

cell stages. The model was built in the 20% O2 cohort and validated in the 5%<br />

O2 cohort. For the transfer experiments, 2PN embryos were collected<br />

following mating of SO females to transgenic male mice heterozygous for<br />

green-fluorescent protein (GFP) and cultured in the EevaÔ system. Blasts<br />

were rated as having ‘‘optimal’’ (Opt) or ‘‘suboptimal’’ (Subopt) timing based<br />

on the CART model. 10 blasts were transferred into each of 10 pseudo-pregnant<br />

recipients, using GFP status to tag the embryos as Opt or Subopt. Pregnancy<br />

outcomes were assessed at mid-gestation (E10.5). Differences in<br />

outcomes were assessed using Fisher’s exact test. Single-embryo microarray<br />

analysis was performed using the Affymetrix GeneChip Mouse Gene ST 1.0<br />

Array on Opt and Subopt blasts.<br />

RESULTS: The final model utilized 2 parameters, cleavage times from 2-3<br />

cell and 3-4 cell. The model predicted blast formation with a 97.5% sensitivity<br />

and 62.5% specificity in 5% O2. Following transfers, 8 mice became<br />

pregnant and were included for analysis. The overall implantation rate per<br />

embryo transferred did not differ significantly between groups (86% Opt<br />

vs. 77% Subopt, p¼0.3). However, the ongoing pregnancy rate was higher<br />

for the Opt compared to Subopt embryos [60% (35/58) vs. 32% (7/22)<br />

(p¼0.03)]. The resorption rate (implantation sites with placental tissue<br />

only) among implanted embryos was significantly higher in the Subopt<br />

compared to Opt group [59% (10/<strong>17</strong>) vs. 30% (15/50)(p¼0.04)]. Gene<br />

expression differences were seen in 74 genes between the Opt and Subopt<br />

groups (FD


Comparison according to blastocelic expansion<br />

Gardner 4-5<br />

(n¼112)<br />

Gardner 6<br />

(n¼122)<br />

Chi<br />

Square<br />

Oocyte Age 36.64.2 35.95.0 NS<br />

Day 3 FSH 5.72.9 6.13.2 NS<br />

End. thickness 9.71.9 9.91.9 NS<br />

(mm) at surge<br />

Peak estradiol 2292.71072.2 2453.91243.3 NS<br />

Oocytes retrieved <strong>17</strong>.29.0 <strong>17</strong>.89.6 NS<br />

MII 13.18.0 13.97.8 NS<br />

Number of 2PN 10.66.8 11.97.7 NS<br />

Implantation rate 56.3% (63/112) 51.2% (66/129) NS<br />

Biochemical PR 65.2% (73/112) 64.8% (79/122) NS<br />

Miscarriage rate 54.5% (61/112) 53.3% (65/122) NS<br />

Miscarriage rate 30.1% (22/73) 27.8% (22/79) NS<br />

O-147 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 11:45 AM<br />

BLASTOCYST TRANSFER IN WOMEN AGED 40 YEARS OR<br />

OLDER: HOW MANY IS TOO MANY?. K. S. Richter, J. E. O’Brien,<br />

M. J. Tucker, R. J. Stillman. Shady Grove Fertility, Rockville, MD.<br />

OBJECTIVE: To evaluate blastocyst transfer outcomes for women aged<br />

40 to 44 years according to the number and quality of embryos transferred.<br />

DESIGN: Retrospective review.<br />

MATERIALS AND METHODS: From 2008-2013 all autologous embryo<br />

transfers performed on day 5 or 6 after oocyte retrieval among women 40 to<br />

44 years were reviewed. Embryos were categorized as ‘‘A’’ blastocysts (grade<br />

AA or BA), ‘‘B’’ blastocysts (grade AB or BB), ‘‘C’’ blastocysts (ICM and/or<br />

trophectoderm grade C or insufficiently expanded to grade), or ‘‘D’’ embryos<br />

(pre-cavitation). Generalized estimating equations (GEE) analysis was used<br />

to account for multiple cycles per patient and adjust for age and transfer day.<br />

RESULTS: 1,145 cycles by 932 patients were available for analysis.<br />

Adjusted GEE analysis of the independent contributions of embryos in<br />

each category estimated that live born children resulted from <strong>21</strong>.2% of<br />

‘‘A’’ blastocysts (p20 weeks of<br />

pregnancy).<br />

DESIGN: Prospective randomized controlled trial (RCT) started in June<br />

2012 with recruitment closed in December 2014. Patients were allocated<br />

through computer-generated randomization into two groups: conventional<br />

blastocyst transfer or preimplantation genetic screening (PGS). Sample<br />

size calculated for the endpoint of ongoing pregnancy rates per cycle and delivery<br />

rates was 120 patients per arm with ovum pick-up.<br />

MATERIALS AND METHODS: Women between 38-41 years, %3 implantation<br />

failures and %3 previous miscarriages, and R5 MII oocytes<br />

from one or two stimulation cycles. Embryo culture and fresh blastocyst<br />

transfer was performed in the non-intervention arm. Cleavage stage embryo<br />

biopsy and array-CGH were performed in the intervention arm with fresh<br />

blastocyst transfer of euploid embryos. Individual blastomeres underwent<br />

whole genome amplification using the Sureplex Kit (Illumina). Amplified<br />

DNA and reference DNAwere labeled and co-hybridized in 24sure arrays (Illumina).<br />

After washing, slides were scanned and analyzed by BlueFuse Multi<br />

software. Statistical comparisons were performed using two-sided Fishersexact<br />

test and t-Student test.<br />

RESULTS: A total of <strong>21</strong>5 cycles have been completed, 113 cycles in the<br />

conventional blastocyst transfer group and 102 in the PGS group. Mean female<br />

age (39.42.8 and 39.53); and mean number of MII oocytes<br />

(9.63.9 and 9.83.5) were similar in both groups. Mean number of transferred<br />

embryos was significantly higher in the blastocyst group compared to<br />

the PGS group (1.80.6 vs. 1.30.7; p


effective for good prognosis patients. This study aim to answer the question:<br />

Does multiple pregnancy rate decrease in good prognosis patients who failed<br />

in fresh elective SET (eSET) and received an elective double frozen embryo<br />

transfer (eDFET) compared with patients receiving fresh elective double embryo<br />

transfer (eDET)?<br />

DESIGN: Retrospective observational study.<br />

MATERIALS AND METHODS: This study evaluated 723 ICSI cycles using<br />

standard conventional protocol, at a private Assisted Reproduction Center<br />

during last 10 years. eDETwas carried out in 639 patients, and 84 patients<br />

received eSET and failed in first fresh cycle and had an subsequent eDFET<br />

(eSET+eDFET). It was defined as eSET and eDET patients who transferred<br />

one or two top quality embryos, respectively, and had at least two spared top<br />

quality embryos cryopreserved.<br />

RESULTS: Patients results for eDET and eSET+eDFET groups were: age<br />

(35.24.1 x 34.03.1; p¼0.016), basal FSH (6.24.7 x 7.110.6;<br />

p¼0.207), FSH dose administered (1881.1516.4 x <strong>17</strong>08.2238.9;<br />

p¼0.003) and oocytes collected (13.05.8 x 10.75.3; p¼0.001). Groups<br />

presented similar pregnancy (36.2% x 31.2%; p¼0.385) and multiple pregnancy<br />

rates (31.1% x 25.0%; p¼0.541).<br />

CONCLUSIONS: Limiting the number of embryos transferred can benefit<br />

the bottom line by decreasing the incidence of multiple pregnancy. We hypothesized<br />

that patients who failed in fresh eSET could present worst prognosis<br />

and hence had lower multiple pregnancy rates in a second attempt with<br />

two embryos transferred. However, in this sample, the second transfer with<br />

two elective embryos, after a failed elective transfer of one embryo, was<br />

not effective to protect this population of multiple gestation, suggesting<br />

that for this purpose the appointment of new elective transfer of an embryo<br />

would be the most appropriate for good prognosis patients, even after a failure<br />

Supported by: Not applicable.<br />

O-150 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:30 PM<br />

THE IMPACT OF TRANSFER TIME ON CLINICAL PREGNANCY IN<br />

FROZEN EMBRYO TRANSFER CYCLES. M. S. Lee, E. R. Cardozo,<br />

A. E. Karmon, D. L. Wright, T. Toth. Vincent Obstetrics and Gynecology,<br />

Massachusetts General Hospital, Boston, MA.<br />

OBJECTIVE: To identify the critical aspects of embryo transfer that<br />

impact clinical pregnancy rates, specifically the impact of transfer time.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: All consecutive frozen embryo transfer<br />

(FET) cycles using cryopreserved blastocysts from 2007 - 2014 were reviewed.<br />

All blastocysts were of high quality, all cycles used a uniform<br />

cryopreservation technique and controlled hormone replacement protocol,<br />

and all transfers were done under ultrasound guidance. Each transfer was<br />

scored on a set of standardized metrics including cohort score (morphologic<br />

appearance at transfer), ease of mock transfer, bends placed in catheter,<br />

presence of blood or mucus, and overall difficulty of transfer.<br />

Transfer time (total seconds from when the catheter was loaded until<br />

the embryo(s) was expelled into the uterine cavity) was recorded.Logistic<br />

regression models using generalized estimating equations were fit to<br />

investigate the relationship between transfer time and clinical pregnancy<br />

rate (CPR) while controlling for confounders and within-person correlations<br />

in cycle outcome. Multivariate results were adjusted for age, diagnosis,<br />

cohort score, number of embryos transferred, use of an outer<br />

sheath, type of catheter used, blood inside the catheter, and physician performing<br />

the transfer.<br />

RESULTS: 467 women undergoing 571 FET cycles were included. Mean<br />

age was 35.9 years (22-51 4.4). Overall CPR was 47.1% . The cohort was<br />

divided into tertiles by transfer time (seconds) (T1: 33-55; T2: 57-81; T3: 82-<br />

582) with mean transfer times (SD) of 47.4 (5.7), 67.1 (7.3) and 1<strong>21</strong>.9 (55.1)<br />

seconds. Crude CPR was 43.9% , 48.7% and 48.7% among respective tertiles.In<br />

univariate analysis, worse cohort score (p


MATERIALS AND METHODS: One-cell embryos from F1 mice were<br />

cultured in 20% oxygen either individually, or in groups of ten, in media<br />

G-1/ G-2 with 5mg/ml HSA, in the presence or absence of 10mM carnitine/10mM<br />

cysteine/5mM lipoate. Embryo development was analysed<br />

through time-lapse and embryo transfers performed. Intracellular levels of<br />

reduced glutathione (GSH) were assessed using fluorometric analysis.<br />

RESULTS: There was no effect of treatment on blastocyst formation.<br />

However, compared to controls, the three antioxidants significantly<br />

increased embryo cell numbers when used individually (P


O-155 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:15 PM<br />

CRYOPRESERVATION METHODS AND COMPARISON OF OUT-<br />

COMES IN THE HUMAN OOCYTE PRESERVATION EXPERIENCE<br />

(HOPE) REGISTRY. Z. P. Nagy, a T. Elliott, a C. Chang, a B. Hayward, b<br />

M. C. Mahony. b a Reproductive Biology Associates, Atlanta, GA; b EMD<br />

Serono, Inc., Rockland, MA.<br />

OBJECTIVE: To compare IVF outcomes by cryoprotectant use and cryopreservation<br />

methodologies for donor oocytes cryopreserved by either slowfreezing<br />

(SF) or vitrification (VIT) in the HOPE Registry.<br />

DESIGN: Post hoc analysis of data from the Phase IV, prospective, multicenter<br />

(16 US centers), observational HOPE Registry.<br />

MATERIALS AND METHODS: Clinical and laboratory data were<br />

collected on ART cycles that used cryopreserved oocytes (autologous and<br />

donor) from women aged 18-50 years enrolled between June 2008 and<br />

September 2010. Six centers (involving 85/145 [59%] patients) were audited;<br />

regular monitoring across all centers ensured clean patient data. We<br />

compared implantation rates (IR) using analysis of variance and clinical<br />

pregnancy rates (CPR) with Fisher’s exact tests in SF and VIT cycles using<br />

donor oocytes for various cryopreservation parameters.<br />

RESULTS: Data from 136 patients receiving donor oocytes cryopreserved<br />

by either SF (n¼41, 302 oocytes) or VIT (n¼95, 704 oocytes) were analyzed.<br />

For VIT cycles, container type had a significant effect on IR (p¼0.0194) and<br />

CPR (p¼0.0374); CryotopÒs were used most often (n¼71) and associated<br />

with the highest CPR (70.4%). Culture media used in the cryosolution also<br />

had a significant effect on IR for VIT cycles (p¼0.0200); homemade<br />

(n¼30) and protein (20%)/homemade/HEPES-based/amino-acid based<br />

(n¼38) were used most often and were associated with IRs of 63.9% and<br />

58.3%, respectively. For SF cycles, 1-2 oocytes/container vs R3 (both<br />

n¼20) was associated with respectively: significantly greater IR (25.1% vs<br />

2.9%, p¼0.0049) and CPR (55.0% vs 5.0%, p¼0.0012); a cooling rate<br />

( C/minute) of 2vs 3or 0.3 was associated with IRs 22.8% vs 6.3%<br />

vs 0% (p¼0.0511) and CPR 50% vs 10% vs 0% (p¼0.0088); a plunging temperature<br />

( C) of < 40 vs 30 to 35 vs 36 to 40 was associated with<br />

greater IR (23.9% vs 3.3% vs 0%, p¼0.0352) and CPR (52.4% vs 5.9% vs<br />

0%, p¼0.0041).<br />

CONCLUSIONS: Container type (CryotopÒ best) and culture media type<br />

in the cryosolution had significant effects on VIT outcomes. SF outcomes<br />

were affected by number of oocytes per container (1-2 best), cooling rate<br />

( 2 C/minute best) and plunging temperature (< 40 C).<br />

References: An abstract reporting data on VIT vs SF use and outcomes<br />

(oocyte survival rate and implantation rates) will be reported at the ESHRE<br />

<strong>2015</strong> Congress (Nagy ZP, Hayward B, Mahony MC. Evaluation of the effect<br />

of various cryoprotectants and protocols on donor oocyte survival and embryo<br />

viability based on HOPE Registry data. Accepted for presentation at<br />

ESHRE <strong>2015</strong>). This ASRM abstract includes important different data on<br />

cryoprotectant methodology and ART outcomes.<br />

Supported by: The study and abstract development were Supported by<br />

EMD Serono, Inc., Rockland, MA, USA, a subsidiary of Merck KGaA,<br />

Darmstadt, Germany.<br />

O-156 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:30 PM<br />

THE EFFICACY OF SYSTEMIC ADMINISTRATION OF GRANU-<br />

LOCYTE COLONY STIMULATING FACTOR (GCSF) ON THE IN<br />

VITRO FERTILIZATION (IVF) SUCCESS IN WOMEN WITH<br />

REPEATED IMPLANTATION FAILURE. Z. Abedi Asl. Shariati IVF<br />

Center, Tehran, Iran, Islamic Republic of.<br />

OBJECTIVE: To evaluate the efficacy of systemic subcutaneous Granulocyte<br />

Colony Stimulating Factor (GCSF) administration on In Vitro Fertilization<br />

(IVF) successin infertile women with repeated IVF failure.<br />

DESIGN: Multi-centric, prospective, randomized, open label, controlled,<br />

trial.<br />

MATERIALS AND METHODS: 100 infertile women with normal endometrial<br />

thicknesswho had R 2 implantation failure after IVF cycles who<br />

were referred for IVFto Infertility Departments of Shariati Hospital, Bahman<br />

Hospital (two tertiary referral centers) and Omid infertility clinic, Tehran,<br />

Iran. Sealed, numbered, opaque envelopes assigned 50 patients to receive<br />

subcutaneous 300mg GCSF before implantation and 50 in the control group.<br />

The outcomes were the implantation (number of gestational sacs on the total<br />

number of transferred embryos) chemical pregnancy (positive serum b-<br />

HCG), and clinical pregnancy (gestational sac and fetal heart)rates.<br />

RESULTS: The implantation rate in the intervention group was significantly<br />

higher than the control group (15.3% vs. 7.2%, p ¼ 0.04),the chemical<br />

pregnancy rate was significantly higher in the intervention group than the<br />

control group (40% vs. 20%, p ¼ 0.02),and the clinical pregnancy rate was<br />

higher in the intervention group than the control group, however this was<br />

not statistically significant (28% vs. 16%, p ¼ 0.14). There were no adverse<br />

events in any of the study groups.<br />

CONCLUSIONS: Administration of single-dose systemic subcutaneous<br />

GCSF before implantation significantly increases the IVF success, implantation<br />

and pregnancy rates in infertile women with repeated IVF failure.<br />

ASSISTED REPRODUCTIVE TECHNOLOGY - GENERAL 1<br />

O-157 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 11:15 AM<br />

GNRH AGONIST INDUCES TH1 AND TH<strong>17</strong> IMMUNITY, WHILE<br />

GNRH ANTAGONIST SUPPRESSES TH1 IMMUNITY. N. Sung, a<br />

M. D. Salazar Garcia, a S. Dambaeva, b J. Kwak-Kim. a a Rosalind Franklin<br />

University of Medicine and Science, Vernon Hills, IL; b Microbiology and<br />

Immunology, Rosalind Franklin University of Medicine and Science, North<br />

Chicago, IL.<br />

OBJECTIVE: Inflammatory immune responses have been reported in<br />

women with multiple implantation failures after IVF cycles. In this study,<br />

we aim to investigate the effect of GnRH analogues to peripheral blood lymphocytes,<br />

in-vitro.<br />

DESIGN: Experimental study.<br />

MATERIALS AND METHODS: PBMCs were treated with various concentration<br />

of GnRH agonist (Buserelin acetate) and antagonist (Cetrorelix<br />

acetate) of 0.1, 1, 5 and 10mM for 4 hours to 48 hours. Th1, Th2 and Th<strong>17</strong><br />

cells were analyzed by flow cytometry. Anti-CD45, anti-CD3 and anti-<br />

CD8 were used to isolate T cell subpopulation and intracellular cytokine<br />

staining was made to analyze Th1, Th2 and Th<strong>17</strong> cell population using<br />

anti-TNF-a, anti- INF-g, anti-IL-10 and IL-<strong>17</strong>.<br />

RESULTS: The proportion of TNF-a and IFN-g producing Th cells were<br />

increased in dose dependent manner with GnRH agonist culture for 24 hours.<br />

The proportions of TNF-a producing Th cells with 5 and 10 mM GnRH<br />

agonist (P¼0.046 each) were significantly higher than that with 1 mM<br />

GnRH agonist. There were no significant differences in IL-10 producing<br />

Th cells after 24 hour culture with GnRH agonist. The proportion of Th<strong>17</strong><br />

cells treated with 1 mM GnRH agonist were significantly higher than those<br />

of controls (without GnRH agonist) (P¼0.008), with 5 mM (P¼0.028) and<br />

10 mM (P¼0.028) GnRH agonist. TNF-a/IL-10 and IFN- g/IL-10 ratios<br />

treated with GnRH agonist for 24 hours (0.1, 1 and 5 mM) were not different<br />

from those of controls. Th<strong>17</strong> cells treated with 1 mM GnRH agonist was<br />

significantly higher than those without GnRH agonist (p¼0.008) after 48<br />

hour culture. TNF-a/IL-10 and IFN-g/IL-10 secreting Th cell ratios in<br />

PBMCs treated with GnRH agonist for 4 hours (0.1, 1, 5 and 10 mM) were<br />

increased as compared to controls, and those treated with GnRH antagonist<br />

(0.1, 1 and 5 mM) were decreased as comparison to controls. Th<strong>17</strong> cells<br />

cultured with 1 mM GnRH agonist was significantly decreased in comparison<br />

to controls (P¼0.019). Th<strong>17</strong> cells treated with 10 mM GnRH antagonist was<br />

significantly increased as compared to controls (P¼0.023).<br />

CONCLUSIONS: GnRH analogues affect T helper cell subpopulations<br />

including Th1 and Th<strong>17</strong> cells. GnRH agonist seems to induce Th1 and<br />

Th<strong>17</strong> immunity. Further investigation of GnRH analogues will be necessary<br />

to understand inflammatory immune responses during ovulation induction<br />

cycles.<br />

O-158 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 11:30 AM<br />

INHIBITION OF AUTOPHAGY BY SERA FROM WOMEN WHOSE<br />

IVF CYCLE RESULTED IN A SPONTANEOUS ABORTION OR<br />

BIOCHEMICAL PREGNANCY. G. Sisti, T. Kanninen, I. Ramer,<br />

S. Witkin, S. D. Spandorfer. Cornell University Medical Center, New<br />

York City, NY.<br />

OBJECTIVE: Autophagy is an intracellular process whereby degraded or<br />

foreign proteins and defective organelles become surrounded by a structure<br />

called an autophagosome. By subsequent fusion with a lysosome, the enclosed<br />

macromolecules are degraded and its component parts returned to<br />

the cytoplasm for reutilization. Induction of autophagy at various stages of<br />

FERTILITY & STERILITY Ò<br />

e61


pre- and post-implantation fetal growth in animal models has been shown to<br />

enhance embryo development and survival. We previously reported that sera<br />

from pregnant women induced autophagy in peripheral blood mononuclear<br />

cells (PBMCs). In the present study we evaluated whether the extent of autophagy<br />

induction by sera of pregnant women would differ according to the<br />

outcome of their IVF cycle.<br />

DESIGN: We evaluated the extent of autophagy induction by available<br />

sera from a retrospective cohort of 94 women who completed an IVF cycle<br />

at our institution. The breakdown was 28 women with a term pregnancy,<br />

16 who delivered preterm, 11 with a SAB, 12 with a BC pregnancy, 16<br />

who did not become pregnant and 11 who had an ectopic pregnancy (EP).<br />

MATERIALS AND METHODS: Donor PBMCs were isolated from<br />

reproductive age healthy female volunteers by Ficoll-Hypaque gradient<br />

centrifugation and added to wells of a sterile microtiter plate (5x105 cells<br />

per well) that contained 35 ml RPMI culture medium and 15 ml sera from<br />

the study subjects. All sera were from day 28 of the IVF cycle, obtained<br />

at the time of the initial pregnancy test. As a positive control parallel cultures<br />

also contained 4800 mM rapamycin, an autophagy inducer. Following<br />

incubation for 48 hours (37 C, 5% CO2) the cells were centrifuged, lysed in<br />

the presence of protease inhibitors and stored at -80oC until assayed. The<br />

extent of autophagy was determined by measuring the concentration of<br />

p62 by ELISA. p62 is a protein whose intracellular concentration is<br />

inversely proportional to the extent of autophagy induction. The non-parametric<br />

Kruskal Wallis test with Dunn’s correction was used to determine<br />

associations.<br />

RESULTS:<br />

Autophagy induction in donor PBMCs by sera from women with different IVF<br />

outcomes<br />

OUTCOME<br />

Median ng/mL p62<br />

No rapamycin<br />

(range)<br />

+ rapamycin<br />

Term birth 7.5 (


O-161 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:15 PM<br />

AN FSH-LOWERING ACTIVIN DISRUPTING THERAPY PRE-<br />

VENTS EGG CHROMOSOME AND SPINDLE MISALIGNMENTS<br />

THAT PREDISPOSE TO ANEUPLOIDY, AND INCREASES<br />

FERTILITY, IN A MOUSE MODEL OF MIDLIFE REPRODUCTIVE<br />

AGING. L. R. Bernstein, a,b,c,d A. C. Mackenzie, b C. L. Chaffin, e<br />

I. Merchenthaler. b a Pregmama, LLC, Montgomery Village, MD; b Epidemiology<br />

and Public Health, University of <strong>Maryland</strong> School of Medicine, <strong>Baltimore</strong>,<br />

MD;<br />

c Gynecology and Obstetrics, Johns Hopkins School of<br />

Medicine, <strong>Baltimore</strong>, MD; d Veterinary Integrative BioSciences, Texas A &<br />

M College of Veterinary Medicine, College Station, TX; e Obstetrics, Gynecology,<br />

and Reproductive Sciences, University of <strong>Maryland</strong> School of Medicine,<br />

<strong>Baltimore</strong>, MD.<br />

OBJECTIVE: Women of advanced maternal age women (AMA, >age 35)<br />

have increased risk of oocyte & embryo aneuploidy, infertility, miscarriages,<br />

and trisomic pregnancies (collectively ‘‘egg infertility.’’). Egg infertility increases<br />

markedly with age due to elevated rates of egg aneuploidy. It is a significant<br />

public health problem, with 1 in 5 US women now attempting her<br />

first pregnancy after 35. Elevated FSH is one of the first signs of ovarian aging.<br />

We hypothesize that high FSH is a cause of egg infertility, that elevating<br />

FSH activity for the period of oocyte growth will increase egg infertility, and<br />

that lowering FSH for the period of oocyte growth will prevent egg infertility.<br />

DESIGN: We developed SAMP8 mice as model with human-like midlife<br />

female reproductive aging characteristics, including elevated FSH, increased<br />

rates of oocyte spindle misalignments, and diminished fertility by midlife<br />

(age 7 months). A regimen to raise FSH activity with chronic PMSG treatment<br />

was given to one test group of midlife SAMP8 for 3 weeks, the period<br />

of oocyte growth. An FSH lowering regimen was developed using ActRIIB:Fc,<br />

an activin decoy receptor that sequesters activin and suppresses activin<br />

signaling. This was given for 3 weeks to a second test group. A third test<br />

group was comprised of untreated control mice.<br />

MATERIALS AND METHODS: Chromosome and spindle misalignments<br />

of ovulated oocytes are highly predictive of impending aneuploidy,<br />

were scored in fluorescence microscopy. Fertility was compared between ActRIIB:Fc-treated<br />

and untreated midlife SAMP8 groups by quantitation of<br />

litter sizes after mating with young proven SAMP8 males.<br />

RESULTS: PMSG increased rates of chromosome misalignments from<br />

32/193 oocytes (16.5%) to 38/1<strong>21</strong> oocytes (31.4%; P¼0.0013), and<br />

increased rates of spindle misalignments from 13/192 (6.77%) to 14/98<br />

(14.3%; P¼0.0331). ActRIIB:Fc lowered FSH in midlife SAMP8 to the<br />

level of young SAMP8. ActRIIB:Fc decreased chromosome misalignments<br />

from 32/193 (16.5%) to 11/159 (6.9%; P¼0.0030), and decreased spindle<br />

misalignments from 13/192 (6.77%) to 4/155 (2.58%; P¼0.0182). Rates<br />

of chromosome and spindle misalignments were lowered to those of young<br />

mice. ActRIIB:Fc restored nearly 40% of the fertility lost with age,<br />

increasing litter sizes from 5.06 to 6.29 pups/litter (P¼0.0305, vs. 8.22 in<br />

young SAMP8).<br />

CONCLUSIONS: These data provide supportive evidence that FSH may<br />

play a role in egg aneuploidy and infertility. Hormone normalization therapy<br />

(‘‘HNT’’) to lower FSH and disrupt activin signaling shows promise as a<br />

novel therapeutic intervention to prevent oocyte aneuploidy and infertility,<br />

miscarriages, and trisomies.<br />

Supported by: 1. <strong>Maryland</strong> Industrial Partnerships Grant (MIPS). 2. Technology<br />

Development Corporation of <strong>Maryland</strong> (TEDCO). 3. Max and Victoria<br />

Dreyfus Foundation Grant. 4. Bernstein and Pine Families. 5. Pregmama.<br />

6. University of <strong>Maryland</strong> School of Medicine Funds. 7. Indiegogo<br />

campaign.<br />

O-162 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:30 PM<br />

GNRH AGONIST LEUPROLIDE ACETATE NEITHER ACTIVATES<br />

ANTI-APOPTOTIC GENES NOR PROTECTS HUMAN OVARY<br />

AND GRANULOSA CELLS FROM DNA DAMAGE AND APOPTOSIS<br />

INDUCED BY CYCLOPHOSPHAMIDE. G. Bildik, a N. Akin, a<br />

F. Senbabaoglu, a Y. Guzel, b U. Ince, c B. Balaban, b B. Urman, d,b<br />

O. Oktem. d,b a Reproductive Biology, Koc University Graduate School of<br />

Health Sciences, Istanbul, Turkey; b Women’s Health Center Asssited Reproduction<br />

Unit Fertility Preservation Program, American Hospital, Istanbul,<br />

Turkey; c Pathology, Acibadem University, Istanbul, Turkey; d Obstetrics<br />

and Gynecology, Koc University School of Medicine, Istanbul, Turkey.<br />

OBJECTIVE: Inconsistent results of randomized controlled trials (RCTs)<br />

and lack of a proven molecular mechanism of action with ovarian protection<br />

with co-administration GnRH agonists (GnRHa) with chemotherapy places<br />

GnRHa under scrutiny as a fertility preservation strategy. We aimed in this<br />

study to provide molecular evidence for-or-against the role of GnRHa in<br />

the prevention of cyclophosphamide induced damage in human ovarian tissue<br />

samples and granulosa cells.<br />

DESIGN: A translational research study.<br />

MATERIALS AND METHODS: Ovarian cortical pieces (n¼15, age<br />

14-37) and human mitotic non-luteinized (COV434, HGrC1) and nonmitotic<br />

luteinized (HLGC) granulosa cells were treated with 4-hydroperoxy<br />

cyclophosphamide (in vitro active metabolite of cyclophosphamide<br />

used at 50 and 100 mM) with and without GnRHa leuprolide acetate<br />

(50 ng/mL: peak intraovarian concentration of the drug) for 24 hrs.<br />

Cell proliferation (real-time quantitative assessment by xcelligence system),DNA<br />

damage (p-histone H2AX), apoptosis (cleaved caspase-3,<br />

YO-PRO-1), follicle counts, hormonal markers of ovarian function and<br />

reserve (estradiol, progesterone and AMH), and the expression of antiapoptotic<br />

genes (bcl-2, bcl-xL, bcl-2L2, Mcl-1, BIRC-2 and XIAP)<br />

were compared among control, chemotherapy and chemotherapy+GnRHa<br />

groups.<br />

RESULTS: GnRH receptor expression and its activation by GnRHa were<br />

validated with qRT-PCR and measuring intracellular cAMP level, respectively.<br />

Exposure to cyclophosphamide resulted in massive follicle loss, arrested<br />

cell growth, increased DNA damage/apoptosis and decreased<br />

hormone productions in the tissue samples and granulosa cells. The coadministration<br />

of GnRHa with cyclophosphamide did not prevent or attenuate<br />

any of these cytotoxic effects. Furthermore, GnRHa did not up-regulate<br />

the anti-apoptotic genes compared to control and cyclophosphamide treated<br />

samples.Mcl-1 and BIRC2 expressions were further decreased after cyclophosphamide+GnRHa<br />

(Table).<br />

CONCLUSIONS: GnRH agonist leuprolide acetate does not offer any protection<br />

against cyclophosphamide induced damage in human ovary and granulosa<br />

cells via its cognate receptors.<br />

The Impact of Cyclophosphamide (Cyc)GnRHa on Ovarian Tissue Samples<br />

and Granulosa Cells<br />

Control Cyc Cyc+GnRHa P value<br />

Ovarian Tissue: Follicle reserve, anti-apoptotic gene expression and<br />

hormone productions<br />

Primordial/mm2<br />

Preantral/mm2<br />

2.540.5<br />

0.60.2<br />

0.330.2<br />

0.120.05<br />

0.330.1<br />

0.140.1<br />

p¼0.006 control vs.<br />

cyc+GnRHa<br />

p¼0.3 cyc vs.<br />

Bcl-2<br />

Bcl-xL<br />

Bcl-2L2<br />

Mcl-1<br />

BIRC2<br />

XIAP<br />

AMH(ng/mL)<br />

E2(pg/mL)<br />

P(ng/ml)<br />

10.06<br />

10.03<br />

10.04<br />

10.02<br />

10.04<br />

10.01<br />

1.20.09<br />

78898<br />

1.760.4<br />

0.600.07<br />

0.970.03<br />

0.600.03<br />

0.930.03<br />

0.750.01<br />

0.860.02<br />

0.460.01<br />

0.920.01<br />

0.510.01<br />

0.810.01<br />

0.590.01<br />

0.720.05<br />

0.30.02<br />

14843<br />

0.310.06<br />

p<br />

cyc+GnRHa<br />

0.30.03<br />

p¼0.004<br />

18516<br />

control vs.<br />

0.30.02<br />

cyc+GnRHa<br />

p¼0.2<br />

cyc vs. cyc+GnRHa<br />

Luteinized Granulosa Cells (HLGCs): Hormone productions and apoptosis rate<br />

E2(pg/mL)<br />

P(ng/mL)<br />

Apoptosis (%)<br />

1560112<br />

59695<br />

3%<br />

27128<br />

15018<br />

89%<br />

20731<br />

10516<br />

91%<br />

Mitotic granulosa cells: Proliferative index and apoptosis rate<br />

Proliferative<br />

index<br />

Apoptosis(%)<br />

1.20.2<br />

3%<br />

0.140.01<br />

88%<br />

p¼0.009 control vs.<br />

cyc+GnRHa<br />

p¼0.3 cyc vs.<br />

cyc+GnRHa<br />

p<br />

p¼0.9 cyc vs.<br />

cycGnRHa<br />

(for proliferative<br />

index)<br />

p<br />

p¼0.8 cyc vs.<br />

cyc+GnRHa<br />

(for apoptosis<br />

rate)<br />

FERTILITY & STERILITY Ò<br />

e63


GENETIC COUNSELING 1<br />

O-163 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 11:15 AM<br />

ADVANTAGES OF TRIPLET REPEAT EXPANSION DETECTION IN<br />

BLASTOCYST BIOPSY FOR PREIMPLANTATION GENETIC<br />

DIAGNOSIS OF FRAGILE X SYNDROME. R. Prates, a S. Jaroudi, b<br />

A. Jordan, a N. Goodall, a B. Chu, a V. Tecson, a A. Hershlag, c<br />

M. Garrisi, d F. Licciardi, e B. Witt, f M. Konstantinidis. a a Reprogenetics, Livingston,<br />

NJ; b Reprogenetics Middle East, Al Ain, United Arab Emirates;<br />

c North Shore LIJ Health System, Cold Spring Harbor, NY; d IRMS-NJ, Livingston,<br />

NJ; e New York University Langone Medical Center, New York, NY;<br />

f Greenwich Fertility and IVF Center, Greenwich, CT.<br />

OBJECTIVE: To evaluate the utilization of a comprehensive diagnosis<br />

strategy for preimplantation genetics diagnosis (PGD) of fragile X syndrome<br />

(FXS) in order to enhance selection of embryos suitable for transfer and increase<br />

therefore, the chances of patients achieving a pregnancy.<br />

DESIGN: Trophectoderm biopsies obtained from blastocysts preimplantation<br />

embryos were processed for direct FXS repeat size determination and<br />

linkage analysis via Karyomapping (Illumina, USA). When requested,<br />

comprehensive chromosome screening (CCS) via array comparative<br />

genomic hybridization testing was also performed in parallel to PGD for<br />

FXS.<br />

MATERIALS AND METHODS: Trophectoderm biopsies were sent from<br />

27 different IVF clinics between December 2013 and March <strong>2015</strong> to a single<br />

PGD laboratory for analysis. A total 168 blastocysts from 34 PGD cycles (4.9<br />

embryos per cycle) were assessed for FXS direct mutation test (average<br />

maternal age: 34.8).<br />

RESULTS: Diagnostic results were obtained for 96.4% of the samples<br />

(162/168). The proportion of embryos found to be available for transfer based<br />

on the inheritance of the normal allele via linkage analysis was 53.7% (87/<br />

162). Among the 46.3% (75/162) categorized as having inherited the affected<br />

allele, direct triplet repeat sizing allowed the identification of 24.1% (39/162)<br />

additional blastocysts suitable for transfer (carriers of premutation or intermediate<br />

expansions) (p43). The aneuploidy rate for the each male-female age group interaction<br />

was computed, with 95% confidence intervals calculated by Clopper-Pearson<br />

method. The aneuploidy rate was modeled by logistic regression. The youngest<br />

age groups were considered reference factors. The model was assessed<br />

by chi-square of ANOVA with significance at p


Table 1: Incidence and types of chromosome errors in aneuploid blastocysts<br />

(*p


for 8511 patients. Individuals were screened for up to <strong>21</strong>3 recessive genetic<br />

diseases using Illumina’s Infinium HD Genotyping Platform. Carrier counts<br />

were tallied for each disease screened, and overall number of carriers was<br />

analyzed for 3 disease panels of increasing scope: (1) diseases currently recommended<br />

for screening by professional society guidelines; (2) diseases<br />

considered to be high-impact for reproductive decision making; and (3) a<br />

broad panel of <strong>21</strong>3 diseases that range in severity.<br />

RESULTS: On the most limiting panel, which only screened for diseases<br />

currently included in professional society guidelines, 10.7% of patients were<br />

identified as carriers. When compared to a larger panel inclusive of 188 high<br />

impact diseases, our results indicated that limited screening failed to identify<br />

26.5% of carriers in our patient population. An additional 5% of patients<br />

were identified as carriers when the panel was expanded further to include<br />

a total of <strong>21</strong>3 diseases of varying impact.<br />

CONCLUSIONS: New technologies have provided the ability to simultaneously<br />

screen for a multitude of mutations at a reduced cost. Concurrently,<br />

genetic diversity has increased due to the admixture of different ethnic groups.<br />

These cultural and technological changes have altered the clinical approach to<br />

genetic carrier screening, and expanded carrier panels with greater than 100<br />

diseases are increasingly being offered to patients. Data from our ethnically<br />

diverse patient population demonstrate that screening for only traditionally<br />

recommended diseases failed to identify over a quarter of carriers. Such oversight<br />

might contribute to decreased detection of carrier couples, and thus the<br />

inability for patients to make fully informed reproductive decisions. Furthermore,<br />

an additional 5% of patients were identified as carriers when including<br />

moderate impact conditions, which may not affect reproductive decisions but<br />

may provide helpful information to couples about their future child’s health.<br />

Overall, our data support that expanded carrier panels are more effective at<br />

identifying patients at risk of having children with genetic conditions who<br />

thus may benefit from increased reproductive options. The continued support<br />

of genetic counselors will be critical to guiding all patients through the<br />

screening process, regardless of carrier status or disease severity.<br />

References:<br />

1. Edwards JG, Feldman G, Goldberg J, et al. Expanded Carrier Screening<br />

in Reproductive Medicine_Points to Consider: A Joint Statement of the<br />

American College of Medical Genetics and Genomics, American College<br />

of Obstetricians and Gynecologists, National Society of Genetic<br />

Counselors, Perinatal Quality Foundation, and Society for Maternal-<br />

Fetal Medicine Obstet Gynecol. <strong>2015</strong>;125(3):653-662.<br />

MALE REPRODUCTION<br />

O-169 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 11:15 AM<br />

INTRACYTOPLASMIC SPERM INJECTION REDUCES TOTAL<br />

FAILED FERTILIZATION RATE BUT DOES NOT IMPROVE PREG-<br />

NANCY OR LIVE BIRTH RATES IN UNEXPLAINED INFERTILITY:<br />

ANALYSIS OF OVER 20,000 CYCLES FROM THE SART<br />

DATABASE. L. Johnson, a,b M. D. Sammel, c A. Dokras. b a Reproductive<br />

Endocrinology Associates of Charlotte, Charlotte, NC; b University of Pennsylvania,<br />

Philadelphia, PA; c Univ. of Pennsylvania, Perelman School of<br />

Medicine, Philadelphia, PA.<br />

OBJECTIVE: In vitro fertilization (IVF) with conventional insemination<br />

is an effective therapy for unexplained infertility, but total fertilization failure<br />

(TFF) remains high. A recent meta-analysis suggests that intracytoplasmic<br />

sperm injection (ICSI) improves fertilization rates and decreases TFF in couples<br />

with unexplained infertility. (1) However, pregnancy data are lacking.<br />

We aim to determine if ICSI improves clinical pregnancy rate and live birth<br />

rate in couples with unexplained infertility using the SART Database.<br />

DESIGN: Retrospective cohort.<br />

MATERIALS AND METHODS: All IVF cycles reported to the Society<br />

for Assisted Reproductive Technology (SART) between 2004 to 2011 were<br />

examined. IVF cycles were included if unexplained infertility was the only<br />

diagnosis. The first fresh cycle for each patient was analyzed. Exclusion<br />

criteria included: first fresh cycle before 2004, both ICSI and conventional<br />

insemination in the same cycle, second-day ICSI, and missing data for<br />

age, BMI, and smoking. Linear and logistic regression models were utilized<br />

to estimate fertilization rate, TFF rate, transfer rate, number of embryos<br />

transferred and cryopreserved, clinical pregnancy rate, miscarriage rate,<br />

and live birth rate. Adjusted models controlled for age, BMI, smoking,<br />

FSH dose, and number of oocytes retrieved. This study has 95% power to<br />

detect a 10% increase in clinical pregnancy and live birth rates. A p value<br />

of


O-<strong>17</strong>1 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 11:45 AM<br />

DELETERIOUS EFFECTS OF MALE OBESITY ON REPRODUC-<br />

TIVE PARAMETERS AMONG A LARGE MULTI-INSTITUTIONAL<br />

COHORT. J. M. Bieniek, a J. A. Kashanian, b C. M. Deibert, c<br />

E. D. Grober, a R. Brannigan, b J. I. Sandlow, c K. A. Jarvi. a a University of<br />

Toronto, Toronto, ON, Canada; b Northwestern University, Chicago, IL;<br />

c Medical College of Wisconsin, Milwaukee, WI.<br />

OBJECTIVE: Obesity has been suggested as a contributing factor in male<br />

subfertility but correlations with semen parameters have been mixed, therefore,<br />

this study was designed to analyze the relationships between body mass<br />

index (BMI) and measures of male fertility among a large multi-institutional<br />

cohort.<br />

DESIGN: Retrospective review of prospectively-collected demographic<br />

data to calculate measures of obesity from three North American male<br />

infertility clinics with correlation to semen and reproductive hormonal<br />

parameters.<br />

MATERIALS AND METHODS: Self-reported or measured height and<br />

weight were captured in review board-approved databases at male infertility<br />

clinics since 2002. Men with semen analysis or reproductive hormone parameters<br />

at initial evaluation were included to reduce effects of infertility-related<br />

treatments. BMI was calculated for all patients with comparisons to reproductive<br />

parameters performed utilizing ANOVA, chi squared, and Spearman’s<br />

rank correlations as appropriate with p< .05 reported as significant.<br />

RESULTS: Complete height and weight data to calculate BMI was available<br />

for 4440 patients with a mean age of 36.1 (7.6) years. Based on WHO<br />

definitions, 30.9% of the cohort was normal weight (BMI 18.5-24.9), 45.1%<br />

overweight (25-29.9), and 23.3% obese (>30). Semen analysis and reproductive<br />

hormone results were collected for 4236 (95.4%) and 2973 (67.0%) men,<br />

respectively. The gonadotropins FSH and LH demonstrated weak positive<br />

correlations with BMI with only LH reaching significance (r¼0.04,<br />

p¼0.06; r¼0.06, p¼0.01). Testosterone (r¼ -0.27, p


discarded (egg donation) and male factor is diagnosed with abnormal protamine<br />

ratio compared to previous ejaculate egg donation cycles.<br />

DESIGN: Unicentric, retrospective study (Jan.2014-Apr.<strong>2015</strong>) including<br />

couples who underwent at least one previous egg donation cycle with sperm<br />

ejaculate. Protamine ratio was studied when poor or null blastocyst rate was<br />

observed and no pregnancy was achieved. Cases with altered protamine ratio<br />

were indicated for TESA in the following cycle.<br />

MATERIALS AND METHODS: 38 couples from egg donation treatment<br />

were included. Gardner’s classification was used for embryo<br />

grading. Protamine ratio was measured through FertiCertTM. TESA<br />

was performed in subsequent cycle if altered protamine ratios were diagnosed.<br />

To compare cycle outcome, fertilization, blastocyst, AA blastocyst<br />

and pregnancy rates were studied in both. Accepted by our Institutional<br />

Review Board.<br />

RESULTS: Fertilization rate was significantly higher in TESA compared<br />

to ejaculate (85.0% vs 69.3%; p¼0.001). Blastocyst formation rate was also<br />

significantly higher in TESA group than in ejaculate (47.6% vs <strong>21</strong>.3%;<br />

p¼0.000). Regarding AA quality blastocyst formation rate, this was significantly<br />

different as well (30.9% vs 9.3%; p¼0.001). When implantation rate is<br />

analyzed, it increases from ejaculate (12.5%) to TESA group (75%) being<br />

this difference significantly different (p¼0.0018). T-Student and c2 tests<br />

with SPSS were used to analyze data.<br />

CONCLUSIONS: Improved cycle outcomes were observed in terms of<br />

fertilization, blastocyst formation and implantation rates when TESA was<br />

used for fertilization, compared to previous ejaculate egg donation cycles<br />

in patients with altered protamine ratios. More cases are needed. A randomized<br />

controlled trial would be difficult due to TESA is an invasive technique<br />

and should not be performed without firm indication. The improvement<br />

observed is difficult to determine whether it comes from unprotected DNA<br />

or epigenetic factors. This study provides a new proceeding strategy for couples<br />

who are not prepared for donor sperm. These data corroborate the relationship<br />

between altered protamine ratios and infertility, but also suggest that<br />

a different sperm sources can result in an improvement of fertilization and<br />

blastocyst rates and thus in the cycle outcome.<br />

O-<strong>17</strong>4 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:30 PM<br />

PREGNANCY LOSS IS SIGNIFCANTLY INCREASED IN MALE<br />

FACTOR OLIGOASTHENOZOOSPERMIA PATIENTS DESPITE<br />

THE TRANSFER OF A EUPLOID BLASTOCYST. J. M. Stevens, a<br />

S. McCormick, a A. Schneiderman, a W. B. Schoolcraft, b M. Katz-Jaffe. b<br />

a Fertility Laboratories of Colorado, Lone Tree, CO; b Colorado Center for<br />

Reproductive Medicine, Lone Tree, CO.<br />

OBJECTIVE: Male factor (MF) infertility contributes to half of all infertility<br />

cases worldwide. Compromised quality of the male gamete is known<br />

to impact fertilization potential, embryo quality and recurrent miscarriage<br />

(RM). Studies have shown that men with MF or RM have increased sperm<br />

aneuploidy compared to controls. This study explored the influence of oligoasthenozoospermia<br />

(sperm motility


OBJECTIVE: The role of the endocannabinoid system (ECS) in the pathogenesis<br />

of endometriosis is still under investigation. In this study, the<br />

changes on ECS related to apoptosis and autophagy processes have been<br />

analized in human granulosa cells (GCs) from women affected by endometrioma<br />

DESIGN: This prospective non-randomized study has been carried out<br />

from January to September 2014 on women undergoing a controlled ovarian<br />

hyperstimulation for an IVF treatment. In particular, GCs were collected<br />

from both ovaries of 10 women with a diagnosis of unilateral ovarian endometrioma<br />

at the time of oocytes retrieval. 9 women with male, idiopathic or<br />

tubal-factor infertility diagnosis were selected for the control group. The<br />

three experimental groups matched for female age (36.24.1 vs 35.42.6)<br />

MATERIALS AND METHODS: GCs obtained from follicles aspirates<br />

were isolated from red blood cells and follicular fluid by density gradient<br />

centrifugation. Q-PCRs were performed with the SYBR green method in<br />

an iQ5 iCycler thermal cycler using bactin and GAPDH as reference genes.<br />

Data are presented as mean S.D. Two-Way ANOVA followed by Tukey<br />

test as Multiple comparisons test, was used for comparison among experimental<br />

groups. All statistical analyses were performed using the statistical<br />

software package Prism5 (Graphpad Software, Inc. USA) with significance<br />

accepted at P


O-<strong>17</strong>9 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:15 PM<br />

STEROIDOGENESIS IN ENDOMETRIOSIS: THE IMPORTANCE OF<br />

GATA6. L. A. Bernardi, S. E. Bulun, M. T. Dyson. Obstetrics and Gynecology,<br />

Feinberg School of Medicine, Northwestern University, Chicago, IL.<br />

OBJECTIVE: The production of steroids from cholesterol is initiated by<br />

the concerted effort of multiple factors coordinated in a tissue specific<br />

pattern. The same steroidogenic factors are also essential for the development<br />

of endometriosis. The PKA pathway and signaling from steroidogenic factor<br />

1 (NR5A1) critically regulate steroidogenic gene expression, and these along<br />

with GATA6 have been implicated in the pathogenesis of endometriosis. This<br />

study evaluated the combinatorial impact of GATA6 and NR5A1 on steroidogenesis<br />

in diseased cells, and expanded these findings to demonstrate how<br />

these pathways permit healthy cells to acquire the steroidogenic potential<br />

that drives endometriosis.<br />

DESIGN: Basic science, in-vitro.<br />

MATERIALS AND METHODS: Ectopic endometriotic fibroblasts<br />

(OSIS), obtained from the cyst walls of endometriomas, and normal eutopic<br />

endometrial fibroblasts (NoEM), obtained from the endometrium of subjects<br />

without endometriosis, were isolated in accordance with IRB-approved protocols.<br />

Human foreskin fibroblasts (BJ) were purchased commercially.<br />

NR5A1 and GATA6 were depleted in OSIS by the reverse transfection of<br />

gene-specific silencer select siRNAs in comparison to a scrambled siRNA.<br />

NoEM and BJ cells, which express little endogenous NR5A1 or GATA6,<br />

were transduced with adenoviruses bearing CMV-driven NR5A1, GATA6,<br />

or an empty CMV-null control. Expression of steroidogenic genes including<br />

steroidogenic acute regulatory protein (StAR), cholesterol side chain cleavage<br />

enzyme (CYP11A1), 3-beta-hydroxysteroid dehydrogenase type 2<br />

(HSD3B2), <strong>17</strong>a-hydroxylase (CYP<strong>17</strong>A1), and aromatase (CYP19A1),<br />

were assessed using RT-qPCR and immunoblotting.<br />

RESULTS: Depleting NR5A1 in OSIS significantly reduced the expression<br />

of StAR and CYP11A1, and attenuated HSD3B2, CYP<strong>17</strong>A1 and<br />

CYP19A1 as compared to controls. Depleting GATA6 in OSIS attenuated<br />

the PKA-dependent induction of CYP<strong>17</strong>A1 and CYP19A1 compared to untreated<br />

controls. In both NoEM and BJ cells, the expression of StAR and<br />

CYP11A1 was induced by NR5A1 alone; however, expression was significantly<br />

higher in NoEM than in BJ cells. In contrast, HSD3B2, CYP<strong>17</strong>A1,<br />

and CYP19A1 were synergistically increased by NR5A1 and GATA6<br />

compared to either gene alone in NoEM, while in BJ cells expression of<br />

HSD3B2, CYP<strong>17</strong>A1 and CYP19A1 was minimal.<br />

CONCLUSIONS: Silencing of NR5A1 and GATA6 in endometriotic tissue<br />

leads to decreased expression of key steroidogenic enzymes. In non-endometriotic<br />

cells the ability of NR5A1 and GATA6 to work synergistically to<br />

increase steroid production depends on a tissue’s inherent potential for steroidogenesis.<br />

In conclusion, GATA6 appears to play an important role in<br />

the regulation of steroid production and ultimately in the pathogenesis of<br />

endometriosis.<br />

Supported by: Research Supported by NIH R37HD038691-12S1 and a<br />

grant from the Friends of Prentice at Northwestern Memorial Hospital.<br />

O-180 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:30 PM<br />

PATIENTS WITH ENDOMETRIOSIS DO NOT HAVE A HIGHER<br />

RATE OF ANEUPLOIDY. C. R. Juneau, J. M. Franasiak,<br />

M. D. Werner, E. J. Forman, T. Molinaro, R. T. Scott. RMA, NJ, NJ.<br />

OBJECTIVE: The diminution in outcomes which impact women with<br />

endometriosis undergoing IVF have variably been attributed to alterations<br />

in endometrial receptivity, decreased oocyte quality, and premature depletion<br />

of follicular reserve. Little data are available to determine if the changes in<br />

ovarian reserve experienced by women with endometriosis adversely impacts<br />

oocyte maturation ultimately resulting in an increased risk of embryonic<br />

aneuploidy. This study seeks to address that question.<br />

DESIGN: Retrospective cohort.<br />

MATERIALS AND METHODS: Patients participating in their first IVF<br />

cycle at a single infertility clinic from 2009-2014 with comprehensive chromosome<br />

screening (CCS) were included. Patients with endometriosis were<br />

identified through surgical diagnosis or by ultrasound findings consistent<br />

with persistent space-occupying disease whose sonographic appearance is<br />

consistent with endometriosis. Patients were divided into the SART age<br />

groups, and aneuploidy rates in each group were compared to those in<br />

3301 age-matched patients in the general IVF population undergoing CCS.<br />

While the control group included all patients in the general IVF population,<br />

including patients with endometriosis, the proportion of patients with endometriosis<br />

only accounts for 1% of this population and is therefore not likely<br />

to confound results. Statistical analysis was performed using c2 test of proportions.<br />

An alpha error of 0.05 was accepted.<br />

RESULTS: There were 253 patients with endometriosis who produced<br />

1283 blastocysts that met criteria for inclusion. Mean age of patients with<br />

endometriosis was 35.7 4.0 years (range 24-46 years). The median antimullerian<br />

hormone level was 1.70 ng/mL (range 0.16-15.36 ng/mL). When<br />

aneuploidy rates in patients with endometriosis and aneuploidy rates in the<br />

general IVF population were stratified by SART age groups and compared,<br />

there was no statistical difference (table 1).<br />

CONCLUSIONS: Patients with endometriosis undergoing IVF have aneuploidy<br />

rates equivalent to their age-matched peers in the general IVF population.<br />

The rate of aneuploidy is not increased in patients with endometriosis<br />

undergoing IVF.<br />

SART age<br />

group<br />

Rate of aneuploidy<br />

in patients with<br />

endometriosis<br />

Rate of aneuploidy<br />

in general IVF<br />

population<br />

p-value<br />

26.7% (160/600) 28.9% (1974/6819) 0.2556<br />

35-38 years 33.3% (128/384) 37.4% (1369/3659) 0.1285<br />

38-40 years 54.0% (129/239) 52.7% (1629/3084) 0.6581<br />

41-42 years 72.9% (35/48) 71.2% (857/1203) 0.9288<br />

>42 years 75.0% (9/12) 84.6% (342/404) 0.6141<br />

FIBROIDS<br />

O-181 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 11:15 AM<br />

ELAGOLIX FOR THE MANAGEMENT OF HEAVY MENSTRUAL<br />

BLEEDING (HMB) ASSOCIATED WITH UTERINE FIBROIDS<br />

(UF): RESULTS FROM A PHASE 2A PROOF-OF-CONCEPT<br />

STUDY. D. F. Archer, a K. Chwalisz, b R. Feldman, c E. A. Stewart, d<br />

A. S. Lukes, e J. North, b J. Gao, b L. A. Williams, b R. Jain. b a Department<br />

of Obstetrics & Gynecology, Eastern Virginia Medical School, Norfolk,<br />

VA; b AbbVie, Inc., North Chicago, IL; c Miami Research Associates, Pinecrest,<br />

FL; d Mayo Clinic, Rochester, MN; e Consulting Teaching Professor,<br />

Duke University Med, Chapel Hill, NC.<br />

OBJECTIVE: Evaluate safety and efficacy of elagolix (an oral GnRH antagonist)<br />

alone and combined with low-dose estrogen + progestin add-back therapy<br />

regimens in premenopausal women with HMB associated with UF.<br />

DESIGN: Phase 2a study with 3 randomized, double-blind, placebo<br />

(PBO)-controlled (RPC) and 3 open-label (OL) cohorts. Three elagolix<br />

dosing regimens were used, alone and in combination with 2 low-dose<br />

add-back therapy regimens for 3 mo (Table). Primary efficacy endpoint<br />

was mean change in menstrual blood loss (MBL) from baseline to the final<br />

month (last 28 d on treatment); percentage change from baseline in MBL<br />

was a secondary HMB efficacy endpoint.<br />

MATERIALS AND METHODS: Premenopausal women aged 20-49 y<br />

with UF (R2 cm), regular menses every 24-35 d, and HMB (MBL >80<br />

mL during 2 screening cycles) were eligible. MBL was assessed at screening<br />

and monthly during treatment as measured with the alkaline hematin method.<br />

Between-group comparisons were tested using analysis of covariance, with<br />

treatment as a factor and baseline as a covariate.<br />

RESULTS: 271 women were randomized; 228 completed. Approximately<br />

75% of women were black, with a mean age of 41.8 y. A substantial reduction<br />

in MBL (ie, improvement in HMB) at the final month was observed in all elagolix-containing<br />

treatment groups (Table). Changes from baseline in MBL<br />

with all elagolix doses were statistically significant (P%0.001) vs PBO in<br />

RPC cohorts. Overall, adverse event (AE) rates were similar across treatment<br />

groups, except hypoestrogenic events such as hot flush. Hot flush was the<br />

most frequent AE in elagolix-only treatment groups (45%-63%) vs PBO<br />

(6%-19%); rates were 26% (E2:NETA) and 19% (E2:P) with add-back therapy.<br />

AEs resulted in study drug discontinuation in 24 women (elagolix alone,<br />

18/160 [11%]; elagolix + add-back therapy, 2/61 [3%]; PBO, 4/50 [8%]).<br />

CONCLUSIONS: All elagolix-containing treatment regimens substantially<br />

improved HMB associated with UF; 300 mg BID was the most<br />

efficacious dose. Both low-dose add-back regimens reduced hot flush by<br />

40%-50% vs corresponding elagolix-only dose groups, with minimal<br />

impact on HMB efficacy endpoints.<br />

e70 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


Table. Key Efficacy Parameters by Total Daily Dose of Elagolix<br />

Elagolix<br />

Daily Dose<br />

0mg<br />

(PBO)<br />

Study<br />

Type/Elagolix<br />

Dosing/<br />

Add-Back<br />

Therapy/n<br />

Mean<br />

(SD) BL<br />

MBL, mL<br />

n¼50 251.7<br />

(160.3) to<br />

200 mg RPC/100<br />

mg BID/<br />

none/n¼33<br />

400 mg RPC/200 mg<br />

BID/none/<br />

n¼35<br />

400 mg RPC/400<br />

mg QD/none/<br />

n¼32<br />

400 mg OL/200<br />

mg BID/<br />

E2:NETA/<br />

n¼34<br />

600 mg RPC/300<br />

mg BID/<br />

none/<br />

n¼30<br />

600 mg OL/600 mg<br />

QD/none/<br />

n¼30<br />

600 mg OL/300 mg<br />

BID/E2:P/<br />

n¼27<br />

Mean<br />

(SE) MBL, Change From<br />

Baseline to Final Month<br />

(Last 28 d on Treatment)<br />

mL %<br />

349.2 (424.1)<br />

-114.8<br />

(32.6) to<br />

8.9 (36.7)<br />

-7.7<br />

(11.1) to<br />

-41.2 (6.6)<br />

269.4 (163.2) -181.0 (25.2) -72.0 (7.7)<br />

335.1 (322.7) -252.8 (24.4) -80.2 (13.6)<br />

<strong>21</strong>3.7 (108.1) -197.0 (25.5) -83.3 (7.7)<br />

247.7 (<strong>17</strong>7.7) -192.3 (191.5) -79.6 (43.6)<br />

206.3 (125.1) -237.0 (<strong>21</strong>.3) -98.0 (4.6)<br />

<strong>21</strong>5.6 (122.8) -189.1 (151.2) -88.6 (39.6)<br />

258.9 (207.5) -2<strong>17</strong>.1 (157.5) -85.4 (28.1)<br />

exosome treatment to measure level of proliferation of the treated cells vs<br />

vehicle control.<br />

RESULTS: Hypoxia induced dramatic increase in exosomes production<br />

by hFSC (up to 6 fold increase) compared to normoxic-grown hFSC (P ¼<br />

0.005). The hypoxic and normoxic exosomes were similar in size about<br />

200 nm. The hypoxic exosomes induced significantly higher rate of proliferation<br />

in primary fibroid cell proliferation compared to normoxic exosomes<br />

(P ¼ 0.02) after 48 hours of treatment. Further characterization of effect of<br />

hypoxia on exosome contents (such as miRNA profile) is ongoing in our laboratory.<br />

CONCLUSIONS: Our data presents, for the first time, a mechanism by<br />

which hypoxic-exposed tumor forming fibroid stem cells would enhance tumor<br />

formation. Hypoxic hFSCs exert pro-proliferative paracrine stimulatory<br />

effect on primary fibroid cells by producing abundant amount of exosomes<br />

with robust oncogenic potential. Further understanding of the cross-talk between<br />

human fibroid tumor-forming stem cells and surrounding differentiated<br />

tumor cells can provide novel therapeutic targets for treatment of<br />

symptomatic uterine fibroids.<br />

O-183 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 11:15 AM<br />

WITHDRAWN<br />

BID¼twice daily; BL¼baseline; E2:NETA¼continuous/combined estrogen<br />

(0.5 mg) + norethindrone acetate (0.1 mg); E2:P¼1 mg continuous estrogen<br />

+ 200 mg oral progesterone on days<br />

<strong>17</strong>-28; MBL¼menstrual blood loss; OL¼open label; PBO¼placebo;<br />

QD¼once daily; RPC¼randomized PBO-controlled; SD¼standard deviation;<br />

SE¼standard error.<br />

Supported by: AbbVie, Inc.<br />

O-182 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 11:30 AM<br />

EXOSOMES FROM HYPOXIA-DRIVEN HUMAN FIBROID STEM<br />

CELLS ACCELERATE TUMOR GROWTH. S. Brakta,<br />

S. M. Shalaby, M. P. Diamond, A. Zimmerman, I. Helwa, Y. Liu,<br />

S. A. Mohamed, L. Gavrilova-Jordan, A. Al-Hendy. GRU, Augusta, GA.<br />

OBJECTIVE: Exosomes are nanovesicles, the smallest extracellular vesicles<br />

identified so far. They are secreted from a variety of cell types and play<br />

an important role in intercellular communication, often inducing physiological<br />

changes in recipient cells by transferring bioactive lipids, nucleic acids,<br />

and proteins. They involve many physiologic processes such as cell growth,<br />

neuronal communication, immune response activation, and cell migration. In<br />

the case of cancer, exosomes may transfer angiogenic proteins, oncogenes or<br />

oncogenic miRNA from tumor forming stem cells to surrounding cells and<br />

promote tumor growth. We have recently characterized human tumor forming<br />

stem cells from uterine fibroid lesions using specific surface markers. In<br />

this work, we wanted to investigate if human fibroid stem cells do indeed produce<br />

exosomes, factors that regulate this process (such as hypoxia) and<br />

possible effect on surrounding differentiated fibroid cells.<br />

DESIGN: Laboratory research studies using human cells.<br />

MATERIALS AND METHODS: For exosome isolation, human fibroid<br />

stem cells were cultured in media with exosome-free FBS for 72 hours at normoxic<br />

or hypoxic (2% O2) conditions. Ten ml of culture media was collected<br />

from each flask and Exoquick TC reagent was added to isolate exosomes according<br />

to manufacturer’s protocol. Electron microscopy and ZetaView were<br />

used to characterized human fibroid stem cell (hFSC) derived hypoxic and<br />

normoxic exosomes. .For proliferation studies, primary human (differentiated)<br />

fibroid cells were cultured in 96 well plates. Media was then changed<br />

to exosome free-FBS and the cells were treated with hypoxic versus<br />

normoxic exosomes. MTT assay was carried out at 24 and 48 hours post<br />

O-184 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:00 PM<br />

EXTRACELLULAR MATRIX PRODUCTION DECREASES IN ULI-<br />

PRISTAL ACETATE (UPA) TREATED HUMAN LEIOMYOMAS IN<br />

VITRO AND IN VIVO RESULTING IN DECREASED FIBROID<br />

SIZE. J. Cox, a M. Malik, b J. L. Britten, b A. Patel, b L. Nieman, a<br />

W. H. Catherino. b a National Institutes of Health, Bethesda, MD; b Uniformed<br />

Services University of the Health Scienc, Bethesda, MD.<br />

OBJECTIVE: Multiple prospective, randomized trials have demonstrated<br />

that UPA significantly reduced fibroid volume. In our trial, study patients underwent<br />

hysterectomy and surgical specimens were collected. Using these<br />

tissues, we examined the impact of UPA treatment on extracellular matrix<br />

FERTILITY & STERILITY Ò<br />

e71


(ECM) expression in leiomyomas compared to placebo treated patients. To<br />

further characterize the impact of UPA, we used 3D leiomyoma and myometrial<br />

cultures to assess ECM.<br />

DESIGN: Sub-analysis of randomized placebo controlled trial and laboratory<br />

analysis of clinically-relevant pharmacologic concentrations of UPA on<br />

3D cell cultures.<br />

MATERIALS AND METHODS: A total of 10 patients (5 placebo, 5<br />

treated with 10mg UPA) that underwent hysterectomy and tissue preservation<br />

were identified. Pre and post fibroid volume was evaluated by MRI measurements.<br />

Proteins related to ECM production were evaluated: fibronectin<br />

(FN), versican (VER), and collagen 1A1 (COL). For in vitro analysis, leiomyoma<br />

cells were grown in collagen gel and treated with different concentrations<br />

of UPA for 48hrs. RNA was extracted for qRTPCR analysis.<br />

Protein concentration and distribution was analyzed using immunohistochemistry<br />

(IHC) and Western Blot.<br />

RESULTS: Of treated patients, 80% had decrease or no change in fibroid<br />

volume over 3 months compared to placebo, with a mean decrease of 16%.<br />

Though increased expression of ECM genes was observed in treated surgical<br />

specimens, 80% of the treated specimens demonstrated a significant decrease<br />

in VER protein as compared to placebo. FN protein was decreased in 60% of<br />

treated samples but no consistent alteration in COL was found compared to<br />

placebo surgical specimens. 3D leiomyoma cultures exhibited a UPA<br />

concentration-dependent decrease in VER (1.62+/- 0.13-fold) and FN<br />

(2.58+/-0.22-fold) proteins as compared to untreated cells. This effect was<br />

highlighted by IHC, where both surgical specimens and 3D cell cultures<br />

demonstrated decreased amount of FN and VER in UPA treated leiomyoma<br />

samples.<br />

CONCLUSIONS: UPA treatment decreased fibroid volume in placebocontrolled,<br />

randomized trials. Gene expression and protein production are<br />

altered in leiomyoma tissue in vivo and in vitro compared to placebo controls.<br />

Proteoglycans like VER are known to contribute to the increased water content<br />

of tissue and the large decrease of VER may thus contribute to decreased<br />

fibroid volume on UPA treatment.<br />

Supported by: Intramural grant from Uniformed Services University of<br />

Health Sciences, R<strong>21</strong> grant R085193713, and intramural research of the Program<br />

in Reproductive and Adult Endocrinology, NICHD, NIH.<br />

O-185 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:15 PM<br />

ALTERED DNA REPAIR GENES IN HUMAN UTERINE FIBROIDS<br />

ARE EPIGENETICALLY REGULATED VIA EZH2 HISTONE<br />

METHYLTRANSFERASE. Q. Yang, L. Elam, A. Laknaur,<br />

L. Gavrilova-Jordan, J. Lue, M. P. Diamond, A. Al-Hendy. Georgia Regents<br />

University, Augusta, GA.<br />

OBJECTIVE: Uterine fibroids (UFs) are benign smooth muscle neoplasms<br />

affecting up to 80% of women. Treatment of symptomatic UFs is a significant<br />

economic burden. Etiology and pathogenesis of UFs is complex. Several genetic<br />

abnormalities related to pathogenesis of UFs have been investigated<br />

including deletions in 7q, trisomy of chromosome 12, rearrangements in<br />

the HMGA2 gene, and mutations in exon 1 and 2 of the MED12 gene. Previous<br />

findings suggest that a low DNA damage repair function may be<br />

responsible for UF formation. The objective of this study is to investigate<br />

whether the DNA damage repair system is altered in UFs and thereby<br />

contribute to an increased risk of UFs development.<br />

DESIGN: Laboratory research studies using human tissues and human<br />

cell lines.<br />

MATERIALS AND METHODS: Surgically removed fresh human<br />

fibroid and adjacent myometrial tissues were collected and primary cells<br />

isolated from fibroids and adjacent myometrial tissues were used for in<br />

vitro studies. The primary cells were grown in media containing inhibitors<br />

of DNMT, HDAC, and EZH2. RNA was isolated using Trizol extraction<br />

method and subjected to cDNA synthesis using Superscript III cDNA<br />

synthesis kit. cDNA was loaded onto the DNA damage signaling pathway<br />

PrimerPCR array plate (H96). RNA levels of over 80 genes were detected<br />

on Bio-Rad CFX96 real-time PCR system. Chromatin Immunoprecipitation<br />

assay was performed to determine the enrichment of H3K27me3<br />

and H3K4me3 in the promoter region of MSH2. Quantitative PCR was<br />

performed to determine the mRNA expression levels of DNA damage<br />

repair genes.<br />

RESULTS: PrimePCR array assay indicated that the expression levels of<br />

over 20 DNA repair genes are altered between fibroid and adjacent myometrial<br />

tissue. Among them, the expression levels of MSH2, a DNA mismatch<br />

repair gene, were significantly downregulated in human fibroid tissue as<br />

compared to the adjacent myometrial tissue. A similar finding was<br />

observed in primary fibroid cells as compared with adjacent myometrial<br />

cells. The fibroid primary cells treated with DZNep (inhibitor of EZH2) exhibited<br />

a marked increase in MSH2 expression in a dose dependent manner.<br />

ChIP assay demonstrated that the enrichment of H3K27me3 in the promoter<br />

region of MSH2 were significantly decreased in DZNep-treated<br />

fibroid cells as compared to vehicle control. In contrast, the recruitment<br />

of H3K4me3 in MSH2 promoter region was not altered in response to<br />

DZNep treatment. These data suggest that EZH2 decreases expression<br />

levels of DNA repair gene MSH2 through epigenetic mark H3K27me3 in<br />

uterine fibroids.<br />

CONCLUSIONS: Identification of the novel EZH2 targeted gene and<br />

MSH2 expression downregulation in fibroid tissue may explain the impaired<br />

DNA repair and high frequency of genetic abnormalities found in fibroid tissue<br />

comparing to normal myometrium. We found that EZH2 decreases the<br />

expression of MSH2 through epigenetic mark H3K27me3. Understanding<br />

of a complex pathogenesis of UF may lead to effective targeted treatment<br />

of symptomatic in the future.<br />

References:<br />

Yang Q, Mas A, Diamond MP, et al. The Mechanism and Function of Epigenetics<br />

in Uterine Leiomyoma Development. Reprod Sci. <strong>2015</strong> Apr<br />

28. pii: 1933719115584449.<br />

Supported by: This work was Supported in part by the Georgia Regents<br />

University Startup package and the National Institutes of Health grant<br />

HD04622811.<br />

O-186 Tuesday, <strong>October</strong> 20, <strong>2015</strong> 12:30 PM<br />

SYNERGISTIC EFFECT OF ESTROGEN AND PROGESTERONE<br />

ON MYOMETRIAL STEM CELL EXPANSION IN VIVO. A. Mas, a<br />

L. Elam, a M. P. Diamond, a C. Simon, b A. Al-Hendy. a a Department of Obstetrics<br />

and Gynecology, Georgia Regents University, Augusta, GA; b Fundacion<br />

Instituto Valenciano de Infertilidad/University of Valencia, INCLIVA,<br />

Paterna (Valencia), Spain.<br />

OBJECTIVE: Uterine fibroids are common gynecologic tumors stimulated<br />

by estrogen (E2) and progesterone (P4), and likely originate from<br />

abnormal myometrial stem cells. We have recently characterized human<br />

myometrial stem cells using specific surface markers, and we demonstrated<br />

lower levels of E2 receptor a and P4 receptors A/B in these cells<br />

compared to differentiated myometrial cells. However, it has also been reported<br />

that these myometrial stem cells are still influenced by E2 and P4<br />

effects in vitro, via a paracrine pathway by the surrounding differentiated<br />

myometrial cells. Our goal in this work is to evaluate the effect of both<br />

E2 and P4 on the number and distribution of myometrial stem cells in<br />

Long Evans rats.<br />

DESIGN: Laboratory research studies using a murine model.<br />

MATERIALS AND METHODS: Steroid hormone deprivation was<br />

achieved by the ovariectomy of 16 female Long Evans rats at 4 months of<br />

age, followed by a 4 weeks rest period, when part of the myometrial tissues<br />

was collected for further analysis. For the hormone supplement experiments,<br />

ovariectomized rats were implanted with hormone pellets subcutaneously for<br />

4 weeks (60-d time- release, 1.5mg b-estradiol pellet; and/or 60-d timerelease,<br />

200mg progesterone pellet). In addition, 4 hormonally intact rats<br />

were used as controls. Stem cell isolation was performed by digesting endometrium-free<br />

myometrial tissues collected from different locations (cervix<br />

and horns) followed by flow cytometry (FACS) analysis. At that time, the percentage<br />

(%) of Stro1/CD44 stem cells was measured in rats supplemented<br />

with E2 or P4 alone or combined, versus control rats.<br />

RESULTS: Hormonal deprivation (by ovarectomy) caused an 8-fold drop<br />

in the number of myometrial stem cells in the cervical region and a 6-fold<br />

drop in the horns, compared to hormonally intact control (p


REPRODUCTIVE ENDOCRINOLOGY: CLINICAL<br />

O-187 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:15 AM<br />

DOES A FROZEN EMBRYO TRANSFER AMELIORATE THE EF-<br />

FECT OF ELEVATED PROGESTERONE ON THE ENDOMETRIUM<br />

SEEN IN FRESH TRANSFER CYCLES: A PAIRED AND UNPAIRED<br />

ANALYSIS. M. W. Healy, a G. Patounakis, a M. T. Connell, a K. Devine, b<br />

A. DeCherney, a M. Levy, b M. J. Hill. a a NIH, Bethesda, MD; b Shady Grove<br />

Fertility Reproductive Science Center, Rockville, MD.<br />

OBJECTIVE: Many ART programs are cryopreserving embryos when<br />

progesterone (P) is elevated in a fresh IVF cycle, but there is minimal evidence<br />

of the effect of P on oocyte quality and thus embryo performance in<br />

subsequent frozen embryo transfer (FET) cycles. Our objective was to<br />

demonstrate the effect of P on the day of trigger in fresh IVF transfer cycles<br />

compared to subsequent FET cycles.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Autologous IVF cycles with fresh transfers<br />

and FET cycles from 2011-2013 were included if P was measured on the<br />

day of trigger in the initial fresh cycle. The primary outcome was clinical<br />

pregnancy. Analysis 1 was a paired analysis of patients who underwent<br />

both a fresh transfer and subsequent FET. Analysis 2 included unpaired comparisons<br />

of data from all fresh transfers and all FET transfers. GEE modeling<br />

was performed to account for patients with multiple cycles and to control for<br />

age, embryo quality, embryo stage, and number of transferred embryos.<br />

Interaction testing was performed to determine if the effect of P was similar<br />

in fresh transfer and FET cycles. P was treated as a continuous variable<br />

except in additional analyses comparing P threshold of 2ng/mL. Subgroup<br />

analysis was performed in good quality blastocyst transfers.<br />

RESULTS: P was negatively associated with pregnancy in fresh transfers<br />

but not FET in both analysis 1 and 2. Interaction testing of P and cycle type<br />

indicated P had a different association with pregnancy outcomes in fresh<br />

transfers versus frozen transfers. When P was R 2ng/mL, pregnancy was<br />

more likely in FET versus fresh transfers in analysis 1 (56% vs 5%,<br />

P


draw and 52% completed a post-test survey. The mean AMH was 4.3 3.5<br />

ng/mL, range


DESIGN: A prospective, randomized controlled trial.<br />

MATERIALS AND METHODS: A prospective, randomized controlled<br />

trial was conducted to evaluate the outcomes in three different hormone<br />

replacement protocols for thawed blastocyst transfer. A total of 330<br />

women (median age 38.2 years) who were undergoing IVF at our clinic<br />

were enrolled. The trial registration number was UMIN000016919. Participants<br />

were separated into groups via computer-generated randomization<br />

as follows. Premarin (Ò) group (n¼110; 2.49 mg/day; from the<br />

second day of menstruation to the fifth day and 4.98 mg/day after the<br />

sixth day); Estrogel (Ò) group (n¼110; 2.16 mg/day in Estrogen conversion<br />

after the second day of menstruation); and Estrana tape (Ò) group<br />

(n¼110; 0.72 mg, 1 piece per 2 days). Statistical analyses were investigated<br />

using univariate regression. Statistical significance was defined<br />

as p


O-195 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:45 AM<br />

HIGH FREQUENCY OF INTERMEDIATE ALLELES OF FRAGILE<br />

X MENTAL RETARDATION 1 GENE (FMR1) IN CANDIDATES<br />

FOR OOCYTE DONATION. J. Guillen, R. Vassena, V. Vernaeve,<br />

A. Rodriguez. Clinica EUGIN, Barcelona, Spain.<br />

OBJECTIVE: The availability of preconceptional screening for X-<br />

linked and recessive diseases offers the opportunity to lower the risk to<br />

conceive an affected child. Genetic matching of donors and patients<br />

further improves ART service; however, the donor selection process might<br />

alter expected frequencies of certain diseases, potentially leading to variations<br />

of pretest risk. Fragile X syndrome (FXS) is the most common<br />

cause of inherited intellectual disability affecting approximately 1:4000-<br />

1:6000 births. FXS is caused by the expansion of CGG repeats in the<br />

FMR1 gene, and depending on CGG repeats number the population has<br />

been classified into: normal (N) 5-44 repeats, intermediate allele (IA)<br />

45-54, premutation (PM) 55-200, and full mutation (FM) >201. People<br />

with PM or FM could develop clinical symptoms and transmit the disease<br />

to the next generation. IA alleles do not confer an increased FXS risk for<br />

the very next generation, however, the 50-54 range may show instability,<br />

with potential expansions in subsequent generations. The aim of the study<br />

is to assess carrier status and ethnic variation of the FMR1 gene in oocyte<br />

donation (OD) candidates.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: 336 consecutive OD candidates aged 18<br />

to 35 years were tested for FXS (Asuragen Amplidex FMR1) in January-<br />

April <strong>2015</strong>. Women with a family history of FXS, mental retardation, chromosomal<br />

abnormalities, genetics disorders or neurological conditions were<br />

excluded from further screening. All women diagnosed with 45 o more<br />

CGG repeats were referred to genetic counseling. Relative frequency of IA<br />

status by ethnicity was assessed by Chi2 test, while the association between<br />

IA status and ovarian reserve was assessed using a logistic regression model,<br />

adjusted by age.<br />

RESULTS: Women ethnicities were: Mediterranean 127 (37.8%), European<br />

103 (30.6%), Latino American 61 (18.2%), Caribbean 27 (8.0%), other<br />

ethnicities 18 (5.4%). There were <strong>17</strong> carriers (1:20 overall), all of them IA; 13<br />

carriers had 45-49 CGG, while 4 had 50-54 CGG. The frequency of IA was<br />

7.4%, 7.8%, and 5.5% in women of Caribbean, European, and Mediterranean<br />

ethnicity, respectively (p>.05). IA status was not associated with the woman<br />

ovarian reserve (p>.05).<br />

CONCLUSIONS: Screening of potential donors for family history of<br />

FXS-like phenotypes lowers the frequency of PM and FM alleles in accepted<br />

donors. The high frequency of IA alleles in Caribbean, Europeans, and Mediterranean<br />

ethnicity is likely due to ethnic and geographical variability, a phenomenon<br />

described in other subgroup population.<br />

O-196 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:00 PM<br />

DO YOU REALLY KNOW YOUR ETHNICITY? COMPARING<br />

SELF-REPORTED AND GENETICALLY-PREDICTED ETHNICITY:<br />

CLINICAL IMPLICATIONS. R. Shraga, a S. L. Bristow, a<br />

A. Manoharan, a S. Yarnall, a A. Bisignano, a N. Kumar, a J. L. Frattarelli, b<br />

S. Ghadir, c J. Grifo. d a Recombine, New York, NY; b Fertility Institute of Hawaii,<br />

Honolulu, HI; c Southern California Reproductive Center, Beverly Hills,<br />

CA; d NYU Langone Medical Center, New York, NY.<br />

OBJECTIVE: Current guidelines published by professional societies<br />

recommend genetic carrier screening be offered on the basis of ethnicity.<br />

However, as the genetic pool homogenizes people are less aware of, or less<br />

likely to identify with, a specific ethnicity. Our goal was to investigate the accuracy<br />

of self-reported ethnicity as a basis for making clinical decisions.<br />

DESIGN: Retrospective.<br />

MATERIALS AND METHODS: Self-reported ethnicity was evaluated in<br />

1442 patients who received expanded carrier screening. Reports were gathered<br />

from patient requisition forms and during genetic counseling appointments.<br />

Comparisons were made to ancestral origin as predicted by a<br />

statistical model based on 672 SNPs validated using samples from the<br />

1000 Genomes Project. Documented informed consent was obtained from<br />

all patients.<br />

RESULTS: We found several discrepancies when comparing self-reported<br />

ethnicities on requisition forms, self-reported ethnicity during genetic counseling<br />

consults, and genetic ancestry. For example, only 33.3% of individuals<br />

who would be considered to be of Mediterranean ancestry based on genetic<br />

counseling consults self-reported this background. Further, in 27.2% of<br />

cases, patients predicted to be of South Asian descent by the statistical model<br />

and confirmed during consults self-reported a different ancestry. Finally, individuals<br />

who reported Latin American ancestry demonstrated a high degree<br />

of genetic admixture; the 3 highest contributing ancestral origin groups were<br />

European (0.5090.126), Native American (0.2340.185), and African<br />

(0.1010.183).<br />

CONCLUSIONS: Our comparison has demonstrated inconsistencies between<br />

self-reported ethnicities on requisition forms, self-reported ethnicity<br />

during genetic counseling consults, and genetic ancestry, highlighting the unreliability<br />

of patient self-reports. Basing carrier screening on patient-reported<br />

ethnicity may result in failure to screen for all appropriate genetic conditions<br />

and thus failure to identify individuals at high reproductive risk. This emphasizes<br />

the importance of offering pan-ethnic expanded carrier screening to all<br />

patients. Moreover, as identification of patient ethnicity is critical to the accurate<br />

calculation of residual reproductive risk, it may be worthwhile to factor<br />

both self-reported and genetic ancestry in these calculations. Finally, in<br />

the clinical setting, considerations should be made in how ethnic groups<br />

are defined for patient reporting purposes.<br />

O-197 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:15 PM<br />

REPRODUCTIVE DECISION-MAKING IN PATIENTS DIAGNOSED<br />

WITH BRCA MUTATIONS. J. Chan, a L. N. Johnson, a L. DiGiovanni, b<br />

C. Voong, a M. D. Sammel, c S. M. Domchek, b C. Gracia. d a Reproductive<br />

Endocrinology and Infertility, Hospital of the University of Pennsylvania,<br />

Philadelphia, PA; b Hematology Oncology, Hospital of the University of<br />

Pennsylvania, Philadelphia, PA; c Univ. of Pennsylvania, Perelman School<br />

of Medicine, Rose valley, PA; d University of Pennsylvania.<br />

OBJECTIVE: Despite our understanding of the medical implications of a<br />

BRCA mutation, there is little data on how knowledge of carrier status influences<br />

decisions about reproduction and parenthood. We aim to investigate<br />

how knowledge of BRCA carrier status impacts women’s decisions about relationships,<br />

childbearing and the use of preimplantation genetic diagnosis<br />

(PGD) and prenatal diagnosis.<br />

DESIGN: Cross-sectional survey.<br />

MATERIALS AND METHODS: BRCA mutation carriers were identified<br />

through the University of Pennsylvania’s Cancer Risk Program and through a<br />

national advocacy group for hereditary ovarian and breast cancers. Demographics,<br />

medical and social history, pregnancy and fertility history, and<br />

age at BRCA testing were obtained from a survey. Participants answered<br />

questions regarding how BRCA status influenced decisions about marriage<br />

and relationships, childbearing and fertility treatment, as well as attitude towards<br />

technologies such as PGD and prenatal diagnosis. Demographic characteristics<br />

were summarized and questionnaire responses compared using<br />

ranksum and Chi-square for continuous and categorical variables as appropriate.<br />

RESULTS: In the 1081 women who completed the survey, the mean age at<br />

the time of BRCA diagnosis was 44.1. 387 (36%) had a history of cancer. The<br />

majority was partnered at the time of learning of their BRCA mutation<br />

(84%). Amongst those who were not partnered, 37% felt pressure to find a<br />

partner after learning of their carrier status, 24% had more desire to get married<br />

and 46% reported that carrier status influenced what characteristics they<br />

were looking for in a partner. 66% had biological children and 64% reported<br />

that their families were complete at the time of diagnosis. Amongst women<br />

whose families were not complete, 38% reported that knowledge of their<br />

BRCA status impacted their decision to have biological children. Younger<br />

women (


O-198 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:30 PM<br />

PREIMPLANTATION GENETIC DIAGNOSIS (PGD) FOR SINGLE<br />

GENE DEFECTS (SGD): THE SIGNIFICANT VALUE OF CONCUR-<br />

RENT ANEUPLOIDY SCREENING. T. G. Nazem, a K. N. Goldman, a<br />

A. S. Berkeley, a J. Grifo. b a NYU School of Medicine, New York, NY;<br />

b NYU Langone Medical Center, New York, NY.<br />

OBJECTIVE: To evaluate outcome differences between patients undergoing<br />

trophectoderm (TE) biopsy and PGD for SGD with and without 24-chromosome<br />

aneuploidy screening.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Patients undergoing their first cycle of<br />

blastocyst culture, TE biopsy, and PGD with and without aneuploidy screening<br />

(array comparative genomic hybridization, aCGH) from July 2010 to August<br />

2014 were included. Cycles were excluded if performed for gender selection,<br />

HLA matching, or translocation. Primary outcomes included number (no.) and<br />

percentage (%) of blastocysts affected/unaffected/carrier of SGD, aneuploid/<br />

euploid blastocysts, and blastocysts eligible for transfer. FET cycle outcomes<br />

were analyzed: implantation rate (IR), spontaneous abortion rate (SABR),<br />

and live birth rate (LBR). Data were analyzed using student’s t-test and Fisher’s<br />

exact t-test and presented as mean S.D. (p


O-201 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:45 AM<br />

MALE OBESITY IS AN INDEPENDENT RISK FACTOR FOR<br />

TREATMENT FAILURE AMONG COUPLES WITH UNEX-<br />

PLAINED INFERTILITY IN THE ASSESSMENT OF MULTIPLE IN-<br />

TRAUTERINE GESTATIONS FROM OVARIAN STIMULATION<br />

(AMIGOS) TRIAL. A. J. Polotsky, a A. A. Allshouse, a S. Krawetz, b<br />

N. Santoro, a E. Eisenberg, c H. Zhang, d M. P. Diamond. e a University of Colorado,<br />

Aurora, CO; b Wayne, Detroit, MI; c NICHD, Bethesda, MD; d Yale,<br />

New Haven, CT; e Georgia Regents University, Augusta, GA.<br />

OBJECTIVE: Couples with unexplained infertility often seek advice on<br />

lifestyle changes but evidence-based medicine is limited. We sought to determine<br />

the impact of male body mass on the chance of pregnancy and live birth<br />

after fertility treatment.<br />

DESIGN: Secondary analysis of a randomized clinical trial (1).<br />

MATERIALS AND METHODS: 900 couples with unexplained infertility<br />

were randomized to ovulation induction with gonadotropins, or letrozole, or<br />

clomiphene citrate, followed by intrauterine insemination (IUI) for up to 4<br />

cycles. Females were 18-40 years old, had regular manses, at least 1 patent<br />

fallopian tube, normal uterine cavity and a male partner with at least 5 million<br />

total motile sperm (TMS) count. Analysis was limited to subjects for whom<br />

complete male partner information was available (n¼836). Logistic regression<br />

was conducted with clinical pregnancy and live birth as the outcomes.<br />

Each model was adjusted for male and female body mass index (BMI), treatment<br />

arm, and TMS.<br />

RESULTS: Two hundred and forty-five couples achieved clinical pregnancy<br />

(29.3%) and <strong>21</strong>4 women had a live birth (25.6%) in our analytic sample.<br />

Primary trial results have been presented (2). 272 male partners were<br />

obese with BMI greater than 30 k/m2 (32.5%) and 134 couples were<br />

concordant for male and female obesity (16.2%). Overall, there was a significant<br />

but very weak inverse correlation between male BMI and TMS<br />

(rho¼-0.08, p¼0.02) and a significant and moderate positive correlation between<br />

male and female BMI (rho¼0.41, p 30), n¼272 0.58 (0.37-0.91) 0.60 (0.38-0.96)<br />

CONCLUSIONS: In the AMIGOS trial, male obesity decreased the likelihood<br />

of live birth after fertility treatment by 40% as compared to couples<br />

with normal male weight. This effect was seen independent of female obesity,<br />

treatment arm and sperm parameters. Patients with unexplained infertility<br />

should be informed that obesity in the male partner reduces success of<br />

fertility treatment.<br />

References:<br />

1. Diamond et al. Fertil Steril <strong>2015</strong>;103:962-973.<br />

2. Diamond et al Fertil Steril 2014;102, p. e39.<br />

Supported by: This work was Supported by National Institutes of Health<br />

NIH)/Eunice Kennedy Shriver National Institute of Child Health and Human<br />

Development (NICHD) Grants U10 HD39005 (to M.P.D.), U10 HD38992 (to<br />

R.S.L.), U10 HD27049 (to C.C.), U10 HD38998 (to R.A.), HD055944 (to<br />

P.R.C.), U10 HD055936 (to G.M.C.),U10HD055925 (to H.Z.); and U10<br />

U54-HD29834 (to the University of Virginia Center for Research in Reproduction<br />

Ligand Assay and Analysis Core of the Specialized Cooperative<br />

Centers Program in Reproduction and Infertility Research). Most importantly,<br />

this research was made possible by the funding by American Recovery<br />

and Reinvestment Act. The content is solely the responsibility of the authors<br />

and does not necessarily represent the official views of the NICHD or NIH.<br />

O-202 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:00 PM<br />

PREVALENCE OF TESTICULAR MICROLITHIASIS IN 382 NON-<br />

VASECTOMIZED, AZOOSPERMIC MEN. J. Fedder. Centre of Andrology<br />

& Fertility Clinic, Odense University Hospital, Odense, Denmark.<br />

OBJECTIVE: To determine the prevalence of testicular microlithiasis<br />

(TM) in unselected azoospermic men and to relate the findings to specific<br />

etiological subgroups and to risk of malignancy.<br />

DESIGN: A retrospective study including 382 nonvasectomized, azoospermic<br />

men consecutively referred to our fertility clinic.<br />

MATERIALS AND METHODS: All ultrasonographic examinations were<br />

performed by JF and the prevalence of TM summarized in total and for specific<br />

etiological subgroups of azoospermia. TM was classified as extensive<br />

TM, universally distributed (uTM) or collected in plaques (pTM), and finally<br />

borderline TM (bTM) with about 5 (3 to 7) TM elements in one or both testicles.<br />

Representative testicular biopsies were taken in 300 men using a<br />

TruCut needle, Ch.14 (Angiotech, USA). Frequencies of Carcinoma In<br />

Situ (CIS) testis/malignancy in men with different categories of TM were<br />

compared to men without TM using Fisher’s exact test.<br />

RESULTS: UTM was found in 11 men (2.9%). In four of these (36%), the<br />

pattern was found bilaterally. In 3 cases (0.8%) including 2 with Klinefelter’s<br />

syndrome (KS), bTM was detected in the contralateral testis. PTM was detected<br />

in 8 cases (2.1%), and except for one KS man, the condition was in<br />

all cases unilaterally. BTM was detected in 32 cases (8.4%), bilaterally in<br />

<strong>17</strong> (53%). Unilateral bTM was found in further 3 cases with uTM in the other<br />

testis, and these cases were categorized as having uTM. A frequent overlapping<br />

between the most relevant etiological groups was found, e.g. 10 (23%)<br />

of the 44 men with KS and 4 (20%) of the 20 men with Y microdeletions had<br />

a history of cryptorchidism. Of the men with KS, 5 had extensive TM, 3 (7%)<br />

uTM and 2 (5%) pTM, while 14 (32%) showed bTM. Of 101 men with a history<br />

of cryptorchidism, 3 (3%) had uTM, 4 (4%) pTM, and 13 (13%) bTM.<br />

Compared to a frequency of CIS testis/testicular malignancy of 1 of 266<br />

found in men without TM, the frequencies of CIS testis/testicular malignancy<br />

in men with uTM was 1 of 6 (p¼ 0.04), in men with pTM 1 of 7 (p¼0.05), and<br />

in men with bTM 1 of <strong>21</strong> (p¼0.14).<br />

CONCLUSIONS: A total prevalence of TM, extensive and borderline, was<br />

found in 13.4% of 382 unselected, nonvasectomized, azoospermic men. The<br />

relatively high frequency of bilateral uTM and bTM supports a hypothesis of<br />

a general defective spermatogenesis as background of TM. The risk of testicular<br />

malignancy was increased in men with extensive TM.<br />

O-203 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:15 PM<br />

A CASE FOR BANKING SPERM DUE TO INCREASED AGE:<br />

DECLINE IN SPERM COUNT AND MOTILITY IN YOUNG ADULT<br />

MEN. G. M. Centola, a A. Blanchard, b J. L. Demick, c S. Li, d<br />

M. L. Eisenberg. d a New England Cryogenic Center, Newton, MA; b New England<br />

Cryogenic Center, Brookline, MA; c Laboratory, New England Cryogenic<br />

Center, Brookline, MA; d Urology, Stanford University School of<br />

Medicine, Stanford, CA.<br />

OBJECTIVE: Controversy exists as to the stability of semen quality over<br />

the past half century with numerous papers reporting a decrease, increase or<br />

stable parameters in heterogeneous populations. Varied methodology and geography<br />

of studies make definitive conclusions challenging. The purpose of<br />

the current study was to examine semen parameters of young adult men who<br />

provided multiple semen specimens over a 10 year period at the same facility.<br />

DESIGN: Retrospective analysis of semen data for the years 2003 - 2013.<br />

MATERIALS AND METHODS: Semen parameters, including volume,<br />

count, and motility prior to and after cryopreservation were analyzed for a<br />

total of 9425 semen specimens from 489 individuals. Demographic information<br />

was obtained from an extensive social and medical history questionnaire<br />

for all participants. Following 2-3 days of abstinence, the specimens were<br />

collected by masturbation at the lab facility, and were assessed by the<br />

same two laboratory technicians, using a standard Makler semen analysis.<br />

Specimens were frozen with a glycerol-based cryopreservative in liquid nitrogen.<br />

A vial was thawed after 48 hours and count and motility determined<br />

post-thaw. The data was analyzed using generalized linear regression after<br />

adjustment for age, days of abstinence and for repeated semen samples, as<br />

well as by the Cochran-Armitage trend test. All p values were two sided<br />

with p < 0.05 considered statistically significant.<br />

e78 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


RESULTS: There was a significant decline in initial sperm count (p <<br />

0.001), sperm motility (p < 0.001), total count (p < 0.001) and total motile<br />

count (p < 0.001) during the period 2003-2013. There was no significant<br />

change in semen volume (p ¼ 0.2). Following cryopreservation, the mean<br />

post-thaw motility significantly (p < 0.001) decreased with time. There<br />

was a significant decline in age (p trend ¼ 0.003) and alcohol use (p trend<br />

¼ 0.005), as well as an increase in college GPA (p trend ¼ 0.02). However,<br />

BMI (p trend ¼ 0.73) educational attainment (p trend ¼ 0.2), race/ethnicity<br />

(p trend ¼ 0.53), and lifestyle habits (weekly exercise, p trend ¼ 0.<strong>21</strong>;<br />

smokers p trend ¼ 0.99; marital status, p trend ¼ 0.85) remained constant<br />

over the period of study.<br />

CONCLUSIONS: There has been a decline in semen quality among young<br />

men presenting for sperm donation during the past 11 years. Uniform technicians<br />

and techniques over the study period make measurement bias unlikely.<br />

Subject characteristics also remained constant during the study<br />

period. The decrease in sperm count and motility following cryopreservation<br />

suggests that the sperm were not as vital since cryosurvival appears to be<br />

compromised over the ten-year poor of time. This report demonstrates a<br />

decline in sperm count in a young adult male population. Although absolute<br />

values did not decline into the abnormal or sub fertile range, it may be prudent<br />

for men to consider sperm cryopreservation if fertility is being delayed<br />

until later years/advanced age.<br />

O-204 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:30 PM<br />

MALE PARTNER SEXUAL ABSTINENCE LESS THAN 2 OR<br />

GREATER THAN 7 DAYS IS NOTASSOCIATED WITH INCREASED<br />

ANEUPLOIDY RATE. J. Gingold, a,b M. C. Whitehouse, c M. Gerson, d<br />

S. Parsons, d J. A. Lee, d A. B. Copperman, d,a N. Bar-Chama. d,e a Obstetrics,<br />

Gynecology and Reproductive Science, Icahn School of Medicine at Mount<br />

Sinai, New York, NY; b Obstetrics/Gynecology and Women’s Health Institute,<br />

Cleveland Clinic Foundation, Cleveland, OH; c Reproductive Medicine<br />

Associates of New York, New York City, NY; d Reproductive Medicine Associates<br />

of New York, New York, NY; e Urology, Icahn School of Medicine at<br />

Mount Sinai, New York, NY.<br />

OBJECTIVE: Males are routinely counseled to maintain 2-7 days sexual<br />

abstinence prior to processing and use during an in vitro fertilization (IVF)<br />

cycle(s) in order to enhance semen quality. Although obtaining an adequate<br />

concentration of motile sperm is an important objective, an adverse deviation<br />

rarely limits a couple from moving forward in their cycle. The ultimate goal<br />

is to maximize the number of euploid embryos prior to transfer selection.<br />

This study evaluates if duration of male sexual abstinence is associated<br />

with an increase in embryonic aneuploidy.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Couples who underwent an IVF cycle<br />

and utilized pre-implantation genetic screening (PGS) from June 2010 -<br />

March <strong>2015</strong> were included. Oocyte age was recorded (A: %35; B: (35-<br />

38]; C: (38-41]; D: (41-43]; and E: >43). Duration of male sexual abstinence<br />

was recorded. Aneuploidy rate for each group was computed, with 95% confidence<br />

intervals calculated by Clopper-Pearson method. Aneuploidy rate<br />

was modeled by logistic regression using oocyte age and days of abstinence.<br />

The youngest oocyte age group (


Supported by: European Union Seventh Framework Programme DEER<br />

grant <strong>21</strong>2844, and NIH grants P30 DK046200 and T32 DK007703-16.<br />

O-206 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:30 AM<br />

WITHDRAWN<br />

OBJECTIVE: To develop an anatomically accurate fluid dynamics model<br />

in order to assess the anatomic causes of elevated testicular vein pressures.<br />

DESIGN: Finite element analysis was used to model the fluid dynamics of<br />

human testicular venous drainage.<br />

MATERIALS AND METHODS: A two dimensional model was created to<br />

represent the inferior vena cava, gonadal and renal veins. The geometry was<br />

constructed to correspond to a representative cross-section of a male patient’s<br />

CT scan that was obtained for non-urologic causes. The geometry was subsequently<br />

meshed into contiguous triangular elements. Finite element analysis<br />

was employed to solve for the stationary Navier-Stokes equations<br />

using Comsol Multiphysics 3.3 (Comsol, Inc., Burlington, MA). Boundary<br />

conditions, such as renal vein and inferior vena cava flow rates and pressures,<br />

were set to known physiologic values reported in the literature. The dynamic<br />

viscosity of blood was set to 0.01 Pa-sec. Gravity was simulated to correspond<br />

to an upright posture of the patient. The final solution was subsampled<br />

to provide for flow rates and pressures along both testicular veins.<br />

RESULTS: The model accurately predicted physiologic flow rates within<br />

both the renal and testicular veins. Without left renal vein impingement by<br />

the superior mesenteric artery (SMA), the left and right venous pressures<br />

at the inguinal canals were predicted to be 41 and 40 cmH2O respectively<br />

- approximately half the pressure that has previously been reported. Turbulence<br />

due to the perpendicular insertion of the left gonadal vein into the<br />

left renal vein contributed minimally to elevated left sided pressures. Simulated<br />

SMA impingement was modeled with up to 75% occlusion of the left<br />

renal vein. Left gonadal vein pressures rose to only 47 cmH2O despite the<br />

significant change in renal vein diameter.<br />

CONCLUSIONS: Gonadal vein pressures can be accurately modeled using<br />

finite element analysis. The perpendicular insertion of the left gonadal vein is a<br />

negligible source of flow resistance to testicular drainage. Marked compression<br />

(beyond a 75% decrease in diameter) of the left renal vein may be required to<br />

manifest a clinically meaningful pressure elevation (e.g., Nutcracker syndrome)<br />

in the distal left testicular vein. The principle contributor to elevated<br />

left testicular vein pressures, and possibly to the greater prevalence of left sided<br />

varicoceles, is the added flow resistance of the longer left gonadal vessel.<br />

O-208 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:00 PM<br />

MELANOMA ANTIGEN PROTEIN MAGEC1 MUTATION<br />

IDENTIFIED IN FAMILIAL NON-OBSTRUCTIVE<br />

AZOOSPERMIA. A. W. Pastuszak, a c. cengiz, b M. Bekheirnia, c<br />

D. J. Lamb. d a Scott Department of Urology, Baylor College of Medicine,<br />

Houston, TX; b Technician, Houston, TX; c Department of Genetics, Baylor<br />

College of Medicine, Houston, TX; d Baylor College of Medicine, Houston,<br />

TX.<br />

O-207 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:45 AM<br />

HYDRODYNAMICS OF THE HUMAN GONADAL VEIN: A FINITE<br />

ELEMENT MODELTO EXPLORE ELEVATED LEFT TESTICULAR<br />

VENOUS PRESSURES RELATED TO VARICOCELE. R. P. Hayden,<br />

C. Tanrikut. Massachusetts General Hospital, Boston, MA.<br />

OBJECTIVE: Many currently unrecognized genetic etiologies of male<br />

infertility likely exist, and we sought to identify genes influencing male infertility<br />

using next-generation sequencing approaches. The melanoma antigen<br />

(MAGE) family of genes may function in germ cell maintenance and fertility,<br />

and herein we present genomic data associating the melanoma antigen gene<br />

MAGEC1 with male fertility.<br />

DESIGN: Genomic study comparing infertile and fertile men.<br />

MATERIALS AND METHODS: Whole-exome sequencing of genomic<br />

DNA from 2 Saudi Arabian brothers with non-obstructive azoospermia<br />

(NOA) and Sertoli cell only (SCO) histology was performed. Sequence data<br />

were filtered against several genomic databases and deleterious SNPs common<br />

to both brothers identified. Sanger sequencing confirmed SNPs of interest and<br />

assessed their presence in other NOA men. Array comparative genomic hybridization<br />

(aCGH) analysis of 22 NOA men with normal karyotypes and Y-<br />

chromosome microdeletion assays assessed for gene dosage changes due to<br />

copy number variants (CNVs), and the presence of additional CNVs in genes<br />

of interest was evaluated using qPCR. Immunohistochemical staining of testis<br />

sections used commercially available antibodies and standard protocols.<br />

RESULTS: An A>T SNP at position 3530 in exon 4 of the MAGEC1 gene<br />

resulting in a potentially deleterious R1082W mutation was identified in a set<br />

of 6 SNPs mapping to 6 unique genes selected from over 300,000 SNPs shared<br />

between the two brothers’ genomes. Sanger sequencing confirmed the SNP in<br />

the two brothers but did not identify it in <strong>17</strong>1 additional NOA and 34 fertile<br />

control men. aCGH of 22 NOA men did not identify CNVs in MAGEC1,<br />

although qPCR based analysis of 57 additional NOA men identified microduplications<br />

in 2/57 (3.5%). In contrast, the frequency of MAGEC1 CNVs in the<br />

general population (including fertile and infertile individuals) is approximately<br />

0.2%. Immunohistochemical localization of MAGEC1 in human testis<br />

from a male with normal spermatogenesis revealed cytoplasmic staining of<br />

spermatocytes, which was less prominent in testis from a male with hypospermatogenesis,<br />

and was absent in the testis from an SCO male.<br />

CONCLUSIONS: A potentially damaging SNP and microduplications in<br />

MAGEC1 were identified in a cohort of infertile NOA men. Future work<br />

elucidating the molecular role of MAGEC1 in male fertility using cell-based<br />

assays and animal models will further our understanding of the MAGE genes<br />

in male germ cell maintenance and fertility.<br />

Supported by: AWP is a National Institutes of Health (NIH) K12 Scholar<br />

Supported by a Male Reproductive Health Research Career (MRHR) Development<br />

Physician-Scientist Award (HD0739<strong>17</strong>-01) from the Eunice Kennedy<br />

Shriver National Institute of Child Health and Human Development<br />

(NICHD) Program (to DJL). MRB is a National Institutes of Health (NIH)<br />

K12 Scholar Supported by a Multidisciplinary K12 Urologic Research<br />

(KURe) Career Development Physician-Scientist Award (K12<br />

DK0083014) (to DJL).<br />

e80 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


O-209 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:15 PM<br />

DIETARY COENZYME Q10 INTAKE AND SEMEN PARAMETERS<br />

IN A SUBFERTILE POPULATION. B. C. Tiseo, a A. J. Gaskins, b<br />

J. E. Chavarro, b R. Hauser, b C. Tanrikut. c a Division of Urology, Hospital<br />

das Clinicas, University of Sao Paulo Medical School, S~ao Paulo, Brazil;<br />

b Harvard T.H. Chan School of Public Health, Boston, MA; c Massachusetts<br />

General Hospital, Boston, MA.<br />

OBJECTIVE: To investigate whether dietary coenzyme Q10 (CoQ10)<br />

intake is associated with semen parameters in a cohort of men from a fertility<br />

clinic.<br />

DESIGN: Prospective cohort study.<br />

MATERIALS AND METHODS: Dietary intake data were obtained via<br />

validated food frequency questionnaires from 155 male partners in subfertile<br />

couples. CoQ10 intake was derived by summing specific contributions across<br />

all food items. A total of 338 semen samples were evaluated. Multivariable<br />

linear mixed models were used to examine the relation between quartiles<br />

of CoQ10 intake and semen quality parameters while adjusting for age,<br />

race, smoking, body mass index, abstinence time and male factor diagnosis<br />

as well as for caloric intake, caffeine, alcohol and total fat intakes and accounting<br />

for within-person correlations.<br />

RESULTS: Mean dietary CoQ10 intake was 7.1mg/day (range¼1.5 to<br />

20.2mg/day). No subjects were taking supplemental CoQ10. The adjusted<br />

mean sperm motility percentage was calculated as 43.1% for the lowest quartile<br />

of CoQ10 intake and 39.3% for the highest quartile(p-trend: 0.489). Evaluating<br />

specifically progressive sperm motility the adjusted means were<br />

24.2%, 25.7%, 25.5% and <strong>21</strong>.3% for increasing quartiles of CoQ10 intake<br />

(p-trend: 0.365). There were no associations between dietary CoQ10 intake<br />

and sperm concentration, total motility or morphology (See Table 1).<br />

CONCLUSIONS: Our findings indicate that dietary CoQ10 intake is not<br />

related to semen quality parameters. The mean dietary intake of CoQ10<br />

was 25-fold lower than the supplemental 200-300 mg daily dose used in<br />

controlled trials which have demonstrated improved sperm motility. The<br />

relative low dietary intake of CoQ10 in our cohort may explain the discrepancy<br />

between our results and the findings of the supplemental trials. Dietary<br />

CoQ10 intake alone may not be sufficient to optimize semen parameters.<br />

Seminal parameters marginal means by CoQ10 intake groups.<br />

Quartile 1<br />

(Lowest)<br />

N ¼ 38<br />

Quartile 2<br />

N ¼ 39<br />

Quartile 3<br />

N ¼ 39<br />

Quartile 4<br />

(Highest)<br />

N ¼ 39<br />

CoQ10 intake (mg/day) 1.45-4.63 4.80-6.36 6.56-8.83 8.99-20.15<br />

Semen Samples (N) 83 85 86 84<br />

Mean Mean Mean Mean p-trend<br />

Abstinence time (hrs) 149.6 133.4 133.8 126.9 0.989<br />

Sperm concentration (x10⁶/mL) 28.2 36.8 34.4 31.2 0.518<br />

Sperm motility (%) 43.1 46.0 45.8 39.3 0.365<br />

Progressive sperm Motility (%) 24.2 25.7 25.5 <strong>21</strong>.3 0.489<br />

Normal sperm morphology (%) 6.4 6.7 7.0 5.5 0.129<br />

Note: N: number of subjects<br />

Results shown as marginal means of the population adjusted for age, race, smoking,<br />

body mass index, abstinence time, male factor diagnosis, calorie intake,<br />

caffeine intake, alcohol intake and total fat intake using linear mixed models.<br />

Supported by: National Institutes of Health grants ES009718, ES022955,<br />

ES000002, P30 DK046200 and T32 DK007703-16.<br />

O-<strong>21</strong>0 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:30 PM<br />

SPERM QUALITY AND FERTILITY ARE COMPROMISED IN<br />

BRCA1 MUTANT MALE MICE. R. Stobezki, a S. Titus, a B. Musul, a<br />

F. Moy, a K. H. Oktay. a,b a Obstetrics & Gynecology, New York Medical College,<br />

Valhalla, NY; b Innovation Institute for Fertility Preservation and IVF,<br />

New York, NY.<br />

OBJECTIVE: BRCA1 and 2 are DNA double strand break (DSB) repair<br />

genes and their mutations result in cancer predisposition. Presence of<br />

BRCA mutations have been reported to affect ovarian reserve. We have previously<br />

shown that BRCA1 gene mutations may also affect sperm DNA<br />

integrity. In this study, we sought to determine if BRCA1 mutations affected<br />

fertility and embryo quality in a transgenic mouse model.<br />

DESIGN: Experimental study.<br />

MATERIALS AND METHODS: We inbred a transgenic BRCA1-mutant<br />

mouse colony, which carries a deletion of 330 base pairs (bp) in intron 10 plus<br />

407 bp in exon 11 of the BRCA1 gene. This deletion causes the gene product<br />

to be retained in the cytoplasm, which severely compromises its nuclear functions.<br />

BRCA1 mutant mice were studied at three months of age. Sperm were isolated<br />

from the vas deferens to observe sperm concentration. Litter size data were<br />

determined from mating BRCA1-mutant and wild type (WT) males with WT female.<br />

The same crosses were performed to determine embryos reaching the blastocyst<br />

stage. Prior to the setup of these breeding crosses, WT females received<br />

PMSG (pregnant mare’s serum gonadotropin, 5 IU) followed by HCG (human<br />

chorionic gonadotropin, 5 IU) 72 hrs later. The morning after the setup, vaginal<br />

plugs were checked and those females who have plugs were separated. On day 5,<br />

embryos were flushed from the uterus and observed for morphology.<br />

RESULTS: BRCA1-mutant male mice had lower sperm counts than the<br />

WT (1.8x10 6 0.1x10 6 vs. 2.1 x10 6 0.07, n¼8; p¼0.045). Upon mating<br />

with WT females, BRCA1-mutant males resulted in a smaller litter size<br />

than when the WT males which were mated with the same (3.22.3 vs.<br />

6.91.3, n¼9; p¼0.0008). Embryos produced by crossing BRCA1 mutant<br />

male mice with WT females had a higher percentage of arrested embryos<br />

(32.14%0.128 vs. 8.33%0.066, n¼7; p¼0.026) as compared to the WT.<br />

CONCLUSIONS: These results indicate that BRCA1 mutations may be associated<br />

with decline in sperm quantity and quality in mice, which results in compromised<br />

embryo development and implantation. Our findings also underscore the<br />

importance of DNA DSB repair in maintenance of sperm quality. Further molecular<br />

studies are needed to understand the mechanism of arrest in embryos resulting<br />

from BRCA1-mutant male mice. Translational studies are warranted in men with<br />

BRCA mutations to determine the clinical significance of these findings.<br />

Supported by: NIH R01HD053112.<br />

HEALTH DISPARITIES 2<br />

O-<strong>21</strong>1 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:15 AM<br />

THE ONTOGENY OF MYOMETRIAL STEM CELLS IN OCT4-GFP<br />

TRANSGENIC MICE. S. Brakta, A. Mas, M. P. Diamond,<br />

S. M. Shalaby, A. Al-Hendy. GRU, Augusta, GA.<br />

OBJECTIVE: Somatic stem cells (SSC) are master stem cells that through<br />

asymmetric division retain their ability to self-renew while producing<br />

daughter cells. Similarly, tumor initiating cells are a subset of cells within<br />

a tumor cell population, which retain the ability to sustain tumors through<br />

asymmetric division. It is thus believed that tumors such as uterine fibroids<br />

originate from a single transformed SSC of the myometrium followed by<br />

expansion and propagation in a steroid-dependent manner. Our group has<br />

recently identified two specific cell surface markers (CD44 & Stro1) as human/rat<br />

myometrial stem cell markers. In this work, we aim to identify myometrial<br />

stem cells from mice at different age points by using stem cell<br />

markers Oct4 and CD44.<br />

DESIGN: Laboratory studies using transgenic mice.<br />

MATERIALS AND METHODS: Twenty four female mice, B6, CBA-Tg<br />

(Pou-5fl-EGFP) 2Mnn/j mouse strains, 1 week to 24 weeks old were purchased<br />

from Jackson laboratory. These mice were homozygous for the<br />

Pou5fl/OCT4 transgenic insert and expressed enhanced green fluorescent<br />

protein (EGFP) in the uterus under the control of POU domain, class 5, transcription<br />

factor 1 (Oct4) promoter and distal enhancer. Uterine blocks were<br />

shipped at ages 1, 3, 4, 8, 12 and 24 weeks. Immunohistochemistry (IHC) and<br />

immunofluorescence studies were performed in order to co-localize the specific<br />

cell surface marker CD44 with the well-known stem cell marker Oct4.<br />

Image J was used to quantify the stem cells as a percentage of total myometrial<br />

cells. Three random high power fields were selected from each uterine<br />

age and the average of stem cell percentage was determined. To correlate<br />

these numbers with possible response to the effects of ovarian sex steroid hormones,<br />

we also evaluated the uterine expression of estrogen receptor a and<br />

progesterone receptors A&B at ages 1, 3 and 4 using inverted microscopy.<br />

RESULTS: The frequency of myometrial stem cells was significantly<br />

lower (2.0%1.2) at 1 week of age compared to sexually more mature<br />

mice (P < 0.05). The frequency of myometrial stem cells at ages 3, 4, 8,<br />

12, and 24 weeks was 13.0%5.4, 11.3%1.5, 10.4%4.5, 6.6%0.6,<br />

and 10.4%3.7 respectively. Interestingly, abundant positive staining for estrogen<br />

receptor a and progesterone receptors A&B were detected in the myometrium<br />

ages 1, 3, and 4 weeks.<br />

FERTILITY & STERILITY Ò<br />

e81


CONCLUSIONS: Myometrial stem cells are present in the myometrium at a<br />

significantly lower percentage at the pre-sexual age of 1 week than at the sexually<br />

mature ages of 3 to 24 weeks. This is likely due to lack of the estrogen hormone<br />

ligand at that early age rather than the availability of the steroid receptors<br />

which are similarly expressed in neonatal and adult myometrium. These results<br />

suggest that stem cells are steroid dependent and increase in number with reproductive<br />

maturity at 3-4 weeks of age in mice. Our findings emphasize the<br />

vulnerability of neonatal myometrium to environmental xeno-estrogen exposure<br />

which can potentially lead to permanent reprogramming of myometrial<br />

stem cells leading to adult onset of diseases such as uterine fibroids.<br />

O-<strong>21</strong>2 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:30 AM<br />

FREE, BIOAVAILABLE, AND TOTAL 25 HYDROXY VITAMIN D IN<br />

VIABLE PREGNANCIES: COMPARISONS ACROSS RACE AND<br />

OVER TIME. S. Senapati, a T. Adedji-Fajobi, b M. D. Sammel, c<br />

C. Coutifaris, a M. S. Bartolomei, a S. Butts. c a University of Pennsylvania,<br />

Philadelphia, PA; b Obstetrics and Gynecology, Perelman School of Medicine,<br />

Philadelphia, PA; c Perelman School of Medicine, University of Pennsylvania,<br />

Philadelphia, PA.<br />

OBJECTIVE: Mounting evidence supports an association between<br />

vitamin D deficiency and reproductive outcomes, however, the optimal<br />

method for characterizing maternal vitamin D status, especially in African<br />

Americans, is controversial. Our objective was to characterize differences<br />

in total, free, and bioavailable vitamin D at peri-conception (PC) and midgestation<br />

(MG) in women with viable pregnancies by race.<br />

DESIGN: Prospective cohort of women attempting to conceive naturally<br />

or following IVF treatment.<br />

MATERIALS AND METHODS: Total 25 (OH)D was measured by liquid<br />

chromatography-tandem mass spectrometry at two timeponts [1] within one<br />

month of conception (PC) and [2] at 18-22 weeks gestational age (MG). Free<br />

(non-protein bound) and bioavailable (free plus albumin-bound fraction)<br />

25(OH)D was calculated with total 25(OH)D, vitamin D binding protein<br />

(VDBP), albumin and relevant binding constants by established methods.<br />

Subjects were classified by Vitamin D status (VitD) as sufficient (¼> 75<br />

nmol/L), insufficient (¼50nmol/L) or deficient (< 50 nmol/L). Kruskal<br />

wallis, Wilcoxon Rank Sum, c2, and Fisher exact testing was used to evaluate<br />

differences in demographic characteristics by VitD status, and the association<br />

between race and change in markers of vitamin D during pregnancy.<br />

RESULTS: Of the 110 women studied, 74.5% were White, 15.5% were African<br />

American, 8.2% were Asian, and 1.8% were other/mixed race. 33.6%<br />

were VitD sufficient, 45.5% were VitD insufficient and <strong>21</strong>% were VitD deficient<br />

peri-conception. PC VitD deficiency was associated with non-White<br />

race (p¼0.01) and median PC total 25(OH)D was lower in African American<br />

(55.4 nmol/L) than White subjects (70.6 nmol/L) (p


95% CI 1.05-15.2, p


(evaluated by HSCORE) and real-time RT-PCR (evaluated for relative<br />

expression using the delta-delta Ct method). An optimal HSCORE threshold<br />

of 1.4 was previously established by ROC analysis. Differences in proportions<br />

of positive staining were analyzed by Chi-square, while RNA data<br />

were evaluated by ANOVA and T-test.<br />

RESULTS: Age and BMI were similar between the three groups.<br />

HSCORE was significantly different (p < 0.0001) (Table 1). Endometrial<br />

BCL6 immunostaining was positive (above the 1.4 HSCORE threshold) in<br />

2/28 (7%) fertile control subjects as compared to 103/116 (88.7%) UI subjects<br />

and 32/43 (74.4%) uRPL subjects. Mean and median expression of<br />

BCL6 RNA was lowest in the fertile controls, intermediate in the uRPL<br />

and highest in UI, with ANOVA analysis approaching significance<br />

(p¼0.07). In simple pairwise comparison, mRNA levels in UI were 2-fold<br />

of those in controls (p¼0.04). Subsequent surgical exploration of a subset<br />

of subjects demonstrated endometriosis or hydrosalpinges in 98% and 89%<br />

in those with positive BCL6 immmunostaining and UI or uRPL, respectively.<br />

CONCLUSIONS: Endometrial immunostaining for BCL6 closely correlated<br />

with the fertility status of those tested, with fertile women demonstrating<br />

low levels of nuclear BCL6 and the majority of women with UI<br />

and uRPL expressing high levels. Endometrial BCL6 mRNA expression<br />

showed similar trends, but differences between groups were smaller. Additionally,<br />

high BCL6 immunostaining strongly predicted the presence of<br />

endometriosis or, occasionally, hydrosalpinges. Early pregnancy loss and<br />

Infertility appear to share a common pathophysiological feature, but may<br />

differ by severity of the defect.<br />

Demographics<br />

Characteristic<br />

Normal Fertile<br />

Controls<br />

(n¼ 28)<br />

Unexplained<br />

Infertility<br />

(n ¼ 116)<br />

Unexplained<br />

Recurrent<br />

Pregnancy<br />

Loss (n ¼ 43)<br />

AGE<br />

32.78 2.6 33.04 4.2 34.65 4.7<br />

(mean STD)<br />

BMI 25.6 4.7 25.23 5.5 25.23 6.6<br />

HSCORE 0.43 0.69 2.8 1.1 2.4 1.4<br />

Supported by: NIH R01 HD0677<strong>21</strong> (S.L.Y. and B.A.L.).<br />

O-<strong>21</strong>8 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:30 AM<br />

SIZE MATTERS: SINGLE NUCLEOTIDE POLYMORPHISM (SNP)<br />

BASED CHROMOSOME ANALYSIS OF PRODUCTS OF CONCEP-<br />

TION (POC) SAMPLES IDENTIFIES CLINICALLY SIGNIFICANT<br />

DELETIONS/DUPLICATIONS BELOW KARYOTYPE<br />

RESOLUTION. M. K. Maisenbacher, a K. Merrion, b S. Sigurjonsson, a<br />

K. G. Paik, b M. J. Young, a B. Pettersen. c a Natera, San Carlos, CA; b Natera,<br />

Inc., San Carlos, CA; c Genetic Counseling, Bend, OR.<br />

OBJECTIVE: Report on products of conception (POC) samples with deletions<br />

and duplications below the resolution of traditional karyotyping (<<br />

10 Mb). These segmental abnormalities are clinically relevant based on<br />

size, location and/or correlation with a known genetic syndrome.<br />

DESIGN: Retrospective analysis.<br />

MATERIALS AND METHODS: Review of <strong>17</strong>,424 consecutive fresh<br />

POC samples sent to a reference lab along with maternal blood samples. Genotyping<br />

was performed using Illumina CytoSNP-12b microarray and bioinformatics.<br />

RESULTS: 2600 cases (14.9%) had maternal cell contamination, <strong>17</strong> cases<br />

(0.01%) had incomplete results, and 14,807 cases (85%) had fetal results. Of<br />

these, 100 cases (0.7%) had deletions/duplications less than 10 Mb without<br />

aneuploidy or uniparental disomy (UPD) of other chromosomes identified:<br />

22 cases (22%) had a single isolated deletion or duplication (11 deletions<br />

and 11 duplications), 43 cases (43%) had an additional deleted or duplicated<br />

segment that was either >10 Mb or


O-<strong>21</strong>9 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:45 AM<br />

LIFESTYLE AND PREGNANCY LOSS AMONG WOMEN USING<br />

HOME PREGNANCY TEST KITS, LIFE STUDY. G. M. Buck Louis, a<br />

K. J. Sapra, a E. Schisterman, b C. D. Lynch, c J. Maisog, a K. L. Grantz, a<br />

R. Sundaram. b a Division of Intramural Population Health Research, NICHD,<br />

Rockville, MD; b Eunice Kennedy Shriver National Institute of Child, Rockville,<br />

MD; c Department of Reproductive Medicine, Ohio State University<br />

Medical Center, Columbus, OH.<br />

OBJECTIVE: Recent cohorts of reproductive aged women now detect<br />

their pregnancies early using sensitive home pregnancy tests, replacing clinical<br />

diagnosis as the gold standard. This prompted us to assess whether the<br />

incidence or risk factors for pregnancy loss have changed relative to guidance<br />

based upon clinically recognized pregnancies.<br />

DESIGN: Prospective cohort with preconception enrollment.<br />

MATERIALS AND METHODS: 501 couples were recruited upon discontinuing<br />

contraception and followed for a year of trying and through pregnancy.<br />

Couples were interviewed upon enrollment and completed daily journals about<br />

usage of vitamins, cigarettes, and caffeinated and alcoholic beverages until 7<br />

post-conception weeks, then monthly until delivery. Women used ClearblueÒ<br />

digital pregnancy tests on the day of expected menses or with bleeding. Pregnancy<br />

loss was identified by conversion to a negative test, onset of menses or<br />

clinical confirmation depending upon gestation. Using proportional hazards<br />

regression and accounting for truncation and right censoring, we estimated hazard<br />

ratios and 95% confidence intervals (HR; 95% CI) for partners’ ages and<br />

BMI, females’ history of pregnancy loss and couples’ lifestyle (alcohol,<br />

caffeine, smoking, vitamins) for two time intervals: 1) enrollment through early<br />

pregnancy and 2) only during early pregnancy.<br />

RESULTS: Among 344 women with a singleton pregnancy, 98 (28%)<br />

experienced a pregnancy loss. In adjusted models for enrollment through<br />

early pregnancy, we observed significant associations: female age R35 years<br />

(2.19; 1.28,3.76), female smoking or alcohol usage (4.63; 2.23, 9.60), female<br />

>2 daily cups of caffeinated beverages (1.86; 1.02, 3.42), and female vitamin<br />

usage (0.42; 0.22, 0.80). When restricting time to only during pregnancy, female<br />

age (2.04; 1.<strong>21</strong>,3.44) and smoking or alcohol usage (5.00; 2.57, 9.73)<br />

remained significant.<br />

CONCLUSIONS: Lifestyle during sensitive windows was associated with<br />

pregnancy loss corroborating earlier results, although with higher risks. Of<br />

note was the reduced risk associated with vitamin use, a finding not previously<br />

reported. These findings support encouraging preconception guidance<br />

for reproductive aged couples.<br />

Supported by: Intramural Research Program of the Eunice Kennedy<br />

Shriver National Institute of Child Health and Human Development.<br />

O-220 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:00 PM<br />

IS THYROID AUTOIMMUNITYASSOCIATED WITH PREGNANCY<br />

LOSS?. T. Plowden, a E. Schisterman, a S. Zarek, a R. M. Silver, b<br />

N. Galai, c A. DeCherney, a S. L. Mumford. a a NICHD, NIH, Bethesda,<br />

MD; b Unversiy of Utah, Salt Lake City, UT; c Haifa University, Haifa, Israel.<br />

OBJECTIVE: Overt thyroid dysfunction has been associated with infertility,<br />

early pregnancy loss and other adverse obstetrical outcomes. However,<br />

results of studies assessing the relationship between thyroid antibodies and<br />

pregnancy loss have varied. Thus, our objective was to examine the association<br />

between pre-pregnancy anti-thyroid antibodies and pregnancy loss.<br />

DESIGN: Prospective cohort study from a multi-center randomized, placebo-controlled<br />

trial of preconception low-dose aspirin to prevent pregnancy<br />

loss among healthy fertile women (n¼1228). Women with abnormal fT4 at<br />

baseline were excluded.<br />

MATERIALS AND METHODS: At baseline, TSH, fT4, anti-thyroglubulin<br />

antibody (anti-TG) and anti-thyroid peroxidase antibody (anti-TPO)<br />

levels were measured. Relative risk (RR) and 95% confidence intervals<br />

(CIs) for pregnancy loss were estimated using generalized linear models adjusting<br />

for age and body mass index.<br />

RESULTS: 155 women in our study had thyroid autoimmunity (+anti-TG<br />

antibodies and/or +anti-TPO antibodies) compared to 902 without thyroid<br />

immunity. During the study period, 114 women with thyroid autoimmunity<br />

became pregnant while 639 women without thyroid immunity conceived. After<br />

adjusting for age and BMI, women with thyroid autoimmunity who<br />

became pregnant during the study had similar rates of clinical pregnancy<br />

loss (RR 0.80; 95% CI 0.50, 1.27) and live birth (RR 1.05; 95% CI 0.95,<br />

1.16) compared to those without thyroid autoantibodies.<br />

CONCLUSIONS: Among healthy fertile women with a history of 1 or 2<br />

losses, presence of anti-TG or anti-TPO antibodies was not associated with<br />

pregnancy loss.<br />

Clinical Pregnancy Loss and Live Birth Rate in Women With and Without<br />

Thyroid Autoimmunity.<br />

Absence of<br />

Thyroid Antibodies<br />

Presence of<br />

Thyroid Antibodies<br />

N(%) 649 (85.1) 114 (14.9)<br />

Clinical Pregnancy Loss 1.00 (ref) 0.80 (0.50,1.27)<br />

Live Birth Rate 1.00 (ref) 1.05 (0.95,1.16)<br />

Supported by: Intramural Research Program, DIPHR, PRAE, NICHD,<br />

NIH<br />

O-2<strong>21</strong> Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:15 PM<br />

GALECTIN 14 PROTECTS HUMAN TROPHOBLAST CELLS<br />

FROM OXIDATIVE STRESS AND PROMOTES EXTRAVILLOUS<br />

DIFFERENTIATION. R. Fritz, a B. Kilburn, a H. Kohan-Ghadr, a<br />

S. Drewlo, a D. Armant. a,b a Wayne State University, Detroit, MI; b National<br />

Institute of Child Health & Human Development, Bethesda, MD.<br />

OBJECTIVE: In the first trimester, invasion of extravillous trophoblast<br />

(EVT) cells into maternal spiral arteries with subsequent endovascular<br />

plug formation is necessary for healthy placentation by maintaining low oxygen<br />

concentrations at the implantation site (1). Failure to form an EVT plug<br />

within the spiral arteries is strongly linked to early pregnancy loss (EPL) (2).<br />

Hypoxia/re-oxygenation injury (H/R) in the first trimester placenta could<br />

contribute to adverse pregnancy outcomes, including EPL. The aim of this<br />

study was to evaluate the effects of galectin 14, a rarely studied placental specific<br />

galectin, on EVT cell survival, invasion, and proliferation using a human<br />

first trimester trophoblast cell line.<br />

DESIGN: Laboratory prospective study<br />

MATERIALS AND METHODS: The HTR-8/SVneo first trimester human<br />

trophoblast cell line was cultured at 20% oxygen (O2), at 2% O2,<br />

or exposed to H/R injury, induced by culture for 2 hours at 2% O2 followed<br />

by 6 hours at 20% O2. Expression of galectin 14 was quantified by immunocytochemistry<br />

(ICC) and image analysis. To evaluate rescue from<br />

apoptosis, 10 ng/ml recombinant galectin 14 was added during re-oxygenation.<br />

Cell death was quantified using terminal deoxynucleotidyl transferase<br />

dUTP nick end labeling (TUNEL). Effects of galectin 14 on<br />

invasion were measured using a Matrigel-based assay, counting trophoblast<br />

cells in the lower chamber after culture for 72 hours. Integrin switching<br />

from a6b4 toa1b1 is associated with extravillous differentiation to an<br />

invasive phenotype, and was assessed by ICC, evaluating the percentage<br />

of cells positive for a6b4 and a1b1 . Proliferation was analyzed using<br />

ICC for Ki-67. All experiments were conducted in triplicates and alongside<br />

negative controls. Statistical analysis was performed with the non-parametric<br />

Kruskal-Wallis test, using open-source R software (http://www.rproject.org/).<br />

P < 0.05 was considered significant.<br />

RESULTS: Expression of galectin 14 significantly increased after treatment<br />

with 2% O2, and significantly decreased after H/R injury. H/R injury<br />

significantly increased cell death; however, galectin 14 rescued trophoblasts<br />

from cell death. Supplementation with galectin 14 significantly increased<br />

trophoblast invasion through Matrigel and caused a significant switch from<br />

integrin a6b4toa1b1. Galectin 14 had no effect on trophoblast proliferation.<br />

CONCLUSIONS: Galectin 14 contributes to trophoblast survival under<br />

hypoxic conditions, and rescues cell death due to H/R injury. Furthermore,<br />

galectin 14 promotes trophoblast invasion and differentiation to the extravillous<br />

phenotype. Galectin 14 could serve as a therapeutic agent for early pregnancy-related<br />

placental disorders, such as recurrent EPL.<br />

References:<br />

1. Burton, G.J. and E. Jauniaux, Placental oxidative stress: from miscarriage<br />

to preeclampsia. J Soc Gynecol Investig, 2004. 11(6): p. 342-52.<br />

2. Hempstock, J., et al., The contribution of placental oxidative stress to<br />

early pregnancy failure. Hum Pathol, 2003. 34(12): p. 1265-75.<br />

Supported by: Grants from the NIH (HD071408), the March of Dimes, the<br />

W.K. Kellogg Foundation, PerkinElmer Health Sciences, and, in part, by the<br />

Intramural Research Program of the NICHD, NIH.<br />

FERTILITY & STERILITY Ò<br />

e85


O-222 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:30 PM<br />

DISCRIMINANT ANALYSIS FORECASTING MODEL OF FIRST<br />

TRIMESTER PREGNANCY OUTCOME FOLLOWING 5054 INFER-<br />

TILE PATIENTS AFTER IVF-ET. Y. Yi, a X. Li, b Y. Ouyang, c G. Lu. c<br />

a Institute of Reproduction & Stem Cell Engineering, Central South University,<br />

Changsha, China; b Reproductive and Genetic hospital of CITIC-Xiangya,<br />

Chang-sha, China; c Central South University, Changsha, China.<br />

OBJECTIVE: To investigate whether ultrasound indicators of embryo on<br />

27th -29th day after IVF-ET are different between 1st trimester miscarriage<br />

group and ongoing pregnancy group and to build discriminant analysis forecasting<br />

model to predict 1st trimester pregnancy outcome.<br />

DESIGN: Prospective study.<br />

MATERIALS AND METHODS: 5054 single pregnancy female were recruited<br />

after IVF-ETover a period of 18 months. Maternal age, infertility duration<br />

and number of IVF cycles of these patients were collected. Transvaginal<br />

ultrasound was performed for these patients to observe MSD (Mean Sac Diameter<br />

), CRL (Crown-Rump Length), YSD (Yolk Sac Diameter), embryonic cardiac<br />

activity and fluid collection around GS (Gestational Sac) on 27th -29th day<br />

after IVF-ET. Then 1st trimester pregnancy outcome of these patients were followed<br />

up. Characteristics and Ultrasound parameters of these patients were<br />

compared by Two Sample t Test or Chi-Squared Test between 1st trimester<br />

miscarriage group (n¼679) and ongoing pregnancy group (n¼4375). Forecasting<br />

model was built by discriminant analysis to predict 1st trimester pregnancy<br />

outcome. The difference was statistically significant when P < 0.05.<br />

RESULTS: Maternal age (32.575.06 VS 30.364.29 years ), infertility<br />

duration (6.014.13 VS 5.273.50 years) and number of cycle<br />

(1.330.81 VS 1.190.55) were all significant higher in 1st trimester<br />

miscarriage group than those in ongoing pregnancy group (P


DESIGN: Prospective cohort study.<br />

MATERIALS AND METHODS: The Longitudinal Investigation of<br />

Fertility and Environment Study (2006-2010) enrolled 501 couples attempting<br />

pregnancy. Couples were followed until they had a positive human chorionic<br />

gonadotropin pregnancy or for up to 12 months of trying. Residences were geocoded<br />

and the distance to the nearest major roadway was determined in meters<br />

(


DESIGN: Cross sectional observational study.<br />

MATERIALS AND METHODS: Women aged between 20-38 years (300<br />

infertile couples) presented to a university affiliated fertility center were approached<br />

to participate in the study. Only 150 couples that underwent ICSI<br />

for male factor infertility accepted to participate and 94 of them had retrieved<br />

enough FF samples suitable for laboratory testing. The FF sample was obtained,<br />

centrifuged, and stored in liquid nitrogen. The concentrations of<br />

four PCBs 28, 52, 138, 180 were estimated in the obtained FF samples using<br />

Gas Chromatography/ Mass Spectrometry. SPSS statistical analysis program<br />

(version <strong>17</strong>) was used. Multiple regression analysis was used to correlate the<br />

PCBS to ICSI outcomes.<br />

RESULTS: The mean age of study participants was 31.5 6.2 years. Their<br />

body mass index (BMI) was 26.19 6.84, baseline FSH was 7.6 2.65, and<br />

Estradiol 59.3 10.5. There were significant negative correlations between<br />

the concentrations of the 4 PCBS and the endometrial thickness (Adjusted r<br />

0.2, P ¼ .0001). The higher PCB 28 concentration in the FF was associated<br />

with lower number of retrieved oocytes (adjusted r 0.07, P ¼ 0.0001). Moreover,<br />

fertilization and embryo cleavage rates were negatively correlated with<br />

PCB 180 FF concentration (adjusted r 0.07, P ¼ 0.001 and adjusted r 0.1,<br />

P¼0.002, respectively). The number of the implanted embryos was negatively<br />

correlated with PCB 52 FF concentration (adjusted r 0.2, P¼ 0.02).<br />

However the clinical pregnancy rate did not reach the level of significance.<br />

CONCLUSIONS: Higher concentrations of any type PCBs are associated<br />

with thinner endometrium. The higher the level of PCB 28 and 180 the lower<br />

the retrieval, fertilization and embryo cleavage rates. High PCBs concentration<br />

in the FF adversely affect embryological ICSI outcome. However, more<br />

data is needed to evaluate the PCBS effect on the clinical outcome.<br />

FERTILITY PRESERVATION<br />

O-229 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:15 AM<br />

CO-TRANSPLANTATION OF ENDOTHELIAL CELLS WITH XEN-<br />

OTRANSPLANTED HUMAN OVARIAN TISSUE IMPROVES PRI-<br />

MIRDIAL FOLLICLES SURVIVAL. L. Man, a Z. Rosenwaks, b<br />

G. L. Schattman, c D. James. a a CRMI, Weill Cornell Medical College,<br />

NYC, NY; b Weill Cornell Medical College, New York, NY; c Weill Medical<br />

College/Cornell University, New York, NY.<br />

OBJECTIVE: Improved treatments have raised cancer survival rates and<br />

increased the number of patients needing fertility preservation. A major<br />

obstacle to efficient ovarian tissue autotransplantation is graft ischemia.<br />

Endothelial cells (ECs) exhibit the unique capacity to auto-assemble neovascular<br />

networks following in vivo transplantation of single cell suspensions.<br />

Here, we show that co-transplantation of exogenous ECs improve survival<br />

and follicular reserve of xenografted ovarian tissue by accelerating anastomosis<br />

of blood vessels at the interface of host and graft.<br />

DESIGN: Xenograft of human ovarian tissue into NSG mice with cotransplantation<br />

of endothelial cells.<br />

MATERIALS AND METHODS: Tissue-specific mouse ECs, were<br />

labeled with green fluorescent protein and modified by the adenoviral gene<br />

fragment E4-ORF1. Human ovarian tissue was co-transplanted with GFP-<br />

ECs into the gluteus maximus of NSG oopherectomized mice aged 10-12<br />

weeks. Each mouse served as its own control with a contralateral site of<br />

engraftment without ECs. Engrafted tissue was harvested 1-2 weeks after<br />

transplantation. The contribution of exogenous ECs to functional vessels,<br />

amount of surviving follicles as well as the extent of necrosis were analyzed<br />

in histologic sections using light and confocal microscopy.<br />

RESULTS: At 1 and 2 weeks we demonstrated the same proportion of<br />

follicular survival between the ovarian tissue co-transplanted with ECs, in<br />

comparison to ovarian tissue without ECs. At 2 weeks after co-transplantation<br />

we found a three-fold increase in the survival of primordial follicles at<br />

the ECs group in comparison to the control; 73% vs 27%, P


give them the greatest chance of a live-born baby. We set out to answer this<br />

question by calculating the theoretical number of MII oocytes necessary to<br />

result in one baby born (BB).<br />

DESIGN: Retrospective cohort study and theoretical model.<br />

MATERIALS AND METHODS: Data from 160 autologous OC thaw<br />

(AOCT) cycles were reviewed. Cycles did not include medically-indicated<br />

or those with Day-3 transfer. To calculate the no. of oocytes needed to result<br />

in one BB, the following data were assessed: total no. of thawed and survived<br />

MII oocytes, 2-pronuclear (2PN) fertilization, and development of morulae<br />

and blastocysts (BL). BL formation yield (BFY) was calculated as total<br />

no. BL/no. MII thawed, and total embryo rate (TER) was calculated as total<br />

no. morula+BL./no. MII thawed. Rate of babies born (RBB) was calculated<br />

as no. BB from morula or BL ET/total no. embryos transferred. The theoretical<br />

number of oocytes needed to result in one BB was calculated as 1/<br />

(TERxRBB). To calculate the RBB, data from PGS cycles was excluded.<br />

RESULTS: See Table. Of 1,438 AOCT oocytes that survived, the 2PN<br />

fertilization rate was 68% and of those, 41% formed BL. The estimated<br />

no. of MII needed to achieve a BB was 8, 10, 14, and 50, and the likelihood<br />

of BB per MII oocyte was 13%, 10%, 7%, and 2% for the age groups 25-34y,<br />

35-37y, 38-40y, and 41-42y respectively.<br />

CONCLUSIONS: As OC is increasingly relied upon for fertility preservation,<br />

our findings provide realistic and important counseling data for women<br />

electing to freeze gametes for usage at a later time. Women in the 25-34, 35-<br />

37, and 38-40 age groups may only require one OC cycle, whereas women in<br />

the 41-42 age group should be counseled extensively and provided realistic<br />

expectations regarding the number of cycles needed to achieve success.<br />

Overall, the calculations likely underestimate the efficiency of each oocyte,<br />

as cycles involving PGS were excluded from final RBB calculations and PGS<br />

cycles resulted in an implantation rate of 64% and live birth rate of 58%.<br />

TABLE. MII oocyte development following AOCT by age.<br />

Age (y)<br />

25-34<br />

(mean 31.5y;<br />

n¼22)<br />

35-37<br />

(mean 36.3y;<br />

n¼39)<br />

38-40<br />

(mean 39.1y;<br />

n¼74)<br />

41-42<br />

(mean 41.6y;<br />

n¼25)<br />

Oocytes retrieved 500 751 1185 367<br />

MII thawed 343 431 735 234<br />

MII survived 286 348 610 194<br />

No. 2PN fertilized 228 259 358 132<br />

No. Morula + 40 + 94 26 + 99 72 + 153 13 + 56<br />

No. BL<br />

BFY 94/343<br />

(27%)<br />

99/431<br />

(23%)<br />

153/735<br />

(<strong>21</strong>%)<br />

56/234<br />

(24%)<br />

TER 134/343<br />

(39%)<br />

125/431<br />

(29%)<br />

225/735<br />

(31%)<br />

69/234<br />

(29%)<br />

OP+LB from<br />

morula ET<br />

1/7 (14%) 1/8<br />

(13%)<br />

5/27<br />

(19%)<br />

0/7<br />

(0%)<br />

OP+LB from<br />

BL ET<br />

11/29<br />

(38%)<br />

5/9<br />

(56%)<br />

3/10<br />

(30%)<br />

1/10<br />

(10%)<br />

RBB 12/36 6/<strong>17</strong> 8/37 1/<strong>17</strong><br />

No. MII oocyte to<br />

achieve BB<br />

(33%) (35%) (22%) (6%)<br />

8 10 14 50<br />

O-232 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:00 PM<br />

CURRENT TRENDS IN PUBLIC SUPPORT FOR ELECTIVE<br />

OOCYTE CRYOPRESERVATION. E. I. Lewis, a,b S. A. Missmer, a,b,c<br />

L. V. Farland, a,b,c E. Ginsburg. a,b a Center for Infertility and Reproductive<br />

Surgery, Dept of OBGYN, Brigham and Women’s Hospital, Boston, MA;<br />

b Harvard School of Medicine, Boston, MA; c Harvard School of Public<br />

Health, Boston, MA.<br />

OBJECTIVE: As oocyte cryopreservation (OC) is now an established<br />

method to preserve fertility for women due to health, aging or other factors,<br />

our aim was to determine whether there is public support for this treatment<br />

and if support varies by demographic factors.<br />

DESIGN: Cross-sectional electronic survey.<br />

MATERIALS AND METHODS: A nationally representative sample of<br />

1064 people aged 18-75 years completed an online questionnaire. Supporters<br />

of OC for various indications were compared with those participants who<br />

were neutral or in opposition using log binomial regression. Statistical<br />

models were adjusted for confounding factors including gender, ethnicity,<br />

age, income, sexual orientation, education, marital status, state political party<br />

affiliation and history of being a parent, to yield relative risk ratios (95% confidence<br />

intervals; Wald 2-sided p-values) of support.<br />

RESULTS: Of the 1,383 recruited, <strong>17</strong>% declined to take the survey and 5%<br />

were against any in-vitro fertilization and were disqualified. OC for cancer patients<br />

was the indication most Supported (89%), followed by delayed childbearing<br />

for career advancement (72%), current lack of a partner (63%), and<br />

insufficient funds for child-rearing (58%). Despite considerable support for<br />

OC, only 37% of participants agreed that employers should fund egg freezing<br />

for female employees, with those never having a child more likely to be in favor<br />

(p¼0.03). Younger participants (18-44 years old) were more likely to support<br />

OC for women currently with insufficient funds for child-rearing (p¼0.03).<br />

Younger age (p


Effectiveness of this method was tested by histology and whole ovary transplantation.<br />

MATERIALS AND METHODS: 1) Ovary preparation. Umodl1-Tg<br />

ovaries aged 3 weeks were collected and encapsulated in 0.5% alginate in<br />

aMEM/ITS+0.25mM ascorbic acid. Cross-linking was done in 50mM<br />

CaCl 2 /130mM NaCl solution. The aggregates were then grafted subcutaneously<br />

under abdominal wall of SCID mice.2) Histological and marker gene<br />

analyses. Preserved ovaries (Pr.) were dissected out at 2, 4 and 6 months after<br />

grafting for H&E or IHC-staining. Antibodies against GCNA, FSHR, Laminin<br />

and CD34 were used to examine ovarian ultrastructures.3) Fertility test.<br />

Ovaries collected at 4 months after grafting were first treated with alginate<br />

lysase. Ovaries from one side of the CD1 females were removed from bursa<br />

sacs and replaced with the Pr. ovaries. Undisturbed ovaries on the other side<br />

were used as internal fertility control. After one-month recovery, CD1 hosts<br />

were primed with PMSG/hCG. Ovulated eggs were collected for lacZ-staining<br />

to distinguish Umodl1-Tg from WT oocytes.Equal numbers of agematched<br />

WT ovaries (nR 3) were included as controls in each assay. Statistical<br />

significance was analyzed by unpaired t-test.<br />

RESULTS: Extensive vascular network had already been established between<br />

transplants and host subcutaneous tissue to support transplant survival<br />

among 2-month old transplants. All of the preserved ovaries examined exhibited<br />

excellent morphologies with distinct structures including ovarian follicles<br />

and corpus luteum. Moreover, these cells showed correct expression of<br />

their lineage-specific markers. Follicle counts on 4-month old tissue sections<br />

(n¼8) showed no significant difference between WT and the Pr. groups<br />

(WT,13745 vs. Pr, 11936/section; p¼0.40). Unlike the WT controls in<br />

which follicles of various stages were observed, the preserved follicles in<br />

the cortex seems to be synchronized at primordial/early preantral stages.<br />

When re-transplanted back to the host bursa sacs, these quiescent ovaries<br />

could be reactivated and produce comparable numbers of oocytes as the<br />

WT controls (Pr., 7.251.28 vs. WT, 8.882.85/mouse; p¼0.16).<br />

CONCLUSIONS: Ectopic transplantation of alginate-enclosed ovaries<br />

can minimize the impact from the disturbed Umodl1-Tg endocrine system,<br />

and thus maintain the ovaries at a quiescent state. Once implemented into<br />

an appropriate milieu, their ovarian functions can be fully restored.<br />

O-234 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:30 PM<br />

EFFICACY OF STANDARDIZED NURSING FERTILITY COUN-<br />

SELING ON SPERM BANKING RATES IN CANCER<br />

PATIENTS. K. L. Rotker, a H. T. Vigneswaran, b D. Omil-Lima, b<br />

G. Baird, c M. Sigman, d K. Hwang. b a Urology, Brown University, Cranston,<br />

RI; b Urology, Brown University, Providence, RI; c Lifespan Biostatistics<br />

Core, Brown University, Providence, RI; d Brown University and Lifespan,<br />

Providence, RI.<br />

OBJECTIVE: Chemotherapy often damages spermatogenesis, leading to<br />

transient or permanent infertility. However, many clinicians fail to consistently<br />

incorporate fertility preservation prior to treatment. The current study<br />

examines sperm banking amongst patients undergoing formalized fertility<br />

counseling prior to initiating chemotherapy.<br />

DESIGN: A retrospective chart review was performed for two institutions<br />

in Providence, RI. Men aged 18-50 with newly diagnosed cancer, from 1998<br />

to 2003, prior to initiation of chemotherapy were evaluated. A standardized<br />

nursing education session including directed fertility counseling was implemented<br />

at one of the institutions in 2007.<br />

MATERIALS AND METHODS: Patient characteristics including age,<br />

cancer type, chemotherapy and/or radiation regimen, marital status, offspring<br />

status, desire for information on sperm banking were recorded. Records of<br />

the area’s only sperm bank were reviewed. Using bivariate statistical analysis,<br />

rates of sperm banking amongst patients who received counseling<br />

were compared to those without counseling at institution A. Multiple logistic<br />

regression was used to model the relationship between time, offspring status<br />

and age at both institutions.<br />

RESULTS: 775 male patients were identified with a mean age of 40.98<br />

years old. The most common cancer type treated was lymph node (19.3%).<br />

At the time of diagnosis, 59.7% of patients already had children, 55.4%<br />

were married, 30.6% were single and 7.5% were divorced. Of the 775 patients<br />

studied, 411 (53%) were treated at institution A and of those 90<br />

(23.4%) received fertility counseling. Of those who received counseling 15<br />

(16.67%) underwent sperm banking compared to 20 (6.2%) of the 3<strong>21</strong><br />

who did not receive counseling. This represented a significant increase in<br />

sperm banking rates (p


Toronto, ON, Canada; f Department of Gynecology, Women’s College Hospital,<br />

Toronto, ON, Canada.<br />

OBJECTIVE: The ability to expand testicular biopsy-derived spermatogonial<br />

stem cells (SSC) ex-vivo may be required for the development male fertility<br />

preservation strategies. Our aim was to characterize and assess the potential use<br />

of first trimester human umbilical cord perivascular cells (HUCPVC) as a source<br />

of human feeder cells for the xeno-free ex-vivo expansion of SSCs.<br />

DESIGN: Paracrine testicular niche-related growth factors were measured<br />

in HUCPVC cultures. The ability of HUCPVC feeders to support ex-vivo<br />

murine germ cell expansion was compared to mouse embryonic fibroblast<br />

(MEF) feeders. In addition, the molecular properties of HUCPVCs were<br />

compared to human testicular somatic cells (TSC), to evaluate the capability<br />

of HUCPVCs to support human SSC propagation.<br />

MATERIALS AND METHODS: We established mitotically inactivated<br />

HUCPVC feeder cultures from previously established male lines in a manner<br />

analogous to MEF feeder cultures. Paracrine factors were measured in<br />

HUCPVC-conditioned media (CM) using ELISA (n¼3 independent lines).<br />

CD-1 male murine neonatal germ cells were seeded on HUCPVCs or<br />

MEFs (7500 cells/cm2) and expanded in complete SSC media. Colony number<br />

and size were compared between the 2 feeder conditions. A BrdU incorporation<br />

assay and immunocytochemistry was used to confirm the presence<br />

of proliferating GPR125+ve and DAZL+ve germ cells. Targeted RNA<br />

AmpliseqÔ was used to compare levels of 89 testicular cell lineage-associated,<br />

growth factor and ECM transcripts in CD90+ve HUCPVCs and human<br />

testicular somatic cells (TSCs) derived from orchiectomies.<br />

RESULTS: FGF2, GDNF, LIF and BMP4 levels were detected at high<br />

levels in HUCPVC-derived CM. A 3-fold increase in number SSC-like colonies<br />

was observed in conditions where germ cells were seeded on HUCPVCs<br />

when compared to MEFs after 2 passages (representing > 28 days in culture,<br />

P


O-239 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:15 PM<br />

CUMULUS-DERIVED FEEDER CELLS MAINTAIN THE PLURIPO-<br />

TENCY POTENTIAL AND IMPROVE SELF-RENEWAL PERFOR-<br />

MANCE OF HUMAN PLURIPOTENT STEM CELLS. O. Ait-<br />

Ahmed, a S. Assou, a E. Pourret, a A. Ferrieres-Hoa, b S. Hamamah. b a Human<br />

Early Embryonic Development and Pluripotency, INSERM U1203, Montpellier<br />

University (UM), Montpellier University Hospital (CHRU), Montpellier,<br />

France; b Human Early Embryonic Development and Pluripotency,<br />

INSERM U1203, Montpellier University (UM), Montpellier University Hospital<br />

(CHRU), ART/PGD Department, Montpellier, France.<br />

OBJECTIVE: Based on the assumption that the microenvironment is<br />

essential in pluripotency induction and maintenance, our objective was to<br />

examine the capacity of cells cultured from the human metaphase II oocyte<br />

CCs niche (hCC), to effectively support the propagation of human pluripotent<br />

stem cells (PSC) in comparison to human foreskin fibroblasts (hFF).<br />

DESIGN: Human cumulus cells collected from patients referred to our<br />

center for intra-cytoplasmic sperm injection (ICSI) for male infertility<br />

were cultivated in xeno-free medium and used as feeders for the growth of<br />

hiPS cells.<br />

MATERIALS AND METHODS: Human cumulus cells were dissociated<br />

mechanically from the MII oocyte on collection day and adapted to growth<br />

on a human collagen substrate for a long period in a xeno-free defined medium.<br />

Affymetrix Human Genome U133 Plus 2.0 DNA chips were performed<br />

to analyze the transcriptome of hCC and hiPS cells. Pluripotency/<br />

differentiation markers were examined by RT-qPCR, immunofluorescence<br />

and flow cytometry. Chromosomal abnormalities were analyzed using array<br />

comparative genomic hybridization (aCGH).<br />

RESULTS: We have optimized isolation and long-term culture of hCC in<br />

xeno-free conditions. This new feeder is able to support growth of hiPS cells<br />

during long term passaging. The Flow cytometric analysis showed that hiPShCC<br />

retained expression of surface markers (CD24, SSEA4, TRA-1-81 and<br />

TRA-1-60), demonstrating successful preservation of stemness. Moreover,<br />

hiPS-hCC also improved the competence to differentiate into the three<br />

germ layers in vitro (embryoid bodies) as well as in vivo (teratoma formation).<br />

Interestingly the self-renewal rate was higher for hiPS-hCC than for<br />

hiPS-hFF, suggesting that hCC has a growth-promoting effect. Additionally,<br />

hiPS-hCC did not exhibit detectable sub-chromosomal aberrations, confirming<br />

their genetic stability. A comparative gene expression study of hCC and<br />

hFF revealed significant differences in expression of cellular matrix components<br />

and an up regulation in hCC of genes known to be important players in<br />

cell proliferation such as interleukin 6 gene (IL6).<br />

CONCLUSIONS: This study shows that hCC have intrinsic properties that<br />

make them better feeders to support pluripotent stem cells growth. The upregulation<br />

of secreted factors encoding genes in hCC opens new perspectives<br />

for investigating the impact of the microenvironment in the pluripotency<br />

competence of the oocyte.<br />

Supported by: This work was partially Supported by Ferring Pharmaceuticals.<br />

O-240 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:30 PM<br />

ESTABLISHMENT OF GYNOGENETIC HAPLOID EMBRYONIC<br />

STEM CELLS AND GENERATION OF CLONED OOCYTES IN<br />

MICE. M. Kobayashi, A. Yoshida. Kiba Park Clinic Research Center, Tokyo,<br />

Japan.<br />

OBJECTIVE: The establishment of embryonic stem cells (ESCs) with one<br />

set of maternal genome will be able to provide the valuable cell source for the<br />

genetic analysis of female genome, the production of genetically modified<br />

animals, and the regenerative medicine. However, available information<br />

about these ESCs is very limited. The aim of this study was to establish gynogenetic<br />

haploid ES cells (GHESCs) with maternal haploid genome in<br />

mice. In addition, we performed the generation of cloned oocytes by using<br />

nuclear transfer with GHESCs.<br />

DESIGN: ES-like cell lines from mouse GH embryos were established and<br />

characterized. The developmental ability of cloned oocytes reconstituted<br />

with GHESCs was studied.<br />

MATERIALS AND METHODS: GH embryos were produced by removal<br />

of male pronucleus from fertilized BDF1 oocytes. Proliferated cells from inner<br />

cell mass (ICM) were picked, dispersed and cultured on fresh feeder cells.<br />

Once ESC-like colonies were formed, they were passaged by Trypsin-EDTA<br />

treatment. After several passages, marker expressions for ESCs were examined<br />

with immunohistochemical staining. The DNA content of the cells was<br />

measured by flow cytometry, and karyotype analysis was also performed.<br />

Next, for reconstitution of cloned oocytes, GHESCs were synchronized the<br />

cell cycle at G2/M phase with Nocodazoleand transferred into the perivitelline<br />

space of enucleated mouse oocyte with Sendai virus. After the membrane<br />

fusion between GHESCs and enucleated oocytes, the cloned oocytes were<br />

fertilized by ICSI and cultured in KSOM.<br />

RESULTS: Blasturation rate for GH embryos was 18.9%, and the ICM<br />

outgrowth was observed in approximately 60% of these. In these cells, the<br />

spherical colony formation and proliferation were maintained up to 4<br />

months. By immunohistchemical staining, the ESC-like cells were positive<br />

for ES cell markers, such as Oct4, Nanog, and Rex1. Further, flow cytometric<br />

analysis and karyotyping revealed that these cells had a haploid set of 20<br />

chromosomes (19 autosomes and the X chromosome). Next, in generation<br />

of cloned oocytes, high rate (95.7%) of membrane fusion between GHESC<br />

and enuclated oocyte was observed. After ICSI to the cloned oocytes reconstituted<br />

by using nuclear transfer with GHESCs, 88.5% of these cleaved to 2-<br />

cells and <strong>17</strong>.3% formed blastocyst.<br />

CONCLUSIONS: ESC-like cells with a maternal haploid chromosome set<br />

derived from GH embryos were able to establish and maintain them for a<br />

long-term period. Further, we showed clearly that the cloned oocytes reconstituted<br />

by nuclear transfer with GHESCs were able to fertilize and developed<br />

to blastocyst stage.<br />

OUTCOME PREDICTORS - CLINICAL: ART 3<br />

O-241 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:15 AM<br />

PERINATAL OUTCOMES WITH AND WITHOUT ART: A POPULA-<br />

TION-BASED STUDY OF LINKED SIBLINGS IN 12<br />

STATES. B. Luke, a M. B. Brown, b L. G. Spector. c a Obstetrics, Gynecology,<br />

and Reproductive Biology, Michigan State University, East Lansing,<br />

MI; b Biostatistics, University of Michigan, Ann Arbor, MI; c Pediatrics, University<br />

of Minnesota, Minneapolis, MN.<br />

OBJECTIVE: To evaluate perinatal outcomes of siblings, conceived with<br />

and without ART.<br />

DESIGN: Longitudinal cohort study.<br />

MATERIALS AND METHODS: Cycles during 2004-09 from the Society<br />

for Assisted Reproductive Technology Clinic Online Reporting System<br />

(SART CORS) resulting in live births were linked to their respective birth<br />

certificates in 12 States (CA, CO, CT, FL, MI, NC, NJ, NY, OH, PA, TX,<br />

VA) (ART children). All other live births to the same woman were identified<br />

(ART siblings), as well as a 10:1 sample of non-ART deliveries (Control children).<br />

All children were also linked to their respective State Cancer Registries,<br />

and death certificates through one year of age. Pairs of singleton<br />

siblings were identified in which one child was in SART CORS and the other<br />

was not. Differences in length of gestation and birthweight were compared<br />

between siblings using Student’s t-test, after adjustment for the effects of<br />

mother’s age, weight, weight gain, parity and infant gender observed in the<br />

controls. Percent distribution of ART factors were compared using the chisquare<br />

test.<br />

RESULTS: The study population included 3,<strong>21</strong>3 pairs in which the ART<br />

child was born after the sibling (ART later) and 4,597 pairs in which the sibling<br />

was born later (SIB later). The difference between siblings was 3.01.4<br />

years (meanSD) for ART later and 1.91.0 years for SIB later. The ART<br />

birth was 33512 grams heavier in ART later pairs and 83488 grams heavier<br />

in SIB later pairs. The ART birth was 2.1<strong>17</strong>.2 days earlier in ART later<br />

pairs but there was no difference in SIB later pairs. Analysis of a subset of<br />

pairs in which the sibling had no indicators of infertility treatment on their<br />

birth certificate (1,085 pairs with the ART child was born later and 2,881<br />

pairs with the sibling was born later) showed similar results. Factors used<br />

in ART that were associated with significantly less likelihood of having a sibling<br />

after an ART birth included the diagnoses of diminished ovarian reserve<br />

and tubal ligation, the use of donor oocytes, thawed embryos, and the use of<br />

ICSI or assisted hatching. The occurrence of childhood cancer or infant mortality<br />

in either the ART child or the sibling was associated with a significantly<br />

greater likelihood of having a subsequent child.<br />

CONCLUSIONS: Differences in birthweight and length of gestation between<br />

siblings were very small, regardless of method of conception. The<br />

ART factors were associated with a less optimal environment for a spontaneous<br />

birth. This suggests that factors inherent in the woman, rather than<br />

those attributable to infertility treatments, may explain the lower birthweight<br />

and shorter gestation associated with ART in prior studies which compared to<br />

the general population.<br />

e92 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


O-242 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:30 AM<br />

DO SERUM ANTI-MULLERIAN HORMONE LEVELS CORRE-<br />

LATE WITH PREGNANCY OUTCOMES IN PATIENTS WITH<br />

DIMINISHED OVARIAN RESERVE UNDERGOING IN VITRO<br />

FERTILIZATION?. N. Pereira, a J. Lekovich, a Z. Rosenwaks, b<br />

S. D. Spandorfer. c a The Ronald O. Perelman and Claudia Cohen Center for<br />

Reproductive Medicine, New York, NY; b Weill Cornell Medical College,<br />

New York, NY; c Cornell University Medical Center, New York, NY.<br />

OBJECTIVE: Previous studies have suggested that patients with serum<br />

markers of diminished ovarian reserve (DOR) have higher rates of embryonic<br />

aneuploidy and lower rates of clinical pregnancy. To test this hypothesis, we<br />

investigate the correlation between pregnancy outcomes in patients < 35<br />

years of age with good quality embryos and serum anti-mullerian hormone<br />

(AMH) levels as a marker for DOR.<br />

DESIGN: Retrospective cohort study. Patients were sub-grouped, a priori,<br />

based on serum AMH levels: < 1or> 1 ng/mL and < 0.5 or > 0.5 ng/mL.<br />

MATERIALS AND METHODS: All patients < 35 years of age undergoing<br />

fresh in vitro fertilization (IVF) - embryo transfer (ET) cycles between<br />

January 2006 and June 2013 were analyzed for potential inclusion. Patients<br />

who had 2, 8-cell, day-3 embryos transferred with grades 1, 1.5 or 2 were<br />

only included in the analysis. Cycles cancelled prior to oocyte retrieval, or<br />

those utilizing donor oocytes were also excluded. Within each AMH subgroup,<br />

demographic, baseline IVF characteristics and controlled ovarian<br />

stimulation (COS) response parameters were compared. Demographic characteristics<br />

included age, gravidity, parity, and body mass index (kg/m 2 ). COS<br />

parameters included protocol type (GnRH antagonist vs. GnRH agonist), total<br />

days of ovarian stimulation, total dosage of gonadotropins administered<br />

(IU), peak estradiol (E 2 ) level (pg/mL), peak endometrial stripe (mm), total<br />

number of oocytes retrieved, and total number of mature oocytes. Clinical<br />

pregnancy (CP), spontaneous abortion (SAB) and live birth (LB) rates<br />

were compared between AMH sub-groups. Student’s t-tests and Chi-square<br />

(c2) tests were used as indicated. Statistical significance was set at P < 0.05.<br />

RESULTS: A total of 1005 patients met inclusion criteria. Overall, there<br />

was no difference in baseline demographic or IVF characteristics. Patients<br />

in the > 1 ng/mL group required lesser gonadotropins (1952.1 968.4<br />

IU) compared to the < 1 ng/mL group (25<strong>17</strong>.1 1056.8 IU) and the < 0.5<br />

ng/mL group (3328.2 <strong>17</strong>65.9 IU). More oocytes were retrieved from the<br />

same group 13.8 (5.51) compared to the latter two (P < 0.001). The CP,<br />

SAB and LB rates in the < 1 ng/mL group was 49.2%, 5.97% and 43.2%,<br />

respectively. Similarily, the corresponding rates were 50.5%, 6.93%,<br />

43.6% in the < 0.5 ng/mL group. Overall, these rates were comparable to<br />

the CP (50.2%), SAB (6.86%), and LB (43.3%) rates in the > 1 ng/mL group.<br />

CONCLUSIONS: Serum AMH levels correlate well with number of oocytes<br />

retrieved in patients < 35 years of age with DOR. However, in patients<br />

with DOR who have good quality embryos, the correlation between serum<br />

AMH levels and clinical pregnancy, spontaneous miscarriage, and live birth<br />

rates is very limited.<br />

O-243 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:45 AM<br />

MONOZYGOTIC TWINNING IN IVF: WHY DO THEY<br />

CLUSTER?. D. A. Vaughan, a R. Ruthazer, b A. Penzias, c E. Norwitz, a<br />

D. Sakkas. d a OBGYN, Tufts Medical Center, Boston, MA; b Clinical and<br />

Translational Research Department, Tufts Medical Center, Boston, MA;<br />

c Beth Israel Deconness Hospital/Boston IVF, Waltham, MA; d Boston IVF,<br />

Waltham, MA.<br />

OBJECTIVE: The incidence of monozygotic twinning (MZT) is greatly<br />

increased among IVF patients compared to the general population (0.7-<br />

13% vs. 0.45%, respectively), but the reason for this remains unclear1.To<br />

further investigate this association, we sought to ascertain whether MZT<br />

events occur in clusters and, if so, to explore possible explanations for this<br />

clustering.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Using an established and validated IVF<br />

database from a single large urban IVF center, consecutive fresh IVF cycles<br />

(autologous and oocyte donation) resulting in a viable clinical pregnancy<br />

(confirmation of a gestational sac(s) and presence of a fetal pole with a heartbeat(s)<br />

on ultrasound) from Jan 2002 to Dec 2013 were retrospectively reviewed.<br />

The incidence of MZT overall and separately for each 6-month<br />

interval was calculated as the total number of MZT events divided by the<br />

number of viable clinical pregnancies. 6-month intervals with a MZT incidence<br />

rate >2 standard deviations (SD) higher than the overall rate were regarded<br />

as a high risk interval. Logistic regression modeling was then used to<br />

adjust for both time and non-time related risk factors for MZT events within<br />

our cohort (Table 1).<br />

RESULTS: Over the 12 year study period, 25,502 fresh IVF cycles were<br />

performed, resulting in 8,598 clinical pregnancies. Of these, 95 cycles<br />

(1.1%) resulted in MZ twins. Median patient age (SD) in the MZT cohort<br />

was 35.4 years (4.4). The % of MZTwas >2SD higher than the overall % of<br />

MZT in 4 of the 24 6-month intervals. PGD, extended embryo culture (R4<br />

days), and more recent cycle (2005 or later) were independent risk factors<br />

for MZT. Conversely, increasing numbers of embryos transferred appeared<br />

to decrease the risk of MZT. Use of multivariable logistic regression<br />

modeling to control for risk factors for MZT did not correct for this clustering<br />

effect, with both high-risk interval (clustering) and extended embryo culture<br />

remaining significant after adjustment (Table 1.)<br />

CONCLUSIONS: This study supports our hypothesis that MZ occurs in<br />

clusters, and that this clustering effect could not be explained by demographics<br />

and cycle characteristics alone. Our study would suggest that<br />

external factors both clinical (such as type of stimulation) and laboratory<br />

(type or lots of cultured media, days in culture) may be involved.<br />

Table 1. Risk Factors for MZT events among 8598 fresh IVF cycles from Jan<br />

2002 through Dec 2013.<br />

MZT<br />

(N¼95)<br />

Non-MZT<br />

(n¼8503)<br />

Un-adjusted<br />

p-value<br />

Adjusted<br />

OR<br />

(95% CI)<br />

and p-value<br />

High-risk<br />

6-month interval<br />

35.8% (34) 15.5% (13<strong>21</strong>) < .0001 2.37 (1.53, 3.66)<br />

p¼ .0001<br />

Embryo Biopsy 12.6% (12) 3.5% (298) < .0001 1.63 (0.82, 3.23)<br />

p¼ .16<br />

Assisted Hatching 7.4% (7) 11.1% (940) 0.2537 –<br />

Donor cycle 8.4% (8) 8.7% (743) 0.9133 –<br />

ICSI 31.6% (30) 38.5% (3277) 0.1655 –<br />

Year 2005 or later 86.3% (82) 72.9% (6197) 0.0033 1.56 (0.84, 2.91)<br />

p¼0.16<br />

Extended Culture 45.3% (43) 18.1% (1541) < .0001 2.68 (1.66, 4.33)<br />

p< .0001<br />

Oocyte provider<br />

>35 years<br />

46.3% (44) 48.6% (4136) 0.6519 –<br />

Age at cycle start 35.4 4.4 35.9 4.6 0.3753 –<br />

>2 embryos<br />

transferred<br />

20.0% (19) 31.7% (2692) 0.015 0.75 (0.45, 1.40)<br />

p¼0.42<br />

References:<br />

1. Vitthala S et al. The risk of monozygotic twins after assisted reproductive<br />

technology: a systematic review and meta-analysis. Hum Reprod<br />

2009;15:45-55.<br />

O-244 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:00 PM<br />

KNOWLEDGE OF EMBRYONIC PLOIDY STATUS IS ASSOCIATED<br />

WITH DECREASED TRANSFER ORDER AND INCREASED ELEC-<br />

TIVE SINGLE EMBRYO TRANSFER (ESET). J. M. Franasiak,<br />

M. D. Werner, C. R. Juneau, R. Barnett, K. H. Hong, E. J. Forman,<br />

R. T. Scott. RMA, NJ, NJ.<br />

OBJECTIVE: To evaluate how the knowledge of embryonic ploidy status<br />

affects embryonic transfer order and the rate of elective single embryo transfer.<br />

DESIGN: Retrospective cohort.<br />

MATERIALS AND METHODS: All first time IVF cycles from 1/2012 to<br />

12/2014 resulting in either a fresh or frozen embryo transfer were selected for<br />

review. Patients with recurrent pregnancy loss or undergoing PGD for single<br />

gene disorders or translocations were excluded. During this period of time<br />

Comprehensive Chromosomal Screening (CCS) was routinely offered to<br />

all patients as a way to increase pregnancy rates and decrease miscarriage<br />

rates. Data collected included use of CCS or not, number of aneuploid/<br />

euploid embryos if CCS was performed, number of embryos transferred,<br />

and number of embryos cryopreserved. eSET was designated when there<br />

were embryos cryopreserved in excess of those transferred. As per practice<br />

standards, maximum transfer order was two embryos. Thus functionally, patients<br />

selected between a one or two embryo transfer.<br />

RESULTS: During the study timeframe there were 3683 unique patients<br />

identified. Of these <strong>17</strong>78 (48%) underwent CCS and 1905 (52%) did not.<br />

Of those undergoing CCS, the average number of embryos biopsied was<br />

FERTILITY & STERILITY Ò<br />

e93


5.6 (1-46), number euploid was 3.7 (0-27), number aneuploid was 1.8 (0-<strong>21</strong>),<br />

and 114 (6.4%) had no euploid for embryo transfer. Of the 3569 who underwent<br />

transfer, the mean number of embryos transferred was 1.3 with CCS and<br />

1.6 without CCS (p


OUTCOME PREDICTORS - LAB: ART<br />

O-247 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:15 AM<br />

CIRCULATING MICRORNAS, AS POWERFUL TOOLS TO PRE-<br />

DICT IVF/ICSI OUTCOMES. E. Scalici, a,b,c T. Mullet, a,c A. Gala, a<br />

C. Vincens, a T. Anahory, a S. Belloc, d S. Hamamah. a,b,c a ART-PGD<br />

Department, Montpellier, France; b INSERM U1203, Montpellier, France;<br />

c Montpellier 1 University, UFR of Medicine, Montpellier, France; d Eylau-<br />

Unilabs Laboratory, Paris, France.<br />

OBJECTIVE: Our aim was to investigate if circulating microRNAs (miR-<br />

NAs) in human follicular fluid (FF) could be used as helpful biomarkers for<br />

predicting IVF/ICSI outcomes.<br />

DESIGN: In this prospective study, a pool of FF was retrieved for each patient<br />

during IVF/ICSI procedure. A total of 91 FF samples from women with<br />

normal ovarian reserve (n¼91) were collected at oocyte retrieval day.<br />

MATERIALS AND METHODS: For each patient, all follicles were aspirated<br />

and all FF samples were pooled. MicroRNAs were extracted from each<br />

FF pool and quantified by RT-qPCR, using TaqMan technology. The expressions<br />

of miR-320a, let-7b and miR-29a were analyzed in FF and related to<br />

IVF/ICSI outcomes.<br />

RESULTS: FF pools related to low number of mature oocytes (%2) contained<br />

significant lower miR-320a expression levels than those related to high<br />

number of mature oocytes (>2), respectively (p¼0.03). Moreover, significant<br />

high let-7b levels were found in FF pools related to embryo cohorts with a<br />

high total blastomere number/total embryo number ratio at day 3 (>8, ie<br />

accelerated development) than in those with normal developmental kinetics<br />

(ratio between 6 and 8) (p¼0.02). Interestingly, we found a significant and<br />

negative correlation between FF let-7b expression levels and blastulation<br />

rate (r¼-0.33, p¼0.003). The Receiving Operator Curve (ROC) analysis indicated<br />

that the performance of FF let-7b in predicting the expanded blastocyst<br />

development was 0.67 (0.54-0.79), with 70% sensitivity and 64.3% specificity<br />

(p¼0.02). In addition, the area under the ROC curve (AUC), evaluating<br />

the potential of FF miR-29a in predicting clinical pregnancy outcome<br />

reached 0.68 [0.55-0.79] with a high sensitivity of 83.3% and a specificity<br />

of 53.5% (p¼0.01).<br />

CONCLUSIONS: Our results suggest that circulating miRNAs constitute<br />

non-invasive powerful tools in IVF process to predict embryo development<br />

and clinical pregnancy outcomes, in order to promote personalized IVF strategy.<br />

Supported by: This work was Supported by the University-Hospital of<br />

Montpellier, INSERM and by a grant from the Ferring Pharmaceutical Company.<br />

The authors of the study have no competing interests to report.<br />

O-248 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:30 AM<br />

CELL-FREE DNA AND PREGNANCY OUTCOME: NEW BIO-<br />

MARKER OF FOLLICULAR MICRO-ENVIRONMENT. T. Mullet, a<br />

E. Scalici, a A. Gala, b A. F. Ferrieres-Hoa, c V. loup, a C. Brunet, d<br />

S. Hamamah. a a ART PGD Department, Montpellier, France; b CHU Montpellier,<br />

Montpellier, France; c CHRU Montpellier, Montpellier, France; d Endocrinologist,<br />

Montpellier, France.<br />

OBJECTIVE: The aims of this study were to investigate if cell-free DNA<br />

(cfDNA) in follicular fluid (FF) could be related to women’s ovarian reserve<br />

status and to their response to controlled ovarian stimulation (COS) protocols,<br />

and thus constitute a potential biomarker of oocyte micro-environment<br />

to predict clinical pregnancy outcome.<br />

DESIGN: This prospective study included 94 women with normal ovarian<br />

reserve, 6 with ovarian insufficiency (OI) and 27 with polycystic ovary syndrome<br />

(PCOS). For each patient, FF samples were collected during IVF/<br />

ICSI procedure.<br />

MATERIALS AND METHODS: At oocyte retrieval day, the FF samples<br />

of the same patient were collected and pooled. CfDNA concentration in<br />

pooled FF samples was evaluated by ALU quantitative PCR.<br />

RESULTS: CfDNA level was significantly higher in follicles from women<br />

with impaired ovarian reserve (OI and PCOS), compared to women with<br />

normal ovarian reserve (2.7 2.7 ng/mL versus 1.7 2.3 ng/mL, p¼0.03,<br />

respectively). Likewise, FF cfDNA levels were significant more elevated in<br />

women who received long ovarian stimulation protocols (> 10 days) or<br />

high total dose of gonadotropins (R 3000 IU/L) than in women who<br />

received short stimulation protocols (7-10 days) or total dose of<br />

gonadotropins < 3000 IU/l (2.4 2.8 ng/mL versus 1.5 1.9 ng/mL,<br />

p¼0.008; 2.2 2.3 ng/mL versus 1.5 2.1 ng/mL, p¼0.01, respectively).<br />

FF cfDNA level was an independent and significant predictive factor for<br />

pregnancy outcome (adjusted odds ratio¼0.69 [0.5; 0.96], p¼0.03). In<br />

multivariate analysis, the Receiving Operator Curve (ROC) analysis showed<br />

that the performance of FF cfDNA in predicting clinical pregnancy reached<br />

0.73 [0.66-0.87] with 88% specificity and 60% sensitivity.<br />

CONCLUSIONS: Our results suggest that cfDNA quantification in FF<br />

pools could provide a new non-invasive and easy method to assess the quality<br />

of the oocyte microenvironment and to predict the clinical pregnancy<br />

outcome. Moreover, in the case of IVF failure, FF cfDNA quantification<br />

could be used to improve the personalized patient’s care and thus to increase<br />

the chance of success of the next IVF cycle by selecting embryo for replacement<br />

issue from follicle which contains low cfDNA level.<br />

O-249 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:45 AM<br />

IMPACT OF TIME-LAPSE AND REDUCED OXYGEN CULTURE<br />

ON LIVE BIRTH RATE AND ITS CORRELATION WITH INFER-<br />

TILITY DIAGNOSIS. N. Zaninovic, Q. Zhan, R. Clarke, Z. Ye,<br />

N. Pereira, Z. Rosenwaks. CRM, Weill Cornell Medical College, New<br />

York, NY.<br />

OBJECTIVE: To evaluate the impact of implementing time-lapse with<br />

reduced oxygen for all IVF patients on live birth (LB) rate. To determine<br />

the correlation between culture conditions with specific patient clinical diagnosis.<br />

DESIGN: A retrospective analysis of IVF patient outcomes from 2012-<br />

2013 cultured in standard incubator (S; n¼1199) vs. time-lapse incubator<br />

(E; n¼1318) was performed. LB rate was used to calculate ongoing pregnancy<br />

(OP) and implantation (IMPL) rates for D3 and D5 ET patients; pregnancy<br />

data was analyzed by maternal age groups (SART). Logistic regression<br />

analysis was used to correlate patient diagnosis with clinical outcome.<br />

MATERIALS AND METHODS: Patients were cultured either in standard<br />

incubator (S, 20% oxygen, Thermo Forma, USA) or EmbryoScopeÒ (E, 5%<br />

oxygen, Vitrolife) using in-house sequential medium.<br />

RESULTS: A significant increase in OP, IMPL and live birth cycles (LBC)<br />

was observed In EmbryoScopeÒ vs. standard incubator (38.7% vs. 29.6%;<br />

19.5% vs. 13.4%; 38.1% vs. 29.3%, respectively; p


MATERIALS AND METHODS: ICSI videos were quantified for the<br />

following parameters: time of needle exposure to the ooplasm, intracellular<br />

needle movement, cytoplasmic leakage, needle touching the membrane adjacent<br />

to the holder, sperm position, volume of cytoplasm removed during injection,<br />

as well as oocyte morphological characteristics. Various endpoints<br />

were evaluated [degeneration, fertilisation, two-pronucleate(2PN), selection<br />

for embryo transfer, blastocyst formation, embryo quality from day 2 to 6,<br />

known implantation data]. Parameters found to significantly affect outcome<br />

were compared between four ICSI practitioners, five needle types (Humagen<br />

SI,SLM,35 or Wallace 45V,50V) and two ICSI rigs (Integra-Ti and Integra-<br />

3). Data were analysed using ANOVA, Kruskal Wallis or Chi-square as<br />

appropriate.<br />

RESULTS: Fertilisation rates were not significantly affected by<br />

practitioner(range71-78%), equipment(74-74%) or needle type(75-78%).<br />

Of all the endpoints assessed, only degeneration, 2PN and<br />

fertilisation rates were affected by ICSI events. ICSIs associated with<br />

intracellular needle movement(66%, n¼ 153) were significantly less<br />

likely to fertilise than controls(77%, n¼373, p


CLINICAL FEMALE INFERTILITY AND GYNECOLOGY<br />

O-253 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:15 AM<br />

AN INTERDISCIPLINARY LIFESTYLE INTERVENTION IMPROVES<br />

CLINICALLY RELEVANT FERTILITY OUTCOMES IN OBESE<br />

INFERTILE WOMEN - PRELIMINARY RESULTS. K. Duval, a,b<br />

M. Belan, a,b F. Jean-Denis, b J. Baillargeon. a,b a Medicine, Division of Endocrinology,<br />

Universite de Sherbrooke, Sherbrooke, QC, Canada; b Centre de<br />

Recherche du Centre Hospitalier Universitaire de Sherbrooke, Sherbrooke,<br />

QC, Canada.<br />

OBJECTIVE: To assess the impact of a lifestyle intervention targeting<br />

obese infertile women on fertility outcomes.<br />

DESIGN: Randomized controlled trial.<br />

MATERIALS AND METHODS: Participants were followed for 18<br />

months or until the end of pregnancy. They were randomized into the intervention<br />

(interdisciplinary lifestyle intervention without fertility treatments<br />

for the first six months) or control group (standard fertility treatments<br />

only). The lifestyle intervention consisted in individual’s encounters with a<br />

dietician and a kinesiologist every 6 weeks, and weekly group sessions (at<br />

least 12 different ones). Pregnancy was defined by a positive serum pregnancy<br />

test. A total of 105 women were randomized, <strong>21</strong> were excluded/withdrawn<br />

(13 intervention; 8 control), 29 are active in the study and 55<br />

completed (24 intervention; 31 control). Only the results of participants<br />

who completed the study are presented.<br />

RESULTS: Mean age of participants was 30.2 4.7 yrs, BMI is 40.1 7.5<br />

kg/m 2 , waist circumference is 118 16 cm, and 67% had polycystic ovary<br />

syndrome. As compared with the control group, women in the intervention<br />

group had significantly higher pregnancy rate (79.2% vs. 41.9%, p¼0.003)<br />

and spontaneous pregnancy rate (50.0% vs. 12.9% of all women,<br />

p¼0.003). There was no significant difference between groups for pregnancy<br />

following assisted reproductive technology (29% in both groups). Importantly,<br />

a tendency to a higher rate of life birth was observed in the intervention<br />

group (62.5 % vs. 38.7%, p¼0.08). At any time during the study, 42% of<br />

women in the intervention group and 36% in the control group lost R5%<br />

of their initial body weight (p¼NS). Furthermore, 50% of women in the intervention<br />

group and 36% in the control group lost R5 cm of their initial waist<br />

circumference (p¼NS). The median weight loss at time of pregnancy in<br />

women who delivered (n¼28) was -2.5% [-6.9 - -1.4%] (p¼0.0001) in the<br />

intervention group and -1.6% [-2.4 - +1.0] (p¼0.22) in controls (p¼0.05 between<br />

groups).<br />

CONCLUSIONS: These results demonstrate that even a modest weight<br />

loss following an interdisciplinary lifestyle intervention in obese infertile<br />

women could improve significantly their pregnancy rate, mainly the occurrence<br />

of a spontaneous pregnancy. They also suggest that the rate of living<br />

babies may also be improved. Such interdisciplinary lifestyle intervention<br />

may improve fertility by other factors than weight or waist loss alone.<br />

Supported by: Quebec Ministry of Health and Social Services, and Canadian<br />

Institutes of Health Research.<br />

O-254 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:30 AM<br />

CLOMID STAIR-STEP PROTOCOL MAY SHORTEN THE TIME TO<br />

OVULATION BUT NOT TO PREGNANCY: A RANDOMIZED CLIN-<br />

ICAL TRIAL. L. B. Craig, a J. D. Peck, b D. Zhao, b K. R. Hansen. a a Dept<br />

Ob/Gyn, Section Reproductive Endocrinology & Infertility, University of<br />

Oklahoma Health Sciences Center, Oklahoma City, OK; b Dept of Biostatistics<br />

& Epidemiology, University of Oklahoma Health Sciences Center, Oklahoma<br />

City, OK.<br />

OBJECTIVE: A novel dosing regimen for clomiphene citrate that<br />

involved increasing the dose (stair-step protocol) without inducing a<br />

period with progestin has been retrospectively described. The objective of<br />

this study was to prospectively compare length of time to ovulation and<br />

pregnancy with escalating doses of clomiphene administration in the<br />

traditional protocol versus the stair-step approach in women with ovulatory<br />

dysfunction.<br />

DESIGN: Unblinded, randomized clinical trial<br />

MATERIALS AND METHODS: Women ages 18-45 with anovulatory<br />

infertility were randomized in a 1:1 ratio at baseline ultrasound to two<br />

protocols for ovulation induction. The traditional protocol included<br />

clomiphene followed by progestin withdrawal if anovulatory before<br />

increasing the clomiphene dose. Under the stair-step protocol, clomiphene<br />

dose was increased cycle day 11-14 without a progestin withdrawal if no<br />

follicle >12mm. Clomiphene dose started at 50mg and was increased up<br />

to 150mg per protocol. Primary outcome was time to ovulation. Secondary<br />

outcome was time to pregnancy. Outcomes were compared between the<br />

two treatment groups using the log-rank test and the proportions of ovulation/pregnancy<br />

versus time were plotted using the Kaplan-Meier method.<br />

The proportion of ovulation/pregnancy between the two treatment groups<br />

was analyzed with the chi-square test.<br />

RESULTS: A total of 120 patients (60 stair-step, 60 traditional) were<br />

randomized. The groups were similar in age, BMI, ethnicity, years of<br />

infertility, and smoking. There was no significant difference in proportion<br />

that ovulated between the stair-step and traditional groups (p¼0.77, 90.0%<br />

vs 88.3%). In addition, there was no significant difference in time to first<br />

ovulation between the two groups (p¼0.65). However, when comparisons<br />

were limited to the 49 patients (34 stair-step, 15 traditional) who did not<br />

ovulate at clomiphene 50mg, time to first ovulation was significantly<br />

decreased (p $50,000 USD<br />

(OR¼2.07, CI¼1.27-3.38), duration of infertility of >24 months<br />

(OR¼0.65, CI¼0.47-0.89), and history of a prior pregnancy loss<br />

(OR¼1.59, CI¼1.12-2.27) were significantly associated with the probability<br />

of live-birth. Other baseline demographic, biochemical, and lifestyle characteristics<br />

were not associated with outcomes.<br />

CONCLUSIONS: While age and duration of infertility were significant<br />

predictors of all pregnancy outcomes, most other baseline characteristics<br />

FERTILITY & STERILITY Ò<br />

e97


were not. The identification of income as a predictor of outcomes independent<br />

of race and education may be reflective of differences in the underlying<br />

etiologies of infertility between the groups as well as disparities in access to<br />

fertility and/or obstetrical care.<br />

References: The content is solely the responsibility of the<br />

authors and does not necessarily represent the official views of the NICHD<br />

or NIH.<br />

Supported by: The Eunice Kennedy Shriver National Institute of<br />

Child Health and Human Development (NICHD) Grants for AMIGOS:<br />

U10 HD39005, U10 HD38992, U10 HD27049, U10 HD38998, U10<br />

HD055942, HD055944, U10 HD055936, U10HD055925 PPCOSII: U10<br />

HD27049, U10 HD38992, U10HD055925, U10 HD39005, U10 HD38998,<br />

U10 HD055936, U10 HD055942, U10 HD055944. This research<br />

was made possible by the funding by American Recovery and Reinvestment<br />

Act.<br />

O-256 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:00 PM<br />

AUTOIMMUNE PROGESTERONE DERMATITIS: CLINICAL PRE-<br />

SENTATIONS AND MANAGEMENT. THE BRIGHAM AND<br />

WOMEN’S HOSPITAL EXPERIENCE. D. Foer, a K. Buchheit, b<br />

D. Lynch, b A. R. Gargiulo, c M. Castells, b P. Wickner. b a Internal Medicine,<br />

Brigham and Women’s Hospital, Boston, MA; b Rheumatology, Immunology<br />

and Allergy, Brigham and Women’s Hospital, Boston, MA; c Brigham and<br />

Women’s Hospital, Boston, MA.<br />

OBJECTIVE: Autoimmune progesterone dermatitis (APD) is a poorly<br />

recognized complex syndrome that can lead to difficulty tolerating fertility<br />

treatment. It is commonly associated with hypersensitivity reaction to exogenous<br />

or endogenous progesterone, often with cyclical symptoms correlating<br />

with the menstrual cycle. Symptoms can range from dermatitis to anaphylaxis,<br />

and an increasing number of cases have been reported after the advent<br />

of in vitro fertilization (IVF), during which women are exposed to supraphysiological<br />

levels of progesterone. We present the largest case series of<br />

APD with successful outcomes following progesterone desensitization<br />

including pregnancy and symptom resolution.<br />

DESIGN: Case series from a large academic hospital Immunology<br />

and Allergy practice, managed in conjunction with a reproductive<br />

endocrinologist.<br />

MATERIALS AND METHODS: Twenty-two cases of APD seen at BWH<br />

were evaluated. Symptom presentation, diagnostic modalities, desensitization<br />

protocols, and outcomes were analyzed. Patient follow-up was conducted<br />

to assess for long-term effects of desensitization including fertility<br />

treatment and quality of life.<br />

RESULTS: Symptoms were heterogeneous and included cyclical dermatitis,<br />

infertility, urticaria, angioedema, and bronchospasm. In nine patients<br />

(41%) the symptoms followed menarche or pregnancy, while in 13 patients<br />

(59%), these seem to have been triggered by exogenous progesterone/progestin<br />

used for contraception or fertility treatment. Nine patients were desensitized<br />

to progesterone. All nine patients had successful desensitizations, 3<br />

patients were desensitized prior to embryo transfer with an IM progesterone<br />

protocol. Desensitization resulted in significant symptom relief in 6 cases and<br />

at least 2 pregnancies through IVF.<br />

CONCLUSIONS: APD presents heterogeneously and may be misdiagnosed.<br />

Women with cyclical allergic symptoms and infertility should be evaluated<br />

jointly by an allergist and a reproductive endocrinologist, for possible<br />

diagnosis of APD and desensitization/endocrine manipulation. This is the<br />

largest case series of APD reported to date, and demonstrates that progesterone<br />

desensitization is successful, reproducible, and can result in pregnancy<br />

via IVF and symptom resolution with improved quality of life.<br />

References:<br />

1. Jenkins, J, Geng, A, Robinson-Bostom. Autoimmune progesterone<br />

dermatitis associated with infertility treatments, Journal of American<br />

Academy of Dermatology, (2008) 58: 353-354.<br />

2. Farah, FS, Shbakli, Z. Autoimmune progesterone urticaria, Journal of<br />

Allergy and Clinical Immunology (1971) 48:257-261<br />

3. Snyder, JL, Krishnaswamy, G. Autoimmune progesterone dermatitis<br />

and its manifestations as anaphylaxis: a case report and literature review,<br />

Annals Allergy, Asthma, and Immunology (2003) 90:469-477.<br />

4. Prieto-Garcia, A, Sloane, DE, Gargiulo, AR, Feldweg, AM, Castells,<br />

M. Autoimmune progesterone dermatitits:clinical presentation and<br />

management with desensitization for successful in vitro fertilization,<br />

Fertility and Sterility (2011).<br />

O-257 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:15 PM<br />

PASSIVE FERTILITY PREDICTION USING A NOVEL VAGINAL<br />

RING AND SMARTPHONE APPLICATION. W. Webster, a<br />

E. M. Godfrey, b L. Costantini, c J. Katilius. c a Emergency Department, EvergreenHealth<br />

Medical Center, Kirkland, WA; b Obstetrics & Gynecology, University<br />

of Washington, Seattle, WA; c Prima-Temp, Inc, Boulder, CO.<br />

OBJECTIVE: Continuous core temperature (CCT) monitoring provides a<br />

reliable marker of circadian rhythm and a statistical model for identifying the<br />

circamensal nadir (CN) in a woman’s menstrual cycle. We have developed a<br />

self-inserted flexible vaginal ring that continuously monitors true core body<br />

temperature and passively communicates CCT to a smartphone app. Diurnal<br />

changes in CCT associated with the transition of the follicular phase to luteal<br />

phase include a transient drop in the mesor (mean temperature) with an<br />

altered period and acrophase (time of peak temperature) for each cycle<br />

that characterizes the CN. The CN occurs within 1 to 2 days before the LH<br />

surge. In this study, we evaluate the safety and reliability of this sensor,<br />

and collect data to further define the predictive algorithm for identifying a<br />

woman’s fertile window.<br />

DESIGN: The study compares the ovulation predicted by this<br />

temperature sensor compared with: 1) daily oral basal body temperature<br />

(BBT) measures, 2) urine LH ovulation prediction kit, 3) serum progesterone<br />

levels, and 4) transvaginal ultrasound. The study also identifies user<br />

acceptance and ease of use. Throughout the menstrual cycle, the sensor<br />

measures CCT every 6 minutes, and data are transferred wirelessly to an<br />

iPhone app every 2 hours. This observational, multi-site, randomized study<br />

is designed to evaluate the sensor in healthy volunteers or women who are<br />

currently trying to conceive.<br />

MATERIALS AND METHODS: 20 female subjects use the sensor for 3<br />

menstrual cycles. Subjects keep an electronic diary of daily oral BBT and<br />

urinary LH tests. Statistical analysis is performed by paired t-test and<br />

multilevel model approach. Ultrasound is used to ascertain ovulation. The<br />

resulting algorithm both predicts and identifies ovulation in comparison<br />

to once a day oral temperature, urinary LH detection and progesterone<br />

methods.<br />

RESULTS: Comparison between all methods is analyzed by measuring the<br />

temporal relationship of a positive test given by each method from the true<br />

day of ovulation as recorded by ultrasound. The resulting novel algorithm<br />

correctly predicts ovulation.<br />

CONCLUSIONS: This innovative vaginal ring and its associated smartphone<br />

app are a novel method to address the needs for passively predicting<br />

ovulation. CCT will provide an improved algorithm for ovulation prediction<br />

from a complete model of circadian rhythm resulting in a safe, convenient<br />

and more accurate method of ovulation prediction.<br />

Supported by: Prima-Temp, Inc.<br />

O-258 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:30 PM<br />

UTILITY OF SONOHYSTEROGRAPHY FOR TUBAL PATENCY<br />

ASSESSMENT IN THE PREGNANCY IN POLYCYSTIC OVARY<br />

SYNDROME II TRIAL. M. S. Christianson, a R. Legro, b S. Jin, c<br />

E. Eisenberg, d M. P. Diamond, e K. R. Hansen, f W. Vitek, g A. K. Styer, h<br />

P. R. Casson, i C. Coutifaris, j G. M. Christman, k R. Alvero, l<br />

E. E. Puscheck, m A. Christy, n H. Zhang, c A. J. Polotsky, l N. Santoro. l<br />

a Johns Hopkins University School of Medicine, <strong>Baltimore</strong>, MD; b Penn State<br />

University College of Medicine; c Yale School of Public Health, New Haven,<br />

CT; d NICHD, Bethesda, MD; e Georgia Regents University, Augusta, GA;<br />

f University of Oklahoma Health Sciences Center, Oklahoma City, OK; g University<br />

of Rochester Medical Center, Rochester, NY; h Massachusetts General<br />

Hospital/Harvard Medical Sch, Boston, MA; i Northeastern Reproductive<br />

Medicine, Colchester, VT; j University of Pennsylvania, Philadelphia, PA;<br />

k University of Florida School of Medicine, Gainesville, FL; l University of<br />

Colorado, Aurora, CO; m Wayne State University School of Medicine, Detroit,<br />

MI; n NIH, Rockville, MD.<br />

OBJECTIVE: To compare saline infusion sonohysterography (SIS) versus<br />

hysterosalpingogram (HSG) for confirmation of tubal patency.<br />

DESIGN: Secondary analysis of the randomized controlled trial, Pregnancy<br />

in Polycystic Ovary Syndrome II (PPCOS II).<br />

MATERIALS AND METHODS: 750 infertile women (18-39 years old)<br />

with polycystic ovary syndrome (PCOS) were randomized to up to 5 cycles<br />

of letrozole or clomiphene citrate. Tubal patency was determined by HSG,<br />

e98 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


the presence of pelvic free fluid during SIS, laparoscopy, or evidence of an<br />

intrauterine pregnancy within the previous 3 years. Logistic regression was<br />

conducted with clinical pregnancy as the outcome and HSG or SIS as the<br />

key independent variable. Results are reported as adjusted odds ratios<br />

(OR) and 95% confidence intervals.<br />

RESULTS: Tubal patency was confirmed in 511 subjects (68.1%) with SIS<br />

and 2<strong>17</strong> subjects (28.9%) with HSG (minimum of 1 patent tube). Unilateral<br />

occlusion was observed in 32 subjects (2.9%) during HSG. After adjustment<br />

for treatment arm, body mass index, duration of infertility, smoking and education,<br />

the likelihood of clinical pregnancy was similar with HSG or SIS<br />

(OR 1.16, 95% CI 0.80,1.70, p¼0.440). Clinical pregnancy rates remained<br />

similar between groups following sensitivity analysis for ovulatory cycles<br />

(n¼619). Ectopic pregnancy occurred more often in subjects with tubal<br />

patency confirmed by HSG compared to SIS (2.8% versus 0.6%, p¼0.02).<br />

Two out of 32 women (6.3%) with unilateralocclusion on HSG were diagnosed<br />

with ectopic pregnancy.<br />

CONCLUSIONS: In this large cohort of women with PCOS, there was no<br />

significant difference in clinical pregnancy rate between subjects who had<br />

tubal patency confirmed by SIS versus HSG. SIS is an acceptable imaging<br />

modality for assessment of tubal patency in this population. The increased<br />

incidence of ectopic pregnancy with unilateral occlusion on HSG merits<br />

further evaluation.<br />

Supported by: NICHD R25HD075737<br />

IN VITRO FERTILIZATION 3<br />

O-259 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:15 AM<br />

COST ANALYSIS OF GLOBAL VERSUS SELECTIVE SCREENING<br />

STRATEGIES FOR CAVITARY LESIONS IN WOMEN SCHEDULED<br />

FOR IN-VITRO FERTILIZATION (IVF). A. M. Abdelmagied, a,b<br />

M. A. Kamel, b A. M. Abuelhasan, b T. A. Farghaly, b A. A. Nassr, a,b<br />

S. A. Shazly, a,b M. H. Makarem. b a Obstetrics and Gynecology, Mayo Clinic,<br />

Rochester, MN; b Obstetrics and Gynecology, Women Health Hospital, Assiut<br />

University, Assiut, Egypt.<br />

OBJECTIVE: Although it is a common practice to screen all IVF women<br />

by office hysteroscopy (OH), it is not always a cost effective approach especially<br />

in low resource settings. The objective of this study was to provide a<br />

cost analysis to support global versus selective use of OH to detect cavitary<br />

lesions (CL) in women scheduled for IVF.<br />

DESIGN: Prospective cohort study and subsequent cost modeling<br />

MATERIALS AND METHODS: Infertile women scheduled for IVF were<br />

examined by trans-vaginal ultrasound (TVUS) then by OH, the gold standard,<br />

to evaluate the endometrial cavity. To assess the utility of OH after<br />

TVUS, two suggested OH screening strategies were compared in terms of<br />

diagnostic accuracy and cost benefits: global for all women and selective.<br />

The selective strategy included women with abnormal TVUS and/or significant<br />

clinical predictors for CL. Statistical methods used for comparisons<br />

included t-test, Chi-square test, Wilcoxon rank sum test and logistic regression.<br />

The diagnostic accuracy of the selective strategy (index method) was<br />

compared to the global strategy (reference method) using McNemar’s test.<br />

A Cost-by-Detected Case Cost Model was utilized to compare the cost benefits<br />

of both strategies.<br />

RESULTS: Out 120 women included, 34 (28.3%) had CL by OH. TVUS<br />

detected only <strong>21</strong> CL with sensitivity, specificity and accuracy: 62%, 93% and<br />

77% respectively. Women with CL had older maternal (Median(IQR),<br />

31.5[5.25] vs. 29[7], P¼0.009) and paternal (MeanSD, 38.8 6.5 vs<br />

35.5 5.6, P¼ 0.007) age and were more likely to have metrorrhagia<br />

(<strong>17</strong>.6% vs.1.2%, P¼0.002) and a previous significant endometrial<br />

procedure (23.6 vs 3.5%, P¼0.002) than women with no CL. Using<br />

multivariate regression analysis, only metrorrhagia (aOR ¼ 20.15, 95%<br />

CI¼2.15- 189.08, P¼0.009) and a previous procedure (aOR ¼ 9.16, 95%<br />

CI¼2.13 - 39.3, P¼0.003) were significant. Based on that, a predictive model<br />

including abnormal TVUS, metrorrhagia and/or a previous endometrial<br />

procedure was used for the selective screening strategy. Selective screening<br />

using this model would have had a Negative Predictive Value (NPV) of 93%<br />

and would have missed only 6 cases (<strong>17</strong>.6 %). These 6 cases would be 4<br />

arcuate uteri and 2 small endometrial polyps. Both selective and global<br />

strategies were comparable in terms of diagnostic accuracy (P¼0.33).<br />

Considering OH cost of $<strong>21</strong>53 (according to Medicare <strong>2015</strong> National Fee<br />

Estimate), the cost of global screening was $7598.8/detected case, and the<br />

cost of selective screening would be $2998.8/detected case.<br />

CONCLUSIONS: Selective screening based on abnormal TVUS, metrorrhagia<br />

and/or past significant endometrial procedure was comparable to<br />

global screening in terms of diagnostic accuracy with cost benefits utilizing<br />

the selective strategy. The suggested selective strategy is particularly valuable<br />

in low resource setting for CL screening before IVF.<br />

Supported by: Research support in university hospitals<br />

O-260 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:30 AM<br />

ABNORMAL IMPLANTATION IN DAY 3 VERSUS DAY 5 EMBRYO<br />

TRANSFERS IN FRESH IVF CYCLES: AN ANALYSIS FROM THE<br />

SOCIETY FOR ASSISTED REPRODUCTIVE TECHNOLOGY<br />

DATABASE. A. Kathiresan, a E. T. Wang, b N. Greene, b<br />

C. J. Alexander, c M. D. Pisarska. b a UCLA, Los Angeles, CA; b Cedars Sinai<br />

Medical Center, Los Angeles, CA; c Southern California Reproductive Center,<br />

Beverly Hills, CA.<br />

OBJECTIVE: Abnormal implantation may be an indicator of a non-receptive<br />

uterine environment. This could be on the spectrum of placentation abnormalities<br />

that lead to adverse pregnancy outcomes associated with IVF. In<br />

order to better understand if timing of embryo transfer plays a role, our objective<br />

was to determine if there was a difference in abnormal implantation in<br />

day 3 (D3) versus day 5 (D5) fresh embryo transfers (ET) as reported to<br />

the Society of Assisted Reproductive Technology (SART).<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Data was obtained from the SART national<br />

database. First autologous non-PGD IVF cycles involving D3 or D5<br />

fresh embryo transfers in the United States between 2004 and 2009 were<br />

included in the analyses. A total of 339,837 cycles fulfilled the inclusion<br />

criteria. Parametric and nonparametric statistical analyses were used to evaluate<br />

differences between D3 and D5 ET in all cycles. In a subset analysis of<br />

193,<strong>21</strong>8 cycles in which conception occurred, logistic regression was used to<br />

determine if day of transfer was associated with a higher risk of abnormal implantation.<br />

Abnormal implantation was defined as biochemical pregnancy,<br />

ectopic pregnancy, or pregnancy loss.<br />

RESULTS: When divided into D3 and D5 ET groups, the D3 ET group<br />

consisted of women who were older (35.6 4.7 vs. 33.7 4.3, p


clinic, or if embryo transfer occurred on days other than 3 or 5. Among the 40<br />

states with more than one clinic, 6 have legislation requiring insurance<br />

coverage for at least one IVF cycle and were designated ‘‘mandated:’’ CT,<br />

HI, IL, MA, MD, and NJ. The remaining 34 states were designated ‘‘nonmandated.’’<br />

Regression models with a mandate*year of cycle interaction<br />

term were used to examine the effect of mandate status on changes in multiple<br />

birth rate per live birth, the proportion of transfers with elective single<br />

embryo transfer (eSET), and the mean number of embryos transferred over<br />

time.<br />

RESULTS: A total of <strong>17</strong>3,968 cycles were included in the analysis. The<br />

multiple birth rate was lower in mandated than non-mandated states (P <<br />

0.001) and decreased over time in both groups (P < 0.001). Similarly, the<br />

proportion of transfers with eSET was higher in mandated than nonmandated<br />

states (P < 0.001), and increased over time in both groups (P <<br />

0.001). The mean number of embryos transferred was lower in mandated<br />

than non-mandated states (P < 0.001), and decreased along a similar trajectory<br />

over time when embryo transfer was performed on day 5 (P < 0.001).<br />

Conversely, there was a significant interaction between mandate status and<br />

time among cycles with day-3 embryo transfer (P < 0.001). The mean number<br />

of embryos transferred decreased over time in both groups (P < 0.001),<br />

but the trajectory of decline was steeper for non-mandated states between<br />

2007 and 2009 (P < 0.001). Between 2010 and 2011, the trajectory of decline<br />

was similar between the groups (P ¼ NS).<br />

CONCLUSIONS: Multiple birth rates have decreased over time in both<br />

mandated and non-mandated states. Although the gap between these two<br />

groups has narrowed with respect to the mean number of embryos transferred<br />

on day 3, providers in mandated states still transfer fewer day-5 embryos and<br />

are more likely to perform eSET.<br />

O-262 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:00 PM<br />

DIFFERENCE IN BIRTH WEIGHT OF CONSECUTIVE SIBLING<br />

SINGLETONS IS NOT FOUND IN OOCYTE DONATION<br />

COMPARING FRESH VERSUS FROZEN EMBRYO<br />

REPLACEMENTS. D. Galliano, a N. Garrido, b V. Serra, c A. Pellicer. c<br />

a IVI Foundation. IVI Barcelona, Barcelona, Spain; b Istituto Universitario<br />

IVI, Valencia, Spain; c IVI Valencia, Valencia, Spain.<br />

OBJECTIVE: First, to assess if there are any differences in birth weight<br />

or gestational length in newborns from egg donation pregnancies delivering<br />

singletons, originating from either fresh or frozen/thawed embryos, when<br />

they were developed and delivered within the same mothers. Second, to<br />

determine if there are any clinical, phenotypical or laboratory factors influencing<br />

this relationship, including the origin of the oocyte (same or<br />

different donor), the order of the children (first fresh or first frozen/thawed<br />

embryo transfer), embryo freezing technique (vitrification or slow<br />

freezing), the in-vitro embryo culture length, and the duration that embryos<br />

remained frozen.<br />

DESIGN: Retrospective cohorts study.<br />

MATERIALS AND METHODS: 360 women undergoing oocyte donation<br />

(OD), delivering (>28 weeks) at least two babies, each one from a single<br />

pregnancy, originating from at least one fresh and one frozen-thawed embryo<br />

transfer, controlling maternal and laboratory characteristics, to test the effect<br />

of embryo freezing on children size (n¼731).<br />

RESULTS: From fresh vs. thawed embryos, respectively, mean birth<br />

weight of children was 3183.7g (95%CI: 3115.0-3252.4) vs. 3226.4g (95%<br />

CI: 3166.3-3243.2), gestational age (days) was 272.1 days (95%CI: 270.1-<br />

274.0) vs 268.8 days (95%CI: 263.1-274.5), and mean percentiles were<br />

47.6th (95%CI: 44.5-50.8) vs 50.1th (95%CI: 46.8-53.3). The proportion<br />

and corresponding odds ratio (OR) from fresh vs. thawed embryos,<br />

respectively, were for children LGA: 13.6% vs. 11.3%<br />

(OR¼0.81(0.52-1.27)); SGA: 9.4% vs 12.5% (OR¼1.37(0.85-2.2)); ONR<br />

23.1% vs. 23.8% (OR¼1.04(0.74-1.47)); and macrosomy 0.3% vs 0.8%<br />

(OR 3.1(0.3-29.7)). Mean Z-scores were 0.28(-0.13 to 0.09) vs 0.04(-0.06<br />

to 0.14), respectively. After adjusting for clinically relevant variables, the<br />

Adj(OR)LGA¼0.96(0.50-1.87); Adj(OR)SGA¼1.40(0.72-2.71); Adj(OR)<br />

ONR¼1.20(0.73-1.97), and Adj(OR)MS¼not computable. None of the<br />

stated measures were significantly different. Also, independent analyses<br />

run on the origin of the oocytes, cryopreservation technique, cleavage stage<br />

of the embryos, and time range of embryos remaining frozen did not reveal<br />

any significant trends.<br />

CONCLUSIONS: This study comparing siblings from OD cycles, and<br />

eliminating the independent variables which affect early events in pregnancy,<br />

revealed no difference in duration of gestation and live birth weights between<br />

fetuses obtained after the replacement of fresh or frozen embryos. Moreover,<br />

no clinical, phenotypical or laboratory factors appeared relevant, once statistically<br />

controlled.<br />

O-263 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:15 PM<br />

TRENDS AND OUTCOMES OF ASSISTED REPRODUCTIVE<br />

TECHNOLOGY (ART) CYCLES USING GESTATIONAL CARRIERS<br />

IN THE UNITED STATES, 1998-2012. K. Perkins, S. Boulet,<br />

D. M. Kissin, D. J. Jamieson. Centers for Disease Control and Prevention,<br />

Atlanta, GA.<br />

OBJECTIVE: To determine trends in gestational carrier (GC) cycles from<br />

1998 to 2012 and compare reproductive outcomes of GC and non-GC cycles<br />

DESIGN: Retrospective cohort study<br />

MATERIALS AND METHODS: We used data from the National ART<br />

Surveillance System and included all ART cycles with transcervical embryo<br />

transfer performed during 1998-2012. We calculated the absolute number<br />

and percent of all ART cycles using GCs over the study period. Use of<br />

GCs among non-US residents was assessed due to high utilization in this population.<br />

Temporal trends were assessed using the Cochran-Armitage trend<br />

test. For the most recent years of data (2007-2012), we used log binomial<br />

regression with generalized estimating equations for correlated outcomes<br />

within clinics to calculate adjusted relative risks (aRRs) and 95% confidence<br />

intervals (CIs) for the association between reproductive outcomes and use of<br />

a GC among fresh cycles, stratified by oocyte status (donor versus nondonor).<br />

RESULTS: Of 1,595,089 cycles performed from 1998 to 2012, 28,228<br />

(1.8%) used a GC, and increased from 733 (1.1%) in 1998 to 3,200 (2.4%)<br />

in 2012 (p


consistent when analyzed per cycle start and per embryo transfer. Using<br />

AMH as a sole predictor of pregnancy outcome, a sensitivity analysis demonstrated<br />

lower clinical pregnancy rates in those with AMH


the ultimate outcome being a child born alive and surviving the neonatal<br />

period. Survival rates and pregnancy lengths were calculated for white<br />

and non-white infants. Differences were determined using the Wilcoxon<br />

signed rank test.<br />

RESULTS: Compared with other months, infant survival was highest in<br />

non-whites (p


matrix scaffolds are reflective of the native tissue. These scaffolds have the<br />

potential to serve as a novel 3-dimensional platform for leiomyoma and myometrium<br />

research in the future.<br />

Supported by: National Institutes of Health Grant R<strong>21</strong>HD077479-01; National<br />

Institutes of Health Grant K12HD0501<strong>21</strong>; Northwestern University<br />

Women’s Reproductive Health (WRHR) Scholar Award; Harold Amos Medical<br />

Faculty Development Award; Robert Wood Johnson Foundation<br />

O-270 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:30 PM<br />

LEUKEMIA INHIBITORY FACTOR (LIF) GENE POLYMOR-<br />

PHISM T>G (RS929271) PREDICTS IMPLANTATION AND PREG-<br />

NANCY AFTER IVF/ICSI INDEPENDENT OF TP53 GENE<br />

POLYMORPHISM. J. A. Oliveira, a,b L. D. Vagnini, b A. Renzi, b<br />

G. R. Oliveira-Pelegrin, b C. G. Petersen, a,b A. L. Mauri, a,b<br />

F. C. Massaro, a,b F. Dieamant, a,b M. Cavagna, a,b,c R. Baruffi, a,b<br />

J. G. Franco, Jr., a,b a Center for Human Reproduction Prof. Franco Jr, Ribeirao<br />

Preto, Brazil; b Paulista Center for Diagnosis Research and Training, Ribeirao<br />

Preto, Brazil; c Women’s Health Reference Center Perola Byington Hospital,<br />

Sao Paulo, Brazil.<br />

OBJECTIVE: Leukemia inhibitory factor (LIF) plays a critical role in<br />

embryo development and implantation. The literature provides evidence<br />

that a LIF T/G gene polymorphism (rs929271) is associated with female<br />

fertility. However, studies on LIF polymorphisms are still scarce. On the<br />

other hand The TP53 gene plays a critical role in regulating blastocyst implantation<br />

function mediate by genes involved in the TP53 pathway,<br />

including the LIF gene. Studies have demonstrated that the genotype of<br />

the TP53 codon 72 polymorphism (rs1042522; encoding arginine [Arg]<br />

or proline [Pro]) is associated with the expression of LIF. This study<br />

analyzed whether the LIF T/G gene can predict pregnancy outcomes in<br />

ART and, if so, whether this is independent of the TP53 codon 72 polymorphism.<br />

DESIGN: Prospective<br />

MATERIALS AND METHODS: A total of 411 women who had undergone<br />

cycles of IVF/ICSI were recruited. DNA was extracted from peripheral<br />

blood samples that were taken from each participant, and LIF T/G TP53<br />

codon 72 Arg/Pro single nucleotide polymorphisms (SNP) were genotyped<br />

using real-time PCR. All procedures were performed under the same clinical/laboratory<br />

conditions. The cumulative results, including fresh and frozen<br />

cycles, were analyzed.<br />

RESULTS: Hardy-Weinberg genotype distributions of the entire sample<br />

indicated concordance between the observed and expected frequencies.<br />

Characteristics such as age, infertility, etiology, number of<br />

transfers and number of transferred embryos were not significantly<br />

different (P>0.05) between the groups. The distribution of SNP genotypes<br />

at codon 72 of the TP53 gene was similar among the three genotypes<br />

of LIF SNP T/G (P>0.05).The G/G LIF genotype was associated<br />

with increased implantation and ongoing pregnancy rates after IVF/<br />

ICSI. However, no correlation was observed between TP53 codon 72<br />

polymorphism genotypes and clinical outcomes after IVF/ICSI.The table<br />

below summarizes the data.<br />

CONCLUSIONS: LIF SNP T/G (rs929271) appears to be a susceptibility<br />

biomarker that is capable of predicting implantation efficiency and pregnancy<br />

outcomes after IVF/ICSI. This influence appears to be independent<br />

of TP53 codon 72 polymorphism (rs1042522) genotypes.<br />

Supported by: Merck Serono grant (GFI-2014-16). The funders had no<br />

role in study design, data collection and analysis, or preparation of the abstract<br />

NUTRITION<br />

O-271 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:15 AM<br />

MALE DIETARY TRANS FAT INTAKE IS INVERSELY ASSOCI-<br />

ATED TO FERTILIZATION RATES. M. Arvizu, a C. Tanrikut, b<br />

R. Hauser, c M. Keller, c J. E. Chavarro. c a Nutrition, Harvard T.H. Chan<br />

School of Public Health, Boston, MA; b Massachusetts General Hospital,<br />

Boston, MA; c Harvard T.H. Chan School of Public Health, Boston, MA.<br />

OBJECTIVE: Saturated and trans fat intake has been related to lower<br />

sperm counts, but it is not known whether this leads to lower reproductive<br />

success among couples undergoing infertility treatment. To address this<br />

issue, we evaluated the association between dietary fat intake and fertilization<br />

rate among couples undergoing assisted reproductive technologies<br />

(ART).<br />

DESIGN: Prospective cohort study<br />

MATERIALS AND METHODS: We followed 141 men from couples presenting<br />

to an academinc fertility center whose partners underwent ART<br />

(n¼246 cycles). Diet was estimated before treatment with a validated food<br />

frequency questionnaire and outcome data were extracted from medical records.<br />

Generalized linear mixed models with random intercepts to account<br />

for multiple cycles per woman were used to evaluate the association of fat<br />

intake with fertilization rate while adjusting for male and female age, total<br />

Fresh Cycles<br />

Cumulative(Fresh+frozen/thawed) Cycles<br />

GENOTYPE<br />

Implantation rate<br />

Ongoing<br />

pregnancy<br />

rate/transfer<br />

Ongoing pregnancy<br />

rate/patient<br />

Implantation rate<br />

Ongoing<br />

pregnancy<br />

rate/transfer<br />

Ongoing<br />

pregnancy<br />

rate/patient<br />

WOMEN’S GENOTYPES<br />

LIF T/G (RS929271) GROUPS<br />

T/T<br />

n:168<br />

16.8% a<br />

(87/519)<br />

19.6%a<br />

(44/225)<br />

26.2%a<br />

(44/168)<br />

15.9%a<br />

(100/630)<br />

18.5%a<br />

(53/287)<br />

T/G<br />

<strong>17</strong>.8%b 22.4%b 31.7%<br />

16.2%b<br />

20.2%b<br />

n:202 (107/601) (64/286) (64/202)<br />

(1<strong>21</strong>/747) (73/362)<br />

G/G<br />

30.1% a,b 38.8%a,b 46.3%a<br />

27.0%a,b<br />

36.7%a,b<br />

n:41<br />

(31/103) (19/49) (19/41)<br />

(34/126)<br />

(22/60)<br />

P<br />

0.002 a<br />

0.007a 0.02a 0.004a<br />

0.003a<br />

0.006 b 0.01b<br />

0.005b<br />

0.007b<br />

GENOTYPE Fresh Cycles Cumulative<br />

(Fresh+frozen/thawed)<br />

Cycles<br />

Implantation rate<br />

WOMEN’S GENOTYPES<br />

TP53 CODON 72 ARG/<br />

PRO (RS1042522) GROUPS<br />

Ongoing<br />

pregnancy<br />

rate/transfer<br />

Ongoing<br />

pregnancy<br />

rate/patient<br />

Implantation<br />

rate<br />

Ongoing pregnancy<br />

rate/transfer<br />

Ongoing pregnancy<br />

rate/patient<br />

31.5%a<br />

(53/168)<br />

36.1%<br />

(73/202)<br />

53.7%a<br />

(22/41)<br />

0.01a<br />

Arg/Arg<br />

n:198<br />

18.8%<br />

(112/595)<br />

22.5%<br />

(63/280)<br />

31.8%<br />

(63/198)<br />

<strong>17</strong>.9%<br />

(130/728)<br />

<strong>21</strong>.1%<br />

(74/350)<br />

37.4%<br />

(74/198)<br />

Arg/Pro<br />

n:182<br />

18.5%<br />

(100/540)<br />

23.8%<br />

(56/235)<br />

30.8%<br />

(56/182)<br />

16.8%<br />

(111/661)<br />

<strong>21</strong>.7%<br />

(65/300)<br />

35.7%<br />

(65/182)<br />

Pro/Pro<br />

n:31<br />

14.8%<br />

(13/88)<br />

<strong>17</strong>.8%<br />

(8/45)<br />

25.8%<br />

(8/31)<br />

12.3%<br />

(14/114)<br />

15.3%<br />

(9/59)<br />

29.0%<br />

(9/31)<br />

P ns ns ns ns ns ns<br />

Values within columns containing the same letter were significantly different. ns: not significant.<br />

FERTILITY & STERILITY Ò<br />

e103


calories, male body mass index, race, male and female smoking status, infertility<br />

diagnosis, protocol type and dietary patterns.<br />

RESULTS: Men’s median age was 36.9 years and median total fat intake<br />

was 32% of calories/day. Fertilization rates were lowest in couples with the<br />

highest male partner intake of trans fats (median 1.20 % calories/day, IQR<br />

1.<strong>21</strong>,1.46%). The multivariate-adjusted fertilization rates (95% Confidence<br />

Interval) in couples in increasing tertiles of men’s trans fat intake were<br />

77% (70-84%), 78% (72-84%) and 64% (55-72%) (p, linear trend ¼ 0.05).<br />

This association was stronger in cycles with conventional insemination<br />

where the corresponding adjusted fertilization rates were 83% (69-91%),<br />

70% (54-83%), and 47% (30-64%) (p, linear trend ¼ 0.005), and less pronounced<br />

in ICSI cycles (p, linear -trend¼0.11).<br />

CONCLUSIONS: Higher male-partner intake of trans fat was associated<br />

with lower fertilization rates in couples undergoing ART, particularly in conventional<br />

insemination cycles.<br />

Supported by: NIH grants R01ES009718, R01ES022955, P30ES000002,<br />

P30DK46200.<br />

Ln mice (Table 1). There were no differences in the levels of IL-6, IL-10, and<br />

interferon-gamma between the groups.<br />

CONCLUSIONS: 10-week exposure to HFD causes significant reduction<br />

in primordial follicles, compromised fertility, higher pro-inflammatory cytokine<br />

levels, and increased ovarian macrophage infiltration, independent of<br />

obesity. In addition, obesity worsens the effects of HFD alone. The negative<br />

effects of HFD on primoridal follicles may be mediated by increased ovarian<br />

inflammation. To the best of our knowledge, this is the first time that HFD<br />

was found to be detrimental to fertility and ovarian function independent<br />

of obesity in an interventional study. Further studies are needed to elucidate<br />

the mechanisms behind these findings.<br />

References:<br />

1. Norman JE. The adverse effects of obesity on reproduction. Reproduction.<br />

2010;140:343-345.<br />

Supported by: 5T32HD040135-13A National Training Program in Reproductive<br />

Medicine (MSW)<br />

O-272 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 11:30 AM<br />

HIGH-FAT DIET CAUSES COMPROMISED FERTILITY AND<br />

INCREASED PRO-INFLAMMATORY CYTOKINES INDEPEN-<br />

DENT OF OBESITY. M. E. Skaznik-Wikiel, A. J. Polotsky,<br />

J. L. McManaman. OB/GYN, University of Colorado Denver, Aurora, CO.<br />

OBJECTIVE: Female obesity is associated with ovarian dysfunction and<br />

subfertility (1). However, there are no studies to date to assess whether<br />

high-fat diet (HFD) alone (without obesity) causes reproductive dysfunction.<br />

The goal of this study was to determine if HFD impacts ovarian function,<br />

fertility, and markers of inflammation independent of obesity.<br />

DESIGN: Prospective animal study.<br />

MATERIALS AND METHODS: 5-week old mice were fed either low fat<br />

diet containing 10% fat (control group-LF-Ln) or HFD containing 60% fat.<br />

After 10 week feeding trial HFD-fed mice were divided into three groups<br />

based on body weight (BW): group 1: BW >25 g - high fat obese (HF-<br />

Ob), group 2: BW


increased with aging from 20w to 32w for both the NC (2.5 0.48 vs. 45.9 <br />

4.7, p¼0.01 and 1.4 0.18 vs 10.2 1.4, p¼ 0.02, respectively) and the HF<br />

mice converted to NC (5.0 0.75 vs. 27.9 4.7, p¼ 0.01 and 3.6 0.67 vs<br />

11.6 2.6, p¼ 0.02, respectively).<br />

CONCLUSIONS: Mouse ovarian macrophage marker gene expression<br />

increases with high fat diet, obesity and aging. The pro-inflammatory M1<br />

macrophage phenotype expression significantly increased with HF diet<br />

while there was no change in the anti-inflammatory M2 macrophage phenotype.<br />

As the mice aged, there was an increase in both overall macrophage<br />

marker expression and M1 macrophage marker expression. This increase<br />

in overall and proinflammatory macrophage infiltration in the ovary may<br />

contribute to the reproductive dysfunction seen with obesity, high fat diet<br />

and aging.<br />

O-274 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:00 PM<br />

OMEGA 3 FATTY ACID SUPPLEMENTATION DOES NOT<br />

IMPROVE RELATIVE HYPOGONADOTROPIC HYPOGONADISM<br />

IN OBESE WOMEN. Z. A. Al-Safi, a H. Liu, b J. Chosich, a<br />

M. Harris, c A. P. Bradford, a C. Robledo, a R. E. Eckel, d N. Carlson, b<br />

A. J. Polotsky. a a Obstetrics and Gynecology, University of Colorado, Aurora,<br />

CO; b Biostatistics and Informatics, Colorado School of Public Health,<br />

Aurora, CO; c Food Science and Human Nutrition, Colorado State University,<br />

Fort Collins, CO; d Medicine, University of Colorado, Aurora, CO.<br />

OBJECTIVE: Female obesity is associated with relative hypogonadotropic<br />

hypogonadism and chronic pro-inflammatory state. Dietary omega-3<br />

fatty acids (FA) exhibit strong anti-inflammatory properties. We hypothesized<br />

that dietary supplementation with omega-3 FA would result in normalization<br />

of obesity-linked hypogonadotropism via improvement in chronic<br />

inflammation.<br />

DESIGN: Prospective interventional study.<br />

MATERIALS AND METHODS: Regularly menstruating obese women<br />

(34.8 +/-1.2 year-old) underwent frequent blood sampling q 10 min for 8<br />

hours. A 75 ng/kg intravenous bolus of gonadotropin releasing hormone<br />

(physiologic dose) was given at 6 hours [1]. Testing was done in the early<br />

follicular phase BEFORE and AFTER 1 month of omega-3 FA supplementation.<br />

At the completion of baseline studies, a 4 g/day omega-3 FA supplementation<br />

was started on day 1 of the subsequent menses. Luteinizing (LH)<br />

and follicle-stimulating hormones (FSH) were assayed by immunofluorometry<br />

(DELFIA). Plasma FA composition were analyzed by gas liquid chromatography.<br />

A custom 10-cytokine array kit (Raybiotech) was used. Paired t and<br />

Wilcoxon tests were used for normally distributed and skewed data, respectively.<br />

RESULTS: 15 women (BMI 38 +/- 2 kg/m2) completed the study. While<br />

the omega-6/omega-3 ratio has been significantly reduced after the intervention,<br />

it still remained higher than 1:1 ratio that was previously shown to be<br />

beneficial for reproductive function in mice [2]. While pro-inflammatory cytokines<br />

were significantly lower after the intervention, no change was<br />

observed for LH or FSH parameters.<br />

Omega-3 FA Supplementation Improves Inflammation but Not Reproductive<br />

Hormones in Obese Women.<br />

Baseline<br />

Following 1 month<br />

of omega-3 FA<br />

supplementation<br />

LH mean serum level, IU/L 3.3 (0.3) 3.3 (0.3) 0.84<br />

FSH mean serum level, IU/L 4.4 (0.4) 4.2 (0.2) 0.51<br />

LH response to GnRH, 6.6 (1.0) 6.7 (1.2) 0.82<br />

mean serum level, IU/L<br />

FSH response to GnRH, 5.1 (0.4) 4.9 (0.3) 0.71<br />

mean serum level, IU/L<br />

Plasma<br />

8.6 (0.4) 3.9 (0.4)


M. S. Christianson, b W. Shen. a a Obstetrics and Gynecology, Johns Hopkins<br />

O-276 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong> 12:30 PM<br />

TRANSITION. T. S. Robinson, a T. Gaines, a J. Wu, a MCM8[a-b]. The primary endpoint was time to menopause, defined as R<br />

University School of Medicine, <strong>Baltimore</strong>, MD; b Johns Hopkins University<br />

VITAMIN D ATTENUATES THE ADVERSE EFFECT OF<br />

ADVANCED GLYCATION END PRODUCTS ON HUMAN GRANU-<br />

School of Medicine, Lutherville, MD.<br />

LOSA CELLS: IMPLICATIONS FOR WOMEN WITH OBJECTIVE: To evaluate the quality of life and experiences of female patients<br />

PCOS. Z. Merhi, a A. Fadiel, a E. Buyuk, b F. Naftolin, a M. Cipolla. c a Obstetrics<br />

with HIV (human immunodeficiency virus) during the menopause tran-<br />

and Gynecology, NYU School of Medicine, New York, NY; b Obstetrics<br />

& Gynecology and Women’s Health, Albert Einstein College of<br />

Medicine / Montefiore M, Bronx, NY; c Obstetrics, Gynecology and Reproductive<br />

Sciences, University of Vermont, Burlington, VT.<br />

sition.<br />

DESIGN: Interview survey of patients at a HIV-clinic.<br />

MATERIALS AND METHODS: We conducted an in-person survey of female<br />

patients at a HIV clinic over 18 months. Inclusion criteria included<br />

women 40 to 50 years old who had at least one menstrual period within<br />

OBJECTIVE: Women with PCOS commonly have elevated levels of the the last 6 months. Interviews utilized the Greene Climacteric scale, a validated<br />

menopausal questionnaire that surveys <strong>21</strong> items on a 4 point Likert<br />

pro-inflammatory advanced glycation end products (AGEs) and low levels<br />

of vitamin D (VD). AGEs are highly reactive molecules that form when lipids scale (0-3) covering 3 domains, including psychological, somatic, and vasomotor<br />

symptoms. Additionally, we queried patients on: (1) if they were in-<br />

or proteins become glycated after exposure to glucose. AGEs accumulate in<br />

the serum and ovaries of women with PCOS potentially altering ovarian forming their primary care physician of menopausal symptoms and (2) if<br />

function. We have shown that: 1) 1,25-dihydroxyvitamin D3 (VD3) supplementation<br />

increases serum levels of the ‘‘anti-inflammatory’’ soluble receptor demographic data was obtained from patient chart review including age,<br />

these symptoms were being addressed and treated by their providers. Basic<br />

for AGEs (sRAGE) in women with PCOS, and 2) VD3 induces a state of race, time of diagnosis and marital status.<br />

luteinization in human cumulus granulosa cells (CCs). We thus hypothesized RESULTS: A total of 23 women ages 40-50 years old agreed to participate<br />

that AGEs adversely affect ovarian function and induce abnormal steroidogenesis<br />

and follicular development. We further hypothesized that VD3 will viewed were African American. The average length of HIV diagnosis was<br />

in an interview. The mean age was 47.1 (SD 2.48). All of the patients inter-<br />

attenuate the AGE-induced ovarian dysfunction.<br />

13.4 years (SD 7.28). The majority of patients, 87% (n¼20), reported experiencing<br />

at least one menopausal symptom with intense frequency, having<br />

DESIGN: Basic science.<br />

MATERIALS AND METHODS: Experiment #1: 72 women who underwent<br />

IVF were enrolled. Follicular fluid (FF) was collected from the first viewed, 100% percent (n¼23), reported experiencing hot flashes to some de-<br />

extreme detrimental effects on their quality of life. Of the women inter-<br />

large follicle and was tested by ELISA for sRAGE, AGEs (pentosidine and gree, ranging from infrequent to persistent episodes. Difficulty in sleeping<br />

N ε-carboxymethyllysine [CML]), 25 hydroxyvitamin D (25 OHD), insulin, was reported by 78% (n¼18) of interviewed women, and on average was<br />

glucose, and SHBG. Experiment #2: CCs (n¼6) were cultured in media (control)<br />

human glycated albumin (HGA; 0.2-0.4 mg/mL) as a source of AGEs of women, 78% (n¼18), reported feeling tired or lacking in energy with mod-<br />

experienced with moderate frequency, 2-3 occurrences a week. A majority<br />

VD3 (50-100 nM) for 48 h after which mRNA was compared using RT- erate frequency. When asked if they told their primary care physician about<br />

PCR for LH receptor (LHR), anti-Mullerian hormone (AMH) and its receptor their menopausal symptoms, 87% (n¼20), reported doing so. Of these patients,<br />

only 20% (n¼4) received treatment for their symptoms, whereas<br />

(AMHR-II), RAGE (pro-inflammatory membranous receptor for AGEs), and<br />

VD receptor (VDR). In addition, RAGE protein was assessed by immunofluorescence.<br />

Experiment #3: KGN granulosa cell line was treated with recom-<br />

CONCLUSIONS: These findings suggest that HIV-infected women under-<br />

80% (n¼16) of these patients went without treatment.<br />

binant AMH (rAMH) with or without HGA VD3 after which going the menopause transition experience intense symptoms that severely<br />

immunofluorescence for SMAD 1/5/8 phosphorylation (AMH signaling impact their quality of life. Although the majority of women reported experiencing<br />

menopausal symptoms to their providers, many of these patients<br />

pathway) was assessed. Data were expressed as % change.<br />

RESULTS: In FF, sRAGE correlated positively with pentosidine (r¼0.24), remain untreated. An opportunity exists to educate providers for this population<br />

on menopause medicine to effectively treat this group and enhance their<br />

CML (r¼0.32), 25 OHD (r¼0.27), SHBG (r¼0.37), and negatively with insulin<br />

(r¼-0.28) and glucose (r¼-0.39) (p


12 months of amenorrhea after the end of chemotherapy. For survivors who<br />

did not undergo chemotherapy, time to menopause was defined as R 12<br />

months of amenorrhea after diagnosis. The association between SNPs and<br />

time to menopause was assessed using Cox proportional hazards models.<br />

An omnibus variable summing the total number of risk alleles that affected<br />

each participant was created for analysis.<br />

RESULTS: Our cohort had a median enrollment age of 40.5 years [range<br />

20.6 - 46] with 76.2% receiving chemotherapy. Median follow up time was<br />

1.77 years [range 0.06 - 5.8 years], and 39 (23%) of participants met criteria<br />

for amenorrhea. Time to menopause was associated with age (HR 1.2 [95%<br />

CI 1.05, 1.26]), BMI (HR 0.92 [0.85, 0.99]) and cyclophosphamide-based<br />

chemotherapy (HR 3.31 [1.<strong>17</strong> - 9.40]). No significant association with<br />

time to menopause was found for any SNP (p-value range: 0.22 - 0.88) (Table<br />

1). Restricting to white participants or testing the omnibus SNP variable did<br />

not change findings.<br />

CONCLUSIONS: Thirteen previously identified SNPs associated with<br />

natural time to menopause in genome wide association studies were not<br />

related to timing of menopause in this limited cohort of breast cancer patients.<br />

Supported by: MRSG-08-110-01-CCE, HD058799; T32 HD007203<br />

FEMALE REPRODUCTIVE ENDOCRINOLOGY<br />

P-3 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

GDF-8 DECREASES STAR EXPRESSION THROUGH ALK5-MEDI-<br />

ATED SMAD3 AND ERK1/2 SIGNALING PATHWAYS IN HUMAN<br />

GRANULOSA CELLS. L. Fang, a H. Chang, b J. Cheng, b Y. Yu, a<br />

P. C. Leung, b Y. Sun. a a Reproductive Center, The First Affiliated Hospital<br />

of Zhengzhou University, Zhengzhou, China; b Department of Obstetrics<br />

and Gynaecology, University of British Columbia, Vancouver, BC, Canada.<br />

OBJECTIVE: The aim of this study was to investigate the effects of GDF-<br />

8 on steroidogenic enzyme expression and investigate the potential mechanisms<br />

of action in human granulosa cells.<br />

DESIGN: An established immortalized human granulosa cell (SVOG) was<br />

used as the study model to examine the steroidogenic enzyme expression,<br />

phosphorylation of Smad2/3 or ERK1/2 and progesterone accumulation<br />

following exposure to GDF-8. Signaling pathway involvement was investigated<br />

with inhibitors (SB431542 and U0126) and small interfering RNAs targeting<br />

activin receptor-like kinase (ALK)4, ALK5, Smad2, Smad3 and<br />

Smad4. The concentrations of GDF-8 and progesterone were examined in<br />

the follicular fluid or conditioned medium.<br />

MATERIALS AND METHODS: RT-qPCR and Western blot were used to<br />

measure mRNA and protein levels, respectively. GDF-8 concentrations and<br />

progesterone accumulation were measured by enzyme immunoassay<br />

(ELISA).<br />

RESULTS: Treatment of GDF-8 significantly down-regulated steroidogenic<br />

acute regulatory protein (StAR) expression and decreased progesteron<br />

production. The suppressive effect of GDF-8 on StAR expression was abolished<br />

by the inhibition of TGF-b type I receptor, SB431542. In addition, the<br />

treatment of GDF-8 activated both Smad2/3 and ERK1/2signaling pathways.<br />

Furthermore, knockdown of activin receptor-like kinase (ALK)5 reversed the<br />

effects of GDF-8 on Smad2/3 phosphorylation and StAR down-regulation.<br />

The inhibition of Smad3 or ERK1/2 signalings attenuated the GDF-8-<br />

induced down-regulation of StAR and production of progesterone. Interestingly,<br />

the concentrations of GDF-8 were negatively correlated with those of<br />

progesterone in follicular fluid.<br />

CONCLUSIONS: GDF-8 down-regulates StAR expression and decreases<br />

progesterone production in human granulosa cells, most likely through<br />

ALK5-mediated Smad3 and ERK1/2 signaling pathways.<br />

Supported by: This work was Supported by an operating grant from the<br />

National Natural Science Foundation of China (31471404) to Y.P.S. and<br />

the Canadian Institutes of Health Research to P.C.K.L.<br />

P-4 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ABORTIFACIENT EFFECT OF THE AQUEOUS EXTRACT OF<br />

STRIGA HERMONTHECA IN FEMALE WISTAR<br />

RATS. I. G. Bako, a H. N. Madugu, b A. Adamu, a I. M. Maje. c a Human<br />

Physiology, Ahmadu Bello University, Zaria, Nigeria; b Obstetrics & Gynaecology,<br />

Ahmadu Bello University, Zaria, Nigeria; c Pharmacology, Faculty of<br />

Pharmaceutical Sciences, Ahmadu Bello University, Zaria, Nigeria.<br />

OBJECTIVE: Large percentage of the world population still rely on herbs<br />

for their primary reproductive health care needs. Striga Hermontheca which<br />

is readily available, is used locally among the rural dwellers for abortion or<br />

induction of labour.<br />

DESIGN: The abortifacient effect of Striga Hermontheca was investigated<br />

by evaluating the serum progesterone level in female pregnant wistar rats after<br />

treatment with the aqueous extract of Striga Hermontheca in varying<br />

doses for three days. The toxicity of the herbs was accessed to ascertain its<br />

safety for consumption.<br />

MATERIALS AND METHODS: The study was compiled with ethical<br />

committee guidelines of Ahmadu Bello University Teaching Hospital, Zaria<br />

with registration number ABUTH/PGO/COMM/0138. The sample of Striga<br />

Hermontheca was collected in Katsina in August 2014 and it was identified<br />

in the Herbarium section of the Department of Biological sciences, with<br />

voucher number 1845. The herbs was extracted using maceration method in<br />

the Department of Pharmacognosy and drug development. Twenty female wistar<br />

rats were randomly grouped into four groups of five rats each (n¼5). Female<br />

wistar rats in their pro-estrous cycle were mated for twenty four (24) hours and<br />

presence of sperm cells in vaginal smear confirmed copulation and day one of<br />

pregnancy. Group I received (control) 10 mg/kg normal saline, Group II<br />

received 200 mg/kg Misoprostol, Groups III and IV received 250 mg/kg and<br />

500 mg/kg of the aqueous extract of Striga Hermontheca respectively. Normal<br />

saline, Misoprostol and aqueous extract of Striga Hermontheca were administered<br />

orally from day five of conception. All data are expressed as Mean SEM<br />

and analyzed using t - test Student’s test, SPSS package 20.0 and post hoc test<br />

for multiple comparison. The (P14 mm in diameter. Categorical variables were assessed by chi-square or<br />

FERTILITY & STERILITY Ò<br />

e107


Fisher’s exact test for small frequencies, with significance at a p-value of<br />

14mm<br />

(p5mg>2.5mg). Because the small number of cases in Group A, statistical<br />

comparison was only carried out between Group B and Group C. Both the<br />

biochemical (15.5% vs. 11.6%) and the clinical (12.8% vs. 9.9%) PRs<br />

were statistically significant increase in Group B when compared to Group<br />

C. The multiple PR and the miscarriage rate were similar between groups.<br />

CONCLUSIONS: Letrozole has been shown to be an efficient and effective<br />

agent in inducing both ovulation and superovulation. There has been<br />

debate as to the optimal starting dose of letrozole. The 5mg-daily yields<br />

higher clinical PRs (p


for treatment management. We embarked on a multi-center initiative to identify<br />

possible subclinical, molecular drivers of idiopathic infertility and treatment<br />

failure.<br />

DESIGN: Multi-center retrospective cohort study.<br />

MATERIALS AND METHODS: Genomic DNA was extracted from<br />

whole blood collected at geographically distinct infertility centers from<br />

women with idiopathic infertility, who had attempted but not achieved a<br />

pregnancy that had progressed beyond week 12 with either timed intercourse<br />

or non-IVF fertility treatments. Medical history and treatment-level clinical<br />

metrics were linked to whole-genome sequence data (30x mean coverage)<br />

generated using Illumina HiSeq sequencers. Novel and rare genetic variants<br />

predicted to be deleterious were identified and filtered using a specially designed<br />

bioinformatics pipeline, incorporating a novel fertility-centric knowledgebase<br />

of >3,000 genetic loci categorized by their role in particular<br />

reproductive processes. Principal component analysis and sequence kernel<br />

association tests were used to analyze the variant signatures of patients in<br />

relation to their infertility treatment histories.<br />

RESULTS: We detected associations between the genetic signatures of patients<br />

and different features of their infertility treatment journeys. For<br />

example, mutations in genes regulating egg quality and early embryo development<br />

were among those most strongly associated with increases in the<br />

number of non-IVF and IVF treatment cycles undertaken before achieving<br />

live birth (LB). We also defined a subset of patients who, after several unsuccessful<br />

non-IVF cycles, achieved live birth after just one IVF treatment.<br />

Analysis revealed that these patients were genetically distinguishable from<br />

those requiring multiple rounds of IVF to achieve LB, as well those experiencing<br />

multiple pregnancy loss.<br />

CONCLUSIONS: Genetic signatures provide novel, subclinical insights<br />

into fertility potential and treatment response that could be a powerful counseling<br />

tool for helping patients progress more quickly to treatment strategies<br />

that will result in a LB outcome.<br />

Supported by: Celmatix Inc<br />

P-9 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

RELATIONSHIP BETWEEN EXPRESSION OF SIRT1 AND SIRT6<br />

GENES AND THE RESPONSE TO OVARIAN<br />

STIMULATION. A. Palumbo, a,b D. Rotoli, c,d R. Gonzalez-Fernandez, d<br />

J. Hernandez, a J. Avila. d a Reproductive Endocrinology, Centro de Asistencia<br />

a la Reproduccion Humana de Canarias, La Laguna, Spain; b Department of<br />

Obstetrics and Gynecology, New York University, New York, NY; c Istituto di<br />

Endocrinologia ed Oncologia Sperimentale (IEOS), National Research<br />

Council (CNR), Naples, Italy; d UDI Bioquimica y Biologia Molecular, Universidad<br />

de La Laguna, La Laguna, Spain.<br />

OBJECTIVE: Sirtuins (SIRTs) are a family of proteins with NAD+-dependent<br />

deacetylase activity with an important function in the response to<br />

different types of metabolic stress or genome injury. SIRT1 and SIRT6<br />

play an important role in basic cellular functions such as genomic stability,<br />

metabolic homeostasis, survival in stress conditions and cellular differentiation.<br />

The objective of this work was to study the expression of Sirt1 and Sirt6<br />

genes in granulosa-lutein cells (GL cells) from IVF patients and its relationship<br />

with the infertility diagnosis and ovulation induction.<br />

DESIGN: Analysis of Sirt1 and Sirt6 gene expression levels by qRT-PCR<br />

in GL cells of young egg donors and IVF patients.<br />

MATERIALS AND METHODS: 82 women were studied, including: <strong>17</strong><br />

egg donors (ED); 15 women >40 years old (yo) with tubal or male factor<br />

and no ovarian pathology (>40); 16 poor responders (PR); 18 endometriosis<br />

cases (EM) and 16 polycystic ovarian syndrome patients (PCOS). After ultrasound<br />

guided egg retrieval, GL cells were isolated from pooled follicular<br />

fluids from each woman using a percoll gradient and anti-CD45 immunobeads<br />

to eliminate WBCs; viability was assessed by trypan blue. After total<br />

RNA isolation and cDNA synthesis, Sirt1 and Sirt6 gene expression levels<br />

were measured by RT-PCR. Expression data were analyzed using the DCt<br />

method and statistical analysis was performed with SPSS using Pearson’s<br />

Correlation coefficient and Student t-test.<br />

RESULTS: Expression levels of Sirt1 in GL cells of patients >40 yo were<br />

higher than in ED and EM, PR and PCOS patients (P40 yo (P40<br />

yo compared to young ED and other diagnostic groups may support a role of<br />

Sirt1 gene in aging. 2) The increased expression level of Sirt6 in EM and PR<br />

may be an indicator of a much higher level of DNA damage and oxidative<br />

stress in both groups. 3) The positive correlation of Sirt1 and Sirt6 with gonadotropins<br />

doses observed only in EM suggests a role of DNA damage and<br />

oxidative stress in the lower response to FSH stimulation observed in EM patients.<br />

Supported by: FIS PI12/00729, Instituto Salud Carlos III, Spain<br />

P-10 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE EFFECT OF ORAL TESTOSTERONE UNDECANOATE THER-<br />

APY ON CONTROLLED OVARIAN STIMULATION AND IVF<br />

OUTCOME IN POOR RESPONDERS. G. Mskhalaya, a E. Eltsova, a<br />

V. Zaletova, b E. Lubimkina, b S. Kalinchenko. c a Andrology and Endocrinology,<br />

Center for Reproductive Medicine MAMA, Moscow, Russian Federation;<br />

b Gynecology, Center for Reproductive Medicine MAMA, Moscow,<br />

Russian Federation;<br />

c Endocrinology, People’s Friendship University of<br />

Russia, Moscow, Russian Federation.<br />

OBJECTIVE: To evaluate the efficacy of treatment with oral form of<br />

testosterone undecanoate (TU) before controlled ovarian stimulation<br />

(COS) in poor ovarian responders undergoing in vitro fertilization (IVF)/intracytoplasmic<br />

sperm injection (ICSI).<br />

DESIGN: Prospective controlled trial<br />

MATERIALS AND METHODS: 89 women with poor ovarian response,<br />

defined based on ESHRE consensus/the Bologna criteria of low ovarian<br />

response were included. Patients were randomly divided into 2 groups: TU<br />

treatment group (44 women) or control group (45 women). For TU group<br />

Testosterone undecanoate oral form 40 mg was given daily for at least 40<br />

days (mean - 48 days) preceding COS for IVF. Primary outcome measures<br />

were clinical pregnancy and live birth rates. Statistical research was made using<br />

a software package statistics (StatSoft Inc. U.S., version 12). Quantitative<br />

data is presented as median and quartile range. When comparing the quantitative<br />

data of two independent groups Mann-Whitney U test and Fisher exact<br />

two-tailed test were used. Values were considered statistically significant if p<br />


symptoms were assessed up to 4 times per cycle via questionnaires. Linear<br />

mixed models evaluated the association between serum vitamin A, C, and<br />

E concentrations and individual and grouped PMS symptoms scores and<br />

severity, and generalized linear models were used to evaluate associations<br />

with classification of PMS, adjusted for age, body mass index, and total energy<br />

intake.<br />

RESULTS: Increased levels of serum antioxidant vitamins were not significantly<br />

associated with decreased prevalence or severity of PMS symptoms<br />

either individually or by category (e.g. vitamin C (mg/dL): psychological<br />

symptoms, beta: 0.13, 95% CI: -0.95,1.<strong>21</strong>; hydration and craving, beta:<br />

-0.11, 95% CI: -0.88,0.67; physical pain, beta: 0.05, 95% CI: -0.73,0.83).<br />

Similarly, increased levels of antioxidant vitamin concentration did not<br />

decrease the chances of moderate or severe PMS classification (e.g. vitamin<br />

C: 5+ moderate/severe symptoms, OR: 1.07, 95% CI: 0.68, 1.69). We also<br />

observed no significant associations between vitamins A and E and PMS<br />

symptoms.<br />

CONCLUSIONS: These data do not support the hypothesis that antioxidant<br />

vitamin concentrations are associated with presence or severity of<br />

PMS symptoms. Therefore, information indicating dietary changes to control<br />

PMS symptoms may be unjustified.<br />

Supported by: Intramural Research Program, DIPHR, NICHD, NIH.<br />

number of metaphase II oocytes per oocyte pick-up. Cancellation rate was<br />

similar in all groups (p¼0.787); however, the probability of cancelling a cycle<br />

was a little bit higher with the corifollitropin alfa (OR¼1.463; 95%IC<br />

0.484-4.424).<br />

CONCLUSIONS: Stimulation with a long-acting gonadotropin yields<br />

similar results in a stimulation protocol in terms of metaphase II oocytes.<br />

The observed differences in the endocrine profile of follicular fluid could<br />

be explained by the structure of the corifollitropin alfa, which induces high<br />

levels of FSH, resulting in a higher concentration of estradiol. In serum,<br />

the higher estradiol concentrations observed in the HP-hMG group might<br />

be explained by the presence pf LH. The absence of significant differences<br />

in the rate of apoptosis confirms the suitability of this treatment in assisted<br />

reproduction cycle<br />

P-13 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

WITHDRAWN<br />

P-12 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

IS THERE ANY DIFFERENCES IN SERUM AND FOLLICULAR<br />

FLUID ENDOCRINE PROFILE AS WELL AS APOPTOSIS RATE<br />

BETWEEN A LONG-ACTING GONADROTOPIN AND HP-<br />

HMG? M. Cruz, a A. Pacheco, a D. Collado, a M. Munoz, b P. Alama, c<br />

A. Requena. a a IVI Madrid, Madrid, Spain; b IVI Alicante, Alicante, Spain;<br />

c IVI Valencia, Valencia, Spain.<br />

OBJECTIVE: To determine the endocrine profile in follicular fluid and the<br />

effectiveness in terms of metaphase II oocytes of a long-term gonadotropin<br />

versus a conventional dose daily in women participating in an oocyte donation<br />

program.<br />

DESIGN: Multicentre prospective randomized study<br />

MATERIALS AND METHODS: 120 oocyte donors have been included.<br />

Subjects were assigned to receive 100 mg of corifollitropin alfa which was<br />

potentially followed by daily administration of recombinant FSH (rFSH)<br />

from day 8 if instructed by the researcher (n¼40) or daily doses of 150 UI<br />

rFSH (n¼40), or 225 UI HP-hMG (n¼40). Daily doses of GnRH antagonist<br />

(0.25 mg/day) were started on day 5 of stimulation in all groups and a single<br />

dose of GnRH agonist (0.2 mg) was administered for triggering oocyte final<br />

maturation. Additionally, it was determined the apoptosis rate by flux cytometry<br />

in cumulus cells as well as the hormonal profile (estradiol and progesterone)<br />

in follicular fluid.An ANOVA test were applied to detect statistical<br />

differences; p


P-14 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

GENETIC INFLUENCES ON ENDOCRINE SIGNALING IN<br />

WOMEN. S. L. Bristow, a R. Shraga, a N. Kumar, a S. Yarnall, a<br />

S. Rodriguez, a A. Bisignano, a S. Munne, b R. Allen, c S. H. Chen. d a Recombine,<br />

New York, NY; b Reprogenetics, Livingston, NJ; c Santa Barbara Fertility<br />

Center, Santa Barbara, CA; d IRMS at Saint Barnabas, Livingston, NJ.<br />

OBJECTIVE: The hypothalamic-pituitary-ovarian (HPO) axis is a critical<br />

signaling pathway in female reproduction, including reproductive hormones<br />

such as gonadotropin releasing hormone (GnRH), follicle stimulating hormone<br />

(FSH), luteinizing hormone (LH), estradiol, and progesterone. Previous<br />

studies have shown that endocrine levels in women are affected by<br />

single nucleotide polymorphisms (SNPs). This study aimed to identify<br />

SNPs impacting the HPO by exploring links between SNPs and baseline<br />

serum hormone levels in women referred from fertility centers.<br />

DESIGN: Retrospective<br />

MATERIALS AND METHODS: Clinical and genetic data was collected<br />

for 143 women less than 40 years of age. Day 2/3 serum FSH, LH, estradiol,<br />

and progesterone levels and AMH levels were collected by chart review. Genetic<br />

polymorphisms within FSH receptor (FSHR), FSH beta polypeptide<br />

(FSHB), anti-Mullerian hormone (AMH), and AMH receptor type II<br />

(AMHR2) were measured using Illumina’s Infinium HD Genotyping assay.<br />

Welch’s t-test was used to test 150 associations between hormone levels<br />

and SNPs. A p-value of pT<br />

(rs10835638; p¼0.034). Finally, LH levels were associated with two AMH variants:<br />

c.-649C>T (rs4807<strong>21</strong>6; p¼0.019) and p.S49I (rs10407022; p¼0.03).<br />

CONCLUSIONS: This study has demonstrated the impact of SNPs on the<br />

HPO signaling pathway, particularly FSHR, FSHB, and AMH. The link between<br />

FSH levels and FSHR SNPs and between progesterone levels and<br />

FSHB SNPs have been shown before, demonstrating our ability to validate previous<br />

findings. We also identified a novel interaction between LH levels and<br />

AMH SNPs, indicating a potential interaction between AMH and LH that has<br />

yet to be characterized. While individual interactions are significant, the combination<br />

of these and other SNPs have not been studied. This endocrine signaling<br />

pathway is highly interconnected. Thus, multiple SNPs inherited together may<br />

influence signaling efficiency in more complex ways than previously thought.<br />

Further and larger studies designed to investigate the possibility of multifactorial<br />

influences on endocrine signaling in female reproduction will lead to a deeper<br />

understanding of the clinical implications of hormone levels on fertility.<br />

P-15 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

WHOLE EXOME SEQUENCING IN MONOZYGOTIC TWINS<br />

DISCORDANT FOR PREMATURE OVARIAN<br />

INSUFFICIENCY. N. Banks, a A. Martinez, b L. Brown, c J. Hughes, c<br />

A. DeCherney, a D. Page, c S. J. Silber, d M. Muenke. b a National Institute<br />

of Child Health and Human Development, National Institutes of Health, Bethesda,<br />

MD; b National Human Genome Research Institute, National Institutes<br />

of Health, Bethesda, MD; c Whitehead Institute, Howard Hughes<br />

Medical Institute, Cambridge, MA; d St. Luke’s Hospital, St. Louis, MO.<br />

OBJECTIVE: Premature Ovarian Insufficiency (POI) affects 1% of<br />

women under the age of 40. Evidence from animal models, the high prevalence<br />

of familial POI (15% of total POI), and the identification of several genetic<br />

causes, all support the hypothesis that POI is largely a genetic disease.<br />

Monozygotic twins (MZ) discordant for the POI phenotype allow for study of<br />

somatic mutations in an otherwise identical genetic background. Novel causative<br />

de novo somatic mutations may be detected by analyzing exome<br />

sequencing in discordant MZ twin pairs.<br />

DESIGN: Whole Exome Sequencing (WES) analysis<br />

MATERIALS AND METHODS: Seven monozygotic twin pairs with one<br />

sibling with POI were consented for WES to IRB approved protocol 11-HG-<br />

0093. Exome sequencing was carried out by the National Intramural<br />

Sequencing Center (NISC, NIH/NHGRI). Capture utilized the SureSelect<br />

Human All-Exon System. Flow cell preparation and 76-bp paired end read<br />

sequencing were performed as per the GAIIx Sequencer protocol. Exome<br />

analysis was conducted using the DNA-Seq module of Golden HelixÒ<br />

SVS 8.3.1 software. Data were filtered for exonic or splicing variants, read<br />

depth greater than 10 and quality score greater than 20.<br />

RESULTS: Each twin pair contained two to 23 variants per pair, with a total<br />

of 81 variants identified. Six variants were in pseudogenes. Twenty-nine<br />

variants were found in known highly polymorphic genes. Forty-six remaining<br />

variants require further analysis and confirmation via Sanger sequencing.<br />

CONCLUSIONS: Based upon our study design and inclusive filtering<br />

strategy, we anticipate a high false positive rate for gene variants detected<br />

with WES. Further analysis and variant confirmation are required. WES of<br />

seven MZ twin pairs discordant for POI offers a unique opportunity to<br />

analyze the role of somatic mosaicism in the pathogenicity of POI and the<br />

potential to identify novel genetic causes.<br />

Supported by: NICHD, NHGRI Intramural Research Programs<br />

P-16 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

RESISTIN DECREASES THE EXPRESSION OF ENDOMETRIAL<br />

RECEPTIVITY RELATED FACTORS THROUGH BINDING TO<br />

TOLL-LIKE RECEPTOR 4 IN ENDOMETRIAL EPITHELIAL<br />

CELLS. J. Yang, a Y. Lu, b Q. Lyu, a Y. Kuang. a a The Ninth Hospital,<br />

Shanghai Jiao Tong University School of Medicine, Shanghai, China; b Department<br />

of Obstetrics and Gynecology, Women’s Hospital, Zhejiang University<br />

School of Medicine, Shanghai, China.<br />

OBJECTIVE: To study the impact of resistin, an adipose-derived peptide<br />

hormone, on the expression of endometrial receptivity related factors in human<br />

endometrial epithelial cells (HEECs) in vitro.<br />

DESIGN: Cross-sectional study and in vitro experiment.<br />

MATERIALS AND METHODS: We detected serum resistin, an adiposederived<br />

peptide hormone, in a group of 93 clinically well-defined infertile patients<br />

with pathology of the fallopian tubes: 42 patients with BMI > 24 and<br />

51 patients with 18.5%BMI %24. The HEECs were isolated from hysterectomy<br />

specimens of 15 premenopausal women undergoing hyerterectomy for<br />

benign reasons, treated with both estrogen and progesterone in vitro and incubated<br />

with or without resistin. Resistin and leukocyte inhibitor factor(LIF) in<br />

serum and supernatant were measured with enzyme-linked immune sorbent<br />

assay (ELISA), and the expression of integrin b3, claudin-4 and dickkopf-<br />

1(DKK-1) were detected with quantitative real-time PCR and western-blotting.<br />

The expression of toll-like receptor 2 (TLR2), toll-like receptor 4<br />

(TLR4) and adenylate cyclase associated protein 1(CAP1) were suppressed<br />

by using specific siRNA transfection.<br />

RESULTS: The clinical data demonstrated that serum elevated resistin<br />

was associated with high BMI in patients who always have lower fertility.<br />

The estrogen combined with progesterone induced expression of endometrial<br />

receptivity related factors including LIF, integrin b3, claudin-4 and DKK-1<br />

could be inhibited by resistin. And TLR4 knockdown also inhibited the<br />

expression of endometrial receptivity related factors, whereas suppression<br />

of TLR2 or CAP1 had no impact on the expression of endometrial receptivity<br />

related factors in HEECs.<br />

CONCLUSIONS: Resistin might bind to TLR4 to perform its inhibition<br />

effect on the expression of endometrial receptivity related factors including<br />

LIF, integrin b3, claudin-4 and DKK-1 in HEECs.<br />

P-<strong>17</strong> Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

PREDICTING PREGNANCY AMONG WOMEN WITH NO KNOWN<br />

FERTILITY PROBLEMS. K. A. Hahn, a L. A. Wise, b E. E. Hatch, b<br />

K. J. Rothman. c a Epidemiology, Boston University School of Public Health,<br />

Boston, MA; b Boston University School of Public Health, Boston, MA; c RTI<br />

International, Newton, MA.<br />

OBJECTIVE: To create a statistical model for women with no known<br />

fertility problems that provides individual predictions of the probability of<br />

conception based on the women’s characteristics and habits.<br />

DESIGN: Prospective cohort study of women trying to conceive from the<br />

U.S., Canada, and Denmark.<br />

MATERIALS AND METHODS: We pooled participants from two prospective<br />

Internet-based cohort studies of pregnancy planners, including<br />

835 women from the Pregnancy Study Online (PRESTO) and 3785 women<br />

from Snart Gravid and Snart Foraeldre (SG/SF). Eligible women had been<br />

trying to conceive for less than 3 months and had no history of infertility.<br />

We included only those terms in a logistic regression model that made a substantial<br />

difference to goodness of fit, assessed using the likelihood ratio test<br />

p


women aged 25 years who had intercourse 4 or more times a week, were<br />

timing their pregnancy attempts, were non-smokers without a history of vaginitis,<br />

had a cycle length of 28 days, and had normal BMI and were not using<br />

hormonal birth control as their last method of conception to about 8% for nulligravid<br />

women aged 35 years who have intercourse


with accompanying histone modification, we examined C/EBPb binding to<br />

the promoter region of Cyp11a1. The binding of C/EBPb which was analyzed<br />

by ChIP assay was significantly higher at 12 h than 0 h.<br />

CONCLUSIONS: Changes of histone modification status and chromatin<br />

structure in the Cyp11a1 promoter region in addition to DNA hypomethylation<br />

status of the promoter are closely associated with the rapid increase of Cyp11a1<br />

mRNA expression in GCs undergoing luteinization during ovulation. In addition,<br />

it’s suggested that accompanied by binding of C/EBPb to chromatin structural<br />

change, the change of Cyp11a1 mRNA expression is controlled.<br />

P-22 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

A RANDOMIZED TRIAL OF WEB-BASED FERTILITY-TRACKING<br />

SOFTWARE AND FECUNDABILITY. L. A. Wise, a E. E. Hatch, b<br />

J. Stanford, c C. J. McKinnon, a A. Wesselink, b K. J. Rothman. d a Department<br />

of Epidemiology, Boston University School of Public Health, Boston, MA;<br />

b Boston University School of Public Health, Boston, MA; c University of<br />

Utah, Salt Lake City, UT; d RTI International, Newton, MA.<br />

P-<strong>21</strong> Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

CHANGES IN HISTONE MODIFICATION, DNA METHYLATION<br />

AND C/EBPB BINDING OF THE CYP11A1 PROMOTER REGION<br />

IN RAT GRANULOSA CELLS UNDERGOING LUTEINIZATION<br />

DURING OVULATION. M. Okada, a H. Asada, b H. Tamura, c<br />

N. Sugino. a a Yanaguchi University Graduate School of Medicine, Yamaguchi,<br />

Japan; b Department of Obstetrics and Gynecology, Yamaguchi, Ube,<br />

Japan; c Yamaguchi University Graduate School of Mrdicine, Ube, Japan.<br />

OBJECTIVE: The ovulatory LH surge rapidly alters the expression of steroidogenesis-related<br />

genes such as StAR, Cyp19a1 and Cyp11a1 in granulosa<br />

cells (GCs). Cyp11a1 codes P450scc and plays an important role in<br />

production of progesterone. The ovulatory LH surge induces rapid up-regulation<br />

of Cyp11a1 in granulosa cells (GCs) undergoing luteinization during<br />

ovulation. Recent evidence has shown that epigenetic mechanisms such as<br />

histone modification and DNA methylation are essentially involved in the<br />

regulation of gene expression.<br />

DESIGN: In this study, we investigated whether epigenetic mechanisms<br />

including histone modifications and DNA methylation are involved in the<br />

rapid change of Cyp11a1 expression after LH surge and also investigated<br />

whether transcription factor is associated with the change.<br />

MATERIALS AND METHODS: <strong>21</strong>-day-old immature rats were injected<br />

with eCG followed by hCG injection 48 h later. The ovaries were removed<br />

and luteinized GCs were collected before hCG (0 h), and 4 h, 8 h, and 12<br />

h after hCG injection.<br />

RESULTS: 1) In RT-PCR analysis, Cyp11a1 mRNA levels rapidly increased<br />

after hCG injection, reached the peak at 4 h, and remained high level until 12 h.<br />

2) Histone modification status in the Cyp11a1 promoter regionwas analyzed by<br />

chromatin immunoprecipitation (ChIP) assay. The level of trimethylation of<br />

histone-H3 lysine-4, which is an active chromatin marker, was increased and<br />

significantly higher at 12 h than 0 h. The level of trimethylation of histone-<br />

H3 lysine-9 and trimethylation of histone-H3 lysine-27, which is a suppressive<br />

chromatin marker, was decreased and significantly lower at 4 h and 12 h than<br />

0 h. 3) DNA methylation status was analyzed from -1427 bp to -73 bp around<br />

Cyp11a1 promoter region by sodium bisulfite sequencing. 5 CpGs were demethylated<br />

while the other 3 CpGs, which locate at relatively distal region, were<br />

methylated. This DNA methylation profile did not change during luteinization<br />

induced by hCG injection. 4) Chromatin accessibility assay showed that the<br />

chromatin condensation of the Cyp11a1 promoter region decreased after<br />

hCG injection. 5) Since we previously found C/EBPb, which is transcription<br />

factor, is involved in the expression of other steroidgenesis-associated genes<br />

OBJECTIVE: To evaluate whether randomization to use of web-based<br />

fertility-tracking software is associated with improved fecundability.<br />

DESIGN: Randomized Trial.<br />

MATERIALS AND METHODS: The Boston University Pregnancy Online<br />

Study (PRESTO) is a prospective cohort study of female pregnancy planners<br />

aged <strong>21</strong>-45 years in the U.S. and Canada. Women were recruited using<br />

internet-based advertisements promoting an incentive to win a premium<br />

membership to FertilityFriend.com (FF), a web-based software application<br />

that records data on menstrual cycles and fertility signs. At baseline, participants<br />

completed a health history questionnaire and were randomized with<br />

50% probability to FF. Women were followed every 8 weeks for up to 12<br />

months or until reported pregnancy. The analysis was restricted to women<br />

who were eligible for FF randomization (i.e., non-users of FF) and who<br />

had been attempting to conceive for %6 cycles at study entry (N¼1,238). Using<br />

an intent-to-treat analysis, we estimated the proportion of women who<br />

became pregnant over follow-up and used proportional probabilities regression<br />

to estimate fecundability ratios (FR) and 95% confidence intervals (CI).<br />

RESULTS: Baseline characteristics were evenly distributed between the<br />

two randomization groups. Median follow-up among women randomized<br />

to FF was 6.6 months (interquartile range: 3.5-9.4 months) and 6.6 months<br />

among non-randomized women (interquartile range: 3.4-10.3 months).<br />

Among the 629 women who were randomized to FF, 392 pregnancies<br />

(62.3%) were reported during follow-up. Among the 609 women who were<br />

not randomized to FF, 379 (62.2%) pregnancies were reported. The overall<br />

FR comparing randomized vs. not randomized to FF was 0.95 (95% CI:<br />

0.84-1.08). Among women who had been trying for 0-2, 3-4, and 5-6 cycles<br />

at study entry, proportions conceiving among randomized and non-randomized<br />

women were 66.4% vs. 69%, 58.7% vs. 57.3%, and 45.6% vs. 30.1%,<br />

respectively. FRs for the association between randomization to FF and fecundability<br />

were 0.90 (CI: 0.78-1.04), 0.97 (CI: 0.72-1.30), and 1.80 (CI:<br />

1.09-2.99), respectively. Of those randomized to FF, 53.7% actually used<br />

the software. The prevalence of FF use was similar across strata of attempt<br />

time at entry (55.2%, 47.9% and 54.4%, respectively), but there were differences<br />

in time to first use, number/type of features used, and intensity of use.<br />

CONCLUSIONS: Overall, there was little evidence that randomization to<br />

the FF menstrual cycle charting software program influenced fecundability<br />

among pregnancy planners participating in an incentive-based internet study.<br />

However, among those who had already been trying to conceive for 5-6 cycles<br />

at enrollment, assignment to FF was associated with faster conception.<br />

Supported by: This research was Supported by NICHD (R<strong>21</strong>-HD072326).<br />

P-23 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ARE PLACENTA SPECIFIC PROTEIN 1 LEVELS ASSOCIATED<br />

WITH RECURRENT PREGNANCY LOSS OR IN VITRO FERTIL-<br />

IZATION FAILURE. N. Yilmaz, a H. Timur, a S. Yilmaz, a<br />

A. S. Erdinc, a S. Erkilinc, b H. A. Inal. c a Reproductive Endocrinology<br />

Department, ZTB, Dr Zekai Tahir Burak Women’s Health Research and Education<br />

Hospital, Ankara, Turkey; b Dr Zekai Tahir Burak Education and<br />

Research Hospit, Ankara, Turkey; c Konya Education and Research Hospital,<br />

Konya, Turkey.<br />

OBJECTIVE: To investigate the relationship between the placenta specific<br />

protein-1( PLAC 1) levels associated and recurrent pregnancy loss or in vitro<br />

fertilization failure.<br />

FERTILITY & STERILITY Ò<br />

e113


DESIGN: A total of 28 patients with recurrent pregnancy loss (group 1), 30<br />

unexplained infertile patients with IVF failure ( group 2), 29 fertil patients (<br />

group 3) were included in this cross- sectional study, from January 2013 to<br />

June 2014.<br />

MATERIALS AND METHODS: Recruited patients were younger than 40<br />

years and had no systemic disease. Antecubital venous samples were obtained<br />

from each woman in the morning. Serum samples were separated<br />

and Placenta Specific Protein-1 levels were determined with Human Placenta<br />

Specific Protein-1 (Plac-1) (Cusabio) ELISA KIT method.<br />

RESULTS: There was no significant difference in terms of age and BMI<br />

(p¼ 0.93; 0.20). Serum PLAC 1 levels were significantly higher in group1<br />

when compared to group 3 and 2 (Median of the serum PLAC1 levels:<br />

10.5 vs 1; 10.5 vs 4 ng/ml, respectively) (p


69 minor surgeries and 130 IVF retrievals per year. Average practice size was 5<br />

(range 1 - 50), with an average of 480 fresh IVF cycles (per program) per year.<br />

28% worked with a fellow. 60% were salaried and 40% were equity partners.<br />

Compensation was higher than listed by the American Medical Group Association<br />

and was greatly skewed. Greater than 85% had a positive morale and view of<br />

the future, 92% would choose REI as a career again and 79% would recommend<br />

this career to their children. The most satisfying areas of reproductive medicine<br />

were; patient interactions, intellectual stimulation, interactions with colleagues,<br />

and work schedule. The least satisfying areas were work schedule and financial<br />

compensation. Training was felt to be too focused on female infertility and basic<br />

research with insufficient training on embryology, genetics, male infertility and<br />

clinical research. In the next 5 years, 49% of respondents believed practice size<br />

would increase, and 53% responded someone in their practice would retire. The<br />

average age of planned retirement was 68. 55% suggested that the need for specialists<br />

would stay the same (20% decrease, 18% increase). A total of 58% felt we<br />

were training the right amount of fellows (36% felt we were training a surplus).<br />

Compared to academia, those in private practice reported: higher compensation,<br />

less major surgery, more IVF, less endocrinology and less research. Men worked<br />

more hours, conducted more surgery (and IVF cycles), had higher compensation<br />

and planned to retire later than women.<br />

CONCLUSIONS: Our subspecialty has an extremely high morale. We are<br />

a middle aged subspecialty with disparate compensation and a focused practice.<br />

Some sense a need for a change in our training and most anticipate only<br />

mild growth in our field.<br />

Supported by: SREI<br />

P-27 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

SHOULD ICSI BE RECOMMENDED ROUTINELY IN PATIENTS<br />

WITH SINGLE OOCYTE RETRIEVED? L. Ma, a L. Cai, b J. Dong, b<br />

M. Xia, c J. Liu, b R. Chian. d a Reproductive Medicine, The Center of Reproductive<br />

Medicine, Nanjing, China; b The First Affiliated Hospital of Nanjing<br />

Medical University, Nanjing, China; c Laboratory of Reproductive Medicine,<br />

Nanjing, China; d McGill University, Nanjing, China.<br />

OBJECTIVE: To compare clinical outcomes following in vitro fertilization<br />

(IVF) and intracytoplamisc sperm injection (ICSI) when only a single<br />

oocyte retrieved with non-male factor.<br />

DESIGN: Retrospective study was performed in this study and included<br />

2,133 patients with 3,310 completed treatment cycles from January 2007<br />

to June 2014, who underwent either stimulated or spontaneous cycles that resulted<br />

in only a single oocyte was retrieved.<br />

MATERIALS AND METHODS: Fertilization, clinical pregnancy, miscarriage<br />

and live birth rates per embryo transfer were compared with two insemination<br />

methods and further analyzed whether the clinical outcomes were<br />

affected by the patient’s diagnosis with primary infertility, secondary infertility<br />

and age (38 years). Statistical analyses were performed using SPSS,<br />

version 13.0 (SPSS, Inc), and the clinical outcomes between groups were<br />

compared using the X2 test. The differences in means of demographic data<br />

were calculated by t test, P38 years-old groups.<br />

CONCLUSIONS: There were similar clinical outcomes with IVF or ICSI<br />

as insemination method when the patients with only a single oocyte retrieved.<br />

The choice for insemination method should be based primarily on sperm<br />

quality. Therefore, the low number of eggs retrieved is not an indication<br />

for performance of ICSI.<br />

P-28 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

EXTERNAL VALIDATION AND CALIBRATION OF IVFPREDICT -<br />

CONFIRMATION THAT PROGNOSTIC STRATIFICATION PRIOR<br />

TO RCT ENTRY IS FEASIBLE. S. M. Nelson a A. D. Smith. b a School<br />

of Medicine, University of Glasgow, Glasgow, United Kingdom; b MRC integrative<br />

Epidemiology Unit, University of Bristol, Glasgow, United Kingdom.<br />

OBJECTIVE: IVFpredict a model for prediction of live birth after in vitro<br />

fertilization has been developed and externally validated in UK population<br />

data. Whether IVFpredict is effective in predicting the outcome of multicenter<br />

international randomized controlled trials (RCTs) or whether IVFpredict<br />

could facilitate stratification of patient prognosis prior to RCT entry is<br />

unknown.<br />

DESIGN: Cohort study using participants from the ENGAGE, ENSURE<br />

and PURSUE randomized controlled trials which compared corifollitropin<br />

alpha and recombinant follicle-stimulating hormone.<br />

MATERIALS AND METHODS: The IVFpredict tool uses patient age,<br />

duration of infertility, cause of infertility, treatment, pregnancy and live birth<br />

history, source of oocytes, and use of ICSI as predictors. The discriminatory<br />

power (ability of the prediction tool to discriminate between cycles resulting<br />

and not resulting in live birth) was assessed using the area under the receiveroperator<br />

curve (AUROC). Calibration (how close the tool’s predictions match<br />

observed live birth rates) was assessed by comparing predicted with observed<br />

live birth rate in cycles with different patient and treatment characteristics.<br />

Differential calibration was assessed using Pearson’s chi-square test.<br />

RESULTS: 3292 fresh cycles of IVF were eligible for inclusion, of which<br />

complete data was available for 2591 cycles. The AUROC was 0.599 (95%<br />

confidence interval: 0.576 to 0.623). The predicted live birth rate in the validation<br />

sample was 27.3% compared to an observed rate of 29.3%. There was<br />

no evidence (all p R 0.25) of differential calibration with respect to different<br />

patient ages, duration of infertility, causes of infertility, number of previous<br />

cycles, or use of ICSI.<br />

CONCLUSIONS: External validation, the assessment of a prediction tool<br />

in a population that is different from that in which the tool was developed, is<br />

an essential step towards analyzing a prediction tool’s impact. IVFpredict has<br />

been validated in an international multicentre RCT population and demonstrates<br />

acceptable discrimination and calibration. IVFpredict may be utilized<br />

for stratification of patient prognosis prior to RCT entry.<br />

P-29 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

EFFECT OF BODY MASS INDEX (BMI) ON OOCYTE QUALITY IN<br />

IVF CYCLES. L. Aghajanova, M. Noel, C. Kao, M. Cedars. Obstetrics,<br />

Gynecology and Reproductive Sciences, University of California, San Francisco,<br />

San Francisco, CA.<br />

OBJECTIVE: Data on the impact of BMI on IVF outcomes and human<br />

oocyte quality are conflicting. We aimed to evaluate the effect of BMI on<br />

measures of oocyte quality and treatment outcomes in women undergoing<br />

IVF.<br />

DESIGN: Retrospective cross-sectional study<br />

MATERIALS AND METHODS: Clinical records were examined for all patients<br />

undergoing their first autologous IVF cycle between January 2009-<br />

<strong>October</strong> 2014 at a single academic institution. Cancelled and egg freezing cycles<br />

were excluded. Patients were grouped by BMI into underweight (30). Basic demographic<br />

characteristics, IVF cycle, oocyte and embryology parameters, and<br />

pregnancy outcomes including implantation rate (IR), pregnancy rate (PR), clinical<br />

pregnancy rate (CPR), miscarriage rate (MR) and live birth rate (LBR) were<br />

evaluated. SAS v9.3 was used for statistical analysis. Data was analyzed with<br />

either an ANCOVA or logistic regression controlling for age. Pairwise comparisons<br />

were performed and adjusted with the Bonferroni method.<br />

RESULTS: A total of 2309 patients met inclusion criteria. 2.8% were underweight<br />

(n¼65), 69.1% normal weight (n¼1596), 18.7% overweight (n¼432)<br />

and 9.4% obese (n¼<strong>21</strong>6). When comparing cycle parameters, peak estradiol<br />

levels were significantly lower in overweight patients compared to those with<br />

normal BMI (difference 239.473.6, p


OBJECTIVE: Obstetricians often recommend waiting 3 months after a<br />

pregnancy loss before attempting to conceive. However, little data supports<br />

this recommendation and previous research has focused on answering the<br />

question ‘‘When should couples achieve pregnancy after a loss?’’ instead<br />

of the more relevant public health question: ‘‘When should couples start<br />

trying to achieve pregnancy after a loss?’’ We compared time to pregnancy<br />

(TTP) leading to a live birth among varying inter-trying intervals (ITI)<br />

defined as time from last pregnancy loss to time starting conception attempt.<br />

DESIGN: Secondary analysis of a multicenter, block-randomized, doubleblind,<br />

placebo-controlled trial to evaluate the effect of preconceptioninitiated<br />

daily low dose aspirin on reproductive outcomes in women with a<br />

history of pregnancy loss.<br />

MATERIALS AND METHODS: 1083 women, aged 18-40 and actively<br />

trying to conceive with 1-2 prior losses and whose last pregnancy outcome<br />

was a loss, were included. Participants were followed for up to 6 menstrual<br />

cycles, and for women who became pregnant, until pregnancy outcome.<br />

TTP was defined as time from starting to try to conceive until date of conception.<br />

Cox models, accounting for left truncation and right censoring, estimated<br />

fecundability odds ratios (FOR) adjusting for age, race, BMI,<br />

education, and subfertility. Multiple imputation addressed missing data.<br />

Sensitivity analyses assessed the potential biasing effects of additional confounding<br />

factors or ITI exposure measurement error.<br />

RESULTS: Couples with a 0-3 month versus >3 month ITI were more<br />

likely to achieve a pregnancy leading to a live birth (53.2% versus 36.1%)<br />

and have a shorter, albeit non-significant, TTP (FOR: 1.<strong>17</strong> [95% CI: 0.86,<br />

1.59]). Findings remained robust after adjustment for additional potential<br />

confounding factors (Table). Sensitivity analyses testing the potential impact<br />

of measurement error corroborated our findings.<br />

CONCLUSIONS: Our findings suggest that the traditional recommendation<br />

to wait at least 3 months after a pregnancy loss before attempting to<br />

conceive may be unwarranted.<br />

Sensitivity analyses assessing robustness of findings.<br />

Inter-trying<br />

Interval<br />

Live Birth<br />

n (%)<br />

Sensitivity<br />

Analysis 1<br />

FOR (95% CI)<br />

Sensitivity<br />

Analysis 2<br />

FOR (95% CI)<br />

0-3 months 407 (53.2) 1.22 (0.87, 1.71) 1.16 (0.86, 1.58)<br />

>3 months 84 (36.1) 1.0 1.0<br />

1. Adjusted for age, race, BMI, education, subfertility, partner’s age, smoking,<br />

alcohol intake, parity, previous number of losses, recency of loss, gestational<br />

age of last loss, age of first intercourse, age of menarche, and dilation<br />

and curettage performed for last loss.<br />

2. Monte Carlo sampling applied to assess potential ITI exposure measurement<br />

error. Models were adjusted for age, race, BMI, education, and subfertility.<br />

The average FOR (95% CI) reported for 500 simulations.<br />

Supported by: Intramural Research Program, DIPHR, PRAE, NICHD,<br />

NIH<br />

P-31 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

PERINATAL OUTCOME IN PREGNANT WITH SUBCLINICAL<br />

HYPOTHYROIDISM. S. Agrawal. Obstetrics & Gynecology, Sudha<br />

Hospital Med Research Center, Kota, India.<br />

OBJECTIVE: Pregnancy is associated with major changes in the thyroid<br />

function. The thyroid function is very closely related to the reproductive performance<br />

in women. Hypothyroidism is one of the most common endocrinopathy<br />

seen during pregnancy. Approximately 5% of all pregnancies are<br />

affected by hypothyroidism. Majority of such cases have subclinical hypothyroidism,<br />

which is defined by an elevated serum thyroid stimulating hormone<br />

(TSH) concentration R 3.00 mU/L and a serum free thyroxine (FT4) in the<br />

normal range, i.e. between 0.80 to 1.90 ng/dl. Hypothyroidism has been associated<br />

with pregnancy complications such as preeclampsia, preterm labor,<br />

low birth weight, placental abruption, recurrent abortions, perinatal death<br />

and congenital hypothyroidism in the newborn. There has been a lot of debate<br />

on the impact of subclinical hypothyroidism and pregnancy outcome. Therefore,<br />

we did this large scale, multi-institutional prospective study to evaluate<br />

the consequences of subclinical hypothyroidism on pregnancy outcome.<br />

DESIGN: A multi-instituitional prospective study conducted in the Northern<br />

part of India between April 2010 and December 2014.<br />

MATERIALS AND METHODS: All pregnant women registered at our<br />

antenatal clinics were screened for thyroid dysfunction by serum TSH and<br />

free serum T4 using an immunometric TSH assay. Out of them, 489 women<br />

with TSH concentration R 3.00 mU/L and a serum free thyroxine (FT4) in<br />

the normal range, i.e. between 0.80 to 1.90 ng/dl were recruited as cases and<br />

similar number of controls were recruited with normal TSH and free T4<br />

values. These patients were followed up, delivered at the same hospitals<br />

and their maternal and perinatal outcomes were compared.<br />

RESULTS: Of all patients screened 489 subclinically hypothyroid pregnant<br />

women fulfilled our inclusion criteria were compared with control group.<br />

There was a significantly higher risk of spontaneous abortions, placental<br />

abruption, preterm labor, abruption placentae, low birth weight and neonatal<br />

hypothyroidism in pregnancies with subclinical hypothyroidism.<br />

CONCLUSIONS: Pregnancies with subclinical hypothyroidism have<br />

significantly higher number of adverse outcomes which can be potentially<br />

be prevented by thyroxine supplementation. So routine screening for hypothyroidism<br />

should be initiated for all pregnant women, however more such<br />

studies are needed before making any conclusion.<br />

References:<br />

1. Leung AS, Millar LK, Koonings PP, Montoro M, Mestman JH (1993)<br />

Perinatal outcome in hypothyroid pregnancies. Obstet Gynecol 81:<br />

349-353.<br />

2. Casey BM, Dashe JS, Wells CE, McIntire DD, Leveno KJ, Cunningham<br />

FG 2006 Subclinical hyperthyroidism and pregnancy outcomes.<br />

Obstet Gynecol 107:337-341<br />

3. Abalovich M , Gutierrez S , Alcaraz G , Maccallini G , Garcia A , Levalle<br />

O 2002 Overt and subclinical hypothyroidism complicating pregnancy.<br />

Thyroid 12:63-68<br />

P-32 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

FREQUENCY OF LOW SERUM LH LEVEL DURING OVARIAN<br />

STIMULATION IS ASSOCIATED WITH EARLY PREGNANCY<br />

LOSS IN IVF/ICSI CYCLES WITH A GNRH ANTAGONIST<br />

PROTOCOL. S. Chen, a Y. Chiang, b G. Jou, a C. Chang, c C. Chen. a a Department<br />

of Obstetrics and Gynecology, National Taiwan University Hospital,<br />

Taipei, Taiwan;<br />

b School of Medicine, National Taiwan University<br />

Hospital, Taipei, Taiwan;<br />

c Department of Medical Research, National<br />

Taiwan University Hospital, Taipei, Taiwan.<br />

OBJECTIVE: The influence of low serum LH level in the antagonist protocol<br />

for controlled ovarian stimulation (COS) remained controversial. Previous<br />

studies focused on the LH level on a certain day of COS. Our study is to<br />

evaluate the association between frequencies of low serum LH levels and<br />

pregnancy outcome in IVF/ICSI cycles with a GnRH antagonist protocol.<br />

DESIGN: Retrospective cohort study<br />

MATERIALS AND METHODS: A total of 695 patients underwent COS<br />

for IVF/ICSI cycles using GnRH antagonists and recombinant FSH-only<br />

from January 2011 to December 2012. Serial serum LH levels on day 2 of<br />

menstruation, before and during the administration of GnRH antagonist were<br />

measured until the day of hCG. We documented the lowest serum LH level<br />

in the cycle for each patient, and a threshold value of 0.8 mIU/mL was determined<br />

from logistic regression. Our control group were women with their<br />

lowest serum LH above 0.8 mIU/mL. Women with one or more low serum<br />

LH levels (LH ¼ 2 times) and early pregnancy loss were evaluated.<br />

RESULTS: During COS, there were 535 patients with their lowest serum<br />

LH above 0.8 mIU/mL. In the study group (n¼160), serum LH 0.8 mIU/mL, the incidence<br />

of early pregnancy loss was increased by 1.6 folds in patients with<br />

one episode of LH level


Incidence of early pregnancy loss in various episodes of low LH levels<br />

during ovarian stimulation.<br />

Episodes of LH<br />

¼ 2 2.354 (1.076-5.150) .032<br />

P-33 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ACTIVIN-A PROMOTES LUTEAL REGRESSION BY DOWN-<br />

REGULATING THE EXPRESSION OF STEROIDOGENIC EN-<br />

ZYMES AND UP-REGULATING BMP-6 AND ACTIVIN-A SUBUNIT<br />

IN HUMAN LUTEAL GRANULOSA CELLS. N. Akin, a G. Bildik, a<br />

Y. Guzel, b A. Seyhan Ata, b B. Balaban, b B. Urman, c,b O. Oktem. c,b a Reproductive<br />

Biology, Koc University Graduate School of Health Sciences, Istanbul,<br />

Turkey;<br />

b Women’s Health Center Assisted Reproduction Unit,<br />

American Hospital, Istanbul, Turkey; c Obstetrics and Gynecology, Koc University<br />

School of Medicine, Istanbul, Turkey.<br />

OBJECTIVE: Activin-A up-regulates the expression of aromatase enzyme<br />

and FSH receptor, and has mitogenic effect on the granulosa cells at early<br />

stage of follicle growth. But its roles on corpus luteum are not clear.<br />

DESIGN: A translational research study<br />

MATERIALS AND METHODS: Human luteal granulosa cells (HLGCs)<br />

retrieved from the patients undergoing natural (n¼10), GnRH antagonist<br />

(n¼10) and agonist (n¼10) IVF cycles were incubated with recombinant activin-A<br />

(60 ng/mL) for 72hrs. The expressions of activin-A beta-A subunit<br />

(INHBA) and activin receptors (ACVR1B, ACVR2A and ACVR2B), steroidogenic<br />

enzymes (stAR, Scc, 3-b HSD and aromatase), LH receptor<br />

and bone morphogenetic proteins (BMPs) 4, 6, 7, 15, and growth and differentiation<br />

factor- 9 (GDF-9) were quantitatively compared by qRT-PCR between<br />

control and activin-A treated cells.<br />

RESULTS: Activin-A and its receptors were expressed by HLGCs of natural<br />

cycles at a significantly higher level than those of GnRH agonist and<br />

antagonist cycles. Compared to control, activin-A significantly decreased<br />

E2 and P production of the cells, reduced their expression of the steroidogenic<br />

enzymes, LH receptor, BMP-4 and 15 and GDF-9; and increased activin-A<br />

beta-A subunit and BMP-6 in both natural and GnRH agonist and<br />

antagonist IVF cycles.<br />

CONCLUSIONS: These results indicate that activin-A expedites luteal<br />

regression in HLGCs. This action of activin-A may open a new avenue in<br />

the treatment of ovarian hyperstimulation syndrome (OHSS).<br />

Granulosa cells before and after activin-A treatment antagonist IVF cycle.<br />

Gene Expression Control Activin-A P-Value<br />

BMP-4 1.000.45 0.640.13


6.4% (3.7-10.9), and 9.8% (6.0-15.4) for AMH quartiles 1-4, respectively (ptrend¼0.52).<br />

Cycle cancellation was also not associated with AMH quartiles<br />

in a multivariate model (p¼0.91).<br />

CONCLUSIONS: Our analysis suggests that AMH may not provide additional<br />

benefit for predicting clinical pregnancy in both FSH and clomiphene<br />

citrate IUI cycles among our patient population.<br />

P-36 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE HETEROGENITY OF ANTI-MULLERIAN HORMONE<br />

DEPLETION MODELS ACROSS ASSAYS IN 129<strong>17</strong><br />

WOMEN. K. Lukaszuk, a,b L. Plociennik, a A. Lukaszuk, a<br />

D. Bialobrzeska. a a INVICTA Fertility and Reproductive Centre, Gdansk,<br />

Poland; b Gynecology and Oncological Gynecology, Medical University of<br />

Gdansk, Gdansk, Poland.<br />

OBJECTIVE: A regular Anti-M€ullerian hormone diagnosis obtained with<br />

different kits is still an open problem to opinion makers in the field of reproductive<br />

medicine. As far now the literature points out that the profile of the<br />

rate of AMH change varies for available assays. Thus, predominately important<br />

is to measure the heterogeneity among kits.<br />

DESIGN: The study was prospectively conducted on 129<strong>17</strong> women (aged<br />

between 20 and 50 years) examined with four AMH assays. AMH measurements<br />

were performed during the time period between January 2007 and<br />

December 2012 in Invicta Fertility Centre, Poland.<br />

MATERIALS AND METHODS: 129<strong>17</strong> women were examined with four<br />

AMH assays: Immunotech I generation kit (ImI 4016 samples), Beckman<br />

Coulter II generation kit RUO (BCII RUO 3430 samples), Beckman Coulter<br />

II generation kit with IVD certificate (BCII IVD 830 samples) and Ansh Labs<br />

I generation kit (AnshLabs 4641 samples).<br />

RESULTS: The data revealed the lack of parallelism in AMH depletion<br />

between assays at P< .05, when pairwise comparisons of assays were performed.<br />

Also the distribution of AMH concentration in quartiles between<br />

kits was statistically different (P< .05). Consequently, calculations proved<br />

an interaction between age variable and assays.Based on evidences we<br />

concluded that heterogeneity in the rate of AMH change exists and the central<br />

tendency for AMH levels across assays is substantially different. Finally,<br />

AMH diagnosis in women is dependent on an assay-age effect.<br />

CONCLUSIONS: We claim future studies should endeavour to create a<br />

profiles for heterogeneity and homogeneity of slopes between kits along<br />

age periods. This will give a certain medical aid to decision makers (clinicians),<br />

who perform opinions on infertility problem in patients. Those gynaecologists<br />

concerned with programming options for fertility preservation<br />

might benefit as well.<br />

P-37 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

CREATION & VALIDATION OF THE FERTILITY AND INFER-<br />

TILITY TREATMENT KNOWLEDGE SURVEY (FIT-<br />

KS). R. Kudesia, a E. Chernyak, b B. McAvey. c a Department of Obstetrics<br />

& Gynecology and Women’s Health, Albert Einstein College of Medicine,<br />

New York, NY; b Albert Einstein College of Medicine, New York, NY; c Icahn<br />

School of Medicine at Mount Sinai, Reproduct, New York, NY.<br />

OBJECTIVE: Reproductive decision-making is complicated by a preponderance<br />

of fertility misconceptions. Low fertility knowledge has been<br />

demonstrated across educational levels, in the general population and even<br />

among physicians. We therefore sought to validate a novel instrument assessing<br />

fertility and infertility treatment knowledge for use in both patients and<br />

providers.<br />

DESIGN: Internet-based survey<br />

MATERIALS AND METHODS: An internet-based survey regarding<br />

fertility knowledge and treatment, along with relevant demographics, was<br />

constructed by the research team and examined for clarity and thoroughness<br />

by a panel of reproductive endocrinologists. The pilot survey was distributed<br />

to female students and OB/GYN housestaff at two academic hospitals, as<br />

well as to females in the general population via Amazon Mechanical Turk.<br />

Item analysis and assessment for convergent and divergent validity was<br />

then performed. The study was IRB-exempted.<br />

RESULTS: 127 medical trainees (99 students, 28 housestaff) and 118<br />

women participated. They represented a broad spectrum of geographic,<br />

ethnic and faith backgrounds. The mean scores were 18.8/29 (64.9%) among<br />

medical trainees (MT) and 16.2/29 (55.9%) among the general population<br />

(GP). Among both groups, a majority underestimated the miscarriage rate<br />

(50.4% MT, 59.3% GP), and over-estimated success in delayed-childbearing<br />

scenarios relating to IVF for a 44-year old woman (73.2% MT, 68.6% GP)<br />

and oocyte cryopreservation (96.1% MT, 91.5% GP). There were additional<br />

inaccuracies relating to fertility-impacting lifestyle choices, such as the<br />

impact of moderate alcohol consumption or lubricants. The instrument<br />

demonstrated an appropriate profile in item analysis and validity testing.<br />

CONCLUSIONS: This study demonstrated that the FIT-KS instrument is<br />

an appropriate tool to evaluate understanding of fertility and infertility treatment<br />

knowledge in both patients and providers. The validation process also<br />

reaffirmed that significant gaps remain among medical trainees and within<br />

the general population, and greater efforts to improve reproductive education<br />

must be undertaken.<br />

Supported by: Joan F Giambalvo Fund for the Advancement of Women,<br />

Harold and Muriel Block Institute for Clinical and Translational Research<br />

at Einstein and Montefiore (UL1 TR001073)<br />

P-38 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE EFFECTS OF FORMOTEROL ON THE SERUM, PERITO-<br />

NEAL VEGF, AND MDA AND VEGF LEVELS IN THE OVARIES<br />

AND ENDOMETRIUM OF RATS WITH OHSS. H. A. Inal, a<br />

Z. O. Inal, a N. Yilmaz, b H. Timur, c A. Sargin Oruc, d M. N. Kalem, e<br />

O. Han. f a IVF Unit, Konya Education and Research Hospital, Konya, Turkey;<br />

b Reproductive Endocrinology Department, ZTB, Ankara, Turkey; c IVF Unit,<br />

Dr. Zekai Tahir Burak Women’s Health Research and Education Hospital,<br />

Ankara, Turkey; d Zekai Tahir Burak Women’s Health Education and Res,<br />

Ankara, Turkey; e Turgut Ozal University Hospital, Ankara, Turkey; f Guven<br />

Hospital, Ankara, Turkey.<br />

OBJECTIVE: To investigate the effects of formoterol (a beta2-adrenoreceptor<br />

agonist) on serum and peritoneal fluid VEGF, and MDA levels and<br />

on VEGF-stained cell counts in the ovaries and endometrium of rats with<br />

OHSS within the framework of immunohistochemical analysis.<br />

DESIGN: A total of 28 immature female Wistar rats were randomly<br />

divided into four groups.<br />

MATERIALS AND METHODS: Three groups were given 10 IU pregnant<br />

mare serum gonadotropin/day on days 22-25 of life. They were administered<br />

30 IU hCG on day 26 of life to mimic OHSS. On days 26 and 27 of life, 24<br />

mcg/kg/day formoterol in group 3 and 48 mcg/kg formoterol in group 4 were<br />

Comparison of the ovarian cortical and stromal VEGF, endometrial VEGF, serum and peritoneal VEGF lev.<br />

Group<br />

Group1<br />

(Control)<br />

Group 2<br />

(OHSS-Placebo)<br />

Group 3<br />

(OHSS-Formoterol 24 mcg)<br />

Group 4<br />

(OHSS-Formoterol 48 mcg)<br />

p value<br />

Ovarian cortex VEGF* 99.14+8.07 a 156.57+47.80 a 142.86+30.39 1<strong>21</strong>.43+26.24 0.013<br />

Ovarian stromal VEGF 74.42+15.60 a,b,c <strong>17</strong>9.00+<strong>17</strong>.37 a <strong>17</strong>1.85+<strong>17</strong>.90 b 168.42+10.98 c 0.001<br />

Endometrial VEGF 76.71+12.47 a 125.29+27.47a,d,e 82.86+<strong>21</strong>.10d 57.00+25.41e 0.001<br />

Blood serum VEGF (mcg/mL) 19.53+6.47 25,83+7,22 25.30+6.00 24.67+6,69 0.281<br />

Peritoneal VEGF (mcg/mL) 22.35+18.16 28.93+7.<strong>17</strong> 23.52+8.16 22.97+6.98 0.674<br />

Blood serum MDA** (ng/mL) 11.94+2.66 13.37+2.57 13.31+1.88 12.30+1.58 0.543<br />

Peritoneal MDA (ng/mL) 0.72+0.68 1.59+1.38 1.22+1.20 0.91+0.91 0.546<br />

*VEGF: vascular endothelial growth factor; **MDA: malondialdehyde. a¼ The difference between groups 1 and 2 was statistically significant, b¼ the difference<br />

between groups 1 and 3 was statistically significant, c¼ the difference between groups 1 and 4 was statistically significant, d¼ the difference between<br />

groups 2 and 3 was statistically significant, and e¼ the difference between groups 2 and 4 was statistically significant.<br />

e118 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


administered intraperitoneally per animal. The control group (group¼1) was<br />

given the same dosage of 0.9% saline solution (ip) on days 22-26 day of life.-<br />

The main outcomes were weight gain, ovarian volumes and weights, serum<br />

and peritoneal fluid VEGF and MDA levels and VEGF-stained cell count<br />

in the ovaries and endometrium were evaluated.<br />

RESULTS: The weight gain and ovarian volumes were highest in the<br />

OHSS-Placebo group (p¼0.001). The ovarian weights were lower in the control<br />

and formoterol treatment groups than in the OHSS-Placebo group<br />

(p¼0.001). Although, there were no statistically significant differences between<br />

the groups in terms of serum and peritoneal fluid VEGF or MDA levels<br />

(serum VEGF p¼ 0.281, peritoneal VEGF p¼ 0.674, serum MDA p¼ 0.543,<br />

peritoneal MDA p¼ 0.506), there was a significant difference between the<br />

control and the OHSS-Placebo groups (p¼0.013) regarding the VEGF in<br />

the ovarian cortex. There was a significant difference between the control<br />

and the other groups in terms of ovarian stroma (p¼0.001), and there was<br />

also a statistically significant difference between the OHSS-Placebo and<br />

the other groups regarding VEGF in the endometrium (OHSS-Placebo vs.<br />

control group p¼0.002, OHSS-Placebo vs. the formoterol 24 mcg/kg group<br />

p¼0.008, and OHSS-Placebo vs. the formoterol 48 mcg/kg group p¼0.001).<br />

There were no significant differences between the control and formoterol<br />

treatment groups (p¼0.955 and p¼0.372).<br />

CONCLUSIONS: Formoterol represents a potential novel strategy for the<br />

management of OHSS. Further studies, including those examining the<br />

dosage of formoterol, are warranted.<br />

P-39 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ELEVATED ESTRADIOL LEVELS POST-TRIGGER ARE ASSOCI-<br />

ATED WITH INCREASED ANEUPLOIDY RATES. A. P. Melnick,<br />

R. Setton, E. M. Murphy, Z. Rosenwaks, S. D. Spandorfer. The Ronald<br />

O. Perelman and Claudia Cohen Center for Reproductive Medicine, Weill<br />

Cornell Medical College, New York, NY.<br />

OBJECTIVE: To determine whether elevated estradiol levels post-hCG<br />

trigger are associated with increased aneuploidy rates in patients undergoing<br />

in vitro fertilization with preimplantation genetic screening (IVF-PGS).<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Patients undergoing IVF-PGS with trophectoderm<br />

biopsy and 24-chromosome screening over a three year period at<br />

an academic fertility center were included. Patients were stratified into two<br />

groups by estradiol level post-HCG per mature oocyte (MII) retrieved. The<br />

primary outcome assessed was aneuploidy rate. Secondary outcomes<br />

included total and mature oocyte yield, total days of stimulation, lead follicle<br />

size at trigger, and estradiol level at trigger and one day post-trigger. Statistical<br />

analysis included student’s t-test, chi-square test, and logistic regression.<br />

P


P-42 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

OBESITY AND METABOLISM<br />

OXIDATIVE STRESS AND QUALITY DEGRADATION IN MOUSE<br />

OOCYTES WERE INDUCED BY MATERNAL<br />

OBESITY. M. Kobayashi A. Yoshida. Kiba Park Clinic Research Center,<br />

Tokyo, Japan.<br />

OBJECTIVE: Obesity and overweight increase the risk of many health<br />

problems. In reproductive outcomes of obese women, rates of infertility<br />

and pregnancy loss are higher than normal women. Recently, the relationship<br />

between obesity and oxidative stress is reported and such studies are progressing.<br />

However, the mechanisms of the defects on oocytes from obese female<br />

have not yet been cleared. The aim of this study was to estimate the<br />

effect of maternal obesity on the function of organelles, embryonic development,<br />

and redox status of oocyte.<br />

DESIGN: The meiotic spindle morphology, mitochondrial membrane potential<br />

(MMP), and developmental potency of ovulated oocytes derived from<br />

obese mice were studied. Redox state in ovulated oocytes and sera were<br />

compared between normal and obese mice.<br />

MATERIALS AND METHODS: Female C57BL/6J mice were fed either a<br />

control diet (control) or a high-fat diet (obesity) for 8 weeks. Then, matured oocytes<br />

were retrieved from each group. Morphology of meiotic spindle was<br />

examined immunohistochemically, and mitochondrial membrane potency<br />

(MMP) was evaluated with JC-1 dye. For evaluation of the developmental ability,<br />

oocytes were fertilized by ICSI and cultured in KSOM until to blastocyst<br />

stage. Redox state of ovulated oocytes and sera derived from normal and obese<br />

mice were assessed by dihydroethidium (DHE) staining and d-ROMs/BAP test,<br />

respectively.Reactive oxygen species (ROS) in ovulated oocytes were assessed<br />

by dihydroethidium (DHE) staining. Further, systemic redox states in normal<br />

and obese mice were evaluated with d-ROMs and BAP test of sera.<br />

RESULTS: After 8 weeks, mice fed high-fat diet were significantly heavier<br />

than control (28.32.8g vs. <strong>21</strong>.<strong>21</strong>.2g, p


attributes of a health promotion class (HPC) would be attractive to infertility<br />

patients.<br />

DESIGN: Cross-sectional survey study.<br />

MATERIALS AND METHODS: Patients aged <strong>21</strong>-35 presenting to an<br />

infertility clinic were invited to complete an anonymous survey regarding demographics,<br />

perceptions of health, weight status, the effect of excess preconception<br />

weight and potential weight optimization interventional components.<br />

Data was analyzed using regression models.<br />

RESULTS: To date, 24 of the 50 eligible women completed the survey.<br />

Women were typically young (30.1 6.5 years), married (87%), Caucasian<br />

(78%), and college educated (61%) with average BMI of 32 kg/m2. On a 5-<br />

point Likert scale, most consider themselves ‘‘somewhat healthy’’ (64%),<br />

feel ‘‘very/completely confident’’ (64%) that they could perform moderate<br />

physical activity (PA) for 30 minutes 5 times/week, correctly identify PA recommendations<br />

(71%) but are aware that they do not meet these recommendations<br />

(68%). In preparation for pregnancy, patients indicated they should<br />

lose weight (65%) and increase PA (74%), and were aware that excess preconception<br />

weight increases the risk of infertility (86%), pregnancy complications<br />

(86%), and difficult childbirth (82%), however, only 50% were aware<br />

of the increased risk of having an overweight/obese child. Of the 67% of patients<br />

who indicated they would attend a health promotion class (HPC), many<br />

would attend weekly or monthly meetings (58%) lasting 30-60 minutes<br />

(63%) in a community room (42%), gym (42%), online (58%), or via email<br />

(46%), without monetary incentive (85%). Most preferred a HPC that would<br />

provide tips for cheap, healthy eating including recipes and progress<br />

tracking.<br />

CONCLUSIONS: Our study demonstrates that infertile patients are aware<br />

of the risks of excess weight as well as PA recommendations, but do not meet<br />

PA recommendations despite indicating they are physically capable of doing<br />

so. A majority of patients would participate in a HPC without incentive. Preferences<br />

for frequency, duration, location and type of class identified by this<br />

study can be used to develop a HPC tailored to infertility patients.<br />

Supported by: Study funded by Virginia Tech Fralin Life Sciences Institute<br />

P-46 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ADOLESCENT BODY MASS INDEX IS DECREASED IN CHIL-<br />

DREN CONCEIVED BY INFERTILE COUPLES, REGARDLESS<br />

OF FERTILITY TREATMENT. M. Link, a H. Hanson, b<br />

J. M. Hotaling, c K. Smith, d K. Aston, e D. T. Carrell, f E. B. Johnstone. c a Obstetrics<br />

and Gynecology, University of Utah, Salt Lake City, UT; b Family and<br />

Preventive Medicine, University of Utah, Salt Lake City, UT; c University of<br />

Utah, Salt Lake City, UT; d Population Sciences, University of Utah, Salt<br />

Lake City, UT; e Andrology and IVF Laboratories, University of Utah, Salt<br />

Lake City, UT; f University of Utah School of Medicine, Salt Lake City, UT.<br />

OBJECTIVE: To determine whether birth weight and adolescent body<br />

mass index (BMI) differ among offspring of infertile couples conceived<br />

with and without the use of assisted reproductive technologies and whether<br />

they differ from offspring of fertile controls.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Children of infertile Utah couples born<br />

prior to 1999 were selected for study inclusion (n¼6,251). Data include 165 children<br />

conceived through intrauterine insemination (IUI), 224 children conceived<br />

through in vitro fertilization (IVF) and 5,862 natural conceptions (Infertile - NC).<br />

Offspring of infertile couples were matched to offspring of fertile controls by sex,<br />

birth year (+/- 1 year), year of BMI measurement (+/- 2 years), and multiplicity<br />

(singleton, multiple birth). Primary outcomes of interest were birth weight and<br />

BMI at first driver’s license record. Multivariate generalized linear models<br />

were run adjusting for birth order, gestational age, maternal age, maternal/<br />

paternal education, socioeconomic status, and maternal/paternal BMI.<br />

RESULTS: Table 1 shows a summary of the results. For singleton births,<br />

there were no significant differences in birth weight among IUI/IVF infants,<br />

infertile-NC children, and children of fertile couples. BMI in adolescence<br />

was lower in singletons conceived through IVF and Infertile-NC children<br />

compared to children of fertile controls. For twin births, birth weight was<br />

lower for IVF children but there were no other significant differences in birth<br />

weight between children of infertile and fertile couples. In adolescence, all<br />

twins of infertile couples, regardless of route of conception, had lower<br />

BMIs than children of fertile controls.<br />

CONCLUSIONS: Twins, and to a lesser degree, singletons born to infertile<br />

couples have similar birth weights but slightly lower adolescent BMI than<br />

children born to fertile couples, regardless of the route of conception. Factors<br />

leading to infertility may contribute to in utero stress that is compounded by<br />

twin gestation and may ultimately result in decreased BMI for offspring.<br />

Table 1 - Birth Weight and Adolescent BMI of Children Born to Infertile and<br />

Fertile Couples.<br />

SINGLETONS<br />

Birth Weight<br />

(grams)<br />

Adolescent<br />

BMI<br />

TWINS<br />

Birth Weight<br />

(grams)<br />

Adolescent<br />

BMI<br />

Children of Fertile 3248 <strong>21</strong>.4 2701 20.0<br />

Controls<br />

Infertile - IUI 3255 <strong>21</strong>.0 2588 18.7*<br />

Infertile - IVF 3245 20.6* 2467* 18.5*<br />

Infertile - NC 3242 <strong>21</strong>.0* 2607 19.2*<br />

P-47 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

OBESITY AND HIGH FAT DIET IMPAIR RESPONSE TO SUPER-<br />

OVULATION BY SUPPRESSING GENES INVOLVED IN OVARIAN<br />

FOLLICULOGENESIS. K. Thornton, a O. A. Asemota, b S. K. Jindal, a<br />

M. Charron, a E. Buyuk. a a Obstetrics and Gynecology, Albert Einstein College<br />

of Medicine/ Montefiore Medical Center, Bronx, NY; b Advanced<br />

Fertility Center of Chicago, Gurnee, IL.<br />

OBJECTIVE: Obese women undergoing controlled ovarian hyperstimulation<br />

(COH) have higher cycle cancellation rates, fewer oocytes collected and<br />

lower pregnancy rates. The exact mechanism of how obesity causes reproductive<br />

dysfunction is not well understood. We hypothesize that obesity<br />

and high fat diet are associated with altered ovarian gene expression, which<br />

interferes with folliculogenesis and/or ovulation.<br />

DESIGN: Prospective controlled study.<br />

MATERIALS AND METHODS: Mice were subjected to dietary manipulation<br />

starting at 6w to develop 2 female rodent models: Group 1: C57BL/6J normal<br />

chow fed mice (NC) (N¼12) and Group 2: C57BL/6J high fat diet fed mice (HF)<br />

(N¼12). At 20w of age, 6 mice in each group were sacrificed while another 6 underwent<br />

superovulation (SO) with gonadotropins and then were sacrificed. At the<br />

time of sacrifice, one ovary was snap frozen in liquid nitrogen and used for RNA<br />

extraction and RT-PCR for detecting the following genes: Gdf-9, Bmp-15, Amh,<br />

Amhr (Amh receptor), Foxl2, Gja1 (mouse homolog for human connexin 43),<br />

Fshr and Lhcgr. Ribosomal protein 36b4 was used as loading control. Data given<br />

as units of expression SEM and p


MATERIALS AND METHODS: We reviewed IVF cycles during the<br />

period from 2010 through 2014 at a university IVF center. We included women<br />

with PCOS under the age of 40 years who underwent autologous cleavage stage<br />

(day 2 or 3) fresh embryo transfer. Women treated with metformin during the<br />

IVF cycle were excluded. The included cycles were classified into 2 groups according<br />

to their body mass index (BMI): Normal (BMI %25) and Overweight/<br />

obese (BMI >25). The primary outcomes were clinical pregnancy and life birth<br />

rates. Secondary outcomes included number of oocytes retrieved, number and<br />

quality of embryos, and abortion rates. Good quality embryos included grade A<br />

(cells 8 and 7, fragmentation 0-10% and symmetry perfect to moderate) and<br />

grade B (cells 8 and 7, fragmentation up to 25% and symmetry perfect to moderate)<br />

or (cells 6 and 5, fragmentation 0-10% and symmetry perfect to moderate).<br />

The control group consisted of women without PCOS under the age of 40<br />

undergoing IVF/ET for unexplained infertility.<br />

RESULTS: 90 PCOS cycles were included in the analysis, 56 cycles in the<br />

normal BMI group and 34 in the overweight and obese group. 152 non-PCOS<br />

cycles were used as control. Overweight/obese women with PCOS compared<br />

to normal weight women with PCOS were found to have an increased number<br />

of poor quality embryos and decreased percentage of good quality embryos<br />

(see table). Overweight/obese women with PCOS also had reduced clinical<br />

pregnancy rate, increased abortion rate and a reduced live birth rate compared<br />

to normal weight women with PCOS. In women without PCOS, overweight/<br />

obesity did not have measurable adverse effects on IVF outcome.<br />

CONCLUSIONS: Overweight and Obesity are associated with adverse IVF<br />

outcomes in women with PCOS but not in those with unexplained infertility.<br />

Outcomes by groups.<br />

BMI%25<br />

(n¼56)<br />

PCOS infertility<br />

BMI>25<br />

(n¼34)<br />

p value<br />

Unexplained infertility<br />

BMI%25<br />

(n¼86)<br />

BMI>25<br />

(n¼66)<br />

p value<br />

BMI (Mean SD) 20.94 2.11 31.02 6.69


gas anesthesia. Isotopic tracers of 13C, D3-Met, 13C3-Ser and D2-Cys were<br />

administered as a 4 hour prime-constant rate infusion. Blood samples were<br />

obtained at 3, 3.5, 3.75 and 4 hours. After sacrifice, germinal vesicle stage<br />

oocytes were isolated and denuded. The rate of appearance of each amino<br />

acid being traced was calculated using the tracer dilution method.<br />

RESULTS: Dietary protein restriction caused differential responses in<br />

the kinetics and concentrations of plasma serine and glycine, and folate<br />

supplementation did not ameliorate the observed responses. The total<br />

flux of serine, a methyl group donor, was higher in the LP and LPF groups<br />

(P


dopamine. It has been reported that dopamine negatively regulates follicle<br />

stimulating hormone (FSH) and insulin secretion through the D2 receptor,<br />

and this suggests that SLC18A2 might play a role in polycystic ovary syndrome<br />

(PCOS). Until now, however, there have been no studies on the association<br />

between SLC18A2 and PCOS. We investigated whether naturally<br />

occurring genetic variations at the human SLC18A2 locus are associated<br />

with the etiology of PCOS and/or with the FSH levels and insulin secretion<br />

seen in PCOS pataints.<br />

DESIGN: We undertook a systematic search for polymorphisms in<br />

SLC18A2 by resequencing the gene and investigated whether the common<br />

genetic variants are associated with PCOS in a large cohort<br />

(n¼539). And we also investigated the relationship between genetic variants<br />

and various kinds of metabolic traits in patients. Finally, we studied<br />

function of corresponding genetic variant by luciferase expression experiment<br />

in vitro.<br />

MATERIALS AND METHODS: We genotyped common single-nucleotide<br />

polymorphisms across the locus in 319 PCOS patients who were well<br />

phenotyped for several metabolic traits to determine associations between<br />

SLC18A2 variants and PCOS. We constructed wild-type and mutant plasmids<br />

into HeLa cell and investigated functional of different allele by<br />

measuring luciferase activity.<br />

RESULTS: We found two common genetic variants in the 3’-untranslated<br />

region(rs363282 and rs363238) that are strongly associated with serum FSH<br />

concentration (P¼0.0004 and P¼0.0001, respectively), insulin level at 0.5h<br />

after an oral glucose tolerance test (P¼0.032 and P¼0.009, respectively),<br />

and insulin level at 1h after the test (P¼0.007 and P¼0.005, respectively).<br />

Finally, we performed a functional study that showed that minor alleles of<br />

the two variants decreased expression from a transfected luciferase reporter/SLC18A23’-UTR<br />

expression plasmid.<br />

CONCLUSIONS: Our results strongly suggest that common genetic variants<br />

in SLC18A2 contribute to the phenotypic expression of PCOS and<br />

suggest novel pathophysiological links between the SLC18A2 locus and<br />

PCOS.<br />

Table1.The association of the phenotype and genotype of rs363282 and<br />

rs363238 in patients with PCOS.<br />

Rs363282 Rs363282 Rs363282 Rs363238 Rs363238 Rs363238<br />

AA+AG GG P-value CC+AC AA P-value<br />

Number 245 72 240 67<br />

Age (years) 26.02 0.40 24.66 0.68 0.135 25.94 0.41 24.95 0.74 0.236<br />

BMI (kg/m2) 25.60 0.33 25.15 0.67 0.667 25.43 0.33 25.56 0.72 0.833<br />

E2 (pmol/L) 188.39 9.10 190.89 <strong>17</strong>.98 0.930 195.25 19.33 195.25 19.33 0.784<br />

T (nmol/L) 2.83 0.05 2.72 0.09 0.<strong>21</strong>1 2.84 0.05 2.70 0.10 0.<strong>17</strong>3<br />

LH (IU/L) 8.91 0.35 10.31 0.87 0.102 8.91 0.35 10.41 0.95 0.074<br />

FSH (IU/L) 7.14 0.16 8.<strong>17</strong> 0.25 0.0004 7.10 0.15 8.33 0.26 0.0001<br />

FINS (mU/L) 16.57 0.88 <strong>17</strong>.62 1.53 0.294 16.52 0.89 <strong>17</strong>.97 1.65 0.288<br />

0.5hINS (mU/L) 89.48 3.16 109.12 12.82 0.032 88.63 3.15 114.46 14.09 0.009<br />

1hINS (mU/L) 110.97 3.70 135.74 13.84 0.007 109.95 3.44 139.92 15.22 0.005<br />

Supported by: This work was Supported by the National Natural Science<br />

Foundation of China (Grants 81270747) and the Shanghai City Board of<br />

Education Scientific Research Innovation Key Projects (Grant 13ZZ001).<br />

P-54 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

IMPACT OF OVARIAN VOLUME-BASED ADJUSTED THERMAL<br />

DOSE VERSUS FIXED-PUNCTURE DOSAGE IN LAPAROSCOPIC<br />

OVARIAN DRILLING ON OVARIAN RESERVE IN CLOMIPHENE<br />

CITRATE-RESISTANT PCOS WOMEN. A. NASR. Obstetrics and Gynecology,<br />

Women’s Health Center, Assiut University, Assiut, Egypt.<br />

OBJECTIVE: To evaluate the impact of adjusted thermal dose on the basis<br />

of ovarian volume versus fixed-puncture dosage in laparoscopic ovarian drilling<br />

(LOD) on reproductive outcome and ovarian reserve assessed by serum<br />

anti-Mullerian hormone (AMH) in women with clomiphene citrate(CC)-<br />

resistant polycystic ovary syndrome (PCOS).<br />

DESIGN: A randomized controlled clinical trial.<br />

MATERIALS AND METHODS: Eighty women with CC-resistant PCOS<br />

who underwent LOD were recruited. They were randomized into two groups:<br />

group A [40 women; mean age: (SD) 27.7 (2.1) years], received an adjusted<br />

thermal dose based on ovarian volume with use of a new model for dose<br />

calculation (60 J/cm3 of ovarian tissue); group B [40 women; mean age:<br />

(SD) 28.5 (1.9) years], received 600 J per ovary through four ovarian holes<br />

regardless of size. A group of normally ovulating women; group C [40<br />

women; mean age: (SD) 27.9 (2.2) years], served as controls. Basal serum<br />

AMH levels were measured in all three groups and six months after LOD<br />

in groups A and B. Women were followed-up for six months. Assuming a<br />

20% difference between the groups, with an a of 5% and a b of 20%, it<br />

was calculated that forty women are required in each arm of the study to<br />

detect a true difference at the 95% confidence level with 80% power. Statistical<br />

analysis was performed using SPSS software, version <strong>17</strong>.0 (SPSS, Chicago,<br />

IL, USA). P


of OHSS, especially in patients with PCO. IVM is a relatively new option for<br />

ART promising significant benefits, but still controversial. However, we have<br />

been using IVM as a first choice of ART mainly for PCO patients for more<br />

than 10 years. The present study was conducted to validate if IVM was clinically<br />

useful for infertility by analyzing all pregnant cases of the last 10 years.<br />

DESIGN: Retrospective clinical study at private setting fertility clinic.<br />

MATERIALS AND METHODS: Three hundred and thirty six cycles of<br />

IVM performed from January 2004 to December 2013 were analyzed retrospectively.<br />

Ultrasound monitoring was started from day 7 of menstrual cycle<br />

and oocyte retrieval was performed when the smallest follicle reached 7 mm<br />

in diameter. Regardless of the choice of priming, HCG 10,000 units was<br />

administered and followed by oocyte retrieval 38 hours later. When endometrium<br />

was thicker than 8mm at oocyte retrieval, fresh embryo transfer was<br />

performed. Otherwise, all embryos were vitrified for frozen transfer cycle.<br />

Frozen cycles were not included in the present study. Clinical outcomes<br />

such as number of oocytes retrieved, maturation rate, embryo transfer rate,<br />

and clinical pregnancy rate (GS positive) were analyzed. Moreover, appropriate<br />

timing for oocyte retrieval was evaluated.<br />

RESULTS: Average number of oocyte retrieved was 10.4 (3498 oocytes<br />

per 336 retrievals) and maturation rate was 47.8 % (1683 metaphase two/<br />

3498 retrieved oocytes). Fresh cycle embryo transfer rate was 47.8% (<strong>21</strong>6<br />

transfers per 336 retrievals). Oocyte retrievals on PCO patient rate was<br />

66.9% (225 cycle per 336 cycles). Clinical pregnancy rate with gestational<br />

sac was 31.5% (68 pregnancies per <strong>21</strong>6 transfers) and 53 babies were born<br />

without any congenital abnormality. Appropriate cycle day for IVM oocyte<br />

retrieval was day 15.4 (between day 9 and 28) for fresh cycle. However,<br />

PCO patients have wide ranges of menstrual cycle days. Therefore, we determined<br />

appropriate timing of retrieval by the percentage of individual menstrual<br />

cycle length. Thirty three percent of cycle length is appropriate for<br />

retrievals in fresh cycle. Average estradiol value of pregnant patients on<br />

the day of HCG was 192 pg/ml.<br />

CONCLUSIONS: Average clinical pregnancy rate of IVM was 31.5% and<br />

it was considered acceptable for clinical validity not only because of less<br />

physical and financial burdens, but also absolute no risk of OHSS. Overall,<br />

IVM is a valid treatment of ART especially for PCO patients when appropriately<br />

performed.<br />

References: Validation of clinical IVM.<br />

P-57 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

CLINICAL OUTCOME OF PCOS PATIENTS UNDERGOING ASSIS-<br />

TED REPRODUCTIVE TECHNOLOGY: THE ROLE OF IN VITRO<br />

MATURATION. T. Takeuchi, a N. Aono, a N. Oka, a R. Obata, a<br />

N. Okuyama, a S. Yanagihori, a T. Okuda, a K. Kyono. a,b a Kyono ART Clinic<br />

Takanawa, Minatoku, Tokyo, Japan; b Kyono ART Clinic, Sendai, Miyagi,<br />

Japan.<br />

OBJECTIVE: Assisted reproductive technology in PCOS patients is often<br />

associated with ovarian hyperstimulation syndrome (OHSS). The severe<br />

form of OHSS is an iatrogenic complication and possibly a life-threatening<br />

condition resulting from ovarian stimulation. In vitro maturation (IVM) techniques<br />

have been developed and employed in clinical settings to circumvent<br />

OHSS by minimizing or eliminating gonadotropin administration. The aim<br />

of this study was to assess the efficacy of IVM in PCOS patients by<br />

comparing its clinical outcome with that of conventional ovarian stimulation<br />

protocols.<br />

DESIGN: Retrospective review of clinical outcome in PCOS patients.<br />

MATERIALS AND METHODS: The average maternal age of the patients<br />

involved was 32.6 3 years. In IVM cycles, no gonadotropin was administered,<br />

but 10,000 IU of hCG 36 hours prior to oocyte retrieval. Oocytes were<br />

cultured in an IVM medium for 24 hours and were inseminated by ICSI. The<br />

endometrium was prepared by administering estradiol from the day of oocyte<br />

retrieval, followed by luteal support. When IVM cycles were not successful,<br />

same patients underwent controlled ovarian stimulation (COS) for harvesting<br />

mature oocytes. Embryological parameters, clinical outcome and complications<br />

were compared between IVM and COS cycles.<br />

RESULTS: A total of 20 consenting PCOS patients underwent 22 IVM<br />

treatment cycles. Following unsuccessful IVM treatment 14 patients underwent<br />

24 COS cycles. The average number of retrieved oocytes in IVM cycles<br />

was 16.9 7, with the maturation and fertilization rate being 41.7% and<br />

49.0%, respectively, while these were 10.4 3, 80.8% and 52.7%, respectively<br />

in COS cycles. IVM yielded more oocytes; however, the number of<br />

fertilized oocytes was similar between the two. Only mild or moderate<br />

form of OHSS was observed in 15 out of 24 (62.5%) COS cycles, while<br />

none developed in IVM cycles. Following fresh and frozen transfers, 8 out<br />

of 20 (40.0%) became pregnant in IVM; similarly, 26 transfers resulted in<br />

10 (41.7%) pregnancies in COS. Pregnancy rate per patient was 33.3% (6/<br />

20) in IVM, and 57.1% (8/14) in COS, cumulatively 70% (14/20).<br />

CONCLUSIONS: Overall clinical pregnancy rate of PCOS patients<br />

following ART was 70%. IVM completely eliminated OHSS and reduced<br />

costs while maintaining a satisfactory pregnancy rate. Thus, IVM can be<br />

offered to PCOS patients undergoing ART as a first line treatment.<br />

P-58 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ANTI MULLERIAN HORMONE (AMH) LEVELS PREDICT CAR-<br />

DIOVASCULAR RISK ASSESSMENT IN YOUNG WOMEN WITH<br />

PCOS. R. Feldman, a,b S. Butts, c A. Dokras. a,b a University of Pennsylvania,<br />

Philadelphia, PA; b Obstetrics and Gynecology, University of Pennsylvania,<br />

Philadelphia, PA; c Perelman School of Medicine, Philadelphia, PA.<br />

OBJECTIVE: Serum levels of anti-mullerian hormone (AMH), a member<br />

of the TGFb family, are elevated in women with PCOS. There is conflicting<br />

information in women with and without PCOS regarding the association between<br />

low AMH levels and increased cardiometabolic risk. Young women<br />

with PCOS have an increased risk of metabolic syndrome and we hypothesized<br />

that serum AMH levels would predict risk of metabolic syndrome in<br />

this population.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Women seen at the Penn PCOS center<br />

from 2010-2014 with complete metabolic work up were included in this<br />

study (n¼469). Metabolic syndrome (Met Syn) was defined by modified<br />

NCEP-ATP III criteria (BMIR30, sBP/dBPR135/85mmHg or taking antihypertensives<br />

medications, fasting glucose R100mg/ml or taking medications<br />

for diabetes, TGR150mg/ml, HDL-C %50mg/ml). AMH was<br />

measured using the Gen II ELISA, lipids were measured by standard enzymatic<br />

methods and total testosterone was measured by radioimmunoassay.<br />

Spearman correlation coefficients were used to determine relationships between<br />

continuous variables. Linear regression was used to model associations<br />

between AMH and lipids adjusting for confounders. AMH was<br />

dichotomized based on tertiles and univariate tests and logistic regression<br />

modeling were used to evaluate associations of selected variables with tertiles<br />

of AMH.<br />

RESULTS: The median AMH level in the entire cohort was 5.06ng/ml<br />

(IQR 3.08-7.97) . The mean age of the group was 27.66years, 42.1%<br />

were non-white and 8.1% were smokers. The overall prevalence of Met<br />

Syn was 19%. AMH levels positively correlated with total testosterone<br />

(p


ovarian response to gonadotropin stimulation, cannot be a definitive marker<br />

for OHSS. In addition to the desired cohort of leading follicles, gonadotropins<br />

also tend to recruit several AMH producing secondary upcoming follicles<br />

in the mid-follicular phase (around day7) of the stimulation cycle.<br />

Therefore, it would be more prudent to estimate serum AMH level on day<br />

7 of stimulation to get a forewarning of the risk of OHSS in that particular<br />

cycle.<br />

DESIGN: Prospective study of PCOS women (n¼240) undergoing IVF/<br />

ICSI procedure using standard protocol involving COH with 225 IU daily<br />

of recombinant FSH and antagonist.<br />

MATERIALS AND METHODS: Day of commencement of stimulation<br />

protocol was considered day 1. After six days of ovarian stimulation, overnight<br />

fasting serum AMH on day seven (D7) was measured. Antral follicle<br />

count and follicular development was continuously monitored by TVS<br />

from day 7 until at least R 3 lead follicles reached 18 mm diameter size<br />

on attainment of which injection hCG (5000IU) was administered. Serum<br />

Estradiol levels were measured on day of hCG administration. Oocyte<br />

retrieval was done 34 - 36 hours of injection hCG. PCOS women were classified<br />

on the basis of dhCG serum E2 level. group A: no OHSS: E2 8000pg/<br />

ml) was observed in this study. Embryo transfer (ET) was done in all women<br />

except 12 patients of group C.<br />

RESULTS: There was parity in all women w.r.t. age, infertility period,<br />

body mass index, and waist: hip ratio. No significant difference was found<br />

in baseline serum AMH (p¼0.09), and AFC on day1 of stimulation (p¼<br />

0.11) respectively. However, compared to group A, D7 AMH level was<br />

significantly higher in group B (p ¼ 0.0024) and group C (p¼0.00<strong>21</strong>). There<br />

was also a highly significant difference in D7 AMH levels between group B<br />

and group C (p¼0.0002). Interestingly, the 12 patients in this study where no<br />

ET was done had significantly elevated D7 AMH level compared to the other<br />

30 women of the same group i.e. group C (p


aimed to investigate the diagnostic value of the levels of Progranulin in the<br />

clinical setting of PCOS, and its metabolic effects.<br />

DESIGN: Prospective case-control study.<br />

MATERIALS AND METHODS: Forty-one adolescent patients with<br />

PCOS and 39 adolescent patients as a control group were recruited for<br />

this study in a tertiary referral center in the capital of Turkey. The diagnosis<br />

of PCOS was dependent upon the recent Amsterdam ESHRE/ASRM proposal<br />

and the presence of all three of the Rotterdam criteria for diagnosing<br />

PCOS in adolescents was required (1). Progranulin levels and individual<br />

characteristics of the two groups were compared. Indices of insulin sensitivity,<br />

metabolic variables, circulating androgen levels, lipidemic markers<br />

were measured and blood pressures (BP) were also assessed. To diagnose<br />

the cases with metabolic syndrome (MetS), Cook/Ford modified criteria<br />

were used and patients who had at least 3 of the 5 criteria were diagnosed<br />

with MetS (2).<br />

RESULTS: Progranulin levels in patients with PCOS (7.481.93 ng/mL)<br />

were significantly higher than in the age-BMI matched controls (6.251.98<br />

ng/mL) (p ¼ 0.006). Luteinizing hormone (LH) levels, LH / Follicle stimulating<br />

hormone (FSH) ratios, free testosterone, dehydroepiandrosterone<br />

sulfate (DHEA-S), C-reactive protein (CRP) levels in patients with PCOS<br />

were significantly higher than in the control group (p 0.05). The MetS<br />

was present in 8 (19.5%) of the patients in the study group and in 1<br />

(2.3%) of the patients in the control group (p ¼ 0.029). There was significant<br />

inverse correlation between serum High-density lipoprotein cholesterole<br />

(HDL-C) and Progranulin levels in patients diagnosed with PCOS (p ¼<br />

0.008).<br />

CONCLUSIONS: The MetS is more common among the adolescent patients<br />

with PCOS. There was a significant association between Progranulin<br />

and HDL-C levels. Progranulin may be a novel biomarker for cardiovascular<br />

risk in patients with PCOS, thus these cases should be directed to close<br />

follow-up for possible cardiovascular diseases. Future larger studies should<br />

focus on this entity, thus providing a further vantage point into the role of Progranulin<br />

in the pathogenesis of PCOS.<br />

Demographic and laboratorial features of the subjects.<br />

Variables<br />

PCOS group<br />

(n:41)<br />

Control<br />

group (n:39)<br />

p value<br />

Age (years) 18.92.3 18.72.5 0.629<br />

BMI(kg/m2) 22.73.1 22.22.3 0.440<br />

Waist to Hip<br />

0.800.08 0.790.07 0.436<br />

Circumference Ratio<br />

Progranulin (ng/mL) 7.481.93 6.251.98 0.006<br />

LH/FSH 1.30.6 1.00.6 0.040<br />

Free Testosterone (pg/mL) 1.5 (0.8-4.0) 1.2 (0.6-2.7) 0.014<br />

DHEA-S (mg/dL) 329.811.3 259.0101.8 0.004<br />

CRP (mg/L) 2.2 (0.5-9.9) 1.9 (0.6-6.9) 0.025<br />

natural killer cells and they have been shown to play important roles in cardiovascular<br />

diseases (CVD) and diabetes. We aimed to determine serum perforin<br />

and granzyme B levels in adolescent polycystic ovary syndrome<br />

(PCOS) patients, and to investigate whether they are associated with insulin<br />

sensitivity, obesity and CVD risk markers.<br />

DESIGN: Case-control study.<br />

MATERIALS AND METHODS: A total of <strong>17</strong>2 adolescents (83 PCOS patients<br />

and 89 healthy controls) were recruited consecutively. Homeostasis<br />

model assessment (HOMA-IR), lipid parameters, anthropometric measurements<br />

were determined. Serum perforin and granzyme B levels were<br />

measured by commercially available ELISA kits. HOMA-IR>2.5 was<br />

considered to indicate the presence of insulin resistance. Multiple Logistic<br />

Regression Analyses were applied for each clinical condition after adjustment<br />

for body mass index.<br />

RESULTS: Adolescents with PCOS had significantly higher levels of fasting<br />

glucose, insulin, HOMA-IR, and total cholesterol/HDL ratio when<br />

compared with controls. Granzyme B levels were also significantly higher<br />

in PCOS patients (median 3.09pg/mL vs 2.<strong>21</strong> pg/mL, P¼0.002). No statistically<br />

significant differences were determined among serum perforin levels.<br />

According to the ROC analysis performed for the diagnostic performance<br />

of granzyme B levels in PCOS patients for body mass index (BMI) of ><br />

25 kg/m2, waist circumference (WC) >80 cm, waist to hip ratio (WHR)><br />

0.8 and HOMA-IR >2.5, the AUC values were all statistically significant.<br />

The best granzyme B cut-off values for distinguishing the groups, and also<br />

the sensitivity, specifity, positive and negative predictive values are given<br />

in Table. Multivariate logistic regression analysis was then used to determine<br />

if a relationship between groups and the defined cut-off levels of granzyme B<br />

was present. Granzyme B levels were found to be independently associated<br />

with increased WHR (OR¼4.256, 95%CI: 1.550-11.687, p¼0.005) and<br />

HOMA-IR (OR¼46.748, 95%CI: 12.431-<strong>17</strong>5.801, p2.06 85.4% 50.0% 59.3% 80.0%<br />

Waist circumference >2.06 80.4% 52.6% 69.5% 66.7%<br />

R80 cm<br />

Waist-to-hip ratio >1.865 84.3% 47.4% 68.3% 69.2%<br />

R0.80<br />

HOMA-IR >2.5 >3.595 87.5% 89.5% 82.4% 92.7%<br />

References:<br />

1. Fauser BC, Tarlatzis BC, Rebar RW, Legro RS, Balen AH, Lobo R, et<br />

al. Consensus on women’s health aspects of polycystic ovary syndrome<br />

(PCOS): the Amsterdam ESHRE/ASRM-Sponsored 3rd PCOS<br />

Consensus Workshop Group. Fertil Steril 2012;97:28-38.<br />

2. Cook S, Auinger P, Li C, Ford ES. Metabolic syndrome rates in United<br />

States adolescents, from the National Health and Nutrition Examination<br />

Survey, 1999-2002. J Pediatr. 2008;152(2):165-70.<br />

P-63 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

INCREASED GRANZYME B LEVELS ARE ASSOCIATED WITH IN-<br />

SULIN RESISTANCE IN ADOLESCENT POLYCYSTIC OVARY<br />

PATIENTS. E. Oztas, a S. Ozler, b A. Tokmak, a N. Yilmaz, c<br />

N. Danisman, a M. Ergin, d H. I. Yakut. a a Zekai Tahir Burak Women’s Health<br />

Education and Res, Ankara, Turkey; b Obstetrician, Perinatolog, Ankara,<br />

Turkey; c Reproductive Endocrinology Department, ZTB, Ankara, Turkey;<br />

d Ataturk Training and Research Hospital, Ankara, Turkey.<br />

OBJECTIVE: Perforin, pore-forming protein and granzyme B, a serin protease<br />

are stored in secretory granules inside the cytotoxic T lymphocytes and<br />

P-64 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

CHOLESTEROL HAS ASSOCIATION WITH BIRTH WEIGHT IN<br />

POLYCYSTIC OVARIAN SYNDROME. A. K. Ahmad, a C. Kao, b<br />

M. Cedars, b H. Huddleston. c a Obstetrics, Gynecology & Reproductive Sciences,<br />

University of California San Francisco, San Francisco, CA; b University<br />

of California San Francisco, San Francisco, CA; c UCSF, Woodside, CA.<br />

OBJECTIVE: Low birth weight (LBW) has been reported to associate<br />

with insulin resistance and increased cardiovascular risk later in life. We assessed<br />

whether birth weight had an association with the development of insulin<br />

resistance and cardiovascular risk in a polycystic ovarian syndrome<br />

(PCOS) population.<br />

DESIGN: Cross sectional.<br />

MATERIALS AND METHODS: Women seen in a multidisciplinary<br />

PCOS clinic from 2006 to 2014 who were diagnosed with PCOS by 2003<br />

Rotterdam criteria were consented and considered for inclusion. PCOS patients<br />

who were born pre-term were excluded from the study population.<br />

Data were systematically collected and included the following metabolic<br />

FERTILITY & STERILITY Ò<br />

e127


and cardiovascular parameters: body mass index (BMI), waist circumference,<br />

cholesterol, low density lipoprotein (LDL), high density lipoprotein<br />

(HDL), triglycerides (TG), aspartate aminotransferase (AST), alanine<br />

aminotransferase (ALT), C-reactive protein, fasting & two-hour insulin, fasting<br />

& two-hour glucose, homeostasis model assessment of insulin resistance<br />

(HOMA-IR), systolic and diastolic blood pressure (SBP, DBP). Birth weight<br />

was analyzed as a predictor of metabolic and cardiovascular parameters as a<br />

continuous variable using multiple and simple regression. Logistic regression<br />

was used to assess the relationship between biochemical hyperandrogenism<br />

(yes or no) and BW as a continuous variable. P-value was set to 0.05. BMI<br />

and age were controlled for as confounding factors.<br />

RESULTS: 230 full-term PCOS patients were identified with mean birth<br />

weight of 3367g (SD 497.5) and mean age at evaluation of 28.43years<br />

(SD 5.64). Of metabolic parameters reviewed, BW was a significantly associated<br />

with cholesterol (-0.0148, p¼0.0043), meaning for each 100g decrease<br />

in BW, cholesterol would increase by 1.48. There was a trend for an association<br />

of BW on hyperandrogenism (OR 0.9995, p¼0.0698).<br />

CONCLUSIONS: There was a significant association between low birth<br />

weight and higher cholesterol levels. In a population already at increased cardiovascular<br />

risk, PCOS patients with LBW should be followed closely with<br />

more frequent screening.<br />

BW as a predictor of metabolic & cardiovascular parameters.<br />

Metabolic &<br />

Cardiovascular<br />

Parameters<br />

Simple<br />

Regression<br />

Coefficient<br />

P-value<br />

Multiple<br />

Regression<br />

Coefficient<br />

(Controlling<br />

for BMI & Age)<br />

P-value<br />

BMI -0.0014 0.<strong>21</strong>90 — —<br />

Cholesterol -0.0183 0.0005 -0.0148 0.0043<br />

LDL -0.0085 0.0564 -0.0062 0.1523<br />

HDL -0.0024 0.2998 -0.0028 0.1873<br />

TG -0.0130 0.1645 -0.0067 0.4475<br />

CRP -0.0001 0.8994 0.0006 0.5751<br />

Fasting Insulin -0.0007 0.8664 0.0010 0.7764<br />

Fasting Glucose -0.0027 0.1677 -0.0016 0.4092<br />

Two-hour Glucose -0.0122 0.0314 -0.0082 0.1<strong>21</strong>8<br />

HOMA-IR -0.0004 0.7100 0.0003 0.7681<br />

DBP 0.0004 0.8077 0.0014 0.3291<br />

P-65 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

PREDICTIVE MARKERS OF ABNORMAL GLUCOSE INTOLER-<br />

ANCE IN POLYCYSTIC OVARY SYNDROME. D. Lee, J. Jeon,<br />

S. Park, H. Chung, K. Jeong. Department of Obstetrics and Gynecology,<br />

School of Medicine, Ewha Womans University, Seoul, Korea, Republic of.<br />

OBJECTIVE: To identify predictive markers of abnormal glucose metabolism<br />

(IGT/NIDDM) in Korean women with polycystic ovary syndrome(P-<br />

COS).<br />

DESIGN: Prospective cohort study.<br />

MATERIALS AND METHODS: A total of 312 PCOS women were evaluated.<br />

All patients underwent 75g-oral glucose tolerance tests (OGTTs). The<br />

2 hour plasma glucose level was used to categorize subjects as IGT or<br />

NIDDM. Areas under the curve (AUC) of the receiver operating characteristic<br />

curves (ROC) were used to compare the power of serum markers. Multiple<br />

linear regression analysis was used to evaluate the contribution of each<br />

confounding factor for 2hr post-load glucose value.<br />

RESULTS: Two hundred eighty five (91.3%) PCOS women with normal<br />

glucose tolerance and 27 (8.7%) PCOS women with abnormal glucose metabolism<br />

(IGT/NIDDM) were evaluated in this study. AUC of HbA1c, high<br />

sensitivity CRP, lipid accumulation product (LAP) index, and triglycerides<br />

were 0.780, 0.772, 0.762, and 0.758 respectively. ROC analysis suggested<br />

a threshold value of 5.45 in HbA1c (71.4% sensitivity and 70.0% specificity),<br />

a value of 1.16 in high sensitivity CRP (70.3% sensitivity and 80.1% specificity),<br />

a value of 12.98 in LAP index (88.5% sensitivity and 52.3% specificity)<br />

and a value of 88.0 in triglycerides (77.8% sensitivity and 63.5%<br />

specificity) for prediction of abnormal glucose metabolism. However,<br />

glucose to insulin ratio (G:I ratio) and quantitative insulin sensitivity check<br />

index (QUICKI) could not replace the role of OGTT in screening of IGT<br />

and NIDDM.<br />

CONCLUSIONS: HbA1C, high sensitivity CRP, LAP index, and triglycerides<br />

can be useful predictive markers of abnormal glucose metabolism<br />

(IGT/NIDDM) in Korean PCOS women.<br />

Comparison of areas under the ROC curves(95% CI) for potential markers<br />

for normal and abnormal(IGT/T).<br />

Areas under<br />

ROC(95%-CI)<br />

Supported by: Korea Centers for Disease Control and Prevention.<br />

P-66 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ABERRANT EXPRESSION OF MICRORNAS IN CUMULUS CELLS<br />

ISOLATED FROM PCOS PATIENTS. X. Huang C. Hao. Reproductive<br />

Medicine Centre, Affiliated Hospital of Qingdao Medical University, Yantai<br />

Yuhuangding Hospital, Yantai, China.<br />

OBJECTIVE: Recent data has pointed to the importance of miRNAs in<br />

disease and embryo development. The developmental competence of oocytes<br />

and embryos in PCOS patients is reduced to a certain extent during the in vitro<br />

fertilization (IVF) process. Cross-talk between the oocyte and cumulus<br />

cells is critical for oocyte maturation and embryo competence. The aim of<br />

this study is to determine if microRNAs are differentially expressed in the<br />

cumulus cells of PCOS patients compared to non-PCOS patients and identify<br />

associated altered gene expression.<br />

DESIGN: Descriptive study.<br />

MATERIALS AND METHODS: MicroRNA analysis was performed using<br />

Human miRNA Microarray Release 18.0 microarray between cumulus<br />

cells isolated from PCOS (n¼5) and non-PCOS (n¼6) patients. Several miR-<br />

NAs were selected to validate the microarray results using quantitative RT-<br />

PCR on additional samples. Potential gene targets were identified and their<br />

expression analyzed by quantitative RT-PCR.<br />

RESULTS: The microarray profiling of cumulus cells revealed expression<br />

of 386 miRNAs, <strong>17</strong> of them were differentially expressed between the two<br />

groups. Annotation of predicted gene targets for these differentially expressed<br />

miRNAs included genes involved in several pathways as focal adhesion,<br />

regulation of actin cytoskeleton, MAPK signaling pathway, and Wnt<br />

signaling pathway. Several predicted miRNA target genes were selected<br />

for analysis and demonstrated significant altered expression consistent<br />

with aberrant miRNA profiles.<br />

CONCLUSIONS: This study describes for the first time that aberration of<br />

cumulus cells microRNA expression is associated with PCOS. With growing<br />

evidence indicating the importance of miRNAs during oocyte development<br />

of PCOS patients, new molecular biomarkers or a new basis for the diagnosis<br />

of PCOS would be provided and the etiology of PCOS might be improved.<br />

Supported by: This study was Supported by the National Natural Science<br />

Foundation of China (Grant 81401<strong>17</strong>2 and 81<strong>17</strong>0622) and the Natural Science<br />

Foundation of Shandong Province (Grant ZR2013HQ004).<br />

P-67 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

p-value<br />

(asymptotic<br />

significance<br />

Cut-off<br />

(calculated by Sensitivity Specificity<br />

of area¼0.5) Youden’s index) (at cut-off) (at cut-off)<br />

HbA1C 0.780(0.66-0.89)


OBJECTIVE: The benefit of choosing GnRH agonist versus antagonist in<br />

women with PCOS has yet to be determined. The existing literature does not<br />

reach a firm consensus. The objective of this study was to compare reproductive<br />

outcomes between agonist and antagonist protocols in PCOS patients undergoing<br />

IVF.<br />

DESIGN: Retrospective cohort study in an academic setting.<br />

MATERIALS AND METHODS: Women undergoing IVF from 2007 to<br />

<strong>2015</strong> and meeting Rotterdam PCOS criteria were included. Data regarding<br />

baseline demographics, clinical parameters, and treatment outcomes was<br />

collected. Provider choice determined participants’ IVF protocol. Descriptive<br />

statistics were computed. Bivariate analysis was performed using Student’s<br />

t, Mann-Whitney U, Pearson’s chi-squared, and Fisher’s exact tests<br />

as appropriate. Multivariate analysis was completed by stratifying outcomes<br />

by demographic and baseline clinical parameters when p


explore the relationship between sRAGE and PCOS and the relationship between<br />

sRAGE and IVF-ET Results in PCOS.<br />

DESIGN: Prospective study.<br />

MATERIALS AND METHODS: A total of 80 patients of childbearing age<br />

were observed and 30 PCOS and 28 non-PCOS women were enrolled into the<br />

final analyses for parameters of IVF-ET results. Measurements of sRAGE,<br />

VEGF in plasma at early follicular phase and leading follicular fluid aspirated<br />

without blood at the time of egg collection by ELISA to examine the<br />

difference and the correlation. Parameters of IVF-ET results including the<br />

number of the oocyte retrieved, the number of the mature follicles, the normal<br />

fertilization rates, and the high-quality embryos.<br />

RESULTS: 1 Compared with the non PCOS group, the concentration of<br />

sRAGE in plasma and follicular fluid in the PCOS group both were significantly<br />

higher (plasma, 1351.79290.62 vs 1185.00205.71, t¼2.148,<br />

P¼0.046. follicle fluid, 1650.23239.711 vs 1034.18156.010, t¼11.510,<br />

P¼0.000). In PCOS group, the follicular fluid concentration of sRAGE<br />

was positively related to the VEGF (r¼0.988, P¼0.000). 2 In PCOS, the concentration<br />

of sRAGE in follicular fluid has no relationship with the number of<br />

the oocyte retrieved(r¼0.375 ,p¼0.103) and the number of the mature follicles<br />

(r¼0.409,P¼0.074), but it was positively related to the IVF-ET parameters<br />

including the normal fertilization rates(r¼0.540,P¼0.014), normal<br />

cleavage rates(r¼0.489,P¼0.029)and the high-quality embryos(r¼0.539,<br />

P¼0.014 ) when the age,BMI,T and LH/FSH were adjusted.<br />

CONCLUSIONS: sRAGE may affect the development of inflammatory<br />

factor generated in PCOS. The concentration of sRAGE in the follicle fluid<br />

was positively related to the IVF-ET parameters,and could be a biomarker to<br />

predicate IVF-ET results with PCOS.<br />

P-71 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE CORRELATION BETWEEN AMH AND LABORATORY PA-<br />

RAMETERS IN PCOS WOMEN ACCORDING TO SUBTYPE: A PI-<br />

LOT-STUDY. Y. Kim, S. Lee, K. Yi, H. Park, J. Shin, T. Kim,<br />

J. Hur. Department of Obstetrics and Gynecology, Korea University Guro<br />

Hospital, SEOUL, Korea, Republic of.<br />

OBJECTIVE: Do AMH level correlate with hormonal and metabolic parameters<br />

in women with PCOS according to subtype?<br />

DESIGN: A university hospital-based observational study.<br />

MATERIALS AND METHODS: The observational study of women with<br />

PCOS between January 2013 and November 2014 in University-based medical<br />

center. Seventy of women with PCOS were classified according to insulin<br />

resistance(IR) by the homeostasis model assessment-estimated insulin<br />

resistance(HOMA-IR) or androgen excess(AE) by free androgen index(FAI).<br />

The following parameters were analyzed; age, BMI, basal and 2 hour-post<br />

prandial(PP2) insulin level, basal and PP2 glucose, HbA1c, testosterone,<br />

DHEA, SHBG, AMH, cholesterol, triglyceride, LDL, HDL, and CRP. Between<br />

IR(-, n¼33) and IR(+, n¼37), and AE(-, n¼25) and AE(+, n¼45),<br />

the correlation between AMH and analyzed parameters were investigated.<br />

RESULTS: In overall analysis, AMH had significant positive correlation<br />

with SHBG (r¼0.385, P¼ .001), testosterone (r¼0.349, P¼ .004), and<br />

HDL (r¼0.340, P¼ .006). In IR (+) type, AMH showed correlation positive<br />

with testosterone (r¼0.337, P¼ .041), HDL (r¼0.486, P¼.003) and negative<br />

with CRP (r¼-0.376, P¼ .024), basal and PP2 glucose (r¼-0.343, P¼ .037;<br />

r¼-0.347, P¼.035) whereas not in IR (-) type. In AE (+) type, AMH showed<br />

correlation positive with HDL (r¼0.372, P¼ .014) and negative with PP2<br />

glucose (r¼-0.303, P¼ .038) although not in AE (-) type.<br />

CONCLUSIONS: AMH may show positive correlation with SHBG and<br />

testosterone regardless of PCOS subtype, and have different correlation<br />

with metabolic parameters according to IR and AE. To clarification of its<br />

clinical meanings, further study with large scale could be necessary.<br />

DESIGN: Cross-sectional study with control group.<br />

MATERIALS AND METHODS: 242 patients with PCOS (Rotterdam<br />

criteria, 2003) aged 16-34 years and 166 women (control group) aged 16-<br />

35 years were studied between 01.01.12 and 07.31.14.Prior institutional<br />

ethical approval and informed written consent (from all participants/ legal<br />

guardians) were obtained. Body mass index (BMI; Kg/M2), abdominal<br />

circumference (AC, cm), waist-hip ratio (WHR), Ferriman Galwey score,<br />

presence of acne (%) and acanthosis nigricans (AN, %) and blood pressure<br />

were assessed. Serum testosterone (T), sex hormone binding globulin<br />

(SHBG), post-prandial serum glucose (PPG) and insulin (PPI) levels 2 hours<br />

after 75 glucose intake were measured. FAI was calculated as [T (ng/ml) x<br />

100 x 3.47] / SHBG (nmol/l). PPG R 140 mg% was considered as AGT.<br />

PPG: PPI % 1.0 was considered as indicative of IR 1.<br />

RESULTS: Prevalence of AGT was 11.9% (95% CI ¼ 7.9%, 16.1%)<br />

among PCOS patients and 12.6% (95% CI ¼ 7.6%, <strong>17</strong>.7%) in the control<br />

group (not significantly different). PCOS group was younger and had higher<br />

BMI and AC. In regression analysis, after controlling for the covariates,<br />

PCOS status was not associated with AGT in our study population; BMI<br />

(p¼ 0.011) and WHR (p¼0.027) were the only significant positive predictors<br />

(for AGT). Among the PCOS patient population (n¼242), univariate analysis<br />

demonstrated that those with IR were significantly younger (p¼0.003) with<br />

higher BMI (p


P-74 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

EXCESSIVE BODILY RETENTION OF ORGANOCHLORINE<br />

PESTICIDE IS ASSOCIATED WITH ENERGY IMBALANCE AND<br />

INFLAMMATION IN WOMEN WITH PCOS: A CASE CONTROL<br />

STUDY. Y. Zhao. Peking University Third Hospital, Beijing, China.<br />

OBJECTIVE: Our aim was to investigate the relationship of the serum<br />

levels of organochlorine pesticides (OCPs) and energy homeostasis factor<br />

adropin with the risk of polycystic ovary syndrome (PCOS), as well as<br />

explore the association between bodily retention of OCPs and metabolic homeostasis<br />

and inflammation of PCOS patients.<br />

DESIGN: This was a preliminary case-control study undertaken at the Division<br />

of Reproductive Center, Peking University Third Hospital. A total of<br />

50 participants affected by PCOS and 30 normal controls were recruited between<br />

August and <strong>October</strong> 2012 from Northern China. All participants were<br />

Han women.<br />

MATERIALS AND METHODS: PCOS participants were diagnosed according<br />

to the 2003 Rotterdam criteria. The control participants were nonpregnant<br />

females unable to conceive solely due to male azoospermia. Serum<br />

levels of three typical OCPs were analyzed using gas chromatographic mass<br />

spectrometry (GC-MS), including p,p’-dichlorodiphenyldichloroethylene<br />

(p,p’-DDE), hexachlorobenzene (HCB), and b-hexachlorocyclohexane (b-<br />

HCH). Serum levels of adropin and C-reactive protein (CRP) were measured<br />

by ELISA kits (R&D System).<br />

RESULTS: Serum levels of three OCPs were significantly higher in the<br />

PCOS group than the control group. After adjustment for age and BMI, concentrations<br />

of p,p’-DDE and the sum of all OCPs ( P OCPs) in serum above<br />

median levels remained significantly associated with PCOS, and ORs were<br />

2.94 [95% confidence interval (CI), 1.11-7.78] and 5.03 [95% confidence interval<br />

(CI), 1.79-14.11], respectively. Partial least-squares-discriminant analysis<br />

(PLS-DA) confirmed that serum levels of OCPs were risk factors of<br />

PCOS, especially for p,p’-DDE. Moreover, dramatically reduced serum level<br />

of adropin and enhanced serum level of CRP was detected in PCOS group<br />

compared with the control group, indicating the energy metabolism imbalance<br />

and inflammatory status in PCOS women. Furthermore, serum levels<br />

of adropin correlated negatively with p,p’-DDE (R¼0.<strong>21</strong>9, P¼0.035), b-<br />

HCH (R¼0.450,P


similar cycle outcomes can be achieved in comparison to their normal baseline<br />

E2 counterparts.<br />

DESIGN: Retrospective analysis.<br />

MATERIALS AND METHODS: 110 patients undergoing oocyte cryopreservation<br />

or IVF at our center from 2013-2014 were included. Estrogen<br />

prime patients were excluded. Patients were separated into two groups.<br />

Group 1 included 31 patients with E2 levels greater than 70 pg/ml on day<br />

2 with subsequent ganirelix acetate given for 2-3 days and repeat E2 levels<br />

of less than 70 pg/ml. Group2 included 79 patients with standard day 2 E2<br />

levels of less than 70 pg/ml whom initiated their cycles immediately without<br />

the use of ganirelix acetate. Groups were compared with t-tests of the means<br />

as well as chi-squared analysis for cycle outcomes.<br />

RESULTS: Group 1 was composed of 12 egg freeze patients and 19 IVF<br />

patients. Group 2 was composed of 1 egg freeze patient and 78 IVF patients.<br />

Table 1 shows the difference in baseline characteristics between group 1 and<br />

group 2. Specifically, there were no significant differences seen in age, FSH<br />

levels, as well as historic high FSH levels and AMH levels. Day 2 E2 levels<br />

were significantly different between the groups, as seen in table 1. There was<br />

also no difference in cycle stimulation parameters. Group 1 had an average<br />

egg yield of 9.2, as opposed to 14.6 in group 2. This was statistically significant<br />

with a p-value of 0.01. Of patients who underwent trophoectoderm (TE)<br />

biopsy (53 patients in both groups), group 1 had 33% (8/24 total biopsied embryos)<br />

euploid embryos as compared to group 2 which had 36.6% euploid<br />

embryos (91/249 total biopsied embryos). This was not significant, p-value<br />

of 0.75. A total of 6 transfers in group 1 and 25 transfers in group 2 were performed.<br />

66.67% of patients in group 1 had a positive beta HCG after transfer<br />

as opposed to 76% in group 2 which was not significant, p-value of 0.49.<br />

Clinical pregnancy rates in group 1 were 66.67% in group 1 and 72% in group<br />

2, also not significant, p-value of 0.58.<br />

CONCLUSIONS: Patients with high baseline E2 levels on day 2 of cycle<br />

demonstrate lower egg yields despite adequate suppression of E2 with ganirelix<br />

acetate however, this did not translate to lower clinical pregnancy rates or<br />

lower euploid embryo rates. Given the lower egg yields in the high day 2 E2<br />

group this may be of clinical importance in oocyte cryopreservation cycles.<br />

Cycle characteristics and outcomes.<br />

Group 1<br />

(E2>70 pg/ml)<br />

Group 2<br />

(E2)<br />

p-value (*all t-tests<br />

of the mean,<br />

**chi-square)<br />

Age (years) 36 (+/- 6.2) 37 (+/- 4.9) 0.22*<br />

Day 2 E2 (pg/ml) 106 (+/- 45) 44 (+/- 14.3)


ulation but is a relatively poor test for prediction of pregnancy and live birth.<br />

Patients with low levels of AMH still can achieve reasonable treatment outcomes<br />

and should not be precluded from attempting IVF solely on the basis<br />

of an AMH value.<br />

Supported by: Natural Science Foundation of Guangdong Province (Grant<br />

No. S2013040013849).<br />

P-80 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE EFFECT OF POSTPARTUM PROGESTOGEN ON CARDIAC<br />

FUNCTION, NT-PROBNP, AND INFLAMMATORY BIOMARKERS<br />

IN WOMEN WITH GESTATIONAL HYPERTENSION. J. S. Yeh, a<br />

M. Daubert, b M. Kuchibhatla, c F. Z. Stanczyk, d T. M. Price. a a Department<br />

of Obstetrics & Gynecology, Division of Reproductive Endocrinology &<br />

Infertility, Duke University Medical Center, Durham, NC; b Department of<br />

Medicine, Division of Cardiology, Duke University Medical Center,<br />

Durham, NC; c Department of Biostatistics & Bioinformatics, Duke University<br />

Medical Center, Durham, NC; d Department of Obstetrics and Gynecology,<br />

University of Southern California, Los Angeles, CA.<br />

OBJECTIVE: The discovery of a mitochondrial progesterone receptor<br />

(PR-M) established a mechanism by which progesterone increases cardiac<br />

cellular respiration. Because human myocardium contains the most mitochondria<br />

per volume than any other tissue, progesterone may have a role<br />

in increasing energy production to meet the greater metabolic demands of<br />

pregnancy. Furthermore, transgenic mice over-expressing PR-M show<br />

higher levels of cardiac cellular respiration and are less susceptible to<br />

heart failure after aortic constriction. Therefore, our primary objective<br />

was to investigate if a progestogen can affect human heart function by<br />

determining if postpartum (PP) administration of Depot medroxyprogesterone<br />

acetate (DMPA) has an effect on cardiac performance, NT-ProBNP,<br />

and/or inflammatory biomarkers in women with gestational hypertension<br />

(GH).<br />

DESIGN: Prospective longitudinal cohort pilot study.<br />

MATERIALS AND METHODS: Ten patients with GH participated in<br />

this study. The DMPA group (DG) consisted of five subjects given<br />

DMPA for contraception on PP day 1, while the control group (CG)<br />

received no exogenous hormones. Subjects underwent echocardiograms<br />

and serum collection for IL-6, TNFa, and NT-ProBNP at three testing sessions<br />

(TS): 2 weeks prior to delivery, 1 day PP, and 2 weeks PP. Absolute<br />

values, net change and percent change were calculated for and between<br />

each TS. Data were analyzed using Student’s t-test, Mann-Whitney U<br />

test, and Spearman’s correlations.<br />

RESULTS: Subjects completed the three TS at an average of 37.5 weeks,<br />

1.0 day PP, and 15.0 days PP. All patients showed evidence of at least mild<br />

left ventricular (LV) hypertrophy by the first TS. By the third TS, the DG<br />

had significantly better diastolic function than the CG, with function positively<br />

correlated to serum levels of medroxyprogesterone acetate (Spearman’s<br />

r(3)¼1.000, p


Comparison of groups related to cycle outcomes.<br />

Group A (P4)<br />

Group B<br />

(P4>1.5 ng/ml)<br />

p-value<br />

(*t-test,<br />

**chi-sqaure)<br />

Age (years) 38.4 +/- 4.1 38.4 +/-3.7 0.89*<br />

Total Gonadotropins (IU) 3795.34 +/- 1539 4395.2 +/- 15<strong>21</strong>.7 0.02*<br />

Estradiol at trigger (pg/ml) 2483.8 +/- 1048.1 3034.2 +/- 1495.3 0.02*<br />

# Eggs retrieved 14.5 +/- 7.9 15.7 +/-7.0 0.34*<br />

# Embryos biopsied 5.0 +/- 3.97 5.2 +/-3.5 0.73*<br />

% Euploid Embryos 32.7 +/-31.5 29.4 +/- 29.2 0.51*<br />

% No diagnosis embryos 10.4 +/- 26.4 11.2 +/-26.4 0.86*<br />

% No normal embryos 32.1 (26-38.9) 25.5 (15.1-39.6) 0.19**<br />

Clinical pregnancy rate (%) 72 (62.1-79.5) 66.7 (47.7-81.5) 0.31**<br />

Implantation rate (%) 59.8 (50.6-68.4) 66.7 (47.7-81.5) 0.43**<br />

Ongoing/Live birth rate (%) 65.6 (56-74.2) 66.7 (47.7-81.5) 0.46**<br />

P-83 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

OVARIAN FUNCTION<br />

EXPRESSION OF HUMAN TELOMERASE REVERSE TRAN-<br />

SCRIPTASE (TERT) IS STAGE SPECIFIC IN THE OVARIAN<br />

FOLLICLE. J. A. Dundee, a N. Esfandiari, b M. M. Blanchette Porter, c<br />

E. McGee. d a ObGyn, University of Vermont, Burlington, VT; b ObGyn, Dartmouth<br />

Hitchcock Medical Center, Lebanon, NH; c Dartmouth Hitchcock<br />

Medical Center, Lebanon, NH; d University of Vermont, Burlington, VT.<br />

OBJECTIVE: The TERT gene encodes telomerase, the activity of which is<br />

required to maintain telomere length. Shorter telomeres and decreased telomerase<br />

activity are associated with cellular dysfunction, ovarian insufficiency<br />

and poorer ART outcomes. In this study TERT expression was<br />

evaluated in human mural granulosa cells and follicles at varying stages in<br />

ovarian tissue.<br />

DESIGN: basic science research using human cells and tissues.<br />

MATERIALS AND METHODS: Mural granulosa cells were isolated<br />

from follicular fluid pooled from women undergoing IVF. Samples<br />

were used to determine expression of mRNA encoding TERT and c-<br />

Myc in both fresh and cultured granulosa cells. Immunohistochemistry<br />

was performed using a TERT antibody on paraffin sections from four<br />

normal ovaries from a pathology tissue bank. TERT protein was also<br />

localized in cultured human granulosa cells. Telomerase activity<br />

(TRAPeze assay; Millipore) was measured in DNA extracted from fresh<br />

mural granulosa cells.<br />

RESULTS: Unexpectedly, TERT mRNA was not detected in human mural<br />

granulosa cells by PCR. To verify the samples and the technique, control<br />

studies were performed. The same technique amplified TERT from other tissues.<br />

C-Myc, which controls TERT gene transcription in breast cancer cells,<br />

did amplify in the same samples for which no TERT was amplified. To<br />

further characterize the timing and nature of TERT expression in granulosa<br />

cells, we performed IHC for TERT protein in ovarian sections from 4<br />

different women. TERT protein is expressed in the nucleus of granulosa cells<br />

from the primary stage onward, in a stage specific pattern. Expression of<br />

TERT protein is fairly uniform in granulosa cell nuclei in pre-antral and<br />

small antral follicles. Expression in large antral follicles is strongest in the<br />

granulosa cell layer adjacent to the antrum. In cultured granulosa cells,<br />

TERT protein is distributed in the cytoplasm. However, in freshly isolated<br />

granulosa cells functional TERT protein was demonstrated by measurable<br />

telomerase activity.<br />

CONCLUSIONS: The TERT message ceases to be expressed in luteal<br />

granulosa cells, though the protein is still present. This is consistent with<br />

the role of TERT to preserve DNA telomere length and the nature of granulosa<br />

cells to undergo controlled senescence. Healthy follicle growth and<br />

development early in folliculogenesis requires the elaboration of numerous<br />

generations of granulosa cells, but continued survival of granulosa cells<br />

beyond the necessary cell divisions could result in pathologic follicle cysts<br />

or tumors. On the other hand, early suppression of TERT could result in<br />

inadequate follicle growth and fewer cells able to express aromatase and<br />

produce estrogen needed for normal folliculogenesis and fertility. Further<br />

study of TERT regulation could provide important clues for conditions<br />

with aberrant folliculogenesis such as PCOS and diminished ovarian<br />

reserve.<br />

Supported by: New England Fertility Society/ Ferring REI Fellow grant.<br />

P-84 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

AKAP13 IS REQUIRED FOR NORMAL MURINE OVARIAN<br />

DEVELOPMENT. X. Wu, a,b K. Devine, c,b C. Quaglieri, b P. Driggers, a<br />

J. Segars. a a Department of Gyn/OB, Johns Hopkins School of Medicine, <strong>Baltimore</strong>,<br />

MD; b Program in Reproductive and Adult Endocrinology, NICHD,<br />

NIH, Bethesda, MD; c Shady Grove Fertility Center, Washington, DC.<br />

OBJECTIVE: Induction of aromatase by FSH is a key step in ovarian folliculogenesis.<br />

We previously reported that AKAP13 siRNA transfection was<br />

accompanied by a significant reduction in aromatase transcripts in COV434<br />

cells, suggesting a role for AKAP13 downstream of FSH. Here we sought to<br />

test whether AKAP13 might be involved in induction of aromatase and<br />

ovarian growth in vivo.<br />

DESIGN: Murine Cre/loxP conditional tissue-specific knockout model.<br />

MATERIALS AND METHODS: To examine the role of AKAP13 in<br />

ovary we used a Cre/LoxP tissue-specific knockout approach with Cre driven<br />

by Amhr2 and a floxed Akap13 gene (CKO¼Akap13Flox/Flox/Cre+). Quantitative<br />

real-time polymerase chain reaction (qRT-PCR) was used to measure<br />

akap13 mRNA in CKO ovaries. qRT-PCR was also used to quantify the aromatase<br />

transcripts and 18S rRNA was used for normalization. To assess effects<br />

on ovary size, CKO and wild type (WT) murine ovaries were freshly<br />

collected, fixed, sectioned and stained with Weigert’s iron/picric blue. Images<br />

of stained sections were imported to ImageJ and section surface area<br />

was measured. Differences between groups was tested with a student t-test<br />

or ANOVA and p


PGRMC1, PGRMC2, and PAQR7 attenuated the P4’s ability to suppress cell<br />

cycle entry. However, this was not seen with siRNA treatment for PGR.<br />

Further, it was found that PGRMC1 interacts with PGRMC2 and PAQR7,<br />

in addition to PGR in the cytoplasm; however, there was limited interaction<br />

between PGR and PGRMC1 in the nucleus, despite having a nuclear localization<br />

on immunocytochemistry. Finally, PGR was not found to interact with<br />

PGRMC2 or PAQR7.<br />

CONCLUSIONS: The present experiments are the first to reveal the<br />

complex interaction that occurs among PGRMC1, PGRMC2, and<br />

PAQR7 as well as their involvement in P4’s ability to regulate cell cycle<br />

entry. These mediators may be acting, at least in part, as a complex<br />

through a common mechanism of action; that does not involve PGR. In<br />

addition, PGR and PGRMC1 were found to interact in the cytoplasm,<br />

although what role if any this interaction plays in regulating granulosa/<br />

luteal cell function is currently unknown. Further studies are now needed<br />

to assess the downstream processes of these P4 mediators, for a more<br />

complete understanding of the potentially implicated pathways that<br />

mediate their action.<br />

P-86 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

KDM4A AND KDM4B EXPRESSION IN OVARIAN CUMULUS<br />

CELLS IN WOMEN UNDERGOING IN VITRO<br />

FERTILIZATION. F. Grimstad, a S. Krieg, b K. Roby, c K. Foster, a<br />

A. Krieg, a K. Marquis, a S. Mullinax. a a Department of Obstetrics and Gynecology,<br />

University of Kansas, Kansas City, KS; b University of Kansas Medical<br />

Center, Kansas City, KS; c Department of Anatomy and Cell Biology,<br />

University of Kansas, Kansas City, KS.<br />

OBJECTIVE: In vitro fertilization (IVF) requires superovulation to<br />

create adequate gametes for transfer. Several groups have suggested that<br />

this process may cause epigenetic changes that could have negative health<br />

outcomes for offspring. However, studies have also demonstrated epigenetic<br />

modifications are important to the process of normal gametogenesis.<br />

The study objectives were to assess expression of the histone demethylases<br />

KDM4A and KDM4B in granulosa collected from women undergoing<br />

oocyte retrieval and to determine if expression correlated with pregnancy<br />

outcome.<br />

DESIGN: Prospective Laboratory Study.<br />

MATERIALS AND METHODS: Patients undergoing IVF by a single provider<br />

were stimulated using microdose flare or antagonist protocols and<br />

monitored by serial ultrasound and estradiol levels. At time of egg retrieval<br />

mural and cumulus granulosa were collected and prepared for RNA isolation.<br />

Total RNAwas isolated using TRIzol reagent, reverse transcribed, and cDNA<br />

subjected to quantitative real-time PCR using SyBR Green with specific<br />

primers and the ABI Prism 7500 System. The comparative DCT method<br />

was used for relative quantitation of specific mRNA normalized to 18S<br />

RNA. The mean DCT was determined with relative expression (2DDCt)<br />

calculated for each patient sample compared to the overall mean. The<br />

average relative expression for each gene and cell type was compared with<br />

pregnancy outcome, significance was tested by Mann-Whitney. KDM4B protein<br />

was measured by immunostaining in sections from normal ovaries of<br />

reproductive age women. Tissues were obtained from the KUCC Biospecimen<br />

Facility.<br />

RESULTS: KDM4A and KDM4B mRNA expression was overall higher<br />

in cumulus compared to mural granulosa. When comparing cumulus granulosa<br />

demethylase gene expression, KDM4B mRNA expression was 1.8-<br />

fold higher in granulosa from not-pregnant patients compared to pregnant<br />

patients (P¼0.05). KDM4A displayed a similar trend (P¼ 0.057). In<br />

contrast, there was no significant difference in expression of KDM4A<br />

and KDM4B in mural granulosa from the same patient populations. Immunohistochemistry<br />

detected KDM4B protein in granulosa cells at all stages<br />

of follicle development with more robust expression observed in less<br />

mature follicles.<br />

CONCLUSIONS: The present studies are the first to demonstrate<br />

expression of the histone demethylases KDM4A and KDM4B in granulosa<br />

cells. Data reveal a differential expression where mRNA expression<br />

was higher in cumulus compared to mural granulosa. Expression of<br />

both KDM4A and KDM4B was lower in cumulus granulosa from<br />

pregnant compared to not-pregnant patients. Ongoing studies are<br />

exploring specific genes affected by KDM4A and KDM4B in cumulus<br />

and mural granulosa. These findings suggest that altered expression of<br />

histone demethylases may impact epigenetic changes associated with<br />

pregnancy.<br />

P-87 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

CHARACTERIZATION OF THE HUMAN OVULATORY NON-COD-<br />

ING TRANSCRIPTOME REVEALS MIRNAS AS NEW REGULA-<br />

TORS OF THE OVULATORY CASCADE. A. Hourvitz, a L. Ophir, b<br />

Y. Yung, c R. Orvieto, d G. M. Yerushalmi. e a IVF Unit, Sheba Medical Center,<br />

Ramt-Gan, Israel; b Sheba Medical Center, Petah Tiqva, Israel; c Embryology,<br />

Sheba Medical Center, Ramat Gan, Israel; d Sheba Medical Center, Ramat Gan,<br />

Israel; e Sackler Faculty of Medicine, Tel Aviv University, Tel Hashomer, Israel.<br />

OBJECTIVE: Ovarian follicular development and ovulation in mammals<br />

is a highly complex and tightly regulated process. Using whole transcriptome<br />

sequencing, we previously identified mRNAs that are differentially expressed<br />

between immature early antral follicles and mature preovulatory follicles.<br />

To complete the identification of factors involved in the ovulatory<br />

process, we aim to generate a library of global miRNAs involved in this process.<br />

Using advanced bioinformatics tools, this library will enable us to identify<br />

the ovulatory miRNA-regulated gene and networks and link them to our<br />

previously described library of mRNA ovulatory genes.<br />

DESIGN: Experimental study.<br />

MATERIALS AND METHODS: Cumulus granulosa cells (CGCs) obtained<br />

from GV cumulus-oocyte complex (COC) aspirated from IVM, MII<br />

COC aspirated from IVF and MII COC obtained from 30 h in vitro matured<br />

oocytes. Nanostring nCounter miRNA expression assay was performed in order<br />

to generate regulated miRNA library. Cross analysis between mRNA and<br />

miRNA libraries was performed using miRTrail bioinformatics tool.<br />

RESULTS: Altogether, we found 124 miRNAs expressed in at least one of the<br />

three groups tested. When comparing miRNA from CGCs of MII IVF COC<br />

(mature oocytes) and GV IVM COC (immature oocytes), we found 16 upregulated<br />

and 1 down-regulated miRNAs. Using miRTrail software, we found <strong>17</strong>24<br />

potential miRNA regulated target genes by the <strong>17</strong> differentially expressed<br />

miRNA. One hundred and forty one of these potential target genes were also<br />

found in our previously generated ovulatory gene library. Pathway analysis found<br />

4 significant KEGG pathways categories including Cell cycle, ECM-receptor<br />

interaction, Hedgehog signaling pathway and Focal adhesion. These pathways<br />

are probably of major importance for ovulation and corpus luteum formation.<br />

When comparing the miRNA expression pattern between the three groups we<br />

observed that the MII IVM COC miRNA profile was close to the GV COC<br />

miRNA profile, but differ from the MII IVF COC transcriptome. This may correspond<br />

to the observed dysregulated coding gene expression reported in IVM<br />

matured COC and may help explain the lower efficacy of IVM treatments.<br />

CONCLUSIONS: The linkage bioinformatics analysis between the libraries<br />

of the coding genes from our preliminary study with the newly generated<br />

library of regulatory miRNAs enable us to better understand the<br />

regulation of coding ovulatory genes by non-coding transcripts and their<br />

function from the single molecule level to the whole pathways.<br />

P-88 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE DETECTION AND CLINICAL UTILIZATION OF ANTI-<br />

MULLERIAN HORMONE (AMH) IN POSTMENOPAUSALWOMEN<br />

USING THE NEW HIGHLY SENSITIVE ELISA. J. Buckbinder, a<br />

S. Mucowski, b H. Burks, c F. Z. Stanczyk, c H. N. Hodis, c W. Mack, d<br />

K. A. Bendikson, c D. Shoupe. c a LAC-USC Medical Center, Los Angeles,<br />

CA; b Dallas Infertility, Frisco, TX; c USC Keck School of Medicine, Los Angeles,<br />

CA; d University of Southern California, Los Angeles, CA.<br />

OBJECTIVE: To determinewhether Anti-M€ullerian Hormone (AMH) levels<br />

as measured by a highly sensitive AMH assay have clinical utility in postmenopausal<br />

women as a marker of low levels of ovarian function that may relate to<br />

other markers of health status in aging women, such as assessments of cardiovascular<br />

and bone health. AMH has been used as a biomarker of ovarian aging<br />

but it has not been useful after onset of menopause as levels are unmeasurably<br />

low. Given that postmenopausal ovaries often contain primordial follicles histologically,<br />

use of a new assay with a very low threshold of detection (3 pg/mL) [a<br />

picoAMH ELISA from AnshLabs] may allow measurement of very low levels<br />

of AMH in menopausal women that are undetectable using earlier versions of<br />

the AMH assay (threshold of detection¼73 pg/mL).<br />

DESIGN: Secondary analysis of a prospective randomized controlled trial,<br />

the Women’s Isoflavone Soy Health Study (WISH) and measurement of<br />

AMH in serum samples of participants with intact ovaries.<br />

MATERIALS AND METHODS: Healthy postmenopausal women participating<br />

in the WISH trial without history of oophorectomy and not using hormone<br />

replacement therapy were included in the study. All participants<br />

underwent baseline venipuncture upon enrollment, and serum was stored<br />

FERTILITY & STERILITY Ò<br />

e135


at -80 C and available for analysis. The picoAMH ELISA from Ansh Labs<br />

was used to measure serum levels of AMH using these stored samples.<br />

RESULTS: Of the 133 samples, the mean age of the postmenopausal<br />

women was 57 6.9 years with a range of 45-78 yrs. The mean time from<br />

the onset of menopause was 7.9 years (range 1-31 years) with 9 of the participants<br />

only being 1 year out from menopause. The mean BMI was 26.1 5.1<br />

kg/m 2 . The ethnicity of the patients was as follows: 55.6% Caucasian, 19.6%<br />

Hispanic, 18.1% Asian or Pacific Islander, and 6.8% African American. In all<br />

133 serum samples, AMH levels were < 3 pg/ml, below the detection limit of<br />

the assay.<br />

CONCLUSIONS: Despite several participants’ recent menopausal transition,<br />

serum AMH was undetectable by a highly sensitive AMH assay in the<br />

entire cohort of postmenopausal women. Measurement of AMH in postmenopausal<br />

women is not clinically useful as levels are unmeasureable, even using<br />

a highly sensitive assay.<br />

Supported by: AnshLabs provided the assays at no cost to investigators.<br />

P-89 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE PROSTAGLANDIN TRANSPORTER (PGT): A NOVEL INDIS-<br />

PENSABLE MEDIATOR OF OVULATION. G. M. Yerushalmi, a<br />

Y. Yung, a R. Orvieto, a E. Y. Adashi, b A. Hourvitz. c a Department of Obstetrics<br />

and Gynecology, Sheba Medical Center, Ramat Gan, Israel; b The Warren<br />

Alpert Medical School Brown University, Providence, RI; c Sheba Medical<br />

Center, Ramt-Gan, Israel.<br />

OBJECTIVE: To investigate the physiological ovarian role of the prostaglandin<br />

transporter (PGT).<br />

DESIGN: Experimental study.<br />

MATERIALS AND METHODS: PGT expression in vivo and in vitro in<br />

human mural granulosa cells (MGC) and in the mouse ovary was determined<br />

by qPCR, WB and IHC. MGCs were treated with human chorionic gonadotropin<br />

(hCG) and different activators and inhibitors of the LH pathway and<br />

PGT expression was analyzed by qPCR. in vitro transport assays of PGT in<br />

MGCs using EIA and radiolabeled PGE2. in vivo PGT function studies<br />

were performed in female mice, subjected to a superovulation protocol.<br />

The effect of PGT inhibitors, DIDS (4,4 0 -Diisothiocyanatostilbene-2,2 0 -disulfonic<br />

acid) and Bromocrestol Green (BCG), on ovulation was examined<br />

by histological evaluation, oocyte number and maturation, serum progesterone<br />

levels and gene expression.<br />

RESULTS: The expression of PGT, a transmembrane PG carrier protein,<br />

is markedly upregulated in preovulatory MGCs. Treatment with hCG, an<br />

ovulatory trigger, significantly increases the expression of PGT mRNA<br />

and protein both in vivo and in vitro. The hCG-induced increase in PGT<br />

expression is mediated via protein kinase A and protein kinase C by way<br />

of the extracellular signal-regulated kinase (ERK) pathway. PGT also mediates<br />

the uptake of PGE2 in cultured MGCs, thereby regulating its extracellular<br />

concentration. in vivo treatment with PGT inhibitors effectively<br />

inhibits ovulation in mice. Inhibition of PGT was also associated with inhibition<br />

of luteal function and a marked downregulation of key ovulatory<br />

genes including AREG, EREG, PR, TNFAIP6, StAR, ADAM-TS1, and<br />

CSTL.<br />

CONCLUSIONS: We hypothesize that the inhibition of PGT in GCs increases<br />

the extracellular concentrations of PGE2, the ability of which to exert<br />

its ovulatory effect is compromised by the desensitization of its cognate receptors.<br />

These findings suggest a new mechanism for the regulation of PGE2<br />

during ovulation, and prove PGT as an indispensable mediator of the ovulatory<br />

process, the inhibitors of which constitute potential novel candidates for<br />

non-hormonal contraception.<br />

P-90 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE MRNA EXPRESSION OF MT-ENCODED GENES OF OXPHS<br />

IN OVARIAN GCS OF INFERTILE WOMEN AND THE EFFECT<br />

OF NMN ON EXPRESSION OF MT-ENCODED GENES IN SVOG<br />

CELLS. Y. Li X. Liang. Reproductive Medicine Research Center, Sixth<br />

Affiliated Hospital, Sun Yat-sen University, Guangzhou, China.<br />

OBJECTIVE: To detect the mRNA expression of mitochondria(mt)-encoded<br />

genes in oxidative phosphorylation system(OXPHS) in ovarian granulosa<br />

cells(GCs) of infertile women with advanced age or poor ovarian<br />

response(POR), and to explore the relationship between mRNA expression<br />

with in vitro fertilization(IVF) outcomes. In addition, effect of nicotinamide<br />

mononucleotide(NMN)-a NAD+ precursor on mRNA expression of mt-encoded<br />

genes in OXPHS is to be detected.<br />

DESIGN: Retrospective and experimental study.<br />

MATERIALS AND METHODS: 45 infertile women were divided into the<br />

old group(>¼ 37 years old), the young group(< 37 years old) with POR and<br />

the young group with normal ovarian response(NOR). On the day of oocyte<br />

retrieval, GCs were purified from follicular fluid in all the patients for use. In<br />

separate experiments, immortalized human granulosa(SVOG) cells were<br />

treated with nicotinamide mononucleotide(NMN) or NMN adenylyltransferase-1(NMNAT-1)<br />

small interfering RNA(siRNA) to see effect of NMN on<br />

expression of mt-encoded genes in OXPHS. Real-time quantitative PCR<br />

was conducted to detect the mRNA expression of mt-encoded genes CYB,<br />

ND1, ATP6, CO1 and nuclear(nu)-encoded genes ATP5A1, SDHB,<br />

UQCRC1, NDUFS8 in GCs or SVOG cells. mRNA expression in different<br />

groups and the relationship with IVF outcomes, such as number of retrieved<br />

oocytes, achievement of good quality embryos, clinical pregnancy or live<br />

birth were evaluated by Kruskal-Wallis H test, Mann-Whitney or one-way<br />

ANOVA as appropriate.<br />

RESULTS: The mRNA expression of the mt-encoded genes in young<br />

POR group and old group were significantly lower than young NOR group,<br />

besides patients who retrieved more than 3 oocytes or had high quality embryos<br />

showed higher mt-encoded genes expression in GCs. In SVOG cells,<br />

the expression of mt-encoded genes was significantly decreased in the<br />

NMNAT-1 gene knockdown group compared with control and control+NMN<br />

group. Additionally, there were no significant differences in<br />

the mRNA expression of nu-encoded genes among these groups in GCs<br />

or SVOG cells.<br />

CONCLUSIONS: Young POR and old infertile patients may have<br />

impaired mitochondrial function in GCs. Higher expression of mt-encoded<br />

genes in GCs implies good ovarian response and high quality embryos. Besides,<br />

supplementation of NAD+ precursors such as NMN may restore mitochondrial<br />

function and prevent ovarian cell aging.<br />

P-91 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ANTI-MULLERIAN HORMONE (AMH) MAY INHIBIT OOCYTE<br />

MATURATION AND FOLLICULAR VASCULARIZATION IN HU-<br />

MAN OVARIAN CORTEX. L. Detti, a L. J. Williams, a<br />

N. M. Fletcher, b G. M. Saed. b a Obstetrics and Gynecology, University of<br />

Tennessee, Memphis, Memphis, TN; b Wayne State University, Detroit, MI.<br />

OBJECTIVE: AMH inhibits follicle recruitment [1, 2]. In fact, in the<br />

absence of AMH, primordial follicles are recruited at a faster pace, a common<br />

complication after transplantation of ovarian cortex blamed on a defective<br />

re-vascularization. VEGF stimulates neovascularization and GDF9 is<br />

essential for oocyte-dependent ovarian follicle development [3]. We tested<br />

the hypothesis that administration of recombinant AMH to ovarian cortex<br />

fragments would inhibit follicular development by modulating GDF9 and<br />

VEGF expression.<br />

DESIGN: Pilot experimental study with ovarian cortex obtained from 3<br />

patients.<br />

MATERIALS AND METHODS: Immediately after explant the ovarian<br />

cortex specimens were divided into 5 equal fragments. One fragment was<br />

flash-frozen (untreated) and four were incubated for 48 hours 37 Cina<br />

pH-adjusted gamete buffer media with increasing AMH concentrations of<br />

0-5-25-50 ng/ml. After incubation, all specimens were rinsed and flashfrozen<br />

for PCR analyses, which were executed in triplicates. We utilized<br />

real time RT-PCR to determine mRNA levels for VEGF and GDF9 in ovarian<br />

cortex tissue . We performed ANOVA with Tukey post hoc tests to evaluate<br />

changes in mRNA levels among the five different fragments. A p


TABLE. VEGF and GDF9 mRNA expression in the different treatment groups.<br />

Treatment<br />

VEGF espression<br />

(pg/mcg)<br />

GDF9 espression<br />

(pg/mcg)<br />

Untreated 8.71.3 18.<strong>21</strong>.9<br />

Media Control<br />

826.08.3<br />

45.27.5<br />

(no AMH)<br />

5 ng AMH 706.256.2 49.41.0<br />

25 ng AMH 422.234.0 18.40.6<br />

50 ng AMH 194.314.8 14.92.7<br />

References:<br />

1. Durlinger ALL, Kramer P, Karels B, et al. Control of primordial follicle<br />

recruitment by anti-Mullerian hormone in the mouse ovary. Endocrinol<br />

1999;140:5789-96.<br />

2. Durlinger ALL, Visser JA, Themmen APN. Regulation of ovarian<br />

function: the role of anti-Mullerian hormone. Reprod<br />

2002;124:601- 9.<br />

3. Dong J, Albertini DF, Nishimori K, Kumar TR, Lu N, Matzuk MM.<br />

Growth differentiation factor-9 is required during early ovarian folliculogenesis.<br />

Nature 1996; 383: 531-535.<br />

in the pups born from natural breeding may indicate that further investigation<br />

is warranted. If appropriately powered the results may reveal a benefit<br />

IVF/ICSI.<br />

CONCLUSIONS: A single nonablative dose of CYC does not immediately<br />

alter semen parameters. IVF/ICSI with CYC exposed sperm decreases<br />

embryo development but does not affect the quality of<br />

surviving embryos. CYC exposure resulted in decreased litter size but<br />

not birth rate following IVF/ICSI-ET or NB. IVF/ICSI with CYC exposed<br />

sperm does not lead to increased genetic variants in the resulting 8 cell<br />

embryos compared with embryos generated using PreCYC or VEH<br />

treated sperm.<br />

Supported by: Eunice Kennedy Shriver National Institutes of Health and<br />

Childhood Development grants HD055475, HD076412, HD075795, Magee-Womens<br />

Research Institute and Foundation and gift funds from Sylvia<br />

Bernassoli.<br />

P-93 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

WITHDRAWN<br />

Supported by: UTHSC Departmental Funds.<br />

ENVIRONMENT AND TOXICOLOGY<br />

P-92 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

EXPOSURE OF TERMINALLY DIFFERENTIATED SPERM TO A<br />

SINGLE NONABLATIVE DOSE OF CYCLOPHOSPHAMIDE<br />

DOES NOT INCREASE THE NUMBER OF GENETIC VARIANTS<br />

IN EMBRYOS OR LIVE BORN OFFSPRING PRODUCED BY IVF/<br />

ICSI. M. D. Johnson, a,b M. Sukhwani, b K. Peters, b S. Malik, a,b<br />

J. S. Sanfilippo, a K. E. Orwig. a,b a Department of Obstetrics, Gynecology<br />

& Reproductive Sciences, University of Pittsburgh School of Medicine, Pittsburgh,<br />

PA; b Magee-Womens Research Institute, Pittsburgh, PA.<br />

OBJECTIVE: Determine the effect of cyclophosphamide on sperm, embryo<br />

development after in vitro fertilization (IVF)/intracytoplasmic sperm injection<br />

(ICSI), birth rate after IVF/ICSI-embryo transfer (ET) or natural<br />

breeding, and genetic integrity of embryos and offspring.<br />

DESIGN: Prospective Laboratory Study.<br />

MATERIALS AND METHODS: Sperm was collected from the epididymis<br />

of male B6D2 mice. 7 days later intraperitoneal cyclophosphamide<br />

(CYC, 300 mg/kg) (n¼10) or vehicle (VEH) (n¼10) were administered.<br />

Sperm were recovered from the contralateral epididymis 7 days later. A<br />

post thaw semen analysis was performed on one straw from each Pre<br />

and Post treatment sample. Pre and Post treatment sperm from CYC<br />

and VEH treated animals were thawed and morphologically normal,<br />

motile sperm were selected to fertilize eggs from B6D2 females by<br />

ICSI. Embryos were monitored for in-vitro development for two days<br />

then collected for exome sequencing. In a separate experiment, birth<br />

rate and genetic integrity of live born offspring was examined after ET<br />

of day two embryos produced by IVF/ICSI with CYC exposed and pre<br />

exposed sperm. These results were compared with offspring produced<br />

by natural breeding (NB) of males before and after exposure to CYC. Genetic<br />

integrity was assessed by exome sequencing of tail DNA from IVF/<br />

ICSI-ET and NB offspring.<br />

RESULTS: CYC did not affect sperm count (P¼.15), motility (P¼.86) or<br />

morphology (P¼.98). 983 embryos from 22 IVF/ICSI cycles were monitored<br />

for preimplantation development. The number of 8 cell embryos produced<br />

using PostCYC sperm was reduced compared to those using PreCYC<br />

sperm (47% vs 18%, P92% of 8 cell embryos in all groups<br />

were Grade A (P>.50). Birth rates after ET were similar for the PreCYC<br />

and PostCYC groups (100% vs 60%, P¼.7) but litter size (3.8 vs 0.8,<br />

P


P-94 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

DIETARY BISPHENOL-A EXPOSURE ALTERS THE METHYL-<br />

OME OF RAT TROPECTODERM AS DETERMINED BY REDUCED<br />

REPRESENTATION BISULFITE SEQUENCING. A. E. Batcheller, a<br />

G. L. Christensen, a Y. Leung, b J. Biesiada, b X. Zhang, b<br />

M. Medvedovic, b M. A. Thomas, a S. Ho. b a Division of Reproductive Endocrinology<br />

and Infertility, Department of Obstetrics and Gynecology, University<br />

of Cincinnati, West Chester, OH; b Department of Environmental Health,<br />

University of Cincinnati, Cincinnati, OH.<br />

OBJECTIVE: To determine if maternal dietary bisphenol-A (BPA) exposure<br />

alters the rat trophectoderm methylome as determined by reduced representation<br />

bisulfite sequencing (RRBS).<br />

DESIGN: Basic research comparative animal study.<br />

MATERIALS AND METHODS: Female Sprague Dawley rats were fed<br />

either AIN-93G control (CTL n¼7) or 250 mg/kg body weight/day BPA<br />

(BPA n¼8) diets during the periconceptional time period. Dams were sacrificed<br />

on gestational day 5, and embryos were flushed from the uterine<br />

horns. Blastocyst trophectoderm (TE) biopsies were obtained using an embryo<br />

splitter (Bioniche). TE biopsies (CTL n¼42, BPA n¼48) were pooled<br />

to make 5 samples in the CTL group, and 6 samples in the BPA group, then<br />

lysed. Genomic DNA was digested with Msp I, size selected and bisulfite<br />

modified. The library was enriched and amplified, then sequenced using Illumina<br />

HiSeq1000. Data were trimmed with Trim Galore. Overall read<br />

quality was determined with FastQC. Read alignment, methylation, and<br />

CpG information was extracted with Bismark. CTL and BPA samples<br />

were collapsed into two groups for analysis. For sites with at least 10<br />

fold coverage, the methylation change was calculated as bBPA(methylated<br />

reads/total reads)-bCTL(methylated reads/total reads). Differences greater<br />

than 60% were considered important. Subsequent statistical analyses<br />

were conducted.<br />

RESULTS: 20464 sequence reads exhibited at least 10 reads per<br />

sequence in both groups, of which 1963 reads demonstrated a difference<br />

in methylation of more than 60%. These reads were located in 230 introns,<br />

five 3’untranslated regions, 16 coding sequences, 625 intragenic regions,<br />

and 16 promoters. Five genes showed more than one differentially methylated<br />

CpG site in the promoter: ERRFI1,Gak, Vom1r90, Rn5-8s, and<br />

Tmem<strong>17</strong>5.<br />

CONCLUSIONS: RRBS can be used to identify differentially methylated<br />

genes in rat trophectoderm. We noted hypomethylation of ERRFI1 promoter<br />

CpG sites resulting from dietary BPA exposure. ERRFI1 is involved in the<br />

embryo/maternal interaction and represents an interesting direction for future<br />

research.<br />

Supported by: Center for Environmental Genetics Innovator Award NIH/<br />

NIEHS P30ES006096. Center for Environmental Genetics New Investigator<br />

Scholar Award NIH/NIEHS P30ES006096.<br />

region as a determinant of LBR after adjusting for recipient’s age, BMI,<br />

race, day and number of donor egg embryos transferred in fresh and FET cycles.<br />

RESULTS: Of 76,296 donor egg IVF cycles reported to SART, regional<br />

information was available for 71,182. Specified regions differed significantly<br />

in: number of donor egg IVF cycles (with highest number of cycles being undertaken<br />

in NE at 20,351, followed by the W (n¼9,051), S(n¼8,054), and<br />

MW (least at 12,241), recipient’s age (p< 0.001), BMI (p< 0.001) and<br />

race (p< 0.001). Regional differences in LBR following fresh and FET<br />

were noted, with lowest LBR’s reported for cycles undertaken in the NE<br />

(46.12% for fresh ET and 26.02% for FET) whereas highest LBR’s were<br />

noted for cycles undertaken in the W (54.23% for fresh ET and 30.12% for<br />

FET, p< 0.001). On adjusted analyses, compared to the NE, donor egg<br />

IVF-ET cycles undertaken in the W region were significantly more likely<br />

to yield LB following fresh ET (OR1.32, 95% CI 1.24-1.41) and FET<br />

(1.18, 95% CI 1.06-1.30) (Table 1).<br />

CONCLUSIONS: Accounting for regional differences in the demographics<br />

of the donor egg recipients, our current analyses (undertaken on<br />

an extended dataset of donor egg IVF cycles) reaffirms our earlier observation.<br />

There are significant regional differences in LBR’s for both fresh and<br />

FET donor egg IVF cycles. Consistent with our earlier observation, highest<br />

LBR’s were observed for donor egg IVF cycles undertaken in the West region<br />

of the USA. Potential influences of ecological nuances on IVF outcome are<br />

suggested.<br />

Association between region and live birth outcome for frozen and fresh embryo<br />

transfer cycles.<br />

Frozen Embryo Transfer<br />

Region Unadjusted OR p Adjusted OR p<br />

(95% CI)<br />

(95% CI)<br />

Northeast Reference — Reference —<br />

South 1.14 (1.05, 1.23) 0.003 1.15 (1.03, 1.27) 0.010<br />

Midwest 1.08 (0.99, 1.18) 0.093 1.08 (0.96, 1.<strong>21</strong>) 0.189<br />

West 1.23 (1.14, 1.32)


were whether or not a state had mandated coverage for either IVF or some<br />

aspect of infertility care.<br />

RESULTS: Our findings were that classical economic drivers relating to<br />

wealth or economic well-being were not predictive of IVF demand. Rather,<br />

IVF demand was best predicted by general medical insurance coverage,<br />

specific IVF coverage, and the proportion of the population with a selfidentified<br />

Catholic background in a state. The model constructed was:<br />

Log(IVF Cases/Pop.) ¼ -2.96 + 2.41 Catholic + 2.88 Insurance Coverage<br />

+ 0.43 IVF Coverage (se) (1.05) (0.72) (1.33) (0.<strong>17</strong>) (p-value) (0.007)<br />

(0.0018) (.036) (0.015) The F-statistic was 16.2; and the R-squared was<br />

0.53.<br />

CONCLUSIONS: Independent variables associated with various measures<br />

of income and economic outlook were not helpful in explaining<br />

the variability of IVF utilization among states. The number of IVF cases<br />

done in a state was not related to the effectiveness of IVF in that state nor<br />

the convenience of having an increased number of IVF programs. The<br />

most unexpected finding was that a 1% increase in the proportion of<br />

self-identified Catholics in a state was associated with a 2.4% increase<br />

in the number of IVF cases per 1000 reproductive aged women. Given<br />

the Donum Vitae, one might have expected the impact of an increased<br />

proportion of Catholics to have a negative impact on IVF utilization.<br />

We hypothesized that the variable of being a self-identified Catholic encompassed<br />

microeconomic factors that played a major role in the decision<br />

to undertake IVF. These might include love of family and an increased<br />

desire to have children.<br />

P-97 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ASSOCIATION OF PREPUBERTAL SERUM DIOXIN CONCEN-<br />

TRATIONS WITH SEMEN QUALITY IN A PROSPECTIVE<br />

COHORT OF YOUNG RUSSIAN MEN. L. Minguez-Alarcon, a<br />

O. Sergeyev, b,c J. S. Burns, a P. L. Williams, a M. M. Lee, d<br />

S. A. Korrick, e L. Smigulina, c B. Revich, b R. Hauser. a a Harvard T H<br />

Chan School of Public Health, Boston, MA; b Russian Academy of Science,<br />

Moscow, Russian Federation; c Chapaevsk Medical Association, Moscow,<br />

Russian Federation;<br />

d University of Massachusetts Medical School,<br />

Worcester, MA; e Brigham and Women’s Hospital, Harvard Medical School,<br />

Boston, MA.<br />

OBJECTIVE: To explore associations of prepubertal serum concentrations<br />

of dioxins, furans and PCBs with semen parameters.<br />

DESIGN: A prospective cohort study conducted in Chapaevsk,<br />

Russia.<br />

MATERIALS AND METHODS: From 2003 to 2005, 499 boys were<br />

enrolled at age 8-9 years and their growth and development was assessed<br />

annually for ten years. Serum dioxins, furans and PCBs were measured<br />

by the U.S. CDC using samples collected at enrollment. At ages 18-19<br />

years, 133 young men provided up to two semen samples collected<br />

one week apart (257 samples) which were analyzed for volume, sperm<br />

concentration and motility according to NAFA and ESHRE-SIGA<br />

manual. Linear mixed models with random intercepts were used to<br />

examine the relation between quartiles of serum concentrations of dioxins,<br />

dioxin toxic equivalents (TEQs), furans and PCBs with semen parameters.<br />

Models were adjusted for BMI, season, and abstinence time,<br />

with further adjustment by time from ejaculation to start of semen analysis<br />

for sperm motility.<br />

RESULTS: Men had a median sperm concentration and motility of 52<br />

mill/mL and 64%, respectively. Higher prepubertal serum TCDD levels<br />

were associated with lower semen parameters. The adjusted difference<br />

(95% CI) in percent change in sperm concentration, total sperm count, and<br />

total motile sperm count between the lowest and highest quartile was<br />

-23.8% (-33.6, -10.5) (p trend¼ .008), -41.5% (-68.9, 1.9) (p trend¼.06)<br />

and -27.6% (-45.2, -1.2) (p trend¼ .06), respectively. In addition, there was<br />

a suggested negative association of PCDD TEQs and P PCDD concentrations<br />

with sperm concentration (p trend¼ .06) and motility (p trend¼.10),<br />

respectively.<br />

CONCLUSIONS: Our results were consistent with the Seveso study that<br />

showed an association of early life exposure to TCDD with poorer semen<br />

quality, indicating that the prepubertal period is a sensitive window of exposure.<br />

Supported by: NIH grants R01ES0014370, P30ES000002, and Russian<br />

Science Foundation grant 14-45-0065.<br />

P-98 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

NEW METHODS IN QUALITY CONTROL: DETECTION OF SUB-<br />

LETHAL TOXICITY. A. Ainsworth, a J. Fredrickson, a D. Morbeck. b<br />

a Obstetrics and Gynecology, Mayo Clinic, Rochester, MN; b Mayo Clinic,<br />

Rochester, MN.<br />

OBJECTIVE: Success of in vitro fertilization (IVF) depends on sound laboratory<br />

methods, culture conditions, and quality supplies, which must be<br />

monitored using sensitive quality control (QC) testing. Current QC testing<br />

using mouse embryo assays (MEA) detects overt toxicity but fails to detect<br />

sublethal toxicity. Sublethal toxicity has been associated with suboptimal<br />

embryo development and occasional manufacturer recalls. This study aimed<br />

to improve the sensitivity of routine MEA in detection of sublethal mineral<br />

oil toxicity.<br />

DESIGN: Two experiments were conducted to examine the detection of<br />

toxicity with sublethal levels of peroxide. Each experiment used fresh 1-<br />

cell mouse embryos, cultured individually, in previously recalled IVF-grade<br />

mineral oil. In experiment 1, an extended MEA (144 h) was compared to<br />

routine MEA (96 h) and cell counts. In experiment 2, sensitivity of the<br />

extended MEA was compared to morphokinetics using serial dilutions of<br />

oil high in peroxides.<br />

MATERIALS AND METHODS: In experiment 1, embryos were cultured<br />

under mineral oil with known toxicity (Oil 31). After 96 hours, half of the<br />

embryos were removed for cell counts. The remaining embryos continued<br />

in culture until 144 hours. In experiment 2, embryos were cultured in serial<br />

dilutions of mineral oil with known toxicity (Oil 44). Morphokinetic data<br />

from the Embryoscope was compared to blastocyst rates at 96 and 144<br />

hours. The results were compared to control oil outcomes using Dunnett’s<br />

t-test.<br />

RESULTS: Extended MEA and cell counts, from experiment 1, identified<br />

sublethal peroxide toxicity not detected by evaluation of blastocyst rates at 96<br />

hours. Blastocyst rates were similar between control and toxic oil at 96 hours,<br />

86% vs. 90%, but significantly different at 144 hours, 30% vs. 0% (p < 0.01).<br />

Cell counts at 96 hours also reflected statistically significant differences, 61<br />

cells in control oil vs. 52 cells in toxic oil (p < 0.01). Morphokinetic data,<br />

from experiment 2, showed delayed cell division with small dilutions of toxic<br />

oil (Table 1). Extended MEA demonstrated increased sensitivity in detection<br />

of trace toxicity when compared to blastocyst rates at 96 hours (Table 1).<br />

CONCLUSIONS: Current methods of QC testing are not sensitive enough<br />

to detect sublethal embryo toxicity. Extended culture to 144 hours, cell<br />

counts at 96 hours, and time-lapse morphokinetic assessment are more effective<br />

than current methods in detection of peroxide toxicity and should be used<br />

to improve reproductive outcomes.<br />

Table 1: Blastocyst Rates and Morphokinetics in Extended MEA.<br />

Oil Blastocyst<br />

Dilutions Rate at 96h<br />

Blastocyst<br />

Rate at 120h<br />

Blastocyst<br />

Rate at 144h<br />

Morphokinetics<br />

i3(h)<br />

Morphokinetics Morphokinetics<br />

cc2(h) tB(h)<br />

1:20 <strong>21</strong>.88%<br />

(p


MATERIALS AND METHODS: This study included 20 adult men who<br />

reported smoking only cigarettes, referred to the Human Reproduction Section<br />

of the Sao Paulo Federal University for andrological evaluation. Twenty<br />

adult men without semen alterations were included as controls. Men with<br />

confounding factors of male infertility and those reporting fever in the 90<br />

days preceding semen collection were excluded. Semen was collected and<br />

analyzed as per the WHO 2010 guidelines. DNA fragmentation was assessed<br />

using an alkaline Comet assay, classified as grade I (no DNA fragmentation)<br />

to class IV (high DNA fragmentation); mitochondrial activity was analyzed<br />

by a reagent which stains only active mitochondria (3,3’-diaminobenzidine<br />

[DAB]), and classified as DAB I (all mitochondria active) to DAB IV (all<br />

inactive); and acrosome integrity was evaluated by a fluorescent stain<br />

(PNA-FITC). The seminal plasma was used to assess oxidative stress by<br />

TBARS quantification and for label-free quantitative proteomics, performed<br />

by liquid chromatography - tandem mass spectrometry (LC-MS/MS).<br />

Groups were compared using a Student’s T-test (p


increase was maintained with increasing concentrations of up to 1000 ng/ml.<br />

A time-dependent increase in sperm motility was observed in Ritalin-treated<br />

samples until 240 minutes at all concentrations. Thereafter, a decrease was<br />

observed at 300 minutes, even though motility remained significantly higher<br />

compared to basal levels. No significant changes were seen in sperm motion<br />

kinetic parameters.<br />

CONCLUSIONS: Ritalin increased sperm motility at a concentration<br />

of 1 ng/mL. It had no toxic effect on sperm motility even at a concentration<br />

of 1000ng/mL. More research is needed to evaluate whether Ritalin<br />

can be used to improve sperm quality in asthenozoospermic specimens<br />

from patients undergoing assisted reproductive procedures. Further<br />

investigation is ongoing to evaluate the effect of Ritalin on other sperm<br />

parameters such as acrosome reaction, DNA fragmentation and membrane<br />

integrity.<br />

In vitro effect of Ritalin on motility of spermatozoa (%, MeanSEM, n¼10).<br />

Time (min)/<br />

Ritalin (ng/ml)Y 0’ 60’ 150’ 240’ 300’<br />

Control 58.33.79 a 58.44.18 a 59.14.06 a 58.24.14 a 56.34.52 a<br />

1 57.04.25 a 63.93.59* b 66.33.64* b 68.93.83* b 60.53.65* b<br />

10 58.03.69 a 63.23.87* b 66.53.82* b 68.84.<strong>21</strong>* b 63.23.98* b<br />

100 59.03.55 a 63.44.07* c 66.43.40* c 68.73.90* c 62.93.69* c<br />

1000 59.24.18 a 65.04.02* b 66.53.66* b 68.53.83* b 63.13.61* b<br />

OBJECTIVE: Among Endocrine Disruptors, Bisphenol A (BPA), a<br />

monomer widely used as plasticizer, showed adverse effects on reproductive<br />

system. BPA it was previously foundinovarianfollicularfluid<br />

at approximately 1-2 ng/ml but the direct effects on ovarian somatic<br />

cells are still poor elucidated. In this light, the aim of this study was<br />

to evaluate by FT-IR microspectroscopy and qPCR how BPA affects<br />

metabolism and steroidogenic activity of living human Granulosa cells<br />

(GCs).<br />

DESIGN: This study has been conducted between June and Dicember<br />

2014 on women undergoing a COH for an IVF treatment (n ¼ 20), referring<br />

male factor infertility, (age¼364). GCs have been in vitro exposed<br />

for 48h to different concentrations of BPA (1ng/ml, 10ng/ml and 100ng/<br />

ml).<br />

MATERIALS AND METHODS: FT-IR analysis will be carried out on<br />

living GCs. At this regard, a specific microfluidic device was designed and<br />

used. Q-PCR was utilised to detect gene expression modulation induced by<br />

BPA.<br />

RESULTS: Results from both FTIR and QPCR analises are showed in the<br />

following table.<br />

CONCLUSIONS: The results obtained clearly evidenced a detrimental<br />

effect of BPA on GCs seriously compromising metabolism and AMH<br />

synthesis activity and inducing lipid peroxidation and cell death via<br />

apoptosis and autophagy process. Particularly, the lowest dose of BPA,<br />

the environmentally relevant one, induced the highest negative impact<br />

on GCs.<br />

FTIR RESULTS Control BPA (1ng/ml) BPA (10ng/ml) BPA (100ng/ml)<br />

P-102 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

GYNECOLOGIC DISEASE RISK OF EXPOSURE TO BISPHENOL<br />

A AND PHTHALATE. J. Jeon, S. Park, D. Lee, K. Jeong, H. Chung.<br />

Obstetrics & Gynecology, Ewha Womans University College of Medicine,<br />

Seoul, Korea, Republic of.<br />

OBJECTIVE: To investigate the risk of gynecologic disease is related to<br />

endocrine disrupting chemicals exposure in Korean reproductive women.<br />

DESIGN: Prospective cohort study.<br />

MATERIALS AND METHODS: A total of 307 women aged 30 to 49,<br />

who have lived in the same area for more than 18 months were evaluated.<br />

Anthropometric measurements, laboratory tests with urine and blood sampling<br />

and pelvic ultrasound examinations were performed. Logistic regression<br />

analysis was used to evaluate the contribution of each endocrine<br />

disrupting chemicals exposure for myoma, infertility, endometriosis, obesity,<br />

and metabolic syndrome.<br />

RESULTS: The mean age was 36.764.39 years old and the highest proportion<br />

(45.6%) was between 30 and 35 years old. The mean body mass index<br />

(BMI) was 22.43.1 kg/m2 and 50 women (16.3%) were diagnosed with<br />

obesity (BMI>25.0 kg/m 2 ). Bisphenol A was significantly higher<br />

(1.992.02 ug/g crea) in women with overweight (BMI>23.0 kg/m 2 ) and<br />

obesity. Logistic regression analysis suggested infertility was increased in<br />

high bisphenol A exposure group and the oods ratio (4.248) was significant<br />

statistically after adjusted for age, birth control pills, age of menarche, parity,<br />

and waist circumference. High phthalate exposure was associated to endometrial<br />

polyp after adjustment (OR: 2.742). However, exposure to these endocrine<br />

disrupting chemicals may not lead to myoma, endometriosis, and<br />

metabolic syndrome.<br />

CONCLUSIONS: Bisphenol A exposure is associated with overweight<br />

and infertility and endometrial polyp is increased in Korean reproductive<br />

women with high phthalate exposure.<br />

P-103 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

HUMAN GRANULOSA CELLS IN VITRO EXPOSURE TO BISPHE-<br />

NOL A: FTIR MICROSPECTROSCOPY AND MOLECULAR<br />

EVIDENCES. G. Gioacchini, a,b E. Giorgini, a L. Vaccari, c E. Sereni, b<br />

C. Zaca, b O. Carnevali, a A. Borini. b a Universita Politecnica delle Marche,<br />

Ancona, Italy; b Tecnobios Procreazione, Bologna, Italy; c Elettra Synchrotron,<br />

Trieste, Italy.<br />

LIPIDS/CELL(%) 14.661.25% 25.072.97%*** 22.122.64%* <strong>21</strong>.332.85%<br />

PHOSPHOLIPIDS/LIPIDS(%) 15.990.69% <strong>21</strong>.080.12%*** 19.560.16% 19.260.16%<br />

LIPIDS PEROXIDATION(%) 1.250.01% 2.040.01%** 1.460.01% 1.630.02%*<br />

PROTEIN/CELL(%) 30.540.83% 20.120.28%** 22.610.36%** 19,870.31%**<br />

UNFOLDED PROTEIN (%) 15.<strong>21</strong>3.65% 43.094.71%*** 35.545.01%** 39.734.13%***<br />

RNA/DNA (%) 10.512.46% 4.600.36%*** 8.981.68% 9.582.01%<br />

QPCR RESULTS Control BPA (1ng/ml) BPA (10ng/ml) BPA (100ng/ml)<br />

Peroxisome proliferator 1.260.27 10.190.86*** 1.230.39 3.330.78*<br />

activated receptorg<br />

(lipid metabolism)<br />

CASPASE3 (apoptosis marker) 1.230.29 8.700.97** 3.301.31* 1.370.45*<br />

BECLIN1 (autophagy marker 1.130.40 5.000.88* 4.<strong>17</strong>0.83* 4.500.90*<br />

Anti Mullerian Hormone 6.231.29 1.400.11 ** 2.130.12** 1.600.48**<br />

*¼P


mean cumulative pre-test scores. At curriculum completion 89% reported<br />

they would adopt or modify strategies for managing patients’ fish intake;<br />

at follow-up 79% had adopted new practices such as providing lists of fish<br />

to avoid and information about adding fish to diet; improving screening for<br />

at-risk groups based on ethnicity and fish consumption levels; adding questions<br />

to electronic medical record screens; and providing patients with magnets<br />

and placing posters on the walls of their waiting rooms.<br />

CONCLUSIONS: Prenatal providers completing the Healthy Fish<br />

Choices CME curriculum can help patients maximize the benefits of fish consumption<br />

while minimizing their exposure to contaminants in fish.<br />

Supported by: This project was Supported by grant #GL-00E00536-0 from<br />

the Great Lakes Restoration Initiative of the USEPA.<br />

P-105 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

GESTATIONAL HYPERTENSION COULD LED TO ADULT<br />

OFFSPRING VASCULAR DYSFUNCTION IN RATS. M. Guo a<br />

H. Huang. b a The Key Laboratory of Reproductive Genetics, Ministry of Education,<br />

Zhejiang University School of Medicine, Hangzhou, China; b Shanghai<br />

Jiao Tong University School of Medicine, Shanghai, China.<br />

OBJECTIVE: To evaluate if offsprings from gestational hypertension<br />

(GH) has vascular dysfunction long term in life.<br />

DESIGN: Fetal origins adult disease has become a lively topic of reproductive<br />

science recently and epidemiology research has proved that intrauterine<br />

environment could impact growth and health of offspring. Our previous<br />

proteomics study on umbilical artery detected several altered proteins<br />

expression during GH related to cardiovascular genesis. As the umbilical artery<br />

could appropriately represent fetal vascular, we assumed that the general<br />

cardiovascular system of offspring could also been affected during this disorder<br />

both short term and long term. To investigate long term effect of offsprings<br />

vascular function changes, we built an animal model use Sprague-<br />

Dawley rats. Body weight and blood pressure were measured continuously.<br />

Futhermore, mesentric arteries were isolated from both groups for contraction<br />

and relaxation function test and morphology analysis were also carried<br />

out use same arteries.<br />

MATERIALS AND METHODS: Animal model were built through L-<br />

NAME application. Maternal blood pressure were measured 2 days before<br />

mate till 2 days after delivery using tail clamp. Body weight and blood pressure<br />

of offsprings were continuously measured from birth to one and half<br />

year-old. 6 offsprings from each group were sacrificed at 8 week-old and 1<br />

year-old for the wire myography (model 620A, DMT). Excess mesentric arteries<br />

from same area were collected for the HE staining. The independentsamples<br />

t test were used to evaluate the statistical significance between two<br />

groups.<br />

RESULTS: Offspring birth weight from GH were lower but become identical<br />

after 3 weeks compared with the control group. GH offspring blood<br />

pressure increase begin at 6 month-old and mesentric artery contraction reaction<br />

is greater than control through same stimulation at 8 week-old. At 1<br />

year-old, not only contraction, the relaxation ability of GH offspring has<br />

also became obtuse to the vasodilator. Through HE staining, mesentric<br />

wall from GH group were much thicker than the control.<br />

CONCLUSIONS: GH is one of the major cause of maternal and neonatal<br />

mortality and morbidity. Besides the negative effects we observe short term<br />

at birth, it can also led to long term increasing risk of cardiovascular diseases<br />

in offspring. In this research, we revealed that the resistance arteries<br />

has both contraction and relaxation ability dysfunction from adult offsprings<br />

of GH and these arteries has a thicker artery wall. Blood pressure<br />

was also increased accompany with the function loss. Further investigation<br />

should be proceed to find out what the mechanism underling these phenomenon.<br />

Supported by: This work was Supported by the National Basic Research<br />

Program of China (No.2012CB944900 to H.F.H); the National Natural Science<br />

Foundation of China (No.81270708 to J.Z.S.).<br />

P-106 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

WITHDRAWN<br />

P-107 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

HEALTH DISPARITIES<br />

IVF OUTCOMES DISPARITIES IN THE UNITED STATES: DO<br />

WOMEN OF MIDDLE EASTERN/NORTH AFRICAN ETHNICITY<br />

EXPERIENCE IN VITRO FERTILIZATION (IVF) DISPARITIES<br />

COMPARED TO A CAUCASIAN COHORT?. W. H. Salem, a<br />

F. Sharara, b O. Abuzeid, c T. I. Abozaid, c M. Ashraf, c M. I. Abuzeid. c<br />

a OB/GYN and Reproductive Sciences, University of California, San Francisco,<br />

San Francisco, CA; b Virginia Center for Reproductive Medicine, Reston,<br />

VA; c IVF Michigan, Rochester Hills, MI.<br />

OBJECTIVE: Health disparities have been demonstrated for ethnic groups<br />

undergoing IVF in the United States however outcomes remain contradictory<br />

(1-3). There are no studies to our knowledge that address the Middle Eastern/<br />

North African (MENA) ethnic group in the U.S. Under SART reporting data,<br />

this group is generally categorized under the Caucasian ethnic group. This<br />

study investigates IVF outcomes and demographic metrics of MENA women<br />

in comparison to a control group of Caucasian women.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: A group of 50 randomly selected<br />

MENA women were compared to a control group of 100 randomly<br />

selected Caucasian women undergoing IVF at a private IVF clinic in<br />

Michigan over the period 2011-2013. Inclusion criteria were met if the<br />

subjects were undergoing an IVF cycle with a fresh day 5 blastocyst,<br />

morula or mixed transfer. Patients were excluded if they had incomplete<br />

data, underwent preimplantation genetic screening or utilized donor gametes.<br />

49 MENA patients and 99 Caucasian patients were included in the<br />

analysis. All continuous variables were analyzed with a two sample t<br />

test while categorical information was analyzed with a chi squared analysis.<br />

RESULTS: This pilot study suggests that there is no difference in IVF<br />

pregnancy or live birth rates between MENA and Caucasian women (67%<br />

vs 72% p¼0.5 and 57% vs 59% p¼0.87). There was a trend toward a higher<br />

miscarriage rate among MENAwomen (10% vs 4% p¼0.14). MENAwomen<br />

also had a trend toward a lower BMI than Caucasian women (25.0 vs 27.1,<br />

p¼0.06). This may be reflective of a lower prevalence of polycystic ovarian<br />

syndrome in the MENA group (26.5% vs 41.4%, p¼0.13). There were no differences<br />

in age, etiology of infertility and the rate of primary infertility.<br />

Embryological data demonstrated no significant differences in the number<br />

of fertilized oocytes, day 5 morulae, blastocysts or total embryos transferred.<br />

e142 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


MENA women required a lower number of stimulation days (8.7 vs 9.4,<br />

p¼0.01).<br />

CONCLUSIONS: This preliminary study suggests that there are no differences<br />

in pregnancy outcomes between MENA and Caucasian women in a<br />

sample U.S. population. This is the first study to support the categorization<br />

of MENA women under the Caucasian ethnic group for SART outcomes.<br />

A more robust analysis of MENA women with regards to IVF outcomes disparities<br />

is forthcoming.<br />

Demographics, Embryological and IVF Outcome Data of MENA vs Caucasian<br />

Women.<br />

MENAN¼49 Caucasian N¼99 P Value<br />

Age (years) 31.8 32.7 0.14<br />

BMI (kg/m^2) 25.0 27.1 0.06<br />

Male Factor 57.1% N¼28 54.5%N¼54 0.09<br />

PCOS/Ovulatory 28.6%N¼14 41.4%N¼41 0.13<br />

Tubal Factor 26.5%N¼13 28.3%N¼28 0.82<br />

Uterine 34.7%N¼<strong>17</strong> 36.4%N¼36 0.84<br />

Unexplained Infertility 4.1%N¼2 0%N¼0 0.04<br />

Antral Follicle Count 11.8 15.1 0.10<br />

Total Embryo Transfer 2.18 2.<strong>17</strong> 0.92<br />

Miscarriage 10.2%N¼5 4.0%N¼4 0.14<br />

Live Birth 57.1%N¼28 58.6%N¼58 0.87<br />

rate was higher in the Caucasian group but not statistically significant likely<br />

secondary to inadequate power (46% vs. 31%; OR 1.95; CI [0.87,4.36]). Live<br />

birth rate after a clinical pregnancy was documented was similar between<br />

groups (86% vs. 81%; OR 0.72; CI [0.14,3.81]).<br />

CONCLUSIONS: Asian ethnicity is associated with a lower clinical pregnancy<br />

rate even when donor oocytes are used. However, they have similar<br />

live birthrates after clinical pregnancy is documented. Differences in IVF<br />

pregnancy outcomes between Asian and Caucasian patients, even with<br />

oocyte donation, are likely secondary to impaired implantation in women<br />

who self declare Asian ethnicity.<br />

Asian (N¼39)<br />

Median<br />

(interquartile range)<br />

Caucasian (N¼84)<br />

Median<br />

(interquartile range)<br />

P value<br />

Donor Age 25 (6) 25 (6) 0.31<br />

Recipient Age 44 (8) 43 (5) 0.75<br />

Donor BMI 20 (2) <strong>21</strong> (3) 0.01<br />

Recipient BMI 22 (3) 22 (4) 0.52<br />

Total Gonadotropin (ampules) 30 (12) 30 (29) 0.53<br />

Peak E2 (pg/mL) 3815 (2476) 3650 (1854) 0.62<br />

Oocytes retrieved <strong>21</strong> (13) 24 (13) 0.06<br />

Mature oocytes <strong>17</strong> (11) 18 (10) 0.39<br />

Oocytes fertilized 15 (12) 16 (10) 0.38<br />

E2/oocyte retrieved (pg/mL) <strong>17</strong>4 (89) 150 (70) 0.03<br />

References:<br />

1. Dayal MB, Gindoff P, Dubey A, Spitzer TL, Bergin A, Peak D, et al.<br />

Does ethnicity influence in vitro fertilization (IVF) birth outcomes?<br />

Fertil Steril. 2009;91:2414-8.<br />

2. Sharara FI, McClamrock HD. Differences in in vitro fertilization (IVF)<br />

outcome between white and black women in an inner-city, university<br />

based IVF program. Fertil Steril 2000;73:1<strong>17</strong>0.<br />

3. Wellons MF, Fujimoto VY, Baker VL, et al. Race Matters: A Systematic<br />

Review of Racial/Ethnic Disparity in Society for Assisted Reproductive<br />

Technology(SART) Reported Outcomes Fertil Steril. 2012<br />

Aug;98(2):406-409.<br />

P-108 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

WHY ARE IVF PREGNANCY RATES LOWER IN WOMEN OF<br />

ASIAN ETHNICITY: AN ANALYSIS OF ETHNICITY-MATCHED<br />

OOCYTE DONOR CYCLES. K. Louie, a L. Ross, a T. Jones, a<br />

B. Rudick, b K. Chung, c K. Bendikson. d a Division of Reproductive Endocrinology<br />

and Infertility, University of Southern California, Los Angeles, CA;<br />

b Division of Reproductive Endocrinology and Infertility, Columbia University,<br />

New York, NY; c USC Keck School of Medicine, Los Angeles, CA;<br />

d USC Fertility, Los Angeles, CA.<br />

OBJECTIVE: To resolve controversies regarding decreased pregnancy<br />

rates in women of Asian ethnicity in comparison to Caucasian women, we<br />

sought to evaluate clinical pregnancy rates between ethnically matched Asian<br />

and Caucasian egg donation cycles.<br />

DESIGN: A retrospective cohort study.<br />

MATERIALS AND METHODS: First-time anonymous oocyte donors<br />

of Asian or Caucasian ethnicity who underwent oocyte stimulation between2006and2013thatresultedinafreshembryotransferwereidentified.<br />

Only the first stimulation of egg donors and first transfers of<br />

recipients were considered. Ethnicity was self-reported by both the<br />

donor and the intended recipient. Only ethnicity matched donor-recipient<br />

pairs were included. Fisher’s exact, Wilcoxon rank-sum, binomial<br />

and multivariate logistic regression were used as appropriate for analysis.<br />

RESULTS: 39 Asian and 123 Caucasian ethnically matched oocyte donor<br />

and recipient pairs meeting inclusion criteria were identified. Baseline characteristics<br />

were similar between Asian and Caucasian donors and recipients<br />

except for BMI in donors. There were no differences in stimulation parameters.<br />

E2 levels per oocyte retrieved were higher in Asian donors. The number<br />

of embryos transferred was the same between groups (2 1 vs. 2 1;<br />

p¼0.64). The Caucasian group had significantly higher implantation rates<br />

(42% 41% vs. <strong>21</strong>% 29%; p¼0.01) and clinical pregnancy rates (57%<br />

vs. 36%; OR 2.38; CI [1.09,5.22]) compared to the Asian group. Live birth<br />

P-109 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

OVERVIEW OF 2012 U.S. ASSISTED REPRODUCTIVE TECHNOL-<br />

OGY (ART) TREATMENT OUTCOMES AND CONTRIBUTION TO<br />

MULTIPLE BIRTH AND PRETERM INFANTS. S. Sunderam, a<br />

D. M. Kissin, b S. Crawford, a S. Folger, a D. Jamieson, a L. Warner, a<br />

W. Barfield. a a Division of Reproductive Health, Centers for Disease Control<br />

and Prevention, Chamblee, GA; b Centers for Disease Control and Prevention,<br />

Atlanta, GA.<br />

OBJECTIVE: To report U.S. ART statistics and compare ART infant outcomes<br />

to all U.S. infant outcomes.<br />

DESIGN: Population-based retrospective analysis.<br />

MATERIALS AND METHODS: Data were obtained from CDC’s National<br />

ART Surveillance System and National Vital Statistics System. The<br />

following were calculated for each state, the District of Columbia, and Puerto<br />

Rico: number of ART procedures performed per million women of reproductive<br />

age (ART utilization), average number of embryos transferred, rates of<br />

elective single embryo transfers (eSET), as well as rates of ART-conceived<br />

multiple-birth, low birth weight, and preterm infants. The proportion of<br />

ART infants among all infants, multiple-birth, low birth weight and preterm<br />

infants was also calculated.<br />

RESULTS: Among 3,991,741 infants born in 2012 in the U.S. and Puerto<br />

Rico, 1.5 % (61,432) were conceived with ART (range: 0.2% in Puerto Rico<br />

to 5% in Massachusetts). ART utilization ranged from 323 to 7,2<strong>21</strong> procedures<br />

per million women aged 15-44 years, and was higher than the national<br />

average of 2,483 in 13 reporting areas, many of which were located in the<br />

northeast. The national eSET rate among women


P-110 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

OXIDATIVE STRESS<br />

FOLLICULAR FLUID FROM INFERTILE WOMEN WITH MILD<br />

ENDOMETRIOSIS IMPAIRS IN VITRO BOVINE EMBRYO DEVEL-<br />

OPMENT AND N-ACETYL-CYSTEINE STIMULATES THE EM-<br />

BRYO HATCHING. V. S. Giorgi, a B. T. Jianini, a M. G. Da Broi, a<br />

C. C. De Paz, b R. Ferriani, a P. A. Navarro. a a Department of Obstetrics<br />

and Gynecology, Faculty of Medicine of Ribeirao Preto, University of S~ao<br />

Paulo, Ribeirao Preto, Brazil; b Department of Genetics, Faculty of Medicine<br />

of Ribeirao Preto, University of S~ao Paulo, Ribeirao Preto, Brazil.<br />

OBJECTIVE: To assess the effect of the addition of follicular fluid (FF)<br />

from infertile women with and without mild endometriosis (ME) and of<br />

the antioxidant N-acetyl-cysteine (NAC) on embryo development, using<br />

bovine model.<br />

DESIGN: Experimental study.<br />

MATERIALS AND METHODS: FF samples were obtained from 22 infertile<br />

women undergoing ovarian stimulation for intracytoplasmic sperm injection<br />

[11 with ME (EFF) and 11 with tubal and/or male factor of infertility<br />

(CFF)], pooled, and utilized in 5 in vitro maturation (IVM) experiments<br />

with immature bovine oocytes (IBO). IBO were submitted to IVM divided<br />

in 5 groups: without FF (NF), with 1% of FF from ME patients (EFF) or control<br />

patients (CFF), EFF + 1.5mM of NAC (ENAC), CFF + 1.5mM of NAC<br />

(CNAC). Then, in vitro fertilization (IVF) was performed and embryos were<br />

in vitro cultured. We analyzed cleavage rate, blastocyst production, and<br />

hatched blastocysts rate. Data were analyzed by gamma distribution.<br />

RESULTS: The cleavage rate was similar comparing NF with CFF<br />

(p¼0.606), CNAC (p¼0.206) and ENAC (p¼0.<strong>17</strong>9); and lower comparing<br />

EFF group with NF (p¼0.007) and CFF (p¼0.029) groups. The blastocyst production<br />

rate was similar between NF and ENAC (p¼0.544) groups, and lower<br />

in CFF (p¼0.019), CNAC (p¼0.003) and EFF (p¼0.004) compared to NF. The<br />

hatched blastocyst rate was similar between NF and CFF (p¼0.837) groups; the<br />

EFF had the lowest hatched blastocyst rate compared to all other groups<br />

(p


Mitochondrial thioredoxin dependent peroxide reductase and thioredoxin<br />

related transmembrane protein 4 were uniquely expressed in high ROS<br />

group providing additional defense. Overexpression of Apolipoprotein D<br />

is suggestive of better androgen response despite oxidative stress in this<br />

group.<br />

CONCLUSIONS: Based on the comparative proteome profiles, we<br />

conclude that fertility in men with high seminal ROS may be related to better<br />

stress response and enhanced antioxidant defense.<br />

P-113 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

MYELOPEROXIDASE SERVES AS A REDOX SWITCH THAT REG-<br />

ULATES APOPTOSIS IN HUMAN LEIOMYOMAS. M. S. Abusamaan, a<br />

N. M. Fletcher, a M. G. Saed, a A. Al-Hendy, b M. P. Diamond, c J. M. Berman, a<br />

G. M. Saed. a a Obstetrics and Gynecology, Wayne State University, Detroit, MI;<br />

b Medical College of Georgia, Georgia Regents Univer, Augusta, GA; c Georgia<br />

Regents University, Augusta, GA.<br />

OBJECTIVE: To determine the expression of myeloperoxidase (MPO)<br />

and nitric oxide (NO), modulators of oxidative stress, and their impact on<br />

apoptosis of leiomyoma cells under normal and hypoxic conditions.<br />

DESIGN: Prospective experimental study.<br />

MATERIALS AND METHODS: We have utilized an immortalized human<br />

leiomyoma and a matched myometrial cell line. Cells were exposed<br />

to normal (20% O2) and hypoxic (2% O2) conditions for 24 hours in Smooth<br />

Muscle Medium-2. Bax, Bcl-2 and caspase-3, markers of apoptosis, and<br />

MPO were measured by real-time RT-PCR. The Nitrate/Nitrite Colorimetric<br />

Assay was used to measure NO levels. Data were analyzed by Student’s t-<br />

test.<br />

RESULTS: MPO and NO levels were higher in leiomyoma (0.043 0.006<br />

fg/ug RNA and 7.0 0.09 mM) than myometrial cells (0.020 0.006 fg/mg<br />

mRNA and 6.3 0.18 mM) (p¼0.07 and p


P-116 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ESTABLISHING THE OXIDATION-REDUCTION POTENTIAL IN<br />

SEMEN AND SEMINAL PLASMA. A. Agarwal, a S. S. Du Plessis, a,b<br />

R. Sharma, a L. Samanta, a,c A. Harlev, a,d G. Ahmad, a,e S. Gupta, a<br />

E. S. Sabanegh. f a Center For Reproductive Medicine, Cleveland Clinic,<br />

Cleveland, OH; b Medical Physiology, Stellenbosch University, Tygerberg,<br />

South Africa; c Redox Biology Laboratory, School of Life Sciences, Ravenshaw<br />

University, Orissa, India; d Soroka Medical Center, Ben-Gurion University,<br />

Beer Sheva, Israel; e Physiology and Cell Biology, University of Health<br />

Sciences, Lahore, Pakistan; f Urology, Cleveland Clinic, Cleveland, OH.<br />

OBJECTIVE: Oxidation-reduction potential (ORP) is a novel measure of<br />

oxidative stress or redox imbalance in biological samples. Static ORP<br />

(sORP) provides an integrated measure of the balance between total oxidants<br />

and reductants in a biological system, wheras capacity ORP (cORP) equates<br />

to the amount of antioxidant reserves. sORP has been shown to correlate well<br />

with illness and injury severity that accompanies the state of oxidative stress;<br />

cORP correlates with the ability to respond to illness or injury. Our objectives<br />

were to evaluate whether 1) ORP can be measured in semen and seminal<br />

plasma samples and 2) ORP levels correlate with sperm motility.<br />

DESIGN: Prospective study measuring ORP in both semen and seminal<br />

plasma.<br />

MATERIALS AND METHODS: Semen samples (n¼18) from normal<br />

control subjects were divided into two fractions and the seminal plasma<br />

was isolated from one fraction (300 x g, 7min). Sperm count and motility<br />

were assessed manually. sORP (mV/10 6 sperm) and cORP (mC/10 6 sperm)<br />

were measured in both fractions (RedoxSYSÒ, Aytu BioScience). Values<br />

are reported as Mean SEM. Spearman correlation and Receiver Operating<br />

Characteristic curves (ROC) were used for statistical analysis.<br />

RESULTS: sORP and cORP levels in semen correlated significantly with<br />

the levels in seminal plasma. A significant negative correlation existed between<br />

sperm motility and sORP in both semen (r¼-0.609; p¼0.004) and seminal<br />

plasma (r¼-0.690; p¼0.002). Furthermore, a sORP cutoff of 4.73mV/<br />

10 6 sperm in semen (sensitivity ¼ 100%, specificity ¼ 89.5%,<br />

AUC¼0.947) and 4.65mV/10 6 sperm in seminal plasma (sensitivity ¼<br />

100%, specificity ¼ 93.8%, AUC ¼ 0.969) was highly predictive of abnormal<br />

sperm motility.<br />

CONCLUSIONS: RedoxSYSÒ accurately measured sORP and cORP in<br />

both semen and seminal plasma samples. Based on high sensitivity as assessed<br />

by ROC analysis, sORP levels can be used to screen infertile men<br />

with oxidative stress. These results are being validated in a larger cohort of<br />

infertile men.<br />

ORP levels in semen and seminal plasma.<br />

sORP<br />

(mV/10 6 sperm)<br />

cORP<br />

(mC/10 6 sperm)<br />

Semen 2.790.66 0.200.08<br />

Seminal Plasma 2.670.71 0.310.16<br />

Correlation (r) 0.994 0.959<br />

P


iogenesis (WT, 3.420.54 x 10 5 vs Tg 2.480.64 x 10 5 copies/blastocyst;<br />

p


showed that there was no DNA methylation in SOD1 CpGs in promoter and<br />

intron regions, as well as SOD2 promoter, while SOD2 intron 2 displayed<br />

moderate methylation. Moreover, the methylation level of SOD2-intron in<br />

preeclampsia placentae increased significantly compared with that in normal<br />

third trimester (P < 0.05), which was confirmed by BSP. The methylation<br />

level of SOD1 was inhibited dose-dependently by ADC, and the mRNA<br />

expression of SOD2 elevated following the decrease of methylation.<br />

CONCLUSIONS: The present study demonstrates that the reduced<br />

expression of both SOD1 and SOD2 may be engaged in the increased oxidative<br />

stress in preeclampsia, which is a crucial pathogenesis in the progress of<br />

preeclampsia. In addition, SOD2 expression was down-regulated by elevated<br />

DNA methylation in intron 2 region. This helps us to understand the etiology<br />

of preeclampsia better and to develop new agents to restrain the progress of<br />

preeclampsia.<br />

Supported by: National Basic Research Program of China<br />

(2012CB944900), the Natural Science Foundation of China (81200446,<br />

31<strong>17</strong>1444 and 30973209), the National Science and Technology Support<br />

Program (2012BAI32B01).<br />

P-122 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

REVISITING AN OLD SEMINAL BIOMARKER TO GAUGE<br />

GAMATE MICROENVIRONMENT AND PREGNANCY<br />

OUTCOME. L. Park, Q. V. Neri, L. Reisman, T. Paniza,<br />

Z. Rosenwaks, G. D. Palermo. Reproductive Medicine, Weill Cornell Medical<br />

College, New York, NY.<br />

OBJECTIVE: To assess the efficacy of additional seminal biomarkers in<br />

predicting pregnancy outcome with IUI.<br />

DESIGN: In a time interval of 12 months, we assessed gamete microenvironment<br />

by measuring total antioxidant capacity (TAC) and seminal fructose<br />

(FRU) on 134 men. To account for fructose consumption by spermatozoa, the<br />

fructose value was corrected (cFRU) by multiplying it with the logarithm of<br />

sperm concentration. Correlation between TAC and cFRU as well as with<br />

clinical outcome were assessed.<br />

MATERIALS AND METHODS: FRU (mg/ml) and TAC (nmol/ml) was<br />

assessed using a colormetric assay on an automated microplate reader<br />

(DOD405/570). cFRU was calculated by multiplying the seminal fructose<br />

by the logarithm of the sperm concentration (for men with >1 million<br />

sperm/ml). IUI pregnancy outcomes were assessed.<br />

RESULTS: A total of 135 men (38.5 8 years old) had a sperm concentration<br />

of 26.8 29.4million/ml with a motility of 39.4 14.2% and<br />

morphology of 2.3 1.0%. The average TAC was 1834.5 278.4nmol/ml<br />

and significantly correlated with fructose (1.97 1.13mg/ml, P <<br />

0.000005). Once fructose was corrected for sperm concentration, the average<br />

cFRU was 2.531.61mg/ml and also correlated with TAC (P < 0.0000001),<br />

as well as with male age (P ¼ 0.03) and sperm motility (P ¼ 0.009). A group<br />

of 41 men were treated in 100 IUI cycles with their female partner (35.9 5<br />

years) and reported a clinical pregnancy rate of 15.0%. The average TAC was<br />

1862.6193.6mM and the average cFRU was 3.02 1.58mg/ml. There were<br />

no significant differences among male and female ages as well as sperm parameters<br />

between couples that achieved a clinical pregnancy (n ¼ 15) and<br />

those who did not (n ¼ 26). However, TAC and cFRU in the successful<br />

outcome group was both remarkably higher than those that failed (P ¼<br />

0.002 and P ¼ 0.02 respectively), even after controlling for an eventual female<br />

factor (%35 years; P ¼ 0.02).<br />

CONCLUSIONS: Information acquired from standard semen analysis has<br />

a limited predictability on the fertile status of men with adequate semen parameters.<br />

Additional seminal vesicle biomarkers such as TAC and corrected<br />

fructose level provide information on the male genital tract microenvironment<br />

and male gamates’ ability to achieve pregnancy.<br />

Supported by: WCMC.<br />

OBJECTIVE: Does Fertilix, a novel antioxidant formulation designed to<br />

alleviate oxidative stress in the male reproductive tract, reduce Sperm<br />

DNA Damage (SDD) and increase pregnancy rates in mouse models of<br />

Sperm Oxidative Stress (SOS)?<br />

DESIGN: Fertilix efficacy was evaluated in two, well-established mouse<br />

models of SOS; Glutathione Peroxidase 5 knockout (GPX-5 KO) mice and<br />

Scrotal Heat Shock (SHS) treatment, each with n¼12, by independent laboratories.<br />

Mice were given Fertilix in their drinking water for 2 months (GPX-<br />

5 KO) and 2 weeks (SHS) and then compared with control groups for SDD<br />

and pregnancy rates.<br />

MATERIALS AND METHODS: In GPX-5 KO mice, SDD was evaluated<br />

by immunocytochemical detection of 8-Hydroxy-deoxy Guanosine (8-<br />

OHdG) residues, a biomarker of DNA oxidation. In the SHS model, each<br />

male’s fertility was tested by partnering with three females for 5 days. The<br />

percentage of pregnant females, number of vaginal plugs, resorptions per<br />

litter, and litter size were recorded.<br />

RESULTS: The average background levels of 8-OHdG in WT mice is<br />

around 30%. This level doubles up to about 60% in transgenic male mice<br />

deficient in the antioxidant enzyme GPX-5. Our results indicate that a 2<br />

month pretreatment of GPX-5 KO mice with Fertilix provides complete protection<br />

of sperm DNA against oxidation. In male mice exposed to the SHS<br />

model, only 35% (19/54) of partnered female mice got pregnant resulting<br />

in 169 fetuses. In contrast, male mice pretreated with Fertilix for 2 weeks<br />

yielded 74% (42/57) of partnered female mice becoming pregnant, resulting<br />

in 427 fetuses. The role of chance in obtaining supporting results for the efficacy<br />

of Fertilix in both models is minimal although it was not possible to<br />

ensure that every mouse took 100% of the product for the treatment period.<br />

CONCLUSIONS: Oral administration of Fertilix significantly reduces<br />

SDD in GPX-5 knockout mice and restores pregnancy rates almost back to<br />

normal levels in mice subjected to SHS. These results, if confirmed in humans,<br />

will impact clinical fertility practice. Current clinical studies confirm<br />

moderate to severe SDD in about 60% of all men visiting IVF centers and in<br />

about 80% of men diagnosed with idiopathic male infertility, figures which<br />

are gravely concerning. Antioxidant supplementation will be an adjuvant<br />

therapy prior to undertaking ART procedures to improve fertilization rates,<br />

maintain a healthy pregnancy, and reduce de novo sporadic mutations being<br />

passed onto children.<br />

Supported by: The study was funded by the University of Clermont and the<br />

University of Madrid.<br />

P-124 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

WITHDRAWN<br />

REPRODUCTIVE IMMUNOLOGY<br />

P-123 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

FERTILIX, A NOVEL ANTIOXIDANT FORMULATION DESIGNED<br />

TO TREAT MALE INFERTILITY EMANATING FROM SPERM<br />

OXIDATIVE DNA DAMAGE: PROMISING PRECLINICAL<br />

EVIDENCE FROM MOUSE MODELS. A. Moazamian, a<br />

P. Gharagozloo, a J. Drevet, b A. Gutierrez-Adan, c A. Kocer, b A. Calle, c<br />

E. Pericuesta, c A. M. Polhemus, a R. J. Aitken. d,a a CellOxess LLC, Princeton,<br />

NJ; b University Blaise Pascal, Clermont-Ferrand, France; c INIA Animal<br />

Reproduction, Madrid, Spain; d The University of Newcastle, Callaghan,<br />

NSW, Australia.<br />

e148 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


4. Meuchel LW, Stewart A, Smelter DF, Abcejo AJ, Thompson MA, Zaidi<br />

SL, et al. Neurokinin-neurotrophin interactions in airway smooth muscle.<br />

Am J PHysiol Lung Cell Mol Physiol 2011;301:L91-8.<br />

5. Argawal A, Saleh Ra, Bedaiwy A. Role of reactive oxygen species in<br />

the pathophysiology of human reproduction. Fertil Steril<br />

2003;79:829-43.<br />

P-126 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

CONCENTRATIONS OF INSULIN-LIKE GROWTH FACTOR (IGF)-<br />

1 AND 2 AT THE BEGINNING OF A MATCHED DONOR CYCLE<br />

BEFORE STIMULATION PREDICT OUTCOME IN<br />

RECIPIENTS. T. Kanninen, a I. Ramer, a G. Sisti, a S. S. Witkin, a<br />

S. D. Spandorfer. b a Weill Cornell Medical College, New York, NY; b Cornell<br />

University Medical Center, New York, NY.<br />

P-125 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

SERUM LEVELS OF BRAIN-DERIVED NEUROTROPHIC FACTOR<br />

PRIOR TO INITIATION OFAN IN VITRO FERTILIZATION CYCLE<br />

PREDICT OUTCOME. I. Ramer, a T. T. Kanninen, a G. Sisti, a<br />

S. S. Witkin, a S. D. Spandorfer. b a Cornell University Medical College,<br />

New York City, NY; b Cornell University Medical Center, New York City, NY.<br />

OBJECTIVE: Prior research has shown that brain-derived neurotrophic<br />

factor (BDNF) levels in sera from women on day 28 of an in vitro fertilization<br />

(IVF) cycle were highly predictive of a live birth. Most BDNF that is present<br />

in sera is derived from activated platelets and recent data has shown that<br />

platelet activation is essential to the initiation of innate and acquired immune<br />

system activation in response to non-physiological conditions. Thus, the<br />

release of BDNF by platelet activation may serve as a sensitive signal for<br />

the presence of physiologic stress. The goal of this study was to determine<br />

if the concentration of circulating BDNF prior to cycle initiation (day 2) predicts<br />

outcome in women undergoing IVF.<br />

DESIGN: This is a cross-sectional, retrospective study of women undergoing<br />

IVF at Weill Cornell Medical College.<br />

MATERIALS AND METHODS: Sera was obtained on day 2 of an IVF<br />

cycle from 54 (23.9%) women with a live birth, 45 (19.9%) with a spontaneous<br />

abortion, 38 (16.8%) with a biochemical pregnancy, 54 (23.9%) who<br />

did not become pregnant, and 35 (15.5%) with an ectopic pregnancy.<br />

BDNF concentration was determined by ELISA. Associations with clinical<br />

parameters were assessed by the Kruskall-Wallis and Mann-Whitney tests.<br />

The ability to predict an ectopic pregnancy was determined by receiver operator<br />

curve analysis.<br />

RESULTS: Median serum concentrations of BDNF were lowest in women<br />

who had a subsequent live birth (3.6 ng/ml), compared to women with an<br />

ectopic pregnancy (7.3 ng/ml), those who did not become pregnant (5.5<br />

ng/ml), women with a biochemical pregnancy (3.8 ng/ml) or a spontaneous<br />

abortion (4.2 ng/ml) (p37 weeks) in<br />

women who were recipients.<br />

DESIGN: Sera from 59 oocyte recipients who had a liveborn, collected at<br />

the initiation of their recipient cycle and frozen, were thawed and assayed for<br />

concentrations of mediators reported to be involved in determining pregnancy<br />

outcome. All of the patients underwent fresh anonymous oocyte donation.<br />

After down regulation with OCPs and leuprolide acetate, serum was<br />

collected. This was before any therapy with estradiol in the egg donation cycle.<br />

MATERIALS AND METHODS: IGF-1, IGF-2, IGF binding protein-1,<br />

pro-inflammatory cytokines (Interleukin-1beta, interleukin-6, tumor necrosis<br />

factor alpha) and anti-inflammatory cytokines (interleukin-13, interleukin-<br />

<strong>17</strong>) in sera were assayed in duplicate by commercial ELISA kits. Values<br />

were converted to ng/ml or pg/ml by reference to a standard curve generated<br />

in parallel to each assay. Differences in biomarker levels between women<br />

who delivered at term (>37 weeks) or preterm were analyzed by the<br />

Mann-Whitney. Associations between gestational age at delivery and either<br />

biomarker concentration or maternal age were analyzed by the Spearman<br />

rank correlation test.<br />

RESULTS: Ten women delivered preterm (32-36.5 weeks) while the remaining<br />

49 had a term birth (37-41.6 weeks). The median concentration of<br />

IGF-1 was <strong>17</strong>.1 ng/ml in women who delivered at term as compared to <<br />

0.03 ng/ml in women with a preterm birth (p ¼ 0.0014). IGF-1 levels were<br />

positively associated with gestational age at delivery (p ¼ 0.0345) but not<br />

associated with maternal age. Conversely, median IGF-2 levels were higher<br />

in women with a preterm delivery (452 pg/ml) than in those with a term birth<br />

(365 pg/ml) (p ¼ 0.0090). IGF-2 levels were inversely correlated with gestational<br />

age at delivery (p ¼ 0.0055) but not with maternal age.<br />

CONCLUSIONS: Determination of IGF-1 and IGF-2 levels in sera of<br />

oocyte donor recipients at the start of their IVF cycle can predict who will<br />

most likely have a preterm birth and who will deliver at term.<br />

P-127 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

INTRAUTERINE HCG INFUSION AFFECTS THE DISTRIBUTION<br />

OF NATURAL KILLER CELLS IN THE ENDOMETRIUM OF<br />

FERTILE OOCYTE DONORS. E. Giuliani, a M. Olson, b M. Strug, b<br />

J. Young, c V. Shavell, c W. Dodds, c R. Leach, b A. Fazleabas. b a Department<br />

of Obstetrics and Gynecology, Grand Rapids Medical Education Partners/<br />

Michigan State University, Grand Rapids, MI; b Department of Obstetrics,<br />

Gynecology and Reproductive Biology, Michigan State University, Grand<br />

Rapids, MI; c The Fertility Center, Grand Rapids, MI.<br />

OBJECTIVE: Reports on intrauterine (IU) human chorionic gonadotropin<br />

(hCG) administration prior to embryo transfer suggest a potential positive effect<br />

on pregnancy rates possibly by promoting the expression of genes and<br />

proteins that are critical for implantation (1,2,3). A prior study by our group<br />

demonstrated that patients with infertility and endometriosis have dysregulated<br />

expression of endometrial CD16 and CD56 natural killer cells (NK)<br />

(4). NK cells play a key role in implantation since they can produce angiogenic<br />

factors that promote spiral artery remodeling and cytokines (LIF, leukemia<br />

inhibitory factor) that direct the migration of trophoblasts (4). The goal<br />

of this study was to investigate if a single IU hCG infusion during the period<br />

FERTILITY & STERILITY Ò<br />

e149


corresponding to embryo transfer could induce histological changes in the<br />

distribution of NK cells in the uterine endometrium.<br />

DESIGN: Case-control study.<br />

MATERIALS AND METHODS: Three days following controlled ovarian<br />

hyperstimulation and oocyte retrieval, fertile oocyte donors were randomly<br />

assigned to receive an IU infusion of either 500 IU hCG (n¼7) or Vehicle<br />

(n¼8). Endometrial biopsies were performed 48h post infusion. Subsequent<br />

analysis included blinded histological staining and quantification (Image J<br />

software) of CD56+ and CD16+ stroma NK cells in the endometrial biopsies.<br />

Paired T-test was used to analyze the results between the two treated groups.<br />

RESULTS: No statistically significant differences in the age, parity, oocytes<br />

retrieved and estradiol levels before retrieval were found between the<br />

two study groups. The percentage of CD16+ stroma cells was higher after<br />

the infusion of hCG (1.320.05 vs 1.180.13, p4 fold; P


OBJECTIVE: Advanced ovarian age is often implicated in implantation<br />

failure and early pregnancy loss in women aged 40 and over. The role of<br />

the uterus in reproductive aging has been investigated, but with conflicting<br />

conclusions. We seek to investigate the impact of uterine age on reproductive<br />

capacity while controlling for oocyte quality by using Comprehensive Chromosomal<br />

Screened Embryos.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Women undergoing in vitro fertilization<br />

(IVF) with preimplantation genetic screening (PGS) from September 2010 to<br />

March <strong>2015</strong>, who used either donor or autologous oocytes, were included.<br />

Fresh or thawed euploid blastocysts, confirmed by PGS, were selected for<br />

embryo transfer (ET). Student’s t-test calculated mean values and a Poisson<br />

regression analysis examined the effect of uterine and oocyte age on IVF outcomes.<br />

RESULTS: Eight hundred and sixty nine women’s (range of 23-54 years<br />

(36.75+/-4.3)) cycles were analyzed. Autologous and donor oocytes ranged<br />

from 23-39 years (35.88+/-4.46). Per each additional year of embryo recipient<br />

age, the endometrial thickness decreased by 0.028mm (p¼0.086). Per<br />

each additional year of oocyte age, the number of eggs retrieved significantly<br />

decreased by a factor of 2.5% (p


P-133 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE PREVALENCE AND MANAGEMENT OF SMALL TESTIC-<br />

ULAR MASSES INCIDENTALLY DISCOVERED ON ULTRASOUND<br />

EVALUATION OF MALE INFERTILITY. J. M. Bieniek, a T. Juvet, a<br />

M. Margolis, b E. D. Grober, a K. C. Lo, a K. A. Jarvi. a a Urology, University<br />

of Toronto, Toronto, ON, Canada; b Radiology, University of Toronto, Toronto,<br />

ON, Canada.<br />

OBJECTIVE: Scrotal ultrasonography (US) is easily employed in the<br />

evaluation of men in infertile couples but increases the prevalence of incidental<br />

testicular lesions, therefore, this study was performed to assess the natural<br />

history of these small testicular masses (STMs) and possible indications<br />

for intervention.<br />

DESIGN: Retrospective review of patients presenting for male infertility<br />

evaluation who were identified to have an incidentally-found small testicular<br />

mass with analysis of patient variables and management strategies.<br />

MATERIALS AND METHODS: After obtaining institutional review<br />

board approval, patients presenting to a high-volume infertility clinic were<br />

reviewed to identify those who had tumor markers drawn or completed<br />

more than one scrotal US between 2001 and July 2014. Charts were reviewed<br />

to identify patients with one or more incidentally discovered small (10mm or<br />

less) hypoechoic masses. All other testicular lesions were excluded. Patient<br />

demographics, further imaging, laboratory evaluation, and treatment were<br />

collected and statistical analysis performed with Student’s t tests and chi<br />

square tests.<br />

RESULTS: Of 4088 men completing a scrotal US for evaluation of infertility<br />

during the study time period, 120 (2.9%) were found to have small<br />

testicular masses. The mean age of men with a STM finding was 36.7 (range<br />

23-62) years. Patients were followed with serial imaging for these lesions unless<br />

concerns existed regarding the US appearance, prior history of testis cancer,<br />

or due to patient concerns. Average follow-up duration was 1.3 years<br />

(0.1-16.9) with 18 men (15.0%) having more than two years of follow-up.<br />

Mean initial maximum mass dimension was 4.14mm (2.0) with vascular<br />

flow noted in 38 patients (42.2%). Tumor markers were obtained in 54<br />

(45.0%) with only one positive result (1.9%). Among patients followed at<br />

least one month, interval mass growth averaged 0.1mm/yr. Eighteen patients<br />

(15.0%) underwent extirpative surgery with 6 (33.3%) having malignant lesions<br />

and the remaining 12 (66.7%) benign pathologies. All cases of malignant<br />

disease were pure seminomas, measured greater than 5.2mm on US, and<br />

demonstrated vascularity. Significant differences were found between the<br />

surgery and surveillance groups in initial max lesion size (5.38 vs 3.92mm,<br />

p¼ .004) and presence of vascularity (81.3% vs 33.8%, p¼ .001). Due to relatively<br />

small numbers, no significant differences were found between men<br />

with benign and malignant lesions on pathology.<br />

CONCLUSIONS: This series represents the largest of its kind with nearly<br />

3% of men undergoing scrotal US for infertility evaluation being diagnosed<br />

with a STM. While the majority of incidental small testicular masses did not<br />

demonstrate significant growth, lesions presenting at greater than 5mm in<br />

size with vascularity were more likely to undergo surgery and potentially harbor<br />

a malignancy necessitating urologic evaluation in this population.<br />

P-134 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

NON-INVASIVE TECHNOLOGY COMBINING TIME-LAPSE IM-<br />

AGING AND STATISTICAL MODELING: BRINGING AUTOMA-<br />

TION INTO THE LAB TO IMPROVE BLASTOCYST<br />

SELECTION. B. Behr, a L. Tan, b J. Conaghan, c J. Liebermann, d<br />

A. Bartolucci, e A. A. Chen. b a Stanford Fertility and Reproductive Medicine<br />

Center, Palo Alto, CA; b Progyny, Menlo Park, CA; c PFC, San Francisco, CA;<br />

d Fertility Centers of Illinois, Chicago, IL; e CARS, Farmington, CT.<br />

OBJECTIVE: The Eeva Test combines automated time-lapse analysis<br />

with statistical modeling to predict continued embryo development. It automatically<br />

captures quantitative image features that the human eye cannot<br />

detect, and incorporates statistical modeling of dynamic parameters to maximize<br />

the test’s predictive power. A new predictive algorithm includes parameters<br />

of patient prognosis, morphology, early cleavage timings and<br />

quantitative image features to generate a 5-category score for developmental<br />

potential. The objective of this multi-center study was to validate this new<br />

algorithm against embryo implantation following blastocyst transfer.<br />

DESIGN: Retrospective multi-center study<br />

MATERIALS AND METHODS: The study included a total of 151 patients<br />

from 9 centers who consented to use the Eeva Test. Embryos were<br />

transferred on Day 5, and embryos with known implantation data were<br />

analyzed. Images and clinical data were processed by the Eeva Test’s new<br />

Xtend algorithm, which utilizes automated image analysis software to classify<br />

embryos into five categories by incorporating key cell division timing<br />

parameters, Day 3 cell#, age and quantitative image features reflective of activity<br />

during the third cell cycle.<br />

RESULTS: The overall known implantation rate (IR) for the study population<br />

was 39% (84/<strong>21</strong>6). The 5-category output of the extended algorithm<br />

correlated positively with blastocyst IR: 51% (36/71), 45% (33/74), 25%<br />

(10/40), 24% (5/<strong>21</strong>), 0% (0/10). In patients younger than 35yo, the overall<br />

known IR was 50% (62/123); however, the IR for ‘‘Category 1’’ embryos<br />

was 60% (32/53). In the same patient group, IR for the good morphology<br />

blastocysts was 56% (54/97); however, embryos with good morphology<br />

and Category 1 achieved a 61% (31/51) IR.<br />

CONCLUSIONS: The novel aspect of the new algorithm is including<br />

computer-extracted quantitative image attributes, patient characteristics<br />

and traditional morphology parameters in the predictive model. Although<br />

the model was built using blastocyst as outcome variables, in current study<br />

we were able to demonstrate that 5-category output in the new algorithm<br />

correlated positively with blastocyst implantation. Specifically, in young patients<br />

who are more likely in need of embryo selection, Category 1 identified<br />

a subset of embryos with 60% implantation rate, which approaches the implantation<br />

rate of euploid blastocysts (Forman et al, 2013). The algorithm<br />

also identified embryos with better implantation potential among good<br />

morphology blastocysts from young patients. Combining extended algorithm<br />

and traditional morphology may present a non-invasive approach assisting<br />

embryologists to select the best blastocyst(s) for transfer, and encourage<br />

the practice of elective SET.<br />

Supported by: Progyny (formerly Auxogyn).<br />

P-135 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

IMPACT OF SALINE INFUSION SONOGRAM PERFORMED DUR-<br />

ING CONTROLLED OVARIAN STIMULATION ON IN VITRO<br />

FERTILIZATION-EMBRYO TRANSFER OUTCOMES OF THE<br />

SAME CYCLE. N. Pereira, a A. P. Hutchinson, b J. Lekovich, a<br />

R. T. Elias, a Z. Rosenwaks. b a The Ronald O. Perelman and Claudia Cohen<br />

Center for Reproductive Medicine, New York, NY; b Weill Cornell Medical<br />

College, New York, NY.<br />

OBJECTIVE: To compare the impact of saline infusion sonogram (SIS)<br />

performed after initiation of gonadotropin stimulation during an ongoing<br />

in vitro fertilization (IVF) cycle to SIS performed in the follicular phase of<br />

an earlier menstrual cycle.<br />

DESIGN: Retrospective single-center cohort study.<br />

MATERIALS AND METHODS: All patients undergoing fresh in vitro<br />

fertilization (IVF) - embryo transfer (ET) cycles between January 2008<br />

and June 2013. Patients were stratified into 2 groups based on whether<br />

they underwent SIS during gonadotropin stimulation of an ongoing IVF cycle<br />

or during the follicular phase of an earlier menstrual cycle. Demographic and<br />

baseline characteristics were extracted from patient charts and included age,<br />

gravidity, parity, body mass index (kg/m 2 ), total days of ovarian stimulation,<br />

total dosage of gonadotropins administered (IU), number of oocytes<br />

retrieved, and number of embryos transferred. Clinical pregnancy, biochemical<br />

pregnancy, spontaneous miscarriage and live birth rates were compared<br />

between the two groups. Tests of equivalence i.e., student’s t-tests and Chisquare<br />

(c2) tests were used as indicated. Statistical significance was set at<br />

P


Group 1: SIS performed during same IVF cycle; Group 2: SIS performed in a<br />

previous menstrual cycle.<br />

Parameter Group 1 (n¼152) Group 2 (n¼7648) P<br />

Age (years) 35.7 (4.99) 35.9 (4.95) 0.62<br />

BMI (kg/m2) 23.2 (5.99) 23.1 (5.89) 0.84<br />

Total stimulation days 9.28 (1.90) 9.13 (1.54) 0.24<br />

Total gonadotropins (IU) 3161.7 (1433.4) 3118.6 (1498.1) 0.72<br />

Number of oocytes retrieved 12.1 (7.61) 11.7 (6.42) 0.45<br />

Number of embryos transferred 2.34 (1.16) 2.41 (1.19) 0.47<br />

Clinical pregnancy rate 64 (42.1%) 2967 (38.8%) 0.41<br />

Biochemical pregnancy rate 12 (7.89%) 578 (7.56%) 0.88<br />

Spontaneous miscarriage rate 8 (5.26%) 275 (3.60%) 0.28<br />

Live birth rate 55 (36.2%) 2687 (35.1%) 0.79<br />

P-136 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ARE MORPHOKINETIC PARAMETERS AFTER THAWING<br />

RELATED TO IMPLANTATION IN DAY 3 CRYOPRESERVED<br />

EMBRYOS?. E. Fernandez Gallardo, C. Spiessens, T. D’Hooghe,<br />

S. Debrock. Leuven University Fertility Centre, Leuven University Hospital,<br />

Leuven, Belgium.<br />

OBJECTIVE: To study, for the first time, the relation between morphokinetic<br />

parameters after thawing and implantation in intact and non-intact vitrified/warmed<br />

and slow frozen/thawed embryos.<br />

DESIGN: Presence of mitosis, presence of compaction, time to mitosis<br />

and time to compaction were analyzed retrospectively for 82 embryos (35<br />

slow frozen and 47 vitrified). After survival, embryos were cultured overnight<br />

in ASTECÒ time lapse incubator (TLI) and transferred. Implantation<br />

was defined as the presence of fetal sac (intra- or extrauterine) at 6-8 weeks<br />

pregnancy after transfer. All transfers included in the study had either 0% or<br />

100% embryos implanted.<br />

MATERIALS AND METHODS: Embryos were vitrified (EmbryoStore,<br />

GynemedÒ) or slow frozen (Vitr Kit Freeze, Irvine ScientificÒ) on day 3<br />

if R6 cells and


concerns (75%) in comparison to other women in the study- 15% Caucasian,<br />

5% Hispanic and 5% Asian.<br />

CONCLUSIONS: Most of the women in this qualitative study raised financial<br />

concerns relating to their fibroids in their interviews. African-American<br />

women, however, brought up financial issues substantially more than White,<br />

Hispanic or Asian women. This is particularly concerning given that fibroids<br />

disproportionately affect African-American women. For the women who expressed<br />

financial challenges, they were mainly related to an inability to pay<br />

for diagnostics and treatment needed to alleviate symptoms or for definitive<br />

treatment. More work needs to be done to improve insurance coverage for<br />

fibroid diagnosis and treatment as well as to quantitatively define the financial<br />

impact of these prevalent tumors on individual women.<br />

Supported by: NIH WRHR Program K12HD0501<strong>21</strong>; RWJ Foundation;<br />

NMH; Evergreen Foundation.<br />

P-139 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ROLE OF DOPAMINE AGONISTS IN UTERINE<br />

MYOMA. B. Bhagavath, a H. Prizant, b S. R. Hammes. c a OBGYN, Strong<br />

Fertility Center, University of Rochester, Rochester, NY; b OBGYN, University<br />

of Rochester, Rochester, NY; c University of Rochester, Rochester, NY.<br />

OBJECTIVE: 1. Study the expression of dopamine receptors, prolactin<br />

and prolactin receptors in leiomyoma in an animal model for uterine fibroids<br />

(Uterine-specific Tsc2 knockout (Tsc2 KO) mice)2. Study the effect of dopamine<br />

agonists on the growth of leiomyoma in Tsc2 KO mice<br />

DESIGN: Interventional study using a mouse model for leiomyoma<br />

MATERIALS AND METHODS: Tsc2 KO mice are known to develop<br />

leiomyoma spontaneously. Previously performed experiments have established<br />

that these tumors are estrogen sensitive and respond well to treatment<br />

with aromatase inhibitor, Letrozole, which decreases circulating levels of estrogen.<br />

Real time quantitative polymerase chain reaction (qPCR) was performed<br />

on mRNA from the uterine tissue to compare prolactin (PRL),<br />

prolactin receptor (PRLR) and dopamine receptor (DRD2) expression with<br />

that of wild type (wt) mice. Serum PRL levels were compared between wt<br />

mice and Tsc2 KO mice. Ovaries were surgically removed (ovariectomy)<br />

from 18 week wt and TSC2 KO mice and mice were treated with estradiol<br />

for 1 week. Uterine PRL, PRLR and DRD2 expressions were studied in these<br />

mice. To study the effect of dopamine, four knockout animals were injected<br />

with cabergoline (dopamine agonist) subcutaneously three times a week for<br />

12 weeks from 18 weeks age. At 30 weeks age, these animals were sacrificed<br />

and their uteri subjected to gross examination. In addition, the uteri were<br />

sectioned, fixed and stained with hematoxylin and eosin, and compared<br />

with uteri of wt mice.<br />

RESULTS: In Tsc2 KO mice at 18 weeks age, PRLR but not DRD2<br />

expression markedly and significantly increased after administration of estradiol<br />

for one week. Local PRL levels are undetectable in the uteri of wt mice<br />

and detected in small quantities in Tsc2 KO mice at 18 and 30 weeks age.<br />

Ovariectomy at 18 weeks age for a period of 12 weeks significantly reduced<br />

PRLR expression and PRL is no longer detectable in the uteri of uterine-specific<br />

Tsc2 KO mice at 30 week age. Serum prolactin levels were significantly<br />

suppressed in Cabergoline treated mice. Gross pathological examination did<br />

not show any improvement in the leiomyoma in Tsc2 KO mice treated with<br />

Cabergoline. Examination of fixed and stained sections of uteri are indistinguishable<br />

from those of untreated Tsc2 KO mice.<br />

CONCLUSIONS: Presence of local uterine PRL production in Tsc2 KO<br />

mice along with PRLR suggests a complex, estrogen dependent role in leiomyoma<br />

development. Although DRD2 agonist did not show a significant<br />

clinical effect on leiomyoma development, it is possible that a PRLR antagonist<br />

might have a clinical effect on leiomyoma development.<br />

Supported by: This work was Supported by a grant from the Richard W.<br />

Goode and Mae Stone Goode Foundation.<br />

P-140 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

HISTORY OF UTERINE LEIOMYOMATA AND TIME-TO-PREG-<br />

NANCY IN BLACK WOMEN. L. A. Wise, a R. Radin, b<br />

L. Rosenberg. c a Department of Epidemiology, Boston University School of<br />

Public Health, Boston, MA; b NIH, Rockville, MD; c Slone Epidemiology<br />

Center, Boston University, Boston, MA.<br />

OBJECTIVE: To evaluate the extent to which a history of uterine leiomyomata<br />

(UL) and its characteristics (tumor location, size, and number) are associated<br />

with delayed time-to-pregnancy (TTP).<br />

DESIGN: Prospective cohort study.<br />

MATERIALS AND METHODS: We assessed the association between UL<br />

and fecundability among participants aged <strong>21</strong>-40 years from the Black<br />

Women’s Health Study. In 2008, women with incident diagnoses of UL<br />

(1995-2008) reported detailed data on their date of UL diagnosis, symptoms,<br />

tumor characteristics and treatment via supplemental questionnaire. In 2011,<br />

women reported the time-to-pregnancy (TTP) in months for each of their<br />

planned pregnancies resulting in a live birth during their reproductive life.<br />

The incident period for the analysis was restricted to 1995-2011 and only<br />

women with UL diagnoses prior to a given pregnancy attempt were considered<br />

‘‘exposed.’’ Proportional probabilities regression models were used to<br />

estimate fecundability ratios (FRs) and 95% confidence intervals (CI), with<br />

adjustment for age and calendar year at pregnancy attempt, prepregnancy<br />

BMI, physical activity, education, and smoking history.<br />

RESULTS: During 1995-2011, 1,353 women contributed 1,572 births and<br />

8,250 months of attempt time. A history of UL was associated with a slight<br />

delay in conception: the FR for a history of UL diagnosis was 0.92 (95% CI:<br />

0.81-1.05) relative to no UL diagnosis (referent for all comparisons). The FR<br />

for submucosal UL was 0.71 (95% CI: 0.45-1.12) and 0.91 (95% CI: 0.63,<br />

1.31) for intramural UL. FRs for women with 1, 2, and R3 UL tumors<br />

were 0.89 (95% CI: 0.63-1.24), 0.89 (95% CI: 0.65-1.<strong>21</strong>), and 0.87 (95%<br />

CI: 0.68-1.10), respectively. FRs for women with UL tumors


References:<br />

1. Kho K, Nezhat C. Evaluating the Risks of Electric Uterine Morcellation.<br />

JAMA 2014 (311) 9: 905-906.<br />

2. Seidman MA, Oduyebo T, Muto MG, Crum CP, Nucci MR, Quade BJ.<br />

Peritoneal dissemination complicating morcellation of uterine mesenchymal<br />

neoplasms. PLoS One. 2012;7(11):e50058.<br />

3. Steiner RA, Wight E, Tadir Y, et al. Electrical cutting device for laparosopic<br />

removal of tissue from the abdominal cavity. Obstet Gynecol.<br />

1993;81:471-474.<br />

4. Milad MP, Milad EA. Laparoscopic morcellator-related complications.<br />

J Minim Invasive Gynecol. 2014;<strong>21</strong>(3):486-491.<br />

5. American Association of Gynecologic Laparoscopists-Tissue Extraction<br />

Task Force. Morcellation during uterine tissue extraction. Available at:<br />

www. Aagl.org/wp-content/uploads/2014/05/Tissue_Extraction_TFR.pdf.<br />

P-142 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

PARADOXIC EFFECT OF MIFEPRISTONE ON LEIOMYOMA<br />

EXTRACELLULAR MATRIX COMPONENT DERMATOPONTIN<br />

EXPLAINED. A. Patel, a M. Malik, a J. L. Britten, a J. Cox, b<br />

W. H. Catherino. a a Obstetrics and Gynecology, Uniformed Services University<br />

of the Health Sciences, Bethesda, MD; b Program in Reproductive and<br />

Adult Endocrinology, Eunice Kennedy Shriver National Institute of Child<br />

Health and Human Development, Bethesda, MD.<br />

OBJECTIVE: Progesterone stimulates leiomyoma growth, but the impact of<br />

this hormone on the extra cellular matrix (ECM) genes remains unknown. Dermatopontin<br />

is a unique ECM gene that is known to be decreased in leiomyomas<br />

compared to patient-matched myometrium, in contradistinction to other<br />

biomarker ECM genes which are elevated. If progesterone promotes the leiomyoma<br />

phenotype of excessive and aberrant ECM, then treatment should result<br />

in decreased dermatopontin, while treatment with the anti-progestin mifepristone<br />

should increase expression. Our objective is to characterize the mechanism<br />

of progesterone on the expression of the ECM protein dermatopontin.<br />

DESIGN: Laboratory setting<br />

MATERIALS AND METHODS: Immortalized leiomyoma cells were<br />

treated with the progestin R5020 and the anti-progestin mifepristone at clinically-relevant<br />

concentrations and various time-points up to 120 hours. Dermatopontin<br />

protein concentration was quantified using Western blot.<br />

RESULTS: At 1 and 6 hours exposure, paradoxically both progestin and<br />

mifepristone-treated cultures resulted in a reduction of dermatopontin protein<br />

concentrations (0.50.23 fold and 0.7 0.08 fold at 1 hour and<br />

0.40.19 fold and 0.6 0.14 fold at 6 hours) respectively. At 72 and 120<br />

hours, dermatopontin expression remained decreased (0.<strong>17</strong> 0.2 fold) with<br />

mifepristone treatment. However, in the same time-frame, progestin impact<br />

on dermatopontin expression was lost (1.0 0.25 fold at 72 hours) and (0.6<br />

0.34 fold at 120 hours ).<br />

CONCLUSIONS: Our results demonstrate that mifepristone has a progestational<br />

effect on dermatopontin protein concentration, and this effect is<br />

maintained despite a diminution of effect by a standard progesterone receptor<br />

agonist. These findings suggest the production of specific signaling pathway<br />

proteins by progestins that regulate dermatopontin expression. Better understanding<br />

of this pathway would result in improved therapeutic options for<br />

women suffering with symptomatic uterine leiomyomas.<br />

Supported by: This research was Supported by Intramural grant from Uniformed<br />

Services University of the Health Sciences, QP85GF13 and R<strong>21</strong>,<br />

R085193713.The research was also Supported, in part, by the intramural<br />

research Program in Reproductive and Adult Endocrinology, NIH.<br />

P-143 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

SEEKING THE TRUTH: A QUALITATIVE ASSESSMENT OF<br />

WOMEN’S EXPERIENCES SEEKING AND OBTAINING KNOWL-<br />

EDGE OF UTERINE FIBROIDS. M. Ghant, a K. Sengoba, b<br />

G. Mendoza, b E. E. Marsh. b a University of Illinois at Chicago College of<br />

Medicine, Chicago, IL; b Northwestern University Feinberg School of Medicine,<br />

Chicago, IL.<br />

OBJECTIVE: To identify thematic content on health information-seeking<br />

amongst women with fibroids.<br />

DESIGN: Qualitative semi-structured interviews and demographic surveys<br />

MATERIALS AND METHODS: Women between the ages of 25-55, who<br />

either currently had symptomatic uterine fibroids or received treatment for<br />

their uterine fibroids within the past 12 months, were recruited from an urban<br />

academic medical center and community-based organizations. Women<br />

completed in-depth, one-on-one recorded interviews, a demographic survey<br />

and a health literacy assessment tool. Interviews were transcribed verbatim<br />

and uploaded to NVivo version 10 for data management and thematic coding.<br />

Three coders used a grounded theory approach to identify major themes and<br />

subthemes.<br />

RESULTS: Sixty women completed the study, resulting in a total of 35<br />

hours of interviews yielding 1,357 transcribed pages. The k across coders<br />

was 0.94. The mean age of participants was 43.0 6.8 (mean SD).<br />

61.7% of participants self-identified as African-American, 25.0% as Caucasian,<br />

8.3% as Hispanic and 5.0% as Asian. 68.3% of the participants had at<br />

least a 4-year college degree. The average health literacy score was 6.8 on<br />

a 7-point scale. Four themes were identified regarding information seeking<br />

and sharing amongst participants. Theme 1 focused on sources of information,<br />

with many women reporting that they used the Internet, family and<br />

friends, physicians, books, the radio and church to obtain information on<br />

uterine fibroids. Theme 2 focused on women’s dissatisfaction with available<br />

information. Despite utilization of multiple resources, the majority of women<br />

(78.3%) were dissatisfied with either the clarity, quality, or quantity of available<br />

information on fibroids for patients. Theme 3 focused on hindsight, as<br />

many women expressed that if they had more knowledge regarding fibroids,<br />

they would have made different decisions regarding their treatment. A fourth<br />

and final theme highlighted what subjects felt that women with fibroids need<br />

to know about the tumors, and centered on the promotion of self-education<br />

and the importance of finding a trustworthy physician.<br />

CONCLUSIONS: A diverse cohort of well educated, health literate<br />

women with a history of symptomatic uterine fibroids felt that there is not<br />

an existing source of high-quality patient-centered information available.<br />

Despite utilizing multiple resources and venues to educate themselves,<br />

women felt that more needed to be done to educate patients and ensure<br />

that they had access to unbiased counseling and trustworthy providers.<br />

Increasing women’s knowledge on fibroids empowers them in the face of<br />

their diagnosis, allowing them to make informed decisions regarding their<br />

treatment options without regret.<br />

Supported by: NIH WRHR Program K12HD0501<strong>21</strong>; RWJ Foundation;<br />

NMH; Evergreen Foundation (EEM).<br />

P-144 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

TISSUE-SPECIFIC EXPRESSION OF ESTROGEN RECEPTOR 1 IS<br />

REGULATED BY DNA METHYLATION IN A T-DMR (TISSUE-<br />

DEPENDENT AND DIFFERENTIALLY METHYLATED<br />

REGION). H. Asada, M. Okada, H. Tamura, N. Sugino. Obstetrics<br />

and Gynecology, Yamaguchi University School of Medicine, Ube, Japan.<br />

OBJECTIVE: The mechanism controlling tissue-specific expression of<br />

ESR1 is still unclear. It has been reported that DNA methylation of a specific<br />

region of the gene has an important role in determining tissue- and cell-specific<br />

gene expression. This region is called the T-DMR (tissue-dependent and<br />

differentially methylated region). We previously found a possible link between<br />

the mRNA expression of ESR1 and the DNA methylation status of<br />

ESR1 promoter, suggesting that the region includes the T-DMR for tissuespecific<br />

ESR1 expression. It is known that ESR1 has several transcription<br />

starts sites (TSS) and corresponding upstream exons (upstream Exon-A to<br />

-E1). The transcription of ESR1 starts from any of these upstream exons,<br />

and the upstream exons are used in a tissue-dependent manner. Three upstream<br />

exons, upstream Exon-A, -B and -C, are often used in the tissues<br />

with high ESR1 expression. In the present study, we investigated whether human<br />

ESR1 has a T-DMR and whether DNA methylation of the T-DMR regulates<br />

its expression. We also investigated whether T-DMR is present in each<br />

upstream exon of ESR1.<br />

DESIGN: basic research<br />

MATERIALS AND METHODS: We obtained informed consent from patients<br />

and approval by Yamaguchi University. DNA methylation profiles and<br />

mRNA expression profiles of ESR1 were analyzed in the endometrium,<br />

mammary gland, placenta, skin, and breast cancer tissues by sodium bisulfite<br />

sequencing. T-DMR-methylated reporter assay was performed to examine<br />

whether DNA methylation at the T-DMR actually suppresses transcription<br />

of ESR1.<br />

RESULTS: ESR1 expression was tissue-specific, being high in the endometrium<br />

and mammary gland and low/nil in the placenta and skin. In all<br />

of the tissues, the proximal promoter regions were unmethylated. On the<br />

other hand, the distal regions were unmethylated in the endometrium and<br />

mammary gland, but were hypermethylated in the placenta and skin,<br />

FERTILITY & STERILITY Ò<br />

e155


indicating this region is a T-DMR. T-DMR-methylated reporter assay revealed<br />

that DNA methylation of the T-DMR suppressed ESR1 transcription.<br />

DNA methylation analysis around upstream Exon-A, -B and -C showed each<br />

upstream exon has its own T-DMR and DNA methylation of the region is<br />

associated with transcriptional regulation in each upstream exon. In some<br />

breast cancer cases, T-DMRs regulate ESR1 transcription via DNA methylation<br />

in a manner similar to normal tissues, while in other cases, ESR1 transcriptional<br />

regulation deviates from the regulation seen in normal tissues.<br />

CONCLUSIONS: This is the first report to demonstrate that ESR1 has T-<br />

DMRs, and that the T-DMRs regulate tissue-specific ESR1 expression via<br />

DNA methylation in normal tissues. We also found that each upstream<br />

exon has a corresponding T-DMR, of which DNA methylation status is<br />

involved in regulating transcription of the upstream exon. Furthermore, our<br />

results show some breast cancer cases deviate from the normal regulatory<br />

mechanism of the transcription regulation of ESR1.<br />

P-145 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE ROLE OF TYPE 2 DIABETES IN MODIFYING THE RISK FOR<br />

FIBROIDS. D. R. Velez Edwards, a M. Wellons, b K. Hartmann, a<br />

T. L. Edwards. c a Obstetrics and Gynecology, Vanderbilt University, Nashville,<br />

TN; b Division of Diabetes, Endocrinology, and Metabolism, Department<br />

of Medicine, Vanderbilt University, Nashville, TN;<br />

c Medicine,<br />

Vanderbilt University, Nashville, TN.<br />

OBJECTIVE: Uterine fibroids (UF) affect 77% of women by menopause,<br />

and account for $9.4 billion in annual healthcare costs. Type-2-diabetes<br />

(T2D) diagnosis has inconsistently associated with UF risk in prior studies.<br />

Differences in results across studies may be due to differences in T2D<br />

severity and treatment. To further evaluate the relationship between T2D<br />

and UFs we tested for association between T2D and UF risk before and after<br />

adjustment for T2D severity and stratifying by T2D treatments.<br />

DESIGN: This nested case-control study of a clinical cohort limited to<br />

women a UF diagnosis after T2D diagnosis and who had a pelvic imaging<br />

both prior to T2D diagnosis and after T2D diagnosis.<br />

MATERIALS AND METHODS: UF cases were those diagnosed with a<br />

UF using pelvic imaging after T2D diagnosis. T2D outcome and treatments<br />

were assessed using inpatient records and natural language processing of<br />

clinical notes implemented in MedEx software. Logistic regression, adjusted<br />

for covariates, was used to test for association between T2D outcome and UF<br />

risk. Secondary analyses were also performed adjusting models for HgBA1C<br />

levels and evaluating diabetics on specific treatments (metformin, thiazolidinedione,<br />

or insulin) for association with UF risk.<br />

RESULTS: We identified 3,789 subjects with T2D outcome, 714 were diabetics<br />

and 3,075 were T2D controls. Among these 16% (N¼608) had UFs.<br />

In adjusted models we observed that T2D decreased risk for UFs (adjusted<br />

odds ratio [OR] 0.61, 95% confidence interval [CI] 0.47 to 0.80). However,<br />

when models were further adjusted for HgbA1C effects were inversed<br />

(adjusted OR¼1.55, 95% CI 0.96 to 2.49), suggesting the models were<br />

confounded by T2D severity. Additional analyses stratifying by diabetics<br />

with HgbA1C


increasingly in cancer genomics research. We use the NGS to investigate the<br />

process of malignant transformation of endometriosis.<br />

DESIGN: We collected four parts of pathological specimen, consisting of<br />

eutopic endometrium, non-atypical endometriosis tissue, atypical endometriosis<br />

and ovarian carcinoma tissue from EAOC patients. NGS was applied to<br />

investigate the gene mutations during the transformation process.<br />

MATERIALS AND METHODS: In this project, we use the Ion AmpliseqTM<br />

Comprehensive Cancer Panel (CCP) which targets the exons of<br />

409 tumor suppressor genes and oncogenes in order to identify pathogenic<br />

mutations associated with EAOC. Ten cases of EAOC were enrolled in this<br />

study. For each case, macrodissection was performed to separate different<br />

four types of cells from formalin-fixed paraffin-embedded (FFPE) sections.<br />

RESULTS: The DNA of endometriosis patients was extracted from<br />

formalin-fixed paraffin embedded (FFPE) tissue. Samples include four transformation<br />

processes that are normal eutopic endometrium, nonatypical endometriosis,<br />

atypical endometriosis and carcinoma tissue from the same<br />

patients of endometriosis associated with ovarian cancer. The NGS direct<br />

sequencing found that coding sequence of exon 1 in ARID1A gene and exons<br />

20 in PIK3CA gene were screened for mutations. Our results showed that<br />

ARID1A and PIK3CA mutation (c.6488delG and c.3140A>G) were identified<br />

in the carcinoma tissue, but not in endometrium, non-atypical endometriosis<br />

and atypical endometriosis.<br />

CONCLUSIONS: Our study aims to develop and validate NGS platform<br />

for identifying the critical factor in the early event of the malignant transformation<br />

process of endometriosis. ARID1A and PIK3CA mutations<br />

contribute to the transformation process in our study. We believe this study<br />

will shed new light on fundamental aspects in the understanding the molecular<br />

pathogenesis of malignant transformation of ovarian endometriosis.<br />

Supported by: This work is Supported by MST Taiwan [grant number 102-<br />

2628-B-037-011-MY3 and 102-2632-B-037-001-MY3].<br />

P-148 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

COMPARATIVE ANALYSIS OF MEMBRANE MICROPARTICLES<br />

(MP) EXPRESSION IN WOMEN WITH EPITHELIAL OVARIAN<br />

CANCER (EOC) AND ENDOMETRIOMAS. J. A. Falcao, Jr., a<br />

F. F. Nunes, b J. Marinuzzi, c O. A. Martins-Filho, b A. T. Carvalho, b<br />

R. M. Lamaita, a M. M. Carneiro, a A. L. Silva-filho. a a Obstetrics and Gynecology,<br />

Federal University of Minas Gerais, Belo Horizonte, Brazil; b Centro<br />

de Pesquisas Rene Rachou - Fiocruz, Belo Horizonte, Brazil; c UNESP Botucatu,<br />

Botucatu, Brazil.<br />

OBJECTIVE: Microparticles (MP) are small vesicles derived from cell<br />

membranes which have been recognized as important mediators of cell activity<br />

and may associated with pathological and physiological processes such as<br />

immune response, cell differentiation, vascular disorders and cancer. The aim<br />

of this study is to compare circulating MP according to their specific cellular<br />

origin in women with endometrioma, EOC and controls.<br />

DESIGN: Prospective evaluation of 60 women from March 2010 to<br />

<strong>October</strong> 2013 divided into three groups: EOC (n¼26); endometriomas<br />

(n¼18); control (n¼16; women operated for benign gynecologic disease<br />

with normal ovaries).<br />

MATERIALS AND METHODS: A convenience sample was used due to<br />

the low prevalence of EOC and the strict inclusion and exclusion criteria.<br />

Surgical staging was performed according to the FIGO classification or<br />

ASRM endometriosis staging as appropriate. Exclusion criteria were: previous<br />

chemotherapy and/or radiotherapy; immune system diseases diagnosis<br />

and / or use of immunosuppressive drugs within the past 6 months, acute infections,<br />

identification of distinct malignancy from EOC in the histopathological<br />

exam. Serum levels of CD66 +/neutrophils, CD45 +/leukocytes,<br />

CD14 +/monocytes, CD235 +/erythrocytes, CD51 +/endothelium, CD41<br />

+/platelets, CD3 +/lymphocytes were performed by flow cytometry. The differences<br />

between groups were evaluated by Mann-Whitney or Kruskal-<br />

Wallis tests. P < 0,05 was considered statistically significant.<br />

RESULTS: Mean patient age was: EOC (62 14.06), endometrioma<br />

(3710.31) and control (408,8) years. Ten cases were identified as FIGO<br />

stage I and II, and 16 as III/IV. Fifteen women were classified as ASRM stage<br />

III and 3 as stage IV. Women with endometrioma were associated with higher<br />

circulating levels of CD45 +/leukocytes (p¼ 0.0292), CD14 +/monocytes<br />

(p¼0.012), CD235 +/erythrocytes (p¼0.0341), CD51 +/endothelium<br />

(p¼0.0228) and CD41 +/platelets (p¼0.0464) compared to patients with<br />

EOC and control group. No association was found between age of the patients<br />

of the 3 groups and the MP assay neither with maximum diameter of<br />

the lesion in the endometrioma group.<br />

CONCLUSIONS: Our data shows for the first time that women with endometrioma<br />

present with higher MP circulating levels compared EOC patients.<br />

MP have potential use in understanding the cellular microenvironments associated<br />

with these diseases and may contribute to the improvement in<br />

screening, diagnosis and therapeutic strategies.<br />

P-149 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE LONG TERM RECURRENCE RATE AFTER CONSERVATIVE<br />

SURGICAL TREATMENT OF ENDOMETRIOSIS IN<br />

ADOLESCENTS. Y. Cho, a S. Lee, b M. Kim, c J. Bae, d M. Han, d<br />

J. Park. d a Department of Obstetrics and Gynecology, Dong-A University<br />

Medical Center, Dong-A University, College of Medicine, Busan, Korea, Republic<br />

of; b Dankook University, School of Medicine, Cheil Gene, Seoul, Korea,<br />

Republic of; c Department of Obstetrics and Gynecology, CHA Gangnam<br />

Medical Center, CHA University, Seoul, Korea, Republic of; d Dong-A University<br />

Hospital, Busan, Korea, Republic of.<br />

OBJECTIVE: Endometriosis, while generally considered as a disease that<br />

affects adult women, has become increasingly recognized as a chronic illness<br />

that can begin during adolescent and young adulthood. This patient group<br />

presents particular challenges in terms of differential diagnosis, variable presentation<br />

and symptoms, and choice of treatment. There is very limited<br />

research in adolescents with endometriosis and long term studies about the<br />

recurrence or progression are not well understood. We aimed to evaluate<br />

the long term recurrence rate of ovarian endometriomas in adolescents<br />

following the first conservative surgical treatments.<br />

DESIGN: Multicenter retrospective cohort study<br />

MATERIALS AND METHODS: Patients % 20 years of age who were<br />

surgically treated with laparoscopic enucleation of ovarian endometrioma<br />

were selected. We included patients only who were followed up more than<br />

36 months. We excluded patients who had reproductive tract anomalies,<br />

those who underwent non-conservative procedures, such as oophorectomy,<br />

those who underwent cyst aspiration. Recurrence of the endometrioma was<br />

established on the basis of transvaginal or transrectal sonography documenting<br />

the presence of a cystic mass with a diameter of R 20 mm. Baseline surgical<br />

characteristics were analyzed.<br />

RESULTS: We recruited 51 adolescent patients who were followed up<br />

more than 36 months. The mean age of patients was 19.0 1.1 years<br />

(range, 16-20 years). According to our definition of recurrence, 15 patients<br />

(29.4%) experienced recurrence of ovarian endometrioma after first<br />

laparoscopic cyst enucleation. The overall cumulative recurrence rates of<br />

ovarian endometrioma per patient at 24, 36, 60, and 96 months after<br />

first-line surgery as 8.2%, 10.2%, 20.8%, and 37.4%, respectively.<br />

Surgical characteristics, such as the diameter of the cyst, disease stage,<br />

unilateral or bilateral involvement, and co-existence of deep endometriosis,<br />

and postoperative medical therapy were not associated with recurrence in adolescents.<br />

CONCLUSIONS: The long-term recurrence rate in adolescent after the<br />

conservative surgery was dependent on the months that had elapsed since<br />

treatment, and this value increased over time. Long-term and continuous<br />

follow up is needed who have undergone surgical treatment for endometriosis<br />

in adolescent period.<br />

P-150 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

EXPRESSION OF HOXB4 IN ENDOMETRIAL TISSUES FROM<br />

WOMEN WITH OR WITHOUT ENDOMETRIO<br />

SIS. G. M. Alkusayer, a,b B. Peng, c C. Klausen, d S. Lisonkova, d<br />

M. Kinloch, e P. Yong, f M. A. Bedaiwy. g a Department of Clinical Sciences,<br />

College of Medicine, Princess Nourah Bint Abdulrahman University, Riyadh,<br />

Saudi Arabia; b Department of Obstetrics and Gynaecology, University of British<br />

Columbia, Vancouver, BC, Canada; c University of British Columbia, CFRI,<br />

Vancouver, BC, Canada; d University of British Columbia, Vancouver, BC, Canada;<br />

e Vancouver General Hospital, Vancouver, BC, Canada; f Assistant Professor,<br />

Vancouver, BC, Canada; g BC Women’s Hospital, Vancouver, BC, Canada.<br />

OBJECTIVE: HOX genes play important roles in the functional differentiation<br />

of adult tissues by regulating proliferation, angiogenesis, adhesion<br />

and motility. The aim of our study was to examine the expression and localization<br />

of HOXB4 in normal human endometrial tissues as well as eutopic<br />

endometrium and ectopic implants from women with endometriosis<br />

throughout the menstrual cycle.<br />

DESIGN: Case-control study.<br />

FERTILITY & STERILITY Ò<br />

e157


MATERIALS AND METHODS: Normal endometrial tissues were<br />

collected from 20 pre-menopausal women with no history of endometriosis.<br />

Additionally, matched eutopic endometrium and ectopic endometrial implants<br />

were obtained from 40 women with endometriosis and no recent history<br />

of hormonal therapy [deep infiltrating endometriosis (DIE), n¼20;<br />

endometrioma, n¼20]. Using the Histoscore method, a pilot immunohistochemical<br />

analysis of HOXB4 was performed on 9 normal endometrial tissues<br />

and 12 eutopic/ectopic endometrial tissue pairs. Kruskal-Wallis non-parametric<br />

test was used to compare multiple categories, Wilcoxon test was<br />

used for pair-wise comparisons.<br />

RESULTS: HOXB4 was immunolocalized in the nuclei of glandular<br />

epithelial cells in normal endometrium and eutopic/ectopic endometrial tissues<br />

from women with endometriosis. HOXB4 immunoreactivity was absent<br />

from endometrial stromal cells in all tissues examined. No significant difference<br />

in HOXB4 immunostaining intensity was found between controls and<br />

eutopic endometrium from endometriosis. Interestingly, HOXB4 immunoreactivity<br />

was significantly reduced in ectopic DIE compared to eutopic endometrium<br />

[P¼ 0.041] and endometrioma [P¼ 0.05] in women with<br />

endometriosis. Moreover, HOXB4 immunoreactivity in eutopic endometrium<br />

of disease and control groups combined was significantly higher in<br />

the proliferative phase than in the secretory phase [P¼ 0.026]. Conversely,<br />

HOXB4 immunoreactivity in ectopic implants was lower in proliferative<br />

phase than in the secretory phase [P¼ 0.0<strong>21</strong>].<br />

CONCLUSIONS: During the menstrual cycle the expression of HOXB4 in<br />

endometrial glandular epithelial cells is higher in the proliferative phase than<br />

the secretory phase. On the whole, HOXB4 expression does not appear to<br />

differ between normal endometrium and eutopic endometrium from endometriosis.<br />

However, its expression is reduced in DIE, but not endometrioma, and<br />

may be dysregulated in ectopic implants depending on the phase of the menstrual<br />

cycle.<br />

Supported by: Internal fund, Department of Obstetrics and Gynecology,<br />

University of British Columbia, Vancouver, BC, Canada. Graduate Sponsorship<br />

Program,Princess Nourah bint Abdulrahman University,Riyadh,<br />

Kingdom of Saudi Arabia.<br />

P-151 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

COULD THE UNILATERAL OVARIAN ENDOMETRIOSIS<br />

AFFECT THE CONTRALATERAL OVARY? NEW INSIGHTS<br />

FROM FOURIER TRANSFORM INFRARED (FTIR)<br />

SPECTROSCOPY. G. Gioacchini, a E. Sereni, b C. Zaca’, b<br />

E. Giorgini, c V. Notarstefano, a L. Vaccari, d O. Carnevali, a A. Borini. b<br />

a Universita Politecnica delle Marche, Ancona, Italy; b Tecnobios Procreazione,<br />

Bologna, Italy; c Polytechnic University of Marche, Ancona, Italy; d E-<br />

lettra-Sincrotrone Trieste, Trieste, Italy.<br />

OBJECTIVE: Up to date, endometriosis’ possible impact on follicle cells<br />

metabolism has never been highlighted. In this light, the aim of this study was<br />

to evaluate, by FT-IR microspectroscopy, metabolic changes on granulosa<br />

cells (GCs) isolated from endometriosis affected ovaries and contralateral<br />

healthy ones<br />

DESIGN: This prospective non-randomized study has been conducted<br />

from January to September 2014 on women undergoing a controlled ovarian<br />

hyperstimulation for an IVF treatment. In particular, GCs were collected<br />

from both ovaries of 10 women with a diagnosis of unilateral ovarian endometrioma<br />

at the time of oocytes retrieval. 9 women with male, idiopathic or<br />

tubal-factor infertility diagnosis were selected for the control group. The<br />

three experimental groups matched for female age (36.24.1 vs 35.42.6)<br />

MATERIALS AND METHODS: GCs obtained from follicles aspirates<br />

were isolated from red blood cells and follicular fluid by density gradient<br />

centrifugation. FTIR analysis was performed by using a Bruker Vertex 70<br />

Interferometer with a Hyperion 3000 Vis-IR microscope. QPCR analysis<br />

Supported FTIR results. . Data are presented as mean S.D. Two-Way AN-<br />

OVA followed by Tukey test as Multiple comparisons test, was used for comparison<br />

among experimental groups. All statistical analyses were performed<br />

using the statistical software package Prism5 (Graphpad Software, Inc. USA)<br />

with significance accepted at P


Cornell University, New York, NY; c The Ronald O. Perelman and Claudia<br />

Cohen Center, New York, NY.<br />

OBJECTIVE: Mechanism of infertility in endometriosis potentially involves<br />

oocyte/embryo defects and lower fertilization and implantation rates.<br />

Cumulus complex in endometriosis patients demonstrates mitochondrial<br />

dysfunction and higher oxidative stress. We sought to investigate if ICSI,<br />

which diminishes oxidative stress when compared to insemination, improves<br />

IVF outcomes in patients with endometriosis.<br />

DESIGN: Retrospective cohort<br />

MATERIALS AND METHODS: 2<strong>21</strong> patients who had surgically or sonographically<br />

confirmed endometriosis were compared with 150 patients with<br />

unexplained infertility. Patients with endometriosis as a sole cause of infertility<br />

were included. Exclusion criteria [1. > 41yo; 2. Male factor (sperm<br />

count < 15 x 106/cc, motility less than 40%, older than 55yo]. Primary outcomes:<br />

1. Number of retrieved eggs; 2. Number of 2 pronuclear zygote (2PN)<br />

stage embryos; 3. Proportion of day 5 transfers; 4. Number of frozen blastocysts.<br />

Secondary outcomes: 1. Clinical intrauterine pregnancy [IUP] rate; 2.<br />

Live birth rate. Student’s t-test and ANOVA were used to continuous and<br />

Fisher’s exact and Chi-square tests for categorical variables.<br />

RESULTS: Within the endometriosis group, 124 patients utilized standard<br />

insemination, whereas 97 utilized ICSI. Endometriosis-insemination, endometriosis-ICSI<br />

and the control groups were matched for age, BMI, anti-mullerian<br />

hormone (AMH) and the number of eggs retrieved [11.95.3 vs. 13.13.1 vs.<br />

13.1 5.9, P¼0.07]. The control and endometriosis-ICSI groups did not differ<br />

in terms of number of 2PNs, number of day 5 transfers, IUP and live birth rates<br />

(Table). Endometriosis-ICSI group demonstrated significantly higher percentage<br />

of 2PN embryos, percentage of day 5 transfers, number/proportion of clinical<br />

IUP as well as live birth rates when compared with the endometriosisinsemination<br />

group. Additionally, endometriosis-ICSI group had a lower proportion<br />

of cycles with no normal fertilization [0 (0%) vs 4 (3.2%), but this difference<br />

did not reach statistical significance.<br />

CONCLUSIONS: Our findings suggest that endometriosis might be an<br />

indication for ICSI, as patients with endometriosis who utilized ICSI<br />

achieved better outcomes (higher proportion of normal fertilization, higher<br />

number of day 5 transfers, clinical pregnancy and live birth rates) when<br />

compared to standard insemination cycles.<br />

Table 1<br />

Endometriosis- Endometriosis-<br />

Insemination(124) ICSI(97) P 1 Controls(150) P 2 P 3<br />

Mean # eggs retrieved 11.95.3 13.13.1 0.5 13.15.9 0.2 0.8<br />

Total# of retrieved eggs(%) 1301 1269 1954<br />

Mean # of 2PN 6.93.9 8.44.9 0.01 8.85.1


Obstetrics and Gynecology, West China Second University Hospital of Sichuan<br />

University, Chengdu, Sichuan, China.<br />

OBJECTIVE: Endometriosis-related infertility is common in women of<br />

reproductive age. Progesterone resistance, especially aberrantly expressed<br />

progesterone receptor in the eutopic endometrium, is considered as a key<br />

causal factor. Our objective is to explore microRNA-mediated mechanism<br />

controlling aberrant progesterone receptor expression in infertile women<br />

with minimal or mild endometriosis.<br />

DESIGN: microRNA Array, combined with Human Gene Expression microarrays,<br />

was used to screen eutopic endometrium of infertile women with<br />

minimal or mild endometriosis. Bioinformatics analysis predicted that microRNA-196a<br />

target the PGR 3 0 UTR. We studied the relationship between<br />

microRNA-196a level and PGR expression in endometrial stromal cells<br />

(ESCs). According to mRNAs microarray result, MAPK pathway is activated<br />

in eutopic mid-secretory endometrium. The role of MAPK pathway involved<br />

regulation of miR-196a on PGR was also investigated.<br />

MATERIALS AND METHODS: The study was conducted at West China<br />

Second University Hospital of Sichuan University. 22 infertile women with r-<br />

AFSI-IIendometriosis and 20 disease-free control subjects were enrolled.<br />

MiRCURY LNATM microRNA Array(v.18.0)(Exiqon), Human 444K<br />

Gene Expression Microarrays v2(Agilent), qRT-PCR, cell culture, transfections,<br />

luciferase reporter assays, and western bolt were used in this study.<br />

RESULTS: 66 differently expressed microRNAs (fold changeR2.00 and<br />

p12wks, FSH > 12). The number of oocytes retrieved, fertilization<br />

rate, clinical pregnancy rate and miscarriage rate were compared. Statistical<br />

comparisons were done with Stata10.0 software.<br />

RESULTS: The clinical pregnancy rate was 114/355(32.1%) in women<br />

with only endometriosis, <strong>17</strong>/88(19.3%) in endometriosis with adenomyosis<br />

and 12/64(20.3%) in only adenomyosis. The odds ratio for clinical pregnancy<br />

showed that the adenomyosis groups (Group B & C) had significantly lower<br />

pregnancy rate (Avs B; OR: 0.51, 95% CI : 0.28-0.89, p< 0.02; A vs. C; OR:<br />

0.49, 95% CI:0.25-0.94, p< 0.03).Miscarriage was observed in 26/355 in<br />

Group A,13/88 in Group B and 10/64 in Group C. Miscarriage rate was<br />

significantly higher in pregnancy associated with adenomyosis (Avs.B;<br />

OR: 0.45, 95% CI : 0.22-0.92, p< 0.03;A vs.C; OR: 0.42, 95% CI:0.19-<br />

0.93, P< 0.03).Number of oocytes retrieved and fertilization rates were comparable<br />

in all the three groups.<br />

CONCLUSIONS: Presence of adenomyosis appeared to have adverse effects<br />

on IVF/ICSI outcomes in terms of lower pregnancy rate and higher<br />

miscarriage rate in women pretreated with long-term GnRH-agonist.<br />

Screening for adenomyosis before going for IVF/ICSI may be considered,<br />

so that counseling can be done regarding decreased probability of viable<br />

pregnancy in women having adenomyosis.<br />

P-158 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ANGIOGENIC ACTIVITY OF PGF2A IN WOMEN WITH<br />

ENDOMETRIOSIS. H. Rakhila, a M. Bergeron, b M. Daris, b<br />

M. Leboeuf, b M. Lemyre, b C. Rheaume, b M. Pouliot. a a Department of<br />

Microbiology and Immunology, Centre Hospitalier Universitaire de Quebec,<br />

Faculty of Medicine, Universite Laval, Quebec, QC, Canada; b Department of<br />

Obstetrics and Gynecology, Centre Hospitalier Universitaire de Quebec, Faculty<br />

of Medicine, Universite Laval, Quebec, QC, Canada.<br />

OBJECTIVE: To investigate interleukin-8 (CXCL8) and vascular endothelial<br />

growth factor (VEGF) expressions (potent angiogenic factors) in<br />

endometrial stromal cells of women with endometriosis compared with controls<br />

in response to Prostaglandin F2a (PGF2a), an important regulators of<br />

female reproductive function.<br />

DESIGN: Primary cultures of eutopic and ectopic endometrial stromal<br />

cells isolated from endometrial biopsies of endometriosis patients (n ¼ 9),<br />

as well as eutopic endometrium of control patients (n¼5) in proliferative<br />

and secretory cycle phases exposed to PGF2a, with assessment of CXCL8<br />

and VEGF expression.<br />

MATERIALS AND METHODS: Cells were cultured in DMEM-F12 medium<br />

with 10% FBS, 1% antibiotics, 5 mg/mL insulin and 5mg/mL transferrin.<br />

Cells grown to confluence were preincubated for 1 h with PGF2a-<br />

FP receptor antagonist (AL8810) (50ug/mL) or cyclo-oxygenase-2 (cox-2)<br />

inhibitor (NS398) (50 ug/mL) and stimulated with different concentrations<br />

of PGF2a (0-1000ng/mL) or PGF2a analog (Fluprostenol) (0-1000 ng/mL)<br />

for different periods of time (0-48 h). PGF2a biosynthetic pathways were<br />

analyzed by RT-qPCR and Western blot. While, CXCL8 and VEGF secretions<br />

were analyzed by ELISAs. An unpaired t test and one-way ANOVA followed<br />

by the Dunnett’s test were performed for this study.<br />

RESULTS: Our study showed that PGF2a stimulated CXCL8 and VEGF<br />

expressions in ectopic endometrial stromal cells of women with endometriosis<br />

in the proliferative and the secretory phases of the menstrual cycle,<br />

without noticeable change in eutopic endometrial stromal cells of women<br />

with or without endometriosis. Cell exposure to PGF2a analog (Fluprostenol)<br />

confirmed these previous observations. After 24h of stimulation, diminutions<br />

of CXCL8 and VEGF were significant when PGF2a receptor (FP) antagonist<br />

was added with either PGF2a or Fluprostenol. Meanwhile, we observed that<br />

PGF2a increased significantly cox-2 expression while its other biosynthetic<br />

enzymes showed no statistical difference. We, therefore, used cox-2 inhibitor<br />

(NS398) which correlated a statistical decrease for both VEGF and IL-8 secretions.<br />

CONCLUSIONS: This study showed that angiogenic pathways are<br />

involved in PGF2a-mediated activation through FP and Cox-2 in ectopic<br />

endometrial stromal cells of women with endometriosis. This is in keeping<br />

with our previous data showing multiple abnormalities in PGF2a pathways<br />

in ectopic endometrial tissue, leading to an abnormal behavior. Expression<br />

of CXCL8 and VEGF in ectopic endometrial cells throughout the menstrual<br />

cycle of women with endometriosis may, considering the role of these cytokines<br />

in cell growth and angiogenesis, play an important role in the capability<br />

of endometrial cells to develop and to survive in ectopic locations.<br />

Supported by: grant MOP-123259 to MP from the Canadian Institutes for<br />

Health Research.<br />

e160 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


P-159 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ENDOCERVIX MICRORNA EXPRESSION IN PATIENTS WITH<br />

DEEP ENDOMETRIOSIS. C. V. de Carvalho, P. Rozenchan,<br />

T. Bonetti, A. Kopelman, M. B. Girao, E. Schor. Gynecology, Universidade<br />

Federal de S~ao Paulo - UNIFESP/EPM, S~ao Paulo, Brazil.<br />

OBJECTIVE: The identification of molecular differences in eutopic endometrium<br />

of women with endometriosis is an important step toward understanding<br />

it pathogenesis and developing novel strategies for the treatment<br />

of associated infertility and pain. A number of differences in gene expression<br />

were found in patients with endometriosis. miRNA are small molecules and<br />

may act as post-transcriptional gene expression regulator. Many miRNAs are<br />

tissue specific and play a significant role in oncogenes, metastasis and tumor<br />

invasion. It has also been found that most miRNAs have dual nature, targeting<br />

both oncogenes and tumor suppressor genes. Endocervix could represent a<br />

less-invasive approach to detect patients with endometriosis, so the aim of<br />

this study was to perform miRNA expression screening in endocervix of patients<br />

with deep endometriosis compared to healthy controls.<br />

DESIGN: Experimental transversal study<br />

MATERIALS AND METHODS: It were included six patients with deep<br />

endometriosis and six control patients without endometriosis or other previous<br />

gynecological illness. Patients underwent an endocervical brush during<br />

the proliferative phase of menstrual cycle. The samples were submitted to<br />

RNA purification and were analyzed by miRNA PCR-array (miScript<br />

miRNA PCR Array, Qiagen), containing 86 different miRNAs.<br />

RESULTS: We were able to found 7 miRNAs significantly upregulated:<br />

Has-miR<strong>21</strong>0-3p (fold: 3.4, p: 0.04), Has-miR23b-3p (fold: 3.5, p: 0.01),<br />

Has-miR151a-5p (fold: 8.2, p: 0.01), Has-miR222-3p (fold: 2.6, p: 0.04)<br />

Has-let7e-5p (fold: 3.4, p 0.05), Has-let7b-5p (fold: 4.6, p: 0.02), Haslet7c-5p<br />

(fold: 3,0, p: 0.04) and one miRNA downregulated in endometrium:<br />

Has-miR142-3p (fold: -8.8, p: 0.004), in patients with deep endometriosis<br />

when compared to control group.<br />

CONCLUSIONS: We found eight microRNAs differentially expressed in<br />

endometrium of patients with deep endometriosis when compared to control<br />

group, highlightening those miRNAs as biomarkers candidates for endometriosis.<br />

This study is a screening of miRNA expression in endocervix of patients<br />

with deep endometriosis and our next step is the validation of these<br />

miRNAs in a different set of patients.<br />

References: Pelvic Pain Unit - UNIFESP-EPM<br />

Supported by: Fundaç~ao de Amparo a Pesquisa do Estado de S~ao Paulo<br />

(FAPESP) Coordenaç~ao de Aperfeiçoamento de Pessoal de Nıvel Superior<br />

(CAPES)<br />

P-160 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

EFFECT OF LUPRON VS NORETHINDRONE TREATMENT ON<br />

LIPID PROFILE OF WOMEN WITH SYMPTOMATIC<br />

ENDOMETRIOSIS. C. Charles, a O. Muneyyirci-Delale, a,b N. Sinaii, c<br />

M. Dalloul, a P. Stratton. d a OB/GYN, SUNY Downstate Medical Center,<br />

Brooklyn, NY; b Ob/Gyn, Kings County Hospital Center, Brooklyn, NY; c National<br />

Institutes of Health, Bethesda, MD; d NICHD, NIH, Bethesda, MD.<br />

OBJECTIVE: To assess changes in the lipid profiles of women with symptomatic<br />

endometriosis who underwent treatment with Leuprolide Acetate<br />

Depot form (LD) vs Norethindrone Acetate (NA).<br />

DESIGN: Prospective, randomized, double-masked clinical trial.<br />

MATERIALS AND METHODS: 62 women with endometriosis-associated<br />

pain were randomized to receive LD 11.25mg every 3 months or NA<br />

5mg daily for 24 weeks in Phase I, then all were given NA for another 28<br />

weeks in Phase II. After 52 weeks of treatment, women were followed for<br />

up to one year (PTFU). Lipid testing included Total Cholesterol (CHOL), Triglycerides<br />

(TRIG), High Density Lipoprotein (HDL) and Low Density Lipoprotein<br />

(LDL). All parameters were assessed at entry, end of Phases I and II<br />

as well as during PTFU. Data were analyzed using paired or two-sample t-<br />

tests, or Wilcoxon signed rank and Wilcoxon rank-sum tests. Regression<br />

modeling adjusted for continued NA use in PTFU.<br />

RESULTS: Women were predominantly Black (82%), between age <strong>21</strong> and<br />

47 years with 31 NA and 31 LD. A decrease in HDL and an increase in LDL<br />

values were noted during Phase I in NA (p


Medical School, Boston, MA; d Brigham and Women’s Hospital and Harvard<br />

Medical School, Boston, MA.<br />

OBJECTIVE: Develop an in vitro model of human disease for endometriosis<br />

DESIGN: Basic science<br />

MATERIALS AND METHODS: G4 mouse embryonic stem cells<br />

(mESCs) were cultured andallowed to differentiate for <strong>21</strong> days as colonies<br />

or as suspended embryoid bodies (EBs). Differentiation of G4 ESCs was<br />

induced by transferring the cells to mouse embryonic feeder cell culture media<br />

(Dulbecco’s modified Eagle’s Medium (DMEM) with 10% fetal bovine<br />

serum (FBS)) for 3 weeks. G4 ESCs were also differentiated with differentiation<br />

media (DMEM, FBS, Bone morphogenetic protein (BMP)-4 50ng/ul<br />

and Activin-A 20ng/ul). Post-differentiation analysis profiles of the mESCs<br />

and EBs were assessed for the endometrium and endometriosis markers<br />

CD9 / CD13 by immunocytochemistry and transcriptional analysis using<br />

reverse and quantitative PCR. Differentiated CD9 + CD13 + were then prepared<br />

for sorting using FACS for further characterization.<br />

RESULTS: Colonies of mESCs showed progressive differentiation of CD9<br />

immunoreactive cells over a 3 week period (26.6%, 48.9%,and 76.4%, weeks<br />

1-3, respectively), while populations of differentiated CD13 cells appeared<br />

relatively constant over this same duration (40.6%, 47.0% and 47.4%).<br />

40% of EBs displayed CD9 immunoreactivity with 61.2% also immunolableing<br />

for CD13. These observations were confirmed using quantitative PCR analyses.<br />

Additional markers of endometrium (CD146, PGDF-b) were also<br />

identified in these differentiating cultures.<br />

CONCLUSIONS: The results from this study demonstrate the generation<br />

of presumptive endometrial precursors in differentiating mouse ESCs. The<br />

ability to produce CD9 + CD13 + cells in vitro, isolate the same by FACS and<br />

continue their in vitro growth shows the utility of ESCs for preclinical<br />

modeling of a human disease, endometriosis. Furthermore this system affords<br />

the unique opportunity to investigate novel therapeutic approaches for the<br />

management of this debilitating disease using a stem cell disease platform.<br />

CONCLUSIONS: We observed that the levels of serum anti-endometrial<br />

antibodies to SLP2, TMOD3 and TPM3 did not appear to elevate further<br />

in advanced stages (Stage III-IV) of endometriosis.<br />

Supported by: Department of Biotechnology, Government of India (BT/<br />

PR4589/MED/30/731/2012) and ICMR (NIRRH/MS/RA/184/09-2014).<br />

P-164 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

WITHDRAWN<br />

P-163 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

STAGE-WISE COMPARISON OF ANTI-ENDOMETRIAL-ANTI-<br />

BODIES AGAINST PEPTIDES OF SLP2, TMOD3 AND TPM3 IN<br />

DIAGNOSIS OF ENDOMETRIOSIS. T. Bendigeri, a A. Ghuge, a<br />

K. Bhusane, a S. Begum, b N. Warty, c R. Sawant, d P. Dasmahapatra, e<br />

K. Padte, f A. C. Humane, g A. Chauhan, h R. Gajbhiye. a a Department of Clinical<br />

Research, National Institute for Research in Reproductive Health, Mumbai,<br />

India; b Department of Biostatistics, National Institute for Research in<br />

Reproductive Health, Mumbai, India; c Sanjivani Diagnostic Centre and General<br />

Maternity, Mumbai, India; d Sanjivani Diagnostic Centre and General<br />

Maternity Home, Mumbai, India; e Spectrum clinic & Endoscopy Research<br />

Institute, Kolkata, India; f Dr Kedar’s Maternity, Infertility and Surgical Hospital,<br />

Endoscopy and IVF Center, Panjim, India; g Goverment Medical College,<br />

Nagpur, India; h Department of Obstetrics and Gynecology, Seth GS<br />

Medical College and King Edward Memorial (KEM) Hospital, Mumbai, India.<br />

OBJECTIVE: The lack of non-invasive diagnostic test for early diagnosis<br />

of endometriosis results in 8-11 years delay in diagnosis leads to deterioration<br />

of quality of life. Previously, we identified anti-endometrial-antibodies<br />

against peptides of Stomatin like protein 2 (SLP2), Tropomodulin 3<br />

(TMOD3) and Tropomyosin 3 (TPM3), and further proposed their utility<br />

as non-invasive biomarkers for early diagnosis of endometriosis.The aim<br />

of the present study was to investigate whether the levels of biomarkers<br />

vary with disease progression and to compare the sensitivity, specificity<br />

and diagnostic accuracy of the biomarkers in diagnosis of endometriosis in<br />

early stages (Stage I-II) versus advanced stages (Stage III-IV).<br />

DESIGN: Multi-centre, cross sectional study<br />

MATERIALS AND METHODS: Women with endometriosis (Stage I-II<br />

n¼133, Stage III-IV n¼133)and healthy controls (n¼104) were screenedfor<br />

eleven novel autoimmune markers (anti-endometrial-antibodies of SLP2a,<br />

SLP2b, SLP2c, TMOD3a, TMOD3b, TMOD3c, TMOD3d, TPM3a,<br />

TPM3b, TPM3c and TPM3d) using the peptide ELISA.The statistical analysis<br />

was performed using STATA software (version 8.2, Texas, USA).<br />

RESULTS: The mean serum levels of anti-endometrial-antibodies of<br />

SLP2a, SLP2b, SLP2c, TMOD3a, TMOD3b, TMOD3c, TMOD3d,<br />

TPM3a, TPM3b, TPM3c and TPM3d) in early stages (Stage I-II) were significantly<br />

higher than that of advanced stages (Stage III-IV) of endometriosis<br />

(p


P-167 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

DOES THE RECURRENCE OF OVARIAN ENDOMETRIOMA<br />

AFFECT THE PREGNANCY RATES IN IVF?. K. Aslan, a<br />

I. Kasapoglu, b B. Avci, c B. Ata, d G. Uncu. e a Uludag University School<br />

of Medicine Depart, Gynecology, Bursa, Turkey; b Obstetrics and Gynecology,<br />

Uludag University School of Medicine, Bursa, Turkey; c Histology/<br />

Embryology, Uludag University Faculty of Medicine, Bursa, Turkey; d Reproductive<br />

Endocrinology and Infertility, _Istanbul, Turkey; e Uludag University,<br />

Bursa, Turkey.<br />

P-166 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

DOWN REGULATION OF BIOMARKERS AND INCREASED PREG-<br />

NANCY RATE FOLLOWING GNRH AGONIST TREATMENT IN<br />

PATIENTS WITH ENDOMETRIOSIS. E. Rahmawati, a<br />

P. K. Maurya, b A. Kao, c H. Chen, d C. Tzeng. e a Clinical Medicine, Taipei<br />

Medical University, Taipei, Taiwan; b Amity Institute of Biotechnology, Amity<br />

University, Uttar Pradesh, India; c Obstetrics and Gynecology, Taipei<br />

Medical University, Taipei, Taiwan; d Graduate Institute of Toxicology, College<br />

of Medicine, National Taiwan University, Taipei, Taiwan; e Taipei Medical<br />

University, Taipei, Taiwan.<br />

OBJECTIVE: The present study was undertaken to evaluate the effect of<br />

GnRHa in endometriosis patients undergoing IVF. We further investigated<br />

the effect of GnRHa on the genes expression and proteins level in endometriotic<br />

tissues. We hypothesized that some protein related with pathogenesis of<br />

endometriosis are down regulated in the patients following GnRHa treatment.<br />

DESIGN: Retrospective case control clinical study combined with analysis<br />

of gene expression profile and the protein level in patients with endometriosis<br />

after GnRHa treatment.<br />

MATERIALS AND METHODS: This study recruited 282 endometriosis<br />

women without GnRHa treatment and 96 endometriosis patients with<br />

GnRHa treatment. IVF outcome was measured by evaluation on oocyte number,<br />

embryo transfer, pregnancy rate, implantation rate and abortion rate. Biopsy<br />

of endometriotic tissues was done to evaluate distinction of genes<br />

expression with and without GnRHa treatment using affymetrix microarray<br />

and independently validated by q-RT PCR, western blot and IHC. The candidate<br />

biomarkers from serum were measured by ELISA. Primary culture of<br />

endometriotic stroma cells to identify the role of estrogen in the expression<br />

of candidate proteins. Statistical analysis was performed using Mann-Whitney<br />

U test. P < 0.05 is considered significant.<br />

RESULTS: IVF outcome shown that implantation rate increased significantly<br />

in endometriosis patients with GnRHa treatment (25.4%) compared<br />

with untreated group (18.7%) (p¼0.001) and the pregnancy rate increased<br />

significantly in treated patient (39.3%) compared with untreated group<br />

(19.7%) (p¼0.003). This data was Supported by microarray data that approximately<br />

65 genes exhibited alterations in expression following GnRHa treatment.<br />

Validation using q-RT PCR, WB, IHC and ELISA indicated that TNS1,<br />

MMP 14, CAV 2, NRN1 and ATP2A3 were significantly lower in patients<br />

with GnRHa treatment. P values are in the following manner: Tensin 1<br />

(p¼ 0.0100), MMP 14 (p¼ 0.013), Caveolin 2 (p¼ 0.023), Neuritin 1 (p¼<br />

0.040) and ATP2A3 (p¼ 0.044). Immunohostochemistry indicated that stromal<br />

cells expressed the five candidates proteins. Treatment with estradiol on<br />

stroma endometriotic cells increased the expression of those proteins. Tensin<br />

1 and MMP 14 play important role in cell migration. The role of Caveolin 2 is<br />

poorly understood. Neuritin 1 is associated with pain. ATP2A3 is involved in<br />

calcium sequestration associated with muscular excitation and contraction.<br />

CONCLUSIONS: GnRHa treatment in endometriosis patients increased<br />

implantation and pregnancy rate in women undergoing IVF and down regulated<br />

the promising prognostic biomarkers: Tensin 1, MMP 14, Caveolin 2,<br />

Neuritin 1 and ATP2A3.<br />

Supported by: This work was Supported by the funding from the<br />

Academia Sinica (BM1040101<strong>17</strong>) and MOST(103-2314-B-038-054-MY2)<br />

of Taiwan<br />

OBJECTIVE: To determine whether recurrence of endometrioma affects<br />

ongoing pregnancy rates in IVF cycles.<br />

DESIGN: Retrospective Study<br />

MATERIALS AND METHODS: This is a retrospective study conducted<br />

at Uludag University Faculty of Medicine, Department of Obstetrics and Gynecology.<br />

Electronic data of the years between 2011-<strong>2015</strong> was screened and<br />

IVF patients with endometriosis were selected. Patiens were classified into<br />

three subgroups; patients with recurrence of endometrioma (Group 1), patients<br />

with primary endometrioma (Group-2) and patients without recurrence<br />

of endometrioma after surgery(Group-3) . Baseline characteristics, embroloygy<br />

laboratory parameters and pregnancy outcomes were analyzed and<br />

compared.<br />

RESULTS: Total 62 infertile patients with endometriosis were selected<br />

from electronic database. Baseline characteristics were similar in each group.<br />

Median duration after surgery was 2 years for IVF. 24 of these 42 patients<br />

with surgical history have had postoperative recurrence of endometrioma<br />

during ovarian stimulation. There have been 42 patients with endometrioma<br />

during ovarian stimulation. Mean endometrioma size was 3,48 cm (+/- 2,1).<br />

24 of 42 patients had recurrence of endometrioma and 18 of 42 patients had<br />

primary endometrioma. The rest of 62 patients had been surgery before and<br />

there were no recurrence of endometrioma in these patients. Embryologic<br />

laboratory parameters and ongoing pregnancy rates (OPR) were similar in<br />

three groups. Number of picked up oocytes were 7,8 (+/-4,8), 12,3(+/-8,7)<br />

and 10,2(+/-6,6), p:0,258, respectively group 1,2,3. Number of metaphase-<br />

2 oocytes were 5,6(+/-3,7), 7,89(+/-5,1) and 7,9(+/-6,1), p:0.164, respectively<br />

group 1,2,3. Pregnancy outcomes were analyzed in two parameters;<br />

positive B-hCG and ongoing pregnancy rate. Positive B-hCG rates were %<br />

58.3, %33.3, %55, respectively group 1,2,3, p:0.244. Ongoing pregnancy<br />

rates were %33.3, %22.2, %35, respectively group 1,2,3, p:0.655.<br />

CONCLUSIONS: Presence of endometrioma during ovarian stimulation<br />

in infertile patients does not affect embryology laboratory parameters and<br />

pregnancy outcomes. After surgical removal, presence of recurrence does<br />

not decrease pregnancy outcomes. Thus in conclusion, there is no need to<br />

re-operate the patients for removal of recurrence for better IVF outcomes.<br />

Embryology and Pregnancy Outcomes.<br />

Group 1<br />

Surgery (+)<br />

Endometrioma (+)<br />

Group 2<br />

Surgery (-)<br />

Endometrioma (+)<br />

Group 3<br />

Surgery (+)<br />

Endometrioma (-)<br />

p value<br />

Number of oocyte (sd) 7,8 (+/-4,8) 12,3(+/-8,7) 10,2(+/-6,6) 0,164<br />

Number of metaphase 5,6(+/-3,7) 7,89(+/-5,1) 7,9(+/-6,1) 0.258<br />

2 oocyte (sd)<br />

Number of 2PN (sd) 4,<strong>17</strong>(+/-2,8) 5,4(+/-4) 5,2(+/-3,9) 0.518<br />

Positive b-hCG (%) 14/24 (%58.3) 6/18 (%33.3) 11/20 (%55) 0.244<br />

Ongoing Pregnancy<br />

Rate (%)<br />

8/24 (%33.3) 4/18 (%22.2) 7/20 (%35) 0.655<br />

P-168 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

IMPROVEMENT IN ENDOMETRIOSIS-RELATED PELVIC PAIN<br />

WITH LEUPROLIDE OR NORETHINDRONE TREAT-<br />

MENT. O. Muneyyirci-Delale, a,b C. Charles, a N. Sinaii, c M. Dalloul, a<br />

P. Stratton. d a OB/GYN, SUNY Downstate Medical Center, Brooklyn, NY;<br />

b Ob/gyn, Kings County Hospital Center, Brooklyn, MD; c National Institutes<br />

of Health, Bethesda, MD; d NICHD, NIH, Bethesda, MD.<br />

OBJECTIVE: To compare the long-term effectiveness of 24 weeks Leuprolide<br />

Acetate Depot form 11.25mg (LD) vs Norethindrone Acetate 5mg<br />

(NA) followed by NA alone for 28 weeks in relieving pain from endometriosis.<br />

DESIGN: Prospective double-masked randomized clinical trial<br />

FERTILITY & STERILITY Ò<br />

e163


MATERIALS AND METHODS: 62 women with symptomatic endometriosis<br />

were randomized to LD (N¼31) or NA (N¼31) for 24 weeks in Phase<br />

I; then all were given NA for 28 weeks in Phase II. After 52 weeks of treatment,<br />

women were followed for up to one year (PTFU) with the option to restart<br />

NA. Visual analog scale (VAS) was used to assess menstrual (MP),<br />

coital (CP), and non-menstrual pelvic pain (NMP), evaluated at entry and<br />

compared to Phase I, Phase II and PTFU to assess improvement. Data<br />

were analyzed as intent-to-treat; VAS deltas were compared within phases<br />

using paired tests and between groups by two-sample tests. PTFU data<br />

considered NA use.<br />

RESULTS: Women were predominantly Black (81%) and single (66%),<br />

had a mean age of 34.1 6.7 yr and reported pelvic pain for 14.3 8.0 yr.<br />

Groups did not differ in these characteristics, stage of endometriosis,<br />

gravidity, BMI, or pain scores upon entry.In Phase I, women using NA had<br />

a significant reduction in VAS for MP (p


compared to controls. DNA methylation arrays showed heightened methylation<br />

of this gene’s promoter region in tissues from endometriosis patients<br />

with pain compared to controls.<br />

CONCLUSIONS: Our results thus far suggest that the upregulation of polycomb<br />

proteins in endometriotic tissues modulates DNA methylation of the<br />

forkhead protein resulting in its decreased expression. This likely leads to<br />

increased growth of the endometriotic tissue in the peritoneal environment,<br />

as well as increased pain. Our results have uncovered a network of proteins<br />

that may aid in the progression of endometriotic lesions. This underlying<br />

mechanism has yet to be exploited for therapeutic value in endometriosis.<br />

Supported by: Marshall University/University of Kentucky CCTS (NIHfunded).<br />

P-<strong>17</strong>3 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

EFFICACY OF NATURAL CYCLE ENDOMETRIAL PREPARA-<br />

TION FOR FROZEN-THAWED EMBRYO TRANSFER IN PATIENTS<br />

WITH ENDOMETRIOSIS. H. Guo. Shanghai Ninth People’s Hospital,<br />

Shanghai, China.<br />

P-<strong>17</strong>2 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE EPIGENETIC ROLE OF EZH2-FOXP3 CROSSTALK IN ENDO-<br />

METRIOSIS AND ITS ASSOCIATED PAIN. K. Wright N. Santanam.<br />

Pharmacology, Physiology, and Toxicology, Marshall University, Huntington,<br />

WV.<br />

OBJECTIVE: Determine the epigenetic mechanisms involved with polycomb<br />

group proteins and associated genes in the development and progression<br />

of endometriosis and its associated pain.<br />

DESIGN: Patient tissues classified by location (eutopic or ectopic) and the<br />

presence of pain were studied to determine trends in genetic expression,<br />

interaction, and methylation.<br />

MATERIALS AND METHODS: This study involved tissues from control<br />

(non-endometriosis) subjects without pain (n¼4), as well as eutopic (nonlesion,<br />

n¼7) and endometriotic (ectopic, n¼4) tissues from patients with<br />

endometriosis and pain. All specimens were obtained from IRB-approved<br />

and consented subjects. Real-time PCR was used to analyze gene expression<br />

of polycomb repressor complex 2 (PRC2) components and Western blot was<br />

used to determine the relative protein levels. Interaction of tumor suppressor<br />

Forkhead box P3 (FoxP3) as a potential target for the PRC2 catalytic subunit,<br />

Enhancer of zeste 2 (EZH2), was also measured using chromatin immunoprecipitation<br />

and immunoblotting. Lastly, Qiagen EpiTect Methyl II Array<br />

was also used to identify methylation patterns of genes associated with<br />

inflammation and autoimmunity in endometriotic fluids compared to controls.<br />

RESULTS: Expression of EZH2 was elevated in ectopic tissues compared<br />

to controls, while lowered in eutopic tissue in endometriosis patients (fold<br />

change ¼ 3.54, 0.11, respectively). Western blots also revealed increased<br />

levels of the tri-methylation of lysine 27 on histone 3 (H3K27me3). FoxP3<br />

gene expression was much lower in both ectopic and eutopic tissues<br />

OBJECTIVE: To assess the efficiency of natural cycle(NC) endometrial<br />

preparation for frozen-thawed embryo transfer (FET) in women with endometriosis(EMS).<br />

DESIGN: Retrospective study<br />

MATERIALS AND METHODS: This was a retrospective study of the use<br />

of the natural cycle for endometrial preparation following FET cycles in patients<br />

with endometriosis from March, 2011 to August, 2013. The study<br />

groups included 97 patients with stage I-II endometriosis who underwent a<br />

total of 120 FET cycles(group A, included cases of mild and minimal endometriosis)<br />

and <strong>17</strong>9 patients with stage III-IVendometriosis who underwent a<br />

total of 233 FET cycles(group B, included cases of moderate and severe<br />

endometriosis). The control group included 258 patients with tubal factor<br />

infertility who underwent a total of 300 FET cycles(group C). In stage III-<br />

IV endometriosis , 40 patients (47 FET cycles,groupB1) had endometrial<br />

cyst recurrence after laparoscopic or laparotomy treatment. All patients<br />

enrolled in the study met the following inclusion criteria: (i) normal uterine<br />

cavity as assessed by ultrasonography, hysterosalpingography or hysteroscopy;<br />

(ii) high-quality frozen embryos were transplanted. (iii)normal male<br />

semen. The exclusion criteria comprised: (i) presence of hydrosalpinges;<br />

(ii) presence of adenomyosis and (iii) past history of myomectomy;(iv) polycystic<br />

ovary syndrome(PCOS); (v)repeated failed intrauterine insemination<br />

(IUI ). For group A , B and C ,we used the natural cycle for endometrium<br />

preparation. The patients’ characteristics, including age, duration of infertility,<br />

body mass index, number of embryos transferred, average endometrial<br />

thickness of ET (mm), basal FSH, serum E2 and P of transplantation day, No.<br />

of previous failed embryo transfer cycle, the rates of basal FSH>10 were recorded.<br />

Secondary measures included clinical pregnancy rate, live birth rate,<br />

ongoing pregnancy rate, implantation rate, and pregnancy complication rate.<br />

RESULTS: There were comparable clinical pregnancy rate (43.33% stage<br />

I-II, 50.<strong>21</strong>% stageIII-IV and 45.33% tubal factor )and live birth rate(36.67%<br />

stage I-II, 41.63% stageIII-IV and 40.00% tubal factor ) among the three<br />

groups, not depending on severity of endometriosis . No differences were<br />

found in other pregnancy parameters , in terms of ongoing pregnancy<br />

rate(36.67% stage I-II, 41.63% stageIII-IVand 40.00% tubal factor ) , miscarriage<br />

rate(14.89% stage I-II, 16.24% stageIII-IV and 10.29% tubal factor ),<br />

pregnancy complication rate(7.69% stage I-II, 10.26% stageIII-IV and<br />

12.50% tubal factor ). No congenital birth defects were found in the endometriosis<br />

groups. In addition, when high-quality embryos are transferred, the<br />

pregnancy results are not affected by endometriomas.<br />

CONCLUSIONS: Natural cycle endometrial preparation for FET obtains<br />

similar pregnancy outcomes in patients with endometriosis compared with<br />

tubal infertility, and don’t increase the risk of birth defects and other complications.<br />

P-<strong>17</strong>4 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

SURGICAL MANAGEMENT OF ENDOMETRIOSIS DOES NOT<br />

AFFECT OXIDATIVE DAMAGE IN ENDOMETRIOTIC TISSUE<br />

OVER TIME. I. Al-Aref, L. R. Goodman, R. Flyckt, A. Goyal,<br />

T. Falcone. Cleveland Clinic, Cleveland, OH.<br />

OBJECTIVE: Oxidative stress has been considered as a potential factor<br />

involved in the pathology and progression of endometriosis. The objective<br />

of this study was to evaluate the natural progression of endometriosis by<br />

FERTILITY & STERILITY Ò<br />

e165


comparing oxidative damage in endometriotic tissue at two distinct time<br />

points.<br />

DESIGN: Retrospective cohort trial<br />

MATERIALS AND METHODS: Women who had presented between<br />

1999-2009 with pelvic pain and/or infertility between the ages of 18-45 years<br />

with a history of two distinct surgeries with biopsy proven endometriosis<br />

were included if there was banked blood and tissue samples available for review.<br />

Histopathochemistry was used to evaluate oxidative damage by<br />

measuring hydroxy deoxyguaninine (8OHdG) in endometriotic cells and to<br />

evaluate the cellular ability for DNA repair with 8-oxoguanine glycosylase<br />

1 (OGG1). The markers were measured in a biopsy sample of pathology<br />

proven endometriosis for each patient at each surgery and compared to<br />

each other using Wilcoxan rank-sum tests. The pathologist was blinded to<br />

whether the samples were from the primary or secondary surgery. All surgeries<br />

were performed by the same surgeon.<br />

RESULTS: There were seven patients who met criteria and were included<br />

in the study. The average age at first surgery was 32.4 +/- 4.4 years. The<br />

average time between surgeries was 2.0 +/- 1.1 years. Stage of disease and<br />

marker intensity and presence in percentage of cells at first and second surgery<br />

are shown in table 1.<br />

Table 1: Stage of endometriosis and markers at first and second surgery.<br />

Stage<br />

8-OHdG<br />

Intensity<br />

8-OHdG<br />

% of cells<br />

OGG1<br />

Intensity<br />

OGG1<br />

% of cells<br />

1 IV/IV 0/1 0/10 2/2 90/80<br />

2 IV/IV 2/3 80/90 3/2 90/30<br />

3 IV/IV 3/3 100/80 2/3 10/80<br />

4 IV/IV 2/2 80/60 2/2 50/80<br />

5 IV/IV 3/2 75/75 3/3 95/95<br />

6 IV/IV 2/2 40/5 3/3 90/90<br />

7 III/III 2/3 60/80 3/3 95/90<br />

Despite complete excision of all visible endometriosis at the time of first<br />

surgery, all patients had return of the disease to the same stage at their second<br />

surgery. Temporally, there was no significant difference between percentage<br />

of cells that expressed 8-OHdG (p ¼ 0.79) or OGG1 (p ¼ 0.81) between the<br />

two surgeries. There was a correlation of 8-OHdG percent of cells and intensity<br />

(r ¼ 0.80, p < 0.01) and OGG1 percent of cells and intensity (r ¼ 0.65, p<br />

< 0.01), but no temporal relationship between surgeries.<br />

CONCLUSIONS: Overall, There was no relationship of progression of<br />

oxidative damage over time. Based on these results, it appears that baseline<br />

oxidative damage in endometriotic cells is predictive of future cell damage<br />

and that surgical excision of endometriosis has no benefit in changing the underlying<br />

pathophysiology leading to oxidative damage.<br />

P-<strong>17</strong>5 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

WITHDRAWN<br />

P-<strong>17</strong>6 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE RELATIONSHIP OF CASPASE-3, CASPASE-9, MMP-9<br />

EXPRESSION AND C-1562T MMP-9 GENE POLYMORPHISM IN<br />

MENSTRUAL BLOOD AS THE ETIOPATHOGENESIS MARKER<br />

TO CLINICAL ENDOMETRIOSIS MANIFESTATION. T. Madjid.<br />

Faculty of Medicine, University of Padjadjaran, Bandung, Indonesia.<br />

OBJECTIVE: The aim of this study is to discover non-invasive diagnostic<br />

method of endo-metriosis using menstrual blood, and also to reveal clearer<br />

understanding in patho-genesis of endometriosis<br />

DESIGN: A case control study involving 149 women who visited the<br />

Reproductive Endocrinology and Fertility Clinic, FKUP/RSHS and RSHS<br />

network hospitals was performed, from February 2007 to February 2008.<br />

MATERIALS AND METHODS: Screening for suspected endometriosis<br />

was performed by history taking, physical examination, and additional examination.<br />

Diagnostic laparoscopy or laparotomy and biopsy was performed afterward.<br />

The immunocytochemical examination on caspase-3, caspase-9,<br />

MMP-9, and mmp-9 gene polymorphism of menstrual blood were performed.<br />

Based on the microscopic confir-mation of histopathological result,<br />

the relationship of endometriotic and non-endometriotic clinical manifestation<br />

with menstrual blood biomolecular marker was assessed.<br />

RESULTS: Sixty-three (42.28%) endometriosis cases and 86 (57.78%)<br />

non-endometriosis cases were found. From those subjects, 34 endometriosis<br />

cases and 48 non-endometriosis cases with complete data were enrolled in<br />

this study. The endometrial cells were successfully isolated using preservative<br />

solution, and with immunocytochemical assay, all samples from 34<br />

endometriosis subjects could be analyzed for the expression of caspase-3,<br />

caspase-9, and MMP-9. The decreased expression of caspase-3 and caspase-9,<br />

and increased expression of MMP-9 in endometriosis group were<br />

higher than those in non-endometriosis group (82.4% vs 77.1%, p¼0.562;<br />

97.1% vs 87.5%, pE-F¼0.129; and 85.3% vs 85.4%, p¼0.988 respectively).<br />

The frequency of allele C in -1562T region of mmp-9 gene was significantly<br />

increased in endometriosis group (p


P-<strong>17</strong>7 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

INCREASED HIPPO SIGNALLING PATHWAY PROMOTES CELL<br />

PROLIFERATION AND ANTI-APOPTOSIS IN<br />

ENDOMETRIOSIS. Y. Song, a W. Huang, b M. Zhou, a L. Xiao, a<br />

X. Feng. a a Obstetrics and Gynecology, West China Second University Hospital<br />

of SCU, Chengdu, China; b West China Second University Hospital of<br />

Sichuan University, Chengdu, China.<br />

OBJECTIVE: Endometriosis is one of the common reasons that results in<br />

female infertility and pelvic pain in reproductive women. However, the pathogenesis<br />

is unclear. The imbalance in proliferation and apoptosis of endometrial<br />

stromal cells(ESCs) is considered as an important role in pathogenesis of endometriosis.<br />

Several signalling pathways involve in these processes and exact<br />

regulation need to be confirmed in endometriosis. A newly established signalling<br />

pathway–Hippo signalling pathway participates in organ development and<br />

tumorigenesis which indicates that it plays a critical role in proliferation and<br />

apoptosis processes. However, the study about Hippo pathway and endometriosis<br />

is lack. Our objective is to explore the mechanism of Hippo signalling<br />

pathway controlling ESCs proliferation and apoptosis in endometriosis.<br />

DESIGN: We investigated expression levels of Hippo signaling components<br />

in the endometrium of women with or without endometriosis. We identified<br />

YAP regulated proliferation and apoptosis ability of endometrial<br />

stromal cells by cell culture and transfections. And endometriosis animal<br />

model of nude mice and verteporfn treatments were used to identify the intervention<br />

effect of Hippo signaling on endometriotic lesions.<br />

MATERIALS AND METHODS: Study population, Quantitative real-time<br />

RT-PCR, Western blotting, immunohistochemistry , cell culture, transfections,<br />

BrdU labeling, TUNEL assay, immunoprecipitation-quantitative<br />

PCR and animal model, verteporfn treatments were used in this study.<br />

RESULTS: Our data showed that increased expression of YAP and<br />

decreased expression of p-YAP in ectopic and eutopic endometrium, compared<br />

with normal endometrium. Further, we demonstrated that knockdown of YAP<br />

in eutopic ESCs could decrease proliferation and enhance apoptosis companied<br />

with decreased expression of CTGF and increased BAX/BCL-2 ratio. In addition,<br />

overexpression of YAP could make normal ESCs cells get higher proliferation<br />

and lower apoptosis ability, with increased CTGF and decreased<br />

BAX/BCL-2 ratio. Furthermore, We identified, by chromatin immunoprecipitation-quantitative<br />

PCR, that CTGF, Bcl-2 were direct downstream target genes<br />

of YAP in the regulation of proliferation and apoptosis, and the expression of<br />

CTGF, Bcl-2 were through the YAP-TEAD complex. At last, endometriosis animal<br />

model of nude mice were treatment with verteporfin–specific inhibitor of<br />

Hippo pathway, which disrupts the formation of the YAP-TEAD complex,<br />

significantly reduced endometric lesions and cell proliferation.<br />

CONCLUSIONS: These data indicate that hippo signalling pathway play<br />

a critical role in the pathogenesis of endometriosis and could represent as a<br />

novel direction for specific diagnosis and therapeutic method.<br />

References:<br />

1. Nasu K, Yuge A, Tsuno A, Nishida M, Narahara H.Involvement of<br />

resistance to apoptosis in the pathogenesis of endometriosis. Histol Histopathol.<br />

2009 ;24:1181-92.<br />

2. Zhao B, Li L, Guan KL.Hippo signaling at a glance. J Cell Sci.<br />

2010;123:4001-6.<br />

P-<strong>17</strong>8 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE USE OF LEVONORGESTREL-RELEASING INTRAUTERINE<br />

SYSTEM FOR LONG-TERM MAINTENANCE AFTER CONSERVA-<br />

TIVE LAPAROSCOPIC SURGERY FOR ENDOMETRIOSIS: IN THE<br />

RESPECT OF RECURRENT OVARIAN ENDOMETRIOMA. M. Kim, a<br />

Y. Cho, b S. Seong. a a Department of Obstetrics and Gynecology, CHA Gangnam<br />

Medical Center, CHA University, Seoul, Korea, Republic of; b Department of Obstetrics<br />

and Gynecology, Dong-A University Medical Center, Dong-A University,<br />

College of Medicine, Busan, Korea, Republic of.<br />

OBJECTIVE: A levonorgestrel-releasing intrauterine system (LNG-IUS)<br />

can be used for the secondary prevention of endometriosis-associated dysmenorrhea<br />

after surgery for endometriosis. Although the LNG-IUS effectively reduces<br />

endometriosis-associated pain, its efficacy in preventing ovarian<br />

endometrioma recurrence is questionable.In this study, we evaluated recurrence<br />

rate of ovarian endometrioma and surgical characteristics in patients who used<br />

LNG-IUS after the conservative laparoscopic surgery for endometriosis.<br />

DESIGN: Multicenter retrospective cohort study<br />

MATERIALS AND METHODS: We performed a retrospective review of<br />

women who underwent conservative laparoscopic enucleation of ovarian endometrioma<br />

and were treated with LNG-IUS at the CHA Gangnam Medical<br />

Center and Dong-A University Medical Center between January 2007 and<br />

September 2014. The inclusion criteria were as follows: (1) pathologically<br />

proven ovarian endometrioma; (2) undergone conservative laparoscopic surgery;<br />

(3) premenopausal status; (4) no residual ovarian lesions, as confirmed<br />

by transvaginal ultrasonography (TVS) before insertion of the LNG-IUS. Patients<br />

were excluded if they had undergone laparotomy or laparoscopic hysterectomy<br />

with or without oophorectomy. Endometrioma recurrence was<br />

defined as a cystic mass with a diameter of R20 mm on TVS.<br />

RESULTS: A total 61 patients were included in this study. The mean age of<br />

the patients was 36.2 5.9 years (range, 23-48 years), and the mean follow-up<br />

duration was 41.6 <strong>21</strong>.4 months (range, 8-98 months). Ovarian endometrioma<br />

recurrence was detected on TVS in 7 of the 61 (11.5%) patients. The cumulative<br />

recurrence rates of ovarian endometrioma per patient at 24, 36, and 60 months<br />

were 5.7%, 11.9%, and 23.1%, respectively. Surgical characteristics, such as the<br />

diameter of the cyst, disease stage, unilateral or bilateral involvement, and coexistence<br />

of deep endometriosis, and postoperative medical therapy were not<br />

associated with recurrence. Nulliparity at surgery was a significant risk factor<br />

for endometrioma recurrence during the treatment with LNG-IUS. When we anlayzed<br />

the recurrence rate according to the age, there was no recurrence of<br />

ovarian endometrioma in patients over 35 years old during the follow up period<br />

CONCLUSIONS: Postoperative long-term maintenance therapy with<br />

LNG-IUS could be a treatment option for prevention of recurrent endometrioma<br />

after conservative laparoscopic surgery for endometriosis, especially<br />

for women in older age.<br />

P-<strong>17</strong>9 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

EVALUATION OF FSHR GENE POLYMORPHISMS IN INFERTILE<br />

WOMEN WITH AND WITHOUT ENDOMETRIOSIS AND ITS COR-<br />

RELATION WITH HUMAN ASSISTED REPRODUCTION<br />

OUTCOMES. B. Bianco, a C. M. Trevisan, b R. Fernandes, a<br />

R. Oliveira, a C. Peluso, a D. M. Christofolini, b C. P. Barbosa. a a Faculdade<br />

de Medicina do ABC, Santo Andre, Brazil; b FMABC, Santo Andre, Brazil.<br />

OBJECTIVE: We aimed to characterize the current relationship between<br />

FSHR gene polymorphisms (Ala307Thr and Asn680Ser) with FSH serum<br />

levels and assisted reproduction outcomes [controlled ovarian hyperstimulation<br />

response, antral follicle counting (AFC), MII (metaphase II oocytes) and<br />

embryos] in Brazilian women that underwent assisted reproduction treatments<br />

with and without endometriosis.<br />

DESIGN: Cross-sectional study comprising 277 infertile women (136<br />

with endometriosis and 141 controls) submitted to assisted reproduction<br />

treatment.<br />

MATERIALS AND METHODS: All patients (cases and controls) were<br />

%38 years old, had the presence of both ovaries without morphological abnormalities,<br />

regular ovulatory cycle, body mass index %30, and no evidence<br />

of endocrine disorders. The control group was formed by patients with tubal<br />

obstruction or women with male factor involved in infertility causes. The<br />

endometriosis group included women with diagnosis of endometriosis<br />

confirmed by laparoscopy and/or laparotomy, with the disease staging according<br />

to American Society for Reproductive Medicine (ASRM, 1996)<br />

and histological evidence of disease. Genotyping of FSHR polymorphisms<br />

were performed using TaqMan methodology by real time PCR. FSH was<br />

measured by Enzyme-linked fluorescence assay. The data was analyzed statistically<br />

(p-value


P-180 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ENDOMETRIUM<br />

THE MIXTURE OF RECOMBINANT HUMAN EPIDERMAL<br />

GROWTH FACTOR, POLOXAMER AND SODIUM ALGINATE<br />

PROMOTES ENDOMETRIAL GROWTH IN INFERTILE WOMEN<br />

WITH THIN ENDOMETRIUM. C. Kim, a J. Moon, a B. Kang, a<br />

W. Cho, b I. Park, b H. Kim. c a Obstetrics and Gynecology, College of Medicine,<br />

University of Ulsan, Asan Med, Seoul, Korea, Republic of; b R&D<br />

Center, Genewel, Co., Seongnam, Korea, Republic of; c R & D Center, Genewel,<br />

Co., Seoul, Korea, Republic of.<br />

OBJECTIVE: To evaluate the effect of the mixture of recombinant human<br />

epidermal growth factor (rhEGF), poloxamer and sodium alginate on endometrial<br />

growth and subendometrial artery (SEA) blood flow in infertile<br />

women with poorly developed thin endometrium.<br />

DESIGN: Prospective observational study.<br />

MATERIALS AND METHODS: A total of 68 infertile patients with thin<br />

endometrium less than 6mm were recruited for this study between March<br />

2013 and September 2014. All subjects aged 28-41 years underwent intrauterine<br />

treatment with either the mixture of rhEGF, poloxamer and sodium<br />

alginate (study group, n¼38) or rhEGF alone solution (control group,<br />

n¼30) every 2-3 days in the follicular phase. Six hundred IU of vitamin E<br />

was orally administered daily to all subjects throughout the whole treatment<br />

period.Mean values were expressed as meanstandard deviation (SD). Student’s<br />

t-test was used to compare mean values. Chi-square test and Fisher’s<br />

exact test were used to compare fraction. Statistical significance was defined<br />

as P


P-183 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

EMBRYO-ENDOMETRIUM ASYNCHRONY MAY AFFECT<br />

THAWED EMBRYO CYCLES USING GENETICALLY SCREENED<br />

BLASTOCYSTS. B. S. Shapiro, S. Daneshmand, F. Garner,<br />

M. Aguirre, C. Hudson. Fertility Center of Las Vegas, Las Vegas, NV.<br />

OBJECTIVE: Compare implantation rates of day 5 blastocysts transferred<br />

on day 6 and day 6 blastocysts transferred on day 7 after embryos were held<br />

overnight in order to obtain results of genetic screening.<br />

DESIGN: IRB-approved retrospective analysis.<br />

MATERIALS AND METHODS: This study included cycles in which<br />

thawed 2pn oocytes were cultured to the blastocyst stage before trophectoderm<br />

biopsy, genetic testing, and transfer in a 3-year time period (2012-<br />

2014). All cycles had artificial endometrial preparation with exogenous estradiol.<br />

Progesterone injections were initiated the evening before 2pn oocyte<br />

thaw. Embryos were cultured to the blastocyst stage, biopsied, and genetically<br />

screened for aneuploidy by array CGH. Transfers occurred the day after<br />

blastocysts formed and were biopsied, as genetic test results arrived the<br />

following morning, so that blastocysts that formed on day 5 of embryo development<br />

were transferred on day 6 and blastocysts that formed on day 6 were<br />

transferred on day 7 of embryo development. These corresponded with transfers<br />

on days 6 and 7 of progesterone exposure, respectively. Implantation and<br />

clinical pregnancy calculations were based on fetal hearts observed on ultrasound<br />

at 6-7 weeks gestation.<br />

RESULTS: The study included 39 day 6 transfers and 28 day 7 transfers<br />

following biopsy and genetic screening. The two groups did not differ significantly<br />

in age at retrieval or number of embryos transferred. The number of<br />

hours of pre-transfer progesterone exposure was significantly lower in the<br />

day 6 transfer group when compared to the day 7 transfer group (mean<br />

137 hours vs 160 hours, P


P-186 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

USE OF CLOMIPHENE CITRATE IN MINIMAL OVARIAN STIMU-<br />

LATION CYCLES NEGATIVELY IMPACTS ENDOMETRIAL<br />

THICKNESS, INDEPENDENT OF PEAK ESTRADIOL LEVELS:<br />

AN ARGUMENT FOR A FREEZE-ALL APPROACH. B. G. Reed,<br />

M. Ezzati, S. N. Babayev, V. Libby, B. Carr, O. Bukulmez. UT Southwestern<br />

Medical Center, Dallas, TX.<br />

OBJECTIVE: Minimal ovarian stimulation cycles use daily clomiphene<br />

citrate (CC) and a small amount of gonadotropin on days 5, 7, and 9 of the<br />

ovarian stimulation. Extrapolating from data on intrauterine insemination cycles<br />

that indicate CC may have a detrimental effect on the endometrial stripe<br />

(ES), we avoid fresh embryo transfer in patients undergoing minimal stimulation<br />

despite the lack of data from IVF cycles to support this freeze-all practice.<br />

We sought to determine if ES is negatively affected in minimal<br />

stimulation as compared to other types of ovarian stimulation that do not utilize<br />

CC. In addition, we sought to compare each patient’s ES during her minimal<br />

stimulation cycle(s) with her ES during her subsequent frozen embryo<br />

transfer (FET) cycle.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: A total cohort of 141 cycles (75 patients)<br />

were analyzed: 78 minimal ovarian stimulation cycles and their 31 subsequent<br />

FETs, 13 mild stimulation cycles, and 19 high dose gonadotropin<br />

GnRH antagonist cycles. One way ANOVA and Tukey’s test were used for<br />

continuous variables. Chi square test was used to compare the cumulative<br />

live birth rate per patient. Paired t-test was used to compare the ES in the<br />

same patient during the minimal stimulation and her subsequent FET cycle.<br />

RESULTS: Therewas no statistically significant difference in age, body mass<br />

index (BMI), antimullerian hormone (AMH) level, or cumulative live birth rates<br />

between the study groups (Table 1). Maximum ES in the minimal stimulation<br />

group was significantly lower than the other two groups (8.1 mm versus 13.6<br />

mm and 13.9 mm). Peak estradiol level was significantly higher in the high<br />

dose gonadotropin group as compared to both the minimal or mild stimulation<br />

groups. Despite a significant difference in peak estradiol levels, there was no difference<br />

in ES between the mild and high-dose stimulation groups. In patients<br />

who underwent minimal stimulation IVF followed by FET, significantly thicker<br />

ES was observed during their FET cycles (7.9 mm vs 10.3 mm).<br />

CONCLUSIONS: Based on our data, endometrial thickness is negatively<br />

impacted during minimal stimulation IVF cycles, independent of peak estradiol<br />

levels. The negative effects on the endometrial thickness are not<br />

observed in the subsequent FET cycles and, hence, a freeze-all approach<br />

can be adopted to mitigate potential adverse endometrial effects in minimal<br />

stimulation IVF cycles.<br />

Table 1. Demographics and outcomes.<br />

Minimal<br />

stimulation<br />

(Included use<br />

of CC)<br />

Mild<br />

stimulation<br />

(Did not<br />

include<br />

the use of CC)<br />

High dose<br />

gonadotropin<br />

GnRH antag<br />

stimulation<br />

(Did not include<br />

the use of CC)<br />

Frozen embryo<br />

transfer following<br />

minimal<br />

stimulation<br />

(Did not include<br />

the use of CC)<br />

p-value<br />

Age (y) 384 38.83.5 37.64.5 - NS<br />

BMI 264.8 25.65.7 25.24.1 - NS<br />

AMH (ng/ml) 0.60.7 0.50.5 1.00.7 - NS<br />

Maximum ES (mm) 8.12 13.63.5 13.93.8 -


P-189 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

GENE EXPRESSIONS OF L-SELECTIN LIGANDS ARE<br />

DECREASED AT MID-SECRETORY PHASE IN WOMEN WITH<br />

ADENOMYOSIS. T. Lai, a,b,c F. Lee, a,b Q. Ling. c,d a Obstetrics and Gynecology,<br />

Cathay General Hospital, Taipei, Taiwan; b School of Medicine, Fu<br />

Jen Catholic University, New Taipei, Taiwan; c Institute of Systems Biology<br />

and Bioinformatics, National Central University, Taoyuan, Taiwan; d Cathay<br />

Medical Research Institute, Cathay General Hospital, Taipei, Taiwan.<br />

OBJECTIVE: To evaluate the gene expression of L-selectin ligands’ subtypes<br />

during different phases of the menstrual cycles in women with adenomyosis.<br />

DESIGN: A prospective study on endometrial samples from reproductiveaged<br />

women with adenomyosis at various phases in the normal menstrual cycles.<br />

MATERIALS AND METHODS: The Institutional Review Board of Cathay<br />

General Hospital approved this study. Tissue samples of endometrium<br />

were collected from women (age 35-49 y) with adenomyosis undergoing hysterectomy.<br />

Forty-two endometrial biopsies included 12 at proliferative (day7-<br />

14), 10 at early secretory (day15-19), 9 at mid-secretory (day20 to 24) and 11<br />

at late secretory phase (dayS 25). Immunohistochemistry was performed using<br />

the antibody, MECA-79, which recognizes the ligands for the L-selectin<br />

adhesion molecule. The mRNA expression profiles of L-selectin ligand’s<br />

genes were analyzed using RT-PCR for the 4 L-selectin ligand’s subtypes: podocalyxin,<br />

endomucin, GlyCAM-1 and CD34. Western blot were used to<br />

quantify protein levels. Kruskal-Wallis (K-W) analysis with multiple comparisons<br />

was performed to examine differences between different phases in the<br />

menstrual cycle. Findings with a two-sided P value < 0.05 were considered<br />

to indicate statistically significant differences between the phases.<br />

RESULTS: Mean age and BMI of the patients among 4 phases were no<br />

different (P>0.05). Immunohistochemistry showed L-selectin ligands were expressed<br />

in the luminal and glandular epithelium of the endometrium at all the<br />

phases. The intensity of L-selectin ligands was greatest at mid-secretory phase<br />

which was at the stage of implantation window. However, the gene expression<br />

profiles of L-selectin ligands including endomucin, GlyCAM-1 and CD34 revealed<br />

gradually decreased from proliferative to mid-secretory phase. Endomucin,<br />

GlyCAM-1 and CD34 had the lowest expression during mid-secretory<br />

phase. Interestingly, the highest gene expression of podocalyxin, endomucin,<br />

GlyCAM-1 and CD34 were found at late secretory phase (P


OBJECTIVE: It has been reported that exogenous hCG from a trigger injection<br />

impairs embryo implantation and endometrial receptivity. This study<br />

was performed to determine if a relationship exists between exogenous hCG<br />

exposure (dose, serum level) and clinical pregnancy in fresh autologous embryo<br />

transfer cycles, while controlling for confounding variables.<br />

DESIGN: IRB-approved retrospective study.<br />

MATERIALS AND METHODS: Cycles in which both hCG and GnRH<br />

agonist were used in combination for oocyte maturation were selected so<br />

that (1) a broad range of hCG doses would be available, and (2) the effect<br />

of hCG in oocyte maturation would be mitigated by the redundant LH<br />

response to the agonist. This study included all such cycles (agonist+hCG)<br />

that had fresh autologous embryo transfer. hCG doses were varied among patients<br />

and were steadily decreased as the protocol matured over the 9-year<br />

study period (2006-2013). GnRH agonist dose was 4mg leuprolide acetate<br />

(off-label use) throughout the study period. Serum levels of hCG were<br />

measured approximately 12 hours after trigger. Fresh embryos were transferred<br />

at the blastocyst stage. Stepwise logistic regression was used to identify<br />

significant predictors of clinical pregnancy. Chi-square analysis was used<br />

for univariate comparison of clinical pregnancy rates.<br />

RESULTS: There were 349 included cycles. hCG doses varied from 500IU<br />

to 10,000IU (<strong>17</strong>7<strong>21</strong>501 IU), and post-trigger hCG serum levels varied from<br />

11 to 354 IU/L (51.539.7 IU/L). Logistic regression identified day of transfer<br />

(P


P-195 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

DONOR TREATMENT WITH MYCOPHENOLATE MOFETIL<br />

ATTENUATES UTERINE ISCHEMIA/REPERFUSION<br />

INJURY. G. Sahin Ersoy, a M. Eken, b O. Cevik, c O. T. Cilingir. d a Obstetrics<br />

and Gynecology, Kartal Dr Lutfi Kirdar Research and Education Hospital,<br />

Istanbul, Turkey; b Zeynep Kamil Education and Research Hospital,<br />

Istanbul, Turkey;<br />

c Cumhuriyet University, Faculty of Pharmacy, Sivas,<br />

Turkey; d Marmara University School of Medicine, Istanbul, Turkey.<br />

OBJECTIVE: Uterine transplantation is a potential option to treat uterine<br />

factor infertility. Mycophenolate mofetil (MMF) is a powerful immunosuppressive<br />

agent that is currently used in uterine transplantation. However<br />

ischemia/reperfusion (I/R) injury has a detrimental effect on early allograft<br />

dysfunction in organ transplanted patients. Therefore we aimed to investigate<br />

whether MMF could attenuate uterine I/R injury.<br />

DESIGN: Experimental study.<br />

MATERIALS AND METHODS: 28 female Wistar rats used for all experiments.<br />

Group I (M): underwent sham surgery and received 20 mg/kg/day<br />

MMF (0.5% sodium carboxymethylcellulose), Group 2 (C): underwent<br />

sham surgery and received vehicle control, Group 3 (IR+V): underwent uterine<br />

I/R procedure and received vehicle control, Group 4 (IR+M): underwent<br />

uterine I/R procedure and received 20 mg/kg/day MMF. All treatments were<br />

conducted by gavage and started 5 days before surgery and maintained until<br />

uteri were removed. Ischemia reperfusion was conducted for 30 minutes.<br />

Specifically ischemia was created by clamping the distal abdominal aorta followed<br />

by 4 hours reperfusion. Uterine tissue ischemia modified albumin<br />

(IMA), malondialdehyde (MDA), glutathione (GSH), superoxide dismutase<br />

(SOD) and miyeloperoxidase (MPO) levels were determined as markers of<br />

oxidative injury. 8 hydroxydeoxyguanosine (8-OHdG) was investigated to<br />

evaluate the levels of oxidative DNA injury. Histopathologic examination<br />

were conducted following hematoxylin-eosin staining.<br />

RESULTS: Histologic examination of IR+M group had more regular<br />

endometrial contours, glands and less polymorphonuclear cell infiltration<br />

than the IR+V group.<br />

CONCLUSIONS: Pretreatment with MMF attenuates I/R injury. This protection<br />

may have occurred by an anti-inflammatory and antioxidant effect.<br />

Therefore MMF used as part of an immunosuppressant regimen may also<br />

provide a cytoprotective effect against uterine warm ischemia injury.<br />

Table 1. The levels of oxidative injury and DNA injury markers.<br />

Groups C IR+V IR+M<br />

IMA (U/mL) 0.23 0.05 0.56 0.08* 0.41 0.07*#<br />

GSH (umol/mg protein) 2.31 0.35 1.74 0.23* 1.88 0.20<br />

MDA (mmol/mg protein) 10.5 1.50 15.96 1.<strong>21</strong>**12.13 1.14##<br />

MPO (U/mg protein) 1.60 0.43 6.27 2.26* 3.05 0.70*#<br />

SOD (U/mg protein) 4.95 1.39 1.98 0.45** 3.76 1.30<br />

8OHdG (ng/ug DNA) 1.06 0.12 1.82 0.40* 1.22 0.12#<br />

Data are shown mean SD. * P


P-198 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

EVALUATION OF OVARIAN RESERVE BEFORE AND AFTER<br />

SURGERY IN WOMEN WITH OVARIAN ENDOMETRIOMA<br />

COMPARED TO PELVIC ENDOMETRIOSIS AND NEGATIVE<br />

CONTROLS: A PROSPECTIVE COHORT TRIAL. L. R. Goodman,<br />

J. M. Goldberg, R. Flyckt, M. Gupta, N. Gueye, k. J. Holoch,<br />

T. Falcone. Cleveland Clinic, Cleveland, OH.<br />

OBJECTIVE: To determine the impact of surgical excision of endometriosis<br />

and endometrioma compared with controls on ovarian reserve.<br />

DESIGN: Prospective cohort study.<br />

MATERIALS AND METHODS: Women aged 18-43 years presenting<br />

with pelvic pain and/or infertility undergoing surgical management of suspected<br />

endometriosis or endometrioma between <strong>October</strong> 2013 and April<br />

<strong>2015</strong>. Patients were excluded if they had prior ovarian surgery. Statistical<br />

analysis included student t-tests, paired t-tests and ANOVA where appropriate.<br />

RESULTS: A total of 116 patients were included with 58 suspected endometriomas<br />

and 58 controls with suspected pelvic endometriosis but no evidence<br />

of ovarian involvement on pelvic imaging. All surgeries were<br />

performed laparoscopically. All of the suspected endometriomas were<br />

completely removed by cystectomy and confirmed by pathology. Of the 58<br />

controls, 31 had biopsy confirmed pelvic endometriosis and 27 patients<br />

had no evidence of endometriosis on laparoscopy with pathology negative<br />

for endometriosis on biopsy. We prospectively evaluated ovarian reserve<br />

measured by Anti-Mullerian Hormone (AMH) prior to surgery, at one month<br />

and six months post-operatively. Age, body mass index and proportion of patients<br />

with pain and/or infertility did not differ significantly between groups.<br />

Table 1: Means (95% Confidence Intervals), { } are p-values compared to preoperative<br />

value.<br />

Endometrioma<br />

(n ¼ 56)<br />

Pelvic<br />

Endometriosis<br />

(n ¼ 31)<br />

Negative<br />

Laparoscopy<br />

(n ¼ 27) P-Value<br />

Age (years) 31.5 (30.1 - 33.3) 31.1 (28.9 - 33.3) 30.9 (28.8 - 33.1) 0.55<br />

Length of 119.8 (105.8 - 128.1) 87.7 (72.6 - 102.7) 69.7 (53.6 - 85.9)


P-201 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ASSOCIATION BETWEEN STATE-MANDATE INSURANCE<br />

COVERAGE FOR INFERTILITY TREATMENT AND SURGICAL<br />

APPROACH FOR MANAGEMENT OF ECTOPIC<br />

PREGNANCY. R. Schickler, a E. Mikhail, b J. Salemi, c A. Imudia, d<br />

S. Plosker, d H. Salihu. c a Department of Obstetrics and Gynecology, University<br />

of South Florida Morsani College of Medicine, Tampa, FL; b Department<br />

of Minimally Invasive Surgery and Department of Obstetrics and Gynecology,<br />

University of South Florida Morsani College of Medicine, Tampa,<br />

FL; c Department of Family and Community Medicine, Baylor College of<br />

Medicine, Houston, TX; d Department of Reproductive Endocrinology and<br />

Department of Obstetrics and Gynecology, University of South Florida Morsani<br />

College of Medicine, Tampa, FL.<br />

OBJECTIVE: To estimate the association between trends of surgical management<br />

of ectopic pregnancy and the availability of state-mandated infertility<br />

insurance coverage (SIC) including in-vitro fertilization.<br />

DESIGN: A cross-sectional analysis of the Healthcare Cost and Utilization<br />

Project’s Nationwide Inpatient Sample (HCUP-NIS), the largest allpayer<br />

inpatient database in the United States, was conducted for all patients<br />

who underwent surgical management of ectopic pregnancy between 1998<br />

and 2011.<br />

MATERIALS AND METHODS: All community hospitals in the US<br />

participating in HCUP, excluding rehabilitation and long-term acute care<br />

hospitals were included. The study includes all inpatient hospitalizations<br />

for women receiving treatment for ectopic pregnancy as identified using<br />

ICD-9-CM diagnosis codes. Joinpoint regression was used to estimate temporal<br />

trends (including the annual percent change [APC]) in salpingectomy<br />

versus salpingostomy during the study period. Stratified analyses were performed,<br />

investigating trends in states with and without SIC (including invitro<br />

fertilization). Those states with SIC were Connecticut, Illinois, Massachusetts,<br />

New Jersey, and Rhode Island.<br />

RESULTS: Between 1998 and 2011, there were over 419,000 inpatient<br />

hospitalizations for ectopic pregnancy, approximately 14% of which<br />

occurred in states with SIC. We observed a gradual increase in the rate of salpingectomy<br />

in states with SIC from 69.3% in 1998 to 80.9% in 2011 (APC:<br />

1.5%; 95% CI: 1.2, 1.8). Conversely, among ectopic pregnancies in states<br />

without SIC, there were no statistically significant changes in temporal trends<br />

of salpingectomy. Overall, there were significant decreasing temporal trends<br />

in both groups when it comes to salpingostomy. The APC for salpingostomy<br />

during the study period in states with SIC was -7.7% (95% CI: -9.0, -6.3). In<br />

non-mandated states, the APC for salpingostomy changed from -2.9% (95%<br />

CI: -5.7, -0.1) from 1998-2004 to -9.8% (95% CI: -12.9, -6.7) from 2004-<br />

2011.<br />

CONCLUSIONS: The availability of state mandated insurance coverage<br />

for infertility treatment is correlated with the choice of surgical management<br />

approach for ectopic pregnancy. In states with SIC, the rate of salpingectomy<br />

increased significantly over time compared to states without such mandate.<br />

Irrespective of infertility treatment mandate status, there was a decline in<br />

the rate of salpingostomy as a surgical management approach for ectopic<br />

pregnancy.<br />

P-202 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

REVIEW AND OUTCOMES OF POWER MORCELLATION USING<br />

AN INNOVATIVE CONTAINED BAG SYSTEM. C. E. Miller, a<br />

C. Steller, b K. Sasaki, c A. Cholkeri-Singh. d a The Advanced IVF Institute,<br />

Naperville, IL; b Advocate Lutheran General Hospital, Park Ridge, IL; c Advanced<br />

Gynecologic Surgery Institute, Naperville, IL; d The Advanced Gynecologic<br />

Surgery Institute, Naperville, IL.<br />

OBJECTIVE: To assess feasibility of an innovative method of power morcellation<br />

within a specimen bag.<br />

DESIGN: Retrospective Chart Review from a specialty practice in suburban<br />

Chicago.<br />

MATERIALS AND METHODS: The study included patients who underwent<br />

power morcellation during a laparoscopic or robotic-assisted hysterectomy<br />

or myomectomy from May 2014 through March <strong>2015</strong>. Exclusion<br />

criteria were known uterine malignancy or successful specimen removal<br />

without morcellation.The procedure was performed using the Espiner Eco-<br />

Sac 230. The bag was inserted through the umbilical port and the specimen<br />

placed inside. The cinched bag edge was pulled out through a 15mm umbilical<br />

incision. The bag was insufflated to 25mm Hg allowing a lateral trocar to<br />

be placed through the bag for laparoscope access. Morcellation was performed<br />

through the umbilical incision.<br />

RESULTS: Of the 93 procedures performed, 72% were myomectomies<br />

and 27% were hysterectomies. The patients’ mean age was 39 years old<br />

and mean BMI was 28.5 (range <strong>17</strong>.2-49.9). The average specimen weight<br />

was 288 grams, with the largest weighing 2,134 grams. Estimated blood<br />

loss (EBL) averaged 77 milliliters. The average operating time was 2.6 hours.<br />

The postoperative admission rate was 11%, majority of which were due to<br />

nausea. Four patients (4.6%) had minor postoperative complications and 1<br />

patient was readmitted to an outside hospital with vomiting and constipation<br />

which resolved with conservative treatment. There were no bag failures or<br />

complications that were due to use of the specimen bag or due to morcellation.<br />

CONCLUSIONS: Using the Espiner EcoSac 230 specimen bag was successfully<br />

performed in 93 patients with minimal complications. This is a<br />

feasible, reliable and reproducible method of contained power morcellation,<br />

even for a large specimen.<br />

P-203 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

SURGICAL OUTCOMES AND COST ANALYSIS OF ABDOMINAL,<br />

MINI-LAPAROTOMY, AND TRADITIONAL AND ROBOTIC-ASSIS-<br />

TED LAPARASCOPY WITH AND WITHOUT TANDEM MINI-<br />

LAPAROTOMY: A COMPARISON OF MYOMECTOMY<br />

TECHNIQUES. J. Stanhiser, B. Mouille, J. M. Goldberg, T. Falcone,<br />

L. R. Goodman. Women’s Health Institute, Cleveland Clinic, Cleveland, OH.<br />

OBJECTIVE: With the recent FDA warning against laparoscopic power<br />

morcellation necessitating larger incisions for leiomyoma extraction at the<br />

time of myomectomy, the objective of this study was to compare the surgical<br />

outcomes and cost of myomectomy by abdominal, mini-laparotomy, and<br />

traditional and robotic-assisted laparoscopy with and without tandem minilaparotomy<br />

techniques.<br />

DESIGN: Retrospective cohort study of patients undergoing myomectomy<br />

January 2010 to February <strong>2015</strong> in a tertiary care center.<br />

MATERIALS AND METHODS: Surgical outcomes included myomectomy<br />

type, operative time, estimated blood loss (EBL), myoma weight,<br />

length of stay (LOS) and postoperative complications. Hospital costs were<br />

assessed. Statistical analysis was by Student’s t-test and Wilcoxon rank<br />

sum test.<br />

RESULTS: 274 myomectomies were included. The average age was 38.7<br />

5.7 years and did not differ between groups. Body mass index (BMI) and<br />

outcomes are listed in Table 1. In laparoscopic cases, the addition of a minilaparotomy<br />

significantly lengthened operative time (p < 0.01) and increased<br />

EBL (p < 0.01), but the myoma weight was significantly higher (p < 0.01).<br />

These differences were not observed in robotic-assisted cases (p¼0.69,<br />

p¼0.07, p¼0.57 respectively), however, robotic cases were significantly<br />

longer and had a higher EBL than laparoscopic cases (p


procedures; however when combined, LOS significantly increased (p <<br />

0.01). The average robotic myomectomy spent one night in the hospital<br />

and the addition of a mini-laparotomy did not significantly increase LOS<br />

(p¼0.41). Complications and blood transfusions were significantly more<br />

common in the abdominal and traditional laparoscopic with tandem minilaparotomy<br />

techniques (p < 0.01). Cost was lowest with solo mini-laparotomy,<br />

higher in the laparoscopic and highest with the robotic group. The<br />

cost of traditional laparoscopy significantly increased 40% when a tandem<br />

mini-laparotomy was performed (p¼0.04), however cost was not significantly<br />

increased in robotic-assisted cases (p¼0.88).<br />

CONCLUSIONS: The addition of tandem mini-laparotomy to laparoscopic<br />

myomectomy increases operative time, EBL, hospital length of<br />

stay, postoperative complications, and cost. This study supports the need<br />

for further investigation of minimally invasive myoma extraction techniques.<br />

P-204 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

VALUE OF HERBAL MEDICINE IN PREVENTION OF POSTOPER-<br />

ATIVE INTRAUTERINE ADHESIONS (MISURATA<br />

EXPERIENCE). A. M. Elbareg. Obstetrics & Gynaecology, Misurata<br />

University, Misurata, Libyan Arab Jamahiriya.<br />

OBJECTIVE: Intrauterine adhesions (IUAs) are the major long-term complications<br />

of operative hysteroscopy(OH). It may result in poor reproductive<br />

outcomes, menstrual abnormalities and pelvic pain. Several methods have<br />

been tried to prevent post-surgical IUAs but all without significant benefits.<br />

Moist exposed burn ointment (MEBO) is a herbal product, containing phytosterols<br />

with anti-inflammatory, anti-bacterial and analgesic effects mainly<br />

used for wound healing and management of burns and ulcers, such properties<br />

might be useful to avoid IUAs, therefore, aim of this study is to assess the<br />

efficacy and safety of MEBO in prevention of de novo IUAs following OH.<br />

DESIGN: A randomized controlled University hospitals-based study.<br />

MATERIALS AND METHODS: Selected 72 patients during a period of<br />

12 months were enrolled in this study. Inclusion criteria: hysteroscopic diagnosis<br />

of submucous myomas, polyps, uterine septa or DUB requiring endometrial<br />

resection. Exclusion criteria were: age > 50, obesity, pregnancy,<br />

prolapse, malignancy, and presence of IUAs. Patients were randomized<br />

into 2 groups, and without significant differences in age, weight, uterine<br />

size and party. Group (A) of 40 patients underwent hysteroscopic surgery(HS)<br />

plus intrauterine application of MEBO. Group (B), the remaining 32<br />

patients, underwent (HS) only (control). Follow-up diagnostic hysteroscopy<br />

was performed for each patient, 6 weeks later. Adhesion score assessment according<br />

to the American Fertility Society (AFS) classification system. Statistical<br />

analysis performed using SPSS package. P-value considered to be<br />

significant if (< 0.05).<br />

RESULTS: Eleven women ( 4 from group A & 7 from group B) did not<br />

attend for follow-up. At 6 weeks follow-up, a highly significant lower rate<br />

of post-surgical IUAs was found in group (A): only one out of 36 women<br />

compared with group (B): 8 out of 25 patients {2.8 % versus 32 % ; P <<br />

0.05}. Evaluation of IUAs staging by AFS system, showed a significant<br />

decrease in adhesions severity in the patient from group A ( stage I ) when<br />

compared with women from group B: 6 patients (stage II) & 2 patients (stage<br />

III ){P < 0.05}. No complications or adverse MEBO-related effects were detected<br />

in group A patients.<br />

CONCLUSIONS: (MEBO), is safe, non-expensive and with a highly significant<br />

reduction rate in the incidence and severity of de-novo formation of<br />

IUAs after HS.<br />

P-205 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE IMPACT OF THE TECHNIQUES OF HYSTEROSCOPIC SEP-<br />

TOPLASTY ON THE REPRODUCTIVE OUTCOMES _IN PATIENTS<br />

WITH SEPTATE UTERUS. O. Dural, C. Yasa, E. Bastu, F. Gungor<br />

Ugurlucan, S. Can, G. Yilmaz, F. Buyru. Department of Obstetrics and Gynecology,<br />

Istanbul University School of Medicine, Istanbul, Turkey.<br />

OBJECTIVE: The study was aimed to determine the reproductive outcomes<br />

of two different techniques of hysteroscopic septoplasty.<br />

DESIGN: We retrospectively studied 49 patients who underwent hysteroscopic<br />

septoplasty (HS) for symptomatic septate uteri between January 2010<br />

and April 2014.<br />

MATERIALS AND METHODS: We divided the patients into two groups<br />

on the basis of the technique of HS. Group I involved 27 patients with a history<br />

of HS which was performed by monopolar hook using the operating hysteroscope<br />

8 mm in diameter. Group II involved 22 patients with a history of<br />

HS which was performed by scissors using office hysteroscopy 5mm in diameter.<br />

We evaluated pregnancy outcomes within first year after HS for Group I<br />

and II.<br />

RESULTS: Reproductive outcomes were obtained from 44 patients who<br />

tried to get pregnant after HS within first year. Grup I and II include 25<br />

and 19 patients with median age 27,5 (19-38) and 28 (<strong>21</strong>-39), respectively.<br />

There were a total of 23 pregnancies in Group I, of which 15 (65.2%) were<br />

carried to term, 3 (13%) resulted in preterm live birth, 5 (<strong>21</strong>.7%) resulted<br />

in pregnancy loss in the first or second trimester. There were a total of <strong>17</strong><br />

pregnancies in Group II, of which 11 (64.7%) were carried to term, 2<br />

(11.7%) resulted in preterm live birth, 4 (23.5%) resulted in pregnancy<br />

loss in the first or second trimester. The overall life birth rate was 78.2 %<br />

in Group I and 76.4% in Group II (p¼0.85). One case of uterine rupture occured<br />

at 37 week gestation in a subsequent pregnancy in Group I.<br />

CONCLUSIONS: There is a scarcity of data regarding whether using of a<br />

specific instrument during HS may improve reproductive outcome. Our data<br />

support that the rates of pregnancy to term and live birth are similar between<br />

two techniques of HS. In accordance with a few previous case reports , uterin<br />

rupture occured after HS with monopolar diathermy. Additional studies are<br />

needed to evaluate the impact of the techniques of HS on the reproductive<br />

outcomes.<br />

P-206 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

LAPAROSCOPIC AND ROBOTIC MYOMECTOMY: COMPARISON<br />

OF COST AND PERIOPERATIVE OUTCOMES FOR COMPLEX<br />

CASES. M. V. Vargas. Department of Obstetrics and Gynecology, Division<br />

of Gynecology, George Washington University Medical Center, Washington,<br />

DC.<br />

OBJECTIVE: To compare cost and outcomes between laparoscopy and robotic-assistance<br />

in women undergoing complex myomectomy.<br />

DESIGN: Retrospective chart review (Canadian Task Force Classification<br />

II-2).<br />

MATERIALS AND METHODS: We reviewed charts of women laparoscopic<br />

and robotic myomectomy by two expert surgeons from January<br />

2014 to December 2014. Each surgeon utilized exclusively one modality.<br />

Characteristics of patients with largest myoma size > 8cm and/or specimen<br />

weight > 250g and/or > 5 myomas removed were compared by surgical modality.<br />

RESULTS: The cohort consisted of 70 patients with 54% laparoscopic and<br />

46% robotic myomectomies. Aside from higher rates of Medicaid insurance<br />

in the robotic group (28% versus 8%, P¼.00<strong>17</strong>), there were no differences in<br />

demographics and medical/surgical history. The robotic group had significantly<br />

lower rates of concomitant operative hysteroscopy (32% versus 2%,<br />

P¼.014), mean (SD) specimen weights in grams (351.5 (4<strong>17</strong>.6) versus<br />

574.0 (525.5), P¼.014), and mean (SD) operative time in minutes (150<br />

(62.9) versus <strong>21</strong>6.7 (84.8), P¼.0006), but no difference in number of myomas<br />

removed (median (range) 4 (1-16) versus 4 (1-20), P¼.057) or size of dominant<br />

myoma (median (range) 8 (2-20 versus 10 (2-20)), P¼.18). There was no<br />

difference in estimated blood loss or rate of complications. The mean (SD)<br />

direct in dollars for robotic cases was lower (5861.3 (2273.9) versus<br />

7081.7 (3373.5), P¼.0402) but indirect costs did not significantly differ<br />

9833.5 (3749.7) versus 12048.6 (6396.3), P¼.058). Cost was significantly<br />

increased by operative time, specimen weight, estimated blood loss, prior<br />

laparotomy, complications, and length of stay.<br />

CONCLUSIONS: In this pilot study of expert surgeons using their<br />

preferred modality for complex myomectomy, outcomes and costs were<br />

comparable. Costs were most influenced by factors that increased surgical<br />

complexity and were contained in the robotics group. This preliminary<br />

data demonstrates a potential capacity for cost containment with the use of<br />

robotic surgery.<br />

P-207 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

WITHDRAWN<br />

e<strong>17</strong>6 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


P-209 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

REPRODUCTIVE OUTCOMES AFTER LAPAROSCOPIC MYO-<br />

MECTOMY (LM) AND LAPAROSCOPIC RADIOFREQUENCY<br />

VOLUMETRIC THERMAL ABLATION UTERINE<br />

LEIOMYOMAS. K. Isaacson. Gynecology, Harvard Medical School,<br />

Newton, MA.<br />

P-208 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

WITHDRAWN<br />

OBJECTIVE: To evaluate reproductive outcomes in women who were<br />

randomized to laparoscopic ultrasound-guided LM and RFVTA for symptomatic<br />

myomas.<br />

DESIGN: Single-center, prospective, randomized controlled trial.<br />

MATERIALS AND METHODS: Fifty women with self-reported heavy<br />

menstrual bleeding and/or bulk symptoms consented to intraoperative<br />

randomization (1:1) to LM or RFVTA for treatment of their symptomatic<br />

myomas. All participants desired uterine and reproductive conservation.<br />

RESULTS: Eight women ages 28 to 35 years (median, 33 years) conceived<br />

9 times between 3.2 to 23.5 months after excision or ablation of intramural<br />

and subserosal myomas. Of the RFVTA cohort (n ¼ 25), 3 women, ages<br />

31 to 35 years old, underwent ablation of 1 or 2 fibroids (intramural and/or<br />

subserosal) measuring 0.5 to 5.7 cm in greatest diameter. No intraoperative<br />

or postoperative complications were reported, and the patients conceived<br />

at 7 to 23.5 months post procedure. Two of the pregnancies resulted in<br />

full-term spontaneous vaginal births of healthy infants, and the third pregnancy<br />

is expected to result in a full-term vaginal birth in August <strong>2015</strong> for<br />

a pregnancy rate of 12% (3/25). Five women in the LM cohort (n ¼ 25)<br />

conceived 6 times for a pregnancy rate of 24% (6/25). One 28-year-old patient<br />

conceived at 3.2 months post myomectomy, but the pregnancy ended<br />

in a therapeutic abortion. She conceived again at 13.8 months post myomectomy<br />

and delivered a healthy infant by C-section. Three myomectomy patients<br />

conceived and delivered healthy infants by vaginal (n ¼ 2) and<br />

Cesarean delivery (n ¼ 1). The fifth myomectomy patient’s pregnancy is<br />

ongoing.<br />

CONCLUSIONS: Viable, full-term pregnancies culminating in vaginal<br />

births of healthy infants are possible after either RFVTA of symptomatic<br />

uterine myomas or laparoscopic myomectomy. Patients in this study will<br />

be followed for 5 years.<br />

P-<strong>21</strong>0 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

PERINATAL RISK IN MICROTUBAL REANASTOMOSIS VERSUS<br />

IN VITRO FERTILIZATION IN PATIENTS WITH TUBAL FACTOR<br />

INFERTILITY. S. A. Greenberg, a S. O’Gradney, a L. Davis, b<br />

E. A. Zbella, a M. Sanchez. a a Bayfront Health St Petersburg, St Petersburg,<br />

FL; b OBGYN, Bayfront Health St Petersburg, St Petersburg, FL.<br />

OBJECTIVE: To assess the perinatal outcome and evaluate the risk of preterm<br />

delivery, low birth weight, multiple gestation, spontaneous abortion,<br />

and ectopic pregnancy in women with tubal factor infertility who underwent<br />

either Microtubal Reanastomosis (MTR) versus In Vitro Fertilization (IVF).<br />

DESIGN: Retrospective Cohort.<br />

FERTILITY & STERILITY Ò<br />

e<strong>17</strong>7


MATERIALS AND METHODS: We analyzed the perinatal outcome of 74<br />

IVF and 39 MTR patients who had tubal factor infertility with documented<br />

fetal cardiac activity between 2009 and 2013. Preterm delivery, low birth<br />

weight, multiple gestation, spontaneous abortion and ectopic pregnancy<br />

were then analyzed between IVF and MTR. Preterm delivery and low birth<br />

weight were only compared amongst cohorts with singleton pregnancies.<br />

We performed multivariable analyses using Fisher’s exact and calculated<br />

P-value and adjusted risk ratio at a 95% confidence intervals.<br />

RESULTS: IVF had an increased risk of preterm delivery (27.7%<br />

compared with 3.33%, p-value 0.0153, adjusted RR 1.92, 95% CI), spontaneous<br />

abortion (24.3% compared with 7.63%, p-value .0485, adjusted RR<br />

1.4, 95% CI) and multiple gestation (35.7% compared with 3.22%, p-value<br />

.0067, adjusted RR 1.74, 95% CI). Although there was an increased risk of<br />

low birth weight, this was not statistically significant (22.2% compared<br />

with 6.66%, p-value .2863, adjusted RR 1.42, 95% CI). MTR had an<br />

increased risk of ectopic pregnancy (15.3% compared with 0%, p-value<br />

0.0031, adjusted RR 3.18, 95% CI).<br />

CONCLUSIONS: Patients who have undergone IVF have an increased<br />

risk for preterm birth, multiple gestation, and spontaneous abortion when<br />

compared to those who have undergone MTR for tubal factor infertility.<br />

P-<strong>21</strong>1 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

EFFECT OF HYSTEROSCOPIC SURGERY FOR UTERI WITH<br />

CAVITARY LESIONS ON CLINICAL PREGNANCY RATE IN<br />

WOMEN SCHEDULED FOR IVF. A. M. Abdelmagied, a,b<br />

M. A. Kamel, b A. M. Abuelhasan, b I. Elnashar, b A. A. Abdelaleem, b<br />

T. A. Farghaly, b A. A. Nassr, a,b M. H. Makarem. b a Obstetrics and Gynecology,<br />

Mayo Clinic, Rochester, MN; b Obstetrics and Gynecology, Women<br />

Health Hospital, Assiut University, Assiut, Egypt.<br />

OBJECTIVE: Adoption of routine office hysteroscopy (OH) before In-Vitro<br />

Fertilization (IVF) to diagnose and treat cavitary lesions (CL) is still<br />

debatable both from evidence and cost perspectives. Our objective was to<br />

evaluate the role of OH in detecting CL and the effect of correction of these<br />

lesions before IVF on clinical pregnancy rate (CPR).<br />

DESIGN: Prospective cohort study.<br />

MATERIALS AND METHODS: Women with primary infertility, before<br />

their enrollment in In-Vitro Fertilization (IVF), were examined by 2 diagnostic<br />

modalities to detect possible CL: trans-vaginal ultrasound (TVUS)<br />

and OH, as a gold standard,. All CL diagnosed by OH were treated before<br />

IVF. TVUS and OH findings, IVF related data and CPR were reported for<br />

all women. The t-test, Wilcoxon rank sum test, Chi-square test, Fisher exact<br />

test and logistic regression were used for comparisons.<br />

RESULTS: Sixty six women were included. In all, 20 women (30.3 %) had<br />

CL, 8 (40 %) of them were missed by TVU. The overall implantation rate<br />

(IR) and CPR were 16% and 36.4% respectively. Women with (n¼20) and<br />

without (n¼46) CL were similar in age, BMI, duration of infertility, antral follicle<br />

count, AMH, number of oocytes retrieved, number and quality of embryos transferred<br />

and IR. However, women with CL were more likely to have previous<br />

failed IVF attempts (55% vs. 26%, P¼0.02), and yet higher CPR (55% vs 28,<br />

P¼0.04) than women without CL. Univariate analysis was conducted for potential<br />

confounders that may influence CPR. CL treated uteri, adjusted for these confounders,<br />

were comparable to uteri with no CL in terms of CPR (aOR ¼ 1.79,<br />

95% CI¼0.52-6.<strong>17</strong>, P>0.05). A similar trend for CPR (58% vs 50%, P>0.05)<br />

was observed when comparing TVUS suspected CL group (n¼12) to missed<br />

CL group (n¼8). Of the overall CPR, 16.7% was attributed to the OH diagnosis<br />

and treatment of the missed lesions. Accordingly, 14 women with normal TVUS<br />

were needed to be screened by OH to achieve one clinical pregnancy.<br />

CONCLUSIONS: Cavitary lesions were common among women scheduled<br />

for IVF and a considerable percentage of these lesions were missed<br />

on TVUS. Women with treated CL and those with no CL had comparable<br />

CPR. The role of OH in recognizing TVUS missed CL should not be underestimated<br />

as considerable increase in CPR could be achieved by their correction.<br />

Supported by: Research support in university hospitals.<br />

P-<strong>21</strong>2 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE SURGICAL INDICATIONS FOR SEPTATE UTERUS IN INFER-<br />

TILE POPULATION BEFORE IN VITRO FERTILIZATION. Y. Yi, a<br />

X. Li, b Y. Ouyang, c G. Lu. a a Institute of Reproduction & Stem Cell Engineering,<br />

Central South University, Chang-sha, China; b Reproductive and Genetic<br />

hospital of CITIC-Xiangya, Chang-sha, China;<br />

c Central South<br />

University, Changsha, China.<br />

OBJECTIVE: To investigate whether the surgical indications are reasonable<br />

for infertile patients with septate uterus before IVF-ET (in vitro fertilization-embryo<br />

transfer) and whether there are differences about pregnancy<br />

outcome between infertile patients with septate uterus and infertile patients<br />

with normal uterus after IVF-ET.<br />

DESIGN: Retrospectively analyzed pregnancy outcome in patients with<br />

normal uterus and patients with septate uterus after IVF-ET.<br />

MATERIALS AND METHODS: 2637 infertile patients who got clinical<br />

pregnancy after IVF-ET from April 2012 to March 2014 were identified.<br />

There were <strong>21</strong>00 patients (A group) with normal uterus and 537 patients<br />

with septate uterus. Hysteroscopic metroplasty was performed before IVF-<br />

ET in 231 patients (B group) with septate uterus who met surgical indications<br />

that the depth of septum >10mm or depth of septum ranges from 5 to 10mm<br />

accompanied by history of unexplained recurrent miscarriage. 306 patients<br />

(C group) with septate uterus were managed expectantly because the depth<br />

of septum < 10mm and without history of unexplained recurrent miscarriage.<br />

Pregnancy outcome including miscarriage rate in 1st and 2nd trimester, preterm<br />

delivery rate, term delivery rate and ectopic pregnancy rate were<br />

compared among above 3 groups by Chi-Square Test. The difference was statistically<br />

significant when P < 0.05.<br />

RESULTS: Though the miscarriage rate in 1st trimester (14.1%) in C<br />

group (Managed expectantly) was higher than that in other two groups, the<br />

difference was not significant (P¼0.130). And miscarriage rate in 2nd<br />

trimester, preterm delivery rate, term delivery rate and ectopic pregnancy<br />

rate indicate no significant difference in these three groups (Table).<br />

Pregnancy outcome in 3 groups (P¼0.130).<br />

Pregnancy outcome Total<br />

A group (Normal) <strong>21</strong>00<br />

(100%)<br />

B group 231<br />

(Hysteroscopic (100%)<br />

metroplasty)<br />

C group (Managed<br />

expectanly)<br />

306<br />

(100%)<br />

Miscarriage<br />

in first<br />

trimester<br />

<strong>21</strong>4<br />

(10.2%)<br />

27<br />

(11.7%)<br />

43<br />

(14.1%)<br />

Miscarriage<br />

in second<br />

trimester<br />

91<br />

(4.3%)<br />

15<br />

(6.5%)<br />

12<br />

(3.9%)<br />

Preterm<br />

delivery<br />

Full term<br />

delivery<br />

276 1467<br />

(13.1%) (69.9%)<br />

37 150<br />

(16.0%) (64.9%)<br />

32 <strong>21</strong>2<br />

(10.5%) (69.3%)<br />

Ectopic<br />

pregnancy<br />

52<br />

(2.5%)<br />

2<br />

(0.9%)<br />

7<br />

(2.3%)<br />

CONCLUSIONS: For infertile patients with septate uterus, the indication<br />

of surgery that the depth of septum>10mm or depth of septum ranges<br />

from 5 to 10mm accompanied by history of unexplained recurrent miscarriage<br />

were reasonable before IVF-ET. According to above surgical<br />

indications, the infertile patients with septate uterus could get similar<br />

pregnancy outcome compared with infertile patients with normal uterus<br />

after IVE-ET.<br />

P-<strong>21</strong>3 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

SMALLER ENDOMETRIOTIC CYST REMOVAL ALONG WITH<br />

LONGER DURATION OF THE CYST HAS HIGHER IMPACT ON<br />

OVARIAN RESERVE-ASSESSMENT OF ANTIMULLERIAN<br />

HORMONE AND ANTRAL FOLLICULAR COUNTS, POST<br />

CYSTECTOMY - A PROSPECTIVE OBSERVATIONAL<br />

STUDY. N. K. Pratap, R. Sandya. Obstetrics and Gynecology, Kasturba<br />

Medical College, Manipal, India.<br />

OBJECTIVE: The damage inflicted to the ovary by cystectomy is a major<br />

concern in balancing the fertility risks and benefits. This study<br />

compared the cyst of less than 5 cms to more than 5 cms post laparoscopic<br />

cystectomy with Antimullerian hormone (AMH) and Antral Follicular<br />

Counts (AFC).<br />

DESIGN: Prospective observational study was done following endometrioma<br />

cystectomy and compared post operatively the cyst size with AMH<br />

and AFC for ovarian reserve after one month.<br />

MATERIALS AND METHODS: Group A less than 5 cms cyst and Group<br />

B with more than 5 cms cyst were compared. Statistical Mauchly’s test of<br />

sphericity was done. AMH was grouped as I, II and III depending on the<br />

values of less than 1,1-3.5, 3.5ng/ml and further analyzed comparing the<br />

cyst size by Pearson Chi square test.Ovarian sections were taken from the<br />

thickest areas of the cyst wall to analyze histologically the amount of primodial,<br />

primary and secondary follicles.<br />

RESULTS: There was an overall drop of AMH from 3.83.01ng/ml to<br />

2.671.92ng/ml (p¼ less than 0.001). AMH was clearly low in Group III<br />

following surgery (46.4% to 26.0%). AMH dropped from 4.53.4ng/ml to<br />

e<strong>17</strong>8 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


3.192.18 with less than 5 cms cyst (Group A), compared to 2.41.2 to<br />

1.70.85 with more than 5 cms cyst (Group B). A mean drop of 1.31 and<br />

0.7ng/ml in Group A and B respectively (p value less than 0.01) was seen.<br />

Overall AFC was 5.<strong>17</strong>1.44 and 3.611.61 pre and post operatively in<br />

the ovary operated. By the Statistical test - Paired t test it was significant.<br />

Drop in AFC was 1.2 and 2.2 in Group A and B. With unilateral cyst, 5 of<br />

the 15 in Group A had very low AMH post operatively (less than 1ng/ml)<br />

with the duration of the cyst being more than four years. Histopathological<br />

analysis of cyst wall based on semi quantitative scale which had grades 0-<br />

3 depending on the loss of follicles were analyzed. Irrespective of the loss<br />

of follicles or not, there was decline in both AMH and AFC. Of the total<br />

cyst wall analyzed 76% did not have loss of follicles. However, with more<br />

loss of follicles, there was more decline of AFC and AMH.<br />

CONCLUSIONS: Smaller the endometriotic cyst and longer the duration,<br />

more is the damage to the ovarian reserve. Hence, with smaller cysts meticulous<br />

surgical techniques are needed for better prognosis compared to the<br />

larger cyst.<br />

P-<strong>21</strong>4 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

PREDICTORS OF HEMORRHAGE AND TRANSFUSION FOR<br />

WOMEN UNDERGOING LAPAROSCOPIC AND ROBOTIC-ASSIS-<br />

TED MYOMECTOMY. M. V. Vargas. Department of Obstetrics and Gynecology,<br />

Division of Gynecology, George Washington University Medical<br />

Center, Washington, DC.<br />

OBJECTIVE: To identify predictors of hemorrhage and/or transfusion for<br />

women undergoing minimally invasive myomectomy.<br />

DESIGN: Retrospective chart review (Canadian Task Force Classification<br />

II-2).<br />

MATERIALS AND METHODS: We reviewed the charts of women who<br />

underwent minimally invasive myomectomy by three expert surgeons from<br />

April 2011 to December 2014. Characteristics of patients with estimated<br />

blood loss (EBL) > 1000mL and/or transfusion were compared to characteristics<br />

of all other patients in the cohort.<br />

RESULTS: The cohort consisted of 220 patients with 100 laparoscopic and<br />

120 robotic myomectomies. The median (range) number of myomas<br />

removed was 3 (1-20) and dominant myoma size in cm was 8 (4-20). The<br />

mean (SD) specimen weight in grams was 409 (396) and ranged from 4 to<br />

2894 grams. The mean (SD) operative room time (ORT) in minutes was<br />

190 (80) and EBL in mL was 352 (494). The complication rate was<br />

13.6%. Twenty-four patients (11%) experienced hemorrhage and/or transfusion,<br />

accounting for 80% of complications. These patients experienced<br />

longer ORT (mean in minutes (SD), 272 (100) versus 180 (73), P¼.0007)<br />

and length of stay (median days (range), 2 (0-6) versus 1 (0-3), P¼.0001),<br />

as well as higher rates of conversion (13% versus 0%, P¼0.0025) and ICU<br />

admission (13% versus 0%, P¼0.0025). The probability of hemorrhage/<br />

transfusion rose significantly with dominant myoma > 10cm (OR 2.83;<br />

95% CI 1.34 to 5.87), removal of > 5 myomas (OR 3.33; 95% CI 1.1 to<br />

10.5), transmural location 2.94 (95% CI 1.27 to 6.81), and specimen weight<br />

> 500g (OR 5.84; 95% CI 2.70 to 12.67).<br />

CONCLUSIONS: In this surgically complex cohort, minimally invasive<br />

myomectomy had a complication rate that was comparable to other reports<br />

of high volume surgeons. Hemorrhage and transfusion were the most common<br />

complications and were associated with case complexity as well as<br />

increased surgical morbidity. Anticipating these complications can facilitate<br />

preoperative planning, patient counseling, and the implementation of preventative<br />

interventions.<br />

P-<strong>21</strong>5 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

SHIFTING ANAEROBIC TO AEROBIC METABOLISM STIMU-<br />

LATES APOPTOSIS IN ADHESION FIBROBLASTS THROUGH<br />

THE MODULATION THE CELLULAR REDOX<br />

HOMEOSTASIS. N. M. Fletcher, a M. G. Saed, a B. R. Neubauer, a<br />

H. Abu-Soud, a A. O. Awonuga, b M. P. Diamond, c G. M. Saed. a a Obstetrics<br />

and Gynecology, Wayne State University, Detroit, MI; b Wayne State University,<br />

Southfield, MI; c Georgia Regents University, Augusta, GA.<br />

OBJECTIVE: To compare the effect of shifting anaerobic to aerobic metabolism<br />

on key regulators of oxidative stress, including extracellular superoxide<br />

dismutase (SOD3), inducible nitric oxide synthase (iNOS) and its<br />

product, nitric oxide (NO), as well as mitochondrial potential (Djm) and<br />

apoptosis in normal peritoneal and adhesion fibroblasts.<br />

DESIGN: Prospective experimental study.<br />

MATERIALS AND METHODS: The normal peritoneal and adhesion fibroblasts<br />

established from normal peritoneal and adhesion tissues from the<br />

same patient(s) were treated with dichloroacetate (DCA 0, 20, 40, 80 mg/<br />

ml, 24 hrs). Additionally, cells were treated with hypoxia (2% O2, 24 hrs)<br />

in the presence of DCA (80 mg/ml). Expression of SOD3 and iNOS and<br />

NO levels were determined by a combination of real-time RT-PCR and<br />

Greiss assay. The Djm was evaluated by the JC-1 Mitochondrial Membrane<br />

Potential Assay. Apoptosis was determined by caspases-3 activity and TU-<br />

NEL assays. Data were analyzed using SPSS 22.0 for Windows. Mixed<br />

model repeated measures ANOVA was used with treatment as the within factor<br />

and cell type as the between factor. Paired comparisons with a Bonferroni<br />

correction were used to compare pairs of treatments. Significant interactions<br />

between treatment and cell type were analyzed with independent sample t-<br />

tests by cell type on each treatment.<br />

RESULTS: Levels of iNOS, SOD3, NO (nitrate/nitrite), and caspase-3 activity<br />

are summarized in Table 1. There was a marked increase in TUNEL<br />

staining in both normal peritoneal and adhesion fibroblasts in response to<br />

DCA (80 mg/ml). There was also enhanced Djm in adhesion fibroblasts,<br />

indicative of high oxidative stress, as compared to normal peritoneal fibroblasts.<br />

Further enhancing oxidative stress by hypoxia treatment markedly<br />

increased Djm in normal peritoneal to levels observed in adhesion fibroblasts.<br />

Treatment with DCA (80 mg/ml) was found to protect against the effects<br />

of hypoxia.<br />

CONCLUSIONS: DCA modulates the redox homeostasis protecting cells<br />

against anaerobic metabolism and the associated oxidative stress. These findings<br />

may provide targets for therapeutic intervention for the reduction of profibrotic<br />

disorders, including postoperative adhesions.<br />

Levels of iNOS, SOD3, NO, and activity of caspase-3 in normal peritoneal and<br />

adhesion fibroblasts.<br />

iNOS<br />

(fg/mg RNA)<br />

SOD3<br />

(fg/mg RNA)<br />

Nitrate/<br />

Nitrite (mM)<br />

Caspase-3<br />

Activity (mM)<br />

Normal fibroblasts - 1.8 0.3 81.4 9.0 7.0 0.4 12.6 0.6<br />

Untreated<br />

Normal fibroblasts - 2.3 0.6 111.7 2.6* 7.0 0.1 15.5 0.9*<br />

20 mg/ml DCA<br />

Normal fibroblasts - 4.5 0.8* 97.9 2.1 9.4 0.04* 18.3 1.2*<br />

40 mg/ml DCA<br />

Normal fibroblasts - 16.8 0.1* 91.8 5.8 10.3 0.5* <strong>21</strong>.6 0.7*<br />

80 mg/ml DCA<br />

Adhesion Fibroblasts - 2.0 0.2 27.9 0.6* 4.4 0.5** 8.1 0.8*<br />

Untreated<br />

Adhesion Fibroblasts - 2.4 0.3 116.2 4.8** 4.2 0.1 11.5 0.6**<br />

20 mg/ml DCA<br />

Adhesion Fibroblasts - 4.4 0.5** 65.9 2.3** 5.7 0.4** <strong>17</strong>.5 0.8**<br />

40 mg/ml DCA<br />

Adhesion Fibroblasts - <strong>21</strong>.1 0.3** 82.9 0.2** 13.2 0.4** <strong>21</strong>.4 1.1**<br />

80 mg/ml DCA<br />

*p<br />

P-<strong>21</strong>6 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

30-DAY POSTOPERATIVE OUTCOMES OF MINIMALLY INVA-<br />

SIVE VERSUS ABDOMINAL MYOMECTOMY: AN ANALYSIS OF<br />

THE NATIONAL SURGICAL QUALITY IMPROVEMENT PRO-<br />

GRAM (NSQIP) DATABASE 2005-2013. S. Moustafa, a C. M. Duke, b<br />

A. Sood, c E. C. Dun, d A. Dabaja. c a Department of Obstetrics, Gynecology,<br />

& Reproductive Sciences, Yale School of Medicine, New Haven, CT; b Yale<br />

University School of Medicine, New Haven, CT; c Vattikuti Urology Institute,<br />

Henry Ford Health System, Southfield, MI; d Yale School of Medicine, New<br />

Haven, CT.<br />

OBJECTIVE: To compare complication/adverse event rates between<br />

minimally invasive (MI; laparoscopic and robotic-assisted) myomectomy<br />

and abdominal myomectomy. Further, to assess the independent effect of<br />

MI myomectomy on complications/adverse events in these patients.<br />

DESIGN: Retrospective cohort study using the National Surgical Quality<br />

Improvement Project Database (NSQIP).<br />

MATERIALS AND METHODS: Patients undergoing either MI or abdominal<br />

myomectomy, from 2005-2013, were identified using their CPT codes<br />

[58140, 58146, 58545, 58546]. The primary endpoint was postoperative<br />

complications. Secondary endpoints included operative time, length of stay<br />

FERTILITY & STERILITY Ò<br />

e<strong>17</strong>9


(LOS), re-intervention, and readmission rates. Multivariable logistic regression<br />

models assessed the independent effect of MI on postoperative outcomes.<br />

RESULTS: 5338 patients were identified: 2261 (42.4%) underwent MI<br />

myomectomy and 3074 (57.6%) underwent abdominal myomectomy. The<br />

mean age of the patients in the MI group and abdominal myomectomy group<br />

were 38.4 (SD¼7.3) years and 37.4 (SD¼6.4) years (p


References:<br />

1. Committee on Adolescent Health C. Committee opinion: no. 562:<br />

mullerian agenesis: diagnosis, management, and treatment. Obstetrics<br />

and gynecology. May 2013;1<strong>21</strong>(5):1134-1137.<br />

2. Smith NA, Laufer MR. Obstructed hemivagina and ipsilateral renal<br />

anomaly (OHVIRA) syndrome: management and follow-up. Fertility<br />

and sterility. Apr 2007;87(4):918-922.<br />

3. Cooper AR, Merritt DF. Novel use of a tracheobronchial stent in a patient<br />

with uterine didelphys and obstructed hemivagina. Fertility and<br />

sterility. Feb 2010;93(3):900-903.<br />

P-<strong>21</strong>9 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ASSOCIATION BETWEEN BODY MASS INDEX (BMI), UTERINE<br />

SIZE, AND OPERATIVE MORBIDITY IN WOMEN UNDERGOING<br />

MINIMALLY INVASIVE HYSTERECTOMY. D. K. Shah, a B. Van<br />

Voorhis, a A. Vitonis, b S. A. Missmer. b,c,d a Division of Reproductive Endocrinology<br />

and Infertility, University of Iowa, Iowa City, IA; b Department of<br />

Obstetrics, Gynecology, and Reproductive Biology, Brigham and Women’s<br />

Hospital, Boston, MA; c Harvard Medical School, Boston, MA; d Harvard<br />

School of Public Health, Boston, MA.<br />

OBJECTIVE: To examine the association of surgical approach and operative<br />

morbidity after minimally invasive hysterectomy and determine<br />

whether the association varies with uterine size and patient BMI.<br />

DESIGN: Retrospective cohort.<br />

MATERIALS AND METHODS: Data abstracted from the American College<br />

of Surgeons National Safety and Quality Improvement Project registry<br />

included 36,855 women who underwent vaginal (VH), laparoscopic-assisted<br />

vaginal (LAVH), or total laparoscopic (TLH) hysterectomy between January<br />

2005-December 2012. Associations between surgical approach, BMI, and<br />

operative morbidity were examined, stratifying based on uterine size and adjusting<br />

for age, race, ethnicity, year of surgery, smoking, diabetes, and American<br />

Society for Anesthesiology physical classification. Adjusted means, rate<br />

ratios, or odds ratios (OR) with 95% confidence intervals (CI) were calculated<br />

using linear, Poisson, or logistic regression.<br />

RESULTS: Operative times were uniformly shorter in women undergoing<br />

VH as compared to LAVH or TLH (p 40 kg/m 2 had 76% lower<br />

odds of blood transfusion (CI¼0.1-0.6; unadjusted rates 0.8% vs. 2.2%)<br />

and 39% lower odds of surgical site infections (CI 0.4-1.0; unadjusted rates<br />

2.4% vs. 3.9%) after TLH as compared to VH. The association between<br />

wound infection and BMI was attenuated or absent in women with large uteri<br />

or in those that underwent LAVH.<br />

CONCLUSIONS: When comparing minimally invasive approaches to<br />

hysterectomy, operative times are shortest with VH, particularly in obese<br />

women with small uteri. Rates of blood transfusion and wound infection<br />

are lowest after TLH but absolute risks are low regardless of surgical<br />

approach. Laparoscopic assisted vaginal hysterectomy appears to confer no<br />

specific advantage over VH or TLH.<br />

Supported by: Departmental funding, University of Iowa.<br />

P-220 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ART - CLINICAL<br />

A RANDOMIZED CONTROLLED TRIAL OF ORAL ACETAMINO-<br />

PHEN FOR ANALGESIC CONTROL AFTER TRANSVAGINAL<br />

OOCYTE RETRIEVAL. S. Kassira, a G. Bates, b M. Powell, c<br />

J. M. Bouknight, b M. McLean. b a School of Medicine, UAB, Birmingham,<br />

AL; b Division of Reproductive Endocrinology and Infertility, UAB, Birmingham,<br />

AL; c Dept of Anesthesiology, UAB, Birmingham, AL.<br />

OBJECTIVE: To evaluate the efficacy of 1000mg of oral acetaminophen<br />

onpost-procedure pain after transvaginal cyst puncture (TVCP).<br />

DESIGN: Double-blind randomized controlled trial of acetaminophen vs<br />

placebo in women undergoing IVF at an academic infertility clinic. Women<br />

with an allergy to acetaminophen, narcotic dependence or use of acetaminophen<br />

within 24 hours were excluded.<br />

MATERIALS AND METHODS: All eligible patients presenting for<br />

TVCP were approached for enrollment. Subjects were randomized at the<br />

time of TVCP to acetaminophen vs placebo using a permuted block<br />

schedule. Pills were taken under supervision of study personnel 1 hour prior<br />

to TVCP. Demographics, IVF cycle characteristics and a baseline pain<br />

score using the visual analogue scale (VAS) were collected. Conscious<br />

sedation was administered by anesthesiology per routine. Procedure time,<br />

intraoperative medications, and number of oocytes retrieved were recorded.<br />

PACU pain scores (15, 30, 45, 60 minutes) and pain medication use in recovery<br />

was collected by the nurse. At 24 hours, post-procedure pain scores<br />

and use of pain medication was assessed. Total pain score was a sum of<br />

PACU pain scores and 24 hour pain score. Wilcoxon rank-sum test was<br />

used to compare pain scores between groups. Chi squared and t-tests<br />

were used to assess factors associated with pain scores. For multivariable<br />

analysis ‘‘moderate pain’’ was considered a cumulative pain score >8 for<br />

PACU pain scores and >10 for 24-hr pain scores. Logistic regression was<br />

used to compare the incidence of moderate pain while controlling for confounders.<br />

Planned enrollment of 100 subjects will provide 80% power to<br />

detect a 2-point difference in pain on the VAS.<br />

RESULTS: To date 32 subjects have been enrolled with mean age<br />

32.8+/- 5.0 years. Mean duration of oocyte retrieval was 23 +/- 10min.<br />

9 subjects (28%) required pain medication in the PACU. 23 subjects<br />

(72%) used an opioid in the 24 hours post-procedure. 16 subjects<br />

(50%) did not undergo fresh embryo transfer, including 5 oocyte donors.<br />

There was no difference in the total pain score (median[IQR]) between<br />

placebo and acetaminophen groups (5.0 [1-13] vs 5.5 [0-10.5] p ¼0.7)<br />

nor in PACU pain scores (6.5 [1.5-12] vs 7.0 [4-15] p ¼0.7). There was<br />

no difference in the use of pain medication post procedure. Baseline<br />

pain score, BMI, or diagnosis of endometriosis was not associated with<br />

pain scores. Oocyte donors had similar pain scores to those undergoing<br />

autologous IVF (8.0 vs 7.0; p ¼ 0.87). Procedure time and ibuprofen<br />

use prior to TVCP was associated with moderate pain scores. There<br />

was no difference in PACU or total pain scores between groups after controlling<br />

for these factors(p¼0.89; 0.55).<br />

CONCLUSIONS: Preoperative administration of oral acetaminophen<br />

does not reduce PACU or 24-hour cumulative pain scores after TVCP. Procedure<br />

time and ibuprofen use prior to TVCP are associated with post-operative<br />

pain after TVCP.<br />

P-2<strong>21</strong> Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

BLASTOCYST VERSUS CLEAVAGE-STAGE ELECTIVE SINGLE<br />

EMBRYO TRANSFER - A COMPARATIVE RETROSPECTIVE<br />

STUDY. Z. Veleva, a S. Boulet, b S. Makinen, c H. Martikainen, d<br />

A. Tiitinen, a J. Tapanainen, a D. M. Kissin. b a Ob/Gyn, University of Helsinki<br />

and Helsinki University Central Hospital, Helsinki, Finland; b Centers for<br />

Disease Control and Prevention, Atlanta, GA; c Family Federation of Finland,<br />

Helsinki and Oulu Infertility Clinics, Helsinki, Finland; d Ob/Gyn, University<br />

of Oulu and Oulu University Hospital, Oulu, Finland.<br />

OBJECTIVE: To compare cumulative term live birth rates (ctLBR) after<br />

elective single embryo transfer (eSET), using data from two countries –<br />

Finland where cleavage stage (day 2/3) eSET is almost exclusively used,<br />

and the U.S. where blastocyst (day 5/6) eSET is more common.<br />

DESIGN: Retrospective cohort study comparing data from Finland’s<br />

largest infertility clinics (LUMI database, cleavage stage eSET performed<br />

in 50.5% of all fresh cycles) with data from the U.S. National Assisted Reproductive<br />

Technology (ART) Surveillance System (blastocyst eSET performed<br />

in 3.9% of all fresh cycles).<br />

MATERIALS AND METHODS: The analysis included all fresh, nondonor<br />

eSET (n¼16,276) and subsequent frozen embryo transfer cycles<br />

(n¼11,625) performed during 2008-2012 in the U.S. and Finland among<br />

women aged %40 years with no prior use of ART. U.S. clinics were classified<br />

into those with high (R52 cases/year) and low (


OR 1.91, 95% CI 1.68-2.16). Gonadotropin dose per oocyte was higher in<br />

cleavage stage Finnish cycles (184.3141.0 IU), compared with U.S. blastocyst<br />

cycles (low eSET use: 154.<strong>21</strong>29.5, P


We observed that the overweight sample showed despite their efforts, an increase<br />

in weight compared to their reported weight one year earlier (mean ¼<br />

0.3kg SD 8.75 kg).<br />

CONCLUSIONS: The majority of the women had made one or more<br />

health promoting changes, but a considerable proportion of patients kept lifestyle<br />

behaviors not optimal for conception, such as obesity and use of tobacco.<br />

Therefore women, who have problems in conceiving, should be<br />

encouraged to seek help earlier than did women in this sample. Fertility<br />

care should include, as a routine, both assessments for lifestyle factors and<br />

interventions to lifestyle changes.<br />

Reference: Key Words: infertility, lifestyle, obesity, alcohol, diet.<br />

Supported by: Funding of the study was received from the Foundation<br />

Family Planning Fund in Uppsala, the College of Medicine at Uppsala University,<br />

Sweden, V€astmanland county council, Sweden and the Uppsala-<br />

€Orebro Regional Research Council, Sweden. None of the authors have any<br />

conflict of interest to declare.<br />

P-225 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

PROGESTERONE LEVEL ON THE DAY OF HCG TRIGGERING: A<br />

NEW TOOL FOR A SUCCESSFUL MODIFIED NATURAL IN VITRO<br />

FERTILIZATION CYCLE? C. Kamga-Ngande, a,b I. Kadoch, a,b<br />

L. Lapensee. a,b a Reproductive Endocrinology and Infertility, Centre Hospitalier<br />

Universitaire de Montreal, Montreal, QC, Canada; b Ovo Fertility<br />

Clinic, Montreal, QC, Canada.<br />

OBJECTIVE: Premature progesterone elevation before hCG triggering<br />

is known to occur occasionally in stimulated IVF although causes for<br />

this elevation are still unclear. The negative effect of this rise on the<br />

endometrium therefore affecting IVF outcomes is acknowledged. However,<br />

this is not documented in mnIVF condition. The aim of this study<br />

is to determine whether PG levels on the day of hCG could predict no<br />

oocyte at egg collection and both biochemical and clinical pregnancy<br />

rates (PR).<br />

DESIGN: We conducted a quality control (from 01-01-2012 to 30-06-<br />

2013) involving 581 women and 1074 mnIVF cycles.<br />

MATERIALS AND METHODS: Progesterone (PG) and Estradiol (E2)<br />

values on the day of hCG triggering were available. Multivariate analyses<br />

were performed tacking into account diagnosis, days of GnRH antagonist,<br />

antimullerian hormone (AMH) level, and age >35 years.<br />

RESULTS: Female were aged 32.43.5years,withAMH,E2andprogesterone<br />

levels at 1.71.9 ng/mL, 916303 pmol/L, and 1.6 0.8 nmol/<br />

L, respectively. The ‘‘no oocyte retrieved at egg collection per cycle’’ rate<br />

was 9.5%. Per egg collection, 94% of oocytes were mature. Biochemical<br />

and clinical PR were respectively <strong>17</strong>.8% and 16.2% per cycle and 35.9%<br />

and 32.8% per embryo transfer. Associations were found between PG<br />

level and no oocyte retrieved at egg collection (OR ¼ 1.61 (1.2-2.16);<br />

p¼0.001), biochemical PR (OR ¼ 0.62 (0.41-0.93); p¼0.02), and clinical<br />

PR (OR ¼ 0.63 (0.42-0.95); p¼0.02). Each increase in PG significantly<br />

increased the risk of not finding oocytes at egg collection by 61%, and<br />

decreased chance of biochemical and clinical PR by 38% and 37%,<br />

respectively.<br />

CONCLUSIONS: PG appears to have a detrimental effect on mnIVF<br />

outcomes. It has been proposed that PG increase in the late follicular<br />

phase influences endometrial maturation. This may lead to decreased<br />

endometrial receptivity and asynchrony between endometrium and the<br />

embryo. More research is needed to determine PG threshold allowing<br />

clinical decisions: cancel the cycle or freeze embryo and transfer in<br />

another cycle.<br />

P-226 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

CLINICAL OUTCOMES OF INTRAFALLOPIAN TRANSFER OF<br />

CRYOPRESERVED-THAWED EMBRYOS IN A NATURAL MEN-<br />

STRUAL CYCLE IN OLDER FEMALES. A. Tanaka, M. Nagayoshi,<br />

I. Tanaka, S. Ikuma, T. Miki, T. Yamaguchi. Saint Mother Hospital, Kitakyusyu,<br />

Japan.<br />

OBJECTIVE: Assisted reproductive technology (ART) in females aged 40<br />

years or older is characterized by low pregnancy and high miscarriage rates.<br />

The impaired quality of the oocytes is thought to be the main cause, but the<br />

environment of the in vitro culture may be another factor. Considering this,<br />

we have carried out thawed embryos intrafallopian transfer (EIFT) in a natural<br />

cycle to expect embryogenesis under the most natural environment. In<br />

this study, we aimed to investigate whether intrafallopian transfer of thawed<br />

embryos in a natural menstrual cycle would improve the implantation rate in<br />

older females.<br />

DESIGN: Retrospective cohort study of intrafallopian transfer of cryopreserved-thawed<br />

embryos in a natural menstrual cycle in older females.<br />

MATERIALS AND METHODS: The subjects were females aged 40 years<br />

or older that had received in vitro fertilization (IVF) and embryo transfer<br />

(ET), three or more times unsuccessfully, but had no passage obstruction<br />

in fallopian tubes. When there were 3 or more antral follicles on day 3 of<br />

the menstrual cycle, a GnRH agonist (the short method) was given. Meanwhile,<br />

when there were 2 or less, the regimen of clomiphene citrate (SerofeneÒ)<br />

plus recFSH every other day was employed. A 4 - 6-cell embryo 2<br />

days after egg collection was frozen by the slow freezing method (1.8M<br />

ethylene glycol + 0.2M sucrose + 20% fetal bovine serum-supplemented<br />

D-PBS). In the group with a regular menstrual cycle, natural ovulation was<br />

first confirmed. Then, the frozen embryo was thawed two days after the<br />

ovulation and transferred into the fallopian tube laparoscopically under the<br />

intravenous anesthesia with Propofol and Ketalar. In principle, the fallopian<br />

tube on the ovulation side was selected for the transfer. For supporting luteal<br />

function, a progesterone vaginal suppository (1000mg) was administered for<br />

two weeks.<br />

RESULTS: ET was carried out in 195 cycles from January 2010 to<br />

November 2013. 30 females became pregnant (pregnancy rate per cycle,<br />

15.4%) and 3 had a miscarriage (miscarriage rate, 10.0%). As a control,<br />

the pregnancy rate in the group of over 40 years and miscarriage rate by intracytoplasmic<br />

sperm injection and ETwere 13.0% (198/1527) and 60.6% (120/<br />

198), respectively.<br />

CONCLUSIONS: In older females cases when in vitro fertilization repeatedly<br />

failed or had miscarriages, the results showed that intrafallopian transfer<br />

of embryos in a natural menstrual cycle was useful.<br />

P-227 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

EVALUATING OOCYTE YIELD, QUALITY & FERTILIZATION<br />

FOLLOWING A CHANGE IN ANESTHESIA PROTOCOL. E. Li, a<br />

B. M. Hanson, a J. Gordon, b M. DiMattina, b G. Celia, b M. Payson. b a Inova<br />

Fairfax Hospital, Falls Church, VA; b Dominion Fertility, Arlington, VA.<br />

OBJECTIVE: Compare oocyte yield, quality, and fertilization rates between<br />

two anesthesia protocols (Demerol/Versed vs. Propofol) for transvaginal<br />

oocyte retrieval in stimulated IVF (SIVF) and natural cycle IVF (NCIVF).<br />

Previous studies have shown that Propofol accumulates in follicular fluid and<br />

may impact fertilization rates (Piroli 2012, Christiaens 1999).<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: SIVF and NCIVF patients at a private<br />

fertility clinic who underwent transvaginal oocyte retrieval with Demerol/<br />

Versed from 2012 to 2013 were compared to a similar cohort of patients<br />

from 2013 to 2014 who had the same procedure using Propofol. Number<br />

of oocytes retrieved, percent of MII oocytes, and fertilization rates were<br />

compared.<br />

RESULTS: <strong>21</strong>6 SIVF transvaginal oocyte retrievals were performed under<br />

conscious sedation with Versed/Demerol. 2,677 total oocytes were retrieved<br />

in this group. After a change in anesthesia protocol, 259 SIVF patients underwent<br />

the same procedure with general anesthesia using Propofol, yielding<br />

3,053 oocytes. No significant difference was noted in number of oocytes<br />

retrieved per patient (12.39 in the Versed/Demerol group; 11.79 in the Propofol<br />

group). A similar percentage of oocytes were noted to be in the MII stage<br />

in both anesthesia groups (78.1% in the Versed/Demerol group; 79.2% in the<br />

Propofol group). The number of oocytes transferred per patient was 1.7 in<br />

both groups. Fertilization rates were similar in both protocols (74.6% in<br />

the Versed/Demerol group; 77.1% in the Propofol group). 543 NCIVF patients<br />

underwent transvaginal oocyte retrieval with Versed/Demerol. 488 total<br />

oocytes were retrieved. 549 NCIVF patients underwent the same<br />

procedure with Propofol, resulting in 477 oocytes. Similar to findings in<br />

the SIVF cohort, no significant difference was noted in number of oocytes<br />

retrieved per patient (0.90 in the Versed/Demerol group; 0.87 in the Propofol<br />

group). A similar percentage of oocytes were noted to be in the MII stage in<br />

both groups as well (82.2% in the Versed/Demerol group; 83.9% in the Propofol<br />

group). Number of oocytes transferred per patient was 1.0 in both<br />

groups. Fertilization rates were similar in both protocols (84.5% in the<br />

Versed/Demerol group; 84.8% in the Propofol group).<br />

FERTILITY & STERILITY Ò<br />

e183


CONCLUSIONS: There was no significant difference observed in number<br />

or quality of oocytes retrieved after the use of Versed/Demerol versus Propofol.<br />

There was also no difference noted in fertilization rates. The use of Propofol<br />

was not found to be detrimental to oocyte quality when compared to the<br />

use of Versed/Demerol for retrieval anesthesia.<br />

References:<br />

1. Piroli A, Marci R, Marinangeli F, et al. Comparison of different anaesthetic<br />

methodologies for sedation during in vitro fertilization procedures:<br />

effects on patient physiology and oocyte competence. Gynecol<br />

Endocrinol. 2012;28(10):796-9.<br />

2. Christiaens F, Janssenswillen C, Verborgh C, et al. Propofol concentrations<br />

in follicular fluid during general anaesthesia for transvaginal<br />

oocyte retrieval. Hum Reprod. 1999;14(2):345-8.<br />

P-228 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ENDOMETRIAL BIOPSY PRIOR TO IVF STIMULATION SHOWS<br />

IMPROVED RATES OF IMPLANTATION. H. Ahmed. Princeton<br />

IVF, Lawrenceville, NJ.<br />

OBJECTIVE: To evaluate the benefit of endometrial injury in an IVF cycle<br />

and to study in a set patient population the implantation rates, biochemical<br />

pregnancy rates, first trimester pregnancy loss, second/third trimester loss,<br />

and live birth rates.<br />

DESIGN: Retrospective data analysis.<br />

MATERIALS AND METHODS: Local endometrial injury, by way of<br />

endometrial biopsy prior to initiation of gonadotropins for controlled ovarian<br />

hyperstimulation for IVF in 120 patients that were randomly selected from<br />

234 patients.<br />

RESULTS: Local endometrial injury in cycles with the transfer of either 2-<br />

3 embryos (experimental group ¼ 120 and control group ¼ 114) resulted in a<br />

significantly increased implantation rate (53.3% vs. 38.6%, p 0.025) and clinical<br />

pregnancy rate (43.3% vs. 29.0%, p 0.022) compared to controls. There<br />

was no statistical significance found in any of the negative outcomes which<br />

included overall pregnancy loss, overall spontaneous miscarriage, biochemical<br />

pregnancy, first trimester pregnancy loss, second/third trimester pregnancy<br />

loss.<br />

CONCLUSIONS: There is a clinically relevant improvement in implantation<br />

and clinical pregnancy rates when local endometrial injury by way of<br />

endometrial biopsy was performed prior to COH for IVF-ET, although this<br />

did not result in a significant increase in live birth rates. There was also no<br />

significant increase found in the rates of spontaneous miscarriage after local<br />

injury to the endometrium by way of endometrial biopsy prior to COH for<br />

IVF-ET. Further investigation is needed to determine what patients may<br />

best benefit from this treatment modality.<br />

P-229 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

IS LUTEAL PHASE SUPPORT FOR IUI CYCLES WITH GONADO-<br />

TROPIN REALLY NECESSARY TO IMPROVE<br />

OUTCOME? N. Karaca, a P. Ozcan, b E. E. Kovalak, c G. Batmaz, d<br />

S. Ates. e a Bezmi Alem University Hospital, Istanbul, Turkey; b Yeditepe University<br />

Hospital, Istanbul, Turkey; c Bagcilar Research and Training Hospital<br />

Obstetrics, Istanbul, Turkey; d MD, _Istanbul, Turkey; e Bezmialem Vakif University,<br />

Faculty of Medicine, Istanbul, Turkey.<br />

OBJECTIVE: to evaluate the effectiveness of vaginal micronized progesterone<br />

for luteal phase support in improving clinical PRs in IUI cycles with<br />

OS using gonadotropin and to compare the results of progesterone supplementation<br />

between recombinant follicle stimulating hormone (rFSH) and human<br />

menopausal gonadotropin (hMG) for IUI cycles.<br />

DESIGN: retrospective case-control study.<br />

MATERIALS AND METHODS: A total of 579 women underwent OS<br />

with gonadotropins who had admitted at IVF Center and Training and<br />

Research Hospital between January 2011 and December 2014 were included<br />

in this retrospective case-control study. Of 579 patients, 284 cycles receive<br />

luteal phase support in the form of vaginal micronized progesterone capsules<br />

(progesterone group), or not to receive luteal support (no progesterone<br />

group).<br />

RESULTS: Of IUI cycles, 132 (22,7%) resulted in a clinical pregnancy.<br />

Between no progesterone group (n¼59 (20%) and progesterone group<br />

(n¼73 (25.7%)) there was no significant difference in terms of clinical<br />

PRs (p ¼ 0.055). Clinical PRs were similar to in women underwent OS<br />

with rFSH or hMG (p ¼ 0.41 and 0,06, respectively) in progesterone group<br />

and no progesterone group .<br />

CONCLUSIONS: The results of study do not support to efficacy of routine<br />

supplementation of the luteal phase with vaginal progesterone in IUI cycles<br />

with OS to improve PRs. Further randomized trials are needed to evaluate if<br />

luteal support is necessary in selected patients.<br />

P-230 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

DOES FOLLICLE FLUSHING DURING OOCYTE RETRIEVAL<br />

IMPROVE FERTILITY PRESERVATION CYCLE<br />

OUTCOMES? V. Emirdar, a V. Turan, a F. Pacheco, a K. H. Oktay. a,b<br />

a New York Medical College, Valhalla, NY; b Innovation Institute for Fertility<br />

Preservation, New York, NY.<br />

OBJECTIVE: Given that cancer patients are at risk for losing their ovarian<br />

reserve post-chemotherapy, maximizing oocyte recovery is crucial in fertility<br />

preservation (FP) cycles. Our aim was to determine the usefullness of follicular<br />

flushing in maximization of oocyte recovery and embryo yield in women<br />

undergoing controlled ovarian stimulation to preserve oocytes or embryos<br />

with letrozole plus gonadotropins (LG) for FP.<br />

DESIGN: Retrospective controlled cohort study of women with breast<br />

cancer who underwent fertility preservation.<br />

MATERIALS AND METHODS: Eleven cycles where oocytes were aspirated<br />

using a double lumen flushing approach (DLF) were compared to 31<br />

age-matched cycles where oocytes were retrieved via single lumen aspiration<br />

(SLA) . All follicles that were visible by ultrasound, including those that are<br />

10-mm on the pre-trigger ultrasound,<br />

a similar number of follicles was detected (DLF vs. SLA: 16.<strong>21</strong>.2<br />

vs. 12.51.3, p¼0.16). A significantly higher total number of oocytes (<strong>17</strong>.4<br />

2.3 vs. 10.9 1.2, respectively; p ¼ 0.015) and a significantly higher number<br />

of immature oocytes were recovered with DLF (8.36 1.12 vs. 3.53 <br />

0.45, p < 0.001). However the difference in the number of cryopreserved embryos<br />

did not reach statistical significance (DLF vs. SLA: 7.2 1.94 vs. 5.2<br />

0.79, p ¼ 0.279).<br />

CONCLUSIONS: Utilization of DLF in FP cycles may facilitate the recovery<br />

of immature oocytes from small follicles and increase the initial<br />

oocyte yield. The value of DLF in enhancing final embryo yield should be<br />

studied in larger prospective studies among women undergoing fertility preservation.<br />

Cycle characteristics and fertility preservation outcomes.<br />

DLF (n¼11) SLA (n¼31) P value<br />

Age (years) 39.8 0.48 37.9 0.74 0.163<br />

AMH (ng/ml) 1.6 0.37 1.4 0.25 0.635<br />

Total FSH dose (IU) 5345 344 50<strong>17</strong> 108 0.236<br />

Total Letrozole dose (mg) 59.09 2.84 57.90 1.08 0.633<br />

Peak E2 level (pg/mL) 1022.46 205.45715.51 33.69 0.026<br />

Number of follicles on 16.2 1.2 12.5 1.4 0.166<br />

trigger day<br />

No. of oocytes retrieved <strong>17</strong>.4 2.34 10.9 1.26 0.015<br />

No. of immature oocytes 8.36 1.12 3.53 0.45 p < 0.001<br />

No. of embryos frozen 7.2 1.94 5.2 0.79 0.279<br />

References:<br />

1. Mok-Lin E, Brauer AA, Schattman G, Zaninovic N, Rosenwaks Z,<br />

Spandorfer S.Follicular flushing and in vitro fertilization outcomes in<br />

the poorest responders: a randomized controlled trial.Hum Reprod.<br />

2013;28:2990-5.<br />

e184 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


2. Levy G, Hill MJ, Ramirez CI, Correa L, Ryan ME, DeCherney AH,<br />

Levens ED, Whitcomb BW.The use of follicle flushing during oocyte<br />

retrieval in assisted reproductive technologies: a systematic review<br />

and meta-analysis. Hum Reprod. 2012;27:2373-9.<br />

P-231 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

CERVICAL MUCUS SCORE AT THE TIME OF INTRAUTERINE<br />

INSEMINATION AND ITS ASSOCIATION WITH CONCEPTION<br />

PROBABILITY. E. A. Evans-Hoeker, a,b A. Liberty, c A. Z. Steiner. b a Reproductive<br />

Medicine and Fertility, Carilion Clinic, Virginia Tech Carilion<br />

School of Medicine, Roanoke, VA; b Obstetrics and Gynecology, University<br />

of North Carolina, Raleigh, NC; c University of North Carolina School of<br />

Medicine, Chapel Hill, NC.<br />

OBJECTIVE: Studies demonstrate that cervical mucus scores at time of<br />

ovulation predict fecundability among couples attempting to conceive naturally.<br />

It is unknown if this is due to the cervical mucus (CM) itself or associated<br />

factors. The aims of this study were to determine 1) the prevalence of<br />

fertile type CM, 2) factors associated with fertile type CM, and 3) whether<br />

fertile type CM predicts pregnancy following intrauterine insemination (IUI).<br />

DESIGN: Clinical cohort study.<br />

MATERIALS AND METHODS: All clinical providers were trained on<br />

CM scoring. Providers recorded the CM score at time of insemination. CM<br />

was classified as absent, type 3 or type 4. Fertile type CM was defined as<br />

types 3 and 4. This analysis includes all cycles with CM type documented<br />

at the time of IUI performed at our academic center between 8/2012-7/<br />

2013. Data was collected regarding patient demographics, medical history,<br />

treatment types and outcome, and were analyzed using regression models<br />

with a cluster term.<br />

RESULTS: One hundred and seventy nine women contributed 3<strong>21</strong> IUI cycles<br />

during the study time period. One hundred eighty cycles (56%) had<br />

documented CM type and were included in our analysis. There were no differences<br />

between cycles with and those without documented CM. Women<br />

(n¼125) were typically Caucasian (76%), with an average age of 34.4 years,<br />

body mass index (BMI) of 25.4 kg/m2 and a diagnosis of unexplained (29%)<br />

or male factor (22%) infertility. A majority of cycles utilized oral ovulation<br />

induction agents (73%) and ovulation prediction kit monitoring for IUI<br />

timing (67%). Gonadotropins were used in 18% of cycles and 9% were<br />

un-medicated. Fertile type CM was present in 91% of cycles. There were<br />

no differences between fertile versus non-fertile type CM in regards to patient<br />

age, race, BMI, AMH, number of follicles on midcycle ultrasound, endometrial<br />

thickness, method used for IUI timing, fertility medications, or infertility<br />

diagnosis. Cycle pregnancy rate did not differ between fertile type CM (<strong>17</strong>%)<br />

and non-fertile type CM (19%), p¼0.75.<br />

CONCLUSIONS: Most patients undergoing IUI demonstrate fertile type<br />

cervical mucus on the day of IUI. Type of cervical mucus does not predict<br />

probability of conceiving following IUI. Therefore, it is likely the cervical<br />

mucus itself that facilitates fertilization during procreative intercourse.<br />

P-232 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

TO WHAT DEGREE SHOULD THE ZONA PELLUCIDA BE CUT<br />

OPEN IN ASSISTED HATCHING FOR BEST CLINICAL<br />

RESULTS? H. Watanabe, R. Suzuki, M. Kobayashi, H. Hasegawa,<br />

K. Tsukamoto, S. Saitou, J. Kobayashi. Kanagawa Ladies Clinic, Yokohama,<br />

Japan.<br />

OBJECTIVE: Assisted hatching is widely used to improve implantation<br />

by opening or thinning of Zona Pellucida artificially. However, there is no<br />

consensus to what degree the Zona Pellucida should be cut open for best clinical<br />

results. We compared the size of the opening of the Zona Pellucida with<br />

clinical results, using Saturn 5 ActiveTM (Research Instruments Ltd. UK) for<br />

lazar assisted hatching (LAH) and EmbryoScopeBR(Vitrolife AB. Sweden)<br />

for continuous observation.<br />

DESIGN: prospective randmaized study.<br />

MATERIALS AND METHODS: Cryopreserved blastocysts (SBB<br />

grade) which were donated by patients for research were used for this<br />

study. LAH was performed on warmed blastocysts after one hour of<br />

culturing. The blastocysts were divided into 5 groups by the size of the<br />

opening for assisted hatching: Group1 has been performed 50% opening<br />

(n¼9) and Group2 has been performed 30% opening (n¼12). Group3 has<br />

been conducted 12 micrometers opening (n¼12) so as Group4 has been<br />

done the thinning (n¼12). For Group5 no LAH has been performed<br />

(n¼12).We observed the start timing of hatching, time of hatching,<br />

hatched rate and frequency of contraction until blastocysts finished hatching<br />

by Time Lapse after LAH.<br />

RESULTS: The time until the initiating of hatching was significantly<br />

longer for groups 4 and 5 when compared to other groups. In fact, blastocysts<br />

of groups 4 and 5 needed 1-3 times of contraction to start hatching as found<br />

by Time Lapse observation. Time of hatching for Group 3 was significantly<br />

slower than for other groups. In that group, embryos were stuck by the small<br />

hole of the Zona Pellucida and repeated contraction occurred during hatching.<br />

The hatched rates were significantly lower in groups 3, 4 and 5 when<br />

compared to groups 1 and 2. Frequency of contraction until the blastocysts<br />

finished hatching was higher for groups 3, 4 and 5. There was no significant<br />

difference between groups 1 and 2.<br />

CONCLUSIONS: From this study, we found that adequate LAH improve<br />

subsequent hatching rates. In fact, the Zona Pellucida should not merely be<br />

thinned by rather cut open, at least 30%.<br />

P-233 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ART - IN VITRO FERTILIZATION<br />

NOVEL APPROACH FOR MANAGING DECREASED FERTILIZA-<br />

TION WITH SEQUENTIAL ARTIFICIAL OOCYTE ACTIVATION:<br />

PROSPECTIVE RANDOMIZED CLINICAL TRIAL. M. Fawzy, a<br />

H. Abdelghaffar, b A. Alaboudy, a M. Sabry, b H. Kasem, a<br />

M. Y. Abdel-Rahman, b M. Gad, a A. Metwallley, c E. Othman, d<br />

A. Mahran, d S. M. Rasheed. b a Ibnsina IVF Center, Sohag, Egypt; b Sohag<br />

Faculty of Medicine, Sohag, Egypt; c AlBarka IVF Center, Manama, Bahrain;<br />

d Banon IVF Center, Assiut, Egypt.<br />

OBJECTIVE: To compare outcomes between two different strategies of<br />

artificial Oocyte activation: Calcium Ionophore activation protocol versus<br />

a novel sequential protocol using Ca Ionophore followed by Strontium Chloride,<br />

In cases of previously decreased fertilization ICSI cycles.<br />

DESIGN: Prospective Randomized Clinical Trial.<br />

MATERIALS AND METHODS: Patients with previous low fertilization<br />

who presented to Ibnsina IVF Center, Sohag, Egypt (private center), between<br />

January <strong>2015</strong> to March <strong>2015</strong> were recruited to undergo ICSI into<br />

this study. Mature Oocytes (840) collected from 69 patients injected and<br />

randomly assigned to two groups (420 each). Group I Oocytes activated artificially<br />

for 15 min in ready made Ca Ionophore (GM508 Cult-Active, GY-<br />

NEMED, Germany) immediately after ICSI followed by culture in<br />

(GLOBAL TOTALÒ, LIFEGLOBALÒ, Canada). Group II Oocytes underwent<br />

sequential activation using Ca ionophore (GM508 Cult-Active, GY-<br />

NEMED, Germany) same protocol followed by another activation step<br />

by washing and incubation for 60 min in Strontium Chloride (Sigma<br />

69042). The Strontium Chloride Prepared in our laboratory at a concentration<br />

of 10 micromol/ml in culture media (GLOBAL TOTALÒ, LIFEGLO-<br />

BALÒ, Canada). The activated oocytes cultured traditionally according to<br />

our protocol. All patients underwent Day 5 extended culture, and the embryos<br />

selected had been randomized either from group I or II to be transferred<br />

using closed envelope method.Outcome measures included:<br />

Fertilization rate, top quality embryos day at 3, blastocyst formation rate<br />

and finally pregnancy rate for each group. We compare the data of our<br />

two group with the Chi-square test.<br />

RESULTS: Study groups patients were similar regarding mean age, BMI,<br />

the dose of FSH/HMG used, number of oocytes collected and number of cycle<br />

days. There were a statistically significant increase in the rate of fertilization<br />

in group II versus group I (group II 69% versus group I 46% P value <<br />

0.01). We documented higher blastocyst formation rate in group II versus<br />

group I (group II 59% compared to 38% group I P


assessment of clinical efficacy regarding multiple other clinical parameters<br />

in a larger study in our center (Clinical Trial Registration:<br />

NCT02418416).<br />

P-234 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

DUAL TROPHECTODERM BIOPSY ON THE SAME BLASTOCYST<br />

DOES NOT IMPAIR CLINICAL OUTCOMES. J. E. Swain, a,b<br />

W. B. Schoolcraft, c M. Katz-Jaffe. c a Fertility Lab Sciences, Lone Tree,<br />

CO; b Fertility Labs of Colorado, Lone Tree, CO; c Colorado Center for<br />

Reproductive Medicine, Lone Tree, CO.<br />

OBJECTIVE: Embryo biopsy has become common procedure for many<br />

in vitro fertilization (IVF) laboratories. However, as one of the most invasive<br />

techniques used, care must be taken to not compromise embryo<br />

viability. Taking two blastomeres from cleavage stage embryos has been<br />

shown to be detrimental compared to a single blastomere biopsy. Trophectoderm<br />

(TE) biopsy has now become the approach of choice due to reduced<br />

embryo damage as well as other advantages in analysis. Situations exist<br />

where two TE biopsies may be warranted. However, it is unknown if a double<br />

TE biopsy is detrimental to reproductive outcomes compared to a single<br />

TE biopsy.<br />

DESIGN: Retrospective data analysis.<br />

MATERIALS AND METHODS: Per standard laboratory protocol for<br />

patients undergoing simultaneous preimplantation genetic diagnosis<br />

(PGD) for single gene defects coupled with comprehensive chromosome<br />

screening (CCS), good quality blastocysts R3BB were subjected to two<br />

TE biopsies (3-4 cells per biopsy). One biopsy was subjected to targeted<br />

PGD for single gene defects, while the other biopsy was subjected to<br />

WGA for CCS. Clinical outcomes were compared to age-matched patients<br />

undergoing only CCS who had blastocysts subjected to one TE biopsy<br />

(4-6 cells). All blastocysts were vitrified/warmed a single time prior<br />

to transfer in a subsequent frozen cycle. This approach ensured similar<br />

quality blastocysts were compared. Differences were determined using<br />

Fisher’s Exact test, p


DESIGN: Prospective, randomized trial.<br />

MATERIALS AND METHODS: From June 2014 to November 2014 data<br />

was collected from 51 patients undergoing IVF treatment whose oocytes<br />

were split and randomly allocated between sequential and 1-step mediums.<br />

Fertilized oocytes were cultured up to day 6. A modified Gardner grading<br />

system was used to score embryos. Good quality (GQ) blastocysts are those<br />

with at least an ICM and TE of grade B. A total of 140 embryos (from <strong>17</strong> patients)<br />

underwent a biopsy procedure on day 5/6 and aCGH testing. Statistical<br />

analysis was performed using Fisher’s exact test, and the results were<br />

considered significant if P


Table 1.<br />

embryos<br />

Y Chromosome<br />

Day 3 biopsy<br />

Trophectoderm<br />

biopsy<br />

Analysis<br />

References:<br />

1. Orzack et al,<br />

2. Tarin et al, Reproductive<br />

3. Maalouf et al,<br />

P-240 Tuesday, <strong>October</strong><br />

TIME-LAPSE VIDEO<br />

TED HATCHING<br />

BLASTOCYST WARMING<br />

BRYO’S ABILITY<br />

C. Underberger, a<br />

W. Chang, b C. Alexander,<br />

Hills, CA; b Southern<br />

OBJECTIVE: The<br />

(ZP) hardening due<br />

quence of vitrification.<br />

ZP post-warming,<br />

receptivity. We set<br />

were able to fully escape<br />

hatching (LAH) of<br />

ing compared to embryos<br />

DESIGN: Prospective<br />

placed in an Embryoscope<br />

MATERIALS AND<br />

blastocysts vitrified<br />

with a range of 23-46.<br />

protocol on the same<br />

ditions. Embryos were<br />

ablated using LAH<br />

All embryos were incubated<br />

bator in GTL medium<br />

the time it took to fully<br />

embryo underwent<br />

RESULTS: Assessment<br />

hatched out. In Group<br />

ing within 24 hours,<br />

an average of 2.1 expansions.<br />

complete hatching<br />

expansions. Two of<br />

at 37 and 48 hours with<br />

plete hatching within<br />

CONCLUSIONS:<br />

tion regarding the<br />

cantly higher percentage<br />

the hatching process<br />

that were not LAH.<br />

completely escape<br />

receptivity. The LAH<br />

ergy to complete the<br />

mized, thus ensuring<br />

Conversely, an overwhelming<br />

have LAH prior to embryo<br />

process, even within<br />

Overall, LAH of 33%<br />

bryo’s escape and could<br />

P-241 Tuesday, <strong>October</strong><br />

A TIME-LAPSE<br />

FREEZING<br />

EMBRYOS. M.<br />

G. R. Navarrete, d Total Total<br />

with without<br />

Chromosome<br />

1893 1966<br />

1243 1291<br />

_<br />

www.pnas.org/cgi/doi/10.1073/pnas.1416546112,<br />

Biology<br />

F & S 101:<br />

20, <strong>2015</strong><br />

ANALYSIS<br />

OF 33%<br />

TO FULLY<br />

D. L. Hill, J. Barritt.<br />

California<br />

human embryo<br />

both to extended<br />

This may<br />

especially within<br />

out to determine<br />

the ZP<br />

33% of the<br />

that were<br />

analysis<br />

with<br />

METHODS:<br />

on Day 5<br />

Warming<br />

day, at the<br />

randomized<br />

on a Zilos laser<br />

for<br />

(Vitrolife).<br />

hatch from<br />

to escape the<br />

was performed<br />

1, 11 of 14<br />

with an average<br />

In<br />

within 24 hours,<br />

the Group 2<br />

an average<br />

24 hours was<br />

Time lapse<br />

hatching process<br />

of warmed<br />

when 33%<br />

Importantly,<br />

the ZP was within<br />

embryos also<br />

hatching process,<br />

the energy<br />

percentage<br />

transfer<br />

48 hours, which<br />

of the ZP<br />

significantly<br />

20, <strong>2015</strong><br />

EVALUATION<br />

AND VITRIFICATION<br />

Meintjes, a<br />

B. Tilley, e embryos<br />

Y Total embryo<br />

gender ratio<br />

0.95<br />

0.96<br />

Not significant<br />

and Endocrinology<br />

13<strong>21</strong>, 2014.<br />

DEMONSTRATES<br />

OF THE ZONA-PELLUCIDA<br />

SIGNIFICANTLY<br />

HATCH.<br />

M. Surrey, ART<br />

Reproductive<br />

is known<br />

culture to<br />

affect the blastocysts<br />

an optimal<br />

if the percentage<br />

post-warming<br />

ZP was performed<br />

not LAH.<br />

of donated<br />

and without LAH.<br />

Overall,<br />

and 6 were used.<br />

of embryos was<br />

same time, under<br />

to either a group<br />

at 400 ms (Group<br />

48 hours in an<br />

We recorded<br />

the ZP and the<br />

ZP.<br />

to determine<br />

embryos (78.6%)<br />

of <strong>17</strong>.0 hours<br />

Group 2, none<br />

after having<br />

embryos (20%)<br />

of 5 expansions.<br />

significantly<br />

video evidence<br />

of blastocysts<br />

embryos<br />

of the ZP was<br />

the average<br />

the optimal<br />

demonstrated<br />

through<br />

stores are not<br />

of<br />

are unsuccessful<br />

is outside optimal<br />

post-warming<br />

increase implantation<br />

OF CONVENTIONAL<br />

S. Purcell,<br />

K. L. Lee, f Euploid Euploid<br />

XY XX gender<br />

554 632<br />

649 608<br />

_ _<br />

12:56,<br />

LASER-ASSIS-<br />

IMPROVES<br />

K. Salmon, a D.<br />

H. Danzer, b S.<br />

Reproductive Center,<br />

Center, Beverly Hills,<br />

to undergo Zona-Pellucida<br />

the blastocyst and<br />

ability to<br />

window of endometrial<br />

of blastocysts<br />

was different if laser<br />

immediately after<br />

vitrified/warmed<br />

24 research good/fair<br />

The mean age<br />

performed per established<br />

the same incubation<br />

that had 33%<br />

1) or no LAH<br />

Embryoscope time<br />

if the embryo escaped<br />

number of expansions<br />

if and when<br />

achieved complete<br />

elapsed to fully<br />

of the 10 embryos<br />

undergone an average<br />

achieved complete<br />

Group 1 vs. Group<br />

different (p


sperm collected from cauda epididymis (C57BL6 inbred and B6D2F1<br />

outbred) for 4-6h. Zygotes were cultured either in suboptimal Whitten’s<br />

medium and 20% O2 (IVFWM) or in optimal KSOM medium with amino<br />

acids and 5% O2 (IVFKAA) for 96h in 37 C. Control blastocysts were<br />

flushed from the uterus of mated mice (FB mice). Blastocysts were transferred<br />

to pseudo-pregnant recipients. Pups were reared to adulthood and peripheral<br />

tissues were collected and serum corticosterone levels measured.<br />

Total RNA and protein were isolated from fat, muscle and liver. mRNA<br />

expression of GR and a selected group of GR-downstream target genes<br />

were analyzed by qPCR; GR protein level by Western Blot. One-way AN-<br />

OVA with post hoc correction was used as appropriate; p


RESULTS: The majority of PGS cases involve biopsy and cryopreservation<br />

following an oocyte retrieval. In these patients we obtained a 68% implantation<br />

rate and 62% ongoing pregnancy. The number of patients was<br />

low and patients were younger in the re-frozen group, and these patients<br />

had higher implantation (89%) and ongoing pregnancy rates (71%).<br />

Table 1. Control and Re-frozen PGS FET results.<br />

Group<br />

Control<br />

PGS<br />

Re-frozen<br />

PGS<br />

N<br />

Patients<br />

Average<br />

age<br />

N<br />

Embryos<br />

Implantation<br />

(%)<br />

Ongoing<br />

pregnancy (%)<br />

78 36.1 98 68 62<br />

7 34.7 9 89 71<br />

CONCLUSIONS: Cryopreserved embryos that had not been previously biopsied<br />

can be thawed, biopsied, and re-frozen effectively. This process does not<br />

appear to have a detrimental effect on their subsequent post-thaw success. It is<br />

not necessary transfer these embryos back immediately after thaw and biopsy,<br />

which allows for greater flexibility to the patient and laboratory.<br />

P-246 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

EVALUATION OF THE USEFULNESS OF REFROZEN-THAWED EM-<br />

BRYO TRANSFER (R-FET). K. Takeuchi, Y. Homan, Y. Fukumoto,<br />

Y. Kuroki, M. Tokudome, H. Setoyama, S. Awata, M. Takeuchi. Takeuchi<br />

Ladies Clinic/Infertility Center, Aira-shi, Kagoshima, Japan.<br />

OBJECTIVE: Multiple embryos are occasionally thawed for frozen<br />

-thawed embryo transfer (FET). In the case that all the thawed embryos<br />

are not used, the surplus embryos need to be refrozen. However, the number<br />

of reports on R-FET is limited, and its usefulness is poorly understood. Due<br />

to this, data on FET at our clinic was collected, and the clinical outcomes of<br />

the embryo transfers that were frozen and thawed twice or three times, were<br />

compared with those that were frozen and thawed once.<br />

DESIGN: Retrospective study.<br />

MATERIALS AND METHODS: The data for 590 frozen-thawed blastocyst<br />

transfer cycles, performed from April 2012 to March 2014, was subjected<br />

to assessment. To make a comparison, with regard to clinical outcomes, they<br />

were classified into the following three groups: Embryos frozen once (A), embryos<br />

frozen twice (B) and embryos frozen three times (C). There was no statistically<br />

significant difference between these groups in terms of patients’ age,<br />

the number of embryos transferred. Cryotip Safety Kit (Kitazato BioPharma<br />

Co., Ltd.) was used for freezing/thawing of embryos.<br />

RESULTS: The pregnancy rate was 45.2%(200/442) in A, 45.7%(64/140)<br />

in B, and 37.5%(3/8) in C, while the abortion rate was 19.5%(39/200) in A,<br />

18.8%(12/64) in B, and 0.0%(0/3) in C. There was no significant difference<br />

between A and B. The pre-freeze embryo grade was compared with the postfreeze<br />

embryo grade and the rate of decline in embryo grade was 19.2%(85/<br />

442) in A, <strong>21</strong>.4%(30/140) in B, and 12.5%(1/8) in C. There was no significant<br />

intergroup difference. The pregnancy/abortion rates of the embryos that<br />

showed elevation in grade were 45.7%(163/463) in A, 46.4%(51/110) in B<br />

and 28.6%(2/7) in C, and 19.0%(13/163) in A, <strong>17</strong>.6%(9/51) in B and<br />

0.0%(0/2) in C respectively. The pregnancy/abortion rates of the embryos<br />

that showed decline in grade were 43.5%(37/85) in A, 43.3%(13/30) in B<br />

and 100.0%(1/1) in C, and <strong>21</strong>.6%(8/37) in A, 23.1%(3/13) in B and<br />

0.0%(0/1) in C respectively. In the assessed embryos that showed elevation<br />

in grade and those that showed decline in grade, no significant difference<br />

was seen between A and B in terms of the pregnancy/abortion rates. C was<br />

cut from assessment because the total number of cycles was limited. However,<br />

both the pregnancy and abortion rates obtained from C suggest the usefulness<br />

of R-FET.<br />

CONCLUSIONS: In this study, the pregnancy rate achieved by R-FETwas<br />

comparable to that of pregnancy achieved by FET. Also no difference was<br />

observed in terms of the abortion rate between the two treatments. These results<br />

suggest the usefulness of R-FET.<br />

P-247 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

UTILIZATION OF CRYOPRESERVED EMBRYOS FOR A SECOND<br />

PREGNANCY. L. Kroener, a A. l. Akopians, a M. D. Pisarska, b<br />

D. L. Hill, c J. Barritt, c E. T. Wang. b a Obstetrics and Gynecology, University<br />

of California Los Angeles, Los Angeles, CA; b Cedars Sinai Medical Center,<br />

Los Angeles, CA; c ART Reproductive Center, Beverly Hills, CA.<br />

OBJECTIVE: As laboratory technology improves and elective single embryo<br />

transfer rises, an increasing number of embryos are being cryopreserved.<br />

However, there is little data on the utilization and fate of this<br />

growing number of cryopreserved embryos. Our aim is to investigate the<br />

use of cryopreserved embryos for a second pregnancy, as well as the overall<br />

utilization and disposition of cryopreserved embryos on a per cycle basis.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: We reviewed 1,131 IVF cycles with at<br />

least one embryo resulting in a clinical pregnancy and remaining cryopreserved<br />

embryos between 1/2009 and 12/2012. Both autologous and donor oocyte cycles<br />

were included. The primary outcome was use of embryos for a second<br />

pregnancy over a mean follow-up time of 51.4 months. Disposition of cryopreserved<br />

embryos was defined as continued cryopreservation, discarded, or<br />

donated to research. Analysis was performed to assess the impact of age<br />

(42) on utilization of cryopreserved embryos for<br />

a second pregnancy, length of time between the first and second pregnancy<br />

transfer, and disposition of cryopreserved embryos. Statistical analysis was performed<br />

using the Chi-squared test and Kruskal Wallis test as appropriate.<br />

RESULTS: A total of 3,994 embryos were cryopreserved from 1,131 IVF<br />

cycles with at least one embryo transfer resulting in a clinical pregnancy. The<br />

breakdown of cycles by age was 453 patients 42, with 46.8% and 94.3%<br />

donor cycles in the 41-41 and >42 age groups respectively. Only 18.0% (204/<br />

1,131) of patients attempted to use their remaining cryopreserved embryos for<br />

a second pregnancy. There was no difference in the likelihood of returning for<br />

a second pregnancy transfer across age groups (p¼0.136). There was an inverse<br />

relationship between patient age and length of time between their first<br />

and second pregnancy transfers (31.4 months for age 42, p¼0.001). There was<br />

no impact of age on the disposition of embryos (p¼0.41). Of the 927 cycles<br />

that patients have not come back to use their embryos for a second pregnancy,<br />

12.9% chose to discard or donate their embryos (no difference across age<br />

groups, p¼0.587) and 87.1% have continued cryopreservation.<br />

CONCLUSIONS: Use of cryopreserved embryos for a second pregnancy<br />

is lower than expected and is not influenced by age. Despite this low utilization<br />

of embryos, the vast majority of patients continue to preserve rather than<br />

discard or donate their embryos. This data provides evidence of an increasing<br />

number of unused cryopreserved embryos and yields important insights for<br />

counseling patients on use of cryopreserved embryo use for second pregnancy.<br />

Studies with a longer follow-up time are needed to capture patients<br />

coming back after more than 4 years.<br />

P-248 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

DECREASED RATE OF EUPLOID EMBRYOS BY PREIMPLANTA-<br />

TION GENETIC SCREENING IS ASSOCIATED WITH LOWER<br />

LIVE BIRTH RATES. J. Lekovich, a P. Chung, a A. Lobel, a<br />

N. Pereira, b J. Stewart, c Z. Rosenwaks. a a Weill Cornell Medical College,<br />

New York, NY; b The Ronald O. Perelman and Claudia Cohen Center for<br />

Reproductive Medicine, New York, NY; c Weill Cornell Medical Center,<br />

New York, NY.<br />

OBJECTIVE: To investigate whether the rate of euploid embryos by preimplantation<br />

genetic screening (PGS) affects the outcome of pregnancy.<br />

DESIGN: Retrospective cohort.<br />

MATERIALS AND METHODS: In a one-year period, data of 102 pregnant<br />

patients who underwent frozen embryo transfer (FET) of PGS-proven<br />

euploid embryos were analyzed. They were divided into two groups based<br />

on the cycle outcome: one with live births and the other whose cycle resulted<br />

in a pregnancy loss (miscarriage or a biochemical pregnancy). Patients with a<br />

history of recurrent miscarriages, thrombophilia, and uterine factor were not<br />

included in the study. Main parameter of interest was to compare the rate (%)<br />

of aneuploid embryos over total number of biopsied embryos (per patient)<br />

between the two groups. Other confounding factors were also compared:<br />

age, gravidity, parity, antimullerian hormone (AMH) levels, mid-cycle endometrial<br />

thickness (ES), number of 2 pronuclear zygote (2PN) stage embryos,<br />

the rate of progression to blastocyst by day 5, and the number of embryos<br />

transferred.<br />

RESULTS: Eighty one patients with live births were compared to <strong>21</strong> patients<br />

who had a pregnancy loss (either a biochemical pregnancy or first<br />

trimester miscarriage). The miscarriage group demonstrated significantly<br />

lower AMH levels. All other confounding factors were comparable between<br />

e190 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


Table 1<br />

the two groups (Table 1). However, aneuploidy rate of all biopsied embryos<br />

was significantly higher in the miscarriage group (0.62+0.28 vs. 0.49+0.28,<br />

p¼0.03) (Table 1).<br />

CONCLUSIONS: Despite the transfer of similar number of PGS-euploid<br />

embryos and other confounding factors being controlled for, it appears that<br />

the pregnancy loss is associated with decreased proportion of euploid embryos<br />

among all embryos biopsied for PGS.<br />

Supported by: Institutional.<br />

P-249 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

Live birth<br />

group<br />

(81)<br />

Miscarriage<br />

group (<strong>21</strong>)<br />

Age (years) 36.5 4.2 36 4.8 0.8<br />

Gravidity 2 1.4 1.8 1.6 0.6<br />

Parity 0.6 0.9 0.6 1 0.8<br />

Mean endometrial stripe (mm) 9.7 þ 1.9 9.6 þ 2 0.7<br />

AMH (ng/mL) 3.7 3.8 2.1 1.6 0.004<br />

Mean # of 2PN 10.5 þ 6 9.6 þ 6.3 0.5<br />

Rate of progression to 0.7 0.2 0.6 0.2 0.6<br />

blastocyst stage by<br />

day 5 (# day 5<br />

blastocyst/# 2PN)<br />

Mean # of embryos transferred 1.3 0.5 1.5 0.4 0.4<br />

Aneuploidy rate<br />

(#aneuploid embryos/total<br />

# of biopsied day<br />

5 embryos per patient)<br />

0.49 0.25 0.62 0.28 0.03<br />

EVALUATION OF EMBRYO IMPLANTATION POTENTIAL AFTER<br />

VITRIFIED AND WARMED SINGLE BLASTOCYST<br />

TRANSFER PER EMBRYO CATEGORIZATION USING THE<br />

EEVAÔ TEST, AN AUTOMATED TIME-LAPSE QUANTITATIVE<br />

ANALYSIS. S. De Gheselle, a B. Heindryckx, a P. De Sutter, b E. Van<br />

den Abbeel. a a Department for Reproductive Medicine, University Hospital<br />

Ghent, Ghent, Belgium; b Head Department for Reproductive Medicine, University<br />

Hospital Ghent, Gent, Belgium.<br />

OBJECTIVE: The Eeva Test provides a set of categories predicting embryo<br />

developmental potential based on an automated analysis of early cell division<br />

timings. The aim of this study was to evaluate the percentage of<br />

embryos vitrified at the blastocyst stage and to compare the implantation<br />

rate of the vitrified and warmed blastocysts in each Eeva Test category<br />

(HIGH, MEDIUM, and LOW), in a group of supernumerary embryos after<br />

single embryo transfer (SET) on day 3.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: This study included 76 patients who<br />

underwent day 3 SET (Feb-Dec 2013). All embryos received Eeva Test<br />

scores on day 3. Supernumerary embryos (N¼540) were cultured to the<br />

blastocyst stage (day 5), and blastocysts exhibiting expansion status 1-2<br />

and 3-5 with quality Inner cell mass (ICM) and Trophectoderm (TE)<br />

grade A, B, C were vitrified (N¼209). Single vitrified and warmed blastocysts<br />

were transferred and implantation potential was evaluated<br />

(N¼95). The percentage of embryos vitrified and the implantation rate<br />

of warmed blastocysts was calculated for each Eeva Test category and<br />

compared.<br />

RESULTS: Pregnancy rate for fresh day 3 SET was 31.6% . The percentage<br />

of good and fair embryos on day 3 in each category was 53.4%<br />

(78/146) for Eeva HIGH, 48.1% (39/81) for Eeva MEDIUM and 27.1%<br />

(85/313) for Eeva LOW, respectively (p 0.05<br />

re-expansion<br />

Survival rate 58.1 63.2 33.3 > 0.05<br />

Implantation rate 22.6 14 5.6 > 0.05<br />

FERTILITY & STERILITY Ò<br />

e191


P-251 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

CUMULATIVE PREGNANCY RATE BASED ON THE NUMBER OF<br />

EMBRYO TRANSFERS IN ASSISTED REPRODUCTIVE<br />

TECHNOLOGIES. Y. Mio, a K. Iwata, b K. Yumoto, b C. Mizoguchi, a<br />

M. Sugishima, b M. Tsuneto, b Y. Iba. a a Reproductive Centre, Mio Fertility<br />

Clinic, Yonago, Japan; b Mio Fertility Clinic, Yonago, Japan.<br />

OBJECTIVE: The pregnancy rate (PR) of assisted reproductive technologies<br />

(ART) was calculated as the number of pregnancies divided by the number<br />

of embryo transfers (ET) cycles and/or patients treated within a specified<br />

period. However, the population of these cycles and/or patients included<br />

several unsuitable cases involving cancellation, failed oocyte pick-up<br />

(OPU) or ET, and/or freeze-all cycles. Thus, it was difficult to evaluate the<br />

actual success rate of ART programs in those couples having continued<br />

ET. Considering that the goal of ART is to achieve successful pregnancy,<br />

the most important issues on which to evaluate success are 1) the number<br />

of ET required for successful pregnancy and 2) the cumulative pregnancy<br />

rate (PR) with repeated ET. This study reviewed these data based on the number<br />

of ET per couple.<br />

DESIGN: Retrospective study based on the ART database in our Reproductive<br />

Centre from January 2006 to December 2014.<br />

MATERIALS AND METHODS: For the <strong>21</strong>92 patients participating in<br />

our ART programs, we reviewed the individual clinical outcomes of<br />

ART, including the number of cycles required to achieve pregnancy, the cumulative<br />

number of pregnancies and take-home babies, and the number of<br />

ART cancelations. We then revised the data against each couple and their<br />

age.<br />

RESULTS: Of the <strong>21</strong>92 study participants, 81.9% (n ¼ <strong>17</strong>97) had ET and<br />

73.8% (n ¼ 1327) achieved successful pregnancy, and of these, 79.1% (n ¼<br />

1050) achieved successful pregnancy within three ET, 94.7% (n ¼ 1257)<br />

within six ET, and 98.5% (n ¼ 1307) within nine ET. The number of patients<br />

who experienced cancellation and no ETwere 3<strong>17</strong> and 395, respectively. The<br />

revised PR [successful pregnancy (n ¼ 1327) divided by the revised population<br />

(n ¼ 1480)] was 89.7%, and of these patients, 95.7% (201/<strong>21</strong>0) were<br />

aged under 30 years, 95.9% (523/547) 30-34 years, 88.1% (474/538) 35-39<br />

years, and 71.9% (128/<strong>17</strong>8) were aged 40-44 years. The number of takehome<br />

babies in this period was 1<strong>17</strong>1 (65.2%/ET and 88.2%/successful pregnancy).<br />

CONCLUSIONS: The PR from individual ET cycles was limited and<br />

declined dramatically with respect to advanced maternal age (32.6% in <<br />

30 year-olds, 30.9% in 30-34 year-olds, 25.6% in 35-39 year-olds, and<br />

12.1% in 40-44 year-olds). In contrast, the miscarriage rate increased markedly<br />

depending on advanced maternal age (14.2% in < 30 year-olds, <strong>17</strong>.8%<br />

in 30-34 year-olds, 23.0% in 35-39 year-olds, and 38.8% in 40-44 year-olds).<br />

This study of revised PR demonstrates clearly that ART provides a great opportunity<br />

to achieve successful pregnancy by repeating the ET. To this end, it<br />

is crucial to obtain a suitable number of cryopreserved embryos, for both<br />

reducing the cost and the psychophysical burden of such techniques, and<br />

increasing the expectation of successful pregnancy in participating ART patients.<br />

P-252 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

DOES EGG VITRIFICATION DISPLACE THE FIRST POLAR<br />

BODY - MEIOTIC SPINDLE AXIS? S. Phillips, a J. Graham, b<br />

A. Coutinho, a M. Racicot, a C. Beauchamp. a a OVO Fertility, Montreal,<br />

QC, Canada; b Shady Grove Fertility Centre, Rockville, MD.<br />

OBJECTIVE: To assess whether the polar body-meiotic spindle axis is<br />

affected by vitrification.<br />

DESIGN: A prospective observational study carried out in one centre of a<br />

donor egg bank programme.<br />

MATERIALS AND METHODS: Vitrification, transport, storage and<br />

warming are all carried out according to strict egg bank protocols and regularly<br />

validated. We received 9 donor egg lots from 8 different clinics around<br />

the US between November 2014 and January <strong>2015</strong>. At the time of ICSI, all<br />

oocytes were assessed for the presence and location of the meiotic spindle in<br />

relation to the first polar body using a polscope. The degree of displacement<br />

was noted. Embryo transfer was carried out on day 3 or day 5 according to the<br />

number and quality of the embryos available. One embryo on day 3 or day 5<br />

was transferred in all cases except one where two day 3 embryos were transferred.<br />

Clinical pregnancy was confirmed by the presence of an intrauterine<br />

fetal heart on ultrasound at 7 weeks.<br />

RESULTS: A total of 65 MII eggs were warmed with a survival rate of<br />

80%. The meiotic spindle was seen using a polscope in 84% of the eggs .<br />

The spindle was within 45 of the first polar body in 34% of cases and was<br />

more than 45 displaced in 50% of eggs. The fertilisation rate was 64%<br />

with 7 cases reaching embryo transfer and 5 of these obtaining a clinical<br />

pregnancy (71%).<br />

CONCLUSIONS: In the literature, the spindle is visualised in approximately<br />

80% of fresh oocytes (1) and this compares with the 80% visualisation<br />

rate that we saw in cryopreserved-warmed oocytes. The proportion<br />

of spindles from fresh oocytes which are within 45 of the first polar body<br />

have been reported as approximately 70% (2,3). In our study it was noted<br />

that only 34% of spindles were within 45 of the first polar body so half of<br />

what would be expected from previous studies on fresh oocytes. It could<br />

be suggested that the method of cumulus cell removal prior to ICSI can<br />

have an effect on the displacement of the polar body, however in our<br />

study oocytes came from 8 different centres all using their own technique<br />

for denudation and displacement was seen from all centres. In addition,<br />

the previous fresh oocyte studies all required cumulus cell removal prior<br />

to polscope assessment. The process of oocyte vitrification causes the ooplasm<br />

to shrink and re-expand within the zona pellucida which could<br />

cause a mechanical displacement of the polar body as cryoprotectants<br />

are introduced and then withdrawn. It has been shown that high doses<br />

of FSH can affect the spindle (4) and it is therefore possible that strong<br />

ovarian stimulation of egg donors attempting to maximize oocyte recovery<br />

could have an impact. Unfortunately in these cases, spindle assessment<br />

prior to vitrification was not possible. Further studies to eliminate<br />

these variables are necessary to confirm that the process of vitrification<br />

itself can impact on the placement of the meiotic spindle post warming<br />

and therefore the use of a polscope when performing ICSI on these<br />

eggs would be highly recommended.<br />

References:<br />

1. Rama Raju G. et al. Meiotic spindle and zona pellucida characteristics<br />

as predictors of embryonic development: a preliminary study using Pol-<br />

Scope imaging. RBMOnline. 2007. 14(2):166-74.<br />

2. Rienzi L et al. Relationship between meiotic spindle location with regard<br />

to the polar body position and oocyte developmental potential after<br />

ICSI. Human Reproduction. 2003. 18(6):1289-93.<br />

3. Moon JH et al. Visualization of the metaphase II meiotic spindle in<br />

living human oocytes using the Polscope enables the prediction of embryonic<br />

developmental competence after ICSI.Human Reproduction.<br />

2003. 18(4):8<strong>17</strong>-20.<br />

4. Madaschi C et al. Zona pellucida birefringence score and meiotic spindle<br />

visualization in relation to embryo development and ICSI outcomes.RBMOnline<br />

2009. 18(5):681-6.<br />

P-253 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

CRYOPRESERVED DONOR OOCYTES YIELD SIGNIFICANTLY<br />

LOWER BIRTH RATES THAN FRESH OOCYTES. V. A. Kushnir, a<br />

D. H. Barad, b D. F. Albertini, c S. K. Darmon, d N. Gleicher. b a Center for Human<br />

Reproduction & Wake Forest Univer, New York, NY; b Center for Human<br />

Reproduction & Foundation for Reproductive Medicine, New York,<br />

NY; c Center for Human Reproduction & University of Kansas Medical Center,<br />

New York, NY; d Center for Human Reproduction, New York, NY.<br />

OBJECTIVE: To assess whether, as has been claimed, IVF live birth rates<br />

(LBRs) in fresh (FDO) and thawed (TDO) donor oocyte cycles are similar.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: We reviewed aggregate national oocyte<br />

donation data for 2013 in the annual Society for Assisted Reproductive<br />

Technology (SART) outcome report for the U.S. IVF centers. LBRs per cycle<br />

and per transfer in FDO and TDO were compared using the chi-square<br />

test.<br />

RESULTS: Of a total of 11,148 oocyte donation cycles, 2227 (20.0%)<br />

involved use of TDO. Recipient LBRs, whether per cycle start (intent to treat)<br />

or with reference point embryo transfer, were significantly higher for FDO<br />

than TDO. Indeed, per cycle start there was a 6.4% absolute difference in<br />

LBR, representing a 12.9% relative difference (P < 0.0001). This difference<br />

further increased with reference point embryo transfer (absolute 9.0%; relative<br />

16.0%; P < 0.0001). The cycle cancelation rate was higher for cycles utilizing<br />

FDO than TDO (P < 0.0001) (Table).<br />

e192 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


Donor oocytes<br />

Fresh<br />

Thawed<br />

P-value<br />

Number of cycles 89<strong>21</strong> 2227<br />

Recipient starts resulting 4425/49.6 962/43.2


(53.8% vs. 22.5%; p¼ .027). In 19 cases, one or more contractions of the blastocyst<br />

were observed but no relation with implantation was confirmed.<br />

CONCLUSIONS: The analysis of warmed blastocysts by time lapse imaging<br />

is providing new promising markers for blastocyst implantation potential<br />

and represents a novel alternative to morphological evaluation, establishing<br />

objective quantitative values linked with clinical outcome. However, further<br />

studies are required to validate the predictive power of this technology.<br />

P-257 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

LIPID PROFILE ANALYSIS FROM MICE EMBRYOS<br />

VITRIFIED. A. A. de Melo, a J. Camillo, a T. D. Oleinki, a<br />

F. B. Cordeiro, a E. G. Lo Turco, b T. G. Santos. c a Department of Surgery, Division<br />

of Urology, Human Reproduction Section, S~ao Paulo Federal University,<br />

S~ao Paulo, Brazil; b Head Embryologist, Sao Paulo, Brazil; c Student, Sao<br />

Paulo, Brazil.<br />

OBJECTIVE: This study aimed to identify potential lipid biomarkers<br />

related to the embryo vitrification, since this procedure is extremely important<br />

for assisted reproduction and may affect the development and the lipid<br />

profile of embryos.<br />

DESIGN: Prospective study.<br />

MATERIALS AND METHODS: Mice females C57BL/6J were superovulated,<br />

the embryos were collected after 60 hours in 8-cell stage and were<br />

divided into two groups: the control group, in wich embryos were cultured<br />

in 5% CO2 incubator until the blastocyst stage and evaluated based on their<br />

development time, and the experimental group, composed of vitrified embryos,<br />

that were thawed after 5 days, cultured in 5% CO2 until the blastocyst stage<br />

and also evaluated based on their development time. The lipids were extracted<br />

using the Bligh & Dyer method and were analyzed by Matrix-assisted laser<br />

desorption/ionization (MALDI-TOF/MS). Statistical analysis was performed<br />

using Principal Component Analysis (PCA) and partial least squares - discriminant<br />

analysis (PLS-DA), and the 15 ions corresponding to the potential lipid<br />

biomarkers were identified by the variable importance in projection (VIP).<br />

RESULTS: There was a lower rate of vitrified blastocyst group of 75.60%<br />

against 94.82% in the control group and variation in the rate at 24.40% of degenerated<br />

embryos vitrified group and 5.18% in the control group and rate<br />

variation in the degenerate embryos vitrified group of 24.40% and 5.18%<br />

in the control group. The PCA demonstrated that the principal component<br />

five better indicates the variance between the groups, presenting a prediction<br />

model with 81% of accuracy. The experimental group presented ganglioside,<br />

anthocyanins, phosphatidylglycerol, phosphatidylcholine, phosphatidylinositol<br />

as higher represented lipids and the experimental group presented flavonoids,<br />

phosphatidylserine and anthocyanins hyper-represented.<br />

CONCLUSIONS: There was a significant difference between the development<br />

rate after vitrification and the lipid profile differs significantly between<br />

embryos in the blastocyst stage and embryos subjected to cryopreservation.<br />

The lipids found in this study are related to good embryo quality in the control<br />

group and may present a broadly correlation with altered metabolism in<br />

the experimental group, indicating that despite the vitrification is essential,<br />

this is a procedure that affects the lipids embryo quality. This lipids are potential<br />

biomarkers since they may help in the development of procedures that<br />

may improve vitrification and reduce embryos damage.<br />

Supported by: Capes.<br />

type A (N¼567; 927 embryos), type B (N¼360; 575 embryos) and type C<br />

(N¼1<strong>21</strong>;196 embryos) were compared to fresh embryo transfers (fET) of<br />

embryos type A (N¼667;1099 embryos), type B (N¼337; 568 embryos)<br />

and type C (N¼84;135 embryos). Outcomes were compared by ANOVA<br />

test or chi-square as appropriate. Logistic regression analysis including<br />

age, own or donated cycles, fresh or vitrified oocytes and natural or HRT cycle<br />

for endometrial preparation as potential confounders, was used to adjust<br />

the Odds ratio (OR) for implantation.<br />

RESULTS: Survival rate was not altered by embryo morphology in CT:<br />

97.3% (96.3-98.3); 94.1% (92.2-97.1); and 95.9% (92.4-99.4) for type A,<br />

B and C embryos respectively (NS). IR was higher for type A embryos in<br />

both CT and fET and type B showed higher IR when compared to type C:<br />

40.4% (36.9-43.9); 32.7% (28.9-36.5); 24.4% (18.1-28.3) and 41.1%<br />

(38.5-44.3); 34.1% (30.2-38.0); <strong>21</strong>.2% (14.4-28.1) for types A, B and C in<br />

CT and fET respectively. IR was similar for each embryo category when<br />

compared between CT vs fET (NS). Similarly CPR was comparable between<br />

CTand fET groups (not shown). The OR for IR (0.932 (95% CI 0.789-1.090))<br />

showed no impact of vitrification on implantation. Adjusted OR showed that<br />

the only factor affecting IR was the embryo morphology (OR for IR (type A<br />

with respect to type C)¼ 2.<strong>21</strong>0 (1.669-2.926) and OR for IR (type B with<br />

respect to type C)¼ 1.600 (1.194-2.143)).<br />

CONCLUSIONS: Implantation potential is related to embryo quality<br />

regardless whether they are fresh or vitrified, thus confirming the safety of<br />

the vitrification procedure. This finding is of particular interest when evaluating<br />

the prognosis of CTs especially when considering the ‘‘Freeze all’’ strategy.<br />

P-259 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE EFFECT OF THE DAY OF CRYOPRESERVATION ON FET SUC-<br />

CESS RATES IN PGS AND NON-PGS BLASTOCYSTS. G. Navarrete, a<br />

S. Purcell, b B. Tilley, a S. J. Chantilis, a K. Lee, a M. Thomas, a R. Gada, a<br />

M. Meintjes. c,a a Dallas-Fort Worth Fertility Associates, Dallas, TX; b Reproductive<br />

Medicine and Surgery Center of Virgin, Charlottesville, VA; c Frisco<br />

Institute for Reproductive Medicine, Frisco, TX.<br />

OBJECTIVE: To determine if the day of blastocyst vitrification will have a<br />

different impact on euploid blastocysts when compared with genetically unscreened<br />

blastocysts.<br />

DESIGN: A retrospective analysis.<br />

MATERIALS AND METHODS: All biopsies and blastocyst transfers took<br />

place over a 2-year period. Embryos were cultured in G-1/G-2 media, supplemented<br />

with 10% SSS at 37.1 Cin6%CO 2 and 5% O 2 before vitrification on<br />

day 5 (d5) and/or day 6 (d6). Vitrification and warming were performed with<br />

Irvine’s vitrification/warming kit and a Rapid-I vitrification device. For PGS<br />

patients, assisted hatching was done on day 3, followed by trophectoderm biopsy<br />

on d5 and/or d6 and immediate vitrification. Standard 24-chromosome<br />

aneuploidy screening was performed by genetic reference laboratories. Clinical<br />

outcomes from mixed d5/d6 transfers were excluded from this study.<br />

RESULTS: A total of 439 embryos were biopsied for the <strong>17</strong>9 patients evaluated<br />

in this study. The proportion of euploid embryos was not different for<br />

embryos biopsied on d5 (48%) compared with those biopsied on d6 (53%).<br />

Similarly, an equivalent number (2.5%) of samples resulted in no DNA<br />

amplification when comparing d5 and d6 biopsies. Degraded DNA was<br />

observed in 5/203 d6 embryos but 0/236 d5 embryos.<br />

P-258 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

IMPLANTATION POTENTIAL IS RELATED TO EMBRYO QUAL-<br />

ITY REGARDLESS WHETHER THEY ARE FRESH OR<br />

VITRIFIED. A. Cobo, a A. Coello, a M. Meseguer, b J. Remohi. c a IVI Valencia,<br />

Valencia, Spain; b Clinical Embryology, Valencia, Spain; c Reproduction,<br />

Valencia, Spain.<br />

OBJECTIVE: Vitrification allows to the recovery of a high percentage of<br />

fully intact embryos in which the pre-freeze morphological characteristics<br />

are entirely preserved. The remaining question is if the ability to implant after<br />

vitrification also remains unaltered. This study was aimed to evaluate the<br />

impact of vitrification of day-3 embryos according to their morphological parameters<br />

on survival implantation (IR) and clinical pregnancy (CPR).<br />

DESIGN: Retrospective analysis.<br />

MATERIALS AND METHODS: Embryo grading was addressed in accordance<br />

with the Istanbul consensus on embryo assessment. All embryo transfers<br />

(ET) in which the embryos belonged to the same morphological category<br />

were selected from January 2012 to <strong>October</strong> 2013. ETs of mixed morphological<br />

categories were excluded. Results after cryotransfers (CT) of embryos<br />

Clinical Outcomes for Day of Cryopreservation With or Without PGS.<br />

Treatment<br />

Group<br />

Number<br />

of Patients<br />

Implantation<br />

Rate SEM<br />

Ongoing Pregnancy<br />

Rate SEM<br />

Day-5 PGS 27 79.6 a 9.3 70.4 a 8.9<br />

Day-6 PGS 20 67.5 a 8.9 65.0 a 9.4<br />

Day-5 no PGS 73 62.3 A 6.1 63.0 A 5.7<br />

Day-6 no PGS 59 47.4 B 5.4 50.8 A 5.9<br />

CONCLUSIONS: Previous studies have consistently demonstrated a significant<br />

increase (10-15%) in FET success when transferring non-PGS blastocysts<br />

vitrified on d5 compared with blastocysts vitrified on d6. Similarly,<br />

the implantation rate (IR) in this study was significantly higher for blastocysts<br />

vitrified on d5 compared with blastocysts vitrified on d6. This difference<br />

in IR between d5 and d6 frozen blastocysts was not observed for<br />

confirmed euploid embryos. The equivalent aneuploidy rates for d5 and<br />

e194 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


d6 biopsied blastocysts suggest that the difference in IR seen for non-PGS<br />

blastocysts is not likely due to an increase in the aneuploidy rate of d6 embryos.<br />

However, blastocysts vitrified without PGS is typically selected<br />

based on a minimum morphology grade; whereas, blastocysts biopsied<br />

for PGS are vitrified with much less emphasis on standard morphology.<br />

The aneuploidy rate of d5 and d6 blastocysts that met minimum<br />

morphology criteria (non-PGS blastocysts) may not necessarily be the<br />

same as those of PGS blastocysts when less attention is paid to standard<br />

morphology. Further studies are needed to confirm that there are also no<br />

difference in aneuploidy rates of d5 and d6 blastocysts that meet a minimum<br />

morphology score.<br />

P-260 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

COMPARISON OF VITRIFICATION MEDIUM WITH TREHALOSE<br />

AND HYDROXYPROPYLCELLULOSE TO THAT WITH SUCROSE<br />

AND COMPLEX PROTEIN SUPPLEMENTATION. T. Schlenker, a<br />

S. McCormick, a R. Smith, a W. B. Schoolcraft, b R. L. Krisher. c a Fertility<br />

Labs of Colorado, Lone Tree, CO; b Colorado Center for Reproductive Medicine,<br />

Lone Tree, CO; c National Foundation for Fertility Research, Lone<br />

Tree, CO.<br />

OBJECTIVE: Human oocytes have historically been difficult to cryopreserve.<br />

Optimizing vitrification solutions is one method to improve outcomes.<br />

Our objective was to compare the efficacy of two solutions for the vitrification<br />

of donor eggs, one containing trehalose and hydroxypropylcellulose<br />

(HPC) and the other with standard sucrose and complex protein supplementation.<br />

DESIGN: Prospective randomized donor egg split.<br />

MATERIALS AND METHODS: Eggs from each donor (n¼33) were<br />

randomly assigned to vitrification using either Cryotech (CT; n¼244) or an<br />

In House prepared (IH; n¼260) medium. IH vitrification solution consisted<br />

of standard Tissue Culture Medium 199 (TCM199) with HEPES, bicarbonate,<br />

serum protein substitute (SPS; Origio), sucrose, DMSO and ethylene glycol<br />

(EG). Cryotech vitrification solution also contains DMSO and EG, but<br />

substitutes trehalose for sucrose and HPC rather than SPS. Eggs were vitrified<br />

using the CryoTop (Kitazato).<br />

RESULTS: There was no difference in egg survival post-warm (CT,<br />

95.5%; IH, 96.2%). Fertilization was significantly higher for eggs vitrified<br />

and warmed in IH solutions (2PN/eggs survived, 85.2%) compared to those<br />

in CT (74.2%). There were no differences between CT and IH in the percentage<br />

of good quality blastocysts (GQ, R3BB) produced per 2PN (CT,<br />

54.3%; IH, 51.3%) or per total eggs vitrified/warmed (CT, 41.8%; IH,<br />

44.0%), or GQ blastocysts produced on D5 per 2PN (CT, 29.3%; IH,<br />

29.0%). Pregnancy data was analyzed from those cases in which only embryos<br />

produced from a single treatment were transferred (CT, n¼10, 1.3<br />

embryos/ET; IH, n¼14, 1.4 embryos/ET). There were no significant differences<br />

between % positive hCG (CT, 80%; IH, 64.3%) or % implantation<br />

(CT, 84.6%; IH, 52.6%).<br />

CONCLUSIONS: Both commercially available (Cryotech) and In House<br />

prepared vitrification media are equally effective for egg vitrification, despite<br />

differences in non-permeating cryoprotectants and protein content. Although<br />

IH solutions resulted in higher fertilization success, the proportion of blastocysts<br />

produced per fertilized zygote and per vitrified/warmed egg was equivalent.<br />

In summary, trehalose with HPC or sucrose with complex protein can<br />

both be used for egg vitrification, resulting in successful embryo development<br />

and pregnancy.<br />

P-261 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

INFLUENCE OF SEMINAL QUALITY IN DONOR EGG IVF PRO-<br />

GRAM USING VITRIFIED OOCYTES. B. Barros, a<br />

T. S. Domingues, b A. S. Belo, a R. Mazetto, b A. P. Aquino, b<br />

E. L. Motta. c a Embryology, Huntington Medicina Reprodutiva, Sao Paulo,<br />

Brazil; b Huntington Medicina Reprodutiva, Sao Paulo, Brazil; c Medical,<br />

Huntington Medicina Reprodutiva, Sao Paulo, Brazil.<br />

OBJECTIVE: To evaluate the influence of different seminal parameters to<br />

fertilize donor banking oocytes and correlate to blastocyst development, implantation<br />

and pregnancy rates.<br />

DESIGN: Retrospective study.<br />

MATERIALS AND METHODS: From July 2013 to April <strong>2015</strong>, 320 cycles<br />

of donor banking oocytes were fertilized by ICSI with ejaculated sperm and split<br />

into three different groups: G1: Normozoopermia (n¼77); G2: Oligoteratozoospermia<br />

(n¼34); G3: Teratozoospermia (n¼209). All oocytes were vitrified using<br />

open system and warmed following standard protocols. After fertilization,<br />

embryos were cultured as routine and blastocyst transfer placed on days 5 or<br />

6. Endometrium preparation was performed with 4 mg of estradiol valerate<br />

plus 800mg of micronized progesterone according to standard protocols.<br />

RESULTS: A total of 2846 vitrified donor oocytes were used. The mean<br />

numbers for sperm concentration were respectively (x106/mL): (107.4 vs 8.0<br />

vs 64.8, p


Technology. Role of assisted hatching in in vitro fertilization: a guideline.<br />

Fertil Steril. 2014 Aug;102(2):348-51. http://dx.doi.org/10.1016/<br />

j.fertnstert.2014.05.034. Epub 2014 Jun 18.<br />

3. Sunkara SK1, Siozos A, et al.The influence of delayed blastocyst formation<br />

on the outcome of frozen-thawed blastocyst transfer: a systematic<br />

review and meta-analysis. Hum Reprod. 2010 Aug;25(8):1906-15.<br />

P-263 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

SINGLE EMBRYO TRANSFER OF FROZEN-THAWED EMBRYOS<br />

IS ASSOCIATED WITH INCREASED MATERNAL<br />

COMPLICATIONS. T. Shavit, a G. Oron, b T. Tulandi, c W. Son, b<br />

H. Holzer, b W. Buckett. c a McGill University- MUHC Reproductive Center,<br />

Montreal, QC, Canada; b Department of Obstetrics and Gynecology, McGill<br />

University, Montreal, QC, Canada; c McGill University, Montreal, QC, Canada.<br />

OBJECTIVE: Cryopreservation of embryos allows transfer of a single embryo<br />

(SET) and storage of supernumerary embryos maximizing the cumulative<br />

pregnancy rates. It has been reported that IVF conceived singletons are<br />

prone to pregnancy complications including low birth weight (LBW), preterm<br />

deliveries (PTD) and small for gestational age (SGA). The purpose of<br />

our study was to compare the pregnancy outcome in singletons born after<br />

fresh or frozen-thawed single blastocyst transfer (SBT).<br />

DESIGN: A single center retrospective cohort study, a reproductive unit of<br />

a tertiary university health center.<br />

MATERIALS AND METHODS: We compared singleton live births resulting<br />

from transfer of fresh or frozen-thawed single blastocyst embryo (SBT).<br />

The primary outcomes were perinatal outcomes including SGA, LBW, very<br />

LBW, PTD, early PTD, large for gestational age (LGA), hospitalization at<br />

the neonatal intensive care unit, respiratory and gastrointestinal complications<br />

and congenital malformations. Maternal complications included preeclampsia,<br />

placenta previa, placental abruption, gestational diabetes mellitus (GDM) and<br />

chorioamnionitis. Adjustment for confounding factors was done.<br />

RESULTS: We studied 1886 fresh-SBTand 1200 FET-SBT cycles. SBTof<br />

fresh embryo resulted in a clinical pregnancy rate of 52.2% and live birth rate<br />

of 31.3% per embryo transfer (ET). These were significantly higher than<br />

34.4% clinical pregnancy rate and 13.7% live birth rate per ET in the FET<br />

group (p


OBJECTIVE: No significant differences in outcomes have been found using<br />

various protocols of endometrial preparation for frozen embryo transfer<br />

(FET), however existing data have not accounted for infertility diagnoses.<br />

This study compares clinical outcomes of women with different diagnoses<br />

in FET cycles using leuprolide acetate (L) suppression versus use of a<br />

mid-cycle antagonist (Ant), both with estrogen/progesterone supplementation.<br />

We hypothesized that prior L suppression versus mid-cycle Ant<br />

blockade may benefit women with polycystic ovarian syndrome (PCOS) or<br />

endometriosis.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: FET cycles performed at a single academic<br />

center from January 2012 to December 2014 were reviewed; previous<br />

FET cycles of identified women were also included. Patient demographics,<br />

infertility diagnosis, in-vitro fertilization and FET cycle information were<br />

collected. Characteristics and clinical outcomes of L versus Ant FET cycles<br />

were compared. We controlled for patient age, use of donor oocytes, BMI,<br />

use of ICSI, day of transfer and number of embryos transferred.<br />

RESULTS: FET cycles (n¼872) were identified and detailed data was<br />

available in 670 cycles of which L (n¼491) and Ant (n¼<strong>17</strong>9) approaches<br />

were prescribed. Further patient follow up and data collection continues to<br />

date. Patients undergoing Ant cycles were older (406.7 vs. 374.9 years,<br />

p


P-269 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

OOCYTE BIOLOGY<br />

SIMILAR LEVELS OF PREMATURE SEPARATION OF SISTER<br />

CHROMATIDS IN THE MOUSE AND HUMAN OOCYTE RE-<br />

VEALED BY NEXT GENERATION SEQUENCING. N. R. Treff, a<br />

R. L. Krisher, b X. Tao, c H. Garnsey, c C. Bohrer, c E. Silva, b J. Landis, c<br />

R. T. Scott, a T. K. Woodruff, d F. E. Duncan. e a RMA, NJ, NJ; b National Foundation<br />

for Fertility Research, LoneTree, CO; c FAEEC, Basking Ridge, NJ;<br />

d Northwestern University, Chicago, IL; e University of Kansas Medical Center,<br />

Kansas City, KS.<br />

OBJECTIVE: Accumulating evidence supports the relevance of the mouse<br />

model to human age-related oocyte aneuploidy. Still, many questions remain<br />

including whether there is chromosome-specific susceptibility and a higher<br />

incidence of premature separation of sister chromatids (PSSC) compared<br />

to classical nondisjunction (ND) in the mouse as in the human. The present<br />

study develops a method to address these and other ongoing questions associated<br />

with the etiology of aneuploidy.<br />

DESIGN: Observational.<br />

MATERIALS AND METHODS: GenomePlex WGA4 (Sigma) and next<br />

generation sequencing (NGS) based (Ion Proton) comprehensive chromosome<br />

screening (CCS) was performed on known trisomy 16 single mouse<br />

embryonic fibroblasts (positive controls), and paired 1st polar body (PB)<br />

and remaining oocytes (to evaluate reciprocal aneuploidies) and whole<br />

blastocysts from aged mice.<br />

RESULTS: Trisomy 16 was clearly observed with NGS CCS in all positive<br />

control (fibroblast) samples tested. 40 paired PB and oocyte samples from<br />

mice aged 15-19 months displayed an aneuploidy incidence of 15% (chromosomes<br />

2, 5, 7, 9, 11, 12, and 15), and PSSC in 5 of the 6 cases (83%). Reciprocal<br />

errors were observed in the paired PBs and oocytes in all cases. 10 of 30<br />

blastocysts (33%) from mice aged 13-14 months displayed aneuploidy<br />

involving chromosomes 1, 2, 3, 5, 8, 10, 12, 13, 14, <strong>17</strong>, and 19.<br />

CONCLUSIONS: This study confirms the validity of a novel method for<br />

aneuploidy screening in mouse polar bodies, oocytes, and embryos based<br />

on detection of trisomy 16 in a positive control and reciprocal errors in sister<br />

PBs and oocytes. It was also possible to distinguish PSSC from ND, which<br />

indicated percentages similar to those observed in human oocytes. Increased<br />

prevalence of aneuploidy in the blastocysts in this study may be related to<br />

additional contribution of meiosis II errors or strain specific differences in<br />

susceptibility. This model system and NGS CCS method of analysis represents<br />

an important new tool to better understand the etiology of and factors<br />

that may influence aneuploidy.<br />

P-270 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

AN FSH-LOWERING ACTIVIN DISRUPTING THERAPY IN-<br />

CREASES YIELD OF HEALTHY OOCYTES IN A MOUSE MODEL<br />

OF FEMALE MIDLIFE REPRODUCTIVE<br />

AGING. L. R. Bernstein, a,b,c,d A. Mackenzie, e C. L. Chaffin, f<br />

I. J. Merchenthaler. g a Pregmama, LLC, Montgomery Village, MD; b Epidemiology<br />

and Public Health, University of <strong>Maryland</strong> School of Medicine,<br />

<strong>Baltimore</strong>, MD; c Gynecology and Obstetrics, Johns Hopkins School of Medicine,<br />

<strong>Baltimore</strong>, MD; d Veterinary Integrative BioSciences, Texas A & M<br />

College of Veterinary Medicine, College Station, TX; e University of <strong>Maryland</strong>,<br />

<strong>Baltimore</strong>, <strong>Baltimore</strong>, MD; f University of <strong>Maryland</strong> School of Medicine,<br />

<strong>Baltimore</strong>, MD; g Epidemiology, <strong>Baltimore</strong>, MD.<br />

OBJECTIVE: Women with diminished ovarian reserve often exhibit poor<br />

response to ovarian stimulation when given FSH stimulation. Alternative<br />

treatments that stimulate oocyte production may reduce cycle cancellation<br />

and improve pregnancy rates.<br />

DESIGN: We developed SAMP8 mice as model with human-like reproductive<br />

aging attributes, including diminished oocyte yield, elevated FSH,<br />

increased rates of oocyte spindle misalignments, and diminished fertility<br />

by midlife (age 7 months). ActRIIB:Fc is an activin decoy receptor that sequesters<br />

activin and suppresses activin signaling. It was given to midlife<br />

SAMP8 for 3 weeks, the duration of oocyte growth. A second test group<br />

was treated for 3 weeks with high FSH activity in the form of PMSG.<br />

MATERIALS AND METHODS: Freshly ovulated oocytes were collected<br />

after hCG injection. Total yields of oocytes and of viable oocytes were scored.<br />

RESULTS: ActRIIB:Fc lowered FSH in midlife SAMP8 to the levels of<br />

young SAMP8. ActRIIB:Fc increased oocyte yield by 3.7/mouse, from<br />

10.64 to 14.3/mouse (P¼0.0031), an increase of 34.7%. The yield was<br />

greater than for young SAMP8. ActRIIB:Fc did not affect the yield of<br />

non-viable oocytes. ActRIIB:Fc increased the yield of viable oocytes by<br />

2.9/mouse, from 9.86 to 12.7 oocytes/mouse (P¼0.0051), an increase of<br />

29.4%, a number greater than that of young mice.PMSG had no effect on<br />

the total yield of oocytes in midlife SAMP8 (10.64 vs. 13.28/mouse;<br />

P¼0.9829, NS), though it increased the total yield in young SAMP8 by 7.4<br />

oocytes, from 12.1 to 19.5/mouse (P¼0.0063). This is analogous to DOR<br />

AMA women, who are often more refractory to FSH stimulation than their<br />

younger peers. PMSG caused a redistribution of oocytes from 9.86/10.64<br />

(92.7%) viable oocytes in untreated mice, to just 4.24/13.28 viable oocytes<br />

(31.9%; P 0.05 in bivariate analysis). Moreover,<br />

there was no significant correlation between blood and oocyte telomere length<br />

from the same patient (R2 < 0.01). No difference was detected in mean telomere<br />

length from blood nor oocytes with pregnancy outcomes.<br />

CONCLUSIONS: Telomere DNA content from single eggs was associated<br />

with ovarian function independent of age, suggesting telomere DNA may be<br />

a useful biomarker of reproductive age. This study was limited by the use of<br />

spare, immature oocytes, and pregnancy outcomes could not be directly<br />

e198 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


elated to telomere length from the transferred embryos. This is the first study<br />

to compare telomere length in human oocytes with somatic tissues. Somatic<br />

tissues from the same individual exhibit a consistent decline in telomere<br />

length with age, but we did not find such an association between oocytes<br />

and leukocyte telomere DNA. Like sperm, oocytes likely represent a unique<br />

niche in telomere biology.<br />

P-272 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

TELOMERE ATTRITION IN GERMINAL VESICLE ARRESTED<br />

HUMAN OOCYTES. K. Kalmbach D. L. Keefe. New York University<br />

Langone Medical Center, New York, NY.<br />

OBJECTIVE: Telomere length is maintained in the ‘immortal’ germline<br />

yet evidence from mouse and human studies suggests that telomere length<br />

in oocytes is significantly shorter than in somatic tissues. We have proposed<br />

that telomere attrition in oocytes contributes to infertility in women, but the<br />

mechanisms underlying the effects of telomere length on reproduction in<br />

women are poorly understood. Telomere shortening arrests mitotic cells so<br />

we examined its effects during meiosis.<br />

DESIGN: Prospective observational study.<br />

MATERIALS AND METHODS: Human germinal vesicle oocytes (GVs)<br />

were collected from patients at an academic fertility center and in-vitro<br />

matured for up to 48 hours following oocyte retrieval. Samples were assessed<br />

for polar body (PB) extrusion and germinal vesicle breakdown (GVBD) at 24<br />

and 48 hours, then fixed for telomere length analysis. If no PB was observed<br />

at 48 hours, the sample was processed and staged as meiosis I (M1) or GV by<br />

absence or presence of GV nuclear structure respectively. Oocytes were<br />

denuded, treated with pronase and washed to exclude granulosa cells. One<br />

oocyte was delivered to each tube assuring no contamination with follicular<br />

cells. Telomere length was measured by single cell telomere pre-PCR amplification<br />

qPCR (SCT-pqPCR) and reported as telomere DNA (T) normalized<br />

to reference DNA (R) in a T/R ratio. T/R measures were log-transformed for<br />

statistical analyses to obtain a normal distribution.<br />

RESULTS: Mean telomere length of GV arrested oocytes (n ¼ 35; 0.09 <br />

0.44) was significantly less than that of oocytes that had matured to metaphase<br />

II (n ¼ 74; 0.50 0.54) oocytes (p < 0.001, paired t-test). Averaged<br />

per patient, mean telomere length of GV (n ¼ 23; 0.00 0.43) and M2 (n<br />

¼ 32; 0.50 0.43) also differed significantly (p < 0.001). Telomere length<br />

of oocytes that had reached M1 by 48 hours was similar to that of M2 oocytes<br />

(n ¼ 9; 0.28 0.68; p ¼0.38).<br />

CONCLUSIONS: Arrested germinal vesicle oocytes exhibited less telomere<br />

DNA content than MI or MII oocytes. Telomere length influences progression<br />

through meiotic cell cycle, and this pilot study suggests oocyte<br />

telomere length may be an important regulator of oocyte maturation. Future<br />

studies will determine whether telomeres regulate mitotic and meiotic cell<br />

cycle via common regulatory pathways.<br />

P-273 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

OOCYTE SECRETION IS REGULATED BY SOMATIC CELLS DUR-<br />

ING OOCYTE MATURATION AND MEDIATES CUMULUS CELL<br />

FUNCTION. H. Cakmak, F. Franciosi, M. Cedars, M. Conti. UCSF<br />

Dept. of Ob/Gyn, San Francisco, CA.<br />

OBJECTIVE: De novo mRNA synthesis ceases during the final stages of<br />

oocyte maturation and previously synthesized mRNAs are translated according<br />

to a well-orchestrated program of recruitment to the polysomes. The aim<br />

of this study is to investigate the regulation of oocyte mRNA translation and<br />

protein secretion during maturation, using interleukin (IL)-7 as the prototypic<br />

secreted factor, and to determine the role of this oocyte secreted factor<br />

in cumulus cell (CC) function.<br />

DESIGN: Prospective.<br />

MATERIALS AND METHODS: Studies involving in vivo/in vitro maturation<br />

and fertilization of oocytes from <strong>21</strong> day old C57BL/6 mice, and human<br />

follicular fluid before and after hCG administration.<br />

RESULTS: Microarray analysis of oocyte polysomal fraction showed that<br />

191 of 7600 polysome-bound mRNAs encoded for secreted proteins. Specifically,<br />

72 transcripts were constitutively present in the polysome fraction<br />

throughout in vivo oocyte maturation, while 80 were decreased and 39 were<br />

increased. IL-7 was selected as the prototype because its translation increased<br />

the most during oocyte maturation. In qPCR analysis, the level of polysomebound<br />

IL-7 mRNA enhanced in MI oocyte compared to GVoocyte, and further<br />

increased with maturation to MII. Total IL-7 mRNA levels did not change during<br />

oocyte maturation. Renilla luciferase reporter under the control of IL-7<br />

3’UTR was injected into cumulus enclosed oocytes (CEOs). Translation of<br />

the IL-7 reporter significantly increased as the oocytes progressed from GV<br />

to MII, and was further stimulated by amphiregulin (AREG), a somatic cellderived<br />

EGF-like growth factor that accumulates in the follicle after LH surge.<br />

AREG-induced effect was not detected when the oocytes were denuded prior<br />

to stimulation. IL-7 protein secretion increased with oocyte maturation in<br />

CEOs and, to a less extent in denuded oocytes (DOs). AREG further enhanced<br />

IL-7 secretion in CEOs, but not in DOs. IL-7 was not detectable in spent media<br />

of CC only cultures. In humans, IL-7 levels were significantly higher in the<br />

follicular fluid containing MII oocyte (n¼42) compared to those containing<br />

GV oocyte (n¼8). IL-7 levels positively correlated with AREG concentration<br />

in the follicular fluid. After fertilization, IL-7 secretion diminished in both<br />

mouse and human embryos. In mouse, IL-7 receptor mRNA expression in<br />

CCs increased during oocyte maturation. IL-7 increased the proliferation of<br />

CCs, but did not affect CC expansion.<br />

CONCLUSIONS: Oocyte secretion during maturation is highly dynamic<br />

and regulated mRNA translation is the molecular mechanism underlying<br />

this timed secretion. The oocyte secretion is sensitive to somatic cues and<br />

modulates CC function supporting the concept that these regulated secretions<br />

are the part of cross-talk between the oocyte and surrounding CCs, which is<br />

crucial for the oocyte developmental competence.<br />

Supported by: RO1-GM097165 and T32 HD007263-28.<br />

P-274 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

IQGAP3 IS A CANDIDATE PROTEIN FOR CORRECT SPINDLE<br />

POSITIONING DURING MEIOTIC MATURATION IN MAMMA-<br />

LIAN OOCYTES. L. A. McGinnis J. P. Evans. Biochemistry and Molecular<br />

Biology, Johns Hopkins Bloomberg School of Public Health, <strong>Baltimore</strong>,<br />

MD.<br />

OBJECTIVE: The progression of the mammalian oocyte through its two<br />

meiotic divisions is accompanied by organization and orientation of the<br />

meiotic spindle, as well as remodeling of the actin-rich cortex. These events<br />

are crucial for oocyte quality because spindle positioning is key for equal<br />

segregation of chromosomes and unequal distribution of cytoplasm to<br />

daughter cells.<br />

DESIGN: IQ motif-containing GTPase activating protein 3 (IQGAP3)<br />

is a protein of interest because it is implicated in cytokinesis and contractile<br />

ring organization in model organisms. The domain structure of IQ-<br />

GAP3 includes an actin-binding domain at the C-terminus and a<br />

microtubule-binding domain at the N-terminus, leading us to hypothesize<br />

that IQGAP3 may serve as a scaffold between the microtubule-based<br />

meiotic spindle and the actin-rich cortex. Mammals have three IQGAPs.<br />

Transcriptome databases suggest that Iqgap3 mRNA is enriched in mouse<br />

oocytes relative to other tissues, prompting us to examine IQGAP3 function<br />

in oocytes.<br />

MATERIALS AND METHODS: Mouse oocyte collection and in vitro<br />

maturation: Prophase I oocytes are collected from ovaries of 6-8-week-old<br />

CF1 female mice and cultured in medium containing dibutyryl-cAMP (dbcAMP)<br />

to maintain meiotic arrest. Oocytes are then allowed to mature to<br />

met I or to met II. Microinjection for RNAi-mediated knockdown: Prophase<br />

I oocytes are microinjected with Iqgap3-targeting siRNA. Oocytes are<br />

cultured in dbcAMP to maintain meiotic arrest for 48 hours to allow sufficient<br />

time for RNA degradation and protein turnover. Knockdown is assessed<br />

by RT-PCR, immunofluorescence or immunoblotting. Immunofluorescence:<br />

The zona pellucida is removed through a brief exposure to acidic medium.<br />

Oocytes and eggs are fixed in paraformaldehyde and stained using anti-IQ-<br />

GAP3 and anti-tubulin antibodies.<br />

RESULTS: IQGAP3 is localized to the cortex of prophase I oocytes, and<br />

associated with the spindle in met II eggs. IQGAP3-deficient oocytes have<br />

a range of phenotypes including failure to recruit the metaphase I spindle to<br />

the cortex. This phenotype supports our hypothesis that IQGAP3 plays a<br />

scaffolding role between the spindle and the cortex such that in an IQ-<br />

GAP3-deficient oocyte, correct positioning of the spindle adjacent to the<br />

cortex does not occur. Studies in HeLa cells show that IQGAP3 interacts<br />

with anillin and this interaction is required for the localization of IQGAP3<br />

to the contractile ring. Ongoing experiments are investigating anillin in IQ-<br />

GAP3-deficient oocytes.<br />

CONCLUSIONS: It is crucial that the meiotic spindle is positioned<br />

correctly through the two rounds of cytokinesis to ensure equal segregation<br />

of chromosomes and unequal segregation of cytoplasmic stores between<br />

the egg and polar body. Our work shows that IQGAP3 plays an important<br />

role in spindle positioning.<br />

Supported by: NIH R03 HD074773 to JPE.<br />

FERTILITY & STERILITY Ò<br />

e199


P-275 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

NON-INVASIVE PREDICTION OF BLASTOCYST FORMATION BY<br />

DAY THREE EMBRYO CULTURE MEDIUM MASS SPECTROM-<br />

ETRY LIPID FINGERPRINTING. D. P. Braga, a,b,c A. S. Setti, a,b,d<br />

E. C. Cabral, e M. N. Eberlin, e E. G. Lo Turco, c A. Iaconelli, Jr., f,b<br />

E. Borges, Jr., f,b a Scientific Department, Fertility Medical Group, Sao Paulo,<br />

Brazil; b Scientific Department, Instituto Sapientiae - Centro de Estudos e<br />

Pesquisa em Reproduç~ao Assistida, Sao Paulo, Brazil; c Urology Deparment,<br />

Universidade Federal de S~ao Paulo (UNIFESP), Sao Paulo, Brazil; d Faculdade<br />

de Ciencias Medicas da Santa Casa de Sao Paulo, Sao Paulo, Brazil;<br />

e Chemistry Institute - ThoMSon Mass Spectrometry Laboratory, Universidade<br />

Estadual de Campinas, Campinas, Brazil;<br />

f Clinical Department,<br />

Fertility Medical Group, Sao Paulo, Brazil.<br />

OBJECTIVE: To identify lipid markers of blastocyst formation by day<br />

three culture medium mass spectrometry (MS) fingerprinting.<br />

DESIGN: Case-control study.<br />

MATERIALS AND METHODS: For this study 50 embryo culture media<br />

samples were harvested on the day three, from patients undergoing day five<br />

embryo transfers. Embryos were split into groups based on their degree of<br />

expansion and hatching status on day five (Complete-Blastocysts, n¼25<br />

and No-Blastocysts, n¼25) and its secretomes were analysed by MS. Mass<br />

spectra fingerprinting was acquired using a Q-Tof spectrometer (LC-MS,<br />

Agilent 6550 iFunnel Q-TOF) equipped with an automated injector. Data<br />

were analysed using the principal component analysis (PCA) followed by<br />

a partial least square discrimination analysis (PLS-DA), combined with variable<br />

influence in the projection (VIP) scores. The statistical analysis was<br />

performed using Metabo-Analyst 2.0 (http://www.metaboanalyst.ca).<br />

RESULTS: Overall, 1657 ions were observed. When the univariate analysis<br />

was performed, 165 ions were observed to be differentially expressed between<br />

the groups, with a fold chance R 4x and p


DESIGN: Basic research comparative study.<br />

MATERIALS AND METHODS: Cumulus cells clumps (CCs) were isolated<br />

from MII oocytes collected from patients &lt 35 years (‘‘younger’’,<br />

n¼10) and &gt 40 years old (‘‘older’’, n¼11) undergoing ICSI for male factor<br />

infertility. CCs were individually processed for RNA extraction, library preparation<br />

and sequenced on Illumina HiSeq 2000 platform. Gene enrichment<br />

and gene ontology analysis were used to define upregulated gene networks<br />

and interactions in the two cohorts. Significance of differentially expressed<br />

genes was assigned when both p value and false discovery rate were<br />


de Ciencias Medicas da Santa Casa de Sao Paulo, Sao Paulo, Brazil; f IVF<br />

Laboratory, Fertility Medical Group, Sao Paulo, Brazil; g Biotechnology<br />

institute, Universidade de Caxias do Sul, Caxias do Sul, Brazil.<br />

OBJECTIVE: To investigate whether patients presenting with endometriosis<br />

are more likely to have oocyte with dysmorphisms, embryos with poor<br />

development and decreased assisted reproductive technologies (ART) cycles’<br />

outcome.<br />

DESIGN: Case-Control study.<br />

MATERIALS AND METHODS: The influence of the presence of endometriosis<br />

on oocyte and embryo quality and blastocyst formation chance was evaluated.<br />

Moreover, the presence of endometriosis was correlated with cycles’<br />

characteristics. To avoid any bias concerning the age of the female, in the first<br />

analysis only patients % 36 years old were included and the cycles were split<br />

into: endometriosis infertility cycles (n¼431 and 3<strong>17</strong>2 oocytes/embryos) and<br />

others (n¼2510 cycles and 24480 oocytes/embryos). In a second analysis,<br />

endometriosis infertility cycles (n¼669 and 4993 oocytes/embryos) and tubal<br />

factor infertility cycles (n¼380 and 5029 oocytes/embryos) were compared.<br />

RESULTS: For the first analysis, the number of retrieved oocytes<br />

(10.6<strong>21</strong>.2 vs 14.6<strong>21</strong>.1, p


Vitrified OVA Banking Success Rates.<br />

2011 2012 2013 2014 OVERALL<br />

NUMBER 5 19 34 38 96<br />

OVA SURVIVAL RATE (%) 59.0 89.5 93.7 93.8 90.5<br />

FERTILIZATION RATE (%) 73.9 73.6 75.8 79.5 76.8<br />

CLINICAL PREGNANCY 20.0 63.2 52.9 63.2 57.3<br />

RATE (%)<br />

NO ET RATE (%) 40.0 0.0 11.8 7.9 9.4<br />

Reference:<br />

1. Apollini Lm, Wyatt Sl, Collazo I, Et al. Preparation Of Warming Medium<br />

For Vitirfied Specimens - Effects On The Final Warming Temperature.<br />

2014 Aab Conference And Crm Symposium; Abstracts:15.<br />

P-285 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

MICROFILAMENTS PLAY CRITICAL ROLE ON MITOCHON-<br />

DRIAL TRAFFIC IN PORCINE GV OOCYTE. T. Yamochi,<br />

S. Hashimoto, A. Amo, H. Goto, M. Yamanaka, M. Inoue, Y. Nakaoka,<br />

Y. Morimoto. IVF Namba Clinic, Osaka, Japan.<br />

OBJECTIVE: Oocyte maturation requires a variety of ATP-dependent reactions,<br />

such as germinal vesicle breakdown, spindle formation, and polar<br />

body extrusion, which is required for fertilization. Mitochondria are accordingly<br />

expected to be localized to subcellular sites of ATP utilization.<br />

Although cytoskeleton-dependent traffic for mitochondria has been studied<br />

extensively in somatic cells, the mechanism of mitochondrial dynamics in<br />

mammalian oocytes remains obscure. This study describes dynamic aspects<br />

of mitochondria in porcine oocytes at germinal vesicle stage.<br />

DESIGN: Basic research study.<br />

MATERIALS AND METHODS: Mitochondria in donor oocytes were<br />

stained with MitoTracker-Orange (MTO). Donor oocytes were centrifuged<br />

at 10,000 x g and 37 C for 15 min. Mitochondria-enriched ooplasm were micropunctured<br />

and injected into either central or subcortical area of recipient<br />

oocytes. Mitochondria-injected oocytes were cultured with or without colcemid,<br />

cytochalasin B or cytochalasin D. The image of mitochondrial dynamics<br />

in the recipient oocytes was captured every 15 min using a confocal microscopy<br />

for 15 hours, and analyzed quantitatively with ImageJ. Relative area<br />

and distribution of fluorescent mitochondria in recipient oocytes were calculated<br />

on the basis of their initial values at 0 h of culture. Total of 227 mitochondria-injected<br />

oocytes were observed.<br />

RESULTS: Mitochondria injected centrally moved unidirectionally to<br />

subcortical area and those injected subcortically moved along plasma membranes.<br />

The area occupied with MTO fluorescence increased significantly<br />

during 15 hours. The area occupied with MTO fluorescence at both sites of<br />

injection was not affected by colcemid, but inhibited strongly by cytochalasin<br />

B and cytochalasin D.<br />

CONCLUSIONS: This study shows that central mitochondria move from<br />

central to subcortical area and those in subcortical area move along plasma<br />

membranes, and suggests that microfilaments play critical role in mitochondrial<br />

traffic in porcine GV oocyte. The method established in this study may<br />

permit studies of the pathophysiology of intracellular traffic of mitochondria<br />

and other organelles in oocytes from patients with infertility.<br />

P-286 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

OTHER: ART - CLINICAL<br />

EXPANDING ESET USE TO ALLWOMEN


and the odds of having a blastocyst transfer, high quality embryos, or ongoing<br />

pregnancy.<br />

TABLE 1. Repeat length and age adjusted associations (n¼4590).<br />

Variable<br />

Adjusted Association Per<br />

Single Increase in CGG<br />

Repeat Length<br />

P value<br />

FSH (U/L) -0.26 (-0.46 - -0.01) 0.01<br />

AMH (ng/ml) 0.02 (-0.03 - 0.06) 0.36<br />

AFC 0.05 (-0.02 - 0.12) 0.16<br />

Estradiol at trigger (pg/ml) 10.0 (3.1 - <strong>17</strong>.0) 0.005<br />

Total gonadotrophins (IU) -12.1 (-25.2 - 1.0) 0.07<br />

Total Follicles 0.04 (0.02 - 0.07) 0.03<br />

Oocytes retrieved 0.04 (-0.01 - 0.09) 0.07<br />

Mature oocytes 0.04 (0.01 - 0.08) 0.04<br />

# of vitrified embryos 0.02 (0.01 - 0.03) 0.03<br />

# of good quality blastocysts 0.02 (0.01 - 0.04) 0.05<br />

CONCLUSIONS: In a population undergoing autologous IVF, increasing<br />

FMR1 repeat size was positively associated with ovarian reserve and<br />

response to stimulation for repeats less than 200. However, FMR1 repeat<br />

number was not associated with pregnancy. These findings suggest that previously<br />

described diminished ovarian reserve with premutation may not<br />

apply in all populations. Instead of a ‘‘dose response continuum’’ of toxic<br />

FMR1 mRNA, there may instead be a threshold effect or influence from as<br />

yet uncharacterized factors, such as favorable X-inactivation patterns within<br />

the ovary.<br />

Supported by: Work Supported in part by the NICHD intramural research<br />

program.<br />

P-288 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

COMPUTER-AUTOMATED TIME-LAPSE TEST RESULTS ARE<br />

PREDICTIVE OF PREGNANCY FOLLOWING BLASTOCYST<br />

TRANSFER. B. Behr, a A. N. Beltsos, b B. Yee, c C. Kingsland, d<br />

C. A. Benadiva, e J. Liebermann. b a Stanford Fertility and Reproductive Medicine<br />

Center, Palo Alto, CA; b Fertility Centers of Illinois, Chicago, IL; c Reproductive<br />

Partners, Redondo Beach, CA;<br />

d Hewitt Fertility Center,<br />

Liverpool, United Kingdom; e University of Connecticut, Farmington, CT.<br />

OBJECTIVE: The Eeva Test is an automated time-lapse enabled test that<br />

generates predictions based on cell division timings. As blastocyst development<br />

was used to build the model, Eeva Test has been shown to improve Day<br />

3 embryo selection (1). However, there is evidence suggesting by incorporating<br />

key cell division timings Eeva results reflect underlying molecular<br />

health of embryos (2,3). The objective of the study is to assess whether<br />

Eeva Test can be applied to blastocyst selection.<br />

DESIGN: Retrospective multi-center study.<br />

MATERIALS AND METHODS: Total of 342 patients underwent blastocyst<br />

transfers from 7 centers consented to use Eeva Test in IVF. Patients were<br />

separated into three groups, at least 1 Eeva High (H) transferred, at least 1<br />

Medium (M) transferred and only Low (L) transferred, and clinical pregnancy<br />

and implantation rates were compared. Spearman rank were used to<br />

check correlations among the variables of interest (age, #eggs, #2PN,<br />

morphology, Eeva results, #embryos transferred). Univariate logistic regression<br />

was used to assess predictive power of the variables against pregnancy.<br />

Multivariate logistic regression with backward selection was used to test the<br />

predictive value of Eeva results with other key pregnancy predictors<br />

included.<br />

RESULTS: Patients with at least 1 Eeva H blastocyst transferred had<br />

significantly higher pregnancy and embryo implantation rates than patients<br />

with only L blastocysts transferred (54% vs 34%, p¼0.002 and 46% vs<br />

29%, p¼0.002), while patient age and #embryos transferred was similar between<br />

groups. Univariate logistic regression showed that age, #2PN and<br />

Eeva results (H vs L) are significantly correlated with pregnancy<br />

(p¼0.008, 0.02, 0.02), while Spearman correlations showed Eeva results<br />

were not correlated with the other variables. Multivariate model revealed<br />

that Eeva results remain as significant predictor of pregnancy after adjusting<br />

for age and blastocyst morphology while the other variables were not<br />

significant.<br />

Logistic regression result.<br />

Variable Odds ratio P-value 95% CI (odds ratio)<br />

Eeva results (H vs. L) 2.15 0.007 (1.23,3.74)<br />

Age 0.92 0.005 (0.86,0.97)<br />

Blastocyst morphology 0.46 0.002 (0.29,0.75)<br />

#embryos transferred 1.41 0.14 (0.88,2.42)<br />

#eggs 0.98 0.69 (0.92,1.05)<br />

#2PN 1.03 0.49 (0.94,1.12)<br />

CONCLUSIONS: It is well known that age and blastocyst morphology are<br />

predictors of pregnancy. The current study is the first to show early cell division<br />

timing parameters have independent predictive value for blastocyst<br />

transfer outcome. Not only does it indicate Eeva Test may be used to improve<br />

blastocyst selection, but it also supports the hypothesis that normal embryo<br />

development may follow programmed timings at early stage (2).<br />

References:<br />

1. Diamond et al, Using the Eeva TestÔ adjunctively to traditional day 3<br />

morphology is informative for consistent embryo assessment within a<br />

panel of embryologists with diverse experience. J Assist Reprod Genet.<br />

<strong>2015</strong>;32(1):618.<br />

2. Wong et al, Non-invasive imaging of human embryos before embryonic<br />

genome activation predicts development to the blastocyst stage. Nature<br />

Biotech. 2010; 28(10):1115-<strong>21</strong>.<br />

3. Chavez et al, Dynamic blastomere behaviour reflects human embryo<br />

ploidy by the four-cell stage.Nature Commun. 2012;3:1251.<br />

Supported by: Progyny (formerly Auxogyn).<br />

P-289 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE IMPACT OF COMPREHENSIVE CHROMOSOME<br />

SCREENING ON EMBRYO SELECTION OF DAY 5<br />

BLASTOCYSTS. M. Schweitz, a S. McReynolds, a M. McGarvey, a<br />

J. M. Stevens, b W. B. Schoolcraft, c M. Katz-Jaffe. c a Fertility Genetics,<br />

Lone Tree, CO; b Fertility Labs of Colorado, Lone Tree, CO; c Colorado Center<br />

for Reproductive Medicine, Lone Tree, CO.<br />

OBJECTIVE: Embryo selection is a critical component of successful<br />

infertility treatment. With the development of robust blastocyst culture, biopsy<br />

and vitrification techniques, comprehensive chromosome screening<br />

(CCS) is being applied for embryo selection with significant increases in<br />

live birth rates following a euploid blastocyst transfer. The aim of this study<br />

was to examine the relationship between D5 blastocyst morphology, chromosome<br />

constitution, and the impact on embryo selection.<br />

DESIGN: Research study.<br />

MATERIALS AND METHODS: A large consecutive cohort of 2,670 blastocyst<br />

CCS cycles performed from 2010-2014 were included for analysis. Blastocysts<br />

were graded using a standard morphological system on D5 of<br />

development prior to trophectoderm biopsy for CCS using either qPCR<br />

(RMA-NJ) or aCGH (Illumina). Chi-Square analysis for independence was performed<br />

to determine the relationship between the highest grade D5 blastocyst in<br />

an embryo cohort and its chromosome constitution, with significance at P


P-290 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

WITHDRAWN<br />

While IVF is more cost effective for the general population, cost effectiveness<br />

is most comparable in the less than 35 year old age group, in which the net monetary<br />

benefit of the strategies is balanced. Although more expensive, the IVF<br />

strategy is favored as cost effective due to decreased cycle cancellations and<br />

the ability to freeze surplus embryos for an additional transfer.<br />

CONCLUSIONS: Although IVF is a more cost effective ART strategy, the<br />

discounted patient charge of the MET protocol makes it a reasonable alternative<br />

to traditional IVF in younger women where clinical pregnancy rates are<br />

highest with MET.<br />

Table 1. Characteristics of patients undergoing MET and IVF cycles.<br />

Characteristic<br />

MET less<br />

than<br />

35 years<br />

(N¼108)<br />

IVF less<br />

than<br />

35 years<br />

(N¼146)<br />

MET 35-40<br />

years<br />

(N¼101)<br />

IVF 35-40<br />

years<br />

(N¼115)<br />

MET<br />

greater<br />

than 40<br />

years<br />

(N¼73)<br />

IVF greater<br />

than 40 years<br />

(N¼54)<br />

Age - mean (SD) 29.8 2.7 30.8 3.0 37.7 1.8 37.4 2.0 42.9 1.8 41.9 1.0<br />

Oocytes retrieved 2.6 1.6 13.6 7.7 2.2 1.3 10.8 6.4 2.5 1.8 9.6 6.5<br />

- mean (SD)<br />

Embryos transferred 1.7 0.7 1.9 0.6 1.7 0.7 2.5 0.9 1.8 1.0 3.2 1.4<br />

- mean (SD)<br />

Clinical pregnancy 26 (24.1) 53 (36.3) <strong>17</strong> (16.8) 46 (40.0) 3 (4.1) 11 (20.4)<br />

per cycle (%)<br />

Live birth per 28 (25.9) 68 (46.6) 11 (10.9) 40 (34.8) 1 (1.4) 9 (16.7)<br />

cycle (%)<br />

# Multiples (5 / 0) 18 / 2) (0 / 0) (8 / 0) (0 / 0) (2 / 0)<br />

(twins/triplets)<br />

Cancelled - 15 (13.8) 0 (0) 18 (<strong>17</strong>.8) 0 (0) 12 (16.4) 0 (0)<br />

no retrieval (%)<br />

Cancelled -<br />

no transfer (%)<br />

15 (13.8) 4 (2.7) 18 (<strong>17</strong>.8) 3 (2.6) 10 (13.7) 1 (1.9)<br />

P-292 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

P-291 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

IVF PROVES COST EFFECTIVE COMPARED TO MINIMAL STIM-<br />

ULATION EMBRYO TRANSFER, BUT MINIMAL STIMULATION<br />

IS A COST NEUTRAL ALTERNATIVE FOR WOMEN UNDER AGE<br />

35. N. Resetkova, a K. C. Humm, b A. Penzias, c D. Sakkas, d<br />

B. M. Lannon. e a Obstetrics and Gynecology, Beth Israel Deaconess Medical<br />

Center, Boston, MA; b Obstetrics and Gynecology, George Washington University<br />

School of Medicine, Washington, DC; c Boston IVF / Harvard Medical<br />

School, Waltham, MA; d Boston IVF, Waltham, MA; e Boston IVF, South<br />

Portland, ME.<br />

OBJECTIVE: In vitro fertilization (IVF) using injectable gonadotropins is<br />

an effective albeit expensive treatment for infertility. A minimal stimulation<br />

embryo transfer (MET) protocol may offer a less costly alternative. Our<br />

objective was to compare treatment outcomes and cost stratify by age for<br />

MET and traditional IVF cycles. Decision analysis was performed to identify<br />

which subset of patients may benefit most from MET.<br />

DESIGN: Retrospective cohort study, Decision analysis.<br />

MATERIALS AND METHODS: We collected data on demographics,<br />

treatment protocols, and outcomes for women who underwent oocyte<br />

retrieval from January 2011 through April 2013. Cost data for IVF and<br />

MET cycles included stimulation medications and were expressed in terms<br />

of typical patient charges expressed in <strong>2015</strong> US dollars. Decision analysis<br />

was performed through the design of a decision tree using TreeAge Pro (Williamstown,<br />

MA).<br />

RESULTS: A total of 597 cycles were included in the study, composed of<br />

282 MET and 315 IVF cycles. Patient ages and cycle characteristics are reported<br />

in Table 1. Cancelled cycles were approximately six times more likely<br />

in MET cycles (0.312) as compared to IVF cycles (0.057). Awillingness to pay<br />

of $50,000 per live birth was applied to our model for the purpose of analyzing<br />

cost effectiveness. The average patient charge for a cycle of METwas $4710 for<br />

an expected likelihood of delivery of 0.12, while the expected total charges of<br />

an IVF cycle was $16,824 for a 0.40 likelihood of delivery. A frozen embryo<br />

transfer cycle was permitted if excess embryos were present from the prior cycle.<br />

Looking at the entire cohort (age independent), one-way sensitivity analysis<br />

demonstrated that IVF is cost effective compared to MET when it is<br />

priced less than 2.86 times that of an MET cycle (including medications).<br />

VALIDATION OF BIRTH DEFECTS DATA IN THE SART CORS TO<br />

BIRTH DEFECTS REGISTRY DATA IN<br />

MASSACHUSETTS. J. E. Stern, a D. Gopal, b M. Kotelchuck, c<br />

B. Luke. d a Obstetrics and Gynecology, Geisel School of Medicine at Dartmouth,<br />

Lebanon, NH; b BUSPH, Boston, MA; c MassGeneral Hospital for<br />

Children, Boston, MA; d Michigan State University, East Lansing, MI.<br />

OBJECTIVE: To validate data on birth defects in the Society for Assisted<br />

Reproductive Technology Clinic Outcome Reporting System (SART CORS)<br />

with birth defects data in Massachusetts Birth Defects Registry (MBDR) reported<br />

for the same children.<br />

DESIGN: Retrospective validation.<br />

MATERIALS AND METHODS: Singleton live births and stillbirths reported<br />

to the SART CORS from July 1, 2004-Dec 31, 2008 were linked to birth<br />

and fetal death certificates in MA with an overall linkage rate of 89.7%, which<br />

was performed using mothers’ names, dates of birth, fathers’ names, and dates<br />

of delivery. The birth and fetal death certificates are linked to the MBDR within<br />

the Pregnancy to Early Life Longitudinal database. Data from the Congenital<br />

Abnormalities field in SART CORS were compared with data in the MBDR;<br />

sensitivity and specificity were calculated using the MBDR as the reference.<br />

In addition, death certificates for any neonatal death were linked to the birth certificates<br />

and compared to the neonatal death field in SART CORS.<br />

RESULTS: There were 6,503 linked singleton births (6,484 live births and<br />

19 stillbirths) in the SART CORS dataset that reported on birth defects. Of<br />

the defects reported, the sensitivity ranged from <strong>21</strong>.4% to 61.5%; specificity<br />

was 98.7% for all defects and >99% for each defect due to the low absolute<br />

percent occurrence. Neonatal deaths were reported for 13 infants in both systems,<br />

but 11infant deaths were reported in SART CORS that did not appear in<br />

the vital record system.<br />

SART<br />

CORS<br />

Mass<br />

BDR<br />

Reported<br />

in Both Sensitivity (%)<br />

Any birth defect 132 132 51 38.6<br />

Cleft palate 6 4 2 50.0<br />

Genetic 18 13 8 61.5<br />

Cardiac 25 42 10 23.8<br />

Limb 7 13 4 30.8<br />

Other 54 103 22 <strong>21</strong>.4<br />

Unknown 25 - - -<br />

FERTILITY & STERILITY Ò<br />

e205


CONCLUSIONS: The SART CORS field for birth defects contains information<br />

that, for Massachusetts deliveries, does not agree with the gold standard.<br />

Although under-reporting of some birth defects may be due to<br />

clinicians primarily reporting only defects discovered within days of delivery,<br />

the reason for over-reporting of other defects is unclear. These findings<br />

indicate that the SART CORS data field for birth defects is unreliable for use<br />

in studies of ART outcome.<br />

Supported by: NIH grants: R01HD064595 and R01HD067270.<br />

P-293 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

FOLLICULAR LYSOPHOSPHATIDIC ACID MAY BE IMPORTANT<br />

FOR FOLLICULOGENESIS IN OBESE WOMEN UNDERGOING<br />

CONTROLLED OVARIAN HYPERSTIMULATION. J. K. Riley a<br />

E. Jungheim. b a Obstetrics and Gynecology, Washington University School<br />

of Medicine, St. Louis, MO; b Obstetrics and Gynecology, Washington University,<br />

St. Louis, MO.<br />

OBJECTIVE: Studies suggest lysophosphatidic acid (LPA) may influence<br />

ovarian folliculogenesis, but the clinical significance is largely unknown.<br />

Thus we investigated associations between LPA species in ovarian follicular<br />

fluid (FF) and important clinical parameters in women undergoing in vitro<br />

fertilization (IVF).<br />

DESIGN: Cross-sectional.<br />

MATERIALS AND METHODS: FF was collected from women ages 18-<br />

44 during oocyte retrieval for IVF. Women using gestational carriers or<br />

whose cycles were cancelled were excluded. FF LPA species were quantified<br />

by liquid chromatography-mass spectroscopy. Clinical parameters investigated<br />

included patient age, body mass index (BMI), infertility diagnosis,<br />

antral follicle count, length of controlled ovarian hyperstimulation (COH)<br />

required to achieve mature follicle size, number of oocytes retrieved, percent<br />

mature oocytes, oocyte fertilization rate, and percent good quality cleavage<br />

stage embryos. Associations between LPA species and clinical parameters<br />

were made using appropriate bivariate statistics and stratified analysis.<br />

RESULTS: 197 women were included. A negative correlation was noted<br />

between amounts of FF LPA (C14:0, C16:0, and C18:0 species) and duration<br />

of COH (r¼-.2, r¼-.2, r¼-.2 respectively, p< .002 for all), whereas a positive<br />

correlation was noted between FF LPA and BMI (r¼.22, r¼.3, r¼.3 respectively,<br />

p< .002 for all). FF LPA species were not associated with any other variables.<br />

Women were stratified by BMI into normal (BMI


of identified articles, according to pre-specified criteria, using EROS software.<br />

As a second step, two randomly selected independent reviewers assessed<br />

each included study, to finally include them in the analysis and to<br />

do the data extraction. We analyzed separately in the abstract and in the results<br />

section of the full text, if the authors mentioned the p-value and a confidence<br />

interval for the main outcome and for secondary outcomes. For<br />

binary outcomes, we evaluated if any relative measure (i.e. RR, OR or<br />

RRR) or absolute measure (i.e. ARR or NNT) was used or mentioned. For<br />

continuous measures, we evaluated if the mean difference and its confidence<br />

interval was used. We analyzed if the MCID or any reference related to any<br />

clinically relevant result was mentioned in the material and methods section.<br />

We also evaluated the discussion and conclusions and described the interpretation<br />

that the authors did from the results that they published.<br />

RESULTS: In the abstracts of the evaluated RCTs, neither the p-value nor<br />

the confidence interval of the central estimation was mentioned in 25% of the<br />

studies. In the material and methods section, the minimal clinically important<br />

difference appeared only in 50% of the articles. In the results, both the p-<br />

value and the confidence intervals were used in 57% of the studies, while absolute<br />

measures were used in 42%. Finally, the confidence intervals were interpreted<br />

in the discussion in only 42% of the cases.<br />

CONCLUSIONS: Our data showed that there is an underreporting of the<br />

most objective statistics such as the absolute differences and the confidence<br />

intervals. It is the responsibility of authors and editorial boards to emphasize<br />

the implementation of adequate reporting guidelines by the use of CON-<br />

SORT in these specific type of studies.<br />

Reference:<br />

1. Glujovsky D, Boggino C, Riestra B, Coscia A, Sueldo CE, Ciapponi A.<br />

Quality of reporting in infertility journals. Fertil Steril. <strong>2015</strong><br />

Jan;103(1):236-41.<br />

P-297 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

PUBLIC FUNDING OF IVF WITHOUT AGE LIMITS: A<br />

CAUTIONARY TALE FROM QUEBEC. S. Ouhilal, a H. Lachgar, a<br />

F. Bissonnette, b B. Buckett, c I. Kadoch, d P. St-Michel, e N. Mahutte. a a Reproductive<br />

Endocrinology and Infertility, Montreal Fertility Centre, Montreal,<br />

QC, Canada; b Reproductive Endocrinology and Infertility, OVO, Montreal,<br />

QC, Canada; c Reproductive Endocrinology and Infertility, McGill University,<br />

Montreal, QC, Canada; d Reproductive Endocrinology and Infertility, Centre<br />

Hospitalier Universitaire de Montreal, Montreal, QC, Canada; e Reproductive<br />

Endocrinology and Infertility, Procrea Clinics, Montreal, QC, Canada.<br />

OBJECTIVE: In August 2010 the Quebec Government began covering the<br />

costs of in vitro fertilization. The goals of the program were to facilitate access<br />

to IVF independent of socioeconomic status and to reduce the multiple<br />

pregnancy rate below 10%. The Quebec program had no upper age limit<br />

eligibility criteria. We sought to determine the outcome and costs of IVF<br />

treatment in women over the age of 40 who used their own eggs.<br />

DESIGN: Cohort study.<br />

MATERIALS AND METHODS: We used the Quebec data from the<br />

CARTR database coupled with Quebec IVF reimbursement rates.<br />

RESULTS: Most clinics have an age cut off between age 42 and age 44.<br />

We looked at data with live birth outcomes between August 2010 and<br />

December 2012. During that time period Quebec had 3685 fresh IVF cycle<br />

starts in women over age 40. There were 3385 egg retrievals performed<br />

and 2681 embryos transferred. The mean number of embryos transferred<br />

was 1.9. The live birth rate per transfer was 9.2%. There were 246 babies<br />

born. The table below shows the outcomes and cost per birth when broken<br />

down by age between age 40 and 44.<br />

CONCLUSIONS: For women over the age of 40, live birth rates are low and<br />

come at a substantial financial cost in a public program. Age eligibility criteria<br />

should be considered by any government planning to introduce public funding.<br />

Outcomes at 40 and older.<br />

Age 40 41 42 43 44<br />

Number of cycle starts 1049 1005 944 488 141<br />

Number of egg retrievals 972 922 865 447 128<br />

Number of embryo 770 722 692 356 99<br />

transfers<br />

Number of live births 105 69 51 20 0<br />

Live birth rate per 10% 6.9% 5.4% 4.1% 0%<br />

cycle start<br />

Mean treatment cost $43,153 $62,290 $79,100 $103,994 $597,800<br />

per birth<br />

(excluding<br />

medications)<br />

P-298 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

VALIDATION OF INFERTILITY TREATMENT AND ASSISTED<br />

REPRODUCTIVE TECHNOLOGY USE ON THE BIRTH CERTIFI-<br />

CATE: A US STUDY IN EIGHT STATES. B. Luke, a M. B. Brown, b<br />

L. G. Spector. c a Obstetrics, Gynecology, and Reproductive Biology, Michigan<br />

State University, East Lansing, MI; b Biostatistics, University of Michigan,<br />

Ann Arbor, MI; c Pediatrics, University of Minnesota, Minneapolis, MN.<br />

OBJECTIVE: To evaluate the accuracy of infertility treatment and assisted<br />

reproductive technology (ART) reported on the birth certificate.<br />

DESIGN: Cohort linkage study.<br />

MATERIALS AND METHODS: ART cycles from the Society for Assisted<br />

Reproductive Technology Clinic Online Reporting System were linked to<br />

certificates of live birth in Florida, Massachusetts, New York, and Pennsylvania<br />

(2004-09), Texas (2005-09), California and Ohio (2006-09), and Colorado<br />

(2007-09) (ART children). All other live births to the same woman<br />

were also identified (ART siblings), as well as a 10:1 sample of deliveries<br />

of non-ART children (controls). Three questions on the most recent version<br />

of the birth certificate were evaluated: Q1) Pregnancy resulted from infertility<br />

treatment; Q2) Fertility-enhancing drugs, artificial insemination (AI) or intrauterine<br />

insemination (IUI); Q3) Assisted reproductive technology (e.g.,<br />

in vitro fertilization (IVF), gamete intrafallopian transfer (GIFT)). Since<br />

not all items were included by each State, we created a summary item:<br />

Any infertility question checked ‘Yes’. Information on the birth certificate<br />

was evaluated for each of the three groups of children, overall and by plurality<br />

(singleton vs multiple birth).<br />

RESULTS: The distribution (%) by study group, plurality, and checkbox<br />

item on the birth certificate is shown below.<br />

CONCLUSIONS: The sensitivity of Q3 was 28.2% and the specificity was<br />

99.7%. Only 36.5% of births of ART children were identified by any checkbox<br />

on the birth certificate; multiple pregnancies were more likely to be indicated<br />

as the result of infertility treatment than singletons (43.4% vs 33.3%). If<br />

this undercount is applied to the ART siblings, about 1/3 of the singleton and<br />

most of the multiple births also resulted from some type of infertility treatment.<br />

Q1: Pregnancy<br />

resulted from<br />

infertility<br />

treatment<br />

Q2: Fertilityenhancing<br />

drugs, AI, or IUI<br />

Q3: Assisted<br />

reproductive<br />

technology,<br />

IVF, or GIFT<br />

Summary: Any<br />

infertility item<br />

checked ‘Yes’<br />

ART Children ART Siblings Controls<br />

Overall 69,969 9,489 636,645<br />

Singleton Births 47,737 8,890 623,030<br />

Multiple Births 22,232 599 13,615<br />

Overall 36.8 14.1 0.8<br />

Singleton births, 33.8, 43.0 12.3, 38.4 0.6, 8.9<br />

multiple births<br />

Overall 11.5 6.2 0.4<br />

Singleton births, 10.1, 14.4 5.2, 20.6 0.3, 5.3<br />

multiple births<br />

Overall 28.2 5.5 0.3<br />

Singleton births, 26.0, 33.0 4.9, 15.0 0.2, 4.0<br />

multiple births<br />

Overall 36.5 12.8 0.7<br />

Singleton births,<br />

multiple births<br />

P-299 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

33.3, 43.4 11.1, 37.1 0.5, 8.7<br />

SEX RATIO AT BIRTH IN RELATION TO USE OF ASSISTED<br />

REPRODUCTIVE TECHNOLOGIES (ART). D. Kumar, a<br />

K. Hoeger, b E. S. Barrett, c D. Li, d T. Dye. d a University of Rochester, Henrietta,<br />

NY; b University of Rochester Medical Center, Rochester, NY; c University<br />

of Rochester, Rochester, NY;<br />

d ObGyn, University of Rochester,<br />

Rochester, NY.<br />

OBJECTIVE: To examine secondary sex ratio, the ratio of males to females<br />

at birth, in relation to use of Assisted Reproductive Therapy (ART)<br />

to conceive.<br />

DESIGN: Retrospective cohort study using birth registry data collected<br />

from 22 counties in Central and Upstate New York.<br />

FERTILITY & STERILITY Ò<br />

e207


MATERIALS AND METHODS: From 2004-2013, data on 322,397<br />

singleton births was collected as part of the Central and Upstate New York<br />

Perinatal Data System. At birth women completed questionnaires related<br />

to their pregnancies, including an item on the use of ART to conceive the index<br />

pregnancy. Multivariable logistic regression models were used to<br />

examine the association between ART use and secondary sex ratio, adjusting<br />

for maternal (age, race, education, weight gain, employment, Medicaid status),<br />

and paternal (age and ethnicity) characteristics.<br />

RESULTS: In total, there were 322,397 singleton births during the study<br />

period, of which 1978 (0.6%) were conceived by ART. The secondary sex ratio<br />

of infants conceived using ARTwas 0.956, whereas the sex ratio of infants<br />

conceived naturally was 1.054. In unadjusted models, the odds of having a<br />

male infant was reduced among ART users compared to non-ART users<br />

(OR: 0.907; 95% CI: 0.831, 0.991). The results of adjusted models were<br />

similar (OR: 0.908; 95%, CI: 0.828, 0.999).<br />

CONCLUSIONS: Our analyses demonstrate that after adjusting for covariates,<br />

ART is associated with a lower secondary sex ratio, indicating relatively<br />

fewer males at birth. The results of this large, population-based<br />

study agree with some (but not all) previous work on this topic. A notable<br />

limitation is that this data registry did not have information on the indications<br />

for use of ART nor the particular procedures used (IVF versus ICSI). Future<br />

research should consider these factors as well as the mechanisms underlying<br />

the skewed sex ratio following ART.<br />

Supported by: UL1 TR000042.<br />

P-300 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

COMPARISON OF OOCYTES DERIVED FROM NON-DOMINANT<br />

SMALL FOLLICLES COLLECTED 24 AND 48 HOURS AFTER<br />

TRIGGERING THE LUTEINIZING HORMONE SURGE IN NATU-<br />

RAL CYCLE IN-VITRO FERTILIZATION. O. Miyauchi, T. Ueno,<br />

T. Okubo, T. Hayashi, M. Kuroda, K. Omi, Y. Watanabe, T. Segawa,<br />

H. Osada, S. Teramoto. Shinbashi Yume Clinic, Tokyo, Japan.<br />

OBJECTIVE: In natural cycle in-vitro fertilization (IVF), when the endogenous<br />

luteinizing hormone (LH) surge has started, oocytes from the dominant<br />

follicles (DF) are collected 24 hours after the administration of gonadotrophin-releasing<br />

hormone-agonist (GnRHa) to enhance the LH surge. However,<br />

the rate of mature oocytes from non-dominant small follicles (SF)<br />

increases if they are collected 48 hours after the GnRHa administration. In<br />

this study, the efficacy of the delayed collection of SF oocytes was investigated<br />

by comparison with SF oocytes collected 24 hours after the GnRHa<br />

administration.<br />

DESIGN: Prospective cohort study from February 2011 to April 2013.<br />

MATERIALS AND METHODS: A total of 340 patients (mean age 35.6<br />

3.7 years) in whom the endogenous LH surge had started in the natural<br />

cycle gave written informed consent to be included in the study. The study<br />

was approved by Shinbashi Yume Clinic’s institutional review board. Both<br />

DF and SF (diameter 3-10 mm) oocytes were collected 24 hours after<br />

GnRHa administration in 185 patients (group A). DF and SF oocytes<br />

were collected 24 hours and 48 hours after GnRHa administration, respectively,<br />

in 155 patients (group B). There was no statistical difference in<br />

average age between the two groups. Mature oocytes and in-vitro-matured<br />

oocytes were cultured to blastocyst stage and vitrified. A thawed single<br />

blastocyst was transferred in the subsequent natural cycle. The vitrified<br />

blastocyst rate (VBR) and the live birth rate (LBR) were compared between<br />

the DF and SF oocytes of the two groups. For statistical analysis, we used a<br />

Chi square for an independence test. The difference was considered significant<br />

at P < 0.05.<br />

RESULTS: In group A, there was no difference in the VBR between the<br />

DF and SF oocytes (DF: 45 blastocysts, 45 patients, 24.3% ¼ 45/185 vs<br />

SF: 51 blastocysts, 42 patients, 22.7% ¼ 42/185). There was no difference<br />

in the LBR (DF: 45 transfers, 19 live births, 10.3% ¼ 19/185 vs SF: 41 transfers,<br />

12 live births, 6.5% ¼ 12/185). The cumulative LBR of the DF and SF<br />

oocytes was 16.8% (31/188), showing no difference compared to the LBR of<br />

the DF-only group (p¼0.189). In group B, there was no difference in the<br />

VBR between the DF and SF oocytes (DF: 36 blastocysts, 36 patients,<br />

23.2% ¼ 36/155 vs SF: 51 blastocysts, 36 patients, 23.2% ¼ 36/155). There<br />

was no difference in the LBR (DF: 35 transfers, 16 live births, 10.3% ¼ 16/<br />

155 vs SF: 40 transfers, 15 live births, 9.7% ¼ 15/155). The cumulative LBR<br />

of the DF and SF oocytes was 20.0% (31/155), showing a significant difference<br />

compared to the LBR of the DF-only group (p ¼ 0.018).<br />

CONCLUSIONS: SF oocyte collection 48 hours after GnRHa administration<br />

compared with 24 hours after in natural cycle IVF, when the endogenous<br />

LH surge had started, was shown to be a more efficient infertility treatment.<br />

P-301 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE LACK OF IMPACT OFAGE ON TREATMENT TERMINATION<br />

IN INSURED IVF PATIENTS. D. Sakkas, a L. Dodge, b M. R. Hacker, b<br />

A. D. Domar. a a Boston IVF, Waltham, MA; b Department of Obstetrics and<br />

Gynecology, Beth Israel Deaconess Medical Center, Boston, MA.<br />

OBJECTIVE: To determine if the age of the female patient influences<br />

treatment termination among insured patients undergoing in vitro fertilization<br />

(IVF).<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: All cycles among insured women who<br />

were age 18-


donor (27.94.3 vs. 37.22.0; p


the factors that influenced the most the ICER. Sensitivity analyses (one-way<br />

and probabilistic), performed to evaluate inaccuracies, corroborate these<br />

findings and attest their robustness.<br />

CONCLUSIONS: Whereas drugs used for COS have different prices,<br />

there are still not enough medico-economic analyses. These results indicate<br />

that HP-hMG is the dominant option for IVF/ICSI in France. They corroborate<br />

data from previous published data from international studies and should<br />

raise awareness of health authorities on cost-effectiveness ratio of different<br />

gonadotropins for COS in the setting of their health policy on IVF/ICSI reimbursements.<br />

Finally, results are consistent when using different sources<br />

of effectiveness data (published sources and French data).<br />

Supported by: This study was sponsored by Ferring SAS, Gentilly, France.<br />

EA is Ferring SAS employee (Medical Advisor, Medical Affairs Dept). CA,<br />

PB, RF, SH and GPB were consulted as experts of ART.<br />

P-306 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ASSISTED REPRODUCTIVE TECHNOLOGIES (ART) AFFECT<br />

EARLY CHILDHOOD GROWTH IN LOWER-BIRTHWEIGHT<br />

CHILDREN. A. Butz, a R. S. Weinerman, a S. Senapati, a C. Sapienza, b<br />

C. Coutifaris, a M. A. Mainigi. a a Division of Reproductive Endocrinology<br />

and Infertility, University of Pennsylvania, Philadelphia, PA; b Fels Institute<br />

for Cancer Research & Molecular Biology, Temple University, Philadelphia,<br />

PA.<br />

OBJECTIVE: A pilot study to determine the effect of ART on childhood<br />

growth and to determine whether ART-associated placental DNA methylation<br />

differences impact childhood growth.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Using parental survey data, current<br />

birthweight (BW) and body mass index (BMI) percentiles (%) were<br />

computed using World Health Organization and Centers for Disease Control<br />

standards, adjusting for age, sex, and gestational age at birth (GA), for children<br />

(age 1-8) born from singleton pregnancies conceived via ART or natural<br />

conception (controls). Birth information had been collected prospectively<br />

during a previous study analyzing placental DNA methylation. Placental<br />

DNA methylation of the ART subjects had been analyzed using a custom-designed<br />

Illumina VeraCode array containing 96 cytosine-phosphate-guanine<br />

(CpG) sites in 82 genes known or thought to play a role in fetal growth.<br />

The ratio of current BMI% to BW% was computed as a measure of a child’s<br />

growth trajectory and was modeled as a function of BW% in control and ART<br />

children. This was also done for ART subjects with normal and abnormal<br />

placental DNA methylation at specific CpG sites. Between-group differences<br />

were assessed using Mann-Whitney and t-tests. Non-linear regression<br />

modeling was assessed using the extra sum of squares F test.<br />

RESULTS: ART (n¼45) and control (n¼54) children did not differ in sex,<br />

current age, mean GA, BW% or current BMI%. The mean BMI:BW% ratio<br />

did not differ between ART and control children. However, when BMI:BW%<br />

ratio was modeled as a function of BW% using non-linear regression, there<br />

were differences in the curves between ART and control children that were<br />

most notable in children with BW


696), the primary outcome of this study. In a post-hoc descriptive summary of<br />

vital pregnancy rates, 298 women with embryo transfer enrolled in the<br />

follow-up FTET study demonstrated per-cycle vital pregnancy rates of<br />

32.0% with CFA (58/181) versus 32.7% with rFSH (54/165). There were<br />

346 cycles among the 298 women, translating to per subject rates of<br />

37.5% for CFA (57/152) versus 36.3% for rFSH (53/146).<br />

CONCLUSIONS: The cumulative vital pregnancy and live-birth rates (from<br />

fresh cycles and FTET) were similar in older women treated with CFA and<br />

rFSH. No new safety signals were detected in this follow-up FTET study.<br />

Supported by: Merck & Co., Inc., Kenilworth, NJ, USA.<br />

P-309 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

DOES ULTRASOUND MONITORING AND OVULATION TRIGGER<br />

WITH HCG IMPROVE OUTCOMES OF INTRAUTERINE INSEMI-<br />

NATIONS (IUI) PERFORMED IN NATURAL CYCLES (NC)? H. El<br />

Hachem, a,b P. E. Bouet, a,b L. Lapensee, a,b F. Bissonnette, a,b J. Benoit, a,b<br />

R. Antaki. a,b a Ovo Fertility Clinic, Montreal, QC, Canada; b Obstetrics and<br />

Gynecology, University of Montreal, Montreal, QC, Canada.<br />

OBJECTIVE: To compare two methods of timing IUI in NC: spontaneous<br />

triggering of ovulation by detecting LH surge with urinary ovulation kits (u-<br />

LH) and ultrasound monitoring of follicular growth followed by ovulation<br />

trigger with hCG (US/hCG).<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: All women %40 years with no history of<br />

infertility who underwent donor sperm IUI (DS-IUI) in NC from Janauray 2011<br />

to June 2014 were included. The indications for DS-IUI were the absence of a<br />

male partner (single women and same-sex couples) and azoospermia. All<br />

women had ovarian reserve testing (FSH, antral follicle count, AMH) and at<br />

least one patent tube. None of the patients received any treatment. Women in<br />

the u-LH group self-monitored ovulation at home using urinary ovulation<br />

kits starting cycle day 10, and IUI was performed the day following the LH<br />

surge. Women in the US/hCG group had serial US starting cycle day 10, and<br />

when a leading follicle R<strong>17</strong> mm was noted, ovulation was triggered with recombinant<br />

hCG (Choriogonadotropin a,250mg) and IUI performed 36 hours<br />

later. The choice of protocol was left to the physician’s discretion. Primary outcomes<br />

were live birth rate per started cycle (LBR) and cumulative LBR. Secondary<br />

outcomes were clinical pregnancy rate per started cycle (CPR) (fetal<br />

heartbeat at 7 weeks), cycle cancellation and miscarriage rates.<br />

RESULTS: 538 DS-IUI in 232 women were included: 267 u-LH in 113<br />

women and 271 US/hCG in 119 women. The two groups were comparable<br />

for age, Body Mass Index, ovarian reserve and number/mobility of inseminated<br />

sperm. There were no significant differences in the primary and secondary<br />

outcomes (Results are shown in table 1).<br />

CONCLUSIONS: In women with no history of infertility undergoing<br />

donor sperm IUI in a natural cycle, US monitoring of follicular growth followed<br />

by ovulation trigger with hCG does not improve CPR and LBR<br />

when compared with u-LH monitoring. The latter has the advantage of being<br />

non-invasive and easy to perform at home and significantly reduces costs by<br />

limiting hospital visits. Therefore, u-LH monitoring might be the best option<br />

available to these women.<br />

Outcomes of DS-IUI.<br />

u-LH group<br />

(n¼267)<br />

US/hCG group<br />

(n¼271)<br />

p value*<br />

Cycle Cancellation 1.1% (3/267) 1.4% (4/271) p¼ 0.71<br />

CPR 16.1% (43/267) 11.8% (32/271) p¼ 0.15<br />

Miscarriage 22.7% (10/44) <strong>21</strong>.8% (7/32) p¼ 0.93<br />

LBR 12.4% (33/267) 9.2% (25/271) p¼ 0.24<br />

Cumulative LBR 29.2% (33/113) <strong>21</strong>% (25/119) p¼ 0.15<br />

* X2 tests were used to assess for differences between groups.<br />

P-310 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

SCREEN MORE OR SCREEN LESS? CARRIER SCREENING IN<br />

GAMETE DONORS: IT’S TIME FOR A NEW<br />

PARADIGM. S. Rodriguez, a N. Kumar, a S. Yarnall, a R. Shraga, a<br />

S. S. Chuan, b C. Pascale, c F. Licciardi. d a Recombine, New York, NY;<br />

b San Diego Fertility Center, San Diego, CA; c Mental Health, Livingston,<br />

NJ; d New York University Langone Medical Center, New York, NY.<br />

OBJECTIVE: Carrier screening in gamete donors varies. Protocols are<br />

based on ASRM guidelines, positive screening rates, and recipient testing<br />

protocols. We assessed screening practices of varying scope and analyzed<br />

the number of donors identified as carriers. With this data, our goal is to propose<br />

a carrier screening protocol for gamete donors.<br />

DESIGN: Retrospective.<br />

MATERIALS AND METHODS: Genotyping was performed for over<br />

1500 mutations associated with over 200 autosomal recessive and X-linked<br />

genetic diseases. Custom testing panels were developed for referring institutions;<br />

this allows for results reporting of only those diseases within a specific<br />

panel. The analysis includes data obtained from 6<strong>21</strong> gamete donors referred<br />

by reproductive endocrinologists and donor banks/agencies. Documented<br />

informed consent was obtained. Frequently ordered testing panels were identified.<br />

Overall genotyping results were compared to reported results, as determined<br />

by inclusion in the ordered testing panel.<br />

RESULTS: Of 78 custom donor screening panels, 2 panels were used to<br />

screen 86% of donors. A limited panel of 10 diseases and


was made 30 minutes after IVI (IVI +30). Each recording was accelerated 10-<br />

fold and UC frequencies at the isthmic and fundal parts of the uterus were<br />

rated by 3 independent observers who were blinded for study design. Overall<br />

UC frequency was obtained by the average frequency between isthmic and<br />

fundal contractions and average ratings of observers.<br />

RESULTS: Median UC frequency (min-max) increased slightly but not<br />

significantly over the 3 observation points (baseline, IUI +15, and IVI +30):<br />

3.2 (2.7-4.1), 3.3 (2.5-4.9), and 3.4 (2.4-4.8) UC/minute, respectively. Further,<br />

UC frequency was not influenced significantly by semen volume nor the magnitude<br />

of ovarian response. UC frequency was negatively correlated (Spearman<br />

test) with women’s ages at IUI +15 (r¼-0.42, P< 0.04) and IVI +30 (r¼-<br />

0.51, P< 0.009) but not at baseline (r¼-0.20, NS). Yet, the % of increase of<br />

UC frequency over the 3 observation points was not related to women’s ages.<br />

Finally, pregnancy rates tended to be higher (30% vs. 15%) in patients showing<br />

R3.5 UC/minute (n¼13) than those with


outcomes such as birth weight (g) 3163 (CI95%3035-3292) vs. 3074(CI95%<br />

2913-3236), low birth weight (0.8-%1.2 ng/mL (group 3) and >1.2 ng/mL (group 4). Low birth weight<br />

was defined as


MATERIALS AND METHODS: Included in the study were agonist<br />

(n¼206) and antagonist (n¼43) down regulated cycles. All patients received<br />

both HP-hMG (MenopurÒ) and HP-FSH (BravelleÒ) in an approximate 1:1<br />

ratio (i.e an LH/FSH ratio of 0.5) from day one of stimulation. Only cycles<br />

with blastocyst transfers were included.<br />

RESULTS: Characteristics of the study population were as follows (mean <br />

SD): age ¼ 33.2 years 4.2 (range <strong>21</strong>-42 years), BMI ¼ 24.0 4.3, stimulation<br />

days ¼ 9.7 0.7, total FSH dose (IU) ¼ 1541 5<strong>21</strong>, total HP-hMG dose<br />

(IU) ¼ 1359 582, oocytes ¼ 13.7 5.5, MII ¼ 10.2 4.2, 2PN ¼ 8.5 3.2,<br />

peak E2 ¼ 2356.35 1888.73 pg/ml, and peak P4 ¼ 0.96 0.63. Cycles were<br />

divided into peak (day of hCG administration) P4 %1.5 ng/ml compared to P4<br />

>1.5 ng/ml. The incidence of PPR was 16.4%. The implantation rate, clinical<br />

and ongoing pregnancy rates were comparable between groups. To analyze the<br />

association among variables associated with increased P4 levels, linear regression<br />

was performed. BMI, peak E2 levels on the day of hCG administration, and<br />

number of retrieved oocytes were associated with PPR.<br />

CONCLUSIONS: Despite a 16.4% incidence of PPR, elevated peak P4<br />

levels were not associated with a negative effect on IVF outcomes following<br />

a stimulation strategy employing HP-hMG and HP-FSH at a 0.5 LH/FSH ratio.<br />

The use of HP-hMG may protect against a potential negative effect of<br />

elevated P4 on implantation in fresh autologous blastocyst transfers.<br />

Table 1.<br />

%1.5 ng/ml<br />

(n¼208)<br />

>1.5 ng/ml<br />

(n¼41)<br />

Embryos transferred 1.9 0.7 1.8 0.5 0.254<br />

Implantation rate 229/406 (57%) 45/75 (59%) 0.753<br />

Clinical pregnancy/ 154/208 (74%) 28/41 (68%) 0.572<br />

cycle started<br />

LB/cycle started 144/208 (69%) 27/41 (66%) 0.809<br />

Reference:<br />

Warner et al, Fertil Steril 2014.<br />

Supported by: VCRM.<br />

P-318 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

WITHDRAWN<br />

P-319 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

CHANGES IN SINGLETON LIVE BIRTH WEIGHTS IN A LARGE<br />

IVF PRACTICE OVER AN 18 YEAR PERIOD. K. Maas, a,b<br />

E. Galkina, c K. Thornton, a,b D. Sakkas. b a Beth Israel Deaconess Medical<br />

Center, Boston, MA; b Boston IVF, Waltham, MA; c Biomedical Engineering<br />

and Biotechnology, University of Massachusetts Lowell, Lowell, MA.<br />

OBJECTIVE: The fetal origins hypothesis suggests that some diseases<br />

originate in utero owing to adaptations made by the fetus to the environment<br />

it encounters. This has allowed live birth weights (LBW) to be used as a surrogate<br />

marker of these in utero environmental encounters. It has been shown<br />

that babies born from IVF in a thaw cycle have higher LBW on average<br />

compared with those born from a fresh cycle. It has also been hypothesized<br />

that embryo culture media can impact LBW. Clinical IVF practices including<br />

stimulation protocols, medication dosing, medication types, transfer day, and<br />

cycle monitoring have evolved significantly since the inception of IVF. The<br />

IVF laboratory including incubators, culture media, culture devices and the<br />

introduction of micromanipulation techniques has also changed significantly<br />

over time. The objective of this study is to investigate the association between<br />

singleton LBW over an 18 year period in a large academic IVF clinic in the<br />

US over time after autologous in vitro fertilization (IVF) in both fresh and<br />

frozen cycles.<br />

DESIGN: Retrospective cohort study of 7332 singleton live births from<br />

patients who underwent autologous fresh or frozen IVF cycles at Boston<br />

IVF between 1996 and 2013.<br />

MATERIALS AND METHODS: 6265 fresh and 1067 frozen cycles were<br />

analyzed. One way ANOVA and t-tests were performed to compare average<br />

LBW in autologous fresh and frozen cycles as well as average LBW per cycle<br />

type over time. Six month increments were compared over the study period.<br />

P<br />

RESULTS: A total of 7332 singleton deliveries were included from the<br />

fresh and frozen cycles. The mean LBW SD in fresh cycle and frozen cycle<br />

cohorts were 3282 620g (3267-3298g) and 3456 600g (3420-3492g)<br />

respectively. ANOVA and t-tests demonstrated a statistically significant<br />

mean difference (<strong>17</strong>3g, p


OVERALL<br />

HIGH FSH &<br />

NORMAL AMH<br />

NORMAL<br />

FSH & LOW AMH<br />

HIGH FSH<br />

& LOW AMH<br />

NORMAL FSH &<br />

NORMAL AMH<br />

P-VALUE<br />

N(%) 33 (2) 647 (33) 166 (8) 1111 (57)<br />

Total gonadotropins Predicted Mean SE 5761 288 a 5297 66 b 6075 129 b 3582 50 ab


D3eSET and 153 had elective single blastocyst transfer (eSBT). Embryology<br />

profiles and clinical outcomes were assessed for both groups. Primary<br />

outcome measures were implantation rate (IR), clinical pregnancy rate<br />

(CPR), and spontaneous abortion rate (SAB). Categorical data are presented<br />

as percentages and compared using the Fisher’s exact test. Continuous variables<br />

are described as mean +/- SD and compared using the Students t-test,<br />

assuming equal variances.<br />

RESULTS: Patients undergoing D3eSET had significantly fewer total oocytes<br />

retrieved, mature oocytes, fertilized oocytes, top quality embryos for<br />

transfer and cryopreservation. However, patients in the D3eSET group<br />

achieved equally high clinical pregnancy and implantation rates as patients<br />

in the eSBT group.<br />

D3eSET vs. eSBT Cycle Outcomes for Patients < 35 years of age.<br />

D3eSET<br />

N¼ 25<br />

eSBT<br />

N¼153<br />

P value<br />

Mean Age 31.12.1 31.02.3 NS<br />

Mean No. Oocytes 10.04.8 16.66.3


OBJECTIVE: To assess trends in embryo transfer practices and outcomes<br />

of assisted reproductive technology (ART) cycles performed in the United<br />

States.<br />

DESIGN: Retrospective population-based cohort study.<br />

MATERIALS AND METHODS: We analyzed ART cycles initiated between<br />

1996 and 2013 in the United States and reported to the National<br />

ART Surveillance System (2013 data are preliminary). For each reporting<br />

year, we calculated: (a) average number of embryos transferred and percentage<br />

of transfers that involved one embryo, (b) percentage of singleton ART<br />

live births, (c) percentage of ART-conceived infants born with normal birth<br />

weight (R2,500 grams), and (d) average gestational age and percentage of<br />

term (R37 weeks) births among all ART births. Linear regression was<br />

used to test for the significance of linear and quadratic trends in the average<br />

number of embryos transferred and average gestational age. Linear probability<br />

models (with a binomial distribution and an identity link) were used to test<br />

for the significance of linear and quadratic trends in the proportions of all<br />

other outcomes. Generalized estimating equations were used to accommodate<br />

clustering by clinic.<br />

RESULTS: The average number of embryos transferred decreased from<br />

3.9 to 1.9 and the proportion of single embryo transfers increased from<br />

5.5% to 30.7% between 1996 and 2013. The all-time largest increase in<br />

the proportion of single embryo transfers (23.8% increase) was observed between<br />

2012 and 2013. The percentage of singleton live births among all ART<br />

births increased from 62.6% to 75.4%. The percentage of normal birth weight<br />

infants among ART-conceived infants increased from 58.0% to 71.5%. Similarly,<br />

the average gestational age of ART births increased from 37.6 weeks to<br />

37.8 weeks, and the percentage of term births increased from 68.0% to<br />

73.8%. All trends were significant (P


in blastocyst development, aneuploidy, pregnancy or implantation rate between<br />

patients with Low or High BMI.<br />

CONCLUSIONS: The impact of maternal BMI on the kinetics of embryo<br />

development, determined by time-lapse image analysis, is dependent on<br />

maternal age. Maternal BMI does not influence the success of ART in patients<br />

35 years or older, possibly due to the overriding negative impact of<br />

maternal age on oocyte quality.<br />

Table 1. Embryo morphokinetics analyzed by AGE, BMI and the interaction<br />

AGE x BMI.<br />

Age BMI t2-t4 t8 tM tSB tB tEB<br />

Young Low 11.6 58.1 92.2 101.6 106.2 108.3<br />

Young High 11.3 56.0 87.6 98.9 103.6 105.9<br />

AMA Low 11.0 59.4 92.4 102.2 107.3 109.9<br />

AMA High 10.4 58.2 93.6 104.7 110.0 113.0<br />

Age 0.06 NS 0.01 0.01 0.005 0.01<br />

pvalue BMI NS NS NS NS NS NS<br />

AgexBMI NS NS 0.0<strong>17</strong> 0.04 0.04 0.10<br />

P-330 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

DOES PREVALENCE AND EFFECT OF ELEVATED PROGESTER-<br />

ONE ON THE DAY OF OOCYTE MATURATION DIFFER BY<br />

ETHNICITY? G. D. Royster, IV, a S. Zarek, a A. Christy, a<br />

A. DeCherney, a K. Devine, b M. J. Hill. a a National Institutes of Health, Bethesda,<br />

MD; b Shady Grove Fertility Center, Washington, DC.<br />

OBJECTIVE: To compare the effect of elevated day-of-trigger progesterone<br />

concentration on live birth rate in various ethnic groups.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Analysis included all fresh autologous<br />

ART cycles from 2009-2013 in which progesterone concentration was<br />

measured on the day of oocyte trigger. Generalized estimating equations<br />

(GEE) were utilized with nesting for patients with multiple cycles and to account<br />

for confounding variables. The primary outcome was clinical pregnancy.<br />

RESULTS: 3289 fresh autologous ART cycles were analyzed. Caucasians<br />

had a similar clinical pregnancy rate (46.2%) to Hispanics (44.7%) and<br />

higher pregnancy rates than African American (41.0%) and Asians<br />

(38.3%) (P


Table-1. CYCLE OUTCOME OF DIFFERENT GROUPS OF EMBRYO.<br />

Groups (n) Pregnancies (n) PR(%) OR (95% CI) IR (sacs/total Et)(%) OR (95% CI)<br />

Total cycles (744) 326 43.8 13.9 (292/2093)<br />

1-Easy-E T (543) 248 45.6 1vs2<br />

1.23<br />

(0.9 to 1.8) (p>0.05)<br />

14.7 (236/1604) 1 vs. 2<br />

1.3<br />

(0.9 to 1.8)(p>0.05)<br />

2-All difficut-ET- (201) 78 39.8 11.4<br />

3-Cx - Tr (83) 42 50.6 1<br />

vs3<br />

0.8<br />

4-Bl-OS (39) <strong>17</strong> 43.6 1<br />

vs 4<br />

1.1<br />

(0.5 to 1.3)(p>0.05)<br />

(0.5 to 2.1)(p>0.05)<br />

5-Bl-TC (45) 11 24.4 1<br />

vs.5<br />

2.6<br />

(1.3<br />

to 5.2)(P0.05)<br />

1 vs. 5<br />

1.7<br />

(0.9 to 3.2)(p>0.05)<br />

1 vs. 6<br />

1.9<br />

(0.8 to 4.2)(p>0.05)<br />

OBJECTIVE: Easy embryo transfer ( E-ET) in contrast to difficult (D- ET)<br />

is thought to result in better cycle outcome. Although definition of E- ET is<br />

agreed upon ,there is disagreement about what constitutes D- ET. We aimed<br />

to compare the impact of individual elements of D-ET : cervical traction (Cx-<br />

Tr), blood on outer sheath ( Bl-OS) , blood on transfer catheter (Bl-TC) and<br />

the need for sounding (Snd) individually and in combination on clinical pregnancy<br />

rate (CPR) and implantation rates (IR) of ICSI/IVF cycles.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: A cohort of 744 ICSI cycles were<br />

included for women who fulfilled the following inclusion criteria :age <<br />

38 years, fresh first ICSI trial, number of transferred embryos R 2 high quality<br />

.The technique of ET used was the tactile method and the same type of<br />

catheter (Labottect ). ET was considered easy if no resistance was encountered<br />

during passing the preloaded TC and sheath through cervical canal<br />

without Cx -Tr ,Bl-OS or Bl-TC nor Snd . It was considered difficult if<br />

Cx-Tr or Snd were needed and /or , Bl-OS, Bl-TC were present . Cycle<br />

outcome (CPR , IR) were compared for cycles with easy ET and those<br />

with D- ET as a whole and individually with subgroups of Cx-Tr, Bl-<br />

OS,Bl-TC & Snd, using Odds ratio (OR) and 95% CI. The differences<br />

were considered significant if the P value was< 0.05.<br />

RESULTS: As shown in the table CPR for E-ET (45.6%) and D- ET<br />

(39.8%) are not statistically significantly different. Comparisons of E- ET<br />

with cx-Tr, Bl-OS,Bl-TC, Snd components of D-ET showed significantly<br />

lower CPR with Bl-TC and Snd transfer subgroups only. Although IR showed<br />

no statistically significant differences between E- ET and overall and components<br />

of D- ET, the Bl-TC, and Snd subgroups tended to have significantly<br />

lower IR but owing to the small number of cycles in these subgroups the differences<br />

did not reach level of statistical significance (type II error).<br />

CONCLUSIONS: Cervical traction and /or blood on outer sheath do not<br />

compromise cycle pregnancy rate or implantation rate. Only when blood<br />

on transfer catheter and/ or sounding the uterus at time of ET is the CPR<br />

significantly undermined and implantation rate tends to be lower.<br />

P-334 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

SHOULD EMBRYO TRANSFER NUMBER BE UPGRADED IN<br />

ETHNIC MINORITY GROUPS? C. Chatzicharalampous, a<br />

J. Stelling, b,c J. Jenkins, c M. Saketos, c,b L. Sung, c,d M. A. Bray. c,e a Obstetrics<br />

and Gynecology, The Brooklyn Hospital Center, Brooklyn, NY; b SUNY<br />

Stony Brook School of Medicine, Stony Brook, NY; c Reproductive Specialists<br />

of NY, Mineola, NY; d Obstetrics and Gynecology, Winthrop Univ Hosp,<br />

Mineola, NY; e Obstetrics and Gynecology, The Brooklyn Hosp Center,<br />

Brooklyn, NY.<br />

OBJECTIVE: To determine if the number of transferred embryos influences<br />

IVF clinical pregnancy rates (CPR) in different ethnic groups.<br />

DESIGN: A retrospective study of 565 fresh embryo transfer cycles: <strong>21</strong>4<br />

single and 351 double. We compared ethnicity, age, BMI, FSH, gravidity,<br />

parity,infertility diagnoses, previous fresh cycles, oocytes retrieved, fertilization<br />

rate, total and quality blastocysts, CPR and twin rate by number of embryos<br />

transferred.<br />

MATERIALS AND METHODS: Women undergoing IVF or ICSI from<br />

2010 to 2013 who had an autologous, fresh, elective single (eSET) or elective<br />

double (eDET) day 5 blastocyst embryo transfer were included in the study.<br />

Clinical Pregnancy Rate by ethnic group and embryo transfer count<br />

Single Embryo Transfer<br />

Double Embryo Transfer<br />

Caucasian % (N) Non-Caucasian % (N) Caucasian % (N) Non-Caucasian % (N)<br />

Non- Pregnant Pregnant Non-Pregnant Pregnant Non-Pregnant Pregnant Non-Pregnant Pregnant<br />

51 (88) 49 (85)* 73 (30) 27 (11)* 55 (144) 45 (118) 56 (50) 44 (39)<br />

Twin rate: 3.5%<br />

Twin rate: 0%<br />

Twin rate: 35%<br />

Twin rate: 26%<br />

Ectopic rate: 2.3%<br />

Ectopic rate: 0%<br />

Ectopic rate: 0%<br />

Ectopic rate: 10%**<br />

* p ¼ 0.02<br />

** p < 0.001<br />

FERTILITY & STERILITY Ò<br />

e<strong>21</strong>9


All women were under 38 years of age and all had < 2 previous fresh IVF<br />

cycles. We included four ethnic groups: Caucasian, African American, Asian<br />

and Hispanic, which were dichotomized into two main groups: Caucasian<br />

and non-Caucasian. Variables were analyzed using Chi-square, Fisher’s exact<br />

test, Student’s t-test and logistic regression. p < 0.05 was considered statistically<br />

significant.<br />

RESULTS: Age, BMI, day 3 FSH, gravidity, parity, number of previous<br />

fresh cycles, number of oocytes retrieved, fertilization rate and number of total<br />

and quality blastocysts were similar in both ethnic groups. Peak estradiol<br />

levels were higher in non-Caucasians (p¼0.02). Infertility diagnoses were<br />

similar in both groups with the exception of more tubal factor in non-Caucasian<br />

women (p¼0.001) and more cases of unexplained infertility in Caucasian<br />

women (p¼0.02). In the eSET groups, Caucasian women had a<br />

significantly higher CPR (49%) compared to their ethnic minority counterparts<br />

(27%) (p ¼ 0.02). Differences in CPR between patient groups persisted<br />

after accounting for differences in infertility diagnosis. For the eDET groups,<br />

the CPR did not differ by ethnicity: Caucasian (45%) vs. non-Caucasian<br />

(44%) (p ¼ N.S.). The multiple gestation rates in both groups were similar<br />

and not statistically significant. Ectopic rate was markedly higher (10%) in<br />

the DET ethnic minority group, likely due to the higher tubal factor infertility<br />

diagnosis.<br />

CONCLUSIONS: In spite of similar measured demographic variables, minority<br />

women have a significantly lower CPR in eSET cycles. Further<br />

research is needed to identify confounding factors and to determine if minority<br />

women would benefit from an upgrade to eDET in selected cases.<br />

P-335 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

USE OF ICSI IN IVF CYCLES IN WOMEN WITH TUBAL LIGA-<br />

TION DOES NOT IMPROVE PREGNANCY OR LIVE BIRTH<br />

RATES. F. Grimstad, a A. Nangia, b B. Luke, c J. E. Stern, d W. Mak. e a Obstetrics<br />

and Gynecology, University of Kansas, Kansas City, KS; b University<br />

of Kansas Medical Center; c Michigan State University, East Lansing, MI;<br />

d Geisel School of Medicine at Dartmouth, Lebanon, NH; e Yale School of<br />

Medicine, New Haven, CT.<br />

OBJECTIVE: To compare the fertilization, pregnancy, and live birth rates<br />

of in vitro Fertalization (IVF) cycles with and without intracytoplasmic<br />

sperm injection (ICSI) among patients with tubal ligation only.<br />

DESIGN: Retrospective cohort.<br />

MATERIALS AND METHODS: Assisted reproductive technology<br />

(ART) cycles from patients with tubal ligation only and no other male or<br />

female factor diagnoses from the Society for Assisted Reproductive Technology<br />

Clinic Outcome Reporting System (SART CORS) database (2004-<br />

2012) were compared according to use of ICSI (some or all) versus conventional<br />

IVF (ICSI-none). Adjusted odds ratios and 95% confidence intervals<br />

(AOR, 95%CI) were calculated using logistic regression with tubal ligation<br />

patients having no ICSI as the reference group, and adjusted for year of<br />

treatment, maternal age, race and ethnicity, gravidity, ovarian hyperstimulation<br />

syndrome, number of oocytes retrieved, day of transfer, and number<br />

of embryos transferred. Outcome measures included fertilization rate (2PN/<br />

total oocytes); clinical intrauterine gestation (CIG/cycle); and live births<br />

(LB/cycle).<br />

RESULTS: Of 1,099,978 fresh autologous cycles performed, 11,198 were<br />

for tubal ligation only (1%); of these, 5,629 cycles (50%) used some or all<br />

ICSI versus 5,569 (50%) conventional IVF. The mean fertilization rate was<br />

49.0% and 57.2% for ICSI-none versus ICSI-some/all respectively. The clinical<br />

intrauterine gestation rate and live birth rate are listed in the table. There<br />

were no significant differences in birth weight or gestational age.<br />

Outcomes of all/some ICSI compared to no ICSI.<br />

N, Cycles<br />

Clinical Intrauterine<br />

Gestation<br />

Live Birth<br />

% AOR 95% CI % AOR 95% CI<br />

No ICSI 5,629 45.0 1.00 Reference 37.5 1.00 Reference<br />

All/Some 5,569 40.3 0.81 0.75, 0.88 32.2 0.79 0.73, 0.86<br />

ICSI<br />

CONCLUSIONS: Use of ICSI versus conventional IVF did not improve<br />

outcomes for tubal ligation patients. These results suggest that the use of<br />

ICSI is generally not warranted in this population.<br />

P-336 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

IDENTIFICATION OF OVARIAN RESPONSE PREDICTORS FOR A<br />

NEWLY AVAILABLE RECOMBINANT FSH<br />

PREPARATION. C. M. Howles, a J. Jenkins, b G. Griesinger, c<br />

D. W. Warne. d a Aries Consulting, Geneva, Switzerland; b R&D and Medical<br />

Affairs, Finox Biotech A.G., Kirchberg, Switzerland;<br />

c Sektion f€ur<br />

Gyn€akologische Endokrinologie und Reproduktionsmedizin, Universit€atsklinikum<br />

Schleswig-Holstein – Campus L€ubeck, L€ubeck, Germany; d Consultant<br />

Biostatistician, Thonex, Switzerland.<br />

OBJECTIVE: Predictors of ovarian response have been reported both for<br />

GnRH agonist and antagonist treatment cycles using established FSH stimulation<br />

agents. Various predictors have been identified in different analyses,<br />

but typically the most reported are Age, FSH, AMH, BMI. With the availability<br />

of newly registered recombinant FSH (r-FSH) preparations, it is important<br />

to establish if the same predictors of ovarian response can be utilized with<br />

similar predictive capacity in order to allow treating clinicians to effectively<br />

and safely perform ovarian stimulation.<br />

DESIGN: A post-hoc analysis was conducted on data obtained from a randomized<br />

controlled Phase III registration trial comparing 2 r-FSH preparations<br />

(clinicaltrials.gov; NCT011<strong>21</strong>666). As recently described (1), a low<br />

response was defined less than 6 oocytes and a high response more than 18<br />

oocytes retrieved. Prognostic models were constructed using data from all<br />

treatment cycles (n¼482).<br />

MATERIALS AND METHODS: Patients aged 20-38 years were treated<br />

in a long luteal phase start GnRH agonist protocol with a fixed-dose daily<br />

regimen, of 150 IU r-FSH s.c. The dose could be reduced after day 6 of<br />

FSH stimulation only if there was a risk of OHSS. 15 centres in six European<br />

countries participated. Recruited subjects underwent up to 2 treatment cycles<br />

(total cycle 1 n ¼ 372; cycle 2 n ¼ 110). The trial was conducted between<br />

July 2010 and April 2012. Logistic and linear regression were used to investigate<br />

factors which influence low and high response, and number of oocytes<br />

retrieved.<br />

RESULTS: 1) 81 cycles were classified as a low response (16.8%) whilst<br />

44 (9.1%) were classified as a high response. AMH was the best performing<br />

predictor for both low and high response in both treatment groups (AMH low:<br />

ROC AUC ¼ 0.751; AMH high: ROC AUC ¼ 0.700). 2) multivariate analysis<br />

showed that the choice of gonadotrophin treatment had no impact on the<br />

predictive model hence results were pooled irrespective of gonadotrophin<br />

used to refine predictive value of other variables. 3) adding further biomarkers,<br />

such as baseline FSH and AFC did not significantly increase the<br />

AUC of prediction models.<br />

CONCLUSIONS: The results demonstrated that for both the new and<br />

reference r-FSH products, there were common predictors of ovarian response<br />

(AMH, AFC, FSH). The new r-FSH product can be clinically used in the<br />

same way as the reference product in ART practice. Additionally the findings<br />

support the utility of AMH as the most relevant marker of ovarian response<br />

and provides physicians with a reliable marker of ovarian response when using<br />

these FSH preparations.<br />

Reference:<br />

1. Broekmans FJ, Verweij PJ, Eijkemans MJ, Mannaerts BM, Witjes H.<br />

Prognostic models for high and low ovarian responses in controlled<br />

ovarian stimulation using a GnRH antagonist protocol. Hum Reprod.<br />

2014 Aug;29(8):1688-97.<br />

Supported by: Finox AG, Technikumstrasse 1, CH-3401 Burgdorf,<br />

Switzerland.<br />

P-337 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

LIVE BIRTH CORRELATION OF EEVA TEST RESULTS IN PA-<br />

TIENTS WITH DAY 5 TRANSFER: A PROSPECTIVE BLINDED<br />

STUDY. M. VerMilyea, J. T. Anthony, M. A. Mainigi. University of Pennsylvania,<br />

Philadelphia, PA.<br />

OBJECTIVE: Time-lapse-based embryo selection algorithms have been<br />

extensively assessed against blastocyst formation, aneuploidy and implantation.<br />

However, the single most important outcome of assisted reproductive<br />

e220 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


technology (ART) is live birth. The objective of this prospective blinded<br />

study was to perform the first examination of the correlation between an automated<br />

time-lapse algorithm (the Eeva Test) and live birth outcome.<br />

DESIGN: Prospective blinded study.<br />

MATERIALS AND METHODS: The study included 69 patients (Jan<br />

2013-2014) who consented to have embryos imaged using the Eeva System,<br />

a platform technology that automatically measures early cleavage timing parameters<br />

and generates an Eeva Test result regarding developmental potential.<br />

For this non-selection, blinded study, embryos were selected for fresh<br />

transfer using only morphology evaluation on Day 5. Embryos of known<br />

live birth were those from patients with #live births equal to #embryos transferred,<br />

and those who did not have a live birth.<br />

RESULTS: Of the 69 patients, 58 patients had embryos with known live<br />

birth rate. A total of 80 Day5 embryos were transferred, resulting in 26 babies<br />

born (overall known live birth rate 32%). When live birth rate was analyzed<br />

based on Eeva Test results, we observed a 38% relative difference in live birth<br />

rates between Eeva High vs. Eeva Low embryos, while egg age was similar<br />

across the categories.<br />

CONCLUSIONS: To our knowledge, this is the first prospectively designed<br />

blinded study of time-lapse analysis results and their correlation to<br />

live birth outcomes. Previous studies have shown correlation of Eeva Test results<br />

to aneuploidy risk and Day 5 implantation, indicating that the Eeva Test<br />

results may reflect embryo quality in a manner that is independent of age. The<br />

study is currently ongoing and increased sample size will allow meaningful<br />

evaluation of statistical significance.<br />

Live Birth Rate<br />

Age<br />

Eeva High 36% (14/39) 33.1 4.8<br />

Eeva Medium 36% (5/14) 34.4 3.7<br />

Eeva Low 26% (7/27) 33.7 4.6<br />

p-value NS NS<br />

Supported by: Progyny (formerly Auxogyn).<br />

P-338 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

PROGESTERONE (P4) ELEVATION ON DAY OF HUMAN CHORI-<br />

ONIC GONADOTROPIN (HCG) TRIGGER PREDICTS LOWER<br />

PREGNANCY RATE DURING IN VITRO FERTILIZATION (IVF)<br />

CYCLES. S. Chang, K. Thornton, H. Lieman, S. K. Jindal, E. Buyuk.<br />

Albert Einstein College of Medicine / Montefiore Medical Center, Bronx,<br />

NY.<br />

OBJECTIVE: The effect of late follicular P4 elevation on IVF success<br />

has been previously evaluated, with conflicting results. Supraphysiologic<br />

estradiol (E2) and P4 elevations may be detrimental to endometrial receptivity,<br />

resulting in premature closure of the implantation window, and a<br />

subsequent decrease in pregnancy rates. Though there is growing evidence<br />

for the negative effect of elevated P4 on day of hCG trigger, fewer<br />

studies have looked at P4 elevations leading up to day of trigger, which<br />

may more accurately reflect changes in the endometrium than a single<br />

value on day of trigger. We aimed to determine whether elevations in<br />

P4 leading up to, and on day of hCG trigger were associated with<br />

decreased pregnancy rates.<br />

DESIGN: Historical cohort.<br />

MATERIALS AND METHODS: The first IVF/intracytoplasmic sperm<br />

injection (ICSI) cycles of 238 women from January 2009 to August 2014<br />

were included for analysis. Demographic data, stimulation method, infertility<br />

etiology, baseline follicle-stimulating hormone (FSH) and E2, antimullerian<br />

hormone (AMH), number of eggs/embryos, embryo quality,<br />

transfer day, progesterone levels, and clinical pregnancy data were<br />

collected. Main outcome measure was clinical pregnancy rate. Student’s<br />

t-test, Mann-Whitney U, chi-square and logistic regression analyses were<br />

used. Data was expressed as mean SD or as % and p


questionnaires, including report of usual alcohol consumption, before the<br />

first IVF cycle. Women were categorized as non-drinkers, social drinkers,<br />

or daily drinkers. Poisson regression with robust variance adjusted for female<br />

age and the number of embryos transferred and was used to quantify the relation<br />

between alcohol and pregnancy and live birth outcomes of the first IVF<br />

cycle.<br />

RESULTS: In total, 1,961 women reported their typical alcohol consumption<br />

prior to their first IVF cycle. There were 536 (27%) non-drinkers, 1,352<br />

(69%) social drinkers, and 73 (4%) daily drinkers. Nearly half (44%) of social<br />

drinkers reported consuming R5 drinks per week. Among non-drinkers,<br />

34% achieved clinical pregnancy, and 29% had a live birth. These proportions<br />

were greater than what we observed among social drinkers (32% and<br />

26%, respectively), but were not statistically significant (both PR0.32).<br />

Among daily drinkers, 30% achieved clinical pregnancy, but only 19% had<br />

a live birth, which though not statistically significantly lower than that among<br />

non-drinkers (P¼0.18), is clinically meaningful. Compared to non-drinkers,<br />

the relative risk of live birth did not differ among social drinkers (RR¼0.97,<br />

95% confidence interval [CI] ¼0.80-1.<strong>17</strong>) or daily drinkers (RR¼0.70,<br />

CI¼0.40-1.<strong>21</strong>).<br />

CONCLUSIONS: While alcohol consumption prior to IVF treatment was<br />

not statistically associated with differences in the proportion of cycles that<br />

achieved clinical pregnancy or live birth, the decreased proportion of daily<br />

drinkers who achieved a live birth compared to non-drinkers appears clinically<br />

meaningful. Ongoing analyses will examine intermediate outcomes<br />

within the IVF treatment response, and future work should examine cumulative<br />

pregnancy rates over additional cycles and patterns of alcohol consumption<br />

during IVF treatment.<br />

P-341 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

HIGH ESTRADIOL AND ICSI INCREASE THE RISK OF<br />

PLACENTAL COMPLICATIONS IN PREGNANCY. G. D. Royster,<br />

IV, a J. Csokmay, b B. Yauger, b A. DeCherney, a E. F. Wolff, a M. J. Hill. a<br />

a National Institutes of Health, Bethesda, MD; b Walter Reed National Military<br />

Medical Center, Bethesda, MD.<br />

OBJECTIVE: Higher peak estradiol (pE2) during ART has been associated<br />

with lower birth weight, but the role of ICSI is not known. Here we<br />

assess the relationships among pE2 levels, ICSI, and adverse placental obstetric<br />

outcomes.<br />

DESIGN: Retrospective cohort.<br />

MATERIALS AND METHODS: All fresh IVF cycles from 2010-2013<br />

were analyzed. Inclusion criteria were singleton pregnancy. Exclusion<br />

criteria were prior history of placental complication, diabetes or hypertension.<br />

Hypertension disorders include gestational hypertension (GHTN),<br />

mild and severe pre-eclampsia. Growth disorders were defined as small for<br />

gestational age (SGA) or intrauterine growth restriction (IUGR). Composite<br />

placental outcome was any growth or hypertensive disorder. Generalized<br />

estimating equations (GEE) models were utilized to analyze associations<br />

of estradiol on the day of trigger (pE2) and ICSI use with pregnancy complications<br />

and control for confounding variables. Receiver operating characteristics<br />

(ROC) curves and greater than efficiency curves were generated to<br />

evaluate estradiol thresholds for predicting adverse outcomes.<br />

RESULTS: 1595 consecutive ART cycles starts resulted in 466 singleton<br />

live births with 393 patients meeting inclusion criteria. pE2 was positively<br />

associated with SGA, GHTN, severe pre-eclampsia, and composite placental<br />

outcomes. ROC analysis showed an area under the curve (AUC) of 0.65 and<br />

0.66 for pE2 predicting HTN and growth disorders respectively. ICSI was<br />

positively associated with SGA and composite placental complications. In<br />

adjusted GEE models, only pE2 (OR 1.33, 95% CI 1.13-1.52) and ICSI<br />

(OR 4.<strong>21</strong>, 95% CI 1.73-10.25) were associated with composite placental<br />

complications, whereas male factor infertility and female age were not. Interaction<br />

testing of ICSI versus IVF with pE2 and composite placental complications<br />

showed P


499 IU/L were at 9.7-fold greater odds of failing single-dose<br />

MTX than those with peak hCG 372 IU/L were at 4.7-fold greater odds of<br />

requiring additional treatment than those with hCG


OBJECTIVE: To determine whether antim€ullerian hormone (AMH) predicts<br />

good quality supernumerary blastocyst cryopreservation.<br />

DESIGN: Retrospective study.<br />

MATERIALS AND METHODS: First, fresh IVF cycles (n¼247) from<br />

two fertility ceneters, grouped as follows: women < 35 years with AMH<br />

< 1 ng/mL (n¼40) or AMH R 1 ng/mL (n¼77); women R 35 years with<br />

AMH < 1 ng/mL (n¼62) or AMH R 1 ng/mL (n¼68). AMH level was<br />

measured prior to IVF. Ovarian stimulation protocols based on patient age<br />

and AMH level included short Gonadotropin Releasing Hormone (GnRH)<br />

agonist, GnRH antagonist, or GnRH agonist microdose flare. Supernumerary<br />

good quality blastocysts were cryopreserved on days 5 or 6 post-retrieval.<br />

Primary outcomes measure was supernumerary good quality blastocyst cryopreservation<br />

in relation to AMH levels. Logistic regression was used for statistical<br />

analyses.<br />

RESULTS: Among women < 35 years of age, there was a significant difference<br />

in supernumerary good quality blastocyst cryopreservation between<br />

groups of AMH < 1 ng/mL and AMH R 1 ng/mL (30.0 % vs 58.4 %) when<br />

adjusted for age. Among women R 35 years of age, there was a significant<br />

difference in supernumerary good quality blastocyst cryopreservation between<br />

groups of AMH < 1 ng/mL and AMH R 1 ng/mL (16.1 % vs 42.6<br />

%), when adjusted for age.<br />

CONCLUSIONS: Low AMH levels are associated with a significantly<br />

lower likelihood of blastocyst cryopreservation as compared to higher<br />

AMH levels. This effect was seen among women < 35 years of age and those<br />

R 35 years of age. Patient counseling should include realistic expectations<br />

for the probability of good quality supernumerary blastocyst cryopreservation.<br />

P-347 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

IS ADDING ESTRADIOLE (E2) TO PROGESTERONE (P) LUTEAL<br />

SUPPORT IN HIGH RESPONDER LONG GNRH AGONIST ICSI CY-<br />

CLES DETRIMENTAL TO OUTCOME? : RANDOMIZED<br />

CONTROLLED TRIAL (RCT). M. Ghanem, a M. H. Bedairy, a<br />

A. S. Helal, a A. Shaaban. b a Mansoura Faculty of Medicine, Mansoura,<br />

Egypt; b Ob.Gyn, Al-Azhar Faculty of Medicine, Cairo, Egypt.<br />

OBJECTIVE: Retrospective studies have shown that high response in<br />

terms of excess egg production and high levels of sex steroids are detrimental<br />

to cycle outcome. The primary aim of this RCT is to find if adding E2 to P<br />

luteal support in high responders ICSI cycles impairs cycle pregnancy rate<br />

(CPR) and implantation rate (IR) Secondary aim is to find predictors of<br />

outcome.<br />

DESIGN: RCT.<br />

MATERIALS AND METHODS: Using computer generated random<br />

numbers, group allocation was blindly done at time of ovum pick-up.Arm<br />

I (102 cases) received E2 valerate 6 mg /day plus daily IM injection of 100<br />

P from day of OPU to the time of pregnancy test , arm II (112 cases) received<br />

only daily IM100 mg P. Inclusion criteria: Female age %39 ,fresh first ICSI<br />

trial, basal FSH< 10 mIU/mL, low risk of OHSS, number of eggs retrieved<br />

>15. E2,P assay on days of hCG, mid-luteal were done .Means were<br />

compared by using T-test , ratios by Chi square tests, and predictive values<br />

were measured by ROC curve. The differences were considered significant<br />

if P was < 0.05.<br />

RESULTS: Baseline features were comparable : female age, basal FSH,<br />

number of eggs retrieved, hormone profile on day of hCG-day, number of<br />

embryos transferred and proportions of blastocyst and cleavage embryoes<br />

and ease of transfer (ET).The two arms had comparable CPR<br />

(48.4,44.2& 44.1,42,4) and implantation rates(18.3,18.5) respectively<br />

(p>0.05). Hormonal profiles showed no significant differences in E2 and P<br />

levels on hCG and mid-luteal days. The only significant difference between<br />

groups was significantly lower mean drop of E2 level in arm I from hCG-day<br />

to midluteal level (p¼0.028). Study of predictors of CPR revealed that only<br />

lower level hCG-day P {(AUC curve ¼0.357 ,(95% CI ¼0.270 to<br />

0.445)(p¼0.003)} and lower mean E2 drop from hCG -day to midluteal<br />

(AUC ¼ 0.390, 95% CI¼0.297 to 0.484) ,p¼0.024) predict pregnancy.<br />

CONCLUSIONS: E2 luteal support in high responders does not impair<br />

outcome , on the contrary pregnant outcome is more linked to lower drop<br />

of E2 in luteal phase implying that E2 luteal support may be beneficial in<br />

high responders contrary to what was expected. While lower P level on<br />

hCG-day is significantly linked to pregnant outcome , no association was<br />

found with the number of eggs retrieved, nor E2 level on hCG-day or midluteal<br />

E2.<br />

P-348 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

IN A YOUNG COHORT OF PATIENTS, NUMBER OF EMBRYOS<br />

FROZEN INDICATES BETTER PROGNOSIS FOR FUTURE<br />

FROZEN CYCLE. E. M. Murphy, S. Amrane, A. P. Melnick,<br />

O. K. Davis, Z. Rosenwaks. The Ronald O. Perelman and Claudia Cohen<br />

Center for Reproductive Medicine, Weill Cornell Medical College, New<br />

York, NY.<br />

OBJECTIVE: To compare in vitro fertilization (IVF) cycles of patients<br />

who had a negative pregnancy test in both fresh and frozen cycles to those<br />

who had a negative pregnancy test in the fresh cycle with a live birth in a subsequent<br />

frozen cycle.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: All patients undergoing a fresh IVF cycle<br />

with a subsequent frozen IVF cycle between January 2001 and December 2013<br />

were included. Patients who did not have an embryo transfer, had genetic<br />

testing on the embryos, and/or were using donor embryos were excluded. Patients<br />

meeting criteria were separated into two groups. Patients in group one<br />

(n¼261) had a negative pregnancy test in the fresh and frozen cycle. Patients<br />

in group two (n¼2<strong>21</strong>) had a negative pregnancy test in the fresh and a live birth<br />

in frozen cycle. Statistical analysis using STATA software included t-test and<br />

Chi-square. P value


ate than those with non-eSET (79/115 (68.7%) vs 99/<strong>21</strong>4 (46.3%), respectively,<br />

P


cycles and stored at -20 C. Media samples were collected, shipped, stored<br />

and analyzed between December 2014 and April <strong>2015</strong>. Blastocyst grading<br />

was carried out according to Gardner’s criteria before ET. The media samples<br />

were shipped frozen to the analytical laboratory and subjected to matrix assisted<br />

laser desorption ionization (MALDI), time of flight (ToF) mass spectrometry<br />

(MS). MALDI ToF mass spectrometry to identify pattern<br />

differences in aneuploid and non-aneuploid blastocysts. Data from spectra<br />

were collected from the region 3,000 to 30,000 m/z, smoothed and normalized.<br />

Qualitative characteristics of the data from spectral hotspots in 6000 m/<br />

z to 9500 m/z regions were examined for difference.<br />

RESULTS: After blastocyst grading 19 were assigned 5AA grades and<br />

correlated well with positive pregnancy (16/19, 84.2%), 18 were given mixed<br />

grades but resulted in a negative pregnancy, 3 had an abnormal PGS and were<br />

not graded or transferred. Secretome patterns following MALDI ToF MS<br />

analysis in the 6000 m/z to 9500 m/z regions show distinct pattern differences<br />

between high grade embryos, embryos which resulted in a negative pregnancy,<br />

and aneuploid embryos. In particular, aneuploid blastocysts give<br />

rise to significant profile differences in the 8500-9000 m/z regions such<br />

that they are completely distinct from either high grade embryos or embryos<br />

which resulted in a negative pregnancy.<br />

CONCLUSIONS: Non-invasive analysis of spent blastocyst culture media<br />

by MALDI ToF MS can identify aneuploid blastocysts amongst other high<br />

and low quality blastocysts.<br />

P-353 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

PRESENCE OF DARK GRANULOSE CELLS FOLLOWING PRO-<br />

LONGED ANTAGONIST ADMINISTRATION AND ITS IMPACT<br />

ON EMBRYO QUALITY. S. Ghosh, a G. B.S., a R. Chattopadhyay, b<br />

S. K. Goswami, b G. Bose, b M. Goswami, b B. Chakravarty. b a Assisted<br />

Reproduction, Institute of Reproductive Medicine, Kolkata, India; b ART,<br />

Institute of Reproductive Medicine, Kolkata, India.<br />

OBJECTIVE: To evaluate the relation between duration of antagonists to<br />

levels of reactive oxygen species in the follicular fluid containing dark granulosa<br />

cells, and to examine its association with embryo quality.<br />

DESIGN: Prospective study done between December 2013 to November<br />

2014.<br />

MATERIALS AND METHODS: A total of 796 women undergoing antagonist<br />

cycle in vitro fertilization (IVF) for tubal factor were divided into two<br />

groups. Women with polycystic ovary syndrome (PCOS), and endometriosis<br />

were excluded from the study. All patients underwent standard controlled<br />

ovarian stimulation with recombinant FSH from day 2 of the cycle and<br />

were given antagonist as in flexible protocol. 254 women (Group A) received<br />

antagonist (Inj. Orgalutran, 0.25mg s.c) for more than 4 days before ovulation<br />

trigger. The rest 542 women (Group B) were administered antagonist<br />

in same dosage for 1-3 days before hCG was given. In group A, 58 women<br />

(22.83%) and in group B, 26 women (4.8%) showed dark granulosa cells<br />

along with oocyte-corona-cumulus complex in the follicular aspirate. Such<br />

follicular fluid was collected for estimation of reactive oxygen species<br />

(ROS) and it was assessed by chemiluminescence method in these patients.<br />

Levels of ROS and embryo quality were subsequently correlated. Statistical<br />

comparisons were performed using student’s t test. ’P’ value < 0.05 was<br />

considered statistically significant. Study was approved by Institutional Review<br />

Board.<br />

RESULTS: Women who had dark granulosa cells following antagonist for<br />

more than 4 days showed significantly higher level of ROS and fragmentation<br />

compared to women who had received antagonist for less than four days.<br />

CONCLUSIONS: Increased duration of exposure to antagonist may have<br />

detrimental effect on the follicular fluid milieu, as evidenced by high ROS<br />

level, resulting in poor quality embryos.<br />

Results.<br />

Parameter Group A Group B P Value<br />

Reactive oxygen<br />

species (ROS) cpm/106<br />

Significant fragmentation<br />

(RGrade 3 embryo)<br />

95.38 7.9 48.84 4.89 0.001<br />

54.46 2.<strong>17</strong> 14.<strong>17</strong> 2.18 0.001<br />

P-354 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

BLASTOCYST IMPLANTATION IS CORRELATED WITH OUT-<br />

PUTS FROM AUTOMATED TIME-LAPSE ANALYSIS BY THE<br />

EEVA TEST. J. Liebermann, a A. Bartolucci, b S. Troup, c C. Wagner<br />

Coughlin, d B. Yee, e B. Behr. f a Fertility Centers of Illinois, Chicago, IL;<br />

b CARS, Farmington, CT;<br />

c Hewitt Fertility Center, Liverpool, United<br />

Kingdom; d Aparent IVF Laboratory, Highland Park, IL; e Reproductive Partners,<br />

Redondo Beach, CA; f Stanford Fertility and Reproductive Medicine<br />

Center, Palo Alto, CA.<br />

OBJECTIVE: The automated, time-lapse enabled Eeva Test provides test<br />

scores that predict blastocyst formation (1) and correlate to implantation (2).<br />

Specifically, positive correlation was demonstrated between Eeva Test results<br />

and implantation/pregnancy following Day 3/Day 5 transfer (2). However,<br />

the relationship between Eeva Test results and blastocyst implantation has<br />

not yet been examined. Furthermore, there is no data to inform the implication<br />

for patients in different age groups.<br />

DESIGN: Retrospective multi-center study.<br />

MATERIALS AND METHODS: The study included a total of 342 blastocyst<br />

transfer patients from 7 centers (2012-2014) who consented to have<br />

embryos imaged using the Eeva System, a platform technology that automatically<br />

measures P2 (time between 2- and 3- cell) and P3 (time between 3- and<br />

4- cell) and generates an Eeva Test score of High, Medium, or Low to indicate<br />

the embryo developmental potential. Embryos with known implantation data<br />

were included in this study. P-values were calculated using c2 or student’s t-<br />

test (SAS 9.3).<br />

RESULTS: The overall known implantation rate for the study was 38%<br />

(148/386). Eeva High blastocysts had a relative 40% higher implantation<br />

rate than Eeva Low blastocysts (p¼0.002), while the age was similar between<br />

the two groups. Further analysis revealed that in older patients (ageR35),<br />

Eeva High blastocysts had over twice the implantation rate than Eeva Low<br />

blastocysts (p¼0.002).<br />

Eeva<br />

results<br />

Implantation<br />

rate<br />

Age<br />

Implantation<br />

rate (Age


of this study were infertile couples underwent IVF. The measurement of<br />

serum AMH and estradiol level was performed at the beginning of IVF cycles,<br />

while follicular fluid AMH, were measured on the day of oocytes retrieval.<br />

Oocytes quality was measured using Xia morphology criteria, including the<br />

assessment of polar bodies, perivitelline space and cytoplasmic granulation.<br />

Pearson correlation and linear regression statistical analyses were performed<br />

to determine the predictive value of AMH for oocytes quality.<br />

RESULTS: We obtained 102 IVF patients with antagonist protocol. Serum<br />

AMH, follicular fluid AMH, serum estradiol, number of mature oocytes, and<br />

oocytes morphological score were assessed. Serum AMH has better correlation<br />

to oocytes morphological score compare to follicular fluid AMH and<br />

serum estradiol (r¼0.804 vs r¼0.525 vs r¼0.278). On the other hand, age<br />

has negative correlation to oocytes morphological score (r¼-0.389). Based<br />

on multivariate analysis, we found that age and serum AMH level are the<br />

best predictor for oocytes quality.<br />

CONCLUSIONS: Serum AMH, but not the follicular AMH, can be used to<br />

predict oocytes quality.<br />

P-356 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

INVESTIGATION OF ZONA PELLUCIDATHICKNESS VARIATION<br />

AND IMPLANTATION RATE - WITH AND WITHOUT ASSISTED<br />

HATCHING. E. I. Lewis, a,b R. Farhadifar, c D. Needleman, c<br />

S. A. Missmer, a,b,d L. V. Farland, a,b,d C. Racowsky. a,b a Center for Infertility<br />

and Reproductive Surgery, Dept of OBGYN, Brigham and Women’s Hospital,<br />

Boston, MA; b Harvard School of Medicine, Boston, MA; c School of Engineering<br />

and Applied Sciences, Dept of Molecular and Cellular Biology,<br />

Harvard University, Cambridge, MA; d Harvard School of Public Health,<br />

Boston, MA.<br />

OBJECTIVE: Recent literature has proposed that zona pellucida thickness<br />

variation (ZPTV) in cleavage stage embryos is associated with implantation<br />

rate (IR), but studies have been small and have only used a crude ZPTV<br />

assessment. We utilized novel image processing software to assess ZPTV<br />

to evaluate the relation between ZPTV and IR. In addition, we investigated<br />

whether this relation was affected by assisted hatching (AH).<br />

DESIGN: Pilot retrospective cohort study.<br />

MATERIALS AND METHODS: 349 embryos (56% fertilized with ICSI)<br />

with known implantation results (i.e. 0% or 100%) were assessed for ZPTV<br />

from single and double day 3 embryo transfers performed between 3/2014-<br />

12/2014. Embryo photographs were analyzed using an automated image processing<br />

platform to segment the zona pellucida (ZP) with an active contour<br />

technique. From each image, 100 data points were obtained of ZP thickness<br />

(ZPT) equidistant around the perimeter of each embryo to calculate ZPTV<br />

(¼ [maximum ZPT-minimum ZPT]/mean ZPT). Logistic regression adjusted<br />

a priori for female age, embryo quality, ICSI, Day 3 FSH, male factor infertility,<br />

IVF attempt number, and AH was used to calculate the odds ratio (OR) of IR by<br />

tertile of ZPTVand 95% confidence intervals (CI) / 2-sided Wald p-values (p).<br />

RESULTS: In the adjusted model, there was no significant association between<br />

ZPTV and Sac-IR (sacs/embryos transferred). The highest tertile of<br />

ZPTV (>30%) compared to the lowest tertile ZPTV (22%), was associated<br />

with 57% lower odds of Fetus-IR (viable fetuses @ >7 wks/embryos transferred)<br />

(p¼0.02). Minimum, maximum, and mean ZPT were not associated<br />

with IR. Among embryos without AH, those in the highest tertile of ZPTV<br />

had 88% lower odds of Fetus-IR compared to those in the lowest tertile of<br />

ZPTV (p¼0.01). However, among embryos with AH, ZPTV was not associated<br />

with Fetus-IR.<br />

CONCLUSIONS: As high ZPTV was associated with lower Fetus-IR,<br />

ZPTV measurement may be a useful adjunctive variable for selecting day<br />

3 embryos for ET. In addition, AH may negate the inverse relation between<br />

high ZPTVon Fetus-IR. Further investigation is warranted to establish the association<br />

between ZPTV and IR and to evaluate the impact of AH.<br />

Associations between ZPTV and IR with and without AH OR (95% CI).<br />

ZPTV Tertile 1(T1) ZPTV Tertile 2(T2) ZPTV Tertile 3(T3)<br />

Sac-IR 30 (26)1.00 (ref) 31 (27)1.06 (0.60-1.88)<br />

p¼0.83<br />

Fetus-IR 28 (24)1.00 (ref) 27 (23)0.90 (0.52-1.59)<br />

p¼0.73<br />

Fetus-IR: <strong>17</strong> (37)1.00 (ref) 16 (35)0.75 (0.33-1.70)<br />

No AH<br />

p¼0.49<br />

Fetus-IR: AH 11 (16)1.00 (ref) 11 (16)1.07 (0.45-2.53)<br />

p¼0.88<br />

22-30% (n¼1<strong>17</strong>) >30% (n¼115)<br />

19 (<strong>17</strong>)0.56 (0.28-1.11)<br />

p¼0.10<br />

15 (13)0.43 (0.22-0.85)<br />

p¼0.02<br />

7 (15)0.22 (0.07-0.67)<br />

p¼0.008<br />

8 (12)0.76 (0.34-1.68)<br />

p¼0.49<br />

Supported by: Ferring REI Research Grant awarded by New England<br />

Fertility Society.<br />

P-357 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

RETROSPECTIVE CASE CONTROL STUDY TO EVALUATE THE<br />

EFFECT OF ENDOMETRIOSIS AND RELATED CLINICAL CON-<br />

DITIONS WITH ELEVATED TNFA LEVELS ON EMBRYO PRO-<br />

DUCTION AND IMPLANTATION DURING CYCLES OF<br />

ASSISTED REPRODUCTION. M. Paczkowski, a,b T. Wincek, a,b<br />

J. F. Pliego, a,b T. J. Kuehl. a,b a Obstetrics and Gynecology, Baylor Scott &<br />

White Healthcare, Temple, TX; b College of Medicine, Texas A&M Health<br />

Science Center, Temple, TX.<br />

OBJECTIVE: Conditions that lead to elevated peritoneal levels of tumor<br />

necrosis factor alpha (TNFa) can result in ovarian dysfunction and impaired<br />

implantation. This study aimed to test the hypothesis that patients with clinical<br />

conditions associated with increased TNFa have reduced development<br />

of in vitro fertilized embryos to the blastocyst stage and reduced implantation<br />

and pregnancy.<br />

DESIGN: Retrospective study.<br />

MATERIALS AND METHODS: A retrospective case-controlled study<br />

with 2 matched controls for each of 43 cases with endometriosis, Crohn’s disease,<br />

or arthritis undergoing IVF/ICSI cycles for infertility was performed.<br />

Controls matched affected patients for age within 3 years and were performed<br />

close to the same time. Groups were compared for age, days of<br />

FSH stimulation, total dose of FSH, number of ova retrieved, number of<br />

mature ova, percent maturation, percent fertilized per mature ova, percent<br />

developing to the blastocyst stage, number of embryos transferred, number<br />

of embryos implanting, number of embryos cryopreserved, types of implantation<br />

(singleton, twin or triplet), percent clinical pregnancies, and percent<br />

maintaining pregnancy to 30 wks by March <strong>2015</strong>. Comparisons used Student’s<br />

t test, chi-square test and for pregnancy, Mantel-Haenszel matched<br />

pairs comparison.<br />

RESULTS: Affected patients and controls did not differ for any of the variables<br />

related to treatment, ova maturation, fertilization or development (p ><br />

0.19). Number of embryos implanting (p ¼ 0.06), type of implantation (p ¼<br />

0.07), and patients with gestational sacs (p ¼ 0.08) approached significance<br />

and tended to decrease in affected patients. Mantel-Haenszel odds ratio for<br />

presence of gestational sacs for affected patients versus controls was 0.48<br />

(95% CI of 0.22 to 1.05). However, as pregnancies progressed, more gestational<br />

sacs in affected patients were lost so that 47% of affected patients<br />

versus 58% in controls (p ¼ 0.<strong>21</strong>) maintained pregnancies to beyond 30 wks.<br />

CONCLUSIONS: There was no difference in stimulation, oocyte recovery,<br />

oocyte maturation, fertilization, and blastocyst development for affected<br />

patients compared to control patients. When similar numbers of embryos<br />

were transferred in the fresh cycle the pregnancy outcomes were similar between<br />

affected patients and controls suggesting that diagnosis and treatment<br />

for conditions with abnormal immune mediators is not a risk factor for<br />

impaired outcome from ART cycles. While there is a tempting trend that patients<br />

with endometriosis, Crohn’s disease, or arthritis could have reduced<br />

implantation, this was not significantly different and the outcomes converged<br />

as the pregnancies progressed.<br />

Supported by: Baylor Scott & White Healthcare, Department of Obstetrics<br />

and Gynecology.<br />

P-358 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

PERINATAL OUTCOME USING TIME-LAPSE SYSTEM AND<br />

REDUCED OXYGEN CULTURE IN IVF PATIENTS. N. Zaninovic,<br />

Q. Zhan, R. Clarke, Z. Ye, J. Malmsten, Z. Rosenwaks. CRM, Weill Cornell<br />

Medical College, New York, NY.<br />

OBJECTIVE: To investigate and compare the perinatal outcomes of patients<br />

using time-lapse system with reduced oxygen vs. standard incubator<br />

with ambient oxygen.<br />

DESIGN: A retrospective analysis of IVF patients with fresh ET resulting<br />

in live births from 2012 to 5/2014 was performed. Frozen ET and PGD/S cycles<br />

were excluded. The incidence of term (TB), pre-term (PTB,


(S; 20% O2, Thermo Forma, USA). In total, 1661 babies were born; 947 from<br />

EmbryoScopeÒ and 714 from standard incubator. Statistical analyses were<br />

performed using X2, ANOVA and logistic regression.<br />

RESULTS: Live birth implantation rate of D3 transferred embryos was<br />

significantly higher in EmbryoScopeÒ vs. standard incubator (48.9% vs.<br />

44.6%, p


Twenty embryos did not survive the warming (96% survival rate). Of 508<br />

transferred embryos, there were 291 pregnancies by serum hCG (57%).<br />

The implantation rate was 250/502 (50%, with 6 cases lacking implantation<br />

data). Blastocyst scoring was determined by Gardner’s criteria before trophectoderm<br />

biopsy. For statistical analysis, the qualities of TRO or ICM<br />

were semi-quantitatively assigned: A¼6, B¼4,C¼2, unclear quality¼1.<br />

The quality of EXP was assigned as its number, from 1 to 6, i.e., 1¼very early<br />

blastocyst to 6¼hatched blastocyst. TQS was the sum of all 3 attributes.<br />

Receiver Operator Characteristic Curve (ROC) and Chi-square tests were<br />

used for statistical analysis for correlation of attributes with pregnancy. Cochran-Mantel-Haenszel<br />

Armitage Tests used to examine the linear trend of<br />

TQS and pregnancy, also analyzed the cut line of TQS for significant better<br />

pregnancy outcome. Due to space limitation, only the set of hCG analysis is<br />

presented.<br />

RESULTS: The significant level of each attributes is summarized in Table<br />

1.TQS is the most sensitive predictor (p < 0.0001). There is a significant<br />

linear trend of TQS with pregnancy (p < 0.0001). When TQS is more than<br />

12, the pregnancy rate is significantly (p


maintain potency and block differentiation. Here we test viable first lineage<br />

differentiation reporter ESCs for use in HTSs for IVF/ART stress and toxicological<br />

stress.To validate the effectiveness of a first differentiated lineage<br />

(Pdgfra-GFP) reporter in embryonic stem cells (ESC) by detecting stress<br />

dose- and time-dependent responses, and detecting additional first lineage<br />

and later lineage markers (Dab2, laminin, AFP, LRP2).<br />

DESIGN: Experimental.<br />

MATERIALS AND METHODS: We obtained published mouse (m)<br />

ESCs transgenic with green fluorescent reporter (GFP) knocked into a<br />

Pdgfra loci and regulated by endogenous factors binding the Pdgfra promoter.<br />

These were tested for stress responses in the presence of LIF (a<br />

potency-maintaining growth factor) or for normal differentiation by<br />

removing LIF. Assays used to test Pdgfra-GFP ESCs included microplate<br />

reader, FACS, cell counts, immunofluorescence, and immunoblots. Dose<br />

and kinetic responses were tested in reporter ESCs for hyperosmotic sorbitol<br />

as we have done previously as this is a widely-used positive control<br />

for stress. Retinoic acid (RA) -a specific first lineage inducer was used as<br />

a positive control.<br />

RESULTS: Hyperosmotic stress in the presence of LIF, or LIF removal<br />

(+/-RA), induces Pdgfra-GFP in a time- and dose-dependent manner as assayed<br />

by immunofluorescence and microplate reader. These stress doses<br />

decreased stem cell population expansion. Only Pdgfra-GFP positive cells<br />

co-expressed first lineage marker DAB2 and stress induced only Pdgfra-<br />

GFP+/DAB2+ cells. With LIF removal there were DAB2+/Pdgfra-GFP+ coexpressors<br />

and also Pdgfra-GFP+/DAB2- cells. This suggests that stress<br />

overrides potency to cause a limited, ‘‘prioritized’’ differentiation but LIF<br />

removal enables a wider array of first lineage and later derivatives of first<br />

lineage that are Pdgfra-GFP+.<br />

CONCLUSIONS: The time and dose-dependence of stress induced differentiation<br />

are a proof-of-concept that the Pdgfra-GFP reporter ESCs can be<br />

used as an HTS for fluids derived from IVF-ART protocols (follicular fluid,<br />

spent media, uterine fluids) and from toxicological substances or new Pharma<br />

encountered by the embryo in vivo. The Pdgfra-GFP differentiation reporter<br />

ESCs should complement the previously reported and patented Rex1-RFP<br />

ESCs and reduce the frequency of false positive and false negative outcomes.<br />

The more limited early lineages induced by hyperosmotic stress - compared<br />

with LIF removal - supports the theory of stress induced prioritized differentiation.<br />

P-365 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

THE COMMON OR CELL-SPECIFIC TRANSCRIPTION FACTORS<br />

BETWEEN OOCYTES AND EMBRYONIC STEM CELLS IN<br />

MICE. K. Kim, S. Lee, K. Park, J. Ko, K. Lee. CHA University,<br />

Gyeonggi-do, Korea, Republic of.<br />

OBJECTIVE: By Yamanaka and Daley group, the six genes (Oct4, Sox2,<br />

c-Myc, Klf4, Nanog, and Lin28) have been shown to induce pluripotency in<br />

somatic cells. Based on the fact that both oocytes and embryonic stem cells<br />

(ESCs) have the potential to reprogram differentiated somatic cells into<br />

pluripotent cells, we conducted the present study to compare gene expression<br />

profiles, especially transcription factors, between oocytes and ESCs for<br />

finding common or cell-specific genes as candidates of novel reprogramming<br />

factors.<br />

DESIGN: This study was conducted to compare gene expression profiles<br />

between oocytes and ESCs for finding safer natural novel regulators of reprogramming.<br />

MATERIALS AND METHODS: Previously, we identified a list of upand-down<br />

regulated genes during oocyte maturation using AB <strong>17</strong>00 Full<br />

Genome Expression Mouse Array. Among these genes, 111 transcription factors<br />

were selected and homemade SuperArray was made on which these transcription<br />

factors were blotted. Total RNA was purified from the GV and MII<br />

oocytes, mESCs, STO and NIH3T3 cells with ArrayGradeTM Total RNA<br />

Isolation Kit, cRNA was synthesized and labeled with biotinylated-UTP<br />

and TrueLabeling-AMPTM 2.0, and finally hybridized with SuperArray. After<br />

selection of a list of common or cell-specific transcription factors, mRNA<br />

expression was confirmed by RT-PCR. The common or specific factors expressed<br />

in oocytes and/or stem cells were identified.<br />

RESULTS: From the SuperArray data, we can select 7 oocyte-specific, 3<br />

mESC-specific, and 6 genes common between oocytes and mESCs. Interestingly,<br />

mESC-specific genes were not found. After confirmation by RT-PCR,<br />

expression of Lhx8, Nobox, Isl2, Nr2e1 and H1foo mRNA was detected only<br />

in oocyte. Three (Lhx8, Nobox and H1foo) were already well-known oocytespecific<br />

genes that have an impact upon oocyte maturation and female<br />

fertility, while two genes (Isl2 and Nr2e1) were firstly confirmed. Exciting<br />

finding was the list of genes detected in both oocytes and mESCs, such as<br />

Klf11, Spic, Spry1, Gbx2, Ell3, and Grhl1, but not in differentiated cells.<br />

Among them, Gbx2 has already reported to have LIF/Stat3-mediated selfrenewal<br />

capacity in stem cells, and its gain of function facilitates reprogramming<br />

of differentiated cells.<br />

CONCLUSIONS: Taken together, list of transcription factors of the present<br />

study per se would be insightful in studying oocyte maturation as well<br />

as reprogramming for stem cell research. Identification of potential key or<br />

master regulators of reprogramming and their functional characteristics in<br />

the oocytes and/or ESCs will not only provide novel way for studying reprogramming<br />

and stem cell characteristics, but also greatly contribute to generating<br />

induced pluripotent stem (iPS) cells that are more relevant and safer<br />

for the future cell replacement therapy.<br />

Supported by: This research was Supported by the Basic Science Research<br />

Program through the National Research Foundation of Korea (NRF) funded<br />

by the Ministry of Education (2009-00938<strong>21</strong>).<br />

P-366 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

NURSING<br />

ESTRADIOL AND PROGESTERONE SUPPLEMENTATION IN<br />

UNITED STATES DONOR OOCYTE AND EMBRYO PROGRAMS:<br />

A PRACTICE SURVEY OF FACTORS INFLUENCING ROUTE<br />

AND TIMING OF ADMINISTRATION. K. R. Hammond,<br />

N. A. Cataldo, M. P. Steinkampf. Alabama Fertility Specialists, Birmingham,<br />

AL.<br />

OBJECTIVE: Estradiol (E) and progesterone (P) supplements for recipients<br />

of donor oocytes and embryos are available in various formulations.<br />

The objective of this study was to survey practice patterns among donoroocyte<br />

programs in regard to timing and route of hormone administration.<br />

DESIGN: Cross-sectional survey.<br />

MATERIALS AND METHODS: A one-page questionnaire was offered to<br />

attendees at a <strong>2015</strong> national third-party reproduction symposium. Descriptive<br />

statistics were compiled, and agreement between practices expressed<br />

by Cohen’s kappa and Bowker’s test.<br />

RESULTS: There were 48 unique responses from 22 states; 2 self-identified<br />

as academic programs and 43 as private practices. The majority of<br />

respondents (85%) were nurses. Programs annually performed 40 (median)<br />

fresh donor egg cycles (range 1-400), 25 (0-150) cycles using cryopreserved<br />

(cryo) embryos from donor oocytes, and 8 (0-24) cycles using<br />

cryo donor oocytes.The modal preferred routes of administration were<br />

oral for E and intramuscular (IM) for P, with little difference in preference<br />

for route among fresh and cryo cycle types. In fresh cycles, oral E was<br />

preferred by 54% over IM, vaginal, patch, and combinations for its ease<br />

(92%), convenience (62%), effectiveness (65%), and cost (58%). In fresh<br />

cycles, IM P was preferred by 66% over vaginal and combinations for its<br />

effectiveness (97%), while programs preferring vaginal P (29%) cited its<br />

effectiveness (71%) and ease (57%). There was strong concordance between<br />

fresh and cryo ET cycles as to preferred E route (k¼0.96), but<br />

only fair concordance between cycle types as to preferred P route<br />

(k¼0.51). Changes in preference for IM in the last year were reported to<br />

be uncommon for either E or P. Patients’ preferred E and P routes were<br />

honored by 24% and 45% of respondents, respectively. Modal hormone<br />

supplement discontinuation was at 10 weeks’ gestation for both E (52%)<br />

and P (47%). Time of E discontinuation ranged from 6-13 weeks, while<br />

P discontinuation ranged from 8-13 weeks. P was continued longer than<br />

E by 11 programs, but shorter than E by none (p¼0.01).<br />

CONCLUSIONS: Route of administration and timing of recipient E and P<br />

supplementation in US third-party reproduction programs are moderately<br />

consistent, with the modal prescription of oral E and IM P until 10 weeks’<br />

gestation. The greatest divergence in practice was found between the IM<br />

and vaginal route for P. While little difference was seen within programs<br />

in E or P route between fresh and cryopreserved embryo/oocyte cycles, programs’<br />

preferences appear entrenched.<br />

e230 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


P-367 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

GENETIC COUNSELING<br />

DEVELOPMENT, VALIDATION AND CLINICAL USE OF AN<br />

EXPANDED PAN-ETHNIC PRECONCEPTION CARRIER GENETIC<br />

SCREENING TEST IN ASSISTED REPRODUCTIVE TECHNOL-<br />

OGY PATIENTS AND DONORS. J. Martin, a B. Rodriguez-Iglesias, a<br />

A. San, b Y. Yuting, b J. Jimenez, a Q. Li, c H. Du, b C. Simon. d a Carrier Genetic<br />

Test, Igenomix, Valencia, Spain; b Binhai Genomics Institute, Tianjin,<br />

China; c Binhai Genomics Institute, Yantian, China; d FIVI Universidad de Valencia,<br />

Igenomix-Valencia, Paterna, Spain.<br />

OBJECTIVE: Development, preclinical validation and accuracy assessment<br />

of NGS-based test for expanded pan-ethnic preconception carrier genetic<br />

screening test for use in assisted reproductive technology patients<br />

and donors.<br />

DESIGN: Retrospective analysis of results obtained from 2,570 analysis<br />

for 1,<strong>17</strong>0 individuals from the gamete donor programs; 1,124 individuals corresponding<br />

to the partner of the patient receiving the donated gamete; and<br />

276 individuals from 138 couples seeking ART using their own gametes.<br />

In addition we performed a short comparative study between our NGS-based<br />

test and a commercially available array-based test when patients moved on to<br />

our screening test.<br />

MATERIALS AND METHODS: NGS of 549 recessive and X-linked<br />

genes involved in severe childhood phenotypes reinforced with 5 complementary<br />

tests covering high prevalent mutations not detected by NGS. Use<br />

of a blinded matching system to achieve a very low risk match for patientdonor.<br />

RESULTS: Preclinical validation included 48 DNA samples carrying<br />

known mutations for 27 genes resulting in a sensitivity of 99%. In the clinical<br />

dataset, 2,161 (84%) samples tested positive with an average carrier burden<br />

of 2.3 per sample. Five percent of the couples using their own gametes were<br />

found to have pathogenic variants conferring high risk for 6 different diseases.<br />

These high-risk couples and patients received genetic counselling<br />

and recommendations for preimplantation genetic diagnosis. For patients<br />

receiving gamete donation, we applied a genetic testing and blinded matching<br />

system to avoid high risk combinations regardless of their carrier burden.<br />

For female donors, 1.94% were positive for X-linked conditions; they<br />

received genetic counselling and were discarded. Interestingly, during our<br />

unintended platform-based test comparative study we detected discordant results,<br />

including an array-based test result positive for the common mutation<br />

in the CF gene (F508del); a separate blinded test was performed always<br />

confirmed NGS-based results.<br />

CONCLUSIONS: We have developed a NGS-based expanded carrier genetic<br />

screening test that, combined with our matching system and genetic<br />

counselling, constitutes a powerful tool to avoid more than 600 Mendelian<br />

diseases in the offspring of patients undergoing ART, achieving enhanced<br />

precision than alternative array-based assays.<br />

P-368 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

ASSESSING TRENDS IN EMBRYO GENDER AFTER PGS:<br />

ARE THERE MALE OR FEMALE PROGENY-DOMINANT<br />

COUPLES? J. Gingold, a,b M. C. Whitehouse, c J. A. Lee, c<br />

T. A. VanWort, c M. Daneyko, c T. Mukherjee, c,a A. B. Copperman. c,a a Obstetrics,<br />

Gynecology and Reproductive Science, Icahn School of Medicine at<br />

Mount Sinai, New York, NY; b Obstetrics/Gynecology and Women’s Health<br />

Institute, Cleveland Clinic Foundation, Cleveland, OH; c Reproductive Medicine<br />

Associates of New York, New York, NY.<br />

OBJECTIVE: Normal spermatogenesis produces equal frequencies of<br />

male and female sperm(1). Couples using in vitro fertilization (IVF) and preimplantation<br />

genetic screening (PGS) for gender selection have expressed<br />

concern that their embryos may have a gender bias. The study aims to identity<br />

whether clinical data support the existence of predominantly male or female<br />

embryo-producing couples.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: IVF couples (n¼116) treated between<br />

February, 2006 -November, 2014 who had R10 embryos (range 10-34)<br />

screened by PGS were analyzed by a two-sided binomial test to calculate<br />

the probability (p) of a comparably or more extreme embryo gender imbalance<br />

due to chance. The male to female embryo ratio was assumed to be<br />

1:1. P-values were adjusted for the false discovery rate (FDR) by the Benjamini-Hochberg<br />

method with significance at p


P-370 Tuesday, <strong>October</strong> 20, <strong>2015</strong><br />

DOES THE PARENTAL ORIGIN OF TRANSLOCATION IMPACT<br />

RATES OF ABNORMAL EMBRYOS? C. K. Celestine, a<br />

M. D. Werner, b J. M. Franasiak, b C. R. Juneau, b N. R. Treff, b<br />

T. Molinaro, b R. T. Scott. b a JSUMC, Neptune, NJ; b RMA, NJ, NJ.<br />

OBJECTIVE: Phenotypically normal patients with balanced translocations<br />

often seek pre-implantation genetic diagnostic testing to prevent transmission<br />

of an unbalanced translocation to offspring. This is a unique<br />

subgroup of patients who are at increased risk for suboptimal outcomes<br />

including clinical pregnancy loss which may be associated with chromosomally<br />

abnormal embryos. The goal of the present analysis is to determine<br />

whether the parental origin of the translocation impacts the degree of embryonic<br />

aneuploidy.<br />

DESIGN: Retrospective Cohort Study.<br />

MATERIALS AND METHODS: Couples undergoing IVF with Preimplantation<br />

Genetic Diagnosis at our center were identified where one partner<br />

carried a balanced translocation. All embryos underwent trophectoderm biopsy<br />

with single nucleotide polymorphism microarray-based screening for<br />

detection of unbalanced translocations and 24-chromosome aneuploidy<br />

screening. Possible outcomes were recorded and included: Balanced/<br />

Euploid, Balanced/Aneuploid, Unbalanced/Euploid, Unbalanced/Aneuploid.<br />

The term ‘balanced’ refers to an embryo that has either a balanced<br />

translocation or no translocation as it was not possible to distinguish these<br />

outcomes. Abnormal embryos comprised the groups of embryos which<br />

were either aneuploid and/or unbalanced. Demographics and outcomes of<br />

maternal and paternal translocation carriers were compared using chi squared<br />

and Mann Whitney U tests. Logistic regression utilizing Generalized Estimating<br />

Equations was used to control for between patient differences and<br />

reduce confounding from maternal and paternal age differences.<br />

RESULTS: 1501 embryos from 271 patients were included in this study<br />

with 36.98% of tested embryos found to be normal or balanced.<br />

734(48.9%) embryos were analyzed from couples with maternal translocation<br />

carriers and 767 (51.1%) from paternal translocation carriers. There<br />

was no significant difference in patient or partner ages or the number of<br />

mean embryos between both groups. When controlling for patient identity,<br />

patient age and partner age, maternal carriers of a chromosomal translocation<br />

had an increased risk of having an unbalanced embryo (OR 1.36; 95%CI<br />

1.04-1.77; p¼0.02). However, due to other aneuploidies, the probability of<br />

being normal for both aneuploidy and translocation screening was equivalent<br />

between maternal and paternal carriers (OR 0.87; 95%CI 0.69-1.15;<br />

p¼0.38).<br />

CONCLUSIONS: Significant difference exists between the number of unbalanced<br />

embryos when translocation carrier origin is paternal compared to<br />

maternal. However, no difference exists in the number of normal embryos.<br />

.<br />

Table 1: Cycle Characteristics.<br />

Paternal (n¼138)<br />

Mean SD<br />

Maternal (n¼133)<br />

Mean SD<br />

p value<br />

Patient age 36.074.71 35.974.46 0.77<br />

Partner age 37.325.<strong>17</strong> 37.184.86 0.92<br />

Biopsied Embryos 5.584.48 5.563.42 0.41<br />

Embryos N (%) N (%)<br />

Abnormal 475 (61.93) 471 (64.<strong>17</strong>) NS<br />

Normal 292 (38.07) 263 (35.83) NS<br />

CONTRACEPTION/FAMILY PLANNING<br />

P-371 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

PHASE III STUDY OFA NEW, LOW-DOSE LEVONORGESTREL IN-<br />

TRAUTERINE CONTRACEPTIVE SYSTEM (LNG-IUS19.5MG)<br />

OVER 5 YEARS OF USE: EVALUATION OF BLEEDING AND<br />

DYSMENORRHEA. T. Faustmann, a A. Nelson, b K. Rosen, c<br />

T. Schmelter, a D. Apter. d a Bayer Pharma AG, Berlin, Germany; b Los Angeles<br />

Biomedical Research Institute, Los Angeles, CA; c Bayer HealthCare<br />

Pharmaceuticals, Whippany, NJ; d Sexual Health Clinic, Helsinki, Finland.<br />

OBJECTIVE: To evaluate the bleeding profile and dysmenorrhea associated<br />

with LNG-IUS19.5mg over 5 years of use.<br />

DESIGN: Phase III study. Women used a new, intrauterine contraceptive<br />

system (IUS; LNG-IUS19.5mg) for 3 years; thereafter they had the option<br />

to continue using it in an extension study for up to 2 more years (5 years<br />

in total).<br />

MATERIALS AND METHODS: Nulliparous or parous women aged 18-<br />

35 years with regular menstrual cycles (<strong>21</strong>-35 days) were recruited.<br />

RESULTS: The full analysis set (n¼1452; mean age 27.1 years; 39.5%<br />

nulliparous) included all women for whom at least one LNG-IUS19.5mg<br />

placement attempt was made. Between the 1st and 2nd 90-day reference intervals<br />

(RIs), the mean number of bleeding/spotting days decreased with<br />

LNG-IUS19.5mg use (39.7 to <strong>21</strong>.1 days); the number gradually declined<br />

thereafter (9.3 days at the 20th 90-day RI). Between the 1st and 20th 90-<br />

day RIs, the mean number and length of bleeding/spotting episodes<br />

decreased from 3.6 to 2.2 and 10.14 to 4.15 days, respectively. Bleeding patterns<br />

(World Health Organization criteria) are shown in the Table.<br />

Table. World Health Organization criteria for menstrual bleeding patterns.<br />

Amenorrhea*<br />

(%)<br />

Infrequent<br />

bleeding y<br />

(%)<br />

Frequent<br />

bleeding z<br />

(%)<br />

1st 90-day RI 0.2 10.4 24.6 60.2<br />

2nd 90-day RI 5.1 <strong>21</strong>.3 10.2 16.8<br />

4th 90-day RI 12.7 26.7 4.2 7.4<br />

(end of Year 1)<br />

12th 90-day RI 20.1 25.7 2.2 2.2<br />

(end of Year 3)<br />

20th 90-day RI<br />

(end of Year 5)<br />

22.6 26.4 2.3 1.5<br />

Prolonged<br />

bleeding x<br />

(%)<br />

* No bleeding/spotting throughout RI; y One or two bleeding/spotting episodes;<br />

z >5 bleeding/spotting episodes; x Bleeding/spotting episodes lasting<br />

>14 days.<br />

Between the 1st and 20th 90-day RIs, the mean number of days with<br />

dysmenorrhea (any severity) declined (14.7 to 3.3 days). The percentage of<br />

women reporting R1 day of moderate/severe dysmenorrhea declined over<br />

time with LNG-IUS19.5mg use (68.8% to 16.9% from the 1st to 20th 90-<br />

day RI, respectively). Thirteen women (0.9%) discontinued due to dysmenorrhea<br />

and 76 women (5.2%) due to bleeding/non bleeding problems.<br />

CONCLUSIONS: Women using a next-generation,low-dose IUS (LNG-<br />

IUS19.5mg) generally experienced shorter, less-frequent bleeding over<br />

time. By the end of 5 years of use, almost a quarter of women were amenorrheic;<br />

a further quarter had infrequent bleeding. In addition, dysmenorrheasubstantially<br />

decreased over time.<br />

Supported by: Study and abstract funded by Bayer HealthCare.<br />

P-372 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

INTERPREGNANCY INTERVAL AND SUBSEQUENT PREG-<br />

NANCY OUTCOMES AFTER DILATION AND<br />

EVACUATION. M. K. Kuwahara. University of Hawaii - OGBYN, Honolulu,<br />

HI.<br />

OBJECTIVE: There is little published on the ideal interconceptual period<br />

following a dilation and evacuation (D&E). The purpose of this study is to<br />

compare outcomes for pregnancies conceived 6 months after D&E.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: The University of Hawaii Abortion<br />

Database was used to identify women who underwent D&E between 14<br />

and 26 weeks gestation and were readmitted with a subsequent pregnancy between<br />

2008 to 2014. The first subsequent pregnancy after D&E was identified<br />

through International Classification Diagnosis-9 (ICD-9) codes. Additional<br />

demographic and clinical data were gathered through chart review. The primary<br />

outcome was the rate of preterm deliveries less than 37 weeks gestational<br />

age. Secondary outcomes included cervical insufficiency,<br />

placentation abnormalities, intrauterine growth restriction, postpartum hemorrhage,<br />

and mode of delivery. Chi-square test and student t-tests were used<br />

to compare categorical and continuous variables, respectively. To demonstrate<br />

a difference of 8.8% in preterm delivery with 80% power and a significance<br />

of 0.05, a sample size of 246 patients in each study arm was required.<br />

RESULTS: A total of 737 D&Es were performed during the study interval<br />

and <strong>21</strong>4 subsequent pregnancies were identified. Outcomes were available<br />

for 184 (86%) of these pregnancies, of which 90 (49%) resulted in live births<br />

>20 weeks gestation. Among the live births, 32 (36%) were conceived 6<br />

e232 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


months from the time of D&E. No differences in demographic or clinical variables<br />

were found between the two groups, nor between those who delivered<br />

preterm or term. The incidence of preterm delivery at less than 37 weeks<br />

gestational age was 9.4% (n¼3) with an interpregnancy interval 6 months<br />

(p¼0.168). No differences in preterm delivery below 34 weeks, postpartum<br />

hemorrhage, placentation abnormalities, intrauterine growth restriction, cervical<br />

insufficiency, or mode of delivery emerged between groups.<br />

CONCLUSIONS: Adverse pregnancy outcomes were not higher in the<br />

group of women who became pregnant less than 6 months after D&E<br />

compared to those who waited longer than 6 months to become pregnant.<br />

Larger studies are needed to confirm these findings and provide evidencebased<br />

recommendations for women desiring pregnancy after D&E.<br />

P-373 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

NO SCALPEL VASECTOMY: 20 YEAR OUTCOMES UTILIZING<br />

COMBINED CAUTERY, CLIP AND FACIAL<br />

INTERPOSITION. K. Chiles, M. Feliciano, M. Goldstein. Urology,<br />

Weill Cornell Medical College, New York, NY.<br />

OBJECTIVE: Vasectomy failure is common (1-10%) and a major cause of<br />

malpractice suits. No-scalpel vasectomy (NSV) is a minimally invasive technique,<br />

but failure rates are dependent on occlusion techniques. In this report,<br />

a combination of 4 occlusion techniques were employed.<br />

DESIGN: Retrospective chart review.<br />

MATERIALS AND METHODS: All men who underwent NSV by a single<br />

surgeon within the past 20 years were included. For NSV the vas was<br />

delivered through a single midline puncture hole under local anesthesia,<br />

excluding vasal vessels and nerves. The vas was hemi-transected in two places<br />

½-1 cm apart using thermal cautery. Intraluminal cautery was performed<br />

on both ends for a distance of 1cm until smoke was observed and the wire tip<br />

then rotated for 10 seconds to ensure a 360o burn. A hemoclip was lightly<br />

placed on the testicular end to prevent sperm leakage until cautery causes<br />

a permanent seal. The abdominal end was completely transected and allowed<br />

to retract into the vasal sheath. The sides of the sheath were grasped with a<br />

hemostat and sealed over the abdominal end with a hemoclip, accomplishing<br />

fascial interposition. The intervening vas segment was excised and the ends<br />

were dabbed with Betadine before retraction into the scrotum. The contralateral<br />

vas was accessed through the same puncture hole and occluded identically.<br />

Post-vasectomy semen analysis (PVSA) is requested after 6 weeks<br />

or 15 ejaculations, and a PVSA after the same interval.<br />

RESULTS: 646 vasectomies were performed over 20 years. Mean patient<br />

age in years was 41.55.8 and partner 38.73.7. Median # of children was 3.<br />

377/646 men (58%) of men had at least one PVSA performed a median of 55<br />

days after NSV. <strong>17</strong>7/377 (47%) required a second PVSA, of which 138/<strong>17</strong>7<br />

(78%) complied. No pregnancies were reported. One patient had redo NSV<br />

for persistent non-motile sperm on PVSA. One abscess required I&D. Two<br />

hematomas (2-3 cm) were managed conservatively. No chronic (>6 months)<br />

post-vasectomy pain was reported.<br />

CONCLUSIONS: Vasectomy failure can be virtually eliminated utilizing<br />

a combination of 4 different occlusion techniques: 1) intraluminal cautery for<br />

10 seconds; 2) testicular end clip occlusion; 3) facial interposition; and 4)<br />

resection of a ½ - 1 cm segment. This approach can be employed using the<br />

no-scalpel technique through a single midline puncture hole. Delivery of<br />

vas cleanly from its sheath, excluding vasal vessels and nerves may minimize<br />

the incidence of chronic post-vasectomy pain. This belt, suspenders, rope and<br />

wire approach to vasal occlusion minimizes failure and complications.<br />

P-374 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

UNDERSTANDING HOW WOMEN WITH CYSTIC FIBROSIS<br />

MAKE DECISIONS ABOUT FAMILY PLANNING. S. A. Traxler, a<br />

C. A. Schreiber, a V. Chavez, a D. Hadjiliadis, b C. Mollen. c a Obstetrics<br />

and Gynecology, University of Pennsylvania, Philadelphia, PA; b Pulmonary,<br />

Allergy, and Critical Care Division, University of Pennsylvania, Philadelphia,<br />

PA; c Pediatrics/Division of Emergency Medicine, The Children’s Hospital<br />

of Pennsylvania, Philadelphia, PA.<br />

OBJECTIVE: Based on a few reports, contraceptive use in women with<br />

cystic fibrosis (CF) appears to be well below that seen in the general population<br />

of women in the United States. Given that more women with CF are<br />

reaching reproductive age and pregnancy has health complications for this<br />

population, our objective is to explore contraceptive decision-making and<br />

investigate ways to meet contraceptive needs for women with CF.<br />

DESIGN: Qualitative methods via one-on-one, semi-structured interviews.<br />

MATERIALS AND METHODS: We conducted an in-depth interview<br />

study. Participants were included if they were female between the ages of<br />

18 and 45 with a diagnosis of CF. Purposive sampling was used to balance<br />

the sample age range until thematic saturation was reached. Interviews<br />

were conducted in person or on the phone by a trained interviewer. The interview<br />

guide included the domains of pregnancy attitudes, contraception attitudes<br />

and perceptions of fertility. Interviews were audio-taped and then<br />

transcribed, coded independently by two study team members and analyzed<br />

using a modified grounded theory approach.<br />

RESULTS: Twenty-four women were interviewed. Participants reported<br />

pregnancy intentions to be influenced by a sense of urgency to become pregnant<br />

due to a shortened life span and reported that pregnancy intentions are<br />

impacted by personal health status as well as ethical issues concerning impact<br />

of CF consequences on a potential child. Participants reported misconceptions<br />

about contraception such as interactions with other medications and pervasive<br />

skepticism about long-acting reversible contraceptive (LARC) methods. Many<br />

reported that beliefs around CF-related infertility led to non-use of highly<br />

effective methods of contraception. While medical providers were described<br />

as valuable participants in shared decision-making about family planning<br />

and contraception, many reported that providers lack knowledge and miss critical<br />

periods for communication, such as during transitions of care.<br />

CONCLUSIONS: This sample of women with CF described thoughtful<br />

considerations about pregnancy intentions and desire pregnancies to be<br />

planned during times when health is optimal, but misconceptions about<br />

fertility and contraception interfere with their ability to use highly effective<br />

methods of contraception. Medical providers, valued for their expertise in<br />

shared decision-making, can bridge this gap, facilitate fully informed decisions<br />

about contraception, and determine the best recommendations for<br />

women with CF around childbearing.<br />

P-375 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

RELATIONSHIP BETWEEN COPPER IUD COMPLICATIONS AND<br />

ULTRASOUND FINDINGS. S. Fadiloglu, a B. Dilbaz, b E. Fadiloglu, a<br />

S. Dilbaz. a a Etlik Zubeyde Hanim Kadin Hastaliklari ve Dogum Hastanesi,<br />

Ankara, Turkey; b Adiyaman University School of Medicine, Adiyaman,<br />

Turkey.<br />

OBJECTIVE: IUD is a widely used long acting contraceptive, however the<br />

side-effects related to IUD lead to method discontinuation. The aim of this<br />

study is to evalute the relationship between the most common side-effects<br />

of IUD use such as dysmenorrhea, menorrhaghia, abdominal cramping and<br />

the position of the IUD device and uterine dimensions evaluated by transvaginal<br />

ultrasonography.<br />

DESIGN: Prospective Trial.<br />

MATERIALS AND METHODS: Two hundred and eighty four patients<br />

who had IUD insertion at the Family Planning Clinic of a tertiary center<br />

were evaluated at insertion, 6th and 12th week after the insertion. Demographic<br />

characteristics, medical history, symptoms and findings of the gynaecological<br />

examination were recorded. Trnasvaginal ultrasonographic<br />

measurement of uterine dimensions, the distance between the tip of the<br />

Cu-IUD and fundus, endometrium and myometrium were measured in order<br />

to evaluate the displacement of the IUD and the relationship between<br />

displacement and the symptoms and complaints were investigated.<br />

RESULTS: No pregnancy occured during the study period and one patient<br />

had the IUD removed due to menorrhgia. The patients who complained<br />

of menorrhgia had statistically significantly shorter uterine length<br />

in comparison to the ones who did not have this complaint (54.27 6.11<br />

vs 60.25 10.52 mm p¼0.02). There was no difference in the uterine length<br />

between the patients who had dysmenorrhea and who had not at 12 weeks<br />

(59,60 10,25 vs 60,33 10,68mm, p ¼ 0.71 ). There was a significant<br />

difference between the IUD tip-fundus, endometrium and myometrium<br />

measurements of the patients who experienced abdominal cramping in<br />

comparison to the ones who had not after 3 months confirming the downward<br />

IUD movement. ( Endometrium; 0,29 0,72 vs 0,45 0,35 p ¼<br />

0.02, Myometrium; 1,25 1,39 vs 2,38 2,26 p


P-376 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

BARRIERS AND MYTHS THAT LIMIT THE USE OF INTRAUTER-<br />

INE CONTRACEPTION IN NULLIPAROUS WOMEN: A SURVEY<br />

OF BRAZILIAN GYNECOLOGISTS. M. M. Carneiro, a J. Lira, b<br />

A. L. Silva-filho. a a Obstetrics and Gynecology, Federal University of Minas<br />

Gerais, Belo Horizonte, Brazil; b Department of Adolescent Gynecology, Instituto<br />

Nacional de Perinatologia, Universidad Nacional Autonoma de<br />

Mexico UNAM, Mexico City, Mexico.<br />

OBJECTIVE: To understand the extent to which barriers and misperceptions<br />

about intrauterine contraception (IUC) remain among Brazilian gynecolgysts,<br />

particularly for nulliparous women.<br />

DESIGN: Prospective web based survey.<br />

MATERIALS AND METHODS: An online survey was developed to<br />

assess Brazilian gynecologists’ knowledge and attitudes towards IUC. Data<br />

collected included demographic and professional data, main barriers when<br />

considering IUC for women in general, and or nulliparous women, attitudes<br />

towards inclusion of IUC in contraceptive counselling, opinions on what<br />

could increase IUC prescription for nulliparous women. A question<br />

regarding the knowledge about the World Health Organization Medical<br />

Eligibility Criteria (WHO MEC) was also included in the survey.<br />

RESULTS: 101 respondents completed the survey. The insertion rate in<br />

nulliparous women was 79.2%. Brazilian gynaecologists considered IUC<br />

in counseling or provided it upon request more for parous than for nulliparous<br />

women (p


DESIGN: A prospective three-group randomized controlled trial was performed<br />

with three assessment points. Participants (n ¼265) were randomly<br />

allocated to the intervention group (IG, fertility awareness video), control<br />

group 1 (CG1, no intervention) and control group 2 (CG2, work-family conflict<br />

video). All participants were assessed before the intervention (T0),<br />

immediately after (T1) and one week after (T2).<br />

MATERIALS AND METHODS: Data were collected Oct-Dec 2014 from<br />

Portuguese University students. After excluding participants who were married,<br />

had children or disclosed a fertility problem, the final sample had 254<br />

participants (96 IG, 76 CG1, and 82 CG2). The questionnaire included sociodemographic<br />

and fertility knowledge questions. The video gave information<br />

about women’s fertility decline, infertility definition and risk factors, probability<br />

of getting pregnant according to women’s age. Mixed Anovas for<br />

repeated measures tested interaction, time and group effects on fertility<br />

knowledge variables.<br />

RESULTS: The large majority (83%) of participants (84% female, Mage ¼<br />

19.7) expressed a desire to have children. There were no pretest differences<br />

among groups concerning sociodemograhic characteristics. A significant interaction<br />

was found between assessment time and group regarding knowledge<br />

variables (infertility definition, women’s fertility decline, probability of pregnancy<br />

depending on age, risk factors). Medium effect sizes of 0.33 were found<br />

between IG and CG2 and 0.39 between IG and CG1. The IG had significantly<br />

higher knowledge than both control groups in all variables between T0 and T1<br />

and T0 and T2, but not between T1 and T2, indicating that knowledge remains<br />

a week later. There were no significant differences over time within each one of<br />

control groups, excepting for women’s most fertile age, women’s fertility<br />

decline, and risk factors. However, these differences were smaller (higher<br />

eta for IG) and in some cases indicated a decrease in knowledge.<br />

CONCLUSIONS: Results Supported the efficacy of a low-cost intervention<br />

in increasing fertility knowledge. Given the increasing childbearing<br />

postponement and the reduced fertility knowledge of young adults who<br />

desire to have children, this intervention can be of great use in western countries<br />

to enable an informed and conscious decision-making process. Future<br />

research should focus on wider longitudinal studies.<br />

Supported by: Study Supported by FEDER/COMPETE and FCT (PTDC/<br />

MHC-PSC/4195/2012 and SFRH/BPD/85789/2012).<br />

MALE FACTOR<br />

P-380 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

COMPARISON BETWEEN THE SEMEN PARAMETERS ACCORD-<br />

ING TO 1999 WORLD HEALTH ORGANIZATION CRITERIA AND<br />

2010 WHO CRITERIA IN MALES OF SUBFERTILE COUPLES IN<br />

KOREA. E. Lee, a J. Shin, a,b C. Joo, c H. Kim, b,a J. Lee, d,b B. Jee, d,b<br />

C. Suh, b,a S. Kim. a,b a Seoul National University Hospital, Seoul, Korea, Republic<br />

of; b Seoul National University College of Medicine, Seoul, Korea, Republic<br />

of; c Maria Fertility Hospital, Seoul, Korea, Republic of; d Seoul<br />

National University Bundang Hospital, Seoul, Korea, Republic of.<br />

OBJECTIVE: To assess the effects of the new 2010 WHO criteria for<br />

semen analysis in comparison with the previous 1999 WHO criteria.<br />

DESIGN: Retrospective study.<br />

MATERIALS AND METHODS: We retrospectively reviewed the results<br />

of semen analyses of men of 660 subfertile couples from April 2014 until<br />

December 2014. The study was performed at the 6 fertility clinics in Korea.<br />

All semen parameters were reviewed and compared according to 1999 WHO<br />

and 2010 WHO criteria.<br />

RESULTS: The mean values of semen analyses in 660 men of subfertile<br />

couples were 3.6ml of volume, 102.3x10^6/mL of concentration, 54.9% of<br />

motility and 4.5% of strict morphology. Two hundred and seventeen men<br />

(32.8%) who were considered to be infertile by the 1999 WHO criteria<br />

were reevaluated to be normal according to 2010 WHO reference values.<br />

Of the men with semen analyses, 11.2%, 3.0%, 16.7% and 42.7% of men<br />

were reclassified as normal for volume, concentration, motility and strict<br />

morphology, respectively.<br />

CONCLUSIONS: About one-third of men of subfertile couples who were<br />

thought to be infertile by the old criteria now considered normal by the new<br />

2010 WHO criteria. Many men who were considered infertile according to<br />

the old criteria now should be reevaluated normal by the new criteria. Strict<br />

morphology was the main cause of abnormal semen parameters.<br />

References:<br />

1. Baker K, Li J, Sabanegh E Jr, Edmund Sabanegh Jr. Analysis of<br />

semen parameters in male referrals: impact of reference limits, stratification<br />

by fertility categories, predictors of change, and comparison<br />

of normal semen parameters in subfertile couples. Fertil Steril<br />

<strong>2015</strong>;103(1):59-65.<br />

2. Catanzariti F, Cantoro U, Lacetera V, Muzzonigro G, Polito M. Comparison<br />

between WHO (World Health Organization) 2010 and WHO<br />

1999 parameters for semen analysis - interpretation of 529 consecutive<br />

samples. Arch Ital Urol Androl. 2013 Sep 26;85(3):125-9.<br />

Supported by: This research was Supported by a fund(2014ER630500) by<br />

research of Korea Centers for Disease Control and Prevention.<br />

P-381 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

COMPARATIVE PROTEOMIC ANALYSIS INDICATES UNDEREX-<br />

PRESSION OF MOLECULAR CHAPERONES IN SPERMATOZOA<br />

OF INFERTILE MEN. R. Sharma, a A. Agarwal, a<br />

D. Durairajanayagam, a,b L. Samanta, a,c M. Assidi, d S. Gupta, a<br />

E. S. Sabanegh. e a Center for Reproductive Medicine, Cleveland Clinic,<br />

Cleveland, OH; b Physiology, Universiti Teknologi MARA, Sungai Buloh,<br />

Malaysia; c Redox Biology Laboratory, School of Life Science, Ravenshaw<br />

University, Orissa, India; d Center of Excellence in Genomic Medicine,<br />

King AbdulAziz University, Jeddah, Saudi Arabia; e Urology, Cleveland<br />

Clinic, Cleveland, OH.<br />

OBJECTIVE: Functional spermatozoa are produced by a highly regulated<br />

process of sperm maturation. Human semen used in assisted reproductive<br />

procedures is traditionally separated by density gradient to provide morphologically<br />

and functionally viable spermatozoa. Often these spermatozoa are<br />

unable to achieve fertilization. Chaperones are involved in processing of<br />

spermatozoal proteins required for motility, zona penetration and stress<br />

response. We therefore studied a comparative proteome profile of these proteins<br />

to understand their role in sperm function.<br />

DESIGN: Protein profiles from spermatozoa in fertile and infertile men<br />

were compared.<br />

MATERIALS AND METHODS: Proteins were extracted and separated<br />

by 1-D gel. Bands were digested with trypsin and analyzed on a LTQ-Orbitrap<br />

Elite hybrid mass spectrometer system. Protein identification was done<br />

using Mascot (Matrix Science, London, UK; version 2.3.02), SEQUEST<br />

(Thermo Fisher Scientific, San Jose, CA, USA; version 1.4.0.288) and X!<br />

Tandem (TheGPM, thegpm.org; version CYCLONE (2010.12.01.1).<br />

Mascot, Sequest and X! Tandem were set up to search the human reference<br />

with database assuming trypsin as the digestion enzyme. Functional annotations<br />

of proteins were obtained using bioinformatic tools and pathway databases.<br />

RESULTS: Proteomic analysis revealed a total of 35 focus chaperone<br />

proteins involved in post-translational modification and protein folding<br />

that were underexpressed in the spermatozoa of infertile men. The pivotal<br />

chaperones downregulated were HSPA4L, HSPA 1A/1B, mitochondrial<br />

HSP90, subunits of the T-complex proteins 1 (Tcp1a, Tcp 1g, Tcp1d,<br />

Tcp 1ε, Tcp 1h, Tcp 1z) as well as the sperm surface protein <strong>17</strong> responsible<br />

for zona binding. The BAG family regulator 5 was upregulated in these patients,<br />

indicating a negative regulation of oxidative stress and ubiquitination.<br />

CONCLUSIONS: Our results suggest that there is a differential expression<br />

of sperm chaperones between fertile and infertile males. These may be<br />

responsible for defective expression of sperm surface proteins leading to<br />

failed fertilization.<br />

P-382 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

SEMEN PARAMETERS DO NOT HAVE A CLINICALLY SIGNIFI-<br />

CANT EFFECT ON ICSI CYCLE OUTCOMES: AN ANALYSIS US-<br />

ING 2,861 SHARED OOCYTE DONOR RECIPIENT CYCLES TO<br />

CONTROL FOR EGG FACTORS. G. Patounakis, a K. Devine, b<br />

B. W. Whitcomb, c A. DeCherney, a M. J. Levy. b a National Institutes of<br />

Health, Bethesda, MD; b Shady Grove Fertility, Rockville, MD; c University<br />

of Massachusetts, Amherst, MA.<br />

OBJECTIVE: The true impact of male factor on ICSI cycle outcomes is<br />

complicated by concurrent female variables. Our objective was to determine<br />

the effect of semen parameters on ICSI cycle outcomes using a cohort of<br />

shared oocyte donation cycles to completely control for oocyte factors.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Anonymous oocyte donation cycles<br />

from January 2010 through December 2014 using non-donor sperm<br />

FERTILITY & STERILITY Ò<br />

e235


were included. Only the first cycle was included for recipients with multiple<br />

cycles. Prewash volume, concentration, motility, and total motile<br />

sperm count from the ICSI specimen and morphology from the baseline<br />

specimen were analyzed. The change in morphology criteria to WHO<br />

2010 in September 2010 was accounted for in the statistical model.<br />

The primary outcome was percent top quality blastocyst (AA, AB,<br />

BA) conversion on the last day of embryo evaluation per inseminated<br />

oocyte. Secondary outcomes were fertilization, overall blastocyst conversion,<br />

implantation, clinical pregnancy, and ongoing pregnancy. Generalized<br />

estimating equations (GEE) were used to account for donor<br />

oocyte sharing and repeated donor cycles. Outcomes were controlled<br />

for donor age, oocytes retrieved, recipient female age, male age, recipient<br />

endometrial thickness, embryo stage at transfer, embryo quality at<br />

transfer, and number of embryos transferred.<br />

RESULTS: There were 2861 recipients with inseminated oocytes in the<br />

cohort of 1057 oocyte donors undergoing 1659 retrievals. The percent of<br />

oocyte donors shared with 3, 2, and 1 recipients were 57%, 26%, and <strong>17</strong>%,<br />

respectively. Only WHO 2010 normal morphology showed a significant association<br />

with the primary outcome. For every 1% increase in WHO 2010<br />

normal morphology the top quality blastocyst conversion increased by<br />

0.4% (p¼0.04). A 6% higher top quality blastocyst conversion was observed<br />

when morphology was >2% relative to %2% (p¼0.001). No other effects of<br />

normal morphology for both WHO 2010 criteria and pre-WHO 2010 were<br />

statistically significant. Pre-wash motility was significantly associated with<br />

secondary outcomes. For every 10% increase in pre-wash motility, fertilization<br />

increased by 0.7% (p¼0.001), implantation increased by 1.6%<br />

(p¼0.004), odds of clinical pregnancy increased by 1.07 (p¼0.004), and<br />

odds of ongoing pregnancy increased by 1.06 (p¼0.01). Pre-wash motility<br />

did not have a significant effect on top quality blastocyst conversion or overall<br />

blastocyst conversion. All other semen parameters lacked statistically significant<br />

associations.<br />

CONCLUSIONS: Sperm motility and morphology have statistically<br />

significant but clinically insignificant effects on ICSI outcomes. Couples<br />

with male factor infertility can be reassured that currently measured<br />

bulk semen parameters are not clinically useful outcome predictors of<br />

ICSI.<br />

Supported by: This work was Supported, in part, by the Program in Reproductive<br />

and Adult Endocrinology, NICHD, NIH, Bethesda, MD.<br />

P-383 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

ACHIEVING PREGNANCY IN MEN WITH ISOLATED<br />

ASTHENOSPERMIA. C. Deibert, K. M. Zeeck, J. Sandlow. Medical<br />

College of Wisconsin, Milwaukee, WI.<br />

OBJECTIVE: Men with isolated asthenospermia present a clinical challenge<br />

regarding treatment. To date no studies have evaluated the clinical outcomes<br />

for couples with isolated asthenospermia.<br />

DESIGN: This is a retrospective chart review study.<br />

MATERIALS AND METHODS: This is an IRB-approved review of infertile<br />

men who presented to our Reproductive Medicine Center with isolated<br />

asthenospermia, defined as motility


DESIGN: A prospective study.<br />

MATERIALS AND METHODS: Sperm from individuals diagnosed as<br />

normozoospermia (n¼43) or asthenozoospermia (n¼32) were collected.<br />

Liquid chromatography-tandem mass spectrometry were used to detect<br />

levels of total 5-methylcytosine (5-mdC) and 5-hydroxymethylcytosine<br />

(5-hmdC) in sperm DNA as well as N6-methyl-adenosine (m6A) and<br />

5-methylcytosine (5-mC) in RNA. Levels of the above four indicators<br />

were compared. Linear and logistic regression models were performed to<br />

determine the factors affecting IVF outcomes, adjusting for confounders.<br />

RESULTS: Demographic characteristics including fertilization rate (FR),<br />

cleavage rate (CR), good quality embryo rate (GQER) and pregnancy rate<br />

(PR) were similar between asthenozoospermia (AS) group and normozoospermia<br />

(NM) group, except for that sperm motility and concentration<br />

were decreased in AS group. The level of 5-mdC (representing global<br />

DNA methylation status) were found to be significantly decreased in AS<br />

group (3.430.65% versus 3.730.57% in NM group, P¼0.042). Linear<br />

regression analysis revealed that 5-mdC level was positively correlated<br />

with FR (b 8.11, 95% CI (1.02,15.20), P¼0.043) and GQER in AS group<br />

(b 16.18, 95% CI (2.294,30.067), P¼0.024), but was not related to cleavage<br />

rate (P¼0.084). In contrast, level of 5-hmdC (standing for DNA demethylation<br />

status) was negatively correlated with FR in AS group (b -25.12, 95% CI<br />

(-49.82,-0.42), P¼0.047)). The other two indicators were found to have no<br />

influence on IVF outcomes in our study. Either of the four indicators had significant<br />

influence on clinical pregnancy as logistic regression analysis revealed.<br />

CONCLUSIONS: Global sperm DNA methylation level was significantly<br />

decreased in asthenozoospermia samples. Besides, DNA methylation/demethylation<br />

status posed a potential influence on fertilization and embryo development<br />

in these patients. This indicates that even though current ART<br />

techniques help to select highly motile sperm in asthenozoospermia patients,<br />

epigenetic alterations may still exist in these sperm, which may affect the<br />

fertilization process and embryo development in IVF.<br />

P-386 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

ASSESSING A COMPREHENSIVE CHROMOSOMAL ANALYSIS<br />

OF HUMAN SPERMATOZOA BY NEXT GENERATION<br />

SEQUENCING. S. Cheung, Q. V. Neri, Z. Rosenwaks,<br />

G. D. Palermo. Reproductive Medicine, Weill Cornell Medical College,<br />

New York, NY.<br />

OBJECTIVE: To carry out a complete molecular karyotype on sperm cells<br />

by unraveling, extracting and amplifying the compacted male genome in<br />

adequate quantity and quality in comparison to FISH.<br />

DESIGN: Following extraction and amplification of good quality sperm<br />

DNA copy number variations were compared between specimens to assess<br />

the occurrence of spermatogenetic meiotic errors. Aneuploidy outcome<br />

was compared to standard FISH assessment.<br />

MATERIALS AND METHODS: DNA extraction was achieved by a commercial<br />

kit. We processed 48 ejaculates where we extracted DNA and processed<br />

for DNA amplification. In 12 specimens where only few<br />

spermatozoa were processed, satisfactory DNA was obtained following<br />

PCR-based random hexamer amplification. Fifteen of these specimens<br />

were processed by NGS and the copy number variations were recorded<br />

and compared using CASAVA and VarScan2 software programs. FISH on<br />

chromosomes X, Y, 13, 15, 16, <strong>17</strong>, 18, <strong>21</strong>, and 22 was carried out on spermatozoa<br />

from fertile men as well as men with history of recurrent pregnancy<br />

loss. Following NGS, a patient sample was compared to that of a fertile anonymous<br />

donor’s.<br />

RESULTS: We extracted good quality DNA ranging from 500,000 to as<br />

low as 250 spermatozoa.Assessment of 9 chromosome FISH on spermatozoa<br />

obtained from fertile men (n¼7) the total aneuploidy rate was only 0.2%,<br />

whereas men with recurrent ART failure (n¼36) had a total aneuploidy of<br />

3.92.5%. When we ranked the data according to paternal age, we established<br />

that the sperm aneuploidy rate progressively increased with advancing<br />

age to reach 10.4% at the average age of 6811yrs particularly for chromosomes<br />

15 and <strong>17</strong>.Following NGS, our data presented a total percentage of<br />

<strong>21</strong>.47% of aneuploidy rate. The most represented was gonosomal with Y<br />

monosomy and disomy at 4.2%, followed by chromosome 15 at 3.3%, and<br />

chromosome 14 at 1.6%. On the other extreme chromosomes 18 and <strong>21</strong><br />

had the lowest disomy at 0.10 and 0.15, respectively.When compared to standard<br />

FISH assessment, the NGS aneuploidy results depicted higher aneuploidy<br />

than FISH results overall. Moreover, NGS copy number variations<br />

for individual chromosomes indicated higher aneuploidy levels for all chromosomes<br />

screened for by FISH, with the exception of chromosomes 13, 18,<br />

and <strong>21</strong>. The remaining chromosomes X, Y, 15, 16, <strong>17</strong>, and 22, were actually<br />

found to have more than twice the aneuploidy value when assessed by NGS.<br />

CONCLUSIONS: Sperm aneuploidy assessment help in counseling male<br />

factor infertile couples or those with recurrent pregnancy losses. FISH has<br />

been the procedure used for this assessment, however, it has a limited number<br />

of chromosomes evaluated and not to mention the inherent accuracy concerns<br />

related to the procedure and the inability when assessing haploid specimen to<br />

discern through nullisomy. NGS proved to be the procedure capable of assessing<br />

aneuploidy. It can assess a wide range of cells from 500,000 to<br />

10million, it assess all chromosomes including nullisomy. In addition, there<br />

is the ability to assess the copy number variations.<br />

Supported by: WCMC.<br />

P-387 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

HIGH THROUGHPUT INTEGRATED PROTEOMIC ANALYSIS OF<br />

SPERMATOZOAL PROTEINS IN PATHOPHYSIOLOGY OF VARI-<br />

COCELE ASSOCIATED MALE INFERTILITY. A. Agarwal, a<br />

D. Durairajanayagam, a,b R. Sharma, a L. Samanta, a,c R. F. Turki, d<br />

E. Sabanegh. e a Center for Reproductive Medicine, Cleveland Clinic, Cleveland,<br />

OH;<br />

b Physiology, Universiti Teknologi MARA, Sungai Buloh,<br />

Malaysia; c Redox Biology Laboratory, School of Life Sciences, Ravenshaw<br />

University, Orissa, India; d Ob/Gyn Specialist, King AbdulAziz University,<br />

Mobil, AL; e Urology, Cleveland Clinic, Cleveland, OH.<br />

OBJECTIVE: Varicocele appears to affect later stages of spermatogenesis.<br />

It causes scrotal hyperthermia, hypoxia, hormonal imbalances, and re-flow of<br />

metabolites from renal and/or adrenal glands leading to oxidative stress. The<br />

objective was to study the major differences in the distribution of spermatozoal<br />

proteins in infertile men diagnosed with varicocele compared to fertile<br />

men.<br />

DESIGN: Prospective proteomic study.<br />

MATERIALS AND METHODS: Proteins were extracted from infertile<br />

men with unilateral and bilateral varicocele (n¼5) and men with proven<br />

fertility (n¼5). 1-D gel electrophoresis followed by LC/MS-MS (LTQ-Orbitrap<br />

Elite hybrid mass spectrometer) was used for protein identification.<br />

Mascot (Matrix Science, London, UK), SEQUEST (Thermo Fisher Scientific,<br />

San Jose, CA, USA) and X! Tandem (TheGPM, thegpm.org) were set up<br />

to search the human reference with database assuming trypsin as the digestion<br />

enzyme. Functional annotations of proteins were obtained using bioinformatics<br />

tools and pathway databases.<br />

RESULTS: Of the 99 proteins that were differentially expressed (DEP) in<br />

the varicocele group, 9 were uniquely expressed in the fertile group<br />

compared to 2 proteins that were unique to the varicocele groups. Over<br />

87% of the DEP involved in major energy metabolism and key sperm functions<br />

were underexpressed in varicocele group. Key protein functions<br />

affected in the varicocele group were spermatogenesis, sperm motility<br />

(ACRBP, SPA<strong>17</strong>, AKA7), and mitochondrial dysfunction (NDUFS1,<br />

UQCRC2).<br />

CONCLUSIONS: We have identified proteins that are underexpressed in<br />

varicocele group. These proteins may be key players involved in the pathology<br />

of varicocele and in the onset of infertility.<br />

P-388 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

CLINICAL OUTCOME OF TREATMENTS FOR AZOO-<br />

SPERMIA. A. Tanaka, a M. Nagayoshi, a I. Tanaka, a S. Ikuma, a<br />

T. Miki, a T. Yamaguchi, a H. Kusunoki, b S. Watanabe. c a Saint Mother Hospital,<br />

Kitakyusyu, Japan; b Faunal Diversity Sciences, Graduate School of<br />

Agriculture, Kobe University, Kobe, Japan; c Anatomical Science, Hirosaki<br />

University Graduate School of Medicine, Hirosaki, Japan.<br />

OBJECTIVE: Azoospermia is found in approximately one out of one hundred<br />

men and the ratio of obstructive azoospermia to non-obstructive one is<br />

about 3:7. The sole treatment for azoospermia is the collection of sperms or<br />

spermatids surgically. We analyzed the clinical outcome following each procedure<br />

over a period of 14 (1999-2013) years. Microsurgical epididymal<br />

sperm aspiration (MESA) for obstructive azoospermia and microscopic<br />

testicular sperm extraction (Micro-TESE) for non-obstructive azoospermia<br />

are generally recommended. However there has not been enough information<br />

available for the clinical data following each procedure.<br />

FERTILITY & STERILITY Ò<br />

e237


DESIGN: Retrospective cohort study to treatments for azoospermia.<br />

MATERIALS AND METHODS: MESA was performed to azoospermic<br />

men who had normal FSH concentration (15mIU/ml).<br />

RESULTS: Clinical outcome was summarized in Table. Clinical data<br />

following MESAwas significantly higher than all other results following Micro-TESE<br />

except miscarriage rates using spermatid. Total of 1<strong>17</strong>1<br />

(M:F¼563:608) babies were born (778 from MESA, 393 from Micro-<br />

TESE) and rate of congenital defect and perinatal complication was 2.7%<br />

(32/1<strong>17</strong>1) (12 cases of heart diseases, 3 cases of annal atresia, 11 cases of<br />

palatal cleft, 3 cases of still birth and 3 cases of <strong>21</strong> trisomy).<br />

CONCLUSIONS: MESA showed a satisfactory clinical outcome. However,<br />

Micro-TESE resulted in lower collection rate and clinical success rate<br />

than reported in other parts of the world. This unexpected lower collection<br />

rate might be related to the homogeneous racial background of the population<br />

of our study. And main causes for lower successful rate might be high percentage<br />

of sperm with low or no motility and heads with anomalies and<br />

low survival after freezing / thawing procedures.<br />

Clinical outcome for azoospermia following MESA or Micro-TESE over a 14<br />

years period.<br />

Pregnancy rates (%)<br />

(per ET)<br />

Miscarriage<br />

rates (%)<br />

Birth<br />

rates (%)<br />

MESA 35y < 36.0% (644/<strong>17</strong>87) 25.9% (167/644) 25.5% (456/<strong>17</strong>87)<br />

35-39y 30.0% (393/1308) 14.2% (56/393) 24.6% (322/1308)<br />

TOTAL 33.5% (1037/3095) <strong>21</strong>.5% (223/1037) 25.1% (778/3095)<br />

Micro-<br />

TESE<br />

(sperm)<br />

Micro-<br />

TESE<br />

(spermatid)<br />

35y < 10.3% (41/397) 24.4% (10/41) 8.8% (35/397)<br />

35-39y 11.5% (46/401) 28.3% (13/46) 8.2% (33/401)<br />

TOTAL 10.9% (87/798) 26.4% (23/87) 8.5% (68/798)<br />

35y < 23.0% (526/2291) 59.1% (311/526) 10.3% (236/2291)<br />

35-39y <strong>17</strong>.5% (231/1320) 64.9% (150/231) 6.7% (89/1320)<br />

TOTAL 20.9% (756/3611) 61.0% (461/756) 9.0% (325/3611)<br />

P-389 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

COMPARISON OF SEMEN QUALITY AND FERTILIZATION PO-<br />

TENTIAL BETWEEN CONSECUTIVE<br />

EJACULATES. A. M. Ragheb. Urology, Faculty of Medicine, Beni Suef<br />

University, Beni Suef, Egypt.<br />

OBJECTIVE: There has been a growing body of evidence on the improvement<br />

of semen quality in sequential semen samples. We aimed to verify this<br />

finding and its implication in the assisted reproduction arena.<br />

DESIGN: A retrospective review of records.<br />

MATERIALS AND METHODS: The study included the records of 102<br />

infertile males with idiopathic asthenozoospermia or oligoasthenozoospermia<br />

attending the Egyptian International Infertility and IVF Center from<br />

January 2014 till December 2014. All the patients had been previously asked<br />

to provide two semen samples (1-3 hours apart) after an abstinence period of<br />

3-7 days. The two consecutive semen samples were analyzed according to<br />

2010 WHO criteria for semen analysis and their parameters were compared.<br />

RESULTS: Average subject age was 34.3 years (22-53). Of 36 patients<br />

who had zero sperm concentration in their first sample, 9 of them were<br />

able to show mature sperm in their second sample (25%). Hence, these patients<br />

were spared from performing surgical sperm retrieval whether for diagnostic<br />

or therapeutic purposes. Mean seminal volume among the study group<br />

was 2.9 1.6 mL in the 1st sample, with a mean concentration of 35.4*106<br />

sperm /mL and mean motility of 4.97.8%. Although second sample mean<br />

seminal volume decreased significantly compared to 1st (1.50.9 mL), yet<br />

we detected a statistically significant incline in the mean sperm concentration<br />

(3.6 6.3*106 sperm /mL) and mean motility (6.79.7%) of 2nd sample<br />

sperm (p


(lutenizing hormone (LH), follicle stimulating hormone (FSH), testosterone),<br />

and the period from SCI injury were also obtained and analyzed to<br />

detect any associations with the presence of spermatogenesis.<br />

RESULTS: In this series, SRR and PR were 89% and 62%, respectively.<br />

Univariate analysis identified age, testicular volume, serum LH level and<br />

serum FSH level as significant predictors of TESE outcome, of which only<br />

serum FSH level appeared to be independently related to TESE outcome<br />

on multivariate analysis (Table).<br />

Univariate and multivariate analysis of several parameters as predictors of<br />

micro-TESE.<br />

Univariate Analysis Multivariate Analysis<br />

Hazard Ratio P Value Hazard Ratio P Value<br />

Age 0.89 0.098 - -<br />

Testicular volume 1.56 0.015 0.065 0.10<br />

LH 0.74 0.0019 7.329 1.00<br />

FSH 0.56


2010 morphologic criteria. Beyond these thresholds, there was no additional<br />

increase in post-wash TMSC. For every % increase in morphology<br />

(below the threshold), the post-wash TMSC increased by 1.1 (p0.05<br />

BFR, total <strong>21</strong> 49.01 15.39 0.23 0.29 49 p>0.05<br />

BFR, total X, Y 49.01 15.39 -0.06 0.29 49 p>0.05<br />

BFR, high quality 13 31.77 11.72 -0.57 0.37 46 p0.05<br />

blastocysts<br />

BFR, high quality 18 31.59 12.13 -0.05 0.29 49 p>0.05<br />

blastocysts<br />

BFR, high quality <strong>21</strong> 31.59 12.13 0.24 0.29 49 p>0.05<br />

blastocysts<br />

BFR, high quality<br />

blastocysts<br />

X, Y 31.59 12.13 0.08 0.29 49 p>0.05<br />

P-397 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

LIVE BIRTH RATES AND MISCARRIAGE RATES WITH THE USE<br />

OF CALCIUM IONOPHORE IN PATIENTS WITH PRIOR FAILED<br />

OR POOR FERTILIZATION. E. Wood, a K. O. Pomeroy, a,b<br />

J. Ricard, a N. Montalvo, c I. Collazo, a J. Eisermann. a a South Florida Institute<br />

for Reproductive Medicine, Miami, FL; b The World Egg Bank, Phoenix, AZ;<br />

c Clinica Montalvo, Santa Cruz, Bolivia, Plurinational State of.<br />

OBJECTIVE: A critical step in fertilization is the activation of the oocyte<br />

via a flux of calcium released from the cortical granules of the oocyte. Treatment<br />

with calcium ionophore has been used to treat the ova of patients where<br />

fertilization via ICSI fails. The aim of this study was to determine the efficacy<br />

of calcium ionophore treatment for patients with poor or no fertilization after<br />

ICSI. A literature review of 12 peer review publications and 4 abstracts since<br />

1997 indicated that to date 64 healthy live births have occurred from the use<br />

of calcium ionophore treatment.<br />

DESIGN: This was a retrospective analysis of ICSI cycles where calcium<br />

ionophore was used from 2004 to <strong>2015</strong> at SFIRM. Ionophore treatment was<br />

used on 12 couples with 15 fresh ICSI cycles. (We also had 2 FET cycles using<br />

ionophore-treated ova.) 10 of these couples had 12 prior cycles at our<br />

clinic where 79 mature ova were retrieved and <strong>17</strong> fertilized normally<br />

(<strong>21</strong>.5%).<br />

MATERIALS AND METHODS: Ionophore treatment consisted of<br />

exposing post-ICSI ova for 10 minutes to 10 uM concentration of<br />

A23187 at 37 C. Ionophore was then rinsed out and ova were cultured using<br />

our standard methods for IVF. Embryos were cultured for 5 or 6 days prior<br />

to transfer.<br />

RESULTS: Fertilization after ionophore treatment was 56.8%, compared<br />

to <strong>21</strong>.5% for the prior cycles. (Our normal fertilization rate for our patients is<br />

about 70%.) All 15 cycles resulted in an embryo transfer with an average of<br />

1.9 embryos transferred per patient. 11 patients (73.3%) had a clinical pregnancy<br />

and 8 (53.3%) had a birth or ongoing pregnancy. (One patient that<br />

delivered a healthy baby from the fresh transfer delivered a healthy baby<br />

from a subsequent frozen embryo transfer). There were 3 miscarriages,<br />

one with a normal karyotype, one with trisomy 20 and one with trisomy<br />

<strong>21</strong>. The miscarriage rate was 27.3%. So far, 9 healthy babies have been delivered<br />

from this technique - two sets of twins and 5 singletons. The trisomy <strong>21</strong><br />

came from a patient 38 years old and the trisomy 20 from a 40 year old patient.<br />

CONCLUSIONS: We found that this method performs favorably with<br />

patients with no fertilization and even those with poor fertilization.<br />

Although the procedure resulted in a high degeneration rate of ova (about<br />

15%) and multicellular embryos with fragmentation, fertilization rates<br />

improved with ionophore treatment and live birth/ongoing rates (53.3%)<br />

are near normal for this age group of women in our practice (41.7%;<br />

SART 2012). The clinical use of calcium ionophore is still considered<br />

experimental. World literature has suggested no adverse outcome in the<br />

offspring. Its use for poor fertilization and fertilization failure in ICSI patients<br />

is promising. Calcium ionophore incubation may become a reasonable<br />

and efficient option to assist couples with this rare condition so that<br />

they may have a healthy child.<br />

e240 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


P-398 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

TREATMENT WITH COMBINED ANTIOXIDANT FORMULATION<br />

BEFORE ICSI IMPROVES PREGNANCY RATE IN COUPLES WITH<br />

OBSTRUCTIVE AZOOSPERMIA. S. Gulati, R. Chattopadhyay,<br />

B. Ghosh, S. Yasmin, S. Ghosh, G. Bose, P. Chakraborty, B. Chakravarty.<br />

Assisted Reproduction, Institute of Reproductive Medicine, Kolkata, India.<br />

OBJECTIVE: To evaluate the efficacy if any, of a combined antioxidant<br />

formulation, in couples with obstructive azoospermia after failed intra cytoplasmic<br />

injection (ICSI) cycle.<br />

DESIGN: A single centre prospective cohort study of <strong>21</strong>0 patients with<br />

azoospermia was performed from February 2012 to March <strong>2015</strong>.<br />

MATERIALS AND METHODS: Patients with documented azoospermia<br />

were divided into non-obstructive (NOA) (group A; n ¼ 98) and obstructive<br />

(OA) (group B; n ¼ 112) counterpart. Absence of sperm (n¼15) and unsatisfactory<br />

endometrium (n ¼ 3) were the exclusion criteria in either arm/s.<br />

Testicular sperm extraction (TESE) followed by ICSI confirmed pregnancy<br />

in 25.3% (n ¼ <strong>21</strong>) and 32.11% (n ¼ 32) in group A and B respectively. Patients<br />

who failed TESE/ ICSI cycles were treated with a combined antioxidant<br />

formulation comprising L-carnitine, CoQ10, zinc, folic acid, vitamin<br />

B12 and selenium for 6 months. A placebo-controlled normozoospermic<br />

group comprising 75 patients was maintained as controls. ROS-TAC score<br />

was determined as biomarker of oxidative stress (OS) from homogenized<br />

testicular tissue by standard methods before and after the treatment. Mitochondrial<br />

membrane potential (Djm), sperm DNA fragmentation were evaluated<br />

by flow-cytometry and comet assay respectively. Study was approved<br />

by Institutional Review Board. Statistical comparisons were performed by<br />

Stata10.0. The main outcome was clinical pregnancy rate.<br />

RESULTS: 62 patients with NOA and 77 patients with OA were treated<br />

with combined antioxidant therapy followed by TESE/ICSI. Improvement<br />

in sperm morphology was observed in 8.06% (n ¼ 5) in group A compared<br />

to 63.64% (n ¼ 49) cases in group B. ROS-TAC score decreased significantly<br />

in group B only than controls (33.1 6.2 vs. 52.0 7.1; p < 0.04). Djm and<br />

rate of apoptosis improved considerably after antioxidant regimen. Fertilization<br />

rate was significantly lower (p< 0.04) in NOA group (56.45%) than the<br />

OA group (74.32%). The clinical pregnancy rate was 27.42% (n ¼ <strong>17</strong>) and<br />

41.55% (n¼32) in NOA and OA cohorts respectively. This was statistically<br />

significant adjusting for age with an odds ratio of 1.06 (95% CI 1.20-3.20,<br />

p


CONCLUSIONS: TPMSC is an independent predictor for biochemical<br />

pregnancy rate and LBR in non-donor sperm, but this association is not<br />

observed in donor sperm; this lack of association in donor group may be<br />

due to small sample size. Furthermore, our data suggest that male factor<br />

infertility in cycles using non-donor sperm (despite high sperm counts)<br />

continues to compromise pregnancy outcomes as compared to donor<br />

sperm.<br />

References:<br />

1: Freour T, Jean M, Mirallie S, Langlois ML, Dubourdieu S, Barriere P. Predictive<br />

value of CASA parameters in IUI with frozen donor sperm. Int J<br />

Androl. 2009 Oct;32(5):498-504.<br />

2: Marshburn PB, McIntire D, Carr BR, Byrd W. Spermatozoal characteristics<br />

from fresh and frozen donor semen and their correlation with fertility<br />

outcome after intrauterine insemination. Fertil Steril. 1992 Jul;58(1):<strong>17</strong>9-<br />

86.<br />

3: Byrd W, Bradshaw K, Carr B, Edman C, Odom J, Ackerman G. A prospective<br />

randomized study of pregnancy rates following intrauterine and intracervical<br />

insemination using frozen donor sperm. Fertil Steril. 1990<br />

Mar;53(3):5<strong>21</strong>-7.<br />

4: Tan O, Ha T, Carr BR, Nakonezny P, Doody KM, Doody KJ. Predictive<br />

value of postwashed total progressively motile sperm count using<br />

CASA estimates in 6871 non-donor intrauterine insemination cycles. J<br />

Assist Reprod Genet. 2014 Sep;31(9):1147-53.<br />

P-401 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

TESTICULAR SPERM EXTRACTION FOLLOWED BY INTRACY-<br />

TOPLASMIC SPERM INJECTION CHOICE OF TREATMENT<br />

FOR THE APPREHENSIVE COUPLES HAVING COLLECTION<br />

PROBLEM. N. Prasad, a R. Chattopadhaya, b S. Ghosh, b<br />

S. K. Goswami, b B. Ghosh, b P. Chakraborty, b B. Chakravarty. b a Reproductive<br />

Medicine, Fellow in Reproductive Medicine, Kolkata, India; b Reproductive<br />

Medicine, Institute of Reproductive Medicine, Kolkata, India.<br />

OBJECTIVE: To compare the fertility potential of freshly retrieved testicular<br />

sperm obtained from patients with azoospermia and collection failure<br />

undergoing intracytoplasmic sperm injection (ICSI).<br />

DESIGN: Prospective analysis of 130 testicular sperm extraction (TESE)<br />

/ICSI cycles from January 2011 to March <strong>2015</strong>.<br />

MATERIALS AND METHODS: Eighty four patients with documented<br />

azoospermia underwent micro-dissection TESE with concomitant ICSI at<br />

the same time (Group A). Patients having previous history of failed collection<br />

(n¼53) were also treated with TESE followed by ICSI (Group B). 7 cycles<br />

were cancelled due to unsuccessful sperm retrieval in azoospermic couples<br />

only. Comparison between the two groups in terms of fertilization, cleavage,<br />

implantation, clinical pregnancy and miscarriage rates was done. The Ethical<br />

Committee of Institute of Reproductive Medicine approved the study and all<br />

the participants consented to enroll in this study . Statistical comparisons<br />

were performed using chi-square and student’s T test as required. P value<br />


at a private suburban Fertility Center by a single andrologist (GAA) were reviewed.<br />

Only the first SA was evaluated per patient.<br />

RESULTS: In total, 901 SA including 276 who worked in Business/<br />

Finance/Sales (30.6%), 224 worked in Information Technology (24.9%),<br />

77 were engineers (8.5%), 51 worked in government (5.7%), and 37 in Health<br />

Care (4.1%). There were no differences in any semen parameter among<br />

different occupations; specifically IT use was not associated with any negative<br />

effects on seminal parameters.<br />

CONCLUSIONS: Men working in IT have no differences in seminal parameters<br />

compared to other white collar occupations. There were no differences<br />

among the different occupations evaluated, a reassuring novel<br />

finding. As far as we know, this is the largest study evaluating the effects<br />

of various occupations, including those that require significant sitting such<br />

as IT, on semen parameters.<br />

P-404 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

IMPROVEMENT IN PREGNANCY RATES IN IMMUNOLOGI-<br />

CALLY INFERTILE MEN UNDERGOING PREDNISOLONE AND<br />

IVF PROGRAMS PRECEDED BY SPERM PENETRATION<br />

ASSAY. A. M. Taiyeb, a M. T. Ridha-Albarzanchi, a A. Haji, a<br />

S. N. Mahmood, b M. E. Kjelland, c N. S. Hawa, b S. A. Muhsen-Alansarri. a<br />

a BARZ IVF Center for Infertility Treatment and Embro Research, Erbil, Iraq;<br />

b Department of Obstetrics and Gynecology, Baghdad College of Medicine,<br />

Baghdad, Iraq; c US Army Engineer Research and Development Center,<br />

Vicksburg, MS.<br />

OBJECTIVE: Anti-sperm antibodies (ASA) impair not only human sperm<br />

motility but also capacitation, penetration of cervical mucus, and conception<br />

rate. Many studies have confirmed the clinical therapeutic effect of corticosteroids<br />

(CST) in treatment of men with ASA [1-3]; however, other studies<br />

have not observed a therapeutic effect [4-5]. This controversy is extended<br />

to include the usefulness of assisted reproductive technologies in patients<br />

with ASA. For instance, some studies have shown that pregnancy rates<br />

following IVF or ICSI were similar or not associated in men with or without<br />

ASA [6-7], whereas other reports showed the superiority of ICSI over IVF [8]<br />

and IUI over natural intercourse in men with ASA [9]. In order to address this<br />

controversy, immunological infertile men, treated or not treated with CST,<br />

were admitted to conventional IVF or ICSI as determined by the sperm penetration<br />

assay (SPA) of hamster ovum. The latter assay can identify patients<br />

with compromised fertilization related to sperm binding and fusion to oolema.<br />

It is possible that some patients with ASA have compromised sperm<br />

fertilization as an additional problem, which may interfere with the therapeutic<br />

effect of CST in patients with ASA.<br />

DESIGN: A prospective clinical comparative study.<br />

MATERIALS AND METHODS: SPA was performed for immunologically<br />

infertile men. Positive or negative SPA partners were then admitted<br />

to conventional IVF or ICSI cycles, respectively. Each IVF or ICSI male<br />

group was divided into treated and control groups. The treated male group<br />

received a prednisolone program for three cycles.<br />

RESULTS: Administration of prednisolone in men with good SPA results<br />

improved sperm motility when compared to control men (P


Table 1.<br />

Group 1 Group 2 P<br />

Age 33.7 2.3 29.4 4.8 0.004<br />

FSH 6.1 1.3 6.4 2.2 NS<br />

LH 6.0 3.9 6.4 2.2 NS<br />

E2 33.1 16.2 44.6 25.4 NS<br />

PRL 19.5 10.4 28.2 11.2 NS<br />

FOL 14.3 8.2 10.4 4.6 NS<br />

INY 9.6 7.5 9.2 5.3 NS<br />

FERT 5.0 3.9 4.6 3.0 NS<br />

P-406 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

EJACULATED SPERM MAY NOT RESULT IN THE BEST CLIN-<br />

ICAL OUTCOME FOR ICSI TREATMENT CYCLES. Y. Lin, a<br />

L. Cai, b J. Dong, b J. Liu, c R. Chian. d a reproduction medicine, nanjing,<br />

China; b The First Affiliated Hospital of Nanjing Medical U, Nanjing, China;<br />

c the First Affiliated Hospital of Nanjing Medical U, Nanjing, China; d McGill<br />

University, Montreal, QC, Canada.<br />

OBJECTIVE: To evaluate the impact of different sperm sources on fertilization,<br />

embryo quality, clinical pregnancy and implantation rates in intracytoplasmic<br />

sperm injection (ICSI) treatment cycles.<br />

DESIGN: A retrospective analysis was performed in 1,263 patients undergoing<br />

ICSI treatment cycles in our center from January 2012 to March 2014.<br />

MATERIALS AND METHODS: The patients were classified into 4<br />

groups based on the sources of sperm obtained, namely fresh ejaculated<br />

semen, (n¼950), fresh testicular sperm aspiration (TESA) (n¼12), fresh<br />

percutaneous epididymal sperm aspiration (PESA) (n¼204), and frozen<br />

PESA sperm (n¼97). Fertilization, high-quality embryo, implantation, clinical<br />

pregnancy and live birth rates in the fresh embryo transferred cycle were<br />

compared among the groups.<br />

RESULTS: There were no differences in the fertilization rates among four<br />

groups (74.4%, 75.6%, 75.1% and 74.7%, respectively). The patients with<br />

fresh TESA sperm and frozen PESA sperm showed significantly higher<br />

high-quality embryo rates (75.4 % and 73.9 %) than those of fresh ejaculated<br />

semen (63.8%) and fresh PESA sperm (66.3%) groups (P


2. Ben-Ami I, Raziel A, Strassburger D, Komarovsky D, Ron-El R, Friedler<br />

S. Intracytoplasmic sperm injection outcome of ejaculated versus<br />

extracted testicular spermatozoa in cryptozoospermic men. Fertil.<br />

Steril. 2013;99(7):1867-71.<br />

3. Bendikson KA, Neri Q V, Takeuchi T, et al. The outcome of intracytoplasmic<br />

sperm injection using occasional spermatozoa in the ejaculate<br />

of men with spermatogenic failure. J. Urol. 2008;180(3):1060-4.<br />

4. Hauser R, Bibi G, Yogev L, et al. Virtual Azoospermia and Cryptozoospermia–Fresh/Frozen<br />

Testicular or Ejaculate Sperm for Better IVF<br />

Outcome? J. Androl. 2011;32(5):484-490.<br />

5. Weissman A, Horowitz E, Ravhon A, Nahum H, Golan A, Levran D.<br />

Pregnancies and live births following ICSI with testicular spermatozoa<br />

after repeated implantation failure using ejaculated spermatozoa. Reprod.<br />

Biomed. Online 2008;<strong>17</strong>(5):605-609.<br />

P-409 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

S. Alshahrani. Prince Sattam Uni-<br />

ARE OBESE MEN SUBFERTILE?<br />

versity, Alkharj, Saudi Arabia.<br />

OBJECTIVE: Although it is well known that overweight and obesity can<br />

affect female fertility, in men the negative effects on reproductive system<br />

attributed to obesity are less evident and have been less often studied. The<br />

aim of this study is to evaluate the effect of Body Mass Index (BMI) on<br />

different semen parameters.<br />

DESIGN: prospective study.<br />

MATERIALS AND METHODS: After IRB approval, semen samples<br />

from 439 male partners of couples presenting for evaluation of their infertility,<br />

in the period from 2013-2014, were collected. Men were divided into<br />

three BMI groups; normal , overweight and obese. Meticulous scrotal examination<br />

plus scrotal ultrasound were performed. Hormonal profile, including<br />

testosterone, FSH, LH and prolactin was done to exclude any possible causative<br />

hormonal factor.<br />

RESULTS: Mean BMI was 29.675.89. ANOVA testing revealed no significant<br />

differences in semen parameters between the 3 different BMI groups.<br />

Also, pair-wise multiple comparisons were found to be non-significant (P ><br />

0.05). The distribution of patients with normal sperm concentration per BMI<br />

group was as follows: normal weight, 46/75 (61.33%); overweight, 114/<strong>17</strong>9<br />

(63.69%); and obese, 116/185 (62.70%). BMI had a negative correlation with<br />

semen volume, sperm concentration, sperm motility and sperm morphology.<br />

However, this correlation reached a significant level only between BMI and<br />

sperm concentration (r¼0.101; p¼0.035).<br />

CONCLUSIONS: Although found to have no significant relation with<br />

semen volume, sperm motility or sperm morphology; BMI was proved, in<br />

our study, to have a significant effect on sperm concentration.<br />

The effect of BMI on semen parameters.<br />

Parameters Total patients Normal BMI Over weight Obese p-value<br />

Number (%) 439 (100) 75 (<strong>17</strong>.08) <strong>17</strong>9 (40.77) 185 (42.14)<br />

BMI 29.675.89 23.051.34 27.061.41 34.885.31 ˂0.001<br />

Age (y) 36.856.73 35.646.56 36.836.82 37.376.69 0.<strong>17</strong>3<br />

Volume (ml) 3.071.49 3.1<strong>21</strong>.42 3.181.47 2.941.53 0.297<br />

Concentration (mil/ml) 41.4650.96 44.6457.25 41.3248.96 40.3050.40 0.824<br />

Actively motile sperm (%) 44.5319.54 46.0018.50 44.1120.46 44.3519.10 0.770<br />

Normal morphology (%) 3.<strong>21</strong>2.95 3.202.92 3.243.02 3.152.91 0.957<br />

P-410 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

INDEPENDENT PREDICTORS OF INTRAUTERINE INSEMINA-<br />

TION (IUI) SUCCESS. R. D. Kastury a G. S. Taliadouros. b a Obstetrics<br />

and Gynecology, Inspira Health Network, Vineland, NJ; b Reproductive<br />

Endocrinology and Infertility, Inspira Health Network, Vineland, NJ.<br />

OBJECTIVE: Various causes of infertility are treated with intrauterine<br />

insemination (IUI). Many factors affect IUI outcomes. Previously published<br />

studies have provided conflicting results regarding the independent predictors<br />

for the success of this treatment (1). In this study, we aim to evaluate<br />

these predictors in a well-defined patient population.<br />

DESIGN: A retrospective study.<br />

MATERIALS AND METHODS: Our analysis evaluated 509 couples, between<br />

ages 24 and 44, that underwent infertility assessment and subsequent<br />

intrauterine insemination. Fresh partner’s semen was processed through a<br />

50%/90% discontinuous gradient and one IUI was performed at the time<br />

of ovulation, confirmed by ultrasonography and laboratory values. Females<br />

with tubal factor infertility and patients with severe male factor infertility<br />

were excluded. All other major medical concerns were addressed prior to<br />

initiating treatment. Anti-estrogen and gonadotropins were used to address<br />

ovulatory aberrations. The primary outcome measure was the occurrence<br />

of pregnancy. Univariate (EXPAND) and multivariate logistic regression analyses<br />

were used. The strength of association between pregnancy and each<br />

independent predictor is presented as an odds ratio (OR) with confidence interval<br />

(CI). All statistical tests used a two-tailed alpha of 0.05.<br />

RESULTS: Univariate analysis shows that motile sperm count in the unprocessed<br />

semen (OR ¼ 0.0023, CI ¼ 0.0010 - 0.0037, p ¼ 0.0005) and total<br />

motile sperm inseminated (OR ¼ 0.0064, CI ¼ 0.0029 - 0.0098, p ¼ 0.0003)<br />

were significant factors associated with the occurrence of pregnancy. Sperm<br />

morphology performed using Kruger’s strict criteria, did not have statistical<br />

significance on the occurrence of pregnancy. The age of female was<br />

extremely predictive of pregnancy occurrence (OR ¼ -0.047, CI ¼ -0.0792<br />

to -0.0149, p ¼ 0.0041) and the type of ovarian stimulation, gonadotropins,<br />

also proved to be statistically significant in predicting a positive outcome<br />

(OR ¼ -0.3257, CI ¼ -0.6256 to -0.0248, p ¼ 0.0339). Multivariate analysis<br />

demonstrated that female age and number of motile sperm in the ejaculate are<br />

the strongest predictors of pregnancy, p ¼ 0.0009 & 0.0007 respectively.<br />

Gonadotropin stimulation increased the chance of achieving pregnancy by<br />

30%.<br />

CONCLUSIONS: Our results indicate that the total number of motile<br />

sperm in the ejaculate is a significant predictor for pregnancy occurrence.<br />

As expected, the increased number of oocytes available during controlled<br />

ovarian stimulation will have an additive effect on the results. It has been<br />

widely accepted that female age has an impact on fecundity. These findings<br />

corroborate previously reported data, which can be used to advise future couples<br />

on the likelihood of achieving pregnancy and guide physicians with<br />

appropriate treatment selection.<br />

Reference:<br />

1. Ombelet W, Dhont N, Thijssen A, Bosmans E, Kruger T. Semen quality<br />

and prediction of IUI success in male subfertility: a systematic review.<br />

Reprod Biomed Online. 2014;28:300-309.<br />

P-411 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

IS IT NECESSARY TO DO INTRACYTOPLASMIC SPERM INJEC-<br />

TION ON PATIENTS WITH ABNORMAL SPERM DNA FRAGMEN-<br />

TATION VALUE IN ART? S. M. Kang, J. H. Kim, Y. J. Lee. I-Dream<br />

Center, Mizmedi Women’s Hospital, Seoul, Korea, Republic of.<br />

OBJECTIVE: The purpose of this study is to evaluate how effective intracytoplasmic<br />

sperm injection (ICSI) could be in the pregnancy outcome of patients<br />

with DNA-fragmented sperm by comparison to conventional in vitro<br />

fertilization (IVF).<br />

DESIGN: A retrospective clinical study.<br />

MATERIALS AND METHODS: According to standard protocol, 137<br />

(ICSI and IVF) cycles that were performed with at least 5 retrieved eggs,<br />

normal count and motility of sperms (the World Health Organization criteria)<br />

at each cycle were retrospectively analyzed. Sperms that were used in this<br />

study were classified into three groups depending on the sperm DNA fragmentation<br />

(SDF) value (by Halosperm kit) as follows: Abnormal,<br />

SDF>30%; Sub-normal, 15%


OH; b Case Western Reserve University, Beachwood, OH; c USC Fertility<br />

Center, Los Angeles, CA.<br />

OBJECTIVE: To study the correlation between pre- and post-wash sperm<br />

morphology indices and pregnancy outcomes in intrauterine insemination<br />

(IUI) and in vitro fertilization (IVF).<br />

DESIGN: Retrospective Cohort.<br />

MATERIALS AND METHODS: A total of 305 IUI and 113 IVF fresh and<br />

frozen cycles performed at a single center from January to <strong>October</strong> 2012 were<br />

reviewed. Three morphology indices were completed on pre- and post-wash<br />

semen samples. The three indices included multiple anomalies index (MAI),<br />

teratozoospermia index (TZI), and sperm deformity index (SDI). Index data<br />

were correlated with pregnancy outcomes for IUI and IVF cycles via a principle<br />

component analysis and logistic regression. Student T-tests were used<br />

for comparisons of MAI, TZI and SDI with pregnancy outcomes.<br />

RESULTS: The analysis showed a high correlation between the MAI, TZI,<br />

and SDI in the pre- and post-wash groups. Due to this similarity, the optimal<br />

morphology index could not be determined. A principle component analysis<br />

was utilized to determine the relationship between the group of indices and<br />

pregnancy outcomes. The pre-wash indices were not significant in predicting<br />

pregnancy (p ¼ 0.335), while the post-wash indices were significant<br />

(p ¼ 0.0<strong>21</strong>1). The post-wash indices were more predictive of pregnancy outcomes<br />

in IUI cycles compared to IVF cycles (p ¼ < 0.0001). When MAI,<br />

TZI, and SDI were evaluated independently in IUI cycles, there was a trend<br />

toward statistical significance in pregnancy outcomes in the post-wash group<br />

(p ¼ 0.056, 0.055, 0.051, respectively). However, in IVF fresh and frozen cycles<br />

the three indices had no statistical significance in predicting pregnancy<br />

outcomes (p ¼ 0.698, 0.704, 0.747, respectively).<br />

CONCLUSIONS: Due to the high correlation of MAI, TZI, and SDI, no<br />

optimal morphology index was identified in this study. Although the grouped<br />

post-wash indices predicted pregnancy outcomes, the individual analyses<br />

were not significant. As expected, the combined post-wash indices may<br />

have higher utility in predicating pregnancy outcomes in IUI compared to<br />

IVF cycles. Consistent with prior studies, sperm morphology indices may<br />

have predictive potential in select groups; however, the data from this study<br />

does not support generalized use.<br />

SPERM PREPARATION<br />

P-413 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

THE IMPACT OF SEMEN PROCESSING ON SPERM QUALITY<br />

AND PREGNANCY RATES WITH INTRAUTERINE<br />

INSEMINATIONS. J. J. Ruiter-Ligeti, a C. Agbo, b M. Dahan. a a McGill<br />

University, Montreal, QC, Canada; b Stanford University, Stanford, CA.<br />

OBJECTIVE: Semen processing is routinely performed prior to intrauterine<br />

insemination (IUI). The expectation is that semen processing improves<br />

semen quality parameters. However, no studies have evaluated the impact<br />

of semen processing for IUI on those parameters. This study aims to evaluate<br />

Table: Effect of processing on semen quality parameters and pregnancy rates.<br />

Semen Quality<br />

Parameter<br />

Absolute<br />

change<br />

Mean<br />

+/- SD<br />

Pearson<br />

correlation<br />

coefficient<br />

(r)<br />

% of subjects<br />

p-value of with parameter<br />

correlation improvement<br />

Concentration (M/ml) 66 74 0.043 0.04* 90%<br />

Percent Motile (%) 24 22 -0.012 0.56 80%<br />

Concentration Motile 27 58 0.038 0.08 70%<br />

(M/ml)<br />

Total Motile Sperm -55 81 -0.002 0.97


P-415 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

STATIC MAGNETIC FIELDS ON HUMAN SPERM. A SIDE EFFECT<br />

OF MAGNETIC-ACTIVATED CELL SORTING? C. Avendano,<br />

A. Mata, I. A. Anduaga Marchetti, C. Sanchez Sarmiento. Nascentis Medicina<br />

Reproductiva, Cordoba, Argentina.<br />

OBJECTIVE: Non-damaged sperm DNA is one of the prerequisites for<br />

achieving successful fertilization and embryo development in assisted<br />

reproductive technologies. Magnetic-activated cell sorting (MACS) using<br />

paramagnetic annexin V-conjugated microbeads has been proposed as a<br />

safe method to select non-apoptotic and viable sperm. The procedure involves<br />

the magnetic labeling of sperm with damaged DNA and the passage<br />

through a high power static magnetic field (SMF). Intact living<br />

spermatozoa without DNA fragmentation pass through the column and<br />

are collected for later use, while the fragmented cells are selectively retained.<br />

Although there are some concerns about the possible selection<br />

and use of sperm with paramegnetic beads, the potential impact of<br />

SMF has never been mentioned. Therefore, the objective of this work<br />

was to evaluate the possible effect of SMF on human sperm during in<br />

vitro manipulation.<br />

DESIGN: Prospective study.<br />

MATERIALS AND METHODS: Motile sperm from healthy men (n¼5)<br />

were selected by swim up. Each sperm suspension was separated in 4<br />

fractions (F1, F2, F3 and C). F1, F2, F3 were exposed to a SMF (1.2<br />

Tesla) during 1, 2 and 10 minutes, respectively. Fraction C was kept<br />

as control without SMF exposure. Sperm parameters were evaluated<br />

immediately after SMF exposure on each fraction and then incubated<br />

in capacitating conditions for 3 hours at 37 C. Progressive sperm<br />

motility (PG, %), curvilinear velocity (VCL, mm/s) and rectilinear velocity<br />

(VSL, mm/s) were evaluated using videomicroscopy with computerized<br />

image analysis. Spontaneous (sRA) and induced (iRA) acrosomal<br />

reaction were evaluated using Coomassie brilliant blue staining. Sperm<br />

DNA fragmentation was assessed by TUNEL assay.<br />

RESULTS: PG was increased immediately after exposure to SMF in F1<br />

compared with C (93.5 2.9 vs. 86.9 3.9, p


changes involved with impaired spermatogenesis were observed in<br />

seminiferous tubule including the smaller size of tube, the scarce of<br />

spermatocyte. [3] CASA detection of sperm showed that sperm count<br />

and motility declined with age in Bnc1 -/- male mice. However Bnc1<br />

+/ mouse showed a significant drop in sperm motility by 20 weeks<br />

of age, suggesting a dosage effect of Bnc1 on testis development. [4]<br />

Bnc1 knock down had no effects on proliferation of CRL-<strong>21</strong>96 and<br />

C18-4 while promote the proliferation of CRL-2053 cell. Furthermore,<br />

several spermatogonia stem cell markers such as Magea4,Ret,Plzf were<br />

down-regulated while differentiation marker Kit were up-regulated in<br />

C18-4 cell and CRL-2053 cell after Bnc1 knock down. Those data<br />

suggested Bnc1 may play a role in maintaining undifferentiation of<br />

spermatogonia cell.<br />

CONCLUSIONS: The genetic mutation of human BNC1 results in male<br />

infertility with the reduction of sperm motility and decreased sperm count<br />

and that BNC1 is a functional molecule essential for mammalian sperm formation.<br />

OVARIAN RESERVE<br />

P-418 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

COMPARISON OF ANTRAL FOLLICLE COUNT ACROSS THE<br />

MENSTRUAL CYCLE WITH ANTIMULLERIAN HORMONE AND<br />

OVARIAN RESPONSE TO CONTROLLED HYPERSTIMULATION<br />

RELATED TO OVARIAN RESERVE. N. Massin, G. Amand,<br />

C. Villette, A. Dessapt, C. Pietin Vialle, H. Bry, M. Pasquier,<br />

B. Haddad. Department of Gynecology, Obstetrics and Reproductive Medicine,<br />

Intercommunal Hospital, University Paris 12, Creteil, France.<br />

OBJECTIVE: To compare antral follicle count (AFC) at early follicular<br />

(EF), late follicular (LF) and luteal (L) phases with AFC at day 1 of the<br />

ovarian stimulation (S1), antimullerian hormone (AMH) and oocytes yield,<br />

in women with low, normal or high ovarian reserve.<br />

DESIGN: Monocentric retrospective analysis of consecutive patients undergoing<br />

their first IVF cycle.<br />

MATERIALS AND METHODS: A total of 131 patients aged from 20<br />

to 41 were included in the analysis between 2014, jan and <strong>2015</strong>, feb.<br />

Ultrasound for basal AFC evaluation was performed in our center on<br />

the day of patient’s visit, independent of her menstrual cycle phase.<br />

The women were classified as EF-AFC from day 1 to 6 (N¼44), LF-<br />

AFC from day 7 to 13 (N¼49) and L-AFC thereafter (N¼38). S1-<br />

AFC was performed either with patient on agonist treatment or after<br />

<strong>17</strong>b-estradiol treatment for antagonist protocol scheduling. AMH dosage<br />

was performed from day 2 to 4. Comparison of AFCs, AMH and number<br />

of oocytes retrieved was done with the Pearson correlation coefficient<br />

in the whole population and according to ovarian reserve : low<br />

for AMH


Statistical analysis included student’s t-test and chi-square test. P


Medicine, Feinberg SOM- Northwestern University, Chicago, IL; c NIEHS,<br />

Research Triangle Park, NC.<br />

OBJECTIVE: To determine what medical, lifestyle and reproductive factors<br />

are associated with low AMH levels in young AAW.<br />

DESIGN: Cross-sectional study.<br />

MATERIALS AND METHODS: 1,654 AAW aged 23-35 years who<br />

participated in the Study of Environment, Lifestyle & Fibroids (SELF)<br />

were included. Medical history and lifestyle habit data were collected via<br />

self-report. AMH levels were determined using an ultrasensitive ELISA<br />

assay. AMH (ng/mL) was classified as ‘‘low’’ (2-7) or ‘‘high’’ (>7). Polytomous logistic regression was used<br />

to identify factors that were associated with low AMH levels. The model<br />

adjusted for age, body mass index (BMI) and use of hormonal contraception.<br />

RESULTS: The mean age of the subjects was 28.73.5 years and the mean<br />

AMH was 4.003.49 ng/mL. AMH levels were low, low normal, normal, and<br />

high in 12.2%, 19.5%, 54.5% and 13.8% of women respectively. Hypertension,<br />

present in 10.0% of women, was associated with a higher likelihood of<br />

having a low AMH versus a normal AMH (OR 2.38, 95% CI 1.49-3.79).<br />

AAW with diabetes were not significantly more likely to have reduced<br />

AMH levels. Weight was also associated with AMH levels. As BMI<br />

increased, AMH levels were significantly more likely to be low versus<br />

normal (OR 1.03, 95% CI 1.01-1.04) or high (OR 1.04, 95% CI 1.02-<br />

1.07). There was no significant association between AMH levels and various<br />

smoking related factors. AAW who smoked, were cared for as a child by a<br />

smoker, or had a mother who smoked while pregnant were not more likely<br />

to have low AMH levels. Factors related to reproductive history were also<br />

investigated to assess their relationship with AMH levels. Women with a history<br />

of a tubal ligation were significantly more likely to have a low AMH than<br />

a high AMH (OR 2.28, 95% CI 1.02-5.08). However, there was no significant<br />

association between low AMH levels and a history of becoming pregnant,<br />

having a live birth, or consulting with a physician for difficulty conceiving.<br />

Previous infections due to gonorrhea, chlamydia or pelvic inflammatory disease<br />

were also not significantly associated with low AMH levels.<br />

CONCLUSIONS: The findings from this large study of young AAW suggest<br />

that hypertension and obesity are two common conditions associated<br />

with low AMH levels. Additional research is needed to determine whether<br />

the impact of these factors on AMH is reversible. A history of tubal ligation<br />

was also associated with lower AMH levels. Patients who present for IVF or<br />

tubal reversal should be counseled that they may have a compromised<br />

ovarian reserve due to their history of a tubal ligation.<br />

Supported by: NIH R<strong>21</strong> HD077479, NIH Women’s Reproductive Health<br />

Research Scholar Award, Harold Amos Medical Faculty Development<br />

Award, Robert Wood Johnson Foundation, Friends of Prentice, The Evergreen<br />

Invitational, The Woman’s Board of Northwestern Memorial Hospital.<br />

P-424 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

UTILITY OF ANTI-M€ULLERIAN HORMONE TO IDENTIFY<br />

WOMEN WITH UNDIAGNOSED DIMINISHED OVARIAN<br />

RESERVE. L. C. Grossman, a L. Zakarin Safier, a M. Kline, a C. Chan, b<br />

R. Lobo, a M. V. Sauer, a N. C. Douglas. a a Columbia University Medical Center,<br />

New York, NY; b Level, New York, NY.<br />

OBJECTIVE: To determine the utility of anti-M€ullerian hormone (AMH)<br />

to identify diminished ovarian reserve (DOR) in asymptomatic women<br />

without personal or family risk factors.<br />

DESIGN: Prospective cohort.<br />

MATERIALS AND METHODS: Women ages 27-37 years (yrs) currently<br />

deferring pregnancy were invited to take an AMH blood test and pre- and posttest<br />

surveys. Women were recruited via alumni listservs, social media, and<br />

flyers in gynecology offices. AMH level was determined for 97 women, median<br />

age 31 (interquartile range (IQR) 29-34) yrs, using the Beckman Coulter<br />

Gen II AMH assay with interassay variability of 0.3 ng/mL and lowest level of<br />

detection of 0.<strong>17</strong> ng/mL. For healthy 31 yr old women, the median AMH level<br />

is 3.23 ng/mL, 10th percentile is 1.15 ng/mL, and 90th percentile is 6.59 ng/mL<br />

(1). Thus we defined DOR as AMH


Omega-3 Fatty Acid Supplementation Lowers Serum FSH in Women.<br />

Baseline<br />

Following 1<br />

month of omega-3<br />

supplementation<br />

p<br />

Percent<br />

change<br />

LH mean serum level, IU/L 4.4(0.6) 4.1(0.4) 0.77 -4.9<br />

FSH mean serum level, IU/L 4.8(0.3) 4.0(0.3) 0.06 -16.9<br />

Peak serum FSH, IU/L 5.5(0.4) 4.6(0.4) 0.06 -16.2<br />

LH response to GnRH,<br />

790 (76) 795 (74) 0.96 0.58<br />

area under the curve, IU/L<br />

FSH response to GnRH, area 626 (42) 518 (38) 0.04 -18.8<br />

under the curve, IU/L<br />

Plasma<br />

omega 6/omega 3<br />

FA ratio<br />

7.7 (0.8) 3.2 (0.4) 67.91) 53 32.8 7.4 <strong>17</strong>.0 11.1 5.7 41.2<br />

Optimal(40.03-67.9) 141 32.9 6.0 14.0 13.3 6.4 31.2<br />

Satisfactory(<strong>21</strong>.97-40.03) 362 34.2 6.7 12.0 12.0 4.4 39.5<br />

Low(3.07-<strong>21</strong>.97) 1314 35.9 9.0 8.0 8.8 6.8 36.8<br />

Very Low(0-3.07) 319 37.6 11.4 4.6 3.7 14.7 16.9<br />

CPR/<br />

Cycle<br />

Started (%)<br />

Reference:<br />

1. La Marca A1, Sighinolfi G, Radi D et al. Anti-Mullerian hormone<br />

(AMH) as a predictive marker in assisted reproductive technology (AR-<br />

T).Hum Reprod Update. 2010 Mar-Apr;16(2):113-30.<br />

P-427 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

ANTRAL FOLLICLE COUNT MEASURED AFTER PITUITARY<br />

SUPPRESSION AS PREDICTOR OF OVARIAN RESPONSE AND<br />

PREGNANCY IN CLINICAL PRACTICE: A REVIEW OF 2075 AS-<br />

SISTED REPRODUCTION CYCLES. S. Peralta, a,b F. Fabregues, a,b<br />

J. Penarrubia, c,b G. Casals, a,b M. Creus, a,b D. Manau, a,d I. Gonzalez-Foruria,<br />

e,b A. Borras, a,b J. Balasch. a,b a Institut Clınic d’Obstetrıcia, Ginecologia<br />

i Neonatologia, Hospital Clinic, Barcelona, Spain; b Institut d’Investigacions<br />

Biomediques August Pi i Sunyer, IDIBAPS, Barcelona, Spain; c Institut<br />

Clınic d’Obstetrıcia, Hospital Clinic, Barcelona, Spain; d IDIBAPS, Barcelona,<br />

Spain; e Gynecologist, Barcelona, Spain.<br />

OBJECTIVE: To investigate the usefulness of Antral Follicle Count<br />

measured after pituitary suppression (AFCaps) in predicting ovarian<br />

response and pregnancy in clinical practice.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: All the IVF-ICSI cycles following a long<br />

GnRH agonist protocol in which the AFCaps had been registered in our database<br />

were included. A total of 2075 cycles between January 2011 and<br />

December 2014 were analyzed. Cycles were divided into three subgroups according<br />

to the following ovarian response categories: high (>15 oocytes),<br />

normal (4-15 oocytes) and low (% 3 oocytes or cycle cancellation).<br />

RESULTS: According to the predefined ovarian response categories, <strong>17</strong>0<br />

cycles (8.9%) were classified as high, 1468 (70.7%) as normal and 437<br />

(<strong>21</strong>.1%) as low. Table shows patient characteristic, AFCaps and basal FSH,<br />

ovarian stimulation, ovum retrieval and cycle outcomes in high, normal<br />

and poor respondes. AFC was significantly lower in poor responders and<br />

significantly higher in high responders than normal responders. AFC was<br />

significantly correlated with parameters of ovarian response (E2 on HCGr<br />

day, number of oocytes retrieved, number of embryos obtained/) For the<br />

prediction of high response, AFCaps had an area under the receiver-operating<br />

characteristic curve (AUC) of 0.80. Both female age and basal FSH had lower<br />

accuracy (AUC 0.66 and 0.63, respectively). For low response prediction,<br />

Table: Baseline patient characteristics, stimulation and clinical outcomes by<br />

subgroup.<br />

VARIABLE<br />

Group 1<br />

High<br />

responders<br />

(n¼<strong>17</strong>0)<br />

Group 2<br />

Normal<br />

responders<br />

(n¼1468)<br />

Group 3<br />

Poor<br />

responders<br />

(n¼437)<br />

P value<br />

Age (y) 34.393.54 35.963.40 37.723.03


again AFC had better accuracy (AUC 0.84) than age and basal FSH (AUC<br />

0.67 and 0.62, respectively). When the likehood of pregnancy was analyzed<br />

the AUC for AFCaps was 0.63, similar than AUC for age and basal FSH<br />

(AUC 0.59 and 0.54, respectively).<br />

CONCLUSIONS: AFCaps in a clinical setting is a good predictor of<br />

ovarian response, better than age and basal FSH. However is not a good predictor<br />

of pregnancy.<br />

P-428 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

UTILITY OFAUTOLOGOUS FRESH IVF CYCLES IN EXTREMELY<br />

POOR PROGNOSIS PATIENTS. M. G. Vega, a D. H. Barad, b<br />

S. K. Darmon, c N. Gleicher, b V. A. Kushnir. d a Mount Sinai St. Lukes Roosevelt<br />

Hospital & Center for Human Reproduction, New York, NY; b Center<br />

for Human Reproduction & Foundation for Reproductive Medicine, New<br />

York, NY; c Center for Human Reproduction, New York, NY; d Center for Human<br />

Reproduction & Wake Forest University, New York, NY.<br />

OBJECTIVE: To determine whether there still is utility of performing<br />

autologous fresh in vitro fertilization (IVF) cycles in extremely poor prognosis<br />

patients.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: We identified among 1,418 consecutive<br />

fresh non-donor IVF cycles in our center’s anonymized electronic research<br />

databank 768, which by Bologna Criteria qualified as poor responders<br />

because they produced %3 oocytes, were > 40 years and/or had AMH<br />


evaluated the oxidative DNA injury. Histopathologic examination for follicle<br />

counting was observed following hematoxylin-eosin staining.<br />

RESULTS: AMH levels were significantly decreased compared to the preoperative<br />

and postoperative periods in I/R, I/R+NAC and I/R+Enoxaparine<br />

groups (p¼0.004, p


controlled primordial follicle activation in vivo. Also, PPAR gamma agonist<br />

induced PTEN expression and inhibited Akt phosphorylation in vitro. Ovary<br />

specific PPAR gamma was predominantly expressed in ovarian tissue. The<br />

aim of this study was to investigate whether PPAR gamma modulator might<br />

induce the activation and growth of primordial follicles from 5day female<br />

mouse after birth.<br />

DESIGN: Animal study.<br />

MATERIALS AND METHODS: Ovaries were collected from 5days old<br />

B6D2F1 or ICR female mouse and cultured on six well plate and insert for<br />

10 days in vitro. Ovaries were cultured in DMEM/F12 containing BSA,<br />

ITS-X, and ascorbic acid, with or without PPAR gamma modulator for 2<br />

days and transferred PPAR gamma modulator free medium. To evaluate<br />

the expression of PPAR gamma, PTEN, and Akt1 on mouse ovary, we performed<br />

real-time-PCR, immunohistochemistry, and western blot. Also, end<br />

of culture, ovaries were fixed 10% neutral-buffered formalin and subjected<br />

of HE staining, immunohistochemistry against Ki-67, AMH, PTEN, Akt1<br />

and FOXO-3a, and apoptosis assay with TUNEL. Also, western blot was performed<br />

to evaluate the changes of PTEN, Akt1.<br />

RESULTS: The expression of PPAR gamma, PTEN, Akt1 and FOXO3a<br />

were evaluated in ovaries from 5, 10, 15, 20 days and 8 weeks old mouse<br />

by RT-PCR. PPAR gamma was expressed in ovaries from all the ages, it<br />

increased in ovary in 20 days explosively. PTEN and Akt1 were increased<br />

in ovary form 10 Days. In PPAR gamma antagonist treated ovaries, PTEN<br />

was decreased and Akt1 was activated. Also nuclear exclusion of FOXO3a<br />

was observed in oocytes of primordial follicles at 6hr treatment with PPAR<br />

gamma antagonist. AMH expression was observed in PPAR gamma antagonist<br />

treated ovaries, but was not in PPAR gamma agonist. After 12 days culture,<br />

increases of ovarian sizes of PPAR gamma antagonist treated groups<br />

were observed as compare to non-treated or PPAR gamma agonist treatment.<br />

CONCLUSIONS: From these result, PPAR gamma may participate primordial<br />

follicle activation and further development, possibly mediated in part of<br />

PI3K-PTEN signaling pathway in vitro. Generation of activated primordial follicle<br />

in cancer treatment patients, or other infertile women may allow fertility<br />

preservation. Also Production of a large number of oocytes may facilitate<br />

future derivation of embryonic stem cell for regenerative medicine.<br />

Supported by: This work was Supported by a grant from the Korea Healthcare<br />

Technology R&D Project, Ministry for Health, Welfare & Family affairs,<br />

Republic of Korea (A120080).<br />

P-434 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

THE COMPARISON OF THE EFFECTS OF SINGLE-DOSE METH-<br />

OTREXATE AND SALPINGECTOMY ON OVARIAN RESERVE IN<br />

TERMS OF THE ANTI-MULLERIAN HORMONE LEVEL<br />

MEASUREMENT. G. Sahin Ersoy. Obstetrics and Gynecology, Dr. Lutfi<br />

Kirdar Education and Research Hospital, Istanbul, Turkey.<br />

OBJECTIVE: Ectopic pregnancy constitutes 1-2% of all pregnancies and<br />

has 6% maternal mortality rate. Both surgical treatment and methotrexate<br />

(MTX) administration for ectopic pregnancy is still controversial in terms<br />

of ovarian reserve protection. Anti-mullerian hormone (AMH) is the best current<br />

available measure of ovarian reserve for different clinical conditions.<br />

(ref). To the best of our knowledge there is not any study dealing with the<br />

long term effects of these two treatment methods on AMH levels. Therefore<br />

we aimed to compare the effects of single-dose MTX and salpingectomy on<br />

ovarian reserve in fertile women with ectopic pregnancy in the long term.<br />

DESIGN: Cross-sectional study.<br />

MATERIALS AND METHODS: A total of 181 patients have been<br />

included into the study; 101 of them had received single-dose MTX or salpingectomy<br />

treatment for ectopic pregnancy in the last 12 to 18 months. Patients<br />

who have underwent salpingostomy, tubal milking, fimbriectomy and other<br />

tuba-ovarian surgery were excluded from the study. Patients in the salpingectomy<br />

group did not receive MTX. Their AMH, follicle stimulating hormone<br />

(FSH), estrogen and antral follicle counts were evaluated.<br />

RESULTS: The duration of time passed between application of treatment<br />

and the measurement of AMH levels for MTX and salpingectomy groups<br />

were 15.43 1.77 and 14.91 1.84, respectively (p¼0.144). The average<br />

age was similar in both groups (p¼0.094). None of the three groups displayed<br />

a significant difference in terms of FSH, E2 levels and antral follicle counts<br />

(p¼0.393, p¼0.1<strong>17</strong>, p¼0.528, respectively).There were no statistically significant<br />

difference in AMH levels in all groups (2.96 0.85, 2.82 0.77<br />

and 2.68 0.59 for control, MTX and salpingectomy groups respectively,<br />

P¼0.147). (Table reference:aData are expressed as mean SD. AMH¼<br />

Anti-mullerian hormone, FSH¼ Follicle stimulating hormone, E2¼ Estrogen,<br />

AFC¼ Antral follicle count.<br />

CONCLUSIONS: Neither single-dose MTX nor salpingectomy does have<br />

any detrimental effects on ovarian reserve for the treatment of ectopic pregnancy<br />

in terms of serum AMH levels and does alter antral follicle counts in<br />

the long term.<br />

Comparison of age, measures of ovarian reserve and smoking status.<br />

Control<br />

(n¼80)<br />

Methotrexate<br />

(n¼56)<br />

Salpingectomy<br />

(n¼45)<br />

Age (years) a 27.15 3.49 27.18 2.97 28.33 3.01 0.094<br />

AMH(ng/mL) a 2.96 0.85 2.82 0.77 2.68 0.59 0.147<br />

FSH(mIU/mL) a 6.87 1.79 7.23 1.41 7.10 1.33 0.393<br />

E2(pg/mL) a 44.61 13.94 45.41 13.95 40.11 7.63 0.1<strong>17</strong><br />

AFCa 10.25 1.87 10.10 1.92 9.84 2.01 0.528<br />

Smoking<br />

(user/total)<br />

<strong>17</strong>/80 11/56 10/45 0.949<br />

P-435 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

IS A FREEZE-ALL CYCLE JUSTIFIED IN POOR<br />

RESPONSE? M. Berkkanoglu, K. Coetzee, H. Bulut, K. Ozgur. Antalya<br />

IVF, Antalya, Turkey.<br />

OBJECTIVE: To investigate what the best treatment strategy is, if there<br />

are 4 or less oocytes after an oocyte pick-up (OPU).<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: The study group included 431 women<br />

undergoing an antagonist protocol for intracytoplasmic sperm injection<br />

and had 4 or less oocytes after OPU. Later, they were asked to have embryo<br />

transfer (ET) on day 2 or day 5 or all freeze on day 5 and subsequent transfer<br />

in a thawed cycle (FET).278 women (group A) wanted a cleavage stage day 2<br />

embryo transfer and 153 women (group B) wanted their embryos cultured on<br />

to day 5 to allow for blastocyst formation. 51 (group B1) of the day 5 group<br />

had a fresh blastocyst transfer and 44 (group B2) had all their blastocysts<br />

vitrified and had a vitrified-warmed blastocyst transfer. Blastocyst vitrification<br />

was performed using the Cryotop method and technology.Student’s t<br />

test and Chi-square test were used for statistical comparisons.<br />

RESULTS: There was no significant difference in the mean age between<br />

the groups (p¼0,288). 204 patients of group A and 95 of group B had ETs.<br />

The cancellation ratio of ET after OPU was significantly lower in group A<br />

than group B (26.1% vs 37.9%, p ¼ 0.02), but the pregnancy rates (PR)<br />

per OPU were higher in group B than group A (31.3% vs 18.3%,<br />

p¼0.003). PRs per ET were also higher in group B than group A (50.5%<br />

vs 25%, p


P-436 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DOES AGE MODIFY THE PREDICTIVE ABILITY OF<br />

AMH? P. Kovacs, a B. Buzas, a F. Rarosi. b a Kaali Institute IVF Center, Budapest,<br />

Hungary; b Dept. of Medical Physics, Bolyai Institue, Univ. of<br />

Szeged, Szeged, Hungary.<br />

OBJECTIVE: Anti Mullerian hormone (AMH) is one of the best<br />

markers of ovarian reserve but it does not predict pregnancy well<br />

either. Age is an important factor influencing fertility treatment<br />

outcome. Our aim was to test the predictive ability of AMH in<br />

different age categories.<br />

DESIGN: Single IVF practice observational study.<br />

MATERIALS AND METHODS: All IVf cycles performed between<br />

Jan 1, 2014 and Dec 31, 2014 were considered for the analysis. Cycles<br />

involving donor egg use and cycles that have not progressed to egg<br />

retrieval were excluded. Cycles in which AMH result was not available<br />

were excluded too. Three AMH categories were evaluated (low: under<br />

1.1 ng/ml; normal: 1.11-3.5 ng/ml; high: over 3.5 ng/ml). IVF treatment<br />

outcomes were compared in 3 age categories (under 35 yrs, 35-<br />

40 yrs, over 40 yrs) based on AMH values. Egg yield, amount of gonadotropins<br />

used per oocyte collected (gonadotropin IU/oocyte) and<br />

ongoing pregnancy rate (OPR) were compared. The low AMH category<br />

was further divided based on FSH/estradiol results (FSH>10IU/l and/<br />

or E2>250 pmol/l and FSH


CONCLUSIONS: Compared to conventional IVF, ZACS permits DOR<br />

patients to reap the benefits of trophectoderm biopsy aneuploidy screening<br />

and offers logistical advantages over other embryo banking approaches. Patients<br />

40 years<br />

and/or had AMH


MATERIALS AND METHODS: The study was performed at the<br />

Department of Obstetrics and Gynecology, Hacettepe University, between<br />

November 2013 and March 2014. Total 2085 patients underwent<br />

an ultrasound scan (Voluson 730, GE Healthcare, Istanbul, Turkey) by<br />

one of the two physicians (C.P. or Z.D.) between days 1 and 12 of the<br />

menstrual cycle. Inclusion criteria were (1) age 20 - 50, (2) regular<br />

menstrual bleeding between <strong>21</strong> to 35 days, and (3) optimal visualization<br />

of both ovaries. The exclusion criteria were (1) hirsutism or menstrual<br />

irregularity, (2) any hormonal drug or oral contraceptive pill use<br />

within the last 6 months, (3) history of endometrioma cystectomy or<br />

detection of current endometrioma at the time of ultrasonography,<br />

and (4) pregnancy. The fertility status was not a criterion while<br />

deciding to include or exclude. The sum of both ovaries counts produced<br />

the final AFC. The LMS method was preferred to produce the<br />

smoothed centile curves of antral follicle count by age. For each set<br />

of percentile curves, the initial smoothing methods were applied to<br />

10th, 25th, 50th, 75th and 90th percentiles. For the comparison of<br />

AFC across the age groups, statistical analyses were performed by<br />

SPSS v<strong>21</strong>.0. Significance value was set at 0.05. Approval from institutional<br />

review board was obtained.<br />

RESULTS: Of the 2805 women that had been examined with ultrasonography<br />

during the study period, 381 were appropriate for the final evaluation<br />

according to inclusion and exclusion criterion (Table 1). The mean decrease<br />

of AFC in each year was 0.41. Among the age groups, there were no statistical<br />

significance between 20-24, 25-29 and 30-34, whereas decline in AFC<br />

was obvious after 35 years.<br />

CONCLUSIONS: The documentation of AFC nomogram might present<br />

several benefits in theory. Initially, marking the current AFC of a given<br />

woman across the nomogram might predict a woman’s total fertility potential<br />

not only for assisted reproduction technologies (Polyzos, 2014) but also for<br />

the chance of natural conception and pregnancy outcome. In conclusion,<br />

comparison of our data and others suggest that AFC nomogram from<br />

different reports and countries present similarity regarding the decline by<br />

age. Those nomograms might be used to define age specific thresholds according<br />

to the clinical conditions.<br />

P-442 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

PATIENTS WITH LOW THEN NORMAL AMH MEASUREMENTS<br />

DEMONSTRATE OVARIAN RESERVE COMPARABLE TO THOSE<br />

WITH TWO NORMAL MEASUREMENTS. J. Gingold, a,b<br />

j. M. knopman, c S. Talebian, d M. C. Whitehouse, e J. A. Lee, d<br />

A. B. Copperman. d,a a Obstetrics, Gynecology and Reproductive Science,<br />

Icahn School of Medicine at Mount Sinai, New York, NY; b Obstetrics/Gynecology<br />

and Women’s Health Institute, Cleveland Clinic<br />

Foundation, Cleveland, OH;<br />

c Reproductive Medicine Associates of<br />

New York, NY, NY;<br />

d Reproductive Medicine Associates of New<br />

York, New York, NY;<br />

e Reproductive Medicine Associates of New<br />

York, New York City, NY.<br />

OBJECTIVE: There is a subset of patients who present initially with an<br />

unexpectedly low AMH level. These situations precipitate concern about<br />

short- and long-term female reproductive function. A later discordant confirmatory<br />

test raises questions about whether the lower or the higher level is the<br />

better predictor of ovarian reserve and responsiveness. This study sought to<br />

evaluate whether oocyte yields in patients with discordant AMH levels were<br />

more similar to those of patients with concordantly normal or abnormal<br />

levels.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Patients who underwent two random<br />

AMH measurements separated by > 30 days in the year prior to their<br />

IVF start date from July 2007-March <strong>2015</strong> were included. Patients<br />

with AMH >5 ng/mL were excluded. The first and second AMH<br />

measurements were categorized as abnormal (


OBJECTIVE: This study was for aim, to identify circulating microRNAs<br />

(miRNAs) in human follicular fluid (FF) in relation to<br />

ovarian reserve disorders, providing new biomarkers of female infertility.<br />

DESIGN: This prospective monocentric study included 131 patients<br />

among which 91 women with normal ovarian reserve, 10 with ovarian<br />

insufficiency (OI) and 30 affected by polycystic ovary syndrome<br />

(PCOS). A pool of FF was retrieved for each patient during IVF/ICSI<br />

procedure.<br />

MATERIALS AND METHODS: At oocyte retrieval day, all FF were<br />

collected and pooled for each patient. MicroRNAs were extracted from FF<br />

pools and quantified by RT-qPCR, using TaqMan technology. The candidate<br />

miRNAs, detected in FF samples were let-7b, miR-30a, miR-140 and miR-<br />

191.<br />

RESULTS: The miR-30a expression was significantly increased in FF<br />

pools from women with PCOS compared to those with normal ovarian<br />

reserve (p¼0.006). By contrast, miR-140 and let-7b levels were significantly<br />

expressed at low levels in FF pools from PCOS patients than patients<br />

with normal ovarian reserve (p¼0.01; p¼0.01, respectively). The<br />

Receiving Operator Curve (ROC) analysis showed that the performance<br />

of the combination of these three miRNAs in detecting PCOS reached<br />

0.83 [0.73-0.92] with 83.8% specificity and 70% sensitivity<br />

(p40g/m2,<br />

or bone marrow transplant (n¼20) and others were classified as having<br />

low dose treatment (n¼22). A matched set of randomly timed fingerstick<br />

dried blood spot samples and venipuncture serum samples were<br />

collected from each participant and sent to Ansh Labs for analysis using<br />

the DBS AMH ELISA and picoAMH ELISA. These are three-step, sandwich-type<br />

enzymatic microplate assays, the DBS was developed to measure<br />

AMH levels in two 7.9 mm dried blood spot discs after extracting<br />

the blood from the filter paper. Levels of serum and bloodspot AMH<br />

were compared between groups using t-tests and multivariable linear<br />

regression.<br />

RESULTS: Compared to similar age controls (mean age 25.9), cancer<br />

survivors had significantly lower serum and bloodspot AMH<br />

values (bloodspot AMH 1101pg/ml vs. 2957pg/ml, p¼0.009) in unadjusted<br />

and age adjusted models. High dose survivors had significantly<br />

lower serum and bloodspot AMH values (bloodspot AMH 474pg/ml<br />

vs. 2023pg/ml, p¼0.005) in unadjusted and age adjusted models.<br />

Finally, late reproductive age controls (mean age 42.9) had significantly<br />

lower serum and bloodspot AMH compared to mid-reproductive<br />

age controls (mean age 25.9) (bloodspot AMH 759pg/ml vs.<br />

2957pg/ml, p¼0.003).<br />

CONCLUSIONS: Differences in ovarian reserve can be detected in<br />

cancer survivors and late reproductive women compared to mid reproductive<br />

controls using the Ansh Labs DBS AMH assay on bloodspot<br />

samples, or picoAMH assay on serum samples. Blood spot collection<br />

may be a useful, less invasive, low cost tool in evaluating ovarian<br />

reserve and reproductive potential in populations at risk for ovarian<br />

insufficiency.<br />

Supported by: Ansh laboratories RO1-HD-062797-05.<br />

P-446 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

CLINICAL USEFULNESS OF AMH APPLIED IN POOR OVARIAN<br />

RESERVE PATIENTS IN AN ART PROGRAM. M. Chen, H. Guu,<br />

Y. Chen, Y. Yih, M. Chou. Department of Obstetrics and Gynecology and<br />

Women’ Health Health, Taichung Veterans General Hospital, Taichung,<br />

Taiwan.<br />

OBJECTIVE: The purpose of present study is to investigate the threshold<br />

value of AMH to predict a poor ovarian response after COH and determine<br />

appropriate protocol to maximize the successful rate of treatment in those<br />

POR patients in our ART program.<br />

DESIGN: A retrospective study with the ART database from Jan 2011 till<br />

Oct 2013 was reviewed.<br />

MATERIALS AND METHODS: We had a total 994 start cycles and<br />

908 finished OPU cycles with total 36.8 % CPR and 28.5% OGPR/<br />

LBR. 416 start cycles had AMH value within 6 months of COH. 161<br />

(5 cancelled) long agonist cycles and <strong>21</strong>6 (27 cancelled) antagonist cycles<br />

were eligible for analysis. POR was defined as cycles < 3 (mature)<br />

oocytes retrieved or cancelled due to poor response after COH. ROC<br />

curves for predicting POR by AMH value were created and threshold<br />

levels were determined. Predicting factors for POR in these patients<br />

were searched. The COH responses and the ART outcome of those patients<br />

with AMH value around the threshold value were analyzed according<br />

to the COH protocol and gonadotropin dosage administered.<br />

Clinically useful opinions were derived from all these preliminary analyses.<br />

Statistics were carried out by SPSS-PC ver.11.5 with p


CANCER<br />

P-447 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

TUBAL LIGATION DOES NOT MODIFY OVARIAN CANCER RISK<br />

IN BRCA MUTATION CARRIERS. J. Chan, a M. D. Sammel, b<br />

T. M. Friebel, c C. Gracia, d T. R. Rebbeck. c a Reproductive Endocrinology<br />

and Infertility, Hospital of the University of Pennsylvania, Philadelphia,<br />

PA; b Univ. of Pennsylvania, Perelman School of Medicine, Rose valley,<br />

PA; c Center for Clinical Epidemiology and Biostatistics, University of Pennsylvania,<br />

Philadelphia, PA; d University of Pennsylvania.<br />

OBJECTIVE: Risk-reducing salpingo-oophorectomy (RRSO) has been<br />

shown to be protective against ovarian cancer (OvCa) in the BRCA mutation<br />

carrier population and is recommended in women by age 40 or at completion<br />

of childbearing(1). However, early RRSO is associated with premature<br />

menopause. As tubal ligation (TL) has been demonstrated to reduce the<br />

risk of OvCa in the non-BRCA population(2), we aimed to determine if<br />

TL has a similar protective effect in BRCA carriers.<br />

DESIGN: Cohort study<br />

MATERIALS AND METHODS: Data from 5038 BRCA carriers were<br />

collected between 1997 and 2011 from 23 centers. Demographics and reproductive<br />

history/surgery were obtained. Occurrences of OvCa and vital status<br />

were verified by medical records. Women who underwent TL were compared<br />

with women who did not (non-TL). For failure-time analysis, observation<br />

period started at date of birth and allocated all person-years of observation<br />

(PYO) before TL to the non-TL group. PYO were thereafter allocated to<br />

the TL group. Observation ended at age at OvCa diagnosis, RRSO, last contact,<br />

death or study close, whichever came first. Log-rank test was used to<br />

compare survival distributions. A Cox model with TL as a time-dependent<br />

variable was used to obtain hazard ratios. Type of mutation, parity and oral<br />

contraceptive pill (OCP) use were considered potential confounders. Ranksum<br />

and Chi-square tests were used to compare demographic characteristics.<br />

RESULTS: 392 underwent TL at a median age of 33. Groups were similar<br />

with regards to BRCA mutation type (p¼0.5), Caucasian race (p¼0.7) and<br />

Jewish heritage (p¼0.4). The TL group was older than the non-TL group<br />

at the time of BRCA diagnosis (49 vs. 43, p


Histologic analysis of the tissue samples revealed no sign of microscopic<br />

infiltration. Residual disease was detected by IgH monoclonality expression<br />

in three of the mice from the DXM treated group and in three animals from<br />

the CONTROL group. RCH-ACV cells were detected six months after injection<br />

of amounts as little as 1000 cells, although only animals being injected<br />

with 5x106 developed a clinical disease.<br />

CONCLUSIONS: DXM incubation prior to re-transplantation of ovarian<br />

tissue does not prevent the reintroduction of malignant disease. Although<br />

the amount of cells introduced seem not to cause any relapse, caution in retrasplanting<br />

ovarian tissue from patients with leukemia must be paid until<br />

more secure system are developed.<br />

Reference:<br />

1. Kaspers GJ, Veerman AJ, Popp-Snijders C, et al. Comparison of the<br />

antileukemic activity in vitro of dexamethasone and prednisolone in<br />

childhood acute lymphoblastic leukemia. Med Pediatr Oncol.<br />

1996;27(2):114-1<strong>21</strong>.<br />

P-450 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DOES DIMINISHED OVARIAN RESERVE CONTRIBUTE TO<br />

CANCER SURVIVORS HAVING FEWER CHILDREN THAN<br />

DESIRED? A REPORT FROM THE FUCHSIA WOMEN’S<br />

STUDY. P. P. Howards, a H. B. Chin, a A. Fothergill, a A. C. Mertens, b<br />

J. B. Spencer. c a Epidemiology, Emory University, Atlanta, GA; b Pediatrics,<br />

Emory University, Atlanta, GA; c Gynecology and Obstretrics, Emory University,<br />

Atlanta, GA.<br />

OBJECTIVE: To examine whether cancer survivors are failing to meet<br />

their reproductive goals more than women never diagnosed with cancer,<br />

and to compare antral follicle count (AFC) measurements between these 2<br />

groups.<br />

DESIGN: The FUCHSIA Women’s Study is a population-based cohort<br />

study comparing reproductive outcomes among cancer survivors diagnosed<br />

between 20 and 35 years to comparison women who were never diagnosed<br />

with cancer, matched on age and place of residence.<br />

MATERIALS AND METHODS: Cancer survivors were identified in the<br />

Georgia Cancer Registry. Participants included 1,282 cancer survivors and<br />

1,073 comparison women who were interviewed about their reproductive<br />

history and desire for children. A subgroup of women also completed a clinic<br />

visit where AFC was measured by ultrasound (n ¼ 374 cancer survivors, n ¼<br />

376 comparison women).<br />

RESULTS: Cancer survivors and comparison women were 22-45 years at<br />

the time of the interview. During the interview, both groups of women reported<br />

wanting an average of 2.5 children, but cancer survivors were less<br />

likely to have met their reproductive goals compared with comparison<br />

women (45% vs. 57%). There were also differences in cancer survivors’ ability<br />

to meet their reproductive goals across cancer types. Women diagnosed<br />

with a reproductive or blood cancer were least likely to achieve their desired<br />

family size (42%). In a logistic model adjusting for age at interview, race, income,<br />

and education, cancer survivors were significantly more likely to<br />

report having fewer children than desired compared with comparison women<br />

(OR ¼ 1.72, 95% CI: 1.45, 2.05). Cancer survivors also had a lower total<br />

AFC (median ¼ 8, IQR: 3, 15) than comparison women (median ¼ 10,<br />

IQR: 5, 18). Among women who had fewer children than desired, the median<br />

AFC was 8, IQR: 4, 16 in cancer survivors and 11, IQR: 5, 19.5 in comparison<br />

women. Women who never received chemotherapy or radiation were similar<br />

to other cancer survivors with respect to meeting their reproductive goals<br />

(44%), but had an AFC similar to the comparison women (median ¼ 10,<br />

IQR: 5, 18).<br />

CONCLUSIONS: Diminished ovarian reserve may contribute to cancer<br />

survivors not having as many children as they want although other factors<br />

may be important as well.<br />

Supported by: NICHD 5R01HD066059 NICHD T32HD052460<br />

P-451 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

SPERMATOZOA PROTEIN PROFILES IN CRYOBANKED SEMEN<br />

SAMPLES FROM TESTICULAR CANCER PATIENTS BEFORE<br />

TREATMENT. A. Agarwal, a E. Tvrda, a,b R. Sharma, a S. Gupta, a<br />

G. Ahmad, a,c E. S. Sabanegh. d a Center For Reproductive Medicine, Cleveland<br />

Clinic, Cleveland, OH; b Department of Animal Physiology, Faculty of<br />

Biotechnology and Food Sciences, Nitra, Slovakia; c Physiology and Cell<br />

Biology, University of Health Sciences, Lahore, Pakistan; d Urology, Cleveland<br />

Clinic, Cleveland, OH.<br />

OBJECTIVE: Patients diagnosed with cancer are encouraged to freeze<br />

their sperm prior to the start of treatment. It is not clear how testicular cancer<br />

impacts sperm quality at the proteomic level. Our goal was to identify differentially<br />

expressed proteins in spermatozoa of patients with testicular cancer<br />

who banked their specimens prior to start of their treatment and in spermatozoa<br />

from infertile patients without cancer.<br />

DESIGN: Prospective study examining proteomic profile of spermatozoa<br />

from testicular cancer and non-cancer (infertile) patients.<br />

MATERIALS AND METHODS: Semen samples were taken from 16<br />

testicular cancer patients, prior to the start of their treatment and from 9<br />

non-cancer patients (infertile men). Semen analysis was conducted before<br />

the samples were banked. For proteomic analysis, cryoprotectant was<br />

removed and the samples were pooled after normalizing for protein concentration.<br />

Proteins were extracted and separated by 1-D gel. Bands were digested<br />

and run on a LTQ-Orbitrap Elite hybrid mass spectrometer system.<br />

Functional annotations of proteins were obtained using bioinformatics tools<br />

and pathway databases.<br />

RESULTS: The most prevalent testicular cancer diagnoses included<br />

seminomas, mixed germ cell tumors, embryonal carcinomas and Sertoli<br />

cell tumors. Significant differences in spermatozoa concentration were<br />

observed between the testicular cancer and infertile patients (p


with cancer with and without embryo banking it was 31.0% and 36.7%,<br />

respectively; for women without cancer, with and without embryo banking,<br />

it was 31.0% and 31.3%, respectively. At cycle 2, live birth rates, and mean<br />

birthweight and length of gestation did not differ significantly between<br />

groups.<br />

CONCLUSIONS: Women with cancer utilize embryo banking at higher<br />

rates but are less likely to return for subsequent cycles and do so after longer<br />

periods compared to women without cancer. Women with cancer who return<br />

have comparable birth outcomes (live birth rates, birthweights and lengths of<br />

gestation) to women without cancer. The use of embryo banking after cancer<br />

may depend on several important factors, including cancer diagnosis, woman’s<br />

age, and her partnership status.<br />

P-453 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

TAMOXIFEN METABOLITES IN BREAST CANCER PATIENTS<br />

UNDERGOING CONTROLLED OVARIAN HYPERSTIMULATION<br />

(COH): ARE WE ACHIEVING THERAPEUTIC LEVELS?. L. Ross, a<br />

N. J. Clarke, b F. Z. Stanczyk, a K. Chung. a a University of Southern California,<br />

Los Angeles, CA; b Quest Diagnostics Nichols Institute, San Juan Capistrano,<br />

CA.<br />

OBJECTIVE: In breast cancer patients undergoing COH, tamoxifen is<br />

used as an adjunct to gonadotropins to reduce the risk of stimulation of breast<br />

cancer cells. The active metabolite, endoxifen, can be measured to assess<br />

therapeutic activity. Our objective was to determine if the standard tamoxifen<br />

dose used during COH in breast cancer patients results in therapeutic levels<br />

of the active metabolite.<br />

DESIGN: Retrospective analysis.<br />

MATERIALS AND METHODS: Breast cancer patients with no previous<br />

chemotherapy exposure who received tamoxifen during COH between<br />

2008-2013 were identified. Tamoxifen 20 mg was initiated at<br />

the start of COH and continued until oocyte retrieval. Samples were acquired<br />

from serum stored from routine testing. Tamoxifen and its<br />

metabolite levels were quantitatively measured throughout the COH cycle<br />

using isotope dilution tandem mass spectrometry on a triple quadrupole<br />

instrument. ANOVA and Fisher’s exact tests were used as<br />

appropriate.<br />

RESULTS: 9 women who met inclusion criteria were included in this<br />

analysis. Mean age was 37.34.18. Mean BMI was 24.25.39. Mean number<br />

of days on tamoxifen was 11.71.58. Tamoxifen and endoxifen serum<br />

levels increased with increasing duration of treatment. Mean peak tamoxifen<br />

level was 71.135.4 ng/mL and mean peak endoxifen level was<br />

7.724.14 ng/mL. 7/9 women reached therapeutic endoxifen levels<br />

(>5.97 ng/mL [1]). On average, therapeutic levels were reached by day 9<br />

of stimulation (range day 7-10). Asian women were significantly less likely<br />

than other ethnicities to reach therapeutic endoxifen levels (p¼0.03). While<br />

there was no statistically significant difference in peak tamoxifen levels between<br />

Asians compared to all other ethnicities (54.0vs76.0 ng/mL,<br />

p¼0.47), there was a significant difference in peak endoxifen levels<br />

(2.61vs9.19 ng/mL, p¼0.04).<br />

CONCLUSIONS: Our data indicate that tamoxifen administered at 20mg<br />

daily during COH achieved therapeutic endoxifen levels in the majority of<br />

women, suggesting that this dose should provide adequate protection from<br />

stimulation of breast cancer cells. Of concern, this dose did not reach therapeutic<br />

levels in women of Asian ethnicity. It is possible that higher doses of<br />

tamoxifen may be required in women of Asian ethnicity. Data on 8 more subjects<br />

are currently being analyzed and will be included at the time of abstract<br />

presentation.<br />

Reference:<br />

1. Madlensky L, Natarajan L, Simone T, Pu M, Mortimer J, Flatt S, Nikoloff<br />

D, Hillman G, Fontecha M, Lawrence J, Parker B, Wu A, Pierce<br />

J. Tamoxifen Metabolite Concentrations, CYP2D6 Genotype and<br />

Breast Cancer Outcomes. Clin Pharmacol Ther 2011; 89(5): 718-725.<br />

Supported by: Quest Diagnostics.<br />

P-454 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

PATIENTS WITH LYMPHOMA DEMONSTRATE DIMINISHED<br />

OVARIAN RESERVE BEFORE CANCER<br />

TREATMENT. J. Lekovich, a A. Lobel, a J. Stewart, b N. Pereira, c<br />

I. Kligman, a Z. Rosenwaks. a a Weill Cornell Medical College, New York,<br />

NY; b Weill Cornell Medical Center, New York, NY; c The Ronald O. Perelman<br />

and Claudia Cohen Center, New York, NY.<br />

OBJECTIVE: Evidence shows that men with Hodgkin’s and non-Hodgkin’s<br />

lymphoma have lower sperm quality before the start of chemotherapy.<br />

We sought to investigate if women with lymphoma also demonstrate lower<br />

ovarian reserve.<br />

DESIGN: Retrospective cohort study<br />

MATERIALS AND METHODS: The study group comprised 194 patients<br />

with cancer who underwent fertility preservation before initiating<br />

cancer treatment between 2010 and 2013. The control group consisted of<br />

410 patients undergoing elective fertility preservation. Patients were<br />

excluded if they had ovarian cancer, prior oophorectomy or were older<br />

than 40. Cancer patients were compared to the control group. Then, a<br />

sub-analysis within the cancer group compared breast cancer and lymphoma<br />

patients. Primary outcomes included anti-mullerian hormone<br />

(AMH) levels and number of eggs retrieved and cryopreserved. Secondary<br />

outcomes included number of days of stimulation and the total amount of<br />

gonadotropins used for stimulation. Mann-Whitney test was used for comparison<br />

of continuous variables. Chi-Square and Fisher’s exact tests were<br />

used for comparison of categorical variables as indicated. P


FERTILITY PRESERVATION<br />

P-455 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

OVARIAN TRANSPLANTATION WITH SCAFFOLDS FOR DRUG<br />

DELIVERY: AN IN VIVO TRANSGENIC MOUSE<br />

MODEL. C. Chen, a,b S. Tan, a,c C. Tzeng. a,b a Center for Reproductive<br />

Medicine and Sciences, Department of Obstetrics and Gynecology, Taipei<br />

Medical University Hospital, Taipei, Taiwan; b Department of Obstetrics<br />

and Gynecology, School of Medicine, College of Medicine, Taipei Medical<br />

University, Taipei, Taiwan; c Graduate Institute of Clinical Medicine, College<br />

of Medicine, Taipei Medical University, Taipei, Taiwan.<br />

OBJECTIVE: Ovarian tissue cryopreservation and autotransplantation is a<br />

promising option for fertility preservation of female cancer patients. However,<br />

ischemia limits the life span of the ovarian grafts after grafting. Vascular<br />

endothelial growth factor (VEGF) can promote angiogenesis, and sphingosine-1-phosphate<br />

(S1P) can protect ovarian grafts from ischemic reperfusion<br />

injury. This study aimed to investigate the efficacy of scaffolds for delivering<br />

different drugs in promoting survival of ovarian grafts.<br />

DESIGN: In vivo study with a transgenic mouse model.<br />

MATERIALS AND METHODS: We use scaffolds served as vehicles for<br />

drug delivery to promote the graft survival.Ovaries from 8-week-old FVB/N-<br />

Tg(PolII-Luc)Ltc transgenic mice with or without scaffolds loaded with S1P<br />

(2 mM, 5 mL) or VEGF (0.2 mg/ml, 5 mL) were transplanted into the peritoneum<br />

of wild-type mice. The graft survival was tracked in vivo by bioluminescence<br />

imaging (BLI) for 4 weeks, and histological examination was<br />

performed at the end of the experiment.<br />

RESULTS: Stronger signals of in vivo BLI were observed in the ovaries<br />

with S1P- and VEGF-loaded scaffolds than those in the scaffolds without<br />

drugs and those without scaffolds. Histological examination also showed<br />

more follicles and surrounding vessels in the S1P group compared with other<br />

groups. The above indicated better survival of the grafts.<br />

CONCLUSIONS: We demonstrated that scaffolds loaded with drug can<br />

promote ovarian graft survival.Scaffolds mimicking the structure and biological<br />

function of native extracellular matrix are beneficial for tissue growth,<br />

and applying tissue engineering technology may overcome some limitations<br />

in regenerative medicine.<br />

Supported by: This work was Supported by the grant MOST 103-23<strong>21</strong>-B-<br />

038-008 from the Ministry of Science and Technology, Taiwan, R.O.C.<br />

P-456 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

OBJECTIVE ASSESSMENT OF THE ONCOFERTILITY EDUCA-<br />

TIONAL INFORMATION, AVAILABLE TO WOMEN, ON THE<br />

WEBSITES OF NCI-DESIGNATED CANCER CENTERS IN THE<br />

US: DO SOCIOECONOMIC DEMOGRAPHIC PROFILES BY STATE<br />

MAKE A DIFFERENCE? C. de Haydu, a S. V. Eleswarapu, b<br />

A. A. Dabaja, b C. M. Duke. a a Yale University School of Medicine, New Haven,<br />

CT; b Henry Ford Health System, Detroit, MI.<br />

OBJECTIVE: Current guidelines recommend that reproductive age<br />

women, newly diagnosed with cancer, should be counseled on fertility preservation<br />

(FP). Hospital websites increasingly serve as portals for reliable<br />

web-based resources to supplement the knowledge of patients and families<br />

regarding their diagnoses & treatments. This study aims to assess the quality<br />

of hospital web-based resources which are available to women undergoing<br />

cancer treatment at major cancer centers.<br />

DESIGN: Prospective observational study.<br />

MATERIALS AND METHODS: A validation rubric for FP/oncofertility<br />

content quality standards using a scoring system for commonly accepted definitions<br />

& terminology was developed. The publicly available websites of the National<br />

Cancer Institute Designated Cancer Centers (NCICC) & the Cleveland<br />

Clinic Foundation (CCF) were accessed by independent teams from two different<br />

institutions between Nov. 1, 2014 & April 30, <strong>21</strong>05 & queried in a systematic<br />

fashion. Specific queries included: 1) Does the website discuss the effects of cancer<br />

& cancer treatment on female fertility? 2) Are options for FP for all patients<br />

discussed? 3) Is there a standalone page dedicated to educating patients on FP? 4)<br />

Is parenting-related cancer survivorship addressed? 5) Is there a link to outside FP<br />

information?. State & Region based demographic information on racial makeup,<br />

household income & poverty status were obtained from the 2010 US Census Bureau’s<br />

‘‘Geographic Level of Poverty & Health Estimates’’. Chi-square tests were<br />

performed to assess for differences between FP website scores (individually &<br />

within Regional groups); analysis was also performed to assess for any correlation<br />

between socioeconomic/racial differences within States/Regions where NCICC<br />

are located. Multivariate logistic regression analyses are ongoing.<br />

RESULTS: 62 clinical NCICC were identified, including CFF. 92% of queried<br />

sites were academic institutions. 84% of all websites mention the risk of cancer<br />

treatment on a woman’s fertility potential but 44% do not discuss FP options<br />

for women. 56% of websites have pages dedicated to discussion of non gender<br />

specific FP options & 65% of the websites contain links to further resources. In<br />

population based adjusted analyses, NCCIC in States where 50% of the population<br />

identified as Non-Hispanic White (even after controlling for socioeconomic<br />

status), p-value < 0.04. There were no differences observed when similar adjustments<br />

& analyses were performed by for US census bureau geographic Regions.<br />

CONCLUSIONS: Preliminary data suggest that NCICC websites are<br />

inconsistent in the quality of oncofertility educational information for female<br />

patients. Racial makeup of a State is associated differences in the quality of<br />

patient centered oncofertility web-based resources for women. These findings<br />

are concerning and suggest that more uniformed efforts aimed at attenuating<br />

these racial gaps in patient education are needed.<br />

P-457 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

THE EFFECT OF VITRIFICATION AND PROGRAMMED<br />

FREEZING ON FROZEN-THAWED HUMAN OVARIAN CORTEX<br />

TISSUE. X. Wang, a C. Fang, a C. Di, a H. Liu, b X. Liang. a a Reproductive<br />

Medicine Research Center, The Sixth Affiliated Hospital of Sun Yat-sen University,<br />

Guangzhou, China;<br />

b Obstetrics and Gynecology Department,<br />

Guangzhou Development District Hospital, Guangzhou, China.<br />

OBJECTIVE: It has been long remained controversial whether vitrification,<br />

compared with programmed freezing, is better for ovarian fertility preservation.<br />

Therefore, this study aimed at comparison of the effects of these two techniques<br />

on the frozen-thawed human ovarian tissues, to provide some<br />

experimental evidences to select a better method to preserve women fertility.<br />

DESIGN: Ovarian tissues came from cases of partial ovariectomy, with necessity<br />

of diagnosis or treatment. And all the pathological reports told neither<br />

tumor cell metastasis in the ovarian tissue nor existence of other nidi. In each<br />

case, the cortex was cut into small pieces, followed by randomly divided into<br />

Fresh (F) group, Vitrification (V) group and Programmed Freezing (PF)<br />

group. Morphology changes, apoptosis in situ, follicle viability and secretion<br />

function were compared with self-control method and in-vitro cultivation.<br />

MATERIALS AND METHODS: Ovarian tissues came from 6 cases,<br />

which met the above criteria. The morphology of tissues were examined<br />

by HE staining. The cell apoptosis in situ was analyzed by TUNEL assay.<br />

Neutral red staining and CaAM/EthD-1 staining were conducted respectively<br />

after collagenase-1 digestion of frozen-thawed tissues, and the survival rate<br />

of small follicles was analyzed. During in vitro culture of frozen-thawed<br />

ovarian tissues, the secretion level of estrogen (E2) and progesterone (P4)<br />

in media were examined and analyzed at Day 2, Day 4, Day 6 and Day 8.<br />

RESULTS: The morphology of follicles and stroma cells in both V Group<br />

and PF Group were similar as F Group. The apoptosis in situ showed no significance<br />

between V Group and F Group, while apoptosis in PF Group is<br />

significantly higher than other two groups. After frozen-thawed, no significant<br />

differences were showed in assessment of viability of small follicles between<br />

PF Group and V Group. Analysis of E2 and P4 level showed that secretion of<br />

E2 had a tendency to change over time, and time effect varies with grouping. It<br />

increased significantly higher in PF Group at Day 4 and Day 8, but similar at<br />

Day 6, when compared with V Group. The secretion of P4 also had a tendency<br />

to change over time, but no significant differences were found in the effect of<br />

interaction of time and grouping, neither in the grouping effect.<br />

CONCLUSIONS: Both of vitrification and programmed freezing can well<br />

preserve the morphological characteristics of ovarian cortex tissue, but vitrification<br />

may be better for protection of DNA integrity in cells. Both of vitrification<br />

and programmed freezing can well preserve the viability of small<br />

follicles in frozen-thawed tissues. Both of the frozen-thawed tissues after<br />

vitrification and programmed freezing can recover secretion function. And<br />

functional reconstruction of frozen-thawed ovarian tissues with these two<br />

techniques still needs further researches.<br />

Supported by: This study was Supported by the National Natural Science<br />

Foundation of China (Grant No.81070495), and the Natural Science Foundation<br />

of Guangdong Province (Grant No. S2013010013404).<br />

P-458 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

MENSTRUATION IS AN UNRELIABLE SURROGATE IN THE<br />

ASSESSMENT OF OVARIAN DAMAGE BY CHEMOTHERAPY: A<br />

PROSPECTIVE LONGITUDINAL STUDY WITH AMH LEVELS<br />

AS THE GOLD STANDARD. G. Bedoschi, a,b S. Goldfarb, c<br />

J. Quistorff, c S. Goswami, d F. Moy, a M. Dickler, c K. Oktay. a,b a Obstetrics<br />

and Gynecology, NYMC, Valhalla, NY; b Innovation Institute for Fertility<br />

Preservation and IVF, Rye, NY; c Memorial Sloan Kettering Cancer Center,<br />

New York, NY; d Yeshiva University, New York, NY.<br />

e262 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


OBJECTIVE: The majority of studies assessing the impact of chemotherapy<br />

(CT) on ovarian reserve continue to use menstruation as a surrogate. We aimed<br />

to determine the reliability of menstrual status and pattern in assessing CTinduced<br />

ovarian damage, using the serum AMH level as the gold standard.<br />

DESIGN: Prospective longitudinal study.<br />

MATERIALS AND METHODS: 81 women with breast cancer stage 1-3<br />

were prospectively enrolled and followed up for 18 months (18mo). Sera<br />

were frozen at baseline (BL) and 18mo post-CT, and were assayed for AMH<br />

(ng/ml) at once. Women kept monthly menstrual calendars. Amenorrhea was<br />

defined as no menses for >6 months, and regular periods as those with <strong>21</strong>-35<br />

day intervals. Results were analyzed with t-test or ANOVA for continuous variables<br />

and chi square or Fisher’s exact test for discrete variables.<br />

RESULTS: The median age at CT was 38 (range 27-44); 72.1% received<br />

anthracycline-based, 13.2% received CMF, 13.2% received taxane-based and<br />

1.5% received other CT regimen. Ten women did not complete menstrual calendars<br />

and were excluded from the analysis. Seventeen (23.9%) patients<br />

developed amenorrhea and 54 (76.1%) were menstruating at the end of<br />

18mo follow up. The groups were similar in age, BL AMH , CT protocol,<br />

and adjuvant tamoxifen use. Of those who were menstruating post-CT,<br />

48.5% had RP and 51.5% had irregular periods (IP). Surprisingly, women<br />

who developed amenorrhea post-CT had higher AMH levels than those<br />

who retained menstruation post CT (0.770.34 vs. 0.230.1 ng/ml,<br />

p¼0.049). However, women with RP showed a trend for higher AMH<br />

post-CT than those with IP (0.310.15 vs. 0.050.02, p¼0.076). Of the<br />

women who had detectable AMH levels at 18mo, a similar percentage had<br />

amenorrhea vs. continued menstruation (41.18% vs. 58.82%, p¼0.393).<br />

Moreover, their mean AMH levels were not significantly different.<br />

CONCLUSIONS: Our data indicate that menstrual status has very little<br />

value in assessing CT-induced ovarian damage. Studies that investigate the<br />

impact of CT or interventions to preserve ovarian function should utilize<br />

more reliable markers such as the serum AMH.<br />

Supported by: NIH HD053112 (NICHD & NCI), Jodi Spiegel Fisher Cancer<br />

Foundation and Susan G. Komen Foundation.<br />

P-459 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DEVELOPMENT OF A NEW LOCAL DRUG DELIVERY SYSTEM<br />

FOR THE UTERUS USING BIO-NANOCAPSULE. K. Koizumi, a<br />

H. Nakamura, a T. Matsuzaki, b S. Kuroda, c Y. Yasui, a K. Furuya, a<br />

T. Miyake, a T. Takiuchi, a K. Kumasawa, a T. Kimura. a a Obstetrics and Gynecology,<br />

Osaka University Graduate School of Medicine, Suita Osaka,<br />

Japan; b Cardiovascular Medicine, Osaka University Graduate School of<br />

Medicine, Suita Osaka, Japan; c The Institute of Scientific and Industrial<br />

Research, Osaka University, Suita Osaka, Japan.<br />

OBJECTIVE: Uterus is the applicable organ for local gene therapy<br />

because it is not part of the peritoneum organ and it can be reached directly.<br />

However, we still do not have any drug delivery system for uterus.Bio-nanocapsule<br />

(BNC) containing hepatitis B virus surface antigen consists of<br />

approximately 50-nm hollow particles displaying a human hepatocyte-recognizing<br />

molecule (pre-S1 peptide). BNC has been used as an HB vaccine for<br />

the last three decades. In this study, we optimized the BNC as a new local<br />

drug delivery system for uterus.<br />

DESIGN: Animal experiment.<br />

MATERIALS AND METHODS: The N terminal of Pre-S1 peptide was<br />

replaced with the TAT (trans-activating transcription factor) peptide.The<br />

Cy7 labeled BNC was transferred into the murine uterine cavity. The distribution<br />

of BNC was observed by in-vivo imaging system and also by immunohistochemistry.<br />

The luciferase expression plasmid DNA was incorporated<br />

into BNC using liposome. The luciferase expression plasmid DNAwas transferred<br />

into uterine cavity using TAT-BNC-liposome complex. The efficiency<br />

of gene transfection was analysed by luciferase assay.<br />

RESULTS: BNCs were observed in the luminal and glandular epithelial<br />

cells, but not in stroma and myometrium. The transfection efficiency of the<br />

TAT-BNC-liposome complex was significantly higher than lipofection.<br />

CONCLUSIONS: These results suggest that BNC could be an applicable<br />

DDS for uterus. In this study, we replaced the N terminal of Pre-S1 peptide<br />

with TAT peptide. However, it is replaceable with sugar chains and antibodies.<br />

Recently there are some reports that some of special sugar chains<br />

are expressed in the uterine endometrium during pregnancy and uterine cancer.<br />

If we can find specific sugar chains or cell surface antibodies on uterine<br />

endometrium for reproductive dysfunction and uterine cancer, this DDS system<br />

can be more targetable.<br />

Supported by: Grants-in-Aid for Scientific Research from the Ministry of<br />

Education, Science and Culture of Japan (Tokyo, Japan).<br />

P-460 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

SMALL ANTRAL FOLLICLE RESPONSIVENESS TO FSH, AS-<br />

SESSED BY THE FOLLICULAR OUTPUT RATE (FORT), IS NOT<br />

ALTERED IN CANCER PATIENTS, CANDIDATES FOR FERTILITY<br />

PRESERVATION. S. Duros, a C. Sonigo, b J. Benard, a C. Sifer, c<br />

M. Grynberg. d a Department of Reproductive Medicine, Hopital Jean Verdier,<br />

APHP, BONDY, France; b Department of Reproductive Medicine, H^opital<br />

Jean Verdier, APHP, BONDY, France; c Department of Cytogenetic and<br />

Reproductive Biology, Hopital Jean Verdier, APHP, BONDY, France; d Department<br />

of Reproductive Medicine, BONDY, France.<br />

OBJECTIVE: To evaluate the small antral follicle responsiveness to exogenous<br />

FSH, assessed by the Follicular Output RaTe (FORT), in cancer patients,<br />

candidates for fertility preservation (FP) using oocyte vitrification<br />

after controlled ovarian hyperstimulation (COH).<br />

DESIGN: Prospective study.<br />

MATERIALS AND METHODS: From July 2013 to December 2014, 71<br />

cancer patients, aged 20-40 years, candidates for oocyte vitrification<br />

following COH (FP group) were studied. Ovarian stimulation characteristics<br />

and outcomes were compared with that of 91 infertilewomen (Control group),<br />

included in an in vitro fertilization program in our centre during the same time<br />

frame, matched for age, antral follicle count (AFC) and serum Anti-Mullerian<br />

hormone (AMH) levels measured just before initiation of the stimulation (d0),<br />

as well as FSH starting dose. All patients had 2 ovaries, no previous history of<br />

chemotherapy and underwent COH using GnRH antagonist protocols. Antral<br />

follicles were counted before FSH administration, and on the day of ovulation<br />

triggering (dOT). FORT was determined by the ratio between the pre-ovulatory<br />

follicle count (16-20 mm) on dOT 100/AFC on d0.<br />

RESULTS: By design, mean age, AFC, AMH and FSH starting dose were<br />

similar in FP and Control groups (31.5 3.6 vs. 32.2 4.9 years; <strong>17</strong>.4 9.8<br />

vs. 16.6 8.3 follicles, 2.9 2.4 vs. 3.0 1.7 ng/mL; 269 81 vs. 248 56<br />

IU, respectively, NS). Characteristics and outcomes of the stimulation in both<br />

groups are reported in the Table.<br />

CONCLUSIONS: The present investigation shows that the cancer status<br />

may not impact the responsiveness of small antral follicles to exogenous<br />

FSH, assessed by the FORT, in candidates for oocytes vitrification. However,<br />

alterations in the granulosa cell function in relation with the malignant disease<br />

may account for the significantly lower levels of serum E 2 reached in the end<br />

of the ovarian stimulation in these patients when compared to infertilewomen.<br />

Characteristic and outcome of the stimulation in FP and Control groups.<br />

FP group<br />

(n¼71)<br />

Control group<br />

(n¼91)<br />

Mean total dose of gonadotropin 29311095 26101188 NS<br />

(IU)<br />

Mean duration of stimulation 10.31.7 10.12.2 NS<br />

(days)<br />

E 2 on dOT (pg/mL) 1261 806 1904997 12 mm (pg/mL) 132.593.2 187.694.8 16 mm on<br />

dOT x 100 / AFC on d0) (%)<br />

3622 3720 NS<br />

P-461 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

ANTI-M€ULLERIAN HORMONE PREVENTS CHEMOTHERAPY-<br />

INDUCED FOLLICULAR BURNOUT. S. Tan, a,b C. Chen, a,c<br />

C. Tzeng. a,c a Center for Reproductive Medicine, Department of Obstetrics<br />

and Gynecology, Taipei Medical University Hospital, Taipei, Taiwan; b Graduate<br />

Institute of Clinical Medicine, College of Medicine, Taipei Medical University,<br />

Taipei, Taiwan; c Department of Obstetrics and Gynecology, School<br />

of Medicine, College of Medicine, Taipei Medical University, Taipei,<br />

Taiwan.<br />

OBJECTIVE: Chemotherapeutic drugs may damage the reproductive system<br />

and lead to infertility and premature ovarian failure. How to prevent<br />

follicular loss is the key to preserve ovarian reserve. Anti-M€ullerian hormone<br />

(AMH), which is produced by the granulosa cells of growing follicles, can<br />

inhibit primordial follicle activation and follicle growth stimulated by follicle-stimulating<br />

hormone. This study aimed to investigate the inhibitory effect<br />

of recombinant AMH on chemotherapy-induced follicular burnout.<br />

p<br />

FERTILITY & STERILITY Ò<br />

e263


DESIGN: In vivo mouse study.<br />

MATERIALS AND METHODS: Eighteen 8-week old mice were randomly<br />

divided into three groups: control, cyclophosphamide (Cy) group, and AMH + Cy<br />

group. Mice in the AMH + Cy group were pretreated intraperitoneally with recombinant<br />

AMH (2 mg) 1 hour before intraperitoneal injection of Cy (150 mg/<br />

kg). Histological examination of ovaries was performed three and seven days later.<br />

RESULTS: In the Cy-treated ovaries, the number of primordial follicles<br />

was decreased, and the ratio of growing to primordial follicles was significantly<br />

increased. Mice pretreated with recombinant AMH had similar ratio<br />

of growing to primordial follicles to the controls, indicating that recombinant<br />

AMH prevented the recruitment of follicles triggered by Cy treatment.<br />

CONCLUSIONS: We demonstrated that pretreatment of recombinant AMH<br />

inhibited the transition of the primordial follicles to growing follicles. Other tests<br />

were under investigation to examine its application to fertility preservation.<br />

P-462 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

VALUE OF ANTIMULLERIAN HORMONE AND ANTRAL FOLLI-<br />

CLE COUNT IN PREDICTING FERTILITY PRESERVATION CY-<br />

CLE OUTCOMES. V. Emirdar, a V. Turan, a F. Moy, b G. Bedoschi, a<br />

K. H. Oktay. a,c a Department of Obstetrics&Gynecology, New York Medical<br />

College, Valhalla, NY; b Biostatistics and Epidemiology, New York Medical<br />

College, Valhalla, NY; c Innovation Institute for Fertility Preservation, New<br />

York, NY.<br />

OBJECTIVE: Appropriate strategy to predict the fertility preservation<br />

(FP) cycle outcomes has not been determined. Our aim was to identify the<br />

values of antimullerian hormone measurements (AMH) and antral follicle<br />

counts (AFC) in the prediction of oocyte yield in response to controlled<br />

ovarian stimulation with letrozole plus gonadotropins (LG).<br />

DESIGN: Secondary analysis of prospectively collected database.<br />

MATERIALS AND METHODS: One hundred and fifty two women with<br />

breast cancer who underwent ovarian stimulation with LG protocol for<br />

oocyte and/or embryo cryopreservation were included. Serum AMH and<br />

AFC were determined on cycle day 2/3. The patients who produced %4 oocytes<br />

or %2 embryos were considered low-responders (LoR).<br />

RESULTS: The mean age of the patients was 34.94.5 years. AMH was<br />

better correlated with oocyte yield and the number of embryos cryopreserved<br />

than AFC (r¼0.625, p


significantly affected by the O2 levels (viability %: HV-PD, 68.4; LV-CD,<br />

59.0; LV-PD; 51.6; HV-CD, 16.6). In a further experiment, follicle grading,<br />

staging and viability in HV-PD and HV-CD was assessed at day 9. Results<br />

clearly indicated the advantage of PD over CD in sustaining follicle quality<br />

(grade 1, %: HV-PD, 40; HV-CD, 7. Viability, %: HV-PD, 89; HV-CD, 66),<br />

and the latter also depended on the quality of the ovary.<br />

CONCLUSIONS: Taken together, data indicate that an optimal level of O2<br />

is required to support the quality and viability of follicles cultured in vitro.<br />

The use of gas permeable dishes enabled us to optimally modulate O2 concentrations<br />

in the tissue vicinity and establishing a healthier environment for<br />

the cultured strips. The design of a bioreactor for modulation of O2 tension,<br />

media supply and shear stress is under way to study the effects of fine tuning<br />

the physicochemical parameters on follicle viability and growth.<br />

P-465 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

SAFETY OF CONTROLLED OVARIAN STIMULATION WITH GO-<br />

NADOTROPINS AND LETROZOLE IN WOMEN WITH BRCA<br />

MUTATIONS. J. Kim, a,b V. Turan, a K. H. Oktay. a a Obstetrics & Gynecology,<br />

NY Medical College, Valhalla, NY; b CHA University, Seoul, Korea, Republic<br />

of.<br />

OBJECTIVE: Safety of fertility preservation (FP) via controlled ovarian<br />

stimulation with letrozole supplementation (COSTLES) prior to breast cancer<br />

treatment has not been determined in BRCA mutation carriers. We sought<br />

to determine whether women with BRCA-mutation-positive (BRCA+) cancer<br />

were at higher risk for relapse after COSTLES.<br />

DESIGN: Secondary analysis of a prospective cohort study.<br />

MATERIALS AND METHODS: A total of 89 women diagnosed with<br />

stage %3 breast cancer and underwent BRCA mutation screening test (26<br />

BRCA+ and 63 BRCA-negative) and COSTLES for FP prior to chemotherapy<br />

were enrolled. Follow-up information was collected either during return<br />

visits, by phone interview, and in some cases by also contacting the patient’s<br />

referring oncologist to confirm the information. Primary endpoint was cancer<br />

recurrence defined as the detection of locoregional tumor (chest wall, regional<br />

nodal disease), distant metastases, or contralateral invasive breast cancer.<br />

RESULTS: There were 26 subjects with BRCA mutations (8 with BRCA1, 7<br />

with BRCA2 and onewith both mutations). The BRCA + group was significantly<br />

younger at cancer diagnosis (31.8 3.8 vs. 35.0 4.1 years, P ¼ 0.004) and FP<br />

treatment (32.3 4.0 vs. 35.1 4.2 years, P ¼ 0.016) than the BRCA-negative<br />

group. The BRCA+ women were more likely to have ER-negative breast cancer<br />

than the BRCA-negative (68% vs. 32%, P ¼ 0.004). Consistent with the previous<br />

reports of lower response in BRCA+ women, the peak E2 levels were lower in<br />

BRCA mutation carriers compared to BRCA-negative (461.4306.0 vs.<br />

604.3470.4 pg/ml, P ¼ 0.011), despite the younger age. The mean follow up<br />

was 4.7 years in both groups. There was one recurrence among the BRCA+<br />

women (BRCA1, 448insA) and four in the BRCA-negative. The two groups<br />

had similar relapse-free survival (Kaplan-Meier method P ¼ 0.18).<br />

CONCLUSIONS: This is the first data demonstrating the safety of ovarian<br />

stimulation in BRCA+ women. COSTLES is unlikely to increase the risk of<br />

breast cancer recurrence in women with BRCA mutations.<br />

Supported by: NIH RO1 HD053112.<br />

P-467 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

WITHDRAWN<br />

P-466 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

WITHDRAWN<br />

FERTILITY & STERILITY Ò<br />

e265


P-468 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

FERTILITY PRESERVATION IN FEMALE PEDIATRIC PA-<br />

TIENTS PRIOR TO GONADOTOXIC THERAPY: PREDICTIVE<br />

FACTORS FOR RECEIVING FERTILITY PRESERVATION<br />

TREATMENT. J. H. Selter, C. N. Cordeiro, F. S. Chuong, J. E. Garcia,<br />

L. A. Kolp, M. S. Christianson. Johns Hopkins University School of Medicine,<br />

<strong>Baltimore</strong>, MD.<br />

OBJECTIVE: To identify predictive factors for receiving fertility preservation<br />

treatment among female pediatric patients.<br />

DESIGN: Retrospective analysis.<br />

MATERIALS AND METHODS: We performed an analysis of female pediatric<br />

patients seen for fertility preservation consultation at an academic<br />

fertility center prior to gonadotoxic treatment from 2002-2014. Multivariate<br />

logistic regression models were constructed to identify factors associated<br />

with receiving fertility preservation treatment. Adjusted odds ratios (OR)<br />

with 95%confidence intervals (CI) were calculated.<br />

RESULTS: Over 12 years, 102 females ages 6-<strong>21</strong> years old were seen for<br />

fertility preservation consultation with a mean age of 16.7 years old; 14.7%<br />

were prepubertal. Diagnoses included the following malignancies: hematologic<br />

(61.8%, n¼63), neurologic (10.8%, n¼11), orthopedic tumors or sarcomas<br />

(9.8%, n¼10), ovarian (8.8%, n¼9) and gastrointestinal (4.9%, n¼5). Remaining<br />

patients had medical conditions requiring gonadotoxic treatment (3.9%,<br />

n¼4). Of those seen for consultation, 25.5% (n¼26 pursued fertility preservation.<br />

The most utilized treatment modality was ovarian tissue cryopreservation<br />

(65.4%, n¼<strong>17</strong>) followed by oocyte cryopreservation (34.6%, n¼9). In<br />

comparing patients who pursued ovarian tissue cryopreservation and oocyte<br />

cryopreservation, there was no significant difference in age or pubertal status.<br />

After adjustment, a history of prior chemotherapy was associated with<br />

decreased odds of receiving fertility preservation treatment (OR¼0.03, 95%<br />

CI 0.01-0.04). Age, race, pubertal status and type of cancer were not statistically<br />

associated with odds of receiving fertility preservation treatment.<br />

CONCLUSIONS: Ovarian tissue cryopreservation emerged as the most<br />

utilized fertility preservation treatment modality in this patient population.<br />

Future work is needed to investigate techniques to use cryopreserved ovarian<br />

tissue for fertility for this group as they reach adulthood. The finding that<br />

prior chemotherapy is negatively associated with fertility preservation in pediatric<br />

patients undergoing gonadotoxic chemotherapy underscores the<br />

importance of earlier referral and education regarding fertility preservation<br />

in newly diagnosed pediatric cancer patients.<br />

P-469 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

PATIENTS WITH SYSTEMIC CANCER HAVE LOWER OVARIAN<br />

RESERVE AND REQUIRE HIGHER GONADOTROPIN<br />

DOSES FOR FERTILITY PRESERVATION (FP). A. V. Dolinko, a,b<br />

L. V. Farland, a,b,c S. A. Missmer, a,b,c C. Racowsky, a,b S. Srouji, a,b<br />

E. Ginsburg. a,b a Center for Infertility and Reproductive Surgery, Dept of Ob-<br />

Gyn, Brigham and Women’s Hospital, Boston, MA; b Harvard Medical<br />

School, Boston, MA; c Harvard School of Public Health, Boston, MA.<br />

OBJECTIVE: To evaluate ovarian stimulation (OS) outcomes in women<br />

with systemic or local cancer<br />

DESIGN: Retrospective cohort study of patients undergoing OS from June<br />

2007 to <strong>October</strong> 2014<br />

MATERIALS AND METHODS: The responses of cancer patients to their<br />

1st OS cycle for FP were compared to patients without cancer undergoing<br />

IVF/ICSI for male factor infertility. Multivariable linear, Poisson, and logistic<br />

regressions were applied to calculate b-coefficients, relative rates, and<br />

odds ratios, respectively, and 2-sided Wald p-values (p).<br />

RESULTS: Of the 147 cancer patients, 105 had local cancer (stage I-III<br />

solid malignancy) and 42 had systemic cancer (hematologic or stage IV solid<br />

malignancy). Women with systemic cancer were significantly younger than<br />

women with no cancer or local cancer. Adjusting for age and BMI, women<br />

with systemic cancer had significantly lower baseline antral follicle count<br />

(AFC) than women with no cancer or local cancer. Also adjusting for tamoxifen<br />

or letrozole use, women with local cancer or systemic cancer required<br />

higher total doses of FSH, and had lower serum E2 levels and total number<br />

of follicles at hCG trigger than women without cancer. Women with systemic<br />

cancer had 14.4 and <strong>17</strong>.0 times greater odds of having a cycle cancellation as<br />

compared to women with no or local cancer, respectively. 14 cancer patients<br />

returned to use frozen embryos in <strong>17</strong> cycles; 7 used gestational carriers (GC)<br />

in 9 cycles. Of these cycles, 7 resulted in live births, 2 as ongoing pregnancies,<br />

and 1 SAB. To date, 19 of the 147 cancer patients have died; 1 of these<br />

had a live birth via GC prior to death.<br />

Characteristic<br />

No Cancer<br />

(n¼664)<br />

Local Cancer<br />

(n¼105)<br />

Systemic Cancer<br />

(n¼42)<br />

Age at Stimulation (y) 34.6 (4.2) 33.6 (4.8) 27.1 (6.4)*^<br />

BMI at Stimulation (kg/m2) 25.7 (4.2) 26.0 (6.3) 24.9 (5.9)<br />

Baseline AFC 9.4 (7.2) 10.2 (7.4) 7.3 (7.5)*^<br />

Total FSH dose (IU) 1839.2 (1294.7) 2813.6 (<strong>17</strong>85.6)* 3358.9 (<strong>21</strong>47.3)*^<br />

Duration of Stimulation (days) 11.7 (2.0) 11.7 (2.3) 12.2 (2.2)<br />

Serum E2 at hCG trigger 2229.0 (905.7) 1255.9 (973.4)* 1503.2 (812.6)*<br />

(pg/mL)<br />

Follicles at hCG trigger 12.9 (6.1) 12.2 (8.4)* 13.2 (5.9)*<br />

Oocytes retrieved 15.7 (8.6) 16.8 (13.6) 20.6 (<strong>21</strong>.0)<br />

Mature (MII) oocytes<br />

76 (19) 76 (20) 78 (15)<br />

retrieved (%)<br />

2PN embryos (n) 8.9 (6.3) 8.8 (6.4) 12.3 (11.7)<br />

Cycle cancelled [n, (%)] 14 (2.1%) 2 (1.9%) 9 (<strong>21</strong>.4%)*^<br />

All values are Mean (SD), unless otherwise noted*Significant difference<br />

compared to ‘‘No Cancer’’, all p%0.03^Significant difference compared to<br />

‘‘Local Cancer’’, all p%0.04<br />

CONCLUSIONS: These data suggest that women with cancer whose cycles<br />

are not cancelled achieve similar oocyte and embryo yields compared to<br />

women with infertile male partners, although they require higher FSH doses<br />

to achieve those outcomes. Those with systemic cancer are at greater risk of<br />

cycle cancellation.<br />

P-470 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

VALIDITY OF SELF-REPORTED CANCER DIAGNOSIS AND GONA-<br />

DOTOXIC TREATMENTS AMONG FEMALE YOUNG ADULT CAN-<br />

CER SURVIVORS. A. N. Knight, a B. W. Whitcomb, b J. Gorman, c<br />

I. Su. d a Reproductive Medicine, University of California San Diego, San Diego,<br />

CA; b University of Massachusetts, Amherst, MA; c Reproductive Medicine,<br />

Moore’s Cancer Center, La Jolla, CA; d UC San Diego, San Diego, CA.<br />

OBJECTIVE: For reproductive endocrinologists, accurate cancer treatment<br />

information is needed for fertility counseling of female young adult<br />

cancer survivors. As primary medical records are hard to obtain, providers<br />

rely on self-report by survivors. We assessed the validity of self-reported cancer<br />

diagnosis and gonadotoxic treatments in this population.<br />

DESIGN: Cross-sectional<br />

MATERIALS AND METHODS: Female cancer survivors ages 18-45 were<br />

recruited to a cohort study on reproductive health. Survivors reported cancer<br />

diagnosis and treatments via an online questionnaire. Medical records were<br />

abstracted for diagnosis and treatments. Self-reported cancer diagnosis and gonadotoxic<br />

therapies (alkylating chemotherapy, stem cell transplant, pelvic radiation,<br />

uterine and ovary surgery) were compared with medical record data.<br />

Logistic regression models estimated odds ratios (OR) for characteristics associated<br />

with accuracy of self-reporting for chemotherapy, radiation and surgery.<br />

RESULTS: 101 survivors were included (mean age 28.2, SD 6.3; median 2.4<br />

years since cancer diagnosis, IQR 1.2-4.4). Lymphoma (33%), breast cancer<br />

(26%), and uterine or ovarian cancer (10%) were the most common cancer types.<br />

98% of survivors accurately reported cancer type. Only 45% reported correct<br />

stage. Sensitivities of self-reported gonadotoxic exposures ranged from 38-<br />

77% (Table). Specificities and negative predictive values were higher. In adjusted<br />

models, breast cancer diagnosis was associated with higher odds of accurate<br />

chemotherapy reporting (OR 5.8, 95% CI 1.6, <strong>21</strong>.2), while Hispanic ethnicity<br />

was associated with lower odds (OR 0.14, 95% CI 0.02, 0.81). No demographic<br />

characteristics were related to accuracy of radiation and surgery reporting.<br />

CONCLUSIONS: Young adult female cancer survivors have poor recall of<br />

gonadotoxic treatment exposures. Self report was accurate for type of cancer,<br />

but not for stage, alkylating chemotherapy, pelvic radiation and surgery.<br />

Healthcare providers who need this type of information will require primary<br />

medical records or cancer treatment summaries.<br />

Sensitivity, Specificity, Predictive values for self-reported gonadotoxic cancer<br />

treatments.<br />

Sensitivity<br />

Alkylating<br />

chemotherapy<br />

Pelvic<br />

Radiation<br />

Stem cell<br />

transplant<br />

Uterine and/or<br />

ovarian surgery<br />

Sensitivity 25/39 (64) 8/13 (61) 5/13 (38) 10/13 (77)<br />

Specificity 44/72 (71) 87/88 (99) 88/88 (100) 88/88 (100)<br />

Positive Predictive Value 25/43 (58) 8/9 (89)<br />

Negative Predictive Value 44/58 (76) 87/92 (92) 88/96 (92) 88/91 (97)<br />

Supported by: Funding: UL1 RR024926 pilot, HD-058799.<br />

e266 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


P-471 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

COMPARISON OF RANDOM START CONTROLLED OVARIAN<br />

STIMULATION WITH STANDARD START IN LETROZOLE<br />

GONADOTROPIN CYCLES FOR FERTILITY PRESERVATION IN<br />

WOMEN WITH BREAST CANCER. G. Bedoschi, a V. Turan, a<br />

V. Emirdar, a M. Sonmezer, b K. H. Oktay. a a New York Medical College,<br />

Valhalla, NY; b Ankara University, Ankara, Turkey.<br />

OBJECTIVE: When women are referred for fertility preservation prior to<br />

chemotherapy, there may not be sufficient time to await menstruation to<br />

initiate ovarian stimulation for embryo or oocyte freezing. In these patients,<br />

random start controlled ovarian stimulation (RSCOH) may be an alternative.<br />

Our aim was to compare the cycle characteristics and outcomes of RSCOH in<br />

the late follicular or luteal phase of the menstrual cycle to the outcomes of<br />

standard start (early follicular) controlled ovarian hyperstimulation<br />

(SSCOH) cycles in women with breast cancer.<br />

DESIGN: Secondary analysis of prospectively collected database in an academic<br />

center<br />

MATERIALS AND METHODS: One hundred and fifty women who underwent<br />

SSCOH and 14 who underwent RSCOH for oocyte and/or embryo<br />

cryopreservation before breast cancer chemotherapy were included. Ovarian<br />

stimulation was performed with concurrent gonadotropin and letrozole-5 mg<br />

treatment; an antagonist was added when there was a follicle>13-mm. Either<br />

hCG or leuprolide acetate was used for trigger.<br />

RESULTS: RSCOH was initiated either in the late follicular (50%) or luteal<br />

phase (50%). The mean age was similar between the groups (RSCOH vs<br />

SSCOH: 34.9 4.5 vs. 32.5 5.5 years, respectively; p ¼ 0.09). Total dose<br />

of gonadotropins was significantly higher in the RSCOH group (4070.4 <br />

1968.1 vs. 2735.5 1<strong>21</strong>9.0 IU, respectively; p < 0.041). Maturation and fertilization<br />

rates were similar between the groups. The mean number of mature oocytes<br />

was significantly higher in the RSCOH group (10.7 6.6 vs. 9.9 4.0;<br />

p¼0.039). The mean number of oocytes retrieved (16.1 7.2 vs. 14.3 9.3,<br />

respectively; p ¼ 0.073), and embryos cryopreserved (8.5 2.7 vs. 7.1 5.5,<br />

respectively; p ¼ 0.053) trended higher in women undergoing RSCOH.<br />

CONCLUSIONS: Our data indicate that the fertility preservation cycle<br />

outcomes with RSCOH are not inferior to those with SSCOH. Further follow<br />

up and larger prospective studies are required to determine the pregnancy<br />

success rates of RSCOH in comparison to SSCOH cycles.<br />

Cycle Characteristics and Outcomes.<br />

Random start<br />

(n¼14)<br />

Standart start<br />

(n¼150)<br />

P value<br />

Total letrozole dose (mg) 54.5 15.2 51.4 9.2 0.025<br />

Total FSH dose (IU) 4070 1968 2735 1<strong>21</strong>9 0.041<br />

Ovarian stimulation length (days) 11.0 2.8 11.6 1.7 0.057<br />

E2 level on trigger day (pg/mL) 678.3 434.1 595.9 410.0 0.575<br />

No. of total oocytes 16.1 7.2 14.3 9.3 0.073<br />

No. of mature oocytes 10.7 6.6 9.9 4.0 0.039<br />

Maturity rate (%) 69.2 18.8 73.5 <strong>21</strong>.2 0.329<br />

Fertilization rate (%) 80.1 30.5 72.3 27.6 0.909<br />

No. of embryos frozen 8.5 2.7 7.1 5.5 0.053<br />

P-472 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

FERTILITY PRESERVATION CHOICES & OUTCOMES OF<br />

WOMEN WITH ADVANCED STAGE CANCERS. K. L. Palmerola, a<br />

J. M. Choi, b M. V. Sauer. b a Columbia University Medical Center, New<br />

York, NY; b Columbia University, New York, NY.<br />

OBJECTIVE: To assess fertility preservation (FP) decisions and outcomes<br />

from assisted reproductive technologies (ART) in newly diagnosed advanced<br />

stage (stage III or IV) cancer patients.<br />

DESIGN: Restrospective case control<br />

MATERIALS AND METHODS: Patients (n¼26) presenting to a single<br />

academic center for FP following a diagnosis of advanced stage cancer (stage<br />

III or IV) between 2007 and <strong>2015</strong> were studied. Women identified as cancer<br />

survivors were excluded. Demographic information, ART cycle data, and<br />

ART outcomes were collected. FP decisions and outcomes were compared<br />

with randomly selected, age-matched patients with stage I or II cancer undergoing<br />

ART during the same period. Mann-Whitney rank sum tests were used<br />

for analysis.<br />

RESULTS: 26 patients with stage III or IV cancer presented for fertility<br />

evaluation between September 2007 and March <strong>2015</strong>. 9 (34.6%) patients<br />

were cancer survivors and excluded from analysis. Of the <strong>17</strong> included patients,<br />

9 (52.9%) were diagnosed with a stage III cancer (5 breast, 2 ovarian,<br />

1 hematologic, 1 lung) and 8 (47.1%) with a stage IV cancer (3 hematologic,<br />

3 gastrointestinal/colorectal, 2 breast). 5 (29.4%) patients pursued FP, collectively<br />

undergoing 6 treatment cycles. The stage III/IV patients who pursued<br />

FP demonstrated no statistically significant differences in baseline E2, FSH,<br />

MIS levels, number of days of stimulation, total gonadotropins prescribed,<br />

peak estradiol level, number of oocytes retrieved and fertilized, number of<br />

embryos or oocytes cryopreserved or cycle cancellations compared to stage<br />

I/II cancer patients. There were no complications from FP in these patients.<br />

Advanced directive status was available for 3 of the 5 patients who pursued<br />

FP. All 3 patients chose to discard unfertilized or cryopreserved eggs; 2 patients<br />

opted to donate cryopreserved embryos to research; 1 patient opted to<br />

donate cryopreserved embryos to other patients. Further information was<br />

available for 8 of the 12 patients who did not pursue FP: 6 opted to defer<br />

infertility treatment until after cancer treatment; 1 was recommended by their<br />

oncologist not to delay cancer therapy; 1 had metastases to her ovaries, and<br />

thus a concern that FP may worsen her disease status.<br />

Comparison of Stage I/II vs. Stage III/IV Fertility Preservation Cycle<br />

Outcomes.<br />

Stage I/II<br />

(medians)<br />

Stage III/IV<br />

(medians)<br />

P-value<br />

Age (years) 32 33 0.341<br />

Baseline FSH (mIU/mL) 5.05 4.46 0.367<br />

Baseline E2 (pg/mL) 40.6 46.2 0.<strong>17</strong>6<br />

AMH (ng/mL) 1.275 0.63 0.288<br />

Days of Stimulation 10 11 0.078<br />

Total Gonadotropins (IU) 4050 4500 0.123<br />

Peak E2 (pg/mL) 1122 1429 0.397<br />

Total # oocytes retrieved 13 12 0.484<br />

MII 8 10 0.33<br />

Fertilization Rate 71.4% 76.8% 0.409<br />

Cancellations 0 0 -<br />

CONCLUSIONS: Despite the potential poor prognosis associated with<br />

advanced cancer, this rare cohort of patients may expect similar fertility preservation<br />

outcomes to their earlier staged counterparts. Regardless of cancer<br />

stage, all patients are candidates for fertility preservation, though prognosis<br />

and advanced directives must be discussed before initiating care.<br />

P-473 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

OUTCOMES OF UTILIZED CRYOPRESERVED AUTOLOGOUS<br />

OOCYTES. J. P. Alvarez, a,b A. l. Akopians, a,b E. T. Wang, a<br />

D. L. Hill, c J. Barritt, c M. Surrey, d,c H. Danzer, d,c M. D. Pisarska. a a Cedars<br />

Sinai Medical Center, Los Angeles, CA; b UCLA, Los Angeles, CA; c ART<br />

Reproductive Center, Beverly Hills, CA; d Southern California Reproductive<br />

Center, Beverly Hills, CA.<br />

OBJECTIVE: Oocyte cryopreservation is a rapidly developing technology<br />

with reassuring outcomes from observational studies. However, there are<br />

limited studies on utilization of cryopreserved oocytes. Thus, our objective<br />

was to compare outcomes in women who returned to utilize their previously<br />

cryopreserved oocytes to IVF cycles that utilized frozen embryos.<br />

DESIGN: Retrospective cohort study from a large fertility center.<br />

MATERIALS AND METHODS: Women who underwent autologous<br />

oocyte cryopreservation for both medical and elective indications between<br />

1/2010 and 12/2014 and returned to utilize their cryopreserved oocytes<br />

were selected. Oocyte cryopreservation was performed either by slow freeze<br />

or vitrification. Oocyte donation cycles were excluded. The control group<br />

consisted of women who underwent IVF-freeze all cycles followed by a<br />

frozen embryo transfer (FET) during the same time period. Clinical outcomes<br />

of interest included number of mature oocytes, fertilization rate, blastocyst<br />

progression (number of blastocysts per mature oocyte), and pregnancy<br />

rate. Statistical analysis was performed using a student t test for continuous<br />

variables and a fisher exact test for categorical variables.<br />

RESULTS: Of 523 patients who underwent autologous oocyte cryopreservation,<br />

29 (5.5%) returned to utilize their cryopreserved oocytes. The mean<br />

age at oocyte cryopreservation was 37.3 4.5 years (range of 24-44), which<br />

was comparable to the control group (P ¼0.24). The mean duration of cryopreservation<br />

was 478.9 days with a range of 76-880 days. The oocyte cryopreservation<br />

group had a similar number of mature oocytes (10.5 7.2 vs<br />

13.1 8.1 P¼ 0.08) and fertilization rate (69% 24.3 vs 73.7% 20.14<br />

FERTILITY & STERILITY Ò<br />

e267


P¼ 0.31) compared to the control group. In addition, the blastocyst progression<br />

was not significantly different between the oocyte cryopreservation and<br />

the control group (0.2 0.19 vs 0.2 0.28 P¼ 0.30). The pregnancy rate per<br />

embryo transfer was not significantly different from the IVF/FET group<br />

(45.5% vs 52.3% P ¼ 0.65).<br />

CONCLUSIONS: Utilization of cryopreserved autologous oocytes leads<br />

to similar outcomes, including pregnancy rates compared to women undergoing<br />

IVF with frozen embryo transfer. With greater utilization of cryopreserved<br />

oocytes, larger studies are needed to confirm these findings.<br />

P-474 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

WITHDRAWN<br />

RESULTS: HLGCs rapidly underwent apoptosis after exposure to cyclophosphamide<br />

and cisplatin. But they were resistant to 5-FU and paclitaxel as<br />

evidenced by similar rates of apoptosis and comparable E2 and P productions<br />

among control, 5-FU and paclitaxel groups. By contrast, paclitaxel inhibited<br />

proliferation and induced apoptosis of COV434 and HGrC1 cells. This effect<br />

was milder than cyclophosphamide and cisplatin. 5-FU had the slightest toxic<br />

effects on the mitotic granulosa cells among the drugs tested.<br />

CONCLUSIONS: The magnitude of cytotoxicity of chemotherapy drugs<br />

varies depending upon their classes and the type of granulosa cells.<br />

P-476 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

WITHDRAWN<br />

P-475 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

HUMAN MITOTIC NON-LUTEINIZING AND NON-MITOTIC LU-<br />

TEINIZED GRANULOSA CELLS HAVE DIFFERENT SENSITIVITY<br />

TO CHEMOTHERAPY AGENTS. N. Akin, a G. Bildik, a Y. Guzel, b<br />

B.-. Balaban, b B. Urman, c,b O. Oktem. c,b a Reproductive Biology, Koc University<br />

Graduate School of Health Sciences, Istanbul, Turkey; b Women’s<br />

Health Center Assisted Reproduction Unit, American Hospital, Istanbul,<br />

Turkey; c Obstetrics and Gynecology, Koc University School of Medicine, Istanbul,<br />

Turkey.<br />

OBJECTIVE: Adult human ovary harbors different types of follicles and<br />

granulosa cells. Of these, granulosa cells of preantral and early antral follicles<br />

are proliferative but are not capable of undergoing luteinization whereas<br />

the granulosa cells of corpus luteum are non-mitotic luteinized type. The aim<br />

of the current study was to investigate if these two types of granulosa cells<br />

differ in terms of their sensitivity to cytotoxic actions of chemotherapy drugs.<br />

DESIGN: A translational research study<br />

MATERIALS AND METHODS: Mitotic non-luteinizing cells (COV434<br />

and HGrC1) and non-mitotic luteinized granulosa cells (HLGCs) retrieved<br />

from IVF patients were used for the experiments. The cells were plated at<br />

a density of 5000 cells/well and treated with 5-fluorouracil (50mg/mL), paclitaxel<br />

(2mg/mL), cisplatin (100mg/mL) and cyclophosphamide (4-hydroperoxy<br />

cyclophosphamide, active in vitro form of the drug, 100mM). The<br />

drugs were used at the concentrations corresponding to their therapeutic<br />

serum levels. Cytotoxic effect of these drugs were determined according to<br />

their impact on the proliferation (real-time and quantitative cell proliferation<br />

index xcelligence system; COV434 and HGrC1 cells only), viability/<br />

apoptosis (YO-PRO-1/cleaved caspase-3 expression) and steroidogenic activity<br />

(estradiol (E2) and progesterone (P) productions) of the granulosa cells.<br />

The impact of different chemotherapy drugs on mitotic vs. non-mitotic human<br />

granulosa cells.<br />

Control Cyclophosphamide Cisplatin Paclitaxel 5-FU p-value<br />

Mitotic non-luteinized granulosa cells<br />

Proliferative<br />

index<br />

3.50.3 0.40.2 (a) 0.60.3<br />

(a)<br />

Apoptosis<br />

(%)<br />

4% 91% (a) 78%<br />

(a)<br />

Non-mitotic luteinized granulosa cells<br />

Estradiol<br />

(pg/ml)<br />

1604116 38956 (a) 43767<br />

(a)<br />

Progesterone<br />

(ng/ml)<br />

Apoptosis<br />

(%)<br />

63092 27142 (a) 33838<br />

(a)<br />

2.50.4<br />

(b)<br />

56%<br />

(b)<br />

1458102<br />

(b)<br />

56956<br />

(b)<br />

3.10.2<br />

(c)<br />

12%<br />

(c)<br />

141<strong>21</strong>03<br />

(b)<br />

60098<br />

(b)<br />

a:


acceptability of one of the most effective, reversible forms of contraception,<br />

the intrauterine device (IUD), in women with cancer is lacking, and<br />

there are no studies on placement of an IUD at the time of oocyte retrieval<br />

for fertility preservation. Thus, we surveyed women with cancer on their<br />

satisfaction with their IUD and any noted side effects. The responses of<br />

women with cancer were also compared to a smaller subset of women<br />

without cancer.<br />

DESIGN: A retrospective survey was performed of women age 18-<br />

45 years who had an IUD placed at our institution between 2007<br />

and <strong>2015</strong>.<br />

MATERIALS AND METHODS: Women who met inclusion criteria were<br />

emailed a link to an internet-based survey. A 5-point Likert scale was used.<br />

Standard descriptive statistics and univariate analyses were performed with<br />

chi-square and T-tests as appropriate.<br />

RESULTS: 46 women completed the survey. 29 had a history of cancer<br />

with breast cancer being the most common cancer diagnosis (72.4%). For<br />

the majority of cancer patients, the discussion about IUD placement was<br />

initiated by a fertility specialist (44.8%) or by the patient herself<br />

(27.6%). Women were overall satisfied (75.0%) with the counseling they<br />

received, and 75.9% required no additional problem visits after IUD placement.<br />

The majority of patients had a copper IUD (75.9%). After IUD placement,<br />

the women described their menses as heavier (58.6%), more irregular<br />

(51.7%) and associated with improved or unchanged menstrual cramps<br />

(56.3%). 75% reported their sex life was improved or unchanged. At the<br />

time of the survey, 84.4% still had their original IUD in place or had<br />

kept the IUD in for over a year. 69.0% reported a positive attitude toward<br />

the IUD, and 79.3% would recommend it to a friend. Thirteen women with<br />

cancer had their IUDs placed at the time of oocyte retrieval. Compared to<br />

those who did not have IUDs placed at the time of retrieval, these women<br />

were more likely to report the procedure as not painful (85.7 vs. 34.4%<br />

p¼0.001) and not stressful (84.6 vs. 43.8%, p¼0.024). Compared to<br />

women without cancer, women with cancer were not more likely to report<br />

stressful or painful IUD placement or changes in bleeding patterns or menstrual<br />

symptoms. There were no differences in high levels of satisfaction<br />

(58.1vs. 41.9%, p¼0.32).<br />

CONCLUSIONS: Despite the increased menstrual symptoms compared to<br />

baseline that cancer patients noted after IUD placement, satisfaction and IUD<br />

retention remained high. This makes the IUD an option for women with cancer,<br />

including women who are undergoing fertility preservation prior to initiating<br />

chemotherapy.<br />

CRYOPRESERVATION<br />

P-478 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DOES LASER ZONA BREACHING (LZB) OF WARMED BLASTO-<br />

CYSTS IMPROVE HATCHING RATES? N. M. Sachdev, a<br />

D. H. McCulloh, b C. McCaffrey, c J. Grifo. d a Obstetrics and Gynecology,<br />

New York University Fertility Center, New York, NY; b New York University<br />

Fertility Center, New York, NY; c NYU Fertility Center, New York, NY;<br />

d NYU Langone Medical Center, New York, NY.<br />

OBJECTIVE: To assess using time lapse microscopy (TLM) if post warming<br />

LZB improves the ability of the blastocyst to completely exit the zona<br />

pellucida (hatching).<br />

DESIGN: Prospective Cohort study.<br />

MATERIALS AND METHODS: Three groups were included in the<br />

study; 1) non biopsied slow frozen blastocysts, 2) non biopsied vitrified<br />

blastocysts and 3) biopsied aneuploid blastocysts, as determined by trophectoderm<br />

biopsy (TE) and array comparative genomic hybridization<br />

(aCGH) performed prior to vitrification. All blastocysts in group 3 had<br />

an existing zona breach prior to cryopreservation. All blastocysts were<br />

warmed using standard protocol. Each group was subdivided into a control<br />

arm with no additional zona pellucida (ZP) manipulation, and a test arm<br />

subjected to LZB following warming. This post warming LZB creates an<br />

opening larger (approximately 1/3 of the zona) than that routinely created<br />

for TE biopsy. Time lapse imaging was reviewed by a single observer. Primary<br />

outcomes were number of successfully hatched blastocysts and<br />

timing of hatching events.<br />

RESULTS: A total of 144 blastocysts donated for research (IRB #H6902)<br />

were included in the study. Survival rate post warming was 96%: An additional<br />

7 embryos were zona free following warming and were excluded<br />

from analysis. Hence, 131 blastocysts were observed using TLM. Group 1<br />

consisted of <strong>21</strong> with LZB and 22 controls. Group 2 consisted of <strong>21</strong> with<br />

LZB and 24 controls. Group 3 consisted of 22 with LZB and <strong>21</strong> controls.<br />

A total of 55 (41.98%) warmed blastocysts completed hatching within 81<br />

hours, while 76 (58%) did not complete hatching despite exhibiting viability.<br />

There was no difference in patient ages among the blastocysts subjected to<br />

LZB (34.58 4.89) and controls (35.95 4.24) p¼0.0895 or within groups.<br />

Groups 1& 2 (non biopsied) showed a difference in the LZB groups versus<br />

controls in the number of blastocysts that completed hatching. Group 3 (previously<br />

biopsied) blastocysts showed no difference and were able to complete<br />

hatching, likely due to the preexisting breach in the ZP created for<br />

the TE biopsy. (Table 1).<br />

Outcomes of Post Warm LZB vs Control Groups.<br />

Group<br />

Lazer Zona<br />

Breaching<br />

(LZB)<br />

% Completed<br />

Hatching (N)<br />

Hours to<br />

P Start of<br />

(completed) Hatching P (start)<br />

Hours to<br />

Complete<br />

Hatching<br />

P<br />

(end)<br />

Group 1 + 66.7% (<strong>21</strong>) .00003 22 9 0.154 38 12 0.161<br />

- 4.5% (22) 6 19<br />

Group 2 + 76.2% (<strong>21</strong>) .0000009 18 5 0.140 26 9 0.732<br />

- 4.2% (24) 27 33<br />

Group 3 + 54.5% (22) 0.882 16 13 0.334 20 <strong>21</strong> 0.603<br />

- 52.3% (<strong>21</strong>) 11 9 35 23<br />

CONCLUSIONS: TLM revealed that non-biopsied embryos have difficulty<br />

completing zona hatching, indicating that zona hardening may play a role. In<br />

the absence of LZB, implantation may be impeded due to inability to escape<br />

from the ZP. In contrast, previously biopsied blastocysts did not show a difference<br />

in hatching rates whether subjected to LZB post warming or not.<br />

Reference:<br />

1. Vaccari & Conaghan, Fert & Ster Volume 103, Issue 2, Suppl, Pages<br />

e7-e8.<br />

P-479 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

ELECTIVE EMBRYO CRYOPRESERVATION AFTER PRO-<br />

LONGED PREMATURE PROGESTERONE ELEVATION TO<br />

IMPROVE IN VITRO FERTILIZATION (IVF) PREGNANCY<br />

OUTCOMES. E. Anspach, B. Maslow, A. Bartolucci, C. Benadiva,<br />

J. Nulsen, L. Engmann. University of Connecticut Health Center, Farmington,<br />

CT.<br />

OBJECTIVE: Studies demonstrate that prolonged premature elevations in<br />

progesterone (P) levels during controlled ovarian stimulation have adverse<br />

effects on IVF outcomes, potentially due to alterations in endometrial receptivity.<br />

Cryopreservation with a subsequent frozen embryo transfer (FET) has<br />

been proposed as a method of overcoming embryo-endometrium asynchrony.<br />

The purpose of this study was to evaluate whether elective cryopreservation<br />

(‘‘freeze all’’) improves pregnancy outcomes in patients with prolonged premature<br />

P elevation during an IVF cycle.<br />

DESIGN: Retrospective cohort<br />

MATERIALS AND METHODS: 732 patients


Baseline characteristics and outcome.<br />

Variable<br />

Group 1<br />

Fresh Transfer,<br />

Elevated P<br />

n¼233<br />

Group2<br />

FET<br />

n¼43<br />

Group 3<br />

Fresh<br />

Transfer,<br />

Normal P<br />

n¼456<br />

P<br />

Value<br />

Age (SD) 34.0 3.6 32.9 4.0 33.6 3.8 0.16<br />

AMH (SD) 3.4 3.9 3.9 5.6 3.6 3.9 0.62<br />

Day of Trigger P (SD) 1.4 0.3 1.9 0.8 0.7 0.2


P-482 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

128 AUTOLOGOUS POST-VITRIFICATION OOCYTE THAW CY-<br />

CLES: WARMING, EMBRYONIC DEVELOPMENT, AND PREG-<br />

NANCY OUTCOMES. J. Doyle, K. S. Richter, J. Lim, J. Graham,<br />

M. J. Tucker. Shady Grove Fertility Reproductive Science Center, Rockville,<br />

MD.<br />

OBJECTIVE: To evaluate a single treatment center’s experience with<br />

autologous oocyte vitrification and warming, including embryonic development<br />

and pregnancy outcomes.<br />

DESIGN: Retrospective review<br />

MATERIALS AND METHODS: Patient and cycle characteristics and<br />

treatment outcomes were compared according to indication for autologous<br />

oocyte vitrification. All MII oocytes that survived warming<br />

were fertilized via ICSI. Outcomes were also compared to a reference<br />

group of all fresh autologous ICSI cycles in 2013 (n¼2,963) using<br />

generalized estimating equations analysis to control for repeat<br />

cycles and adjust for age, blastocyst vs cleavage stage transfer, and<br />

numbers of embryos transferred. Clinical pregnancy was defined as<br />

ultrasound confirmation of an intrauterine gestational sac; implantation<br />

rate was calculated as the number of gestational sacs per embryo<br />

transferred.<br />

RESULTS: Through January <strong>2015</strong>, 128 oocyte warming cycles were<br />

conducted following oocyte vitrification for the indications of elective<br />

fertility preservation, unavailability of sperm at oocyte retrieval, or<br />

desire for limited oocyte insemination, which compare as shown in the<br />

table (within rows, values with different superscripts are significantly<br />

different by post hoc Tukey-Kramer HSD, p


micronized progesterone 400 mg b.d. was started after 12 days when the<br />

GnRh agonist was also stopped. Frozen-thawed embryos were transferred<br />

on day+1 of their chronological age and when the endometrium reached<br />

12 mm in thickness. Group B consisted of 100 patients (100 cycles)<br />

who started daily estradiol valerate 6 mg administration from the second<br />

day of the FET cycle and followed the same regimen but without prior<br />

treatment with triptorelin<br />

RESULTS: There was a significant increase in implantation rate in the<br />

GnRh agonist group (group A) compared to the estrogen and progesterone<br />

only group (group B) (44.1% vs <strong>21</strong>.1%; P¼ 0.002*). The pregnancy rate<br />

was also significantly higher in group A compared to group B (65.5% vs<br />

42 %, P¼ 0.013*).<br />

CONCLUSIONS: GnRh agonist administration during endometrial preparation<br />

for FET increases the implantation rate and pregnancy rate.<br />

CONCLUSIONS: The sperm Vitrification and conventional freezing have<br />

negative impact on sperm parameters as well as sperm DNA oxidation, and<br />

mitochondrial activity. Sperm Vitrification presented improved sperm<br />

motility recovery, similar levels of DNA oxidation and a slightly increase<br />

in mitochondrial activity when compared with sperm cryopreservation<br />

with the conventional method. Suggesting Vas an optimal protocol for sperm<br />

cryopreservation.<br />

P-486 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

WITHDRAWN<br />

Effect of GnRh agonist on implantation of FET.<br />

GnRh group A Group B<br />

P<br />

Total number of cases 110 100<br />

Total number of ET 322 285<br />

Total number of implanted embryos 142 60<br />

Implantation rate 44.1 % <strong>21</strong>.1% 0.002*<br />

Number of pregnant cases 72 42<br />

Pregnancy rate 65.5% 42% 0.013*<br />

* Statistically significant.<br />

P-485 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

EFFECT OF HUMAN SPERM FREEZING-THAWING PROCESS ON<br />

VITRIFICATION AND CONVENTIONAL FREEZING: EVALUA-<br />

TION OF SURVIVAL, MOTILITY, DNA OXIDATION AND MITO-<br />

CHONDRIAL ACTIVITY. D. Pabon, a M. Meseguer, a G. Sevillano, a<br />

A. Cobo, a J. Romero, a N. Sota, b A. Mifsud, a J. de los Santos, a<br />

J. Remohi, c M. de los Santos. a a Clinical Embryology-IVI, Valencia, Spain;<br />

b Clinical Andrology-IVI, Valencia, Spain; c Reproduction-IVI, Valencia,<br />

Spain.<br />

OBJECTIVE: Sperm Vitrification is a method for cryopreservation,<br />

without the use of conventional cryoprotectants, by plunging the sperm suspension<br />

directly into liquid nitrogen. We developed and evaluated a new<br />

method of vitrification, using a combination of 0,25 M trehalose-sucrose.<br />

This study aimed to compare the vitrification (V) with conventional freezing<br />

protocol (CF) on several sperm parameters: motility, vitality, mitocondryal<br />

activity and DNA oxidation.<br />

DESIGN: A prospective non-randomized study to evaluate spermatozoa<br />

vitrification with respect to the conventional freezing protocol, using fresh<br />

sperm as reference.<br />

MATERIALS AND METHODS: The sperm vitrification solution was<br />

0,25 M trehalose-sucrose and plunge directly in liquid nitrogen in microdroplets<br />

of 5-10ml. Sperm mobility and viability rates were calculated<br />

before and after cryopreservation. Mitochondrial function was evaluated<br />

using JC-1 (fluorescent cationic dye, 5,5’,6,6’-tetrachloro-1-1’,3,3’-tetraethyl-benzamidazolocarbocyanin<br />

iodide). Sperm DNA oxidation was<br />

determined using a fluorescent assay (Oxy-DNA test) for the detection of<br />

8-oxoguanine. The evaluation was carried out before and after cryopreservation<br />

using flow cytometry analysis. Statistical analysis was performed using<br />

ANOVA test and Chi square, p


TUNEL<br />

PATIENTS<br />

TUNEL<br />

DONORS<br />

ANNEXIN V<br />

PATIENTS<br />

ANNEXIN<br />

V DONORS<br />

BODIPY<br />

PATIENTS<br />

BODIPY<br />

DONORS<br />

8OH-dG<br />

PATIENTS<br />

FRESH-BASAL 20.112.8 14.74.2 8.93.2 6.72.8 12.44.8 9.63.1 65.313.6<br />

POST THAWED 36.613.6 25.75.1 18.35.1 11.52.7 27.18.5 <strong>17</strong>.53.3 112.6 22.1<br />

after thawed had higher values in compare to swim-up (fresh). Vitality<br />

decreased after cryopreservation, contrary the control group samples presented<br />

more tolerance to cryoinjury. It was determined that these parameters<br />

significantly increase in men >45 y/o. All the comparisons performed on<br />

these analysis were statistically significant (page 35 at time of treatment.<br />

When successfully combined, these technologies afford postponement of<br />

Total No. MI + MII 559/674 (83%) 238/292 (82%) 0.6<br />

Oocytes Survived<br />

/Thawed<br />

Survived and<br />

534 (79% ) 225 (77%) 0.4<br />

Assessed as MII at<br />

Thaw<br />

2-Pronuclear (2PN) 373 (70%) 167 (74%) 0.3<br />

Fertilization<br />

No. Cycles Achieving 39/47 (83%) 15/22 (68%) 0.2<br />

Blastocysts<br />

Suitable for TEBX<br />

No. Biopsied BL 145 (39%) 54 (32%) 0.2<br />

No. Euploid BL 58 (40% ) 8 (15%) 0.0007<br />

No. Aneuploid BL 86 (60%) 44 (85%) 0.004<br />

No. TEBX Cycles 28/39 (72%) 7/15 (47%) 0.1<br />

with At Least One<br />

Euploid BL<br />

No. TEBX Cycles 11/39 (28%) 8/15 (53%) 0.1<br />

with All Aneuploid<br />

BL<br />

Embryo Implantation 14/22 (64%) 2/3 (67%) 1.0<br />

Rate<br />

Ongoing+Delivered<br />

Rate/Transfer<br />

12/<strong>21</strong> (57%) 2/3 (67%) 1.0<br />

FERTILITY & STERILITY Ò<br />

e273


P-490 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

INCREASED INCIDENCE OF MOSAICISM AMONG BIOPSIED<br />

TROPHECTODERM CELLS ANALYZED BY NEXT GENERATION<br />

SEQUENCING. M. Smith, a D. Johnson, a D. L. Hill, a M. Surrey, b<br />

S. Ghadir, b W. Chang, b H. Danzer, b C. Alexander, b S. Munne, c<br />

J. Barritt. a a ART Reproductive Center, Beverly Hills, CA; b Southern California<br />

Reproductive Center, Beverly Hills, CA; c Reprogenetics, Livingston, NJ.<br />

OBJECTIVE: Preimplantation genetic screening (PGS) using Array CGH,<br />

qPCR or SNP arrays provided very high levels of diagnostic accuracy, however,<br />

very little ability to recognize low-level mosaicism. The newest Next<br />

Generation Sequencing (NGS) technique provides the same level of diagnostic<br />

accuracy and allows for an increased chance to detect chromosomal<br />

mosaics, down to as low as 20%. We set out to determine the presence, quantity<br />

and the types of genetic mosaicism in trophectoderm cells following embryo<br />

biopsy in clinical PGS cases using NGS analysis.<br />

DESIGN: Retrospective data analysis from a single, large, private fertility<br />

clinic.<br />

MATERIALS AND METHODS: Sixteen blastocyst biopsy PGS cycles<br />

using NGS were analyzed. No ovum donor cycles were included. The<br />

mean female age was 36. Trophectoderm biopsy was performed on 128 blastocysts<br />

cultured 5, 6 and/or 7 days. All embryos were vitrified and the biopsied<br />

cells were sent to a genetic testing laboratory (Reprogenetics, Los<br />

Angeles, CA). Results were compiled, including the genetic diagnosis rate,<br />

the quantity of mosaics, the types of mosaics, and the number of occurrences<br />

for each chromosome involved in the mosaicism. Fisher’s Exact Test was<br />

used for statistical analysis.<br />

RESULTS: Genetic results were obtained in 124 of 128 (96.9%) blastocysts,<br />

with 1 failure to amplify DNA and 3 chaotic profiles not allowing diagnoses.<br />

Mosaicism was reported in <strong>17</strong> (13.7%) samples, as compared with<br />

our previous experience with aCGH of 6728 blastocysts reporting only 10<br />

(0.14%) mosaics (p


P-493 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

CLINICAL ERROR RATE OF ARRAY COMPARATIVE<br />

GENOMIC HYBRIDIZATION (ACGH) IN EUPLOID<br />

BLASTOCYSTS. A. W. Tiegs, B. Hodes-Wertz, D. H. McCulloh,<br />

J. Grifo. NYU Langone Medical Center, New York, NY.<br />

OBJECTIVE: Preimplantation genetic screening (PGS) and diagnosis<br />

(PGD) with euploid embryo transfer is associated with improved implantation<br />

and live birth rates as compared to routine in vitro fertilization. However,<br />

misdiagnosis of the embryo is a potential risk. The purpose of this study was<br />

to investigate the clinical misdiagnosis rate associated with transfer of trophectoderm-biopsied<br />

blastocysts deemed to be euploid via array comparative<br />

genomic hybridization (aCGH) at a large university-based fertility center.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: All cycles utilizing PGS or PGD with<br />

trophectoderm biopsy, aCGH and euploid embryo transfer at a large university-based<br />

fertility center with known birth outcomes from 11/2010 through<br />

7/2014 were included (n¼525). The number of embryos transferred, gestational<br />

sacs, clinical pregnancies with a fetal heartbeat (FH), and cycle outcomes<br />

(specifically live birth (LB) and spontaneous abortion (SAB)) were<br />

recorded. Any products of conception (POCs) that were collected from dilation<br />

and curettage of a SAB were reviewed.<br />

RESULTS: There were 525 embryo transfers of 584 euploid embryos as designated<br />

by aCGH. Implantation rate was 61%, clinical pregnancy rate with an FH<br />

was 59%, and the first trimester SAB rate was 13% of pregnancies with a sac<br />

(6% of pregnancies with an FH). There was one intrauterine fetal demise and<br />

one periviable loss. Five clinical misdiagnoses were identified during this time.<br />

Clinical aCGH errors.<br />

Embryo aCGH<br />

result Karyotype Outcome Error<br />

46,XX 47,XX +13 SAB Biologic-Mosaic<br />

46,XX 47,XX +7 SAB Test Error<br />

46,XY 47,XY +<strong>21</strong> SAB Biologic-Mosaic<br />

46,XY 47,XYY LB Test Error<br />

46,XX 46,XY LB Likely Contamination<br />

Error<br />

All aCGH specimens were analyzed again with aCGH and 2/5 were found<br />

to be mosaic embryos. Error rate per embryo transfer cycle was 1.0%, 0.9%<br />

per embryo transferred, 1.5% per pregnancy with a sac and 0.4% per euploid<br />

embryo diagnosed by aCGH. The LB error rate was 0.7% (both sex chromosome<br />

errors) and the SAB POCs error rate was <strong>17</strong>.6% (3/<strong>17</strong> POCs tested). Of<br />

note, single gene disorders sequenced against using PGD were not mistakenly<br />

selected for in any known cases.<br />

CONCLUSIONS: Although aCGH has been shown to be a superior<br />

method of comprehensive chromosome screening, several possible sources<br />

of error still exist. These include contamination of the specimen and sampling<br />

errors when mosaicism exists. Although the overall error rate is low<br />

(60,000 BLASTO-<br />

CYSTS TESTED VIA ARRAY CGH. E. Yeboah, a S. Munne, a<br />

T. Escudero, a C. Wagner Coughlin, b M. Surrey, c S. Ghadir, c R. P. Marrs, d<br />

J. Zhang, e G. Garzo. f a Reprogenetics, Livingston, NJ; b Aparent IVF Laboratory,<br />

Highland Park, IL; c Southern California Reproductive Center, Beverly<br />

Hills, CA; d California Fertility Partners, Los Angeles, CA; e New Hope<br />

Fertility Center, New York, NY; f Reproductive Partners - UCSD, La Jolla, CA.<br />

OBJECTIVE: To assess the results of blastocysts via aCGH by age and<br />

type of abnormality<br />

DESIGN: Retrospective Analysis<br />

MATERIALS AND METHODS: 60,492 blastocyst biopsies were<br />

analyzed from 7/24/2011 to 4/<strong>21</strong>/<strong>2015</strong> from 12,275 cycles of preimplantation<br />

genetic screening (PGS) using blastocyst biopsy and aCGH. Subsequently,<br />

they were categorized by SART age groups to examine the percentage of embryos<br />

that were euploid, monosomic, trisomic, had partial chromosome gains<br />

or losses, double aneuploidies, orcomplex abnormal (more than two abnormalities),<br />

using the PGS specific eIVF database (PracticeHwy, Dallas, TX).<br />

RESULTS: The proportion of full monosomies to trisomies were 50:50,<br />

however, for partial abnormalities it was skewed 69:31<br />

CONCLUSIONS: As it is well known, chromosome abnormalities<br />

increased with advanced maternal age, from 31% in egg donors to 85% in<br />

women older than 42. The complexity of the abnormalities also increased<br />

with advancing maternal age, with double and multiple abnormalities<br />

increasing and single full aneuploidies and single partial aneuploidies<br />

decreasing. Partial aneuploidies, that is, the presence or absence of a fragment<br />

of a chromosome, had a ratio of partial monosomies to partial trisomies that<br />

was significantly skewed, 69:31, while full aneuploidies had a 50:50 ratio of<br />

monosomies to trisomies. In addition, the total rate of partial abnormalities<br />

(alone or in combination with other abnormalities) does not change with<br />

advancing maternal. These two very different tendencies indicate different origins<br />

of partial and full aneuploidies, with full aneuploidies being mostly<br />

meiotic in origin, while partials probably mitotic (non-age related) or paternal<br />

in origin with the resulting chromosome fragments being lost more easily than<br />

gained.<br />

The results are shown in the below table:<br />

EGD < 35 35-37 38-40 41-42 > 42<br />

Embryos Analyzed 12832 15103 10302 11530 5744 3090<br />

No Results 3% 4% 3% 3% 3% 4%<br />

Euploid 69% 63% 55% 39% 23% 15%<br />

Abnormal 31% 37% 45% 61% 77% 85%<br />

1 Aneuploidy 52% 56% 54% 50% 39% 26%<br />

2 Aneuploidies <strong>17</strong>% 16% 22% 25% 29% 26%<br />

>2 Abnormalities 16% 16% 16% 20% 30% 47%<br />

Partial aneuploidy 15% 12% 8% 4% 2% 1%<br />

(alone)<br />

Partial total (as % of<br />

abnormal)<br />

15% 13.6% 16.3% 12% 15% 22%<br />

FERTILITY & STERILITY Ò<br />

e275


P-496 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DOES BLASTOCYST MORPHOLOGY PREDICT PREGNANCY<br />

OUTCOMES IN EUPLOID SINGLE EMBRYO<br />

TRANSFERS? M. Shah, a J. Kort, b R. Lathi. c a Stanford University,<br />

Palo Alto, CA; b Stanford University, Menlo Park, CA; c Stanford University,<br />

Stanford, CA.<br />

OBJECTIVE: Comprehensive chromosome screening (CCS) is a powerful<br />

tool which allows for selection of euploid embryos to improve pregnancy<br />

outcomes in IVF cycles. There is limited data suggesting a correlation between<br />

blastocyst morphology of euploid embryos and pregnancy outcomes.<br />

The objective of this study was to determine whether blastocyst morphology<br />

could be used to predict implantation potential of a euploid embryo.<br />

DESIGN: We performed a retrospective observational study performed<br />

between January 2012 and December 2014 at the Stanford University<br />

Fertility and Reproductive Health Center. The study included fresh and<br />

frozen embryo transfer cycles of single euploid blastocysts.<br />

MATERIALS AND METHODS: Trophectoderm biopsy was performed<br />

for CCS analysis on day 5 blastocysts. Morphology, as defined by the Gardner<br />

criteria, was assessed on the day of biopsy for frozen IVF cycles, and on<br />

the day of transfer for fresh IVF cycles. The ongoing pregnancy rate (OPR)<br />

was defined as presence of fetal cardiac activity beyond 20 weeks of gestation.<br />

Chi squared analysis was used to test the relationship between blastocyst<br />

morphology and OPR.<br />

RESULTS: There were 133 single euploid embryo transfers included for<br />

analysis from 125 unique patients. The mean patient age was 37.3 years<br />

old and similar among patients with and without an ongoing pregnancy.<br />

Fifty-five percent of these cycles employed a frozen-thaw embryo transfer,<br />

while 45% were fresh embryo transfers. Forty-percent of euploid embryo<br />

transfers resulted in an ongoing pregnancy or live birth. Chi-squared analyses<br />

revealed that blastocysts with a greater expansion grade, and better inner cell<br />

mass and trophectoderm quality grades were more likely to result in an<br />

ongoing pregnancy or live birth, with the trend reaching significance for<br />

the association with trophectoderm quality (Table I).<br />

CONCLUSIONS: All components of the Gardner grading scale of blastocyst<br />

assessment may help to select the embryo most likely to become a successful<br />

pregnancy, even after assisted hatching and pre-implantation genetic<br />

screening for euploid embryos.<br />

Ongoing pregnancy/live birth rate (OPR) stratified by components of the<br />

Gardner grading scale after.<br />

Expansion grade 4 5 6 p-value<br />

OPR 3% 22% 75% 0.06<br />

Inner Cell Mass grade A B C<br />

OPR 59% 41% 1% 0.09<br />

Trophectoderm grade A B C<br />

59% 35% 5% 0.04<br />

P-497 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

SHOULD EGG DONORS BE SCREENED FOR X-LINKED<br />

DISEASE? A. Hershlag, a O. A. Tusheva, a C. Mullin, a Y. Danovitch, b<br />

T. Singer, a S. Munne, c N. Kumar. c a The Center for Human Reproduction,<br />

North Shore LIJ Health System, Manhasset, NY; b Sackler School of Medicine,<br />

New York, NY; c Recombine, New York, NY.<br />

OBJECTIVE: Egg donors with negative family history of X-linked disorders<br />

may still be carriers and thus put male offspring at risk. Carrier screening<br />

for such disorders, with the exception of Fragile X, is currently limited. We<br />

here present a case where an undiagnosed carrier state in the donor led to an<br />

affected male child, review the literature, and propose a novel approach to<br />

prevent transmission of X-linked disorders through egg donation.<br />

DESIGN: A case report and literature review<br />

MATERIALS AND METHODS: Case: An egg donation from a patient’s<br />

sister led to the birth of a male child who experienced excessive bleeding at<br />

circumcision and was diagnosed with Hemophilia A. The donor had a de<br />

novo mutation. Second egg donor cycle using PGD for Hemophilia A was<br />

performed and resulted in a birth of healthy unaffected girl.<br />

RESULTS: Previously reported births of affected males have uncovered<br />

unsuspected mutations for X-linked diseases in egg donors (hemophilia; adrenal<br />

hypoplasia congenita; ocular albinism). While donors are screened for<br />

the most common autosomal recessive disorders such as CF, SMA<br />

(1:2500;1:6000 LB), they are not screened for the most common X-linkeddiseases,<br />

such as Hemophilia A, Duchenne Muscular Dystrophy<br />

(DMD)(1:5000;1:3300 male LB). Gender selection for females can reduce<br />

the risk of symptoms but has ethical issues. Other approaches include molecular<br />

testing (such as targeted mutation analysis or sequencing) in the donor to<br />

determine carrier status and/or in the embryo to determine disease status. In<br />

the case of Hemophilia A, with over 615 point mutations, 57 insertions, 270<br />

deletions ranging from a single nucleotide to over 150 kb, and complex inversions,<br />

testing techniques are cumbersome, technically challenging and<br />

applicable only to a small number of families with a particular mutation.<br />

Similarly, in the case of other X-linked disorders such as DMD, screening<br />

is challenging due to deletions accounting for a significant proportion of disease,<br />

large numbers of ‘‘private’’ mutations, and high de novo mutation rates.<br />

Preimplantation genetic haplotyping (PGH), originally described in 2006,<br />

combines a haplotyping approach and whole genome amplification, and<br />

may be a robust, efficient and successful alternative method to detect X-<br />

linked disorders in embryos.<br />

CONCLUSIONS: It is incumbent upon us to do everything within the<br />

reach of reproductive genetics to prevent X-linked diseases in male offspring<br />

from egg donation, as well as from all ART in general. We need to incorporate<br />

screening for X-linked disorders as part of regular carrier screening practice,<br />

and overcome technical challenges to ensure we are effectively able to<br />

test for the wide heterogeneity of mutations that cause X-linked disorders.<br />

Preimplantation genetic haplotyping (PGH) may be a promising method<br />

for X-linked disease screening in embryos.<br />

P-498 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

CCS IMPROVES PREGNANCY OUTCOMES IN EGG DONOR FET<br />

CYCLES. A. Coates, a C. Welch, b K. Ketterson, b S. Munne. b a Oregon<br />

Reproductive Medicine, Portland, OR; b Reprogenetics, Livingston, NJ.<br />

OBJECTIVE: Donor egg recipients are offered CCS as part of their IVF<br />

cycle to optimize embryo selection. Aneuploidy rates of blastocyst embryos<br />

from Egg donor cycles are around 25% and this alone may justify selection<br />

by CCS. The purpose of this study is to determine if CCS can improve treatment<br />

outcome in frozen /thawed egg donor cycles.<br />

DESIGN: Retrospective data and sample analysis.<br />

MATERIALS AND METHODS: Blastocyst embryos were biopsied on<br />

day 5 or 6 and the cells were analyzed by array CGH (Illumina) or Next Generation<br />

Sequencing (NGS, Life Technologies). All embryos were vitrified for<br />

future use post biopsy. Egg donor cycles of frozen / thawed untested embryos<br />

were used as a control group.<br />

RESULTS: Results are shown in the table below:<br />

CONCLUSIONS: 79% of frozen embryo transfer cycles resulted in<br />

ongoing pregnancies after transfer of CCS selected embryos compared to<br />

59% of transfers without selection. This significant difference justifies the extra<br />

cost of the biopsy and testing procedure (compared to the cost of a subsequent<br />

egg donor cycle) and helps reduce the risk of aneuploid<br />

conceptions, which is still a significant concern after compensating for<br />

advanced maternal age by using young donor eggs.<br />

Frozen donor egg embryo transfers with and without CCS tested embryos.<br />

CCS tested FET<br />

Non tested<br />

FET<br />

P value<br />

Average age(range) 25(<strong>21</strong>-27) 25(22-28)<br />

# Frozen transfers 166 61<br />

Total # embryos<br />

246(1.5) 92(1.5)<br />

transferred(ave/FET)<br />

# positive tests (%/FET) 159*(96%) 52*(85%) *p¼0.02<br />

# ongoing pregnancies 131*(79%) 36*(59%) *p¼0.006<br />

(%/FET)<br />

# embryos implanting (%<br />

+ sac/embryo)<br />

201*(82%) 54*(59%) *P¼0.0001<br />

e276 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


P-499 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

TROPHECTODERM BLASTOCYST (TE) BIOPSY RESULTS IN<br />

HIGHER PREGNANCY RATE, LOWER MULTIPLE RATE, AND<br />

FEWER CYCLES TO PREGNANCY IN GESTATIONAL<br />

CARRIERS. A. Nasseri, a K. Gleason, a C. Zollinger, a M. Post, a<br />

D. Chen, a J. Grifo. b a The Valley Hospital Fertility Center, paramus, NJ;<br />

b The New York University Fertility Center, New York, NY.<br />

OBJECTIVE: To determine the usefulness of comprehensive chromosome<br />

screeniong (CCS) by TE biopsy in gestational surrogacy<br />

DESIGN: Retrospective case control study<br />

MATERIALS AND METHODS: Data from all cycles involving nondonor<br />

egg embryo transfers to gestational carriers (GC) at our center from<br />

April 2008 to January <strong>2015</strong> were reviewed. FDA screening guidelines<br />

were adhered to in all cases. Embryos of the intended parents (IP) were<br />

created as per routine IVF protocol. There were a total of 88 embryo transfer<br />

cycles. Twenty eight cycles involved the transfer of frozen eupolid blastocysts<br />

after TE biopsy. Embryos were screened for all 24 chromosomes by<br />

array CGH (group A). Of the remaining 60 cycles (group B), there were<br />

35 fresh, and 25 frozen transfers. One fresh transfer was canceled due to embryo<br />

arrest. Paired student t-test and the Chi-squared test of independence<br />

were used for statistical analysis. P value


CONCLUSIONS: Aneuploidy screening was the most common indication<br />

for PGD. Use of PGD was not associated with increased rates of pregnancy<br />

and live birth for women 37 years, thereby suggesting<br />

that more viable or euploid embryos were transferred after<br />

screening.<br />

P-502 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

WITHDRAWN<br />

P-503 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

ELEVATED SERUM PROGESTERONE REDUCES THE LIKELI-<br />

HOOD OF LIVE BIRTH AFTER FROZEN TRANSFER OF EUPLOID<br />

EMBRYOS: TOO MUCH OF A GOOD THING? L. A. Kondapalli, a<br />

L. Munkwitz, a R. Collins, a D. A. Minjarez, b W. B. Schoolcraft, a<br />

M. Katz-Jaffe. a a Colorado Center for Reproductive Medicine, Lone Tree,<br />

CO; b Colorado Center for Reproductive Medicine, Denver, CO.<br />

OBJECTIVE: Although comprehensive chromosomal screening (CCS) allows<br />

for selection of euploid embryos with high implantation and live birth<br />

rates, IVF failures still occur. The objective of this study was to investigate<br />

factors associated with negative pregnancy after frozen transfer.<br />

DESIGN: Case control study<br />

MATERIALS AND METHODS: Patients undergoing euploid frozen embryo<br />

transfers (FET) between 2013-2014 were included (n¼441). All embryos<br />

were biopsied at the blastocyst stage for CCS prior to vitrification.<br />

Standard protocols were used for hormone supplementation prior to FET.<br />

Patients were stratified by cycle outcome: successful live birth versus<br />

non-pregnant (including biochemical pregnancy). Data on demographics,<br />

ovarian reserve markers, IVF stimulation, and cycle outcomes were obtained<br />

from medical records. Appropriate parametric and non- parametric<br />

statistical analyses were performed using STATA version 13.1 (Stata<br />

Corp; College Station, TX, USA) with P value


from a second biopsy for each embryo. Phase 2 involved clinical utilization<br />

of combined SGD and CCS testing with follow-up.<br />

RESULTS: Workup time approximated 1 month for each case. Phase 1<br />

testing examined 152 embryos and demonstrated 99% concordance with<br />

reference lab data with all discrepancies confirmed as an error with the reference<br />

labs results. Phase 2 involved clinical application of these methods in 43<br />

patients (304 embryos). A definitive result was reported for 99.7% (303/304)<br />

embryos, with 0.3% (1/304) having an inconclusive result likely due to<br />

recombination. All cases for which a sample was available in a newborn child<br />

confirmed the PGD result. In patients receiving a transfer with follow-up,<br />

clinical outcomes were excellent with an implantation rate of 75% (12/16),<br />

with 92% of patients achieving pregnancy.<br />

CONCLUSIONS: This methodology has demonstrated excellent concordance<br />

with current methods of SGD and provides a unique opportunity to<br />

avoid pitfalls of WGA when targeting additional loci in the genome in parallel<br />

with CCS. Additional advantages of the method include the ability to<br />

manage microdeletion and duplication cases and the potential for a 4 hour<br />

turnaround time. In addition, it is well established that SNP markers are<br />

denser than STRs thus generally reducing the distance of markers from mutations<br />

and the potential impact of recombination.<br />

P-506 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

THE CLINICAL OUTCOME OF PREIMPLANTATION GENETIC<br />

DIAGNOSIS (PGD) CYCLES UNDER PRIVATELY FUNDED AND<br />

PUBLICLY FUNDED PERIODS IN ONE CENTER. A. Ao, a<br />

X. Zhang, b L. Zhang, b W. Buckett. b a Obstetrics and Gynecology, McGill<br />

Univerisity, Montreal, QC, Canada; b McGill University, Montreal, QC,<br />

Canada.<br />

OBJECTIVE: To compare the overall clinical outcome of IVF-PGD cycles<br />

during privately funded period and publicly funded period in one center.<br />

DESIGN: A single center retrospective study.<br />

MATERIALS AND METHODS: All couples that underwent IVF-PGD<br />

cycles within specific study periods were divided under two groups: Privately<br />

funded period (PRP) where the IVF and PGD treatment cycles were privately<br />

funded and publicly funded period (PUP), where the treatment cycles were<br />

funded by the provincial government healthcare system. Under Privately<br />

funded period, 298 cycles (195 PGS and 42 translocation and 61 SGD)<br />

and under publicly funded period, 2<strong>21</strong> cycles (59 PGS and 72 translocation<br />

and 90 SGD) underwent IVF-PGD procedure. The outcomes were assessed<br />

for pregnancy rate and live birth rate.<br />

RESULTS: The average age of female patients for both private and<br />

publicly funded periods were similar (35 vs 36 years old). The fertilization<br />

rate (73% vs 72%), embryo biopsy rate (80% vs 80%) and genetic<br />

diagnosis rate (89% vs 94%) were not significantly different between<br />

these two groups. More embryos were transferred per cycle during<br />

PRP than PUP (2.34 vs 1.12, p


RESULTS: The diagnostic rates were 98.7% for Karyomapping and<br />

98.6% for CCS. The average numbers of embryos analyzed and eligible<br />

for transfer per cycle were 5.3 and 2.5 for PGD alone, and 6.2 and 2.3 for<br />

PGD+CCS. Batching of embryos prior to analysis was carried out for 11%<br />

of PGD+CCS cases (2% for PGD alone).The CCS improved clinical<br />

outcome. The implantation and clinical pregnancy rates were 70.5% (43/<br />

61) and 64% (34/53) for PGD+CCS cycles, versus 54.5% (6/11) and<br />

55.6% (5/9) for PGD alone. Most transfers involved a single thawed blastocyst,<br />

except for ten patients (16%) who had two embryos transferred at once.<br />

The average maternal ages were 32.3 and 34.2 years for PGD and PGD+CCS,<br />

respectively. There are 13 live births and 36 ongoing pregnancies to date. Prenatal<br />

and perinatal testing was performed in four cases confirming PGD in all<br />

instances.<br />

CONCLUSIONS: The high clinical pregnancy rates were similar to outcomes<br />

achieved by the most successful in-vitro fertilization (IVF) programs.<br />

As lengthy patient-specific test developments were unnecessary, greater<br />

numbers of patients were treated with faster time to diagnosis, transfer and<br />

pregnancy. SNP arrays thus allow PGD laboratories to deal with the rapidly<br />

growing demand for embryo testing for inherited disorders. The application<br />

of Karyomapping with simultaneous CCS speeds up the testing process and<br />

further improves clinical outcomes.<br />

P-509 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

BLASTOCYST MORPHOLOGY AND CHROMOSOME STATUS:<br />

CORRELATION IN WOMEN < AGE 35. C. F. Boylan,<br />

L. S. Morrison, S. M. Carney, G. Kovalevsky, A. B. Neithardt,<br />

R. F. Feinberg. Reproductive Associates of Delaware, Newark, DE.<br />

OBJECTIVE: To determine if day 5 and 6 blastocyst morphology correlate<br />

with ploidy status in women < 35.<br />

DESIGN: A retrospective study that compiled morphology and ploidy status<br />

from day 5 and 6 blastocysts created from 131 women under age 35.<br />

MATERIALS AND METHODS: All blastocysts were cultured in vitro to<br />

day 5 or 6 of development. When stage appropriate, trophectoderm biopsy<br />

was performed. Approximately 2-10 cells were removed by laser excision<br />

and sent to a PGS/PGD laboratory. The inner cell mass (ICM) and trophectoderm<br />

cells of each embryo were graded prior to biopsy according to Gardner’s<br />

grading scale. Each embryologist in the laboratory was extensively<br />

trained on the scale to lessen subjectivity. Blastocysts were grouped into<br />

‘‘AA’’, ‘‘BB’’ or ‘‘CC’’ according to their overall quality. They were then<br />

grouped exclusively by their ICM and trophectoderm. Chi Square analysis<br />

was performed to test significance between the grade and ploidy status.<br />

RESULTS:<br />

Overall Grade Euploid Aneuploid<br />

‘‘AA’’ Blastocysts 81/99 (81.8%) 1 18/99 (18.2%)<br />

‘‘BB’’ Blastocysts 306/454 (67.4%) 1 148/454 (32.6%)<br />

‘‘CC’’ Blastocysts 34/79 (43.0%) 1 45/79 (57.0%)<br />

ICM Grade<br />

‘‘A’’ ICM 123/158 (77.8%) 2 35/158 (22.2%)<br />

‘‘B’’ ICM 382/582 (65.6%) 2 200/582 (34.4%)<br />

‘‘C’’ ICM 52/111 (46.8%) 2 59/111 (53.2%)<br />

Trophectoderm Grade<br />

‘‘A’’ Trophectoderm 94/115 (81.7%) 3 <strong>21</strong>/115 (18.3%)<br />

‘‘B’’ Trophectoderm 365/543 (65.6%) 3 <strong>17</strong>8/543 (34.4%)<br />

‘‘C’’ Trophectoderm 98/193 (50.8%) 3 95/193 (49.2%)<br />

1,2,3 p-value¼


well characterized. The present study sought to develop a working model of<br />

mosaicism through the use of aneuploid cell lines and to evaluate detection<br />

limits of an NGS based CCS methodology.<br />

DESIGN: Observational<br />

MATERIALS AND METHODS: Cell lines with single chromosomal aneuploidies<br />

(+15 or +18) were mixed with a total of 6 cells in various ratios<br />

(16%, 33%, 50%, 66%, 83% and 100%) with euploid cells in order to model<br />

a range of mosaicism levels that might be observed in a typical trophectoderm<br />

biopsy. NGS based CCS was performed using a previously validated<br />

custom amplification protocol and an Ion Proton. Statistical outlier analysis<br />

of chromosome specific copy number assignments, with values above the 3<br />

interquartile range (IQR) being identified as outliers, was used to identify<br />

mosaic chromosomes.<br />

RESULTS: Examination of mosaicism with artificial 6-cell mixtures revealed<br />

that NGS could detect mosaicism at 16%, was relatively reliable at<br />

50%, and showed high specificity throughout each mixture (Table 1).<br />

CONCLUSIONS: This is study represents the most extensive evaluation<br />

of CCS sensitivity to mosaicism reported to date. Most importantly, this<br />

study defined the limits of detection of aneuploidy mosaicism within a 6-<br />

cell sample, and can be used to set evidence-based expectations when counseling<br />

patients. Future work will involve evaluating whether detection<br />

criteria established in this model may be predictive of actual clinical outcomes<br />

when applied to trophectoderm biopsies.<br />

TABLE 1. Results of NGS CCS detection of mosaicism in a model system.<br />

Mosaicism Spike-in 16% 33% 50% 66% 83% 100%<br />

(n) 12 12 12 12 12 12<br />

Sensitivity 8% 33% 67% 75% 75% 100%<br />

Specificity 100% 91% 91% 100% 91% 100%<br />

P-512 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

PUSHING THE LIMITS OF PREIMPLANTATION GENETIC DIAG-<br />

NOSIS (PGD): A KARYOMAPPING AND NEXT GENERATION<br />

SEQUENCING (NGS) STRESS TEST. M. R. Lindeman,<br />

E. Czuprenski, J. Kitchen, T. K. McWilliams, M. Wyatt, T. T. Gordon,<br />

M. Hughes. Genesis Genetics, Plymouth, MI.<br />

OBJECTIVE: We sought to examine the limits of embryo genetic analysis<br />

by performing PGD and aneuploidy screening (PGS) in 1) a consanguineous<br />

mating for, 2) two disease-causing point mutations and 3) HLA-genotyping<br />

to match 3 affected children while 4) detecting genetic recombination and<br />

potential maternal cell contamination, and 5) 24-chromosome aneuploidy<br />

screening with NGS.<br />

DESIGN: Karyomapping utilizes 300,000 Single Nucleotide Polymorphisms<br />

(SNPs) across the human genome and allows high-resolution analysis<br />

for nearly every gene. We used karyomapping to rapidly design a PGD test<br />

targeting the HBB gene and the highly recombinant HLA locus to identify<br />

embryos with one or both healthy HBB alleles and an HLA genotype matching<br />

the children in need of transplantation.<br />

MATERIALS AND METHODS: Cheek-swab kits were used to collect<br />

DNA samples from the parents and affected children. Mutation status of<br />

each individual was confirmed by direct sequencing. DNA samples were<br />

analyzed using the Infinium HumanKaryomap-12 DNA Analysis Kit from Illumina.<br />

BluFuse Multi software was used to compare the SNP profiles of<br />

each family member to identify informative SNPs, regions of consanguinity,<br />

and determine HLA phase for each affected child. NGS was used for 24-chromosome<br />

aneuploidy screening.<br />

RESULTS: We identified sufficient informative SNPs surrounding the HBB<br />

gene and HLA region to track the inheritance of familial haplotypes. As expected,<br />

the same HBB haplotypes were observed in each affected child. Two children<br />

were found to have the same set of HLA haplotypes, while the third inherited<br />

the opposite paternal HLA haplotype. This is consistent with the HLA-typing reported<br />

by an external laboratory. Recombination and consanguineous regions<br />

were observed but did not impact our ability to determine genetic phase.<br />

CONCLUSIONS: We successfully used karyomapping to build a complex<br />

PGD test, which would have been highly challenging using previous methods.<br />

Moreover, we determined the HLA phases of all three children simultaneously<br />

and confirmed that recombination and consanguinity would not interfere<br />

with diagnosis or HLA matching. Our success addresses concerns about<br />

using karyomapping for HLA matching and diagnosing embryos of consanguineous<br />

families. Karyomapping combined with PGS by Next-Generation<br />

Sequencing represents a great advance to the field, 25 years after its inception.<br />

P-513 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

REPRODUCTIVE OUTCOME FOLLOWING PRE-IMPLANTATION<br />

GENETIC DIAGNOSIS (PGD) - AN ANALYSIS OF UK NATIONAL<br />

DATABASE OVER TWO DECADES. A. N. Sharpe a M. Choudhary. b<br />

a Newcastle Fertility Centre at Life, Newcastle-upon-Tyne, United Kingdom;<br />

b Consultant Gynaecologist & Sub Specialist in Reproduction, Newcastle<br />

upon Tyne, United Kingdom.<br />

OBJECTIVE: In 2013, as a drive to ensure an equal and consistent policy,<br />

the National Health Service Commissioning board centralised the funding in<br />

England for up to 3 PGD cycles for couples who have or are carriers of a specific<br />

genetic disorder and wish to avoid having an affected child. This study<br />

aims to determine the UK incidence of PGD over the last 20 years and associated<br />

clinical outcome.<br />

DESIGN: Retrospective national database cohort.<br />

MATERIALS AND METHODS: Data from 1991 to 2012 was analysed<br />

from the Human Fertilisation Embryology Authority (HFEA) database to<br />

determine the incidence and clinical outcomes of couples who had PGD. Binary<br />

logistic regression was used to compare PGD cycles versus non-PGD<br />

cycles for age matched women over the years.<br />

RESULTS: Of a total 1,071,040 cycles of assisted reproduction techniques<br />

(ART) from 1991 to 2012, 2974 had PGD (0.28%). The incidence has risen<br />

127 fold from 1/15953 cycles in 1991 to 530/66553 cycles in 2012.There was<br />

2443 PGD cycles from 1999 to 2011, where 585 resulted in live birth (23.9%)<br />

compared to 137588 live births from 659622 non-PGD cycles (20.9%,<br />

p¼0.15). The live birth rate increased over time and was greater in women<br />

aged 18-34 years for both PGD and non-PGD cycles (p


P-514 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

LASER-ASSISTED HATCHING IN EARLY BLAST IS OPTIMAL<br />

FOR TROPHECTODERM BIOPSY OF MOUSE<br />

EMBRYOS. Y. Chiang, D. Russell, M. Rosario, S. Wang. Obetetric<br />

and Gynecology, University of Colorado Denver, Aurora, CO.<br />

OBJECTIVE: To determine which strategy of laser-assisted hatching is<br />

optimal for trophectoderm biopsy in mouse embryos<br />

DESIGN: Prospective in vitro study<br />

MATERIALS AND METHODS: Totally 280 mouse embryos were randomized<br />

to laser-assisted hatching in different cleavage stages, including<br />

89 embryos without laser-assisted hatching, 76 embryos with laser-assisted<br />

hatching in 8 cell stage, 46 embryos with laser-assisted hatching in morula<br />

stage and 69 embryos with laser-assisted hatching in early blast stage. The<br />

morphology of embryo in trophectoderm biopsy, biopsy time and atretic<br />

rate in 24 and 48 hours after biopsy were analyzed.<br />

RESULTS: Significant higher percentages of hatching in trophectoderm<br />

was noted in embryos with laser-assisted hatching in early blast stage than<br />

embryos without laser-assisted hatching, embryos with laser-assisted hatching<br />

in 8 cell and morula stages (94.2% versus 18.0%, 43.4%, 32.6%, Chisquare<br />

test, p


CONCLUSIONS: Our data suggest that mosaicism may be present in the trophectoderm,<br />

at a rate of approximately 20.0%. More importantly, although mosaicism<br />

is present, this phenomena only resulted in a clinically significant rate of<br />

approximately 2.0% (euploid vs. aneuploid). Therefore, mosaicism may be present<br />

in the human blastocyst; however at a low rate of clinical significance.<br />

P-5<strong>17</strong> Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

ANEUPLOIDY PARENT OF ORIGIN IN BLASTOCYST BIOPSIES<br />

USING KARYOMAPPING TECHNOLOGY. K. McWilliams,<br />

T. K. McWilliams, M. Wyatt, M. Hughes. Genesis Genetics, Plymouth, MI.<br />

OBJECTIVE: Characterization of parental derivation of whole chromosome<br />

aneuploidies by the single nucleotide polymorphism (SNP)-based karyomapping<br />

platform.<br />

DESIGN: Retrospective, multi-faceted analysis of blastocyst biopsies submitted<br />

for single-gene preimplantation genetic diagnosis (PGD) and chromosome<br />

screening (PGS) analysis.<br />

MATERIALS AND METHODS: 90 blastocyst biopsies from 54 families<br />

were tested by PGD and PGS by comparative genomic hybridization microarray<br />

(aCGH) or Next Generation Sequencing. Maternal ages ranged from<br />

27-39 (mean of 32). Paternal ages ranged from 27-54 (mean of 36). Aneuploidies<br />

detected by PGS were analyzed by Illumina’s karyomapping platform,<br />

which contains 300,000 genome-wide SNPs on a BeadChip array.<br />

Embryos with segmental duplications/deletions or greater than 2 whole chromosome<br />

aneuploidies were excluded from analysis. BlueFuse Multi software<br />

data analysis revealed familial haploblock inheritance. Monosomies showed<br />

lack of inheritance from one parent; whereas trisomies demonstrated inheritance<br />

of 3 parental haploblocks. Data was collected on the basis of aneuploidy<br />

type, maternal vs. paternal inheritance, and parental age.<br />

RESULTS: To date, 49.6% of embryo biopsies studied by karyomapping<br />

and PGS combined testing have been aneuploid. Ninety-eight aneuploid<br />

events were studied by karyomapping. The majority (85.7%) of all detectable<br />

whole chromosome aneuplodies were maternally-derived (p¼35 years) and embryo euploidy, and suggest<br />

that variation in maternal genes may explain differences in aneuploidy incidence.<br />

24-chromosome SNP-based PGS helps infertility patients achieve<br />

high rates of implantation, and viable, healthy pregnancies.<br />

Supported by: This work was partly Supported by National Institutes of<br />

Health grants R01 GM100366, R01 GM097415, and R01 GM089926.<br />

P-520 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DESCRIPTION OF EUPLOID EMBRYO IMPLANTATION<br />

OUTCOME BY MORPHOKINETIC INVESTIGATION. A. Tejera, a<br />

M. Stoppa, b M. Meseguer, a A. Capalbo, b M. Florensa, c F. Ubaldi, b<br />

Y. Galiana, a L. Rienzi. d a IVI Valencia, Valencia, Spain; b Genera, Rome,<br />

Italy; c IVF Lab Director, Barcelona, Spain; d Genera Centres for Reproductive<br />

Medicine, Rome, Italy.<br />

FERTILITY & STERILITY Ò<br />

e283


OBJECTIVE: To define euploid implanted morphokinetic parameters and<br />

propose an algorithm for embryo selection based on this analysis.<br />

DESIGN: Retrospective observational study.<br />

MATERIALS AND METHODS: Embryo development from 265 patients<br />

undergoing pre-implantation genetic screening, euploid transferred embryos<br />

(n¼342) were analyzed with time-lapse imaging (Embryoscope, Vitrolife,<br />

Denmark). Chromosomal analysis was performed through array-comparative<br />

genome hybridization and qPCR. Timing of each cellular event was annotated<br />

along embryo development (tn) in hours post ICSI, also we<br />

calculated cell cycle duration (ccn¼tn+1-tn) and synchrony between blastomeres<br />

(sn¼tn+1-tn). In addition, timing of compaction (tM), blastocyst formation<br />

(tB) and expansion (tEB) were recorded. The data to define<br />

morphokinetic behaviour was obtained from a total of transferred embryos,<br />

and embryos with known implantation data (KID). Only KID embryos<br />

with either full implantation (number of sacs matches the number of transferred<br />

embryos n¼130) or no implantation (no biochemical pregnancy<br />

achieved n¼169) were included in this analysis.<br />

RESULTS: A logistic regression analysis identified tM (


In this study we found out that patients older than 36 years old with AMH<br />

level less than 1 ng/ml, have significantly lower number of blastocyst suitable<br />

for biopsy, compared to the patients with AMH level great than 1 ng/ml<br />

(0.270.20 vs. 2.470.32). Patient in the lower AMH group also has significantly<br />

higher BMI compared to the group with higher AMH group<br />

(28.311.44 vs. 24.850.63).<br />

P-523 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

USING 24-CHROMOSOME ANEUPLOIDY TEST AND FROZEN<br />

BLASTOCYST TRANSFER CYCLES SIGNIFICANTLY IMPROVES<br />

PREGNANCY RATE AND DECREASES MISCARRIAGE RATE IN<br />

IVF PROGRAMS. K. Kirienko, a A. Strashnova, a E. Uvarova, a<br />

N. Voronich, a Z. Zlatopolsky, b O. Verlinsky, b T. Pakhalchuk, b<br />

V. Apryshko, a S. Yakovenko. a a Altravita IVF Clinic, Moscow, Russian<br />

Federation; b Reproductive Genetic Institute, Chicago, IL.<br />

OBJECTIVE: The purpose of this study was to compare clinical outcomes<br />

(pregnancy and miscarriage rates) between frozen blastocyst transfer (FBT)<br />

cycles without preimplantation genetic screening (PGS), FBT cycles with<br />

24-chromosome aneuploidy test and FBT cycles with 9-chromosome<br />

screening for infertile patients with normal karyotype aged 25-45.<br />

DESIGN: Retrospective comparative study.<br />

MATERIALS AND METHODS: Embryos were cultured to the blastocyst<br />

stage and vitrified shortly (Cryotech, Japan) after trophectoderm (TE)<br />

mechanical biopsy. Mechanically biopsied cells were analyzed by using 24-Chromosome<br />

Microarray - aCGH or by QF-PCR with the focus on 9-chromosomes<br />

(13,14,15,16,18,<strong>21</strong>,22,X,Y). A total of 503 couples were involved in the study.<br />

The experimental group included 203 couples (average female age is 37), 1022<br />

blastocyst stage (BS) embryos were obtained from this group. Of these embryos,<br />

489 were analyzed by aCGH and 533 were analyzed by QS-PCR. The control<br />

group included 300 couples (average female age is 33) undergoing IVF program<br />

without PGS. Ongoing pregnancies after single embryo transfer (SET) were<br />

confirmed by one fetal sac and heartbeat. Data was analyzed using Student’s t-test.<br />

RESULTS: Results are shown in the table. In IVF with 24-chromosome<br />

aneuploidy test group the risk of aneuploidies was strongly age-dependent<br />

(p39<br />

SET cancellation in %39<br />

connection with >39<br />

abscence of<br />

PGS-normal<br />

embryos, %<br />

Clinical pregnancy %39<br />

rate/SET, % >39<br />

Miscarriage rate/ %39<br />

pregnancy, %<br />

The average cost of<br />

the program,<br />

USD (% of IVF)<br />

>39<br />

%39<br />

>39<br />

IVF<br />

-<br />

-<br />

-<br />

-<br />

35.0<br />

20.0<br />

34.8<br />

57.5<br />

65<strong>21</strong><br />

6911<br />

IVF+PGS 9-<br />

chromosome<br />

75.6<br />

72.5<br />

6.4<br />

10.4<br />

35.5<br />

16.0<br />

22.0<br />

43.0<br />

7911 (1<strong>21</strong>%)<br />

7184 (104%)<br />

IVF+24-<br />

chromosome<br />

aneuploidy test<br />

50.5<br />

24.5<br />

11.8<br />

45.0<br />

58.7<br />

26.0<br />

11.4<br />

16.7<br />

8147 (125%)<br />

8562 (124%)<br />

OBJECTIVE: To identify if pregnancy outcomes differ in patients that<br />

make multiple euploid embryos from a single in vitro fertilization (IVF) cycle<br />

in comparison to patients who make only one.<br />

DESIGN: Retrospective Cohort Study<br />

MATERIALS AND METHODS: Global retrospective IRB was obtained.<br />

All patients who underwent IVF with preimplantation genetic screening<br />

(PGS) with array comparative genomic hybridization (aCGH), with at least<br />

one euploid embryo were included. The first single euploid frozen embryo<br />

transfer cycle following the IVF cycle was included. All causes of infertility,<br />

all indications for PGS, autologous and donor cycles were included. Patients<br />

who elected to undergo two or more euploid embryo transfer cycles were<br />

excluded. The primary outcome of the study included pregnancy (presence<br />

of a gestational sac) and clinical pregnancy (presence of fetal heart beat<br />

through cycle day 63 or a live birth). Statistical analysis was done using AN-<br />

OVA, t-test and a linear regression model.<br />

RESULTS: A total of 327 single euploid embryo transfers were included.<br />

The average age of patients included in the study was 37.37 4.90. The overall<br />

pregnancy rate was 68.50%(224/327), of which 61.47% (201/327) led to<br />

ongoing clinical pregnancies or live births. The most common number of<br />

euploid embryos per cohort was one (n¼108) with the highest number being<br />

FERTILITY & STERILITY Ò<br />

e285


20 (n¼1). There was no difference in pregnancy rates when embryos from<br />

cycles with multiple euploid embryos were compared to cycles with only<br />

one euploid embryo (RR 0.9965, CI 0.8872-1.119, p 0.9457). Linear regression<br />

showed no difference was found in pregnancy rates when the percentage<br />

of euploid embryos per cohort was compared as well (R ¼0.0905). There was<br />

a significant difference in pregnancy rates in embryos of higher-grade inner<br />

cell mass, however when controlled for embryo grades no difference was<br />

noted in pregnancy rates in lower numbered euploid embryo cohorts.<br />

Pregnancy Outcomes by Number of Euploid Embryos per IVF cycle.<br />

Number<br />

of Euploid<br />

Tota<br />

Patients<br />

(n ¼ 327)<br />

Average<br />

Number of<br />

Eggs<br />

Retrieved<br />

Total<br />

Pregnant<br />

(presence of<br />

gestational<br />

sac)<br />

Percent<br />

Total<br />

Clinical<br />

Pregnancies<br />

(Fetal heart<br />

beat or live<br />

birth)<br />

Percent<br />

1 108 12.32 +/- 5.75 70 65 65 60<br />

2 84 15.60 +/- 7.23 62 74 55 65<br />

3 42 15.79 +/- 6.91 29 69 25 60<br />

4 30 19.30 +/- 7.63 22 73 <strong>21</strong> 70<br />

5 16 22.56 +/- 9.92 10 63 10 63<br />

6 13 22.38 +/- 8.91 9 69 7 54<br />

7 13 27.50 +/- 5.09 6 46 5 38<br />

8 or more <strong>21</strong> 30.0 +/- 10.78 16 76 13 62<br />

p value p 0.552 p 0.552 p 0.676<br />

CONCLUSIONS: The number or percentage of euploid embryos per IVF<br />

cycle, is not associated with the pregnancy outcome of STEET.<br />

P-526 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

FIRST VALIDATED METHOD FOR DISTINGUISHING NORMAL<br />

FROM BALANCED TRANSLOCATION CARRIER<br />

EMBRYOS. N. R. Treff, a K. Thompson, a M. Rafizadeh, a X. Tao, b<br />

H. Garnsey, b C. V. Reda, a T. Metzgar, a E. Forman, a R. T. Scott. a a RMA,<br />

NJ, NJ; b FAEEC, Basking Ridge, NJ.<br />

OBJECTIVE: Carriers of balanced translocations often wish to prevent inheritance<br />

in their offspring in order for their children to eventually avoid the<br />

same reproductive challenges. Although identifying unbalanced embryos has<br />

been possible for many years, current PGS methods are unable to reliably<br />

distinguish a truly normal embryo from one that carries a balanced translocation.<br />

The present study was conducted to validate a method that provides<br />

the opportunity to make this distinction for the first time.<br />

DESIGN: Prospective blinded.<br />

MATERIALS AND METHODS: A series of translocation carrier couples<br />

that underwent IVF with SNP array based PGD were included. Embryos that<br />

were transferred in these cases could have been balanced or normal. SNP<br />

array analysis was performed on parental DNA and unbalanced embryos<br />

from each case to define informative SNPs whose alleles were linked with<br />

the derivative chromosomes (phasing). The transferred sibling embryo array<br />

data were then evaluated at the informative SNP loci to predict whether they<br />

were truly normal (did not inherit the derivative chromosomes) or translocation<br />

carriers. IRB approval was obtained to perform conventional karyotyping<br />

on the newborns in order to establish the true genetic status of the original<br />

transferred embryo (balanced or normal). Embryonic SNP array predictions<br />

were then compared to the newborn karyotypes obtained.<br />

RESULTS: Phasing SNPs using unbalanced embryos allowed accurate<br />

prediction of whether transferred embryos were balanced translocation carriers<br />

or truly normal in all 10 cases completed to date (100% concordance<br />

with conventional karyotyping of newborns).<br />

CONCLUSIONS: This study demonstrates the validity of the first method<br />

capable of distinguishing normal from balanced translocation carrier embryos.<br />

The only prerequisite is the availability of parental DNA and unbalanced<br />

IVF embryos, making the method applicable to the majority of<br />

carrier couples. In addition, the SNP array platform allows simultaneous<br />

evaluation of comprehensive chromosome screening for aneuploidy, in parallel,<br />

from the same biopsy. Future work will involve prospective predictions<br />

to select normal embryos with subsequent karyotyping of the resulting newborns.<br />

P-527 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

EMBRYOS FROM FEMALE CARRIERS OF BALANCED TRANS-<br />

LOCATIONS DO NOT HAVE A HIGHER RISK OF WHOLE-CHRO-<br />

MOSOME ANEUPLOIDY THAN THEIR AGE-RELATED<br />

RISK. E. J. Forman, a N. R. Treff, b J. M. Franasiak, c R. T. Scott. b<br />

a RMA, NJ, NJ; b RMANJ, Rutgers-RWJ, Basking Ridge, NJ; c RMANJ, NJ,<br />

NJ.<br />

OBJECTIVE: Given the altered chromosomal configuration of tetrads in<br />

the synaptonemal complexes of translocation carriers during meiosis, there<br />

has been concern that these couples may harbor increased risk for aneuploidy<br />

in chromosomes not directly affected by the translocation. The purpose of<br />

this study is to determine whether this interchromosomal effect (ICE) in<br />

translocation carriers increases the risk of aneuploidy beyond their agerelated<br />

risk when compared to a large data set of embryos tested with<br />

comprehensive chromosome screening (CCS) at the blastocyst stage.<br />

DESIGN: Case control.<br />

MATERIALS AND METHODS: The baseline aneuploidy rate per chromosome<br />

was calculated for all patients undergoing IVF with CCS and stratified<br />

into


Frequency distribution analysis was performed. Log-regression analysis was<br />

performed for the pre-treatment variables.<br />

RESULTS: 105 patients met inclusion criteria. Treatment indications<br />

included isolated HG in 7 patients and combined infertility and HG in 98 patients.<br />

22 patients (<strong>21</strong>.0%) had inadequate androgenization after treatment<br />

with CC, of which 13 patients paradoxically experienced a decrease in androgenization.<br />

Pre- and post-treatment characteristics and univariate analyses<br />

are reported in Table 1. The logistic regression model was statistically significant<br />

22 patients (<strong>21</strong>.0%) had inadequate androgenization after treatment<br />

with CC, of which 13 patients paradoxically experienced a decrease in androgenization.<br />

Pre-treatment characteristics and univariate analyses are reported<br />

in Table 1. The logistic regression model was statistically significant (c2(9)¼<br />

45.12, p¼


11. Clark R, Leonessa F, Welch JN, Skaar TC ‘‘Cellular and Molecular<br />

Pharmacology of Antiestrogen Action and Resistance’’ Pharmacol<br />

Rev 53: 2571, 2001<br />

12. DrugBank http://www.drugbank.ca/drugs/DB00675<br />

13. Edwin JR, John PS, ‘‘Structural Features of Alkylphenolic Assoicated<br />

with Estrogenic Activity’’ (1997) The Journal of Biological Chemistry,<br />

Vol 272, No.6, 32803288<br />

14. Jordan VC, Collins MM, Rosby L, and Prestwich G, (1977) J. Endocrinol.<br />

75, 305316<br />

15. Santa Cruz Biotechnology<br />

16. Jin et al. 2005, ‘‘CYP2D6 genotype, antidepressant use, and tamoxifen<br />

metabolism during adjuvant breast cancer treatment’’ J Natl Cancer<br />

Inst 97 (1): 3039.<br />

<strong>17</strong>. Regulation for Registration of Medicinal Products, Ministry of Health<br />

and Welfare in Taiwn ; article 7.1 and 14.1, 05/07/2014 http://law.-<br />

moj.gov.tw/Eng/LawClass/LawAll.aspx?PCode¼L0030057<br />

18. Bonde JP, Ernst E, Jensen TK, et al . (1998) Relation between semen<br />

quality and fertility: a populationbased study of 430 firstpregnancy<br />

planners. Lancet, 352 , 1<strong>17</strong><strong>21</strong><strong>17</strong>7<br />

19. SwedishAmerican Hospital, 1401 East State Street, Rockford, Illiois<br />

61104, USA<br />

20. VITROS Immunodiagnostic Products: REF 135 0198 and REF 109 0133<br />

<strong>21</strong>. VITROS Immunodiagnostic Products: REF 193 1922 and REF 190<br />

1263<br />

22. BECKMAN COULTER Access Immunoassay System: REF 33560<br />

23. Actavis, Morris Corporate Center III, 400 Interpace Parkway, Parsippany,<br />

NJ 07054<br />

24. Teva Pharmaceutical Industries LTD, 5 Basel Street, Petach Tikva<br />

49131, Israel<br />

25. The FDA’s Drug Review Process: Ensuring Drugs Are Safe and Effective<br />

http://www.fda.gov/drugs/resourcesforyou/consumers/ucm143534.htm<br />

26. Zitzmann M, Nieschlag E. ‘‘Hormone substitution in male hypogonadism’’<br />

Mol Cell Endocrinol. 2000;161(12): 7388<br />

27. Han TS, Bouloux PM. ‘‘What is the optimal therapy for young males<br />

with hypogonadotropic hypogonadism?’’ Clin Endocrinol (Oxf).<br />

2010;72(6):73<strong>17</strong><br />

28. K. Singh, D. Jaiswal. ‘‘Human male infertility: a complex multifactorial<br />

phenotype’’ Reprod Sci, 18 (2011), pp. 418-425<br />

29. N. Pemmaraju , M. F. Munsell , G. N. Hortobagyi, and S. H. Giordano,<br />

‘‘Retrospective review of male breast cancer patients: analysis of tamoxifenrelated<br />

sideeffects.’’ Ann Oncol. 2012 Jun; 23(6): 1471-1474.<br />

P-531 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

ROLE OF DIETARY ANTIOXIDANT SUPPLEMENTATION IN<br />

TREATMENT OF IDIOPATHIC MALE INFERTILITY: PROMISING<br />

EVIDENCE FROM A SUB-CONTINENTAL STUDY. S. K. Goswami, a<br />

S. Yasmin, b P. Chakraborty, c R. Chattopadhyay, b S. Ghosh, b M. Goswami, b<br />

B. Ghosh, d B. Chakravarty. b a Reproductive Medicine, Consultant, Kolkata,<br />

India; b Assisted Reproduction, Institute of Reproductive Medicine, Kolkata,<br />

India; c Infertility, Institute of Reproductive Medicine, Kolkata, India; d Andrology,<br />

Institute of Reproductive Medicine, Kolkata, India.<br />

OBJECTIVE: To evaluate the ameliorating potential of a diet rich in antioxidants<br />

in couples with idiopathic male infertility with high reactive oxygen<br />

species in comparison to a standard combined antioxidant formulation.<br />

DESIGN: Prospective observational study from March 2013 to April<br />

<strong>2015</strong>.<br />

MATERIALS AND METHODS: <strong>17</strong>5 patients with idiopathic male infertility<br />

(IMI) with high reactive oxygen species (ROS) were recruited, of which<br />

80 patients were advised to take a diet rich in antioxidants along with few<br />

changes in lifestyle (group A) and remaining 95 were prescribed a combined<br />

oral antioxidant therapy (group B). A placebo-controlled group comprising<br />

75 patients was maintained in parallel. Semen parameters, antioxidant concentrations<br />

(CoQ-10, L-carnitine, zinc) with biomarkers of oxidative stress<br />

including plasma total antioxidant capacity (TAC), total glutathione (GSH)<br />

were analyzed before and after the treatment. Mitochondrial membrane potential<br />

(Djm), ROS and sperm DNA fragmentation were evaluated by flowcytometry<br />

and TUNEL assay respectively. MnSOD and PRDX5 expressions<br />

were determined by real time PCR to evaluate the increased antioxidant effect.<br />

Study was approved by Institutional Review Board. Statistical comparisons<br />

were performed using Stata10.0.<br />

RESULTS: Increased concentrations of plasma TAC and total GSH with a<br />

significant increment in sperm motility (p


chemotherapy agents and the interval from chemotherapy were also<br />

obtained and analyzed to detect any associations with the success in sperm<br />

retrieval.<br />

RESULTS: In a cohort of 66 NOA patients after chemotherapy, sperm<br />

were successfully retrieved in 31 patients, with clinical pregnancy occurring<br />

in 23 couples. Therefore, per-patient SRR was 47% and per-patient pregnancy<br />

rate was 35%. Per-patient live birth rate was 27% (18/66). Our patients<br />

had received chemotherapy for testicular cancer (<strong>21</strong> patients), Hodgkin lymphoma<br />

(9 patients), non-Hodgkin lymphoma (7 patients), acute lymphoblastic<br />

leukemia (9 patients), acute myeloblastic leukemia (7 patients),<br />

rhabdomyosarcoma (7 patients), bladder cancer (3 patients), osteosarcoma<br />

(2 patients), and anaplastic anemia (1 patient). Table shows detailed information<br />

of each cancer type. A multivariate analysis showed no significant predictors<br />

of micro-TESE outcome.<br />

CONCLUSIONS: Micro-TESE is the only method for fertility revival in<br />

chemotherapy induced permanent NOA patients. However, our results suggest<br />

micro-TESE and following intracytoplasmic sperm injection could<br />

rescue only 27% of NOA patients after chemotherapy. No predictive values<br />

of sperm retrieval were determined in this analysis. Sperm banking should be<br />

offered before any chemotherapy even if the possibility of permanent azoospermia<br />

is thought to be low.<br />

P-533 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

SUPPLEMENTATION OF CRYOBUFFER WITH CATALASE AND<br />

N-ACETYL CYSTEINE IMPROVES HUMAN SPERM POST-THAW<br />

MOTILITY, VIABILITY AND DNA INTEGRITY. Y. Kobori,<br />

M. Kathrins, C. Niederberger, G. S. Prins. University of Illinois at Chicago,<br />

Chicago, IL.<br />

OBJECTIVE: Cryopreservation of human sperm is a routine procedure<br />

in assisted reproductive technology. The cryopreservation process can<br />

lead to structural and functional alterations in spermatozoa, impairing<br />

fertilization potential. Reactive oxygen species has been suggested as a<br />

major contributing factor for cryodamage to spermatozoa. Accordingly,<br />

antioxidant supplementation has been used to yield significantly<br />

improved quality of frozen sperm post-thaw. We sought to investigate<br />

improved outcomes with a combination of antioxidants Catalase and<br />

N-Acetyl Cysteine (NAC), which function as intracellular and extracellular<br />

antioxidants, respectively.<br />

DESIGN: Laboratory investigation<br />

MATERIALS AND METHODS: SpermFreeze (Vitrolife) with 15% glycerol<br />

was used as cryobuffer supplemented with 0.0mM (control), 2.5mM,<br />

5mM, and 10mM NAC with/without 200U/ml Catalase. Semen samples<br />

were collected from normospermic men (n¼20) and aliquots frozen in<br />

each dose of antioxidant tested. Post-thaw semen analysis by CASAwas performed<br />

at 30 min and 2 hr. In addition, sperm viability (Eosin-Y staining) and<br />

sperm DNA fragmentation (TUNEL assay) were quantitated at 2 hr postthaw.<br />

RESULTS: A significant increase in total motility, progressive motility<br />

and sperm viability was observed in all samples frozen with antioxidants<br />

as compared to control. The cryobuffer containing 5mM NAC achieved the<br />

highest recovered motility post-thaw among NAC doses. At 30 minutes<br />

post-thaw, recovered motility (post thaw motility /pre-freeze motility <br />

100%) in control, 200U/ml Catalase, 5 mM NAC and combined Catalase/<br />

NAC supplemented buffer of sperm was 40.3%, 44.8%, 46.2%, and 47.8%,<br />

respectively. Similar increased post-thaw progressive motility, viability<br />

was observed with best results observed using the NAC/Catalase combination.<br />

Preliminary results with DNA integrity analysis for double-stranded<br />

breaks using TUNEL show similar trends and complete analysis will be presented.<br />

CONCLUSIONS: Combined Catalase and N-Acetyl Cysteine supplementation<br />

during cryopreservation results in improved post-thaw recovery of total<br />

motility, progressive motility, percentage of sperm vitality and improved<br />

DNA integrity. The results indicate that the combination of intracellular and<br />

extracellular antioxidants results in the most pronounced effect in improving<br />

post-thaw quality of human spermatozoa.<br />

P-534 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

IDENTIFICATION OF SERTOLI CELL MARKERS IN MEN WITH<br />

NON-OBSTRUCTIVE AZOOSPERMIA. A. Agarwal, a R. Sharma, a<br />

Z. Cui, a,b E. S. Sabanegh. c a Center for Reproductive Medicine, Cleveland<br />

Clinic, Cleveland, OH; b Institute of Toxicology, Third Military Medical University,<br />

Chongqing, China; c Urology, Cleveland Clinic, Cleveland, OH.<br />

OBJECTIVE: Dysfunctional spermatogenesis is presumed to be the main<br />

cause of non-obstructive azoospermia (NOA). At present, serum follicle<br />

stimulating hormone (FSH) levels and testicular volume are used to diagnose<br />

NOA, however their sensitivity and specificity are limited as predictors of<br />

testicular sperm extraction outcomes. The development of a reliable, noninvasive<br />

biomarker that can predict NOA, especially in patients with normal<br />

FSH, is the ultimate aim. We utilized proteomic analysis to identify proteins<br />

that are differentially expressed in the seminal plasma of NOA patients with<br />

normal FSH and NOA patients with increased FSH to help our understanding<br />

of the etiology and mechanism of FSH in normal NOA. This might provide<br />

new markers that can help in the diagnosis of NOA patients and better predict<br />

the outcome of testicular biopsy.<br />

DESIGN: Proteomic analysis of seminal plasma proteins from NOA men<br />

with normal and elevated FSH levels in comparison with fertile men.<br />

MATERIALS AND METHODS: Seminal plasma was obtained from 10<br />

NOA patients with a high FSH level, and 4 NOA patients with a normal<br />

FSH level and 7 fertile men. Short and long gels were run. Bands were cut,<br />

and the gels (1mm 3 pieces) were trypsinized and run on a LTQ-Orbitrap<br />

Elite hybrid mass spectrometer system. Functional annotations of proteins<br />

were obtained using bioinformatic tools and pathway databases. Western<br />

Blotting was performed to verify the expression levels of 3 proteins of interest.<br />

RESULTS: We identified 493 proteins in the seminal plasma of fertile controls,<br />

448 in NOAwith normal FSH and 436 in NOAwith high FSH levels. Of<br />

these, 68 were differentially expressed proteins (DEPs) in NOA and normal<br />

FSH and 15 between NOA and high FSH. DEPs that were overexpressed in<br />

NOA and normal FSH levels were involved in glycosylation, cell-adhesion,<br />

cell proliferation regulation, and secretion. DEPs that were underexpressed<br />

were involved in stress response, glycosylation, secretion, spermatogenesis<br />

and sterol transport. DEPs that were overexpressed in NOA with normal<br />

and elevated FSH levels were involved in the processes of glycosylation,<br />

secretion, defense response, cellular homeostasis, response to hormone stimulus,<br />

and response to hypoxia. Proteins that were underexpressed were<br />

involved in the processes of secretion, cell adhesion, cell skeleton, ECM-receptor<br />

interaction, and oxidoreductase activity. We identified 4 proteins that<br />

were related to Sertoli function such as secretion, Sertoli-germ cell junction<br />

and hormone receptor function. These were lactotransferrin, desmoglein,<br />

desmoplakin, and junction plakoglobin.<br />

CONCLUSIONS: We have identified proteins that may serve as markers<br />

of non-obstructive azoospermia. They participate in important Sertoli cell -<br />

related functions and may play an important role in identifying patients<br />

with NOA who are likely to have spermatozoa.<br />

P-535 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DIAGNOSTIC ACCURACY OF PHYSICAL EXAMINATION<br />

COMPARED WITH COLOR DOPPLER ULTRASOUND IN THE<br />

DETERMINATION OF TESTIS SIZE AND VARICOCELE<br />

DIAGNOSIS. B. C. Tiseo, a M. Cocuzza, a V. Srougi, a G. J. Wood, a<br />

S. Esteves, b M. Srougi. a a Division of Urology, Hospital das Clinicas, University<br />

of Sao Paulo Medical School, Sao Paulo, Brazil; b ANDROFERT, Andrology<br />

and Human Reproduction Clinic, Campinas, Brazil.<br />

OBJECTIVE: To evaluate the accuracy of physical examination on the<br />

diagnosis and grading of varicocele and on testicular volume measurement<br />

comparing to ultrasonographic assessment. Varicocele treatment should be<br />

offered to patient with palpable varicocele, provided it is associated with<br />

infertility. Also, varicocele should be repaired in patients with testicular<br />

growth impairment. Therefore, physical examination is the most important<br />

parameter to deselect men with subclinical varicocele or testicular size alteration<br />

to undergo treatment, as shown by the equivocal beneficial effect of<br />

such interventions in such cases.<br />

DESIGN: Prospective cohort.<br />

MATERIALS AND METHODS: We consecutively evaluated 78 patients<br />

attending at the University-based Urology Unit. All patients were examined<br />

for presence of varicocele and testis size estimation by a group of two<br />

non-experienced urologists and two experienced infertility specialists.<br />

Anthropometric data was acquired for each patient. All of them were submitted<br />

to Color Doppler Ultrasonography (CDU) to evaluate the presence of<br />

varicocele and measure testicular volume.<br />

RESULTS: There were 74 patients who completed the protocol with CDU<br />

totalizing 147 testicular units evaluated. Mean testicular volume measured at<br />

CDU was 13.6 5.6. Varicocele was diagnosed in 61.2% of the units by CDU.<br />

Non-experienced urologists diagnosed varicocele in 47.6% and 48.9% and<br />

false positive rate was 25.7% and 31.9%, while experienced urologists had<br />

47.6% and 35.3% varicocele prevalence and false positive rate of 24.3%<br />

FERTILITY & STERILITY Ò<br />

e289


and 11.5%. The ROC curve analysis of physical examination for varicocele<br />

presence revealed area under the curve of 0.570 and 0.631 within non-experienced<br />

urologists and 0.645 and 0.703 within specialists.Regarding to volume<br />

estimation, all examiners did not have correctly assessed testicular<br />

volume with its difference form CDU measurement ranging from 1.30 to<br />

2.41cc (p


P-539 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

OPTIMAL TIME TO BANK SPERM FROM PATIENTS WITH<br />

TESTICULAR TUMORS. M. Ribeiro de Andrade, P. Ferigolo,<br />

M. P. Antoniassi, M. Camargo, P. Intasqui, D. S. Tibaldi, R. Bertolla,<br />

D. M. Spaine. Departament of Surgery, Division of Urology, Sao Paulo Federal<br />

University, Sao Paulo, Brazil.<br />

OBJECTIVE: While testicular tumors present high survival rates, post-orchiectomy<br />

adjuvant treatment often utilizes gonadotoxic agents, which renders<br />

sperm banking of great importance in these patients. We wished to verify<br />

if sperm banking should be performed before or after orchiectomy (before<br />

adjuvant therapy) in patients with testicular tumors.<br />

DESIGN: Prospective Study.<br />

MATERIALS AND METHODS: This prospective study was carried out<br />

including <strong>17</strong> patients with testicular germ cell tumors, who provided one<br />

semen sample before they were submitted to orchiectomy of the affected<br />

testis, and one other semen sample 30 days after the surgery. An aliquot of<br />

each sample was used to seminal analysis and other to determinate sperm<br />

function. Sperm DNA fragmentation was assessed by an alkaline Comet<br />

assay, in which sperm were classified as high DNA integrity (class I) to<br />

high DNA fragmentation (class IV). Acrosome integrity was determined<br />

by a fluorescent probe (PNA-FITC) and assessed by fluorescent microscopy.<br />

Mitochondrial activity was determined by a colorimetric stain (3,3’-diaminobenzidine<br />

- DAB) and classified as DAB I (all mitochondria active) to DAB<br />

IV (all mitochondria inactive). Seminal plasma was also tested for oxidative<br />

stress with a TBARS assay. Pre- and Post-orchiectomy samples were<br />

compared using a paired Wilcoxon test (p


was a trend towards a higher BMI with worsening semen concentration.<br />

When controlled for the multiple factors listed above, we did not find any significant<br />

differences in birth weight or BMI at time of first driver’s license for<br />

children of men with semen analyses relative to fertile controls.<br />

CONCLUSIONS: Although obesity is associated with poor semen quality,<br />

our study demonstrates that the BMI of offspring is not associated with<br />

abnormal semen parameters of the father.<br />

P-542 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

PATERNAL AGE AND VASECTOMY REVERSAL<br />

SUCCESS. B. E. Kahn, a D. L. Davenport, b C. G. Schrepferman. c a Urology,<br />

University of Kentucky, Lexington, KY; b University of Kentucky, Lexington,<br />

KY; c University of Louisville, Louisville, KY.<br />

OBJECTIVE: To evaluate the effect of paternal age on patency rates<br />

following vasectomy reversals.<br />

DESIGN: Retrospective review of all vasectomy reversals performed by a<br />

single surgeon from 2005 to 2014. For inclusion, patients supplied at least<br />

one post-operative semen analysis or a reported a pregnancy. Vasectomy reversals<br />

performed for pain were excluded (n¼3). Patency was defined as the<br />

presence of > 500,000 motile sperm on a post-operative semen analysis or<br />

report of pregnancy.<br />

MATERIALS AND METHODS: Patency rates were assessed by vasectomy<br />

reversal procedure performed, which included microsurgical bilateral<br />

vasovasostomies (VV/VV), unilateral vasovasostomy with contralateral vasoepididymostomy<br />

(VV/VE), bilateral vasoepididymostomy (VE/VE), unilateral<br />

VV, and unilateral VE. The patency rate was stratified by decade of<br />

paternal age (20-29, 30-39, 40-49, 50-59). Statistical analysis was performed<br />

using Fisher’s Exact test, Chi-square test, and ANOVA linear trend tests.<br />

P-values


x10⁶ sperm and 55.028,57x10⁶/ mL, respectively. The lowest and highest<br />

catalase concentration detected was 11.39 (catalase units /total protein)<br />

and 101.84 (catalase units/total protein), respectively. There was no significant<br />

correlation between ROS levels and catalase levels in spermatozoa.<br />

Methylation status of CpG island I and CpG island II of CAT gene was evaluated.<br />

CpG island I hypermethylation was found in all samples, however<br />

CpG island II of CAT gene was not methylated in the 20 samples studied.<br />

CONCLUSIONS: Our results demonstrate hypermethylation of CpG island<br />

I and unmethylation of CpG island II of CAT gene in spermatozoa. In<br />

addition, no significant association was seen between ROS level and methylation<br />

status of CAT gene.<br />

Supported by: Cleveland Clinic.<br />

P-546 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

COMPROMISED TOTAL ANTIOXIDANT CAPACITY SUPPORTED<br />

BY HIGH DFI TO GUIDE SPERM SOURCING. T. Paniza, a L. Park, a<br />

L. Reisman, a Q. V. Neri, a T. Cozzubbo, a M. Goldstein, b Z. Rosenwaks, a<br />

G. D. Palermo. a a Reproductive Medicine, Weill Cornell Medical College,<br />

New York, NY; b Urology, Weill Cornell Medical College, New York Presbyterian,<br />

New York, NY.<br />

OBJECTIVE: To assess the impact of seminal antioxidant capacity on<br />

sperm DNA integrity. To measure embryo developmental competence of<br />

spermatozoa retrieved from a healthier micro-environment through microsurgery.<br />

DESIGN: Men with extremely high DFI in their ejaculates (n¼45) simultaneously<br />

received a TAC analysis on seminal fluid. Following counseling,<br />

men underwent surgical sampling and DNA integrity assessment on spermatozoa<br />

retrieved from vassal fluid, epididymis, and testis. TAC, sperm parameters,<br />

DFI and clinical outcome were recorded and compared for each<br />

individual sperm source.<br />

MATERIALS AND METHODS: Ejaculates were processed in standard<br />

fashion, assessed for TAC, DFI by TUNEL and/or SCSA. TAC was determined<br />

by a colorimetric assay on an automated microplate reader. Surgical<br />

specimens were isolated from different sites of the male genital tract and<br />

DFI measured exclusively by TUNEL then cryopreserved for later use.<br />

RESULTS: In 51 ejaculates the average DFI was 32.0% for TUNEL<br />

and 40.6% with SCSA with a good concordancy between the two assays.<br />

While proximal surgical sampling improved DFI, it inversely yielded<br />

lower concentration of spermatozoa and with somewhat decreased<br />

motility. The antioxidant buffering capacity of the ejaculates inversely<br />

correlated with the DFI (P


formulations. 42 (48.3%) used cutaneous testosterone gels, while 23 (26.4%)<br />

used testosterone injections. 49 (56.3%) respondents found their patients to<br />

be satisfied ‘‘most times’’ and 30 (34.5%) said their patients are ‘‘almost always’’<br />

satisfied with Testopel therapy. When Testopel therapy was discontinued,<br />

cost was cited as the most common reason.<br />

CONCLUSIONS: Testopel implantation appears to be a highly successful<br />

therapy for the treatment of hypogonadism. This study categorizes common<br />

practices that have yet to be standardized. Based on this specialty society<br />

questionnaire, management of hypogonadism may require 10 or more pellets<br />

in the majority of cases.<br />

P-549 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

PREDICTORS FOR SPERM RETRIEVAL IN MICRODISSECTION<br />

SPERM EXTRACTION FOR NON-OBSTRUCTIVE<br />

AZOOSPERMIA. T. Ishikawa, K. Yamaguchi, Y. Takaya,<br />

R. Nishiyama, K. Kitaya, H. Matsubayashi. Reproduction Clinic Osaka,<br />

Osaka, Japan.<br />

OBJECTIVE: Recently, the most popular treatment in patients with nonobstructive<br />

azoospermia (NOA) has been micro TESE with subsequent assisted<br />

fertilization by intracytoplasmic sperm injection (ICSI). With the spread of ICSI,<br />

the presence of a minimum number of spermatozoa is required for fertilization.<br />

Micro TESE and ICSI cycles expose the couple to an emotional and financial<br />

burden, so it would be beneficial to predict the success of sperm retrieval using<br />

noninvasive parameters before attempted procedure. The aim of this study is to<br />

assess the predictors of sperm retrieval by micro TESE in NOA patients.<br />

DESIGN: A retrospective study.<br />

MATERIALS AND METHODS: A total of 1323 micro TESE attempts<br />

were done in 1275 men with confirmed cryptozoospermia and NOA to<br />

recover spermatozoa for ICSI between January 2006 and March <strong>2015</strong> by a<br />

single surgeon. Micro TESE was used in which seminiferous tubule are<br />

directly examined throughout the testicle using an operating microscope<br />

and selectively biopsied for all of the NOA patients. We analyzed sperm<br />

retrieval rate (SRR) of the patients with NOA. Serum follicle stimulating hormone<br />

(FSH), luteinizing hormone (LH), testosterone (T), testicular volume,<br />

age at micro TESE, chromosomal analysis, AZF microdeletions analysis, and<br />

past history were examined as predictive factors for sperm recovery. Chromosomal<br />

analysis was performed on all patients on cultured lymphocytes<br />

from peripheral blood.<br />

RESULTS: Testicular sperm were successfully retrieved by micro-TESE<br />

in 560 of 1323 (42.5%). No correlation was found between serum FSH,<br />

LH, and T level with the success of sperm retrieval. Testicular volume and<br />

patient age also did not affect the SRR for micro-TESE. Good candidates<br />

of sperm retrieval by micro TESE were cryptozoospermia (114/120:<br />

95.0%), AZFc microdeletion (22/28:78.6%), associated with cryptorchidism<br />

(50/69: 72.5%), and non-mosaic Klinefelter syndrome (77/144: 53.5%).<br />

Worse candidates of sperm retrieval were AZFa or b microdeletions (0/<br />

10:0%), 46XY male with NOA without past history (180/643: 28.0%), and<br />

after chemotherapy (25/70: 35.7%).<br />

CONCLUSIONS: A prognostic parameter for successful sperm retrieval<br />

in TESE seems to be decisive for male fertility. FSH is not able to resolve<br />

spermatogenesis on an individual tubule level, and, therefore, they should<br />

not be used as predictors of sperm recovery. We conclude that at the present<br />

time there are no absolute predictors of sperm yield for micro TESE. However,<br />

we could predict good candidates for micro TESE by past history<br />

and genetic analysis.<br />

P-550 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

THE RELATIONSHIP BETWEEN A MAN’S SOMATIC HEALTH<br />

AND ART OUTCOMES. M. Eisenberg, a S. Li, b B. Behr, c<br />

S. Nakajima, a V. L. Baker. d a Urology, Stanford University, Stanford, CA;<br />

b Stanford University, Palo Alto, CA; c Stanford Fertility and Reproductive<br />

Medicine Cente, Pali alto, CA; d Stanford University, Stanford, CA.<br />

OBJECTIVE: As medical comorbidity and medication use increases,<br />

semen quality declines. However, less is known about how a man’s somatic<br />

health may impact the outcomes of assisted reproductive techniques.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: After IRB approval, we identified couples<br />

undergoing assisted reproductive technology (ART) cycles at our center<br />

from 2004 until 2014. We only fresh IVF cycles utilizing fresh ejaculated<br />

sperm from the male partner. We recorded patient and partner demographic<br />

characteristics. The cohort was linked to administrative data to obtain information<br />

on the male partners’ comorbidities identified using ICD-9-CM codes<br />

and limited to men evaluated within the health care system. Cycle outcomes<br />

were queried from our clinical database. We calculated fertilization rate, clinical<br />

pregnancy rate, miscarriage rate, implantation rate, live birth rate, and<br />

singleton birth weight. Regression models were adjusted for male and female<br />

covariates.<br />

RESULTS: In all, we identified 772 men who had outpatient data available.<br />

Those men underwent 1,503 fresh ART cycles - 702 were IVF only<br />

and 801 utilized ICSI. Overall, the mean age of the man was 39.6 and 37.7<br />

for his female partner. 67 % of men had at least one medical diagnosis.<br />

96% of men had a CCI of 0. After stratifying by organ system, differences<br />

were noted for ART outcomes based on any male diagnosis. Men with neurologic<br />

diseases had a lower live birth rate (15% vs 23%, p¼0.02) while men<br />

with endocrine diseases had a higher implantation rate (73% vs 62%,<br />

p¼0.02). The associations were similar for unadjusted and adjusted models.<br />

When examining singleton birth weights after all forms of ART, men with<br />

diseases of the nervous system (3270g vs 2990g, p


function tests. Fourteen of 32 subjects (43.8%) achieved pregnancy,<br />

including 10/13 (76.9%) conceiving through in vitro fertilization, 2/3<br />

(66.7%) through intrauterine insemination and 2/16 (12.5%) through natural<br />

intercourse.<br />

CONCLUSIONS: Significant improvement in testosterone levels and<br />

sperm concentration is seen with the use of anastrozole in the treatment<br />

of the overweight or obese, subfertile male. Anastrozole is well tolerated<br />

with minimal side effects in overweight or obese male subjects.<br />

Concomitant anastrozole therapy in conjunction with assisted reproductive<br />

techniques appears to be associated with successful conception for<br />

infertile couples.<br />

P-552 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

EFFECT OF TIME ON OXIDATION-REDUCTION POTENTIAL<br />

IN SEMEN AND SEMINAL PLASMA. R. Sharma, a S. S. Du Plessis,<br />

a,b A. Agarwal, a A. Harlev, a,c L. Samanta, a,d G. Ahmad, a,e S. Gupta, a<br />

E. S. Sabanegh. f a Center for Reproductive Medicine, Cleveland Clinic,<br />

Cleveland, OH; b Medical Physiology, Stellenbosch University, Tygerberg,<br />

South Africa; c Soroka Medical Center, Ben-Gurion University, Beer Sheva,<br />

Israel; d Redox Biology Laboratory, School of Life Science, Ravenshaw<br />

University, Orissa, India; e Physiology and Cell Biology, University of<br />

Health Sciences, Lahore, Pakistan; f Urology, Cleveland Clinic, Cleveland,<br />

OH.<br />

OBJECTIVE: Oxidation-reduction potential (ORP) is a novel measure of<br />

oxidative stress or redox imbalance in biological fluids. Reactive oxygen species<br />

(ROS) are highly reactive and have a very short half-life. ROS levels in<br />

the seminal ejaculate should be measured within an hour after collection to<br />

prevent a reduction in ROS levels over time. The traditional methods of<br />

measuring seminal ROS are time sensitive and time consuming, making it<br />

difficult to use them for diagnostic purposes. It would be highly advantageous<br />

to employ a method that is independent of semen age and provides results in<br />

real time. The objective was to assess the effect of time on static ORP<br />

(sORP), which provides a snapshot of current redox balance, and capacity<br />

ORP (cORP) which is indicative of the amount of antioxidant reserves available.<br />

DESIGN: Prospective study measuring ORP in semen and seminal plasma<br />

samples at time 0 and 120 minutes.<br />

MATERIALS AND METHODS: The sORP and cORP of both semen<br />

(n¼18) and seminal plasma (n¼15) samples from normal control subjects<br />

were measured after liquefaction (time 0) and after 120 minutes<br />

of incubation at room temperature (RedoxSYSÒ, Aytu BioScience).<br />

Values are mean SEM. Spearman correlation was used for statistical<br />

analysis.<br />

RESULTS: A significant correlation was seen between sORP at time 0 and<br />

120 minutes in semen and seminal plasma. Similar correlations were found<br />

for cORP values at both time intervals.<br />

CONCLUSIONS: ORP values are not affected by the age of semen or seminal<br />

plasma for up to 120 minutes, making it easier to employ this new technology<br />

for diagnostic use.<br />

Effect of time on semen and seminal plasma ORP levels.<br />

Semen<br />

sORP<br />

(mV/10 6<br />

sperm)<br />

Semen<br />

cORP<br />

(mC/10 6<br />

sperm)<br />

Seminal<br />

Plasma<br />

sORP<br />

(mV/10 6<br />

sperm)<br />

Seminal<br />

Plasma<br />

cORP<br />

(mC/10 6<br />

sperm)<br />

Time 0’ 2.790.66 0.200.08 2.670.71 0.310.16<br />

Time 120’ 2.840.67 0.240.11 2.790.79 0.260.11<br />

Correlation (r) 0.97 0.88 0.97 0.98<br />

Median absolute<br />

value of<br />

difference<br />

0.<strong>17</strong>5 0.01 0.13 0.01<br />

P-553 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

THE EFFECT OF VARICOCELE ON EMBRYO DEVELOPMENT<br />

AND IVF OUTCOMES. A. S. Polackwich, a E. S. Sabanegh, b<br />

N. Desai. c a Urology, Cleveland Clinic, Cleveland Heights, OH; b Cleveland<br />

Clinic, Cleveland, OH; c Cleveland Clinic, Beachwood, OH.<br />

OBJECTIVE: The treatment of male factor infertility secondary to varicocele<br />

is controversial in the era of Intracytoplasmic Sperm Injection. Utilizing<br />

this technique, all but the most severe male factor can be overcome. In some<br />

patients, treatment of varicoceles can help a couple avoid IVF, but for those<br />

who require it, the impact of a varicocele on outcomes is not well defined. We<br />

evaluated the influence of a clinically relevant varicocele on IVF.<br />

DESIGN: Retrospective chart review of all couples treated by both Cleveland<br />

Clinic Andrology and Reproductive Endocrinology<br />

MATERIALS AND METHODS: After IRB approval, we cross-referenced<br />

our database of patients who have undergone IVF with all patients seen in our<br />

male infertility clinic. We then extracted data on IVF outcomes, demographics,<br />

seminal parameters, physical exam and surgical history. All patients<br />

who underwent a varicocele ligation prior to IVF were excluded. All<br />

grading of varicoceles was confirmed by one of two specialists in male infertility<br />

(AT and ES). Ultrasound findings were not included. We evaluated the<br />

effect of the presence of a varicocele on IVF outcome parameters. Outcomes<br />

evaluated were: fertilization rate, number of embryos transferred, clinical<br />

pregnancy rate, implantation rate, number of embryos frozen, and blast formation<br />

rate.<br />

RESULTS: There were a total of 194 cycles performed. Of those, 11 had a<br />

grade 3 varicocele, 54 had a Grade 2, 40 had a grade 1 and 89 had no varicocele.<br />

There were no statistically significant differences in seminal parameters.<br />

Comparing those with a varicocele to those without, patient ages were<br />

37.3 +/- 5.4 vs 37.4 +/- 6 yrs. Partner ages were 34.5 +/- 4.1 vs 34.1 +/- 4<br />

yrs. Results are listed in the table below. There were a higher number of embryos<br />

transferred in those with a varicocele, but less embryos frozen.<br />

CONCLUSIONS: Clinically significant varicoceles may affect IVF cycle<br />

outcomes. For those patients who require IVF, varicocele may alter the outcomes,<br />

but significant differences were not seen, though larger studies need<br />

to be conducted to evaluate for more subtle effects.<br />

IVF Cycle outcomes by varicocele presence and grade<br />

Fertilization<br />

Rate (%)<br />

Embryos<br />

Transferred<br />

Clinical<br />

Pregnancy<br />

rate (%)<br />

Implantation<br />

rate (%)<br />

Number<br />

of<br />

Embryos<br />

frozen<br />

Blastocyst<br />

Formation<br />

rate (%)<br />

No Varicocele<br />

76.7 1.8 +/- 0.77 53.9 44.7 2.6 +/- 2.8 49.4<br />

(89)<br />

Any Varicocele<br />

75.1 2.1 +/- 0.62 43.2 30.1 1.8 +/- 2.54 39.4<br />

(105)<br />

Grade 1 (40) 76.1 2.3 43.6 27.4 1.7 37.7<br />

Grade 2 (54) 72.9 2.1 44.4 32.7 1.7 42.6<br />

Grade 3 (11) 80.8 1.9 36.4 27.2 2.9 32.7<br />

P value (Any<br />

v None)<br />

0.69 0.006 0.16 0.11 0.06 0.06<br />

FERTILITY & STERILITY Ò<br />

e295


SPERM BIOLOGY<br />

P-554 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

WITHDRAWN<br />

RESULTS: A total of 1202 proteins were identified in the F1 fraction while<br />

1140, 1025 and 890 proteins were recovered from the three other fractions,<br />

F2, F3 and F4 respectively. With respect to the differentially expressed proteins,<br />

F1 exhibited the highest number (522), followed by F2 (362) and<br />

lowest was observed in F3 (188) as compared to F4. Ingenuity Pathway analysis<br />

revealed 162 pathways showing a decreasing trend from F1 through F4.<br />

The principal pathways were spermatogenesis, protein metabolism, cell cycle,<br />

integration of energy metabolism, regulation of apoptosis, cell redox homeostasis<br />

and translational elongation. The proteins detected in our data set<br />

belonging to the top 15 pathways shared many proteins that were downregulated,<br />

suggesting failure of signaling cascade in the process of maturation in<br />

infertile patients.<br />

CONCLUSIONS: We conclude that a defective signaling cascade is<br />

responsible for the faulty sperm function in infertile patients. Particularly,<br />

a decline in mitochondrial function and oxidative phosphorylation in the<br />

F4 fraction implies an energy deprived hypoxic state. Dysregulated protein<br />

turnover and protein folding may lead to accumulation of defective proteins<br />

and can potentially impair of sperm function.<br />

P-556 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

A COMPREHENSIVE ASSESSMENT OF HISTONE MODIFICA-<br />

TIONS IN HUMAN SPERM. S. B. Schon, a L. J. Luense, b X. Wang, b<br />

G. Donahue, b B. A. Garcia, b M. S. Bartolomei, b S. L. Berger. b a Reproductive<br />

Endocrinology & Infertility, University of Pennsylvania, Philadelphia,<br />

PA; b University of Pennsylvania, Philadelphia, PA.<br />

P-555 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

MORPHOLOGICALLY DISTINCT PHENOTYPES OF SPERMATO-<br />

ZOA IN INFERTILE MEN REVEAL DOWN REGULATION OF<br />

MULTIPLE SIGNALING PATHWAYS. A. Agarwal, a Z. Cui, a,b<br />

R. Sharma, a L. Samanta, a,c R. Turki, d M. Abu-Elmagd. e a Center For Reproductive<br />

Medicine, Cleveland Clinic, Cleveland, OH; b Institute of Toxicology,<br />

Third Military Medical University, Chongqing, China; c Redox Biology Laboratory,<br />

School of Life Science, Ravenshaw University, Orissa, India; d Ob/<br />

Gyn, King AbdulAziz University, Mobil, AL; e CEGMR, Center of Excellence<br />

in Genomic Medicine Research, Jeddah, Saudi Arabia.<br />

OBJECTIVE: Seven morphologically distinct spermatozoal phenotypes<br />

can be detected in human semen under electron microscopy: sperm with<br />

dysplasia of the fibrous sheath, non-specific flagellar defects, immotile cilia<br />

syndrome, acrosomal hypoplasia, defective chromatin condensation and<br />

compaction, pin head and sperm without heads. These conditions cannot<br />

be identified by routine semen analysis because they are but secondary manifestations<br />

of underlying pathology. We hypothesize that proteomics would<br />

enable us to better understand the mechanism(s) underlying these pathologies.<br />

DESIGN: This proteomic study analyzed proteins in spermatozoa from<br />

infertile men fractionated on three layers of density gradient (80, 60 and<br />

40%). Fraction 1 (F1) refers to the least mature stage having the lowest density<br />

whereas fraction 4 (F4) contains the most dense and morphologically<br />

mature motile spermatozoa. Fraction 2 (F2) and fraction 3 (F3) are the intermediate<br />

stages.<br />

MATERIALS AND METHODS: 1-D gel electrophoresis followed by LC/<br />

MS-MS (LTQ-Orbitrap Elite hybrid mass spectrometer) was used for protein<br />

identification. Mascot (Matrix Science, London, UK), SEQUEST (Thermo<br />

Fisher Scientific, San Jose, CA, USA) and X! Tandem (TheGPM, thegpm.org)<br />

were set up to search the human reference with database assuming<br />

trypsin as the digestion enzyme. Functional annotations of proteins were obtained<br />

using bioinformatic tools and pathway databases.<br />

OBJECTIVE: The unique process of spermiogenesis involves a dramatic<br />

reorganization of paternal chromatin, whereby 90% of histones<br />

are evicted and replaced with protamines. Proper exchange is critical<br />

for nuclear compaction and abnormalities in this process have been associated<br />

with male infertility [1,2]. Recent evidence suggests that the 10%<br />

of histones retained in mature sperm are located at specific genomic loci.<br />

Furthermore, selective histone modifications (i.e. acetylation, methylation)<br />

are enriched at genes encoding master regulators of embryo development,<br />

suggesting a role for paternal transmission of epigenetic information<br />

[3,4]. To date, our understanding of histone modifications in human<br />

sperm has largely been limited to descriptive data of specific modifications.<br />

Our objective was to perform a comprehensive evaluation of histone<br />

modifications in normal sperm, to quantify their abundance, and to<br />

assess variation in modifications between individuals and in fresh vs.<br />

frozen sperm.<br />

DESIGN: Descriptive study of human sperm with normal semen analysis<br />

parameters utilizing acid extraction of histones and bottom up mass spectrometry<br />

to identify and quantitate histone modifications.<br />

MATERIALS AND METHODS: Normal semen samples were obtained<br />

from men undergoing routine semen analysis (n¼8). 4 samples were<br />

equally divided with one portion cryopreserved at our Andrology Laboratory<br />

and one half processed immediately. Following acid extraction, histones<br />

were digested into peptides for bottom up LC-tandem mass spectrometry<br />

by treatment with propionic acid and trypsin. The identities and<br />

relative abundance of peptides and associated modifications were analyzed<br />

based on retention time, MS1 and MS2 of ion peaks. Kruskal-Walis and<br />

ANOVA tests were used as appropriate for assessment of inter-individual<br />

variability in modification abundance and to compare fresh vs. frozen<br />

samples.<br />

RESULTS: We identified modifications on all four canonical histones<br />

(H2A, H2B, H3, H4), the linker histone H1 and histone variants. We describe<br />

a total of 102 modifications identified for the first time in human sperm. Variability<br />

in modifications on H3 was noted between individuals, while the<br />

abundance of modifications on H4 was consistent. No differences were<br />

observed between fresh vs. frozen sperm.<br />

CONCLUSIONS: Our study is the first to provide a comprehensive and<br />

quantitative analysis of histone modifications in human sperm. We demonstrate<br />

significant variation between individuals in specific modifications on<br />

H3 and relative conservation in the abundance of H4. The conservation of<br />

specific modifications further supports the hypothesis that the paternal epigenome<br />

plays a role in early developmental programming and we further postulate<br />

that aberrations in histone modifications may play a role in decreased<br />

fertility. Future analysis will focus on modification profiles of abnormal<br />

sperm.<br />

References:<br />

1. Carrell, D.T. and L. Liu, Altered protamine 2 expression is uncommon<br />

in donors of known fertility, but common among men with poor fertilizing<br />

capacity, and may reflect other abnormalities of spermiogenesis. J<br />

Androl, 2001. 22(4): p. 604-10.<br />

e296 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


2. Aoki, V.W., et al., Sperm protamine 1/protamine 2 ratios are related to<br />

in vitro fertilization pregnancy rates and predictive of fertilization ability.<br />

Fertil Steril, 2006. 86(5): p. 1408-15.<br />

3. Hammoud, S.S., et al., Distinctive chromatin in human sperm packages<br />

genes for embryo development. Nature, 2009. 460(7254): p. 473-8.<br />

4. Arpanahi, A., et al., Endonuclease-sensitive regions of human spermatozoal<br />

chromatin are highly enriched in promoter and CTCF binding<br />

sequences. Genome Res, 2009. 19(8): p. 1338-49.<br />

Supported by: T32HD040135, U54HD068157.<br />

P-557 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

SIGNIFICANT CORRELATION BETWEEN SPERM DNA DAMAGE<br />

AND ACHIEVING A PREGNANCY IN 1847 NON-IVF FERTILITY<br />

TREATMENT CYCLES. J. Brachtchenko, S. Moskovstev,<br />

S. Swanson, E. Shlush, L. Kan-Ool, P. Sharma, A. Y. Baratz, K. Glass,<br />

C. L. Librach. CReATe Fertility Centre, Toronto, ON, Canada.<br />

OBJECTIVE: Timed intercourse (TI) or intrauterine insemination (IUI),<br />

with or without ovulation induction (OI), are often attempted before in-vitro<br />

fertilization (IVF) when the tubes are patent. Despite much research, controversy<br />

remains regarding the effectiveness of these treatments as well as the<br />

effect of sperm quality on the outcomes of TI and IUI. The objective of<br />

this study was to identify semen parameters that might predict success of<br />

these first-line fertility treatments.<br />

DESIGN: A retrospective chart review.<br />

MATERIALS AND METHODS: A retrospective chart review of patients<br />

who underwent TI or IUI with or without OI at the CReATe Fertility Centre in<br />

Toronto, Canada, was performed. Only cases where the male partner<br />

completed an extensive laboratory evaluation of semen (defined as standard<br />

semen analysis, computer-aided sperm analyses (CASA; with calculation of<br />

mucus-penetrating kinetics [MPK] and sperm DNA damage (DNA fragmentation<br />

index [DFI]) prior to the treatment were included. The patient clinical<br />

data and cycle information was collected for both partners and analyzed by<br />

SSSP statistical software.<br />

RESULTS: In total, data was collected from 362 couples who underwent<br />

1847 treatment cycles; 354 TI (19%) and 1493 IUI (79%); while<br />

532 (28%) were ‘natural’ cycles and the rest were underwent OI. The<br />

likelihood of pregnancy was not statistically different between OI and<br />

‘natural’ cycles. In total, 132 couples were able to achieve a pregnancy<br />

(36.5%) with 96 couples having a successful birth (26.5%) and 37<br />

(10.2%) experiencing miscarriage(s) only. 230 (63.3%) did not achieve<br />

a pregnancy using TI or IUI. The multiple pregnancy rate was 6.3%<br />

(6/96). The pregnancy and miscarriage rate was stable from 1 to 6 cycles<br />

performed, however, once the number of cycles surpassed 7, the likelihood<br />

of achieving a pregnancy was greatly diminished (P < 0.05),<br />

with no patients who attempted 8-10 cycles achieving a pregnancy.<br />

DFI ranged from 0.8 to 77.6% (18.3% 12) and was significantly correlated<br />

to patient’s age, sperm concentration, motility, MPK and total<br />

motile sperm concentration (TMSC) inseminated during IUI (p 30%<br />

DFI) was significantly less likely to eventually achieve a pregnancy<br />

following TI and IUI when compared to the Low (


organ development, development and function of the reproductive system<br />

were uniquely expressed in fraction F4 of the fertile donors.<br />

CONCLUSIONS: The study suggests that altered protein profiles in infertile<br />

men may be responsible for their sperm dysfunction. Further work is in<br />

progress to validate the expression profile of these proteins.<br />

Proteomic profile of different fractions of spermatozoa from infertile men.<br />

Proteins identified<br />

No. of<br />

Proteins<br />

No. of<br />

Proteins<br />

No. of<br />

Proteins<br />

No. of<br />

Proteins<br />

F1 F2 F3 F4<br />

Total 1366 1285 1325 1100<br />

No. of DEP 119 167 373 262<br />

Overexpressed 38 79 135 106<br />

Underexpressed 37 46 79 82<br />

Unique to<br />

<strong>21</strong> 18 91 44<br />

fertile group<br />

Unique to<br />

infertile group<br />

23 24 68 30<br />

P-560 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

PATERNALTRANSCRIPTOME ANALYSIS BY RNA-SEQUENCING<br />

AS A MEASURE OF EMBRYONIC DEVELOPMENTAL<br />

POTENTIAL. T. Cozzubbo, Q. V. Neri, Z. Rosenwaks,<br />

G. D. Palermo. Reproductive Medicine, Weill Cornell Medical College,<br />

New York, NY.<br />

OBJECTIVE: To classify and predict the value of the contribution by the<br />

paternal gamete to the embryonic epigenome. By identifying vital spermatozoa<br />

RNA transcripts, we can predict their influence upon delivery to the<br />

oocyte and how they link to embryonic development.<br />

DESIGN: Beyond transporting the genetic material to the oocyte, spermatozoa<br />

deliver coding and non-coding RNA that ordain early embryonic development.<br />

We attempted to profile the paternal transcriptome and categorize<br />

key regulating factors involved in syngamy by controlling vital checkpoints<br />

during embryogenesis. Expression of specific genes were related to fertile<br />

and infertile individuals.<br />

MATERIALS AND METHODS: Men undergoing infertility treatment<br />

were screened and semen parameters were assessed according to WHO<br />

2010. RNA was isolated from 25x106 human spermatozoa using a spin column<br />

commercial kit. The nucleic acid quality, RNA integrity number (RIN),<br />

and spermatozoal RNA concentration were assessed. The RNA samples were<br />

then made into paired-end libraries. Pilot paired-end 36bp RNA-Sequencing<br />

(RNA-Seq) using an Illumina platform (NextSeq 500) was carried out and<br />

then expanded to 60M reads.<br />

RESULTS: In 19 consenting infertile men with a mean age of<br />

39.65.1yrs, an average sperm concentration of 47.916.0x106/mL,<br />

motility of 47.15.5%, and normal morphology of 2.91.6%. The RNA<br />

isolated from the samples yielded an average concentration of 14.055.9<br />

ng/mL and a RIN of 5.91.7. BOK, a regulating factor of apoptosis during<br />

the cell cycle, had an excess of transcripts in the sperm of infertile patients,<br />

corresponding to a higher (14.3%) embryonic fragmentation rate. RANBP2<br />

contributes to the reorganization of the paternal genome post-fertilization<br />

but was found to be comparably expressed in both fertile and infertile<br />

men; coinciding with acceptable embryo cleavage (85.7%) in the infertile<br />

cohort. PLK4 and BUB1, which dictate chromosome stability & mitotic<br />

segregation by regulating centriole development, displayed a decreased<br />

expression in the infertile group and indeed this group yielded 66.7%<br />

genetically abnormal embryos as determined by preimplatation genetic<br />

diagnosis.<br />

CONCLUSIONS: Sperm RNA-Seq is an ideal technique capable of reaching<br />

beyond microscopic evaluation, morphometric assessment, and traditional<br />

diagnostic assays to help predict the contribution of the male gamete<br />

to early pre-implantation development. This novel technique, by assessing<br />

coding and non-coding RNA, may further classify the transcripts that play<br />

an influential role in pre and post fertilization and informs on embryo developmental<br />

competence of a specific paternal epigenome.<br />

Supported by: WCMC.<br />

OOCYTE MATURATION<br />

P-561 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

THE FGF 23 LEVELS IN HUMAN FOLLICULAR FLUIDS IN THE<br />

FOLLICULAR PHASE OF PREIMPLANTATION ARE ASSOCI-<br />

ATED WITH THE AGING OR QUALITY OF HUMAN<br />

OOCYTES. F. Saji, S. Taguchi, M. Funabiki, N. Amano, T. Takano,<br />

L. K. Young, T. Hayashi, Y. Tada, Y. Iwaki, M. Karita, Y. Nakamura.<br />

Oak Clinic, Osaka, Japan.<br />

OBJECTIVE: Though approaches to evaluate the aging of human oocytes<br />

have been investigated, a simple biomarker to evaluate the aging or quality of<br />

human oocytes is required. Therefore, the present study aimed to find such a<br />

biomarker.<br />

DESIGN: An experimental prospective cohort study.<br />

MATERIALS AND METHODS: The associations between the fibroblast<br />

growth factor (FGF) 23 levels in human follicular fluids in the follicular<br />

phase of preimplantation and outcome data were investigated from June<br />

2013 to January <strong>2015</strong> in 60 patients (median age 37.3 years) who provided<br />

informed consent at an IVF clinic. The primary outcomes in the present study<br />

were the maturation rates in human oocytes and fertilization rates. The FGF<br />

23 levels in the human follicular fluids obtained during the follicular phase of<br />

preimplantation were evaluated using an FGF-23 ELISA Kit. Statistical analyses<br />

were investigated using ANOVA, Welch’s t test and multivariate analysis.<br />

RESULTS: The FGF 23 levels in human follicular fluids in the follicular<br />

phase of preimplantation were 20.78 6.84 pg/ml (mean SD). Furthermore,<br />

the FGF 23 levels in human follicular fluids in the follicular phase<br />

of preimplantation were elevated by aging (p


(54.2% and 79.2%) in Type C and 223 (71.7% and 75.0%), respectively. The<br />

maturation rate of immature oocytes was significantly different with Type D<br />

and Type A (71.7% vs. 51.0%, P


14.06, respectively. There was a significant inverse relationship between<br />

increasing BMI and lower serum hCG levels in the SC group. However, there<br />

was no relationship between BMI and percent mature oocytes. Because of the<br />

difference in luteal support between the two groups, strict comparisons of<br />

pregnancy rates were not possible, however there were no significant difference<br />

in pregnancy rates.<br />

Subcutaneous<br />

hCG (n¼<strong>21</strong>5)<br />

Intramuscular<br />

hCG (n¼159)<br />

p-value<br />

Age 34.87 35.78 0.08<br />

BMI 25.10 25.91 0.<strong>17</strong><br />

AMH 2.61 2.76 0.70<br />

Total oocytes retrieved 15.<strong>21</strong> 14.06 0.19<br />

Number mature 11.04 10.28 0.26<br />

Percent mature 76.28% 74.37% 0.57<br />

CONCLUSIONS: SC-hCG is as effective as IM-hCG for final follicular<br />

maturation in IVF cycles. There was a significant inverse relationship between<br />

increasing BMI and lower serum hCG levels in the SC-hCG group.<br />

However, there was no relationship between BMI and percent mature oocytes.<br />

P-566 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

THE CORRELATION OF ANTI-MULLERIAN HORMONE CON-<br />

CENTRATIONS TO OOCYTE MATURITY AT<br />

RETRIEVAL. M. Bustillo, a C. Alford, a I. Collazo, a K. O. Pomeroy. b,a<br />

a South Florida Institute for Reproductive Medicine, Miami, FL; b The World<br />

Egg Bank, Phoenix, AZ.<br />

OBJECTIVE: Determine if Anti-Mullerian hormone levels correlate to the<br />

number of mature ooctyes or the percent of mature oocytes at the time of<br />

oocyte retrieval.<br />

DESIGN: Retrospective analysis of 2459 cycles of IVF from 2010 to 2014.<br />

MATERIALS AND METHODS: Spearman’s rank correlation was done<br />

on all IVF cycles where the AMH value was equal to or less than 11.10<br />

ng/ml to determine correlation between total number of mature ova and<br />

percent of mature ova. Cycles where the number of collected mature ova<br />

were less than 6 were classified as poor cycles and an ROC curve was generated<br />

to identify the cutoff value of AMH for ‘‘poor cycles.’’<br />

RESULTS: There was a moderate correlation of AMH value to the total<br />

number of mature oocytes retrieved (Spearman’s rho ¼ 0.578; p


emain unclear (1). Studies exploring this topic, however, do not address<br />

more clinically significant outcomes such as clinical pregnancy rate (CPR)<br />

or live birth rate (LBR). Therefore, the objective of this study was to determine<br />

whether patients with PCOS have better pregnancy outcomes following<br />

ICSI versus IVF.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Data from the 2011-2012 SART registry<br />

were analyzed. First fresh autologous cycles in women under 40 with<br />

the diagnosis of PCOS were included in the study. All other diagnostic<br />

indications for IVF were excluded from this study including male factor.<br />

Groups were further stratified based on the insemination technique: conventional<br />

IVF or ICSI. Cycles using split IVF/ICSI were excluded. Cycles<br />

were divided into two groups: cleavage stage embryo transfer (ET) and<br />

blastocyst stage ET. Main outcomes measured were the number of two<br />

pronuclear (2PN) embryos, LBR, and miscarriage rate. Secondary outcomes<br />

included implantation rate (IR), CPR and the percentage of cycles<br />

that did not result in 2PN embryos. Data were analyzed using two-sided<br />

Welch’s t-test.<br />

RESULTS: 22<strong>17</strong> cycles reported use of conventional IVF and 2462 cycles<br />

reported use of ICSI. In the IVF group, the number of 2 PN oocytes, CPR, and<br />

LBR were statistically higher in both cleavage stage and blastocyst stage<br />

transfer cycles compared to the ICSI group. Miscarriage rate was lower in patients<br />

undergoing ICSI with blastocyst ET but higher in those with cleavage<br />

ET compared to the IVF group. The percentage of cycles with no 2PN oocytes<br />

was statistically lower in the ICSI group with blastocyst ET only (Table).<br />

Although statistical significance was found between these two groups,<br />

the differences were not clinically significant.<br />

CONCLUSIONS: Contrary to previous studies, ICSI does not offer an<br />

advantage over conventional IVF in terms of number of oocytes fertilized,<br />

CPR, or LBR.<br />

MATERIALS AND METHODS: Using an established and validated IVF<br />

database from a single large urban IVF center in an insurance mandated state,<br />

we investigated failed fertilization cases with more than 5 eggs and where<br />


MATERIALS AND METHODS: Sperm samples were produced by<br />

masturbation following 2-5 days abstinence and subjected to computer-assisted<br />

sperm analysis (CASA), Kremer penetration testing, calcium response<br />

recording (FLUOstar), PLCz immunocytochemistry (anti-human PLCz;<br />

Cova-Lab UK), and electrophysiology studies using single-cell patch clamping(4).<br />

DNA was extracted from whole blood samples and subjected to<br />

exome sequencing.<br />

RESULTS: 20% (5/25) had significant sperm pathologies detected,<br />

including abnormal calcium response, reduced Kremer penetration and putative<br />

CatSper / K+ channel abnormalities, most of which can be overcome by<br />

ICSI. Significantly, over half of the men investigated (13/25) were found to<br />

have reduced proportion and / or total relative fluorescence of PLCz.<br />

Following appropriate counselling, 6 of these couples have subsequently undergone<br />

ICSI and AOA using calcium ionophore A23187 (GM508 Cult Activ;<br />

Gynemed) Average M2 oocyte fertilisation rate was 61.2% (54.5% -<br />

70.8%), with 3 biochemical pregnancies and an ongoing intrauterine pregnancy<br />

to date, and further treatment results awaited. Exome sequence analysis<br />

is currently in progress.<br />

CONCLUSIONS: This study supports our hypothesis that sperm dysfunction<br />

commonly underlies unexplained infertility, and unexplained TFF,<br />

which may not be predicted by standard diagnostic semen analysis. Utilising<br />

a variety of laboratory approaches to investigate sperm calcium response,<br />

motility and function can enable application of an individualised approach<br />

to a future cycle of ART, including AOA where PLCz deficiencies are identified.<br />

References:<br />

1. Factors associated with failed treatment: an analysis of 1<strong>21</strong>,744 women<br />

embarking on their first IVF cycles. Bhattacharya S, Maheshwari A,<br />

Mollison J. PLoS One. 2013 Dec 5;8(12):e82249.<br />

2. The cytosolic sperm factor that triggers Ca2+ oscillations and egg activation<br />

in mammals is a novel phospholipase C: PLCzeta. Swann K,<br />

Larman MG, Saunders CM, Lai FA. Reproduction. 2004<br />

Apr;127(4):431-9. Review.<br />

3. Assisted oocyte activation following ICSI fertilization failure. Vanden<br />

Meerschaut F, Nikiforaki D, Heindryckx B, De Sutter P. Reprod Biomed<br />

Online. 2014 May;28(5):560-71. Review.<br />

4. Patch clamp studies of human sperm under physiological ionic conditions<br />

reveal three functionally and pharmacologically distinct cation<br />

channels. Mansell SA, Publicover SJ, Barratt CL, Wilson SM. Mol<br />

Hum Reprod. 2014 May;20(5):392-408.<br />

Supported by: TENOVUS Scotland MRC Chief Scientist Office NHS Tayside<br />

Royal Society UK Oxford University Medical Research Fund.<br />

P-571 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

ARTIFICIAL OOCYTE ACTIVATION WITH CALCIUM IONO-<br />

PHORE (A+23187) FOLLOWING ICSI FERTILIZATION<br />

FAILURE. A. Sdrigotti, a G. J. Rey Valzacchi, b F. A. Leocata Nieto, c<br />

V. E. Canada. d a Reproductive Medicine, Procrearte, Buenos Aires,<br />

Argentina;<br />

b Director, Buenos Aires, Argentina;<br />

c Embriologyst, Buenos<br />

Aires, Argentina; d Procrearte, Capital Federal, Argentina.<br />

OBJECTIVE: To assess the CALCIUM IONOPHORE (A+23187) use in<br />

patients with total failure or low fertilization rate on previous ICSI cycle/s.<br />

DESIGN: Retrospective, case- control study.<br />

MATERIALS AND METHODS: Patients with a history of one or more<br />

ICSI cycle/s with low (


Program through the National Research Foundation of Korea funded by the<br />

Ministry of Education, Science and Technology and a grant (A120080) from<br />

the Korean Healthcare Technology R&D project, Ministry for Health, Welfare<br />

and Family Affairs, Republic of Korea.<br />

P-573 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DELAYED PRONUCLEAR APPEARANCE AS AN INDICATION OF<br />

COMPROMISED OOCYTE REPAIR CAPACITY. T. Cozzubbo,<br />

Q. V. Neri, T. Paniza, L. Reisman, Z. Rosenwaks, G. D. Palermo. Reproductive<br />

Medicine, Weill Cornell Medical College, New York, NY.<br />

OBJECTIVE: To associate the delayed formation of two-pronuclei (2PN)<br />

after ICSI with the integrity of the oocytes capability to cope and eventually<br />

overcome varying levels of genomic error.<br />

DESIGN: In a retrospective study, we investigated the impact of the degree<br />

of sperm DNA fragmentation in relation to oocyte aging. We gauged the efficiency<br />

of oocyte repair mechanisms in occasion of delayed pronuclear<br />

appearance and consequent syngamy was assessed.<br />

MATERIALS AND METHODS: We identified couples undergoing ICSI<br />

treatment (n¼152) that had sperm chromatin assessment by TUNEL on their<br />

ejaculates. TUNEL was performed by adjusting sperm concentration to 5<br />

million, fixation with 4% PFA, exposure to mild permeabilizing solution,<br />

and incubated at 37 C with TUNEL reaction mixture. At least 500 spermatozoa<br />

were scored under fluorescent microscopy. Patients were categorized<br />

by DFI level and grouped according to maternal age. Each subdivision of<br />

DFI was evaluated in relation to clinical pregnancy outcome (presence of a<br />

fetal heartbeat). Standard fertilization is performed 16-18 hours after ICSI<br />

and delayed fertilization was conducted over 18 hours. To determine the significance<br />

of a delayed normal fertilization, arguably linked to an extensive<br />

processing and repair of the male genome following sperm nuclear unraveling,<br />

we assessed cohorts of oocytes that had at least one oocyte with a delayed<br />

appearance of the pronuclei in relation to a matched control.<br />

RESULTS: In a previous study, we established in 315 ICSI cycles which<br />

included 152 women with a mean age of 37.34yrs. In a cohort of women<br />


RESULTS: Following AOA, fertilization rate was remarkably improved<br />

from <strong>21</strong>.4% to 67.8%, and was comparable to that of non-AOA (65.5%) After<br />

transferring 32 AOA embryos, 6 (18.7%) implanted (IMP). Among AOA embryos<br />

considered morphologically suitable for transfer, IMP embryos<br />

showed earlier PN fading and first cell cleavage than non-IMP (P < 0.05).<br />

Interestingly, these significant differences disappeared after 2-cell stage.<br />

When AOA IMP embryos were compared to non-AOA IMP embryos (n ¼<br />

34), no difference was observed in any of morphokinetic events.<br />

CONCLUSIONS: Zygotes fertilized with the aid of activating stimuli<br />

showed morphokinetics in early development similar to those of conventional<br />

ICSI zygotes. After AOA, differences between embryos with or<br />

without the ability of implantation were identified only at PN fading and<br />

the first cleavage, but not at later developmental stages. This phenomenon indicates<br />

that the effect of AOA on morphokinetics is transitory.<br />

P-576 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

CORRELATION OF THE HYALURONAN BINDING ASSAY TO<br />

FERTILIZATION IN VITRO. K. O. Pomeroy, a,b G. De Meglio, a<br />

J. Garcia, c I. Collazo. d a South Florida Institute for Reproductive Medicine,<br />

Miami, FL; b The World Egg Bank, Phoenix, AZ; c Andrology, South Florida<br />

Institute for Reproductive Medicine, Miami, FL; d Embryology, South Florida<br />

Institute for Reproductive Medicine, Miami, FL.<br />

OBJECTIVE: Determine if the hyaluronan binding assay (HBA) can be<br />

used to identify patients with poor fertilization in In Vitro Fertilization.<br />

DESIGN: Retrospective analysis of 905 IVF cases from 2012 to 2014.<br />

MATERIALS AND METHODS: The same sperm used for IVF was<br />

analyzed for hyaluronan binding using the manufacturer’s instructions.<br />

Less than 80% binding was used as the cut-off as recommended by the manufacturer.<br />

100 sperm were counted (or 100 squares when there were few<br />

sperm) and percent binding was determined.<br />

RESULTS: Abnormal HBA results were found in 36.5% of patients. Fertilization<br />

rates for sperm that had low binding (¼80%; n¼<strong>21</strong>2) and poor fertilization (


their own controls. Primary outcomes included fertilization rates, number of<br />

embryos that reached an appropriate stage (2Bc or greater) for trophoectoderm<br />

biopsy (TEBx) and number of euploid embryos. T-tests of means and<br />

paired t-tests were used for statistical analysis.<br />

RESULTS: Average age of the patients in this study was 35.6 years +/- 5.2.<br />

Average day 2 estradiol (pg/ml) and follicle stimulating hormone (mIU/ml)<br />

values were 55.1 +/-56.5 and 6.75 +/- 2.87. Average total gonadotropin<br />

dosage (IU) and estradiol at trigger were 3648.6 +/- 144.6 and 2789 +/-<br />

1418.3, respectively. Approximately 19.8 +/- 9.3 oocytes on average were<br />

retrieved with a maturation percentage of 81%. Insemination fertilization<br />

rates were 69% +/- 23.8 and 75% +/- 13.8 (p-value of 0.28) for the ICSI<br />

cohort. The percentage of embryos available for TEBx for the insemination<br />

cohort was 32% and 38% for the ICSI cohort, with a p-value of 0.15 and RR<br />

0.84 (0.652-1.07). 52.8% of embryos biopsied resulting from ICSI and 47%<br />

of embryos from standard insemination were euploid, and paired t-test analysis<br />

showed no significance between the two groups, p ¼ 0.07.<br />

CONCLUSIONS: Standard insemination techniques and ICSI showed no<br />

difference in fertilization rates, and no difference in percent of chromosomally<br />

competent embryos. In the absence of standard indications for ICSI,<br />

there does not appear to be any benefit or harm in the use of ICSI for<br />

increased development of chromosomally competent embryos.<br />

P-579 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

MULTIPLE PRONUCLEAR ANALYSIS BY IMMUNOFLUORES-<br />

CENCE STAINING IN HUMAN EMBRYOS DERIVED FROM<br />

PATIENTS WITH POSITIVE ANTI-CENTROMERE ANTI-<br />

BODIES. M. Tokoro, H. Ohno, N. Aoyagi, M. Sonohara, M. Tsuiki,<br />

K. Ishihara, Y. Funagayama, Y. Kida, N. Fukunaga, Y. Asada. Asada Ladies<br />

Clinic Medical Corporation, Nagoya, Japan.<br />

OBJECTIVE: Anti-centromere antibody (ACA) is an anti-nuclear antibody<br />

(ANA), and specifically recognizes the centromere. Recently, several<br />

studies have reported that ACA have effects on oocyte maturation and embryo<br />

cleavage. Moreover, the rate of multiple pronuclear (MPN) formation<br />

is higher in patients with ACA. However, the reason for the high rate of<br />

MPN formation in ACA patients is not clear. In this study, we examined<br />

the cause of MPN formation in embryos derived from ACA patients, by performing<br />

immunofluorescence staining of MPN embryos and immature oocytes<br />

of ACA patients.<br />

DESIGN: Retrospective, nested case-control study.<br />

MATERIALS AND METHODS: A total of 1290 patients from our clinic<br />

were tested for ANA before oocyte retrieval from August 2014 to February<br />

<strong>2015</strong>. After ANA testing, patients were classified according to 3 groups;<br />

ACA patients (with only ACA), non-ACA patients (with ANA not included<br />

ACA) and non-ANA patients (without ACA and any other ANA). The rate of<br />

MPN formation after ICSI was compared in the 3 groups. MPN embryos with<br />

over 3 pronuclei (PN) and immature MI oocytes were used for research after<br />

informed consent was obtained from ACA patients. The embryos and oocytes<br />

were stained with H3K9me2 antibody to detect the female chromosome.<br />

Ethics Committee approval was obtained for the study.<br />

RESULTS: Among tested patients, 366 patients were positive for ANA,<br />

and 12 of 366 (3.3%) patients were positive for ACA. The average MPN formation<br />

rate per oocyte retrieval was 51.3% in ACA patients, 3.3% in non-<br />

ACA patients and 4.2% in non-ANA patients, significantly higher in ACA<br />

patients vs other groups (P0.05). Additionally,<br />

44.4% SNT blastocysts showed euploid chromosomes following aCGH<br />

test compared to 53.5% euploidy rate in the non-SNT control group<br />

(p>0.05).<br />

CONCLUSIONS: Using the current nuclear transfer-membrane fusion<br />

system, SNT seems to interfere with normal fertilization of human oocyte,<br />

however, it did not impose adverse impact to the pre-implantation embryonic<br />

development and its chromosomal complements.<br />

P-581 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DOES LASER ASSISTED HATCHING ON THAWED EMBRYO AF-<br />

TER VITRIFICATION IMPROVE IMPLANTATION<br />

RATE? S. A. Hebisha, a A. I. Ahmed, b H. N. Sallam, c M. S. Omran. d<br />

a Lecturer of Gynecology, Alexandria University - Faculty of Medicine, Alexandria,<br />

Egypt; b Obstetrics and gynecology MFM division, Perinatology<br />

research/NICHD/NIH/DHHS, Detroit, MI; c Gynecology, Alexandria University<br />

- Faculty of Medicine, Alexandria, Egypt; d Alexandria University -<br />

Faculty of Medicine, Alexandria, Egypt.<br />

OBJECTIVE: To evaluate the impact of laser assisted hatching on the implantation<br />

rate of vitrified-thawed embryos transferred after IVF/ICSI<br />

DESIGN: Prospective controlled study.<br />

FERTILITY & STERILITY Ò<br />

e305


MATERIALS AND METHODS: One hundred and eighty IVF/ICSI patients<br />

undergoing frozen (vitrified) thawed embryo transfer. All embryos<br />

were frozen on day three by vitrification using the open technique. Patients<br />

were divided into two groups: embryos from group A were subjected to laser<br />

assisted hatching of half thickness of one quarter of the zona two hours after<br />

thawing and two hours before transfer, while embryos from group B patients<br />

were transferred two hours after thawing without assisted hatching.<br />

RESULTS: Implantation rate was significantly higher in group A<br />

compared to group B ( 47.2 % vs 23.4 % respectively; P¼ 0.0041). Pregnancy<br />

rate was also significantly higher in group A compared to group B (61.1 % vs<br />

41.1 % respectively; P ¼ 0.031).<br />

CONCLUSIONS: Laser assisted hatching of half thickness of a quarter of<br />

the zona pellucida increases the implantation and pregnancy rates after the<br />

transfer of thawed embryos vitrified using the open technique.<br />

Effect of LAH on implantation rate.<br />

Group A Group B P value<br />

Total number of cases 90 90<br />

Number of embryo transferred 288 252<br />

Implanted embryos 136/288 59/252<br />

Implantation rate 47.22% 23.41% 0.0041*<br />

Pregnancies 55 37<br />

Pregnancy rate 61.1% 41.1% 0.031*<br />

* Statistically significant.<br />

P-582 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

SPERM DNA DAMAGE IMPAIRS EMBRYO DEVELOPMENTAL<br />

POTENTIAL AND INDUCES EPIGENETIC CHANGES IN THE RE-<br />

SULTING MOUSE OFFSPRING CONCEIVED THROUGH<br />

ICSI. Y. Li, H. Wang, S. Zhang, X. Huang. Reproductive Medical Center,<br />

The First Affiliated Hospital of Wenzhou Medical University, Wenzhou,<br />

China.<br />

OBJECTIVE: To evaluate the effects of sperm DNA damage on fertilization<br />

and embryonic development, and the question whether sperm DNA damage<br />

could introduce epigenetic changes was also investigated.<br />

DESIGN: Sperm DNA damage was triggered by freezing-thawing process,<br />

fertilization were performed by intracytoplasmic sperm injection<br />

(ICSI), in vitro developmental progression was observed and recorded.<br />

Meanwhile, methylation status of differentially methylated region (DMRs)<br />

for selected imprinted genes, a paternally imprinted gene H19 and a maternally<br />

imprinted gene SNRPN, was analyzed in midgestation mouse fetuses.<br />

MATERIALS AND METHODS: B6D2F1 (C57BL/6DBA/2) strain<br />

mice were used as oocytes and semen donors, sperm was freeze-thawed<br />

for ten times, then DNA damage extent was evaluated by sperm chromatin<br />

dispersion (SCD) assay. Oocytes (N¼524) were fertilized and in vitro developmental<br />

progression till to blastocyst formation was observed at different<br />

time stage. Afrer embryo transfer, midgestation mouse fetuses were<br />

collected, followed by DNA extraction, bisulfite-converted, and then amplified,<br />

cloned, sequenced and finally the methylation status of DMRs of H19<br />

and SNRPN was established by DNA methylation analysis.<br />

RESULTS: (1) The treated spermatozoa retained their fertilization potential<br />

and no statistically decrease existed in pronucleus formation (treated<br />

89.1% vs. control 91.5%, P > 0.05), however, sperm DNA damage effectively<br />

blocked fertilized oocytes from reaching 2-cell embryo stage and<br />

decreased the formation rate of morula and blastocyst in vitro (treated<br />

56.5% vs. control 85.3%, P < 0.05).(2) The DMRs for these two imprinted<br />

genes displayed an abnormal methylation status in the fetuses derived<br />

from DNA-damaged spermatozoa. For H19 DMR, there was a substantial increase<br />

in methylation level compared with control (treated 69.9% vs.control<br />

52.5%, P ¼0.0375), showing aberrant hypermethylation of the H19 DMR.<br />

Oppositely, SNRPN DMR showed a considerably reduction of methylation<br />

level, also reaching statistically significance (treated 31.5% vs.control<br />

53.0%, P ¼ 0.0451).<br />

CONCLUSIONS: Sperm DNA damage impairs in vitro embryo developmental<br />

potential, and it could induce epigenetic changes in the resulting F1<br />

mouse offspring, reinforcing the observation that the aberrant epigenetic<br />

modification is a contributory factor in the roles of sperm DNA damage<br />

played, which may associate the declined embryonic development.<br />

Supported by: Department of Health of Zhejiang Province(2014KYB340).<br />

P-583 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

GRANULOCYTE COLONY STIMULATING FACTOR (G-CSF)<br />

LEVEL IN FOLLICULAR FLUID IS A PROGNOSTIC FACTOR<br />

FOR EMBRYO DEVELOPMENTAL POTENTIAL IN IN-VITRO<br />

FERTILIZATION CYCLES. N. M. Chimote, a N. M. Nath, b<br />

B. N. Chimote. b a Embryology/Endocrinology, Vaunshdhara Clinic and Assisted<br />

Conception Centre, Nagpur, India; b Vaunshdhara Clinic and Assisted<br />

Conception Centre, Nagpur, India.<br />

OBJECTIVE: Local production of G-CSF in the human ovary, its expression<br />

in follicular luteal granulosa cells (GCs) as well as presence of higher<br />

concentration of GCSF in follicular fluid (FF) than in serum implies a potential<br />

autocrine/paracrine role of G-CSF within the follicular microenvironment.<br />

Therefore the objective of this study was to evaluate if FF-GCSF<br />

levels have a predictive value on the potential of embryo to develop to blastocyst<br />

stage.<br />

DESIGN: Prospective study of fresh non-donor, no male factor conventional<br />

IVF cycles (n¼44, mean age¼ 30.32 2.70 years) carried out from<br />

July 2014 to December 2014. Standard controlled ovarian stimulation with<br />

r-FSH and antagonist protocol was followed. Fertilization, cleavage and blastocyst<br />

formation rates were assessed. All cycles involved day5 blastocyst<br />

transfer.<br />

MATERIALS AND METHODS: For each patient FF-GCSF levels were<br />

estimated in fluid pooled only from lead follicles (R 18 mm in diameter)<br />

from which oocytes had been retrieved. Levels were reported as a measure<br />

of total protein content of FF. Cycles were divided into Low FF-GCSF group<br />

(n¼22, % 10.33 pg/mg protein) and High FF-GCSF group (n¼22, >10.33<br />

pg/mg protein) depending on the 50th centile value.<br />

RESULTS: Fertilization rate (71.33 0.03 vs. 50.91 0.03%, p lt<br />

0.0001), Cleavage rate (68.53 0.03 vs. 49.09 0.02%, p lt 0.0001) and<br />

blastocyst formation rate (47.55 0.03 vs. 26.13 0.03%, p lt 0.0001)<br />

were significantly higher in High FF-GCSF group compared to Low FF-<br />

GCSF group despite non-significant differences in the total number of eggs<br />

retrieved (286 vs. 222, p¼0.13) in the two groups. Pearson r values for correlation<br />

of FF-GCSF with fertilization, cleavage and blastocyst formation<br />

rates were 0.44, 0.40 and 0.36 respectively. An ROC cutoff of gt12.05 ng/<br />

mg protein increased the likelihood of embryos with good developmental potential<br />

(AUCROC¼ 92.98%, Sensitivity¼ 90.91%, Specificity¼ 72.73%).<br />

Although clinical pregnancy rates were higher in high compared to low<br />

FF-GCSF group, the difference was not significant. However, since live birth<br />

data has yet to be calculated for all patients involved in this study, it would be<br />

interesting to observe the differences in live birth rate.<br />

CONCLUSIONS: Endogenously produced Granulocyte Colony Stimulating<br />

Factor (GCSF) level in pooled follicular fluid (FF) strongly correlates<br />

with fertilization, cleavage and blastocyst formation rate and may be a predictive<br />

marker of embryo developmental potential in IVF cycles.<br />

P-584 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DOES THE TIMING OF BLASTULATION PROGNOSTICATE RISK<br />

OF ANEUPLOIDY? C. R. Juneau, M. D. Werner, J. M. Franasiak,<br />

E. J. Forman, K. H. Hong, T. Molinaro, R. T. Scott. RMA, NJ, NJ.<br />

OBJECTIVE: It has been established the rate of aneuploidy is higher in<br />

embryos which blastulate more slowly (day 6) relative to those which blastulate<br />

within the normal time range (day 5). Interpreting these data is complicated<br />

by the fact that later blastulation is associated with increasing maternal<br />

age and declining ovarian reserve. Thus, it is unknown if, within a single<br />

cohort of embryos from a single treatment cycle, those embryos which blastulate<br />

on day 6 have an increased risk of being aneuploid. This study seeks to<br />

address that question.<br />

DESIGN: Retrospective cohort.<br />

MATERIALS AND METHODS: Patients in their first IVF cycle at a single<br />

center from 2010-2014 using comprehensive chromosome screening<br />

(CCS) were reviewed. Those patients who had a portion of their cohort biopsied<br />

on day 5 and the remaining biopsied on day 6 were included. Blastocyst<br />

assessment was made in the mid-afternoon of day 5 and subsequently on the<br />

morning of day 6, 15 hours later. Blastocysts were deemed appropriate for<br />

biopsy if there was evidence of cells herniating through an artificial opening<br />

in the zona made on day 3 of development. The rate of aneuploidy between<br />

day 5 and day 6 blastocysts was compared using a paired Wilcoxon signed<br />

rank test for non-parametric data. Data were stratified by the Society of Assisted<br />

Reproductive Technologies (SART) age groups, and the aneuploidy<br />

rates between day 5 and day 6 blastocysts were compared using a c2 test<br />

of proportions. Logistic regression was employed to control for confounders,<br />

e306 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


and a generalized estimating equation (GEE) model was created to control<br />

for patient specific differences.<br />

RESULTS: <strong>21</strong>43 blastocysts from 274 patients met inclusion criteria. The<br />

mean patient age was 34.3 3.8 years. 1563 blastocysts were biopsied on<br />

day 5 and 580 blastocysts were biopsied on day 6. In the entire population,<br />

when controlling for basal antral follicle count, follicle stimulating hormone,<br />

age, and IVF stimulation type using a GEE analysis, the odds of aneuploidy<br />

were increased in day 6 blastocysts (OR 1.38, 95% CI 1.01-1.89). However,<br />

in the paired analysis, the rate of aneuploidy per patient on day 6 was similar<br />

to that on day 5 (35.5% versus 29.9%, p¼0.2).<br />

CONCLUSIONS: Given the observed difference in the rate of aneuploidy<br />

between day 5 and day 6 is negated when the patient serves as their own control<br />

in this paired analysis, it is likely the risk of aneuploidy is intrinsic to the<br />

patient and not related to the time of biopsy. Thus, embryos that blastulate<br />

later maintain a potential for euploidy and should be screened.<br />

P-585 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

ANEUPLOIDY RATES IN EMBRYOS PRODUCED BY FERTILE<br />

COUPLES. D. Wells, a K. Ravichandran, b M. Alper, c J. Jain, d<br />

A. Penzias, e C. A. Benadiva, f P. Colls, b M. Konstantinidis, b S. Munne. b<br />

a Reprogenetics, Oxford, United Kingdom; b Reprogenetics, Livingston, NJ;<br />

c Boston IVF, Waltham, MA; d Santa Monica Fertility, Santa Monica, CA;<br />

e Boston IVF / Harvard Medical School, Waltham, MA; f University of Connecticut,<br />

Farmington, CT.<br />

OBJECTIVE: To determine whether the high aneuploidy rates routinely<br />

seen in embryos of patients undergoing infertility treatments are also typical<br />

of couples without fertility problems.<br />

DESIGN: Examination of aneuploidy rates observed during preimplantation<br />

genetic diagnosis (PGD) and preimplantation genetic screening (PGS)<br />

cycles.<br />

MATERIALS AND METHODS: 96 fertile couples underwent PGD for a<br />

single gene disorder. Additionally, PGS was employed in order to enhance<br />

the likelihood of transferring a chromosomally normal embryo. A total of<br />

597 blastocyst stage embryos, derived from these couples, were biopsied<br />

and subjected to gene testing and also comprehensive chromosome screening<br />

using microarray-CGH (aCGH). During the same period of time 31,137 embryos<br />

were produced by couples undergoing assisted reproductive treatments<br />

(ARTs). Once again, embryos were subjected to aCGH at the blastocyst stage<br />

to assist in the selection of viable embryos for transfer. Information for the<br />

study was extracted from the eIVF database (Dallas, TX).<br />

RESULTS: The aneuploidy rate for blastocysts from patients of proven<br />

fertility, undergoing ART in order to generate embryos for PGD of a gene disorder,<br />

was found to be 49% (mean maternal age 33.5). This compares to an<br />

aneuploidy rate of 47% for embryos from infertile/subfertile patients undergoing<br />

PGS (mean age 34.7). These aneuploidy rates are not significantly<br />

different (P¼0.30, Chi square with Yates correction).<br />

CONCLUSIONS: It has long been known that aneuploidy is extremely<br />

common in human preimplantation embryos generated during ARTs. However,<br />

it is unknown whether such abnormalities represent part of the infertile<br />

pathology of the parents. This study assessed a large number of embryos using<br />

a well-validated chromosome screening method, providing robust data on<br />

aneuploidy rates. No difference in the incidence of aneuploidy was observed<br />

in embryos from typical infertile/subfertile patients compared to those from<br />

patients who had previously conceived naturally. Thus, it can be concluded<br />

that in the great majority of cases the high frequency of aneuploidy in embryos<br />

produced using ART is not related to the patient’s underlying infertility.<br />

As all patients in the current underwent ART, we were unable to address the<br />

question of whether the methods used for infertility treatment influence the<br />

incidence of aneuploidy.<br />

P-586 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

A MITOCHONDRIAL D LOOP VARIANT ASSOCIATED WITH<br />

REDUCED RISK OF EMBRYONIC ANEUPLOIDY. M. Olcha,<br />

X. Tao, Y. Wang, T. Xing, Y. Zhan, J. M. Franasiak, R. T. Scott,<br />

N. R. Treff. RMA, NJ, NJ.<br />

OBJECTIVE: Mitochondria are integral to oocyte and embryo energy production,<br />

spindle formation, and chromosomal segregation. Single nucleotide<br />

variants within the mitochondrial genome have been associated with susceptibility<br />

to many disease processes, including autoimmune and inflammatory<br />

processes. These variants could influence functions essential to early embryo<br />

development. The aim of this study was to assess if any association existed<br />

between rates of embryonic aneuploidy and sequence variants in the mitochondrial<br />

genome.<br />

DESIGN: Retrospective genetic association study.<br />

MATERIALS AND METHODS: All patients underwent IVF at a single<br />

center with comprehensive chromosomal screening. 469 patients were identified<br />

as either being young with atypically high embryonic aneuploidy rates<br />

or older with low aneuploidy rates. Patient DNAwas obtained from an on-site<br />

repository and used for analysis. Mitochondrial DNA was amplified with Takara<br />

LA PCR KitsÔ and mitochondria specific primers. Whole genome<br />

sequencing was accomplished using an Ion PGMÔ system utilizing Ion<br />

318Ô Sequencing Chips in order to achieve 20x depth reads. Individual variants<br />

were identified by comparing sequence reads to the Cambridge reference<br />

sequence. Patients were also assigned a haplogroup using the<br />

HaploFind software (https://haplofind.unibo.it). Individual variants and<br />

mitochondrial haplogroup frequencies were calculated and compared between<br />

the two study populations using logistical regression and chi squared<br />

contingency tables.<br />

RESULTS: Sequencing revealed 1,475 SNPs amongst 469 tested patients.<br />

A variant at position 16,390 (G>A) was significantly more common in patients<br />

with low aneuploidy rates (p.0.0008), which survived multiple test<br />

correction. Additionally, 443 individuals were assigned into 1 of 9 known European<br />

haplogroups (H, I, J, K, T, U, V, W, X) and analyzed in this study. The<br />

remaining patients were grouped as other. Haplogroup H contained the<br />

largest proportion of patients (41.4%). No association was found between<br />

aneuploidy rate and haplogroup inheritance.<br />

CONCLUSIONS: Although prior studies suggested an association between<br />

haplogroup variants and the incidence of oocyte chromosomal errors<br />

the present study found no association with a significantly larger sample size.<br />

However, with the availability of sequence data for the entire mitochondrial<br />

genome a significantly associated single nucleotide variant was identified in<br />

patients with lower risk of embryonic aneuploidy. The variant is located<br />

within the D-loop, a non-coding region of the mitochondrial genome<br />

involved with replication and transcription. It is possible that this variant confers<br />

a reproductive advantage given its distinct location. Additional studies<br />

need to be performed in order to further investigate the role of this variant<br />

in embryo development.<br />

P-587 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

TOO SLOW? WE DON’T THINK SO. OUTCOMES RELATIVE TO<br />

EMBRYO DEVELOPMENT IN A BUSY OOCYTE CRYOPRESER-<br />

VATION (OC) PROGRAM. S. Druckenmiller, F. Licciardi, P. Labella,<br />

M. Clarke-Williams, N. Seta, D. H. McCulloh, N. Noyes. NYU Medical<br />

Center, New York, NY.<br />

OBJECTIVE: OC is now a mainstream ART, frequently used to bank<br />

donor oocytes or defer fertility. To date, >2,000 babies have been born as<br />

a result of OC. Thus, we analyzed the specifics of post-fertilization (fert) embryo<br />

development in order to glean outcome differences when using autologous<br />

(Aut; older) vs. donor (OD; younger) frozen-thawed oocytes. We also<br />

compared OC blastocyst (BL) formation rates (BFR) to those of fresh-oocyte<br />

cycles.<br />

DESIGN: Retrospective and case control.<br />

MATERIALS AND METHODS: For OC, 154 Aut (139 pts) & 53 OD (no<br />

commercial-bank oocytes) thaw cycles completed from 2004-<strong>2015</strong> were reviewed.<br />

Cycle data was mined for: oocyte maturity, survival, 2-pronuclear<br />

(2PN) fert, development, embryo implantation (EI) & pregnancy. When assessing<br />

development, the highest stage & grade achieved before embryo<br />

transfer (ET) was used. BL were graded using Gardner’s criteria. When evaluating<br />

EI & pregnancy rates relative to embryo stage, only single ETs &<br />

multi-embryo ETs with same-stage embryos were considered. BFR in 286<br />

age-matched fresh Aut & 186 fresh OD cycles from 2011-2014 were<br />

computed & compared to OC thaw outcomes. Fisher’s exact was used for<br />

stats.<br />

RESULTS: For OC, 1,098 (7/cycle) Aut (mean age 383y; range 25-44y)<br />

& 411 (8/cycle) OD (mean age 275y; range 22-33y) oocytes were thawed,<br />

survived & achieved 2PN fert. The Table shows post-fert development. OC<br />

BFRs were not different when Aut (n¼428 ; 39%) & OD (n¼<strong>17</strong>3; 42%) rates<br />

were compared; however, within the BL-Stage groups, more BL expanded to<br />

Stage 3 & 4 in OD than Aut cycles (p< .02). Conversely, a greater % of BL<br />

only reached Stage 2 in Aut vs. OD thaws (p< .01). When comparing EI &<br />

pregnancy rates between Aut & OD relative to the day (D) of BL ET (i.e.<br />

D-5 vs. D-6/7), no statistical difference between ETs of similarly-staged<br />

BL formed on D-5 (24/53 {45%} vs. 16/28 {57%};p¼0.4) or D-6/7 (1/3<br />

{33%} vs. 1/1 {100%};p¼1) was noted; however, within the Aut & OD<br />

groups, EI was significantly greater with ETs of BL >Stage 1 (p< .04).<br />

FERTILITY & STERILITY Ò<br />

e307


BFRs for fresh Aut (959/1539; 59%) & OD (2578/3658; 71%) cycles were<br />

significantly higher than in age-matched frozen Aut & OD cycles (p28% peaked at morula or Stage-1 BL, both considered early ‘‘usable’’ embryos<br />

resulting in reasonable, albeit lower EI & pregnancy rates. When<br />

comparing BL maturity in Aut vs. OC cycles, OD oocytes achieved >BL<br />

expansion with higher EI & pregnancy rates, most likely reflecting younger<br />

oocyte age.<br />

Embryo development & pregnancy outcomes of OC thaws using Autol vs. OD<br />

oocytes.<br />

Autol OC<br />

(n¼1098)<br />

OD OC<br />

(n¼411)<br />

No. 2PN-Zygotes with No 81 (7%) 39 (9%) 0.2<br />

Division (1-cell)<br />

No. 2PN-Zygotes with<br />

433 (39%) 139 (34%) 0.04<br />

Cleavage Arrest<br />

No. Morula (highest stage 156 (14%) 60 (15%) 0.9<br />

achieved)<br />

No. BL - Stage 1 75 (18%) 22 (13%) 0.2<br />

No. BL - Stage 2 181 (42%) 33 (19%)


Using these SRM assays and aCGH, we performed blastocoel protein<br />

profiling and cytogenetic assessment on 14 embryos donated to research.<br />

Subsequently, we conducted statistical analysis using logistic regression to<br />

reveal any correlations between blastocoel protein profile, embryo<br />

morphology, ploidy status, gender, patient’s age.<br />

RESULTS: Using levels of a metabolic protein and the detection of a nuclear<br />

protein as predictors and embryo’s chromosomal status as class label,<br />

the statistical model was able to predict whether an embryo was euploid or<br />

chromosomally abnormal. This was achieved with 100% accuracy within<br />

the small cohort of embryos investigated. Statistical analysis did not show<br />

any association between blastocoel protein profile and any of the other parameters<br />

considered.<br />

CONCLUSIONS: The application of logistic regression analysis to the<br />

data obtained suggests that it may be possible to distinguish chromosomally<br />

normal embryos from those affected by aneuploidy using a targeted proteomics<br />

approach applied to blastocoel fluid. However, due to the small size<br />

of the sample population investigated and the retrospective nature of the analysis,<br />

further work will be essential to determine sensitivity and specificity of<br />

this promising method of preimplantation aneuploidy detection. If confirmed<br />

in larger studies, this minimally invasive proteomic approach could offer a<br />

novel methodology for the prediction of embryonic viability and developmental<br />

competence.<br />

Supported by: Institutional funding.<br />

P-591 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

SYMMETRY AT THE 4-CELL STAGE USING TIME-LAPSE<br />

IMAGING IS CORRELATED WITH EMBRYO<br />

ANEUPLOIDY. C. C. Shenoy, Z. Khan, C. Coddington, J. Jensen,<br />

G. S. Daftary, E. A. Stewart, D. Morbeck. Mayo Clinic, Rochester, MN.<br />

OBJECTIVE: Embryo selection is important for optimal ART outcomes.<br />

Embryo selection has improved over time and the advent of time-lapse<br />

(TL) monitoring offers increased information to aid in embryo selection.<br />

Nonetheless, TL timing parameters alone do not provide high sensitivity<br />

for determining embryo viability or ploidy. We sought to compare euploid<br />

prediction using TL timings with traditional morphometric parameters that<br />

are easily measured using a TL system.<br />

DESIGN: Embryos undergoing preimplantation genetic screening (PGS)<br />

that were cultured using TL monitoring were examined in this retrospective<br />

study.<br />

MATERIALS AND METHODS: All embryos from 2012-<strong>2015</strong> undergoing<br />

PGS with trophectoderm biopsy were included. The distance between<br />

the second and first polar body was determined (PBD). Zona pellucida thickness<br />

(ZPT) was measured at the pronuclear stage and at the 2-cell stage in<br />

four locations and averaged. Blastomere area was assessed at the 2- and 4-<br />

cell stages. Symmetry at the 2-cell stage was determined by percent difference<br />

between blastomeres (2cSY). Symmetry at the 4-cell stage was the<br />

percent difference between the smallest and largest blastomeres (4cSY). T-<br />

test was used to compare group means.<br />

RESULTS: Embryos (n¼182) from <strong>21</strong> patients were analyzed. Fouty five<br />

percent were euploid. Patient age ranged from 22-43 (avg¼34). The only variable<br />

that differed significantly between euploid and aneuploid embryos was symmetry<br />

at the 4-cell stage. Aneuploidy rates were 45.5% in the lowest symmetry<br />

quartile and 70.5% in the quartile with the most asymmetry at the 4-cell stage.<br />

CONCLUSIONS: Poor symmetry at the 4-cell embryo stage is more predictive<br />

of aneuploidy than TL parameters, suggesting that embryo selection<br />

models using time-lapse parameters should incorporate cleavage-stage<br />

morphology.<br />

Morphologic and TL Differences Between Euploid and Aneuploid Embryos.<br />

Euploid Aneuploid p value<br />

PBD (mm) 43.2 24.8 43.6 26.8 0.91<br />

ZPT at PN stage (mm) 18.0 2.7 18.3 2.4 0.45<br />

ZPT at 4-cell stage (mm) <strong>17</strong>.9 2.6 <strong>17</strong>.9 2.4 0.81<br />

2cSY (% difference) 11.3 7.5 10.0 8.2 0.28<br />

4cSY (% difference) 26.6 11.4 31.3 13.4 0.01*<br />

t2 (h) 26.4 3.7 26.8 2.8 0.41<br />

t5 (h) 51.2 8.3 52.2 6.9 0.39<br />

tsb (h) 101.8 9.4 103.7 7.9 0.15<br />

tb (h) 108.9 9.7 110.7 8.9 0.27<br />

cc2 (h) 11.1 3.1 11.6 2.9 0.20<br />

s2 (h) 2.2 4.2 1.1 2.4 0.06<br />

P-592 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

FUNCTIONAL CHARACTERIZATION OF CULTURED HUMAN<br />

GRANULOSA CELLS IN SERUM FREE CULTURE<br />

SYSTEM. Y. Wu, a D. F. Albertini, b Q. Wang, a D. H. Barad, c<br />

V. A. Kushnir, d E. Lazzaroni-Tealdi, a N. Gleicher. c a Center for Human<br />

Reproduction, New York, NY; b Center for Human Reproduction & University<br />

of Kansas Medical Center, New York, NY; c Center for Human Reproduction<br />

& Foundation for Reproductive Medicine, New York, NY; d Center for<br />

Human Reproduction & Wake Forest University, New York, NY.<br />

OBJECTIVE: In vitro culture of human granulosa cells (GCs) is an important<br />

research technique for reproductive cell biology and endocrinology<br />

studies. However, traditional culture protocols, involving supplementation<br />

of serum to culture medium, result in luteinization of GCs, and changes their<br />

cell functions during culture. Establishment of a GC culture system that does<br />

not cause luteinization would, therefore, be important.<br />

DESIGN: Prospective laboratory study.<br />

MATERIALS AND METHODS: GCs were collected from follicular<br />

fluid after oocytes retrievals. After washed twice by DPBS to remove<br />

blood contamination,GCs were seeded into 12-well plates at density of<br />

10x105/ml in DMEM/F12 containing 10% FBS, 2mg/ml of HSA, 2<br />

mM glutamine and 1x Insulin-Transferin-Selenium X and cultured for<br />

eight hours at 37 C with 5% CO2 to allow cell attachment. Then the culture<br />

medium was replaced by serum free medium (DMEM/F12 with same<br />

supplementation but without FBS). After overnight culture, medium was<br />

replaced once more. GCs were then cultured for another 96 hours<br />

with or without 50ng/ml FSH supplementation. To evaluate GC functions<br />

after culture, we examined aromatase and FSH receptor (FSHR)<br />

mRNA expression by real-time PCR, cell proliferation by MTT assay,<br />

apoptosis by DAPI staining and estradiol (E2) production by analyzing<br />

medium hormone concentration after adding testosterone (1mM) to the<br />

medium.<br />

RESULTS: After 96 hours culture, FSH treatment significantly induced<br />

aromatase and FSHR expression in GCs (P


test for numerical variables and chi square test for categorical variables were<br />

employed as statistical analysis. P-value lower than 0.05 was considered as<br />

statistically significant.<br />

RESULTS: The incidence of embryos that were generated from fresh<br />

donated oocytes 73.28% (192/262) and 26.72% (70/262) from thawed<br />

donated oocytes. PNf asyn rate was 14.5% (38/262) being slightly more<br />

frequent in thawed oocytes (<strong>17</strong>.14% vs 13.54%) without any significant difference<br />

(p¼0.42). The average duration between the disappearance of the<br />

first pronuclei and the second was 0.33h (IC: 0.<strong>17</strong>-0.34). The top quality embryo<br />

rate (A category from ASEBIR) in those embryos showing PNf asyn<br />

was 34.<strong>21</strong>% (13/38), and in synchronical was 48,86%(86/<strong>17</strong>6); p¼0.10. In<br />

Embryos with asynchronical fading, tPNf were significantly later than in<br />

those with synchronical (24.664.16 vs 26.564.75, p


component analysis (PCA) followed by a partial least square discrimination<br />

analysis (PLS-DA), combined with positive correlation to the blastocyst<br />

group.<br />

RESULTS: The PLS-DA analysis demonstrated a significant difference<br />

between groups in which seven metabolites were differently expressed.<br />

The differently expressed metabolites were phosphocholine arachidonic<br />

C36:2 (Corr. 0.8588), lyso phosphocholine arachidonic C18:0 (Corr.<br />

0.83102), GCKR** (Corr. 0.81856), phosphocholine arachidonic C36:3<br />

(Corr. 0.79307), phosphocholine arachidonic ether C30:2 (Corr. 0.74401),<br />

Glutamine (Corr. 0.74384), and Glutamine/Glycine (Corr. 0.7373).<br />

CONCLUSIONS: The day three culture medium quantitative secretomics<br />

provides important information concerning the embryo developmental<br />

competence, therefore may be an useful tool for the selection of embryos<br />

to be culture until day five improving the implantation and single pregnancy<br />

rates.<br />

Supported by: Coordination for the improvement of Higher Level or Education<br />

Personnel - (Capes-Brazil).<br />

P-597 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

MITOCHONDRIAL DNA CONTENT AS A VIABILITY SCORE IN<br />

HUMAN EUPLOID EMBRYOS: LESS IS BETTER. A. Diez Juan, a<br />

C. Rubio, b C. Marin, a S. Martinez, c P. Diaz-Gimeno, d M. Riboldi, e<br />

N. Al-Asmar, f D. Valbuena, g C. Simon. h a Product Innovation, Igenomix,<br />

Rocafort, Spain; b Igenomix SL; FIVI/INCLIVA, Paterna, Spain; c Product<br />

Innovation, Igenomix, Paterna, Spain; d Fundacion IVI / INCLIVA, Paterna<br />

(Valencia), Spain; e Igenomix Brasil, Sao Paulo, Brazil; f IviGen, Miami,<br />

FL; g Igenomix S.L., Paterna, Spain; h FIVI/INCLIVA, Valencia University,<br />

Stanford University, Paterna, Spain.<br />

OBJECTIVE: To investigate the clinical relevance of mitochondrial<br />

mtDNA content as a viability score in human euploid embryos.<br />

DESIGN: Retrospective analysis of mtDNA content of transferred euploid<br />

embryos.<br />

MATERIALS AND METHODS: Setting: Reproductive genetics laboratory.<br />

Patients: Single embryo transfer in 270 patients who underwent preimplantation<br />

genetic screening (205 day-3 blastomere biopsies, and 65 day-5<br />

trophectoderm biopsies), and 10 patients with double embryo transfer<br />

(male-female). Samples: DNA from single blastomere or trophectoderm biopsy<br />

amplified using the Sureplex DNA amplification system. Euploid embryos<br />

were transferred at the blastocyst stage and an aliquot of their<br />

amplified DNA was used for this study. Relative amounts of nuclear DNA<br />

(nDNA) and mtDNA were determined by quantitative real-time (RT) PCR.<br />

The ratio of mtDNA/nDNA was classified as the mitochondrial score (Ms),<br />

and used as an indicator of the mitochondrial copy number per cell. MtDNA<br />

copy number was compared using a non-parametric test, Wilcoxon’s rank<br />

sum test and Shapiro-Wilk test were used to normalize. Ms scores are presented<br />

based on quartiles.<br />

RESULTS: High mtDNA copy number in euploid embryos is indicative of<br />

lower embryo viability and implantation. Using the normalized mtDNA content,<br />

we created the mitochondrial score or Mitoscore (Ms). Day-3 embryos<br />

with less than 34 (MsA) had an implantation rate (IR) of 59% (n¼51); those<br />

with 34 to 52 (MsB) had an IR of 44% (n¼52); those with 52 to 97 (MsC) had<br />

an IR of 42% (n¼50); and those with levels > 97 (MsD) had an IR of 25%<br />

(n¼52). Embryos with Ms>160 (n¼22) never implanted. Day-5 with less<br />

than 18.19 (MsA) had an IR of 81%; those with 18.19 to 24.15 (MsB) had<br />

an IR of 50% (n¼16); those with 24.15 to 50.58 (MsC) had an IR of 62%<br />

(n¼16); and those with levels > 50.58 (MsD) had an IR of 18% (n¼<strong>17</strong>). Embryos<br />

with levels > 60 (n¼7) never implanted.<br />

CONCLUSIONS: An increased amount of mtDNA in euploid embryos is<br />

related to poor implantation potential and may be indicative of reduced metabolic<br />

fuel during oocyte maturation. We are implementing Ms in our PGS<br />

platform to prospectively-analyze its clinical relevance.<br />

P-598 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

WITHDRAWN<br />

P-599 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

PROFILING OF GENDER-SPECIFIC GENE EXPRESSION IN FE-<br />

MALE AND MALE PRIMORDIAL GERM CELLS IN<br />

MICE. T. Kono, A. Sakashita, Y. Kawabata, Y. Jincho, H. Kobayashi.<br />

Bioscience, Tokyo University of Agriculture, Tokyo, Japan.<br />

OBJECTIVE: After complete migration of primordial germ cells (PGCs)<br />

into the genital ridges by E13.5 in mice, they undergo a gender-specific fate:<br />

Male and female PGCs undergo mitotic arrest and meiosis, respectively.<br />

However, differentiation of PGCs on a gender-specific basis remained poorly<br />

understood. The aim of this study was to gain further insight into the genderspecific<br />

features of PGCs in mice.<br />

DESIGN: We performed RNA-Seq analysis of E13.5 female and male<br />

PGCs by using a next-generation sequencer.<br />

MATERIALS AND METHODS: PGCs of E13.5 fetuses were isolated using<br />

a FACSAria II Cell Sorter (BD Bioscience). Total RNA of 1 104 PGCs<br />

was isolated, and cDNA synthesis and pre-amplification were performed using<br />

10 ng of total RNA. Sequencing libraries were created by NEBNext Ultra<br />

DNA Library Prep Kit for Illumina, according to the library construction protocol<br />

(New England BioLab). Indexed libraries were pooled and sequenced<br />

using an Illumina Hiseq 2500 Sequencer. ChIP-seq analysis was performed<br />

for H3K4me3 and H3K27me3. RNA-Seq reads were aligned to the mouse<br />

genome (mm10, Genome Reference Consortium Mouse Build 38) by using<br />

a CLC Genomics Workbench (CLC bio). Aligned reads were subsequently<br />

assembled into transcripts guided by reference annotation (mm10, UCSC<br />

gene annotation).<br />

RESULTS: We identified gender-specific transcripts: 651 and 428 were<br />

specific expressed (>2-fold and P


and significance differences between implantation rates of HIGH and LOW<br />

embryos to the previous analysis (p¼0.019, OR¼1.72CI95% 1.10-2.70).<br />

CONCLUSIONS: Our study presents,to our knowledge, the largest set of<br />

transferred embryos after time lapse analysis using Eeva. It has demonstrated<br />

that embryo selection using the classification provided by Eeva is related<br />

with reproductive outcome. The observed relationship with the implantation<br />

potential reflects a direct link between the parameters provided by the automatic<br />

system and embryo quality and can be used for embryo selection in a<br />

clinical setting. Moreover, our multivariable analysis demonstrate that the<br />

relationship between EEVA and implantation potential is robust and independent<br />

of other clinical variables such oocyte quality or embryo morphology.<br />

P-601 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

TIME LAPSE KINETIC MARKERS ENABLE EARLY IDENTIFICA-<br />

TION OF DEVELOPMENTALLY DELAYED EUPLOID<br />

BLASTOCYSTS. B. Hixon, a J. C. Parks, a J. M. Stevens, b<br />

W. B. Schoolcraft, c M. Katz-Jaffe. c a National Foundation for Fertility<br />

Research, Lone Tree, CO; b Fertility Labs of Colorado, Lone Tree, CO; c Colorado<br />

Center for Reproductive Medicine, Lone Tree, CO.<br />

OBJECTIVE: The introduction of time lapse (TL) technology into the IVF<br />

laboratory has allowed for a more in depth and uninterrupted study of embryo<br />

development. A viable blastocyst requires successful completion of specific<br />

developmental events and all 23 pairs of chromosomes. Research investigating<br />

TL data have indicated that competent blastocysts are identifiable<br />

by early cleavage stage TL kinetic markers. This study examined embryonic<br />

development and TL kinetic markers of euploid blastocysts, and their implantation<br />

outcome.<br />

DESIGN: Research Study.<br />

MATERIALS AND METHODS: Normally fertilized zygotes from infertility<br />

patients (mean maternal age ¼ 37.0 years) were cultured in the EmbryoScopeTM<br />

(Vitrolife) to the blastocyst stage. A trophectoderm biopsy was<br />

performed on either D5 or D6 of development for comprehensive chromosome<br />

screening using qPCR (RMA-NJ) or aCGH (Illumina), to identify<br />

euploid blastocysts (n¼272). Biopsied blastocysts were vitrified to be<br />

warmed at a later date for a frozen embryo transfer. Statistical analysis<br />

included Student’s t-test and Chi-square test for independence, significance<br />

at P


CONCLUSIONS: While ovarian stimulation can negatively impact uterine<br />

receptivity (3), this study addresses the effects of stimulation more<br />

directly on implantation of euploid embryos. Use of more gonadotropin<br />

and retrieval of more oocytes were both associated with decreased IR. Higher<br />

gonadotropin doses could indicate a poorer ovarian response or that excessive<br />

use of gonadotropin is detrimental. Higher numbers of oocytes retrieved<br />

could indicate a more robust ovarian response or aggressive stimulation/<br />

retrieval. Higher IR occurred when FhMG was larger, suggesting that LHlike<br />

activity improves oocyte/embryo quality.<br />

References:<br />

1. Hodes-Wertz, et al. Fertil. Steril. (<strong>2015</strong>) 103:947.<br />

2. Hong et al., Fertil. Steril. (2014) 102:41.<br />

3. Shapiro et al., Fertil. Steril. (2011) 96:344<br />

EMBRYO CULTURE<br />

P-604 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DIPEPTIDE GLUTAMINE FUNCTIONS AS AN ORGANIC OSMO-<br />

LYTE IN MOUSE PREIMPLANTATION EMBRYOS BUT DOES<br />

NOT SUPPORT BLASTOCYST HATCHING AT THE SAME RATE<br />

AS INDIVIDUAL COMPONENT AMINO ACIDS. J. E. Swain. a,b a Fertility<br />

Lab Sciences, Englewood, CO; b Colorado Center for Reproductive<br />

Medicine, Lone Tree, CO.<br />

OBJECTIVE: Dipeptide glutamine is included in human embryo culture<br />

media to reduce ammonia production. However, dipeptides have not been<br />

examined to determine if they can fulfill known physiologic roles of individual,<br />

non-dipeptide amino acids. Our objective was to examine whether dipeptide<br />

alanyl-glutamine (ala-glu) can act as an organic osmolyte in mammalian<br />

embryos as efficiently as its component individual amino acids.<br />

DESIGN: Prospective experimental study.<br />

MATERIALS AND METHODS: Thawed 1-cell mouse embryos were<br />

cultured in groups of 10 in 500ul in 1 of 4 media treatments for 96h in 6%<br />

CO2, 5% O2, 89% N2. Positive controls (Pos Con) consisted of amino acidfree<br />

Human Tubal Fluid (HTF; 280mOsm). Treatment media included HTF<br />

with elevated NaCl to increase osmolality (320mOsm) and either 1mM dipeptide<br />

ala-glu or 1mM alanine + 1mM glutamine (ala+glu). Negative controls<br />

(Neg Con) consisted of 320mOsm with no amino acids. Blastocyst development<br />

was assessed and sizes determined using Cronus software. Data were<br />

collected over 5 replicates and analyzed using ANOVA and Tukey analysis.<br />

RESULTS: Pos Con resulted in 79.0% 5.9 blastocyst development while<br />

Neg Con media resulted in significantly lower rates of blastocyst formation<br />

(20.1% 7.3), p


patients. In contrast, the arrest rate between day 3 and the blastocyst stage<br />

increased dramatically with age. This supports the use of extended culture<br />

to optimize embryo selection. The fact that the blastulation rates were quite<br />

high in even the oldest patient group suggests that clinicians and embryologists<br />

should not forego extended culture out of concern that there will be a<br />

lack of blastulation. This allows the older high risk age groups access to<br />

advance diagnostics such as CCS.<br />

Embryonic performance in extended culture by age.<br />

Age


second mitosis. Time is calculated by hours after insemination. Data were<br />

analyzed using ANOVA and c2 test.<br />

RESULTS: Euploidy is not correlated with the morphokinetic parameters<br />

used, either in BB group or TB group (Table 1). Interestingly, multinucleation<br />

at 2-cell stage did not increase aneuploidy rates in either Day 3 or<br />

Day 5 biopsies. Euploidy rate was significantly higher in the TB group<br />

(33%) compared with BB group (16%, p¼0.005).<br />

CONCLUSIONS: Our results suggest that euploidy (from either BB or<br />

TB) cannot be predicted with the selected morphokinetic parameters using<br />

time-lapse microscopy. Future investigation should focus on different morphokinetic<br />

parameters to predict ploidy status.<br />

Table 1. The Mean Differences of Morphokinetic Parameters within Groups.<br />

Biopsy stage<br />

Trophectoderm<br />

Biopsy<br />

Trophectoderm<br />

Biopsy<br />

Blastomere<br />

Biopsy<br />

Blastomere<br />

Biopsy<br />

Parameters /ploidy Aneuploidy Euploidy p-value Aneuploidy Euploidy p-value<br />

tPB2 3.61.5 3.<strong>21</strong>.2 0.14 4.83.4 4.73.3 0.81<br />

tPNa 8.51.9 8.12.1 0.40 10.83.9 10.53.5 0.63<br />

tPNf 24.53.7 23.63.1 0.27 24.66.5 24.24.6 0.69<br />

tP2 11.24.8 11.14.9 0.49 7.86.5 7.55.8 0.67<br />

tSB 103.79.7 101.56.6 0.26 99.98.9 100.611.6 0.75<br />

Multinucleation<br />

at 2 cell stage<br />

25% 39% 0.24 37% 33% 0.62<br />

Supported by: The authors wish to thank Dr.MH. Fakih, Dr.FN. Shamma,-<br />

Dr.A.Hammoud and Dr. S.Dogan for data mining support.<br />

P-610 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

RECOMBINANT ALBUMIN AS A POTENTIAL ALTERNATIVE TO<br />

BLOOD-DERIVED ALBUMIN FOR USE IN BLASTOCYST CUL-<br />

TURE: A SIBLING EMBRYO STUDY. M. Murakami, a<br />

D. K. Gardner, b S. Mizumoto, a K. Tanaka, a A. Egashira, a T. Kuramoto. a<br />

a Kumamoto Women’s Clinic, Fukuoka, Japan; b University of Melbourne,<br />

Parkville, Australia.<br />

OBJECTIVE: Replacing human serum albumin (HSA) with lower<br />

amounts of recombinant human albumin (rHA) during embryo culture yields<br />

good-quality blastocysts. In this study, the efficacy of using completely<br />

defined media containing rHA for all ART protocols was evaluated using a<br />

sibling embryo model to minimize experimental variation between patients.<br />

Further, we examined the effects of the albumin source on pregnancy rates<br />

following vitrification.<br />

DESIGN: Prospective sibling embryo study.<br />

MATERIALS AND METHODS: Patients undergoing elective embryo<br />

cryopreservation between <strong>October</strong> 2014 and April <strong>2015</strong> were evaluated.<br />

Forty-one patients (aged 35.5 0.6 years) who underwent oocyte<br />

retrieval after hyperstimulation and IVF/ICSI treatment using rHA-containing<br />

defined media and had R4 2PN oocytes at 18 h after insemination<br />

were included in the study. In total, 362 sibling 2PN oocytes from<br />

these patients were randomized for side-by-side culture in G1/G2 media<br />

containing either 0.5 mg/mL rHA or 5 mg/mL HSA. All good-quality<br />

blastocysts (R3BB) were vitrified by day 6. Morphokinetics were evaluated<br />

by time-lapse microscopy in certain patients. Data was analyzed<br />

for a single ET of vitrified blastocysts performed by the end of March<br />

<strong>2015</strong>.<br />

RESULTS: Percentages of good-quality day 3 embryos (6- to 10-cell embryos<br />

with %25% fragmentation) and good-quality blastocysts were<br />

similar in rHA (55.2% and 18.8%, respectively) and HSA (56.9% and<br />

26.5%) groups. The time taken to develop into full blastocysts (with the<br />

blastocoel completely filling the embryo) from insemination was not<br />

different between rHA (108.6 5.6 h) and HSA (100.9 3.1 h) groups.<br />

There were 12 and 11 cycles of vitrified blastocyst ET in the rHA and<br />

HSA groups, respectively. All blastocysts survived after warming (1 embryo/cycle).<br />

Clinical pregnancy rate/ET was 75.0% and 45.5%, and<br />

ongoing pregnancy rate/ET was 75.0% and 18.2% (P < 0.05) in rHA<br />

and HSA groups, respectively.<br />

CONCLUSIONS: We demonstrated the feasibility of using chemically<br />

defined ART systems with rHA to yield good-quality blastocysts, and<br />

subsequently, high pregnancy rates. Studies involving optimization of<br />

rHA concentration and long-term follow-up studies with more participants<br />

are required to validate the efficacy of the procedures. Using<br />

rHA will help eliminate variation and risks associated with blood-derived<br />

products.<br />

P-611 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DOES THE TRANSITION FROM STANDARD INCUBATOR TO A<br />

TIME-LAPSE IMAGING CHAMBER IMPACT EMBRYONIC<br />

DEVELOPMENT AND NEONATAL OUTCOMES? R. W. Goldberg,<br />

P. K. Gill, J. M. Goldberg, N. Desai. Obstetrics & Gynecology / Women’s<br />

Health Institute, Cleveland Clinic, Beachwood, OH.<br />

OBJECTIVE: Time lapse imaging chambers offer a unique environment<br />

for uninterrupted culture while still allowing visualization of embryo development.<br />

The altered culture environment may influence cell division and progression<br />

to the blastocyst stage. Ability to identify abnormal cleavage<br />

patterns as well as specific kinetic data with time lapse systems may further<br />

aid in embryo selection and ultimately pregnancy outcomes. This study compares<br />

embryonic development, clinical and neonatal outcomes after transition<br />

from standard box type incubators to time lapse.<br />

DESIGN: This study retrospectively analyzed IVF cycle data from 2011<br />

when embryo culture was performed in Forma standard incubators (SI) to<br />

2013 when all patient embryos were exclusively cultivated in EmbryoScope<br />

time lapse imaging chambers (TL). For analysis the data was further stratified<br />

by age and restricted to patients using autologous oocytes and having a fresh<br />

transfer.<br />

MATERIALS AND METHODS: Zygotes were individually cultured in<br />

Global medium (LifeGlobal; Guilford, CT) with 10% SPS (SAGE; Trumbull,<br />

CT) at 37o C with 6% CO 2 . In the SI group, dishes were removed once a day<br />

for embryo evaluation. Embryos in the TL group underwent continuous uninterrupted<br />

culture under low oxygen tension. Embryo development was<br />

monitored by viewing time lapse videos. Clinical pregnancy was determined<br />

by the presence of a fetal heart on ultrasound. Implantation rate was calculated<br />

based on gestational sacs. Cycle data and neonatal outcomes were<br />

compared using chi square analysis and the Student’s T-test. P-values of<br />

37<br />

TL<br />

N¼222<br />

p-value<br />

Standard<br />

N¼123<br />

TL<br />

N¼92<br />

p-value<br />

Embryos Transferred 2.06 2.00 0.332 2.69 2.33 0.0014<br />

0.60 0.52<br />

0.80 0.83<br />

Clinical Pregnancy (%) 53.6 56.3 0.564 30.1 34.8 0.465<br />

(126/235) (125/222)<br />

(37/123) (32/92)<br />

Implantation (%) 35.2 42.7 0.0192 15.7 <strong>21</strong>.0 0.113<br />

(<strong>17</strong>0/483) (190/445)<br />

(52/331) (45/<strong>21</strong>4)<br />

Blastocyst Formation (%) 48.7 62.9


Mono media is a single formulation. Mono formulations require less manipulation<br />

and are less expensive. Both formulations have attained good outcomes,<br />

but an adequately powered assessment including usable<br />

blastulation rates (USBR), ploidy risk, and SIR is lacking.<br />

DESIGN: Paired RCT.<br />

MATERIALS AND METHODS: Patients with normal ovarian reserve<br />

were recruited. A paired design allowed each patient to serve as their own<br />

control, eliminating many confounding variables seen in prior studies. After<br />

confirming fertilization, patients’ zygotes were randomized (1:1 ratio) to<br />

either Seq media (Quinn’s Advantage Cleavage Medium, Sage then Blast<br />

Assist, Origio) or Mono media (Continuous Single Culture; Irvine Scientific).<br />

Each culture system used separate incubators. Assessed endpoints for<br />

all embryos included USBR, blastulation timing (arrest vs day 5 vs 6) and<br />

ploidy status. Paired euploid blastocyst transfers, one from each group,<br />

were performed. DNA fingerprinting of concepti was used as needed to<br />

link each embryo to a definitive outcome. SIR was defined as the presence<br />

of a fetal heart beat at 8-9 weeks gestation. Statistical analysis performed<br />

via McNemar’s Chi Square and Wilcoxon sum rank tests.<br />

RESULTS: 186 patients had their 2PN embryos (N¼2257) randomized to<br />

each culture system. Seq media had a higher blastulation rate then Mono media<br />

(p¼0.001, 55.5% vs. 46.0%). No differences were found in the day of<br />

blastulation (p¼0.4063) or in the aneuploidy rate (p¼0.5518). Of the 168 patients<br />

who had euploid blastocysts suitable for transfer, 126 completed a<br />

paired embryo transfer and 42 had a SET. Amongst the SETs, the SIR was<br />

equivalent (p¼1.0). Of the 126 double embryo transfers, 36 patients had<br />

one fetal heart beat at discharge, however there was no statistical difference<br />

in the likelihood of implantation between groups (p¼0.8642).<br />

CONCLUSIONS: This is the first randomized controlled trial to systematically<br />

examine paired euploid transfers of sibling zygotes cultured in in Seq<br />

vs. Mono media. This study demonstrates that the usable blastocyst rate is<br />

greatest after culture in Seq media in comparison to a Mono formulation;<br />

however no difference exists in SIR.<br />

Supported by: Irvine scientific provided Mono media.<br />

P-613 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

SUPPLEMENTING SINGLE STEP CULTURE MEDIA WITH INSU-<br />

LIN FOR CONTINUOUS UNINTERRUPTED IN VITRO CULTURE<br />

OF HUMAN EMBRYOS AND MONITORING THE OUTCOME:<br />

PROSPECTIVE RANDOMIZED CLINICAL TRIAL. M. Fawzy, a<br />

A. Alaboudy, a M. Sabry, b M. Gad, a H. Morsy, a H. Kasem, a<br />

F. Alaboudy, a E. R. Othman, c H. Abdelghaffar, b A. M. Metwalley, d<br />

M. Abdel-Rahman, b S. Roshdy. b a Ibnsina IVF Center, Sohag, Egypt; b Sohag<br />

Faculty of Medicine, Sohag, Egypt; c Assiut University, Assiut, Egypt; d IVF<br />

Lab Director, Bahrain, Bahrain.<br />

OBJECTIVE: To compare embryologically and clinical outcomes between<br />

two different culture strategies: Uninterrupted culture in single step<br />

IVF culture media, versus uninterrupted culture in the same medium with insulin<br />

supplementation.<br />

DESIGN: Prospective Randomized Clinical Trial.<br />

MATERIALS AND METHODS: We recruited 1<strong>17</strong> patients’ Oocytes presented<br />

to Ibnsina IVF Center, Sohag, Egypt, between September 2014 and<br />

March <strong>2015</strong> to this study. Mature Oocytes (1354) underwent Intracytoplasmic<br />

sperm injection (ICSI). We randomly assigned the injected Oocytes<br />

into two groups (sibling Oocytes, 677 each). Group I Oocytes underwent<br />

ICSI and cultured in insulin free single step media (GLOBAL TOTALÒ,<br />

LIFEGLOBALÒ, Canada). Also, incubated in (7% Co2, 5% O2 and 88%<br />

N2) for uninterrupted culture from day 0 to day 5. Group II Oocytes underwent<br />

ICSI then cultured in insulin supplemented single step media<br />

(GLOBAL TOTALÒ, LIFEGLOBALÒ, Canada). We supplemented the culture<br />

media with insulin to reach a concentration of 50 ng/ml (Sigma I9278)<br />

under the same gas phase for uninterrupted culture protocol. We randomized<br />

the transfer of the resulting embryos between the two groups by closed envelope<br />

method. Outcome measures included: Fertilization rate, top quality embryos<br />

and compactionrate at day 3, blastocyst rate and quality at day 5, embryo<br />

utilization and finally pregnancy rate. We used Chi-square test for comparing<br />

ourdata.<br />

RESULTS: The recruited patients were similar regarding mean age, BMI,<br />

the dose of FSH/HMG used, number of oocytes collected and number of cycle<br />

days. There was no difference in fertilization rate (group II 72.6% and a<br />

group I 73%). There was no difference in top quality embryos at day 3 with a<br />

trend toward better compaction of group II. We documented a significantly<br />

higher blastocyst formation rate in group II versus group I (66% versus<br />

44%, respectively. P value < 0.01). We also recorded a higher top quality<br />

blastocyst (better expansion, intact inner cell mass and more trophectoderm<br />

layers) in the group II (64% versus group I 41% P value


and 3 x Zand-air 100C units were used as air filtration. Chi square and student’s<br />

t-tests were used as appropriate. There were no lab protocol changes<br />

during the study period, besides the in-duct air system.<br />

RESULTS: 289 ART cycles were analyzed, 163 cycles pre and 126 cycles<br />

post in-duct air filtration installation. There were no differences between the<br />

two groups in the mean age of the patient, the number of oocytes retrieved,<br />

percentage of mature oocytes per patient, number of fertilized oocytes, or<br />

number of embryos transferred. However, significant increases were seen in<br />

the number of positive pregnancies, ongoing pregnancy rate, and viable blastocyst<br />

development rates. Air quality VOC testing demonstrated improved air<br />

quality after the installation of the in-duct air system (0.0 ppm) as compared<br />

with previous VOC readings prior to installation (0.3 ppm).<br />

Freestanding air<br />

filtration units<br />

alone<br />

In-duct air<br />

filtration<br />

units<br />

p value<br />

Mean age of patient (years) 30.5 30.5<br />

Number of patients (n) 163 126<br />

Mean no. eggs retrieved 12.6 13.3 p¼0.3231<br />

Mean no. mature eggs 8.77 9.77 p¼0.0583<br />

Mean no. fertilized eggs 7.28 7.67 p¼0.4128<br />

Mean no. embryos 1.88 1.79 p¼0.0749<br />

transferred<br />

Mean no. viable blastocysts 1.75 3.50 p¼0.0001<br />

First bhCG (%) 47.2% 63.5% p¼0.0059<br />

Clinical pregnancy rate 45.4% 55.6% p¼0.0868<br />

(%)(gestational sac)<br />

Ongoing pregnancy rate (%)<br />

(fetal heart tones)<br />

34.5% 55.6% p¼0.0003<br />

CONCLUSIONS: Air quality in the IVF laboratory has been shown to have<br />

an impact on embryo development and resultant pregnancy outcomes. In an<br />

urban, multi-office, multiple story building, air quality is often difficult to<br />

control. The results of this study show that even multiple types and numbers<br />

of freestanding air filtration units may not be enough to create an ideal air<br />

quality environment within the IVF laboratory. The observations seen in<br />

this study suggest that the installation of an in-duct, positive pressure air filtration<br />

system does increase viable blastocyst development and pregnancy rates.<br />

P-616 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

PREDICTING BLASTOCYST FORMATION RATE: AN AUTO-<br />

MATIC CELL TRACKING SYSTEM AIDS IN THE SELECTION<br />

AMONG GOOD MORPHOLOGY EMBRYOS. N. Basile, a<br />

I. Cabanes, a M. Testillano, a D. Cernuda, a J. A. Garcia-Velasco, a<br />

M. Meseguer. b a IVI, Madrid, Spain; b Clinical Embryology, Valencia, Spain.<br />

OBJECTIVE: To quantify, by multivariable analysis, the blastocyst formation<br />

rate according to the three predictive categories provided by Eeva (Early<br />

Embryo Viability Assessment).<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Patients undergoing IVF cycles using<br />

their own or donated oocytes. Embryos were cultured in standard incubators<br />

including multiple Eeva systems. The Eeva test utilizes an automatic cell<br />

tracking software to classify embryos into three categories (HIGH-ME-<br />

DIUM-LOW) according to their probability of becoming a blastocyst. For<br />

that aim the system relies on an algorithm based on the variables P2 ¼ t3-<br />

t2 (time to 3 cell - time to 2 cell) and P3 ¼ t4-t3 (time to 4 cell - time to 3<br />

cell). In order to quantify the blastocyst formation rate according to the<br />

Eeva categories, a logistic regression analysis was performed taking in<br />

consideration possible confounding factors: embryo morphology according<br />

to ASEBIR (Spanish Association of Biologists; A-B-C-D), source of oocytes<br />

(own or donated), and number of oocytes.<br />

RESULTS: A total of 494 patients generated 3596 embryos. The overall<br />

blastocyst formation rate was 59.3 % (1347/2269). When categorizing according<br />

to Eeva, we found significant differences in the overall blastocyst formation<br />

rate between LOW vs. MEDIUM (p< 0.001, OR¼ 1.964 CI95% 1.550-<br />

2.489) and LOW vs. HIGH (p< 0.001, OR¼ 3.743 CI95% 2.724-5.143).<br />

Development to ‘‘optimal’’ blastocyst was also analyzed with significant differences<br />

between LOW vs. MEDIUM (p< 0.001, OR¼ 1.634 CI95% 1.248-<br />

2.140) and LOW vs. HIGH (p< 0.001, OR¼ 3.053 CI95% 2.295-4.061). The<br />

only confounding factor presenting a significant effect was embryo<br />

morphology according to ASEBIR. The correlation between Eeva categories<br />

and blastocyst formation rate differs between good quality embryos (A-B) and<br />

poor quality embryos (C-D). Taking this in consideration, we still found significant<br />

differences between LOW vs. MEDIUM (p¼ 0.001, OR¼ 1.554<br />

CI95% 1.207-1.999) and LOW vs. HIGH (p< 0.001, OR¼ 2.505 CI95%<br />

1.792-3.502) for the overall blastocyst formation rate but correlation was<br />

reduced in <strong>21</strong>% and 33% respectively. For ‘‘optimal’’ blastocyst formation<br />

rate, no significant differences were observed between LOW vs. MEDIUM<br />

(p¼ 0.110, OR¼ 1.263 CI95% 0.948-1.682) but we did observe significant<br />

differences between LOW vs. HIGH (p< 0.001, OR¼ 1.950 CI95% 1.437-<br />

2.647). Once again correlation was reduced in 23% and 36% respectively.<br />

CONCLUSIONS: Eeva categories are strongly correlated with blastocyst<br />

formation rates and the prediction is significantly affected by day 3 embryo<br />

morphology. Therefore the best strategy, among good quality embryos, is the<br />

combination of both: morphology and morphokinetics.<br />

P-6<strong>17</strong> Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

EFFECT OFARTIFICIAL OOCYTE ACTIVATION IN INTRACYTO-<br />

PLASMIC SPERM INJECTION USING TESTICULAR SPERMATO-<br />

ZOA ON SIBLING OOCYTES. C. Takahashi, S. Mizuta, R. Nishiyama,<br />

K. Yamaguchi, K. Kitaya, H. Matsubayashi, T. Ishikawa. Reproduction<br />

Clinic Osaka, Osaka, Japan.<br />

OBJECTIVE: Artificial oocyte activation (AOA) has been proposed as a<br />

suitable means to overcome the problem of failed or impaired fertilization after<br />

intracytoplasmic sperm injection (ICSI). To analyze with calcium ionophore<br />

after ICSI using testicular spermatozoa improves fertilization,<br />

embryonic development and pregnancy outcome in patients with obstructive<br />

azoospermia (OA) or non-obstructive azoospermia (NOA).<br />

DESIGN: Prospective clinical analysis on sibling oocytes.<br />

MATERIALS AND METHODS: This prospective study was performed<br />

between <strong>October</strong> 2013 and April <strong>2015</strong>. All patients involved gave written<br />

consent, and institutional review board approval was granted. This study includes<br />

22 OA and 47 NOA couples. We excluded the couples using only<br />

immotile spermatozoa for ICSI. Retrieved oocytes were incubated in culture<br />

medium (Universal IVF Medium: UIM) for 2 hours at 37C and 6% CO₂ and<br />

were underwent ICSI with motile testicular spermatozoa. When eight or more<br />

metaphase M (MII) oocytes were available, AOA was performed on half of<br />

the sibling MII oocytes. After ICSI, oocytes were incubated in UIM for 30<br />

minutes, and exposed to 10mM of calcium ionophore A23187 for 15 minutes.<br />

The oocytes were then washed and placed in UIM. Two pronuclei (2PN) oocytes,<br />

blastocysts development, good-quality blastocysts, biochemical pregnancies,<br />

and clinical pregnancies rates were compared between two groups.<br />

RESULTS: In terms of OA couples, there were no significant difference in<br />

2PN oocytes, blastocysts development, and good-quality blastocysts rates<br />

(73.0%, 56.1%, and 47.8% with AOA and 66.0%, 41.2%, and 45.7% without<br />

AOA, respectively). For NOA couples, 2PN oocytes with AOA (74.0%) was<br />

significantly higher than those without AOA (61.2%). Blastocysts development,<br />

and good-quality blastocysts rates for NOA couples were 53.8% and<br />

38.4% with AOA and 56.7% and 40.5% without AOA, respectively (no significant<br />

differences).<br />

CONCLUSIONS: AOA with calcium ionophore showed favorable effect<br />

on fertilization rate in patients with NOA but OA. The sperm source was<br />

strongly affect the fertility potential or clinical outcomes. Severe male factor<br />

infertility, especially NOA, could be an indication for application of AOA.<br />

P-618 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

AN EVALUATION OF CONTINUOUS HUMAN EMBRYO CULTURE<br />

USING THE WOW DISH. H. Watanabe, N. Fukunaga, K. Nakayama,<br />

M. Shimomura, H. Tsuji, H. Kitasaka, F. Tamura, Y. Konuma,<br />

M. Kojima, Y. Asada. Asada Ladies Clinic Medical Corporation, Nagoya,<br />

Japan.<br />

OBJECTIVE: The WOW dish (LinKIDTMculture dish,DNP) 25 microwells<br />

that allows group culture under a single drop of medium. Through its<br />

design it is possible to manage embryos separately whilst in group culture.<br />

There are several reports (SUGIMURA et al, 2013) suggesting that the<br />

WOW dish improved bovine embryo culture results. Therefore, we investigated<br />

whether the WOW dish is suitable for continuous human embryo culture<br />

without exchange of culture medium.<br />

DESIGN: Retrospective study.<br />

MATERIALS AND METHODS: The study consisted of three experimental<br />

groups. ‘‘15ul drop culture’’ group (control) was the standard embryo<br />

culture method with exchange from single medium to single medium at Day3<br />

FERTILITY & STERILITY Ò<br />

e3<strong>17</strong>


and Day5 using normal 35mm dish. The other 2 groups were ‘‘15ul drop<br />

continuous culture’’ and ‘‘WOW dish continuous culture’’ without exchange<br />

of culture medium. A total of 143 pronuclear embryos from 12 patients who<br />

consented to post-thaw culture were studied. Pronuclear stage frozen embryos<br />

were thawed and cultured across the 3 groups for up to 7days (n¼47, 48, 48).<br />

We compared the incidence of good blastocysts (R 3BB,Gardner) and blastocyst<br />

formation rate at Days 5, 6 and 7 in the 3 groups.<br />

RESULTS: In the control group, blastocyst formation rate at Days 5, 6 and<br />

7 was 46.8%, 57.4% and 59.6%, respectively. The incidence of good blastocysts<br />

was 25.5%, 44.7%, and 48.9% (n ¼ 47) respectively. In the ‘‘15ul drop<br />

continuous culture’’ group, blastocyst formation rate at Days 5, 6 and 7 was<br />

45.8%,54.2% and 54.2%. The incidence of good blastocysts was<br />

14.6%,31.3%,33.3% (n ¼ 48). In the ‘‘WOW dish continuous culture’’ group,<br />

blastocyst formation rate at Days 5, 6 and 7 was 52.1%,58.3% and 58.3%.<br />

Incidence of good blastocysts was 25.0%,43.8% and 47.9% (n ¼ 48). Between<br />

each observation time interval, there was no significant differences<br />

in both the incidence of good blastocysts and the blastocyst formation rate.<br />

CONCLUSIONS: These results suggest that the WOW dish is suitable for<br />

continuous culture without exchange of culture medium compared to the<br />

15ul drop control method.<br />

P-619 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

COMPARISON OF CONTINUOUS AND SEQUENTIAL CULTURE<br />

MEDIA ON BLASTOCYST UTILIZATION AND IMPLANTATION<br />

RATES USING SIBLING OOCYTES. A. E. Fritz, K. Gehrke,<br />

D. Kim, M. Goering. Center for Advanced Reproductive Medicine, Department<br />

of Obstetrics and Gynecology, University of Kansas School of Medicine,<br />

Overland Park, KS.<br />

OBJECTIVE: The broad availability of commercial embryo culture media<br />

has improved performance and consistency in human ART laboratories, however<br />

the variety of media compositions from different manufacturers and culture<br />

conditions between laboratories has left a lack of consensus across the<br />

field. The aim of this study was to provide a carefully controlled comparison<br />

of human preimplantation embryo development in two distinct culture media<br />

formulations, continuous (Global, Life Global Group) versus sequential<br />

(Quinn’s Advantage, Sage).<br />

DESIGN: Retrospective, randomized cohort study.<br />

MATERIALS AND METHODS: A total of 247 fresh non-donor ICSI cycles<br />

(2841 oocytes) from January 2014 through April <strong>2015</strong> were included in<br />

the analysis. Study inclusion required cohorts of at least six sibling oocytes to<br />

be randomly divided into one of two different culture media, continuous<br />

(n¼1451) or sequential (n¼1390). All sibling oocytes and embryos were<br />

treated identically regardless of media formulation, including protein supplementation<br />

(5mg/mL SPS) and media refreshment on D1 and D3. Culture<br />

occurred in Planer mini-benchtop incubators using premixed gas or Thermo<br />

cabinet incubators with 6% CO 2 and 5% O 2. Manufacturer target post-equilibration<br />

pH ranges were carefully monitored and varied only slightly between<br />

formulations (7.25-7.35 +/-0.05). Rates of fertilization, D5 and D6<br />

blastocyst conversion, embryo transfer (D5), cryopreservation, implantation,<br />

and clinical pregnancy were compared.<br />

RESULTS: The rate of fertilization post-injection was higher in the continuous<br />

media (80%) versus the sequential media (73%; p


monitoring, but 50 mL prevents changes in media contained in the SAFE Sens<br />

sensor for weeklong testing.<br />

Osmolality Values.<br />

SAFE Sens<br />

Day 035 mL Oil<br />

268 SAFE Sens Day<br />

335 mL Oil<br />

Culture DishDay 265 Culture Dish Day<br />

0 5.5 mL Oil<br />

3 5.5 mL Oil<br />

Control Culture Dish 268 Control Culture Dish<br />

Day 011 mL Oil<br />

Day 3 11 mL Oil<br />

292 SAFE Sens Day<br />

735 mL Oil<br />

283 Culture DishDay<br />

75.5 mL Oil<br />

270 Control Culture Dish<br />

Day 7 11 mL Oil<br />

Supported by: Thanks to Blood Cell Storage Inc. for use of their evaluation<br />

equipment.<br />

P-622 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

RANDOMIZED COMPARISON OF EMBRYO DEVELOPMENT IN<br />

CLOSED TIME-LAPSE PHOTOGRAPHY SYSTEM WITH TRADI-<br />

TIONAL STANDARD EMBRYOLOGY CULTURE WITH DAY-3 EM-<br />

BRYO TRANSFERS. Y. Wu, a E. Lazzaroni-Tealdi, a Q. Wang, a<br />

D. F. Albertini, b D. H. Barad, c V. A. Kushnir, d N. Gleicher. c a Center for Human<br />

Reproduction, New York, NY; b Center for Human Reproduction & University<br />

of Kansas Medical Center, New York, NY; c Center for Human<br />

Reproduction & Foundation for Reproductive Medicine, New York, NY;<br />

d Center for Human Reproduction & Wake Forest University, New York, NY.<br />

OBJECTIVE: To prospectively compare a time-lapse photographic system<br />

(EmbryoscopeÒ, UnisensFertilitech, Vitrolife, Aarhus, Denmark) to<br />

the traditional standard manual embryology in the management of human<br />

embryos during donor oocyte cycles with day-3 embryo transfers (ET).<br />

DESIGN: Prospectively randomized open label study (PRS).<br />

MATERIALS AND METHODS: This study had two components: In (1)<br />

we compared embryo quality in 76 embryos from 7 oocyte donor cycles during<br />

randomized parallel culture in our standard culture system (SC) and an<br />

EmbryocopeÒ culture system (EC). After three day culture, embryos were<br />

evaluated by our center’s standard morphology grading. In (2), we parallel<br />

cultured 37 embryos from 4 donor cycles in SC and EC dishes, respectively,<br />

in a standard incubator for 3 days before grading to determine whether culture<br />

dishes used in EC and SC may be responsible for outcome differences.<br />

All embryos were produced by ICSI and only normally fertilized embryos<br />

were selected for the study.<br />

RESULTS: The table demonstrates that the SC system produced better<br />

quality embryos than the EC system (P 0.05).<br />

Phase I: Culture system<br />

test (n¼76, N¼7)<br />

Phase II: Culture dish<br />

test in standard<br />

incubator(n¼37, N¼4)<br />

Embryo<br />

quality<br />

n¼number of embryos; N¼number of patients.<br />

EC<br />

system<br />

SC<br />

system<br />

8<strong>17</strong><br />

333<br />

268<br />

P value<br />

Good (%) 55.86.4 81.24.1 0.005<br />

Fair (%) 36.88.5 7.74.1 0.01<br />

Poor (%) 7.35.7 11.04.7 0.62<br />

Embryo EC dish SC dish P value<br />

quality<br />

Good (%) 65.08.6 66.03.3 0.96<br />

Fair (%) 16.79.6 11.45.9 0.67<br />

Poor (%) 23.35.1 27.33.9 0.56<br />

CONCLUSIONS: In oocytes donors (i.e. likely good prognosis patients),<br />

undergoing day-3 embryo transfer, SC produced embryos of better quality<br />

than EC. Since this difference does not appear to be consequence of differences<br />

in culture dishes, the observed discrepancy in outcomes has to be<br />

due to differences in culture conditions. Since here reported results are based<br />

on limited study size and, therefore, do not include pregnancy outcomes,<br />

they should be considered preliminary. They, however, do raise concerns,<br />

and should induce the industry to perform PRSs, utilizing embryo randomization<br />

rather than patient randomization, to assure that EC systems do not<br />

negatively affect IVF outcomes in comparison to standard embryology<br />

practice.<br />

Supported by: We acknowledge the support of UnisenseFertilitech, Vitrolife,<br />

which loaned CHR the instrument. Intramural funds from The Center for<br />

Human Reproduction and grants from The Foundation for Reproductive<br />

Medicine paid for installation fees and reagents.<br />

OVARIAN STIMULATION<br />

P-623 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

COMPARISON OF STANDARD GNRH ANTAGONIST PROTOCOL<br />

AND LUTEAL PHASE ESTRADIOL/GNRH ANTAGONIST PRIM-<br />

ING PROTOCOL IN THE POOR RESPONDERS. A. Erdem, a<br />

M. F. Mutlu, b M. Erdem, a I. Mutlu, c I. Guler. a a Gazi University Faculty<br />

of Medicine, Ankara, Turkey; b Koru Hospital, Ankara, Turkey; c NovaArt<br />

IVF Center, Ankara, Turkey.<br />

OBJECTIVE: To compare in vitro fertilization (IVF) outcomes between<br />

standard GnRH antagonist protocol and luteal phase estradiol/ GnRH antagonist<br />

priming protocol in the poor responders.<br />

DESIGN: Prospective randomized trial.<br />

MATERIALS AND METHODS: One hundred and five poor responders<br />

according to ESHRE Bologna criteria whose ages were between 25 to 45<br />

were included to the study. Based on computer generated randomization<br />

52 patients used luteal phase estradiol/ GnRH antagonist priming protocol<br />

and 53 patients used standard GnRH antagonist protocol prospectively. Patients<br />

with luteal phase estradiol/ GnRH antagonist priming protocol took<br />

0.1 mg transdermal estradiol on alternate days for 3 times starting from<br />

7th day after ovulation in the previous cycle. In addition, on the second<br />

day after application of patch daily subcutaneous cetrorelix 0.25 mg was<br />

started and continued for 3 days. Both of two groups underwent controlled<br />

ovarian hiperstimulation starting on day 3 of menstural cycle with flexible<br />

antagonist protocol Primary outcomes of the study were the number of oocytes<br />

retrieved, clinical pregnancy, implantation and cycle cancellation rates.<br />

Secondary outcomes were peak estradiol levels, the number of mature oocytes,<br />

fertilization rates. Student’s t-test was used for continuous variables,<br />

and the chi-square test was used for categorical variables.<br />

RESULTS: The baseline characteristics as age, body mass index, antral<br />

follicle count, basal FSH, basal E2, the number of prior IVF attempts and<br />

duration of infertility were similar between the two groups. The mean number<br />

of oocytes retrieved (3.7 2.8 vs. 3.8 2.8), the number of mature oocytes<br />

(2.6 2.2 vs. 3 2.3), fertilization rates (61.1% 37.8% vs. 71% <br />

31.3%) and the number of embryos transferred (1.6 0.6 vs. 1.7 0.6) were<br />

also similar. The cancellation rate was not different between groups (15.1%<br />

vs 11.1%). There were no significant differences between groups in terms of<br />

the implantation (8.4% vs 8.8%), clinical pregnancy (13.2% vs. 13.3%) and<br />

ongoing pregnancy rates (9.4% vs. 4.4%) per started cycle.<br />

CONCLUSIONS: Luteal phase estradiol/ GnRH antagonist priming protocol<br />

has no effect to improve the IVF outcomes in terms of improved cycle<br />

cancellation and pregnancy outcome as compared to standard GnRH antagonist<br />

protocol in patients with poor response to gonadotropins after IVF.<br />

P-624 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

LUTEAL PHASE LUTEINIZING HORMONE (LH) LEVELS AFTER<br />

GONADOTROPIN-RELEASING HORMONE AGONIST (GNRH-A)<br />

TRIGGER AND PROBABILITY OF CONCEPTION. C. B. Bartels,<br />

B. L. Maslow, E. Anspach, C. A. Benadiva, J. Nulsen, L. Engmann. University<br />

of Connecticut Health Center, Farmington, CT.<br />

OBJECTIVE: It has been proposed that low luteal phase luteinizing hormone<br />

(LH) levels following gonadotropin-releasing hormone agonist<br />

(GnRH-a) trigger induce abnormal corpus luteum function, which may lower<br />

conception rates. This study aimed to evaluate whether luteal phase serum<br />

LH profile predicts probability of conception after GnRH-a trigger.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: We included all women


continuous variables. Logistic regression models were built to test significant<br />

findings in the univariate analysis.<br />

RESULTS: 375 cycles were included in this study. 255 (68%) were<br />

conception cycles. There were no significant differences in baseline characteristics<br />

between conception and non-conception cycles. There was no significant<br />

difference in use of GnRH-a alone and GnRH-a+HCG trigger in<br />

conception compared to non-conception cycles (p¼0.55).Median LH levels<br />

were 1.6, 65.3, 0.3, 0.1 and 0.2 on trigger +0, +1, +5, +9 and +16 days respectively.<br />

There were no significant differences in median LH levels at any of the<br />

time points between those who received GnRH-a alone and those who<br />

received GnRH-a+HCG trigger. There were also no significant differences<br />

between conception and non-conception cycles in any time points apart<br />

from the day of the pregnancy test. The higher LH levels on the day of pregnancy<br />

test remained significant in the regression model (p¼0.007), even<br />

when controlling for type of trigger.<br />

Luteal phase luteinizing hormone (LH) levels in conception and nonconception<br />

cycles.<br />

Time<br />

Conception<br />

Cycles<br />

LH IU/L<br />

(MedianIQR)<br />

N¼ 255<br />

Non-conception<br />

Cycles<br />

LH IU/L<br />

(MedianIQR)<br />

N¼ 120<br />

p-value<br />

Day of Trigger 1.62.1 1.51.7 0.56<br />

Day after Trigger 67.250.5 63.249.7 0.68<br />

Trigger + 5 days 0.30.4 0.30.4 0.83<br />

Trigger + 9 days 0.10.3 0.10.2 0.85<br />

Trigger + 16 days 0.20.3 0.10.1 0.001<br />

CONCLUSIONS: Luteal phase LH levels are generally low in both<br />

conception and non-conception cycles after GnRH-a trigger. This is the first<br />

study to demonstrate that luteal phase LH levels following GnRH-a trigger do<br />

not significantly predict the probability of conception and therefore its measurement<br />

is not warranted.<br />

P-625 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

WITHDRAWN<br />

P-626 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

COMPARISON OF HORMONAL LEVELS IN POLYCYSTIC OVARY<br />

PATIENTS USING SHORTAND LONG ANTAGONIST PROTOCOLS<br />

DURING CONTROLLED OVARIAN STIMULATION IN IVF-ET<br />

CYCLES. M. Thakur, a A. Bolnick, a O. Abuzeid, b R. Raju, c J. Dai, a<br />

E. E. Puscheck, a M. P. Diamond, d M. I. Abuzeid. e a Division of Reproductive<br />

Endocrinology and Infertility, Department of Obstetrics and Gynecology,<br />

Wayne State University/ Detroit Medical Center, Detroit, MI; b OBGYN,<br />

Dearborn, MI; c Hurley Medical Center/Michigan State University Co, Flint,<br />

MI; d Georgia Regents University, Augusta, GA; e Hurley Medical Center,<br />

Flint, MI.<br />

OBJECTIVE: The usefulness of gonadotropin-releasing hormone (GnRH)<br />

antagonist for the inhibition of premature luteinization during ovarian hyperstimulation<br />

has been studied. GnRH antagonist has been shown to offer<br />

increased safety contrasted with GnRH agonist cycles, with no significant<br />

difference in birth rate. However some studies suggest that longer than 5<br />

days of GnRH antagonists may have a negative impact on the endometrium<br />

and in turn on pregnancy outcome. We set out to study the affect of different<br />

timing of the initiation of GnRH antagonist on hormonal levels in patients<br />

with polycystic ovary syndrome (PCOS) undergoing IVF. We also evaluated<br />

effect of length of GnRH antagonist use on pregnancy outcomes.<br />

DESIGN: Prospective trial (ISRCTN69937<strong>17</strong>9).<br />

MATERIALS AND METHODS: Patients with PCOS undergoing IVF<br />

were prospectively randomized to Group 1, initiating GnRH-antagonist on<br />

day 1 of stimulation and group 2, with GnRH antagonist initiated on day<br />

5. Estradiol (E2) and progesterone (P4) were measured during follicular<br />

phase on cycle day 2-3, 5-6, 7-8 and daily or every other day thereafter.<br />

Serum E2 and P4 were measured on the day of embryo transfer and every<br />

third day thereafter during luteal phase (1st, 2nd, 3rd sample). Analysis of<br />

hormonal values was performed to determine differences in the groups based<br />

on day of antagonist initiation and we also studied the effect of the length of<br />

GnRH used on pregnancy outcomes.<br />

RESULTS: 140 patients with PCOS were recruited with Group 1 (n¼69)<br />

and Group 2 (n¼71). Embryo transfer (ET) was performed in 122 women.<br />

Eighteen patients (9 in each group) were excluded as they did not have ET.<br />

Mean days of GnRH-antagonist used in Group 1 was 10 2 days (range<br />

7-16 days) and for Group 2 was 6 2 days (range 3-15 days) [p % 0.0]. There<br />

were no differences noted in hormonal levels in either the follicular or luteal<br />

phase regardless of the group. In addition there were no significant differences<br />

in the clinical pregnancy rate (68.3 % vs 56.5%) and first trimester<br />

miscarriage rate (7.3 % vs 8.6%) between Group 1 and Group 2, respectively.<br />

CONCLUSIONS: Our data shows that amongst PCOS patients undergoing<br />

IVF-ET there was no difference in hormonal levels during the follicular<br />

and luteal phases regardless of when GnRH antagonist was started. Contrary<br />

to previous reports we did not observe any negative impact on clinical pregnancy<br />

rate in this pilot study.<br />

P-627 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

THE EFFICACY AND SAFETY OF 100 MCG CORIFOLLITROPIN<br />

ALFA IN ART CYCLES FOR PATIENTS WITH HIGH SERUM<br />

ANTI-MULLERIAN HORMONE LEVELS. T. Lee, a,b H. Chen, b<br />

C. Huang, b M. Lee. a,b a Department of Obstetrics and Gynecology, Chung<br />

Shan Medical University Hospital, Taichung, Taiwan; b Department of Infertility<br />

Clinic, Lee Women’s Hospital, Taichung, Taiwan.<br />

e320 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


OBJECTIVE: The selection of 100mcg or 150mcg corifollitropin alfa in<br />

IVF/ICSI cycles depends on the body weight of the patients with good prognosis.<br />

A higher percentage of excessive ovarian response and consequent<br />

cancellation is found in corifollitropin alfa compared to daily FSH with<br />

GnRH antagonist protocols. However, the excessive ovarian response in corifollitropin<br />

alfa protocol did not jeopardize the ongoing pregnancy rate and<br />

might be predicted by serum anti-Mullerian hormone levels. This analysis<br />

tried to prove the concept that 100mcg corifollitropin alfa is adequate for<br />

all patients with a high AMH level (>3.5 ng/ml).<br />

DESIGN: This is a nested case-control analysis of the medical records in a<br />

prospective cohort trial.<br />

MATERIALS AND METHODS: A total of 1<strong>21</strong> patients with an AMH<br />

level >3.5 ng/ml were recruited into this analysis. Forty-nine patients<br />

received 100mcg corifollitropin alfa and 72 patients underwent daily FSH<br />

(150 IU-225 IU) in the initial 5 days. The main outcome was the oocyte<br />

retrieval number, ongoing pregnancy rate, and ovarian hyperstimulation<br />

rates.<br />

RESULTS: The estradiol (23961181 vs. 30031864 pg/ml) and progesterone<br />

levels (0.920.54 vs. 1.220.84 ng/ml) on the day of hCG injection<br />

were significantly lower in corifollitropin alfa group compared to daily FSH<br />

group. The oocyte retrieval number (13.36.3 vs. 14.37.5), fertilization<br />

rates (74.715.6 vs. 74.719.1%), day3 good embryo rates (53.532.6<br />

vs. 58.725.3%), and pregnancy rates (28/48¼58.3% vs. 31/70¼ 44.3%)<br />

are similar between the two groups. The OHSS rate is 2.04% (1/49) and<br />

2.78 (2/72) for corifollitropin alfa and daily FSH group, respectively. All<br />

were moderate OHSS and paracentesis was not necessitated.<br />

CONCLUSIONS: The 100mcg corifollitropin alfa is effective and safe in<br />

IVF/ICSI cycles for patients with high AMH levels. Further studies are<br />

needed to investigate the efficacy of 100mcg corifollitropin alfa for patients<br />

with weight above 90 Kg.<br />

References:<br />

1. Fatemi HM, Doody K,Griesinger G, et al., High ovarian response does<br />

not jeopardize ongoing pregnancy rates and increases cumulative pregnancy<br />

rates in a GnRH-antagonist protocol. Hum Reprod, 2013;28:442-<br />

452.<br />

2. Polyzos NP, Tournaye T, Guzman L, et al., Predictors of ovarian<br />

response in women treated with corifollitropin alfa for in vitro fertilization/<br />

intracytoplasmic sperm injection. Fertil Steril, 2013;100: 430-<br />

437.<br />

Supported by: Grant from Chung Shan Medical University Hospital CSH-<br />

2014-C-012.<br />

P-628 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

MINIMAL STIMULATION IN VITRO FERTILIZATION (IVF) WITH<br />

A FREEZE-ALL APPROACH GIVES ACCEPTABLE LIVE BIRTH<br />

RATES IN AN OVERALL POOR PROGNOSIS<br />

POPULATION. B. G. Reed, M. Ezzati, S. N. Babayev, V. Libby,<br />

B. Carr, O. Bukulmez. UT Southwestern Medical Center, Dallas, TX.<br />

OBJECTIVE: Minimal stimulation IVF typically utilizes daily clomiphene<br />

citrate and 3 low doses of gonadotropin injections on stimulation<br />

days 5, 7, and 9. Minimal stimulation cycles are planned with the intention<br />

to freeze all available embryos exclusively at blastocyst stage. After reasonable<br />

embryo accumulation, frozen embryo transfer (FET) is performed.<br />

There is a paucity of published data on the outcomes of minimal stimulation<br />

IVF with a freeze-all protocol in patients with diminished ovarian reserve<br />

(DOR) and/or a prior history of traditional IVF failures.<br />

DESIGN: Retrospective observational study.<br />

MATERIALS AND METHODS: Institutional Review Board approval<br />

was obtained and the data for all initiated minimal stimulation IVF cycles between<br />

2012-2014 was entered into a database. Patient and cycle characteristics,<br />

pregnancy and live birth data were reviewed and analyzed.<br />

RESULTS: A total of 54 patients initiated 147 cycles (102 stimulations,<br />

and 45 FETs). Age, body mass index (BMI), and antimullerian hormone<br />

(AMH) levels and prior history of failed traditional IVF attempts were suggestive<br />

of an overall poor prognosis group (Table 1). Cancellation rates for<br />

minimal stimulation and FET were 24% and 22% respectively. Stimulation<br />

cycle cancellation reasons included premature ovulation (13), poor response<br />

(8), early follicular growth (1), and non-compliance (2). FET cycle cancellations<br />

were mostly due to either inadequate endometrial development (6) or<br />

inability to be suppressed with a GnRH agonist (3). The methods of fertilization<br />

were conventional IVF in 46 cycles and ICSI in 35 cycles. Despite being<br />

a poor prognosis population and having a high cancellation rate, the cumulative<br />

live birth rate per patient was 24%. The live birth rate per FET was 37%.<br />

Of the women who have not yet achieved a live birth, 32% still have remaining<br />

embryos to transfer. Among many cycle parameters, female age was a<br />

strong predictor of obtaining at least one frozen blastocyst.<br />

CONCLUSIONS: Since there is no data available on minimal stimulation<br />

using a freeze-all approach in poor prognosis patients, this information is<br />

important for patient counseling. Minimal stimulation cycles for embryo<br />

accumulation at blastocyst stage result in acceptable live birth rates in a<br />

poor prognosis population.<br />

Table 1: Demographics and clinical outcome data in completed minimal<br />

stimulation cycles.<br />

Values are expressed<br />

as either percentage<br />

or meanSD. Range<br />

Age (y) 37.894.1 (28-45)<br />

BMI 25.954.6 (18.1-38.4)<br />

AMH (ng/ml) 0.550.67 (0-3)<br />

Patients with prior unsuccessful 44 % (24/54) -<br />

traditional IVF<br />

Number of retrievals per patient 1.690.99 0-4<br />

Number of FETs per patient 0.720.63 0-2<br />

Biochemical pregnancy rate 60% (<strong>21</strong>/35) -<br />

per FET<br />

Clinical pregnancy rate per FET 48.6% (<strong>17</strong>/35) -<br />

Live birth rate per FET 37% (13/35) -<br />

Cumulative live birth rate<br />

per patient<br />

24% (13/54) -<br />

P-629 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

VERY LOW LEVEL OF ANTI-MULLERIAN HORMONE (AMH) IN<br />

A LARGE COHORT OF ASSISTED REPRODUCTIVE TECHNOL-<br />

OGY (ART) PATIENTS. J. Rodriguez-Purata, a M. Luna, a<br />

E. Cervantes, a J. A. Lee, a M. C. Whitehouse, a A. B. Copperman, b<br />

B. Sandler. a a Reproductive Medicine Associates of New York, New York,<br />

NY; b RMANY-Mount Sinai, New York, NY.<br />

OBJECTIVE: The counseling and management of women with low AMH<br />

levels presents a significant challenge where either cycle cancellation or poor<br />

response is foreseen. Prior clinical impression suggests that an extremely low<br />

level of AMH can serve as an alert for a clinician to dissuade a patient from<br />

utilizing certain fertility treatments. The aim of this study was to evaluate<br />

ART laboratory outcomes from patients with extremely low AMH levels.<br />

DESIGN: Retrospective.<br />

MATERIALS AND METHODS: All patients scheduled for a fresh IVF<br />

cycle with ‘‘extremely low’’ (Group A: %0.2) or ‘‘low’’ AMH levels (Group<br />

B: 0.2 - %0.5; Group C: %1 ng/ml) between January 2009 to March <strong>2015</strong><br />

were included. Main outcome measures were total retrieved oocytes and cycle<br />

cancellation rates. Secondary outcomes analyzed were age, day 3 FSH,<br />

basal AFC, number of follicles greater than 14mm at surge, peak estradiol,<br />

cumulative amount of gonadotropins (GND) and days of stimulation.<br />

RESULTS: Five hundred and forty four patients underwent 1647 cycles<br />

(Table 1). The rate of cycle cancellation prior to VOR was significantly<br />

higher in Group A patients (44.2%) when compared to Group B and C patients<br />

(20.8%; 10.0%). In patients that reached VOR, the number of oocytes<br />

retrieved per cycle was significantly lower in Group A (5.13.0) compared to<br />

Groups B and C (6.43.5; 8.64.7). All secondary variables were statistically<br />

different between groups except the average days of stimulation<br />

required.<br />

CONCLUSIONS: AMH is one of the best available tools for the detection<br />

of low ovarian reserve and its role as a clinical test is robust. To our knowledge,<br />

this is one of the largest studies to examine laboratory outcomes in<br />

extremely low AMH level patients. The results demonstrated that patients<br />

with very low AMH levels require a higher amount of gonadotropins than patients<br />

with merely low levels. Although clinicians should inform patients the<br />

potential adverse realities of seeking treatment, our study demonstrated that<br />

patients should not be excluded from pursing an IVF cycle as they still<br />

respond, fairly, to stimulation. With further research, a more personalized<br />

stimulated treatment regimen incorporated genomic and phenotypic variables<br />

will lead to accurate prognostic information and optimal treatment strategies.<br />

FERTILITY & STERILITY Ò<br />

e3<strong>21</strong>


Outcomes in patients with Low and Very Low AMH levels<br />

ANOVA Chi SquareA vs. B Chi SquareA vs. C<br />

Cycles 437 557 653 1647<br />

Age 38.54.5 39.04.0 38.53.9 NS NS NS<br />

Day 3 FSH 8.87.4 7.54.7 7.64.0 p


d Calculated as the number of 2 pronuclei oocytes divided by the total number<br />

of oocytes retrieved.<br />

e Visit 8 corresponds to Day 14 +/- 2 days after embryo transfer.<br />

f Live birth data were collected by each study site after the end-of-study visit.<br />

g Calculated using the total number of neonates for the denominator.<br />

Supported by: This study was sponsored by Ferring Pharmaceuticals, Inc.<br />

P-632 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

EGG YIELD VS. EGG QUALITY. C. Chang, J. M. Linn, D. B. Shapiro,<br />

A. A. Toledo, M. W. Best, Z. Nagy. Reproductive Biology Associates, Atlanta,<br />

GA.<br />

OBJECTIVE: To assess the correlation between donor oocyte yield and<br />

oocyte quality using a cryo-banking model.<br />

DESIGN: Retrospective study.<br />

MATERIALS AND METHODS: Cryopreservation of donor oocytes<br />

derived from 543 egg retrievals was performed using minimum volume vitrification.<br />

All donors were treated with recFSH/GnRH antagonist/GnRH<br />

agonist trigger. Oocytes were warmed for each matched recipient independently<br />

(1334 recipient cycles, total of 8344 oocytes warmed). Outcome<br />

data was grouped by oocyte yield; I)


TRANSCRIPTIONAL COMPARISON OF NATURAL AND STIMULATED<br />

IVF CYCLES.<br />

NATURAL<br />

CYCLE<br />

ANTAGONIST<br />

CYCLE<br />

AGONIST<br />

CYCLE<br />

p-value<br />

Age 33.12.5 34.23.4 32.74.5 a:


P-638 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

THE USE OF COENZYME Q10 AND DHEA DURING COH AND IVF<br />

CYCLES IN PATIENTS WITH DECREASED OVARIAN RESERVE<br />

(DOR). I. Gat, a S. Blanco Mejia, b H. Balakier, a C. L. Librach, a<br />

A. Claessens, b E. A. J. Ryan. b a Create Fertility Centre, Toronto, ON, Canada;<br />

b Toronto West Fertility Center, Etobicoke, ON, Canada.<br />

OBJECTIVE: Treatment of patients with diminished ovarian reserve<br />

(DOR) is one of the biggest challenges in assisted reproductive therapeutics.<br />

Dehydroepiandrosterone (DHEA) and Coenzyme Q10 (CoQ10) are supplements<br />

that have been purported to have a beneficial effect in these patients.<br />

Our objective was to compare the effect of combining DHEA and CoQ10<br />

supplementation with DHEA alone in COH and IVF cycles in patients<br />

with DOR.<br />

DESIGN: Clinical retrospective study.<br />

MATERIALS AND METHODS: We extracted data from patients charts<br />

treated by DHEA (25mg tid) with/without CoQ10 (600mg daily) in a private<br />

infertility clinic between Feb. 2006 to June 2014. Pre-stimulation parameters<br />

analyzed included age, BMI, day 3 FSH and antral follicular count (AFC).<br />

Ovarian response parameters analyzed included total dose of gonadotropins,<br />

peak serum estradiol (E2), follicles number > 16 mm on day of triggering<br />

and fertilization rate for IVF cycles. Clinical outcomes analyzed included<br />

clinical and ongoing pregnancy rates per cycle initiated.<br />

RESULTS: 797 COH cycles and 253 IVF cycles were included. Of these,<br />

330 COH cycles involved both DHEA and CoQ10 (D+C) and 467 cycles of<br />

DHEA (D) alone; 78 IVF cycles involved D+C and <strong>17</strong>5 D. In both COH and<br />

IVF groups, AFC was significantly higher with D+C compared to D alone<br />

(7.45.7 vs. 5.94.7 and 8.26.3 vs. 5.25, respectively, p 16 mm on the<br />

trigger day during COH cycles (3.32.3 vs. 2.92.2, respectively,<br />

p¼0.01), but no difference was observed in IVF cycles (5.53.5<br />

vs.5.94.6, p¼0.4). There was lower gonadotropin consumption during<br />

D+C IVF cycles compared with D cycles (3,4141141 IUs vs.<br />

3,8771143 IUs respectively, p¼0.032) without difference in the COH comparison.<br />

For COH cycles there was no difference in clinical or ongoing pregnancy<br />

rates in the D+C vs. D groups (8.1 vs. 10.9, p¼0.2 and 6.2 vs. 8.2,<br />

p¼0.3, respectively). Similarly, there was no difference regarding clinical<br />

or ongoing pregnancy rates in the D+C vs. D groups in the IVF cycles<br />

(25.1 vs. 29.5, p¼0.5 and <strong>21</strong>.1 vs. 23.1, p¼0.7, respectively).<br />

CONCLUSIONS: Combined DHEA and CoQ10 supplementation significantly<br />

increases the AFC compared to DHEA alone, which lead to a higher<br />

ovarian responsiveness during both COH and IVF, but without a difference in<br />

pregnancy rate.<br />

P-639 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

EARLY GNRH-ANTAGONIST INITIATION DOES NOT IMPROVE<br />

PREGNANCY RATES OF PCOS CASES UNDERGOING IVF/ICSI:<br />

A PROSPECTIVE RANDOMIZED STUDY. C. S. Atabekoglu, a<br />

E. Kocbulut, a E. G. Pabuccu, b B. Ozmen, a B. Berker, a M. Sonmezer. a a Obstetrics<br />

and Gynecology, Ankara University School of Medicine, Ankara,<br />

Turkey; b Ufuk University School of Medicine, Ankara, Turkey.<br />

OBJECTIVE: Is to assess the effect of early GnRH antagonist initiation on<br />

the hormonal environment and pregnancy outcomes of Polycystic Ovary<br />

Syndrome (PCOS) patients undergoing in vitro fertilization cycles due to unexplained<br />

infertility.<br />

DESIGN: Prospective randomized, pilot study.<br />

MATERIALS AND METHODS: Twenty-four women participated in a<br />

prospective randomized trial. Diagnosis of PCOS was based on Rotterdam<br />

Criteria (1). Inclusion criteria were: age < 38 year, no more than two previous<br />

failed IVF attempts, body-mass index between 18-29 kg/m2, no endometriosis<br />

or previous poor response to ovarian stimulation. Following 1 month<br />

of oral contraceptive pill usage, all participants were stimulated with daily<br />

recombinant FSH at doses ranging from 150 IU to maximal of 200 IU according<br />

to ovarian reserve and/or body mass index values, starting on day 2 of the<br />

cycle. Patients received GnRH-antagonist, starting either on day 2 (n ¼ 12,<br />

early antagonist group) or on day 6 of stimulation (n ¼ 12, flexible antagonist<br />

group). Primary outcome was to compare the serum progesteron (P) levels on<br />

the day of hCG administration between groups. Secondary outcome was to<br />

compare implantation and clinical pregnancy rates.<br />

RESULTS: Demographic characteristics of the groups were similar in<br />

terms of age, BMI, previous history and ovarian reserve. Serum P concentrations<br />

were not differed among groups both during early follicular phase and<br />

on the day of hCG. Other outcome measures including mature oocyte yield,<br />

fertilization-implantation and clinical pregnancy rates were also comperable<br />

among groups (p


with 1,2 versus 3,0, and the luteal phase discomfort was rated as 1.2 vs. 2.5<br />

respectively with the patient friendly protocol (corifollitropin alpha+GnRHagonist-triggering)<br />

as compared with the conventional protocol (recFSH+hCGtriggering)<br />

(from 1- low, to 5 - severe).<br />

CONCLUSIONS: This is the first study to compare two different stimulation<br />

regimens in the same donor to avoid inter-patient variation. Taking into<br />

account the low risk of ovarian hyperstimulation syndrome (OHSS), the<br />

lower degree of physical distress during the luteal phase as well as the similar<br />

pregnancy rates and efficacy, the use of corifollitropin alpha combined with<br />

GnRH-agonist for triggering appears to be the ideal standard in oocyte donors.<br />

P-641 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

LOWER MULTIPLE PREGNANCY RATES WITH LETROZOLE<br />

VERSUS CLOMIPHENE IN 16,001 IUI CYCLES OVER 10<br />

YEARS. J. Toner. Atlanta Center for Reproductive Medicine, Atlanta, GA.<br />

OBJECTIVE: To evaluate whether pregnancy and multiple pregnancy outcomes<br />

are different in cycles stimulated with letrozole versus clomiphene citrate.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: All 16,001 IUI cycles at our clinic using<br />

letrozole or clomiphene citrate for ovulation induction or enhancement from<br />

2004 to 2014 were included. Cycles with concomitant gonadotropin use were<br />

excluded. hCG to trigger ovulation was optional. Cycles were subdivided by<br />

infertility diagnosis and female age.<br />

RESULTS: Over all diagnoses and female ages, pregnancy rates were<br />

slightly higher with clomiphene than letrozole (<strong>17</strong>.4% vs. 15.1%), but multiple<br />

pregnancy rates were significantly higher with clomiphene (twins/preg:<br />

7.5% vs. 4.7%; triplets/preg: 1.3% vs. 0.2%). These effects were also seen<br />

across diagnoses: PCO, male, pelvic, endometriosis, unexplained.Unexpectedly,<br />

clomiphene’s higher pregnancy rates relative to letrozole vanished over<br />

age 40, where letrozole had a small advantage (8.2% vs. 5.6%).<br />

CONCLUSIONS: Our large experience with both letrozole and clomiphene<br />

for many types of infertility patients suggests that letrozole produces<br />

fewer multiple pregnancy rates and more pregnancies over age 40 relative to<br />

clomiphene.<br />

P-642 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

COMPARISION OF LUTEAL ESTRADIOL PATCH AND ORAL<br />

CONTRACEPTIVE PILLS IN ANTAGONIST CYCLES FOR PRE-<br />

TREATMENT OF THE NORMAL RESPONDER. N. Pereira, a<br />

A. C. Petrini, b J. Lekovich, a I. Kligman, b Z. Rosenwaks. b a The Ronald<br />

O. Perelman and Claudia Cohen Center for Reproductive Medicine, New<br />

York, NY; b Weill Cornell Medical College, New York, NY.<br />

OBJECTIVE: To investigate the impact of pre-treatment with luteal phase<br />

estradiol (E 2 ) patch compared to both combined oral contraceptive pills<br />

(OCPs) and no pretreatment on ovarian stimulation response in patients undergoing<br />

fresh in vitro fertilization (IVF) - embryo transfer (ET) cycles.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: All patients undergoing fresh IVF-ET<br />

cycles between January 2008 and June 2013 at our center were analyzed<br />

for inclusion. Exclusion criteria included polycystic ovarian syndrome,<br />

oocyte donors, age > 40 and known poor response to ovarian stimulation.<br />

Demographic characteristics recorded were age, gravidity, parity, BMI (kg/<br />

m 2 ), and infertility diagnosis. Ovarian stimulation parameters recorded<br />

were as follows: total days of ovarian stimulation, total dosage of gonadotropins<br />

administered (IU), peak estradiol (E 2 ) level (pg/mL), peak endometrial<br />

stripe (mm), total number of oocytes retrieved, and total number of mature<br />

oocytes. For each ET cycle, the total number of embryos transferred, clinical<br />

pregnancy rate, biochemical pregnancy rate, spontaneous miscarriage rate,<br />

and live birth rate was recorded. Student’s t-tests and Chi-square tests were<br />

used to compare means and percentages, respectively. Statistical significance<br />

was set at P < 0.05.<br />

RESULTS: 4080 patients met inclusion criteria. There was no difference<br />

in the baseline demographics of the two groups. Differences in the ovarian<br />

stimulation parameters are highlighted in the table below. Patients in the<br />

OCP group underwent ovarian stimulation longer than the E2 patch group<br />

i.e.,10.7 (1.63) days vs. 9.92 (1.94) days (P < 0.001). As a result, the<br />

former group required higher doses of gonadotropins (2750.6 1297.1 IU)<br />

compared to the latter group (2550.1 1270.2 IU; P¼0.003). Patients in<br />

the OCP group also had higher E2 levels on the day of and the day after<br />

hCG trigger comapred to the E2 patch group. There were no differences in<br />

pregnancy outcomes after ET when comparing all three groups.<br />

CONCLUSIONS: Compared to E 2 patch pre-treatement, OCPs increase<br />

the total stimulation days, total gonadotropins administered, E2 level on<br />

the day of trigger, while slightly decreasing the total number of mature oocytes<br />

in the normal responder. The overall clinical pregnancy and live birth<br />

rates remain unaffected by E 2 patch, OCP or no pretreatment.<br />

P1: E2 patch-Ant vs. OCP-Ant; P2: E2 patch-Ant vs. No E2 patch/OCP-Ant.<br />

Parameter<br />

E2 Patch-Ant<br />

(n¼1057)<br />

OCP-Ant<br />

(n¼1035)<br />

No E2/OCP-Ant<br />

(n¼1988) P1 P2<br />

Age (years) 36.1 (2.73) 35.9 (3.91) 36.3 (3.08) 0.15 0.08<br />

Basal AMH (ng/mL) 0.98 (0.13) 0.97 (0.<strong>17</strong>) 1.00 (0.14) 0.41 0.73<br />

Total stimulation days 9.92 (1.94) 10.7 (1.63) 9.89 (2.03)


CONCLUSIONS: BRCA status does not appear to correlate with response<br />

to controlled ovarian stimulation, oocyte yield, and cancellation rate in patients<br />

undergoing ovarian stimulation. Prospective investigation in a larger<br />

cohort is needed.<br />

P-644 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

THE TIMING OF ADMINISTRATION OF LETROZOLE SIGNIFI-<br />

CANTLY AFFECTS THE OOCYTE RECOVERY RATE IN BREAST<br />

CANCER PATIENTS UNDERGOING CONTROLLED OVARIAN<br />

STIMULATION FOR FERTILITY PRESERVATION. C. Diaz-Garcia,<br />

a J. Domingo, b A. Romero, a M. Martinez, c J. M. Rubio, d<br />

J. A. Garcia-Velasco, c A. Pellicer. a,e a Woman’s Health Area, La Fe University<br />

Hospital, Valencia, Spain; b IVI-Las Palmas, Las Palmas de Gran Canaria,<br />

Spain; c IVI Madrid, Madrid, Spain; d La Fe University Hospital, Valencia,<br />

Spain; e IVI-Valencia, Valencia, Spain.<br />

OBJECTIVE: Diminished ovarian response to controlled ovarian stimulation<br />

(COH) in breast cancer (BC) patients has been reported by different authors,<br />

although others do not find such differences. COH protocols in patients<br />

suffering from BC have some distinctive features: aromatase inhibitors are<br />

commonly used to avoid an increase in estradiol (E2) levels. To what extent<br />

this is a consequence of using modified stimulation protocols remains unclear.<br />

The aim of this study was to quantify the effect of two different protocols<br />

of COH on the number of oocytes retrieved in BC patients undergoing<br />

COH for fertility preservation.<br />

DESIGN: Observational multicentric prospective cohort study.<br />

MATERIALS AND METHODS: Patients were allocated to receive two<br />

different stimulation protocols: group A-patients started COH using letrozole<br />

5mg/d and hrFSH was added from the third day of stimulation.<br />

Group B-patients started COH using hrFSH and letrozole was added<br />

from the third day of stimulation to prevent the rise of estradiol (E2)<br />

levels. All patients started stimulation during the early follicular phase,<br />

GnRH antagonist was administered when the leading follicle reached<br />

14 mm, and final oocyte maturation was triggered with 0.2 mg of<br />

GnRH agonist. E2 levels the day of triggering, the total number of oocytes<br />

retrieved and the proportions of MII oocytes were compared between<br />

groups. A GLM-based multivariate regression analysis was<br />

applied to control for potential confounders.<br />

RESULTS: Patients in group A (n¼44) and group B (n¼257) had similar<br />

age and antral follicle count (AFC). Duration of stimulation was similar between<br />

groups. Patients in group A yielded higher estradiol level the day of<br />

ovulation triggering and higher numbers of total and MII oocytes retrieved<br />

(Table 1) when compared to patients in group B. The mean differences in<br />

oocytes retrieved (5.2-95%CI: 2.5-7.9) and MII oocytes retrieved (2.9-95%<br />

CI: 0.9-5.0) remained significant after adjustment for doses of rhFSH used<br />

during stimulation (total oocytes: 4.7-95%CI: 1.9-7.5 and MII: 2.5-95%CI:<br />

0.4-4.7).<br />

CONCLUSIONS: The use of letrozole before administration of rhFSH decreases<br />

the number of mature oocytes retrieved after COH in BC patients undergoing<br />

fertility preservation. On the other hand, the estradiol pick at<br />

ovulation triggering is higher when letrozole is administered after rhFSH.<br />

The total dose of rhFSH used during COH acts and effect modifier.<br />

Baseline Characteristics of the patients and COH results.<br />

Group A Group B p-value<br />

Age (years) 32.6 (3.4) 33.7 (3.9) n.s.<br />

AMH (pM)<br />

27.0 (<strong>17</strong>.0-43.0) 25.7 (13.9-35.3) n.s.<br />

Duration of COH (days) 9.3 (1.5) 9.3 (3.8) n.s.<br />

Starting dose of rhRSH (IU) 248.3 (59.2) <strong>17</strong>3.1 (82.0) p


analysis in integrating available literature evidence to support dose selection<br />

of recFSH.<br />

Simulated RecFSH dose-response at clinical relevant doses based on the model<br />

based meta analysis.<br />

recFSH Dose 100 IU 150 IU 200 IU 250 IU 300 IU<br />

Predicted Mean 7.16 9.99 10.6 10.83 10.94<br />

Oocytes Count<br />

(95 % CI)<br />

(4.7 – 8.3) (8.8 – 10.7) (9.7 – 11.3) (9.9 – 11.5) (9.9 – 11.7)<br />

References: Table Note: 10,000 simulations conducted incorporating<br />

parameter uncertainty.<br />

Supported by: Merck & Co., Inc., Kenilworth, NJ, USA.<br />

P-647 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

LUTEAL-PHASE OVARIAN STIMULATION VERSUS CONVEN-<br />

TIONAL OVARIAN STIMUATION IN PATIENTS WITH NORMAL<br />

OVARIAN RESERVE TREATED FOR IVF: A LARGE RETROSPEC-<br />

TIVE COHORT STUDY. N. Wang, a Y. Wang, b Y. Kuang. c a Department<br />

of Assisted Reproduction, Shanghai Ninth People’s Hospital, Shanghai Jiaotong<br />

University School of Medicine, Shanghai, China; b Reproductive Endocrinology,<br />

Shanghai, China; c Shanghai Ninth People’s Hospital, Shanghai,<br />

China.<br />

OBJECTIVE: To systematically assess the efficiency and security of the<br />

luteal-phase ovarian stimulation (LPS) strategy for infertility treatment by<br />

comparing with conventional ovarian stimulation protocols.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Patients with normal ovarian reserve<br />

taking ovum pick-up (OPU) cycles between April 2012 and September<br />

2013 from three ovarian stimulation protocols were enrolled (727 cycles<br />

from the LPS protocol, corresponding to 708 patients; 830 cycles from the<br />

mild treatment protocol, corresponding to 745 patients, and 1385 cycles<br />

from the short-term protocol, corresponding to 1287 patients). Main outcomes<br />

were number of mature oocytes retrieved, number of top-quality embryos<br />

obtained, implantation rate, pregnancy rate, live birth and ongoing<br />

pregnancy rate, and the neonatal outcomes.<br />

RESULTS: Compared with mild treatment protocol, the LPS was associated<br />

with higher number of mature oocytes retrieved per OPU cycle (10.97.6 vs.<br />

3.73.0, P0.05), live birth and<br />

ongoing pregnancy rate (44.4% vs. 41.7%, P>0.05) in the LPS and mild treatment<br />

protocols, respectively. Comparisons between the LPS protocol and<br />

short-term protocol indicated that the LPS could achieve more mature oocytes<br />

and top-quality embryos per cycle (10.97.6 vs. 9.15.5, 4.64.3 vs.<br />

3.73.1, respectively, both P<br />

95%th percentile for the study cohort i.e., > 13 days. Patients with history<br />

of poor response to ovarian stimulation or with PCOS were excluded. Non<br />

GnRH-antagonist-based cycles were excluded. Demographic characteristics<br />

included age, gravidity, parity, body mass index (kg/m 2 ) and infertility diagnosis.<br />

COS parameters recorded were total days of GnRH-antagonist administration,<br />

total dosage of gonadotropins administered (IU), peak E 2 level (pg/<br />

e328 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


mL), number of mature oocytes retrieved, fertilization rate (%) number of<br />

embryos transferred, and day of ET. Clinical pregnancy, biochemical pregnancy,<br />

spontaneous miscarriage, and live birth rates in patients undergoing<br />

COS < 13 days and > 13 days were estimated. Wilcoxon rank sum test<br />

and student’s t-test were utilized for continuous variables. Chi-square (c2)<br />

test with Mantzel-Hansel correction was used for categorical variables.<br />

Odds ratios (OR) with 95% confidence intervals (CI) for live birth were<br />

calculated. Statistical significance was set at P 13 days, respectively.<br />

Overall, there were no differences in demographic characteristics. Patients in<br />

the > 13 day group had more total antagonist days and total gonadotropins<br />

administered. COS < 13 days was associated with increased odds of clinical<br />

pregnancy 1.76 (95% CI 1.33 - 2.34) and live birth 1.70 (95% CI 1.33 - 2.34),<br />

which remained unchanged when adjusting for total antagonist days and total<br />

gonadotropins administered. Patients in the > 13 day group who had live<br />

births were younger (34.6 4.91 vs. 38.2 4.72; P < 0.001).<br />

CONCLUSIONS: Our findings suggest that COS < 13 days is associated<br />

with increased odds of clinical pregnancy and live birth. In patients undergoing<br />

COS > 13 days, younger age is associated with greater odds of live birth.<br />

Comparison of Demographics, COS response and IVF-ET outcomes.<br />

Parameter >13 days (n¼240) P<br />

Age (years) 37.6 (4.57) 37.5 (4.99) 0.67<br />

BMI (kg/m2) 23.0 (6.32) 23.4 (6.68) 0.34<br />

Total antagonist days 4.14 (1.44) 5.85 (2.34)


OBJECTIVE: Women of reproductive age with DOR may have regular<br />

menses but respond poorly to stimulation and/or have suboptimal fecundity<br />

compared to those of similar age. We sought to evaluate DOR patients<br />

engaged in IVF treated with either a luteal estradiol/gonadotropin-releasing<br />

hormone antagonist (antGnRH) (Estradiol-Priming protocol (EPP)) or with<br />

an oral contraceptive pill (OCP) daily micro-dose leuprolide acetate (Microflare)<br />

protocol (MFP). Our goal was to use age, baseline hormone values,<br />

number of injections, and outcome to optimize choice of ovarian stimulation<br />

protocol.<br />

DESIGN: Retrospective.<br />

MATERIALS AND METHODS: Patients who were identified as having<br />

DOR who underwent an IVF utilizing a MFP or an EPP were included.<br />

DOR was defined by: 1) history a canceled IVF; 2) poor response to stimulation<br />

(


significantly higher than for the isolated markers (age ¼ 0.67, AFC ¼ 0.74,<br />

AMH ¼ 0.75, FSH ¼ 0.61). The cut-off values and false positive rates for<br />

obtaining detection rates of 50% and 80% are presented in Table 1.<br />

CONCLUSIONS: The formula combining age, FSH, AMH, and AFC is<br />

more accurate in predicting POR than individual markers. More studies are<br />

still needed before recommending its use in clinical practice.<br />

CONCLUSIONS: IVF protocols that include letrozole yield comparable<br />

fertilization rates and an improved oocyte yield in BRCA-positive women.<br />

The addition of letrozole to the BRCA-positive patient’s IVF protocol should<br />

be considered regardless of cancer or estrogen-receptor status. Further<br />

studies are needed to determine whether these findings hold true in a larger<br />

cohort and to investigate pregnancy outcomes.<br />

Cut-off values and false positive rates (FPR) to predict poor ovarian response.<br />

Detection rate ¼ 50% Detection rate ¼ 80%<br />

Cut-off FPR Cut-off FPR<br />

Age R35 22% R33 57%<br />

AFC %4 13% %8 42%<br />

AMH %1,0 18% %1,5 43%<br />

FSH R7,0 24% R4,3 72%<br />

Formula R0,5 10% R0,37 27%<br />

P-655 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

LETROZOLE PROTOCOLS ARE ASSOCIATED WITH IMPROVED<br />

OOCYTE YIELD IN BRCA POSITIVE PATIENTS. A. P. Melnick,<br />

E. M. Murphy, G. L. Schattman, Z. Rosenwaks. The Ronald O. Perelman<br />

and Claudia Cohen Center for Reproductive Medicine, Weill Cornell Medical<br />

College, New York, NY.<br />

OBJECTIVE: To determine whether letrozole-based protocols improve<br />

ovarian response in BRCA-positive women.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Patients with BRCA mutations undergoing<br />

IVF for fertility preservation (breast cancer patients) or preimplantation genetic<br />

diagnosis (no history of cancer) were included. Patients with breast cancer<br />

who had received prior chemotherapy were excluded. Patients were stratified<br />

into two groups: those who received letrozole continuously throughout their cycle<br />

and those who received conventional GnRH agonist- or antagonist-based<br />

protocols without the addition of letrozole. Outcomes assessed included total<br />

and mature oocyte yield, oocyte maturity rate, fertilization rate, and cancellation<br />

rate. Statistical analysis included student’s t-test, chi-square and Fisher’s<br />

exact tests. P


EMBRYO TRANSFER<br />

Treatment outcomes.<br />

P-657 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

G-HRT<br />

protocol<br />

Previous<br />

protocol<br />

P-value<br />

PITUITARY SUPPRESSION BEFORE FROZEN EMBRYO TRANS-<br />

FER BENEFICIAL FOR THE IDIOPATHIC REPEATED IMPLAN-<br />

TATION FAILURE PATIENTS. X. Yang, R. Huang, X. Liang.<br />

Reproductive Medicine Centre, The Sixth Affiliated Hosptial of Sun Yat-<br />

Sen University, Guangzhou, China.<br />

OBJECTIVE: We were interested in whether long-term GnRHa pretreatment<br />

combine with hormone replacement therapy (G-HRT)could improve<br />

pregnancy outcomes in idiopathic repeated implantation failure(RIF) patients.<br />

DESIGN: 18 idiopathic RIF patients undertaken G-HRT protocol were<br />

included in this retrospective self-control study, compared clinical data<br />

with their 49 previous embryo transfer cycles.<br />

MATERIALS AND METHODS: 18 idiopathic RIF patients undertaken<br />

G-HRT Protocol. The first injection, leuproreline acetate 2.5 mg i.m.,<br />

was administered during the early follicular phase of the menstrual cycle<br />

for G-HRT patients. Twenty eight days later, 1.875 mg of leuproreline acetate<br />

was administered, and 15 days later the HRT protocol was started.<br />

Estradiol valerate was commenced orally for at least 10 days, progesterone<br />

intramuscular injection 40mg twice daily were added when the endometrial<br />

thickness is > 7mm. FETwas performed on the fourth day of administration<br />

of progesterone for day 3 embryo transfer or on the sixth day for blastocyst<br />

embryo transfer. The same doses of estrogen and progesterone were<br />

continued until a serum beta human chorionic gonadotropin(hcg) assay<br />

was performed 14 days after FET. If the assay was positive, HRT was<br />

continued until 10 weeks of gestation, and the patient was followed with ultrasonography<br />

to determine fetal viability until approximately 7 weeks of<br />

gestation. Pregnancy rate was defined by positive assay of hcg 14 days after<br />

FET. The Clinical pregnancy was defined by the presence of an gestational<br />

sac and a live fetus on TVS at 7 weeks of gestation. Implantation rate was<br />

defined as number of gestational sacs observed divided by the number of<br />

embryos transferred. In their previous cycles, HRT was performed identically<br />

as above description but without GnRHa pretreatment. Or embryos<br />

were transferred after ovulation with appropriate luteal support in Natural<br />

cycle.<br />

RESULTS: The data showed that after RIF, patients with advanced age<br />

result in significantly promotion of reproductive outcome in G-HRT protocol,<br />

the pregnancy rate, clinical pregnancy rate , implantation rate and<br />

on-going pregnancy rate were 0.67,0.61,0.44 and 0.33, compared with<br />

same parameters in their previous cycle were 0.<strong>17</strong>,0.11,0.07 and0.04,respectively.<br />

Similar number of embryo were transferred with same endometrium<br />

thickness.<br />

CONCLUSIONS: Hormonally controlled endometrial preparation with<br />

prior GnRHa suppression could be used for patients who have had RIF of<br />

IVF-ET treatment despite having morphologically optimal embryos and<br />

may be useful for increasing receptivity of the endometrium in these patients.<br />

Endometrium thickness on 10.840.52 11.390.35 0.42<br />

progesterone day (mm)<br />

E2 level on progesterone 148.<strong>21</strong>4.87 759.8142.7 0.0037<br />

administration day<br />

(pg/ml)<br />

p Level on progesterone 19.312.58 40.554.93 0.0042<br />

administration day<br />

(ng/ml)<br />

Mean No. of embryo 1.780.13 2.040.10 0.15<br />

transferred<br />

Pregnancy rate 0.67 0.1<strong>17</strong> 0.167 0.0537 0.0012<br />

Implantation rate 0.6111 0.12 0.1163 0.049 < 0.0001<br />

Clinical pregnancy rate 0.44 0.098 0.071 0.033


quality (n¼78), were more likely to be euploid than non-freeze quality blastocysts<br />

(n¼369) (64.1% vs., 43.0%, respectively; P


hatching with a differentiated inner cell mass and trophectoderm. All blastocysts<br />

were collapsed prior to freezing. Vitrifcation and warming were performed<br />

using Irvine Scientific freeze and thaw kits and Cryolocks. Patients<br />

who used donor eggs were excluded from this study. The outcomes of interest<br />

were positive hCG and positive clinical pregnancy rates (defined as the presence<br />

of fetal activity at 7 weeks).<br />

RESULTS: A total of 253 of the 354 (71%) FET patients had a positive<br />

hCG compared to 61 of the 94 (67%) of euploid FET patients. Of the FET<br />

patients, 182 (of 354) developed fetal cardiac activity at their 7 week ultrasound,<br />

resulting in 51% clinical pregnancy rates. In comparison, 35 (of 94)<br />

patients who transferred euploid embryos developed fetal cardiac activity, resulting<br />

in 53% clinical pregnancy rate. Statistical analysis by t-Test demonstrated<br />

no significant difference between standard frozen embryo transfer and<br />

transferring frozen euploid embryos.<br />

CONCLUSIONS: Vitrified blastocyst that reach freeze quality have no<br />

significant difference in positive hCG or clinical pregnancies compared<br />

to transferring vitrified euploid embryos. These results may be indicative<br />

that biopsy may impact the ability of the embryo to implant or develop.<br />

Also, the criteria of the embryos that are biopsied is not as consistent as<br />

what is classified as ‘‘freeze quality’’. This flexibility in grading may be<br />

detrimental to the embryo’s ability to survive the biopsy and subsequent<br />

transfer.<br />

P-662 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

AVOIDING CANCELLATION OF EMBRYO TRANSFER: EM-<br />

BRYOS WITH POOR MORPHOLOGY ON DAY 5 YIELD PREG-<br />

NANCIES AND LIFE BIRTHS. G. Koustas, H. Smith, C. Sjoblom.<br />

Obtetrics and Gynaecology, University of Sydney/ Westmead Fertility<br />

Centre, Sydney, Australia.<br />

OBJECTIVE: The objective of this study was to assess the occurrence<br />

and outcome of clinical pregnancies and live birth following transfer of<br />

embryos on day 5 of development with deviant morphological parameters.<br />

DESIGN: Retrospective analysis of fresh cycles with resulting poor quality<br />

embryos on day 5 for single embryo transfer (SET) and double embryo<br />

transfer (DET) between January 2010 and January <strong>2015</strong>.<br />

MATERIALS AND METHODS: This study included patients having conventional<br />

IVF and/or ICSI in a total of 5019 fresh non-donor transfer cycles<br />

conducted from January 2010 to January <strong>2015</strong>. Embryos were assessed on<br />

day 5 of development based on morphology. A poor embryo was identified<br />

as having inadequate compaction and/or poor or absent inner cell mass<br />

(ICM) and trophectoderm (TE), in accordance with the grading system<br />

defined by Gardner and Schoolcraft (1999).<br />

RESULTS: Out of 5019 fresh cycles, we identified 367 cycles where the<br />

embryos transferred had been assessed as poor. There was a significant difference<br />

in age between patients with poor quality embryos compared to<br />

the overall average age of patients during this period. The proportion of<br />

IVF/ ICSI and number of embryos transferred was found to be significantly<br />

higher in the poor embryo group (P


among women aged 35-39 allows the physician to choose between either<br />

fresh or FET according the patient’s needs. A very low ongoing pregnancy<br />

rate in women R40 suggests that donor oocytes cycles should be highly recommended<br />

for this age group.<br />

P-664 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DOES THE DEGREE OF EXPANSION AT TIME OF FROZEN EM-<br />

BRYO TRANSFER (FET) AFFECT OUTCOMES OF SINGLE<br />

THAWED EUPLOID EMBRYO TRANSFERS<br />

(STEET)? N. M. Sachdev, a D. H. McCulloh, b J. Grifo. c a Obstetrics and<br />

Gynecology, New York University Langone Medical Center Fertili, New<br />

York, NY; b New York University Fertility Center, New York, NY; c NYU Langone<br />

Medical Center, NY, NY.<br />

OBJECTIVE: To evaluate if the degree of blastocyst expansion following<br />

rewarming of euploid embryos in a frozen embryo transfer cycle affects on<br />

pregnancy outcomes.<br />

DESIGN: Retrospective Cohort Study.<br />

MATERIALS AND METHODS: Patients undergoing a frozen euploid<br />

embryo transfer from June 2013 to <strong>October</strong> 2014 were included. Preimplantation<br />

genetic screening using array comparative genomic hybridization<br />

designated embryos to be euploid or aneuploid. All embryos had undergone<br />

trophectoderm biopsy followed by vitrification. Following our standard FET<br />

protocol, vitrified euploid embryos were rewarmed on the day of transfer.<br />

Embryologists documented degree of expansion at time of transfer for all<br />

frozen embryo transfers. Patients with all causes of infertility, and both autologous<br />

and donor cycles were included. Primary outcome included pregnancy,<br />

defined by presence of a gestational sac, and clinical pregnancy, defined by<br />

live birth or fetal heart beat through cycle day 63. Statistical analysis was<br />

done using relative risk and Chi Square.<br />

RESULTS: A total of 228 embryo transfers were included in the study.<br />

The mean age of patients at time of embryo freezing included in the study<br />

was 36.20 +/- 4.59 (range <strong>21</strong>-45). There were 143 expanded, 64 partially<br />

expanded and <strong>21</strong> collapsed blastocysts at time of transfer. The overall<br />

pregnancy rate was 66.2% and clinical pregnancy was 60.1%. There is<br />

a higher rate of clinical pregnancies in embryos that were expanded at<br />

time of transfer (Table 1). The collapsed embryo group had similar pregnancy<br />

rates but was noted to have a lower clinical pregnancy rate.Blastocysts<br />

were biopsied and vitrified on day 5 (166,72.8%) or on day 6<br />

(62,27.2%). No difference was noted on its affect on pregnancy or clinical<br />

pregnancy rates.<br />

unknown whether hatching at this point is associated with poor embryo survival.<br />

This study sought to determine whether single embryo transfers (SETs)<br />

of euploid embryos that hatched during the warming process were associated<br />

with lower pregnancy rates (PRs).<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Patients underwent trophectoderm biopsy<br />

with qPCR-based comprehensive chromosome screening (CCS) from<br />

June 2011 to January <strong>2015</strong>. Only patients who underwent SET of a frozen<br />

euploid embryo were included. All embryos were graded according to the<br />

Gardner scale both prior to freeze by vitrification and after thaw prior to<br />

transfer. Patients were segregated into two groups: A) FETs of embryos<br />

that hatched from expansion 4 or 5 to expansion 6 during the warming process;<br />

and B) FETs of embryos expansion 4 or 5 that remained the same after<br />

re-warming. Contingency tables were generated for PR and clinical PR based<br />

on embryo hatching stage and the PR and clinical PR were computed. Statistical<br />

analysis was calculated by fisher exact test with significance at p< 0.05.<br />

RESULTS: Couples underwent 262 euploid frozen SETs of oocytes aged<br />

23.5-44.3 yo. Embryos that had hatched prior to biopsy to grade 6 at transfer<br />

(n¼86) experienced PRs (0.76, 0.65-0.84 95% CI) that were not significantly<br />

different from those (n¼<strong>17</strong>6) that remained grade 4 or 5 (0.78, 0.71-0.84 95%<br />

CI, p¼0.75). Clinical PRs from euploid SET in embryos that had hatched<br />

from grade 4 or 5 prior to biopsy to grade 6 (0.65, 0.54-0.75 95% CI) were<br />

not significantly different from those that remained grade 4 or 5 (0.66,<br />

0.59-0.73 95% CI, p¼0.89).<br />

Embryo Transfer Outcomes.<br />

Blast 4/5 Embryos at Freeze At Transfer Blast 6 vs 4/5<br />

Embryo Transfer Count <strong>17</strong>6 86<br />

Pregnancy 137 65<br />

Pregnancy Rate 0.78 0.76 0.88 [0.46-1.70]<br />

Clinical Pregnancy 1<strong>17</strong> 56<br />

Clinical Pregnancy Rate 0.66 0.65 0.94 [0.53-1.69]<br />

CONCLUSIONS: Extended culture has been improved and is more<br />

commonly applied in practice, consequently increasing the prevalence of<br />

blastocyst cryopreservation. Embryos hatching from their zona pellucida<br />

during the re-warming process experienced comparable PR and clinical<br />

PRs to those hatching prior to cryopreservation. There is no cause for concern<br />

when upon rewarming an unhatched embryo, complete escape from the zona<br />

pellucida occurs. Further studies utilizing larger subsets of patients are<br />

needed to confirm this study’s findings.<br />

Pregnancy Outcomes by Embryo Expansion at Time of Transfer<br />

N<br />

Total Pregnant (presence<br />

of gestational sac)<br />

Percent<br />

Total Clinical Pregnancy (fetal<br />

heart beat of Live Birth)<br />

Percent<br />

Expanded 143 106 74.12 95 66.43<br />

Partially Expanded 64 34 53.13 32 50.0<br />

Collapsed <strong>21</strong> 11 64.70 10 47.62<br />

Total 228 151 137<br />

P value 0.0047 0.0392<br />

CONCLUSIONS: There is a higher chance of implantation and ongoing<br />

pregnancies that lead to live birth in embryos that are re-expanded at time<br />

of transfer following the thaw of a euploid embryo.<br />

P-665 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DOES COMPLETE HATCHING AT THE TIME OF THAW NEGA-<br />

TIVELY AFFECT PREGNANCY RATES? J. Gingold, a J. Rodriguez-<br />

Purata, b M. C. Whitehouse, b B. Sandler, b A. B. Copperman. c,a a Obstetrics,<br />

Gynecology and Reproductive Science, Icahn School of Medicine at Mount<br />

Sinai, New York, NY; b Reproductive Medicine Associates of New York, New<br />

York, NY; c Reproductive Medicine Associates, New York, NY.<br />

OBJECTIVE: After re-warming for frozen embryo transfers (FET),<br />

embryonic expansion and hatching frequently often observed in biopsied embryos,<br />

even when they were previously unhatched at time of vitrification. It is<br />

P-666 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

ECONOMIC IMPLICATIONS OF THE SART GUIDELINES ON<br />

EMBRYO TRANSFER: HEALTHCARE DOLLARS SAVED BY<br />

REDUCING IATROGENIC TRIPLETS. M. S. Lee, a,b B. T. Evans, b,c<br />

M. D. Hornstein. a,b a Obstetrics and Gynecology, Brigham and Women’s Hospital,<br />

Boston, MA; b Harvard Medical School, Boston, MA; c Harvard Combined<br />

Orthopaedic Residency Program, Massachusetts General Hospital,<br />

Boston, MA.<br />

OBJECTIVE: Since SART first released guidelines in 1998 on number of<br />

embryos to transfer, there has been a consistent decrease in the number of embryos<br />

transferred per cycle and the percentage of higher order multiple (HOM)<br />

gestations resulting from ART in the United States. The purpose of this study<br />

was to estimate the total national cost savings resulting from reductions in<br />

FERTILITY & STERILITY Ò<br />

e335


HOM gestations since the publication of the initial SART guidelines on embryo<br />

transfer (ET).<br />

DESIGN: Descriptive cost analysis and three-point estimation.<br />

MATERIALS AND METHODS: A systematic review of the literature was<br />

conducted to estimate the hospital costs of singleton and triplet pregnancies<br />

and subsequent delivery. National data reported to the CDC from 1996 to<br />

2012 were reviewed. The trends in ET and reductions in resulting HOM gestations<br />

were examined. The number of HOM deliveries prevented following<br />

the introduction of SART guidelines was estimated. The total costs were estimated<br />

based on direct hospital charges for associated obstetric and perinatal<br />

conditions/complications, and projected to 2014 US dollars.<br />

RESULTS: A singleton gestation (including pregnancy, delivery and up to<br />

one year of neonatal care) was estimated to cost between $<strong>17</strong>,112-24,<strong>21</strong>2. A<br />

triplet gestation was estimated at $190,788-453,935. Comparable estimates<br />

of singleton and triplet gestations demonstrated the latter to be between 11<br />

and 27 times as costly. The percentage of HOM gestations amongst all<br />

ART pregnancies decreased from 11.4% in 1997 to 2.0% in 2012, with the<br />

sharpest year-to-year decline of 20.3% occurring from 1998-1999, the year<br />

following the publication of the initial SART guidelines. Similarly, the number<br />

of liveborn HOM infants secondary to ART has decreased from 59.8 per<br />

1,000 fresh non-donor cycles in 1997 to 13.2 in 2012. From 1998 to 2012, the<br />

cumulative number of prevented HOM deliveries was estimated to be 13,512.<br />

This corresponds to an estimated net total savings in direct hospital-related<br />

costs of $4.27B (range $2.35-5.85B, 2014 dollars).<br />

CONCLUSIONS: Iatrogenic HOM gestations represent a substantial economic<br />

burden to our healthcare system. Changes to the practice of ET<br />

following the publication of the initial SART guidelines in 1998 have resulted<br />

in a dramatic decrease in the HOM rate. The associated cumulative<br />

cost savings to the US healthcare system are estimated to be over $4B.<br />

Supported by: No financial support. We are indebted to Drs. Dmitry Kissin<br />

and Sara Crawford at the CDC for their invaluable assistance and support.<br />

P-667 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

FROZEN BLASTOCYST TRANSFER IN NATURAL CYCLE VS<br />

HORMONE REPLACEMENT THERAPY CYCLE. A. Biryukov, a<br />

V. Apryshko, a,b I. Zorina, a E. Osina, a N. Dmitrieva, a S. Yakovenko. a,c a Altravita<br />

IVF clinic, Moscow, Russian Federation; b Faculty of Biology, Moscow<br />

State University, Moscow, Russian Federation; c Faculty of Physics,<br />

Moscow State University, Moscow, Russian Federation.<br />

OBJECTIVE: The introduction of vitrification in assisted reproduction<br />

technology leads to significant increase in embryo survival rates by up to<br />

95-100%. Frozen blastocyst transfer (FBT) treatment can be scheduled in a<br />

natural ovulatory cycle or in a hormone replacement therapy (HRT) in which<br />

the endometrium is prepared with estrogen and progesterone supplementation.<br />

Natural cycle FBT is easy to implement and can be offered to women<br />

that have a regular menstrual cycle and are proven to ovulate. Conversely,<br />

HRT is an option for anovulatory patients and offers flexibility for the doctor<br />

by controlling the timing of the cycle. In this research the hypothesis is that<br />

there is no significant difference in efficiency between natural FBT and HRT<br />

FBT.<br />

DESIGN: Retrospective analysis of 500 natural FBT cycles and 500 HRT<br />

FBT cycles. Clinical pregnancy, miscarriage and live births rates were<br />

compared. Cycles with donor embryo transfers were excluded from this<br />

research.<br />

MATERIALS AND METHODS: Embryos were cultured in Global total<br />

media at 7.3% CO2, 36.6 C for 5-6 days and were vitrified using CryoTech<br />

vitrification method at blastocyst stage. Only AA, AB, BA and BB grades<br />

(Gardner’s system) of blastocysts were frozen. After thawing, blastocysts<br />

were cultured for 120 minutes prior to transfer in order to make sure of survival.<br />

The average number of transferred blastocysts was 1.4 for natural cycle<br />

and 1.3 for HRT. The average age of patients was 34.8 for natural cycle and<br />

35.1 for HRT. Statistical differences between the values were made using<br />

analysis of variance (ANOVA).<br />

RESULTS: Clinical pregnancy rates for natural cycle FBT were greater<br />

than HRT FBT (43.7% vs 39.8%). However this was not statistically significant<br />

(p¼0.26). Miscarriage rates for natural cycle FBT were greater than<br />

HRT FBT (33.7% vs 36.4%). However this was not statistically significant<br />

(p¼0.2). Live birth rates for natural cycle FBT were lower than HRT FBT<br />

(38.7% vs 44.3%). However this was not statistically significant (p¼0.14).<br />

CONCLUSIONS: This retrospective cohort study found no significant<br />

difference in clinical pregnancy, miscarriage and live birth rates between<br />

natural cycle FBT and HRT FBT. At present our success rates for natural<br />

cycle vs HRT FBT suggest that if a patient has a regular menstrual cycle<br />

then they should be offered a natural cycle FBT. This avoids the use of<br />

multiple medications, reduces the cost for patients, and simplifies the<br />

treatment.<br />

P-668 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DEMOGRAPHIC CHARACTERISTICS ASSOCIATED WITH ELEC-<br />

TIVE SINGLE EMBRYO TRANSFER IN IVF: WHO CHOOSES JUST<br />

ONE? E. M. Munch, a K. M. Summers, a G. Ryan, a J. D. Kapfhamer, a<br />

B. Collura, b G. D. Adamson. c a Reproductive Endocrinology and Infertility,<br />

University of Iowa Hospitals and Clinics, Iowa City, IA; b RESOLVE,<br />

McLean, VA; c PAMF Fertility Physicians of Northern California, Saratoga,<br />

CA.<br />

OBJECTIVE: Younger patient age has been found in previous IVF studies<br />

to be associated both with desire for twins and the choice of single embryo<br />

transfer; we sought to examine this and other demographic and geographic<br />

parameters in relation to elective single embryo transfer (eSET).<br />

DESIGN: Descriptive analysis of online survey results from a crosssectional<br />

sample of U.S. community women.<br />

MATERIALS AND METHODS: An online survey, advertised through<br />

RESOLVE (the National Infertility Association) was conducted over 5<br />

weeks in 2014. Interested participants were screened for gender and cycle<br />

eligibility and gave consent by acknowledging an online privacy statement.<br />

Inclusion criteria were age < 40 and the completion of at least one IVF cycle<br />

with embryo transfer. Participants were asked to identify their age, race,<br />

income, highest education completed, and insurance coverage in the survey.<br />

The outcome of interest was the election of single embryo transfer (eSET,<br />

defined as multiple embryos available for transfer but electing a single embryo<br />

transfer) versus multiple embryo transfer (MET) in the first IVF cycle.<br />

We examined geographic characteristics by US Census region and by state<br />

IVF access quartile, as previously published. 1 All variables, including age<br />

and income, were categorical, and X 2 was used to compare proportions<br />

among groups.<br />

RESULTS: Of 888 participants, 587 met age and cycle inclusion criteria.<br />

Participants who chose eSET tended to be younger than those choosing MET,<br />

with 25% of participants under 29 choosing eSET, compared to 12% of those<br />

aged 35-39 (p¼0.014). Education level, race, income, and insurance<br />

coverage for infertility did not differ between eSET and MET groups. There<br />

was no association between eSET or MET with regard to accessibility to IVF<br />

in the participant’s state. When comparing patterns of eSETand METaccording<br />

to US Census region, patients from the Midwest were significantly more<br />

likely to choose MET over eSET (91% for the Midwest vs an average of 80%<br />

for other census regions, p¼0.003).<br />

CONCLUSIONS: Even in patients


DESIGN: Retrospective cohort study of fresh first IVF cycles started May<br />

2012 to November 2013.<br />

MATERIALS AND METHODS: Among 1186 cycles, 857 had a day 3 ET<br />

and 329 a day 5 ET; day 5 ET was based on a clinical algorithm that included<br />

day 3 availability of R3 embryos with at least 8-cells with


alance between the potential cryo-damage to the embryos from the vitrification<br />

procedure and improved endometrial receptivity. FET does not increase<br />

miscarriage rate.<br />

IMPLANTATION<br />

P-673 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

EMBRYO S-PHASE LENGTH ANALYSIS ON FIRST AND SECOND<br />

CELL CYCLE AND ITS RELATIONSHIP WITH REPRODUCTIVE<br />

OUTCOME. J. A. Aguilar, a E. Munoz, b A. Galan, c Y. Motato, c<br />

M. Ojeda, b V. Garcia, c M. Meseguer. d a IVF Laboratory, IVI Vigo, Vigo,<br />

Spain; b IVI Vigo, Vigo, Spain; c IVI Valencia, Valencia, Spain; d Clinical<br />

Embryology, Valencia, Spain.<br />

OBJECTIVE: to analyse the length of the s-phase (synthesis of DNA), in<br />

the first cell cycle (ECC1) and in each blastomere of the second cell cycle<br />

(ECC2) of known implantation data (KID) embryos, and to relate them<br />

with reproductive outcome.<br />

DESIGN: Observational retrospective study of 1679 transferred embryos<br />

from 940 ICSI patients from our egg donation program between January<br />

2011 and January 2014 in IVI Vigo and IVI Valencia, cultured in Embryoscope.<br />

Sperm samples under 1 million/ml were used as an exclusion<br />

criteria.<br />

MATERIALS AND METHODS: Only non-multinucleated embryos<br />

which either failed to implant or fully implanted were included in the study.<br />

The s-phase was calculated as the period while nuclei were visible, being<br />

ECC1 S-phase¼tPNfading-tPNappearance, and ECC2 S-phase¼<br />

t2MONO1(f)- t2MONO1(a). ANOVA test and c2-test were performed<br />

when applicable to assess the influence of the length of S-phase in the implantation<br />

rate.<br />

RESULTS: S-phase in both cell cycles were calculated in 904 KID embryos<br />

out of 1679, and therefore were analyzed. 32,74% (n¼296) fully implanted<br />

(KID+), and 67,25% (n¼608) failed to implant (KID-). The<br />

average length of s-phase in the ECC1 in the KID+ embryos was longer<br />

than in KID-, 15,50h v.s. 14,38h, and slightly shorter in ECC2, 8,35 in<br />

KID+ v.s. 8,60h in KID-. 46,1% (n¼112) of the embryos implanted when<br />

the difference between both s-phases was greater than 7.96h, being 29,1%<br />

for those where the difference is between 5,96h-7,95h; 23,1% for those between<br />

3,61h-5,95h; and 30% for those below 3,6h.<br />

CONCLUSIONS: Results show that the s-phase ECC1, nearly double the<br />

length of s-phase ECC2, and the greater the difference between both s-phase,<br />

the greater the embryo implantation, suggesting that the replication in the<br />

first cell cycle may implicate particularities related to the genome combination,<br />

and that those embryos which do not implant may not complete the<br />

replication of its genome during the first cycle producing a larger reduction<br />

in the duration of the second.<br />

P-674 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

REPEATED IMPLANTATION FAILURE RELATES TO CIRCULA-<br />

TORY ABNORMALITIES. D. A. Pattinaja, M. E. Spaanderman,<br />

C. Ghossein-Doha, R. J. van Golde. Obstetrics & Gynaecology Department,<br />

Maastricht University Medical Centre, Maastricht, Netherlands.<br />

OBJECTIVE: A diminished pre-pregnant plasma volume as marker of<br />

reduced cardiovascular reserves may influence embryo implantation by circulatory<br />

redistribution at the expense of uterine perfusion. In this pilot study<br />

we tested the hypothesis that recurrent implantation failure in in-vitro-fertilization<br />

(IVF) patients is associated with decreased plasma volume and<br />

increased uterine arteries resistance as proxy measures for circulatory anomalies.<br />

DESIGN: In this cross-sectional pilot study we included patients who had<br />

at least 2 embryo transfers after IVF treatment (n¼10) and a control group<br />

consisting of women with a history of uncomplicated reproduction (n¼10).<br />

Both groups were assessed for mean arterial blood pressure (MAP), plasma<br />

volume and the pulsatility index (PI) of both uterine arteries.<br />

MATERIALS AND METHODS: Blood pressure was measured at three<br />

minute intervals during a half hour in sitting position. Plasma volume was<br />

measured by the iodine 125-labeled albumin indicator dilution method and<br />

expressed in ml/ body surface area (m2). Resistance of the uterine arteries<br />

was obtained by measuring the pulsatility index by transvaginal doppler<br />

flow ultrasonography. Data was analyzed using the Mann Whitney U test.<br />

A p-value below 0.05 was considered statistically significant.<br />

RESULTS: The median duration of subfertility in the study group is 39<br />

months [IQR 32-51]. The conventional diagnoses of subfertility were male<br />

factor (n¼6; one women combined with endometriosis and one women combined<br />

with cycle irregularity), unexplained subfertility (n¼3) and tubal pathology<br />

(n¼1). Baseline characteristics were comparable between both<br />

groups (Table 1). Plasma volume was lower in the IVF group (1313 [IQR<br />

1189-1406] ml/m2) compared to the control group (1468 [IQR 1362-1551]<br />

ml/m2). Furthermore the pulsatility indices of the uterine arteries were higher<br />

in the IVF group compared to the control group with 3.5 [IQR 3.0-3.8] vs 2.1<br />

[IQR 1.6-2.7] for the right branch and 3.2 [IQR 2.7-3.7] vs 2.2 [IQR 1.5-2.5]<br />

in the left branch respectively.<br />

CONCLUSIONS: Repeated implantation failure after IVF relates to<br />

diminished plasma volume along with increased uterine vascular resistance.<br />

We speculate that reduced plasma volume relates to circulatory redistribution<br />

at the expense of the uterine perfusion, which in turn negatively affects reproductive<br />

performance.<br />

Baseline characteristics and circulatory variables, presented as median with<br />

interquartile range.<br />

Control<br />

group(n¼10)<br />

Study group<br />

(n¼10)<br />

p-<br />

value<br />

Age (years) 32 (30-33) 31 (29-35) 0.80<br />

Height (metres) 1.7 (1.6-1.7) 1.6 (1.6-1.7) 0.35<br />

Weight (kilograms) 60.6 (56.6-65.3) 67.5 (57.8-75.0) 0.<strong>17</strong><br />

Body mass index (kg/m2) <strong>21</strong> (20-24) 23 (<strong>21</strong>-29) 0.08<br />

Body surface area (m2) 1.7 (1.7-1.8) 1.8 (1.6-1.9) 0.82<br />

Parity 2 (1-3) 0 (0-0) 0.02<br />

Mean arterial pressure 88 (81-108) 88 (82-94) 0.87<br />

(mmHg)<br />

Plasma volume per body 1468 (1362-1551) 1313 (1189-1406) 0.02<br />

surface area (ml/m2)<br />

PI right uterine artery 2.1 (1.6-2.7) 3.5 (3.0-3.8)


47.8%, p


are generated, the implantation rates are independent of female reproductive<br />

age. The application of CCS appears to be useful in the identification of blastocysts<br />

with high implantation potential in women of all reproductive ages.<br />

Implantation rates in each age group after vitrified euploid blastocysts ET.<br />

FET<br />

cycles (n)<br />

Total<br />

number of<br />

blastocysts<br />

transferred<br />

(n)<br />

Mean<br />

number of<br />

blastocysts<br />

transferred<br />

(n)<br />

Heart<br />

Beat# (n)<br />

Implantation<br />

rates (%)<br />

Donor 126 202 1.6 111 55%<br />

42y 14 <strong>17</strong> 1.2 11 65%<br />

P-679 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

EMBRYO CYTOPLASMATIC WAVE AND EVENNESS OF BLASTO-<br />

MERES CHARACTERIZATION AT THE END OF CELL CYCLE<br />

AND THEIR RELATION WITH CLINICAL<br />

OUTCOME. J. A. Aguilar, a E. Munoz, b M. Ojeda, b E. Taboas, b<br />

M. Perez, b M. Meseguer. c a IVF Laboratory, IVI Vigo, Vigo, Spain; b IVI<br />

Vigo, Vigo, Spain; c Clinical Embryology, Valencia, Spain.<br />

OBJECTIVE: Early cell cycle events have recently been studied as<br />

possible markers for embryo selection and related with implantation rate.<br />

The aim of this research is to characterize the cytoplasmic wave and the evenness<br />

of blastomeres at the end of 2 cells and 4 cells stage in embryos from<br />

fresh donated oocytes with known implantation data.<br />

DESIGN: Observational retrospective study conducted in IVI Vigo, between<br />

January 2011 and January 2014. 275 transferred embryos, cultured<br />

in an Embryoscope incubator in a 37 C, 6% CO2 and 20% O2 atmosphere,<br />

from <strong>17</strong>1 oocyte donation cycles were analyzed.<br />

MATERIALS AND METHODS: Cytoplasmic wave was annotated at its<br />

beginning and its end, and its duration was calculated as CWduration¼<br />

tCWend- tCWinitiation. CWduration were categorized and the frequency<br />

of embryos was also registered. The symmetry of the embryo blastomeres<br />

were assessed at the end of the 2cells, and 4 cells stage, in order to avoid<br />

the change in their volume due to the embryo dynamic behaviour. Only those<br />

embryos with which either failed to implant or fully implanted (KID+) were<br />

included in the study. ANOVA test, t-test and c2-test were performed when<br />

applicable to assess the influence of the continuous and categorical variables<br />

in the implantation rate.<br />

RESULTS: 275 embryos were analyzed. CW was always present, 19.2%,<br />

of embryos showed a CWduration shorter than 1.5h, 58,5% between 1,5-3h,<br />

19.6% between 3-4,5h and 2,54% longer than 4.5h. No significant differences<br />

were found between these categories and embryo implantation<br />

(p¼0.13)In KID+ embryos blastomere evenness was more frequent 33.8%<br />

vs 8% (p¼0.26) (KID-) at the end of the 2cells stage. Differences were<br />

greater at the end of the 4cells stage, and statistically significant 37.09%<br />

vs 4,72 (p¼0.025).<br />

CONCLUSIONS: The majority of embryos present a CW grouped between<br />

1,5h and 3h, although it seems not to be related to embryo implantation<br />

The unevenness of the blastomeres at the end of the 4 cells stage is negatively<br />

related to embryo implantation, while it seems not to be at the end of the 2<br />

cells stage. Although numbers from these observations are still reduced,<br />

we suggest to annotate the latter embryo morphological characteristic<br />

following our descriptions, due to the continuous modification of embryo<br />

blastomeres during the cell cycles, enhancing the annotation at the end of<br />

the 4 cells stage, and considering for the future elaboration of embryo selection<br />

algorithms.<br />

P-680 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

HOW WIDE IS THE UTERINE IMPLANTATION<br />

WINDOW? D. H. McCulloh, a C. McCaffrey, b J. Grifo. c a Obstetrics<br />

and Gynecology, New York University Fertility Center, New York, NY;<br />

b NYU Fertility Center, New York, NY; c NYU Langone Medical Center,<br />

New York, NY.<br />

OBJECTIVE: The period of time when embryos implant following natural<br />

cycles with intercourse or following in vitro fertilization is quite<br />

diverse spanning more than one week. Diverse times of implantation<br />

(TIs) are consistent with at least two different models: 1) the uterus is<br />

receptive to implantation for a long period of time following its onset or,<br />

2) the uterus is receptive to implantation for a brief period with diverse<br />

times of onset. Our objective was to determine whether the window of uterine<br />

receptivity is wide (model 1) or narrow (model 2) and to determine just<br />

the window’s width.<br />

DESIGN: Retrospective Analysis.<br />

MATERIALS AND METHODS: Human Chorionic Gonadotropin<br />

(hCG) levels and fetal sacs were assembled for patients who became pregnant<br />

and where all detected fetal sacs delivered. TIs were estimated by<br />

extrapolating the log(hCG) levels on cycle days 28 and 35 to a time<br />

when the hCG was10 mIU/mL per fetal sac. Although we refer to this<br />

time as TI, it, strictly speaking, lags behind implantation with a delay.<br />

Distributions for two TIs were compared: singletons versus twin pregnancies<br />

following fresh, day 3 embryo transfers. The method to determine the<br />

width of the uterine window of receptivity relied upon narrowing of the TI<br />

distribution that occurs when twins implant (Our method estimates only<br />

one ‘‘averaged’’ TI when twins implant at diverse times. Narrowing is<br />

analogous to standard errors with 2 samples being smaller than the standard<br />

deviation.).<br />

RESULTS: TIs for singletons averaged 7.6 +/- 2.1 days (N ¼ 231) after<br />

oocyte retrieval. TIs for twins averaged 8.1 +/- 1.8 days (n ¼ 89 twin<br />

pregnancies). Standard deviation of the distribution of TIs estimated of<br />

the width of the distribution of TIs. The distribution for twins was<br />

0.86 X the width for singletons. When twins implant simultaneously,<br />

the distribution of TIs is expected to be as wide as the distribution of<br />

TIs for singletons. The twin distribution was significantly narrower<br />

than this expectation. Twins implanting randomly throughout the distribution<br />

of times seen for singletons would be expected to lead to a<br />

twin distribution narrowed to 0.71 X the singleton TI distribution. The<br />

twin distribution was significantly wider than this expectation. Modelling<br />

the implantation of twins within a defined interval led to estimates that<br />

the width (standard deviation) of the uterine window of receptivity is<br />

0.6 times as wide as the standard deviation of the singleton implantation<br />

distribution (1.26 days).<br />

CONCLUSIONS: The standard deviation (width) of the window of uterine<br />

receptivity of 1.26 days (0.6 x the width of the singleton distribution) is<br />

consistent with a restricted window of uterine receptivity. Therefore, we<br />

believe that Model 2 is correct, indicating that there is a variable time of onset<br />

with a restricted period of uterine receptivity. This suggests that as many as<br />

20-40% of embryos that are capable of implantation may fail to implant simply<br />

due to asynchrony between embryo and uterus.<br />

P-681 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DOES ENDOMETRIAL INJURY IMPROVE PREGNANCY OUT-<br />

COMES IN OVUM DONATION CYCLES? E. Szlit Feldman, a<br />

E. A. Salama, a A. Torno, b L. Ferle, b G. Arruguete, c A. Sdrigotti. d a Procrearte<br />

Director, Buenos Aires, Argentina; b Gynecologist - Procrearte, Buenos<br />

Aires, Argentina; c Procrearte, Buenos Aires, Argentina; d gynecologist and<br />

obstetrician. fellow in reproduct, Buenos Aires, Argentina.<br />

OBJECTIVE: To explore the possibility that local endometrial injury<br />

(LEI) improves the clinical outcomes in ovum donation recipients.<br />

DESIGN: Retrospective cohort analysis.<br />

Main results.<br />

Study Group<br />

(n¼66)<br />

Control<br />

Group(n¼104)<br />

p Value<br />

Donor age (min-max) 24,3 (<strong>21</strong>-30) 24,8 (<strong>21</strong>-30) 0.5<br />

Recipients age<br />

43,5 (33-55) 41,85 (33-55) 0.13<br />

(min-max)<br />

Severe Male factor 1/66 (1.5%) 4/106 (3.7%) 0.69<br />

Number of embryos 2,01 2,01 NS<br />

transfered<br />

Implantation Rate( %) 32/133 (24%) 43/<strong>21</strong>0 (20.47%) 0.5<br />

Clinical Pregnancy (%) 24/66 (36%) 31/104 (29.8%) 0.47<br />

Multiple pregnancy<br />

Twins<br />

3/24 (12.5%) 8/31 (25.8%) 0.37<br />

e340 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


MATERIALS AND METHODS: A total of <strong>17</strong>0 ovum donation cycles<br />

were included from 2012 to 2014 in a private fertility center: the study group<br />

included 66 patients who underwent the local endometrial injury; the control<br />

group included 104 patients without the experimental treatment. Oocyte donors<br />

were <strong>21</strong>-30 years old with proven fertility, and recipients were women<br />

with indication of ovum donation due to diminished ovarian reserve or<br />

advanced maternal age. The local injury was performed once on day 20-22<br />

of the previous menstrual cycle. The endometrial scratch was done by vacuum<br />

aspiration with a Frydman Catheter and a 5 cc. syringe on the uterine<br />

wall until a sample of tissue was removed through the tube. Statistical analysis<br />

was performed using the SPSS. p< 0.05 was considered statistically significant.<br />

RESULTS: There were no demographic differences between the groups.<br />

The difference between implantation, clinical pregnancy and multiple pregnancy<br />

rates was not statistically significant.<br />

CONCLUSIONS: Our results show that local injury of endometrium by<br />

aspiration did not improve implantation, pregnancy and multiple pregnancy<br />

rates in ovum recipients. To our knowledge this is the second study (1) presented<br />

in order to evaluate the influence of LEI on ovum donation recipients.<br />

The benefit of LEI is controversial in cycles of IVF in patients with recurrent<br />

implantation failure. The advantage of using oocyte donor cycles in the analysis<br />

of LEI is that it eliminates the potential bias of oocyte quality on pregnancy<br />

outcome. A prospective randomized study will be conducted soon in<br />

order to elucidate the controversies of results when using this procedure in<br />

different studies.<br />

Reference:<br />

1. Dain L, Ojha K, Bider D, Levron J, Zinchenko V, Walster S, Dirnfeld<br />

M. Effect of local endometrial injury on pregnancy outcomes in<br />

ovum donation cycles. Fertil Steril. 2014 Oct;102(4):1048-54.<br />

P-682 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

MECHANICAL ENDOMETRIAL INJURY DOES NOT INCREASE<br />

SERUM LEVELS OF PRO-IMPLANTATION<br />

CYTOKINES. A. P. Melnick, a I. Ramer, a E. M. Murphy, a<br />

Z. Rosenwaks, a S. Witkin, b S. D. Spandorfer. a a The Ronald O. Perelman<br />

and Claudia Cohen Center for Reproductive Medicine, Weill Cornell Medical<br />

College, New York, NY; b Obstetrics and Gynecology, Weill Cornell Medical<br />

College, New York, NY.<br />

OBJECTIVE: Mechanical endometrial injury via endometrial biopsy or<br />

hysteroscopy in the menstrual cycle prior to in vitro fertilization (IVF) has<br />

been proposed as a way to increase implantation rates by improving endometrial<br />

receptivity. Specifically, endometrial injury may induce an inflammatory<br />

response with upregulation of cytokines, adhesions molecules, and growth<br />

factors critical to implantation. We have previously demonstrated an association<br />

between cytokine levels and IVF outcome. The goal of this study was to<br />

determine whether mechanical endometrial injury affects serum levels of<br />

pro-implantation cytokines in patients undergoing IVF with autologous<br />

endometrial coculture (AECC).<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: Patients with a history of prior IVF<br />

failure undergoing IVF with autologous endometrial coculture at our center<br />

were included. All patients underwent luteal phase endometrial biopsy<br />

in preparation for IVF. Biopsy samples were either utilized for IVF in the<br />

consecutive menstrual cycle or were frozen and thawed for use in a future<br />

cycle. Cycles were stratified into two groups by biopsy timing: biopsy<br />

performed in the cycle prior to IVF and biopsy performed more than<br />

one cycle prior to IVF. Serum samples were obtained from day of IVF cycle<br />

start and levels of cytokines implicated in implantation were determined<br />

by quantitative ELISA performed by blinded personnel. Primary<br />

outcome was levels of IL-1b, IL-6, IL-11, IL-<strong>17</strong>, IGF-1, IGF-2, TNF-a,<br />

IL-1RA, IFN-g, BDNF, NT-3, and NT-4. Statistical analysis included<br />

Mann-Whitney U, chi-square, and Fisher’s exact tests. P140 pg/mL, Progesterone > 10 ng/mL at the day of transfer),<br />

with good blastocyst (3BB or more with Gardner’s classification), were<br />

included in the study. Pregnancy test (serum hCG), serum copper, zinc and<br />

ceruloplasmin concentrations were measured 16 days after the first date of<br />

progesterone replacement. Hundred women with pregnant without miscarriage<br />

at 10 weeks of gestation (group P) and <strong>17</strong>6 women without pregnant<br />

(group NP) were compared. Student’s t test was used for comparison between<br />

two groups. Cutoff line was determined by receiver-operator characteristic<br />

(ROC) curve, and Chi-squared test was used for 2x2 contingency table. Statistical<br />

significance was defined as P


(embryos were biopsied day 5, vitrified day 6). A logistic regression analysis<br />

was used to determine a correlation of elapsed time in culture from biopsy to<br />

FET and PR.<br />

RESULTS: A total of 362 cycles met criteria for inclusion (SDBV¼325<br />

and NDBV¼37). PR observed for SDBV was 75.2% and for NDBV was<br />

70.3%. No correlation was found between time in culture from biopsy to<br />

FET and pregnancy outcomes (p>0.05).<br />

CONCLUSIONS: Vitrification is an excellent tool in the IVF laboratory,<br />

enabling better clinical flexibility in planning a patient’s IVF protocol.<br />

This study is the first of its kind to analyze the impact of delayed vitrification<br />

of biopsied embryos. We found no correlation between time in culture to<br />

successful ET. Embryologists can be reassured that prolonging culture of<br />

embryos following biopsy does not negatively impact implantation.<br />

P-685 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

PREDICTABILITY OF BLASTOCYST, EUPLOIDY AND IMPLAN-<br />

TATION RATE WITH EARLY MORPHOKINETICS<br />

PARAMETERS. E. Rocafort, A. Leza, M. Guijarro, L. Medrano,<br />

B. Ramos, M. Fernandez, J. Aizpurua. IVF Spain, Alicante, Spain.<br />

OBJECTIVE: To determine if early morphokinetic parameters , such as<br />

P2 (time from 2 to 3 cells) and P3 (time from 3 to 4 cells), are correlated<br />

with the likelihood of blastocyst formation, euploidy and implantation<br />

rate.<br />

DESIGN: Unicentric and restrospective (Sept.2013 - April <strong>2015</strong>). A total<br />

of 76 patients were included to a private clinic Preimplantation Genetic<br />

Screening (PGS) program. 740 embryos were cultured into<br />

EevaÔ. EevaÔ is an automated time-lapse system that provides on Day<br />

3 of embryo culture a prediction (High/Medium/Low) depending on P2<br />

and P3 morphokinetic parameters. Euploid embryos with Known Implantation<br />

Data (KID) and their implantation depending on EevaÔ prediction<br />

were studied.<br />

MATERIALS AND METHODS: All embryos were cultured and biopsied<br />

on Day 5-6 of embryo development and screened by Next Generation<br />

Sequencing. All transfers were differed and under HRT. Embryos selected<br />

for transfer were based on comprehensive-chromosomal-screening results<br />

and morphology criteria. Implantation was calculated using known implantation<br />

confirmed by ultrasound at 6-7 weeks.<br />

RESULTS: In 740 embryos, 29.8% were High, 20.5% Medium and 49.7%<br />

Low. Blastocyst formation rates for Eeva High, Medium and Low embryos<br />

were 80.9% (<strong>17</strong>8/220), 72.4% (110/152) and 40.2% (148/368), respectively.<br />

The difference in blastocyst formation rates between Eeva High, Medium and<br />

Low embryos was statistically significant. Euploidy rates were 35.5% (78/<br />

220) in High embryos, 31.6% (48/152) Medium and 18.5% (68/368) Low<br />

with significant differences from High and Medium to the Low group<br />

(p< .0001). However, no significant difference was found between Blastocyst<br />

Euploidy rates for Eeva High, Medium and Low (43.8% (78/<strong>17</strong>8), 43.6% (48/<br />

110) and 45.9% (68/148), respectively). A total of 62 KID euploid blastocyst<br />

were analyzed, being their implantation rate for High, Medium and Low of<br />

52.6% (20/38), 33.3% (3/9) and 13.3% (2/15), respectively. The difference<br />

in KID euploid blastocyst between High and Low was statistically significant<br />

(p¼0.009).<br />

CONCLUSIONS: Based on our results, P2 and P3 are correlated with blastocyst<br />

formation, euploidy and implantation rates. However, these parameters<br />

are not useful in euploidy predictability when embryo culture is<br />

performed until Day 5. Automated time-lapse systems may be useful in combination<br />

with PGS in terms of selecting the blastocyst with the highest potential<br />

among euploid embryos. This preliminary results may indicate that<br />

implantation rates do not only depend on the ploidy of the embryo as a factor<br />

for a successful pregnancy.<br />

P-686 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

REVISITING THE PREDICTIVE VALUE OF UTERINE ARTERY<br />

PULSATILITY INDEX FOR UTERINE RECEPTIVITY. H. Zaki, a<br />

E. Geneidi, b C. Coulam. c a Director Ganin Fertility Center, Cairo, Egypt;<br />

b Radiology, Faculty of Medicine, Cairo, Egypt; c Fertility & Cryogenics<br />

Lab, Downers Grove, IL.<br />

OBJECTIVE: Key factors for the process of implantation include endometrial<br />

thickness, embryo competence, hormonal milieu and uterine blood<br />

flow.There have been conflicting reports assessing the usefulness of doppler<br />

ultrasound for predicting pregnancy outcome, part of the problem is the interaction<br />

of all key factors. Now that substantial enhancement in embryo selection<br />

with comprehensive chromosome screening has given the chance of<br />

transferring single euploid embryo. Again, the substantial improvement in<br />

embryo freezing with vitrification methods, allows eliminating the negative<br />

effect of supernatural levels oestrogen achieved during ovarian stimulation<br />

upon implantation.There is a renewed focus upon assessing other key factors<br />

such as uterine blood flow. The objective of this study is to assess the role of<br />

pulsatility index (PI) in the process of implantation.<br />

DESIGN: Prospective Study.<br />

MATERIALS AND METHODS: We evaluated 83 patients under the age<br />

of 40 years of age, undergoing thawed single euploid embryo transfer. All<br />

embryos are the result of long protocol of controlled ovarian stimulation,<br />

when embryos reach blastocyst at day 5 or day 6, embryo biopsy were<br />

done and the embryos were vitrified. Comprehensive Chromosome<br />

Screening (CCS) were done using q-PCR platform.Frozen cycles were performed<br />

using oral hormonal replacement therapy with oral <strong>17</strong> beta estradiol<br />

valerate. Progesterone injections of 50 mg were taken when the endometrium<br />

thickness exceeds 7.5 mm and had a trilaminar pattern. Ultrasound doppler<br />

measurements of Pulsatility Index(PI) were done in both uterine arteries during<br />

the last day before starting the progesterone injections. Pregnancy was<br />

diagnosed by positive pregnancy test two weeks after embryo transfer and<br />

fetal heart detection 3 weeks after pregnancy test. T-test was used and statistical<br />

significance was established on p 0.3) were recruited<br />

for the study during actual ICSI cycles. Intervention(s): Patients<br />

self-administered sildenafil citrate (Viagra), vaginal suppositories (100 mg/<br />

once daily) for 7days, starting from the day of ovulation triggering till the embryo<br />

transfer day (ET D5). Endometrial thickness and spiral artery RI & PI<br />

were assessed using trans-vaginal color-pulsed Doppler 2D & 3D ultrasound<br />

on the sub-endometrial zone in two sessions, (day of triggering and day of<br />

embryo transfer). Main Outcome Measure(s): Endometrial thickness and<br />

RI & PI plus reconstructed 3D Doppler images were the primary outcomes<br />

while the implantation rate and pregnancy rate were the secondary outcomes.<br />

RESULTS: Improvement of endometrial thickness and uterine blood flow<br />

with significant reduction of RI & PI were detected in 15 (68%) out of 22 patients.<br />

Implantation rate and pregnancy rate were higher in patients with positive<br />

effect in comparison to the other patients with no effect; (26% vs 7%) &<br />

(40% vs 14%) respectively.<br />

CONCLUSIONS: Vaginal sildenafil might be an interesting therapeutic<br />

option during ICSI cycles as it may improves the uterine perfusion especially<br />

with women with a thin endometrium. Further studies are needed on sildenafil<br />

to determine whether there are other mechanisms for improvement of implantation<br />

rather than its direct effect on the endometrium.<br />

Reference: Thin endometrium.<br />

e342 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


P-688 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

EFFECT OF DURATION OF PROGESTERONE SUPPLEMENTA-<br />

TION BEFORE THE TRANSFER OF VITRIFIED-WARMED<br />

BLASTOCYSTS ON IMPLANTATION RATES IN FROZEN<br />

EMBRYO TRANSFER (FET) ESTROGEN REPLACEMENT<br />

CYCLES. S. H. Anderson, a,b D. Brasile, a,b E. S. Verrecchio, a<br />

H. Pearlstein, a W. P. Haberstroh, a M. J. Glassner. a,b a Main Line Fertility<br />

Center, Bryn Mawr, PA; b Obstetrics and Gynecology, Drexel University College<br />

of Medicine, Philadelphia, PA.<br />

OBJECTIVE: There have been few studies evaluating the optimal timing<br />

of progesterone (P4) supplementation prior to transfer of embryos in FET cycles.<br />

The objective of this study was to determine whether starting P4 supplementation<br />

five days or six days before transfer of vitrified-warmed<br />

blastocysts results in higher implantation rates in estrogen replacement<br />

cycles.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: A total of 229 estrogen replacement FET<br />

cycles were studied. All blastocysts were vitrified-warmed using the same<br />

protocol, and only expanding good or excellent quality blastocysts were<br />

warmed and transferred. Blastocysts were transferred on the day of warming.<br />

Patients were prepared for transfer with oral, intramuscular (IM), and/or<br />

transdermal estradiol to maintain a serum estradiol of 200 pg/ml. After endometrial<br />

thickness of >8 mm and tri-laminar pattern were attained, patients<br />

started P4 in the evening, either five or six days prior to blastocyst transfer.<br />

Patients were given 50 mg of IM P4 every-other-evening, and 400 mg vaginal<br />

P4 daily until at least 8 weeks of pregnancy. The Mann-Whitney non-parametric<br />

test was used to determine if starting P4 supplementation five vs six<br />

days before FET had a statistically significant effect on the blastocyst implantation<br />

rates.<br />

RESULTS: There was no difference in maternal age between the group<br />

of patients that started P4 five days before FET (mean age¼32.3)<br />

compared to the group that started P4 six days before FET (mean<br />

age¼31.8). Based on the non-parametric statistical test, the mean implantation<br />

rate of vitrified-warmed blastocysts was higher (p-value¼0.03)<br />

when P4 was started five days before transfer (36.6%; N¼1<strong>17</strong>) compared<br />

to the implantation rate when P4 was started six days before transfer<br />

(26.3%; N¼112).<br />

CONCLUSIONS: Starting P4 supplementation five days before the transfer<br />

of vitrified-warmed blastocysts resulted in higher implantation rates<br />

compared to starting P4 six days before transfer in estrogen replacement<br />

FET cycles. Prospective, randomized studies should be performed to further<br />

pinpoint the duration of P4 supplementation before FET, as well as the route<br />

of administration, that optimizes pregnancy outcomes.<br />

P-689 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

TO EVALUATE THE PREGNANCY RATE AFTER ENDOMETRIAL<br />

SCRATCHING IN COUPLES WITH UNEXPLAINED INFERTILITY<br />

IN OVULATION INDUCTION AND IUI CYCLES. R. Mahey, T. Goel,<br />

M. Gupta, G. Kachhawa, A. Kriplani. Obstetrics & Gynaecology, All India<br />

Institute of Medical Sciences, New Delhi, New Delhi, India.<br />

OBJECTIVE: Embryo implantation is the rate-limiting step for success of<br />

IVF and IUI cycles and is affected by various hormonal and biological<br />

markers. Endometrial scratching done in same cycle or previous cycles has<br />

been shown to increase endometrial receptivity thus making capable of<br />

receiving and adhering the embryo. The aim of present study was to compare<br />

the pregnancy rates during IUI cycles with or without endometrial scratching<br />

in follicular phase in couples with unexplained infertility.<br />

DESIGN: A prospective randomized study.<br />

MATERIALS AND METHODS: Sixty four couples with unexplained<br />

infertility were recruited during March 2014 to March <strong>2015</strong>. Patients were<br />

randomized into two groups by computer generated random table. Patients<br />

in group I (32 women, 78 cycles) underwent endometrial scratching on<br />

day 8 of the same menstrual cycle after receiving 7 days of ovulation induction<br />

with clomiphene citrate 50mg (day 2 to day 6) and Inj hMG 75 IU on day<br />

6,7. Group II(32 women, 81 cycles) received ovulation induction (CC+hMG)<br />

without endometrial scratching. Transvaginal USG was done to monitor the<br />

follicular growth and endometrial thickness from day 8 till at least one follicle<br />

reached 18mm diameter.Primary outcome was measured by pregnancy<br />

rate and secondary outcome was measured by abortion rate and ongoing<br />

pregnancy rate(>12wks).<br />

RESULTS: Both groups were comparable in regard to mean age, BMI,<br />

duration and type of infertility. There was no difference in the dose of clomiphene<br />

and gonadotropin requirement in two groups. The mean endometrial<br />

thickness in group I was 5.82.3mm on day 8 and 8.73.4mm on day of<br />

hCG trigger. The respective values in group II were 6.11.8mm (day 8)<br />

and 8.52.3mm (day of hCG trigger) (p value-0.9). The pregnancy rate in<br />

group I was 18.7% (6/32) as compared to 9.3% (3/32) in group II (p value-<br />

0.43). The abortion rate was 1/6 (16.7%)in group I and 0 in group II. The<br />

ongoing pregnancy rate was 5/6(83%) in group I and 2/3 (66.7%)in group<br />

II with the one being an ectopic pregnancy in group II. There were no multiple<br />

pregnancies.<br />

CONCLUSIONS: Endometrial scratching is associated with better pregnancy<br />

rates, implantation rate and ongoing pregnancy rates however, further<br />

trials with large sample size are required to establish its role in IUI cycles.<br />

P-690 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

RECURRENT ECTOPIC PREGNANCY HISTORY IS ONE OF THE<br />

RISK FACTORS OF INTRAUTERINE IMPLANTATION FAILURE<br />

IN WOMEN UNDERGOING FROZEN-THAWED EMBRYO TRANS-<br />

FER CYCLES. T. Du, H. Chen, Q. Lyu, Y. Kuang. Reproduction Department,<br />

Shanghai Ninth People’s Hospital, Shanghai, China.<br />

OBJECTIVE: To assess the influence of ectopic pregnancy history on intrauterine<br />

implantation outcome in patients undergoing frozen-thawed embryo<br />

transfer(FET) treatments.<br />

DESIGN: Retrospective cohort study.<br />

MATERIALS AND METHODS: A total of 7,118 patients undergoing<br />

10,043 FETs in the period between September 2013 and April <strong>2015</strong> were<br />

included and classified into two groups: implanted group (n¼4424 FETs)<br />

and non-implanted group (n¼5619 FETs).Endometrial preparation and<br />

FETs were performed in ordinary practice.<br />

RESULTS: Among all subjects, 784 patients had ectopic pregnancy once<br />

before (corresponding to 1082 FETs), 284 patients had twice or more<br />

ectopic pregnancies before (corresponding to 390 FETs). The total times<br />

of ectopic pregnancies (0[range 0-3] vs. 0[range 0-3]), the age<br />

(31.294.05 vs. 32.834.95 years) and the duration of infertility (3[range<br />

0-24] vs. 3[range 0-<strong>21</strong>] years) were significant lower in implanted group<br />

compared with non-implanted group, whereas the total number of embryo<br />

transferred (2[range 1-2] vs. 2[range 1-3]) and the BMI (<strong>21</strong>.703.14 vs.<br />

<strong>21</strong>.662.96 kg/m2) was significantly higher. No differences were detected<br />

regarding the endometrial thickness on embryo transfer (ET) day, estradiol<br />

and progesterone levels on ET day when comparing patients in two groups.<br />

After adjusting confounding factors including the age, BMI, uterine disorders,<br />

duration of infertility, different stage of the embryos transferred, total<br />

number of embryos transferred and times of ectopic pregnancy, the risk of<br />

implantation failure significantly increased for age, uterine disorder history<br />

and the duration of infertility, the adjusted odds ratios were 0.938 (95%<br />

confidence interval [CI]: 0.929-0.947), 0.835 (95% CI: 0.741-0.940) and<br />

0.983 (95% CI: 0.968-0.998), respectively. Compared with those who<br />

had no ectopic pregnancy history, the risk of implantation failure significantly<br />

increased for the patients had twice or more ectopic pregnancies previously,<br />

the adjusted odds ratios was 0.783 (95% CI: 0.637-0.962). The<br />

patients transferred blastula embryo had a significantly increased odds<br />

for implantation compared with patients transferred cleavage stage embryo,<br />

so were the patients with thicker endometrium on ET day and larger total<br />

number of embryo transferred, the adjusted odds ratios were 2.053 (95%<br />

CI: 1.772-2.380), 1.063 (95% CI: 1.045-1.081) and 1.854 (95% CI:<br />

1.629-2.111), respectively.<br />

CONCLUSIONS: Recurrent ectopic pregnancy is one of the risk factors of<br />

intrauterine implantation failure. The declined intrauterine implantation rate<br />

to some extent indicates an altered endometrium, which may play an important<br />

role in the pathogenesis of ectopic pregnancy.<br />

P-691 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

CAN OOCYTE RECIPIENT OUTCOME BE PREDICTED BY<br />

OOCYTE DONOR AGE? E. Cater, a L. Jenner, a A. Campbell, b<br />

S. Fishel. c a Embryology, CARE Fertility, Nottingham, United Kingdom;<br />

b Director of Embryology, CARE Fertility, Nottingham, United Kingdom;<br />

c C.E.O, CARE Fertility, Nottingham, United Kingdom.<br />

OBJECTIVE: It is widely accepted that maternal age is one of the major<br />

determining factors in ART success. It is therefore a consideration that in<br />

oocyte recipient cycles, the age of the donors may also be an influential factor.<br />

The literature is however not conclusive; some articles describing<br />

FERTILITY & STERILITY Ò<br />

e343


predictive power in oocyte donor age for recipient outcome, some showing<br />

no difference and others describing a reduction in outcome from very young<br />

donors. This study aims to establish if oocyte donor age can predict the recipient<br />

outcome.<br />

DESIGN: Standard oocyte recipient cycles from four private ART clinics<br />

between January 2008 and December 2013 were retrospectively reviewed<br />

and analysed for number of oocytes retrieved, maturity, fertilisation, achievement<br />

of an embryo transfer (ET), clinical pregnancy rate/ET (CPR), implantation<br />

rate (IR) and live birth rate/ET (LBR), based on oocyte donor age.<br />

MATERIALS AND METHODS: 1229 cycles were divided into 8 groups<br />

based on oocyte donor age of 2 year intervals; 20-<strong>21</strong>(n¼18), 22-23(n¼37),<br />

24-25(n¼82), 26-27(n¼111), 28-29(n¼148), 30-31(n¼238), 32-<br />

33(n¼266), 34-35(n¼329).Significance was determined using a Z-test of<br />

two proportions; a significance value of 0.05 and two tailed hypothesis.<br />

RESULTS: No significant difference was apparent in embryological<br />

parameters assessed up to ETalong with CPR, IR and LBR within the groups;<br />

a decrease in the three latter was seen in groups 28-29 and 34-35. Combining<br />

the two upper age groups and comparing to the two lower to increase the n<br />

number also did not show significance.CPR; 41.2% (7/<strong>17</strong>), 41.9% (13/31),<br />

36% (27/75), 36.9% (38/103), 30.8% (44/143), 39.5% (88/223), 38.4%<br />

(93/242), 33.4% (100/299) respectively according to the above age groups.<br />

IR; 26.5% (9/34), 28.8% (<strong>17</strong>/59), 25.5% (38/146), 23.5% (47/200), <strong>21</strong>.8%<br />

(60/275), 28.3% (1<strong>21</strong>/427), 24.9% (118/473), 22% (128/582) respectively.<br />

LBR equalled the CPR up until the age of 25, after which miscarriages<br />

occurred. LBR; 34.9% (36/103), 26.6% (38/143), 34.5% (77/223), 33.1%<br />

(80/242), 27.8% (83/299) respectively.<br />

CONCLUSIONS: No statistical difference was seen in outcome parameters<br />

of oocyte recipient cycles based on oocyte donor age. It should be noted<br />

that the 8 groups have an uneven number of cycles included which may affect<br />

data analysis. Those with a greater number of cycles are likely to be more<br />

reflective of the true result. As one of the largest studies so far published,<br />

it is reassuring to note that the age range provided by the HFEA guidelines<br />

shows no significant variation in the incidence of live Birth.<br />

P-692 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

MID-SECRETORY EUTOPIC ENDOMETRIUM IN INTRAMURAL<br />

FIBROIDS AND SEVERE ENDOMETRIOSIS: RELEVANCE TO<br />

FERTILITY. L. Aghajanova, J. Irwin, L. C. Giudice. Obstetrics, Gynecology<br />

and Reproductive Sciences, University of California San Francisco,<br />

San Francisco, CA.<br />

OBJECTIVE: Data are conflicting on the expression of endometrial receptivity<br />

markers in women with intramural uterine fibroids and the effect of on<br />

fertility. We aimed to investigate the mid-secretory phase (MSE) endometrial<br />

transcriptome of women with intramural fibroids and compare them to controls<br />

and those with severe endometriosis.<br />

DESIGN: In silico and laboratory-based study.<br />

MATERIALS AND METHODS: Well-annotated endometrial tissue samples<br />

were obtained through the UCSF/NIH Human Endometrial Tissue Bank.<br />

MSE samples were from 8 women with no uterine/endometrial pathology, 4<br />

with intramural fibroids (no submucosal component) and 8 with severe endometriosis.<br />

Purified total RNAwas subjected to microarray analysis with Gene<br />

1.0 ST Affymetrix platform. Data were analyzed with GeneSpring and Ingenuity<br />

Pathway Analysis. Menstrual cycle phase was assigned by endometrial<br />

histology, estrogen/progesterone levels and bioinformatics methods. Microarray<br />

analysis validation was performed with Fluidigm array and real-time<br />

PCR.<br />

RESULTS: Intramural fibroid MSE samples clustered separately from the<br />

control or endometriosis MSE. Comparison of differentially regulated genes<br />

revealed dysregulation of 1496 genes (989 up and 506 down) in endometrial<br />

samples from women with intramural fibroids vs. controls, 244 genes (139 up<br />

and 105 down) in severe endometriosis vs. controls and 1936 (1232 up and<br />

704 down) genes in fibroid vs. endometriosis samples. Comparison of the<br />

gene lists above with the 238 Endometrial Receptivity Aarray (ERA) genes<br />

(Diaz-Gimeno et al, 2011) showed that only 8 and 1 genes were dysregulated<br />

beyond the ERA 3-fold threshold in samples from women with intramural<br />

fibroids or severe endometriosis respectively vs. controls. IGF-1 and nitric<br />

oxide signaling were the top regulated pathways in the fibroid group, and<br />

planar cell polarity and CRH signaling were the top regulated pathways in<br />

the endometriosis group, while comparison of the two diseases revealed<br />

involvement of hypoxia and oxidative stress pathways.<br />

CONCLUSIONS: While ERA receptivity genes are minimally affected in<br />

the setting of intramural uterine fibroids and severe endometriosis, other<br />

genes are markedly different vs. normal controls, which may contribute to<br />

the known poor pregnancy outcomes in these populations. Further studies<br />

are needed to validate or refute these observations.<br />

Reference:<br />

1. Diaz-Gimeno, Horcajadas JA, Martınez-Conejero JA, et al. A genomic<br />

diagnostic tool for human endometrial receptivity based on the transcriptomic<br />

signature. Fertil Steril 2011; 95(1):50-60.<br />

Supported by: NIH NCTRI P50HD055764 (LCG).<br />

LUTEAL PHASE SUPPORT<br />

P-693 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

IMPACT OF SHORT LUTEAL PHASE ON NATURAL<br />

FERTILITY. N. M. Crawford, K. Chantala, A. Steiner. University of<br />

North Carolina, Chapel Hill, NC.<br />

OBJECTIVE: To determine the impact of a short luteal phase on fecundability.<br />

DESIGN: Prospective, time-to-pregnancy study.<br />

MATERIALS AND METHODS: Women, 30-44 years old with no history<br />

of infertility, who were trying to conceive for less than 3 months, were<br />

enrolled and followed until pregnancy. Each day women recorded bleeding,<br />

ovulation predictor test results, intercourse, and pregnancy test results for<br />

up to 4 months while attempting to conceive. For each cycle in which a<br />

woman did not conceive, the length of the luteal phase was determined<br />

by calculating the number of days from the first positive ovulation test to<br />

the last day of the menstrual cycle. We defined a short luteal phase as<br />

%11 days in length. We subsequently examined the probability of<br />

conceiving in a cycle based on presence or absence of a short luteal phase<br />

in the preceding non-conception cycle. Discrete time models were created<br />

to calculate fecundability ratios (FR) adjusting for maternal age. A FR


Yuden index was applied for identifying the best cut-off value of the progesterone<br />

(P4) level at day 16 in recognizing the 12th ongoing pregnancy status.<br />

Using the best cut-off value of the P4 level at day 16, we categorized our subjects<br />

into ‘‘with CLR (corpus luteum rescue)’’ and ‘‘without CLR’’. Logistic<br />

regression with age and BMI adjustment was utilized to evaluate the association<br />

between the duration of LPS and 12th week ongoing pregnancy in each<br />

group.<br />

RESULTS: A total of 158 (57.9%) women were confirmed as pregnancy at<br />

day16. Among them, 56 (35.4%) women had LPS until 7-12th week, and 102<br />

(64.6%) women had LPS 20.6ng/ml) and 38 (24.1%) women without CLR (P4


Summary of the main findings.<br />

Vaginal<br />

progesterone<br />

capsules Oral dydrogesterone RR (95% CI)<br />

Participants /<br />

Studies<br />

Authors’ Interpretation<br />

Quality of the<br />

evidence<br />

Ongoing pregnancy 20% 25% (<strong>21</strong>-31%) 1.24 (1.02-1.51) 1,611 / 5 Dydrogesterone is better Moderate<br />

Clinical pregnancy 27% 34% (30-39%) 1.24 (1.08-1.41) 2,286 / 6 Dydrogesterone is better Moderate<br />

Miscarriage 23% <strong>17</strong>% (11-26%) 0.75 (0.49-1.14) 438 / 5 No evidence of difference Low<br />

Dissatisfaction 26% 3% (1-10%) 0.10 (0.02-0.39) 430 / 1 Dydrogesterone is better High<br />

Vaginal progesterone gel Oral dydrogesterone<br />

Ongoing pregnancy 27% 27% (23-31%) 0.97 (0.83-1.13) 1,735 / 2 No relevant difference High<br />

Clinical pregnancy 31% 29% (25-34%) 0.95 (0.82-1.09) 1,735 / 2 No relevant difference High<br />

Miscarriage 11% 9% (5-15%) 0.81 (0.48-1.35) 523 / 2 No evidence of difference Low<br />

Dissatisfaction 18% 5% (3-8%) 0.26 (0.16-0.42) 822 / 1 Dydrogesterone is better High<br />

OBJECTIVE: To identify, appraise and summarize the evidence from randomized<br />

controlled trials (RCTs) comparing oral dydrogesterone with progesterone<br />

for luteal-phase support (LPS) in women undergoing IVF/ICSI.<br />

DESIGN: Systematic review and meta-analysis.<br />

MATERIALS AND METHODS: Record screening, study selection, data<br />

extraction, and evaluation of the risk of bias were performed independently<br />

by two authors. The comparison between dydrogesterone and progesterone<br />

were summarized as risk ratio (RR) and the precision of the estimates was<br />

assessed by the 95% confidence interval (CI).<br />

RESULTS: Our electronic search was performed in Mar-<strong>17</strong>-<strong>2015</strong> and 7<br />

studies were included in our quantitative analysis: 5 studies compared oral<br />

dydrogesterone with vaginal progesterone capsules; 1 study compared oral<br />

dydrogesterone with vaginal progesterone gel; and 1 study has three arms:<br />

oral dydrogesterone, vaginal progesterone capsules and vaginal progesterone<br />

gel, being considered in both comparisons. We didn’t find any study<br />

comparing oral dydrogesterone with intramuscular or oral progesterone.<br />

The main results are reported in the table.<br />

CONCLUSIONS: The available evidence from RCTs suggests that oral<br />

dydrogesterone for LPS provides at least similar reproductive outcomes<br />

and less dissatisfaction than vaginal progesterone capsules or gel.<br />

Supported by: Conselho Nacional de Desenvolvimento Cientıfico e Tecnologico<br />

(CNPq); S~ao Paulo Research Foundation (FAPESP).<br />

P-698 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

THE ADDITION OF GNRH AGONIST FOR LUTEAL PHASE SUP-<br />

PORT IN OVUM DONATION CYCLES. P. Casanova, a E. Szlit Feldman,<br />

b G. J. Rey Valzacchi, b L. A. Blanco, c C. A. Carrere, b A. Torno, d<br />

M. A. Rodriguez Kubrusli, a J. I. Mannara, d V. E. Canada. e a Gynecologist<br />

Procrearte, Buenos Aires, Argentina; b Procrearte Director, Buenos Aires,<br />

Argentina; c Reproductive Medicine, Caba, Argentina; d Procrearte, Buenos<br />

Aires, Argentina; e Procrearte, Capital Federal, Argentina.<br />

OBJECTIVE: To evaluate the effect of 0,1mg triptorelin in a single injection<br />

administered to ovum donation recipients the 6th day after donor oocyte<br />

retrieval on pregnancy outcomes.<br />

DESIGN: Prospective trial.<br />

MATERIALS AND METHODS: This study involved 78 ovum donation<br />

cycles between August 2014 and February <strong>2015</strong>, in a private fertility center.<br />

Patients were divided into two groups. Group A (Study group): 43 patients<br />

received one dose of 0,1mg Triptorelin (GonapeptylÒ Ferring) on<br />

the day 6th after donor oocyte retrieval in addition to regular luteal phase<br />

support (vaginal progesterone 600-800 mg/day and oral <strong>17</strong> B estradiol 6-<br />

8 mg/day). Group B (Control group): 35 patients received only regular<br />

phase support. The study group was conformed according to patients’<br />

agreement to participate. Embryos were transfered at cleaving stage<br />

(48-72hs) with Frydman Catheter. Patients with at least one good quality<br />

embryo (class I-II) were included. Severe male factor were excluded<br />

(WHO criteria).Statistical analysis was performed using SPSS. p


supplementation over the period 01.11.2013 to 30.06.2014. Previously, the<br />

patients were enrolled as a vaginal p. group between 01.11.2012 and<br />

01.11.2013. In the combined p. group, IM progesteron 100 mg/day and<br />

vaginal progesteron 90 mg/day were administered from embryo transfer<br />

day to bhCG day (a total of 10 days). 90 mg of vaginal progesteron gel<br />

was given twice daily in the vaginal p. group. Luteal phase support was<br />

continued with vaginal progesteron up to 10 weeks of gestation for pregnant<br />

women.<br />

RESULTS: The clinical pregnancy rate per embryo transfer (ET), ongoing<br />

pregnancy rate/ETand live birth rate/ETwere significantly higher in the combined<br />

p. group than in the vaginal p. group (Table). There was no difference in<br />

the serious side effects between the groups.<br />

CONCLUSIONS: This study reveals a new perspective in the luteal phase<br />

support. Our results indicate that, short-term, intensive luteal phase support<br />

significantly improved live birth rates in intracytoplasmic sperm injection cycles.<br />

Demographic features and cycle outcomes of the groups.<br />

Combined p.<br />

gr.(IM+Vag. p)<br />

n¼386<br />

Vaginal p. gr.<br />

n¼251<br />

P value<br />

Age (y) 305 30.94.6 0.04<br />

Total gonadotrophin 2286979.9 2295.8948 NS<br />

dose (IU)<br />

No. of MII oocytes 8.45.1 9.45.9 NS<br />

Estradiol level on hCG 2549.1<strong>17</strong>68.7 2925.42063.7 0.049<br />

day (pg/ml)<br />

Estradiol level on OPU 1660.31189.8 1839.<strong>21</strong>413.9 NS<br />

day(pg/ml)<br />

Progesterone level on OPU 7.954.4 8.275.6 NS<br />

day (ng/ml)<br />

Estradiol level on ET 1546.71290.6 1900.71512 0.003<br />

day(pg/ml)<br />

Progesterone level on ET 46.728.2 50.843.4 NS<br />

day(ng/ml)<br />

Clinical pregnancy rate/ET 43.2 28.5 0.001<br />

Ongoing pregnancy rate/ET 29.1 20.2 0.001<br />

Live birth rate/ET 28.6 20.2 0.001<br />

EARLY PREGNANCY<br />

P-700 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DO ELEVATED TSH LEVELS PREDICT EARLY PREGNANCY<br />

LOSS IN ART PATIENTS? J. Rodriguez-Purata, a J. Gingold, b<br />

M. C. Whitehouse, a J. A. Lee, a A. B. Copperman. c a Reproductive Medicine<br />

Associates of New York, New York, NY; b Icahn School of Medicine at Mount<br />

Sinai, New York, NY; c RMANY-Mount Sinai, New York, NY.<br />

OBJECTIVE: Infertility related to ovulatory dysfunction is commonly<br />

associated with thyroid disorders such as hypothyroidism. A number of<br />

studies have observed increased miscarriage rates in patients with hypoand<br />

hyperthyroidism, but the basis of this association remains unclear. While<br />

the manufacturer-listed normal range TSH is


P-702 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

WOMEN’S EXPERIENCE POST MISOPROSTOL TREATMENT<br />

FOR FIRST TRIMESTER MISCARRIAGES. B. Zhou. OB GYN,<br />

Houston Methodist Hospital, Houston, TX.<br />

OBJECTIVE: To report the outcomes and patient perspectives of the medical<br />

management of nonviable pregnancies before 12 weeks gestation.<br />

DESIGN: Retrospective cohort analysis.<br />

MATERIALS AND METHODS: The study included women with first<br />

trimester miscarriages who underwent treatment with vaginal misoprostol<br />

from 1/2011 through 12/2014 at Houston IVF Clinic. Gestational age were<br />

calculated by certain menstrual dates or date of embryo transfer and<br />

compared with ultrasound criteria. Pregnancies were considered nonviable,<br />

if no cardiac activity was detect by a CRL greater than 7mm, no fetal pole<br />

detected when mean sac diameter was greater than 25mm, or if the pregnancy<br />

displayed abnormal growth. Non-viability was confirmed with serial ultrasound<br />

scans. Once non-viability was confirmed, patients were given misoprostol<br />

800mcg vaginally and followed up 24 hours post placement. The<br />

patients then filled out a post treatment questionnaire, which assessed onset<br />

of pain, bleeding, passage of tissues, and their overall experience. The primary<br />

outcome measured was whether or not patients would recommend misoprostol<br />

as a treatment to peers in the future. Means and standard deviations<br />

were calculated.<br />

RESULTS: 83 underwent medical treatment with misoprostol for their first<br />

trimester miscarriage. Five patients were excluded because of incomplete<br />

completion of the questionnaire. Complete miscarriage was induced without<br />

the need for surgery in 75 women (96%). Surgical dilation and curettage<br />

(D&C) was performed in 3 patients (4%). Women reported average onset<br />

of pain was 2.9 hours after miso placement, average pain rating was 7/10<br />

and average duration of pain was 7.4 hours. Onset of bleeding after miso<br />

placement was 4.7 hours and 54% (42/78) perceived the bleeding as heavy,<br />

and 91% (71/78) reported noticing passed products of conception. Among<br />

those who returned patient questionnaires, 91% participants (71/78) indicated<br />

they would recommend medical management to peers, while 3% (3/<br />

78) indicated they would undergo surgery next time, and 5% (4/78) were undecided.<br />

40% (31/78) of participants have had a D&C prior to this study. The<br />

major reason participants recommended misoprostol over D&C was that misoprostol<br />

was more simple, fast, and not as painful as expected.<br />

CONCLUSIONS: The medical management of first trimester miscarriage<br />

on the outpatient basis with misoprostol is effective and highly recommended<br />

by patients who experienced it.<br />

P-703 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

PRE-PREGNANCY LOW TO MODERATE ALCOHOL INTAKE<br />

AND RISK OF SPONTANEOUS ABORTION. A. J. Gaskins, a<br />

J. W. Rich-Edwards, b P. Williams, c T. Toth, d S. A. Missmer, e<br />

J. E. Chavarro. f a Department of Nutrition, Harvard T.H. Chan School of Public<br />

Health, Boston, MA; b Connors Center for Women’s Health and Gender<br />

Biology, Brigham and Women’s Hospital and Harvard Medical School, Boston,<br />

MA; c Harvard T. H. Chan School of Public Health, Boston, MA; d Massachusetts<br />

General Hospital; e Brigham and Women’s Hospital and Harvard<br />

Medical School, Boston, MA; f Harvard School of Public Health, Boston,<br />

MA.<br />

OBJECTIVE: To examine the relationship between pre-pregnancy alcohol<br />

intake and risk of spontaneous abortion (SAB).<br />

DESIGN: Prospective cohort study.<br />

MATERIALS AND METHODS: Our prospective cohort study included<br />

27,580 pregnancies reported by <strong>17</strong>,929 women with no history of SAB in<br />

the Nurses’ Health Study II between 1990 and 2009. Pre-pregnancy alcohol<br />

intake over the past year was assessed in 1989, 1991, and every 4 years thereafter<br />

using a validated questionnaire. Pregnancies were self-reported with<br />

case pregnancies lost spontaneously at less than 20 weeks of gestation and<br />

comparison pregnancies ending in ectopic pregnancy, induced abortion, or<br />

live birth. Multivariable log-binomial regression models with generalized<br />

estimating equations were used to estimate the relative risks (RRs) and<br />

95% confidence intervals (CIs) adjusting for age, year, BMI, smoking status,<br />

physical activity, history of infertility, marital status, race, caffeine intake,<br />

and multivitamin use.<br />

RESULTS: Incident spontaneous abortion was reported in 4,326 (15.7%)<br />

pregnancies. Pre-pregnancy alcohol intake was not associated with pregnancy<br />

loss. Compared to women who did not consume alcohol, the multivariable<br />

RRs (95% CIs) for increasing categories of alcohol intake were 1.04<br />

(0.97, 1.12) for 0.1-1.9 g/day, 1.02 (0.94, 1.11) for 2-4.9 g/day, 1.01 (0.92,<br />

1.10) for 5-9.9 g/day, and 0.98 (0.88, 1.09) for R 10 g/day (p-trend¼0.45).<br />

Women who consumed R2 servings of beer per week prior to pregnancy had<br />

a 9% (95% CI 1, <strong>17</strong>%) lower risk of pregnancy loss compared to women<br />

consuming


P-705 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

PRECONCEPTION LOW DOSE ASPIRIN TREATMENT IM-<br />

PROVES CLINICAL PREGNANCY AND LIVE BIRTH IN WOMEN<br />

WITH HIGHER SYSTEMIC INFLAMMATION. L. Sjaarda, a<br />

E. Mitchell, a S. L. Mumford, a R. Radin, a N. J. Perkins, a N. Galai, b<br />

R. M. Silver, c E. Schisterman. a a NICHD, NIH, Rockville, MD; b Haifa University,<br />

Haifa, Israel; c Unversiy of Utah, Salt Lake City, UT.<br />

OBJECTIVE: While some studies of low dose aspirin (LDA) to<br />

improve reproductive outcomes have found a beneficial effect, others<br />

have found none. Recently, we reported that preconception LDA<br />

(81mg) increased fecundability and live birth rates relative to placebo<br />

in a subset of women with a history of one recent (prior 12 months) pregnancy<br />

loss in the EAGeR trial. Here we aimed to evaluate the effect of<br />

LDA, a known anti-inflammatory agent, according to inflammatory status<br />

to explore possible mechanisms and best identify women who may benefit<br />

from treatment.<br />

DESIGN: Multicenter, block-randomized, double-blind, placebocontrolled<br />

trial of 1228 women to evaluate the effect of preconception-initiated<br />

daily LDA on clinical pregnancy and live birth in women with a history<br />

of pregnancy loss.<br />

MATERIALS AND METHODS: We assessed high-sensitivity C-reactive<br />

protein (CRP), a marker of systemic inflammatory status, at baseline. Women<br />

were stratified into tertiles of baseline CRP concentration: lower<br />

(


were included in this study. Exclusion criteria included donor oocyte cycles,<br />

frozen-thawed embryo cycles, pregnancies of unknown location, and treatment<br />

with salpingectomy. LFTs, specifically alanine aminotransferase (ALT), aspartate<br />

aminotransferase (AST), albumin and total bilirubin levels were measured<br />

on day of MTX administration (day 1) and 7 days later (day 7). The change in<br />

LFTs (D) between day 1 and day 7 was calculated for both single and double<br />

dose MTX protocols. Furthermore, the change in LFTs (D) for single dose and<br />

double dose MTX protocols was compared. Continuous variables were checked<br />

for normality and expressed as mean standard deviation. Paired student’s t-tests<br />

were utilized. Statistical significance was set at P


Table 1. Notch Activity and Expression of Notch Proteins and Ligands in the Decidua and Placenta<br />

E8.5 Notch Activity Notch1 Notch2 Notch4 Dll4 Jag1 Jag2<br />

Maternal decidua ECs<br />

[capillaries &<br />

ECs Decidual cells ECs,<br />

Decidual cells<br />

ECs ECs<br />

[spiral arterioles]<br />

spiral arterioles]<br />

Ectoplacental cone TBs, TGCs TBs, TGCs TBs, TGCs TBs, TGCs<br />

E12.5 Notch Activity Notch1 Notch2 Notch4 Dll4 Jag1 Jag2<br />

Maternal decidua ECs<br />

ECs ECs ECs<br />

[capillaries]<br />

Junctional zone Spongio-TBs Spongio-TBs<br />

Labyrinth<br />

Fetal ECs,<br />

TGCs<br />

Fetal ECs TGCs, syncytio-TBs TGCs, syncytio-TBs Fetal ECs TGCs<br />

capillaries and Dll4 activates Notch1 in fetal placental vessels. In the developing<br />

placenta at E8.5, Dll4 may activate Notch2 and Notch4 in TBs, while<br />

Jag2 may activate Notch2 and Notch4 in TGCs in the mature placenta. Taken<br />

together these results suggest Notch signaling dynamically regulates<br />

decidual angiogenesis and placentation.<br />

Supported by: Robert Wood Johnson Foundation.<br />

P-711 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

PREGNANCY-ASSOCIATED PLASMA PROTEIN A (PAPP-A): A<br />

BIOMARKER FOR THE AID IN RISK STRATIFICATION OF<br />

NONVIABLE PREGNANCY. R. J. Batson, a B. B. Mills, b Z. Nagy, c<br />

W. Roudebush. d a Biomedical Sciences, University of South Carolina School<br />

of Medicine Greenville, Greenville, SC; b Obstetrics & Gynecology, Greenville<br />

Health System, Greenville, SC; c Reproductive Biology Associates, Atlanta,<br />

GA; d Biomedical Sciences, University of South Carolina School of<br />

Medicine, G, Greenville, SC.<br />

OBJECTIVE: Pregnancy-associated plasma protein A (PAPP-A) is a<br />

placenta-derived glycoprotein produced by trophoblastic cells, which rises<br />

gradually during the early weeks of a viable pregnancy. Its biological function<br />

is unknown but measuring serum levels has proven useful in a number<br />

of abnormal obstetrical conditions. Currently, serum quantitative bhCG<br />

levels and ultrasound are utilized for following early pregnancies and, ultimately,<br />

separating viable from nonviable pregnancies (miscarriages and ectopics);<br />

however, the diagnosis is often delayed due to limitations of these<br />

tests. The objective of this study is to compare PAPP-A levels in human<br />

sera obtained from women with viable intrauterine pregnancies to those<br />

with nonviable pregnancies following assisted reproductive technology.<br />

DESIGN: Retrospective cohort study in patients undergoing in vitro fertilization<br />

and embryo transfer (IVF-ET) at a tertiary fertility clinic.<br />

MATERIALS AND METHODS: Women with positive bhCG measurements<br />

14 days after IVF transfer werecategorized into two groups: (a)<br />

nonviable pregnancy, including biochemical pregnancy (bhCG<br />

5-50 mIU/mL) and pregnancies with bhCG >50 mIU/mL which did not double<br />

within 72 hours, and (b) viable pregnancy (bhCG >50 mIU/mL which at<br />

least doubled within 72 hours). Serum PAPP-A levels were measured via an<br />

ultrasensitive enzyme-linked immunosorbent assay (Beckman Coulter,<br />

Chaska, MN). PAPP-A levels between 4 weeks-4 days and 6 weeks-3 days<br />

were compared for viable and nonviable pregnancies.<br />

RESULTS: Twenty-five pregnant women undergoing IVF-ET were followed<br />

for this study, 7 resulted with a biochemical pregnancy, 5 with an<br />

ectopic pregnancy, and 13 with a viable intrauterine pregnancy. Three or<br />

four serum samples from each patient (n¼97 samples) were assayed for<br />

PAPP-A. The levels of PAPP-A ranged from 0.60 ng/mL to 3.3 ng/mL<br />

with analytical and functional sensitivities of 0.2 ng/mL and 1.25 ng/mL with a sensitivity of 69.2 and a specificity<br />

of 100.0 at a gestational age of 5 weeks-4days.<br />

CONCLUSIONS: From gestational week 5 onwards in normal pregnancies,<br />

PAPP-A was significantly increased in normal viable pregnancies<br />

over nonviable pregnancies. In nonviable pregnancy, PAPP-A appeared<br />

to remain low for the entire observation period. Pregnancy-associated<br />

plasma protein A (PAPP-A) appears to be a biomarker for nonviable pregnancy.<br />

P-712 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

THE RELATIONSHIP BETWEEN INITIAL HCG LEVELS AND<br />

DEVELOPMENT OF PREECLAMPSIA IN PREGNANCIES<br />

FOLLOWING IN VITRO FERTILIZATION. M. Erdem, a<br />

A. Najafaliyeva, a I. Guler, a A. Erdem, a N. Bozkurt, a M. Oktem, a<br />

M. F. Mutlu. b a Obstetrics and Gynecology, Gazi University Faculty of Medicine,<br />

Ankara, Turkey; b Koru Hospital, Ankara, Turkey.<br />

OBJECTIVE: To evaluate whether initial levels of hCG in pregnant patients<br />

after an IVF-ICSI cycle is related with development of preeclampsia.<br />

DESIGN: Retrospective case-control study.<br />

MATERIALS AND METHODS: IVF and obstetric records of 4<strong>17</strong> patients<br />

who had a live birth after IVF from 2005 to 2012 were evaluated. Serum betahCG<br />

levels on day 12 and 14 after embryo transfer were compared between<br />

patients who were treated with ICSI and fresh or frozen/thawed embryo<br />

transfer cycle and had a singleton live birth that was complicated with preeclampsia<br />

or not.<br />

RESULTS: A total number 168 patients were included into the analysis. In<br />

32 patients, hCG levels were measured other than day 12 or 14 and were<br />

excluded from the study. In 136 patients day 12 and day14 hCG levels<br />

were measured and included to the study. Twenty-five (18%) patients had<br />

preeclamsia. Demographic data of patients were comparable between each<br />

group. Day 12 and day 14 beta hCG values were significantly lower in preeclampsia<br />

group (137, <strong>17</strong> 72,0 IU/l and 254.1 158,7 IU/l, respectively) as<br />

compared to control group (205,1 99,5 and 457.5 267,3 IU/l, respectively)<br />

(p¼0.001 and p¼0.003, respectively).<br />

CONCLUSIONS: Initial plasma levels of hCG were inversely correlated<br />

with development of preeclampsia in pregnancies following embryo transfer.<br />

Our results support the theory of impaired invasion of trophoblasts and<br />

abnormal placentation in the main pathophysiology of preeclampsia.<br />

P-713 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

ASSOCIATION BETWEEN LEPTIN AND PREGNANCY<br />

LOSS. T. Plowden, a S. Zarek, a E. Schisterman, a L. Sjaarda, a<br />

R. M. Silver, b N. Galai, c A. DeCherney, a S. L. Mumford. a a NICHD,<br />

NIH, Bethesda, MD; b Unversiy of Utah, Salt Lake City, UT; c Haifa University,<br />

Haifa, Israel.<br />

OBJECTIVE: Previous research in both mouse and human models suggests<br />

that leptin plays an important role in reproductive health, although<br />

the exact mechanism of action is unclear. Our objective was to explore if<br />

maternal leptin levels are associated with early pregnancy loss.<br />

DESIGN: Prospective cohort study of a multi-center, block-randomized<br />

placebo-controlled trial of pre-conception low dose aspirin to assess its<br />

impact on reproductive outcomes in healthy, fertile women who have had<br />

1 or 2 prior pregnancy losses. Leptin was measured in serum drawn at baseline.<br />

Women who became pregnant during the study were followed prospectively<br />

throughout their pregnancy.<br />

MATERIALS AND METHODS: A total of 1194 women had measured<br />

leptin levels and were included in this analysis. The women were classified<br />

into low (0.007-11.3ng/ml), middle (26.3-97.4 ng/ml) and high<br />

(11.4-26.2 ng/ml) tertiles, with the middle tertile serving as the reference<br />

group. Pregnancy status was established using daily first-morning urine<br />

collection and spot urine clinic pregnancy tests at monthly visits. Chemical<br />

pregnancy loss was defined as positive hCG without clinical evidence of<br />

FERTILITY & STERILITY Ò<br />

e351


pregnancy (n ¼56) and clinical pregnancy loss as a loss following ultrasound<br />

confirmed pregnancy (n¼133). RR and 95% CIs for pregnancy loss were estimated<br />

using generalized linear models adjusted for age and the probability of<br />

confirmed pregnancy using stabilized inverse-probability-weights.<br />

RESULTS: Among women who became pregnant during this study<br />

(n¼771), 279 were in the low tertile, 275 were in the middle tertile and 2<strong>17</strong><br />

were in the high tertile. After adjusting for age, those in the low and high leptin<br />

tertiles did not have a higher risk of chemical pregnancy loss compared to the<br />

reference middle tertile group (low tertile: RR 0.89, 95% CI 0.48, 1.65; high<br />

tertile: RR 1.13 95% CI 0.61, 2.09). Additionally, there was no difference in<br />

clinical pregnancy loss among the groups (low tertile: RR 0.88, 95% CI<br />

0.61, 1.28; high tertile: RR 1.02, 95% CI 0.69, 1.46) compared to the middle<br />

tertile. Similar results were noted after adjusting for body mass index.<br />

CONCLUSIONS: Preconception leptin levels were not associated with an<br />

increased risk of chemical or clinical pregnancy loss among women with a<br />

history of prior pregnancy losses. Our data suggests that leptin levels are<br />

not associated with early pregnancy loss.<br />

Supported by: Intramural Research Program, DIPHR, PRAE, NICHD,<br />

NIH.<br />

P-714 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

C-REACTIVE PROTEIN AND PREGNANCY LOSS: RESULTS<br />

FROM THE EFFECTS OF ASPIRIN IN GESTATION AND REPRO-<br />

DUCTION (EAGER) TRIAL. S. L. Mumford, a L. Sjaarda, b<br />

R. Silver, c R. Radin, d E. Mitchell, b E. Schisterman. e a NICHD, NIH, Rockville,<br />

MD; b NICHD, Rockville, MD; c University of Utah, Salt Lake City, UT;<br />

d NIH, Rockville, MD; e Eunice Kennedy Shriver National Institute of Child,<br />

Rockville, MD.<br />

OBJECTIVE: A relationship between systemic inflammation and pregnancy<br />

loss has been reported. Therefore, we determined whether baseline inflammatory<br />

status, as indicated by C-reactive protein (CRP), was associated<br />

with pregnancy loss.<br />

DESIGN: Secondary analysis of the EAGeR Trial, a multicenter, blockrandomized,<br />

double-blind, placebo-controlled clinical trial to evaluate the effect<br />

of preconception-initiated daily low dose aspirin on reproductive outcomes<br />

in women with a history of pregnancy loss.<br />

MATERIALS AND METHODS: Participants were attempting pregnancy,<br />

aged 18-40 years, with one to two prior pregnancy losses and no history<br />

of infertility or other gynecologic disorders. We assessed highsensitivity<br />

CRP, a marker of systemic inflammatory status, at baseline prior<br />

to randomization to LDA or placebo. Women were allocated to tertiles<br />

based on their CRP concentration at baseline (12 missing CRP): low<br />

(n¼279,


embryo transfer number or eSETwas directly correlated with choosing eSET,<br />

with higher numbers of channels being associated with higher rates of eSET<br />

(p


from off-site work, as shown in the latest 2014 survey. In addition, most survey<br />

responses showed satisfaction with their current jobs and optimism on<br />

the future reproductive laboratory job market.<br />

CONCLUSIONS: Reproductive lab professionals have experienced an<br />

overall steady increase in salaries relative to national labor wage and most<br />

other clinical laboratory wage benchmarks. They also maintain an optimistic<br />

outlook on their future job market. Potential factors and impacts on these<br />

trends warrant further investigation.<br />

P-720 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

FERTILITY-RELATED SMARTPHONE APPLICATION USE<br />

AMONG PATIENTS SEEKING TREATMENT FOR<br />

INFERTILITY. M. Lanham M. A. Christensen. University of Michigan,<br />

Ann Arbor, MI.<br />

OBJECTIVE: To determine how women presenting with infertility to a subspecialty<br />

clinic are using fertility-related smartphone applications ("apps").<br />

DESIGN: Cross-sectional observational study of women presenting as<br />

new patients to an reproductive medicine clinic.<br />

MATERIALS AND METHODS: Paper surveys were offered to 310<br />

consecutive, English-speaking women who presented as new patients to an<br />

outpatient fertility clinic. IRB approval was obtained prior to initiating the<br />

study. The response rate was 89.7% (278/310).<br />

RESULTS: Of those patients completing the survey, the most common<br />

age range was 30 to 34 years (35.0%), with 88.0% of participants’ ages falling<br />

between 25 and 39 years. The majority of patients (81.5%) had an<br />

annual household income greater than fifty-thousand dollars. Thirty-two<br />

percent of women had been attempting pregnancy between 6 and 12<br />

months, while 56.5% had been attempting pregnancy for over 12 months.<br />

Eighty-eight percent of participants were actively trying to become pregnant,<br />

68.0% had regular periods, and 97.4% owned a smartphone. The<br />

vast majority of patients owned either an iPhone (63.5%) or an Android device<br />

(35.0%). The participants who were trying to become pregnant were<br />

more likely to be tracking their periods (odds ratio [OR] 7.0, 95% confidence<br />

interval [CI] 3.2-15.4). Of those trying to become pregnant, those<br />

with regular periods were more likely to be tracking their periods (OR<br />

3.2, 95%CI 1.4-7.7) and those who were tracking periods were more likely<br />

to have fertility apps installed on their smartphone (OR 23.7, 95%CI<br />

5.5-102.6). Of those who were trying to become pregnant and tracking their<br />

periods, age under 35 years was associated with an increased likelihood of<br />

using an app to track periods compared to other tracking methods such as<br />

basal body temperature or ovulation predictor kits (OR 2.8, 95%CI 1.5-<br />

5.3). Out of a total of 37 reported fertility apps, the top three most popular<br />

by reported usage were "Fertility Friend Mobile Smart Ovulation Tracker<br />

Calendar" (used by 10.4% of participants), "My Days - Period & Ovulation"<br />

(9.0%) and "Period Tracker" (7.9%). The top three most desired functionalities<br />

for a fertility app, chosen from a list of 10 options, were an<br />

appointment reminder (92.2% were interested), a way to send cycle data<br />

to a physician (85.1%), and access to medical records (82.1%). The least<br />

desired function was delivery of information on healthy activities (47.2%).<br />

CONCLUSIONS: Smartphones are nearly ubiquitous in the population of<br />

women presenting to the fertility clinic in this study. Women who are under<br />

age 35, trying to become pregnant, and have regular cycles are more likely to<br />

use apps to track their periods than other women. There is demand for<br />

increased interaction with providers through a digital medium and for integration<br />

of consumer health apps with clinical medicine. The possibilities<br />

are numerous and might include, for example, remote clinical monitoring<br />

during cycles of stimulated ovulation. Although the specific benefits or harms<br />

of fertility apps remain to be determined the potential is large.<br />

Reference: Study data were collected and managed using REDCap electronic<br />

data capture tools hosted at the University of Michigan.<br />

Supported by: Michigan Institute for Clinical & Health Research grant<br />

support (CTSA: UL1RR024986).<br />

SEXUALITY<br />

P-7<strong>21</strong> Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

THE PROSPECT OF CHILDBEARING AND FAMILY FORMATION<br />

IN THE EYES OF CHINESE YOUNG LESBIANS: A QUALITATIVE<br />

STUDY. I. P. Lo C. Chan. The University of Hong Kong, Hong Kong,<br />

Hong Kong.<br />

OBJECTIVE: This research aims at exploring Chinese young lesbians’<br />

attitudes and intention towards the use of assisted reproductive technologies<br />

(ART) and the ways in which they construct the meanings of childbearing<br />

and family in a culture where homosexuality still remains<br />

stigmatized.<br />

DESIGN: This is a qualitative study. Twelve lesbian-identified Chinese<br />

women in Hong Kong who aged between 18 and 35 were recruited through<br />

local LGBT organizations. In-depth semi-structured interviews were conducted<br />

in April <strong>2015</strong>.<br />

MATERIALS AND METHODS: Based on grounded theory, this study<br />

investigated Chinese young lesbians’ lived experiences with regard to regulatory<br />

social practices and processes that frame their perceptions of ART,<br />

childbearing, and family formation. Participants were invited to share their<br />

desire and motivation for childbearing and the prospective use of ART, as<br />

well as their coping strategies in a heterosexist society where same-sex marriage,<br />

civil partnership, and ART are not legally allowed. Each interview<br />

lasted around 90 minutes. All interviews were audio-recorded and transcribed<br />

for coding.<br />

RESULTS: The majority of respondents reported that one of the major obstacles<br />

in achieving parenthood was the perceived negative costs of ‘‘being<br />

outed’’ in the public by conceiving children and becoming parents. Even<br />

though they would like to have their own children, they felt reluctant to<br />

disclose their sexual orientation particularly to their parents since being homosexual<br />

would be considered a shame to the family. Most of them also<br />

worried that their children born as a result of ART would face discrimination<br />

due to their non-normative family formation and structure. In this regard, it<br />

was found that they would internalize the feelings of shame and guilt about<br />

their inevitable use of unconventional means of conception. Homophobic<br />

messages in society had probably led them to believe that their sexual identity<br />

was incompatible with reproduction, which is not only practically unachievable<br />

but also socially censured in Hong Kong.<br />

CONCLUSIONS: This study highlights that the nature of involuntary<br />

childlessness among Chinese young lesbians is by no means only a medical<br />

issue but also a psychosocial and ethical issue which is tightly linked with<br />

gender and heterosexual family norms in the Hong Kong Chinese context.<br />

It signifies a major step forward to understand local lesbians’ attitudes and<br />

intentions towards childbearing as well as the role of parental and social pressure<br />

on their fertility plan. It sensitizes mental health practitioners to the<br />

emotional burden of Chinese lesbians in relation to their fertility plan. Public<br />

education is needed to understand the needs and wants of lesbians and to<br />

create social space for same-sex relationships and families to be visible<br />

and legitimized.<br />

Supported by: This research is Supported by the Small Project Funding<br />

from University Research Committee of the University of Hong Kong [Project<br />

No.: 201409<strong>17</strong>6248].<br />

MENTAL HEALTH<br />

P-722 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

AS BOTH PERSON AND A PROFESSIONAL: CULTIVATING<br />

COMPASSION SATISFACTION AMONG HEALTHCARE PRACTI-<br />

TIONERS SPECIALIZED IN ASSISTED REPRODUCTIVE TECH-<br />

NOLOGIES (ART). H. Chan, a S. Wong, b M. Tam. c a Queen Mary<br />

Hospital, Hong Kong, Hong Kong; b The University of Hong Kong, Hong<br />

Kong, Hong Kong; c Department of Social Work and Social Administratio,<br />

Hong Kong, Hong Kong.<br />

OBJECTIVE: Healthcare practitioners in ART setting work under high<br />

pressure environment with heavy caseloads and patients facing a lot uncertainty<br />

and uncontrollability. This study aims at evaluating a professional<br />

development program which was develop to cultivate compassion satisfaction<br />

(CS), which refers to the pleasure derived from doing health care<br />

work effectively, and enhancing their ability to receive gratification from<br />

caregiving.<br />

DESIGN: This is a quasi-experimental study. Twenty-five healthcare practitioners<br />

from 11 licensed ART clinics in Hong Kong, including obstetricians<br />

and gynecologists, nurses, social workers and embryologists, attended a<br />

professional training course on infertility counseling. Effectiveness of the<br />

program was evaluated before and after the implementation of the program.<br />

MATERIALS AND METHODS: The professional development program<br />

was composed of nine three-hour weekly sessions, including components<br />

such as self-reflective activities, mindfulness, compassionate meditation<br />

and personal reflection on vulnerabilities and resilience, life and death, as<br />

e354 ASRM Abstracts Vol. 104, No. 3, Supplement, September <strong>2015</strong>


well as pain and suffering. All of them were invited to complete a self-administered<br />

questionnaire before (T0) and after (T1) the course, which contained<br />

Professional Quality of Life (ProQol) measuring CS, compassion fatigue<br />

(CF), and other validated measures related to psychological wellbeing.<br />

RESULTS: Participants showed significant improvement in CS after<br />

attending the course (Pre-course: 33.94 + 4.4; Post-course: 36.35 + 3.82,<br />

t¼-3.29, p


RESULTS: Correlational analyses showed positive associations between<br />

anxiety and all of the predicted risk variables. To assess the unique<br />

effect of the minority-specific stressor, a multiple regression model was<br />

used. Results showed that, after controlling for the effects of general<br />

stressors, perceptions of heterosexist bias were associated with higher<br />

levels of anxiety.<br />

Table 1: Effect of General and Minority-Specific Stressors on Anxiety.<br />

Anxiety<br />

Step 1: General stressors .<strong>21</strong>*<br />

Length of time trying to conceive -0.06<br />

Cost of treatment 0.18<br />

Perceived lack of control 0.39*<br />

Step 2: Minority-specific stressors .13*<br />

Perceived heterosexist bias 0.33*<br />

Note: *p< .05<br />

b<br />

R 2 Change<br />

CONCLUSIONS: In the context of fertility treatment, sexual minorityspecific<br />

stressors appear to uniquely contribute to anxiety for lesbian women<br />

above and beyond the effect of general stressors. Changes in practice that<br />

reduce perceptions of heterosexist discrimination, or interventions to help<br />

women cope with such experiences, may help optimize the experiences<br />

that lesbian women have in treatment settings and reduce psychological<br />

distress.<br />

P-726 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

MOTHERHOOD AFTER AGE 50: LONG-TERM FOLLOW UP OF<br />

PHYSICAL AND MENTAL WELL-BEING OF WOMEN WHO<br />

BECAME MOTHERS THROUGH OOCYTE<br />

DONATION. E. Davenport, H. Burks, R. Paulson. Obstetrics and Gynecology,<br />

University of Southern California, Los Angeles, CA.<br />

OBJECTIVE: To compare long-term mental and physical health outcomes<br />

of women who became mothers in their 50s to outcomes of younger women<br />

who also became mothers through oocyte donation.<br />

DESIGN: A descriptive study using validated surveys.<br />

MATERIALS AND METHODS: Participants were former infertility patients<br />

at an academic IVF center who became pregnant at any age through<br />

oocyte donation, and who had a live birth of one or more children who are<br />

now adolescents age 11 to 18. Participants completed a demographic survey<br />

and the SF-36-v2 Health Survey, a multi-purpose survey to assess mental and<br />

physical well-being. The SF-36v2 scores for physical and mental well-being<br />

were calculated for each participant, and median scores were calculated for<br />

each age group:


OBJECTIVE: To evaluate parent and donor-conceived child characteristics<br />

in families headed by single mothers (SM) compared with same-sex female<br />

couples (SSC) and heterosexual couples (HC).<br />

DESIGN: Cross-sectional study of <strong>21</strong>6 parents (n ¼ 85 SM, n ¼ 72<br />

SSC, n ¼ 59 HC). Participants were recruited from fertility clinics,<br />

donor programs, and parenting groups. The <strong>21</strong>6 children (4-<strong>17</strong> years<br />

old; M age ¼ 7.56 years, SD age ¼ 3.82 years) were conceived via<br />

IVF (31.0%) or IUI (69.0%) with donor sperm (83.4%), donor egg<br />

(10.6%), or donor sperm and donor egg (6.0%). Participating families<br />

were from the U.S. (77.4%), Canada (14.3%), Europe (7.4%), and<br />

Australia (0.9%).<br />

MATERIALS AND METHODS: Parents completed an online survey assessing<br />

parental well-being (Center for Epidemiological Studies - Depression<br />

- 10 items; Radloff, 1977), parent-child relationship satisfaction<br />

(Marital Opinion Questionnaire; Huston & Vangelisti, 1991), parent-child<br />

communication climate, which facilitates positive, open communication on<br />

various topics (Revised Family Communication Pattern Instrument - Conversation;<br />

Ritchie & Fitzpatrick, 1990), and child psychosocial adjustment<br />

(Strengths and Difficulties Questionnaire; Goodman, 1997). Covariates<br />

included parent age, family income, multiple birth status, child’s sex and<br />

age. ANCOVAs were used to test differences in outcomes across family<br />

types.<br />

RESULTS: After accounting for covariates, parents in SM, SSC and HC<br />

families had similar levels of parental well-being (Msm ¼ 5.76, Mssc ¼<br />

6.20, Mhc ¼ 5.61; clinical cutoff: 10.00), parent-child relationship satisfaction<br />

(Msm ¼ 4.30, Mssc ¼ 3.80, Mhc ¼ 4.29), and child emotional (Msm ¼<br />

2.89, Mssc ¼ 4.07, Mhc ¼ 3.10; normal range: 0.00 - 5.00), behavioral (Msm<br />

¼ 4.22, Mssc ¼ 4.<strong>17</strong>, Mhc ¼ 4.95; normal range: 0.00 - 7.00), and prosocial<br />

adjustment (Msm ¼ 8.54, Mssc ¼ 8.66, Mhc ¼ 8.36; normal range: 6.00 -<br />

10.00). SSC families had a more open parent-child communication climate<br />

(Mssc ¼ 4.42) than SM families (Msm ¼ 4.28) and HC families (Mhc ¼<br />

4.23; F(2, 205) ¼ 3.46, p ¼ .033).<br />

CONCLUSIONS: Families headed by single mothers should not be<br />

perceived as differing substantially from traditional nuclear families. SM<br />

families appear to be functioning well.<br />

References:<br />

1. Goodman R. The strengths and difficulties questionnaire: a research<br />

note. J Child Psychol Psychiatry 1997;38:581-586.<br />

2. Huston TL, Vangelisti AL. Socioemotional behavior and satisfaction in<br />

marital relationships: a longitudinal study. J Pers Soc Psychol<br />

1991;61:7<strong>21</strong>-733.<br />

3. Radloff LS. The CES-D scale: a self-report depression scale for<br />

research in the general. Appl Psych Meas 1977;1:385-401.<br />

4. Ritchie LD, Fitzpatrick MA. Family communication patterns:<br />

measuring interpersonal perceptions of interpersonal relationships.<br />

Comm Res 1990;<strong>17</strong>:523-544.<br />

Supported by: University of Minnesota (UMN) Agriculture Experiment<br />

Station, UMN Grant-in-Aid, UMN College of Education & Human Development<br />

Research Development Investment Grant.<br />

P-729 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

MENTAL HEALTH DISORDERS IN INFERTILE WOMEN: PREVA-<br />

LENCE, PERCEIVED EFFECT ON FERTILITY, AND WILLING-<br />

NESS FOR TREATMENT FOR ANXIETY AND<br />

DEPRESSION. H. S. Hoff, a N. M. Crawford, b J. E. Mersereau. c a Reproductive<br />

Endocrinology and Infertility, University of North Carolina, Chapel<br />

Hill, NC; b University of North Carolina, Chapel Hill, NC; c UNC, Chapel<br />

Hill, NC.<br />

OBJECTIVE: To evaluate whether women with infertility identify with<br />

being anxious or depressed, assess their perception that mental health disorders<br />

affect fertility, and determine their willingness to undergo treatment.<br />

DESIGN: Prospective, observational study.<br />

MATERIALS AND METHODS: An electronic survey was sent to all<br />

new infertile female patients over a period of 2 years at UNC Fertility.<br />

Informed consent was obtained prior to survey initiation. The proportion<br />

of respondents affected by anxiety and depression was evaluated using<br />

the NIH PROMIS, a validated 4 question scale, and we assessed their<br />

views about mental health disorders. Summary statistics were used to<br />

evaluate the population cohort. Data was analyzed using student’s t-<br />

test and Person’s chi square for continuous and categorical variables,<br />

respectively.<br />

RESULTS: A survey response rate of 43% was achieved (414/959).<br />

Overall, the cohort was young (


(SC), Hypomania (MA), and Social Introversion (SI). Cohorts were segregated<br />

by status of cycle obligation (‘‘Complete-Oocytes Retrieved’’; ‘‘Incomplete-Dropout’’).<br />

A random-number generator was used from a larger subset<br />

to isolate the ‘‘Complete-Oocytes Retrieved’’ group.<br />

RESULTS: Sixty oocyte donors were evaluated and MMPI was deemed<br />

valid and acceptable based on program criteria. Of the MMPI taken by<br />

both who completed (n¼45) and did not complete (n¼15) donation cycles,<br />

no statistical difference between the 13 personality characteristics was<br />

observed. A difference in the PA category was detected, although it did not<br />

reach significance.<br />

CONCLUSIONS: The MMPI test does not prospectively predict donor<br />

compliance to protocols and obligations in an ovum donation program.<br />

Donors demonstrated relatively consistent personality traits. Donors who<br />

scored the highest on the ‘‘hysteria’’ and ‘‘paranoia’’ scale were less likely<br />

to complete the process. These characteristics are common in individuals<br />

who display hypervigilance in social interactions, and have a tendency to<br />

be apprehensive to interpersonal communication and resist authorities’ direction.<br />

Although the MMPI did not find a differentiation in donors’ personality<br />

traits, future studies that evaluate alternative personality or psychopathology<br />

could assist in identification of appropriate donors who will be compliant<br />

with all participation requirements.<br />

P-731 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

EFFECTIVENESS OF A SELF-HELP INTEGRATIVE BODY-MIND-<br />

SPIRIT INTERVENTION (I-BMS) IN REDUCING INFERTILE<br />

WOMEN’S ANXIETY DURING THEIR IN VITRO FERTILIZATION<br />

(IVF) TREATMENT RESULT AWAITING PERIOD. C. Chan, a<br />

S. Wong, a M. Tam. b a The University of Hong Kong, Hong Kong, Hong<br />

Kong; b Department of Social Work and Social Administratio, Hong Kong,<br />

Hong Kong.<br />

OBJECTIVE: This study aims at evaluating the effectiveness of a series of<br />

self-help techniques developed based on the Body-mind-spirit intervention<br />

(I-BMS) intervention model to fulfill the psychosocial needs of infertile<br />

women during the most distressful treatment stage, which is the result awaiting<br />

period after embryo transfer.<br />

DESIGN: This is a quasi-experimental study. 20 women who underwent<br />

IVF treatment were recruited and randomly assigned into intervention<br />

group (n¼10) and control group (n¼10). The study adopted the clinical<br />

framework of I-BMS model, which aims at maintaining a dynamic and<br />

harmonious balance between physical, psychosocial and spiritual wellbeing.<br />

MATERIALS AND METHODS: Participants in the intervention group<br />

attended a 3-hour I-BMS workshop organized by a registered social<br />

worker in educating the self-help techniques, and received a self-help exercise<br />

book containing self-help exercises, spiritual proverbs and journal<br />

writing for home practice on a 14-day basis. They were asked to practice<br />

the self-help techniques and report their subjective well-being on a daily<br />

basis during the result awaiting period after embryo transfer. Meanwhile,<br />

participants in control group received educational materials on healthy<br />

diet for home reading.All participants were invited to complete a selfadministered<br />

at three time points: on the day of recruitment (T0), on<br />

the day of embryo transfer (T1), before pregnancy test (T2). The scales<br />

used include Importance of Childbearing Index, Chinese Kansas Marital<br />

Satisfaction Scale (C-KMS) and Chinese State-Trait Anxiety Inventory<br />

(C-STAI).<br />

RESULTS: By using ANOVA, participants in the intervention group<br />

showed significant reduction in the importance in childbearing (T1: 26.90<br />

+ 4.09; T2: 24.30 + 3.98, F¼2.10, p


OBJECTIVE: While the Uterine Fibroid Symptom and Quality of Life<br />

questionnaire (UFS-QOL) is a validated and helpful tool in determining<br />

response to treatment, it does not allow for the texture and intensity of fibroid<br />

related experiences to be conveyed. The purpose of this study is to qualitatively<br />

identify and characterize challenges that impact quality of life and<br />

emotional well-being of women with symptomatic fibroids.<br />

DESIGN: Qualitative semi-structured interviews, health literacy assessments,<br />

and demographic surveys.<br />

MATERIALS AND METHODS: Women with symptomatic uterine<br />

fibroids were recruited from an urban academic medical center and community-based<br />

organizations in a large Midwest metropolitan area. Participants<br />

completed in-depth, one-on-one interviews, a health literacy assessment<br />

and a demographic survey. Interviews were transcribed verbatim and uploaded<br />

to NVivo version 10 for data management and thematic coding. Three<br />

coders identified major themes and subthemes using a grounded theory<br />

approach.<br />

RESULTS: Sixty women completed the study for a total of 35 hours of interviews<br />

yielding 1,357 transcribed pages. The k across coders was 0.94. The<br />

mean age of participants was 43.0 6.8 (mean SD). 61.7% of participants<br />

self-identified as African-American, 25.0% as Caucasian, 8.3% as Hispanic<br />

and 5.0% as Asian. 68.3% of the participants had at least a 4-year college degree<br />

and 55.0% of the women had a total annual household income of less<br />

than $75,000. Two of the major themes identified were impact of fibroids<br />

on work and impact on social life. Work impact examples include women<br />

feeling distressed due to fear of soiling their clothes and furniture. They<br />

also felt frustration and anxiety because heavy bleeding and pain symptoms<br />

frequently led to missed days of work. Women’s social lives were also<br />

impaired by having symptomatic fibroids. Women reported feeling lonely<br />

and socially isolated due to staying at home to avoid embarrassment caused<br />

by unexpected pain or bleeding and expressed guilt about missing out on major<br />

social events and disruption of relationships. Women also expressed<br />

shame about their appearance, difficulty planning for social events, and<br />

limited behavioral coping strategies as a result of having to work around<br />

the symptoms of their fibroids.<br />

CONCLUSIONS: Symptomatic fibroids greatly impair women’s work<br />

and social lives in ways that cannot be fully captured in a survey. These challenges<br />

impact the quality of life in women largely by limiting their work productivity<br />

and social engagement with loved ones. The emotional distress<br />

reported by women in this qualitative study suggests the need for the inclusion<br />

of mental health professionals in improving clinical care and wholehealth<br />

outcomes for women suffering from these prevalent tumors.<br />

Supported by: NIH WRHR Program K12HD0501<strong>21</strong>; RWJ Foundation;<br />

NMH; Evergreen Foundation (EEM).<br />

P-734 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

DONOR EGGS: ON ICE, STILL NICE? J. C. Patel, a<br />

D. L. Cunningham, a M. Y. Fung, b D. A. Sheehan, a M. K. Connerney, a<br />

K. J. Go. b a Embryology, IVF New England, Lexington, MA; b IVF New England,<br />

Lexington, MA.<br />

OBJECTIVE: The vitrification and thawing of mature human eggs are<br />

among the newest of assisted reproductive technologies (ART), and present<br />

the opportunity for creating donor egg banks. Given some of the advantages<br />

to the recipient patient of using vitrified eggs (VE), e.g., greater selection of<br />

donors, lower cost, and more convenience with no need for cycle synchronization<br />

with the donor, ART laboratories may encounter VE cycles more<br />

frequently, requiring development of expertise in using VE. To learn how<br />

VE compare with fresh eggs (FE) and identify the technical differences in using<br />

each egg type, we compared clinical outcomes for recipients of VE and<br />

FE from July 1, 2012 through December 31, 2014. VE were purchased from a<br />

single egg bank in this study.<br />

DESIGN: Retrospective data analysis of IVF cycles of recipients of FE or<br />

VE from a single VE bank. Statistical analyses were conducted by Student’s<br />

t-test and Mann-Whitney U test.<br />

MATERIALS AND METHODS: Recipient cycles of FE or VE were<br />

compared for average number of eggs per cycle, fertilization rate, #embryo<br />

transfers, average # embryos transferred, pregnancy rate,and chance for the<br />

recipient to have supernumerary embryos for cryopreservation. FE were<br />

inseminated conventionally or by ICSI; VE received ICSI exclusively. Assisted<br />

hatching was used electively on cleavage stage embryos (CSE)from FE<br />

but on all VE CSE. Embryologists were trained in the specific thawing protocol<br />

of VE obtained from a single egg bank.<br />

RESULTS: 80 FE and 168 VE cycles were analyzed, with 1595 and 1<strong>17</strong>1<br />

eggs, respectively, available for the recipients (averages of 19.9 FE vs 7 VE<br />

per recipient). Fertilization rates of 76.18.5% and 84.18.0%)of inseminated<br />

FE and VE, respectively, were achieved(significant, p¼0.0011) leading<br />

to averages of 1.2 (FE) and 1.3 (VE) embryos per transfer (no statistical difference)and<br />

49/75 FE and 91/158 VE clinical pregnancies per ET (not significant,<br />

chi-squared¼1.2699, p¼0260).72.5% of FE recipients had frozen<br />

embryos after ET compared to 51.2% of VE.<br />

CONCLUSIONS: Despite the almost 3-fold higher number of eggs available<br />

per treatment cycle for FE vs. VE, comparable fertilization rates, application<br />

of elective single embryo transfer, and clinical pregnancy rates per ET<br />

were obtained. FE patients had an advantage in more frequently having supernumerary<br />

embryos for freezing from their cycles. VE can provide more<br />

cost-effective and convenient treatment to patients but require specific<br />

training in the thawing protocol prescribed by the egg bank to realize this<br />

advantage.<br />

P-735 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

THE PSYCHOSOCIAL IMPACT OF INFERTILITY AMONG<br />

WOMEN SEEKING FERTILITY TREATMENT. W. D. Winkelman, a<br />

P. P. Katz, b J. F. Smith, c T. Rowen. a a Obstetrics, Gynecology and Reproductive<br />

Medicine, University of California, San Francisco, San Francisco, CA;<br />

b Department of Medicine, University of California, San Francisco, San Francisco,<br />

CA; c Department of Urology, University of California, San Francisco,<br />

San Francisco, CA.<br />

OBJECTIVE: To identify factors that are associated with increased psychosocial<br />

impact on women seeking infertility treatment.<br />

DESIGN: Cross sectional study of women seeking infertility treatment<br />

who presented for an initial intake exam.<br />

MATERIALS AND METHODS: Responses to written questionnaire that<br />

included items on the sexual, personal, marital, and social impact of infertility<br />

using a previously validated scale. Respondents also classified the underlying<br />

cause of infertility as female factor, male factor, concurrent male<br />

and female factor or unexplained infertility. Multivariate regression<br />

Table 1: Impact of infertility among respondents for select demographic traits, adjusted model,<br />

Characteristic<br />

Sexual impact<br />

score, mean (p)<br />

Personal impact<br />

score, mean (p)<br />

Marital impact<br />

score, mean (p)<br />

Social impact<br />

score, mean (p)<br />

Perceived Infertility Male factor only 22 (ref) 58 (ref) 27 (ref) 55 (ref)<br />

Etiology<br />

Combined 15 (0.09) 67 (0.1) 29 (0.68) 40 (0.2)<br />

Female factor only 33 (


analyses were used to identify factors independently associated with<br />

increased impact.<br />

RESULTS: A total of 809 women met the inclusion criteria, of which 396<br />

(49%) completed the questionnaire. Most participants (75.5%) reported having<br />

no prior children. The majority (58.8%) attributed infertility to only female<br />

factors, 30.4% to a combination of male and female factors, 7.3% to<br />

only male factor, and 3.5% reported that the etiology had not been determined.<br />

The actual cause of infertility from the medical records was statistically<br />

different from the perceived cause of infertility, p


MATERIALS AND METHODS: Study participants were 152 female<br />

infertility patients who were enrolled in an IVF clinical trial at a fertility<br />

clinic in the United States. To measure anxiety, depression, grief, and<br />

maternal identity centrality, all women completed self-report questionnaires<br />

at the outset of treatment in advance of the start of a first cycle of IVF. Data<br />

were collected between <strong>October</strong> 2012 and June 2014.<br />

RESULTS: Mean anxiety and depression scores for the clinical sample<br />

were significantly lower than for a community sample (p < .001). Mean grief<br />

scores for the clinical sample were significantly lower than for a control sample<br />

(p < .01), although grief symptoms were endorsed with greater frequency<br />

than either anxiety or depression. Predicted interaction effects between<br />

maternal identity centrality and grief on the presence of anxiety and depression<br />

were not Supported. While hypotheses involving maternal identity centrality<br />

were not confirmed, a statistically significant positive relationship was<br />

found between maternal identity centrality and grief.<br />

CONCLUSIONS: The findings of this investigation demonstrated that<br />

clinical symptoms are not reported in the majority of women beginning<br />

IVF treatment. Instead, more women report the experience of grief, which<br />

suggests a normative psychological process in response to loss. The significant<br />

positive relationship found between maternal identity centrality and<br />

grief implicates the role of development and identity in shaping psychological<br />

response to infertility. These findings support approaching subclinical<br />

distress in treatment-seeking women through the grief paradigm and explaining<br />

adverse psychological response to infertility as resulting, in part, from the<br />

normative desire for motherhood as a means of identity fulfillment.<br />

Reference:<br />

1. Levin, K., Samstag, L., & Freeman-Carroll, N. (2014). Psychological<br />

distress in women presenting for first-time in vitro fertilization: Relationships<br />

among maternal identity centrality, grief, and psychopathology.<br />

Available from ProQuest Dissertations and Theses databases<br />

(UMI No. pending).<br />

P-739 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

INTEGRATING DONOR CONCEPTION INTO IDENTITY:<br />

PARENT-CHILD RELATIONSHIPS AND IDENTITY DEVELOP-<br />

MENT IN DONOR-CONCEIVED ADOLESCENTS. J. Slutsky, a<br />

V. Jadva, b T. Freeman, b S. Persaud, a W. Kramer, c M. Steele, a<br />

H. Steele, a S. Golombok. b a Center for Attachment Research, The New<br />

School for Social Research, New York, NY; b Centre for Family Research,<br />

University of Cambridge, Cambridge, United Kingdom; c The Donor Sibling<br />

Registry, Nederland, CO.<br />

OBJECTIVE: This study explores the experiences of donor-conceived adolescents<br />

connecting to same-donor offspring (individuals conceived using<br />

the same donor), and whether the quality of parent-child relationships relates<br />

to how donor-conceived adolescents, informed of their genetic origins in<br />

early development, integrate donor conception into their sense of identity.<br />

DESIGN: During in-home visits, qualitative and quantitative data were<br />

collected from adolescents conceived by sperm donation on their experiences<br />

of being donor conceived, connecting to same-donor offspring, and their relationships<br />

with their parent(s) and siblings.<br />

MATERIALS AND METHODS: Twenty-three adolescents completed<br />

questionnaires and a semi-structured interview including the Friends and<br />

Family Interview (FFI) (Steele & Steele, 2005; Steele et al. 2009), designed<br />

to explore the adolescents’ relationship with their parent(s), siblings, and<br />

same-donor offspring. Participants also completed a questionnaire designed<br />

to evaluate the extent to which donor conception factors into their sense of<br />

identity. Participants were members of the Donor Sibling Registry, an American-based<br />

registry that facilitates contact between same-donor offspring,<br />

their parents, and donors.<br />

RESULTS: Participants, (16 female, 7 male) ages 12 to 19 (mean 14.5<br />

years), had all made contact with at least one same-donor offspring; none<br />

had located their donor. Just over half (13, 56 %) felt ‘neutral/indifferent’ about<br />

their donor conception, 8 (32%) felt ‘positive/interested’, and 2 felt ‘negative/<br />

avoidant’. The importance of knowing the donor varied greatly from those who<br />

felt it was ‘unimportant’ (5, 23%) to those who felt it was ‘extremely important’<br />

(5, 20%), with the remainder seeing it as ‘somewhat important’. Data<br />

will be presented on the FFI and its relationship to donor conception identity.<br />

CONCLUSIONS: This study presents the first in-depth data from donorconceived<br />

adolescents on their experience of being donor conceived and contacting<br />

same-donor offspring, their familial relationships, and their incorporation<br />

of donor conception into their identity. These results enable greater<br />

understanding of the factors that affect identity development in donorconceived<br />

individuals.<br />

References:<br />

1. Steele, H. & Steele, M. (2005). The construct of coherence as an indicator<br />

of attachment security in middle childhood: the friends and family<br />

interview In: Kerns, K.A. & Richardson, R.A. (Eds.) Attachment in<br />

middle childhood (Chapter 7). New York: Guilford Press.<br />

2. 2.Steele, H., Steele, M. & Kriss, A. (2009). The friends and family<br />

interview (FFI): coding guidelines. Unpublished manuscript.<br />

Supported by: This study was Supported by the Wellcome Trust [097857/<br />

Z/11/Z].<br />

P-740 Wednesday, <strong>October</strong> <strong>21</strong>, <strong>2015</strong><br />

PATIENT PERSPECTIVES ON IVF SUCCESS AND LIKELIHOOD<br />

OF MULTIPLE GESTATIONS. S. J. Miller, a K. D. Schoyer, b<br />

E. M. Wozniak, b J. Davis, a J. Sandlow, b E. Y. Strawn, c K. E. Flynn. b<br />

a OB/GYN, Medical College of Wisconsin Affliated Hospital, Wauwatosa,<br />

WI; b Medical College of Wisconsin, Milwaukee, WI; c Medical College of<br />

Wisconsin, Mequon, WI.<br />

OBJECTIVE: To describe couples’ perceptions of fertility treatment success<br />

rates and chance of multiples.<br />

DESIGN: Longitudinal, prospective, mixed-methods study of patients and<br />

their partners who presented for care with a reproductive endocrinologist<br />

(REI) at an academic medical center.<br />

MATERIALS AND METHODS: Both members of 37 couples (2 same-sex<br />

female) separately completed surveys and qualitative interviews before their<br />

initial consultation with an REI and up to 5 additional times over the subsequent<br />

12 months. We present means (SD) and ranges of perceived chances<br />

(%) of success and multiples with in vitro fertilization (IVF) from survey<br />

data pre-consult and at 12 months. For couples with diagnoses of diminished<br />

ovarian reserve (DOR) or anovulation, we present means (SD) of perceived<br />

chances of success specific to the treatment they were planning next, based<br />

on interview data at 2 months, after testing and diagnoses. The study was primarily<br />

qualitative in nature and not powered to make statistical comparisons.<br />

RESULTS: The Table shows the perceived chance (%) of IVF success<br />

and multiples for all respondents (patients and partners) stratified by patient<br />

age. There was little difference pre-consult and 12 months later. At 2<br />

months, couples with a diagnosis of DOR who were planning IVF with<br />

donor oocytes perceived their chance of success to be 72.9% (12.9). Those<br />

with DOR planning ovulation induction with intrauterine insemination<br />

(IUI) perceived their chance of success to be 23.0% (22.8). Couples with<br />

anovulation planning ovulation induction with IUI perceived their success<br />

to be 38.7% (26.2).<br />

CONCLUSIONS: The very wide ranges of perceived likelihood of success<br />

with IVF and other diagnosis-specific treatments confirm the need for patient/partner<br />

education. Additionally, there is significant potential for education<br />

regarding options to minimize the risk of multiple births with IVF.<br />

Table.<br />

Mean (SD)<br />

Range<br />

Patients<br />

Patients 35-40 years old<br />

Prior to consult with REI 12 months later Prior to consult with REI 12 months later<br />

Perceived likelihood of success with IVF 43.1% (23.2)<br />

1.0-90.0<br />

Perceived likelihood of multiples with IVF 30.8% (20.3)<br />

3.0-78.0<br />

41.2% (20.4)<br />

25.0-60.0<br />

32.2% (22.4)<br />

3.0-85.0<br />

50.3% (30.0)<br />

1.0-90.0<br />

32.8% (<strong>21</strong>.8)<br />

4.0-60.0<br />

36.6% (<strong>21</strong>.1)<br />

11.0-86.0<br />

29.8% (<strong>21</strong>.5)<br />

5.0-70.0<br />

FERTILITY & STERILITY Ò<br />

e361


AUTHOR INDEX<br />

Aballa, T., O-19<br />

Aballa, T. C., O-24<br />

Abdalmageed, O. S., P-48, P-568<br />

Abd El Aal, D. M., P-161<br />

Abdelaleem, A. A., O-228, P-<strong>21</strong>1<br />

Abdelaziz, M., O-6<br />

Abdelghaffar, H., P-233, P-613<br />

Abdel-Hamid, S., P-48<br />

Abdelmagied, A. M., O-69, O-259, P-137, P-<strong>21</strong>1<br />

Abdel-Rahman, M., P-613<br />

Abdel-Rahman, M. Y., P-233<br />

Abdo, G. A., P-403<br />

Abdulwahab, F. M., P-283<br />

Abedi Asl, Z., O-156<br />

Abozaid, T. I., P-107, P-695<br />

Abuelhasan, A. M., O-69, O-259, P-<strong>21</strong>1<br />

Abu-Elmagd, M., P-112, P-555<br />

Abusamaan, M. S., P-113, P-115<br />

Abu-Soud, H., P-<strong>21</strong>5, P-282<br />

Abu-Soud, H. M., O-143<br />

Abuzeid, M. I., O-70, P-81, P-107, P-609, P-626,<br />

P-695<br />

Abuzeid, O., O-70, P-81, P-107, P-626, P-695<br />

Acacio, B., O-34, P-516<br />

Acharya, C. R., O-38, O-41, O-118, P-568<br />

Acharya, K. S., O-38, O-41, O-118, P-568<br />

Achour-Frydman, N., O-159<br />

Acton, A. J., O-52<br />

Adamowicz, M., P-516<br />

Adamson, G. David., O-<strong>21</strong>6, P-659, P-668, P-716,<br />

P-7<strong>17</strong><br />

Adamu, A., P-4<br />

Adashi, E. Y., P-89<br />

Adda-Herzog, E., P-311<br />

Adedji-Fajobi, T., O-<strong>21</strong>2<br />

Adir, M., O-136<br />

Agarwal, A., O-111, P-101, P-111, P-112, P-114,<br />

P-116, P-119, P-381, P-387, P-451, P-534,<br />

P-544, P-545, P-552, P-555, P-559<br />

Agarwal, N., P-157<br />

Agbo, C., P-413<br />

Aghajanova, L., P-29, P-692<br />

Agrawal, S., P-31<br />

Aguilar, J. A., O-45, P-593, P-673, P-679<br />

Aguirre, M., P-183, P-192, P-327, P-349<br />

Ahmad, A. K., P-64<br />

Ahmad, G., P-101, P-114, P-116, P-119, P-451,<br />

P-552<br />

Ahmady, A., P-412<br />

Ahmed, A. I., P-484, P-581, P-637<br />

Ahmed, H., P-228<br />

Ahn, J.-H., P-543<br />

Ainsworth, A., P-98<br />

Ait-Ahmed, O., O-239<br />

Aitken, R. J., P-1<strong>17</strong>, P-123<br />

Aizpurua, J., O-128, O-<strong>17</strong>3, P-685<br />

Akbas, H., O-1<strong>17</strong><br />

Akin, N., O-162, P-33, P-475, P-634<br />

Akopians, A. L., P-247, P-473<br />

Akutsu, H., P-407<br />

Alaboudy, A., P-233, P-613<br />

Alaboudy, F., P-613<br />

Alama, P., P-12<br />

AlAnsari, A. A., O-111<br />

Al-Aref, I., P-<strong>17</strong>4<br />

Al-Asmar, N., P-597<br />

Alazami, A. M., P-283<br />

Albahlol, I., P-333<br />

Albert, C., P-256<br />

Albertini, D. F., O-151, P-253, P-592, P-622<br />

Albertsen, H. M., O-<strong>17</strong>5<br />

Alegretti, J. R., P-351<br />

Alexander, C., P-240, P-490<br />

Alexander, C. J., O-260<br />

Alfarawati, S., O-252<br />

Alford, C., P-566<br />

Al-Hassan, S., P-283<br />

Al-Hendy, A., O-6, O-73, O-74, O-75, O-76,<br />

O-182, O-185, O-186, O-<strong>21</strong>1, P-113, P-115<br />

Al-Hussaini, T. K., O-228<br />

Ali, M. K., P-161<br />

Alikani, M., P-323<br />

Alkabra, M., P-283<br />

Alkruaya, F. S., P-283<br />

Alkudmani, B., P-200, P-443<br />

Alkusayer, G. M., P-150<br />

Allen, R., P-14<br />

Allon, R., P-704<br />

Allshouse, A. A., O-201<br />

Al Ma’mari, N., P-302<br />

Alohali, A., O-1<br />

Alper, E., P-634<br />

Alper, M., P-585<br />

AlRumaihi, K. R., O-111<br />

Al-Safi, Z. A., O-274, P-425<br />

AlSaid, S., O-111<br />

Alshahrani, S., P-409<br />

Altman, E., O-94<br />

Alur, S., O-54<br />

Alvarez, J. P., O-120, P-473<br />

Alvarez Sedo, C., O-36, P-487, P-515<br />

Alvero, R., O-258<br />

Alviggi, C., P-630<br />

Amand, G., P-418<br />

Amano, N., P-320, P-561<br />

Amdani, S., P-570<br />

Amo, A., P-285<br />

Ampeloquio, E., O-123, P-489<br />

Amrane, S., O-60, P-348<br />

Anahory, T., O-247, P-360<br />

Anchan, R. M., O-188, P-162<br />

Anderson, A. R., O-5<br />

Anderson, K. N., P-727, P-728<br />

Anderson, M., O-<strong>17</strong>0<br />

Anderson, R. E., P-507, P-541<br />

Anderson, S., P-508<br />

Anderson, S. H., P-688<br />

Andreucci, S., P-296<br />

Andriani, L., P-4<strong>21</strong><br />

Anduaga Marchetti, I. A., P-415<br />

Anspach, E., P-255, P-479, P-624<br />

Antaki, R., O-132, P-309<br />

Anthony, J. T., P-337<br />

Antoine, Y., P-182<br />

Antoniassi, M. P., P-99, P-539<br />

Antunes, D. M. F., P-271<br />

Ao, A., P-506<br />

Aono, N., P-57, P-536, P-575<br />

Aoyagi, N., P-579<br />

Aparicio-Ruiz, B., P-600<br />

Apryshko, V., P-523, P-663, P-667<br />

Apter, D., O-100, P-371<br />

Aquino, A. P., P-261, P-351<br />

Arafa, M. M., O-111<br />

Arai, G., P-391<br />

Arav, A., O-62<br />

Arbo, E., O-129, P-305<br />

Arbona, C., O-235<br />

Archer, D. F., O-181, P-191<br />

Arheart, K. L., P-651<br />

Ariza, M., P-41<br />

Arju, R., P-422<br />

Armant, D., O-89, O-2<strong>21</strong><br />

Armenti, E. M., O-193<br />

Armstrong, A., O-39, O-268<br />

Arora, R., P-196<br />

Arredondo, F., P-8, P-345, P-419<br />

Arruguete, G., P-681<br />

Arthur, R., P-196<br />

Arvas, A., P-645<br />

Arvizu, M., O-271<br />

Asada, H., P-<strong>21</strong>, P-144<br />

Asada, Y., O-64, P-579, P-618<br />

Asemota, O., O-273<br />

Asemota, O. A., P-47<br />

Ashcraft, L., P-402<br />

Ashraf, M., O-70, P-107, P-695<br />

Aslan, K., P-167<br />

Assidi, M., P-111, P-381<br />

Assou, S., O-239<br />

Aston, K., P-46, P-541<br />

Ata, B., P-167, P-250<br />

Atabekoglu, C. S., P-639<br />

Ates, S., P-229<br />

Attar, E., O-51<br />

Attia, G., O-265, P-651<br />

Auer, H., O-47<br />

Ausin, I., P-331<br />

Austin, C. M., O-251<br />

Avci, B., P-167, P-250<br />

Avendano, C., P-414, P-415<br />

Avila, J., P-9<br />

Avril, C., P-305<br />

Awonuga, A. O., P-<strong>21</strong>5<br />

Ayaz, A., P-111, P-112, P-559<br />

Ayers, J., P-440<br />

Ayers, J. W., P-609<br />

Aygun, M., O-1<strong>17</strong><br />

Azab, M., O-237<br />

Azuma, Y., P-<strong>17</strong>0<br />

Babayev, S. N., P-186, P-<strong>21</strong>8, P-400, P-628<br />

Baccarelli, A. A., O-136<br />

Bae, J., P-149, P-187<br />

Bahceci, M., P-303<br />

Baillargeon, J.-P., O-253<br />

Baillet, S., P-254<br />

e362 Author Index Vol. 104, No. 3, Supplement, September <strong>2015</strong>


Baird, D., O-77, P-423<br />

Baird, G., O-234<br />

Baker, V., O-39, O-141, O-268<br />

Baker, V. L., P-26, P-550<br />

Bako, I. G., P-4<br />

Balaban, B., O-162, P-33, P-475, P-634<br />

Balasch, J., P-427<br />

Baldwin, M., P-569<br />

Balthazar, U., O-5<br />

Banchereau, J., O-122<br />

Bankowski, B., O-44<br />

Banks, N., P-15, P-287<br />

Bar, H., P-278<br />

Baracat, E. C., O-149<br />

Barad, D. H., O-151, P-253, P-428, P-439, P-592,<br />

P-622<br />

Barash, O., O-246<br />

Baratz, A. Y., P-196, P-557<br />

Barbato, V., P-464<br />

Barberet, J., P-360<br />

Barberry, C., O-23<br />

Barbosa, C. P., P-<strong>17</strong>9<br />

Barbosa, M. W. P., P-697<br />

Bar-Chama, N., O-112, O-164, O-204<br />

Bardos, J., O-1<strong>21</strong>, O-245, P-131<br />

Bareh, G., P-432, P-483<br />

Barfield, W., P-109<br />

Barnett, R., O-244<br />

Barnhart, K. T., O-88, P-26<br />

Baroni, R. H., O-26<br />

Barragan, M., O-47<br />

Barratt, C. L.R., P-570<br />

Barrera, A. D., P-483<br />

Barrett, C., P-569<br />

Barrett, E. S., P-299, P-377<br />

Barretto, T., O-236<br />

Barriere, P., O-129, P-305<br />

Barrionuevo, M., O-193<br />

Barritt, J., P-240, P-247, P-473, P-490, P-500<br />

Barros, B., P-261, P-351<br />

Bartels, C. B., P-624<br />

Bartmann, A. K., O-159<br />

Bartolomei, M. S., O-<strong>21</strong>2, O-226, P-556<br />

Bartolucci, A., P-134, P-354, P-479<br />

Baruffi, R., O-131, O-270<br />

Bashook, P., P-104<br />

Basile, N., P-616<br />

Bastu, E., O-51, O-127, P-205<br />

Batcheller, A. E., P-94, P-438<br />

Bates, G., P-220, P-279<br />

Batmaz, G., P-229<br />

Batson, R. J., P-711<br />

Battello, N. V., P-414<br />

Batzofin, D., O-84<br />

Bauer, C., O-95<br />

Bauer, J. D., P-3<strong>17</strong>, P-403<br />

Baumgarten, S. C., O-14<br />

Bayer, A. H., O-12<br />

Bayles, A., P-294<br />

Baysal, B., O-127<br />

Beall, S., P-3<strong>21</strong><br />

Beauchamp, C., P-252<br />

Bedairy, M. H., P-333, P-347<br />

Bedaiwy, M. A., P-150, P-155, P-701<br />

Bedard, J., P-511<br />

Bedoschi, G., P-458, P-462, P-471<br />

Begum, S., P-163<br />

Behr, B., P-134, P-288, P-354, P-550<br />

Bekheirnia, M., O-208<br />

Belan, M., O-253<br />

Belisle, S., O-85<br />

Bell, E., O-154<br />

Bellerose, H., O-166<br />

Bello, S., O-154<br />

Belloc, S., O-247<br />

Bellver, J., O-148<br />

Belo, A. S., P-261<br />

Beltsos, A. N., P-288, P-508<br />

Benadiva, C., P-479<br />

Benadiva, C. A., P-255, P-288, P-508, P-585,<br />

P-624<br />

Benages, C. A., P-7<br />

Benard, J., P-460, P-476<br />

Bender, J., P-707<br />

Bendigeri, T., P-163<br />

Bendikson, K., P-108<br />

Bendikson, K. A., P-88<br />

Bennett-Toomey, J. N., O-101<br />

Benoit, J., O-132, P-309<br />

Berga, S. L., P-19<br />

Berger, S. L., P-556<br />

Bergeron, M., P-158<br />

Bergh, C. M., O-83<br />

Bergh, P. A., O-83<br />

Berglund, A., P-224<br />

Berkeley, A. S., O-198<br />

Berker, B., P-639<br />

Berkkanoglu, M., P-435, P-653<br />

Berkowitz, K., P-280<br />

Berman, J. M., P-113, P-115<br />

Bernardi, L. A., O-77, O-<strong>17</strong>9, P-423<br />

Bernstein, L. R., O-161, P-270<br />

Berro, R., P-419<br />

Berry, T., O-76<br />

Bertolla, R., O-24, P-99, P-539<br />

Best, M. W., P-632<br />

Bhagavath, B., P-139<br />

Bhatt, D., O-275<br />

Bhattacharya, S. M., O-11, P-72<br />

Bhusane, K., P-163<br />

Bialobrzeska, D., P-36, P-40<br />

Bianchi, E., P-199<br />

Bianco, B., P-<strong>17</strong>9<br />

Bieniek, J. M., O-<strong>17</strong>1, O-199, P-133<br />

Biesiada, J., P-94<br />

Bildik, G., O-162, P-33, P-475, P-634<br />

Biley, Y., P-82<br />

Biryukov, A., P-667<br />

Bishop, C. V., O-130<br />

Bishop, L. A., P-4<strong>21</strong><br />

Bisignano, A., O-196, P-14<br />

Bissonnette, F., O-132, P-297, P-309<br />

Bissonnette, L., P-193<br />

Blanchard, A., O-203<br />

Blanchette Porter, M. M., P-83<br />

Blanco, L. A., P-698<br />

Blanco Mejia, S., P-638<br />

Blazek, J., P-239, P-507<br />

Bleess, J., O-107<br />

Blesson, C. S., P-50<br />

Blum, K., P-529<br />

Bochkovsky, S., P-82<br />

Boehnlein, L. M., P-346<br />

Boekelheide, K., P-199<br />

Boggino, C., P-636<br />

Bohrer, C., O-134, P-269, P-511<br />

Boivin, J., O-106<br />

Bolkas, M., P-323<br />

Bolnick, A., O-89, P-81, P-626<br />

Bolt, A., P-663<br />

Bond, K. R., O-97, O-99, P-129<br />

Bonetti, T., P-159<br />

Boostanfar, R., P-308<br />

Boots, C. E., O-15, P-78<br />

Borges Jr., E., P-275, P-276, P-277, P-281, P-596<br />

Borghi, C. M., O-36, P-296, P-636<br />

Borini, A., O-<strong>17</strong>6, O-252, P-103, P-151<br />

Bormann, C. L., O-138, P-564, P-669<br />

Borras, A., P-427<br />

Bosch, E., P-656<br />

Bose, G., P-353, P-398<br />

Botes, A., O-84<br />

Botting, B. J., O-93<br />

Boudoures, A., O-15<br />

Bouet, P. E., P-309<br />

Bouknight, J. M., P-220<br />

Boulet, S., O-263, P-2<strong>21</strong>, P-236, P-326, P-501<br />

Boulet, S. L., P-633<br />

Boylan, C. F., P-509<br />

Boynukalin, F. K., P-303<br />

Bozdag, G., O-67, P-441<br />

Bozkurt, N., O-68, P-712<br />

Brachtchenko, J., P-557<br />

Brackett, N., O-24<br />

Brackett, N. L., O-19<br />

Bradford, A. P., O-274, P-425<br />

Brady, P. C., P-343<br />

Braga, D. P. A. F., P-275, P-276, P-277, P-281,<br />

P-596<br />

Brakta, S., O-6, O-74, O-182, O-<strong>21</strong>1<br />

Brannigan, R., O-<strong>17</strong>1<br />

Brasile, D., P-688<br />

Braun, S., P-464<br />

Braverman, J., O-151<br />

Bray, M. A., P-334<br />

Brewer, A., P-238<br />

Bringer, S., P-182<br />

Bristow, S. L., O-168, O-196, P-14<br />

Britten, J., O-4<br />

Britten, J. L., O-184, P-142<br />

Broce, M., P-359<br />

Bronet, F., P-41<br />

Brower, M., P-500<br />

Brown, L., P-15<br />

Brown, M. B., O-115, O-241, P-298, P-452<br />

Brown, S. G., P-570<br />

Browne, A. S., O-5<br />

Brunet, C., O-248, P-193, P-254<br />

Bry, H., P-418<br />

B.S., G., P-353<br />

Buchanan, S., P-104<br />

Buchheit, K., O-256<br />

Buckbinder, J., P-88<br />

Buckett, B., P-297, P-302<br />

Buckett, W., P-263, P-506<br />

Buck Louis, G., O-154<br />

Buck Louis, G. M., O-114, O-<strong>21</strong>9, O-225<br />

Bukulmez, O., P-186, P-628<br />

Bulun, S. E., O-<strong>17</strong>9<br />

Bulut, H., P-435, P-653<br />

Burks, C. A., P-51<br />

Burks, H., P-88, P-726<br />

Burnett, B., P-8<br />

Burns, J. S., P-97<br />

Bush, M., P-519<br />

Bustillo, M., P-566<br />

Butler, S. A., P-352<br />

FERTILITY & STERILITY Ò<br />

e363


Butler, V. L., P-197<br />

Butts, S., O-<strong>21</strong>2, O-226, O-255, P-58<br />

Butz, A., P-306<br />

Buyalos, R. P., P-678<br />

Buyru, F., O-51, O-127, P-205<br />

Buyuk, E., O-273, O-276, P-47, P-338<br />

Buzas, B., P-436<br />

Cabanes, I., P-616<br />

Cabanillas, S., O-235<br />

Cabey, R., O-193<br />

Cabey, R. E., P-508<br />

Cabral, E. C., P-275, P-276, P-277<br />

Cadieux, G., P-402<br />

Cai, L., P-27, P-265, P-406, P-672<br />

Cakmak, H., P-273<br />

Calafat, A. M., O-96, O-223<br />

Calhoun, A. R., O-97<br />

Calis, P., P-441<br />

Calle, A., P-123<br />

Camargo, M., O-24, P-99, P-539<br />

Camillo, J., P-257, P-596<br />

Camlibel, T., P-645<br />

Campbell, A., P-691<br />

Can, S., P-205<br />

Canada, V. E., P-571, P-698<br />

Canas, M. T., O-131<br />

Cao, Y., O-43<br />

Capalbo, A., P-520, P-630<br />

Cardozo, E. R., O-150<br />

Cardozo, K., O-24, P-99<br />

Carignan, C. C., O-223<br />

Carlson, N., O-274<br />

Carlson, N. E., P-425<br />

Carlsson, M., O-<strong>17</strong>2<br />

Carneiro, M. M., P-148, P-376<br />

Carnethon, M., O-77, P-423<br />

Carnevali, O., O-<strong>17</strong>6, P-103, P-151<br />

Carney, S. M., P-509<br />

Carr, B., P-186, P-400, P-628<br />

Carrell, D., P-541<br />

Carrell, D. T., P-46<br />

Carrere, C. A., P-698<br />

Carusi, D. A., P-706<br />

Carvalho, A. T., P-148<br />

Carvalho, V., O-24, P-99<br />

Casals, G., P-427<br />

Casanova, P., P-698<br />

Casson, P. R., O-258<br />

Castells, M., O-256<br />

Castillon, G., O-148<br />

Cataldo, N. A., P-366<br />

Catapano, G., P-464<br />

Cater, E., P-691<br />

Catherino, W. H., O-4, O-184, P-142<br />

Cavagna, M., O-131, O-270<br />

Cearsolo, A., P-331<br />

Cebi, Z., P-645<br />

Cedars, M., P-29, P-64, P-273<br />

Cedenho, A. P., P-99<br />

Celestine, C. K., P-370<br />

Celia, G., P-227<br />

Celik, C., P-303<br />

Celik, H. T., P-62<br />

Celik, S., P-303<br />

Cengiz, C., O-208<br />

Centola, G. M., O-203<br />

Cernuda, D., P-616<br />

Cervantes, E., O-146, P-5, P-629, P-652, P-675<br />

Cevher, F., O-68<br />

Cevik, O., P-195, P-430, P-431<br />

Cha, H.-J., P-543<br />

Cha, J.-H., P-543<br />

Chaffin, C. L., O-161, P-270<br />

Chakraborty, P., P-157, P-243, P-398, P-401,<br />

P-531<br />

Chakravarty, B., P-157, P-243, P-353, P-398,<br />

P-401, P-531<br />

Chan, C., O-189, P-424, P-7<strong>21</strong>, P-731, P-732<br />

Chan, C. H. Y., P-723<br />

Chan, H., P-722, P-723, P-732<br />

Chan, J., O-197, P-447<br />

Chan, P. J., P-432, P-483<br />

Chang, C.-C., O-155, P-237, P-632, P-719<br />

Chang, C.-H., P-32<br />

Chang, E., P-433<br />

Chang, H.-M., P-3<br />

Chang, J., O-90, P-501<br />

Chang, S., O-58, P-69, P-338<br />

Chang, T.-C., P-719<br />

Chang, W., P-240, P-490, P-508<br />

Chantala, K., P-693<br />

Chantilis, S. J., P-241, P-245, P-259, P-480, P-658<br />

Chapa, C. R., P-405<br />

Charles, C., P-160, P-168<br />

Charo, L. M., O-7, P-2<br />

Charron, M., O-273, P-47<br />

Chattopadhaya, R., P-401<br />

Chattopadhyay, R., P-243, P-353, P-398, P-531<br />

Chatzicharalampous, C., P-334<br />

Chauhan, A., P-163<br />

Chauhan, S., P-239, P-322, P-361<br />

Chavarro, J. E., O-205, O-209, O-271, P-703<br />

Chavez, V., P-374<br />

Chavez-Badiola, A., O-63, P-580<br />

Chawla, S., P-650<br />

Chazenbalk, G. D., O-49, O-53, O-120<br />

Chen, A. A., P-134<br />

Chen, C.-D., P-32<br />

Chen, C.-H., P-455, P-461<br />

Chen, D., P-499<br />

Chen, H., P-690<br />

Chen, H.-H., P-627<br />

Chen, H.-W., P-166<br />

Chen, M., P-727, P-728<br />

Chen, M.-J., P-416, P-446<br />

Chen, S., P-79<br />

Chen, S. H., O-44, P-14<br />

Chen, S.-U., P-32, P-694<br />

Chen, X., P-181<br />

Chen, Y.-F., P-416, P-446<br />

Chen, Z., O-114<br />

Cheng, J.-C., P-3<br />

Chenoz, L., P-311<br />

Cherala, G., O-98<br />

Chernyak, E., P-37<br />

Chetani, M., P-157<br />

Chettier, R., O-<strong>17</strong>5<br />

Cheung, S., P-378, P-386<br />

Chi, H., P-267, P-344<br />

Chian, R.-C., P-27, P-265, P-406, P-672<br />

Chiang, Y.-C., P-514, P-522<br />

Chiang, Y.-T., P-32<br />

Child, T., P-590<br />

Chiles, K. C., O-5<br />

Chiles, K., P-373<br />

Chimote, B. N., P-59, P-583<br />

Chimote, N. M., P-59, P-583<br />

Chin, H. B., P-448, P-450<br />

Chinthala, S., P-162<br />

Chiware, T. M., P-440<br />

Cho, S., P-146<br />

Cho, W., P-180<br />

Cho, Y., P-149, P-<strong>17</strong>8, P-187<br />

Choi, D., P-154<br />

Choi, J., P-154<br />

Choi, J. M., P-472<br />

Choi, Y., P-433, P-572<br />

Cholkeri-Singh, A., O-72, P-202<br />

Chong, W., O-19<br />

Chosich, J., O-274<br />

Chosich, J. D., P-425<br />

Choudhary, M., P-513<br />

Chow, V., P-538<br />

Christensen, G. L., P-94<br />

Christensen, M. A., P-720<br />

Christianson, M. S., O-10, O-39, O-258, O-268,<br />

P-1, P-468<br />

Christman, G. M., O-258<br />

Christofolini, D. M., P-<strong>17</strong>9<br />

Christy, A., O-78, O-258, P-330<br />

Chu, B., O-163<br />

Chuan, S. S., P-310<br />

Chulet, S., P-185<br />

Chung, H., P-65, P-102<br />

Chung, K., P-108, P-339, P-453, P-677<br />

Chung, M., P-671<br />

Chung, P., O-60, P-248<br />

Chuong, F. S., O-58, P-468<br />

Chwalisz, K., O-181<br />

Ciapponi, A., P-296<br />

Cil, A., P-268<br />

Cil, A. P., O-1<strong>17</strong><br />

Cilingir, O., P-430<br />

Cilingir, O. T., P-195, P-431<br />

Cimadomo, D., P-630<br />

Cipolla, M., O-276<br />

Claessens, A., P-638<br />

Clapp, M. A., P-67<br />

Clarke, N. J., P-453<br />

Clarke, R., O-249, P-358<br />

Clarke, R. N., P-492<br />

Clarke-Williams, M., P-587<br />

Clemons, J., O-<strong>21</strong>4<br />

Coates, A., P-498<br />

Cobo, A., P-256, P-258, P-314, P-485<br />

Cocuzza, M., O-26, P-535<br />

Coddington, C., O-107, P-591<br />

Coello, A., P-256, P-258<br />

Coetzee, K., P-435, P-653<br />

Cohen, J., O-91<br />

Cole, A., P-548<br />

Collado, D., P-12<br />

Collazo, I., P-284, P-397, P-566, P-576<br />

Collins, G., P-412<br />

Collins, R., O-32, P-503<br />

Colls, P., O-35, O-63, P-585<br />

Collura, B., O-<strong>21</strong>6, P-659, P-668, P-716, P-7<strong>17</strong><br />

Comtet, M., P-476<br />

Conaghan, J., P-134<br />

Conceicao, C., P-379<br />

Confino, E., P-733<br />

Connaughton, H., P-1<strong>17</strong><br />

Connell, M. T., O-61, O-187, P-194<br />

Connerney, M. K., P-734<br />

Connor, J. J., P-727, P-728<br />

Considine, R. V., O-52<br />

Conti, M., P-273<br />

e364 Author Index Vol. 104, No. 3, Supplement, September <strong>2015</strong>


Convissar, S. M., O-14<br />

Conway, D. A., P-294<br />

Cooney, L., P-477<br />

Copperman, A. B., O-57, O-80, O-112, O-1<strong>21</strong>,<br />

O-146, O-160, O-164, O-204, O-245, P-5,<br />

P-8, P-130, P-131, P-345, P-368, P-419,<br />

P-442, P-629, P-652, P-665, P-675, P-684,<br />

P-700, P-730<br />

Cordeiro, C. N., P-468<br />

Cordeiro, F. B., P-257, P-276, P-277, P-596<br />

Corrado, J., O-70<br />

Correa, L. F., O-266<br />

Corselli, J., P-432, P-483<br />

Cosar, E., P-146<br />

Coscia, A., P-636<br />

Coskun, S., P-283<br />

Costantini, L., O-257<br />

Costantini, M., P-660<br />

Coulam, C., P-686<br />

Coutifaris, C., O-3, O-144, O-<strong>21</strong>2, O-226, O-258,<br />

P-306<br />

Coutinho, A., P-252<br />

Coward, K., P-570<br />

Cox, J., O-4, O-184, P-142<br />

Cox, J. M., P-325<br />

Cozzubbo, T., O-2, O-110, P-546, P-560, P-573<br />

Craig, L. B., O-50, O-<strong>17</strong>0, O-227, O-254<br />

Crawford, N. M., O-224, P-693, P-729<br />

Crawford, S., O-90, P-109, P-236, P-326<br />

Creus, M., P-427<br />

Crowe, R., O-2<strong>17</strong><br />

Cruz, M., P-12<br />

Csokmay, J., O-78, P-341<br />

Cui, Z., P-534, P-555, P-559<br />

Cunningham, D. L., O-81, P-734<br />

Czuprenski, E., P-512, P-518<br />

Dabaja, A., P-<strong>21</strong>6<br />

Dabaja, A. A., O-22, O-<strong>21</strong>3, P-456<br />

Da Broi, M. G., P-110<br />

Daftary, G., O-107<br />

Daftary, G. S., O-266, P-591<br />

Dahan, M., O-85, P-413<br />

Dai, J., P-81, P-626<br />

Dalloul, M., P-160, P-168<br />

Dambaeva, S., O-157, P-128<br />

Daneshmand, S., P-183, P-192, P-327, P-349<br />

Daneyko, M., O-80, P-368, P-730<br />

Daniel, C. E., O-250<br />

Danisman, N., P-63<br />

Danovitch, Y., P-497<br />

Danzer, H., P-240, P-473, P-490, P-500, P-677<br />

Daris, M., P-158<br />

Darmon, S. K., P-253, P-428, P-439<br />

Darwish, A. M., P-48<br />

Dasmahapatra, P., P-163<br />

Daubert, M., P-80<br />

Davenport, D. L., P-542<br />

Davenport, E., P-726<br />

Davidson, K., P-241<br />

Davidson, M., P-724<br />

Davie, J., P-295, P-307<br />

Davies, M. C., O-93<br />

Davila, A., P-405<br />

Davis, J., O-264, P-740<br />

Davis, L., P-<strong>21</strong>0<br />

Davis, O. K., P-348, P-420<br />

Davis, P., P-294<br />

Dean, N., O-132<br />

Dean, R., O-25<br />

Debrock, S., P-136<br />

de Carvalho, C. V., P-159<br />

De Chavez, P. D., O-77, P-423<br />

DeCherney, A., O-61, O-78, O-140, O-142, O-187,<br />

O-191, O-220, P-6, P-15, P-194, P-287,<br />

P-3<strong>21</strong>, P-328, P-330, P-341, P-382, P-394,<br />

P-399, P-713<br />

Declercq, E., O-37, O-42<br />

De Gheselle, S., O-55, P-249<br />

de Haydu, C., O-22, P-456<br />

Deibert, C., P-383<br />

Deibert, C. M., O-<strong>17</strong>1<br />

Deimling, T. A., P-141<br />

Delaney, A. A., O-266<br />

Delaroche, L., P-193<br />

Delgado, A., O-45<br />

de los Santos, J., P-485<br />

de los Santos, M., P-485, P-600<br />

De Meglio, G., P-576<br />

de Melo, A. A., P-257, P-596<br />

DeMichele, A., O-7, P-2<br />

Demick, J. L., O-203<br />

Demir, B., O-67, P-350, P-699<br />

Demiral, I., O-51, O-127<br />

Demirci, U., O-188<br />

Demko, Z., P-519<br />

Denes, F. T., O-26<br />

Deng, J., P-710<br />

Denomme Tignanelli, M. M., O-46<br />

De Paz, C. C P., P-110<br />

Desai, N., O-251, P-553, P-611<br />

Dessapt, A., P-418<br />

De Sutter, P., O-55, P-249<br />

Detti, L., O-13, P-91<br />

Devine, K., O-61, O-187, P-84, P-287, P-3<strong>21</strong>,<br />

P-328, P-330, P-382, P-394, P-4<strong>21</strong><br />

Dewailly, D., O-129<br />

Dhanjani, D. G., O-199<br />

D’Hooghe, T., P-136<br />

Di, C.-G., P-457<br />

Dia, F., P-280<br />

Diamond, M., O-6, O-74<br />

Diamond, M. P., O-73, O-75, O-76, O-92, O-182,<br />

O-185, O-186, O-201, O-<strong>21</strong>1, O-258, P-81,<br />

P-113, P-115, P-<strong>21</strong>5, P-626<br />

Diaz-Garcia, C., P-449, P-644<br />

Diaz-Gimeno, P., P-597<br />

Dickler, M., P-458<br />

Dieamant, F., O-131, O-270<br />

Diez Juan, A., P-597<br />

DiGiovanni, L., O-197<br />

Dilbaz, B., P-375<br />

Dilbaz, S., P-375<br />

DiMattina, M., P-227, P-677<br />

Dimitriadis, I., O-200<br />

Ding, J., P-385<br />

Ding, X., O-49<br />

Diop, H., O-37, O-42<br />

Dmitrieva, N., P-663, P-667<br />

Do, L., P-197<br />

Dodds, W., P-127<br />

Dodge, L., P-301<br />

Dodge, L. E., P-340<br />

Dokras, A., O-169, P-58<br />

Dolinko, A. V., P-469<br />

Domar, A. D., O-106, P-301<br />

Domchek, S. M., O-197<br />

Domingo, J., P-644<br />

Domingues, T. S., P-261, P-351<br />

Donahue, G., P-556<br />

Dong, F., O-8<br />

Dong, J., P-27, P-406<br />

Dong, S., P-100<br />

Donjacour, A., P-242<br />

Doody, K., P-400<br />

Douglas, N. C., O-189, P-424, P-710<br />

Doyle, J., P-482<br />

Doyle, N., P-399<br />

Drevet, J., P-123<br />

Drewlo, S., O-89, O-2<strong>21</strong><br />

Driggers, P., O-58, P-84<br />

Druckenmiller, S., O-123, O-231, P-489, P-567,<br />

P-587<br />

Druschel, C., O-154<br />

Du, H., P-367<br />

Du, T., P-690<br />

Dubaut, J. P., O-8<br />

Dudley, P., P-677<br />

Duhamel, A., O-129<br />

Duke, C. M. P., O-22, O-<strong>21</strong>3, P-<strong>21</strong>6, P-456<br />

Duke, J., O-<strong>21</strong>4<br />

Duke, M., O-146, P-684<br />

Dumesic, D. A., O-49, O-53, O-120<br />

Dumitriu, B., O-142<br />

Dun, E. C., O-<strong>21</strong>3, P-<strong>21</strong>6<br />

Duncan, F. E., P-269, P-422<br />

Duncan, J., P-738<br />

Dundee, J. A., P-83<br />

Dunn, R., P-239, P-322, P-361<br />

Duong, J., O-8<br />

Du Plessis, S. S., P-101, P-114, P-116, P-544,<br />

P-552<br />

Durairajanayagam, D., O-111, P-381, P-387<br />

Dural, O., O-51, O-127, P-205<br />

Duros, S., P-460<br />

Duval, K., O-253<br />

Dwan, P. G., O-81<br />

Dye, T., P-299, P-377<br />

Dyson, M. T., O-<strong>17</strong>9<br />

Dzidic, N., P-516<br />

Eaton, J., P-48<br />

Eaton, J. L., O-40, O-119, O-261<br />

Eberlin, M. N., P-275, P-276, P-277<br />

Eckel, R. E., O-274, P-425<br />

Edmonds, J. W., P-279<br />

Edwards, T. L., P-145<br />

Efymow, B., P-477<br />

Egashira, A., P-610<br />

Eid, M. E., P-687<br />

Eid, S., O-102<br />

Eisenberg, E., O-92, O-201, O-258<br />

Eisenberg, M., O-114, P-550<br />

Eisenberg, M. L., O-203<br />

Eisermann, J., P-284, P-397<br />

Eken, M., P-195, P-430, P-431<br />

Elam, L., O-73, O-185, O-186<br />

Elam, L. A., O-75, O-76<br />

El Assal, R., O-188<br />

ElBardisi, H., O-111<br />

Elbareg, A. M., P-204<br />

El-Baz, M. A. H., O-228<br />

Eleswarapu, S. V., O-22, P-456<br />

El Hachem, H., O-132, P-309<br />

El-Husseini, H., O-74<br />

Elias, R., P-649, P-707<br />

Elias, R. T., P-135<br />

El-khayat, W., O-66<br />

Ellen, J., P-540<br />

FERTILITY & STERILITY Ò<br />

e365


Elliot, M. C., P-332<br />

Elliott, T., O-155<br />

Elnashar, I., O-228, P-<strong>21</strong>1<br />

El Sadek, M., O-66<br />

Eltsova, E., P-10<br />

Emerson, G., P-650<br />

Emirdar, V., P-230, P-462, P-471<br />

Engmann, L., O-255, P-255, P-479, P-624<br />

Entezami, F., P-182, P-193<br />

Er, T.-K., P-147<br />

Erdem, A., O-68, P-623, P-635, P-712<br />

Erdem, E., P-645<br />

Erdem, M., O-68, P-623, P-635, P-712<br />

Erdinc, A. S., P-23<br />

Erdogan, D., P-558<br />

Eren, F., P-52<br />

Ergin, M., P-52, P-60, P-63<br />

Erkilinc, S., P-23<br />

Ermilova, I., P-663<br />

Ersoy, A. O., P-62<br />

Ersoy, E., P-62<br />

Escudero, T., O-35, P-495<br />

Esfandiari, N., P-83<br />

Eskew, A. M., P-332<br />

Esteves, S., P-535<br />

Eum, J., P-433<br />

Evans, B. T., P-666<br />

Evans, E. A., O-166<br />

Evans, J. P., P-274<br />

Evans, S., O-105<br />

Evans-Hoeker, E. A., P-45, P-231<br />

Ezzati, M., P-186, P-400, P-628<br />

Fabregues, F., P-427<br />

Fadiel, A., O-276<br />

Fadiloglu, E., P-375<br />

Fadiloglu, S., P-375<br />

Falcao Jr, J. A., P-148<br />

Falcone, T., O-251, P-<strong>17</strong>4, P-198, P-203<br />

Fanchin, R., O-159, P-305, P-311<br />

Fang, C., O-133, P-457<br />

Fang, L., P-3<br />

Farghaly, T. A., O-69, O-228, O-259, P-<strong>21</strong>1<br />

Farhadifar, R., P-356<br />

Farland, L. V., O-232, P-49, P-343, P-356, P-469<br />

Farrington, P., O-<strong>17</strong>5<br />

Faulkner, N., P-307<br />

Faustmann, T., O-100, P-371<br />

Fawzy, M., P-233, P-613<br />

Fazleabas, A., P-127<br />

Fedder, J., O-202<br />

Fedick, A., P-505<br />

Feinberg, E., P-704<br />

Feinberg, R. F., P-509<br />

Feldman, G., P-7<br />

Feldman, R., O-181, P-58<br />

Feliciano, M., P-373<br />

Feng, C., P-75, P-1<strong>21</strong><br />

Feng, R., O-144<br />

Feng, X., P-<strong>17</strong>7<br />

Ferguson, K., P-538<br />

Ferigolo, P., P-539<br />

Ferle, L., P-681<br />

Fernandes, R., P-<strong>17</strong>9<br />

Fernandez, M., O-128, O-<strong>17</strong>3, P-685<br />

Fernandez Gallardo, E., P-136<br />

Ferrando, M., P-314<br />

Ferrer, A., O-47<br />

Ferriani, R., P-110, P-697<br />

Ferriani, R. A., P-654<br />

Ferrieres-Hoa, A., O-239, P-254, P-360, P-444<br />

Ferrieres-Hoa, A. F., O-248<br />

Feskov, O., P-396<br />

Feskov, V., P-396<br />

Feuer, S., P-242<br />

Ficicioglu, C., P-437<br />

Fierro, M. A., O-14<br />

Figueira, R. C. S., P-281<br />

Fino, M., P-25<br />

Fino, M. E., O-104, O-231, P-77<br />

Fiorentino, I., P-464<br />

Fisch, B., O-237<br />

Fischer, E., P-519<br />

Fishel, S., P-691<br />

Fisseha, S., P-736<br />

Fiszbajn, G., O-36, P-636<br />

Flaws, J. A., O-10<br />

Fletcher, N. M., O-13, P-91, P-113, P-115, P-<strong>21</strong>5<br />

Florensa, M., P-520<br />

Flowers, L., P-501<br />

Flyckt, R., P-<strong>17</strong>4, P-198<br />

Flynn, K. E., P-740<br />

Fode, M., O-28, P-537<br />

Foer, D., O-256<br />

Folger, S., P-109<br />

Fontenot, G. K., O-29<br />

Foong, S., O-<strong>21</strong>5<br />

Ford, J. B., O-96<br />

Forman, E., P-526<br />

Forman, E. J., O-180, O-244, P-223, P-527, P-584,<br />

P-606<br />

Foster, K., P-86<br />

Fothergill, A., P-448, P-450<br />

Fox, J. H., P-706<br />

Fragouli, E., O-91, O-252<br />

Franasiak, J. M., O-31, O-135, O-137, O-180,<br />

O-244, P-223, P-370, P-527, P-584, P-586,<br />

P-594, P-595, P-606, P-612, P-660, P-715<br />

Franca, U., P-564<br />

Franciosi, F., P-273<br />

Franco Jr, J. G., O-131, O-270<br />

Frankel, R., P-11<br />

Fraser, B., P-1<strong>17</strong><br />

Frattarelli, J. L., O-196<br />

Fredrickson, J., P-98<br />

Freeman, T., O-103, P-737, P-739<br />

Friebel, T. M., P-447<br />

Friedman, L., P-104<br />

Fritz, A. E., P-619<br />

Fritz, R., O-89, O-2<strong>21</strong><br />

Fru, K. N., P-325<br />

Fuehrer, D., O-107<br />

Fujii, M. G., O-149, P-718<br />

Fukuda, A., P-56, P-563<br />

Fukunaga, N., O-64, P-579, P-618<br />

Funabiki, M., O-192, P-320, P-561<br />

Funagayama, Y., P-579<br />

Fung, J. L., P-120<br />

Fung, M. Y., O-81, P-734<br />

Furuya, K., P-459<br />

Gable, E., P-377<br />

Gabriele, D. A., P-494<br />

Gabrielsen, J., O-205<br />

Gad, M., P-233, P-613<br />

Gada, R., P-245, P-259, P-480, P-658<br />

Gaines, T., P-1<br />

Gajbhiye, R., P-163<br />

Gala, A., O-247, O-248, P-254, P-360, P-444<br />

Galai, N., O-140, O-220, P-6, P-705, P-713<br />

Galal, A. F., P-648<br />

Galan, A., P-256, P-673<br />

Galiana, Y., P-520<br />

Galkina, E., P-319<br />

Galliano, D., O-262<br />

Ganguly, N., P-82<br />

Gantt, P. A., P-359<br />

Ganzabal, T., P-331<br />

Gao, J., O-181<br />

Gaona, M., O-84<br />

Garcia, B. A., P-556<br />

Garcia, D., O-153<br />

Garcia, J., P-576<br />

Garcia, J. E., P-468<br />

Garcia, V., P-673<br />

Garcia-Velasco, J. A., P-41, P-616, P-644<br />

Gardner, D. K., O-152, P-610<br />

Garg, D., P-264, P-709<br />

Gargiulo, A. R., O-256<br />

Garner, F., P-183, P-192, P-327, P-349<br />

Garnsey, H., P-269, P-526<br />

Garrido, N., O-262<br />

Garrisi, M., O-163<br />

Garzo, G., P-495<br />

Gaskins, A. J., O-205, O-209, P-703<br />

Gat, I., O-236, O-238, P-638<br />

Gates, D., P-308<br />

Gauthier-Fisher, A. S., O-236, O-238<br />

Gavrilova-Jordan, L., O-182, O-185<br />

Gayet, V., O-129<br />

Gebhart, M. B., O-82<br />

Gehrke, K., P-619<br />

Gemzell-Danielsson, K., O-100<br />

Geneidi, E., P-686<br />

Genro, V., O-159<br />

Gentry, A. L., P-45<br />

George, J., O-122<br />

Gerami-Naini, B., O-188<br />

Gerkowicz, S. A., P-651<br />

Gerson, M., O-204<br />

Ghadir, S., O-196, P-240, P-490, P-495<br />

Ghanem, M., P-333, P-347<br />

Ghant, M., P-138, P-143, P-733<br />

Gharagozloo, P., P-1<strong>17</strong>, P-123<br />

Ghassabian, A., O-154<br />

Gheyas, F., P-646<br />

Ghosh, B., P-398, P-401, P-531<br />

Ghosh, S., P-157, P-353, P-398, P-401, P-531<br />

Ghossein-Doha, C., P-674<br />

Ghuge, A., P-163<br />

Gibbons, W. E., P-50, P-565, P-607<br />

Gibson-Corley, K. N., P-547<br />

Giles, J., O-148<br />

Gill, P. K., P-611<br />

Gingold, J., O-146, O-164, O-204, P-368, P-442,<br />

P-665, P-700<br />

Ginsburg, E., O-<strong>17</strong>, O-232, P-49, P-343, P-469<br />

Gioacchini, G., O-<strong>17</strong>6, P-103, P-151<br />

Giorgi, V. S I., P-110<br />

Giorgini, E., O-<strong>17</strong>6, P-103, P-151<br />

Girao, M. B., P-159<br />

Giudice, L. C., P-692<br />

Giuliani, E., P-127<br />

Givens, C., P-519<br />

Glass, K., O-<strong>21</strong>5, P-196, P-557<br />

Glassner, M. J., P-688<br />

Gleason, K., P-499<br />

Gleicher, N., O-151, P-253, P-428, P-439, P-592,<br />

P-622<br />

e366 Author Index Vol. 104, No. 3, Supplement, September <strong>2015</strong>


Globus, S. T., O-57<br />

Glujovsky, D., P-296<br />

Go, K. J., O-81, P-734<br />

Godfrey, E. M., O-257<br />

Goel, T., P-689<br />

Goering, M., P-619<br />

Gold, M., O-168<br />

Goldberg, J., O-194<br />

Goldberg, J. D., O-166<br />

Goldberg, J. M., O-251, P-198, P-203, P-611<br />

Goldberg, R. W., P-611<br />

Goldberg-Strassler, D., O-193, P-508<br />

Goldfarb, J. M., O-38, O-41, O-118<br />

Goldfarb, S., P-458<br />

Goldman, K. N., O-12, O-104, O-123, O-198,<br />

O-231, P-25, P-422<br />

Goldman, M. B., P-120<br />

Goldman, R. H., P-49, P-669<br />

Goldstein, M., O-109, P-373, P-546<br />

Golombok, S., O-65, O-103, O-105, P-737, P-739<br />

Gomes, A., P-718<br />

Gomes, A. P., O-149<br />

Goncalves, S. P., O-149, P-718<br />

Gonzalez, F., O-52, O-267<br />

Gonzalez, M., P-331<br />

Gonzalez-Fernandez, R., P-9<br />

Gonzalez-Foruria, I., P-427<br />

Goodall, N.-N., O-163<br />

Goodman, L. R., O-251, P-<strong>17</strong>4, P-198, P-203<br />

Goodrich, D., P-511<br />

Gopal, D., O-37, O-42, P-292<br />

Gordon, J., P-227<br />

Gordon, K., P-308<br />

Gordon, T. T., P-512<br />

Gorman, J., P-470<br />

Goswami, M., P-353, P-531<br />

Goswami, S. K., P-243, P-353, P-401, P-531<br />

Goswami, S., P-458<br />

Goto, H., P-285<br />

Goyal, A., P-<strong>17</strong>4<br />

Gracia, C., O-92, O-197, P-445, P-447, P-477<br />

Graham, J., P-238, P-252, P-482<br />

Grainger, D. A., O-8<br />

Grantz, K. L., O-<strong>21</strong>9<br />

Grazi, R., O-275, P-264, P-709<br />

Green, K. A., O-135, P-594, P-595, P-715<br />

Greenberg, S. A., P-<strong>21</strong>0<br />

Greenblatt, E., P-196<br />

Greene, A. F., O-56<br />

Greene, N., O-260, P-313, P-316<br />

Griesinger, G., P-336<br />

Griffin, D. K., O-48, P-491, P-676<br />

Grifo, J., O-35, O-44, O-91, O-123, O-196, O-198,<br />

P-77, P-82, P-422, P-478, P-489, P-493,<br />

P-499, P-504, P-510, P-524, P-525, P-578,<br />

P-603, P-664, P-680<br />

Grifo, J. A., O-34<br />

Grimstad, F., P-86, P-335<br />

Grober, E. D., O-<strong>17</strong>1, O-199, P-133<br />

Grogan, T. R., O-53<br />

Gronier, H., P-476<br />

Gross, J., O-106<br />

Grossman, L. C., O-189, P-266, P-424<br />

Grunert, G., P-322<br />

Grunert, G. M., P-239, P-361<br />

Grunfeld, L., O-245, P-130, P-131, P-652<br />

Grynberg, M., P-460, P-476<br />

Gualtieri, R., P-464<br />

Guan, Y., P-308, P-646<br />

Guedikian, A. A., O-120<br />

Gueye, N.-A., P-198<br />

Guijarro, M., O-128, O-<strong>17</strong>3, P-685<br />

Guillen, A., O-148<br />

Guillen, J., O-195<br />

Gulati, S., P-398<br />

Guler, I., O-68, P-623, P-635, P-712<br />

Gultomruk, M., P-303<br />

Gunes, S., P-545<br />

Gunes, Z. K., O-44<br />

Gungor Ugurlucan, F., P-205<br />

Guo, H., P-<strong>17</strong>3<br />

Guo, M., P-105<br />

Guo, Y., O-59, P-44, P-70, P-670<br />

Gupta, M., P-198, P-689<br />

Gupta, S., P-101, P-111, P-112, P-114, P-116,<br />

P-119, P-381, P-451, P-544, P-552, P-559<br />

Gurtcheff, S. E., P-294<br />

Gustofson, R. L., O-125<br />

Gutierrez-Adan, A., P-123<br />

Gutmann, J. N., P-8, P-345, P-419<br />

Guven, S., O-188<br />

Guvenir, H., P-350<br />

Guzel, Y., O-162, P-33, P-475, P-634<br />

Gwinnett, D., O-250<br />

Ha, T. Kim., P-400<br />

Haberman, S., P-709<br />

Haberstroh, W. P., P-688<br />

Hacker, M. R., P-301, P-315, P-340<br />

Haddad, B., P-418<br />

Haddad, G., O-34<br />

Hade, E. M., P-34<br />

Haden, S., P-738<br />

Hadjiliadis, D., P-374<br />

Hahn, K. A., P-<strong>17</strong><br />

Haji, A., P-404<br />

Hakim, L., P-548<br />

Halder, S. K., O-75<br />

Hall, S. J., P-199<br />

Halow, N. G., O-101<br />

Haltas, H., P-61<br />

Hamamah, S., O-239, O-247, O-248, P-182,<br />

P-193, P-254, P-305, P-360, P-444<br />

Hamer, J., O-36<br />

Hammes, S., O-54<br />

Hammes, S. R., P-139<br />

Hammond, K. R., P-366<br />

Han, M., P-149, P-187<br />

Han, O., P-38<br />

Han, T., P-399<br />

Hancock, K., P-707<br />

Hanna, C., O-95, O-98<br />

Hannam, T., O-190<br />

Hansen, K. R., O-50, O-<strong>17</strong>0, O-254, O-255, O-258<br />

Hanson, B. M., P-227<br />

Hanson, H., P-46, P-541<br />

Hao, C., P-66<br />

Hao, J., P-44<br />

Haouzi, D., P-182, P-193<br />

Haque, I. S., O-166, O-167, O-194<br />

Harada, T., P-<strong>17</strong>0<br />

Harden, S. A., P-45<br />

Harkins, G., P-141<br />

Harlev, A., P-101, P-116, P-552<br />

Harris, A., P-280<br />

Harris, M., O-274<br />

Harris, M. A., P-425<br />

Harrity, C., P-426, P-574<br />

Hartman, M., P-287<br />

Hartman, T. J., P-120<br />

Hartmann, K., P-145<br />

Harutunian, A., P-507<br />

Hasegawa, E., P-407<br />

Hasegawa, H., P-232<br />

Hashimoto, S., P-285<br />

Hasson, J., P-302<br />

Haswell, C., O-264<br />

Hatch, E. E., P-<strong>17</strong>, P-22, P-384<br />

Hattori, H., P-536<br />

Haunschild, C., O-7, P-2<br />

Hauser, R., O-96, O-136, O-209, O-223, O-271,<br />

P-97<br />

Haviland, M., P-315<br />

Hawa, N. S., P-404<br />

Hawkins Bressler, L., O-77<br />

Hayashi, T., O-192, P-320, P-561<br />

Hayden, R. P., O-207<br />

Hayward, B., O-155<br />

Hazlett, W. D., P-615<br />

He, A. W., O-255<br />

Healy, M. W., O-25, O-61, O-187, P-194, P-328<br />

Hebert, J., O-70, P-695<br />

Hebisha, S. A., P-484, P-581, P-637<br />

Heindryckx, B., P-249<br />

Heiser, P. W., P-631<br />

Helal, A. S., P-333, P-347<br />

Heller, B., P-724<br />

Hellmers, A., P-661<br />

Helwa, I., O-182<br />

Hennebold, J. D., O-101<br />

Henriquez, S., P-651<br />

Hernandez, J., P-9<br />

Herraiz, S., P-449<br />

Hershlag, A., O-44, O-163, P-323, P-497<br />

Hesla, J., O-34<br />

Hesley, G., P-137<br />

Hickman, C., O-250<br />

Hijazi, H., P-283<br />

Hill, D., P-500<br />

Hill, D. L., O-120, P-240, P-247, P-473, P-490<br />

Hill, M., P-287<br />

Hill, M. J., O-61, O-187, P-194, P-328, P-330,<br />

P-341<br />

Hill, W. D., O-6<br />

Hillis, L., O-81<br />

Hinckley, M., O-246<br />

Hindoyan, R. H., P-339<br />

Hines, R. S., O-82<br />

Hipp, H., O-90<br />

Hipp, H. S., P-633<br />

Hirshfeld-Cytron, J., P-724<br />

Hirshfeld-Cytron, J. E., P-51<br />

Hixon, B., P-329, P-601<br />

Ho, S.-M., P-94<br />

Hoang, L., O-37, O-42<br />

Hobeika, E., O-60, P-649<br />

Hodes-Wertz, B., O-104, P-493, P-504, P-603<br />

Hodis, H. N., P-88<br />

Hoeger, K., O-54, P-299, P-377<br />

Hoff, H. S., O-16, P-729<br />

Hoffman, D. I., P-508<br />

Holden, J. P., P-530<br />

Holder, S., O-226<br />

Holland, A. C., O-82<br />

Holley, S. R., P-725<br />

Holmes, R., P-238<br />

Holoch, K. J., P-198<br />

Holzer, H., P-263<br />

FERTILITY & STERILITY Ò<br />

e367


Homel, P., P-264<br />

Homer, M. V., O-7, P-2<br />

Hong, K. H., O-31, O-244, P-584, P-594, P-595,<br />

P-606, P-612, P-715<br />

Hopeman, M. M., P-445<br />

Horne, A. W., P-701<br />

Hornstein, M. D., P-666, P-669<br />

Hotaling, J., P-541<br />

Hotaling, J. M., P-46<br />

Hourvitz, A., P-87, P-89<br />

Howard, B., O-102<br />

Howards, P. P., P-448, P-450<br />

Howles, C. M., P-336<br />

Hsu, J. W., P-50<br />

Hsu, J. Y., P-266<br />

Huang, C.-C., P-627<br />

Huang, H., P-100, P-105, P-190, P-4<strong>17</strong><br />

Huang, J.-Y., P-188<br />

Huang, R., P-657<br />

Huang, W., P-156, P-<strong>17</strong>7<br />

Huang, X., P-66, P-582<br />

Huang, Y., P-4<strong>17</strong><br />

Huang, Y.-H., P-530<br />

Hubbard, J., O-47<br />

Huberlant, S., P-360<br />

Hubert, G., P-678<br />

Huddleston, H., P-64<br />

Hudson, C., P-183, P-192, P-327, P-349<br />

Huffman, C. S., O-79<br />

Hughes, J., P-15<br />

Hughes, M., P-512, P-5<strong>17</strong>, P-518<br />

Huleihel, M., O-237<br />

Humane, A. C., P-163<br />

Humm, K. C., P-291<br />

Hunter Cohn, K., P-345<br />

Hur, J., P-71<br />

Hurd, W. W., O-40, O-119, O-261, P-48, P-568<br />

Hurst, B. S., P-332<br />

Hu-Seliger, T., P-8, P-419<br />

Hutchinson, A. P., P-135, P-649<br />

Hwang, K., O-<strong>21</strong>, O-234, P-199<br />

Iaconelli Jr., A., P-275, P-276, P-277, P-281<br />

Iba, Y., P-251<br />

Ibrahim, E., O-19, O-24<br />

Ibrahim, Y., P-315<br />

Ida, M., P-56<br />

Ikuma, S., P-226, P-324, P-369, P-388<br />

Ilagan, Y., O-<strong>17</strong>8<br />

Iles, R. K., P-352, P-620<br />

Iliodromiti, S., O-9<br />

Ilioi, E. C., O-65<br />

Ilnitsky, S., P-696<br />

Imbalzano, M., O-226<br />

Imrie, S., O-105<br />

Imudia, A., P-201<br />

Inal, H. A., P-23, P-38, P-61<br />

Inal, Z. O., P-38<br />

Ince, U., O-162<br />

Ingles, S., P-339<br />

Inoue, M., P-285<br />

Insua, M. F., P-314<br />

Intasqui, P., P-99, P-539<br />

Irani, M., O-275, P-264, P-709<br />

Irani, S., O-275<br />

Irwin, J., P-692<br />

Isaacs, J. D., P-197<br />

Isaacson, K., P-209<br />

Isaka, K., P-407<br />

Ishihara, K., P-579<br />

Ishikawa, T., O-27, O-113, P-390, P-549, P-6<strong>17</strong>,<br />

P-683<br />

Islas, C. A., P-405<br />

Israel, M. A., P-719<br />

Ito, H., P-407<br />

Ivani, K., O-246<br />

Iwahashi, K., P-262<br />

Iwahata, T., P-391, P-532<br />

Iwaki, Y., O-192, P-320, P-561<br />

Iwata, K., P-251<br />

Jacobs, I., O-93<br />

Jacobson, M. H., P-448<br />

Jacoby, V., P-137<br />

Jadhav, P., P-646<br />

Jadva, V., O-65, O-105, P-737, P-739<br />

Jahoor, F., P-50<br />

Jain, G., P-185<br />

Jain, J., P-585<br />

Jain, R., O-181<br />

Jalas, C., P-494, P-505<br />

James, A. N., O-25<br />

James, D., O-229<br />

Jamieson, D., P-109<br />

Jamieson, D. J., O-90, O-263, P-236, P-326, P-633<br />

Jang, H., P-433<br />

Jarmuz, P., P-569<br />

Jaroudi, S., O-163, P-508, P-590<br />

Jarvi, K., O-236<br />

Jarvi, K. A., O-<strong>17</strong>1, O-199, P-133<br />

Jasani, S., P-51, P-724<br />

Jasulaitis, S., P-724<br />

Jean-Denis, F., O-253<br />

Jeng, G., P-501<br />

Jenkins, J., P-334, P-336<br />

Jenner, L., P-691<br />

Jensen, C. F., O-28, P-537<br />

Jensen, J., O-95, O-98, O-107, P-591<br />

Jeon, J., P-65, P-102<br />

Jeong, H. J., P-671<br />

Jeong, J. H., P-562<br />

Jeong, K., P-65, P-102<br />

Jha, A., O-11, P-72<br />

Ji, D., O-43<br />

Jianini, B. T., P-110<br />

Jimenez, C., P-331<br />

Jimenez, J., P-367<br />

Jin, H., P-670<br />

Jin, S., O-92, O-258<br />

Jincho, Y., P-599<br />

Jindal, S., O-273<br />

Jindal, S. K., P-47, P-338<br />

Jo, J.-D., P-267<br />

Jo, M., P-154<br />

Joelsson, L. S., P-224<br />

Joergensen, N., O-205<br />

Johnson, D., P-240, P-490<br />

Johnson, L., O-169<br />

Johnson, L. N. C., O-197<br />

Johnson, M. D., P-92<br />

Johnson, M. R., P-284<br />

Johnstone, E. B., P-46<br />

Johnston-MacAnanny, E. B., P-19<br />

Jones, C., P-570<br />

Jones, J. M., P-346<br />

Jones, M. E., O-93<br />

Jones, T., P-108<br />

Jordan, A., O-163, O-193, P-677<br />

Jorge, S., O-58<br />

Jou, G., P-32, P-694<br />

Juarez, L., P-724<br />

Julka, N., O-275<br />

Juneau, C. R., O-31, O-137, O-180, O-244, P-370,<br />

P-584, P-594, P-595, P-606, P-660, P-715<br />

Jungheim, E., P-293<br />

Jungheim, E. S., P-43<br />

Juvet, T., P-133<br />

Jwa, J., P-312<br />

Jwa, S., P-312<br />

Kachhawa, G., P-689<br />

Kadoch, I.-J., P-225, P-297<br />

Kahn, B. E., P-542<br />

Kahraman, S., O-1<strong>17</strong><br />

Kahyaoglu, I., O-67, P-350, P-699<br />

Kaing, A., P-500<br />

Kalem, M. N., P-38<br />

Kalgi, B., O-275, P-264<br />

Kaliappan, S., P-264<br />

Kalinchenko, S., P-10<br />

Kalinina, E., P-663<br />

Kalmbach, K., P-271, P-272<br />

Kalra, B., P-445<br />

Kamel, M. A., O-69, O-259, P-<strong>21</strong>1<br />

Kamga-Ngande, C., P-225<br />

Kang, B.-M., P-180<br />

Kang, H. H., P-562<br />

Kang, S. M., P-411<br />

Kanninen, T., O-158, P-126<br />

Kanninen, T. T., P-125<br />

Kan-Ool, L., P-557<br />

Kant, G., P-185<br />

Kao, A.-P., P-166<br />

Kao, C.-N., P-29, P-64<br />

Kapfhamer, J. D., O-<strong>21</strong>6, P-659, P-668, P-716,<br />

P-7<strong>17</strong><br />

Kaplanaoglu, I., P-350<br />

Kaplanoglu, I., P-558, P-699<br />

Karaca, N., P-229<br />

Karacan, M., P-645<br />

Karahasanoglu, A., O-1<strong>17</strong><br />

Karakas, N., P-52, P-60<br />

Karande, V., P-615<br />

Karita, M., O-192, P-320, P-561<br />

Karmon, A. E., O-150, P-35<br />

Karvir, H., O-57, O-160, P-345<br />

Kasapoglu, I., P-167<br />

Kasem, H., P-233, P-613<br />

Kaser, D. J., P-669, P-706<br />

Kashanian, J. A., O-<strong>17</strong>1<br />

Kaskar, K., P-607<br />

Kassira, S., P-220<br />

Kastury, R. D., P-410<br />

Kathiresan, A., O-260<br />

Kathrins, M., P-408, P-481, P-528, P-533<br />

Katilius, J., O-257<br />

Katsoula, E., P-640<br />

Katz, P. P., P-735<br />

Katz-Jaffe, M., O-32, O-46, O-48, O-125, O-165,<br />

O-<strong>17</strong>4, P-234, P-289, P-329, P-438, P-503,<br />

P-589, P-601, P-676<br />

Kavoussi, S. K., P-346<br />

Kawabata, Y., P-599<br />

Kawwass, J. F., O-90, P-633<br />

Ke, Z., P-190<br />

Keefe, D. L., O-168, P-271, P-272, P-422<br />

Keller, M., O-96, O-271<br />

Kellogg, G. R., O-168<br />

Kenigsberg, D., O-193<br />

Kenigsberg, S., O-236<br />

e368 Author Index Vol. 104, No. 3, Supplement, September <strong>2015</strong>


Kennedy, E., O-<strong>21</strong>5<br />

Ketterson, K., P-498<br />

Keyhan, S., O-38, O-41, O-118, P-48, P-568<br />

Khalafalla, K., O-111<br />

Khan, I., P-609<br />

Khan, O., O-28, P-537<br />

Khan, S., O-143, P-282<br />

Khan, Z., O-107, O-266, P-591<br />

Kharitonova, M., P-663<br />

Khater, M. K., O-74<br />

Kholy, A. M., P-333<br />

Khruapenkova, T., P-663<br />

Kida, Y., O-64, P-579<br />

Kiehl, M., P-519<br />

Kieslinger, D. C., O-55<br />

Kilburn, B., O-2<strong>21</strong><br />

Kilburn, B. A., O-89<br />

Kilpatrick, S. J., P-316<br />

Kiltz, R., O-193<br />

Kiltz, R. J., O-34<br />

Kim, B. Y., P-562<br />

Kim, C.-H., P-180<br />

Kim, D., P-619<br />

Kim, H., P-180, P-380<br />

Kim, H.-Y., P-543<br />

Kim, J., P-268, P-465<br />

Kim, J. H., P-411<br />

Kim, J. W., P-671<br />

Kim, J. Y., P-562<br />

Kim, K.-H., P-365<br />

Kim, M., P-572, P-671<br />

Kim, M. J., P-562<br />

Kim, M.-L., P-149, P-<strong>17</strong>8<br />

Kim, R., P-572<br />

Kim, S., O-114, P-267, P-344, P-380<br />

Kim, T., P-71<br />

Kim, T. H., P-562<br />

Kim, Y., P-71<br />

Kim, Y. S., P-562<br />

Kim, Y.-J., P-543<br />

Kimura, T., P-459<br />

Kingsland, C., P-288<br />

Kinloch, M., P-150<br />

Kirienko, K., P-523<br />

Kirkpatrick, G., P-538<br />

Kiskis, E., P-477<br />

Kissin, D., P-501<br />

Kissin, D. M., O-90, O-263, P-109, P-2<strong>21</strong>, P-236,<br />

P-326, P-633<br />

Kitajewski, J., P-710<br />

Kitasaka, H., O-64, P-618<br />

Kitaya, K., O-27, P-390, P-549, P-6<strong>17</strong>, P-683<br />

Kitchen, J., P-512<br />

Kjelland, M. E., P-404<br />

Klausen, C., P-150, P-701<br />

Klebanoff, M. A., P-34<br />

Klepacka, D., O-165<br />

Klepukov, A., P-663<br />

Klett, D., O-22<br />

Kligman, I., P-454, P-642, P-707<br />

Kline, M., O-189, P-424<br />

Klipstein, S., P-615<br />

Knight, A. N., P-470<br />

Knopman, J. M., O-245, P-442<br />

Ko, J.-J., P-365, P-572<br />

Kobayashi, H., P-599<br />

Kobayashi, J., P-232<br />

Kobayashi, M., O-240, P-42, P-232<br />

Kobayashi, R., P-563<br />

Kobayashi, T., P-391, P-532<br />

Kobori, Y., P-391, P-408, P-481, P-532, P-533<br />

Kocbulut, E., P-639<br />

Kocer, A., P-123<br />

Koch, L., P-452<br />

Kofinas, J., P-77, P-567<br />

Kofinas, J. D., P-82, P-578<br />

Kogan, P., P-547<br />

Kohan-Ghadr, H.-R., O-89, O-2<strong>21</strong><br />

Kohlmeier, A., P-440<br />

Koizumi, K., P-459<br />

Kojima, J., P-407<br />

Kojima, M., P-618<br />

Kokeguchi, S., P-24<br />

Kolb, B., O-34<br />

Kolp, L. A., P-468<br />

Komiyama, J., P-393<br />

Kondapalli, L. A., O-32, O-125, P-503<br />

Kono, T., P-599<br />

Konstantinidis, M., O-44, O-163, P-508, P-580,<br />

P-585<br />

Konuma, Y., P-618<br />

Kopelman, A., P-159<br />

Korrick, S. A., P-97<br />

Kort, J., P-496<br />

Kostelijk, E. H., O-55<br />

Kotelchuck, M., O-37, O-42, P-292<br />

Kounogi, S., O-64<br />

Koustas, G., P-662<br />

Kovacs, P., P-436<br />

Kovalak, E. E., P-229<br />

Kovalevsky, G., P-429, P-509<br />

Kovanci, E., P-239, P-322, P-361, P-565<br />

Kramer, W., P-737, P-739<br />

Kramer, Y. G., P-271<br />

Krawetz, S., O-201<br />

Krieg, A., P-86<br />

Krieg, S., P-86<br />

Krikun, G., O-<strong>17</strong>8, P-146<br />

Kriplani, A., P-689<br />

Krisher, R. L., O-56, P-260, P-269, P-329<br />

Krishnamoorthy, K., O-265<br />

Kroener, L., P-247, P-500<br />

Kuang, Y., O-33, P-16, P-647, P-690<br />

Kuchibhatla, M., P-80<br />

Kudesia, R., P-37, P-67, P-69<br />

Kuehl, T. J., P-357<br />

Kuji, N., P-407<br />

Kulkarni, A., P-326<br />

Kumar, A., P-445, P-678<br />

Kumar, D., P-299<br />

Kumar, N., O-168, O-196, P-14, P-310, P-497<br />

Kumasawa, K., P-459<br />

Kumbak, B. A., O-1<strong>17</strong><br />

Kuo, P.-L., P-188<br />

Kuokkanen, S., O-18<br />

Kuramoto, T., P-610<br />

Kuroda, S., P-459<br />

Kus, C., O-154<br />

Kushnir, V. A., O-151, P-253, P-428, P-439,<br />

P-592, P-622<br />

Kusunoki, H., P-222, P-388<br />

Kuwahara, A., P-312<br />

Kuwahara, M. K., P-372<br />

Kwak, S., P-267, P-344<br />

Kwak-Kim, J., O-157, P-128<br />

Kyono, K., P-57, P-536, P-575<br />

Kyrou, D., P-640<br />

Labarta, E., P-656<br />

Labella, P., O-123, O-231, P-489, P-567, P-587<br />

LaBrie, S., P-402<br />

Lachgar, H., P-297<br />

Lagunov, A., O-190<br />

Lai, T.-H., P-189<br />

Laknaur, A., O-75, O-185<br />

Lamaita, R. M., P-148<br />

Lamb, D. J., O-208<br />

Lambalk, C. B., O-55<br />

Landis, J., O-31, O-135, P-269<br />

Landis, J. N., O-137, P-511<br />

Lange, A., O-223, P-35<br />

Lanham, M., P-720, P-736<br />

Lannon, B. M., P-291<br />

Lapensee, L., O-132, P-225, P-309<br />

Large, M., P-507<br />

Larreategui, Z., P-314<br />

Larsen, C., P-237<br />

Lathi, R., O-86, P-496<br />

Laufer, M. R., P-162<br />

Laughlin-Tommaso, S. K., P-137<br />

Lavery, S., O-250<br />

Lavolpe, M., P-515<br />

Lawson, A., P-733, P-736<br />

Lazarin, G. A., O-166, O-167, O-194<br />

Lazzaroni-Tealdi, E., P-592, P-622<br />

Leach, R., P-127<br />

Leboeuf, M., P-158<br />

Lebovic, D. I., P-346<br />

Le Bras, A., P-311<br />

Lechtenberg, L., P-445<br />

Lee, D., P-65, P-102, P-433<br />

Lee, D. M., O-130<br />

Lee, D. R., P-562<br />

Lee, D.-Y., P-154<br />

Lee, E., P-154, P-380<br />

Lee, F.-K., P-189<br />

Lee, H.-L., O-123, P-489<br />

Lee, H. S., P-671<br />

Lee, J., P-380, P-433<br />

Lee, J. A., O-80, O-112, O-1<strong>21</strong>, O-146, O-160,<br />

O-164, O-204, O-245, P-5, P-130, P-131,<br />

P-368, P-442, P-629, P-652, P-675, P-700,<br />

P-730<br />

Lee, J.-W., P-543<br />

Lee, K., P-245, P-259, P-658<br />

Lee, K.-A., P-365<br />

Lee, K. L., P-241, P-480<br />

Lee, K.-H., P-267, P-344<br />

Lee, M. S., O-150, P-666<br />

Lee, M. M., P-97<br />

Lee, M.-S., P-627<br />

Lee, S., P-71, P-149<br />

Lee, S.-K., P-543<br />

Lee, S.-Y., P-365<br />

Lee, T.-H., P-627<br />

Lee, W., P-433, P-572<br />

Lee, W. S., P-562<br />

Lee, Y. J., P-411<br />

Lefebvre, J., O-132<br />

Legro, R., O-92, O-258<br />

Lekovich, J., O-60, O-124, O-242, P-135, P-153,<br />

P-248, P-342, P-454, P-463, P-642, P-649,<br />

P-707<br />

Lemyre, M., P-158<br />

Leocata Nieto, F. A., P-571<br />

Leon, J., P-405<br />

Lerner, V., P-25<br />

Lessey, B. A., O-16, O-2<strong>17</strong><br />

FERTILITY & STERILITY Ò<br />

e369


Leung, P. C. K., P-3, P-701<br />

Leung, Y.-K., P-94<br />

Levens, E., P-287<br />

Levin, H. I., P-710<br />

Levin, K., P-738<br />

Levitas, E., O-237<br />

Levy, M., O-187<br />

Levy, M. J., P-382<br />

Lewis, C. M., O-<strong>17</strong>0<br />

Lewis, E. I., O-232, P-356<br />

Leza, A., O-<strong>17</strong>3, P-685<br />

Li, B., P-363<br />

Li, D., P-299, P-377<br />

Li, E., P-227<br />

Li, F., O-<strong>17</strong>7, O-<strong>17</strong>8, P-609<br />

Li, G., O-59<br />

Li, J., P-190, P-4<strong>17</strong><br />

Li, M., P-678<br />

Li, Q., P-364, P-367<br />

Li, S., O-203, P-550<br />

Li, T., P-181<br />

Li, X., O-49, O-130, O-222, P-<strong>21</strong>2<br />

Li, Y., P-90, P-582<br />

Liang, X., P-90, P-657<br />

Liang, X.-Y., P-457<br />

Libby, V., P-186, P-400, P-628<br />

Liberty, A., P-231<br />

Librach, C., O-238, P-200, P-443<br />

Librach, C. L., O-236, P-196, P-557, P-638<br />

Licciardi, F., O-163, P-310, P-587, P-603<br />

Liebermann, J., P-51, P-134, P-238, P-288, P-354,<br />

P-481<br />

Lieman, H., P-67, P-69, P-338<br />

Lien, Y. Ron., P-694<br />

Light, A., O-54<br />

Lim, J., P-238, P-482<br />

Lin, H., O-133<br />

Lin, J., O-33, P-155<br />

Lin, W., O-<strong>17</strong><br />

Lin, Y., P-406<br />

Linan, A., P-41<br />

Lindeman, M. R., P-512<br />

Lindheim, S. R., O-<strong>21</strong>4<br />

Lindsey, J. S., O-188, P-162<br />

Lindstrom, M. J., P-346<br />

Ling, Q.-D., P-189<br />

Lingaiah, S., P-55<br />

Link, M., P-46<br />

Linn, J. M., P-632<br />

Lira, J., P-376<br />

Lisonkova, S., P-150<br />

Liu, D., O-225<br />

Liu, E., O-35<br />

Liu, H., O-63, O-274, P-425, P-580<br />

Liu, H.-C., O-233, P-118, P-420<br />

Liu, H.-E., P-457<br />

Liu, J., O-33, P-27, P-79, P-265, P-406<br />

Liu, K., P-196<br />

Liu, L., P-181<br />

Liu, X., P-85, P-242<br />

Liu, Y., O-182, O-230, P-4<strong>17</strong><br />

Liu, Z., P-5<strong>21</strong><br />

Lo, I. P. Y., P-7<strong>21</strong><br />

Lo, K., O-236, O-238<br />

Lo, K. C., O-199, P-133<br />

Lo, P., P-429<br />

Lobel, A., P-248, P-454<br />

Lobo, R., O-189, P-266, P-424<br />

Logan, P., O-94<br />

Lonczak, A., P-494<br />

Londra, L., O-87<br />

Lorenzi, D., P-487, P-515<br />

Lo Turco, E. G., P-257, P-275, P-276, P-277,<br />

P-596<br />

Louie, K., P-108, P-244<br />

Loup, V., O-248<br />

Loy, G., P-104<br />

Lu, G., O-222, P-<strong>21</strong>2<br />

Lu, Q.-S., P-1<strong>21</strong><br />

Lu, Y., P-16<br />

Lu, Z., O-63, P-580<br />

Lubimkina, E., P-10<br />

Lue, J., O-185<br />

Luense, L. J., P-556<br />

Lujan, M. E., P-68<br />

Lujan, S., P-528<br />

Lukaszuk, A., P-36<br />

Lukaszuk, K., P-36, P-40<br />

Luke, B., O-37, O-39, O-42, O-115, O-241, O-268,<br />

P-292, P-298, P-335, P-452<br />

Lukes, A. S., O-181<br />

Lukkari-Lax, E., O-100<br />

Luna, M., O-146, P-5, P-629, P-652, P-675<br />

Lunenfeld, E., O-237<br />

Lv, P., P-75, P-100, P-4<strong>17</strong><br />

Lvov, E., O-138<br />

Lynch, C. D., O-<strong>21</strong>9, P-34<br />

Lynch, D.-M., O-256<br />

Lynch, K. A., P-402<br />

Lynne, C., O-19<br />

Lynne, C. M., O-24<br />

Lysiak, J., O-23<br />

Lyu, Q., P-16, P-608, P-690<br />

Ma, L., P-27, P-265<br />

Ma, S., P-244, P-538<br />

Maas, K., P-319<br />

Machtinger, R., O-136<br />

Mack, W., P-88<br />

Mackenzie, A., P-270<br />

Mackenzie, A. C. L., O-161<br />

Madjid, T., P-<strong>17</strong>6<br />

Madrigal, V., O-49, O-53<br />

Madugu, H. N., P-4<br />

Maghen, L., O-236, O-238<br />

Magno, E., P-238<br />

Mahalingaiah, S., P-384<br />

Mahey, R., P-689<br />

Mahmood, S. N., P-404<br />

Mahmoud, S., P-283<br />

Mahony, M. C., O-155<br />

Mahran, A., P-233<br />

Mahutte, N., P-297<br />

Mainigi, M. A., O-3, O-144, P-306, P-337<br />

Maisel, L., P-19<br />

Maisenbacher, M. K., O-<strong>21</strong>8<br />

Maisog, J., O-<strong>21</strong>9<br />

Maje, I. M., P-4<br />

Majzoub, A., O-111<br />

Mak, W., O-133, P-335<br />

Makarem, M. H., O-69, O-259, P-<strong>21</strong>1<br />

Makinen, S., P-2<strong>21</strong><br />

Malik, M., O-4, O-58, O-184, P-142<br />

Malik, S., P-92<br />

Malmsten, J., P-358<br />

Man, L., O-229<br />

Manau, D., P-427<br />

Mangal, R., P-239, P-322, P-361<br />

Manichaikul, A., O-141<br />

Mannara, J. I., P-698<br />

Manoharan, A., O-196<br />

Mansour, A., O-136<br />

Margolis, D. J., O-53<br />

Margolis, M., P-133<br />

Marin, C., P-597<br />

Marin, D., P-656<br />

Marinuzzi, J., P-148<br />

Marquis, K., P-86<br />

Marrs, R. P., P-495<br />

Marsh, E. E., O-77, O-269, P-138, P-143, P-423,<br />

P-733<br />

Marsidi, A. M., P-255<br />

Martikainen, H., P-2<strong>21</strong><br />

Martin, J., P-367<br />

Martinez, A., P-15<br />

Martinez, C., P-577<br />

Martinez, E., P-41<br />

Martinez, M., P-644<br />

Martinez, S., P-597<br />

Martins, M. V., P-379<br />

Martins, W., P-697<br />

Martins, W. P., P-654<br />

Martins Da Silva, S. J., P-570<br />

Martins-Filho, O. A., P-148<br />

Mas, A., O-73, O-76, O-186, O-<strong>21</strong>1<br />

Maslow, B.-S., P-479<br />

Maslow, B.-S. L., O-122, P-624<br />

Massaro, F. C., O-131, O-270<br />

Massin, N., P-418<br />

Mata, A., P-414, P-415<br />

Mataro, D., O-153<br />

Mather, K. J., O-52<br />

Mathews, J., P-51<br />

Matsubayashi, H., O-27, P-390, P-549, P-6<strong>17</strong>,<br />

P-683<br />

Matsukawa, N., P-536<br />

Matsumoto, H., P-563<br />

Matsumoto, Y., P-24<br />

Matsuura, K., P-393<br />

Matsuzaki, T., P-459<br />

Matyas, R. A., O-139<br />

Mauri, A. L., O-131, O-270<br />

Mauricio, R., O-12<br />

Maurya, P. K., P-166<br />

Maxwell, R., O-<strong>21</strong>4<br />

Mayordomo, E., P-449<br />

Mazetto, R., P-261, P-351<br />

Mazur, E., P-322, P-361<br />

Mazur, E. C., P-239<br />

McAvey, B., O-18, P-37<br />

McCaffrey, C., O-91, P-478, P-524, P-603, P-680<br />

McCalla, S., P-709<br />

McCallie, B., O-48, P-589, P-676<br />

McCarthy, K., O-102<br />

McCormick, S., O-<strong>17</strong>4, P-260<br />

McCoy, R., P-519<br />

McCulloh, D. H., O-104, O-231, P-82, P-271,<br />

P-478, P-493, P-504, P-510, P-524, P-525,<br />

P-578, P-587, P-603, P-664, P-680<br />

McCullough, A., P-540<br />

McGarvey, M., P-289<br />

McGee, E., P-83<br />

McGinnis, L. A., P-274<br />

McIntire, D. D., P-400<br />

McKinnon, C. J., P-22<br />

McLean, M., P-220<br />

McManaman, J. L., O-272<br />

McQueen, D., P-704<br />

e370 Author Index Vol. 104, No. 3, Supplement, September <strong>2015</strong>


McReynolds, S., O-165, P-289<br />

McWilliams, K., P-5<strong>17</strong>, P-518<br />

McWilliams, T. K., P-512, P-5<strong>17</strong>, P-518<br />

Medrano, L., O-128, O-<strong>17</strong>3, P-685<br />

Medvedovic, M., P-94<br />

Mehr, H., P-82<br />

Mehta, R., P-241<br />

Meintjes, M., P-241, P-245, P-259, P-480, P-658<br />

Melnick, A. P., O-126, P-39, P-342, P-348, P-420,<br />

P-643, P-655, P-682<br />

Mendiola, J., O-205<br />

Mendola, P., O-225<br />

Mendoza, G., P-138, P-143, P-733<br />

Meng, H., P-265<br />

Meng, Y., P-100<br />

Merchenthaler, I., O-161<br />

Merchenthaler, I. J., P-270<br />

Mergler, R. J., P-68<br />

Merhi, Z., O-276<br />

Merolla, A., P-464<br />

Merrion, K., O-<strong>21</strong>8<br />

Mersereau, J. E., P-729<br />

Mertens, A., P-448<br />

Mertens, A. C., P-450<br />

Mesaros, C., O-226<br />

Meseguer, M., O-45, P-256, P-258, P-314, P-485,<br />

P-520, P-593, P-600, P-616, P-673, P-679<br />

Messerlian, C., O-96<br />

Metwalley, A., P-233<br />

Metwalley, A. M., P-613<br />

Metzgar, T., P-526<br />

Metzgar, T. L., P-612, P-660<br />

Meyer, L. R., P-615<br />

Michau, A., P-311<br />

Michels, K. A., P-11<br />

Mifsud, A., P-485<br />

Mikhail, E., P-201<br />

Miki, T., P-226, P-324, P-369, P-388<br />

Mikkelsen, E. M., P-384<br />

Miller, B. T., P-8, P-345<br />

Miller, C. E., O-72, P-202, P-631<br />

Miller, M., P-279<br />

Miller, S. J., P-740<br />

Mills, B. B., P-711<br />

Milne, P. A., P-570<br />

Minguez-Alarcon, L., O-223, P-97<br />

Minjarez, D. A., O-125, P-503<br />

Minkoff, H., P-709<br />

Mio, Y., P-132, P-251<br />

Miorin, J., O-149, P-718<br />

Mirisol, R., P-718<br />

Mirisol, R. J., O-149<br />

Mironova, A., P-663<br />

Missmer, S. A., O-232, P-49, P-162, P-<strong>21</strong>9, P-340,<br />

P-343, P-356, P-469, P-669, P-703<br />

Mitchell, E., P-30, P-705, P-714<br />

Mitchell-Leef, D., P-237<br />

Miyake, T., P-459<br />

Miyata, A., P-391<br />

Miyauchi, O., P-300<br />

Mizoguchi, C., P-251<br />

Mizumoto, S., P-610<br />

Mizuno, S., P-563<br />

Mizuta, S., O-113, P-390, P-6<strong>17</strong><br />

Mneimneh, A., P-236<br />

Moazamian, A., P-123<br />

Moazamian, R., P-1<strong>17</strong><br />

Mocanu, E. V., P-650<br />

Mohamed, S. A., O-6, O-182<br />

Mohan, A., P-185<br />

Moley, K. H., O-15, P-78<br />

Molinari, E., P-278<br />

Molinaro, T., O-83, O-180, P-370, P-584<br />

Mollamahmutoglu, L., P-350, P-699<br />

Mollen, C., P-374<br />

Montalvo, N., P-397<br />

Montegriffo, E., O-100<br />

Monteleone, P. A., O-149, P-718<br />

Moon, J.-W., P-180<br />

Moon, K., P-4<strong>21</strong><br />

Moore, R., P-279<br />

Mor, A., P-264, P-709<br />

Moravek, M. B., P-736<br />

Morbeck, D., P-98, P-591<br />

Moreau, C., O-87<br />

Mori, R., P-56<br />

Moriguchi, H., O-116<br />

Morimoto, Y., P-56, P-285, P-563<br />

Morin-Papunen, L., P-55<br />

Morris, P. L., O-<strong>21</strong><br />

Morrison, L. S., P-509<br />

Morsy, H., P-613<br />

Moschini, R. A., P-684<br />

Moskovstev, S., P-557<br />

Mostisser, C., O-226<br />

Motan, T., P-696<br />

Motato, Y., P-673<br />

Motta, E. L. A., P-261, P-351<br />

Mouille, B., P-203<br />

Mounce, G., P-570<br />

Moustafa, S., P-95, P-<strong>21</strong>6<br />

Moy, F., O-<strong>21</strong>0, P-458, P-462<br />

Mskhalaya, G., P-10<br />

Muasher, S. J., O-38, O-41, O-118, P-48, P-568<br />

Mucowski, S., P-88<br />

Muenke, M., P-15<br />

Muhsen-Alansarri, S. A., P-404<br />

Mukerji, B., P-280<br />

Mukherjee, T., O-245, P-5, P-130, P-131, P-368,<br />

P-652, P-675<br />

Mulla, Z., P-2<strong>17</strong><br />

Mullen, T., P-295<br />

Mullet, T., O-247, O-248, P-444<br />

Mullin, C., P-323, P-497<br />

Mullinax, S., P-86<br />

Mumford, S. L., O-1, O-139, O-140, O-191,<br />

O-220, P-6, P-11, P-18, P-30, P-3<strong>21</strong>, P-325,<br />

P-705, P-713, P-714<br />

Mumusoglu, S., O-67<br />

Munch, E. M., O-<strong>21</strong>6, P-659, P-668, P-716, P-7<strong>17</strong><br />

Muneyyirci-Delale, O., P-160, P-168<br />

Munkwitz, L., P-503<br />

Munne, S., O-34, O-35, O-36, O-63, O-91, O-193,<br />

P-14, P-490, P-495, P-497, P-498, P-580,<br />

P-585, P-677<br />

Munoz, E., O-45, P-593, P-673, P-679<br />

Munoz, M., P-12<br />

Murakami, M., P-610<br />

Murphy, E. M., P-39, P-348, P-420, P-643, P-655,<br />

P-682<br />

Murphy, M. J., O-101<br />

Murtadi, G., O-200, P-35<br />

Musul, B., O-<strong>21</strong>0<br />

Mutlu, I., P-623, P-635<br />

Mutlu, L., O-<strong>17</strong>7, O-<strong>17</strong>8, P-95<br />

Mutlu, M. F., O-68, P-623, P-635, P-712<br />

Muzzey, D., O-167<br />

Naaman, R. R., P-403<br />

Nachtigall, L. E., O-12<br />

Nachtigall, M., P-25<br />

Nachtigall, M. J., O-12<br />

Naeemi, F., O-20<br />

Naftolin, F., O-12, O-276<br />

Nagayoshi, M., P-222, P-226, P-324, P-369, P-388<br />

Nagras, Z. G., P-537<br />

Nagy, Z., O-34, P-237, P-632, P-711<br />

Nagy, Z. P., O-155<br />

Najafaliyeva, A., P-712<br />

Najari, B. B., O-109<br />

Najdecki, R., P-640<br />

Nakajima, S. T., P-26<br />

Nakajo, Y., P-536<br />

Nakamura, H., P-459<br />

Nakamura, Y., O-192, P-320, P-536, P-561<br />

Nakaoka, Y., P-285<br />

Nakayama, K., O-64, P-618<br />

Nakhuda, G., P-677<br />

Nangia, A., P-335<br />

Naqvi, H., O-<strong>17</strong>8<br />

Nasr, A., P-54<br />

Nassar, J., O-159<br />

Nasseri, A., P-499<br />

Nassr, A. A., O-69, O-259, P-<strong>21</strong>1<br />

Nastri, C. O., P-697<br />

Natan, Y., O-62<br />

Natarajan, L., O-7, P-2<br />

Nath, N. M., P-583<br />

Naumann, P. W., P-547<br />

Naumova, A., P-663<br />

Navarrete, G., P-259, P-480, P-658<br />

Navarrete, G. R., P-241, P-245<br />

Navarro, P. A., P-110, P-654, P-697<br />

Nayar, K. D., P-185<br />

Nazem, T. G., O-198<br />

Needleman, D., P-356<br />

Nehir Aytan, A., O-51<br />

Neisani Samani, E., O-<strong>17</strong>7, O-<strong>17</strong>8<br />

Neithardt, A. B., P-509<br />

Neitzel, D., P-295, P-307<br />

Nelson, A., P-371<br />

Nelson, S. M., P-28<br />

Neri, Q. V., O-2, O-110, P-122, P-378, P-386,<br />

P-546, P-560, P-573<br />

Neubauer, B. R., P-115, P-<strong>21</strong>5<br />

Newell, J., P-141<br />

Ng, N. W., O-188, P-162<br />

NICHD’s Reproductive Med Network, O-255<br />

Niederberger, C., P-408, P-481, P-528, P-533<br />

Nieman, L., O-184<br />

Nieman, L. K., O-4<br />

Nimeh, T., P-528<br />

Nishiyama, R., O-27, P-390, P-549, P-6<strong>17</strong>, P-683<br />

Niu, W., O-59<br />

Noble, L. S., P-2<strong>17</strong><br />

Noblia, F., P-515<br />

Nodar, F., P-487, P-515<br />

Noel, M., P-29<br />

Nogales, M., P-41<br />

North, J., O-181<br />

Norton, H., P-332<br />

Norwitz, E., O-243<br />

Notarstefano, V., O-<strong>17</strong>6, P-151<br />

Novoa Sanchez, Y., P-651<br />

Noyes, N., O-104, O-123, O-168, O-231, P-489,<br />

P-567, P-578, P-587<br />

Nulsen, J., O-122, P-255, P-479, P-624<br />

Nunes, F. F., P-148<br />

FERTILITY & STERILITY Ò<br />

e371


Obata, R., P-57, P-575<br />

O’Brien, J. E., O-20, O-147<br />

Ocali, O., P-569<br />

Odenwald, K., P-346<br />

O’Gradney, S., P-<strong>21</strong>0<br />

Oguz, Y., O-68<br />

Oh, S., P-187<br />

Ohgaki, A., P-563<br />

Ohl, D. A., O-28, O-<strong>17</strong>2, P-537<br />

Ohno, H., O-64, P-579<br />

Ojeda, M., P-593, P-673, P-679<br />

Oka, N., P-57, P-575<br />

Okada, H., P-391, P-532<br />

Okada, K., O-<strong>21</strong><br />

Okada, M., P-<strong>21</strong>, P-144<br />

Okitsu, O., P-393<br />

Oktay, K., P-458<br />

Oktay, K. H., O-<strong>17</strong>, O-<strong>21</strong>0, P-230, P-268, P-462,<br />

P-465, P-471<br />

Oktem, M., O-68, P-635, P-712<br />

Oktem, O., O-162, P-33, P-475, P-634<br />

Okubo, T., P-300<br />

Okuda, T., P-57, P-575<br />

Okuyama, N., P-57, P-536, P-575<br />

Olcha, M., O-134, P-586<br />

Olds, S., P-6<strong>21</strong><br />

O’Leary, T., P-238<br />

Oleinki, T. D., P-257<br />

Oliana, O., O-250<br />

Oliveira, J. A., O-131, O-270<br />

Oliveira, R., P-<strong>17</strong>9<br />

Oliveira-Pelegrin, G. R., O-131, O-270<br />

Olson, M., P-127<br />

Omil-Lima, D., O-234<br />

Omran, M. S., P-484, P-581, P-637<br />

Onalan, G., P-304<br />

Ono, M., P-262<br />

Ophir, L., P-87<br />

Oppenheimer, A., O-159<br />

Ord, T. S., O-3, O-144<br />

Ori, A., P-590<br />

Oron, G., P-263<br />

Ortiz Maffei, N., P-636<br />

Orvieto, R., O-237, P-87, P-89<br />

Orwig, K. E., P-92<br />

Osborne, S. E., O-13<br />

Osina, E., P-667<br />

Othman, E., P-233<br />

Othman, E. R., P-613<br />

Ouhilal, S., P-297<br />

Ouyang, Y., O-222, P-<strong>21</strong>2<br />

Owen, C. M., P-328<br />

Ozcan, P., P-229, P-437<br />

Ozgur, K., P-435, P-653<br />

Ozimek, J. A., P-316<br />

Ozler, S., P-52, P-60, P-62, P-63<br />

Ozmen, B., P-639<br />

Ozsurmeli, M., O-127<br />

Oztas, E., P-52, P-60, P-62, P-63<br />

Pabon, D., P-485<br />

Pabuccu, E. G., P-639<br />

Pacheco, A., P-12<br />

Pacheco, F., P-230<br />

Paczkowski, M., P-357<br />

Padovani, G. P., O-26<br />

Padte, K., P-163<br />

Page, D., P-15<br />

Pagou, E., P-640<br />

Paik, K. G., O-<strong>21</strong>8<br />

Pak, K.-A., P-543<br />

Pakhalchuk, T., P-523<br />

Pal, L., O-<strong>21</strong>3, P-95<br />

Palermo, G. D., O-2, O-110, P-122, P-153, P-378,<br />

P-386, P-546, P-560, P-573<br />

Palhares, M. B., P-654<br />

Palmerola, K. L., P-266, P-472<br />

Palumbo, A., P-9<br />

Pan, Y., P-323<br />

Paniza, T., O-110, P-122, P-546, P-573<br />

Papadakis, M., P-332<br />

Papanikolaou, E. G., P-640<br />

Papier, S., O-36, P-487, P-515, P-636<br />

Papouchis, N., P-738<br />

Parasar, P., O-188, P-162<br />

Pardue, S. L., O-52<br />

Parfitt, D.-E., O-57, P-8, P-345, P-419<br />

Parikh, T. P., O-142<br />

Park, H., P-71, P-2<strong>17</strong><br />

Park, I., P-180, P-267, P-344<br />

Park, J., P-149, P-187, P-267<br />

Park, J. H., P-671<br />

Park, K.-S., P-365<br />

Park, L., P-122, P-546<br />

Park, M., P-572<br />

Park, R., O-26<br />

Park, S., P-65, P-102, P-187, P-392<br />

Park, S.-Y., P-543<br />

Parks, J. C., O-48, P-589, P-601, P-676<br />

Parry, J., P-197<br />

Parsons, S., O-112, O-164, O-204<br />

Pascale, C., P-310<br />

Pasch, L., P-725<br />

Pasqualotto, F. F., P-281<br />

Pasquier, M., P-418<br />

Pastore, L., O-141<br />

Pastuszak, A. W., O-208<br />

Patel, A., O-4, O-184, P-142<br />

Patel, B., O-30, P-412, P-548, P-551<br />

Patel, J. C., O-81, P-734<br />

Patounakis, G., O-61, O-187, P-287, P-328, P-382,<br />

P-394, P-399<br />

Patrizio, P., O-62, P-278<br />

Pattinaja, D. A. P. M., P-674<br />

Patton, A. L., O-48, P-589, P-676<br />

Paulson, R., P-339, P-726<br />

Payson, M., P-227<br />

Pearlstein, H., P-688<br />

Peavey, M., P-565, P-607<br />

Peck, J. D., O-<strong>17</strong>0, O-227, O-254<br />

Pedro, J., P-379<br />

Pellicer, A., O-262, P-449, P-644, P-656<br />

Pelts, E. J., P-51<br />

Peluso, C., P-<strong>17</strong>9<br />

Peluso, J. J., P-85<br />

Penarrubia, J., P-427<br />

Peng, B., P-150, P-155, P-701<br />

Penman, A. D., O-82<br />

Penrose, L., P-605, P-614<br />

Penzias, A., O-243, P-291, P-315, P-585<br />

Peralta, S., P-427<br />

Pereira, N., O-60, O-124, O-242, O-249, P-135,<br />

P-153, P-248, P-454, P-463, P-642, P-649,<br />

P-707<br />

Perez, M., O-45, P-593, P-679<br />

Perez-Albala, S., P-600<br />

Pericuesta, E., P-123<br />

Perkins, K., O-263, P-236<br />

Perkins, N. J., O-139, O-191, P-18, P-705<br />

Perlman, J., P-709<br />

Perreault-Micale, C., P-295, P-307<br />

Perretti, J., P-284<br />

Persaud, S., P-737, P-739<br />

Peters, K., P-92<br />

Petersen, C. G., O-131, O-270<br />

Petersen, P., O-149, P-718<br />

Petrini, A. C., O-124, P-642<br />

Petrov, D., P-519<br />

Petrozza, J. C., P-35<br />

Pettersen, B., O-<strong>21</strong>8<br />

Phan, J. D., O-49<br />

Phillips, S., O-132, P-252<br />

Picou, A., P-661<br />

Pier, B., P-279<br />

Pietin Vialle, C., P-418<br />

Piltonen, T., P-55<br />

Pinkas, H., O-237<br />

Pisarska, M. D., O-260, P-247, P-313, P-316,<br />

P-473<br />

Pliego, J. F., P-357<br />

Plociennik, L., P-36, P-40<br />

Plosker, S., P-201<br />

Plowden, T., O-78, O-140, O-191, O-220, P-6,<br />

P-3<strong>21</strong>, P-713<br />

Pohlmeier, A., O-102<br />

Poizat, C., P-283<br />

Polackwich, A. S., P-553<br />

Polhemus, A. M., P-1<strong>17</strong>, P-123<br />

Poli, M., P-590<br />

Pollard, J., O-18<br />

Polotsky, A. J., O-39, O-201, O-258, O-268,<br />

O-272, O-274, P-425<br />

Pomeroy, K. O., P-284, P-397, P-566, P-576<br />

Pons, J., O-47<br />

Pool, T. B., P-346<br />

Porcu Buisson, G., P-305<br />

Pospisil, C., O-56<br />

Post, M., P-499<br />

Pouliot, M., P-158<br />

Pourret, E., O-239<br />

Powell, M., P-220<br />

Prasad, N., P-401<br />

Pratap, N. K., P-<strong>21</strong>3<br />

Prates, R., O-163, P-508<br />

Price, T. M., P-26, P-80, P-137<br />

Prien, S., P-605, P-614<br />

Prins, G., P-481<br />

Prins, G. S., P-533<br />

Prizant, H., P-139<br />

Proctor, C., O-267<br />

Proctor, G., P-519<br />

Prokai, D., P-19<br />

Provost, M., O-38, O-41, O-118<br />

Provost, M. P., O-40, O-119, O-261<br />

Purandare, N. C., P-650<br />

Purcell, S., P-241, P-245, P-259, P-658<br />

Purcell, S. H., P-480<br />

Puscheck, E., O-143, P-26<br />

Puscheck, E. E., O-258, P-7, P-81, P-364, P-626<br />

Puurunen, J., P-55<br />

Pyle, A., P-278<br />

Qian, Y., P-4<strong>17</strong><br />

Qin, Y., P-241<br />

Quach, K., O-238, P-196<br />

Quaglieri, C., P-84<br />

Quea, G., P-331<br />

Quevedo, S., P-331<br />

Quinteiro Retamar, A., O-36<br />

e372 Author Index Vol. 104, No. 3, Supplement, September <strong>2015</strong>


Quistorff, J., P-458<br />

Rabinowitz, M., P-519<br />

Raburn, D. J., P-48<br />

Racicot, M.-H., P-252<br />

Racowsky, C., O-136, O-138, O-193, P-356,<br />

P-469, P-564, P-669, P-706<br />

Radhakrishnan, G., P-184<br />

Radin, R., P-18, P-30, P-140, P-705, P-714<br />

Rafizadeh, M., P-526<br />

Ragheb, A. M., P-389<br />

Rahil, T., P-402<br />

Rahmawati, E., P-166<br />

Raia, M., O-166<br />

Raju, R., P-81, P-626<br />

Rakhila, H., P-158<br />

Ramalingam, N., P-45<br />

Ramer, I., O-158, P-125, P-126, P-682<br />

Ramirez, D., P-2<strong>17</strong><br />

Ramirez, E., O-53<br />

Ramirez, L. B., O-138, P-564<br />

Ramos, B., O-128, O-<strong>17</strong>3, P-685<br />

Ramos, L., P-316<br />

Ramos, L. G., P-313<br />

Rangarajan, S., P-200<br />

Rappolee, D. A., P-364<br />

Rarosi, F., P-436<br />

Rasheed, S. M., P-233<br />

Rauch, E., O-145<br />

Ravichandran, K., O-44, P-585<br />

Rebbeck, T. R., P-447<br />

Reda, C. V., P-526, P-612, P-660<br />

Reed, B., P-<strong>21</strong>8, P-400<br />

Reed, B. G., P-186, P-628<br />

Reed, S., P-294<br />

Reid, A., O-88<br />

Reid, M., P-78<br />

Reis, H. S., P-730<br />

Reisman, L., O-110, P-122, P-546, P-573<br />

Remohi, J., P-256, P-258, P-314, P-485, P-600,<br />

P-656<br />

Ren, H., P-538<br />

Renzi, A., O-131, O-270<br />

Requena, A., P-12<br />

Resetkova, N., P-291<br />

Revich, B., P-97<br />

Rey Valzacchi, G. J., P-571, P-698<br />

Rheaume, C., P-158<br />

Ribeiro de Andrade, M., P-539<br />

Riboldi, M., P-597<br />

Ribustello, L., O-34, O-35, P-677<br />

Ricard, J., P-397<br />

Riche, D., P-197<br />

Rich-Edwards, J. W., P-703<br />

Richter, K. S., O-20, O-147, P-287, P-3<strong>21</strong>, P-394,<br />

P-399, P-4<strong>21</strong>, P-482<br />

Riddle, M. P., O-108<br />

Ridha-Albarzanchi, M. T., P-404<br />

Rienzi, L., P-520, P-630<br />

Riley, J. K., P-43, P-293<br />

Riley, K., P-141<br />

Rinaudo, P., P-242<br />

Rippon, G., O-<strong>17</strong>2<br />

Risteli, J., P-55<br />

Robertshaw, I., P-708<br />

Robinson, M. K., P-49<br />

Robinson, R., P-432<br />

Robinson, T. S., P-1<br />

Robledo, C., O-274, P-425<br />

Roby, K., P-86<br />

Rocafort, E., O-128, O-<strong>17</strong>3, P-685<br />

Rodolosse, A., O-47<br />

Rodosthenous, R., O-136<br />

Rodrigo, L., O-148<br />

Rodrigues, D., O-149, P-718<br />

Rodriguez, A., O-153, O-195<br />

Rodriguez, S., O-168, P-14, P-310<br />

Rodriguez-Iglesias, B., P-367<br />

Rodriguez Kubrusli, M. A., P-698<br />

Rodriguez-Purata, J., O-80, O-1<strong>21</strong>, O-146, O-245,<br />

P-5, P-131, P-629, P-652, P-665, P-675,<br />

P-684, P-700<br />

Roeca, C., P-706<br />

Rogel, S., O-128<br />

Rogers, K., P-725<br />

Romany, L., P-600<br />

Romero, A., P-644<br />

Romero, J., P-485<br />

Rooney, K., O-106<br />

Rosario, M., P-514, P-522<br />

Rose, S. D., P-96<br />

Rosen, A. M., O-44<br />

Rosen, J., O-94<br />

Rosen, K., P-371<br />

Rosenberg, L., P-140<br />

Rosenblad, A., P-224<br />

Rosenbluth, E., O-246<br />

Rosenwaks, Z., O-2, O-60, O-110, O-124, O-126,<br />

O-229, O-242, O-249, P-39, P-122, P-135,<br />

P-153, P-248, P-348, P-358, P-378, P-386,<br />

P-420, P-454, P-463, P-492, P-546, P-560,<br />

P-573, P-642, P-643, P-649, P-655, P-682,<br />

P-707<br />

Roshdy, S., P-613<br />

Ross, L., P-108, P-453<br />

Roth, E. W., O-269<br />

Rothman, K. J., P-<strong>17</strong>, P-22, P-384<br />

Rotker, K. L., O-234<br />

Rotoli, D., P-9<br />

Roudebush, W., P-711<br />

Roudebush, W. E., P-602<br />

Rowen, T., P-735<br />

Royster IV, G. D., P-330, P-341<br />

Rozenchan, P., P-159<br />

Rubin, R. S., O-20<br />

Rubio, C., O-148, P-597<br />

Rubio, J. M., P-644<br />

Rudick, B., P-108<br />

Rueter, M. A., P-727, P-728<br />

Ruiter-Ligeti, J. J., P-413<br />

Ruman, J., P-631<br />

Russell, D., P-514, P-522<br />

Ryan, A., P-519<br />

Ryan, E. A. J., P-638<br />

Ryan, G., O-<strong>21</strong>6, P-659, P-668, P-716, P-7<strong>17</strong><br />

Ryan, G. A., P-650<br />

Rydze, R. T., P-50<br />

Ryu, S., P-572<br />

Sabanegh, E., P-114, P-119, P-387<br />

Sabanegh, E. S., P-101, P-111, P-112, P-116,<br />

P-381, P-451, P-534, P-544, P-545, P-552,<br />

P-553, P-559<br />

Sabry, M., P-233, P-613<br />

Sachdev, N. M., P-478, P-510, P-525, P-664<br />

Sadeghi-Nejad, H., P-548<br />

Saed, G., P-282<br />

Saed, G. M., O-13, P-91, P-113, P-115, P-<strong>21</strong>5<br />

Saed, M. G., P-113, P-115, P-<strong>21</strong>5<br />

Saha, I., P-157<br />

Sahin Ersoy, G., P-195, P-430, P-431, P-434<br />

Saito, H., P-312<br />

Saitou, S., P-232<br />

Saji, F., P-320, P-561<br />

Sakashita, A., P-599<br />

Saketos, M., P-334<br />

Sakkas, D., O-243, P-291, P-301, P-315, P-319,<br />

P-569<br />

Salama, E. A., P-681<br />

Salazar Garcia, M. D., O-157, P-128<br />

Saldivar, S., P-2<strong>17</strong><br />

Salem, W. H., P-107<br />

Salemi, J., P-201<br />

Salihu, H., P-201<br />

Sallam, H. N., P-484, P-581, P-637<br />

Salmon, K., P-240<br />

Samanta, L., P-101, P-111, P-112, P-116, P-381,<br />

P-387, P-552, P-555, P-559<br />

Sammel, M. D., O-88, O-169, O-197, O-<strong>21</strong>2,<br />

P-447<br />

Samplaski, M. K., O-199<br />

Samstag, L., P-738<br />

San, A., P-367<br />

Sanchez, M., P-<strong>21</strong>0<br />

Sanchez Sarmiento, C., P-414, P-415<br />

Sandler, B., O-44, O-80, O-1<strong>21</strong>, O-146, O-245,<br />

P-5, P-130, P-131, P-629, P-652, P-665,<br />

P-675, P-684<br />

Sandlow, J., P-383, P-740<br />

Sandlow, J. I., O-<strong>17</strong>1<br />

Sandya, R., P-<strong>21</strong>3<br />

Sanfilippo, J. S., P-92<br />

Sankaranarayanan, K., O-<strong>17</strong>0<br />

Santamaria, X., O-235<br />

Santanam, N., P-<strong>17</strong>2<br />

Santoro, N., O-39, O-92, O-201, O-258, O-268<br />

Santos, R., P-405<br />

Santos, T. G., P-257<br />

Sanz, J., P-449<br />

Sapienza, C., P-306<br />

Sapra, K. J., O-114, O-<strong>21</strong>9<br />

Sarac, G. N., P-558<br />

Sargin Oruc, A., P-38<br />

Saribal, S., P-250<br />

Sasaki, C., P-536<br />

Sasaki, K., O-72, P-202<br />

Sasamine, K., P-390<br />

Sato, Y., P-536<br />

Sauer, M. V., O-189, P-266, P-424, P-472<br />

Sawant, R., P-163<br />

Sbracia, M., P-152<br />

Scalici, E., O-247, O-248, P-444<br />

Scarpellini, F., P-152<br />

Schanne, A., P-445<br />

Schattman, G. L., O-124, O-229, P-463, P-643,<br />

P-655<br />

Scheib, J. E., P-727, P-728<br />

Schenk, L., P-239, P-322, P-361<br />

Schenkman, E., P-323<br />

Schickler, R., P-201<br />

Schiewe, M. C., P-507<br />

Schisterman, E., O-139, O-140, O-191, O-<strong>21</strong>9,<br />

O-220, P-6, P-18, P-30, P-705, P-713, P-714<br />

Schlenker, T., P-260<br />

Schliep, K. C., P-30<br />

Schmelter, T., P-371<br />

Schneider, R., P-422<br />

Schneiderman, A., O-<strong>17</strong>4<br />

Schon, S. B., P-556<br />

FERTILITY & STERILITY Ò<br />

e373


Schoolcraft, W. B., O-32, O-46, O-48, O-56,<br />

O-125, O-165, O-<strong>17</strong>4, O-190, P-234, P-260,<br />

P-289, P-329, P-438, P-503, P-589, P-601,<br />

P-676<br />

Schor, E., P-159<br />

Schorsch, K., P-615<br />

Schoyer, K. D., O-264, P-740<br />

Schreiber, C. A., P-374<br />

Schrepferman, C. G., P-542<br />

Schufreider, A., P-704<br />

Schulte, M. B., P-43<br />

Schutt, A., P-50<br />

Schweitz, M., O-165, P-289<br />

Schweitzer, A., O-47<br />

Schymura, M. J., P-452<br />

Scoccia, H., O-14<br />

Scott, R. T., O-31, O-134, O-135, O-137, O-180,<br />

O-244, P-223, P-269, P-370, P-494, P-505,<br />

P-511, P-526, P-527, P-584, P-586, P-594,<br />

P-595, P-606, P-612, P-660, P-715<br />

Sdrigotti, A., P-571, P-681<br />

Seal, P., P-287<br />

Seaman, E., P-529<br />

Segal, T., P-325<br />

Segars, J., O-58, P-26, P-84<br />

Seifer, B., P-146<br />

Seifer, D. B., O-275<br />

Sekhon, L., P-130<br />

Selter, J. H., P-468<br />

Seminsky, M., O-<strong>21</strong>5<br />

Senapati, S., O-88, O-<strong>21</strong>2, P-306<br />

Senbabaoglu, F., O-162<br />

Sengoba, K., P-138, P-143, P-733<br />

Sengul, O., O-67<br />

Seong, S., P-<strong>17</strong>8<br />

Serafini, P. C., P-351<br />

Sereni, E., O-<strong>17</strong>6, P-103, P-151<br />

Sergeyev, O., P-97<br />

Session, D. R., P-633<br />

Seta, N., O-231, P-489, P-567, P-587<br />

Setti, A. S., P-275, P-276, P-277, P-281<br />

Setton, R., O-126, P-39<br />

Seungdamrong, A. M., O-92<br />

Sevillano, G., P-485<br />

Seybold, D. J., P-359<br />

Seyhan Ata, A., P-33, P-634<br />

Seymen, M. C., P-558<br />

Sha, H., O-43<br />

Shaaban, A., P-347<br />

Shaaban, O. M., O-228, P-161<br />

Shaeib, F., O-143, P-282<br />

Shah, D. K., P-<strong>21</strong>9<br />

Shah, M., O-86, P-496<br />

Shah, T., O-28, O-30, P-537, P-551<br />

Shahar, A., O-62<br />

Shalaby, S. M., O-6, O-74, O-182, O-<strong>21</strong>1<br />

Shalamova, E., P-663<br />

Shamonki, M., P-678<br />

Shandley, L. M., P-448<br />

Shapiro, B. S., P-183, P-192, P-327, P-349<br />

Shapiro, D., P-631<br />

Shapiro, D. B., P-632<br />

Shapiro, M., P-408<br />

Sharara, F., P-107, P-286, P-325, P-352, P-403,<br />

P-620<br />

Sharara, F. I., P-3<strong>17</strong><br />

Sharara, N. F., P-403<br />

Sharma, P., P-196, P-200, P-443, P-557<br />

Sharma, R., P-101, P-111, P-112, P-114, P-116,<br />

P-119, P-381, P-387, P-451, P-534, P-544,<br />

P-545, P-552, P-555, P-559<br />

Sharma, S., P-157<br />

Sharpe, A. N., P-513<br />

Shaulov, T., O-85<br />

Shavell, V., P-127<br />

Shavit, T., P-263, P-302<br />

Shawber, C., P-710<br />

Shazly, S. A., O-259<br />

Shazly, S. A. M., O-69<br />

Sheehan, D. A., P-734<br />

Shelinbarger, C. L., P-614<br />

Shen, W., P-1<br />

Shen, Y., P-100<br />

Sheng, J., P-100<br />

Shenoy, C. C., P-591<br />

Shi, H. L., O-138<br />

Shibasaki, S., P-536<br />

Shigeta, M., P-262<br />

Shim, S., P-392<br />

Shimada, H., P-407<br />

Shimomura, M., O-64, P-618<br />

Shimomura, Y., P-391<br />

Shin, D., O-30, P-433, P-551<br />

Shin, J., P-71, P-380<br />

Shin, P. R., O-20<br />

Shin, S.-H., P-543<br />

Shin, T., P-391, P-532<br />

Shipley, S., O-20<br />

Shlush, E., O-236, O-238, P-557<br />

Shoupe, D., P-88<br />

Shraga, R., O-168, O-196, P-14, P-310<br />

Shu, Y., P-19<br />

Shwayder, J. M., P-197<br />

Siegersma, K., O-269<br />

Sifer, C., P-460, P-476<br />

Sigman, M., O-234, P-199<br />

Sigurjonsson, S., O-<strong>21</strong>8<br />

Silber, S. J., P-15<br />

Silva, E., P-269, P-329<br />

Silva, L. R., P-697<br />

Silva-filho, A. L., P-148, P-376<br />

Silver, R. M., O-139, O-140, O-191, O-220, P-6,<br />

P-18, P-705, P-713, P-714<br />

Silverberg, K., P-661<br />

Simbulan, R., P-242<br />

Simon, A., P-519<br />

Simon, C., O-73, O-148, O-186, O-235, P-367,<br />

P-597<br />

Sinaii, N., P-160, P-168<br />

Singer, A., O-88<br />

Singer, T., P-497<br />

Singh, S., P-243<br />

Singh, T., P-280<br />

Sisti, G., O-158, P-125, P-126<br />

Sjaarda, L., O-191, P-6, P-11, P-18, P-705, P-713,<br />

P-714<br />

Sjoblom, C., P-662<br />

Skanes-DeVold, H. D., O-<strong>21</strong>4<br />

Skaznik-Wikiel, M. E., O-272<br />

Skrip, L., P-95<br />

Slayden, O. D., O-95, O-97, O-99, P-129<br />

Slizewski, D., O-2<strong>17</strong><br />

Slutsky, J., P-737, P-739<br />

Smarr, M. M., O-225<br />

Smigulina, L., P-97<br />

Smith, A. D., P-28<br />

Smith, H., P-662<br />

Smith, J. F., O-94, P-735<br />

Smith, K., P-46, P-541<br />

Smith, M., P-490<br />

Smith, R., O-23, P-260<br />

Smith-Harrison, L., O-23<br />

Smotrich, D., O-84<br />

Soh, S., P-391<br />

Soh, Y., O-152<br />

Somova, O., P-396<br />

Son, W.-Y., P-263<br />

Song, S.-H., P-392<br />

Song, Y., P-156, P-<strong>17</strong>7<br />

Sonigo, C., P-460, P-476<br />

Sonksen, J., O-28, P-537<br />

Sonmezer, M., P-471, P-639<br />

Sonohara, M., P-579<br />

Sood, A., O-107, O-<strong>21</strong>3, P-<strong>21</strong>6<br />

Sota, N., P-485<br />

Souter, I., O-200, P-35<br />

Spaanderman, M. E. A., P-674<br />

Spaine, D. M., P-539<br />

Spandorfer, S. D., O-126, O-158, O-242, P-39,<br />

P-125, P-126, P-342, P-682<br />

Spath, K., O-252, P-590<br />

Spector, L. G., O-241, P-298, P-452<br />

Spencer, J., P-448<br />

Spencer, J. B., P-450<br />

Spiessens, C., P-136<br />

Sroga, J. M., P-708<br />

Srougi, M., O-26, P-535<br />

Srougi, V., P-535<br />

Srouji, S., P-469<br />

Srouji, S. S., P-669<br />

Stalzer, A., P-359<br />

Stanczyk, F. Z., P-80, P-88, P-453<br />

Stanford, J., O-191, P-22<br />

Stanhiser, J., P-203<br />

Stapleton, G., P-104<br />

Stecher, V., O-<strong>17</strong>2<br />

Steele, H., P-737, P-739<br />

Steele, M., P-737, P-739<br />

Stegmann, B. J., P-308, P-646<br />

Stein, D. E., O-1<strong>21</strong><br />

Steinberg, M. L., O-269<br />

Steiner, A., O-92, O-224, P-693<br />

Steiner, A. Z., P-231<br />

Steinkampf, M. P., P-366<br />

Steller, C., O-72, P-202<br />

Stelling, J., O-193, P-334<br />

Stemm, K., P-6<strong>21</strong><br />

Stern, J. E., O-37, O-42, O-115, P-292, P-335,<br />

P-452<br />

Stevens, J. M., O-56, O-125, O-<strong>17</strong>4, P-289, P-329,<br />

P-438, P-601<br />

Stewart, E., O-107<br />

Stewart, E. A., O-181, P-137, P-591<br />

Stewart, J., P-248, P-342, P-454<br />

Stillman, R., P-399<br />

Stillman, R. J., O-147<br />

St. Marie, P., P-402<br />

St-Michel, P., P-297<br />

Stobezki, R., O-<strong>21</strong>0<br />

Stocco, C., O-14<br />

Stoppa, M., P-520<br />

Stouffer, R. L., O-101<br />

Strachan, G., O-102<br />

Strashnova, A., P-523<br />

Stratton, P., P-160, P-168<br />

Straub, R. J., P-237<br />

e374 Author Index Vol. 104, No. 3, Supplement, September <strong>2015</strong>


Strawn, E. Y., O-264, P-740<br />

Strobino, D., O-87<br />

Strug, M., P-127<br />

Strumbly, D., P-412<br />

Strynar, M., O-224<br />

Styer, A. K., O-39, O-255, O-258, O-268<br />

Su, I., O-7, P-2, P-470<br />

Su, Y., O-59<br />

Su, Y.-T., P-719<br />

Such, E., P-449<br />

Sudhakaran, S., P-464<br />

Sueldo, C., P-296<br />

Sueldo, C. M., P-85<br />

Sugino, N., P-<strong>21</strong>, P-144<br />

Sugishima, M., P-132, P-251<br />

Suh, C., P-380<br />

Sukhwani, M., P-92<br />

Sullivan-Pyke, C. S., P-710<br />

Sulo, S., O-72<br />

Sultan Ahamed, A. M., P-235<br />

Summers, K. M., O-<strong>21</strong>6, P-659, P-668, P-716,<br />

P-7<strong>17</strong><br />

Summers-Colquitt, R. B., P-346<br />

Sun, H., P-267, P-344<br />

Sun, L., O-225<br />

Sun, Y., O-59, P-3, P-670<br />

Sundaram, R., O-154, O-<strong>21</strong>9, O-225<br />

Sunderam, S., P-109<br />

Sung, L., P-334<br />

Sung, N., O-157, P-128<br />

Surrey, E., O-125<br />

Surrey, M., P-240, P-473, P-490, P-495<br />

Sutcliffe, A. G., O-93<br />

Suzuki, K., P-391, P-407<br />

Suzuki, R., P-232<br />

Swain, J. E., O-56, O-190, P-234, P-346, P-438,<br />

P-604<br />

Swan, S. H., O-205<br />

Swanson, K., O-79<br />

Swanson, S., P-557<br />

Swerdlow, A. J., O-93<br />

Sylvestre, C., O-132<br />

Sylvestre-Margolis, G., O-145<br />

Szlit Feldman, E., P-681, P-698<br />

Taboas, E., O-45, P-593, P-679<br />

Tada, Y., O-192, P-320, P-561<br />

Tadros, T., P-311<br />

Taguchi, S., O-192, P-320, P-561<br />

Taiyeb, A. M., P-404<br />

Takahashi, C., P-6<strong>17</strong><br />

Takahashi, M., P-536<br />

Takano, T., P-320, P-561<br />

Takaya, Y., O-27, P-549, P-683<br />

Take Kaplanoglu, G., P-558<br />

Takemoto, Y., P-222<br />

Takeuchi, K., P-246<br />

Takeuchi, M., O-64<br />

Takeuchi, T., O-113, P-57, P-536, P-575<br />

Takiuchi, T., P-459<br />

Tal, O., O-275<br />

Tal, R., O-275, P-709<br />

Talebian, S., P-442<br />

Talevi, R., P-464<br />

Taliadouros, G. S., P-410<br />

Tam, M., P-722, P-723, P-731, P-732<br />

Tamura, F., P-618<br />

Tamura, H., P-<strong>21</strong>, P-144<br />

Tan, L., P-134<br />

Tan, O., P-400<br />

Tan, S.-J., P-455, P-461<br />

Tanacan, A., P-441<br />

Tanaka, A., P-222, P-226, P-324, P-369, P-388,<br />

P-407<br />

Tanaka, I., P-226, P-324, P-369, P-388<br />

Tanaka, K., P-610<br />

Tang, B., O-102<br />

Tang, Y., O-233, P-118<br />

Taniguchi, F., P-<strong>17</strong>0<br />

Tanrikut, C., O-205, O-207, O-209, O-271<br />

Tao, X., O-31, O-135, O-137, P-269, P-494, P-505,<br />

P-511, P-526, P-586<br />

Tapanainen, J., P-2<strong>21</strong><br />

Tapanainen, J. S., P-55<br />

Tarozzi, N., O-252<br />

Tatpati, L. L., O-8<br />

Tavares, A., O-26<br />

Taylor, D., O-5<br />

Taylor, H. S., O-<strong>17</strong>7, O-<strong>17</strong>8, P-146<br />

Taylor, L., O-88<br />

Taylor, R. N., P-19<br />

Taylor, T. H., P-491, P-516<br />

Tecson, V., O-163<br />

Tejera, A., P-520<br />

Tepper, A. B., O-83<br />

Terakawa, N., P-<strong>17</strong>0<br />

Testillano, M., P-616<br />

Thakore, S., P-412<br />

Thakur, M., O-143, P-7, P-81, P-626<br />

Thomas, A., O-53<br />

Thomas, M., P-259, P-658<br />

Thomas, M. A., P-94, P-708<br />

Thomas, M. R., P-241, P-245, P-480<br />

Thomas, S. M., O-40, O-119, O-261<br />

Thompson, A., O-15, P-78<br />

Thompson, K., P-526<br />

Thompson, W., O-73<br />

Thornton, K., O-273, P-47, P-315, P-319, P-338<br />

Thornton, K. L., P-340<br />

Tibaldi, D. S., P-539<br />

Tiegs, A. W., P-493<br />

Tiitinen, A., P-2<strong>21</strong><br />

Tilley, B., P-241, P-245, P-259, P-480, P-658<br />

Timur, H., P-23, P-38, P-61<br />

Tiseo, B. C., O-26, O-205, O-209, P-535<br />

Titus, S., O-<strong>17</strong>, O-<strong>21</strong>0<br />

Tjaden, B. L., O-8<br />

Tjoa, M., O-102<br />

Tokmak, A., P-52, P-60, P-62, P-63<br />

Tokoro, M., P-579<br />

Tokudome, M., P-246<br />

Toledo, A. A., P-237, P-632<br />

Toner, J., P-641<br />

Toner, J. P., P-326<br />

Torno, A., P-681, P-698<br />

Toth, T., O-150, P-703<br />

Toth, T. L., O-200, O-223<br />

Trabucco, E., P-630<br />

Tran, N. D., O-94<br />

Traxler, S. A., P-374<br />

Treff, N. R., O-31, O-134, O-135, O-137, P-269,<br />

P-370, P-494, P-505, P-511, P-526, P-527,<br />

P-586<br />

Trevino, L. S., O-76<br />

Trevisan, C. M., P-<strong>17</strong>9<br />

Trew, G., O-250<br />

Trofimenko, V., P-541<br />

Troshina, M., P-663<br />

Troup, S., P-354<br />

Truong, T. T., O-152<br />

Tsai, E.-M., P-147<br />

Tsai, S., P-43<br />

Tsuiki, M., P-579<br />

Tsuji, H., P-618<br />

Tsukamoto, K., P-232<br />

Tsuneto, M., P-251<br />

Tucker, M., P-238<br />

Tucker, M. J., O-147, P-482<br />

Tulandi, T., P-263, P-302<br />

Turan, V., P-230, P-462, P-465, P-471<br />

Turki, R., P-555<br />

Turki, R. F., P-387<br />

Turkkani, A., P-350<br />

Turner, T. G., P-661<br />

Tusheva, O. A., P-497<br />

Tvrda, E., P-451<br />

Tyden, T., P-224<br />

Tzeng, C.-R., P-166, P-455, P-461<br />

Ubaldi, F., P-520<br />

Ubaldi, F. M., P-630<br />

Uhler, M. L., P-481, P-704<br />

Ulker, K., O-1<strong>17</strong><br />

Ulug, M., P-645<br />

Umezawa, A., P-407<br />

Unal, S., O-1<strong>17</strong><br />

Uncu, G., P-167, P-250<br />

Underberger, C., P-240<br />

Upham, K. M., O-31, P-594, P-595, P-612<br />

Uquillas, K., P-25<br />

Urich, M., P-609<br />

Uriondo, H., P-487<br />

Urman, B., O-162, P-33, P-475, P-634<br />

Utsunomiya, T., P-407<br />

Uvarova, E., P-523<br />

Vaast, M., P-254<br />

Vaccari, L., O-<strong>17</strong>6, P-103, P-151<br />

Vagnini, L. D., O-131, O-270<br />

Vaiarelli, A., P-630<br />

Valbuena, D., P-597<br />

Valdes, C., P-565<br />

Valdes, C. T., P-50<br />

Vallejo, V., O-145<br />

van den Abbeel, E., O-55, P-249<br />

van Golde, R. J. T., P-674<br />

Vanijgul, C., P-239<br />

van Rijswijk, J., O-55<br />

Van Voorhis, B., P-<strong>21</strong>9, P-326<br />

VanWort, T. A., P-368<br />

Vargas, M. V., P-206, P-<strong>21</strong>4<br />

Vaskivuo, T. E., P-55<br />

Vassena, R., O-47, O-153, O-195<br />

Vaughan, D. A., O-243, P-426, P-574<br />

Vaughan, L. E., P-137<br />

Vega, M. G., P-428, P-439<br />

Vega, R. R., P-405<br />

Veleva, Z., P-2<strong>21</strong><br />

Velez Edwards, D. R., P-145<br />

Vergouw, C. G., O-55<br />

Verlinsky, O., P-523<br />

VerMilyea, M., P-337<br />

Vernaeve, V., O-153, O-195<br />

Verrecchio, E. S., P-688<br />

Victorino, A. B., P-596<br />

Vidal, C., O-148<br />

Vidali, A., O-151<br />

Vigneswaran, H. T., O-234<br />

Villamon, E., P-449<br />

Villeneuve, K., O-190<br />

FERTILITY & STERILITY Ò<br />

e375


Villette, C., P-418<br />

Viloria, T., O-45<br />

Vincens, C., O-247, P-182, P-254<br />

Vinning, T., P-538<br />

Vintejoux, E., P-444<br />

Vitek, W., O-39, O-258, O-268, P-377<br />

Vitiello, D., O-193<br />

Vitonis, A., P-<strong>21</strong>9<br />

Voong, C., O-197<br />

Voronich, N., P-523<br />

Vyas, N., P-313, P-316<br />

Wachs, D., O-246<br />

Wachter, K., P-6<strong>21</strong><br />

Wactawski-Wende, J., O-1<br />

Wagner Coughlin, C., P-354, P-495<br />

Wald, M., P-547<br />

Walker, C., O-73<br />

Walker, C. L., O-76<br />

Wallace, M. E., O-225<br />

Wang, B., P-70<br />

Wang, C., P-181<br />

Wang, E. T., O-260, P-247, P-313, P-316, P-473<br />

Wang, H., P-582<br />

Wang, N., P-647<br />

Wang, Q., P-592, P-622<br />

Wang, R., P-548<br />

Wang, S., P-514, P-522<br />

Wang, W., O-233, P-118<br />

Wang, X., O-84, P-44, P-457, P-556<br />

Wang, Y., P-586, P-647<br />

Wantman, E., O-115<br />

Ward, K., O-<strong>17</strong>5<br />

Warinner, C., O-<strong>17</strong>0<br />

Warne, D. W., P-336<br />

Warner, E., O-<strong>21</strong>5<br />

Warner, L., P-109<br />

Warty, N., P-163<br />

Watanabe, C., P-563<br />

Watanabe, H., P-232, P-618<br />

Watanabe, S., P-222, P-388<br />

Waud, K., P-191<br />

Weaver, A., P-137<br />

Webster, W., O-257<br />

Weckstein, L., O-246<br />

Weedin, E. A., O-227<br />

Weinerman, R. S., O-3, O-144, P-306<br />

Welch, C., P-498<br />

Welliver, C., P-540<br />

Wellons, M., P-145<br />

Wells, D., O-44, O-91, O-252, P-585, P-590<br />

Werland, H., P-661<br />

Werner, M. D., O-31, O-135, O-137, O-180,<br />

O-244, P-223, P-370, P-584, P-594, P-595,<br />

P-606, P-612, P-660, P-715<br />

Wesselink, A., P-22<br />

Wesselink, A. K., P-384<br />

Wessels, C. E., P-605, P-614<br />

Whalley, K., P-570<br />

Wheeler, K., O-23<br />

Whitcomb, B. W., P-287, P-382, P-470<br />

Whitehouse, M. C., O-80, O-1<strong>21</strong>, O-146, O-164,<br />

O-204, P-5, P-130, P-368, P-442, P-629,<br />

P-665, P-675, P-684, P-700<br />

Whiting, S., P-1<strong>17</strong><br />

Wickner, P., O-256<br />

Widra, E., P-328<br />

Widra, E. A., O-61<br />

Wiehle, R., O-29<br />

Wiemer, K., P-6<strong>21</strong><br />

Wild, R. A., O-50<br />

Wilkinson, T., O-250<br />

Willey, R., P-631<br />

Williams, C. L., O-93<br />

Williams, H. L., P-570<br />

Williams, L. A., O-181<br />

Williams, L. J., O-13, P-91<br />

Williams, M., P-452<br />

Williams, P., O-96, O-223, P-703<br />

Williams, P. L., P-97<br />

Williams, S. E., P-602<br />

Willman, S., O-246<br />

Wilson, E. E., P-<strong>21</strong>8<br />

Wilson, L., P-279<br />

Wincek, T., P-357<br />

Winchester, P. D., O-267<br />

Wing, R., P-491<br />

Wininger, J. D., P-602<br />

Winkel, A., P-25<br />

Winkelman, W. D., P-735<br />

Winston, N. J., O-14<br />

Wise, L. A., P-<strong>17</strong>, P-22, P-140, P-384<br />

Witkin, G., P-730<br />

Witkin, S., O-158, P-682<br />

Witkin, S. S., P-125, P-126<br />

Witt, B., O-163<br />

Witz, C. A., P-631<br />

Wiweko, B., P-355<br />

Wolff, E. F., O-142, P-341<br />

Wong, J. M. K., P-736<br />

Wong, K., P-443<br />

Wong, S., P-722, P-723, P-731, P-732<br />

Woo, I., P-339<br />

Wood, E., P-397<br />

Wood, G. J. A., P-535<br />

Woodruff, T. K., P-269<br />

Woodrum, D. A., P-137<br />

Wormer, K., P-332<br />

Wozniak, E. M., P-740<br />

Wright, D. L., O-150<br />

Wright, K., P-<strong>17</strong>2<br />

Wu, H., P-345<br />

Wu, J., P-1<br />

Wu, X., P-84<br />

Wu, Y., P-4<strong>17</strong><br />

Wu, Y.-G., P-592, P-622<br />

Wun, W.-S. A., P-239, P-322, P-361, P-565<br />

Wyatt, M., P-512, P-5<strong>17</strong>, P-518<br />

Wylie, B., O-96<br />

Xia, M., P-27<br />

Xiao, L., P-156, P-<strong>17</strong>7<br />

Xie, C., O-267<br />

Xie, Y., O-154<br />

Xing, T., P-586<br />

Xu, F., O-98<br />

Xu, H., P-190, P-4<strong>17</strong><br />

Xu, J., O-<strong>17</strong>0<br />

Xu, W., O-23<br />

Yagi, H., P-391<br />

Yakin, K., P-634<br />

Yaklic, J., O-<strong>21</strong>4<br />

Yakovenko, S., P-523, P-663, P-667<br />

Yakut, H. I., P-63<br />

Yallampalli, C., P-50<br />

Yamagata, Z., P-407<br />

Yamaguchi, K., O-27, O-113, P-390, P-549,<br />

P-6<strong>17</strong>, P-683<br />

Yamaguchi, T., P-226, P-324, P-369, P-388<br />

Yamanaka, M., P-285<br />

Yamochi, T., P-285<br />

Yan, Z., P-53, P-608<br />

Yanagihori, S., P-57, P-575<br />

Yang, J., P-16, P-385<br />

Yang, M., P-5<strong>21</strong><br />

Yang, N., P-433<br />

Yang, P.-K., P-694<br />

Yang, Q., O-76, O-185<br />

Yang, X., P-657<br />

Yang, Y., P-364<br />

Yang, Z., O-33<br />

Yango, P. L., O-94<br />

Yankov, V. I., P-631<br />

Yao, S., O-95, O-98<br />

Yarnall, S., O-168, O-196, P-14, P-310<br />

Yasa, C., O-127, P-205<br />

Yasmin, S., P-398, P-531<br />

Yasui, Y., P-459<br />

Yauger, B., O-25, P-194, P-341<br />

Ye, A., P-11<br />

Ye, Z., O-249, P-358, P-492<br />

Yeboah, E., P-495<br />

Yee, B., P-288, P-354<br />

Yee, S., O-<strong>21</strong>5<br />

Yegunkova, O., P-396<br />

Yeh, J. S., O-38, O-40, O-41, O-118, O-119,<br />

O-261, P-80<br />

Yelumalai, S., P-570<br />

Yerkes, S. E., O-80, P-730<br />

Yerushalmi, G. M., P-87, P-89<br />

Yesiladali, M., P-437<br />

Yeste, M., P-570<br />

Yeung, E., O-154<br />

Yi, K., P-71<br />

Yi, Y., O-222, P-<strong>21</strong>2<br />

Yih, Y.-C., P-416, P-446<br />

Yildirim, U., P-699<br />

Yilmaz, G., P-205<br />

Yilmaz, N., P-23, P-38, P-61, P-62, P-63<br />

Yilmaz, S., P-23, P-61<br />

Yin, O., O-10<br />

Yin, T., P-385<br />

Yong, P., P-150<br />

Yoon, J.-S., P-543<br />

Yoon, S., P-433, P-572<br />

Yoon, T., P-433, P-572<br />

Yoon, T. K., P-562<br />

Yoshida, A., O-240, P-42<br />

Yoshimura, T., O-64<br />

Young, D. O., O-165<br />

Young, J., P-127<br />

Young, L. K., P-320, P-561<br />

Young, M. J., O-<strong>21</strong>8<br />

Young, N., O-142<br />

Young, S. L., O-16, O-141, O-2<strong>17</strong><br />

Younis, J. S., O-237<br />

Yu, D., P-100<br />

Yu, Y.-P., P-3<br />

Yuksel, S., P-60<br />

Yumoto, K., P-132, P-251<br />

Yung, Y., P-87, P-89<br />

Yurttas Beim, P., O-57, O-160, P-8, P-345, P-419<br />

Yuting, Y., P-367<br />

Zaca, C., O-<strong>17</strong>6, P-103, P-151<br />

Zacur, H. A., O-10<br />

Zadeh, S., O-103<br />

Zaghmout, O., O-70<br />

Zajic, S., P-646<br />

Zakarin Safier, L., O-189, P-424<br />

e376 Author Index Vol. 104, No. 3, Supplement, September <strong>2015</strong>


Zaki, H., P-686<br />

Zaletova, V., P-10<br />

Zamah, A. M., O-14<br />

Zamara, C., O-131<br />

Zaninovic, N., O-249, P-358, P-492<br />

Zarek, S., O-140, O-142, O-191, O-220, P-6, P-30,<br />

P-330, P-713<br />

Zarutskie, P., P-607<br />

Zavy, M. B., O-50<br />

Zavy, M. T., O-50<br />

Zbella, E. A., P-<strong>21</strong>0<br />

Zeadna, A., O-237<br />

Zeeck, K. M., P-383<br />

Zengin, D., O-67, P-441<br />

Zeyneloglu, H. B., P-304<br />

Zgodic, A., O-57<br />

Zhan, H., P-155<br />

Zhan, Q., O-249, P-358, P-492<br />

Zhan, Y., O-137, P-586<br />

Zhang, D., P-4<strong>17</strong>, P-646<br />

Zhang, F., P-670<br />

Zhang, H., O-92, O-201, O-258<br />

Zhang, J., O-33, O-34, O-63, P-495, P-5<strong>21</strong>, P-580,<br />

P-677<br />

Zhang, L., P-506<br />

Zhang, S., P-181, P-582<br />

Zhang, W., O-15<br />

Zhang, X., P-94, P-506<br />

Zhang, Y., P-236, P-385<br />

Zhao, D., O-254<br />

Zhao, Y., O-87, O-230, P-74<br />

Zheng, H., P-79<br />

Zheng, Y., O-266<br />

Zhou, B., P-565, P-702<br />

Zhou, M., P-156, P-<strong>17</strong>7<br />

Zhou, Y., O-<strong>17</strong>8<br />

Zhu, Y., O-225<br />

Zhylkova, L., P-396<br />

Zimmerman, A., O-182<br />

Zimmerman, R., P-505<br />

Ziv-Polat, O., O-62<br />

Zlatopolsky, Z., P-523<br />

Zollinger, C., P-499<br />

Zorina, I., P-667<br />

Zozula, S., P-507<br />

Zullo, F., P-630<br />

FERTILITY & STERILITY Ò<br />

e377


TOPIC INDEX<br />

ART- In Vitro Fertilization: O-037, O-038, O-040,<br />

O-041, O-042, O-061, O-068, O-078, O-081,<br />

O-090, O-115, O-116, O-118, O-119, O-120,<br />

O-127, O-129, O-153, O-169, O-187, O-<strong>21</strong>6,<br />

O-241, O-242, O-247, O-248, O-260, O-261,<br />

O-262, O-264, O-267, O-270, P-023, P-027,<br />

P-028, P-029, P-095, P-096, P-107, P-108,<br />

P-109, P-157, P-196, P-<strong>21</strong>1, P-220, P-2<strong>21</strong>,<br />

P-222, P-223, P-225, P-226, P-227, P-228,<br />

P-230, P-233, P-235, P-236, P-237, P-240,<br />

P-242, P-243, P-251, P-263, P-265, P-286,<br />

P-287, P-288, P-292, P-293, P-294, P-297,<br />

P-298, P-300, P-301, P-302, P-303, P-304,<br />

P-314, P-315, P-316, P-3<strong>17</strong>, P-319, P-320,<br />

P-3<strong>21</strong>, P-324, P-325, P-326, P-327, P-328,<br />

P-329, P-330, P-332, P-333, P-334, P-335,<br />

P-336, P-337, P-338, P-339, P-340, P-341,<br />

P-342, P-345, P-346, P-347, P-348, P-349,<br />

P-351, P-353, P-354, P-355, P-356, P-357,<br />

P-358, P-364, P-385, P-388, P-396, P-398,<br />

P-399, P-406, P-498, P-519, P-578, P-600,<br />

P-626, P-633, P-637, P-645, P-646, P-651,<br />

P-652<br />

ART-Other: O-065, O-083, O-084, O-087, O-103,<br />

O-1<strong>17</strong>, O-154, O-257, O-263, P-224, P-229,<br />

P-231, P-232, P-234, P-239, P-244, P-266,<br />

P-289, P-291, P-295, P-296, P-299, P-305,<br />

P-306, P-307, P-308, P-309, P-310, P-311,<br />

P-312, P-313, P-322, P-323, P-331,<br />

P-343, P-344, P-350, P-352, P-359, P-360,<br />

P-361, P-407, P-543, P-601, P-638, P-719,<br />

P-737<br />

Cancer: O-007, O-0<strong>17</strong>, O-093, P-447, P-448,<br />

P-449, P-450, P-451, P-452, P-453, P-454<br />

Contraception/Family Planning: O-095, O-098,<br />

O-100, O-101, O-102, O-<strong>21</strong>4, P-371, P-372,<br />

P-373, P-374, P-375, P-376, P-377, P-378,<br />

P-379<br />

Cryopreservation: O-003, O-062, O-124, O-155,<br />

O-203, P-238, P-241, P-245, P-246, P-247,<br />

P-248, P-249, P-250, P-252, P-253, P-254,<br />

P-255, P-256, P-257, P-258, P-259, P-260,<br />

P-261, P-262, P-264, P-267, P-268, P-478,<br />

P-479, P-480, P-481, P-482, P-483, P-484,<br />

P-485, P-487, P-533<br />

Early Pregnancy: O-059, O-085, O-086, O-088,<br />

O-089, O-096, O-139, O-140, O-158,<br />

O-194, O-<strong>21</strong>9, O-220, O-2<strong>21</strong>, O-222, P-030,<br />

P-700, P-701, P-702, P-703, P-704, P-705,<br />

P-706, P-707, P-708, P-709, P-710, P-711,<br />

P-712, P-713, P-714, P-715<br />

Embryo Biology: O-044, O-045, O-046, O-048,<br />

O-063, O-091, O-134, O-144, O-243,<br />

O-252, P-579, P-580, P-581, P-582, P-583,<br />

P-584, P-585, P-586, P-587, P-589, P-590,<br />

P-591, P-592, P-593, P-594, P-595, P-596,<br />

P-597, P-599, P-602, P-603<br />

Embryo Culture: O-056, O-138, O-152, O-249,<br />

O-251, P-604, P-605, P-606, P-607, P-608,<br />

P-609, P-610, P-611, P-612, P-613, P-614,<br />

P-615, P-616, P-6<strong>17</strong>, P-618, P-619, P-620,<br />

P-6<strong>21</strong>, P-622<br />

Embryo Transfer: O-005, O-039, O-125, O-126,<br />

O-145, O-146, O-147, O-149, O-150,<br />

O-244, O-245, O-268, P-657, P-658, P-659,<br />

P-660, P-661, P-662, P-663, P-664, P-665,<br />

P-666, P-667, P-668, P-669, P-670, P-671,<br />

P-672<br />

Endometriosis: O-016, O-<strong>17</strong>5, O-<strong>17</strong>6, O-<strong>17</strong>7,<br />

O-<strong>17</strong>8, O-<strong>17</strong>9, O-180, O-266, P-146, P-147,<br />

P-148, P-149, P-150, P-151, P-152, P-154,<br />

P-155, P-156, P-158, P-159, P-160, P-161,<br />

P-162, P-163, P-166, P-167, P-168, P-<strong>17</strong>0,<br />

P-<strong>17</strong>2, P-<strong>17</strong>3, P-<strong>17</strong>4, P-<strong>17</strong>6, P-<strong>17</strong>7, P-<strong>17</strong>8,<br />

P-<strong>17</strong>9<br />

Endometrium: O-018, O-2<strong>17</strong>, P-180, P-181,<br />

P-182, P-183, P-184, P-185, P-186, P-187,<br />

P-188, P-189, P-190, P-191, P-192, P-193<br />

Environment and Toxicology: O-205, O-209,<br />

O-223, O-224, O-225, O-226, O-228, P-092,<br />

P-093, P-094, P-097, P-098, P-099, P-100,<br />

P-101, P-102, P-103, P-104, P-105<br />

Female Reproductive Endocrinology: O-004,<br />

O-092, O-1<strong>21</strong>, O-142, O-162, O-188, O-<strong>21</strong>2,<br />

O-256, O-274, P-003, P-004, P-005, P-006,<br />

P-007, P-008, P-009, P-010, P-011, P-012,<br />

P-013, P-014, P-015, P-016, P-0<strong>17</strong>, P-018,<br />

P-019, P-020, P-0<strong>21</strong>, P-022, P-024, P-025,<br />

P-026, P-031, P-032, P-033, P-034, P-035,<br />

P-036, P-037, P-038, P-039, P-040, P-041<br />

Female Reproductive Surgery: O-067, O-<strong>21</strong>3,<br />

P-194, P-195, P-197, P-198, P-199, P-200,<br />

P-201, P-202, P-203, P-204, P-205, P-206,<br />

P-209, P-<strong>21</strong>0, P-<strong>21</strong>2, P-<strong>21</strong>3, P-<strong>21</strong>4, P-<strong>21</strong>5,<br />

P-<strong>21</strong>6, P-2<strong>17</strong>, P-<strong>21</strong>8, P-<strong>21</strong>9<br />

Female Reproductive Tract: O-097, O-099, O-181,<br />

P-129, P-130, P-131<br />

Fertility Preservation: O-022, O-104, O-<strong>21</strong>5,<br />

O-229, O-230, O-231, O-232, O-233,<br />

O-234, P-455, P-456, P-457, P-458, P-459,<br />

P-460, P-461, P-462, P-463, P-464, P-465,<br />

P-468, P-469, P-470, P-471, P-472, P-473,<br />

P-475, P-476, P-477<br />

Fertilization: O-250, P-153, P-568, P-569, P-570,<br />

P-571, P-572, P-573, P-574, P-575, P-576,<br />

P-577<br />

Genetic Counseling: O-036, O-165, O-166,<br />

O-167, O-193, O-195, O-196, O-197,<br />

O-198, O-204, O-<strong>21</strong>8, P-367, P-368, P-369,<br />

P-370<br />

Imaging: O-055, O-064, O-069, O-070, O-074,<br />

P-132, P-133, P-134, P-135, P-136<br />

Implantation: O-066, O-122, O-156, P-673, P-674,<br />

P-675, P-676, P-677, P-678, P-679, P-680,<br />

P-681, P-682, P-683, P-684, P-685, P-686,<br />

P-687, P-688, P-689, P-690, P-691, P-692<br />

Leiomyoma: O-058, O-071, O-072, O-073, O-075,<br />

O-076, O-182, O-184, O-185, O-186, O-<strong>21</strong>1,<br />

O-269, P-137, P-138, P-139, P-140, P-141,<br />

P-142, P-143, P-144, P-145<br />

Luteal Phase Support: P-693, P-694, P-695, P-696,<br />

P-697, P-698, P-699<br />

Male Factor: O-020, O-110, O-<strong>17</strong>0, O-<strong>17</strong>3, O-<strong>17</strong>4,<br />

O-200, O-271, P-380, P-381, P-382, P-383,<br />

P-384, P-386, P-387, P-389, P-390, P-391,<br />

P-392, P-393, P-394, P-397, P-400, P-401,<br />

P-402, P-403, P-404, P-405, P-408, P-409,<br />

P-410, P-411, P-412<br />

Male Reproductive Endocrinology: P-528, P-529,<br />

P-530, P-531<br />

Male Reproductive Urology: O-019, O-0<strong>21</strong>,<br />

O-024, O-025, O-026, O-027, O-028,<br />

O-029, O-030, O-109, O-111, O-112,<br />

O-113, O-114, O-164, O-199, O-207,<br />

O-208, O-<strong>21</strong>0, P-532, P-534, P-535, P-536,<br />

P-537, P-538, P-539, P-540, P-541, P-542,<br />

P-544, P-545, P-546, P-547, P-548, P-549,<br />

P-550, P-551, P-552, P-553<br />

Menopause: O-008, O-009, O-011, P-001, P-002<br />

Mental Health: O-105, O-106, O-107, O-108,<br />

P-7<strong>21</strong>, P-722, P-723, P-724, P-725, P-726,<br />

P-727, P-728, P-729, P-730, P-731, P-732,<br />

P-733, P-734, P-735, P-736, P-738, P-739,<br />

P-740<br />

Nursing: O-079, O-080, P-366<br />

Obesity and Metabolism: O-015, O-<strong>17</strong>1, O-201,<br />

O-227, O-253, O-273, P-042, P-043, P-044,<br />

P-045, P-046, P-047, P-048, P-049, P-050,<br />

P-051<br />

Oocyte Biology: O-043, O-047, O-133, O-135,<br />

O-143, P-269, P-270, P-271, P-272, P-273,<br />

P-274, P-275, P-276, P-277, P-278, P-279,<br />

P-280, P-281, P-282, P-283, P-284, P-285<br />

Oocyte Maturation: P-561, P-562, P-563, P-564,<br />

P-565, P-566, P-567<br />

Ovarian Function: O-006, O-014, O-136, O-272,<br />

P-083, P-084, P-085, P-086, P-087, P-088,<br />

P-089, P-090, P-091<br />

Ovarian Reserve: O-077, O-131, O-141, O-151,<br />

O-159, O-189, P-418, P-419, P-420, P-4<strong>21</strong>,<br />

P-422, P-423, P-424, P-425, P-426, P-427,<br />

P-428, P-429, P-430, P-431, P-432, P-433,<br />

P-434, P-435, P-436, P-437, P-438, P-439,<br />

P-440, P-441, P-442, P-443, P-444, P-445,<br />

P-446<br />

e378 Topic Index Vol. 104, No. 3, Supplement, September <strong>2015</strong>


Ovarian Stimulation: O-060, O-128, O-130,<br />

O-132, O-160, O-254, O-255, P-623, P-624,<br />

P-627, P-628, P-629, P-630, P-631, P-632,<br />

P-634, P-635, P-636, P-639, P-640, P-641,<br />

P-642, P-643, P-644, P-647, P-648, P-649,<br />

P-650, P-653, P-654, P-655, P-656<br />

Oxidative Stress: P-110, P-111, P-112, P-113,<br />

P-114, P-115, P-116, P-1<strong>17</strong>, P-118, P-119,<br />

P-120, P-1<strong>21</strong>, P-122, P-123<br />

Polycystic Ovary Syndrome: O-010, O-013,<br />

O-049, O-050, O-051, O-052, O-053,<br />

O-054, O-258, O-265, O-275, O-276, P-052,<br />

P-053, P-054, P-055, P-056, P-057, P-058,<br />

P-059, P-060, P-061, P-062, P-063, P-064,<br />

P-065, P-066, P-067, P-068, P-069, P-070,<br />

P-071, P-072, P-073, P-074<br />

Practice Management: O-057, O-259, P-716,<br />

P-7<strong>17</strong>, P-718, P-720<br />

Preimplantation Genetic Diagnosis: O-031,<br />

O-032, O-033, O-034, O-035, O-082,<br />

O-123, O-148, O-163, O-168, O-246, P-489,<br />

P-490, P-491, P-492, P-493, P-494, P-495,<br />

P-496, P-497, P-499, P-500, P-501, P-503,<br />

P-504, P-505, P-506, P-507, P-508, P-509,<br />

P-510, P-511, P-512, P-513, P-514, P-515,<br />

P-516, P-5<strong>17</strong>, P-518, P-520, P-5<strong>21</strong>, P-522,<br />

P-523, P-524, P-525, P-526, P-527<br />

Reproductive Hormones: O-001, O-012, O-161,<br />

O-190, O-191, O-192, P-075, P-076, P-077,<br />

P-078, P-079, P-080, P-081, P-082<br />

Reproductive Immunology: O-137, O-157, P-125,<br />

P-126, P-127, P-128<br />

Sexuality: O-<strong>17</strong>2<br />

Sperm Biology: O-023, P-555, P-556, P-557,<br />

P-558, P-559, P-560<br />

Sperm Preparation: P-413, P-414, P-415, P-416<br />

Stem Cells: O-094, O-235, O-236, O-237, O-238,<br />

O-239, O-240, P-363, P-365<br />

Testis: O-002, O-202, P-4<strong>17</strong><br />

FERTILITY & STERILITY Ò<br />

e379


AUTHOR AND SPOUSE/PARTNER DISCLOSURES INDEX<br />

All speakers at the <strong>2015</strong> ASRM Annual Meeting and Postgraduate Courses were required to complete a disclosure form. These<br />

disclosures were reviewed and potential conflicts of interest resolved by the Subcommittee on Standards of Commercial Support of<br />

the Continuing Medical Education Committee. Each abstract or video author is listed below along with any relationships their<br />

partners/spouses disclosed.<br />

Adamson, G. D.<br />

Advanced Reproductive Care, CEO,<br />

Shareholder; LabCorp, Paid<br />

consultant; Bayer, Paid consultant;<br />

Ferring, Paid consultant; Auxogyn,<br />

Grant recipient; Ziva, Direct<br />

stockholder<br />

Ahmed, H.<br />

Celgene, Full-time company<br />

employee<br />

Albertini, D.<br />

Springer Publishing, Paid consultant;<br />

Italian Ministry of Science, Grant<br />

recipient; State of Oklahoma,<br />

Honoraria; TEDCO, Paid<br />

consultant<br />

Albertsen, H. M.<br />

Juneau Biosciences, LLC, Company<br />

officer; Juneau Biosciences, LLC,<br />

Direct stockholder<br />

Alikani, M. Reprogenetics, Direct stockholder;<br />

Life Global, Paid consultant<br />

Alper, M. M.<br />

EMD Serono, Honoraria; Good Start<br />

Genetics, Advisory Board; EMD<br />

Serono, Honoraria; Reprosource,<br />

Advisor-Board; Optum, Honoraria<br />

Alvero, R.<br />

Cooper Surgical, Paid consultant<br />

Andriani, L.<br />

I own non-controlling interests in<br />

publicly traded stock of<br />

companies in the medical field.<br />

These securities were purchased<br />

and are managed on a fully<br />

discretionary basis by an SEC<br />

registered independent investment<br />

advisor or are owned through<br />

pooled, investment vehicles like<br />

mutual funds and/or exchange<br />

traded funds and represent<br />

substantially less than 1%<br />

ownership interests. Investments<br />

include: Celgene, Cerner, Costco,<br />

Danaher, Gilead, Johnson &<br />

Johnson, Merck, Mylan, Proctor &<br />

Gamble, & Walgreens.<br />

Apter, D.<br />

Bayer, Grant recipient; Merck, Grant<br />

recipient; Exeltis, Grant recipient;<br />

Bayer, Merck and Exeltis,<br />

Speakers bureau; GSK, Grant<br />

recipient<br />

Arbo, E.<br />

Ferring, Full-time company employee<br />

Archer, D. F.<br />

AbbVie, Paid consultant; AbbVie,<br />

Grant recipient; TherapeuticsMD,<br />

Paid consultant; TherapeuticsMD,<br />

Grant recipient; Bayer Healthcare,<br />

Paid consultant; Bayer Healthcare,<br />

Grant recipient; Agile<br />

Pharmaceuticals, Paid consultant;<br />

Exeltis/CHEMO France, Paid<br />

consultant; Endoceutics, Paid<br />

consultant; Endoceutics, Grant<br />

recipient; TEVA/HR Pharma, Paid<br />

consultant<br />

Avril, C.<br />

Merck Serono, Paid consultant;<br />

Ferring, Paid consultant<br />

Baker, V. L.<br />

Good Start Genetics, Advisory Board;<br />

Ovuline, Unpaid consultant<br />

Barad, D. H.<br />

U.S. Patents, DHB is a co-inventor on<br />

a number of FMR1 gene-related<br />

U.S. patents and still pending<br />

patent applications, which claim<br />

diagnostic benefits from<br />

evaluations of the gene. One of<br />

these patents was licensed to<br />

Generation Medical Associates,<br />

PLLC.; Generation Medical<br />

Associates, PLLC, Direct<br />

stockholder; U.S. Patents, DHB<br />

holds patents that claim therapeutic<br />

benefits from androgen<br />

supplementation in women with<br />

LFOR and hypoandrogenism and<br />

receives licensing fees for the<br />

patents from Fertility<br />

Nutraceuticals, LLC.<br />

Barnhart, K.<br />

Bayer, Paid consultant; SPD, Paid<br />

consultant<br />

Barrett, C. B. ReproSource Fertility Diagnostics,<br />

Paid consultant<br />

Barriere, P.<br />

Genevrier France, Honoraria; Merck<br />

Serono France, Paid consultant;<br />

MSD France, Paid consultant;<br />

HRA Pharma, Paid consultant;<br />

Ferring France, Honoraria<br />

Bauer, J. D.<br />

Ferring, Employee<br />

Behr, B.<br />

Auxogyn, Direct stockholder; Ivigen,<br />

Direct stockholder<br />

Bellerose, H.<br />

Counsyl, Inc, part-time contracted<br />

employee<br />

Beltsos, A. N.<br />

Merck Pharmaceuticals, Speakers<br />

bureau; EMD Serono<br />

e380 Author Disclosures Index Vol. 104, No. 3, Supplement, September <strong>2015</strong>


Pharmaceuticals, Speakers bureau;<br />

Ferring Pharmaceuticals, Speakers<br />

bureau; Diclegis, Speakers bureau;<br />

Good Start Genetics, Paid<br />

consultant; Optum, Paid consultant<br />

Benadiva, C. A.<br />

OvaScience, Paid consultant<br />

Bendikson, K.<br />

Theralogix, Paid consultant<br />

Berga, S. L. Ferring Pharmaceuticals, Attended<br />

fundraiser at own expense;<br />

Ferring Pharmaceuticals, Grant<br />

recipient; Pfizer, Paid consultant;<br />

UpToDate, Editor; Pfizer, Grant<br />

recipient<br />

Bergh, C. M.<br />

MedSoftware, Company officer<br />

Berkeley, A. S.<br />

Merck, Direct stockholder; Pfizer,<br />

Direct stockholder; Glaxo, Direct<br />

stockholder; Becton, Dickenson,<br />

Full-time company employee;<br />

Bristol Myers, Direct stockholder<br />

Bernstein, L. R.<br />

Merck, Received free medication.;<br />

Fertility Center of <strong>Maryland</strong>, Paid<br />

consultant<br />

Berro, R.<br />

Celmatix, Inc, Full-time company<br />

employee<br />

Bhagavath, B.<br />

Hologic, Paid consultant; Abbvie,<br />

Paid consultant<br />

Bisignano, A. Recombine, Company officer;<br />

Recombine, Direct stockholder;<br />

Recombine, Full-time company<br />

employee<br />

Bissonnette, F.<br />

YAD, Direct stockholder<br />

Blazek, J.<br />

Genesis Genetics, Full-time company<br />

employee<br />

Boekelheide, K.<br />

Boehringer Ingelheim, Grant<br />

recipient; Tb Alliance, Paid<br />

consultant; Drugs for Neglected<br />

Diseases, Paid consultant;<br />

CytoSolv, Direct stockholder<br />

Boivin, J.<br />

Merck & Co, Paid consultant;<br />

Actavis, Paid consultant; EMD<br />

Ltd, Honoraria<br />

Borini, A.<br />

Merck Serono GFI, Grant recipient;<br />

Merck Serono Poland, Speakers<br />

bureau; Unisense fertiliTech,<br />

Participate advisory board<br />

Brannigan, R. E.<br />

Abbvie, Inc., A grant in support of<br />

Northwestern University’s<br />

Andrology fellowship was<br />

provided to Northwestern<br />

University, Feinberg School of<br />

Medicine. I am the Director of the<br />

Andrology Fellowship.; The<br />

American Urological Association/<br />

The Journal of Urology, I am an<br />

Assistant Editor for The Journal of<br />

Urology<br />

Bristow, S. L.<br />

Recombine, Full-time company<br />

employee; Recombine, Stock<br />

options<br />

Butler, S. A.<br />

MAP Diagnostics Ltd, Company<br />

officer<br />

Carlsson, M.<br />

Pfizer Inc., Full-time company<br />

employee; CSG INC., Full-time<br />

company employee<br />

Carrell, D. T.<br />

Episona, Inc, Company officer<br />

Castells, M.<br />

Merck, Sanofi, Novartis, Paid<br />

consultant<br />

Cataldo, N. A.<br />

CenseoHealth, Independent<br />

contractor; GoodStart Genetics,<br />

Paid consultant<br />

Catherino, W. H.<br />

Abbvie, Paid consultant; Viteava<br />

Pharmaceuticals, Paid consultant;<br />

Medical College of Wisconsin,<br />

Honoraria; Recombine, Full-time<br />

company employee; Bayer, Grant<br />

recipient<br />

Cedars, M. Ferring Pharmaceutical, Research<br />

support - investigator - initiated<br />

Celia, G.<br />

Good Start Genetics, Inc., Providing<br />

research samples to the company<br />

Centola, G. M.<br />

New England Cryogenic Center, Paid<br />

consultant; Cryos International<br />

USA, Paid consultant; Manhattan<br />

Cryobank, Paid consultant;<br />

NYMHB Fertility Services;<br />

Dr. George Kofinas, Paid<br />

consultant; Seattle Sperm Bank/<br />

Phoenix Sperm Bank, Paid<br />

consultant<br />

Chan, C.<br />

Level, Company officer<br />

Chang, C.-C.<br />

MyEggBank, Direct stockholder<br />

Chen, A. A.<br />

Progyny, Full-time company<br />

employee<br />

Chen, D.<br />

Gilead Sciences, Direct stockholder;<br />

Pfizer, Direct stockholder<br />

Chen, S. H.<br />

Ovascience, Paid consultant; Hologic,<br />

Paid consultant; Optum, Paid<br />

consultant; Recombine, Paid<br />

consultant<br />

Chettier, R.<br />

Juneau Biosciences, Full-time<br />

company employee; Affiliated<br />

Genetics; Juneau Biosciences,<br />

Direct stockholder<br />

Cholkeri-Singh, A. Ethicon Endo Surgery, Speakers<br />

bureau; Ethicon Endo Surgery,<br />

Advisory Board Member; Bayer,<br />

Speakers bureau; Bayer, Advisory<br />

Board Member<br />

Christman, G. M.<br />

Abbvie Pharmaceuticals, Grant<br />

recipient; Abbvie Pharmaceuticals,<br />

Honoraria; Bayer Pharmaceuticals,<br />

Grant recipient; Bayer<br />

Pharmaceuticals, Paid consultant<br />

Chwalisz, K.<br />

AbbVie Inc., Full-time company<br />

employee; AbbVie Inc., Direct<br />

stockholder<br />

Clarke, N. J.<br />

Quest Diagnostics, Full-time<br />

company employee<br />

Coddington, C.<br />

PG, Merck, Stock<br />

Cohen, J.<br />

Reprogenetics LLC, Direct<br />

stockholder; Reprogenetics LLC,<br />

Paid consultant; Life Global Inc.,<br />

Paid consultant<br />

Conaghan, J.<br />

Irvine Scientific, Paid consultant;<br />

Auxogyn, Paid consultant<br />

Considine, R.<br />

Eli Lilly Research Labs, Paid<br />

consultant; Merck Research Labs,<br />

Paid consultant<br />

Copperman, A. B.<br />

EMD Serono, Speakers bureau;<br />

Merck, Speakers bureau; Ferring,<br />

Speakers bureau<br />

Corrado, J.<br />

Good Start Genetics, Member of the<br />

Nurse Advisory Panel<br />

Costantini, L.<br />

Prima-Temp, Inc, Company officer<br />

FERTILITY & STERILITY Ò<br />

e381


Craig, L.T. B. Ferring Pharmaceuticals, Principal<br />

Investigator; Roche Diagnostics,<br />

Principal Investigator<br />

Czuprenski, E.<br />

Genesis Genetics, Full-time company<br />

employee<br />

Daneshmand, S.<br />

Actavis Inc., Grant recipient<br />

Daubert, M.<br />

Stealth Peptides, Paid consultant<br />

Davie, J. Good Start Genetics, Inc., Paid<br />

consultant<br />

Davies, M. C.<br />

Centre for Reproductive and Genetic<br />

Health, 256 Grays Inn Road,<br />

London WC1X 8LD UK, Paid<br />

consultant<br />

Demirci, U.<br />

DxNow, Co-founder, Scientific<br />

Advisor; Koek Biotech,<br />

Co-founder, Scientific Advisor<br />

Demko, Z.<br />

Natera, Full-time company<br />

employee; Natera, Direct<br />

stockholder<br />

Desai, N.<br />

Fertilitech, Advisory Board<br />

D’Hooghe, T.<br />

WERF (World Endometriosis<br />

Research Foundation), Board<br />

Membership till July 2011; Bayer<br />

Pharma, Proteomika,<br />

MPharmaplex, Astellas, roche<br />

Diagnostics, Paid consultant;<br />

Ferring, Merk Serono, Merck,<br />

Besins, Pharmaplex, Grant<br />

recipient; Ferring, Merk Serono,<br />

Merk, Travel/accomodations/<br />

meeting expenses unrelated to<br />

activities listed.Money to our<br />

institution, University Hospitals<br />

Leuven, Belgium<br />

Diamond, M. P. NICHD, AbbVie, EMD Serono,<br />

Baxter, Grant recipient; Teijin<br />

Pharmaceuticals, Auxogyn, Paid<br />

consultant; Advanced<br />

Reproductive Care, Board of<br />

Directors and Stockholder<br />

Diaz-Gimeno, P.<br />

Igenomix, ERA Patent authorship<br />

Diez Juan, A.<br />

Igenomix, Full-time company<br />

employee<br />

Domar, A. D.<br />

Merck, Grant recipient;<br />

Johnson&Johnson, Paid<br />

consultant; Merck, Speakers<br />

bureau; Ovascience, Paid<br />

consultant; Ferring, Paid<br />

consultant; Merck, Paid<br />

consultant; UptoDate, Paid<br />

consultant; Nora Therapeutics,<br />

Paid consultant<br />

Doody, K.<br />

Merck Pharmaceutical, Paid<br />

consultant; Ferring<br />

Pharmaceutical, Paid consultant;<br />

Finox Pharmaceutical, Paid<br />

consultant; Serono<br />

Pharmaceutical, Speakers bureau;<br />

Good Start Genetics, Paid<br />

consultant<br />

Drevet, J. R.<br />

Celloxess, Board member<br />

Driggers, P.<br />

Eisai, Inc., Speakers bureau; novonordisk,<br />

Speakers bureau; Eisai,<br />

Inc., Paid consultant; Eisai, Inc.,<br />

Clinical Investigator; novonordisk,<br />

Clinical Investigator<br />

Dudley, P. S.<br />

MyEggBank, N.A., Direct<br />

stockholder; Attain Genetics,<br />

LLC, Direct stockholder<br />

Dundee, J. A.<br />

New England Fertility Society/<br />

Ferring Pharmaceuticals, Grant<br />

recipient<br />

Dunn, R. D.<br />

Finox, PI for one of many sites for<br />

Afolia ART Study<br />

Dwan, P. G.<br />

EMD SERONO, Speakers bureau;<br />

EMD SERONO, Paid consultant<br />

Dye, T. D.<br />

Pfizer, Grant recipient; Humana, Inc.,<br />

Paid consultant<br />

Dzidic, N. Combimatrix, Full-time company<br />

employee<br />

Eid, S.<br />

Teva Pharmaceuticals, Paid<br />

consultant<br />

Eisenberg, M. Sandstone Diagnostics, Direct<br />

stockholder; Reprovantage, Direct<br />

stockholder; Glow, Advisor;<br />

EmbraceHer, Direct stockholder<br />

Evans, E.<br />

Counsyl, Company officer<br />

Fanchin, R. Merck-Serono, Grant recipient;<br />

Ferring, Grant recipient<br />

Farrington, P. Affiliated Genetics, Direct<br />

stockholder; Juneau Biosciences,<br />

Company officer; Juneau<br />

Biosciences; Affiliated Genetics,<br />

Direct stockholder; Juneau<br />

Biosciences, Full-time company<br />

employee<br />

Faulkner, N.<br />

Good Start Genetics, Full-time<br />

company employee<br />

Faustmann, T. A.<br />

Bayer Pharma AG, Full-time<br />

company employee<br />

Feinberg, E.<br />

Abbvie, Data Safety Monitoring<br />

Board; Natera, Advisory Board<br />

Feinberg, R. F.<br />

AbbVie, Clinical Trial; Nora, Clinical<br />

Trial; Ferring, Clinical Trial;<br />

Finox, Clinical Trial<br />

Fode, M.<br />

Eli Lilly, Paid consultant; Astellas,<br />

Paid consultant; Menarini, Paid<br />

consultant; Coloplast, Honoraria<br />

Foong, S.<br />

EMD Serono, Honoraria<br />

Forman, E. J.<br />

Ferring Pharmaceuticals, Speakers<br />

bureau<br />

Gao, J.<br />

AbbVie, Full-time company<br />

employee<br />

Gardner, D. K.<br />

Vitrolife, Grant recipient<br />

Gargiulo, A. R.<br />

OmniGuide, Inc., Paid consultant;<br />

Kawasaki Robotics (USA), Inc.,<br />

Paid consultant<br />

Gates, D.<br />

Merck and Co., Full-time company<br />

employee; Covance, Full-time<br />

company employee<br />

Gayet, V.<br />

Ferring, Honoraria<br />

Gemzell-Danielsson, K.<br />

Bayer AG; MSD/Merck; HRA<br />

Pharma; ExelGyn; Gedeon<br />

Richter, Honoraria; Bayer, Grant<br />

recipient<br />

Gharagozloo, P.<br />

CellOxess LLC, Company officer<br />

Gheyas, F.<br />

Merck, Company officer<br />

Ginsburg, E.<br />

UptoDate, Honoraria; Springer Inc,<br />

Honoraria; Serono: investigator<br />

initiated study, Grant recipient<br />

Giudice, L. ASRM, Company officer; Merck;<br />

Pfizer, Direct stockholder<br />

Givens, C.<br />

Merck, Paid consultant<br />

Gleicher, N.<br />

Fertility Nutraceuticals, LLC, Direct<br />

stockholder; Generation Medical<br />

Associates, PLLC, Direct<br />

stockholder; Fertility<br />

Nutraceuticals, LLC, Receive<br />

e382 Author Disclosures Index Vol. 104, No. 3, Supplement, September <strong>2015</strong>


patent licensing fees; Generation<br />

Medical Associates, PLLC, Patent<br />

licensed; U.S. Patents, NG holds<br />

patents that claim therapeutic<br />

benefits from androgen<br />

supplementation in women with<br />

LFOR and hypoandrogenism.;<br />

U.S. Patents, NG is a co-inventor<br />

on a number of FMR1 gene-related<br />

U.S. patents and still pending<br />

patent applications, which claim<br />

diagnostic benefits.<br />

Globus, S. T.<br />

Celmatix Inc., Full-time company<br />

employee<br />

Go, K. J.<br />

Ferring Pharmaceuticals, Full-time<br />

company employee<br />

Godfrey, E. M.<br />

Prima-Temp, Grant recipient; Bayer<br />

Pharmaceuticals, Grant recipient;<br />

Teva Women’s Health, Grant<br />

recipient; Merck, Sharp & Dohme<br />

Corp, Speakers bureau<br />

Goering, M. C.<br />

MedTech for Laboratory Solutions,<br />

Paid consultant<br />

Gold, M.<br />

Recombine, Full-time company<br />

employee; Recombine, I have<br />

stock options.<br />

Goldberg, J. D.<br />

Counsyl, Inc, Full-time company<br />

employee<br />

Goldfarb, J. M.<br />

Lumara Health – no longer affiliated<br />

with teh company since 11/14,<br />

Direct stockholder<br />

Goldstein, M.<br />

Therologix, advisory board<br />

Gordon, K.<br />

Merck & Co, Full-time company<br />

employee<br />

Gordon, T. T.<br />

Genesis Genetics, Direct stockholder;<br />

Rubicon Genomics, Paid<br />

consultant<br />

Grainger, D. A.<br />

Abbvie, Speakers bureau; Shionogi,<br />

Inc., Speakers bureau<br />

Grantz, K. L.<br />

Mylan, Inc., Direct stockholder<br />

Grifo, J.<br />

Illumina, Speakers bureau;<br />

Ovascience, Medical Advisory<br />

Board<br />

Grifo, J. A.<br />

Illumina, Speakers bureau;<br />

Ovascience, Medical Advisor<br />

Grober, E. D. Abbott, Paid consultant; Paladin,<br />

Grant recipient; Eli Lilly, Paid<br />

consultant<br />

Gross, J.<br />

Novartis, Full-time company<br />

employee<br />

Guan, Y.<br />

Merck&Co., Inc., Full-time company<br />

employee<br />

Gutmann, J.<br />

somalogic, Paid consultant<br />

Hadjiliadis, D.<br />

Gilead, Advisory board; Bayer,<br />

Advisory board; Vertex, Advisory<br />

board<br />

Hakim, L. S.<br />

ENDO/AMS, Speakers bureau<br />

Hamamah, S.<br />

Ferring, Grant recipient<br />

Hammond, K.<br />

Good Start Genetics, Paid consultant;<br />

CenseoHealth, Independent<br />

contractor<br />

Hannam, T.<br />

EMD Serono, Our company sells<br />

pharmaceutical products to<br />

patients; Organon, Our company<br />

sells pharmaceutical products to<br />

patients; Ferring, Our company<br />

sells pharmaceutical products to<br />

patients<br />

Hansen, K. R.<br />

Roche Diagnostics, Grant recipient;<br />

Ferring International<br />

Pharmascience Center US, Grant<br />

recipient<br />

Haque, I. S.<br />

Counsyl, Full-time company<br />

employee<br />

Harkins, G.<br />

Ethicon EndoSurgery, Paid<br />

consultant; Intuitive Surgical, Paid<br />

consultant<br />

Harutunian, A.<br />

Genesis Genetics, Direct stockholder<br />

Hasson, J.<br />

‘‘Temp-drop LTD’’; This is a start-up<br />

company which manufactures<br />

a self monitoring device which<br />

continuously measures body<br />

temperature in sleep., Medical<br />

adviser. The position is unpaid.<br />

Hayward, B.<br />

EMD Serono, Inc., Full-time<br />

company employee<br />

Heiser, P. W.<br />

Ferring Pharmaceuticals, Inc., Fulltime<br />

company employee<br />

Heller, B.<br />

Pulling Down the Moon Inc.,<br />

Company officer<br />

Hennebold, J. D.<br />

AbbVie Pharmaceuticals, Direct<br />

stockholder; Abbott Laboratories,<br />

Direct stockholder; Omnicare,<br />

Direct stockholder; Gedeon<br />

Richter PregLem SA, Grant<br />

recipient<br />

Hesley, G.<br />

Insightec, Grant recipient<br />

Hirshfeld-Cytron, J. E.<br />

Duchesnay, Speakers bureau<br />

Hornstein, M. D. WINFertility, Meidcal Advisory<br />

Board; Up-To-Date, Author<br />

Hotaling, J. M.<br />

Andro360, own equity in early stage<br />

start up company that has not<br />

made a profit; SpermDx, own<br />

equity in early stage start up<br />

company that has not made a profit;<br />

StreamDx INC, own equity in early<br />

stage start up company that has not<br />

made a profit<br />

Howard, B. Teva Pharmaceuticals, Full-time<br />

company employee<br />

Howles, C. M.<br />

Finox Biotech AG., Paid consultant<br />

Hubbard, J. W.<br />

Affymetrix Inc, Full-time company<br />

employee<br />

Hunter Cohn, K.<br />

Celmatix Inc, Full-time company<br />

employee<br />

Hu-Seliger, T.<br />

Celmatix, Full-time company<br />

employee<br />

Iles, R. K.<br />

MAPDiagnostics Ltd, Company<br />

officer<br />

Israel, M. A.<br />

Progyny, Inc., Full-time company<br />

employee<br />

Jacobs, I.<br />

ABCODIA, Company officer<br />

Jacoby, V.<br />

Halt Medical, Grant recipient<br />

Jadhav, P.<br />

Merck, Full-time company employee<br />

Jain, R.<br />

AbbVie, Full-time company<br />

employee<br />

Jasulaitis, S.<br />

Merck Pharmaceuticals, Speakers<br />

bureau<br />

Jenkins, J.<br />

Finox Biotech A.G., Company officer<br />

Kadoch, I.-J.<br />

Clinique Ovo, Clinical director; Yad-<br />

Tech, Shares holder<br />

Kalra, B.<br />

Ansh Labs, Full-time company<br />

employee<br />

Kamga-Ngande, C.<br />

Clinique Ovo, Clinical director<br />

Karvir, H.<br />

Celmatix Inc., Full-time company<br />

employee<br />

FERTILITY & STERILITY Ò<br />

e383


Katilius, J.<br />

Prima-Temp,Inc, Paid consultant<br />

Kellogg, G. R.<br />

Recombine, Full-time company<br />

employee; Recombine, Stock<br />

Options<br />

Ketterson, K.<br />

Recombine, Shareholder<br />

Kimura, T.<br />

Takeda Pharmaceutical Co.Ltd.,<br />

Grant recipient; Nohon Kayaku<br />

Co.Ltd., Paid consultant; Nihon<br />

Shinyaku Co.Ltd, Grant recipient;<br />

Fuji Pharmaceutical Co. Ltd.,<br />

Grant recipient; MSD Co.Ltd.,<br />

Speakers bureau; Chyugai<br />

Pharmaceutics Co. Ltd., Grant<br />

recipient<br />

Kolb, B. A. My Fertility, Company officer;<br />

Ferring Pharmaceuticals, Speakers<br />

bureau; EMD Serono, Paid<br />

consultant; Good Start Genetics,<br />

Paid consultant<br />

Kovalevsky, G.<br />

AbbVie, Clinical research trial<br />

investigator; Nora, Clinical<br />

research trial investigator; Ferring,<br />

Clinical research trial investigator;<br />

Finox, Clinical research trial<br />

investigator<br />

Krawetz, S.<br />

Taylor and Francis, EIC of Systems<br />

Biology in Reproductive Medicine<br />

Krisher, R. L.<br />

Merck Serono, Grant recipient<br />

Kumar, A.<br />

AnshLabs, Full-time company<br />

employee<br />

Kumar, D.<br />

Hologic Inc., Although I am an<br />

employee of the University of<br />

Rochester, part of my salary is paid<br />

by Hologic Inc. as I am<br />

coordinating a study sponsored by<br />

Hologic Inc. This study is<br />

regarding the effectiveness of the<br />

Novasure Global Endometrial<br />

Ablation procedure in t<br />

Kumar, N.<br />

Recombine, Full-time company<br />

employee<br />

Kushnir, V. A.<br />

Generation Medical Associates,<br />

PLLC, Direct stockholder<br />

Kyrou, D.<br />

MSD, Honoraria<br />

Lamb, D. J.<br />

Cellmatrix, Paid consultant<br />

Lannon, B.<br />

Dyax Inc., Direct stockholder<br />

Large, M.<br />

Genesis Genetics, Full-time company<br />

employee<br />

Lazarin, G. A.<br />

Counsyl, Full-time company<br />

employee<br />

Leboeuf, M.<br />

Actavis, Speakers bureau; Bayer,<br />

Speakers bureau; Abbvie, Speakers<br />

bureau<br />

Legro, R. S.<br />

Ferring, Grant recipient; Astra<br />

Zeneca, Grant recipient;<br />

Euroscreen, Paid consultant;<br />

Kindex, Paid consultant; Takeda,<br />

Paid consultant; Clarus<br />

Therapeutics, Paid consultant<br />

Leone Roberti Maggiore, U. DEKA, Speakers bureau<br />

Lessey, B. A.<br />

Pfizer, Paid consultant; Sepal, Inc.,<br />

Licensed technology and potential<br />

royalty payments<br />

Levy, M.<br />

Donor Egg Bank, USA: President,<br />

Company officer<br />

Li, M.<br />

PacGenomics, Company officer<br />

Liebermann, J.<br />

Sage/Origio, Paid consultant; Vivere,<br />

Paid consultant; Irvine Scientific,<br />

Speakers bureau<br />

Lindeman, M. R.<br />

Genesis Genetics, Full-time company<br />

employee<br />

Lindheim, S. R.<br />

Hologic, Speakers bureau; Abbvie,<br />

Speakers bureau; Progenity,<br />

Speakers bureau; Bayer, Grant<br />

recipient; Cooper Surgical,<br />

Non-paid Advisor<br />

Liu, K.<br />

Ferring Canada, Grant recipient<br />

qukaszuk, K.<br />

Ferring Pharmaceuticals, Honraria;<br />

Roche Diagnostics, Honoraria;<br />

Merck, Honoraria<br />

Luke, B.<br />

Society for Assisted Reproductive<br />

Technology, Paid consultant<br />

Lukes, A. S.<br />

Abbvie, Grant recipient; Abbvie, Paid<br />

consultant; Glaxo-Smith Kline,<br />

Paid consultant; Mirabilis Medica,<br />

Paid consultant; Hologic, Paid<br />

consultant; Bayer, Paid consultant;<br />

Agile, Grant recipient; Sequoia,<br />

Grant recipient; Therapeutics,<br />

Grant recipient; Bayer, Grant<br />

recipient; Watson, Grant recipient;<br />

Hologic, Grant recipient; Merck,<br />

Grant recipient; Amgen, Grant<br />

recipient<br />

Lukkari-Lax, E.<br />

Bayer Oy, Full-time company<br />

employee<br />

Mahony, M. C.<br />

EMD Serono, Inc., a subsidiary of<br />

Merck KGaA, Darmstadt,<br />

Germany, Full-time company<br />

employee<br />

Malik, M.<br />

AB Sciex, Full-time company<br />

employee<br />

Manoharan, A. P.<br />

Recombine, Full-time company<br />

employee; Recombine, Stock<br />

options<br />

Martinez, C.<br />

Upward Labs Holdings, Inc.<br />

(www.glowing.com), Company<br />

officer<br />

Massin, N.<br />

MSD, Honoraria; Merck, Honoraria<br />

Mather, K.<br />

Novo Nordisk, Grant recipient; sanofi<br />

aventis, Grant recipient; Merck<br />

Inc, Grant recipient; Abbott, Grant<br />

recipient; Boehringer Ingelheim,<br />

Paid consultant<br />

Matsuura, K.<br />

ZEON Corporation, Speakers bureau<br />

McCoy, R.<br />

Stanford University, Co-inventor on<br />

a provisional patent filed by<br />

Stanford University, related to<br />

work in collaboration with Natera.<br />

McCulloh, D. H. ReproART: Georgian American<br />

Center for Reproductive<br />

Medicine, Tbilisi, Georgia,<br />

Company officer; Biogenetics<br />

Corporation, Mountainside, New<br />

Jersey, USA, Company officer;<br />

NYU Fertility Center, New York,<br />

New York, USA, Full-time<br />

company employee<br />

McCullough, A. Repros, Paid consultant; Repros,<br />

Direct stockholder; Pfizer, Direct<br />

stockholder; Pfizer, DSMB;<br />

Antares, Grant recipient; Repros,<br />

Grant recipient; ISSM, Honoraria;<br />

AUA, Honoraria; Capital Region<br />

Medical Research Fund, Grant<br />

recipient<br />

McWilliams, K.<br />

Genesis Genetics, Full-time company<br />

employee<br />

e384 Author Disclosures Index Vol. 104, No. 3, Supplement, September <strong>2015</strong>


McWilliams, T. K.<br />

Genesis Genetics, Full-time company<br />

employee<br />

Meintjes, M.<br />

Vitrolife AB, Paid consultant<br />

Merrion, K. Natera, Inc., Full-time company<br />

employee; Natera, Inc., Option to<br />

hold stock in Natera, Inc.<br />

Miller, C. E. Covidien, Femasys, Olympus,<br />

Novartis, Abbvie, Intuitive<br />

Surgical, Gynesonics, Grant<br />

recipient; Ethicon, Covidien,<br />

Femasys, Abbvie, Halt Medical,<br />

Intuitive Surgical, Gynesonics,<br />

Paid consultant; Ethicon, Smith &<br />

Nephew, Intuitive Surgical,<br />

Speakers bureau<br />

Minjarez, D. A.<br />

Ferring, Speakers bureau<br />

Moazamian, A.<br />

CellOxess LLC, Full-time company<br />

employee<br />

Moazamian, R. J.<br />

CellOxess, Intern<br />

Moley, K. OvaScience, Scientific Advisory<br />

Board Member<br />

Montegriffo, E.<br />

Bayer HealthCare Pharmaceuticals,<br />

Full-time company employee<br />

Morbeck, D.<br />

Vitrolife, Research equipment loan<br />

Mullen, T.<br />

Good Start Genetics, Full-time<br />

company employee<br />

Munne, S. Reprogenetics, Direct stockholder;<br />

Recombine, Direct stockholder<br />

Muzzey, D.<br />

Counsyl Inc., Full-time company<br />

employee<br />

Nagy, Z. P.<br />

My Egg Bank, Direct stockholder;<br />

Cooper Surgical/Origio, Paid<br />

consultant; Fertilitech, Paid<br />

consultant; MERCK MSD,<br />

Speakers bureau<br />

Natan, Y.<br />

FertileSafe Ltd., Full-time company<br />

employee<br />

Neitzel, D.<br />

Good Start Genetics, Full-time<br />

company employee<br />

Nelson, A.<br />

Agile; Bayer, Grant recipient;<br />

Actavis; Bayer; Merck; Pfizer ;<br />

Teva, Honoraria; Actavis; Agile;<br />

Bayer; ContraMed; Merck; Teva;<br />

MicroCHIPS Biotech; Pharmanest,<br />

Paid consultant<br />

Niederberger, C.<br />

American Urological Association,<br />

Journal section editor; update<br />

series editor; Ferring, Grant<br />

recipient; IBSA, Grant recipient;<br />

NexHand, Company officer<br />

North, J.<br />

AbbVie Inc., Full-time company<br />

employee; Abbott Laboratories,<br />

Prior employee<br />

Norwitz, E. R.<br />

Natera, Unpaid member of Advisory<br />

Board; Hologic, Unpaid member<br />

of Advisory Board; Bayer<br />

Pharmaceuticals, Named inventor<br />

on a patent for the diagnosis of<br />

preeclampsia purchased by Bayer<br />

Ohl, D. A. Pfizer, Scientific Advisor; Endo,<br />

Grant recipient; Coloplast, AMS,<br />

Surgical Consultant and Advisory<br />

Board<br />

Page, D. C.<br />

Counsyl, Scientific Advisory Board<br />

Paik, K. G.<br />

Natera, Inc., Full-time company<br />

employee<br />

Palermo, G. D.<br />

Irvine Scientific, Royalties<br />

Park, S. O.<br />

CHA Gangnam Medical center, CHA<br />

University, Full-time company<br />

employee<br />

Patrizio, P.<br />

Counsyl, Honoraria; FertileSafe,<br />

scientific advisor<br />

Paulson, R.<br />

Ferring Pharmaceuticals, Honoraria;<br />

Origio, Paid consultant<br />

Penrose, L.<br />

Reproductive Solutions Inc.,<br />

Company officer<br />

Penzias, A.<br />

OvaScience, Company Advisor;<br />

ReproSource, Company Advisor;<br />

Nora Therapeutics, Company<br />

Advisor; Ferring Pharmaceuticals,<br />

Honoraria<br />

Peralta, S. Merck-Serono, Attendance at the<br />

meeting and travel expenses<br />

Petrozza, J. C.<br />

Interlace Medical / Hologic,<br />

Scientific Advisory Committee<br />

Pettersen, B.<br />

Natera, Inc., Full-time company<br />

employee<br />

Pohlmeier, A. M.<br />

Teva Pharmaceuticals, Full-time<br />

company employee<br />

Polhemus, A. M.<br />

CellOxess, LLC, Full-time company<br />

employee<br />

Pomeroy, K. O.<br />

The World Egg Bank, Full-time<br />

company employee<br />

Pool, T. B.<br />

Auxogyn, Honoraria<br />

Porcu Buisson, G.<br />

Ferring SA, Honoraria; Merck Serono<br />

SA, Honoraria<br />

Prien, S.<br />

Reproductive Solutions Inc,<br />

Company officer<br />

Rabinowitz, M.<br />

Natera, Inc., Company officer<br />

Racowsky, C.<br />

Life Global Group, Paid consultant;<br />

UpToDate, Honoraria; World<br />

Health Organization, Paid<br />

consultant<br />

Rafizadeh, M. J. Colgate-Palmolive Company,<br />

Summer Internship Offer<br />

Raia, M. H.<br />

Counsyl (Genetic Counselor),<br />

Full-time company employee<br />

Reed, B. G.<br />

Sequenom, Direct stockholder<br />

Riboldi, M.<br />

Igenomix Brasil, Full-time company<br />

employee<br />

Riche, D. M.<br />

Merck, Novo Nordisk, Speakers<br />

bureau<br />

Rippon, G. A.<br />

Pfizer Inc., Full-time company<br />

employee<br />

Robinson, M. K.<br />

Thrive Rx, Paid consultant<br />

Robinson, R. D.<br />

AbbVie, Grant recipient<br />

Rodriguez, S.<br />

Recombine, Full-time company<br />

employee<br />

Rosen, K. A. Bayer Healthcare Pharmaceuticals<br />

Inc., Full-time company employee<br />

Ruman, J. Ferring Pharmaceuticals, Full-time<br />

company employee<br />

Ryan, A.<br />

natera, Full-time company employee<br />

Sadeghi-Nejad, H.<br />

Auxilium (now Endo), Grant<br />

recipient<br />

Sakkas, D. Ferring, Grant recipient; Origio,<br />

Scientific Advisory Board;<br />

Fertilitech, Scientific Advisory<br />

Board; Good Start Genetics, Paid<br />

consultant; INVO Biosciences,<br />

Direct stockholder<br />

Saldivar, S.<br />

Myriad Genetics, Speakers bureau;<br />

Intuitive Surgical Inc, Paid<br />

consultant<br />

FERTILITY & STERILITY Ò<br />

e385


Sammel, M. D.<br />

Swiss Precision Diagnostics, Paid<br />

consultant<br />

Santoro, N.<br />

Bayer Inc, Grant recipient;<br />

Menogenix Inc, Stock Options<br />

Sasaki, K.<br />

E-medicine, Honoraria<br />

Schattman, G. L.<br />

Femasys, medical advisor, clinical<br />

investigator; Theralogix, Paid<br />

consultant; Ferring, Speakers<br />

bureau<br />

Schmelter, T.<br />

Bayer Pharma AG, Full-time<br />

company employee<br />

Schoolcraft, W. B. OvaScience, Advisory Board<br />

Member; Ferring, Paid consultant;<br />

Serono, Paid consultant<br />

Schweitzer, A. C.<br />

Affymetrix, Full-time company<br />

employee<br />

Scott, R. T. Foundation for Assessment &<br />

Enhancement of Embryonic<br />

Competence, Inc.; Neither myself<br />

or my program get any personal<br />

benefit., Company officer; Ferring<br />

Pharmaceutical, Scientific<br />

Advisory Board<br />

Seifer, D.<br />

Rutgers Medical School/ MGH<br />

licensing aggreement with<br />

Beckman-Coulter, Co-inventor of<br />

AMH as a method of determining<br />

ovarian reserve; Beckman-Coulter,<br />

Paid consultant; Women’s<br />

Integrated Network, Paid<br />

consultant; none, none<br />

Shah, D.<br />

Grand Rounds Health, Paid<br />

consultant<br />

Shapiro, B. S.<br />

Actavis Inc., Grant recipient<br />

Shapiro, D. Actavis, Speakers bureau; Merck,<br />

Honoraria; EMD-Serono, Paid<br />

consultant; Ovascience, Paid<br />

consultant<br />

Sharara, F.<br />

MAP Diagnostics, Company officer;<br />

Ferring Pharmaceuticals, Speakers<br />

bureau<br />

Shawber, C. Eisai Pharmaceuticals, Columbia<br />

University has established<br />

a research agreement (CU12-3625<br />

Eisai Research Collaborative<br />

Agreement) with Dr. Kitajewski as<br />

the PI to develop of Notch1 decoys.<br />

This research agreement provided<br />

funds to me as an inventor and<br />

my lab<br />

Shin, D.<br />

Endo Pharmaceutical, Speakers<br />

bureau<br />

Shraga, R.<br />

Recombine, Full-time company<br />

employee; Recombine, Stock<br />

Options<br />

Shwayder, J. M.<br />

Cook Ob-Gyn, Royalties for<br />

SonoBiopsy catheter<br />

Sigurjonsson, S.<br />

Natera, Full-time company employee<br />

Silverberg, K.<br />

Abbvie, Honoraria; Good Start<br />

Genetics, Paid consultant; Actavis,<br />

Paid consultant; Serono, Paid<br />

consultant; Illumina, Paid<br />

consultant; Myriad Genetics, Paid<br />

consultant; Finox, Grant recipient;<br />

Auxogyn, Grant recipient<br />

Simon, A.<br />

Natera, Inc., Full-time company<br />

employee<br />

Simon, C. Equipo IVI, Direct stockholder;<br />

Igenomix SL, Direct stockholder;<br />

OVASCIENCE, Paid consultant;<br />

TEVA; EXCEMED; MSD;<br />

Ferring, Honoraria<br />

Slayden, O. D.<br />

Bayer Pharma AG, Berlin, Grant<br />

recipient<br />

Soh, Y. M.<br />

Vitrolife, Grant recipient<br />

Sood, A.<br />

Global Center for Resiliency and<br />

Wellbeing, Owner<br />

Stanczyk, F. Z. Merck & Co., Paid consultant;<br />

TherapeuticsMD, Paid consultant;<br />

Noven Pharmaceuticals, Paid<br />

consultant; Enteris Biopharma,<br />

Paid consultant; AbbVie, and<br />

Agile Therapeutics, Paid<br />

consultant<br />

Stecher, V.<br />

Pfizer Inc, Full-time company<br />

employee<br />

Stegmann, B. J.<br />

Merck & Co., Full-time company<br />

employee<br />

Steinkampf, M. P.<br />

AbbVie Pharmaceuticals, Speakers<br />

bureau<br />

Stewart, E. A.<br />

AbbVie, Astellas,Bayer,<br />

GlaxoSmithKline, Gynesonics,<br />

Welltwigs, Viteava, Paid<br />

consultant; UpToDate, Honoraria<br />

St-Michel, P.<br />

MD Serono, Paid consultant; Ferring<br />

Canada, Grant recipient<br />

Strachan, G.<br />

Teva Pharmaceuticals, Full-time<br />

company employee<br />

Stratton, P.<br />

Alllergan has provided a grant to my<br />

institution (NIH) for a study on<br />

which I am the medically<br />

responsible investigator. I have no<br />

direct of personal profit from this<br />

grant., Grant recipient<br />

Strumbly, D.<br />

Janssen Pharmaceuticals, Contract<br />

Employee<br />

Sung, L.<br />

Ferring, Grant recipient<br />

Surrey, E.<br />

AbbVie Laboratories, Advisory<br />

board, speaker’s bureau, grant<br />

recipient<br />

Swain, J. E.<br />

Irvine Scientific, royalties from<br />

a previously designed culture<br />

medium<br />

Swerdlow, A. J.<br />

GlaxoSmithKline, Direct stockholder<br />

Tan, L.<br />

Progyny (formerly Auxogyn), Fulltime<br />

company employee; Progyny<br />

(formerly Auxogyn), Direct<br />

stockholder<br />

Tang, B. Teva Pharmaceutical, Full-time<br />

company employee<br />

Taylor, R. N. AbbVie, Paid consultant; Ferring,<br />

Paid consultant<br />

Taylor, T. H. CombiMatrix, Full-time company<br />

employee<br />

Thomas, M. A.<br />

Medicines 360, Grant recipient;<br />

Agile, Grant recipient; Merck,<br />

Grant recipient; Bayer, Grant<br />

recipient<br />

Tjoa, M. L.<br />

Teva Pharmaceuticals, Full-time<br />

company employee<br />

Toth, T. L.<br />

Good Start Genetics, medical<br />

advisory board<br />

Treff, N. R.<br />

EMD Serono, Grant recipient<br />

Truong, T. T.<br />

Vitrolife AB, Grant recipient<br />

e386 Author Disclosures Index Vol. 104, No. 3, Supplement, September <strong>2015</strong>


Tulandi, T.<br />

Actavis Inc, Honoraria; AbbVie<br />

Canada, Honoraria<br />

Valbuena, D.<br />

Igenomix S.L., Full-time company<br />

employee<br />

Vaskivuo, T.<br />

Roche Diagnostics, Speakers bureau;<br />

Siemens Diagnostics, Speakers<br />

bureau<br />

VerMilyea, M.<br />

Irvine Scientific, Paid consultant;<br />

Progyny (formerly Auxogyn), Paid<br />

consultant<br />

Wang, R. AMS, Paid consultant; Coloplast,<br />

Paid consultant; Men MD, Paid<br />

consultant; Eli Lilly, Speakers<br />

bureau<br />

Ward, K. Affiliated Genetics, Inc.; Juneau<br />

Biosciences, Direct stockholder;<br />

Affiliated Genetics, Inc., Full-time<br />

company employee; Affiliated<br />

Genetics, Inc., Direct stockholder;<br />

Juneau Biosciences, Company<br />

officer; Juneau Biosciences, Fulltime<br />

company employee<br />

Warne, D. W. Finox, Paid consultant; PregLem,<br />

Paid consultant<br />

Webster, W.<br />

Prima-Temp, Inc, Company officer<br />

Welch, C. Reprogenetics, Full-time company<br />

employee<br />

Wellons, M.<br />

Pfizer, advisory board<br />

Wells, D. Reprogenetics, Direct stockholder;<br />

Illumina, Paid consultant; Ferring,<br />

Honoraria; Recombine, Direct<br />

stockholder; Merck Serono,<br />

Honoraria<br />

Widra, E. A.<br />

Counsyl, Paid consultant<br />

Willey, R. G.<br />

Ferring International Pharmascience<br />

Center US, Inc., Full-time<br />

company employee<br />

Williams, L. A.<br />

AbbVie,Inc, Full-time company<br />

employee<br />

Winston, N.<br />

Pfizer, Direct stockholder; Merck,<br />

Direct stockholder<br />

Wolff, E. F.<br />

OvaScience, Grant recipient<br />

Woodrum, D. A.<br />

BioSentry, Paid consultant<br />

Wright, K.<br />

Alcon, Full-time company employee<br />

Wu, H.<br />

Celmatix Inc., Full-time company<br />

employee<br />

Wyatt, M.<br />

Genesis Genetics, Full-time company<br />

employee<br />

Yang, Z.<br />

ZytoGen, Company officer<br />

Yankov, V. Ferring Pharmaceuticals, Full-time<br />

company employee<br />

Yarnall, S. Recombine, Full-time company<br />

employee; Recombine, Stock<br />

options<br />

Yurttas Beim, P.<br />

Celmatix Inc, Company officer<br />

Zajic, S.<br />

Merck & Co., Inc., Full-time<br />

company employee; Merck & Co.,<br />

Inc., Direct stockholder<br />

Zgodic, A.<br />

Celmatix, Full-time company<br />

employee<br />

Zhang, H.<br />

BMS, Full-time company employee<br />

Zimmerman, R.<br />

Foundation for Embryonic<br />

Competence, Full-time company<br />

employee; BioReference<br />

Laboratories, Full-time company<br />

employee<br />

FERTILITY & STERILITY Ò<br />

e387

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!