02.11.2014 Views

SNS07 - Abstracts pdf file 35 Mb

SNS07 - Abstracts pdf file 35 Mb

SNS07 - Abstracts pdf file 35 Mb

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.

CONSEJO SUPERIOR DE INVESTIGACIONES CIENTÍFICAS<br />

TRABAJOS<br />

DEL<br />

INSTITUTO CAJAL<br />

TOMO LXXXI<br />

C O N T I N U A C I Ó N D E L A “ R E V I S T A T R I M E S T R A L M I C R O G R Á F I C A ”<br />

F U N D A D A P O R S. RAMÓN Y CAJAL<br />

4 th INTERNATIONAL MEETING<br />

STEROIDS AND NERVOUS SYSTEM<br />

TORINO, Italy, Villa Gualino<br />

February 17 - 21, 2007<br />

ABSTRACTS OF INVITED LECTURES<br />

AND FREE CONTRIBUTIONS<br />

G.C. Panzica and S. Gotti, editors<br />

MADRID – 2007


Organizers<br />

Roberto C. Melcangi<br />

GianCarlo Panzica<br />

(Milano, Italy)<br />

(Torino, Italy)<br />

International Scientific Committee<br />

Jacques Balthazart<br />

Luis M. García-Segura<br />

Allan E. Herbison<br />

Margaret McCarthy<br />

Roberto C. Melcangi<br />

GianCarlo Panzica<br />

Belgium<br />

Spain<br />

New Zealand<br />

USA<br />

Italy<br />

Italy<br />

Educational Committee<br />

Cheryl Frye<br />

USA<br />

Local Organizing Committee<br />

Aldo Fasolo<br />

Carla Viglietti Panzica<br />

Francesca Allieri<br />

Elisabetta Bo<br />

Daniela Grassi<br />

Stefano Gotti<br />

Mariangela Martini<br />

Désirée Miceli<br />

Elena Mura<br />

Honour Committee<br />

Ezio Pelizzetti<br />

Alberto Conte<br />

Aldo Fasolo<br />

Rector of the University of Torino<br />

Dean of the Faculty of Science<br />

President of the Research Committee of the University of Torino<br />

Visit us at our WWW site<br />

http://www.dafml.unito.it/anatomy/panzica/neurosteroids/<br />

The <strong>Abstracts</strong> published here were reproced directly from author’s original text with little or no<br />

alteration. The Editors take no responsibility for their content.


Contents<br />

• Selective estrogen receptors modulators and the brain 5<br />

• New perspectives in the dosage of neuroactive steroids 17<br />

• Plenary lecture: Herbison A.E. 31<br />

• Effects mediated by classical steroid receptors <strong>35</strong><br />

• Round table I: Steroid hormones and sexually dimorphic brain circuits 49<br />

• Neuroactive steroids and neurogenesis 61<br />

• Plenary lecture: Mellon S.H. 71<br />

• Neuroprotective effects 75<br />

• Xenoestrogens and brain circuitries 87<br />

• Young investigators symposium 95<br />

• Effects mediated by membrane receptors 109<br />

• Plenary lecture: Swaab D.F. 123<br />

• Corticosteroid effects and stress 127<br />

• Posters’ exhibition 136


SATURDAY, 17 th February 2007<br />

10.00 – 13.00<br />

Satellite Symposium:<br />

Selective estrogen receptors<br />

modulators and the brain


Satellite Symposium:<br />

Selective estrogen receptors modulators and the brain<br />

(Organizers: Marchetti B., Panzica G.C.)<br />

• R.D. Brinton, Zhao L. (USA) Mechanisms of neuroprotection by SERMs in the<br />

brain<br />

• Di Paolo T., Bourque M., Liu B., Dluzen D.E., Morissette M. (Canada)<br />

Tamoxifen and raloxifene neuroprotection in Parkinson's disease models: examples<br />

of MPTP and methamphetamine toxicities<br />

• Garcia-Segura L.M., Tapia González S, Diz-Chaves Y, Pernía O, Carrero P,<br />

Ciriza I (Spain) Selective estrogen receptor modulators, neuroprotection and glial<br />

cells<br />

• Frye C.A., Walf, A.A. (USA) Estrogen receptor beta is a target of estrogens’ and<br />

androgens’ for affective and cognitive behavior<br />

• Marchetti B, L’Episcopo F, Tirolo C, Testa N, Caniglia S, Giaquinta G, Gennuso<br />

F, Arcieri P, Serra P-A, Desole MS, Miele E, Delitala G, Morale MC. (Italy)<br />

Neuroimmune interactions and estrogen deficiency: innate immunity as a doubleedged<br />

sword in neurodegeneration and repair<br />

• Tena-Sempere M. (Spain) Effects of selective ER ligands and modulators on the<br />

gonadotropic axis


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

MECHANISMS OF NEUROPROTECTION BY SERMS IN THE BRAIN<br />

Brinton R.D. and Zhao L.<br />

Department of Pharmacology and Pharmaceutical Sciences, School of Pharmacy, and<br />

Neuroscience Program, University of Southern California, Los Angeles, CA, USA<br />

Our scientific endeavors are a hybrid of basic science discovery and preclinical<br />

translational research. The goal of our basic science discovery is to elucidate fundamental<br />

cellular mechanisms of 1) neural defense and repair and 2) neural plasticity required for<br />

cognitive function. Our therapeutic development goal is to translate our cellular<br />

mechanistic insights into safe and efficacious therapeutics for the prevention of and<br />

rehabilitation from neurodegenerative diseases, such as Alzheimer’s, Parkinson’s and<br />

stroke. To achieve these goals, we have investigated the mechanisms and neurobiological<br />

outcomes of estrogens, progestins, hormone therapies and neurosteroids. Results of these<br />

analyses have yielded insights into cellular strategies required for neural defense against<br />

degenerative insults that involve multifaceted cytoplasmic and nuclear signaling cascades<br />

that converge upon the mitochondria. Further, these signaling cascades are required for<br />

gonadal hormone regulation of morphogenesis and neurogenesis. Our data indicate a<br />

healthy cell bias of estrogen action for estrogen-inducible neuroprotective and neurotrophic<br />

outcomes. Our translational therapeutic development efforts target sites of estrogen action<br />

that promote neural defense and plasticity in brain while reducing untoward effects in<br />

reproductive tissue. Our novel NeuroSERM molecules are designed to target the<br />

membrane site of estrogen action whereas our natural source PhytoSERMs are designed to<br />

target estrogen receptor beta. Results of our in vitro analyses indicate that our novel<br />

NeuroSERM candidate molecule induces neuroprotective efficacy comparable to 17 β-<br />

estradiol. In addition, NeuroSERM candidate molecule induces neuronal plasticity markers<br />

consistent with 17 β-estradiol. In parallel, PhytoSERM formulations show a high degree of<br />

estrogenic efficacy in cultured hippocampal and cortical neurons and induces both<br />

neuroprotective and neural plasticity reponses. Results of in vitro analyses provide proof of<br />

principle that combination of ERβ selective NeuroSERM and PhytoSERMs provide<br />

neuroprotective efficacy which will be extended into in vivo analyses. The data thus far<br />

suggest that novel NeuroSERM molecules and select combinations of PhytoSERMs have<br />

the potential to be effective and safe estrogen alternative therapies for preventing estrogen<br />

deficiency-related cognitive decline and vulnerability to neurodegenerative insults such as<br />

those leading to Alzheimer’s disease in postmenopausal women.<br />

Acknowledgements: This work was supported by a grant from the Alzheimer’s Association<br />

to LZ and by the Kenneth T. and Eileen L. Norris Foundation to RDB<br />

7


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

TAMOXIFEN AND RALOXIFENE NEUROPROTECTION IN PARKINSON'S<br />

DISEASE MODELS: EXAMPLES OF MPTP AND METHAMPHETAMINE<br />

TOXICITIES<br />

Di Paolo T. 1 , Bourque M. 1 , Liu B. 2 , Dluzen D.E. 2 , and Morissette M. 1<br />

1 Molecular Endocrinology and Oncology Research Center, Laval University Medical<br />

Center, CHUL, Quebec City, G1V 4G2 and Faculty of Pharmacy, Laval University,<br />

Quebec City, G1K 7P4, Quebec, Canada.<br />

Fax: 418-654-2761, e-mail: therese.dipaolo@crchul.ulaval.ca<br />

2 Department of Anatomy, Northeastern Ohio Universities College of Medicine<br />

(NEOUCOM), Rootstown, Ohio 44272, USA.<br />

Parkinson’s disease (PD) is a common neurodegenerative disorder characterized by a<br />

selective depletion of dopamine (DA) neurons in the substantia nigra (SN). Accumulating<br />

evidence support a gender difference in PD with a relative risk 1.5 times greater in men<br />

[3,6]. Gender differences in the evolution of symptoms and responses to L-dopa are also<br />

reported [3]. A neuroprotective role of estrogens on DA activity is likely a contributing<br />

factor but the Women’s Health Initiative (WHI) study reported increased risk over benefits<br />

of equine estrogens in post-menopausal women [4]. Thus, it becomes of great interest to<br />

identify alternative estrogens with neuroprotectant potential. Selective estrogen receptor<br />

modulators (SERMs) could be an alternative for both men and women. The SERM,<br />

tamoxifen, is the most widely prescribed medication for prevention and treatment of breast<br />

cancer; it has estrogen antagonist activity in mammary tissue while it mimics the effects of<br />

estrogen in other tissues [1]. The second generation SERM, raloxifene, is an estrogen<br />

antagonist in mammary and uterine tissues while it is an estrogen agonist in bone and<br />

cholesterol metabolism [1]; it treats osteoporosis and is used in menopausal women.<br />

Tamoxifen and raloxifene have in vivo affinity for both estrogen receptors (ERs) [5].<br />

It is well documented that estrogens modulate rat DA receptors and membrane DA<br />

transporters (DAT). We characterized the estrogenic specificity of this modulation by<br />

testing various estrogenic drugs. Ovariectomy decreased DA D2 receptor and DAT<br />

specific binding in the striatum. Estradiol and raloxifene, but not tamoxifen treatment<br />

prevented this decrease. Estradiol and raloxifene treatment decreased DA D3 receptor<br />

binding in the islands of Calleja, the nucleus accumbens and striatum, compared to<br />

ovariectomized rats. The ERbeta agonist 2,3-bis(4-hydroxyphenyl)-propionitrile (DPN),<br />

but not the ERalpha agonist 4,4’,4’’-(4-Propyl-[1H]-pyrazole-1,3,5-triyl)trisphenol (PPT),<br />

mimicked the estradiol increase of D2 receptor and DAT specific binding. Tamoxifen and<br />

raloxifene also corrected the decrease of DAT specific binding caused by ovariectomy.<br />

Neither ovariectomy nor estrogenic treatments modulated striatal specific binding to the<br />

monoamine vesicular transporters 2 (VMAT2). These results suggest that ERbeta mediates<br />

estradiol and SERMs increase of striatal D2 receptors and DAT.<br />

Similar gender differences are observed in animal models of PD with male mice<br />

displaying more nigrostriatal DA neuronal loss than females after administration of the<br />

toxins 1-methy-4-phenyl-1,2,3,6-tetrahypdropyridine (MPTP) [2] or methamphetamine<br />

(MA) [7]. Estradiol is neuroprotective against these toxicities. The prevention of MPTPinduced<br />

DA depletion in the striatum of male mice by estradiol is mimicked by raloxifene<br />

but not tamoxifen. Moderate doses of MPTP induce a decrease of striatal DAT specific<br />

binding (50% of control) and DAT mRNA in the SN (20% of control), suggesting that loss<br />

of neuronal nerve terminals is more extensive than cell bodies. The MPTP-induced<br />

decrease of striatal DAT specific binding is prevented by estradiol or raloxifene<br />

8


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

(5mg/kg/day) but not by a lower dose of raloxifene (1mg/kg/day). Striatal DAT specific<br />

binding is positively correlated with DA concentrations in intact, saline or hormone treated<br />

MPTP mice. This paradigm models early DA nerve cell damage and is responsive to<br />

hormones.<br />

Protective properties of tamoxifen against MA toxicity were compared to estradiol.<br />

Estradiol or tamoxifen were administered (estradiol: 1, 10 or 40 µg; tamoxifen: 12.5, 125<br />

or 500 µg) 24 hours prior to a MA injection (40mg/kg) in ovariectomized mice. Both<br />

treatments, at all concentrations, prevented the MA-induced decrease of striatal DA<br />

concentrations and VMAT2-binding. Only estradiol prevented the loss of DAT-binding in<br />

the lateral striatum and attenuated the MA-induced increase in striatal preproenkephalin<br />

(PPE) mRNA levels (at 1 or 40 µg). While both treatments prevented the DA decrease,<br />

estradiol protected more efficiently other dopaminergic parameters suggesting that overall<br />

it is more effective than tamoxifen as a neuroprotectant in female mice. By comparison,<br />

male mice are protected against MA toxicity by tamoxifen and not by estradiol. Intact male<br />

mice were administered 12.5 or 50 µg tamoxifen 24 hours before MA (40 mg/kg)<br />

treatment. MA reduced striatal DA concentration and its metabolites 3,4-<br />

dihydroxyphenylacetic acid and homovanillic acid, striatal and SN DAT and VMAT2<br />

specific binding as well SN DA transporter mRNA levels whereas tyrosine hydroxylase<br />

and VMAT2 mRNA levels were not significantly decreased. These MA effects were not<br />

altered by 12.5 µg tamoxifen except for increased striatal DA metabolites and turnover.<br />

Tamoxifen at 50 µg reduced the MA effect on striatal DA concentration and DA<br />

transporter specific binding; in the SN tamoxifen prevented the decrease of DA transporter<br />

mRNA levels. The present results show a dose–dependent protection of tamoxifen against<br />

MA-induced DA toxicity in male mice. MA is a potent and addictive psychostimulant of<br />

increasing use in humans and toxic to striatal nerve terminals. Tamoxifen is the only<br />

known hormonal protection of male mice against MA toxicity thus providing important<br />

new information on specific parameters of nigrostriatal dopaminergic function preserved<br />

by this SERM.<br />

Tamoxifen and raloxifene thus modulate and protect specific parameters of<br />

nigrostriatal dopaminergic function with similarities and differences compared to estradiol.<br />

Reference list<br />

[1] TA Grese, JA Dodge JA Selective estrogen receptor modulators (SERMs). Curr Pharm Des.<br />

4(1998):71-92.<br />

[2] D.B. Miller, S.F. Ali, J.P. O'Callaghan, S.C. Law, The impact of gender and estrogen on striatal<br />

dopaminergic neurotoxicity. Ann N Y Acad Sci, 844 (1998) 153-165.<br />

[3] L.M. Shulman, V. Bhat, Gender disparities in Parkinson’s disease, Expert Rev Neurother 6 (2006)<br />

407-416.<br />

[4] S.A. Shumaker, C. Legault, S.R. Rapp, L. Thal L, R.B. Wallace, J.K. Ockene, S.L. Hendrix, B.N.<br />

Jones 3rd, A.R. Assaf, R.D. Jackson, J.M. Kotchen, S. Wassertheil-Smoller, J. Wactawski-Wende;<br />

WHIMS Investigators. Estrogen plus progestin and the incidence of dementia and mild cognitive<br />

impairment in postmenopausal women : the Women’s Health initiative memory study: a randomized<br />

controlled trial JAMA 289 (2003) 2651-2662.<br />

[5] R.V. Weatherman., N.J. Clegg, T.S. Scanlan, Differential SERM activation of the estrogen receptors<br />

(ERalpha and ERbeta) at AP-1 sites. Chem Biol, 8 (2001) 427-36.<br />

[6] G.F.Wooten, L.J. Currie, V.E. Bovbjerg, J.K. Lee, J. Patrie, Are men at greater risk for Parkinson’s<br />

disease than women?, J Neurol Neurosurg Psychiatry 75 (2004) 637-639.<br />

[7] L. Yu, P.C. Liao, Sexual differences and estrous cycle in methamphetamine-induced dopamine and<br />

serotonin depletions in the striatum of mice. J Neural Transm, 107 (2000) 1139-1147.<br />

9


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

SELECTIVE ESTROGEN RECEPTOR MODULATORS, NEUROPROTECTION<br />

AND GLIAL CELLS<br />

Garcia-Segura L.M., Tapia González S., Diz-Chaves Y., Pernía O., Carrero P., and<br />

Ciriza I.<br />

Instituto Cajal, CSIC, Avenida Doctor Arce 37, E-28002 Madrid, Spain.<br />

E-mail: lmgs@cajal.csic.es; Fax: +34-915854754.<br />

Neuroprotective effects of estradiol are well characterized in animal experimental<br />

models. However, in humans, the outcome of estrogen treatment for cognitive function and<br />

neurological diseases is very controversial. Selective estrogen receptor modulators<br />

(SERMs) may represent an alternative to estrogen for the treatment or the prevention of<br />

neurodegenerative disorders. SERMs interact with the estrogen receptor and have tissuespecific<br />

effects distinct from those of estradiol, acting as estrogen agonists in some tissues<br />

and as antagonists in others. Previous studies have shown that some SERMs are<br />

neuroprotective. However, it is important to understand the mechanisms of action of these<br />

compounds before considering their possible therapeutic use for the treatment of<br />

neurodegenerative diseases. In our laboratory we are currently exploring the cellular<br />

targets that may be related with the neuroprotective actions of SERMs. Astroglia and<br />

microglia seem to play an important role in the neurodegenerative response since both cell<br />

types acquire a reactive phenotype after neural injury. To determine the effect of SERMs<br />

on glia activation we have assessed the effect of tamoxifen, raloxifene, lasofoxifene (CP-<br />

336,156), bazedoxifene (TSE-424) and estradiol in two experimental models: (i), the<br />

systemic administration of kainic acid (KA), an excitotoxin that induces neuronal loss and<br />

glial activation in the hippocampus and (ii), the systemic administration of<br />

lipopolysaccharide (LPS) from Escherichia coli, which induces microglia activation in the<br />

brain. Estradiol exerted dose-dependent neuroprotective effects, preventing neuronal loss<br />

in the hilus of the dentate gyrus of the hippocampal formation of adult ovariectomized rats<br />

injected with KA. In addition, estradiol reduced the activation of astrocytes and microglia<br />

in the hilus, assessed by the expression of vimentin and major histocompatibility complex<br />

class II (MHCII), respectively. Estradiol also reduced the number of MHCII microglia,<br />

assessed in the white matter of the cerebellum, after LPS administration. Tamoxifen,<br />

raloxifene and bazedoxifene prevented neuronal loss in the hilus after the administration of<br />

KA in a dose dependent manner and reduced microglia activation after KA or LPS<br />

administration. In contrast, none of these SERMs had a significant effect on the activation<br />

of astroglia. Lasofoxifene was not neuroprotective and did not affect glial activation. These<br />

findings suggest that SERMs act with cellular specificity in the brain and may<br />

differentially modulate the activation of microglia and astroglia. Reactive glial cells exert a<br />

mixture of positive and negative responses for neuronal survival and regeneration.<br />

Reactive astroglial cells release many different factors that promote neuronal survival and<br />

participate in the formation of the glial scar, which is important to maintain water<br />

homeostasis but prevents axonal regeneration. Reactive microglial cells exert important<br />

positive functions by remodeling the damaged tissue. However, they also release<br />

proinflammatory cytokines and their prolonged activation may exacerbate neuronal<br />

damage. Therefore, the selective action of tamoxifen, raloxifene and bazedoxifene on<br />

astroglia and microglia may be highly relevant for the control of neural damage after<br />

injury.<br />

Supported by Ministerio de Educación y Ciencia, Spain (SAF 2005-00272) and the<br />

European Union (EWA project: LSHM-CT-2005-518245).<br />

10


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

ESTROGEN RECEPTOR BETA IS A TARGET OF ESTROGENS’ AND ANDROGENS’<br />

FOR AFFECTIVE AND COGNITIVE BEHAVIOR<br />

Frye C.A. 1-4 , and Walf A.A. 1<br />

Dept. of Psychology 1 , Biological Sciences 2 , and The Centers for Neuroscience 3 and Life Science 4 Research<br />

The University at Albany-SUNY, Life Sciences 01058, 1400 Washington Avenue, Albany, NY USA 12222;<br />

cafrye@albany.edu, 518-591-8823<br />

Estrogens, such as 17β-estradiol (E 2 ), and androgens, such as the 5α-reduced<br />

metabolite of testosterone, 3α-androstanediol (3α-diol), have anti-anxiety, antidepressant-like,<br />

and cognitive-enhancing effects in female and male rodents. Our<br />

laboratory has been investigating the β isoform of the estrogen receptor (ER) as a putative<br />

target for these effects. To accomplish this, we have utilized three approaches. First, the<br />

effects of systemic or intra-brain administration of selective ER modulators (SERMs) or<br />

selective androgen receptor modulators (SARMs) when administered to female and male<br />

rodents, respectively, on anxiety, depression, and cognitive behavior were assessed.<br />

Second, systemic or intra-brain administration of ER antagonists or ERα and/or ERβ<br />

antisense oligonucleotides to rodents administered SERMs or SARMs was utilized to<br />

determine effects of blocking or knocking down ERs on anxiety, depression, and cognitive<br />

behavior. Third, the effects of SERMs or SARMs administration to mice with targeted<br />

deletions of ERβ (BERKO) on anxiety, depression, and cognitive behavior was<br />

investigated. Utilizing these approaches has revealed the importance of steroids’ actions at<br />

ERβ in the hippocampus for anxiety, depression, and cognitive behavior.<br />

Whether E 2 ’s effects among female rodents require activity at ERβ was<br />

investigated. First, ovariectomized (ovx) rats were administered subcutaneous (SC) 17β-<br />

E 2 (equal affinity for ERα and ERβ), ERα-selective SERMs (PPT), ERβ-selective SERMs<br />

(coumestrol; DPN) or vehicle before testing in anxiety (open field, elevated plus maze) and<br />

depression (forced swim test) tasks, or following training in cognitive (object recognition,<br />

inhibitory avoidance, water maze) tasks. Subcutaneous administration of ERβ-specific<br />

SERMs, compared to vehicle, decreased anxiety (i.e. increased open field central entries,<br />

increased open arm time in the plus maze), decreased depression (decreased time immobile<br />

in the forced swim test) and improved performance in the object recognition, inhibitory<br />

avoidance, and watermaze tasks [6,8,12,13]. Similar anti-anxiety and anti-depressive<br />

effects were observed in ovx rats administered ERβ-selective SERMS directly to the<br />

hippocampus [10]. Second, the effects of blocking or knocking down ERs for anxiety,<br />

depression, and cognitive (conditioned place preference) behavior were determined.<br />

Administration of ER antagonists, SC or to the hippocampus, produce similar effects to<br />

attenuate the anti-anxiety and anti-depressant-like effects of SC 17β-E 2 or ERβ-SERMs<br />

administration [11,12]. SC or intra-nucleus accumbens administration of ER antagonists<br />

or antisense oligonucleotides attenuate E 2 -induced conditioned place preference [8]. These<br />

data suggest that ERs are important for the functional effects of E 2 , but the ER antagonists<br />

utilized in these studies were not ER isoform (ERα or ERβ)-specific. Therefore, the effects<br />

of infusions of ERα and/or ERβ antisense oligonucleotides to the lateral ventricle in ovx<br />

rats administered SC 17β-E 2 for anxiety and depression behavior was investigated.<br />

Infusions of ERβ, but not ERα, antisense oligonucleotides attenuates the anti-anxiety and<br />

anti-depressive effects of SC injections of E 2 to ovx rats and reduces ERβ expression in the<br />

hippocampus [7]. Third, the effects of natural fluctuations in E 2 , and the effects of SC 17β-<br />

E 2 or ERβ-SERMs administration to, BERKO mice for anxiety, depression, and cognitive<br />

performance were investigated. These studies revealed that BERKO mice do not respond<br />

11


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

like WT mice to endogenous variations in, and exogenous administration of, E 2 . Proestrous<br />

(high E 2 ) WT mice had decreased anxiety and depression behavior and improved cognitive<br />

performance across a variety of tasks compared to diestrous (low E 2 ) WT mice. This<br />

estrous cycle effect was not observed in ΒERKO mice. Furthermore, WT, but not BERKO,<br />

mice administered E 2 or DPN had decreased anxiety and depression behavior and<br />

enhanced learning across a number of tasks compared to ovx mice administered vehicle.<br />

Thus, these data suggest that ERβ is required for the anti-anxiety and anti-depressant-like<br />

and cognitive-enhancing effects of E 2 among female rodents.<br />

Whether androgens’s effects for anxiety and cognitive behavior among male<br />

rodents require activity at ERβ was investigated. First, gonadectomized (gdx) rats were<br />

administered SC SARMs that have different affinity for ERβ and anxiety and depression<br />

behavior were assessed. SC SARMs with high affinity for ERβ (3α-diol and 3β-diol)<br />

decreased anxiety in the open field and elevated plus maze, compared to vehicle or a<br />

SARM with low affinity for ERβ (androsterone). SC 3α-diol enhanced performance in the<br />

object recognition, conditioned contextual fear, conditioned place preference, and<br />

inhibitory avoidance tasks [3-5]. There are similar effects of 3α-diol when administered SC<br />

or to the hippocampus of gdx rats [2-3]. Second, the effects of knocking down ERβ for<br />

cognitive performance was assessed. Compared to vehicle, scrambled, or ERα antisense<br />

oligonucleotides, infusions of ERβ antisense oligonucleotides to the hippocampus of 3αdiol-administered<br />

rats attenuated performance in the inhibitory avoidance task [1]. Third,<br />

the effects of SARMs administration to gdx BERKO mice was investigated. SC 3β-diol to<br />

WT, but not BERKO, mice decreased anxiety behavior compared to that seen with vehicle.<br />

Thus, androgens may require ERβ for their anti-anxiety and cognitive-enhancing effects.<br />

In summary, ERβ is a putative target for the anti-anxiety, anti-depressant-like, and<br />

cognitive-enhancing effects of estrogens in females and androgens in males.<br />

Supported by: NSF (IBN03-16083) U.S.A. Army Dept. of Defense (BC051001).<br />

Reference List<br />

[1] K.L., Edinger, C.A. Frye, Androgens' effects to enhance learning may be mediated in part through actions at<br />

estrogen receptor-β in the hippocampus, Neurobiol Learn Mem. 87 (2007) pp. 78-85.<br />

[2] K.L. Edinger, C.A. Frye, Testosterone's anti-anxiety and analgesic effects may be due in part to actions of its 5α -<br />

reduced metabolites in the hippocampus, Psychoneuroendocrinology. 30 (2005) pp. 418-30.<br />

[3] K.L. Edinger, C.A. Frye, Testosterone's analgesic, anxiolytic, and cognitive-enhancing effects may be due in part to<br />

actions of its 5α -reduced metabolites in the hippocampus. Behav Neurosci. 118 (2004) 1<strong>35</strong>2-64.<br />

[4] K.L. Edinger, B. Lee, C.A. Frye, Mnemonic effects of testosterone and its 5α-reduced metabolites in the<br />

conditioned fear and inhibitory avoidance tasks, Pharmacol Biochem Behav. 78 (2004) pp. 559-68.<br />

[5] C.A. Frye, Some rewarding effects of androgens may be mediated by actions of its 5alpha-reduced metabolite 3αandrostanediol,<br />

Pharmacol Biochem Behav. (in press).<br />

[6] M.E. Rhodes, C.A. Frye, ERβ -selective SERMs produce mnemonic-enhancing effects in the inhibitory avoidance<br />

and water maze tasks. Neurobiol Learn Mem. 85 (2006) pp. 183-91.<br />

[7] A.A. Walf , I. Ciriza, L.M. Garcia-Segura, C.A. Frye, , Antisense oligodeoxynucleotides for estrogen receptor β and<br />

α attenuate estrogen’s modulation of affective and sexual behavior, respectively, Neuropsychopharmacology (in<br />

revision).<br />

[8] A.A. Walf, M.E. Rhodes, C.A. Frye, Ovarian steroids enhance object recognition in naturally cycling and<br />

ovariectomized, hormone-primed rats, Neurobiol Learn Mem. 86 (2006) pp. <strong>35</strong>-46.<br />

[9] A.A. Walf, M.E. Rhodes, J.R. Meade, J.P. Harney, C.A. Frye, Estradiol-induced conditioned place preference may<br />

require actions at estrogen receptors in the nucleus accumbens, Neuropsychopharmacology (in press).<br />

[10] A.A. Walf, C.A., Frye, Administration of estrogen receptor β-specific selective estrogen receptor modulators to the<br />

hippocampus decrease anxiety and depressive behavior of ovariectomized rats, Pharmacol Biochem Behav. (in<br />

press).<br />

[11] A.A. Walf, C.A., Frye, A review and update of mechanisms of estrogen in the hippocampus and amygdala for<br />

anxiety and depression behavior. Neuropsychopharmacology 31 (2006) pp. 1097-111.<br />

[12] A.A. Walf, C.A., Frye, ERβ-selective estrogen receptor modulators produce antianxiety behavior when<br />

administered systemically to ovariectomized rats. Neuropsychopharmacology.30 (2005) pp. 1598-609.<br />

[13] A.A. Walf, M.E. Rhodes, C.A. Frye, Anti-depressant effects of ERβ selective estrogen receptor modulators in the<br />

forced swim test. Pharmacol Biochem Behav. 78 (2004) pp. 523-9.<br />

12


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

NEUROIMMUNE INTERACTIONS AND ESTROGEN DEFICIENCY: INNATE<br />

IMMUNITY AS A DOUBLE-EDGED SWORD IN NEURODEGENERATION AND<br />

REPAIR<br />

Marchetti B. 1,2 , L’Episcopo F. 1,2 , Tirolo C. 1 , Testa N. 1 , Caniglia S. 1 , Giaquinta G. 1,2 ,<br />

Gennuso F. 1 , Arcieri P. 1 , Serra P.-A. 1 , Desole M.S. 1 , Miele E. 1 , Delitala G. 3 , Morale<br />

M.C. 1<br />

1 Neuropharmacology, OASI Institute for Research and Care (IRCCS), Troina (EN),<br />

2 Department of Pharmacology and 3 Endocrinologic Clinic, University of Sassari, Italy;<br />

bianca.marchetti@oasi.en.it<br />

Over the past decade neuroinflammation has been increasingly recognized as a crucial<br />

contributory factor to neurodegeneration in normal aging, as well as in age-related<br />

neurodegenerative diseases, including Alzheimer’s and Parkinson’s diseases [1]. Hence, the<br />

therapeutic potential of steroidal and non-steroidal anti-inflammatory drugs (NSAIDs) for treating<br />

various neurodegenerative diseases including PD has recently been emphasized by various<br />

laboratories including our own [1]. Neurodegeneration caused by chronic inflammation involves<br />

activation of astrocytes and the brain’s resident immune cells, the microglia, which produce a large<br />

number of pro-inflammatory factors. Also acute brain insults, e.g. stroke and traumatic brain injury,<br />

are linked to inflammation, which contributes to the propagation of the neuropathological events<br />

[1,2]. Systemic inflammation may also impact on local inflammation in the diseased brain, leading<br />

to over-production of inflammatory mediators, which may, in turn, influence neuron survival,<br />

plasticity and repair [1,2]. Inflammation triggered by brain injury may either activate or impair<br />

neurogenesis, depending on the type of neuronal insult (i.e acute vs chronic) and degree of glia<br />

activation. Of particular mention, there are critical interactions between the hypothalamic-pituitary<br />

gonadal (HPG) and the hypothalamic-pituitary-adrenocortical (HPA) axes in modulating both<br />

central and systemic inflammation, with important implications for the brain’s ability to mount an<br />

efficient beneficial and protective response to injury [3-7]. Then, the net effect of glial-mediated<br />

inflammation may be beneficial, depending on quantitative and qualitative aspects of astrocytes<br />

and microglia activation, specificity of the brain region affected, coupled to sex, age, genetic<br />

predisposition and a number environmentally-mediated factors, which may, in turn dictate the<br />

severity of a neuronal insult, and/or influence the brain’s ability to recover. Indeed, crosstalk<br />

between blood-born cells, astrocytes and microglia uniquely contribute to an inflammatory<br />

signature critically modulated by the sex steroid hormone milieu [3-7]. Both experimental and<br />

clinical evidences clearly document that women are protected from many forms of neurological<br />

injury, in part attributable to endogenous estrogen. Here we emphasize estrogen modulation of<br />

central and peripheral innate mechanisms as potential contributing factors [3-7]. The onset of<br />

menopause is associated with spontaneous increases in cytokine production, whereas cytokine<br />

levels are lower in post-menopausal women on hormone replacement therapy and in estrogentreated<br />

ovariectomized rodents. However, estrogen background may influence different types of<br />

pro- and anti-inflammatory molecules in complex and distinct ways. Thus, whether estrogen and<br />

selective estrogen receptor modulators (SERMs) effects on inflammation have therapeutical<br />

potentials for neuroprotection, still remains to be elucidated. In our laboratory we have focused on<br />

the nigrostriatal dopaminergic (DA) neurons, the cells that degenerate in Parkinson’s disease (PD).<br />

The main pathological hallmark of PD is a selective and progressive death of DA neurons in the<br />

substantia nigra pars compacta (SNpc) leading to loss of DA afferents to basal ganglia and motor<br />

dysfunction, including tremor, rigidity, and bradykinesia. Idiopathic PD is one of the most common<br />

ageing-associated neurodegenerative disorders, with a prevalence in male gender. Using the MPTP<br />

mouse model of PD we documented activated astrocyte, macrophages and microglial cells as key<br />

elements of the neuroendocrine-immune dialogue involved in steroid hormone programming of<br />

brain response to neurotoxic challenge [4-7]. Hence, glia and its pro-/anti-inflammatory mediators<br />

were shown to represent a final common pathway through which genetic, hormonal and<br />

environmental influences modulate resistance or susceptibility to experimental PD [4-7].<br />

13


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

Specifically, we have presented an hypothetical model whereby efficiency of the neuroimmune<br />

dialogue via steroid receptor-nitric oxide crosstalk represents a critical step in dictating<br />

susceptibility versus resistance to Parkinson’s and possibly other degenerative CNS diseases [4-7].<br />

Here we highlight neuroinflammation as a window to develop therapeutical strategies for<br />

neuroprotection and describe our experimental paradigm based on the dualistic role of<br />

inflammation in neurodegeneration and repair. In our approach, we investigate both central and<br />

systemic innate mechanisms underlying the selective vulnerability/resistance of nigral DA neurons<br />

to MPTP-induced PD in wild type and genetically modified mice, a. using different schedules and<br />

dose-regimens of the neurotoxic challenges to more closely mimick the slowly progressing late<br />

onset nature of PD; b. by studying the impact of aging, sex and estrogen deficiency in the crosstalk<br />

between systemic and central innate and adaptive response to brain injury; c. by investigating how<br />

these interactions may impact in the severity of the nigrostriatal lesion, in motor behaviour and<br />

neurogenesis. One aspect regards the identification of immununogenetic and inflammatory markers<br />

for early detection of neuronal damage, as indicators of disease onset and progression. In this<br />

ongoing research program integrating both in vivo findings with in vitro data and molecular<br />

screening in the search for ideal pharmacological targets, which include steroids, NSAIDs and<br />

SERMs, the overall data point to anti-inflammatory pathways amplifying drugs (AIPADs) as<br />

candidate disease modifying therapeutics. Cytokines variations in human peripheral blood<br />

monocytes (PBMC) or cytokine polymorphisms, which may influence susceptibility to neuronal<br />

damage, are also being actively investigated as diagnostic/prognostic indicators of disease<br />

progression [8,9]. In particular, early markers of neuronal deterioration may be exploited for<br />

identifying high-risk subpopulations and for assessing the usefulness of timely (combination)<br />

treatments [8,9]. Efforts aimed at identify integrated approaches will hopefully modify the natural<br />

course of the disease, by either preventing, slowing progression, or halting neuronal degeneration.<br />

Supported by Italian Ministry of Health (National Research Project RF-05/82) and Italian Ministry<br />

of Research.<br />

References List<br />

[1] Marchetti B and Abbracchio MP. To be or not to be (inflamed)--is that the question in anti-inflammatory drug<br />

therapy of neurodegenerative disorders? (2005) Trends Pharmacol Sci. Oct; 26(10):517-25.<br />

[2] Marchetti B, Kettenmann H, Streit WJ (Eds) (2005b) Glia-neuron crosstalk in neuroinflammation,<br />

neurodegeneration and neuroprotection. Brain Res Reviews Special Issue vol 48/2: 129-408.<br />

[3] Morale MC, Gallo F, Tirolo C, Testa N, Caniglia S, Marletta N, Spina-Purrello V, Avola R, Caucci F, Tomasi P,<br />

Delitala G, Barden N, Marchetti B. (2001) Neuroendocrine-immune (NEI) circuitry from neuron-glial interactions<br />

to function: Focus on gender and HPA-HPG interactions on early programming of the NEI system. Immunol Cell<br />

Biol. Aug 79(4):400-17.<br />

[4] Marchetti B, Serra PA, L’Episcopo F, Tirolo C, Caniglia S, Testa N, Cioni S, Gennuso F, Rocchitta G, Desole MS,<br />

Mazzarino MC, Miele E, Morale MC. (2005) Hormones are key Actors in gene x environment interactions<br />

programming the vulnerability to Parkinsons disease: Glia as a common final pathway. Ann. NY Acad. Sci.; 1057:<br />

296-318.<br />

[5] Morale MC, Serra PA, L'episcopo F, Tirolo C, Caniglia S, Testa N, Gennuso F, Giaquinta G, Rocchitta G, Desole<br />

MS, Miele E, Marchetti B. (2006) Estrogen, neuroinflammation and neuroprotection in Parkinson's disease: glia<br />

dictates resistance versus vulnerability to neurodegeneration. Neuroscience. 2006;138(3):869-78. Epub 2005 Dec 5.<br />

[6] Morale MC, Serra PA, Delogu MR, Migheli R, Rocchitta G, Tirolo C, Caniglia S, Testa N, L'Episcopo F, Gennuso<br />

F, Scoto GM, Barden N, Miele E, Desole MS, Marchetti B. (2004) Glucorticoid receptor deficiency increases<br />

vulnerability of the nigrostriatal dopaminergic system: critical role of glial nitric oxide. FASEB J. Jan;18(1):164-6.<br />

Epub 2003 Nov 20.<br />

[7] Marchetti B, Serra P-A, Tirolo C, L’Episcopo F, Caniglia S, Gennuso F, Testa N, Miele E, Desole MS, Barden N,<br />

Morale MC (2005). Glucocorticoid receptor-nitric oxide crosstalk and vulnerability to experimental Parkinsonism:<br />

pivotal role for glia-neuron interactions. Brain Res Reviews 48/2: 302-321.<br />

[8] Wirz SA, Morale MC, Marchetti B, Barr AM, Sotgiu S, Rosati G, Pugliatti M, Sanna MV, Giliberto O, Bartfai T,<br />

Conti B. (2004) High frequency of TNF alleles -238A and -376A in individuals from northern Sardinia. Cytokine<br />

26(4):149-54.<br />

[9] Sotgiu S, Zanda B, Marchetti B, Fois ML, Arru G, Pes GM, Salaris FS, Arru A, Pirisi A, Rosati G. Inflammatory<br />

biomarkers in blood of patients with acute brain ischemia. Eur J Neurol. 2006 May;13(5):505-13.<br />

14


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

EFFECTS OF SELECTIVE ESTROGEN RECEPTOR (ER) LIGANDS AND<br />

MODULATORS ON THE GONADOTROPIC AXIS<br />

Tena-Sempere M.<br />

Physiology Section, Department of Cell Biology, Physiology and Immunology. Faculty of<br />

Medicine, University of Córdoba. Avda. Menéndez Pidal s/n. 14004 Córdoba, Spain<br />

fi1tesem@uco.es<br />

The gonadotropic (or hypothalamic-pituitary-gonadal) axis is extremely sensitive to<br />

the regulatory actions of sex steroids during development and in adulthood [1]. Thus,<br />

acting during critical periods of maturation, estrogen is the molecular signal responsible for<br />

the sexual differentiation of the brain centers governing reproduction. In addition to these<br />

organizing effects, estrogen (either systemically-derived or locally-produced via<br />

aromatization of androgen) conducts a wide diversity of activational effects at the<br />

hypothalamus and pituitary; transient modulatory actions that contribute to the dynamic<br />

regulation of the gonadotropic axis at different physiologic and pathologic conditions.<br />

The complexity of the biological actions of estrogen upon the HPG axis is further<br />

determined by the diversity of mechanisms and receptors mediating these actions, which<br />

involve genomic and non-genomic (i.e., via surface receptors) effects, as well as<br />

interactions between classical (i.e., nuclear) and cell-surface signalling systems. Even<br />

concerning genomic actions, the effects of estrogen might be mediated, in a tissue-specific<br />

manner, by at least two major receptor isoforms (ERα and ERβ) with ability to interact<br />

with multiple co-modulators (either activators or repressors). This degree of complexity<br />

likely contributes to the plasticity of estrogen effects, at different levels of the HPG axis<br />

and during different stages of development.<br />

In order to provide a deeper insight into the mechanisms for the effects of estrogen in<br />

the control of key aspects of the development and function of the HPG axis, the<br />

consequences of in vivo administration of selective ER ligands (specific for ERα or ERβ)<br />

and modulators (SERMs) on different functional parameters of the gonadotropic system<br />

have been evaluated in our laboratory over the last years. Of note, these studies have<br />

targeted both organizing actions (i.e., administration during critical periods) and<br />

activational effects.<br />

Concerning developmental effects, we have evaluated the reproductive consequences<br />

of neonatal administration of the SERM raloxifene in the rat. The aims of these studies<br />

were two-fold: (i) to provide evidence for the mode of action of this SERM on the brain<br />

(i.e., estrogenic vs. anti-estrogenic), taking advantage of the well-known state of sensitivity<br />

to the effects of sex steroids during critical periods of development; and (ii) to illustrate on<br />

the complexity of estrogen actions in the control of related, but distinct, reproductive<br />

functions, such as the neuroendocrine regulation of gonadotropic secretion and sexual<br />

behaviour. Our results indicate that neonatal administration of raloxifene induces an array<br />

of reproductive defects both in male and female rats, which included delayed<br />

balanopreputial separation, reduced body weight and hyper-prolactinaemia in males, and<br />

advanced vaginal opening, decreased body weight, reduced gonadotropin secretion,<br />

blunted positive and negative feed-back between estradiol and LH, anovulation and<br />

infertility in females [2,3]. Such defects are grossly similar to those induced by neonatal<br />

administration of estradiol, thus suggesting an estrogen-like effect of raloxifene upon the<br />

functional organization of the gonadotropic axis. Interestingly, however, raloxifene did not<br />

act as an estrogen upon the organization of the system governing sexual receptivity in the<br />

female rat [4], thus reflecting a high degree of cell-specificity within the hypothalamus for<br />

15


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

the mechanisms of action of estrogen on different reproductive centers. Further studies on<br />

this phenomenon, evaluating the consequences of neonatal administration of estrogen, as<br />

well as selective ER ligands and modulators, on sensitive molecular markers, such as<br />

hypothalamic expression of Kiss1 gene, are currently in progress in our laboratory.<br />

Concerning activational effects, the impact of in vivo administration of selective ER<br />

ligands to adult animals on different molecular and functional parameters of the<br />

gonadotropic axis, at the pituitary and hypothalamic levels, has been evaluated at our<br />

laboratory. At the pituitary, the effects of short-term exposure to selective ERα (PPT) or<br />

ERβ (DPN) ligands on the expression of ER-related transcripts was studied in<br />

ovariectomized (OVX) rats. OVX resulted in increased pituitary expression of ERβ<br />

mRNA, but decreased TERP-1 and -2 levels without affecting those of ERα. Estradiol<br />

benzoate, as agonist of α and β forms of ER, fully reversed the responses to OVX; a<br />

phenomenon which was mimiked by PPT but not by DPN, which failed also to increase the<br />

pituitary expression of progesterone gene (sensitive marker for estrogen action) but evoked<br />

a moderate stimulation of TERP-1 and -2 mRNA levels. Of note, the SERM tamoxifen<br />

behaved as ERα agonist at the pituitary, albeit with a lower magnitude of responses [5]. In<br />

addition to expression tests, functional analyses suggested that ERα is the predominant<br />

subtype mediating the effects of estrogen on the negative feedback control of<br />

gonadotropins (likely via repression of Kiss1 gene expression at the arcuate nucleus),<br />

GnRH self-priming and stimulation of prolactin secretion. Intriguingly, the role of ERβ,<br />

which is the only ER expressed in GnRH neurons and is also present at the gonadotrope, in<br />

the control of the gonadotropic axis remains to be fully elucidated.<br />

In sum, we have summarized herein studies on the effects of in vivo administration of<br />

selective ER ligands and modulators on the maturation and function of the gonadotropic<br />

axis. These analyses have contributed to define the predominant mode of action (agonist<br />

vs. antagonist) of SERMs such as tamoxifen and raloxifene at the pituitary and<br />

hypothalamus, and have provided interesting clues for the receptor mechanisms mediating<br />

some of the pleiotropic effects of estrogen in the control of key aspects of reproductive<br />

function, from sexual differentiation to gonadotropin feedback control.<br />

Reference List<br />

[1] Tena-Sempere et al. Current Topics in Steroid Research (2000) 3:23-37.<br />

[2] Pinilla et al. Reproduction (2001) 121:915-924.<br />

[3] Pinilla et al. Journal of Endocrinology (2002) 172:441-448.<br />

[4] Pinilla et al. Neuroscience Letters (2002) 329:285-288.<br />

[5] Tena-Sempere et al. Biology of Reproduction (2004) 70:671-678.<br />

16


SATURDAY, 17 th February 2007<br />

15.00 - 18.00<br />

Satellite Symposium:<br />

New perspectives in the dosage of neuroactive steroids


Satellite Symposium:<br />

New perspectives in the dosage of neuroactive steroids<br />

(Organizers: Melcangi R.C., Mensah-Nyagan A.G.)<br />

• Schumacher M, Liere P, Labombarda F, De Nicola AF, Guennoun R, Baulieu EE<br />

(France) Analysis of steroids by gas chromatography/mass spectrometry: novel<br />

perspectives for understanding their significance in the nervous system.<br />

• Griffiths WJ, Wang, Y. (UK) Capillary Liquid Chromatography Combined with<br />

Tandem Mass Spectrometry for the Study of Neurosteroids and Oxysterols in Brain<br />

• Higashi T, Nagahama A., Ninomiya Y., Shimada K. (Japan) Analysis of stressinduced<br />

changes in rat brain neuroactive steroid levels using LC/MS coupled with<br />

derivatization.<br />

• Caruso D., Scurati S., Crotti S., Maschi O., Melcangi RC (Italy) Recent advances<br />

in liquid chromatography-mass spectrometry related to the evaluation of plasma<br />

and tissue levels of neuroactive steroids during neurodegenerative events<br />

• Reddy DS (USA) Mass Spectrometric Quantification and Physiological-<br />

Pharmacological Activity of Androgenic Neurosteroids<br />

• Romeo E, Rupprecht R., Pasini A., Manieri G., Bernardi G., Longone P. (Italy)<br />

Neurosteroids Determination in Biological Fluids and their Enzymes Expression in<br />

Lymphocytes of Patients with Neuropsychiatric Disorders


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

ANALYSIS OF STEROIDS BY GAS CHROMATOGRAPHY/MASS<br />

SPECTROMETRY: NOVEL PERSPECTIVES FOR UNDERSTANDING THEIR<br />

SIGNIFICANCE IN THE NERVOUS SYSTEM<br />

Schumacher M. 1 , Liere P. 1 , Labombarda F. 2 , De Nicola A.F. 2 , Guennoun R. 1 , and<br />

Baulieu E.E. 1<br />

(1) UMR 788 Inserm and Univ. Paris 11, 80 rue du Général Leclerc, 94276 Kremlin-<br />

Bicêtre, France. Fax : +33-1-45211940 e-mail :Michael.Schumacher@kb.inserm;fr.<br />

(2) Instituto de Biologia y Medicina Experimetal and University of Buenos Aires,<br />

Argentina. e-mail: denicola@dna.uba.ar<br />

The conventional view that the gonadal steroids just have reproductive functions,<br />

and that adrenal steroids only regulate fluid homeostasis, metabolic pathways and<br />

adaptative responses to stress, has completely changed over the past few years. A large<br />

number of experimental studies, most performed in rodents, have indeed demonstrated the<br />

multiple actions of progestagens, estrogens, androgens, glucocorticoids and<br />

mineralocorticoids throughout the nervous system, and their considerable influence on the<br />

functioning of neurons and glial cells. In particular their neurotrophic and neuroprotective<br />

effects have attracted much attention because of their therapeutic promises. In this respect,<br />

progesterone and its 5α-reduced metabolites are particularly interesting because they not<br />

only promote the viability and regeneration of neurons, but they also act on the<br />

myelinating glial cell and play an important role in the formation of myelin sheaths [6].<br />

What makes steroids particularly attractive molecules for treating lesions and<br />

diseases of the nervous system is their property to easily cross the blood-brain and bloodnerve<br />

barriers and to very rapidly distribute throughout nervous tissues. In addition, some<br />

steroids can also be synthesized within the central nervous system (CNS) and the<br />

peripheral nervous system (PNS) by neurons and by glial cells. Steroids which are<br />

synthesized within the nervous system de novo from cholesterol have been named<br />

“neurosteroids” [1,2]. As both the endocrine glands and local production contribute to the<br />

pool of steroids present within nervous tissues, changes in their circulating levels do not<br />

necessarily reflect changes in their availability for neural cells. The physiological<br />

significance of neurosteroid synthesis in development, regeneration and aging is still<br />

poorly understood, and much remains to be done.<br />

Sensitive methods for the accurate analysis and quantification of steroids in plasma<br />

and nervous tissues have thus become a major issue. Indeed, the reliability of methods<br />

which continue to be used for the analysis of steroids remains a serious problem. This<br />

concerns both the methods for quantitative analysis, such as immunoassays, and the<br />

methods for pretreatment of the biological samples, including steroid extraction and<br />

separation. That is, specificity cannot be guaranteed even with carefully validated<br />

immunoassays, some chromatographic procedures for separating the different categories of<br />

steroids are questionable, and steroid sulfates are usually determined indirectly after<br />

removal of the sulfate group by solvolysis and subsequent analysis of the free steroids. The<br />

dangers of the latter method are well illustrated by our finding that the currently used solidphase<br />

extraction on C18 columns does not remove nonpolar steroid-containing lipids from<br />

the fraction which contains sulfated steroids. Moreover, it was shown that acidic solvolysis<br />

is not specific for 3β-hydroxy-Δ5 steroid sulfates and is able to release free steroids from<br />

other nonpolar components. As a consequence, sulfated steroids have been artefactually<br />

measured in rodent plasma and brain during decades [5].<br />

19


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

Steroid sulfates can now be measured directly by liquid chromatography/mass<br />

spectrometry (LC/MS) and free steroids by gas chromatography/mass spectrometry<br />

(GC/MS) [4,5]. Advantages of these technologies are that : 1) they allow the chemical<br />

identification and accurate quantification of steroids ; 2) they are not dependent on the<br />

availability of antibodies with limited specificity ; 3) several steroids can be analyzed<br />

within a single tissue sample.<br />

We have also recently developed a new sample separation method coupled to<br />

GC/MS analysis for the quantification of sulfate esters of pregnenolone (PREGS) and of<br />

dehydroepiandrosterone (DHEAS) in the rat and human brain and plasma [5]. Using a<br />

solid-phase extraction recycling protocol, the results show that little or no PREGS and<br />

DHEAS is present in rat and mouse brain and plasma. These data are in agreement with<br />

studies in which steroid sulfates were analyzed without deconjugation. We suggest that<br />

discrepancies between analyses with and without deconjugation are caused by internal<br />

contamination of the brain extract fraction, supposed to contain steroid sulfates, by nonpolar<br />

components. However, in contrast to rodents, PREGS and DHEAS were shown to be<br />

present in human brain tissue.<br />

Analysis of free steroids by GC/MS has allowed us to provide reference values for<br />

steroid levels in the different compartments of the nervous system, and to show that an<br />

increase in the synthesis of neuroprogesterone and its 5α-reduced metabolites within the rat<br />

spinal cord is part of the mechanisms by which nerve cells cope with neurodegeneration<br />

[3]. In this study, it was shown that the levels of pregnenolone and progesterone are<br />

increased in the spinal cord of castrated and adrenalectomized male rats in response to<br />

transection. As the animals were deprived of their steroidogenic endocrine glands, these<br />

findings strongly suggest an increase in the local synthesis of the steroids. Moreover,<br />

increased levels of 5α-dihydroprogesterone and allopregnanolone were measured in the<br />

spinal cord of lesioned animals, whithout changes in their plasma levels, consistent with<br />

their local synthesis.<br />

Reference list<br />

[1] E.E. Baulieu, Neurosteroids: of the nervous system, by the nervous system, for the nervous system,<br />

Recent Prog. Horm. Res. 52 (1997) 1-32.<br />

[2] E.E. Baulieu, P.Robel, and M.Schumacher, Neurosteroids: beginning of the story, Int Rev Neurobiol 46<br />

(2001) 1-32.<br />

[3] F. Labombarda, A.Pianos, P.Liere, B.Eychenne, S.Gonzalez, A.Cambourg, A.F.De Nicola,<br />

M.Schumacher, and R.Guennoun, Injury elicited increase in spinal cord neurosteroid content analysed<br />

by gas chromatography mass spectrometry, Endocrinology 147 (2006) 1847-1859.<br />

[4] P. Liere, Y.Akwa, S.Weill-Engerer, B.Eychenne, A.Pianos, P.Robel, J.Sjovall, M.Schumacher, and<br />

E.E.Baulieu, Validation of an analytical procedure to measure trace amounts of neurosteroids in brain<br />

tissue by gas chromatography-mass spectrometry, J Chromatogr B 739 (2000) 301-312.<br />

[5] P. Liere, A.Pianos, B.Eychenne, A.Cambourg, S.Liu, W.Griffiths, M.Schumacher, J.Sjovall, and<br />

E.E.Baulieu, Novel lipoidal derivatives of pregnenolone and dehydroepiandrosterone and absence of<br />

their sulfated counterparts in rodent brain, J Lipid Res 45 (2004) 2287-2302.<br />

[6] M. Schumacher, S.Weill-Engerer, P.Liere, F.Robert, R.J.Franklin, L.M.Garcia-Segura, J.J.Lambert,<br />

W.Mayo, R.C.Melcangi, A.Parducz, U.Suter, C.Carelli, E.E.Baulieu, and Y.Akwa, Steroid hormones<br />

and neurosteroids in normal and pathological aging of the nervous system, Prog Neurobiol 71 (2003) 3-<br />

29.<br />

20


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

CAPILLARY LIQUID CHROMATOGRAPHY COMBINED WITH TANDEM<br />

MASS SPECTROMETRY FOR THE STUDY OF NEUROSTEROIDS AND<br />

OXYSTEROLS IN BRAIN<br />

Griffiths W.J., and Wang Y.<br />

The School of Pharmacy, University of London, 29-39 Brunswick Square, WC1N 1AX,<br />

London, UK. Fax +44 20 7753 5964, e-mail: william.griffiths@pharmacy.ac.uk<br />

Cholesterol is present at a level of 7-8 microgram/milligram wet weight in human<br />

brain, where it is synthesised de novo, and corresponds to about 25% of the total body pool<br />

[1]. In brain, cholesterol is metabolised to oxysterols and neurosteroids. Both of these<br />

steroids are biologically active. Oxysterols can act as ligands to nuclear receptors e.g. liver<br />

X receptors (LXRs), while neurosteroids interact with neurotransmitter gated ion channels<br />

and modulate neural transmission. Neurosteroids exert their effects in seconds or<br />

milliseconds, in contrast to nuclear receptor whose physiological actions occur within<br />

hours or days.<br />

Oxysterols are present in brain at levels ranging from 20 nanogram/milligram to 30<br />

picogram/milligram, while neurosteroids are found at the picogram/milligram level. Both<br />

oxysterols and neurosteroids are present in brain in different isomeric forms, where<br />

different isomers have distinct biological activity. The low level and structural diversity of<br />

these cholesterol metabolites has made their analysis challenging. In this paper we will<br />

discuss the mass spectrometric (MS) methods we have employed for steroid analysis in<br />

brain and present new data recently obtained.<br />

In the early 1980’s Baulieu and colleagues reported the presence of steroid sulphates in rat<br />

brain [2,3]. It was subsequently found that these steroids act on neurotransmitter gated ion<br />

channels, e.g. pregnenolone sulphate inhibits GABA-mediated currents and enhances<br />

NMDA-activated currents in cultured rat hippocampal neurons [1]. Neutral steroids were<br />

also identified and found to be biologically active, e.g. allopregnanolone enhances<br />

GABAergic transmission and decreases NMDA transmission, while progesterone can bind<br />

to the progesterone receptor. In their early work on steroid sulphates, Baulieu and<br />

colleagues used GC-MS for steroid sulphate analysis after removal of the sulphate ester<br />

group. However, intact conjugate were not analysed, and recent data questions the earlier<br />

identification of steroid sulphates in brain [5].<br />

Since the early 1980’s MS in many laboratories has shifted from GC-MS to LC-MS, which<br />

has the advantage of being able to analyse intact conjugates and give direct molecular<br />

weight information, and when in combination with MS/MS provide structural information.<br />

To maximise chromatographic performance and MS sensitivity we have applied a strategy<br />

based on miniaturised chromatography and low-flow rate electrospray (ES)-MS for steroid<br />

analysis in brain. Steroid sulphates are readily ionised by negative-ion ES and give<br />

informative MS/MS spectra with high sensitivity (0.1 pg on column detection limit) [6].<br />

Neutral steroids can be more difficult to analyse by ES-MS/MS. Although steroids with a<br />

3-oxo-4-ene structure e.g. progesterone and testosterone, can be analysed with high<br />

sensitivity by positive-ion ES-MS, steroids with less basic functional groups e.g.<br />

dehydroepiandrosterone, pregnenolone, are ionised poorly. To counter this problem, we<br />

have incorporated a derivatisation step to our analytical method. Initially, we derivatised<br />

oxosteroids with hydroxylamine to give steroid oximes, which are more readily ionised<br />

than their underivatised analogues [6]. Now we use a different derivatisation reagent, the<br />

Girard P (GP) hydrazine reagent, which adds greater sensitivity to the assay and provides<br />

more informative fragmentation spectra. Using the GP derivative neurosteroids can be<br />

21


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

analysed on the sub-picogram level [4]. When derivatised with the GP reagent,<br />

neurosteroids are suitable for analysis at high sensitivity by both MALDI-MS and LC-ES-<br />

MS.<br />

24S-Hydroxycholesterol is the major cholesterol metabolite found in brain [7]. This<br />

oxysterol is difficult to analyse by LC-ES-MS or MALDI-MS, however by using<br />

microchemical oxidation and derivatisation with the GP reagent, this oxysterol and<br />

analougous metabolites can be analysed with high sensitivity [8]. Using a combination of<br />

oxidation and derivatisation we have been able to identify a series of dihydroxycholesterols<br />

with structures of potential ligands to LXRbeta, a nuclear receptor expressed in neurons.<br />

Reference List<br />

[1] M.R. Bowlby, Pregnenolone sulfate potentiation of N-methyl-D-aspartate receptor channels in<br />

hippocampal neurons, Mol. Pharmacol. 43 (1993) 813-9.<br />

[2] C. Corpéchot, P. Robel, M. Axelson, J. Sjövall, E.E. Baulieu, Characterization and<br />

measurement of dehydroepiandrosterone sulfate in rat brain, Proc. Natl. Acad. Sci. U S A. 78<br />

(1981) 4704-7.<br />

[3] C. Corpéchot, M. Synguelakis, S. Talha, M. Axelson, J. Sjövall, R. Vihko, E.E. Baulieu, P.<br />

Robel, Pregnenolone and its sulfate ester in the rat brain, Brain Res. 270 (1983) 119-25.<br />

[4] W.J. Griffiths, S. Liu, G. Alvelius, J. Sjövall, Derivatisation for the characterisation of neutral<br />

oxosteroids by electrospray and matrix-assisted laser desorption/ionisation tandem mass<br />

spectrometry: the Girard P derivative, Rapid Commun. Mass Spectrom. 17 (2003) 924-<strong>35</strong>.<br />

[5] P. Liere, A. Pianos, B. Eychenne, A. Cambourg, S. Liu, W. Griffiths, M. Schumacher, J.<br />

Sjövall, Baulieu E.E. Novel lipoidal derivatives of pregnenolone and dehydroepiandrosterone<br />

and absence of their sulfated counterparts in rodent brain, J. Lipid Res. 45 (2004) 2287-302.<br />

[6] S. Liu, J. Sjövall, W.J. Griffiths, Neurosteroids in rat brain: extraction, isolation, and analysis<br />

by nanoscale liquid chromatography-electrospray mass spectrometry, Anal. Chem. 75 (2003)<br />

58<strong>35</strong>-46.<br />

[7] D. Lutjohann, Cholesterol metabolism in the brain: importance of 24S-hydroxylation, Acta<br />

Neurol. Scand. Suppl. 185 (2006) 33-42.<br />

[8] Y. Wang, M. Hornshaw, G. Alvelius, K. Bodin, S. Liu , J. Sjövall , W.J. Griffiths. Matrixassisted<br />

laser desorption/ionization high-energy collision-induced dissociation of steroids:<br />

analysis of oxysterols in rat brain. Anal. Chem. 78 (2006) 164-73.<br />

22


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

ANALYSIS OF STRESS-INDUCED CHANGES IN RAT BRAIN NEUROACTIVE<br />

STEROID LEVELS USING LC/MS COUPLED WITH DERIVATIZATION<br />

Higashi T., Nagahama A., Ninomiya Y., and Shimada K.<br />

Division of Pharmaceutical Sciences, Graduate School of Natural Science and Technology, Kanazawa<br />

University, Kakuma-machi, Kanazawa, Japan. Fax +81-76-234-4459 e-mail: higashi@p.kanazawa-u.ac.jp<br />

Neuroactive steroids possessing anxiolytic and anti-stress properties act through<br />

membrane receptors, mostly the γ-aminobutyric acid type A (GABA A ) receptor complex.<br />

For example, allopregnanolone (AP) positively modulates the action of GABA at these<br />

receptors to raise the threshold of brain excitability during the response to stressful stimuli.<br />

It is also known that the endogenous AP level in the animal brain is rapidly elevated from<br />

the trace level (practically none) to the ng/g tissue level by several acute stress paradigms.<br />

Thus, the stress-induced changes in the brain levels of AP and also its precursors have been<br />

fairly well studied and a great deal of knowledge has been accumulated regarding this<br />

matter. However, the stress-induced change in the brain and serum levels of<br />

epiallopregnanolone (EAP), the 3β-isomer of AP, has been poorly studied, although it can<br />

antagonize the GABAergic function of AP in rats [1]. Furthermore, several papers<br />

described that some androstane steroids, such as testosterone (T), elicit the anesthetic and<br />

anxiolytic effects in animal models [2], but their brain levels have not been elucidated.<br />

Based on this background information, we performed the following two studies; 1)<br />

analysis of the stress-induced level change of EAP in the rat brain and serum, and 2)<br />

elucidation of the influence of acute stress on the T and 5α-dihydrotestosterone (DHT)<br />

levels in the rat brain and serum. In order to clarify these matters, we chose LC/ESI-<br />

MS/MS as the analytical methodology.<br />

1. Analysis of the stress-induced level change of EAP in the rat brain and serum<br />

We began this study by developing the LC/ESI-MS/MS method for the simultaneous<br />

determination of AP, EAP and their precursor, 5α-dihydroprogesterone (DHP). The<br />

biggest problem encountered in the ESI-MS measurement of these 5α-reduced steroids is<br />

the lack of sensitivity; around 100 pg of steroids were required to obtain a signal to noise<br />

ratio of 5. To overcome this problem, we introduced the derivatization with a permanently<br />

charged reagent, 2-hydrazine-1-methylpyridine (HMP) [3]. This reagent quantitatively<br />

reacts with steroids having an oxo-group to give the ESI-active derivatives. The derivatives<br />

provided only their molecular cations, [M] + , with a high intensity in the positive-ESI-MS<br />

and also provided a characteristic product ion at m/z 108 in MS/MS. When the SRM mode<br />

with the transition from [M] + to this product ion was employed, the low femtomole level of<br />

the derivatives could be readily detected. The developed method was then applied to the<br />

animal experiment. Rats were divided into two groups, control (n=10) and immobilization<br />

stressed rats (n=10), and the latter was immobilized on their backs on a board for 20 min<br />

and then the brain and serum were collected. The brain homogenate corresponding to 20<br />

mg of tissue or a 20 µl of serum was purified using a Strata-X cartridge and the<br />

neuroactive steroid fraction was then treated with HMP. In the brain of the stressed rats,<br />

the concentrations of AP, EAP and DHP were 1.74±0.71, 0.58±0.30 and 2.74±1.12 ng/g<br />

tissue, respectively, but the levels in the control group were less than the quantitation limit<br />

(0.25 ng/g tissue). The serum AP (1.31±0.72 ng/ml) and DHP (0.94±0.36 ng/ml) levels<br />

were also drastically elevated by the stress (control: not detected), while EAP was not<br />

detected in the serum even in the stressed rats. In conclusion, this study demonstrated that<br />

23


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

the brain EAP level was rapidly elevated by immobilization stress like the AP and DHP<br />

levels. The study also found that EAP is the brain-specific product.<br />

2. Influence of acute stress on the T and DHT levels in rat brain and serum<br />

We first analyzed the stress-induced level change of T. Because T has the ESI-responsive<br />

3-oxo-4-ene structure, its quantification was performed without derivatization. As a result,<br />

we could not find a recognizable change in the brain T level between the control and<br />

stressed rats. There was also no statistical difference in the serum T level between the two<br />

groups. However, due to the large individual differences in the experimental animals, we<br />

could not develop a clear answer to the question of whether or not the T level is influenced<br />

by the stress. The fact that most of T found in the brain is derived from the peripheral<br />

organs [4] prompted us to use the ratio of the brain T concentration to the serum T<br />

concentration (BT/ST) as an alternative index for the more detailed analysis of the stressinduced<br />

level change in T. The BT/ST value in the stressed rat was significantly higher<br />

than that of the control rat. Thus, we found that the T level is also influenced by the acute<br />

stress. Although several mechanisms for the elevation of the BT/ST value due to the stress<br />

can be imagined, we thought that the most probable mechanism is the suppression of the T<br />

metabolism in the brains by the stress. T is metabolized into DHT by the catalysis of 5αreductase.<br />

The affinity of 5α-reductase is higher for progesterone than for T, thus<br />

progesterone may competitively inhibit the T metabolism, which may cause the decline of<br />

the DHT level and accumulation of T in the brain. To prove this point, we measured the<br />

brain and serum DHT using LC/ESI-MS/MS after conversion to its HMP derivative. The<br />

measured values obtained from a 100 mg of brain tissue revealed that there was no<br />

significant difference in the DHT level between the control (0.71±0.59 ng/g tissue) and<br />

stressed rats (0.58±0.39 ng/g tissue). In the serum, the DHT peak was observed neither in<br />

the control nor in the stressed rats. These results show that the brain always synthesizes a<br />

fair amount of DHT. Moreover, when the brain T concentration and DHT concentration<br />

were plotted on a scatter diagram, they lay almost on a straight line. These data<br />

demonstrate that the brain DHT level is not influenced by the stress and depends on the<br />

brain T level. Thus, our forecast that the suppression of the T metabolism in the brain<br />

causes elevation of the BT/ST value was found to be wrong. In conclusion, we found that<br />

the BT/ST value was significantly elevated by the immobilization stress, although the<br />

mechanism for this phenomenon is still unclear.<br />

Reference List<br />

[1] Bäckström T., Wahlström G., Wahlström K., Zhu D., Wang M.-D., 2005.<br />

Isoallopregnanolone; an antagonist to the anaesthetic effect of allopregnanolone in<br />

male rats. Eur. J. Pharmacol. 512, 15–21.<br />

[2] Edinger K.L., Frye C.A., 2005. Testosterone’s anti-anxiety and analgesic effects may<br />

be due in part to actions of its 5α-reduced metabolites in the hippocampus.<br />

Psychoneuroendocrinology 30, 418–430.<br />

[3] Higashi T., Yamauchi A., Shimada K., 2005. 2-Hydrozine-1-methylpyridine: a highly<br />

sensitive derivatization reagent for oxosteroids in liquid chromatography-electrospray<br />

ionization-mass spectrometry. J. Chromatogr. B 825, 214–222.<br />

[4] Alomary A.A., Vallée M., O’Dell L.E., Koob G.F., Purdy R.H., Fitzgerald R.L., 2003.<br />

Acutely administered ethanol participates in testosterone synthesis and increases<br />

testosterone in rat brain. Alcohol. Clin. Exp. Res. 27, 38–43.<br />

24


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

RECENT ADVANCES IN LIQUID CHROMATOGRAPHY-MASS<br />

SPECTROMETRY RELATED TO THE EVALUATION OF PLASMA AND<br />

TISSUE LEVELS OF NEUROACTIVE STEROIDS DURING<br />

NEURODEGENERATIVE EVENTS<br />

Caruso D.# 1 , Scurati S. 1 , Crotti S. 1 , Maschi O. 1 and Melcangi R.C. 2<br />

1 Dept. of Pharmacological Sciences, 2 Dept. of Endocrinology, University of Milano,<br />

Milano, Italy.<br />

#Presenting author: Dept. of Pharmacological Sciences, University of Milano, via<br />

Balzaretti 9, 20133, Milano Italy, donatella.caruso@unimi.it<br />

Nervous system represents an important target for the action of neuroactive<br />

steroids. Indeed, these molecules exert a broad spectrum of actions regulating<br />

neuroendocrine events, reproduction, feeding, behavior etc. Moreover, as recently<br />

demonstrated, neuroactive steroids also exert important protective effects at the level of<br />

central (CNS) and peripheral nervous system (PNS), suggesting that they might represent a<br />

new therapeutic strategy to counteract neurodegenerative events. In this context, to know<br />

the levels of the neuroactive steroids present in nervous structures and how these are<br />

modified during neurodegenerative events is absolutely necessary. This means the set up of<br />

a sensitive and specific methodology that gives a complete information of steroid levels in<br />

target nervous structures and in plasma. The advent of robust and analytically reliable<br />

techniques based on the combination of liquid chromatography (LC) and mass<br />

spectrometry (MS) by means of atmospheric pressure ionization (API) techniques (e.g.,<br />

electrospray, ESI and atmospheric pressure chemical ionization, APCI) and in particular<br />

the improvements brought about by tandem MS (MS/MS), has opened new perspectives in<br />

terms of mass spectrometric identification and quantification of steroids that are difficult to<br />

analyse by gas chromatography-MS. With respect to the ionization mode, APCI is mainly<br />

applied to rather less polar compounds than ESI but is less susceptible to ion suppression<br />

due to the presence of several interferences as in biological tissues. For quantitative assays<br />

employing MS detection, triple quadrupole systems are most commonly used, while the<br />

new generation of ion trap, namely the linear trap, exerts similar performance, as<br />

demonstrated also by our results. In addition, when an APCI-linear trap is operated in the<br />

MS/MS mode, the identification and quantification of the analyte are based on both<br />

precursor and product ions, giving higher selectivity and best sensitivity than for any other<br />

MS system. On these remarks, we set up an analytical method mass spectrometry-based for<br />

the identification and quantification of pregnenolone (PREG), progesterone (PROG),<br />

dihydroprogesterone (DHP), tetrahydroprogesterone (THP), testosterone (T),<br />

dihydrotestosterone (DHT), 5α-androstan-3α17β-diol (3α-diol), and 17β estradiol (17β-<br />

E2). After validation of the HPLC-APCI-MS/MS procedure, the method was applied to the<br />

detection and determination of these neuroactive steroids in plasma and nervous structures<br />

of control rat and we have compared these levels to what occurring in an experimental<br />

model of neuropathy, such as streptozotocin (STZ)-induced diabetes. Indeed it is well<br />

known, that diabetes is associated to a spectrum of functional and structural changes in<br />

PNS and CNS.<br />

Neuroactive steroids in plasma, in cerebellum (representative of CNS) and in brachial<br />

nerves (representative of PNS) were extracted by solid phase technique and quantified by<br />

means of calibration curves using deuterated internal standards, since they are expected to<br />

possess similar ionization efficiencies as the correspondent analyte. Our results indicate<br />

that all neuroactive steroids here considered are present in the two nervous structures and<br />

25


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

in plasma, with the exception of THP and DHT, which in plasma are under the detection<br />

limit. Interestingly, tissue and plasma levels of these neuroactive steroids are significantly<br />

modified in the experimental model of STZ-induced diabetes. In particular, PREG, PROG<br />

and T levels are significantly decreased in cerebellum. The same trend is shown by their<br />

metabolites without reaching statistical significance. Moreover, in brachial nerves, not only<br />

PREG and T, but also metabolites of PROG (i.e., DHP and THP) and of T (i.e., DHT and<br />

3α-diol) show a significant decrease. Furthermore, our data indicate that diabetes causes<br />

alterations in the plasma levels of steroids that, in some cases, do not completely reflect the<br />

changes observed in nervous structures. For instance, in contrast to what observed in<br />

nervous structures, plasma PREG, DHP and DHT levels are unaffected by diabetes while<br />

3α-diol levels show a significant increase.<br />

Taken together these data demonstrate that LC-MS/MS method allows the assessment of<br />

neuroactive steroids in plasma and in structures of central and peripheral nervous system<br />

with high sensitivity and specificity and that neurodegenerative processes affect their<br />

levels.<br />

26


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

MASS SPECTROMETRIC QUANTIFICATION AND PHYSIOLOGICAL-<br />

PHARMACOLOGICAL ACTIVITY OF ANDROGENIC NEUROSTEROIDS<br />

Doodipala S.R.<br />

Department of Molecular Biomedical Sciences, North Carolina State University College of<br />

Veterinary Medicine, Raleigh, NC 27606,USA.<br />

Testosterone modulates seizure susceptibility in animals and humans, but the underlying<br />

mechanisms are obscure. Here, I present evidence that testosterone-derived “androgenic<br />

neurosteroids”, 3α-androstanediol and 17β-estradiol, mediate the testosterone effects on<br />

neural excitability and seizure susceptibility. The presentations will primarily be focused<br />

on: (1) development and validation of mass spectrometric determination of the androgenic<br />

neurosteroid 3α-androstanediol; and (2) investigations on the molecular mechanisms<br />

underlying the testosterone modulation of seizure susceptibility in animals models of<br />

epilepsy.<br />

Although 3α-androstanediol (5α-androstane-3α,17-diol) could act as a key<br />

neuromodulator in the brain, there is no specific and sensitive assay for quantitative<br />

determination of the 3α-androstanediol in biological samples. We have established a liquid<br />

chromatography-tandem mass spectrometry assay to measure 3α-androstanediol in rat<br />

plasma. Standard 3α-androstanediol added to rat plasma has been successfully analysed<br />

with excellent linearity, specificity, sensitivity, and reproducibility.<br />

Testosterone modulation of seizure susceptibility is hypothesized to occur through its<br />

conversion to neurosteroids with “anticonvulsant” and “proconvulsant” actions, and hence<br />

the net effect of testosterone on neural excitability and seizure activity depends on the<br />

levels of distinct testosterone metabolites. Testosterone undergoes metabolism to<br />

neurosteroids via two distinct pathways. Aromatization of the A-ring converts testosterone<br />

into 17β-estradiol. Reduction of testosterone by 5α-reductase generates 5αdihydrotestosterone,<br />

which is then converted to 3α-androstanediol, a powerful GABA A<br />

receptor-modulating neurosteroid with anticonvulsant properties. Systemic doses of<br />

testosterone decreased seizure threshold in rats and increased the incidence and severity of<br />

pentylenetetrazol-induced seizures in mice. These proconvulsant effects of testosterone<br />

were associated with increases in plasma 17β-estradiol and 3α-androstanediol<br />

concentrations. The 5α-reduced metabolites of testosterone, 5α-dihydrotestosterone and<br />

3α-androstanediol, had powerful anticonvulsant activity.<br />

The 3α-androstanediol assay is an important tool in this area because of the growing<br />

interest in the potential to use adjuvant hormonal therapy to improve treatment of epilepsy.<br />

The testosterone-derived neurosteroids 3α-androstanediol and 17β-estradiol could<br />

contribute to the net cellular actions of testosterone on neural excitability and seizure<br />

susceptibility. Because of its powerful GABA-A receptor-modulating properties, the<br />

androgenic neurosteroid 3α-androstanediol is proposed as an endogenous modulator of<br />

seizure susceptibility in men with epilepsy.<br />

Acknowledgements: Supported partly by NC State CVM grant.<br />

27


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

NEUROSTEROIDS DETERMINATION IN BIOLOGICAL FLUIDS AND THEIR<br />

ENZYMES EXPRESSION IN LYMPHOCYTES OF PATIENTS WITH<br />

NEUROPSYCHIATRIC DISORDERS<br />

Romeo E. a,b , Rupprecht R. d , Pasini A. a , Manieri G. c , Bernardi G. a,b , and Longone P. b<br />

a Department of Neuroscience Tor Vergata University; b Experimental Neurology Santa<br />

Lucia Foundation; c Clinica Neurologica, La Sapienza University, Rome Italy; d Department<br />

of Psychiatry Ludwig Maximilian University, Munich Germany.<br />

Steroids action is on the basis of concentration, as well as the specificity of tissue<br />

receptors. Steroids affect several different types of body processes, that taken together<br />

lead to a highly specific function. This might explain how steroids came to symbolize<br />

precise body’s events and necessities.<br />

In the nervous system (NS) steroids elicit important glial and neuronal cell responses by<br />

stimulating gene expression and by modulating different neurotransmitters receptors [2].<br />

Nevertheless a precise physiological role in the NS has not yet been established. We and<br />

others have found that steroids levels can be altered in many neuropsychiatric and<br />

neurodegenerative disorders as: depression, anxiety, schizophrenia, panic disorder and<br />

Alzheimer [11,12,13,15]. This indicate that their dysregulation it is maybe related to<br />

psychopathological dimensions present in different pathologies more than being linked to a<br />

particular syndrome. Central and peripheral fluids show the same concentrations trend of<br />

steroids and one of the reasons is that the most of them can easily cross the blood brain<br />

barrier allowing steroids in the two compartments to reach an equilibrium [5].<br />

Yet, if we consider the biosynthesis of steroids, there are some differences between the<br />

periphery and the NS: i) in the NS, steroids biosynthesis is not restricted to specialized<br />

brain areas or cells endowed of their synthesis, as it happens in the peripheral endocrine<br />

organs; ii) steroids, in the NS, can have alternative pathways, like the biosynthesis of<br />

cholesterol [3]; iii) the enzymes isoforms, expressed in the NS, are often present<br />

periferally, as in the case of 3alpha hydroxysteroidehydrogenase (3HSD) [14], that rapidly<br />

clears steroids from the circulation, differently from the NS where 3HSD synthetizes<br />

3alpha,5alpha tetrahydroprogesterone (THP) [9], one of the most potent modulator of<br />

numerous neurotransmitter receptors such as GABA A receptors [8]. Therefore, the<br />

expression of steroidogenic enzymes and the synthesis of steroids in the NS seems to relay<br />

either on the crosstalk with the periphery, and on a direct communication between the NS<br />

cells. Thus, to relate steroids quantification and their enzymes expression, we have<br />

quantified steroids involved in the GABAergic system, in cerebro spinal fluid (CSF) and<br />

plasma of parkinsonian patients along with the mRNA of their enzymes and isoforms in<br />

peripheral blood mononuclear cells (PBMC) [4,7]. We have measured the steroids<br />

[progesterone (PROG), 3alpha, 5alpha THP and 3beta, 5alpha THP, 5alpha<br />

dihydroprogesterone (DHP) and dehydroepiandrosterone(DHEA)] in CSF and plasma of<br />

parkinsonian patients and age matched controls with the GC/MS technique [4]. Moreover<br />

we have quantified by means of comparative RT-PCR the mRNA expression of the<br />

enzymes 5alpha reductase type I (mainly central), 3HSD type I (mainly peripheral) and II<br />

(mainly central) [10] and the peripheral benzodiazepine receptor (PBR) [7]. As a source of<br />

mRNA, for the quantification of these enzymes, we have used PBMC because they express<br />

both central and peripheral isoforms of many neurosteroidogenic enzymes [6,16] and of<br />

many receptors such as GABA A [1]. Steroids levels had a similar trend in plasma and CSF<br />

of parkinsonian patients treated with L-DOPA, showing a decrease of 3alpha, 5alpha THP<br />

and 5alpha DHP and unchanged levels of PROG, 3beta, 5alpha THP and DHEA. The<br />

28


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

5alpha reductase expression was decreased, while PBR expression was unchanged.<br />

Interestingly the two 3HSD isoforms behaved differently, and while the mRNA expression<br />

of HSD I was unchanged compared to control, the HSD II was increased. Our data indicate<br />

that the enzymes isoforms present also in the NS are differently regulated from those only<br />

peripheral and further studies are required.<br />

In conclusion we consider important to evaluate steroids levels in biological fluids along<br />

with the expression of their biosynthetic enzymes, in order to have a more inclusive<br />

comprehension of the role that steroids may have in the NS and in the cross talk with the<br />

periphery.<br />

Reference List<br />

[1] Alam S, Laughton DL, Walding A, Wolstenholme AJ., 2006. Human peripheral blood mononuclear<br />

cells express GABAA receptor subunits. Mol Immunol. 43, 1432-42.<br />

[2] Baulieu EE, Robel P, Schumacher M., 2001. Neurosteroids: beginning of the story. Int Rev<br />

Neurobiol.46,1-32.<br />

[3] Bjorkhem, I., Meaney, S., 2004. Brain cholesterol: long secret life behind a barrier. Arterioscler Thromb<br />

Vasc Biol. 24, 806-815.<br />

[4] di Michele F, Longone P, Romeo E, Lucchetti S, Brusa L, Pierantozzi M, Bassi A, Bernardi G,<br />

Stanzione P., 2003. Decreased plasma and cerebrospinal fluid content of neuroactive steroids in<br />

Parkinson's disease. Neurol Sci. 24, 172-3.<br />

[5] Dubrovsky, B., 2006. Neurosteroids, neuroactive steroids, and symptoms of affective disorders.<br />

Pharmacol Biochem Behav. 84, 644-655.<br />

[6] Leb C.R., Hu F..Y, Murphy B.E., 1997. Metabolism of progesterone by human lymphocytes:<br />

production of neuroactive steroids. J Clin Endocrinol Metab 82, 4064-8<br />

[7] Luchetti S., di Michele F., Romeo E., Brusa L., Bernardi G., Cummings B.J., Longone P., 2006.<br />

Comparative non-radioactive RT-PCR assay: an approach to study the neurosteroids biosynthetic<br />

pathway in humans. J Neurosci Methods. 153, 290-8.<br />

[8] Magnaghi V, Ballabio M, Consoli A, Lambert JJ, Roglio I, Melcangi RC., 2006. GABA receptormediated<br />

effects in the peripheral nervous system: A cross-interaction with neuroactive steroids. J Mol<br />

Neurosci. 28, 89-102.<br />

[9] Martini L., Celotti F., Melcangi R.C., 1996. Testosterone and progesterone metabolism in the central<br />

nervous system: cellular localization and mechanism of control of the enzymes involved. Cell Mol<br />

Neurobiol. 16, 271-82.<br />

[10] Penning TM, Burczynski ME, Jez JM, Hung CF, Lin HK, Ma H, Moore M, Palackal N,Ratnam<br />

K.,2000. Human 3alpha-hydroxysteroid dehydrogenase isoforms (AKR1C1-AKR1C4) of the aldo-keto<br />

reductase superfamily: functional plasticity and tissue distribution reveals roles in the inactivation and<br />

formation of male and female sex hormones. Biochem J. <strong>35</strong>1, 67-77.<br />

[11] Romeo E., Strohle A., Spalletta G., di Michele F., Hermann B., Holsboer F., Pasini A.,Rupprecht R.,<br />

1998. Effects of antidepressant treatment on neuroactive steroids in major depression. Am J Psychiatry<br />

55, 910-3.<br />

[12] Schaeffer V, Patte-Mensah C, Eckert A, Mensah-Nyagan AG., 2006. Modulation of neurosteroid<br />

production in human neuroblastoma cells by Alzheimer's disease key proteins. J Neurobiol. 66, 868-81.<br />

[13] Schumacher M, Weill-Engerer S, Liere P, Robert F, Franklin RJ, Garcia-Segura LM, Lambert JJ, Mayo<br />

W, Melcangi RC, Parducz A, Suter U, Carelli C, Baulieu EE, Akwa Y., 2003. Steroid hormones and<br />

neurosteroids in normal and pathological aging of the nervous system. Prog Neurobiol. 71 ,3-29.<br />

[14] Stoffel-Wagner, B. 2001. Neurosteroid metabolism in the human brain. Eur J Endocrinol. 145, 669-679.<br />

[15] Strohle A., Romeo E., di Michele .F, Pasini A., Hermann B., Gajewsky G., Holsboer F., Rupprecht R.,<br />

2003. Induced panic attacks shift gamma-aminobutyric acid type A receptor modulatory neuroactive<br />

steroid composition in patients with panic disorder: preliminary results. Arch Gen Psychiatry. 60, 161-8.<br />

[16] Zhou Z., Shackleton C.H., Pahwa S., White P.C., Speiser P.W., 1998. Prominent sex steroid metabolism<br />

in human lymphocytes. Mol Cell Endocrinol. 138, 61-9.<br />

29


SUNDAY, 18 th February<br />

12.00 - 13.00<br />

Plenary Lecture:<br />

Herbison A.E. (Dunedin, New Zealand)


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

ESTROGEN REGULATION OF GONADOTROPIN-RELEASING HORMONE<br />

NEURONS<br />

Herbison A.E., Porteous R., Clarkson J., Romano N., and Campbell R.E.<br />

Centre for Neuroendocrinology and Department of Physiology, University of Otago,<br />

Dunedin, New Zealand. Fax +64-34797323 e-mail: allan.herbison@otago.ac.nz<br />

Estrogen exerts critical feedback actions upon multiple neuronal networks within<br />

the brain. One of the most important targets for estrogen is the gonadotropin-releasing<br />

hormone (GnRH) neuronal network that controls pituitary gonadotropin secretion and,<br />

thus, fertility, in all mammalian species. However, the GnRH cell bodies exhibit a<br />

scattered topography within the basal forebrain and this has greatly hindered progress in<br />

understanding their biology. Recent transgenic approaches, targeting GnRH neurons with<br />

fluorescent and calcium-sensitive reporter molecules, have been of benefit in<br />

characterizing the cellular and molecular characteristics of adult GnRH neurons in situ.<br />

Furthermore, mice with cell type-specific knockouts of selective receptors have been of<br />

great use in refining information obtained from earlier global knockout mice. In addition,<br />

exciting new transgenic approaches are now being used to define primary and higher-order<br />

inputs to GnRH neurons within the GnRH neuronal network. The use of these transgenic<br />

approaches in elucidating the mechanisms of estrogen feedback regulation of GnRH<br />

neurons will be presented. Data from global and neuron-specific estrogen receptor (ER)<br />

mutant mice demonstrate that ERalpha-expressing neurons are critical for estrogen positive<br />

feedback. As GnRH neurons do not express ERalpha, estrogen feedback must occur<br />

through ERalpha-expressing neuronal afferents to GnRH neurons. Experiments using Credependent<br />

Pseudorabies virus in GnRH-Cre transgenic mice to trace these afferents,<br />

indicate that ERalpha-expressing neurons innervating GnRH neurons are located<br />

predominantly in the periventricular preoptic area [1]. On-going electrophysiological and<br />

imaging studies suggest that neurons expressing the newly discovered neuropeptide,<br />

kisspeptin, provide a critical estrogen-sensitive neuronal population regulating the activity<br />

of GnRH neurons<br />

Reference List<br />

[1] T.M. Wintermantel, R.E. Campbell, R. Porteous, D. Bock, H.J. Grone, M.G.<br />

Todman, K.S. Korach, E. Greiner, C.A. Perez, G. Schutz and A.E. Herbison,<br />

Definition of estrogen receptor pathway critical for estrogen positive feedback to<br />

gonadotropin-releasing hormone neurons and fertility, Neuron 52 (2006) 271-280.<br />

33


SUNDAY, 18 th February<br />

15.00 - 18.00<br />

Symposium:<br />

Effects mediated by classical steroid receptors


Symposium:<br />

Effects mediated by classical steroid receptors<br />

(Chairs: Mani S., Tena-Sempere M.)<br />

• Smith J (Australia) Steroid regulation of kisspeptin signalling in the brain<br />

• Bass AH (USA) Steroid-dependent modulation of vocal motor systems<br />

• Handa RJ (USA) Estrogen receptor beta in the brain: from form to function<br />

• Bodo C, Rissman EF (USA) A role for the androgen receptor in the sexual<br />

differentiation of the olfactory system in mice.<br />

• Ishunina T.A., Swaab D.F. (Russia) Canonical and alternatively spliced<br />

estrogen receptor α in the human mamillary bodies and hippocampus in aging<br />

and alzheimer’s disease<br />

• Klein S., Grossmann R. (Germany) Female specific activation of galanin in the<br />

supraoptic nucleus of hens after oviposition related upregulation of arginine<br />

vasotocin (AVT)<br />

• Sica M., Martini M., Verzè L., Viglietti-Panzica C., Panzica G.C. (Italy)<br />

Expression of nitric oxide synthase in the male mouse limbic system is mediated<br />

by estrogen receptors.


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

STEROID REGULATION OF KISSPEPTIN SIGNALLING IN THE BRAIN<br />

Smith J.T.<br />

Department of Physiology, Monash University, Building 13F, Clayton, Victoria 3800,<br />

Australia. E-mail: jeremy.smith@med.monash.edu.au Fax: +61 3 99052547<br />

The KiSS-1 gene encodes a family of peptides called kisspeptins. Post-translational<br />

processing of an initial 145 amino acid peptide results in the formation of smaller C-<br />

terminal peptides (kisspeptin-54, -14, -13 and -10), which activate with equal efficacy the<br />

G protein-coupled receptor, GPR54. In both humans and mice, inactivating mutations in<br />

GPR54 result in the failure to initiate puberty and hypogonadotropic hypogonadism [1,7],<br />

indicating that these peptides play a vital role in the regulation of GnRH secretion. In many<br />

species, centrally administered kisspeptin stimulates gonadotrophin secretion in a GnRH<br />

dependant manner [4]. Moreover, virtually all GnRH neurons co-express GPR54 [5,6]. In<br />

the hypothalamus, the vast majority of kisspeptin producing cells (those expressing KiSS-1<br />

mRNA) also express sex steroid receptors, particularly oestrogen receptor alpha [9]. Thus,<br />

sex steroids are able to directly regulate the expression of KiSS-1 mRNA, implicating<br />

kisspeptin as the missing link between sex steroids and GnRH feedback. Kisspeptin<br />

producing neurons (or KiSS-1 neurons) have been localised to various regions of the<br />

forebrain in rodents, primates and most recently sheep [2-4,8]. In the arcuate nucleus<br />

(ARC) of the rodent, sex steroids inhibit the expression of KiSS-1 mRNA, suggesting that<br />

the kisspeptin secreting neurons here are the conduit for the negative feedback regulation<br />

of GnRH secretion [9]. However, in the anteroventral periventricular nucleus (AVPV), sex<br />

steroids induce the expression of KiSS-1 mRNA, implying that these kisspeptin neurons<br />

play a role in the positive feedback regulation of rodent GnRH secretion [9]. In sheep,<br />

KiSS-1 neurons appear robustly within the ARC, and a smaller population is present in the<br />

preoptic area (POA)[2,3]. Recent studies in the ewe demonstrate that KiSS-1 mRNA in the<br />

ARC is inhibited by oestrogen (see Figure 1.) and progesterone, but is stimulated<br />

immediately prior to the preovulatory luteinising hormone surge. Interestingly, the AVPV<br />

of the ewe is void of KiSS-1 mRNA expression and the population of KiSS-1 neurons in<br />

the POA is not regulated by sex steroids. Thus, kisspeptin neurons in the ovine ARC<br />

appear well placed to play a role in the negative and positive feedback regulation of GnRH<br />

exerted by sex steroids.<br />

Intact<br />

OVX<br />

OVX+E<br />

3V<br />

3V<br />

3V<br />

Figure 1<br />

Dark-field photomicrographs of the ovine ARC showing KiSS-1 mRNA-expressing cells<br />

(as shown by the presence of silver grain clusters) from gonad-intact, ovariectomised<br />

(OVX), and OVX & oestrogen replacement (E) ewes. 3V = third ventricle.<br />

37


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

Reference List<br />

[1] de Roux, N., Genin, E., Carel, J.C., Matsuda, F., Chaussain, J.L. and Milgrom, E.,<br />

Hypogonadotropic hypogonadism due to loss of function of the KiSS1-derived peptide receptor<br />

GPR54, Proc Natl Acad Sci U S A, 100 (2003) 10972-6.<br />

[2] Estrada, K.M., Clay, C.M., Pompolo, S., Smith, J.T. and Clarke, I.J., Elevated KiSS-1 Expression in<br />

the Arcuate Nucleus Prior to the Cyclic Preovulatory Gonadotrophin-Releasing<br />

Hormone/Lutenising Hormone Surge in the Ewe Suggests a Stimulatory Role for Kisspeptin in<br />

Oestrogen-Positive Feedback, J Neuroendocrinol, 18 (2006) 806-9.<br />

[3] Franceschini, I., Lomet, D., Cateau, M., Delsol, G., Tillet, Y. and Caraty, A., Kisspeptin<br />

immunoreactive cells of the ovine preoptic area and arcuate nucleus co-express estrogen receptor<br />

alpha, Neurosci Lett, 401 (2006) 225-30.<br />

[4] Gottsch, M.L., Cunningham, M.J., Smith, J.T., Popa, S.M., Acohido, B.V., Crowley, W.F.,<br />

Seminara, S., Clifton, D.K. and Steiner, R.A., A role for kisspeptins in the regulation of<br />

gonadotropin secretion in the mouse, Endocrinology, 145 (2004) 4073-7.<br />

[5] Han, S.K., Gottsch, M.L., Lee, K.J., Popa, S.M., Smith, J.T., Jakawich, S.K., Clifton, D.K., Steiner,<br />

R.A. and Herbison, A.E., Activation of gonadotropin-releasing hormone (GnRH) neurons by<br />

kisspeptin as a neuroendocrine switch for the onset of puberty, J Neurosci (2005).<br />

[6] Irwig, M.S., Fraley, G.S., Smith, J.T., Acohido, B.V., Popa, S.M., Cunningham, M.J., Gottsch,<br />

M.L., Clifton, D.K. and Steiner, R.A., Kisspeptin Activation of Gonadotropin Releasing Hormone<br />

Neurons and Regulation of KiSS-1 mRNA in the Male Rat, Neuroendocrinology, 80 (2005) 264-<br />

272.<br />

[7] Seminara, S.B., Messager, S., Chatzidaki, E.E., Thresher, R.R., Acierno, J.S., Jr., Shagoury, J.K.,<br />

Bo-Abbas, Y., Kuohung, W., Schwinof, K.M., Hendrick, A.G., Zahn, D., Dixon, J., Kaiser, U.B.,<br />

Slaugenhaupt, S.A., Gusella, J.F., O'Rahilly, S., Carlton, M.B., Crowley, W.F., Jr., Aparicio, S.A.<br />

and Colledge, W.H., The GPR54 gene as a regulator of puberty, N Engl J Med, 349 (2003) 1614-27.<br />

[8] Shahab, M., Mastronardi, C., Seminara, S.B., Crowley, W.F., Ojeda, S.R. and Plant, T.M., Increased<br />

hypothalamic GPR54 signaling: a potential mechanism for initiation of puberty in primates, Proc<br />

Natl Acad Sci U S A, 102 (2005) 2129-34.<br />

[9] Smith, J.T., Cunningham, M.J., Rissman, E.F., Clifton, D.K. and Steiner, R.A., Regulation of Kiss1<br />

gene expression in the brain of the female mouse, Endocrinology, 146 (2005) 3686-92.<br />

38


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

STEROID-DEPENDENT MODULATION OF VOCAL MOTOR SYSTEMS<br />

Bass A.H.<br />

Department of Neurobiology and Behavior, Cornell University, Ithaca, New York, 14853,<br />

U.S.A. ahb3@cornell.edu, Fax 607-254-1303<br />

Vocal communication is not unique to humans, but rather is a trait shared with most<br />

non-mammalian vertebrates. A practical way to address questions of vocal signal<br />

production and encoding has been to identify mechanisms in non-mammalian model<br />

systems that use acoustic communication signals in their social behavior. A major goal of<br />

this presentation is take a broad phylogenetic view of How and Why steroid hormones have<br />

been adopted as neuromodulators shaping both long-term and short-term plasticity in the<br />

neural mechanisms of vocal behaviors. It is in this context that I will first review my<br />

laboratory’s studies of neuroendocrine mechanisms leading to vocal-auditory plasticity<br />

among teleost fishes.<br />

Teleosts make up nearly half of all living vertebrate species [7]. Vocal teleosts have<br />

a simple repertoire of acoustic communication signals, and auditory and vocal pathways<br />

organized like those of birds and mammals [4]. They also present the simplest and perhaps<br />

primordial example of how a vertebrate nervous system is organized to both produce and<br />

detect social, context-dependent sounds [2, 4]. Studies of vocal species have recently<br />

begun to elucidate a suite of neuroendocrine adaptations for both the production and the<br />

perception of acoustic signals that are essential to reproductive success [3, 4]. I review our<br />

studies of both auditory and vocal mechanisms because both the production and perception<br />

of vocal signals is “in the time domain … the time varying acoustic waveform is the<br />

physical signal that is actually produced by the temporally patterned action of the motor<br />

system under ongoing control of the central nervous system. …. The two systems [vocal<br />

and auditory] co-evolved and we should expect them to share the same underlying code for<br />

signal generation and recognition” [5].<br />

The closely related midshipman fish and toadfish (same family and order) produce<br />

simple, highly stereotyped vocal signals that are essential to social interactions, including<br />

reproduction and aggression [4]. Behavioral studies demonstrate that the temporal features<br />

within a call, including pulse duration, rate and number, can all be important to a call’s<br />

communicative value; hence, our focus on temporal processing [4]. How does the auditory<br />

system encode vocal signals? Neurophysiological studies show that the saccule division of<br />

the inner ear is the main auditory end organ in most teleosts; single neuron recordings<br />

show that afferents within the saccular branch of the eighth cranial nerve encode acoustic<br />

waveforms in the temporal pattern of their action potentials [4]. We recently discovered<br />

that the temporal encoding of frequency via phase-locking by midshipman saccular<br />

afferents exhibits reproductive state and steroid-dependent plasticity [10]. This leads to<br />

enhanced sensitivity to the higher harmonics of male advertisement calls during the<br />

breeding season that likely also contributes to improved mechanisms of sound localization.<br />

We are now studying the cellular and molecular events leading to steroid modulation of<br />

auditory encoding mechanisms in the context of natural shifts in plasma and brain hormone<br />

concentrations [see 9, 11].<br />

Midshipman and toadfish produce sound by the simultaneous contraction of a pair<br />

of sonic muscles attached to the walls of the gas-filled swim bladder. The physical<br />

attributes of the acoustic waveform are a direct translation of the temporal attributes of a<br />

hindbrain-spinal vocal pattern generator (VPG) that shares developmental origins with the<br />

vocal pacemaker circuits of birds and mammals [1, 2]. The VPG is part of a more<br />

39


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

expansive vocal-acoustic network that most closely resembles the organizational pattern in<br />

mammals [6]. We refer to the rhythmic, oscillatory-like output of the VPG (i.e., the sonic<br />

motor volley) as a “fictive vocalization” because its temporal properties set the<br />

fundamental frequency and duration of natural vocalizations [1, 8]. Fictive calls are easily<br />

monitored using intracranial recordings from occipital nerve roots (homologous to<br />

hypoglossal nerve of tetrapods) that give rise to the sonic nerve that innervates each of the<br />

swim bladder muscles [1]. Several studies now show that steroids can induce changes in<br />

the duration and rate of production of fictive calls within minutes [e.g., 8]. Importantly,<br />

field studies establish a causal relationship between rapid shifts in plasma levels of steroids<br />

and vocal behaviors [8].<br />

In sum, several studies in teleost fish now show how naturally occurring steroids<br />

modulate the response properties of auditory neurons and vocal motor neurons that are<br />

essential to, respectively, the encoding and production of vocalizations. As will be<br />

discussed, the principles identified here can be generalized across vertebrates due to the<br />

conserved pattern of auditory-vocal and neuroendocrine mechanisms between vocal<br />

teleosts and tetrapods [e.g., 4, 6].<br />

Acknowledgements. Research support from the U.S. National Institutes of Health<br />

(DC00092) and National Science Foundation (IBN9987341, IOB-0516748).<br />

Reference list<br />

[1] Bass, A.H. and Baker, R., Sexual dimorphisms in the vocal control system of a teleost fish:<br />

morphology of physiologically identified neurons, J Neurobiol, 21 (1990) 1155-1168.<br />

[2] Bass, A.H. and Baker, R., Phenotypic specification of hindbrain rhombomeres and the origins of<br />

rhythmic circuits in vertebrates. Brain, Behav Evol, 50 (1997) 3-16.<br />

[3] Bass, A.H. and Forlano, P.M., Neuroendocrine mechanisms of alternative reproductive tactics: the<br />

chemical language of social plasticity. In R. Oliveira, M. Taborsky and J. Brockmann (Eds.),<br />

Alternative Reproductive Tactics: An Integrative Approach, Cambridge University Press,<br />

Cambridge, UK, in press.<br />

[4] Bass, A.H. and McKibben, J.R., Neural mechanisms and behaviors for acoustic communication in<br />

teleost fish, Prog Neurobiol, 69 (2003) 1-26.<br />

[5] Capranica, R.R., The untuning of the tuning curve: is it time? Sem Neurosci, 4 (1992) 401-408.<br />

[6] Goodson, J. L., and Bass, A.H., Vocal-acoustic circuitry and descending vocal pathways in teleost<br />

fish: Convergence with terrestrial vertebrates reveals conserved traits. J Comp Neurol, 448 (2002)<br />

298-322.<br />

[7] Nelson J.S., Fishes of the world. (1994) New York: Wiley & Sons.<br />

[8] Remage-Healey, L.H. and Bass, A. H., From social behaviour to neurons: Rapid modulation of<br />

advertisement calling and vocal pattern generators by steroid hormones. Horm Behav 50 (2006)<br />

432-441.<br />

[9] Schlinger, B.A., Greco, C. and Bass, A.H., Aromatase activity in the hindbrain vocal control region<br />

of a teleost fish: divergence among males with alternative reproductive tactics, Proc Roy Soc Lond<br />

B-Biol Sci, 266 (1999) 131-136.<br />

[10] Sisneros, J.A., Forlano, P.M., Deitcher, D.L. and Bass, A.H., Steroid-dependent auditory plasticity<br />

leads to adaptive coupling of sender and receiver, Science, 305 (2004) 404-7.<br />

[11] Sisneros, J.A., Forlano, P.M., Knapp, R. and Bass, A.H., Seasonal variation of steroid hormone<br />

levels in an intertidal-nesting fish, the vocal plainfin midshipman, Gen Comp Endocrinol, 136<br />

(2004) 101-116.<br />

40


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

ESTROGEN RECEPTOR BETA IN THE BRAIN: FROM FORM TO FUNCTION<br />

Handa R.J.<br />

Department of Biomedical Sciences, College of Veterinary Medicine, Colorado State<br />

University, Fort Collins, Colorado, USA<br />

Estrogens have numerous effects on the brain both in adulthood and during<br />

development. These actions of estrogen are mediated by two distinct estrogen receptor<br />

(ER) systems, ERα and ERβ. In brain, ERα plays a critical role in estrogen regulating<br />

reproductive neuroendocrine function and behavior, however, a definitive role for ERβ in<br />

any neurobiological function has been slow in forthcoming. Clues to the function of ERβ<br />

in the CNS can be gleaned from the neuroanatomical distribution of ERβ and the<br />

phenotypes of neurons that express ERβ. ERβ immunoreactivity has been found in<br />

populations of GnRH, CRH, vasopressin, oxytocin and prolactin containing neurons. We<br />

have also utilized subtype-selective estrogen receptor agonists to determine the roles for<br />

ERβ in non-reproductive behaviors in a rat model. ERβ selective agonists were found to<br />

exert potent anxiolytic activity when animals were tested in a number of behavioral<br />

paradigms. Consistent with this, ERβ selective agonists also inhibited the ACTH and<br />

corticosterone response to stress. In contrast, ERalpha selective agonists were found to be<br />

anxiogenic and correspondingly increased the hormonal stress response. Taken together,<br />

our studies implicate ERβ as an important modulator of some non-reproductive<br />

neurobiological systems. The molecular and neuroanatomical targets of estrogen that are<br />

mediated by ERβ remain to be determined.<br />

In the course of these studies we have identified a number of splice variants of<br />

ERβ mRNA in brain tissue. Imaging of eGFP labeled receptor proteins in transfected cell<br />

lines has demonstrated that ERβ splice variation can alter trafficking patterns. The<br />

originally described ERβ (herein termed ERβ1) is characterized by possessing a high<br />

affinity for estradiol. Similar to ERα, it is localized in the nucleus and is trafficked to<br />

nuclear sites termed “hyperspeckles” following ligand binding. In contrast, ERβ2 contains<br />

an 18aa insert within the ligand binding domain and as a result can be best described as a<br />

low affinity form of ERβ. A delta3 variant of ERβ has a deletion of the 3 rd exon (coding<br />

for the second half of the DNA binding domain) and as a result does not bind an estrogen<br />

response element in DNA. Delta3 variants are trafficked to a unique low abundance and<br />

larger nuclear site following ligand binding. A delta 4 variant lacks exon 4 and as a result<br />

is localized to the cytoplasm. The amount of individual splice variant mRNAs varies<br />

depending upon brain region. Examination of neuropeptide promoter regulation by ERβ<br />

and its splice variants demonstrate that ERβ functions as a constitutively active<br />

transcription factor. Moreover, it appears that splice variation of ERβ alters its ability to<br />

regulate transcription in a promoter-dependent and ligand-dependent fashion.<br />

41


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

A ROLE FOR THE ANDROGEN RECEPTOR IN THE SEXUAL<br />

DIFFERENTIATION OF THE OLFACTORY SYSTEM IN MICE<br />

Bodo C. 1 and Rissman E.F. 2<br />

1 Graduate Program in Neuroscience and 2 Department of Biochemistry and Molecular<br />

Genetics, University of Virginia, Charlottesville, VA 22908<br />

Phone: 01 434 982 4742; Fax: 01 434 243 8433<br />

Olfactory signals play a central role in the identification of a mating partner in<br />

rodents, and the behavioral response to these cues varies markedly between the sexes. As<br />

several other sexually dimorphic traits, this response is thought to differentiate as a result<br />

of exposure of the developing individual to gonadal steroids, but both the identity of the<br />

specific steroid signal and the neural structures targeted for differentiation on this<br />

particular case are largely unknown. Using genetic males affected by the testicular<br />

feminization syndrome (Tfm) as our experimental model, we have identified a potential<br />

role for non-aromatized gonadal steroids acting through the androgen receptor (AR) in the<br />

differentiation of olfactory cues processing in mice. In contrast with their WT male<br />

littermates, Tfm males spend more time investigating bedding soiled by other males rather<br />

than by receptive females, and they do not a show a clear preference for females as<br />

potential partners. Moreover, when cFos expression on the central projections of the<br />

accessory olfactory system in response to male odors was measured on these individuals,<br />

they exhibited a feminized pattern of activation. Conversely, when WT females were<br />

treated neonatally with the non-aromatizable androgen dihydrotestosterone, their<br />

preference for soiled bedding was masculinized, confirming the hypothesis that AR<br />

activation is solely responsible for the masculinization of the neural circuit that regulates<br />

the detection and processing of non-volatile olfactory cues in mice. We are currently<br />

working to obtain data on partner preference and c-fos response to olfactory signals on<br />

neonatally-androgenized individuals. Considered together, the pieces of data to be<br />

presented represent a good example of the specificity of function of steroids receptors in<br />

the overall process of sexual differentiation of neural circuits in mammals.<br />

This work was supported by NIH grants R01 MH57759 and K02 MH01349.<br />

42


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

CANONICAL AND ALTERNATIVELY SPLICED ESTROGEN RECEPTOR α IN<br />

THE HUMAN MAMILLARY BODIES AND HIPPOCAMPUS IN AGING AND<br />

ALZHEIMER’S DISEASE<br />

Ishunina T.A. 1,2 , and Swaab D.F. 2<br />

1 Department of Histology, Kursk State Medical University, Kursk, Russia<br />

2<br />

Neuropsychiatric diseases, Netherlands Institute for Neuroscience, Meibergdreef 47, 1105<br />

BA, Amsterdam, The Netherlands. Fax +31 20 6961006, e-mail: T.Ishunina@nin.knaw.nl<br />

Beneficial role of estrogens in brain aging and neurodegenerative diseases remains<br />

a controversial topic of research. In addition to multiple mechanisms of estrogen action,<br />

the subject is complicated by a diversity of estrogen receptor isoforms that may mediate<br />

estrogen effects or influence the wild type receptor function.<br />

In the present study we aimed at determining whether estrogen receptor α (ERα)<br />

mRNA and protein are affected in aging and Alzheimer’s disease (AD) in the human<br />

mamillary bodies (MB) and the hippocampus. These brain areas were chosen for this<br />

investigation since they are involved in the regulation of memory, learning, emotions [1].<br />

They are joined within the mamillothalamic circuit and are severely affected in AD [2].<br />

We also addressed the question whether specific ERα mRNA splice variants are present in<br />

these brain areas and change with advanced age and in AD.<br />

ERα protein levels were determined by quantitative immunocytochemistry with<br />

MC-20 antibody (Santa Cruz) [3]. Wild type (wt) ERα mRNA and ERα splice variants<br />

were amplified in RT-PCR with subsequent sequencing and relative quantification in Q-<br />

PCR [4, 5].<br />

Canonical ERα mRNA and protein levels were increased in the MB and were<br />

decreased in the hippocampus of AD patients compared to elderly control cases [3-5].<br />

Exon-skipping ERα variants were common in both brain areas [4, 5]. In the MB the major<br />

ERα splice form was del. 7 [4], lacking exon 7 that encodes a significant portion of the<br />

ligand-binding domain (LBD) [6]. Del. 4 (missing exon 4 encoding a part of the LBD [7])<br />

and del. 2 (lacking exon 2 leading to alterations in the DNA-binding domain [7]) were<br />

present in the MB to a lesser extent than del. 7. In the hippocampus the most abundant<br />

splice variant was del. 4 [5] that does not bind to estradiol and to the estrogen responsive<br />

elements [8]. The del.7 and del. 2 were expressed in the hippocampus to a significantly<br />

lesser extent (del. 4 > del. 7 > del. 2). Interestingly, we have identified two novel ERα<br />

mRNA splice variants: MB1 (mamillary body, exon 1) [4] and TADDI (from the<br />

hippocampus) [5]. MB1 is characterized by a 168 nucleotide deletion in exon 1 encoding<br />

the major portion of the transactivation function 1 domain of the ERα [4]. TADDI is<br />

missing 31 bp at the junction between exons 3 and 4 with an insertion of 13 nucleotides<br />

from the middle of the exon 2 into this splice site [5]. In general, ERα mRNA splice<br />

variants showed the same direction of changes in AD as the wt ERα mRNA amplicons.<br />

While no aging-related changes were observed for the wild type and alternatively<br />

spliced ERα mRNA in the hippocampus of elderly subjects (≥ 46 years of age), canonical,<br />

del. 7 and del. 2 ERα mRNA amplicons declined during aging in elderly control patients ≥<br />

61 years of age in the MB1.<br />

Interestingly, ERα protein levels as judged from quantitative immunocytochemistry<br />

were increased in the CA4, hilus and dentate granular layer of postmenopausal women (58-<br />

83 years of age) compared to young women (34-50 years old). Moreover, young women<br />

(34-50 years of age) showed significantly more ERα in all hippocampal subregions than<br />

43


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

men of the same age. Since no sex differences were found in aromatase immunoreactivity,<br />

this finding is related to higher circulating estrogen levels in young women rather than to<br />

locally synthesized estrogens.<br />

Our data show that ERα splice variants in the brain are region-specific, that they<br />

may change in aging and AD and that they should be considered as potential mediators of<br />

estrogen effects in the human brain and/or as confounders of the normal receptor function.<br />

Reference List<br />

[1] D.F. Swaab, The human hypothalamus: basic and clinical aspects. Part I. Nuclei of the human<br />

hypothalamus. In M.J. Aminoff, F. Boller and D.F. Swaab (series eds.): Handbook of Clinical<br />

Neurology. Elsevier, 2003, vol.79, 3 rd series, vol. 1, pp.502<br />

[2] D.F. Swaab, The human hypothalamus: basic and clinical aspects. Part II. Neuropathology of the human<br />

hypothalamus and adjacent structures. In M.J. Aminoff, F. Boller and D.F. Swaab (series eds.):<br />

Handbook of Clinical Neurology. Elsevier, 2004, vol.80, 3 rd series, vol. 2, pp.632<br />

[3] T.A. Ishunina, W. Kamphorst, D.F. Swaab, Changes in metabolic activity and estrogen receptors in the<br />

human medial mamillary nucleus: relation to sex, aging and Alzheimer's disease, Neurobiol Aging 24<br />

(2003) 817-828.<br />

[4] T.A. Ishunina, D.F. Swaab, D.F. Fischer, Estrogen receptor-α splice variants in the medial mamillary<br />

nucleus of Alzheimer’s disease patients: identification of a novel MB1 isoform, J Clin Endocrinol<br />

Metab 90 (2005) 3757-3765.<br />

[5] T.A. Ishunina, D.F. Fischer, D.F. Swaab, Estrogen receptor α and its splice variants in the hippocampus<br />

in aging and Alzheimer’s disease, Neurobiol Aging (2006): in press.<br />

[6] J.M. García Pedrero, P. Zuazua, C. Martínez-Campa, P.S. Lazo, S. Ramos, The naturally occurring<br />

variant of estrogen receptor (ER) ERΔe7 suppresses estrogen-dependent transcriptional activation by<br />

both wild type ERα and ERβ, Endocrinology 144 (2003) 2967-2976<br />

[7] S. Hirata, T. Shoda, J. Kato, K. Hoshi, Isoform/variant mRNAs for sex steroid hormone receptors in<br />

humans, Trends Endocrinol Metab 14 (2003) 124-129<br />

[8] S.G. Koehorst, J.J. Cox, G.H. Donker, S. Lopes da Silva, J.P. Burbach, J.H. Thijssen, M.A.<br />

Blankenstein, Functional analysis of an alternatively spliced estrogen receptor lacking exon 4 isolated<br />

from MCF-7 breast cancer cells and meningioma tissue, Mol Cell Endocrinol 101 (1994) 237-245<br />

44


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

FEMALE SPECIFIC ACTIVATION OF GALANIN IN THE SUPRAOPTIC<br />

NUCLEUS OF HENS AFTER OVIPOSITION RELATED UPREGULATION OF<br />

ARGININE VASOTOCIN (AVT)<br />

Klein S., and Grossmann R.<br />

Institute for Animal Science Mariensee (FAL), Dept. Functional Genomics and<br />

Bioregulation, Höltystrasse 10, D-315<strong>35</strong> Neustadt, Germany, klein@tzv.fal.de,<br />

fax: +49-5034-871 247<br />

The neuroendocrine AVT system in birds comprises homeostatic function as it does<br />

argi-nine vasopressin in mammals and was additionally shown to be activated for<br />

reproductive functions. Oviposition in birds is accomanied by about sevenfold increase of<br />

plasma AVT concentrations to the unstimulated values with a short half life time of 13 min<br />

and no influence on plasma osmolality [11] [12]. From the nonapeptides of birds only<br />

AVT, but not mesotocin, was found with oviposition related changes in plasma [5]. Thus,<br />

AVT is proposed to perform functions as they are known from oxytocin in mammals.<br />

However, little is known about AVT functions of magnocellular neurons. Oxytocin<br />

secreted from magnocellular supraoptic neurons (SON) in mammals was shown to be<br />

involved in regulation of female reproductive behaviour [9]. The positive feedback mechanism<br />

of oxytocin during birth and lactation needs additional neuromodulatory regulation to<br />

stabilize the system. Galanin was found colocalized with oxytocin in female rats [8] and is<br />

colocalized with vasopressin and oxytocin in lactating rats [7]. Because of its<br />

hyperpolarizing action, galanin was suggested as neuromodulator to curb estrogenic<br />

stimulated dendritic oxytocin release via GABA-ergic inhibition [10] and by presynaptic<br />

action on synaptic inputs [6]. Variable galanin immunoreactivity was found in hens, those<br />

ovulatory cycle was not determined [3].<br />

Previously, we have shown sex differences for galanin immunoreactivity (ir) in<br />

magnocellular SON neurons in chickens [4]. Fluorescent immunohistochemistry revealed<br />

no galanin in male supraoptic neurons as it was previously reported from male quails [1].<br />

Galanin-ir was found in up to 50% of AVT neurons in the rostral half of the SON in hens<br />

perfused within three hours after lights on [4]. In the current study, we investigated the<br />

time course of AVT and galanin-ir in the SON of female chickens from four hours before<br />

until four hours after oviposition. Multitrack confocal imaging (LSM510, Zeiss, Germany)<br />

was performed to localize FITC labeled guinea pig anti-galanin (Chemicon) and Alexa 555<br />

labeled rabbit anti-AVT [2] in perfused brains of adult hens of 25 to 45 weeks of age. The<br />

rostral half of SON, known from mammals to contain most oxytocin synthesizing neurons,<br />

was investigated for the number of neurons and intensity of ir for galanin and AVT. The<br />

intensity of AVT-ir within neurons increased from four hours to 30 minutes before<br />

oviposition by one third. The significant decrease of AVT intensity at 30 minutes after<br />

oviposition implies an oviposition related AVT release from these neurons. Galanin-ir was<br />

found in about 10% of neurons with AVT labeling at four hours and 30 minutes before<br />

oviposition with low intensities. At 30 minutes after oviposition, a significant increase of<br />

the intensity of galanin labeling in an increased number of AVT neurons was seen. Thus,<br />

the proportion of colocalization for galanin and AVT in SON neurons increased to 25 -<br />

30%. At four hours after oviposition, no galanin was detectable in SON of female chickens<br />

with intensities at least three times above background.<br />

The data suggest that AVT-ir is strongly upregulated in the SON neurons from four hours<br />

before oviposition to 30 minutes before oviposition. The decreased intensities of AVT ir at<br />

30 minutes after oviposition indicate a release of AVT at the time of oviposition from SON<br />

45


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

neurons. Galanin synthesis in magnocellular supraoptic neurons is stimulated after the<br />

release of AVT at oviposition. Taken together, these observations indicate that galanin-ir is<br />

upregulated in SON neurons delayed to previous strong stimulation of the AVT system.<br />

Therefore, the data are in agree- ment to the findings of galanin function in negative<br />

feedback control of the upregulated oxytocinergic system in mammals and indicate similar<br />

mechanisms of oxytocin regulation for reproductive behaviour during birth in mammals<br />

and AVT during oviposition in birds.<br />

Reference list<br />

[1] Y. Azumaya and K.Tsutsui, Localization of galanin and its binding sites in the quail brain, Brain Res.<br />

727 (1996) 187-195.<br />

[2] D.A. Gray and E.Simon, Mammalian and avian antidiuretic hormone: studies related to possible species<br />

variation in osmoregulatory systems., J Comp Physiol [A] 151 (1983) 241-246.<br />

[3] R. Jozsa and B.Mess, Galanin-like immunoreactivity in the chicken brain, Cell Tiss Res. 273 (1993)<br />

391-399.<br />

[4] S. Klein, A.Jurkevich, and R.Grossmann, Sexually dimorphic immunoreactivity of galanin and<br />

colocalization with arginine vasotocin in the chicken brain (Gallus gallus domesticus), J Comp Neurol<br />

499 (2006) 828-839.<br />

[5] T.I. Koike, K.Shimada, and L.E.Cornett, Plasma levels of immunoreactive mesotocin and vasotocin<br />

during oviposition in chickens: relationship to oxytocic action of the peptides in vitro and peptide<br />

interaction with myometrial membrane binding sites, Gen Comp Endocrinol 70 (1988) 119-126.<br />

[6] M.G. Kozoriz, J.B.Kuzmiski, M.Hirasawa, and Q.J.Pittman, Galanin modulates neuronal and synaptic<br />

properties in the rat supraoptic nucleus in a use and state dependent manner, J Neurophysiol 96 (2006)<br />

154-164.<br />

[7] M. Landry, D.Roche, E.Angelova, and A.Calas, Expression of galanin in hypothalamic magnocellular<br />

neurons of lactating rats: coexistence with vasopressin and oxytocin, J Endocrinol 155 (1997) 467-481.<br />

[8] M. Landry, A.Trembleau, R.Arai, and A.Calas, Evidence for a colocalization of oxytocin mRNA and<br />

galanin in magnocellular hypothalamic neurons: a study combining in situ hybridization and<br />

immunohistochemistry, Mol Brain Res 10 (1991) 91-95.<br />

[9] I. Neumann, A.J.Douglas, Q.J.Pittman, J.A.Russell, and R.Landgraf, Oxytocin released within the<br />

supraoptic nucleus of the rat brain by positive feedback action is involved in partuition-related events, J<br />

Neuroendocrinol 8 (1996) 227-233.<br />

[10] S. Papas and C.W.Bourgue, Galanin inhibits continuous and phasic firing in rat hyphothalamic<br />

magnocellular neurocsecretory cells, J Neurosci 17 (1997) 6048-6056.<br />

[11] G.E. Rice, S.S.Arnason, Z.Arad, and E.Skadhauge, Plasma concentrations of arginine vasotocin,<br />

prolactin, aldosterone and corticosterone in relation to oviposition and dietary NaCl in the domestic<br />

fowl, Comp Bioch Physiol A 81 (1985) 769-777.<br />

[12] B. Robinzon, N.Sayag, I.T.Koike, S.Kinzler, and H.L.Neldon, Effects of sex and gonadal steroids on<br />

arginine vasotocin and mesotocin int he pineal gland and neurohypophysis of white leghorn fowls, Brit<br />

Poult Sci 31 (1990) 843-849.<br />

46


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

EXPRESSION OF NITRIC OXIDE SYNTHASE IN THE MALE MOUSE LIMBIC<br />

SYSTEM IS MEDIATED BY ESTROGEN RECEPTORS<br />

Sica M., Martini M., Verzè L., Viglietti-Panzica C., and Panzica G.C.<br />

Dept. Anatomy, Pharmacology and Forensic Medicine. Neuroscience Institute of Torino.<br />

University of Torino. Corso M. D’Azeglio 52 10126 Torino Italy.<br />

Nitric oxide (NO) is involved in the control of reproductive functions, and sexual<br />

behavior [3]. NADPH-diaphorase-positive and neuronal nitric oxide synthase (nNOS)<br />

immunoreactive neurons are located in medial preoptic area (MPOM), bed nucleus of stria<br />

terminalis (BST), paraventricular nucleus (PVN) and ventromedial nucleus of the<br />

hypothalamus (VMH) [1]. nNOS expression in the hypothalamus changes by treatments<br />

with testosterone as well as with estrogens, suggesting that effects of T on NOS expression<br />

are controlled by aromatase, the enzyme that catalyzes the biosynthesis of estrogens from<br />

precursor androgens [2,4,5]. The endocrine specificity of nNOS control could thus be<br />

similar to the specificity of the activation of male sexual behavior: both could depend more<br />

on the action of estrogens produced by aromatization than on the action of testosterone<br />

itself.<br />

So far the majority of the studies concerning the effect of E 2 and/ or T on NO synthase<br />

(NOS) expression have been performed mainly in castrated or ovariectomized subjects. To<br />

better describe how the E 2 modulate the expression of NOS immunoreactive system in<br />

mouse hypothalamus we studied the modifications in NOS expression in two mutant<br />

strains: the estrogen receptor α knock out (ERαKO) and the aromatase knock out (ArKO)<br />

mice.<br />

In both models we have detected variations in the expression of neural NOS (nNOS), as<br />

detected by means of immunocytochemistry and quantitative analysis.<br />

In particular in MPOM and PVN we have observed a significant decrease of<br />

immunoreactivity in both ERαKO and ArKO mice. ERαKO mice show also a significant<br />

decrease in the arcuate nucleus, whereas in ArKO mice a significant decrease was<br />

observed at the level of VMH. No changes have been observed at the level of BST and of<br />

the caudate-putamen (a region lacking of estrogen receptors, that was analyzed as a control<br />

nucleus).<br />

These data are a further confirmation that at least part of the hypothalamic nitrinergic<br />

system of male mice is under the control of brain estrogens and that the decrease of NO in<br />

nuclei involved in the control of sexual behavior could be one of the main factors<br />

responsible of the impairment of sexual behavior displayed by these mutant mice.<br />

Moreover, these data indicate that ER-alpha is involved in the control of nNOS expression<br />

is some limbic regions, but not in the whole system. Differences observed at the level of<br />

VMH may indicate a major involvement of ER-beta in this region.<br />

Acknowledgements – The authors want to acknowledge Julie Bakker (Liege) and Emilie<br />

Rissmann (Charlottesville) for having provided the animals used in this study. Supported<br />

by PRIN and University of Torino grants.<br />

47


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

References list<br />

[1] S. Gotti, M. Sica, C. Viglietti Panzica and G.C. Panzica, The distribution of nitric<br />

oxide sythase immunoreactivity in the mouse brain, Micr. Res. Techn. 68 (2005) 13-<br />

<strong>35</strong>.<br />

[2] M. Martini, M. Sica, C. Eva, C. Viglietti Panzica and G.C. Panzica, Dimorphism and<br />

effects of estrous cycle on the nitrinergic system in mouse hypothalamus, Horm.<br />

Behav. 46 (2004) 95-96.<br />

[3] G.C. Panzica, C. Viglietti-Panzica, M. Sica, S. Gotti, M. Martini, H. Pinos, B. Carrillo<br />

and P. Collado, Effects of gonadal hormones on central nitric oxide producing<br />

systems, Neuroscience 138 (2006) 987-995.<br />

[4] S. Sato, C.S. Braham, S.K. Putnam and E.M. Hull, Neuronal nitric oxide synthase and<br />

gonadal steroid interaction in the MPOA of male rats: co-localization and testosteroneinduced<br />

restoration of copulation and nNOS-immunoreactivity, Brain Res. 1043<br />

(2005) 205-213.<br />

[5] E.M. Scordalakes, S.J. Shetty and E.F. Rissman, Roles of estrogen receptor alpha and<br />

androgen receptor in the regulation of neuronal nitric oxide synthase, J. Comp. Neurol.<br />

453 (2002) 336-344.<br />

48


SUNDAY, 18 th February<br />

21.00 - 23.00<br />

Round table I:<br />

Steroid hormones and sexually dimorphic brain circuits


Round table I:<br />

Steroid hormones and sexually dimorphic brain circuits<br />

(Chair: Guillamon A., Panzica G.C.)<br />

• Guillamon A, Segovia S (Spain) Sex differences in the olfactory system of<br />

mammals<br />

• Swaab DF (Netherlands) Human brain sex differences in relation to gender,<br />

sexual orientation and brain disorders<br />

• Patisaul H.B. (USA) Assessing the functional disruption of brain sexual<br />

differentiation by endocrine disrupting compounds<br />

• Micevych P, Phoebe Dewing P. (USA) Sexual differentiation: it’s not just for<br />

development anymore<br />

• Bass A.H. (USA) Sexually polymorphic neuroendocrine phenotypes: examples<br />

from singing fish<br />

• Bakker J (Belgium) Are estrogens required for the development of the female<br />

brain?<br />

• Balthazart J (Belgium) Sexual dimorphism in the avian limbic system


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

SEX DIFFERENCES IN THE OLFACTORY SYSTEM OF MAMMALS<br />

Guillamon A., and Segovia S.<br />

UNED, Madrid, Spain. aguillamon@psi.uned.es<br />

At the beginning of the eighties, and working with neurohistological techniques and<br />

rats, we found that the rat vomeronasal organ is a sexually dimorphic structure<br />

differentiated by gonadal hormones early after birth ( ) and we suggested that the whole<br />

vomeronasal system (VNS) could be sexually dimorphic. Sex differences in the VNS have<br />

functional importance since VNS structures are implicated in the control of sexual and<br />

maternal behavior. The structures that that receive vomeronasal input, such as the<br />

accessory olfactory bulb, medial amygdala, (Me) posteromedial cortical amygdala (C 3) ,<br />

medial posterior region of the bed nucleus of the stria terminalis, , medial preoptic area<br />

(MPA), the ventromedial hypothalamus (VMH) and the premammillary nucleus (PMV)<br />

have androgen and estrogen receptors and present sexual dimorphism (1, 2). Interestingly,<br />

in all these structures the male shows greater morphological measurements (volume and or<br />

number of neurons) than the female rats. From this view we suggested that the VNS of<br />

Rodent and, probably of Mammals could be sexually dimorphic. Recently, we have<br />

investigated the VNS structures of the rabbit, as representative species of Lagomorpha, and<br />

found sexual dimorphism. Females have greater number of mitral and dark and light<br />

granule cells in the accessory olfactory bulb while the males. In the medial amygdala and<br />

in it dorsal and ventral subdivisions , males show greater values than females in volume<br />

and number of neurons while in the posteromedial cortical amygdala females show greater<br />

density of neurons than males. However, in the posteromedial division of the bed nucleus<br />

of the bed nucleus of the stria terminalis males have more neurons than females (3).<br />

More recently, and using voxel-based morphometry, we have investigated the human<br />

olfactory system (Price, ) that still retains functions related to sexual physiology and<br />

sexual behavior. Women have a higher concentration of gray matter in the orbitofrontal<br />

cortex involving Brodmann’s areas 10, 11 and 25 and temporomedial cortex (bilateral<br />

hippocampus and right amygdala), as well as their left basal insular cortex. In contrast,<br />

men show a higher gray matter concentration in the left entorhinal cortex (Brodmann´s 28),<br />

right ventral pallidum, dorsal left insular cortex and a region of the orbitofrontal cortex (4).<br />

Taking into account the existence of sex differences in the olfactory system of rodents,<br />

lagomorphs and humans, and the fact that some structures that receive olfactory input such<br />

as the sexually dimorphic nucleus of the medial preoptic areas and the posteromedial<br />

division of the bed nucleus of the stria terminalis are sexually dimorphic in a wide range of<br />

species, we suggest that the mammalian olfactory system is a sexually dimorphic network,<br />

with species specifics characteristics. This could provide a theoretical framework for the<br />

morphofunctional approach to sex differences in mammals. Such an approach will help to<br />

exploit the abundant literature of animal data on sex differences and sexual behavior and<br />

facilitate the construction of a neurobiological motivational model that could explain the<br />

function of the sex differences found in the human brain.<br />

51


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

References list<br />

(1) Segovia, S and Guillamon, A., Sexual differences in the vomeronasal pathway and<br />

sex differences in reproductive behavior, Brain Res. Review. 18: 51-74, 1993.<br />

(2) Simerly, R.B., Wired for reproduction: organization and development of sexually<br />

dimorphic circuits in the mammalian brain. Ann. Rev. Neurosci., 25: 507-536,<br />

2002.<br />

(3) Segovia, S., Garcia-Falgueras, A., Carrillo, B., Collado, P., Pinos, H, Perez-Laso,<br />

C., Vinader-Caerols, C., Beber, C., and Guillamon, A., Sexual dimorphism in the<br />

vomeronasal system of the rabbit, Brain Res. 1102: 52-62, 2006.<br />

(4) Garcia-Falgueras, A., Junque, C., Jiménez, M., Caldú, X., Segiovia, S., and<br />

Guillamon, A., Sex differences in the human olfactory system. Brain Res. 1116:<br />

103-111, 2006.<br />

52


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

HUMAN BRAIN SEX DIFFERENCES IN RELATION TO GENDER, SEXUAL<br />

ORIENTATION AND BRAIN DISORDERS<br />

Swaab D.F.<br />

Netherlands Institute for Neuroscience, Amsterdam, The Netherlands.<br />

(d.f.swaab@nin.knaw.nl)<br />

Sex differences in behaviour are present from early childhood onwards, e.g. in our<br />

playing behaviour and drawings, followed by cognition, reproduction, gender (the feeling<br />

to be male or female) and sexual orientation, and the incidence of neurological and<br />

psychiatric disorders. All these differences are presumed to be based upon structural and<br />

functional sex differences that are present all over the human brain. They arise during<br />

development by an interaction of sex hormones and the developing neurons, although<br />

direct genetic effects are probably also involved. Factors influencing structural en<br />

functional sex differences in the brain are genetic factors like mutations or polymorphisms<br />

in the sex hormone receptors, abnormal prenatal hormone levels and compounds that<br />

interact with the action of sex hormones on the brain during early development such as<br />

anticonvulsants, DES and environmental endocrine disrupters. An influence of postnatal<br />

social factors on gender or sexual orientation has not been established. In rodents,<br />

masculinization of the brain in development is due to estrogens that are formed by<br />

aromatization of testosterone. In sexual differentiation of the human brain direct effects of<br />

testosterone seem to be of primary importance as is clear e.g. from subjects with mutations<br />

in the gene for the androgen receptor, estrogen receptor or aromatase.<br />

In transsexuals we observed a reversal of the sex difference in the Bed Nucleus of<br />

the Stria terminalis (BSTc). The size, type of innervation and neuron number agreed with<br />

their gender and not with their genetic sex.<br />

Various brain differences related to sexual orientation have now also been reported.<br />

In addition, there are sex differences present in the way the brain ages and in Alzheimer<br />

neuropathology. The field is becoming extra complex by the presence of splice variants<br />

and isoforms of estrogen receptor-α and the local production of steroid hormones in the<br />

brain. This means that sex differences may be expected in many functions in all stages of<br />

life, in heath as well as in disease.<br />

53


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

ASSESSING THE FUNCTIONAL DISRUPTION OF BRAIN SEXUAL<br />

DIFFERENTIATION BY ENDOCRINE DISRUPTING COMPOUNDS<br />

Patisaul H.B.<br />

North Carolina State University, Department of Zoology, 127 David Clark Labs, Raleigh,<br />

NC 27695 USA, fax: (919) 515-5327, heather_patisaul@ncsu.edu<br />

Developmental disruption of the reproductive neuroendocrine system is likely to be<br />

expressed as subtle changes in adult behaviors and functions, rather than overt changes in<br />

brain anatomy and reproductive physiology. Therefore, in developing predictive strategies<br />

for evaluating the ultimate effects of early endocrine active compound (EAC) exposure, it<br />

is critical to employ a comprehensive approach that assesses neuronal function and<br />

reproductive physiology across the lifespan. We have found that neonatal exposure to<br />

EACs such as Bisphenol-A and genistein can affect sexually dimorphic brain morphology<br />

and neuronal phenotypes in adulthood with regional and cellular specificity. We have also<br />

found that EACs can simultaneously enhance and interfere with the effects of endogenous<br />

estrogen on the developing brain, making it difficult to extrapolate effects seen in one<br />

region to more global inferences about potential EAC effects on sexually dimorphic<br />

development and behavior. However, subtle impairments to sex-specific behaviors,<br />

fertility, and cyclicity could collectively have significant ramifications for the fitness of an<br />

affected population, including human populations. Therefore it is imperative that the<br />

effects of EAC exposure during development, when sexually dimorphic systems are<br />

organizing, are appreciated.<br />

54


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

SEXUAL DIFFERENTIATION: IT’S NOT JUST FOR DEVELOPMENT<br />

ANYMORE<br />

Micevych P., and Dewing P.<br />

Dept of Neurobiology, MRRC and Lab of Neuroendocrinology of the Brain Research<br />

Institute, David Geffen School of Medicine at UCLA, Los Angeles, CA 90095<br />

Brain sex differences are the result of events beginning with sex-determining genes<br />

and continuing with the actions of sex steroids throughout the lifetime of the mammalian<br />

species. In particular, gonadal steroid hormones from the developing embryo are the<br />

primary signals which initiate brain sexual differentiation. As a result, anatomical and/or<br />

morphological dimorphisms within the brain arise, which then translate into differences in<br />

brain function and behavior. Neurochemical sex differences, unlike the permanent<br />

morphological sex differences, co-vary with the steroid hormone environment that<br />

influences reproduction. One example is the neuropeptide cholecystokinin (CCK), which<br />

has been a useful marker to study sex differences and the action of steroids on the brain.<br />

CCK is found in the limbic system and hypothalamic circuitry that regulate reproductive<br />

behavior. Animals are born with a sexual dimorphism in the number of CCK cells that<br />

favor males. This slight sex difference, in the absence of hormones, becomes dramatic with<br />

the rise of sex steroid hormone levels during puberty. Estradiol in male and female rats<br />

increases CCK expression, resulting in a neurochemical sexual dimorphism favoring males<br />

throughout the limbic-hypothalamic lordosis regulating circuit. Despite this sex difference,<br />

CCK does not affect male copulatory behavior but rather mediates female sexual<br />

receptivity. In fact, CCK can augment lordosis behavior in castrated adult male rats treated<br />

with estradiol. In these male rats displaying lordosis, CCK expression in the lordosis<br />

regulating circuitry resembles that of female rats. Treatment with testosterone induces a<br />

greater expression of CCK but no female behavior, suggesting that testosterone, in addition<br />

to its actions on CCK expression, induces a circuit in male brains that is inhibitory to the<br />

display of lordosis behavior.<br />

Although the effects of sex steroid hormones are undeniable, there are instances<br />

where genetic determinants influence the brain causing sex differences that are<br />

independent of hormones. For example, Sry, the critical gene for sex determination in<br />

mammals has been identified in male mouse and human brain, but not in female brain. In<br />

the adult mouse brain, transcripts of Sry have been found specifically in the dopaminergic<br />

neurons of the substantia nigra. Down regulation of Sry expression in the brain<br />

significantly reduces the levels of dopamine in these nigro-striatal neurons ultimately<br />

altering mobility and limb-usage. These studies underscore the action of a gene, Sry,<br />

encoded only in the male genome to mediate a direct male-specific effect on the brain,<br />

independent of gonadal hormones.<br />

Together these studies demonstrate the complex interaction between genes and sex<br />

steroid hormones on the one hand, and the structure and neurochemistry on the other that<br />

determine the sexual dimorphism of the brain. These interactions influence brain functions<br />

as obviously dimorphic as reproduction and as opaque as motor function.<br />

55


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

SEXUALLY POLYMORPHIC NEUROENDOCRINE PHENOTYPES: EXAMPLES<br />

FROM SINGING FISH<br />

Bass A.H.<br />

Department of Neurobiology and Behavior, Cornell University, Ithaca, New York, 14853, U.S.A.<br />

ahb3@cornell.edu, Fax 607-254-1303<br />

Teleost fishes comprise nearly half of all living vertebrate species. They also show<br />

the greatest range of reproductive plasticity, including species with individuals that show<br />

sex and role change during their lifetime to species with two male morphs that permanently<br />

adopt alternative reproductive tactics (ARTs). We have extensively studied behavioral,<br />

neural and endocrine polymorphisms in the midshipman fish, Porichthys notatus, which<br />

has male ARTs [see 1, 2]. While the term sexual dimorphism applies, strictly speaking, to<br />

morphological dimorphisms, we as others have used the term to encompass behavioral and<br />

physiological mechanisms that diverge both within and between the sexes.<br />

Type I male midshipman build nests under rocky shelters in the intertidal zone and<br />

produce very long duration (mins - > 1h) advertisement calls (“hums”) to attract females to<br />

their nest. Smaller, non-humming, non-territorial type II males sneak- or satellite-spawn to<br />

steal fertilizations from a resident type I. Type I males also generate long trains of brief<br />

grunts (msec) during nest defense and egg guarding. So far, type II males and females<br />

(neither of which participate in nest/egg defense) have only been found to produce isolated<br />

grunts. We identified a parallel suite of intra- and intersexual dimorphisms in vocal traits<br />

along multiple dimensions, from neuromuscular junctions to the dimensions of individual<br />

neurons in a hindbrain-spinal vocal pattern generator (VPG) that establishes the temporal<br />

properties of natural vocalizations. Developmental studies of vocal motor traits support the<br />

hypothesis that type I and II males follow have distinct life history trajectories. Type II<br />

males become sexually mature at an earlier age than type I’s that essentially tradeoff early<br />

reproduction and gonadal maturation for a larger body size and an expansive vocal motor<br />

system. Type II males and females are convergent, both of which are divergent from type<br />

I’s, in vocal traits. These and other studies show an uncoupling of gonadal and behavioral<br />

(vocal) sex from neural mechanisms and an evolutionarily adaptable patterning of these<br />

traits (i.e., gonad, behavior and nervous system). Thus, type II males, like females and<br />

juveniles, essentially lack the behavioral, neurophysiological and morphological traits that<br />

allow type I males to produce a more dynamic vocal repertoire that functions in female<br />

courtship and territorial defense. This includes studies of the rapid neuromodulatory-like<br />

actions of neuropeptides [2] and androgenic steroids [see 3] on the vocal system.<br />

Might comparable dimorphisms shape the widespread evolution of ARTs among<br />

teleost fish [4] as well as the expression of intrasexual behavioral and phenotypes among<br />

other vertebrate taxa, including birds and mammals?<br />

Acknowledgements. Research support from U.S. NIH (DC00092), NSF (IOB-0516748).<br />

Reference list<br />

[1] Bass A.H., Dimorphic male brains and alternative reproductive tactics in a vocalizing fish. Trends<br />

Neurosci, 15 (1992)139-145.<br />

[2] Bass, A.H., Shaping brain sexuality. Amer Sci, 84 (1996) <strong>35</strong>2-363.<br />

[2] Goodson, J.L. and Bass, A.H., Forebrain peptide modulation of sexually polymorphic vocal motor<br />

circuitry. Nature, 403 (2000) 769-772.<br />

[3] Remage-Healey, L.H. and Bass, A.H., Rapid, hierarchical modulation of vocal patterning by steroid<br />

hormones. J Neurosci, 24 (2004) 5892-5900.<br />

[4] Mank, J.E. and Avise J.C., Comparative phylogenetic analysis of male alternative reproductive tactics in<br />

ray-finned fishes. Evol, 60 (2006) 1311-1316.<br />

56


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

ARE ESTROGENS REQUIRED FOR THE DEVELOPMENT OF THE FEMALE<br />

BRAIN?<br />

Bakker J.<br />

Center for Cellular & Molecular Neurobiology, University of Liège, Liège, Belgium<br />

The classic view of sexual differentiation in mammalian species holds that sex<br />

differences in the brain and behavior develop under the influence of estrogens derived<br />

from the neural aromatization of testosterone: the brain develops as male in the presence of<br />

estrogens and as female in their absence. In agreement with this view, it has been<br />

proposed 1 that the female brain needs to be protected from estrogens produced by the<br />

placenta and that alpha-fetoprotein (AFP) - a major fetal plasma protein present in many<br />

developing vertebrate species and produced transiently in great quantities by the<br />

hepatocytes of the fetal liver among other sources – is the most likely candidate to achieve<br />

this protection because of its estrogen-binding capacity. However, the idea that the female<br />

brain develops in the absence of estrogens and the role of AFP in protecting the brain<br />

against the differentiating action of estrogens have been challenged. First, there is<br />

accumulating evidence that the normal development of the female brain might actually<br />

require the presence of estrogens. The most recent evidence comes from our study<br />

revealing that female aromatase knockout (ArKO) mice that are deficient in estrogens due<br />

to a targeted mutation in the aromatase gene showed reduced levels of female sexual<br />

behavior 2 . Second, the presence of AFP within neurons in the absence of any evidence for<br />

local AFP synthesis suggests that AFP is transported from the periphery into the brain. It<br />

was thus proposed 3 that AFP acts as a carrier, which actively transports estrogens into<br />

target brain cells and, by doing so, has an active role in the development of the female<br />

brain.<br />

The recent introduction of an AFP mutant mouse model (AFP-KO 4 ) now finally<br />

allowed us to resolve this longstanding controversy concerning the role of AFP in brain<br />

sexual differentiation, and thus to determine whether prenatal estrogens contribute to the<br />

development of the female brain. We 5 found that AFP-KO females showed no female<br />

sexual behavior at all, and that normal levels of this behavior could be induced by blocking<br />

estrogen action during embryonic development, demonstrating that the principal action of<br />

prenatal estrogen exposure is to defeminize individuals (i.e. to decrease their capacity to<br />

display female sexual behavior later in life) and that AFP normally binds estradiol<br />

circulating in the female fetus and thereby protects the developing brain from<br />

defeminization. Thus our findings 5 corroborate the hypothesis originally proposed 1 and<br />

argue against the model that considers AFP a carrier delivering estrogen to the brain 3 . Why<br />

AFP is present in brain cells in some brain regions but not in others remains unclear,<br />

however. Likewise, the question whether estrogens are actually needed for the<br />

development of the female brain has not been resolved yet. Behavioral data from the AFP-<br />

KO 5 and ArKO 2 mouse models suggest that estrogens can have both defeminizing and<br />

feminizing effects on the developing brain mechanisms that control sexual behavior.<br />

Therefore, we suggest here that the defeminizing action of estradiol normally occurs<br />

prenatally in males and is avoided in fetal females because of the protective actions of<br />

AFP. Furthermore, the feminizing action of estradiol normally occurs in genetic females<br />

between birth and the age of puberty, when the ovaries start to produce estrogens and AFP<br />

no longer plays a role of significance.<br />

57


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

Reference List<br />

1. McEwen BS, Plapinger L, Chaptal C, Gerlach J, Wallach G (1975). Brain Res. 96:<br />

400-406.<br />

2. Bakker J, Honda S, Harada N, Balthazart J (2002). J. Neurosci. 22: 9104-9112.<br />

3. Toran-Allerand CD (1984). Prog. Brain Res. 61: 63-98.<br />

4. Gabant P, Forrester L, Nichols J, Van Reeth T, De Mees C, Pajack B, Watt A, Smitz<br />

J, Alexandre H, Szpirer C, Szpirer J (2002). PNAS 95: 6965-6970.<br />

5. Bakker J, De Mees C, Douhard Q, Balthazart J, Gabant P, Szpirer J, Szpirer C (2006).<br />

Nat. Neurosci. 9:220-226.<br />

58


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

SEXUAL DIMORPHISM IN THE AVIAN LIMBIC SYSTEM<br />

Balthazart J.<br />

University of Liege, Center for Cellular and Molecular Neurobiology, Research Group in<br />

behavioral Neuroendocrinology, 1 avenue de l’Hopital (B36), B-4000 Liege, Belgium<br />

Sex differences in reproductive behavior can be activational in nature, i.e. reflect<br />

sex differences in the endocrine milieu of the adult subjects, or organizational, i.e.<br />

represent irreversible effects of sex steroids during a critical period of the early<br />

development. The Japanese quail (Coturnix japonica) provides an excellent model to study<br />

these sex differences because both types co-exist in the same species. Sexually mature<br />

males exhibit a pre- and post-copulatory display called strutting and attract females with<br />

the use of a loud distinctive vocalization, the crow. Females never show these behaviors<br />

but this sex difference reflects exclusively the higher plasma concentration of testosterone<br />

in males as compared to females. Castration eliminates these behaviors in males and<br />

females treated with testosterone will display the behaviors exactly like males. Sexually<br />

mature males exposed to a female will also display a characteristic sequence of behaviors<br />

ultimately leading to the actual copulation, the male-typical copulatory sequence. These<br />

behaviors are never observed in females and cannot be activated in ovariectomized females<br />

by treatments with testosterone even in doses much larger than the minimally active dose<br />

in males. This sex difference in responsiveness to testosterone is organized before day 12<br />

of embryonic life by ovarian steroids. Indeed male embryos treated before day 12 of<br />

incubation with estradiol will be unable as adults to show the copulatory behavior in<br />

response to testosterone and conversely, females treated before day 12 with an aromatase<br />

inhibitor (that prevents secretion of estrogens by the ovary) will display as adult the full<br />

masculine behavioral phenotype in response to testosterone.<br />

If the endocrine control of the ontogeny of these sex difference is well understood, we have<br />

in contrast a very poor understanding of the brain mechanisms that mediate these sex<br />

differences, in other words we do not know what is different between the brain of a male<br />

and of a female that is organized by early estrogen action and justifies the sex difference in<br />

responsiveness to testosterone.<br />

Three lines of investigations have approached this problem. First neuroanatomical<br />

differences were investigated and volumetric differences between corresponding brain<br />

regions in males and females were identified. The medial preoptic nucleus (POM) is for<br />

example larger in males than in females but this difference is essentially activational and<br />

disappears in gonadectomized subjects of both sexes treated with similar doses of<br />

testosterone. The size of neurons in the dorso-lateral part of this nucleus is however also<br />

larger in males than in females and this difference seems to result from organizational<br />

effects of embryonic estrogens.<br />

A second line of research has investigated sex differences in neurochemical features of<br />

specific brain regions. A number of sex differences in peptide distribution were identified<br />

but again most of these differences appear activational in nature and thus cannot explain<br />

sex differences in responsiveness to testosterone. The density of the vasotocinergic<br />

innervation of the POM, bed nucleus of the stria terminalis and lateral septum is, however,<br />

higher in males than in females, demasculinized by embryonic treatment of males with<br />

estrogen and masculinized in females treated in ovo with an aromatase inhibitor. This<br />

neurochemical difference, which is controlled by the same endocrine mechanisms as<br />

sexual behavior is thus a likely candidate to explain the sex difference in behavior. The<br />

59


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

specific role of vasotocin in the control of male sexual behavior is however poorly<br />

understood.<br />

Finally a third line or research has tried to determine whether sex differences in behavior<br />

could result from differences in connectivity between brain regions. The quail POM<br />

projects to a premotor nucleus, the periaqueductal central gray (PAG) and this connection<br />

probably plays an important role in the control of sexual behavior. Retrograde tract tracing<br />

recently showed that males have more aromatase-immunoreactive neurons in the POM<br />

projecting to the PAG than females and this difference was most prominent in the caudolateral<br />

part of the nucleus that has been specifically implicated in the control of male<br />

copulatory behavior. These data therefore support the hypothesis that sex differences in<br />

POM-PAG connectivity are causally linked to the sex difference in the behavioral response<br />

to testosterone.<br />

60


MONDAY, 19 th February 2007<br />

08.30 - 12.00<br />

Symposium:<br />

Neuroactive steroids and neurogenesis


Symposium:<br />

Neuroactive steroids and neurogenesis<br />

(Chairs: Herbison A.E., Micevych P.)<br />

• Galea L.A.M., Barha C., Barker J.M., Pawluski J.L. Spritzer M.D. (Canada)<br />

Gonadal hormone regulation of adult hippocampal neurogenesis<br />

• Wang Z, Fowler CD (USA) Estrogen, amygdala, and adult neurogenesis<br />

• Brinton RD, Wang J, Irwin R, Liu L, Chen S, Chung E (USA) Allopregnanolone<br />

regulation of proliferation of human neural stem cells and neurogenesis in triple<br />

transgenic Alzheimer's disease mice<br />

• Abrous N (France) Neurogenesis and age-related-cognitive functions: implication<br />

of steroids<br />

• Herbert J (UK) control of neurogenesis in the adult hippocampus by corticoids and<br />

serotonin<br />

• Lecanu L., Ibrahim A., Yao W., McCourty A., Greeson J., Papadopoulos V.<br />

(USA) In vitro and in vivo induced neurogenesis by the naturally occurring steroid<br />

solasodine is associated with GAP-43/HuD pathway activation and increase of the<br />

translocator protein (18 kDa) (TSPO) expression


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

GONADAL HORMONE REGULATION OF ADULT HIPPOCAMPAL<br />

NEUROGENESIS<br />

Galea L.A.M., Barha C., Barker J.M., Pawluski J.L. and Spritzer M.D.<br />

Program in Neuroscience, Department of Psychology and Brain Research Centre,<br />

University of British Columbia, 2136 West Mall, Vancouver B.C., V6T 1Z4, CANADA,<br />

Fax: 001-604-822-6923, email: lgalea@psych.ubc.ca<br />

Gonadal hormones modulate neurogenesis in the dentate gyrus differentially in<br />

male and female adult rodents. Neurogenesis is comprised of both cell proliferation (the<br />

production of new cells) and cell survival (the number of new cells that survive to<br />

maturity). Acute estradiol initially enhances and subsequently suppresses cell proliferation<br />

in the dentate gyrus of adult female rodents, but has limited effects in male rodents. Lowlevel<br />

progesterone appears to attenuate the estradiol-induced enhancement of cell<br />

proliferation in female rodents. Intriguingly the estradiol-induced upregulation of<br />

hippocampal neurogenesis seems to be limited to 17-β estradiol and not extended to other<br />

forms of estrogens such as estrone or 17α-estradiol. Chronic levels of estradiol (15-30 d)<br />

also modulate hippocampal neurogenesis and cell death in adult female, but not male,<br />

rodents. However short-term estradiol treatment (5 days) in males enhances new cell<br />

survival in the dentate gyrus but only when administered during the ‘axon-extension’<br />

phase. Testosterone and dihydrotestosterone upregulate hippocampal neurogenesis (via<br />

cell survival), but not cell proliferation, in adult male rodents. These effects of gonadal<br />

hormones on adult neurogenesis are not limited to exogenous manipulations but are also<br />

observed during endogenous fluctuations in hormones, such as during the breeding versus<br />

non-breeding seasons in males and females and during pregnancy and lactation in the<br />

mother. Pregnancy and motherhood differentially regulate adult hippocampal neurogenesis<br />

in the adult female rodent, with primiparous rats displaying lower levels of hippocampal<br />

cell proliferation and survival after parturition and multiparous rats displaying enhanced<br />

levels of hippocampal cell survival at this time. Thus, gonadal hormones and hormones<br />

associated with pregnancy and lactation appear to alter hippocampal neurogenesis. There<br />

are very few studies comparing males and females but those that have indicate that there is<br />

a sex difference in the response to hormone-regulated hippocampal neurogenesis in the<br />

adult. Clearly more work needs to be done to elucidate the effects of gonadal hormones on<br />

neurogenesis in the dentate gyrus of both male and female rodents, especially if we are to<br />

use our knowledge of how adult neurogenesis is regulated to develop strategies to repair<br />

neuron loss in neurodegenerative diseases.<br />

63


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

ESTROGEN, AMYGDALA, AND ADULT NEUROGENESIS<br />

Wang Z.X., and Fowler C.D.<br />

Department of Psychology and Neuroscience Program, Florida State University,<br />

Tallahassee, FL 32306, USA<br />

In addition to the dentate gyrus of the hippocampus (DG) and subventricular zone,<br />

neurogenesis has been documented in other brain regions, including the amygdala. Given<br />

the role of the amygdala in sensory processing and social behavior, our goal was to<br />

examine adult neurogenesis in this brain region in species differing in social behaviors.<br />

Male exposure and mating increased cell proliferation in the amygdala, but not the DG, of<br />

female prairie voles (Microtus ochrogaster). Since such male experience leads to increased<br />

levels of estrogen, we next investigated the effects of estrogen on neurogenesis. Estradiol<br />

treatment was ineffective in female prairie voles, but increased the density of new cells in<br />

the amygdala of female meadow voles (M. pennsylvanicus). The majority of these new<br />

cells also co-expressed estrogen receptor α, suggesting a potential direct effect of estrogen<br />

on proliferation. In male meadow voles, treatment of testosterone or estradiol, but not<br />

dihydrotestosterone, enhanced cell proliferation in the amygdala, further supporting the<br />

notion of estrogen-mediated adult neurogenesis. Finally, preliminary data has shown that<br />

anti-mitotic drug treatment decreases the number of new cells in the amygdala and inhibits<br />

the formation of a pair bond (a behavior mediated by the amygdala) in female prairie voles.<br />

Together, these data indicate that steroid hormones affect adult neurogenesis in the<br />

amygdala in a species-specific manner, and new cells may have functional significance in<br />

the vole’s social behavior.<br />

64


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

ALLOPREGNANOLONE REGULATION OF PROLIFERATION OF HUMAN<br />

NEURAL STEM CELLS AND NEUROGENESIS IN TRIPLE TRANSGENIC<br />

ALZHEIMER'S DISEASE MICE<br />

Brinton R.D., Wang J.M., Irwin R., Liu L., Chen S. and Chung E.<br />

Department of Pharmacology and Pharmaceutical Sciences, School of Pharmacy, and<br />

Neuroscience Program, University of Southern California, Los Angeles, CA, USA<br />

The neuroendocrine status of the brain has been linked to the quality of the aging<br />

process, to risk of Alzheimer’s disease and to progression of neurodegenerative pathology.<br />

Data from multiple levels of analysis ranging from in vitro cellular investigations to in vivo<br />

studies in animal models of aging and disease to observational investigations of health<br />

outcomes in humans, indicate that gonadal steroid hormones and their metabolites can<br />

promote neural health whereas their decline or absence are associated with decline in<br />

neural health and increased risk of neurodegenerative disease including Alzheimer’s.<br />

Among the steroids in decline, is allopregnanolone (APα), a neurosteroid metabolite of<br />

progesterone, which was found to be reduced in the serum and plasma and brain of aged<br />

vs. young subjects. Further, Alzheimer disease (AD) victims showed an even further<br />

reduction in plasma and brain levels of APα relative to age-matched neurologically normal<br />

controls. Our earlier work demonstrated that APα is a neurogenic agent for rodent<br />

hippocampal neural progenitors and for human neural progenitor cells derived from the<br />

cerebral cortex. Our ongoing research seeks to determine the neurogenic potential of APα<br />

in the triple transgenic mouse model of Alzheimer’s disease (3xTgAD) as AD related<br />

pathology progresses from imperceptible to mild to severe. Initial analyses suggest that<br />

neurogenic potential changes with age in nontransgenic mice and that the neurogenic<br />

pro<strong>file</strong> differs between non-transgenic and 3xTgAD mice. Comparative analyses indicate<br />

that APα modifies neurogenesis in both nontransgenic and 3xTgAD mice. Preliminary<br />

data suggest that APα may modify Alzheimer’s pathology progression. Together the data<br />

suggest that APα could maintain the regenerative ability of the brain and modify<br />

progression of AD related pathology.<br />

Acknowledgements: This work was supported by a grant from the Institute for Study of<br />

Aging and the Kenneth T. and Eileen L. Norris Foundation to RDB.<br />

65


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

NEUROGENESIS AND AGE-RELATED COGNITIVE FUNCTIONS:<br />

IMPLICATION OF STEROIDS<br />

Abrous D.N.<br />

INSERM U588, Institut F Magendie, University of Bordeaux 2, 146 rue Léo-Saignat,<br />

Bordeaux Cedex 33077, France.<br />

Aging is associated with cognitive dysfunction, which has been correlated to an<br />

alteration of hippocampal functioning Indeed, the hippocampal formation (HF) plays a<br />

crucial role in controlling cognitive functions, and is the brain region most vulnerable to<br />

ageing processes. The mammalian HF, in particular the dentate gyrus (DG), is an important<br />

site for the production of new neurons during adulthood. The aim of our work is to<br />

determine the role of adult neurogenesis in the appearance of age-related cognitive deficits.<br />

We have found that cognitively-impaired senescent rats display lower levels of<br />

neurogenesis than cognitively-unimpaired old rats. We have further shown that these interindividual<br />

differences result from early deleterious life events. Indeed, prenatal stress<br />

orients neurogenesis in pathological ways for the entire life, and precipitates age-related<br />

cognitive impairments. More importantly, we have recently fond that the consequences of<br />

prenatal stress on hippocampal plasticity can be reversed by a form of postnatal<br />

environmental stimulation, neonatal handling. Finally, we have highlighted that inhibition<br />

or stimulation of neurogenesis is one of the mechanisms by which glucocorticoids may<br />

fragilize and neurosteroids may protect, respectively, the cognitive functions during aging.<br />

Altogether these data strengthen that hippocampal neurogenesis plays a pivotal role in the<br />

development of pathological aging and reinforce the hypothesis of an early<br />

neurodevelopmental origin for psychopathological vulnerabilities in aging.<br />

66


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

CONTROL OF NEUROGENESIS IN THE ADULT HIPPOCAMPUS BY<br />

CORTICOIDS AND SEROTONIN<br />

Herbert J.<br />

University of Cambridge UK<br />

Neurogenesis in the dentate gyrus of the hippocampus in the adult is not only<br />

highly active but also highly labile. Glucocorticoids are a potent regulators of progenitor<br />

cell proliferation, maturation of new neurons, and their survival to form new connections<br />

with the established circuitry that links the dentate gyrus with the entorhinal cortex and the<br />

CA3 region of the pyramidal layer of the hippocampus. Since corticoids are themselves<br />

labile and are driven by events external to the animal, this means that neurogenesis is<br />

responsive to both predictable changes in the environment (eg time of day) as well as to<br />

more adventitious events (eg stressors). Corticoids interact with serotonin, a second<br />

powerful modulator of neurogenesis. The presence of the diurnal rhythm in corticosterone<br />

(in the rat) seems particularly essential for some of the other control systems to regulate<br />

neurogenesis. For example, abolishing this rhythm prevents the ability of fluoxetine (an<br />

SSRI that increases serotonin) stimulating progenitor cell division rates. The mechanism<br />

for this interaction is under investigation. Corticoids also interact with nitric oxide (NO),<br />

an intracellular regulator of neurogenesis, at least partly by altering the activity of nitric<br />

oxide synthases (NOSs). Current studies point to the possibility that BDNF may be a<br />

common endpoint for these control systems. The current interest in linking major<br />

depression in man with rhythmic changes in cortisol, neurogenesis in the dentate gyrus,<br />

activity of serotonin and polymorphisms in BDNF suggest that unraveling the control<br />

systems that regulate neurogenesis may have clinical as well as experimental interest.<br />

67


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

IN VITRO AND IN VIVO INDUCED NEUROGENESIS BY THE NATURALLY<br />

OCCURRING STEROID SOLASODINE IS ASSOCIATED WITH GAP-43/HUD<br />

PATHWAY ACTIVATION AND INCREASE OF THE TRANSLOCATOR PROTEIN<br />

(18 kDa) (TSPO) EXPRESSION<br />

Lecanu L. *§ , Ibrahim A. *§ , Yao W. *§ , McCourty A. *§ , Greeson J. ** and Papadopoulos<br />

V. *§<br />

* Georgetown University Medical Center, Department of Biochemistry, Molecular and<br />

Cellular Biology, 3900 Reservoir Rd NW, Washington, DC 20057, USA. § Samaritan<br />

Research Laboratories, Georgetown University Medical Center, Washington, DC 20057,<br />

USA. ** Samaritan Pharmaceuticals Inc., 101 Convention Drive, Las Vegas, NV 89109,<br />

USA.<br />

Laurent Lecanu, ll55@georgetown.edu, Fax (202) 687-2<strong>35</strong>4<br />

Repairing brain damage by replacing neuronal losses and restoring the associated<br />

functions is an ambitious challenge. In that aspect, the concept of stem cell therapy is<br />

extremely promising. The graft of stem cells differentiated into dopaminergic neurons has<br />

already been successfully applied to treat patients suffering from Parkinson’s disease.<br />

However, in disease or conditions during which the neuronal loss could be much more<br />

important, like Alzheimer’s disease (AD), brain stroke or traumatic brain injury, the<br />

transplantation of differentiated stem cells, although critical, might not be enough to<br />

compensate for the existing brain damage and to restore the hampered functions. Under<br />

these conditions one could hope that for maximal recovery, the transplantation might have<br />

to be associated to a stimulation of the in situ neurogenesis. We present herein the naturally<br />

occurring steroid (20α)-25ξ-methyl-(22R,26)-azacyclofurost-5-en-3ξ-ol (solasodine) was<br />

able to trigger in vitro the differentiation of neural stem cells into cholinergic neurons and<br />

to activate in vivo the neurogenesis from the stem cells present in the subventricular zone<br />

(SVZ) of rat brains.<br />

Solasodine was identified as potential candidate based on is silico screening for<br />

structural homologues of 22R-hydroxycholesterol, an intermediate of steroid biosynthesis<br />

that we found to induce neurogenesis in vitro but as an intermediate of steroid biosynthesis<br />

is rapidly metabolized. Mouse embryonic teratocarcinoma P19 cells were grown on glass<br />

cover-slip. When cells reached 70% confluence, the medium was replaced fresh medium<br />

containing 90 µM solasodine. P19 cells were then incubated for 2 days before solasodine<br />

was washed out and replaced by standard medium. The culture medium was changed every<br />

2 days for 5 days or every 2 days for 30 days before cells were fixed for<br />

immunocytochemistry. For in vivo studies, solasodine at 375 µM was infused for 2 weeks<br />

in the left ventricle of male Long-Evans rats (300-325 g) using an Alzet osmotic. Rats were<br />

sacrificed 3 weeks after the end of the infusion by intracardiac perfusion and brains fixed<br />

before being paraffin embedded for immunohistochemistry. To study neural stem cell<br />

proliferation, rats were injected daily with a BrdU solution at 100 mg/kg. The first<br />

injection took place the day following the surgery and the last injection was performed the<br />

day prior to euthanasia. To determine whether solasodine could be metabolized by<br />

steroidogenic enzymes thus affecting steroid formation mouse MA-10 tumor Leydig cells<br />

were treated with the compound and progesterone formation was monitored by<br />

radioimmunoassay. The affinity of solasodine for the various human steroid receptors was<br />

determined using High Throughput Screening technology in various steroid receptor<br />

specific cell and tissue models.<br />

Solasodine treatment induced P19 cell differentiation into neurons. Differentiated<br />

P19 cells displayed strong choline-acetyltransferase immunoreactivity showing a<br />

68


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

cholinergic phenotype. Significant sprouting processes were formed that expressed the<br />

specific neuronal markers βIII-tubulin and MAP2 suggesting axon formation. Solasodine<br />

also induced the expression of synaptophysin indicating that synaptogenesis was also<br />

triggered. The synapse formation and axonal growth kept evolving even 1 month after the<br />

treatment suggesting that the effect of solasodine was not transient. Neuronal<br />

differentiation was further confirmed by following the expression of doublecortin, a<br />

marker shown to be expressed specifically in newly formed neuroblasts. The continuous<br />

perfusion of solasodine inside rat left ventricle for 2 weeks induced a dramatic increase of<br />

the BrdU uptake by cells in SVZ. An increase of BrdU labeling was observed in the<br />

ependymal and sub-ependymal cells demonstrating that solasodine induced the<br />

proliferation of both cell populations. Although controversial, a theory already proposed is<br />

that the ependymal cells would serve as a progenitor cell reserve that would be activated<br />

only when massive neuronal cell replacement is required. Interestingly, solasodine greatly<br />

increased the expression of doublecortin that co-localized with BrdU immunostaining.<br />

Doublecortin is not expressed in proliferating cells but only in differentiating neural stem<br />

cells suggesting that solasodine increased the proliferating rate of the ependymal and subependymal<br />

cells before pushing them towards neuronal differentiation. These results<br />

support previous data showing that ependymal cells are capable of neurogenesis.<br />

Interestingly, solasodine did not bind to any of the classic steroid receptors and does not<br />

serve as a precursor for steroid synthesis, thus ruling out any contributing direct steroidlike<br />

pharmacological effect to the induced neurogenesis.<br />

In an attempt to identify pathways activated by solasodine, we observed that P19<br />

cells treated with solasodine over-expressed the neurite outgrowth-associated protein<br />

GAP43 and its associated post-transcriptional regulatory element HuD. Another interesting<br />

finding is the presence of a robust translocator protein (18 kDa; TSPO) immunostaining in<br />

the ependymal cells in which BrdU immunostaining was also dramatically increased,<br />

suggesting that solasodine increased the expression of TSPO in the proliferating<br />

ependymal cells. These data are consistent with previously reported findings that TSPO<br />

expression is increased during axonal regeneration and stem cell differentiation and the<br />

role of TSPO in mitochondrial membrane biogenesis, a requirement for rapidly<br />

proliferatings cells.<br />

Drug-induced activation of resident neural progenitor cell differentiation is the less<br />

controversial aspect of the stem cell therapy concept. The results presented herein suggest<br />

that the steroidal compound solasodine, naturally found in plants from the solanaceae<br />

family, induced neurogenesis both in vitro and in vivo. Although further studies are<br />

required to determine its mechanism of action, these data indicate that small molecules like<br />

solasodine could constitute an interesting approach to pharmacologically induce in situ<br />

neurogenesis as part of neuron replacement therapy.<br />

69


MONDAY, 19 th February<br />

12.00 - 13.00<br />

Plenary Lecture:<br />

Mellon S.H. (USA)


NEUROSTEROIDS IN HEALTH AND DISEASE<br />

Mellon S.H., Gong W., Schonemann M.<br />

4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

Department of Obstetrics, Gynecology & Reproductive Sciences, The Center for<br />

Reproductive Sciences, 513 Parnassus Avenue, University of California, San Francisco,<br />

San Francisco, CA 94143-0556 USA. FAX 1-415-502-7866,<br />

email: mellon@cgl.ucsf.edu<br />

The functions for neurosteroids during development and in response to nervous<br />

system injury are beginning to be identified. We took several approaches to studying the<br />

function of some neurosteroids in vivo. Transgenic mice were created by us and by several<br />

other laboratories, in which the genes encoding steroidogenic enzymes were universally<br />

ablated. Unfortunately, none of these transgenic mice provided concrete evidence for<br />

functions of neurosteroids during embryonic development or during development of the<br />

nervous system. Hence, we chose to focus on another mouse model in which we believed<br />

neurosteroid production would be altered, and which had a neurodegenerative phenotype.<br />

Niemann Pick Type-C (NP-C) is an autosomal recessive neurodegenerative disease caused<br />

by mutations in NPC1 (95%) or NPC2 (5%), resulting in lysosomal accumulation of<br />

unesterified cholesterol and glycolipids. The NIH mouse model of NP-C has a mutation in<br />

the NPC1 gene, and exhibits several pathological features of the most severe NP-C<br />

patients. How lysosomal storage and trafficking defects lead to neurodegeneration is<br />

unknown. We found that these mice had normal neurosteroidogenic enzyme activity<br />

during development, but lost this activity in the early neonatal period, prior to onset of<br />

neurological symptoms. The most severely affected enzyme was 3α hydroxysteroid<br />

dehydrogenase, and decreased 5α reductase activity was also seen throughout the neonatal<br />

brain. Neurons that expressed P450scc, 3ß HSD, as well as those that expressed 3α HSD<br />

and 5α reductase were lost in adult NP-C brains, resulting in diminished concentrations of<br />

allopregnanolone. We treated NP-C mice with allopregnanolone using different regimens,<br />

and found that a single dose in the neonatal period resulted in a doubling of lifespan,<br />

substantial delay in onset of neurological symptoms, survival of cerebellar Purkinje and<br />

granule cell neurons, and reduction in cholesterol and ganglioside accumulation. The<br />

mechanism by which allopregnanolone elicited these effects is unknown. Our in vitro<br />

studies showed that Purkinje cell survival promoted by allopregnanolone was lost by<br />

treatment with bicuculline, suggesting GABA A receptors may play a role. Studies in<br />

collaboration with Dan Ory at Washington University using a GABA A -inactive<br />

entantiomer of allopregnanolone demonstrated survival and neurological benefits in NP-C<br />

mice, similar to allopregnanolone, suggesting a lack of GABA A -receptor involvement.<br />

Pregnane-X-receptors may mediate the effects of allopregnanolone and its enantiomer, as<br />

NP-C mice treated with these compounds increased expression of PXR-regulated genes in<br />

their cerebella. We showed that an additional feature of neuropathology of NP-C mice,<br />

region- and time-specific onset of neuroinflammation and its pathologic sequelae, is also<br />

ameliorated by allopregnanolone treatment, suggesting further functions for this<br />

neurosteroid. Thus, mouse models of neurodegeneration may be beneficial in establishing<br />

both physiologic and pharmacologic actions of neurosteroids. These animal models further<br />

establish the wide range of functions of these compounds, which may ultimately be useful<br />

for treatment of human diseases.<br />

73


MONDAY, 19 th February 2007<br />

15.00 - 18.30<br />

Symposium:<br />

Neuroprotective effects


Symposium:<br />

Neuroprotective effects<br />

(Chairs: Garcia-Segura L.M., Mellon S.H.)<br />

• Simpkins JW, Dykens JA (USA) Mitochondrial mechanisms of estrogen<br />

neuroprotection<br />

• Rosario ER, Carroll JC, Oddo S, LaFerla FM, Pike CJ (USA) Androgen<br />

regulation of neuropathology in a triple transgenic mouse model of AD<br />

• Stein DG (USA) Neurosteroids as protective factors in TBI: from laboratory<br />

bench to the bedside<br />

• Mensah-Nyagan AG, Meyer L., Kibaly C, Schaeffer V and Patte-Mensah C.<br />

(France) Neurosteroids and nociceptive sensitivity in neuropathic rats<br />

• Melcangi RC (Italy) Neuroactive steroids and peripheral neuropathy<br />

• Ritz M.-F., Hausmann O. (Switzerland) Neuroprotective effects of estradiol in a<br />

rat model of spinal cord injury.<br />

• Kondo S., Imaizumi K (Japan) BBF2H7, a novel transmembrane bZIP<br />

transcription factor, is a new type of ER stress transducer


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

MITOCHONDRIAL MECHANISMS OF ESTROGEN NEUROPROTECTION<br />

Simpkins J.W. and Dykens J.A.<br />

Department of Pharmacology & Neuroscience<br />

Institute for Aging and Alzheimer’s Disease Research<br />

University of North Texas Health Science Center<br />

Fort Worth, TX 76102<br />

Oxidative stress, bioenergetic failure and mitochondrial dysfunctions are all<br />

implicated in the etiology of neurodegenerative diseases such as Alzheimer’s disease (AD).<br />

The mitochondrial involvement in neurodegenerative diseases reflects the regulatory<br />

position mitochondrial failure plays in both necrotic cell death and apoptosis. The potent<br />

feminizing hormone, 17 β-estradiol (E2), is neuroprotective in a host of cell and animal<br />

models of stroke and neurodegenerative diseases. The discovery that 17α-estradiol, an<br />

isomer of E2, is equally as neuroprotective as E2 yet is >200-fold less active as a hormone,<br />

has permitted development of novel, more potent analogs where neuroprotection is<br />

independent of hormonal potency. Studies of structure-activity-relationships and<br />

mitochondrial function have led to a mechanistic model in which these steroidal phenols<br />

intercalate into cell membranes where they block lipid peroxidation reactions, and are in<br />

turn recycled. Indeed, the parental estrogens and novel analogs stabilize mitochondria<br />

under Ca 2+ loading otherwise sufficient to collapse membrane potential. The<br />

neuroprotective and mitoprotective potencies for a series of estrogen analogs are<br />

significantly correlated, suggesting that these compounds prevent cell death in large<br />

measure by maintaining functionally intact mitochondria. This therapeutic strategy is<br />

germane not only to sudden mitochondrial failure in acute circumstances, such as during a<br />

stroke or myocardial infarction, but also to gradual mitochondrial dysfunction associated<br />

with chronic degenerative disorders such as AD. (Supported by NIH grants AG10485 and<br />

AG22550)<br />

77


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

ANDROGEN REGULATION OF NEUROPATHOLOGY IN A TRIPLE<br />

TRANSGENIC MOUSE MODEL OF AD<br />

a Rosario E.R., a Carroll J.C., c Oddo S., c LaFerla F.M., b Pike C.J.<br />

a Neuroscience Graduate Program, b Davis School of Gerontology, University of Southern<br />

California, c Department of Neurobiology and Behavior, University of California, Irvine<br />

Normal, age-related, testosterone depletion in men is a recently identified risk<br />

factor for the development of AD. To investigate the relationship between testosterone<br />

depletion and the development of AD, we compared indices of pathology under different<br />

androgen conditions using the 3xTg-AD triple-transgenic mouse model. Male 3xTg-AD<br />

mice were gonadectomized (GDX) to deplete endogenous levels of androgens at 3 months<br />

of age, and then exposed to subcutaneous, slow-release drug delivery pellets containing<br />

either 10 mg dihydrotestosterone (DHT) or placebo. After 4 months of androgen depletion<br />

(7 months of age), mice were evaluated for behavioral deficits and severity of AD-like<br />

neuropathology. In comparison to gonadally intact 3xTg-AD mice, GDX mice exhibited<br />

robust increases in accumulation of β-amyloid (Aβ), the protein implicated as the primary<br />

causal factor in AD pathogenesis, in both hippocampus and amygdala. In parallel to<br />

elevated levels of Aβ, GDX mice exhibited significantly impaired performance in a<br />

spontaneous alternation Y-maze task, indicating deficits in hippocampal function.<br />

Importantly, DHT treatment of GDX 3xTg-AD mice prevented both neural accumulation<br />

of Aβ and hippocampal performance deficits. These data provide the first definitive<br />

evidence in an animal model linking androgen depletion to the development of AD<br />

pathology. These findings suggest that androgen-based hormone therapy may be a useful<br />

strategy for the prevention and treatment of AD in aging men.<br />

78


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

NEUROSTEROIDS AS PROTECTIVE FACTORS IN TBI: FROM LABORATORY<br />

BENCH TO THE BEDSIDE<br />

Stein D.G.<br />

Emory University School of Medicine, Department of Emergency Medicine, Atlanta,<br />

Georgia 30322<br />

e-mail: donald.stein@emory.edu<br />

phone: 404-712-9704<br />

Background: At present, there are no clinically effective treatments for traumatic brain<br />

injury (TBI). Many of the clinical trials seeking agents for the treatment of stroke and TBI<br />

have ended in failure (e.g., the CRASH trial with methylprednisolone and more recently<br />

the magnesium sulfate trial, which resulted in a 25% increase in mortality compared to<br />

patients given placebo). There is speculation that treatment with serum albumin may be<br />

efficacious, but the studies are still in progress and nothing definitive can be said yet.<br />

Hypothermia has been given considerable attention in the treatment of TBI, but the results<br />

of several clinical trials have yet to prove any substantial efficacy over controls. So, what<br />

is left? A growing contingent of investigators are now proposing that, after TBI and stroke,<br />

progesterone plays a much more important role in CNS repair, organization and function<br />

than has previously been realized.<br />

Why progesterone? Mainly because after more than 15 years of pre-clinical investigation,<br />

this hormone and its metabolite, allopregnanolone, appear to meet almost all the criteria for<br />

a safe and effective clinical treatment for TBI. The trajectory of research from the<br />

laboratory bench to the patient’s bedside will be the subject of my presentation. For well<br />

over a decade, our laboratory and others have been examining the role of progesterone and<br />

its precursors and metabolites to determine their specific molecular, physiological and<br />

functional mechanisms of action in repair and protection of the damaged central nervous<br />

system. There is now substantial experimental evidence that progesterone plays a vital role<br />

in reducing inflammatory disorders of the brain and that it enhances morphological and<br />

functional repair in the victims of TBI and stroke.<br />

A principal effect of progesterone treatment, given after injury, is to reduce cerebral edema<br />

and the inflammation that accompanies TBI. Cerebral edema also occurs after stroke and<br />

can be very problematic for patients undergoing open-heart surgery. Thus, an agent<br />

capable of resolving inflammation, lipid peroxidation, edema, necrosis and programmed<br />

cell death that also has a good safety pro<strong>file</strong> with few side effects, could be of major<br />

benefit in a clinical setting. After presenting the background experimental data, the results<br />

of a recently completed, National Institutes of Health (NINDS) sponsored, Phase II (a),<br />

single-center trial for safety and efficacy of progesterone will be discussed.<br />

Summary points:<br />

• In laboratory animals, females with traumatic brain injury have better functional<br />

and morphological outcomes than males with the same extent of injury.<br />

79


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

• Treatment with exogenous progesterone and its’ metabolite, allopregnanolone in<br />

both adult males and females, enhances the rate and extent of recovery from<br />

traumatic brain injury and stroke.<br />

• The reparative molecular and receptor mechanisms of action of these neurosteroids<br />

on nerve cells, glia and vascular tissue are becoming better understood and may be<br />

directly applicable to clinical trial and further study.<br />

• Progesterone and its metabolites act by reducing immune-inflammatory reactions,<br />

membrane lipid peroxidation, cerebral edema, apoptosis and necrosis; thus<br />

preventing the slow but steady death of nerve cells long after the initial injury<br />

itself.<br />

• Progesterone and allopregnanolone stimulate soft-tissue wound repair, improve<br />

vascular responses to brain injury and enhance the remyelination of damaged<br />

axons. The neurosteroids regulate gene expression and protein synthesis involved<br />

in glial activity, apoptosis and regenerative repair.<br />

• New research is showing that progesterone and its related metabolites may also<br />

prove effective in the treatment of neurodegenerative disorders such as transient<br />

and ischemic stroke, multiple sclerosis and spinal cord injuries.<br />

Objectives of the Presentation:<br />

1. Provide a history and a better understanding of the role of neurosteroids in the early<br />

treatment of traumatic brain injury and stroke.<br />

2. Consider the data showing that sex differences in CNS functions led the way to<br />

research showing that so-called “sex hormones” can play a critical role in<br />

enhancing brain repair.<br />

3. Recognize that early intervention with neurosteroid treatments during the most<br />

acute phase of the TBI/stroke injury cascade, may enhance the efficacy of<br />

subsequent<br />

80


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

NEUROSTEROIDS AND NOCICEPTIVE SENSITIVITY IN NEUROPATHIC<br />

RATS<br />

Mensah-Nyagan A.G.*, Meyer L., Kibaly C., Schaeffer V. and Patte-Mensah C.<br />

Institut des Neurosciences Cellulaires et Intégratives, UMR7168/LC2 CNRS-ULP, Equipe<br />

« Stéroïdes et Système Nociceptif », 67084 Strasbourg Cedex, France.<br />

* E-mail : gmensah@neurochem.u-strasbg.fr<br />

Chronic neuropathic pain constitutes a major public health concern in many<br />

countries all around the world. This category of pain, which is refractory to the currently<br />

available analgesics including opioids, provokes persistent suffering in several thousands<br />

of patients and socio-economical problems such as substantial costs due to disability,<br />

decreased productivity and medical expenses. Therefore, development of novel therapeutic<br />

strategies against neuropathic pain has become a real challenge for biomedical research.<br />

Neuropathic pain is generated by injuries of the nervous system or by disturbances in the<br />

activity of spinal and supraspinal neural networks controlling nociception. Thus, it appears<br />

that the compounds to be characterized for the treatment of stubborn neuropathic pain must<br />

necessarily be capable of modulating the activity of spinal and supraspinal neuronal<br />

pathways. Various family of molecules are currently explored among which are<br />

neurosteroids that strongly modulates GABA A , NMDA and P2X receptors intervening in<br />

nociception and pain control.<br />

To obtain exploitable results on the role of neurosteroids in neuropathic pain modulation,<br />

we developed a multidisciplinary project allowing integration of data from molecular and<br />

cellular levels to behavioral components using a model of chronic pain generated in rat by<br />

sciatic nerve ligatures. Prior to the study with neuropathic animals, we investigated the<br />

occurrence of neurosteroid biosynthesis in the adult rat spinal cord (SC), a crucial structure<br />

involved in painful message transmission. We observed that the SC dorsal horn (DH), a<br />

pivotal nociceptive center, contains key steroidogenic enzymes such as cytochrome<br />

P450side-chain-cleavage, cytochrome P450c17 (P450c17), 5a-reductase and 3ahydroxysteroid<br />

oxido-reductase (3a-HSOR). Afterwards, molecular analyses using the<br />

real-time polymerase chain reaction after reverse transcription were used to determine<br />

changes occurring in the expression of genes encoding steroid-synthesizing enzymes in the<br />

DH during the chronic neuropathic pain situation. Reversed-phase HPLC analysis was<br />

coupled with flow scintillation detection to compare enzymatic activities leading to<br />

neurosteroid production in SC slices of control and neuropathic-pain rats.<br />

Radioimmunoassays allowed assessments of changes of endogenous neurosteroid levels in<br />

the SC during neuropathic pain state.<br />

The results revealed an up-regulation of enzymatic pathways (P450scc, 5a-reducatse and<br />

3aHSOR) leading to the biosynthesis of allopregnanolone or 3a,5a-THP (a potent allosteric<br />

activator of GABA A receptors) in the DH of neuropathic rats. In contrast, the biosynthetic<br />

pathway responsible for DHEA synthesis (P450c17) was down-regulated in the DH during<br />

the chronic painful state.<br />

Behavioral analyses were performed using the Hargreaves’ method and Von Frey filament<br />

test to assess respectively the thermal and mechanical nociceptive thresholds in naïve,<br />

sham-operated and neuropathic-pain rats. Two main symptoms characterizing chronic<br />

neuropathic pain in humans, namely a thermal hyperalgesia and a mechanical allodynia,<br />

were detected in rats submitted to sciatic nerve-evoked peripheral neuropathy. Intrathecal<br />

injection of 3a,5a-THP in the lumbar SC, which increased the thermal and mechanical<br />

sensitivity thresholds in naïve and sham-operated rats, provoked analgesia in neuropathic-<br />

81


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

pain animals by suppressing the thermal hyperalgesia and mechanical allodynia. Unlike<br />

3a,5a-THP, Provera, a pharmacological inhibitor of 3a-HSOR, decreased both thermal and<br />

mechanical nociceptive thresholds in naïve and sham-operated animals. In neuropathic<br />

rats, intrathecal administration of Provera potentiated both thermal hyperalgesia and<br />

mechanical allodynia.<br />

Subcutaneous administration of DHEA induced a biphasic action on pain sensitivity: a<br />

short term pro-nociceptive effect followed by a late analgesic action probably due to<br />

DHEA conversion into androgenic metabolites. The proper action of DHEA seems<br />

effectively to be pro-nociceptive because in vivo blockade of DHEA biosynthesis in the<br />

SC by intrathecal injection of ketoconazole, a pharmacological inhibitor of P450c17,<br />

induced analgesia in neuropathic rats. Unlike ketoconazole, intrathecal administration of<br />

DHEA rapidly potentiated both thermal hyperalgesia and mechanical allodynia<br />

characterizing the neuropathic pain.<br />

Since neurosteroids are endogenous compounds capable of interacting with various<br />

neurotransmitters involved in pain control, we hope that our results may open new<br />

perspectives for the development of efficient therapy against neuropathic pain based on the<br />

selective modulation of neurosteroidogenic pathways.<br />

82


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

NEUROACTIVE STEROIDS AND PERIPHERAL NEUROPATHY<br />

Melcangi R.C.<br />

Dept. of Endocrinology and Center of Excellence on Neurodegenerative Diseases,<br />

University of Milan, via Balzaretti 9, 20133, Milano Italy.<br />

roberto.melcangi@unimi.it<br />

It is now clearly demonstrated that peripheral nerves are target for the action of<br />

neuroactive steroids. Indeed, both classical steroid receptors, (e.g., progesterone receptor,<br />

PR, androgen receptor, AR, etc.), and non-classical steroid receptors (e.g., GABA-A and<br />

GABA-B receptors, sigma 1 receptor, NMDA receptor 1 subunit, etc.) have been<br />

demonstrated [see for review, 1]. On this basis, we have assessed whether treatment with<br />

neuroactive steroids had protective effects in experimental models of peripheral<br />

neuropathy, such as nerve injury (transection or crush), aging and diabetic neuropathy.<br />

Data we have so far obtained are very promising because they have indicated that<br />

neuroactive steroids, such as progesterone, testosterone and their metabolites (i.e.,<br />

dihydroprogesterone, tetrahydroprogesterone, dihydrotestosterone and 5alpha -androstan-<br />

3alpha, 17beta-diol) exert important protective effects on several components of peripheral<br />

nerve. As will be reported, depending by experimental models considered, morphological<br />

parameters of the nerve, intraepidermal nerve fiber density, expression of myelin proteins,<br />

Na + ,K + -ATPase activity, thermal nociceptive threshold, nerve conduction velocity, which<br />

are affected by neurodegenerative events are a target of protective effects of neuroactive<br />

steroids [2-5]. These observations suggest that neuroactive steroids themselves or synthetic<br />

ligands of their receptors might represent an interesting therapeutic perspective for<br />

acquired peripheral neuropathies. Interestingly, an alternative to this therapeutic strategy<br />

might be represented by ligands of peripheral-type benzodiazepine receptor (PBR). This is<br />

a protein predominantly located in the mitochondrial outer membrane that plays an<br />

important role in the steroidogenesis and neurosteroidogenesis, and in the regulation of cell<br />

survival and proliferation. Indeed, recent observations have indicated that also PBR ligands<br />

may be considered as protective agents for peripheral neuropathy [6]<br />

(PRIN-2005060584_004 and FIRST from University of Milan).<br />

References list<br />

[1] Melcangi R.C., Cavarretta I.T., Ballabio M., Leonelli E., Schenone A., Azcoitia I., Garcia-Segura L.M.,<br />

Magnaghi V. Peripheral nerves: a target for the action of neuroactive steroids. Brain Res. Rev. 48:328-<br />

338, 2005.<br />

[2] Melcangi R.C., Magnaghi V., Galbiati M., Ghelarducci B., Sebastiani L., Martini L. The action of steroid<br />

hormones on peripheral myelin proteins: a possible new tool for the rebuilding of myelin? J.<br />

Neurocytol., 29:327-339, 2000.<br />

[3] Azcoitia I., Leonelli E., Magnaghi V., Veiga S., Garcia-Segura L.M., Melcangi R.C. Progesterone and its<br />

derivatives dihydroprogesterone and tetrahydroprogesterone reduce myelin fiber morphological<br />

abnormalities and myelin fiber loss in the sciatic nerve of aged rats. Neurobiol Aging 24:853-860, 2003.<br />

[4] Veiga S., Leonelli E., Beelke M., Garcia-Segura L.M., Melcangi R.C. Neuroactive steroids prevent<br />

peripheral myelin alterations induced by diabetes. Neurosci. Lett. 402:150-153, 2006.<br />

[5] Leonelli E., Bianchi R., Cavaletti G., Caruso D., Crippa D., Garcia-Segura L.M., Lauria G., Magnaghi<br />

V., Roglio I., Melcangi R.C. Progesterone and its derivatives are neuroprotective agents in experimental<br />

diabetic neuropathy: a multimodal analysis. Neuroscience, in press, 2006,<br />

doi:10.1016/j.neuroscience.2006.11.014.<br />

[6] Leonelli E., Yague J.G., Ballabio M., Azcoitia I., Schumacher M., Garcia-Segura L.M., Melcangi R.C.<br />

Ro5-4864, a synthetic ligand of peripheral benzodiazepine receptor, reduces aging-associated myelin<br />

degeneration in the sciatic nerve of male rats. Mech. Ageing Dev. 126:1159-1163, 2005.<br />

83


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

NEUROPROTECTIVE EFFECTS OF ESTRADIOL IN A RAT MODEL OF<br />

SPINAL CORD INJURY<br />

Ritz M.-F. 1 , Hausmann O. 1,2<br />

1<br />

University Hospital, Department of Research, Neurosurgery Laboratory,<br />

Klingelbergstrasse 50, 4056 Basel, Switzerland. e-mail: marie-francoise.ritz@unibas.ch<br />

Fax:+41 61 267 1628,<br />

2<br />

Hirslanden Clinic St. Anna, Neurosurgery, Lucerne, Switzerland.<br />

Spinal cord injury (SCI) occurs mostly in young people as a result of traffic or sportsrelated<br />

accidents and leads to severe neurological deficits such as paraplegia and<br />

quadriplegia. After the initial mechanical deformation of the spinal cord, a cascade of<br />

biochemical and cellular processes initiate further cellular damage and cell death, known<br />

as the secondary injury. The secondary injury mechanisms include vascular changes<br />

(including ischemia, vasospasms, haemorrhages and thrombosis), ionic disturbances,<br />

neurotransmitter (glutamate) accumulation [1], generation of free radicals (NO) [2],<br />

edema, depletion of energy substrates, and activation of a variety of proteases including<br />

caspases, phospholipases, endonucleases and metalloproteinases [3]. This active and<br />

progressive spread of damage results from a process that begins within minutes and<br />

continues for weeks after the initial injury. Unfortunately, this secondary segmental<br />

neuronal loss is responsible for an often deleterious secondary functional worsening.<br />

Inflammation is one key-player that may exacerbate the spreading of the initial lesion.<br />

After experimental SCI, transcripts of pro-inflammatory cytokines such as interleukin 1<br />

(IL-1 etaand IL-1 lpha and tumor necrosis factor lphaare upregulated within the first<br />

few hours [3-5] in the injured environment. Inhibiting these processes is thought to<br />

potentially lead to a better functional outcome after SCI, however the beneficial effects of<br />

these cytokines are now also considered for improving some protective actions during<br />

secondary injury in SCI.<br />

17beta-Estradiol (E2) has been shown to possess neuroprotective activities and to modulate<br />

brain neurotransmitter transmission. Studies using in vivo and in vitro models of<br />

neurodegenerative disorders such as Alzheimer’s and Parkinson’s diseases suggest that<br />

estrogens provide neuroprotection of the central nervous system (CNS). Protective effects<br />

of E2 in stroke has also been extensively studied and include anti-apoptotic, anti-excitatory<br />

and anti-inflammatory actions.<br />

Therefore, the use of E2 to reduce secondary injury after SCI was performed in this study.<br />

We evaluated the effects of an immediate treatment with a physiological and a supraphysiological<br />

dose of E2 on the rat locomotor function recovery and on the lesion size over<br />

time after a mild spinal cord compression injury. In addition, the influence of this treatment<br />

on the inflammatory response, by measuring the release of various pro- and antiinflammatory<br />

cytokines was assessed. The activation of astrocytes and the size of the<br />

lesions were also followed at the compression site over time.<br />

A low physiological dose (0.1 mg/kg) and a supra-physiological dose (4 mg/kg) of E2<br />

were injected i.p. into male rats immediately after spinal cord compression at the T8-T9<br />

level. Functional outcome was assessed using the BBB score and the narrow beam walk<br />

test from day 3 to 4 weeks post-injury. The expression of both types of IL-1 and IL-6 as<br />

well as the anti-inflammatory cytokines IL-4 and IL-10 in the injured spinal cord and<br />

adjacent rostral and caudal segments were evaluated at 6 hr, 3 days and 1 week post-injury.<br />

The expressions of the glial fibrillary acidic protein (GFAP) and vimentin in astrocytes<br />

84


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

were visualized by immunohistology on spinal cord horizontal sections containing the<br />

compressed spine segment.<br />

After SCI, high-dose E2-treated rats improved more rapidly and had a better final score<br />

than the low-dose E2-treated group and the controls in the BBB test. These rats were also<br />

able to walk on the narrow beam at earlier time points compared to controls or low-dose<br />

treated rats and on a longer distance. However, both E2-treated groups showed the best<br />

performances in the narrow beam test at 4 weeks. In the lesion sites, both doses of E2<br />

enhanced the trauma-induced expression of IL-1alpha, IL-1beta, IL-6 and IL-1ra at 6 hr<br />

post-injury. The releases of anti-inflammatory cytokines were also slightly up-regulated by<br />

both doses of E2 but only after 3 d in the caudal segments.<br />

Astrogliosis, characterized by elevated expressions of GFAP and vimentin in the astrocytes<br />

was followed during the 4 weeks post-SCI period. The increased expression of both<br />

proteins was seen as early as 3 days after SCI in rats treated with 4 mg/kg E2, but only<br />

after 2 weeks in controls and in the low-dose E2-treated rats. Moreover, the expression of<br />

both proteins was higher in the high-dose E2-treated rats until 4 weeks. The sizes of the<br />

spinal lesions were quantified after 2 and 4 weeks post-injury. Rats treated with 4 mg/kg<br />

E2 showed significantly smaller lesions after 2 weeks as the low-dose treated or control<br />

rats. However, the lesion sizes were similar in all groups after 4 weeks, a time point where<br />

acute phase is already finished.<br />

Altogether, these results suggest that the acute treatment with a supra-physiological dose of<br />

E2 at the onset of spinal cord injury activates the early expression of pro-inflammatory<br />

cytokines involved in astroglial activation. In parallel, the lesion sizes were reduced at 2<br />

weeks and an improvement of the locomotor activity was observed with this treatment. A<br />

low physiological showed some effects on the induction of inflammatory cytokines, but the<br />

effects on the lesion size and functional improvement were small. The use of E2 as a<br />

therapy during the acute phase in spinal cord injured patients may be an option to<br />

significantly reduce secondary damage.<br />

Reference List<br />

[1] Liu, D, Thangnipon, W, McAdoo, DJ: Excitatory amino acids rise to toxic levels upon impact<br />

injury to the rat spinal cord. Brain Res 1991;547:344-8.<br />

[2] Diaz-Ruiz, A, Ibarra, A, Perez-Severiano, F, Guizar-Sahagun, G, Grijalva, I, Rios, C:<br />

Constitutive and inducible nitric oxide synthase activities after spinal cord contusion in rats.<br />

Neurosci Lett 2002;319:129-32.<br />

[3] Wang, CX, Olschowka, JA, Wrathall, JR: Increase of interleukin-1beta mRNA and protein in<br />

the spinal cord following experimental traumatic injury in the rat. Brain Res 1997;759:190-6.<br />

[4] Bartholdi, D, Schwab, ME: Expression of pro-inflammatory cytokine and chemokine mRNA<br />

upon experimental spinal cord injury in mouse: an in situ hybridization study. Eur J Neurosci<br />

1997;9:1422-38.<br />

[5] Hayashi, M, Ueyama, T, Nemoto, K, Tamaki, T, Senba, E: Sequential mRNA expression for<br />

immediate early genes, cytokines, and neurotrophins in spinal cord injury. J Neurotrauma<br />

2000;17:203-18.<br />

85


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

BBF2H7, A NOVEL TRANSMEMBRANE BZIP TRANSCRIPTION FACTOR, IS A<br />

NEW TYPE OF ER STRESS TRANSDUCER<br />

Kondo S., and Imaizumi K.<br />

Division of Molecular and Cellular Biology, Department of Anatomy, Faculty of<br />

Medicine, University of Miyazaki, Japan. e-mail: sh-kondo@med.miyazaki-u.ac.jp<br />

Endoplasmic reticulum (ER) stress transducers IRE1, PERK and ATF6 are well<br />

known to transduce signals from the ER to the cytoplasm and nucleus when unfolded<br />

proteins accumulate in the ER. Recently, we identified OASIS as a novel ER stress<br />

transducer expressed in astrocytes.<br />

We report here that BBF2H7, originally identified as a novel human protein whose C-<br />

terminal part was fused to the FUS genes in low grade fibromyxoid sarcoma, is structurally<br />

homologous to OASIS. BBF2H7, an ER-resident transmembrane protein with the bZIP<br />

domain in the cytoplasmic portion, is cleaved at the membrane in response to ER stress.<br />

The cleaved fragments of the BBF2H7 translocate into the nucleus and can bind directly to<br />

CRE site to activate transcription of target genes. Interestingly, although BBF2H7 protein<br />

is not expressed in normal conditions, it is markedly induced at the translational level<br />

during ER stress, suggesting thatBBF2H7 might contribute to only the late phase of UPR<br />

signaling. In a mouse model of focal brain ischemia, BBF2H7 protein is prominently<br />

induced in neurons in the peri-infarction region. Taken together, our results suggest that<br />

BBF2H7 is a novel ER stress transducer and could play important roles in preventing<br />

accumulation of unfolded proteins in damaged neurons.<br />

86


TUESDAY, 20 th February 2007<br />

08.30 - 11.30<br />

Symposium:<br />

Xenoestrogens and brain circuitries


Symposium:<br />

Xenoestrogens and brain circuitries<br />

(Chair: Celotti F., Panzica G.C.)<br />

• Ottinger MA, Lavoie E, Thompson N, Whitehouse K, Barton M, Abdelnabi M,<br />

Quinn M, Jr. (USA) Neuroendocrine and behavioral consequences of embryonic<br />

exposure to endocrine disrupting chemicals<br />

• Patisaul HB, Polston EK (USA) Influence of endocrine active compounds in the<br />

developing brain<br />

• Maggi A (Italy) The ERE-Luc reporter mouse: twenty years of basic research to<br />

be applied to the study of endocine disrupters<br />

• Kawato S, Ogiue-Ikeda M., Tanabe N., Tsurugizawa T., Hojo Y., Mukai H.<br />

(Japan) Rapid modulation of long-term depression and spinogenesis by endocrine<br />

disrupters in adult rat hippocampus<br />

• Corrieri L., Della Seta D., Paola Materazzi, Dessì-Fulgheri F., Farabollini F.,<br />

Fusani L. (Italy) Environmental-like exposure to xenoestrogen affects sexual<br />

differentiation of brain and behavior in female rats<br />

• Ponzi D, Palanza P , Maruniak J, Parmigiani S , Vom Saal F (USA) Sexual<br />

dimorphism in the number of TH-immunostained neurons in the Locus Coeruleus of<br />

young mice is eliminated by prenatal exposure to Bisphenol A


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

NEUROENDOCRINE AND BEHAVIORAL CONSEQUENCES OF EMBRYONIC<br />

EXPOSURE TO ENDOCRINE DISRUPTING CHEMICALS<br />

Ottinger M.A., Lavoie E., Thompson N., Whitehouse K., Barton M., Abdelnabi M.,<br />

and Quinn M. Jr 1 .<br />

Department of Animal and Avian Sciences, University of Maryland College Park,<br />

Maryland 20742 USA and 1 Department of the Army, Aberdeen, MD.<br />

A number of endocrine disrupting chemicals (EDCs) have been characterized and<br />

generally appear to act via steroid-like mechanisms. However, it is not clear if avian<br />

species respond in a parallel manner to effects described in mammals. Moreover, there are<br />

also effects due to ancillary toxic actions which often vary with species. We have<br />

conducted comparative study of a variety of classes of EDCs that have potentially different<br />

mechanisms of action in an embryo bioassay, using the precocial Japanese quail model.<br />

The EDCs examined included estradiol, androgen active compounds, and atrazine and we<br />

studied effects on hypothalamic neuroendocrine systems and behavior. Fertile quail eggs<br />

(n=85-95/group) were injected with 17β estradiol, trenbolone, or DDE into the yolk at<br />

embryonic day 4 and atrazine at embryonic day 0. Quail were sampled at hatch or raised<br />

to maturity to determine consequences of early exposure on later reproductive maturation<br />

and function. Birds were behaviorally tested at 1 week of age on an adapted runway and<br />

adult males were tested at maturity to assess male sexual behavior. Hypothalamic<br />

aromatase (AROM), catecholamines, and GnRH-I were measured. Behavior proved to be<br />

a sensitive index of exposure for all EDCs. In young chicks, trenbolone exposure impaired<br />

vocalization in week old chicks. Male sexual behavior was impaired by estradiol,<br />

androgenic EDCs, and atrazine treatment, especially mount latency. GnRH-I was sexually<br />

dimorphic in adult controls; atrazine affected hypothalamic GnRH-I in hatchlings and<br />

adults; other EDCs had variable effects. Atrazine decreased dopamine in hatchlings and<br />

adults. Hypothalamic neurotransmitters that modulate reproductive function may provide<br />

valuable indices of endocrine disruption associated with later consequences of embryonic<br />

exposure to EDCs. In addition, plasma steroid hormones were affected by some of the<br />

EDCs; atrazine impacted thyroid gland hormone content. Finally, we have evidence for<br />

immune system impacts from embryonic exposure to the EDCs. In summary, the Japanese<br />

quail embryo provides an excellent avian model for determining effects of a variety of<br />

EDCs in development and for ascertaining long term impacts on adults. Attention should<br />

be paid to consequences of embryonic exposure to EDCs on adult health, reproductive<br />

function, and immune function as these types of effects are likely to impair fitness of field<br />

birds.<br />

Research supported by EPA R826134010 (Star Grant), NSF 9817024, and EPA R-<br />

2877801(MAO).<br />

89


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

INFLUENCE OF ENDOCRINE ACTIVE COMPOUNDS IN THE DEVELOPING<br />

BRAIN<br />

Patisaul H.B. 1 , and Polston E.K. 2<br />

1 North Carolina State University, Department of Zoology, 127 David Clark Labs, Raleigh, NC<br />

27695 USA, fax: (919) 515-5327, heather_patisaul@ncsu.edu<br />

2 Howard University College of Medicine, Department of Physiology, 520 W St. NW, Washington,<br />

DC, 20059 USA<br />

Changes in the volumes of sexually dimorphic brain nuclei are often used as a biomarker<br />

for developmental disruption by endocrine-active compounds (EACs). However, these gross,<br />

morphological analyses do not reliably predict disruption of cell phenotype or neuronal function. In<br />

the present experiments, we used a more comprehensive approach to assess whether postnatal<br />

exposure to the EACs genistein (GEN) or bisphenol-A (BPA) affected the development of two<br />

sexually dimorphic brain regions in male and female rats: the anteroventral periventricular nucleus<br />

of the hypothalamus (AVPV) and the sexually dimorphic nucleus of the preoptic area (SDN). The<br />

AVPV and SDN are both structurally and functionally differentiated in rodents. In females, the<br />

AVPV relays hormonal and environmental signals to the gonadotropin-releasing hormone (GnRH)<br />

neurons that regulate ovulation. The female AVPV is larger than that of the male and contains<br />

higher numbers of tyrosine hydroxylase (TH) expressing neurons. The SDN is 2 – 4 times larger in<br />

males than females and is thought to play a role in the display of male sex behaviors. Recently, a<br />

sexually dimorphic subpopulation of neurons within the SDN that expresses calbindin-d28k<br />

(CALB) was described. The volume of this region, the CALB-SDN, is also larger in males. We<br />

hypothesized that neonatal exposure to BPA or GEN in the first few days of life, when both the<br />

AVPV and the SDN are undergoing steroid hormone directed sexual differentiation, would disrupt<br />

the anatomical structure, cellular phenotype, and function of these nuclei. Male and female rat pups<br />

were given 4 subcutaneous injections of sesame oil (control), 50 µg 17β-estradiol (E2), 250 µg<br />

GEN, or 250 µg BPA at twelve hour intervals over postnatal days (PND) 1 and 2. Half of the<br />

animals were sacrificed on PND 19 and the other half were allowed to grow to adulthood. In the<br />

PDN 19 animals, TH expression was unaltered by either BPA or GEN in the female AVPV, but<br />

both compounds demasculinized TH expression in the male AVPV. This observation suggests that<br />

BPA and GEN blocked the masculinizing effect of endogenous estrogen on this endpoint.<br />

Interestingly, in the adult animals GEN and E2, but not BPA, disrupted AVPV volume in both<br />

sexes, and neither EAC feminized GnRH secretion patterns in the adult males. GEN, also<br />

disrupted estrus cyclicity in the females. Over 80% of the GEN and E2 treated females were<br />

acyclic 6 weeks post-puberty, but the BPA females retained normal cycles through this period.<br />

These data demonstrate that the disruption of nuclear volume does not necessarily coincide with<br />

disruption of cellular phenotype or neuroendocrine function. SDN volume and phenotype were<br />

only examined in the adult males. SDN and CALB-SDN volume were unchanged by treatment, but<br />

the number of neurons immunoreactive for CALB in the SDN was significantly increased by both<br />

BPA and GEN. In this case the results suggest that BPA and GEN augmented the masculinizing<br />

effects of endogenous estrogen. Collectively, our results demonstrate that neonatal exposure to<br />

EACs such as BPA and GEN can affect sexually dimorphic brain morphology and neuronal<br />

phenotypes in adulthood with regional and cellular specificity. They also illustrate that EACs can<br />

simultaneously augment and interfere with the effects of endogenous estrogen on the developing<br />

brain. These findings emphasize the need to employ a comprehensive approach that addresses both<br />

anatomical and functional endpoints when evaluating the potential effects of EAC exposure.<br />

90


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

THE ERE-LUC REPORTER MOUSE: TWENTY YEARS OF BASIC RESEARCH<br />

TO BE APPLIED TO THE STUDY OF ENDOCRINE DISRUPTERS<br />

Maggi A.<br />

Center Of Excellence CEND University Of Milan, Via Balzaretti 9, I-20133 Milan, Italy.<br />

E-Mail: adriana.maggi@unimi.It. Fax +39-0250318290<br />

In view of the newly acquired awareness on the complexity and tissue specificity of IR<br />

mechanism of action and functions there is a growing concern on the validity of the<br />

methodologies so far applied to identify EDs and to predict their harmful effects. Current<br />

in vivo and in vitro methodologies restrict the analysis to very specific organs or cell types<br />

and, therefore, are not sufficient to predict the potential consequences of EDs after short or<br />

long-term exposure. The recent emphasis on generating in vitro model systems for<br />

toxicological analysis has certainly discouraged the development of models that are<br />

suitable to envision the whole spectrum of effects on the body, which is required when<br />

dealing with endocrine disruption that is, by definition, a living organism. We took<br />

advantage of novel molecular imaging techniques applied to animal engineering create a<br />

model system enabling to measure in living animals the state of trancriptional activity, the<br />

ERE-Luc reporter mouse. This model represents a first paradigmatic reporter mouse<br />

provided major insights on ER physiology and has been utilized to asses its suitability to<br />

the study of the systemic effects of EDs.<br />

With comparative studies involving the quantitative analysis of photon emission in vivo<br />

and luciferase enzymatic activity ex vivo we demonstrate that imaging is a method<br />

applicable to obtain a map of the effects of different concentrations of a given<br />

xenoestrogen in space and time. The faithfulness of luciferase as reporter of ER<br />

transcriptional activity is shown by comparing the effect of administration of oestrogens of<br />

different origin (the natural hormone 17β-oestradiol; the phytoestrogen genistein and the<br />

synthetic oestrogens p,p’-DDT and BHC) on the accumulation of luciferase or transcripts<br />

from endogenous ER target genes (such as progesterone receptor, CYP17 and the<br />

oestrogen receptors themselves). We also appied imaging-based methodology to the study<br />

of the effects of long-term exposure to food oestrogens like the phytoestrogen genistein or<br />

to complex mixture of oestrogens such as soy milk. Our experiments showed that<br />

phytoestrogens accumulate in the body and in long-term exposures interfere with ER<br />

signalling in a tissue selective manner. Most interestingly, treatment with soy milk<br />

generates a state of activation of ERs which differs significantly from that which was<br />

shown with genistein: in long-term treatment we observed a slight activation of ERs in<br />

liver, but a very significant activation in testis, indicating that the pure substance, genistein,<br />

may have effects significantly different than mixtures of phytoestrogens even if enriched in<br />

genistein itself. The model system proved particularly sensitive to exposure to different<br />

concentrations of xenoestrogens because it could show different responses to<br />

administration of pure or dilute soy milk and the state of ER activity was directly<br />

proportional to the amount of soy milk ingested. These data point to the necessity to better<br />

investigate the potential ED activity of soy milk, particularly when administered to infants<br />

of both sexes.<br />

In our view, the reporter mouse technology is an excellent candidate to REPLACE the<br />

existing tests that, for their nature, are unable to provide an overall view of oestrogenic<br />

activity in the whole organism. By means of non-invasive in vivo imaging technology, the<br />

methods provide the opportunity to REDUCE the number of animals to be used in the in<br />

vivo tests first of all because: animal sacrifice is not needed, and secondly the possibility to<br />

91


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

follow the endocrine effects in time in the same animal reduces the need to use large<br />

numbers of animals in each experimental group to reach significant data and also reduces<br />

the need for control groups (the effects of a treatment is evaluated versus the baseline state<br />

of activity of the receptor on the same animal). Last, but not least, the technology will<br />

REFINE current methods by providing for the first time the possibility to study the effect<br />

of EDs systemically and, after long-term exposure even to low doses, our methodology<br />

will eliminate the pain created by animal testing and abolish the necessity of animal<br />

sacrifice.<br />

Reference list<br />

[1] Ciana P, Brena A, Sparaciari P, Bonetti E, Di Lorenzo D, Maggi A. Estrogenic<br />

activities in rodent estrogen-free diets. Endocrinology. 2005, 146:5144-50. Epub 2005<br />

Sep 8.<br />

[2] Di Lorenzo D, Villa R, Biasiotto G, Belloli S, Ruggeri G, Albertini A, Apostoli P,<br />

Raviscioni M, Ciana P, Maggi Isomer-specific activity of<br />

dichlorodyphenyltrichloroethane with estrogen receptor in adult and suckling estrogen<br />

reporter mice. Endocrinology. 2002, 143:4544-51.<br />

[3] Maggi A, Ciana P. Reporter mice and drug discovery and development. Nat Rev Drug<br />

Discov. 2005, 4:249-55.<br />

[4] Villa R, Bonetti E, Penza ML, Iacobello C, Bugari G, Bailo M, Parolini O, Apostoli P,<br />

Caimi L, Ciana P, Maggi A, Di Lorenzo D. Target-specific action of organochlorine<br />

compounds in reproductive and nonreproductive tissues of estrogen-reporter male<br />

mice. Toxicol Appl Pharmacol. 2004, 201:137-48<br />

92


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

ENVIRONMENTAL-LIKE EXPOSURE TO XENOESTROGEN AFFECTS<br />

SEXUAL DIFFERENTIATION OF BRAIN AND BEHAVIOR IN FEMALE RATS<br />

°Corrieri L., *Della Seta D., *Materazzi P., °Dessì-Fulgheri F., *Farabollini F., and<br />

*Fusani L.<br />

*Dipartimento di Fisiologia, Sezione di Neuroscienze e Fisiologia Applicata, Università di<br />

Siena, via Aldo Moro, 53100, Siena, Italy, e-mail: fusani@unisi.it, fax: +39-0577-234037;<br />

°Dipartimento di Biologia Animale e Genetica, Università di Firenze.<br />

Estrogens act on the brain during development to regulate sexual differentiation of brain<br />

and behavior. In the adult female, estrogens are involved in the regulation of sexual behavior by<br />

acting on estrogen-sensitive brain structures. In particular, the sexually dimorphic nuclei of the<br />

preoptic area (SDN-POA) and the anteroventral periventricular nucleus (AVPV) are sexually<br />

dimorphic and are involved in the regulation of sexual behavior and the estral cycle in female rats.<br />

Environmental estrogens or xenoestrogens, entering the body through contaminated food and<br />

water, have the potential of interfering with the differentiation of brain and behavior and thus<br />

induce permanent physiological, behavioral and neural alterations. There is little experimental<br />

evidence for such effects in mammals, and the biological relevance of several studies has been<br />

question because the modalities of treatment did not mimic environmental exposure. In addition,<br />

most ‘xenoestrogens’ have additional toxic effects, therefore it is difficult to identify true<br />

xenoestrogenic modulation.<br />

We studied the effects of an environmental-like exposure to the pure, synthetic estrogen,<br />

ethynylestradiol (EE), on the sexual differentiation of brain and behavior in Sprague-Dawley<br />

female rats. EE is found in surface waters for its widespread use – it is the main estrogen of<br />

anticonceptional pills. The rats were treated from conception to puberty with either of two doses of<br />

EE: a lower dose (EEL, 4 ng/kg/day) equivalent to concentrations found in contaminated waters,<br />

and a higher dose (EEH, 400 ng/kg/day) equivalent to that of anticonceptional pills. Rats were<br />

treated indirectly from GD 5 to PND 21 by oral administration in peanut oil (EEL, EEH, or vehicle<br />

only = OIL) to the mothers and directly from PND 22 to PND 30. At 12 weeks of age, females<br />

were tested for sexual behavior with a standard sexually active male. After testing, the rats were<br />

deeply anesthetized and perfused with 4% formaldehyde. The brain was dissected, postfixed for 2<br />

hr in 4% formaldehyde and then cryoprotected by impregnation with 10% and then 20% sucrose in<br />

phosphate-buffer saline. Forty-micron sections were cryocut, mounted, and stained with thionin.<br />

We took digital microphotographs of the preoptic-hypothalamic region at 40X and measured the<br />

areas of interest.<br />

Exposure to the higher EE dose (EEH) caused loss of estral cyclicity with permanent<br />

vaginal estrus. Proceptive behavior was altered in females treated with the lower dose of EE (EEL)<br />

compared to controls (OIL). The EEL females also showed a larger volume of the medial preoptic<br />

nucleus compared to both EEH and OIL females. Thus, although exposure to pharmacological<br />

doses of EE throughout development results in major physiological alterations in adulthood, effects<br />

on sexually dimorphic brain structures appears to be limited. On the contrary, an exposure to lower,<br />

environmental-like doses of EE does not disrupt the estral cycle, but both sexual behavior and<br />

sexually dimorphic nuclei are altered. This work shows that xenoestrogen exposure during<br />

development can affect the differentiation of sexual behavior and sexually dimorphic nuclei in<br />

female rats, and that concentrations of xenoestrogen with no evident effects on physiological<br />

markers can in fact induce significant alterations in brain morphology and behavior.<br />

93


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

SEXUAL DIMORPHISM IN EXPLORATION AND IN THE NUMBER OF TH-<br />

IMMUNOSTAINED NEURONS IN THE LOCUS COERULEUS OF YOUNG MICE<br />

IS ELIMINATED BY PRENATAL EXPOSURE TO ENVIRONMENTAL<br />

ESTROGENS<br />

Ponzi D.(1,2), Maruniak J.(2), Parmigiani S.(1), Vom Saal F.S.(2), and Palanza P.(1)<br />

(1) Dipartimento di Biologia Evolutiva e Funzionale, Università di Parma, Parco Area<br />

delle Scienze 11A, 43100 Parma, Italy; (2) Div. Biol. Sci. Univ. of Missouri, Columbia-<br />

Mo, USA<br />

Bisphenol A is a man made chemical used in the production of polycarbonate,<br />

epoxy resins lining food storage cans and dental sealants. Because of its estrogenic<br />

activity, very high production volume, and presence in virtually all humans there is<br />

concern about its ability to disrupt the endocrine system, especially the neuroendocrine<br />

system. During fetal life the intrauterine environment is critical for the normal<br />

development, and even small changes in the levels of hormones, such as estradiol, or<br />

estrogen-mimicking chemicals, such as BPA, lead to changes in the brain structure,<br />

function and consequently in the behavior. It has thus been proposed that exposure during<br />

fetal life to chemicals such as BPA could contribute to mental diseases expressed later in<br />

life. In this experiment we administered to CD-1 pregnant mice 50 micrograms/kg body<br />

weight or to 5 mg/kg body weight doses per day of BPA from day 11-18 of gestation. 0.5<br />

microgram/kg of diethylstilbestrol (DES) and corn oil were used as positive and negative<br />

controls respectively. The aim of this experiment was to examine the effect of the prenatal<br />

exposure of BPA or DES on the patterns of exploration and on the number of neurons<br />

producing tyrosine hydroxylase (TH) in the locus coeruleus of 30 days old mice by<br />

performing an immunohistochemical assay. We choose to examine LC because of its<br />

sexual dimorphism in rodents and because the noradrenergic system (and more widely the<br />

monoaminergic system) is altered in several neurological diseases such as ADHD, anxiety<br />

related disorders and Parkinson's disease. Mice perinatally exposed to BPA and DES<br />

showed altered patterns of exploratory behavior; specifically, sex differences in<br />

exploratory activity were eliminated. We found that control animals showed sex difference<br />

in the number of TH-stained neurons in the LC, with females having significantly more<br />

stained neurons, but the exposure to the 5mg dose of BPA eliminate this difference as did<br />

DES. This study provides further evidence that fetal exposure to doses of BPA that are far<br />

below the “safe” , no effect dose can alter behavioral and brain development.<br />

94


TUESDAY, 20 th February 2007<br />

11.30 - 12.30<br />

Young Investigators Symposium


Young Investigators Symposium<br />

(Chairs: Frye C., Mensah-Nyagan A.G.)<br />

• Belloni V., Alleva E., Dessì-Fulgheri F., Zaccaroni M., Santucci D. (Italy) Effects<br />

of low doses of atrazine on the neurobehavioral development of mice<br />

• Forlano P.M., Bass, A.H. (USA) Substrates for plasticity: brain aromatase,<br />

estrogen and androgen receptors in sexually dimorphic, vocal-acoustic and<br />

auditory pathways in a teleost fish<br />

• Meyer L., Patte-Mensah C., Mensah-Nyagan AG. (France) The enzyme 3alphahydroxysteroid<br />

oxidoreductase is a key regulator of nociceptive mechanisms in the<br />

rat spinal cord<br />

• Romeo, R.D., McEwen, B.S. (USA) Stress during adolescence leads to depressivelike<br />

behaviors and changes in hypothalamic-pituitary-adrenal axis function in<br />

adulthood<br />

• Taziaux M., Keller M., Bakker J., Balthazart J. (Belgium) Brain estradiol rapidly<br />

regulates in a neurotransmitter-like fashion male sexual behavior in mice<br />

• Paris J.J., Rhodes M.E., Frye C.A. (USA) Inhibition of 3α,5α-THP formation<br />

decreases exploratory/anti-anxiety and socio-sexual behavior in sexually receptive<br />

female rats


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

EFFECTS OF LOW DOSES OF ATRAZINE ON THE NEUROBEHAVIORAL<br />

DEVELOPMENT OF MICE<br />

Belloni V. * , Alleva E. § , Dessì-Fulgheri F. * , Zaccaroni M. * , and Santucci D. §<br />

* Department of Animal Biology and Genetics, University of Firenze, Via Romana 17,<br />

I-50125 Firenze, Italy<br />

§<br />

Section of Behavioral Neurosciences, Dep. Cell. Biol. and Neuroscience, Istituto Superiore di<br />

Sanità, viale Regina Elena 299, I-00161 Rome, Italy<br />

Atrazine (ATZ) is a triazine herbicide widely used and frequently detected in ground and<br />

surface water. The extensive use of atrazine has made this compound a focus for environmental<br />

impact studies [6].<br />

Recent studies suggest that ATZ is an environmental risk factor able to affect estrogen production<br />

by inducing aromatase [7], the enzyme that converts androgen into estrogen [2], an essential<br />

transformation occurring at the CNS level for maturation and expression of behavior. Then, since it<br />

is well known that availability of sexual steroids, testosterone and estradiol, at CNS level, is<br />

essential for maturation and expression of sexually dimorphic behaviors [3], we consider that<br />

behavior may be an efficient marker of subtle effects of low ATZ concentrations.<br />

In the present study we evaluated the effects of environmentally-relevant doses of ATZ on somatic<br />

growth and early behavioral ontogeny, a crucial stage in shaping future behavior. For this purpose<br />

we observed mice born to mothers exposed to 1 or 100 µg/kg ATZ during pregnancy and lactation.<br />

We studied, between postnatal day 2 to 15, righting reflex, cliff aversion, forepaw grasping,<br />

auditory startle, eyelid and ear opening [5], and ultrasound vocalizations vocalizations [1, 4]. In<br />

both sexes effects of ATZ were evident on body weight at birth, on the maturation of righting and<br />

grasping reflexes, and on the rate of emission and the spectrographic characteristics of ultrasound.<br />

At birth (PND2) the offspring of ATZ treated mothers was lighter than controls (Fig.1), but this<br />

difference is compensated from PND4 on.<br />

Moreover, ontogenesis of cliff aversion and of auditory startle were not affected by ATZ, while<br />

righting reflex maturation was delaied in both sexes and grasping reflex maturation was accelerated<br />

in male pups (Table 1).<br />

Finally, the rate of emission of vocalizations showed a tendency to decrease at PND9 both in males<br />

and in females (Fig.2). A spectrographic analysis was conducted at PND9, revealing an increase of<br />

the interval between vocalizations in females (Fig.3) and an increase of the peak frequency<br />

(maximum amplitude frequency) in ATZ pups.<br />

These findings are also consistent with an additional study conducted to evaluate long-term effect<br />

of perinatal administration of ATZ on behavior from weaning to the adult. Infact, our results<br />

showed that ATZ is able to affect explorative and social behavior in males and cognitive behavior<br />

in both sexes.<br />

Dosage level appeared also particularly relevant since, in some cases the lowest ATZ exposure was<br />

more effective than the highest one in modifying behavior. This suggest that this compound,<br />

similarly to many others endocrine disruptors [8], does not follow a linear dose-response curve [6],<br />

and that, as a consequence, its effects should be studied by employing low, environmentallyrelevant<br />

exposure levels. Our results, compatible with ATZ properties, suggest caution in the use of<br />

a chemical agent that may, even at low doses, interfere with brain development and differentiation,<br />

inducing alterations of developmental trajectories of behavior.<br />

Reference list<br />

[1] Alleva, E., Laviola, G., Tirelli, E., Bignami, G. Short-, medium-,and long-term effects of prenatal oxazepam on<br />

neurobehavioural development of mice. Psychopharmacology, 1985;87:434-441.<br />

[2] Amateau, S.K., Alt, J.J., Stamps, C.L., McCarthy, M.M. Brain estradiol content in newborn rats: sex differences,<br />

regional heterogeneity, and possible de novo synthesis by the female telencephalon. Endocrinology,<br />

2006;145:2906-2917.<br />

[3] Arnold, A.P., Gorski, R.A. Gonadal steroid induction of structural sex differences in the central nervous system.<br />

Ann Rev Neurosci, 1984;7:413-442.<br />

97


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

[4] Branchi, I., Santucci, D., Alleva, E. Ultrasonic vocalization emitted by infant rodents: a tool for assessment of<br />

neurobehavioral development. Behav Brain Res, 2001;125:49-56.<br />

[5] Fox, M. Reflex-ontogeny and behavioural development of the mouse. Anim Behav, 1965;13:234-241.<br />

[6] Hayes, T.B., Haston, K., Tsui, M., Hoang, A., Haeffele, C., Vonk, A. Feminization of male frog in the wild.<br />

Nature, 2002;419:895-896.<br />

[7] Sanderson, J.T., Letcher, R.J., Heneweer, M., Giesy, J.P., van den Berg, M. Effects of Chloro-s Triazine Herbicides<br />

and Metabolites on Aromatase Activity in Various Human Cell Lines and on Vitellogenin Production in Male Carp<br />

Hepatocytes. Environ Health Perspect , 2001;109:1027-1031.<br />

[8] Welshons, W.V., Nagel, S.C., vom Saal, F.S., Large effects from small exposures. III. Endocrine mechanisms<br />

mediating effects of bisphenol A at levels of human exposure. Endocrinology, 2006;147: 56-69.<br />

Table 1. Assessment of somatic growth and neurobehavioral development (day of maturation) of<br />

atrazine-treated pups and their controls from PND2 to PND15 (n of subjects).<br />

Day of adult-like response<br />

Males<br />

Righting<br />

Cliff<br />

aversion<br />

Forepaw<br />

grasp<br />

Auditory<br />

startle<br />

Eyes<br />

opening<br />

Ears<br />

opening<br />

Control (20) 4,20±0,40 4,20±0,40 7,50±0,53 11,45±0,37 14,90±0,10 14,<strong>35</strong>±0,13<br />

ATL (17) 4,65±0,73 4,12±0,49 5,18±0,53** 11,18±0,42 14,65±0,17 14,05±0,13<br />

ATH (20) 6,<strong>35</strong>±0,69 3,50±0,20 5,25±0,50** 11,25±0,40 14,95±0,08 14,30±0,10<br />

Females<br />

Control (20) 4,70±0,67 4,40±0,37 6,05±0,65 11,72±0,27 14,95±0,11 14,30±0,01<br />

ATL (18) 5,00±0,71 4,22±0,45 6,77±0,60 10,38±0,43 14,89±0,08 14,17±0,09<br />

ATH (20) 6,85±0,68 * 4,<strong>35</strong>±0,51 6,65±0,62 10,71±0,41 15,00±0,00 14,<strong>35</strong>±0,11<br />

ATL, low dose atrazine; ATH, high dose atrazine. **p


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

SUBSTRATES FOR PLASTICITY: BRAIN AROMATASE, ESTROGEN AND<br />

ANDROGEN RECEPTORS IN SEXUALLY DIMORPHIC, VOCAL-ACOUSTIC<br />

AND AUDITORY PATHWAYS IN A TELEOST FISH<br />

Forlano P.M., and Bass A.H.<br />

Department of Neurobiology and Behavior, Cornell University, Ithaca, NY, U.S.A.<br />

pmf4@cornell.edu, Fax 607-254-1303<br />

The plainfin midshipman fish, Porichthys notatus, is a well-established<br />

neuroethological model system that has provided insights into neural and endocrine<br />

mechanisms of vocal-acoustic communication that are conserved throughout vertebrates.<br />

During the breeding season, males emit a mate call generated by a hindbrain-spinal cord<br />

pattern generator which innervates sonic musculature, and this signal is necessary and<br />

sufficient for females to localize potential mates [4]. Furthermore, this vocal-motor system<br />

is inter-and intrasexually dimorphic, i.e. individual neurons and nuclear volume are larger<br />

in courting (type I) males compared to non-courting (type II) males and females [1,3].<br />

Androgens are known to masculinize dimorphic vocal motor circuitry and muscle, and<br />

both androgens and estrogen rapidly modulate vocal patterning, and are elevated during<br />

gonadal recrudescence in males and females and vocal courtship in males [2,5,11,12,15].<br />

Since the enzyme aromatase could provide estrogen and/or regulate how much androgen<br />

(testosterone) reaches specific populations of neurons, we wanted to determine the sites of<br />

action of aromatase as well as estrogen and androgen receptors in the function of this<br />

behaviorally significant circuitry.<br />

We investigated the neuroanatomical localization of aromatase mRNA and protein<br />

using in situ hybridization (ISH) and teleost-specific antibodies [9]. The cellular basis of<br />

aromatase production in the teleost brain was discovered by showing abundant expression<br />

throughout the central nervous system in radial glia rather than neurons. Highest numbers<br />

of cells expressing aromatase were found in the forebrain, along ventricular areas<br />

throughout the brain and in the sexually dimorphic vocal-motor center of the hindbrainspinal<br />

cord. Quantitative ISH showed inter and intrasexual dimorphism in aromatase<br />

mRNA expression within the vocal hindbrain-spinal cord, consistent with differences in<br />

aromatase activity and other dimorphisms in this region between alternative male<br />

reproductive/vocal phenotypes [6,13].<br />

In order to identify potential sites of action of local produced estrogens as well as<br />

sites of interaction between androgens and aromatase, ISH was employed to identify<br />

estrogen receptor alpha (ER) and androgen receptor (AR) mRNA distribution. AR<br />

expression was found to be more widespread than ER alpha in the brain, and robustly<br />

expressed in descending vocal motor nuclei and vocal-acoustic integration centers.<br />

Furthermore, while ER expression in the dimorphic sonic motor nucleus appear to be over<br />

the motor neurons themselves, AR mRNA follows a similar pattern to aromatase<br />

expression in glial cells, surrounding the motor nucleus [8,10]. Thus, while ARs may act<br />

directly on glial cells to regulate aromatase expression, estrogen produced by glia likely<br />

acts in a paracrine fashion to reach nearby motor neurons which express nuclear and/or<br />

membrane ERs. Together, this neuroanatomical evidence demonstrates that aromatase in<br />

glial cells within vocal circuitry can provide a local estrogen source for rapid modulation<br />

of vocal behavior. Furthermore, aromatase may function to regulate male morph-dependent<br />

differences in the availability of estrogen to act as a neurosteroid throughout the forebrain<br />

and in brain regions that support divergent reproductive and, in this case vocal,<br />

phenotypes.<br />

99


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

Males and females have distinct seasonal patterns of circulating steroids which<br />

correspond to changes in gonadal development, and reproductive and vocal behavior [15].<br />

Frequency encoding by the inner ear of females changes seasonally, such that they are<br />

better adapted to detect the mate call of the male during the breeding season. This plasticity<br />

in audition was demonstrated experimentally to depend on elevated plasma levels of either<br />

testosterone or estradiol that naturally accompany the seasonal occurrence of reproduction<br />

[14]. Furthermore, these same manipulations upregulate and mimic seasonal changes in<br />

expression of aromatase, throughout the brain [7]. In the peripheral auditory system,<br />

aromatase-ir was identified in ganglion nerve cells in the branch of the eighth nerve<br />

adjacent to the sensory epithelium of the main auditory end organ (saccule) of the inner ear<br />

of females. Furthermore, ERα mRNA expression was found in unidentified cells just<br />

outside the saccular hair cell layer [8]. Localization of aromatase and ER in the inner ear<br />

support the hypothesis that estradiol alone could account for effects of circulating steroids<br />

on hearing. Thus, the ear itself may provide a local source of estrogen independent of the<br />

gonad to modulate plasticity of hearing.<br />

Research support from the U.S. National institutes of Health (NIDCD DC00092) and National Science<br />

Foundation (IBN9987341, IOB-0516748) to A.H.B., NIMH T32MH015793 to P.M.F.<br />

Reference list<br />

[1] Bass, A.H. and Baker, R., Sexual dimorphisms in the vocal control system of a teleost fish: morphology of<br />

physiologically identified neurons, J Neurobiol, 21 (1990) 1155-1168.<br />

[2] Bass, A.H. and Forlano, P.M., Neuroendocrine mechanisms of alternative reproductive tactics: the chemical<br />

language of social plasticity. In R. Oliveira, M. Taborsky and J. Brockmann (Eds.), Alternative Reproductive<br />

Tactics: An Integrative Approach, Cambridge University Press, Cambridge, UK, in press.<br />

[3] Bass, A.H. and Marchaterre, M.A., Sound-generating (sonic) motor system in a teleost fish (Porichthys<br />

notatus): Sexual polymorphisms and general synaptology of sonic motor nucleus, J Comp Neurol, 286 (1989)<br />

154-169.<br />

[4] Bass, A.H. and McKibben, J.R., Neural mechanisms and behaviors for acoustic communication in teleost fish,<br />

Prog Neurobiol, 69 (2003) 1-26.<br />

[5] Brantley, R.K., Marchaterre, M.A. and Bass, A.H., Androgen effects on vocal muscle structure in a teleost fish<br />

with inter-sexual and intra-sexual dimorphism, J Morphol, 216 (1993) 305-318.<br />

[6] Forlano, P.M. and Bass, A.H., Seasonal plasticity of brain aromatase mRNA expression in glia: Divergence<br />

across sex and vocal phenotypes, J Neurobiol, 65 (2005) 37-49.<br />

[7] Forlano, P.M. and Bass, A.H., Steroid regulation of brain aromatase expression in glia: Female preoptic and<br />

vocal motor nuclei, J Neurobiol, 65 (2005) 50-8.<br />

[8] Forlano, P.M., Deitcher, D.L. and Bass, A.H., Distribution of estrogen receptor alpha mRNA in the brain and<br />

inner ear of a vocal fish with comparisons to sites of aromatase expression, J Comp Neurol, 483 (2005) 91-113.<br />

[9] Forlano, P.M., Deitcher, D.L., Myers, D.A. and Bass, A.H., Anatomical distribution and cellular basis for high<br />

levels of aromatase activity in the brain of teleost fish: aromatase enzyme and mRNA expression identify glia<br />

as source, J Neurosci, 21 (2001) 8943-55.<br />

[10] Forlano, P.M., Marchaterre, M.A., Deitcher, D.L. and Bass, A.H., Distribution of androgen receptor mRNA in<br />

vocal and non-vocal circuitry of a teleost fish. Society for Neuroscience, Society for Neuroscience Online,<br />

Washington, D.C., 2005, pp. 1001.6.<br />

[11] Knapp, R., Marchaterre, M.A. and Bass, A.H., Relationship between courtship behavior and steroid hormone<br />

levels in parental male plainfin midshipman fish, Horm Behav, 39 (2001) 3<strong>35</strong>.<br />

[12] Remage-Healey, L. and Bass, A.H., Rapid, hierarchical modulation of vocal patterning by steroid hormones, J<br />

Neurosci, 24 (2004) 5892-900.<br />

[13] Schlinger, B.A., Greco, C. and Bass, A.H., Aromatase activity in the hindbrain vocal control region of a teleost<br />

fish: divergence among males with alternative reproductive tactics, P Roy Soc Lond B Bio, 266 (1999) 131-<br />

136.<br />

[14] Sisneros, J.A., Forlano, P.M., Deitcher, D.L. and Bass, A.H., Steroid-dependent auditory plasticity leads to<br />

adaptive coupling of sender and receiver, Science, 305 (2004) 404-7.<br />

[15] Sisneros, J.A., Forlano, P.M., Knapp, R. and Bass, A.H., Seasonal variation of steroid hormone levels in an<br />

intertidal-nesting fish, the vocal plainfin midshipman, Gen Comp Endocrinol, 136 (2004) 101-116.<br />

100


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

THE ENZYME 3ALPHA-HYDROXYSTEROID OXIDOREDUCTASE IS A KEY<br />

REGULATOR OF NOCICEPTIVE MECHANISMS IN THE RAT SPINAL CORD<br />

Meyer L., Patte-Mensah C. and Mensah-Nyagan A.G.*<br />

Institut des Neurosciences Cellulaires et Intégratives, UMR7168/LC2 CNRS-ULP, Equipe<br />

« Stéroïdes et Système Nociceptif », 67084 Strasbourg Cedex, France.<br />

* E-mail : gmensah@neurochem.u-strasbg.fr<br />

Neuroactive 5α,3α-reduced steroids are potent activators of GABA-A receptors<br />

which play a crucial role in the modulation of nociceptive sensitivity. The biosynthesis of<br />

these steroids requires the activity of 3α-hydroxysteroid oxidoreductase (3α-HSOR) which<br />

converts 5α-reduced steroids such as dihydroprogesterone (5α-DHP) into 5α,3αmetabolites<br />

as tetrahydroprogesterone (5α,3α-THP) also called allopregnanolone.<br />

Immunohistochemical investigations revealed a strong expression of 3α-HSOR in<br />

the rat spinal cord (SC), particularly in the dorsal horn which controls nociceptive<br />

transmission. Incubation of SC slices with [3H]progesterone yielded the formation of<br />

[3H]5α-DHP and [3H]5α,3α -THP indicating that 3α-HSOR located in the rat SC is an<br />

active form of the enzyme. Moreover, in the SC of rats submitted to neuropathic pain<br />

generated by sciatic nerve ligature, 3α-HSOR activity was up-regulated and enhanced<br />

5α,3α-THP endogenous concentration. Direct intrathecal injection of 5α,3α-THP in the<br />

lumbar SC of naive rats increased the thermal and mechanical sensitivity thresholds which<br />

were assessed by the Hargreaves’ method and the Von Frey filament test, respectively.<br />

These sensitivity thresholds were both decreased by Provera, a pharmacological inhibitor<br />

of 3α-HSOR activity. We observed that the neuropathic-pain rats were characterized by<br />

thermal hyperalgesia and mechanical allodynia. Treatment of these neuropathic animals<br />

with 5α,3α-THP significantly increased both thermal and mechanical thresholds<br />

suggesting a potential analgesic action of neuroactive 5α,3α-reduced steroids in the control<br />

of neuropathic pain.<br />

Taken together, our results provide the first neurochemical and behavioral evidence<br />

for an effective role of 3α-HSOR in the modulation of nociceptive mechanisms in the<br />

spinal cord.<br />

101


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

STRESS DURING ADOLESCENCE LEADS TO DEPRESSIVE-LIKE<br />

BEHAVIORS AND CHANGES IN HYPOTHALAMIC-PITUITARY-ADRENAL<br />

AXIS FUNCTION IN ADULTHOOD<br />

Romeo R.D., and McEwen B.S.<br />

Rockefeller University, Laboratory of Neuroendocrinology, 1230 York Ave., New York,<br />

NY USA, romeor@rockefeller.edu, fax:212-327-8634<br />

Adolescence is a period of significant developmental vulnerabilities [1,2,7], marked by<br />

increases in the morbidity of, and susceptibility to, numerous psychological disorders (e.g.,<br />

anxiety and depression) [4]. The mechanisms mediating the adolescent (pubertal) increase<br />

in these disorders are presently unknown. In adulthood, it is well established that exposure<br />

to stressors can lead to the onset and exacerbation of psychological disorders [5].<br />

Interestingly, recent studies in adolescent boys and girls have suggested that pubertal<br />

exposure to stress may be a particularly relevant environmental factor contributing to an<br />

individual’s vulnerability to various psychopathologies later in life [3,8]. In an effort to<br />

model how adolescent stress exposure affects neurobehavioral function in adult animals,<br />

the present study investigated whether physical and/or psychological stressors (e.g.,<br />

restraint stress and/or social isolation) experienced during puberty leads to changes in<br />

depressive-like behaviors and hypothalamic-pituitary-adrenal (HPA) axis function in<br />

adulthood.<br />

To these ends, 40 male Sprague-Dawley rats were randomly assigned to one of four<br />

experimental groups (n=10): (i) GROUP HOUSED (3 per cage), (ii) GROUP HOUSED +<br />

RESTRAINT (3 per cage + 1 h of restraint stress every other day), (iii) ISOLATION<br />

(housed alone) and (iv) ISOLATION + RESTRAINT (housed alone + 1 h of restraint<br />

stress every other day). Animals were exposed to these conditions from 28 to 50 days of<br />

age, encompassing the pubertal period of development in this species. Animals were<br />

weighed weekly to examine growth rates. At 52 and 53 days of age all animals were<br />

administered the Porsolt forced swim test (FST), a pharmacologically validated behavioral<br />

test to assess learned helplessness and depressive-like behaviors [6]. To investigate<br />

possible changes in basal and stress-induced HPA reactivity, 48 h after the behavioral tests,<br />

each group was further divided into two (n =5) and blood was collected from the animals<br />

either immediately before or after a 30 min session of restraint stress. We found that both<br />

groups of animals exposed to restraint stress during the pubertal period demonstrated less<br />

weight gain during adolescence compared to animals either group housed or socially<br />

isolated without restraint. Our analyses on the FST data revealed that compared to group<br />

housed controls, animals undergoing the pubertal physical and/or social stressors showed a<br />

shorter latency to become immobile and spent significantly less time struggling/swimming<br />

and more time immobile, all indicating greater learned-helplessness behavior. Finally,<br />

basal plasma corticosterone (CORT) levels were significantly elevated in the animals<br />

exposed to the physical and/or social stressors during puberty compared to group housed<br />

controls. These data indicate that pubertal exposure to physical and/or social stressors lead<br />

to reduced weight gain during adolescence and increased depressive-like behaviors and<br />

basal CORT levels in adulthood. Importantly, these results lend construct and face validity<br />

to this model of pubertal stress and depressive-like behaviors in adulthood and recapitulate<br />

three key features of typical, melancholic depression: weight loss, feelings of learnedhelplessness<br />

and elevated basal HPA function. These data highlight how stress and<br />

pubertal development can interact to affect adult behavioral, physiological and endocrine<br />

102


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

function and provide a model to study stress-induced neurobehavioral changes during<br />

adolescent development.<br />

Acknowledgements:<br />

This work was supported by NIH grants MH 065749 and MH 41256<br />

Reference list<br />

[1] S.L. Andersen, Trajectories of brain development: point of vulnerability or window of<br />

opportunity., Neuroscience and Biobehavioral Reviews 27 (2003) 3-18.<br />

[2] R.E. Dahl, Adolescent brain development: a period of vulnerabilities and opportunities.,<br />

Annals of the New York Academy of Sciences 1021 (2004) 1-22.<br />

[3] K.E. Grant, B.E. Compas, A.F. Stuchlmacher, A.E. Thurn, S.D. McMahon, J.A. Halpert,<br />

Stressors and child and adolescent psychopathology: moving from markers to mechanisms<br />

of risk., Psychological Bulletin 129 (2003) 447-466.<br />

[4] A. Masten, Toward a developmental psychopathology of early adolescence. In M. Levin<br />

and E. McArnarny (Eds.), Early adolescent transitions, Heath, Lexington, 1987, pp. 261-<br />

278.<br />

[5] B.S. McEwen, Mood disorders and allostatic load., Biological Psychiatry 54 (2003) 200-<br />

207.<br />

[6] R.D. Porsolt, M. Le Pichon, M. Jalfre, Depression: a new animal model sensitive to<br />

antidepressant treatments., Nature 266 (1977) 730-732.<br />

[7] L.P. Spear, The adolescent brain and age-related behavioral manifestations., Neuroscience<br />

and Biobehavioral Reviews 24 (2000) 417-463.<br />

[8] R.J. Turner, D.A. Lloyd, Stress burden and the lifetime incidence of psychiatric disorder in<br />

young adults., Archives of General Psychiatry 61 (2004) 481-488.<br />

103


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

BRAIN ESTRADIOL RAPIDLY REGULATES IN A NEUROTRANSMITTER-<br />

LIKE FASHION MALE SEXUAL BEHAVIOR IN MICE<br />

Taziaux* M., Keller* 1 M., Bakker* J. and Balthazart* J.<br />

*University of Liège, Center for Molecular and Cellular Neurobiology, Research Group in<br />

Behavioral Neuroendocrinology, 1 Avenue de l’Hôpital (Bat. B36), B-4000 Liège,<br />

Belgium. Fax : 32 (0)4- 366.59.71. E-mail : mtaziaux@student.ulg.ac.be<br />

1 Now at: Laboratoire de Physiologie de la Reproduction et des Comportements,<br />

CNRS/INRA/University of Tours, Nouzilly, France.<br />

Steroid hormones exert their physiological and behavioral effects predominantly by<br />

regulating the transcription of a variety of hormone-sensitive genes. In addition to these<br />

relatively “slow” genomic effects (hours to days), steroid hormones also exert rapid<br />

(within seconds or minutes after administration), non-genomic effects by acting at the<br />

membrane level in many cellular, in particular neural, systems. For example, estradiol-17β<br />

(E 2 ) modulates in this manner brain electrical activity, ion channel function or G-protein<br />

coupled receptor activity (see [7] for review). At present, the impact of these rapid cellular<br />

effects of E 2 on the functioning of the whole-organism is poorly documented.<br />

The presence of estrogen synthase (aromatase) in the brain was demonstrated more than<br />

30 years ago [8] and more recently the enzyme was shown to be present and active at the<br />

pre-synaptic level [9]. Changes in brain aromatase activity (AA) are mediated largely by<br />

slow steroid-dependent changes in enzyme transcription [10]. In addition, brain AA can be<br />

rapidly (within minutes) modulated by non-genomic mechanisms including Ca 2+ -<br />

dependent phosphorylations [1], suggesting that local brain estrogen concentrations could<br />

also be modified more rapidly than previously thought.<br />

Male sexual behavior offers excellent opportunities to investigate the relative<br />

contribution of genomic and non-genomic effects of estrogen. This behavior is activated in<br />

many vertebrate species by estrogens derived from testosterone aromatization in the brain<br />

that act largely by activating the transcriptional activity of nuclear estrogen receptors, but<br />

is also enhanced within 15-25 min in rats and quail by a single injection of a high dose of<br />

E 2 [5, 3]. Conversely, studies in quail indicated that the acute inhibition of AA (by<br />

injection of a specific inhibitor) rapidly decreases aspects of male sexual behavior (15 min<br />

post-injection; [4]) as well as significantly modifies the reaction latency to nociceptive<br />

stimuli (1-5 min post injection; [6]). Although these acute effects of aromatase inhibition<br />

could be partly (sexual behavior) or completely (nociception) reversed by a concomitant<br />

injection of estradiol, their specificity, and more specifically their link to aromatase<br />

inhibition could not be fully established.<br />

In the present experiments, we used male wild-type (WT) and aromatase knock-out<br />

(ArKO) mice to establish this specificity and further investigate rapid behavioral effects of<br />

fast up-and down-regulations of brain estrogen concentrations that should occur following<br />

activation (by dephosphorylation) or inactivation (by phosphorylation) of aromatase<br />

activity.<br />

With the use of ArKO mice that are presumably more sensitive to estrogen action due to<br />

their constitutive lack of exposure to this steroid, we first showed that most aspects of male<br />

sexual behavior are stimulated within 15 min after a single peripheral injection of E 2 . This<br />

rapid behavioral effect of E 2 was best observed if ArKO mice were sexually experienced<br />

and pre-treated for about one week with a small dose of EB, suggesting an interaction<br />

between slow genomic and fast non genomic actions of estrogen in the activation of this<br />

behavior.<br />

104


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

Conversely, a single injection of three different aromatase inhibitors (Vorozole TM , ATD<br />

or its metabolite 17OH-ATD) almost completely suppressed male sexual behavior<br />

(decrease in mount and intromission frequencies and increase in the latency of these<br />

behaviors) expressed 10 to 20 min later by WT mice. ATD and 17OH-ATD had in contrast<br />

no significant effect on the mounts and intromissions activated in male ArKO mice by a<br />

chronic treatment with estradiol benzoate (EB) confirming that the rapid behavioral<br />

inhibitions are, from the neuroendocrine point of view, specifically due to the blockade of<br />

aromatase activity. At the behavioral level, this idea is also supported by the fact that the<br />

behavioral inhibition observed in male WT mice following injection of ATD and 17OH-<br />

ATD could be partly reversed by the simultaneous injection of a single dose of E 2 .<br />

Moreover, the rapid inhibitory effects on male WT sexual behavior following<br />

administration of aromatase inhibitors were not systematically accompanied by a<br />

significant decrease in ano-genital olfactory investigations, suggesting that treated male<br />

mice were still socially and possibly sexually interested in the stimulus females. Finally, no<br />

changes in locomotor activity measured in the open field as well as no changes in odor<br />

preference for estrous females were detected following treatment with the aromatase<br />

inhibitors in male WT mice. Overall, these data strongly indicate that this behavioral<br />

inhibition is highly specific and does not simply result from a non-specific detrimental<br />

effect of the inhibitor.<br />

Taken together, the present experiments demonstrate that both up- and downregulations<br />

of brain estrogen concentrations rapidly affect in parallel the expression of<br />

male sexual behavior in ArKO and WT mice. These results, associated with previous<br />

reports demonstrating that brain E 2 is produced at the pre-synaptic level in a manner that<br />

can be rapidly modulated via calcium-dependent phosphorylations by inputs to aromataseexpressing<br />

cells [1] supports the notions that E 2 produced by aromatase in the brain<br />

displays most if not all the functional characteristics of neurotransmitters, or at least<br />

neuromodulators (see [2] for further elaboration of this notion).<br />

Reference list<br />

[1] Balthazart, J., Baillien, M., Charlier, T. D. & Ball, G. F., 2003. Calcium-dependent phosphorylation<br />

processes control brain aromatase in quail. Eur. J. Neurosci. 17, 1591-1606.<br />

[2] Balthazart, J. & Ball, G. F., 2006. Is brain estradiol a hormone or a neurotransmitter? Trends Neurosci.<br />

29, 241-9.<br />

[3] Cornil, C. A., Dalla, C., Papadopoulou-Daifoti, Z., Baillien, M. & Balthazart, J., 2006. Estradiol rapidly<br />

activates male sexual behavior and affects brain monoamine levels in the quail brain. Behav. Brain Res.<br />

166, 110-23.<br />

[4] Cornil, C. A., Taziaux, M., Baillien, M., Ball, G. F. & Balthazart, J., 2006. Rapid effects of aromatase<br />

inhibition on male reproductive behaviors in Japanese quail. Horm. Behav. 49, 45-67.<br />

[5] Cross, E. & Roselli, C. E., 1999. 17beta-estradiol rapidly facilitates chemoinvestigation and mounting in<br />

castrated male rats. Am. J. Physiol. Regul. Integr. Comp .Physiol. 276, R1346-R1<strong>35</strong>0.<br />

[6] Evrard, H. C. & Balthazart, J., 2004. Rapid regulation of pain by estrogens synthesized in spinal dorsal<br />

horn neurons. J. Neurosci. 24, 7225-7229.<br />

[7] McEwen, B.S., 2001. Invited review: Estrogens effects on the brain: multiple sites and molecular<br />

mechanisms. J. Appl. Physiol. 91, 2785-2801.<br />

[8] Naftolin, F., Ryan, K. J., Davies, I. J., Reddy, V. V., Flores, F., Petro, Z., Kuhn, M., White, R. J.,<br />

Takaoka, Y. & Wolin, L., 1975. The formation of estrogens by central neuroendocrine tissues.Rec. Prog.<br />

Horm .Res. 31, 295-319.<br />

[9] Peterson, R. S., Yarram, L., Schlinger, B. A. & Saldanha, C. J., 2005. Aromatase is pre-synaptic and<br />

sexually dimorphic in the adult zebra finch brain. Proc. Biol. Sci. 272, 2089-96.<br />

[10]Roselli, C.E & Resko, J.A., 1984. Androgens regulate brain aromatase activity in adult male rats through<br />

a receptor mechanism. Endocrinology 114(6):2183-9.<br />

105


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

INHIBITION OF 3α,5α-THP FORMATION DECREASES<br />

EXPLORATORY/ANTI-ANXIETY AND SOCIO-SEXUAL BEHAVIOR IN<br />

SEXUALLY-RECEPTIVE FEMALE RATS<br />

Paris, J.J. 1 , Rhodes, M.E. 1 , Frye, C.A. 1-4<br />

Dept. of Psychology 1 , Biological Sciences 2 , and The Centers for Neuroscience 3 and Life<br />

Sciences 4 Resesarch; The University at Albany-SUNY, Life Sciences 01058, 1400<br />

Washington Avenue, Albany, NY USA 12222; paris.j@gmail.com, 518-591-8823<br />

The progesterone (P 4 ) metabolite and neurosteroid, 5α-pregnan-3α-ol-20-one<br />

(3α,5α-ΤΗP) has actions in the midbrain ventral tegmental area (VTA) to modulate the<br />

intensity and duration of lordosis [3]. Other studies have demonstrated that 3α,5α-THP can<br />

also produce anti-anxiety effects through actions in the hippocampus [9,10]. 3α,5α-THP<br />

administration to VTA or hippocampus increases time spent on the open arms of an<br />

elevated plus maze, latency to bury in response to a shock, and time spent interacting with<br />

a novel conspecific [2,5,6].<br />

3α,5α-THP is increased rapidly in brain in response to stressful stimuli to levels<br />

that produce agonist-like effects at GABA A receptors and dampen hypothalamic-pituitaryadrenal<br />

activity [1]. Administration of 3α,5α-THP in conjunction with an acute stressor<br />

decreases the adrenocorticotropic hormone response and can attenuate overexpression of<br />

corticotrophin releasing hormone mRNA following adrenalectomy [7,8]. 3α,5α-THP<br />

increases in midbrain, hippocampus, diencephalon, and cortex in response to reproductive<br />

behaviors (lordosis, proceptivity, aggression, pacing of sexual contacts) [4]. How blockade<br />

of 3α,5α-THP formation in midbrain VTA would effect exploratory/anti-anxiety and sociosexual<br />

behaviors of sexually-receptive female rats was investigated.<br />

Female, Long-Evans rats (n=33) were implanted with bilateral guide cannulae<br />

aimed at the VTA and were behaviorally screened for sexual receptivity daily. Rats that<br />

demonstrated lordosis in response to male mounting were considered to be in behavioral<br />

estrus. Sexually-receptive rats received infusions (1µg) of either a mitochondrial<br />

benzodiazepine receptor (MBR) blocker (PK11195), which acts by inhibiting translocation<br />

of neurosteroid precursor across the MBR membrane, a 3β-hydroxysteroid oxidoreductase<br />

inhibitor (indomethacin) that blocks P 4 metabolism to 3α,5α-ΤΗP, a combination of both<br />

inhibitors, or an infusion of β-cyclodextrin vehicle to VTA.<br />

As shown in Figure 1, rats infused with PK11195, indomethacin, or both to the<br />

VTA had significantly reduced midbrain 3α,5α-ΤΗP levels compared to vehicle-infused<br />

rats. Infusions of inhibitor, compared to infusions of vehicle, decreased exploratory/antianxiety<br />

behavior, indicated by a significant reduction in central entries in an open field and<br />

time spent on the open arms of an elevated plus maze, reduced social responding in a<br />

partner preference task and significantly decreased social interaction with a conspecific.<br />

Likewise, inhibitor infusions attenuated reproductive behaviors in a paced mating<br />

paradigm compared to rats infused with vehicle. Inhibition of metabolism of P 4 to 3α,5α-<br />

THP (which occurs both centrally and peripherally with indomethacin administration) was<br />

observed to attenuate reproductively-relevant behaviors to a greater degree than did<br />

inhibition of central production alone (PK11195). The greatest decrements in behavior<br />

were observed when rats received a combination of both inhibitors. These findings suggest<br />

that formation of 3α,5α-ΤΗP in midbrain is essential for increased exploration/anti-anxiety<br />

and socio-sexual behaviors associated with sexual receptivity and that expression of these<br />

behaviors may be dependent on both de novo production of 3α,5α-ΤΗP and metabolism<br />

from ovarian sources.<br />

106


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

Supported by: National Institute of Mental Health (MH06769801).<br />

Figure 1: Rats receiving VTA infusions of PK11195 (n=9), indomethacin (n=8), or both<br />

(n=8) demonstrate decreased exploratory/anti-anxiety and socio-sexual behavior compared<br />

to rats receiving vehicle infusions (n=8), * indicates p


WEDNESDAY, 21 th February 2007<br />

8.30 - 12.00<br />

Symposium:<br />

Effects mediated by membrane receptors


Symposium:<br />

Effects mediated by membrane receptors<br />

(Chairs: Melcangi R.C., Schumacher M.)<br />

• Micevych P., Hariri O., Soma K., Sinchak K. (USA) Neuroprogesterone: essential<br />

trigger for estrogen positive feedback?<br />

• Belcher S.M., Le H.H., Spurling L., Zsarnovszky A. (USA) Integrated signaling<br />

mechanisms of estrogen in developing neurons and neuroectodermal derived tumors<br />

• Guennoun R., Meffre D, Labombarda F, Gonzalez S.L, Gonzalez Deniselle M.C,<br />

Stein DG, De Nicola AF, Schumacher M (France) The membrane-associated<br />

progesterone-binding protein 25-Dx: expression, cellular localisation and upregulation<br />

after brain and spinal cord injuries<br />

• Valenzuela C.F., Mameli M., Carta M., Zamudio, P.A. (USA) Modulation of<br />

glutamatergic synaptic transmission by neurosteroids<br />

• Biggio G., Mostallino MC, Pisu MG, Talani G, Carta M, Sanna E, Serra M<br />

(Italy) Neurosteroid responses and GABA A receptor plasticity during chronic stress<br />

• Cambiasso M.J., Gorosito S.V. (Argentina) Axogenic effect of oestrogen in male<br />

rat hypothalamic neurons involves Ca 2+ , PKC and ERK signaling<br />

• Nyberg S., Bäckström T., Zingmark E., Sundström Poromaa I. (Sweden)<br />

Allopregnanolone decrease with symptom improvement during placebo and GnRH<br />

agonist treatment in women with premenstrual dysphoric disorder


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

NEUROPROGESTERONE: ESSENTIAL TRIGGER FOR ESTROGEN POSITIVE<br />

FEEDBACK?<br />

Micevych P. 1* , Hariri O. 1 , Soma K. 2 , and Sinchak K. 3<br />

1* Department of Neurobiology, David Geffen School of Medicine; Laboratory of<br />

Neuroendocrinology, Brain Research Institute at UCLA, 10833 LeConte Ave, Los<br />

Angeles, CA 90095-1763; pmicevych@mednet.ucla.edu, Tel: 310.825.2224, FAX<br />

310.825-2224. 2 Depts. of Psychology & Zoology, University of British Columbia,<br />

Vancouver, BC CANADA. 3 Dept of Biological Sciences, California State University, ,<br />

Long Beach CA.<br />

In the gonadally intact rat, the surge of luteinizing hormone (LH) needed for<br />

ovulation, reproductive behaviors and implantation of fertilized ova, is dependent on<br />

estradiol and progesterone. Estradiol of ovarian origin induces progesterone receptors in<br />

the preoptic area and hypothalamus. Sequential activation of estrogen and progesterone<br />

receptors coordinates reproductive physiology and behavior. In ovariectomized and<br />

adrenalectomized (ovx/adx) rats, exogenous administration of estradiol alone is sufficient<br />

to initiate an LH surge suggesting that an endogenous source of progesterone remains in<br />

these animals. This is idea is supported by the observation that in ovx/adx rats,<br />

progesterone levels in the hypothalamus increase prior to the LH surge and inhibition of<br />

progesterone synthesis prevents the LH surge. These results suggest that the increase in<br />

pre-surge progesterone levels in the hypothalamus is a necessary aspect of the estrogen<br />

positive feedback mechanism that initiates the LH surge. Estradiol increases the expression<br />

of the progesterone synthesizing enzyme 3β-hydroxysteroid dehydrogenase (3β-HSD) in<br />

the hypothalamus. Previous studies had indicated that astrocytes may be the most active<br />

progesterone synthesizing cells in the CNS. Experiments with neonatal cortical and postpubertal<br />

hypothalamic astrocytes in vitro demonstrated that estradiol stimulates<br />

progesterone synthesis. Estradiol stimulates progesterone synthesis through rapidly<br />

increasing in free cytoplasmic calcium ([Ca 2+ ] i ). Astrocytes were demonstrated to express<br />

both membrane and nuclear estrogen receptors (ER), but the effects on [Ca 2+ ] i were<br />

consistent with the activation of a membrane ER in terms of the time course of the<br />

response (seconds) and its activation with a membrane impermeable estradiol construct (E-<br />

6-BSA). Blockade of phospholipase C (PLC) or the inositol trisphosphate (IP 3 ) receptor<br />

prevented the estradiol-induced [Ca 2+ ] i transients, indicating that intracellular stores of<br />

Ca 2+ were mobilized. In neurons, estradiol stimulation of the PLC - [Ca 2+ ] i pathway is<br />

dependent on type 1a metabotropic glutamate receptors (mGluR1a). Similarly, in<br />

astrocytes, blockade of the mGluR1a, prevented estradiol-induced [Ca 2+ ] i transients.<br />

Thapsigargin, a sesquiterpene lactone, which releases IP 3 receptor-sensitive Ca 2+ stores<br />

rapidly increased [Ca 2+ ] i levels and mimicked the effect of estradiol treatment – increasing<br />

[Ca 2+ ] i and progesterone synthesis in astrocytes. One hour treatment with thapsigargin was<br />

as effective as estradiol at increasing progesterone levels in astrocyte cultures suggesting<br />

that estradiol induces progesterone synthesis in the hypothalamus by acting on membrane<br />

ER that act through mGluR1a to activate the PLC-IP 3 pathway stimulating the synthesis of<br />

progesterone. The steroid signals regulating reproductive behavior and ovulation are the<br />

same, estradiol and progesterone. To determine whether blocking neuroprogesterone<br />

synthesis prevented the display of proceptive and receptive sexual behaviors, ovx/adx rats<br />

were treated with aminoglutethimide (AGT), a blocker of P450side chain cleavage enzyme<br />

the first enzymatic step of steroidogenesis. Proceptive behaviors induced by estradiol were<br />

blocked in the AGT treated animals, but sexual receptivity was unaffected. Further, AGT<br />

111


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

infused into the III ventricle of gonadally and adrenally intact, cycling rats disrupted the<br />

cycle apparently by preventing the LH surge and ovulation. However, circulating levels of<br />

estradiol were unaffected by central administration of AGT, indicating that AGT did not<br />

affect peripheral steroidogenesis.<br />

In summary, the estrogen positive feedback mechanism involves estradiol acting on<br />

a membrane ER that requires mGluR1a signaling to increase [Ca 2+ ] i levels and induce<br />

progesterone synthesis in astrocytes. This neuroprogesterone acts on estradiol-induced<br />

progesterone receptors to initiate the cascade of events culminating in the surge release of<br />

LH and facilitation of proceptive behaviors. However, neither progesterone nor<br />

progesterone receptors are needed when lordosis is activated by estradiol alone, suggesting<br />

that estradiol + progesterone activate different circuits regulating sexual receptivity than<br />

estradiol alone.<br />

112


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

INTEGRATED SIGNALING MECHANISMS OF ESTROGEN IN DEVELOPING<br />

NEURONS AND NEUROECTODERMAL DERIVED TUMORS<br />

Belcher S.M., Le H.H., Spurling L., and Zsarnovszky A.<br />

University of Cincinnati College of Medicine, Department of Pharmacology and Cell<br />

Biophysics. 231 Albert Sabin Way, Cincinnati, OH USA 45267-0575. e-mail:<br />

scott.belcher@uc.edu; Fax: (513) 558-4329<br />

The steroid hormone 17β-Estradiol (E2) regulates the normal function and<br />

development of the mammalian nervous system. Many of estradiol’s effects are mediated<br />

via the nuclear hormone estrogen receptor (ER) α and ERβ. In addition to regulating<br />

estrogen-responsive gene expression, E2 also acts in an immediate and cell-specific<br />

fashion to regulate various intracellular signal transduction pathways. Whereas<br />

mechanisms underlying E2-responsive gene expression are fairly well understood, the<br />

signaling mechanisms of rapid estradiol-mediated signal transduction are less clear. In<br />

vitro analysis employing neonatal rat cerebellar granule cell culture, a model of nonreproductive<br />

actions of E2 on neuronal development, was used to investigate the signaling<br />

mechanisms active in developing and mature populations of cerebellar neurons. We<br />

demonstrated that rapid E2 activation of extracellular signal regulated kinase (ERK)<br />

signaling regulates oncotic, but not apoptotic programmed cell death in a subpopulation of<br />

sensitive granule cell precursors. Furthermore, it was demonstrated that exposure to low<br />

concentrations of E2 and/or the xenoestrogen bisphenol A, transiently modulates granule<br />

cell ERK signaling in vitro and in vivo. Pharmacological inhibition of Gαi- or protein<br />

kinase A (PKA)-mediated signaling blocked E2- and ICI182,780-mediated ERK activation<br />

in vitro, revealed a requirement for Gαi and PKA signaling. Moreover, E2 exposure<br />

following pretreatment with pertussis toxin or H-89 depressed basal ERK levels,<br />

suggesting that E2 also rapidly activates a Gαi- and PKA-independent phosphatase<br />

activity. Additional studies revealed that E2 specifically stimulates protein phosphatase 2A<br />

(PP2A) activity by an independent rapid intracellular mechanism. This first demonstration<br />

of a rapid effect on protein phosphatase activity could account for the transient nature of<br />

E2-induced ERK activation in these neurons. Additional molecular and pharmacological<br />

inhibitor studies revealed that activation of c-Src, but not the activation of the epidermal<br />

growth factor receptor (EGFR) is required for E2-mediated ERK signaling. The results<br />

presented reveal a distinctive cell specific mechanism for rapid E2-induced modulation of<br />

ERK signaling and a concerted activation of PP2A.<br />

Because medulloblastomas (MD), the most common malignant brain tumor in<br />

children, arise from cerebellar granule cell-like precursors we investigated the impact of<br />

rapid and nuclear hormone receptor mediated actions of E2 on these invasive<br />

neuroectodermal tumors. As a result of their shared developmental lineage, we<br />

hypothesized that like normal granule cell precursors, malignant MD cells would express<br />

ERβ and their growth or migration would be estrogen-responsive. To test this hypothesis<br />

immunohistochemical studies were done to determine whether ERs were expressed in<br />

human MD tumors from both male and female patients. We found evidence of ERβ<br />

expression in all medulloblastoma tumors analyzed. Highly focal expression of ERα was<br />

detected in 40% of the tumors. Western blot analysis was also used to determine whether<br />

ERs are expressed in a cerebrocortical primitive neuroectodermal tumor (PNET) cell line<br />

(PFSK1), and two MD-derived cell lines (Daoy; D283Med) representing the “glial” and<br />

“neuronal” phenotypic pro<strong>file</strong>s of MD. Results of those studies indicated that specific<br />

113


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

isoforms of ERβ- and low levels of ERα-like proteins are expressed in each cell line. The<br />

physiological significance of the expressed ERs was also examined; physiological<br />

concentrations of E2 stimulated Rho-dependant migration in each MD/PNET line. The<br />

growth of D283Med cells, but not the growth of Daoy or PFSK1 cells, was also stimulated<br />

by E2. In all cases, the ER-antagonist ICI182,780 blocked E2-induced stimulation of<br />

growth and migration. Xenograft studies in nude mice have confirmed that the growth of<br />

medulloblastoma cells is markedly stimulated by the estradiol. These in vitro and in vivo<br />

studies reveal that ERs contribute to the growth and invasive phenotypes of some PNET<br />

and MD tumors, and demonstrate that MD are estrogen-responsive tumors.<br />

114


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

THE MEMBRANE–ASSOCIATED PROGESTERONE BINDING PROTEIN 25-<br />

DX: EXPRESSION, CELLULAR LOCALISATION AND UP-REGULATION<br />

AFTER BRAIN AND SPINAL CORD INJURIES<br />

Guennoun R. 1* , Meffre D. 1 , Labombarda F. 2 , Gonzalez S.L. 2 , Gonzalez Deniselle<br />

M.C. 2 , Stein DG 3 , De Nicola A.F 2 , and Schumacher M. 1<br />

(1) UMR788 Inserm and University Paris 11, Kremlin-Bicêtre, France;<br />

(2) Instituto de Biologia y Medicina Experimental and Department of Human<br />

Biochemistry, University of Buenos Aires, Argentina<br />

(3) Department of Emergency Medicine, Emory University, Atlanta, USA<br />

Phone : 33 1 49 18 95 Fax : 33 1 45 21 19 40 e-mail : Rachida.Guennoun@kb.inserm.fr<br />

Progesterone has neuroprotective effects in the injured and diseased spinal cord and after<br />

traumatic brain injury (TBI). In addition to intracellular progesterone receptors, membrane<br />

binding sites of progesterone may be involved in neuroprotection. A first putative<br />

membrane receptor of progesterone, distinct from the classical intracellular PR isoforms,<br />

with a single membrane spanning domain, has been first cloned from porcine liver.<br />

Homologous proteins were cloned in rat (named 25-Dx), cattle and humans. We shall refer<br />

to this progesterone binding protein as 25-Dx, to distinguish it from the progesterone<br />

membrane receptors (mPRs) which have later been cloned. The distribution and regulation<br />

of 25-Dx in the nervous system may provide some clues concerning its functions.<br />

In brain, 25-Dx is present in the microsomal and mitochondrial fractions. 25-Dx is<br />

particularly abundant in the hypothalamic area, circumventricular organs, ependymal cells<br />

of the ventricular walls, and in the meninges. 25-Dx is co-expressed with vasopressin in<br />

neurones of the paraventricular, supraoptic and retrochiasmatic nuclei. In spinal cord, 25-<br />

Dx is localised in cell membranes of dorsal horn and central canal neurones. In response to<br />

TBI, 25-Dx expression was up-regulated in neurones and induced in astrocytes. The<br />

expression of 25-Dx in structures involved in cerebro spinal fluid production and in<br />

osmoregulation, and its up-regulation after brain damage, point to a potentially important<br />

role of this progesterone-binding protein in the maintenance of water homeostasis after<br />

TBI. A role of 25-Dx in mediating protective effects of progesterone in the spinal cord is<br />

supported by the observation that its mRNA and protein are up-regulated by progesterone<br />

in dorsal horn of the injured spinal cord. On the contrary, the classical PR was downregulated<br />

under these conditions. Our observations point to the possibility that<br />

progesterone actions may involve different signalling mechanisms depending on the<br />

pathophysiological context and that 25-Dx may be involved in the neuroprotective effect of<br />

progesterone in the injured brain and spinal cord.<br />

115


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

MOUDULATION OF GLUTAMATERGIC TRANSMISSION BY<br />

NEUROSTEROIDS<br />

Valenzuela C.F., Mameli M., Carta M., Zamudio, P.A.<br />

Department of Neurosciences, University of New Mexico School of Medicine,<br />

MSC08 4740, Albuquerque, NM, U.S.A.<br />

Fax 1(505) 272-8082 e-mail: FValenzuela@salud.unm.edu<br />

A major question in neuroscience is how neuronal circuits develop. Studies have<br />

demonstrated that neuronal connections undergo profound remodeling after their initial<br />

formation and that this is mediated by activity-dependent synaptic plasticity. Despite the<br />

developmental importance of this process, its cellular mechanisms in immature neurons are<br />

not fully understood. We found that the excitatory neurosteroid pregnenolone sulfate<br />

(PREGS) induces long-term potentiation (LTP) of AMPA receptor (AMPAR)-mediated<br />

postsynaptic currents in hippocampal neurons during a restricted developmental period [3].<br />

We recorded AMPAR miniature excitatory postsynaptic currents (mEPSCs) in the wholecell<br />

patch-clamp configuration from CA1 pyramidal neurons in hippocampal slices from<br />

postnatal day (P) 3-4 rats. Brief (5 min) exposure to 25 µM PREGS, induced LTP of in the<br />

frequency but not the amplitude of these events. This effect was robust in slices from P3-4<br />

rats and gradually decreased at P5 to become undetectable by P6. We evaluated the timecourse<br />

of the PREGS effect on the paired-pulse ratio (PPR) of AMPAR EPSCs evoked by<br />

stimulation of Schaffer collaterals. Application of PREGS significantly decreased the<br />

PPR. However, this effect was transient and EPSC amplitude continued to increase even<br />

after the PPR had returned to baseline. We next evoked single AMPA EPSCs using a<br />

stimulation intensity that yielded a mixture of synaptic transmission successes and failures.<br />

A few minutes after application of PREGS, but not vehicle, we could only record<br />

successes, consistent with an increase in glutamate release probability. At later time points<br />

(~20 min after PREGS application), EPSC amplitude gradually increased and remained<br />

elevated. Currents evoked by exogenous AMPA were dramatically increased ~20 min<br />

after brief exposure to PREGS. Based on these results, we conclude that PREGS induces a<br />

transient increase in glutamate release at the presynaptic level (early phase) that triggers<br />

LTP of postsynaptic AMPAR function (late phase).<br />

Studies have demonstrated that LTP of AMPA receptor-mediated responses<br />

involves changes in phosphorylation and/or trafficking of this receptor that are triggered by<br />

Ca 2+ influx via postsynaptic NMDARs. Therefore, we tested if postsynaptic NMDARmediated<br />

Ca 2+ influx was required for the late phase of PREGS-induced plasticity. We<br />

found that this phase could not be observed in neurons intracellularly dialyzed with the<br />

Ca 2+ chelator BAPTA or the NMDAR blocker MK-801. Bath application of ifenprodil, an<br />

antagonist of NR2B-containing receptors, abolished the late phase of the PREGS effect.<br />

Collectively, these findings indicate that postsynaptic NR2B-subunit containing NMDARs<br />

mediate the late phase of PREGS LTP. We are currently further characterizing these<br />

postsynaptic effects of PREGS.<br />

We next investigated the presynaptic mechanism of action of PREGS. First, we<br />

used BAPTA-AM to chelate Ca 2+ at both the presynaptic and postsynaptic levels. In the<br />

presence of this agent, both the early and late phases of PREGS LTP were abolished.<br />

Second, we assessed the effects of bath application of NMDAR antagonists. The nonselective<br />

antagonist, DL-APV, also blocked the two phases. The antagonist of the glycine<br />

co-agonist binding site, 7-chlorokynurenate, had a similar effect. Importantly, both phases<br />

of the PREGS effect were eliminated by low concentrations of PPDA, which selectively<br />

116


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

antagonizes receptors containing NR2D subunits [2]. These findings indicate that a<br />

presynaptic elevation in [Ca 2+ ] i is required for the early phase of the PREGS-induced LTP<br />

and that this rise is mediated by NR2D-containing presynaptic receptors. The presence of<br />

this receptor in presynaptic terminals was confirmed by bath application of the<br />

nonselective NMDA agonist homoquinolinic acid. Importantly, homoquinolinic acid did<br />

not increase mEPSC frequency in slices from P6-10 animals, indicating that functional<br />

expression of these presynaptic NMDARs is developmentally restricted. The role of<br />

presynaptic NR2D-containing NMDARs in the mechanism of action of PREGS is<br />

currently being further investigated.<br />

Studies have demonstrated that depolarization of immature neurons can increase the<br />

probability of glutamate release at presynaptic terminals. We investigated whether<br />

depolarization could trigger the release of endogenous PREGS. Depolarization (15 sec)<br />

produced a long-lasting increase in mEPSC frequency in slices from P3-4 but not P6-10<br />

rats. Importantly, pre-incubation of slices with PPDA or rabbit anti-PREGS polyclonal<br />

antibodies, but not rabbit IgG, blocked this effect. These findings suggest that a PREGSlike<br />

neurosteroid is an endogenous factor involved in synapse maturation, acting in a<br />

retrograde manner like other lipids such as endocannabinoids and arachidonic acid.<br />

On a related study, we examined the role of PREGS in the developmental actions of<br />

alcohol [4]. We previously showed that chronic prenatal ethanol exposure increases<br />

PREGS levels in the developing brain [1]. We discovered that 50 mM ethanol strengthens<br />

AMPAR-mediated transmission in the CA1 region in a PREGS-like manner. This effect of<br />

ethanol is age-dependent and blocked by application of an anti-PREGS antibody<br />

scavenger. These data suggest that the deleterious effects of ethanol on hippocampal<br />

development are mediated, in part, by alterations in the production and/or release of a<br />

PREGS-like neurosteroid, which might result in premature stabilization of excitatory<br />

synaptic connections.<br />

Supported by grants from the National Institute of Mental Health and the National Institute<br />

of Alcohol Abuse and Alcoholism.<br />

References list<br />

[1] J.C. Caldeira, Y. Wu, M. Mameli, R.H. Purdy, P.K. Li, Y. Akwa, D.D. Savage, J.R.<br />

Engen, C.F. Valenzuela, Fetal alcohol exposure alters neurosteroid levels in the<br />

developing rat brain, J Neurochem 90 (2004) 1530-1539.<br />

[2] B. Feng, H.W. Tse, D.A. Skifter, R. Morley, D.E. Jane, D.T. Monaghan, Structureactivity<br />

analysis of a novel NR2C/NR2D-preferring NMDA receptor antagonist: 1-<br />

(phenanthrene-2-carbonyl) piperazine-2,3-dicarboxylic acid, Br J Pharmacol 141<br />

(2004) 508-516.<br />

[3] M. Mameli, M. Carta, L.D. Partridge, C.F. Valenzuela, Neurosteroid-induced<br />

plasticity of immature synapses via retrograde modulation of presynaptic NMDA<br />

receptors, J Neurosci 25 (2005) 2285-2294.<br />

[4] M. Mameli, C.F. Valenzuela, Alcohol increases efficacy of immature synapses in a<br />

neurosteroid-dependent manner, Eur J Neurosci 23 (2006) 8<strong>35</strong>-839.<br />

117


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

NEUROSTEROID RESPONSES AND GABA A RECEPTOR PLASTICITY<br />

DURING CHRONIC STRESS<br />

Biggio G. 1,2 , Mostallino M.C. 2 , Pisu M.G. 2 , Talani G. 1 , Carta M. 1 , Sanna E. 1 and<br />

Serra M. 1,2<br />

1 Department of Experimental Biology, Center of Excellence for Neurobiology of Drug<br />

Dependence, University of Cagliari, Cagliari, Italy. 2 C.N.R. Institute of Neuroscience,<br />

Italy<br />

Rats deprived of social contact with other rats at a young age experience a form of<br />

prolonged stress that leads to long-lasting alterations in their behavioral pro<strong>file</strong> This<br />

chronic stress paradigm is thus thought to be anxiogenic for these normally gregarious<br />

animals and their abnormal reactivity to environmental stimuli, when reared under this<br />

condition, is thought to be a product of prolonged stress.<br />

We have previously demonstrated that social isolation of rats immediately after<br />

weaning is associated to a reduction in the cerebrocortical and plasma concentrations of<br />

progesterone and its metabolites 3α,5α-TH PROG and 3α,5α-THDOC (Serra et al., 2000).<br />

Moreover, although we found that the basal plasma concentration of adrenocorticotropic<br />

hormone in isolated rats was slightly decreased compared with that in group-housed<br />

animals, the functional response of the hypothalamic-pituitary-adrenal axis HPA axis to an<br />

acute stressful stimulus (foot shock), or to an acute injection of ethanol or isoniazid is<br />

markedly increased in isolated rats. We found that foot shock, used as a novel acute<br />

stressor, or acute intraperitoneally injection of isoniazid or ethanol, increased the<br />

cerebrocortical and plasma concentrations of progesterone, 3α,5α-TH PROG, and 3α,5α-<br />

TH DOC by a greater percentage in isolated rats than in group-housed animals (Serra et al.,<br />

2000; 2003). Moreover, acute administration of ethanol in socially isolated rats increases<br />

the concentration of 3α,5α-TH PROG to a substantially greater degree in the cerebral<br />

cortex than in plasma. This finding is consistent with our observation that ethanol increases<br />

local neurosteroid synthesis in the brain independently of the HPA axis (Sanna et al.,<br />

2004). Social isolation modified the effects of ethanol on the amounts of steroidogenic<br />

acute regulatory protein (StAR) mRNA and protein in the brain. The increased sensitivity<br />

of neuroactive steroid production to ethanol in socially isolated rats may thus be due in part<br />

to an increase in the expression of StAR in brain mitochondria. Ethanol also increased the<br />

amplitude of GABA A receptor–mediated miniature inhibitory postsynaptic currents<br />

(mIPSC) recorded from CA1 pyramidal neurons with a greater potency in hippocampal<br />

slices prepared from socially isolated rats than in those from group-housed. Indeed,<br />

whereas ethanol at a concentration of 50 mM significantly increased mIPSC amplitude in<br />

neurons from isolated animals, it proved ineffective in those from group-housed rats. The<br />

effect of ethanol on mIPSC amplitude was inhibited by finasteride supporting the idea that<br />

this action is mediated by an increased production of 3α,5α-TH PROG. This conclusion is<br />

supported by the lack of a difference in the effect of exogenous 3α,5α-TH PROG at<br />

concentrations of 1 and 3 µM, on mIPSC amplitude in CA1 pyramidal neurons between<br />

isolated and group-housed rats. Behavioral studies have also indicated that the ability of<br />

ethanol to inhibit isoniazid-induced convulsions is greater in isolated rats than in grouphoused<br />

animals and this effect of isolation is prevented by treatment with the 5α-reductase<br />

inhibitor finasteride. These observations and the finding that social isolation rats are more<br />

sensitive to the effects of ethanol on the brain concentrations of 3α,5α-TH PROG and<br />

3α,5α-TH DOC, both of which posses anxiolytic properties and potentiate the central<br />

118


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

actions of ethanol, suggest that chronic stress may induce plastic adaptation of neuronal<br />

systems that contributes to a vulnerability to alcohol abuse.<br />

The selective gene expression of GABA A receptor and its subsequent subunit<br />

composition is affected by chronic stress. In socially isolated rats the level of α 2 α 4 and δ<br />

subunits immunoreactivity was found to be increased, while the amounts of α 1 and γ 2 were<br />

decreased, throughout the hippocampus compared with that apparent in group-housed rats.<br />

Given that δ substitutes for γ 2 and that the latter subunit is essential for synaptic<br />

localization of GABA A receptors, those containing α 4 and δ would be expected to be<br />

extrasynaptic GABA A receptors responsible for tonic inhibition. Accordingly, we found<br />

that the amplitude of GABA A receptor–mediated tonic inhibitory currents in granule cells<br />

of the dentate gyrus was markedly greater in hippocampal slices from socially isolated rats<br />

than in those from group-housed animals. The reduction in tonic current noise induced by<br />

bath application of the GABA A receptor antagonist bicuculline (20 µM) was thus<br />

significantly greater in hippocampal slices than in those from group-housed animals. In<br />

addition, the enhancement of tonic current noise induced by 3α,5α-TH PROG (3 µM) was<br />

markedly greater in granule cells of the dentate gyrus from isolated rats than in those from<br />

group-housed animals. Since enhanced tonic inhibition is an effective means of reduced<br />

neuronal excitability, our data therefore suggest that the augmented tonic inhibitory current<br />

mediated by δ subunit–containing extrasynaptic GABA A receptors in the hippocampus of<br />

socially isolated rats might reflect a compensatory mechanism to counteract the increased<br />

seizure susceptibility due to the increased expression of the α4 subunit.<br />

Neurochemical, molecular and electrophysiological evidence demonstrate that<br />

social isolation is associated with alteration in the structure and function of GABA A<br />

receptors and suggest that endogenous levels of progesterone metabolites such as 3α,5α-<br />

TH PROG may be an important determinant in regulating brain excitability and sensitivity<br />

to stimuli and point out their possible role in psychiatric and neurological disorder.<br />

References list<br />

[1] Sanna, E., Talani, G., Busonero, F., Pisu, M.G., Purdy, R.H., Serra, M. and Biggio,<br />

G. (2004) Brain steroidogenesis mediates ethanol modulation of GABA A receptor<br />

activity in rat hippocampus. J. Neurosci. 24, 6521–6530.<br />

[2] Serra, M., Pisu, M.G., Floris, I., Cara, V., Purdy, R.H. and Biggio, G. (2003) Social<br />

isolation-induced increase in the sensitivity of rats to the steroidogenic effect of<br />

ethanol. J. Neurochem. 85, 257–263.<br />

[3] Serra, M., Pisu, M.G., Littera, M., Papi, G., Sanna, E., Tuveri, F., Usala, L., Purdy,<br />

R.H. and Biggio, G. (2000) Social isolation-induced decreases in both the<br />

abundance of neuroactive steroids and GABA A receptor function in rat brain. J.<br />

Neurochem. 75, 732–740.<br />

119


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

AXOGENIC EFFECT OF OESTROGEN IN MALE RAT HYPOTHALAMIC<br />

NEURONS INVOLVES Ca 2+ , PKC AND ERK SIGNALING<br />

Cambiasso M.J., and Gorosito S.V.<br />

Instituto de Investigación Médica M y M Ferreyra (INIMEC-CONICET), Friuli 2434<br />

(5016) Córdoba, Argentina. Fax: +54-<strong>35</strong>1-4695163. E-mail: jcambiasso@immf.uncor.edu<br />

17-β-estradiol (E2) stimulates the growth of axons in male derived hypothalamic<br />

neurons in vitro [3,4]. This effect is not exerted through the classical intracellular<br />

oestrogen receptor (ER) but depends on a membrane mechanism [2] in which the tyrosine<br />

kinase receptor type B participates [1,5]. More recently, we showed that E2 rapidly<br />

induced phosphorylation of the extracellular signal-regulated kinases 1/2 (ERK) mitogenactivated<br />

protein kinases (MAPK) [6]. In the present study we investigate the intracellular<br />

signaling cascades that mediate the axogenic effect of E2. Treatment with an intracellular<br />

Ca 2+ chelator, a Ca 2+ -dependent PKC inhibitor, or two specific inhibitors of MEK-ERK<br />

pathway completely inhibited the E2-induced axogenesis. E2 and the membrane<br />

impermeant construct E2BSA rapidly induced phosphorylation of ERK, which was<br />

blocked by the specific inhibitor of ERK pathway UO126 but not by the ER antagonist ICI<br />

182,780. Decrease of intracellular free Ca 2+ or disruption of PKC activation by Ro 32-0432<br />

attenuate ERK activation, indicating the confluence of signals in the MAPK pathway. Subcellular<br />

analysis of ERK demonstrated that the phospho-ERK signal is transduced to the<br />

nucleus by E2. We also have shown that E2 increased phoshorylation of CREB via ERK<br />

signaling. In summary, this study demonstrates that E2 induces axogenesis in malehypothalamic<br />

neurons through activation of the calcium-dependent PKC and ERK pathway<br />

leading to increased CREB phosphorylation, events required to induce axon elongation.<br />

Research is supported by CONICET (PIP6238) and FONCyT (PICT26331).<br />

IBRO Travel Grant is gratefully acknowledged.<br />

Reference list<br />

[1] Brito, V.I., Cambiasso, M.J., Carrer, H.F., 2004. Inhibition of TrkB synthesis blocks axogenic effect of<br />

estradiol on hypothalamic neurons in vitro. Eur. J. Neurosci. 20, 331-337.<br />

[2] Cambiasso, M.J., Carrer, H.F., 2001. Nongenomic mechanism mediates estradiol stimulation of axon<br />

growth in male rat hypothalamic neurons in vitro. J. Neurosci. Res. 66, 475-481.<br />

[3] Cambiasso, M.J., Colombo, J.A., Carrer, H.F., 2000. Differential effect of oestradiol and astrogliaconditioned<br />

media on the growth of hypothalamic neurons from male and female rat brains. Eur. J.<br />

Neurosci. 12, 2291-2298.<br />

[4] Cambiasso, M.J., Diaz, H., Caceres, A., Carrer, H.F., 1995. Neuritogenic effect of estradiol on rat<br />

ventromedial hypothalamic neurons co-cultured with homotopic or heterotopic glia. J. Neurosci. Res.,<br />

42, 700-709.<br />

[5] Carrer, H.F., Cambiasso, M.J., Brito, V.I., Gorosito, S.V., 2003. Neurotrophic factors and estradiol<br />

interact to control axogenic growth in hypothalamic neurons. Ann. NY Acad. Sci. 1007, 306-316.<br />

[6] Carrer, H.F., Cambiasso, M.J., Gorosito, S., 2005. Effects of estrogen on neuronal growth and<br />

differentiation. J.Steroid Biochem. Mol. Biol., 93, 319-323.<br />

120


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

ALLOPREGNANOLONE DECREASE WITH SYMPTOM IMPROVEMENT<br />

DURING PLACEBO AND GnRH AGONIST TREATMENT IN WOMEN WITH<br />

PREMENSTRUAL DYSPHORIC DISORDER<br />

Nyberg S. 1 , Bäckström T. 1 , Zingmark E. 1 , and Sundström Poromaa I. 2<br />

1 Department of Clinical Science, Obstetrics and Gynecology, Umeå University Hospital,<br />

Umeå, Sweden. Fax +46-907852277 e-mail:Sigrid.nyberg@obgyn.umu.se<br />

2 Department of Women’s and Children’s Health, University Hospital, Uppsala, Sweden<br />

Background: Neurosteroids such as allopregnanolone and pregnanolone, are<br />

suggested to be of importance for the pathophysiology of premenstrual dysphoric disorder<br />

(PMDD). The aim of this study was to investigate if the luteal phase serum concentrations<br />

of these neurosteroids are associated with improvement of premenstrual symptoms in 12<br />

women with PMDD after treatment with low dose GnRH-agonist (100 µg buserelin) and<br />

placebo.<br />

Methods: Daily ratings for mood and physical symptoms were made prior to treatment and<br />

throughout the study and serum progesterone, allopregnanolone, and pregnanolone was<br />

assessed in the luteal phase (cycle day -1 to cycle day -9). Based on their symptom ratings<br />

prior and throughout the study, subjects were grouped as either buserelin responders (n =<br />

6) or placebo responders (n = 6).<br />

Results: Buserelin responders displayed decreased levels of allopregnanolone (p < 0.05)<br />

and progesterone (p < 0.05) in parallel with improvement of symptoms during buserelin<br />

treatment. During the placebo treatment, the placebo-responders had lower<br />

allopregnanolone serum concentrations compared to the serum concentrations in buserelin<br />

responders (p < 0.05). This was in parallel with improvement in symptoms compared to<br />

pre-treatment ratings.<br />

Conclusion: Treatment response, whether induced by buserelin or placebo, appears to be<br />

paralleled by a decrease in allopregnanolone concentration.<br />

121


WEDNESDAY, 21 th February 2007<br />

12.00 – 13.00<br />

Plenary lecture:<br />

Swaab D.F. (Netherlands)


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

THE STRESS SYSTEM IN THE HUMAN BRAIN IN DEPRESSION AND<br />

NEURODEGENERATION<br />

Swaab D.F.*, Bao A-M*<br />

*Netherlands Institute for Neuroscience, Meibergdreef 47, 1105 BA Amsterdam, The<br />

Netherlands, fax: 0031-20-6961006;<br />

e-mail: d.f.swaab@nin.knaw.nl, a.m.bao@nin.knaw.nl.<br />

The stress system<br />

The stress response is mediated by the hypothalamo-pituitary-adrenal (HPA) system.<br />

Activity of the corticotropin-releasing hormone (CRH) neurons in the hypothalamic<br />

paraventricular nucleus (PVN) forms the basis of the activity of the HPA axis. The CRH<br />

neurons co-express vasopression (AVP) that potentiates the CRH effects. CRH neurons<br />

project not only to the median eminence but also into brain area where they e.g. regulate<br />

the adrenal innervation of the autonomic system. The CRH neurons induce<br />

adrenocorticotropin (ACTH) release from the pituitary, which subsequently causes cortisol<br />

release from the adrenal cortex. In addition, the hypothalamo-neurohypophysial system is<br />

also involved in the stress response. It releases AVP from the PVN and the supraoptic<br />

nucleus (SON) and oxytocin (OXT) from the PVN via the neurohypophysis into the<br />

bloodstream. The suprachiasmatic nucleus (SCN), the hypothalamic clock, is responsible<br />

for the rhythmic changes of the stress system. The stress system is under the influence of<br />

sex hormones. Sustained hyperactivity of the stress system can cause prolonged<br />

overexposure of the brain to aberrant cortisol levels that are presumed to induce psychiatric<br />

disorders and neuropathology conditions such as hippocampal damage in depression and<br />

Alzheimer’s disease (AD) [1,2].<br />

The stress system in depression<br />

Depression results from an interaction between environmental stress and a<br />

genetic/developmental predisposition that cause a permanent activation of the CRH<br />

neurons of the HPA-axis. Stressful life events such as child abuse and early maternal<br />

separation also form risk factors for later depression and anxiety disorder.<br />

Hypertrophy of the adrenals, decreased glucocorticoid receptor (GR) function, enhanced<br />

adrenal response to ACTH, as well as adrenal and pituitary enlargement is observed in<br />

depressed patients. Both centrally released CRH and increased levels of cortisol contribute<br />

to the signs and symptoms of depression. Symptoms such as decreased food intake,<br />

decreased sexual activity, disturbed sleep and motor behavior and increased anxiety can be<br />

induced in experimental animals by intracerebroventricular injection of CRH. Depression<br />

is also a frequent side effect of glucocorticoid treatment and of the syndrome of Cushing's<br />

disease. Inhibitors of cortisol production such as metyrapone can be clinically effective for<br />

treatment of depression, while the GR antagonist mifepristone (RU486) is effective in<br />

treating psychotic depression. The AVP neurons in the hypothalamic PVN and SON that<br />

are projecting to the neurohypophysis are also activated in depression, which contributes to<br />

the increased release of ACTH from the pituitary. Increased levels of circulating AVP are<br />

also associated with the risk for suicide. The increased activity of OXT neurons may be<br />

related to the eating disorders in depression, since OXT acts as a satiety peptide [2].<br />

There is a clear sex difference in depression: the prevalence, incidence and morbidity risk<br />

is higher in females than in males, which is due to both organizing and activating effects of<br />

125


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

sex hormones on the HPA-axis. In particular the fluctuations of sex hormone levels are<br />

considered to be involved in the etiology of depression, e.g. in the premenstrual period,<br />

ante- and postpartum, during the transition phase to the menopause and by the use of oral<br />

contraceptives. About 40% of the activated CRH neurons in mood disorders co-express<br />

nuclear estrogen receptor (ER)-α in the PVN [3], while estrogen-responsive elements have<br />

been found in the CRH gene promoter region and estrogens stimulate CRH production. An<br />

androgen-responsive element in CRH gene promoter region initiates a suppressing effect<br />

on CRH expression [4].<br />

The decreased activity of the SCN is the basis for the disturbances of circadian and<br />

circannual fluctuations in mood, sleep and hormonal rhythms found in depression.<br />

The stress system in AD<br />

CRH neurons are moderately activated in AD resulting in higher plasma cortisol levels.<br />

The SON and PVN remain largely intact , but the SCN is strongly affected from the<br />

earliest AD stages onwards[5]. Diminishment of sex hormones, i.e. of estrogens in<br />

menopause and of testosterone in aging men are considered to be risk factors for AD.<br />

Hippocampal damage in AD and depression<br />

The hippocampus is strongly affected in AD. Also the sensitivity for estrogens is changed<br />

in this disorder because of the changes in the ratio of the various splice variants and<br />

isoforms of the ER-α (see abstract T. Ishunina). Neuronal loss was also reported in the<br />

hippocampus of stressed or corticosteroid-treated rodents and primates. Because of the<br />

inhibitory control of the hippocampus on the HPA-axis, damage to this structure was<br />

expected to disinhibit the HPA-axis, and to cause a positive feedforward cascade of<br />

increasing glucocorticoid levels over time. This ‘glucocorticoid cascade concept of stress<br />

and hippocampal damage’ was hypothesized to be causally involved in an age-related<br />

accumulation of hippocampal damage in disorders like AD and major depression.<br />

However, in postmortem studies we could not find the presumed neuropathological<br />

consequences of steroid overexposure in either depressed patients or in patients treated<br />

with synthetic steroids[6,7]<br />

References list<br />

[1] Swaab DF, Bao AM, Lucassen PJ: The stress system in the human brain in depression and<br />

neurodegeneration. Ageing Res Rev 2005;4:141-194.<br />

[2] Swaab DF: The human hypothalamus. Basic and clinical aspects. Part II:. Handbook of clinical<br />

neurology. Amsterdam, Elsevier, 2004.<br />

[3] Bao AM, Hestiantoro A, Van Someren EJ, Swaab DF, Zhou JN: Colocalization of corticotropinreleasing<br />

hormone and oestrogen receptor-alpha in the paraventricular nucleus of the hypothalamus in<br />

mood disorders. Brain 2005;128:1301-1313.<br />

[4] Bao AM, Fischer DF, Wu YH, Hol EM, Balesar R, Unmehopa UA, Zhou JN, Swaab DF: A direct<br />

androgenic involvement in the expression of human corticotropin-releasing hormone. Mol Psychiatry<br />

2006; 11:567-576.<br />

[5] Wu YH, Fischer DF, Kalsbeek A, Garidou-Boof ML, van der Vliet J, van Heijningen C, Liu RY, Zhou<br />

JN, Swaab DF: Pineal clock gene oscillation is disturbed in alzheimer's disease, due to functional<br />

disconnection from the "master clock". FASEB J 2006;20:1874-1876.<br />

[6] Lucassen PJ, Muller MB, Holsboer F, Bauer J, Holtrop A, Wouda J, Hoogendijk WJ, De Kloet ER,<br />

Swaab DF: Hippocampal apoptosis in major depression is a minor event and absent from subareas at<br />

risk for glucocorticoid overexposure. Am J Pathol 2001;158:453-468.<br />

[7] Muller MB, Lucassen PJ, Yassouridis A, Hoogendijk WJ, Holsboer F, Swaab DF: Neither major<br />

depression nor glucocorticoid treatment affects the cellular integrity of the human hippocampus. Eur J<br />

Neurosci 2001; 14: 1603-1612.<br />

126


WEDNESDAY, 21 th February 2007<br />

15.00 – 18.00<br />

Symposium:<br />

Corticosteroid effects and stress


Symposium:<br />

Corticosteroid effects and stress<br />

(Chairs: Riva M.A., Swaab D.F.)<br />

• Sousa N (Portugal) Corticosteroid receptors and neuroplasticity<br />

• Pryce CR (Switzerland) Postnatal ontogeny of hippocampal expression of the<br />

mineralocorticoid and glucocorticoid receptors in the common marmoset monkey<br />

• Gass P (Germany) Mice with compromised glucocorticoid receptor expression<br />

show behavioural and biochemical features of depression<br />

• Matthews SG (Canada) Maternal adversity, glucocorticoids and programming of<br />

neuroendocrine function and behaviour.<br />

• Darnaudéry M., Morley-Fletcher S. and Maccari S. (France) Long lasting effects<br />

of stress during pregnancy on HPA function and behaviour in mother and offspring<br />

rats


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

CORTICOSTEROID RECEPTORS AND NEUROPLASTICITY<br />

Sousa N.<br />

Life and Health Science Research Institute, University of Minho, Portugal<br />

The balance in actions mediated by mineralocorticoid (MR) and glucocorticoid<br />

(GR) receptors in certain regions of the brain, predominantly in the limbic system, appears<br />

critical for neuronal activity, stress responsiveness, and behavioral programming and<br />

adaptation. Altered MR/GR balance renders nervous tissue into a vulnerable status, with<br />

consequences for the regulation of stress response and enhances susceptibility to<br />

psychopathology, especially in predisposed individuals. The influence of corticosteroids on<br />

limbic functions is now indisputable. However, closer scrutiny of early studies together<br />

with interpretations from newer studies would suggest that the proposition that<br />

corticosteroid-induced neuronal death accounts fully for the associated corticosteroidinduced<br />

cognitive deficits is only partially correct. Firstly, it is now clear that specific subpopulations<br />

of neurons are more sensitive to changes in the corticosteroid environment.<br />

Secondly, from a critical analysis of the available data, the picture that emerges is that<br />

corticosteroids, by acting through two distinct receptors, influence not only cell birth and<br />

death, but probably mainly synaptic plasticity. MR occupation appears to be essential for<br />

the survival of existing and newly generated neurons. In contrast, while GR can induce loss<br />

of neurons in the absence of MR activation, it appears that their occupation usually results<br />

in less drastic effects involving only dendritic atrophy and loss of synaptic contacts.<br />

In this presentation we will build a revised scheme of corticosteroid actions on neuronal<br />

plasticity that ultimately determine brain structure and function.<br />

129


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

POSTNATAL ONTOGENY OF HIPPOCAMPAL EXPRESSION OF THE<br />

MINERALOCORTICOID AND GLUCOCORTICOID RECEPTORS IN THE<br />

COMMON MARMOSET MONKEY<br />

Pryce C.R.<br />

Department of Neuroscience, Novartis Institutes for Biomedical Research, Novartis<br />

Pharma AG, Basel, Switzerland e-mail: christopher.pryce@novartis.com<br />

The few primate studies to-date of the CNS distribution of corticosteroid receptor<br />

expression have been conducted in adults and have demonstrated some marked differences<br />

to the expression distribution in adult rodents, and some important homologies to that in<br />

adult humans. Developmental studies in primates were lacking until recently [1], but are<br />

important in order to understand the roles of the mineralocorticoid receptor (MR) and<br />

glucocorticoid receptor (GR) in CNS maturation, both in a homeostatic environment and as<br />

mediators of the short- and long-term effects of early-life stress. The common marmoset is<br />

a small South American monkey that is suitable for developmental study, e.g. gestation<br />

period of 5 months, infancy period of 3 months, and sexual maturation at 15 months.<br />

Marmosets live in extended family groups, and reproduction is characterized by twinning<br />

and biparental care. Based on the findings obtained in long-term descriptive and<br />

experimental studies conducted in family groups of common marmosets, the following will<br />

be presented: (1) the development pro<strong>file</strong>s of basal and acute-stress ACTH and cortisol<br />

titres from birth until adulthood in undisturbed family groups; (2) the development pro<strong>file</strong>s<br />

of hippocampal and hypothalamic MR and GR expression from birth until adulthood in<br />

family groups.<br />

(1) Basal plasma levels of ACTH were similar in neonates (day 1-2) and infants (week 4)<br />

and elevated at these life stages relative to juveniles (month 4-10) and adults (year 2-6).<br />

Basal plasma and CSF levels of cortisol were markedly (10-fold) elevated in neonates<br />

relative to infants and elevated (2-fold) in infants relative to juveniles and adults. The high<br />

cortisol levels in neonates were associated with large adrenal (fetal zone) glands. Using<br />

blood sampling as an acute challenge, infants (week 8) demonstrated similar ACTH and<br />

cortisol peak stress responses to juveniles and adults; ACTH post-stress recovery was<br />

similar in infants, juveniles and adults, whereas cortisol recovery was retarded, suggesting<br />

slower cortisol metabolism at younger life stages [2].<br />

(2) For MR and GR, the marmoset cDNA sequence exhibited high homology (97%) with<br />

human MR and GR. Using marmoset-specific riboprobes and in situ hybridization, it was<br />

demonstrated that MR mRNA expression in the dentate gyrus and Ammon’s horn (CA1-4)<br />

was greater in marmoset infants (week 4-6) than in neonates (day 1-2), juveniles (month 4-<br />

5) and adults (year 3-6), with expression in the latter three stages being similar. In the same<br />

subjects and ontogenetic stages, GR mRNA expression was developmentally consistent in<br />

DG, CA1-4 and paraventricular nucleus of the hypothalamus (PVNh). Qualitative<br />

immunohistochemical analysis demonstrated that developmental pro<strong>file</strong>s of MR and GR<br />

protein expression reflected those of their respective mRNA pro<strong>file</strong>s [1].<br />

Combining the findings summarized in (1) and (2), infancy is characterized by a<br />

combination of high ACTH and cortisol levels and high hippocampal MR expression,<br />

relative to subsequent life stages.<br />

130


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

Reference list<br />

[1] Pryce CR, Feldon J, Fuchs E, Knuesel I, Oertle T, Sengstag C, Spengler M, Weber E,<br />

Weston A, Jongen-Relo A. Postnatal ontogeny of hippocampal expression of the<br />

mineralocorticoid and glucocorticoid receptors in the common marmoset monkey. Eur<br />

J Neuroscience 2005;21:1521-15<strong>35</strong>.<br />

[2] Pryce CR, Palme R, Feldon J. Development of pituitary-adrenal endocrine function in<br />

the marmoset monkey: Infant hyper-cortisolism is the norm. J Clin Endocrinol Metab<br />

2002;87:691-699.<br />

131


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

MICE WITH COMPROMISED GLUCOCORTICOID RECEPTOR EXPRESSION<br />

SHOW BEHAVIOURAL AND BIOCHEMICAL FEATURES OF DEPRESSION<br />

Gass P.<br />

Central Institute of Mental Health Mannheim (ZI), University of Heidelberg, Germany<br />

Altered glucocorticoid receptor (GR) signalling is a postulated mechanism for the<br />

pathogenesis of major depression. To mimic the human situation of altered GR function<br />

claimed for depression we have generated mouse strains that under- or overexpress GR,<br />

but maintain the regulatory genetic context controlling the GR gene. We generated: i) GR<br />

heterozygous mutant mice (GR +/- ) with a 50% GR gene dose reduction; and ii) mice<br />

overexpressing GR by a yeast artificial chromosome resulting in a 2-fold gene dose<br />

elevation. These strains were subjected to a large battery of basal and stress-related<br />

behavioral tests. Furthermore, they were analysed for neuroendocrinological and<br />

neurochemical alterations. GR +/- mice exhibit normal baseline behaviors, but demonstrate<br />

increased helplessness after stress exposure, a behavioral correlate of depression in mice.<br />

Similar to depressed patients, GR +/- mice have a disinhibited HPA system and a<br />

pathological DEX/CRH test. Thus, they represent a murine depression model with good<br />

face and construct validity. Overexpression of GR in mice evokes reduced helplessness<br />

after stress exposure, and an enhanced HPA system feedback regulation. Therefore they<br />

may represent a model for a stress-resistant strain. These mouse models can now be used to<br />

study biological changes underlying the pathogenesis of depressive disorders. As a first<br />

potential molecular correlate for such changes we identified a downregulation of BDNF<br />

protein content in the hippocampus of GR +/- mice, which is in agreement with the so-called<br />

neurotrophin hypothesis of depression. Furthermore, these strains may represent a tool to<br />

detect pharmacological mechanisms or develop new psychopharmacological principles for<br />

the treatment of depression.<br />

132


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

MATERNAL ADVERSITY, GLUCOCORTICOIDS AND PROGRAMMING OF<br />

NEUROENDOCRINE FUNCTION AND BEHAVIOUR<br />

Matthews S.M.<br />

Departments of Physiology, Obstetrics and Gynecology and Medicine, Faculty of<br />

Medicine University of Toronto, 1 King’s College Circle, Toronto, Ontario, M5S 1A8,<br />

CANADA.<br />

The fetus may be exposed to increased endogenous glucocorticoid (GC) or synthetic<br />

glucocorticoid (sGC) in late gestation. Indeed, approximately 7% of pregnant women in<br />

Europe and North America are treated with sGC to promote lung maturation in fetuses at<br />

risk of pre-term delivery. It is now established that exposure of the fetal brain to excess GC<br />

can have life-long effects on neuroendocrine function and behaviour. Using the guinea pig,<br />

we have shown that both endogenous GC and sGC exposure has a number of rapid effects in<br />

the fetal brain in late gestation, including modification of neurotransmitter systems and<br />

transcriptional machinery. Such fetal exposure permanently alters hypothalamo-pituitaryadrenal<br />

(HPA) function and behaviour in prepubertal, post-pubertal and aging offspring, in a<br />

sex-dependent manner. However, these effects are dependant on the time in gestation at<br />

which GC exposure occurs as well as the age at which endocrine and behavioural outcomes<br />

are assessed. More recently, we have identified transgenerational effects of prenatal exposure<br />

to sGC on HPA function, behaviour and growth. Indeed, the magnitude of effects in the<br />

second generation is greater than that in the first generation. Defining the mechanisms<br />

involved in programming of HPA function and behaviour across generations will have<br />

considerable implications for clinical management of pregnancy.<br />

Supported by: Canadian Institutes of Health Research (CIHR) and Natural Sciences &<br />

Engineering Research Council (NSERC).<br />

133


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

LONG LASTING EFFECTS OF STRESS DURING PREGNANCY ON HPA<br />

FUNCTION AND BEHAVIOUR IN MOTHER AND OFFSPRING RATS<br />

Darnaudéry M., Morley-Fletcher S. and Maccari S.<br />

University of Lille 1, Laboratory of Neurosciences and Adaptive Physiology, Perinatal<br />

Stress Team, Bât. SN4.1, 59655 Villeneuve d’Ascq, France.<br />

E-mail:muriel.darnaudery@univ-lille1.fr; Fax: +33 3 20 43 46 02<br />

Life events occurring during the perinatal period have strong permanent long-term effects<br />

on the behavioural and neuroendocrine response to stressors. In rats, repeated restraint<br />

stress of the pregnant dam during the last week of pregnancy produces long lasting changes<br />

in the hypothalamic-pituitary-adrenocortical (HPA) axis function and behaviours in the<br />

offspring. These changes include a hyperactivity of HPA axis response associated with a<br />

reduction in the number of hippocampal corticosteroid receptors [8]. This can be evidenced<br />

by a more prolonged elevation of plasma ACTH and/or corticosterone after moderate<br />

stressors such as exposure to novelty, restraint stress or exposure to elevated plus<br />

maze[25,25,33,48,48,52,52]. The HPA dysfunctions have been reported in infant, young<br />

adult and aged animals, therefore suggesting a permanent effect of early stress [5, 7, 13,<br />

15]. Interestingly, after the confrontation to an intense inescapable footshock, prenatally<br />

stressed rats durably show a blunted corticosterone secretion after stress. Prenatal stress<br />

also induces a hyporeponse of the HPA axis when animals are exposed to an alcohol<br />

challenge [14]. These results suggest that HPA alterations associated to prenatal stress may<br />

vary according to the nature and/or the intensity of the stressor.<br />

Rats exposed to a prenatal stress also show behavioural disturbances known to be related to<br />

the HPA axis. Indeed, prenatal stress produces high anxiety levels and depressive-like<br />

behaviour during adulthood [10,12]. With ageing, these animals exhibit memory<br />

impairments in hippocampo-dependent tasks [13, 3]. Despite the permanent imprinting<br />

induced by stress in utero, the dysfunctions observed after prenatal stress can be reversed<br />

by environmental or pharmacological strategy. For example, early adoption [7] or<br />

environmental enrichment during adolescence [9], as well as a chronic treatment with<br />

Insulin-like growth factor 1 in aged animals [3] attenuated some HPA dysfunction’s<br />

produced by prenatal stress.<br />

Mechanisms underlying the prenatal stress effects on the offspring remain largely<br />

unknown. However, previous works demonstrated that maternal glucocorticoids during<br />

pregnancy may play an important role in the HPA disturbances reported. Thus, stressed<br />

mothers show high glucocorticoid levels during pregnancy [1, 6]. Furthermore, in the<br />

offspring of stressed mothers, the HPA response to stress is normalised by maternal<br />

adrenalectomy during pregnancy [1]. Recently, our group has reported that repeated<br />

restraint stress during pregnancy leads to a decrease of the placental 11β-HSD2 activity.<br />

Finally, gestational stress has long lasting effects on HPA axis and behaviour in female<br />

dams [2, 11]. Thus, during lactating period, stressed mothers show an impairment of<br />

maternal care and low aggressive behaviour against a male intruder. Moreover, females<br />

stressed during pregnancy show an increase of anxiety-like behaviour several weeks after<br />

the end of the stress period. Given that, several evidences suggest that changes in maternal<br />

care may durably program offspring’s HPA function and behaviours [4], it could be<br />

postulated that the alterations of the maternal behaviour during the early postnatal period<br />

may also strongly contribute to the long-term effect described after prenatal stress.<br />

134


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

Reference list<br />

1. A. Barbazanges, P.V.Piazza, M. Le Moal., and S.Maccari, Maternal glucocorticoid secretion mediates<br />

long-term effects of prenatal stress, J. Neurosci. 16 (1996) 3943-3949.<br />

2. M. Darnaudery, I.Dutriez, O.Viltart, S.Morley-Fletcher, and S.Maccari, Stress during gestation induces<br />

lasting effects on emotional reactivity of the dam rat, Behav. Brain Res. 153 (2004) 211-216.<br />

3. M. Darnaudery, M.Perez-Martin, G.Belizaire, S.Maccari, and L.M.Garcia-Segura, Insulin-like growth<br />

factor 1 reduces age-related disorders induced by prenatal stress in female rats, Neurobiol. Aging. 27<br />

(2006) 119-127.<br />

4. E.W. Fish, D.Shahrokh, R.Bagot, C.Caldji, T.Bredy, M.Szyf, and M.J.Meaney, Epigenetic Programming<br />

of Stress Responses through Variations in Maternal Care, Ann. N. Y. Acad. Sci. 1036:167-80. (2004)<br />

167-180.<br />

5. C. Henry, M.Kabbaj, H.Simon, M.Le Moal, and S.Maccari, Prenatal stress increases the hypothalamopituitary-adrenal<br />

axis response in young and adult rats, J. Neuroendocrinol. 6 (1994) 341-345.<br />

6. M. Koehl, M.Darnaudery, J.Dulluc, O.Van Reeth., M. Le Moal., and S.Maccari, Prenatal stress alters<br />

circadian activity of hypothalamo-pituitary-adrenal axis and hippocampal corticosteroid receptors in<br />

adult rats of both gender, J. Neurobiol. 40 (1999) 302-315.<br />

7. S. Maccari, P.V.Piazza, M.Kabbaj, A.Barbazanges, H.Simon, and M. Le Moal., Adoption reverses the<br />

long-term impairment in glucocorticoid feedback induced by prenatal stress, J. Neurosci. 15 (1995) 110-<br />

116.<br />

8. S. Maccari, M.Darnaudery, S.Morley-Fletcher, A.R.Zuena, C.Cinque, and.O.Van Reeth, Prenatal stress<br />

and long-term consequences: implications of glucocorticoid hormones, Neurosci. Biobehav. Rev. 27<br />

(2003) 119-127.<br />

9. S. Morley-Fletcher, M.Rea, S.Maccari, and G.Laviola, Environmental enrichment during adolescence<br />

reverses the effects of prenatal stress on play behaviour and HPA axis reactivity in rats, Eur. J. Neurosci.<br />

18 (2003) 3367-3374.<br />

10. S. Morley-Fletcher, M.Darnaudery, M.Koehl, P.Casolini, O.Van Reeth., and S.Maccari, Prenatal stress<br />

in rats predicts immobility behavior in the forced swim test. Effects of a chronic treatment with<br />

tianeptine, Brain Res. 989 (2003) 246-251.<br />

11. J.W. Smith, J.R.Seckl, A.T.Evans, B.Costall, and J.W.Smythe, Gestational stress induces post-partum<br />

depression-like behaviour and alters maternal care in rats, Psychoneuroendocrinology. 29 (2004) 227-<br />

244.<br />

12. M. Vallee, W.Mayo, F.Dellu, M. Le Moal., H.Simon, and S.Maccari, Prenatal stress induces high<br />

anxiety and postnatal handling induces low anxiety in adult offspring: correlation with stress-induced<br />

corticosterone secretion, J. Neurosci. 17 (1997) 2626-2636.<br />

13. M. Vallee, S.Maccari, F.Dellu, H.Simon, M.Le Moal, and W.Mayo, Long-term effects of prenatal stress<br />

and postnatal handling on age-related glucocorticoid secretion and cognitive performance: a longitudinal<br />

study in the rat, Eur. J. Neurosci. 11 (1999) 2906-2916.<br />

14. V. Van Waes, M.Enache, I.Dutriez, J.Lesage, S.Morley-Fletcher, E.Vinner, M.Lhermitte, D.Vieau,<br />

S.Maccari, and M.Darnaudery, Hypo-response of the hypothalamic-pituitary-adrenocortical axis after an<br />

ethanol challenge in prenatally stressed adolescent male rats, Eur. J. Neurosci. 24 (2006) 1193-1200.<br />

15. O. Viltart, J.Mairesse, M.Darnaudery, H.Louvart, C.Vanbesien-Mailliot, A.Catalani, and S.Maccari,<br />

Prenatal stress alters Fos protein expression in hippocampus and locus coeruleus stress-related brain<br />

structures, Psychoneuroendocrinology. 31 (2006) 769-780.<br />

1<strong>35</strong>


Posters’ Exhibition


Posters’ Exhibition:<br />

Effects mediated by classical steroid receptors<br />

• Benedusi V., Pozzi S., Maggi A. and Vegeto E. (Italy) The anti-inflammatory<br />

activity of estrogenic compounds in microglia<br />

• Mattsson, A., Mura, E., Halldin, K., Panzica G.C., Brunström, B (Sweden)<br />

Embryonic exposure to an ERα agonist affects reproductive organ development but<br />

does not alter copulatory behavior or the parvocellular vasotocin system in male<br />

Japanese quail<br />

• Mayoral, S.R. and Penn, A.A. (USA) Brain estrogen receptor expression in<br />

perinatal mice<br />

• Pozzi S., Benedusi V., Vegeto E. and Maggi A. (Italy) Anti-inflammatory activity<br />

of estrogen in acute and chronic brain inflammation<br />

• Sanz A., Carrero P., Pernía O., Garcia-Segura L.M. (Spain) Both basal and<br />

estradiol-regulated activation of IGF-I receptor signaling in the rat brain are<br />

affected by the duration of previous ovarian hormonal deprivation<br />

• Walf, A.A., Frye, C.A (USA) Estradiol and selective estrogen receptor modulators<br />

with activity at estrogen receptor beta have dose-dependent effects to reduce<br />

anxiety and depressive behavior


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

THE ANTI-INFLAMMATORY ACTIVITY OF ESTROGENIC COMPOUNDS IN<br />

MICROGLIA<br />

Benedusi V., Pozzi S., Maggi A. and Vegeto E.<br />

Center of Excellence on Neurodegenerative Diseases, Department of Pharmacological<br />

Sciences, University of Milan, Via Balzaretti 9, 20133 Milan, Italy.<br />

Fax: 0039.02.50318284<br />

e-mail: elisabetta.vegeto@unimi.it<br />

The generation of proinflammatory and neurotoxic factors by microglia, the resident brain<br />

immune cells, plays a prominent role in mediating the progressive neurodegenerative<br />

process. Since there is an increasing evidence of beneficial effects of estrogen against<br />

neuroinflammation, we hypothesized that microglia could be a target for estrogen antiinflammatory<br />

activity in brain.<br />

Using primary cultures of microglia we initially found that estrogen blocks the LPSinduced<br />

conversion of microglia cells towards the activated phenoptype and inhibits the<br />

production of inflammatory mediators like MMP-9, iNOS and PGE2 associated with<br />

microglia reactivity. To elucidate the molecular mechanism of estrogen anti.-inflammatory<br />

activity we analysed whether estrogen could modify the activity of p65, a transcription<br />

factor which stimulates the espression of many genes involved in the inflammatory process<br />

(e.g. IL2, TNF, adhesion molecules and acute phase proteins). Using molecular biology<br />

techniques we demonstrated that estrogen prevents p65 translocation into the nucleus<br />

through a non genomic, ERalpha-mediated event which requires PI3K activation. Current<br />

studies are underway to identify which inflammatory genes are modulated by estrogen and<br />

whether isoform-specific or selective-estrogen receptor modulators mimic the antiinflammatory<br />

activity of the endogenous hormone in microglia cells. Recent preliminary<br />

data in our laboratory show that estrogen is able to reduce LPS-induced TNF-alpha<br />

production; TNF-alpha expression is regulated by NF-kB so this observation is in<br />

agreement with our previous findings. Interestingly, selective ERalpha agonists are able to<br />

reduce LPS-induced TNF-alpha production, whilst selective ERbeta agonists do not show<br />

any effect. These preliminary data are further confirming the key role of ERalpha, and not<br />

ERbeta, in the anti-inflammatory effect of estrogen in the brain; further recent results will<br />

also be discussed.<br />

138


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

EMBRYONIC EXPOSURE TO AN ERα AGONIST AFFECTS REPRODUCTIVE<br />

ORGAN DEVELOPMENT BUT DOES NOT ALTER COPULATORY BEHAVIOR<br />

OR THE PARVOCELLULAR VASOTOCIN SYSTEM IN MALE JAPANESE<br />

QUAIL<br />

Mattsson A. a , Mura E. b , Halldin K. c , Panzica G.C. b , Brunström B. a<br />

a Dept. of Environmental Toxicology, Uppsala University, Norbyvägen 18A, SE-75236<br />

Uppsala, Sweden. E-mail: Anna.Mattsson@ebc.uu.se, Fax: +46-18518843<br />

b Dept. of Anatomy, Pharmacology, and Forensic Medicine, Laboratory of<br />

Neuroendocrinology, University of Torino, Torino, Italy<br />

c Inst. of Environmental Medicine, Karolinska Institutet, Stockholm, Sweden.<br />

Estradiol plays an important role in female sex differentiation of the brain and the<br />

reproductive organs during embryonic development in birds. In Japanese quail, the<br />

copulatory behavior and the parvocellular vasotocin (VT) system are sexually dimorphic<br />

traits that are organized during embryonic development. The nucleus of the stria terminalis,<br />

pars medialis (BSTm), the lateral septum and the medial preoptic nucleus (POM) contain<br />

denser VT-immunoreactivity (VT-ir) in males than in females [4]. Xenoestrogen exposure<br />

can demasculinize the brain of male embryos and interfere with the differentiation of the<br />

reproductive organs in both males and females. This may result in long lasting effects that<br />

impair fertility and reproductive success. For instance, the copulatory behavior of male<br />

Japanese quail exposed to estradiol benzoate, diethylstilbestrol, or genistein is reduced or<br />

totally abolished, which is paralleled by a decreased VT-ir [3]. Ethinylestradiol (EE2)<br />

exposure has been shown to result in retention of oviducts, increased testicle weight<br />

asymmetry, and reduced copulatory behavior in males [1,2]. The respective roles of the<br />

two nuclear estrogen receptors, ERα and ERβ, in normal and disrupted sexual<br />

differentiation are not well known. The aim of this study was to elucidate whether ERα<br />

can mediate disrupted differentiation of the brain and the reproductive organs in male<br />

Japanese quail embryos. We injected fertilised Japanese quail eggs with the selective ERα<br />

agonist Propyl pyrazole triol (PPT) on incubation day 3, i.e. well before sexual<br />

differentiation of reproductive organs and copulatory behaviour is completed. A second<br />

group was injected with EE2, a potent agonist to both ERα and ERβ. The males were<br />

tested for copulatory behavior at sexual maturity, and the gross morphology of the<br />

reproductive organs was examined. The brains were processed for VT<br />

immunohistochemistry. The copulatory behaviour was significantly reduced in the EE2<br />

group, whereas PPT had no effect. There was no correlation between behavioral<br />

performance and plasma concentration of testosterone, which confirms that the behavioural<br />

changes were not due to changes in testosterone concentration, but were rather caused by<br />

an organizational effect established during embryonic life. The VT-ir in BSTm, the lateral<br />

septum and POM was not affected by EE2 or PPT treatment. The effects on the<br />

reproductive organs were similar in the PPT-exposed as in the EE2-exposed birds. Effects<br />

included presence of oviducts which frequently were malformed, testicle weight<br />

asymmetry and reduced cloacal gland area. Our results show that treatment with the<br />

selective ERα agonist PPT causes similar adverse effects on reproductive organ<br />

development as treatment with EE2, and hence these effects can be mediated via ERα.<br />

However, PPT exposure was not sufficient to demasculinize the copulatory behavior or the<br />

VT system in male Japanese quail. This indicates that the demasculinizing effect of<br />

estrogens on the male brain is not mediated by ERα alone. Ongoing studies in our<br />

139


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

laboratory aim to resolve whether selective activation of ERβ may result in disrupted sex<br />

differentiation in the quail embryo.<br />

Reference list<br />

[1] Berg, C., Halldin, K., Fridolfsson, A.K., Brandt, I. and Brunstrom, B., The avian egg as a test<br />

system for endocrine disrupters: effects of diethylstilbestrol and ethynylestradiol on sex organ<br />

development, Sci Total Environ, 233 (1999) 57-66.<br />

[2] Halldin, K., Berg, C., Brandt, I. and Brunstrom, B., Sexual behavior in Japanese quail as a test end<br />

point for endocrine disruption: Effects of in Ovo exposure to ethinylestradiol and diethylstilbestrol,<br />

Environmental Health Perspectives, 107 (1999) 861-866.<br />

[3] Panzica, G., Mura, E., Pessatti, M. and Viglietti-Panzica, C., Early embryonic administration of<br />

xenoestrogens alters vasotocin system and male sexual behavior of the Japanese quail, Domestic<br />

Animal Endocrinology, 29 (2005) 436-445.<br />

[4] Panzica, G.C., Balthazart, J., Pessatti, M. and Viglietti-Panzica, C., The parvocellular vasotocin<br />

system of Japanese quail: a developmental and adult model for the study of influences of gonadal<br />

hormones on sexually differentiated and behaviorally relevant neural circuits, Environ Health<br />

Perspect, 110 Suppl 3 (2002) 423-8.<br />

140


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

BRAIN ESTROGEN RECEPTOR EXPRESSION IN PERINATAL MICE<br />

Mayoral S.R. and Penn A.A.<br />

Department of Pediatrics, Stanford University, 300 Pasteur Drive, Stanford, CA 94305-<br />

5208, USA. Fax: 650-725-7724 e-mail: apenn@stanford.edu.<br />

Estrogen is important for neuronal growth and survival [1, 2]. Studies of preterm infants,<br />

who experience an early loss of placental estrogen, show that preterm males suffer worse<br />

brain abnormalities and cognitive outcomes than preterm females [3]. These observations<br />

suggest that estrogen is crucial for the proper development of the brain. However, the role<br />

and the expression of the two principal estrogen receptors, ERα and ERβ, in CNS<br />

development are still not well established. We examined perinatal ERα and ERβ mRNA<br />

expression in CD1 mouse cortex, hippocampus and cerebellum by quantitative real-time<br />

RT-PCR. We found significant gender-dependent differences in ERα expression in the<br />

developing cortex. In embryonic (E18) CD1 mice, ERα levels in males and females in the<br />

cortex were similar (P=.24; t-test). However, after birth (P0 and P7) these values diverge<br />

significantly (P


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

ANTI-INFLAMMATORY ACTIVITY OF ESTROGEN IN ACUTE AND<br />

CHRONIC BRAIN INFLAMMATION<br />

Pozzi S., Benedusi V., Vegeto E. and Maggi A.<br />

Center of Excellence on Neurodegenerative Diseases, Department of Pharmacological<br />

Sciences, University of Milan, Via Balzaretti 9, 20133 Milan, Italy. Fax:<br />

0039.02.50318284<br />

e-mail: adriana.maggi@unimi.it<br />

Activation of microglia cells, the resident macrophage cells of the brain, is the hallmark of<br />

acute and chronic neurodegenerative disorders, such as ischemia, Multiple Sclerosis (MS)<br />

and Alzheimer Disease (AD), characterized by inflammatory events.<br />

Recent studies reported that 17β-estradiol (E 2 ) acts as a neuroprotective agent in brain by<br />

both targeting neurons and inhibiting the brain inflammatory reactions. In our lab we<br />

recently developed an experimental model of brain inflammation, in which LPS<br />

(lipopolysaccharide), a potent inflammatory agent, is injected in the cerebral ventricles,<br />

resulting in a transient and localised acute neuroinflammatory reaction. We used this<br />

system to investigate the E 2 anti-inflammatory activity in the central nervous system and<br />

demonstrated that E 2 is a potent inhibitor of microglia reactivity in several regions of the<br />

brain, including cortex, hippocampus and noncortical areas and that hormone<br />

administration results in a significant reduction of the expression of inflammatory markers,<br />

such as TNF-α, MCP-1 and MIP-2 [1]. Our data thus show that estrogen is able to quench<br />

acute brain inflammation in vivo, in agreement with data published by other groups on<br />

neuroinflammatory processes such as ischemia or experimental allergic encephalomyelitis<br />

(EAE). On the other hand, we used the APP23 transgenic mice, expressing the human<br />

amyloid precursor protein (APP) with a mutation reported in familial AD, in order to<br />

understand the role of E 2 in chronic neurodegenerative diseases. In this animal model of<br />

AD microglia displays the characteristic activated morphology and immunoreactive<br />

phenotype induced by the chronic deposition of the amyloid peptide (Aβ). We observed<br />

that ovariectomy increases microglia activation at Aβ deposits whereas chronic<br />

replacement with E 2 in ovarectomized APP23 mice reduced neuroinflammation. Thus,<br />

chronic neuroinflammatory events, associated with neurodegeneration, are also targeted by<br />

hormone action. Future studies using synthetic estrogenic compounds will be discussed.<br />

Reference list<br />

[1] E. Vegeto, S. Belcredito , S. Ghisletti, C. Meda, S. Etteri and A. Maggi, The endogenous estrogen status<br />

regulates microglia reactivity in animal models of neuroinflammation, Endocrinology 147 (5) (2006), pp.<br />

2263-2272<br />

142


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

BOTH BASAL AND ESTRADIOL-REGULATED ACTIVATION OF IGF-I<br />

RECEPTOR SIGNALING IN THE RAT BRAIN ARE AFFECTED BY THE<br />

DURATION OF PREVIOUS OVARIAN HORMONAL DEPRIVATION<br />

Sanz A.*, Carrero P., Pernía O., Garcia-Segura L.M.<br />

Instituto Cajal, C.S.I.C., Avenida Doctor Arce, 37, E-28002, Madrid, Spain. *E-mail:<br />

sanz_amaya@hotmail.com FAX: +34-915854754<br />

The ovarian hormone 17 beta-estradiol (E2) is neuroprotective in many experimental<br />

models of neurodegeneration. This pro-survival effect of E2 is partially due to its ability to<br />

regulate the activity of some proteins of the signaling pathways of growth factors, such as<br />

the phosphoinositide 3-kinase (PI3K) pathway and the mitogen-activated protein kinase<br />

(MAPK) pathway. However, the neuroprotective effects of E2 have not been always<br />

confirmed in clinical studies. There is evidence that early initiation of hormone therapy,<br />

during the perimenopausal period, may provide cognitive benefits to postmenopausal<br />

women. In contrast, late initiation of hormone therapy, several years after menopause, may<br />

increase cognitive deficits. In addition to possible differences due to age, the discrepant<br />

results of hormone therapy may be explained if brain responsiveness to E2 is sensitive to<br />

the duration of previous ovarian hormonal deprivation. To test this hypothesis, we have<br />

assessed whether different lengths of ovarian hormonal deprivation before E2<br />

administration may affect the basal and E2-regulated phosphorylation of kinases associated<br />

to the IGF-I receptor (IGF-IR) in the brain. Wistar female rats were ovariectomized 10<br />

days (short-term hormonal deprivation) or 30 days (long-term hormonal deprivation)<br />

before the i.p. injection of 300 micrograms E2 and killed at the age of 3 months, 24 h after<br />

the administration of E2. The cerebral cortex, the hippocampus, the hypothalamus and the<br />

cerebellum were dissected out and the levels of phosphorylation of Akt, glycogen synthase<br />

kinase 3 beta (GSK3 beta) and extracellular signal-regulated kinases (ERK1/2), the levels<br />

of expression of the IGF-IR and the levels of expression of estrogen receptors (ER alfa and<br />

ER beta) were assessed by Western blotting. Basal levels of phosphorylation of Akt, GSK3<br />

beta and ERK1/2 showed significant differences depending on the time elapsed after<br />

ovarian removal and the brain region. Long-term hormonal deprivation resulted in<br />

increased basal phosphorylation levels of Akt and ERK-1 in the hippocampus, GSK3 beta,<br />

ERK1 and ERK2 in the hypothalamus and Akt, ERK-1 and ERK-2 in the cerebral cortex.<br />

In contrast, long-term hormonal deprivation reduced basal phosphorylation levels of GSK3<br />

beta, ERK-1 and ERK-2 in the cerebellum. In addition, E2 decreased the phosphorylation<br />

levels of ERK-1 and ERK-2 in the cerebral cortex at 10 days, but not at 30 days, after<br />

ovariectomy and decreased ERK-2 phosphorylation in the cerebellum at 30 days, but not at<br />

10 days, after ovariectomy. The levels of expression of ERs and IGF-IR were also affected<br />

by the duration of ovarian hormonal deprivation and E2 administration. IGF-IR expression<br />

was decreased in the cerebral cortex after long-term hormonal deprivation and by E2<br />

treatment after short-term hormonal deprivation. The expression of ER alpha was<br />

decreased after long-term hormonal deprivation in the hippocampus. In conclusion, our<br />

findings indicate that, within the same age group, the duration of previous ovarian<br />

hormone deprivation differentially affects basal levels of IGF-IR signaling and its<br />

responsiveness to estrogen and is therefore an important parameter to consider when<br />

assessing neuroprotective and cognitive effects of hormone therapy.<br />

Supported by Ministerio de Educación y Ciencia, Spain (SAF 2005-00272) and the<br />

European Union (EWA project: LSHM-CT-2005-518245).<br />

143


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

ESTRADIOL AND SELECTIVE ESTROGEN RECEPTOR MODULATORS WITH<br />

ACTIVITY AT ESTROGEN RECEPTOR BETA HAVE DOSE-DEPENDENT EFFECTS<br />

TO REDUCE ANXIETY AND DEPRESSIVE BEHAVIOR<br />

Walf, A.A. 1 , Frye, C.A. 1-4 ,<br />

Dept. of Psychology 1 , Biological Sciences 2 , and The Centers for Neuroscience 3 and Life Science 4 Research<br />

The University at Albany-SUNY, Life Sciences 01048, 1400 Washington Avenue, Albany, NY USA 12222;<br />

aawalf@yahoo.com, 518-591-8838<br />

Estradiol (E 2 ) alters anxiety and depressive behavior of female rodents. Female rats during<br />

proestrus, which have high physiological E 2 levels, have decreased anxiety and depressive behavior<br />

compared to females with lower E 2 levels and males [3,4,11]. Similarly, subcutaneous (SC) administration<br />

of E 2 that produces physiological E 2 concentrations anxiety and depressive behavior of ovariectomized (ovx)<br />

rats and mice [3-6]. However, E 2 dosing that produces more sustained and/or higher E 2 concentrations does<br />

not consistently decrease anxiety and depressive behavior [11-14]. Indeed, some of the variations that have<br />

been reported for the effects of E 2 on anxiety and depressive behavior among females may be related to E 2<br />

dosing regimen utilized [11-12].<br />

Another factor that may modulate the functional response to E 2 may involve E 2 ’s estrogen receptor<br />

(ER) isoform-specific effects. Administration of ER antagonists systemically or directly to an area of the<br />

brain that is important for E 2 ’s effects on anxiety and depressive behavior, the hippocampus, attenuates E 2 ’s<br />

anti-anxiety and anti-depressive effects in ovx rats [11,13]. Although these data suggest that ERs are<br />

important for the functional effects of E 2 , the ER antagonists utilized in these studies were not ER isoform<br />

(ERα or ERβ)-specific. Indeed, recent studies using knockout mice, antisense oligonucleotides, and<br />

administration of selective estrogen receptor modulators (SERMs) suggest that E 2 ’s effects for anxiety and<br />

depression may be dependent upon ERβ, whereas E 2 ’s effects for reproductive behavior are dependent upon<br />

ERα [1,2,5,11-14]. Whether there are dose-dependent effects of ERα or ERβ agonists for these effects is of<br />

interest.<br />

These data suggesting that there are ERβ-specific effects of E 2 for anxiety and depressive behavior<br />

is intriguing given that estrogens can have robust proliferative effects on peripheral, E 2 -sensitive tissues, such<br />

as mammary glands and uterus, that involve actions of ERα. As such, it is important to further characterize<br />

ER-isoform-specific and dose-dependent effects of SERMs to be able to parse beneficial effects in brain from<br />

possible negative effects in these peripheral tissue. To begin to address this, the effects of SC administration<br />

of different dosages (0, 0.3, 0.6, 0.9, or 1.8 mg/kg) of 17β-E 2 , an ERα agonist (propyl pyrazole triol; PPT, or<br />

an ERβ agonist (diarylpropionitrile; DPN) for behavior in measures of anxiety (elevated plus maze),<br />

depression (forced swim test), and reproduction/sexual receptivity (lordosis) were compared in ovx rats. We<br />

hypothesized that there would be dose-dependent effects of E 2 and DPN for anxiety and depression behavior<br />

and a dose-dependent effects of E 2 and PPT for lordosis.<br />

Results of the present study supported our hypothesis of the dose-dependent, ER-isoform specific<br />

effects of SERMs for anxiety, depression, and reproductive behavior. Administration of 0.9 mg/kg, but not<br />

lower or higher dosages of E 2 , decreased anxiety (i.e. increased open arm time in the elevated plus maze),<br />

decreased depressive behavior (i.e. decreased immobility in the forced swim test), compared to vehicle,<br />

which replicated our previous findings [4]. All E 2 dosages increased lordosis. 0.3 mg/kg DPN, compared to<br />

vehicle, increased open arm time, decreased immobility, and did not alter lordosis. All doses of PPT<br />

enhanced lordosis compared to vehicle, but did not alter affective behavior.<br />

Together, these data suggest that there are dose-dependent effects of ERβ agonists for affect and<br />

ERα agonists for sexual receptivity. Studies are ongoing investigating the proliferative effects of these<br />

compounds concomitant with these behavioral effects.<br />

Reference list<br />

Supported by: NSF (IBN03-16083) U.S. Dept. of Defense (BC051001).<br />

[1] D.B. Imwalle, J.A. Gustafsson, E.F. Rissman, Lack of functional estrogen receptor β influences anxiety behavior<br />

and serotonin content in female mice, Physiol Behav. 84 (2005) 157-63.<br />

[2] W. Krezel, S. Dupont, A. Krust, P. Chambon, P.F. Chapman, Increased anxiety and synaptic plasticity in estrogen<br />

receptor β -deficient mice. PNAS. 98 (2001) pp. 12278-82.<br />

[3] C.A. Frye, S.M. Petralia, M.E. Rhodes, Estrous cycle and sex differences in performance on anxiety tasks coincide<br />

with increases in hippocampal progesterone and 3α,5α-THP, Pharmacol. Biochem. Behav. 67 (2000) pp. 587-596.<br />

[4] C.A. Frye, A.A. Walf, Changes in progesterone metabolites in the hippocampus can modulate open field and forced<br />

swim test behavior of proestrous rats. Horm. Behav. 41 (2002) pp. 306-15<br />

144


% of control- Open Arm Time<br />

% of control- Lordosis Quotient<br />

% of control- Immobility Time<br />

% of control- Open Arm Time<br />

% of control- Lordosis Quotient<br />

% of control- Immobility Time<br />

4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

[5] T.D. Lund, T. Rovis, W.C. Chung, R.J. Handa, Novel actions of estrogen receptor-beta on anxiety-related behaviors.<br />

Endocrinology. 146 (2005) pp. 797-807.<br />

[6] S. Ogawa, J. Chan, A.E. Chester, J.A. Gustafsson, K.S. Korach, D.W. Pfaff, Survival of reproductive behaviors in<br />

estrogen receptor β gene-deficient male and female mice, PNAS 96 (1999) pp. 12887-92.<br />

[7] S. Ogawa, J. Chan, J.A. Gustafsson, K.S. Korach, D.W. Pfaff, Estrogen increases locomotor activity in mice through<br />

estrogen receptor α: specificity for the type of activity. Endocrinology 144 (2003) pp. 230-9.<br />

[8] B.A. Rocha, R. Fleischer, J.M. Schaeffer, S.P. Rohrer, G.J. Hickey, 17β-Estradiol-induced antidepressant-like effect<br />

in the Forced Swim Test is absent in estrogen receptor-β knockout (BERKO) mice. Psychopharmacology. 179 (2005) pp.<br />

637-43.<br />

[9] A.A. Walf , I. Ciriza, L.M. Garcia-Segura, C.A. Frye, , Antisense oligodeoxynucleotides for estrogen receptor beta and<br />

alpha attenuate estrogen’s modulation of affective and sexual behavior, respectively, Neuropsychopharmacology (in<br />

revision).<br />

[10] A.A. Walf, C.A., Frye, Administration of estrogen receptor beta-specific selective estrogen receptor modulators to<br />

the hippocampus decrease anxiety and depressive behavior of ovariectomized rats, Pharmacol Biochem Behav. (in<br />

press).<br />

[11] A.A. Walf, C.A., Frye, A review and update of mechanisms of estrogen in the hippocampus and amygdala for<br />

anxiety and depression behavior. Neuropsychopharmacology 31 (2006) pp. 1097-111.<br />

[12] A.A. Walf, C.A., Frye, Estradiol’s effects to reduce anxiety and depressive behavior may be mediated by estradiol<br />

dose and restraint stress. Neuropsychopharmacology 30 (2005) pp. 1288-301.<br />

[13] A.A. Walf, C.A., Frye, ERβ-selective estrogen receptor modulators produce antianxiety behavior when<br />

administered systemically to ovariectomized rats. Neuropsychopharmacology.30 (2005) pp. 1598-609.<br />

[14]A.A. Walf, M.E. Rhodes, C.A. Frye, Anti-depressant effects of ERβ selective estrogen receptor modulators in the<br />

forced swim test. Pharmacol Biochem Behav. 78 (2004) pp. 523-9.<br />

300<br />

300<br />

200<br />

*<br />

200<br />

100<br />

vehicle<br />

control<br />

100<br />

vehicle<br />

control<br />

0<br />

300<br />

0.3 0.6 0.9 1.8<br />

0<br />

300<br />

0.3 0.6 0.9 1.8<br />

200<br />

*<br />

200<br />

100<br />

vehicle<br />

control<br />

100<br />

vehicle<br />

control<br />

0<br />

300<br />

0.3 0.6 0.9 1.8<br />

0<br />

300<br />

0.3 0.6 0.9 1.8<br />

200<br />

200<br />

* * * *<br />

100<br />

vehicle<br />

control<br />

100<br />

vehicle<br />

control<br />

Figure 1: * vs. vehicle<br />

(p


Posters’ Exhibition:<br />

Neuroactive steroids and neurogenesis<br />

• Alias A.G. (USA) Does 5α-reductase stimulation improve cognitive functions,<br />

while inhibition improve immunity?<br />

• Eser D., Schüle C., Romeo E., Uzunov DP., di Michele F., Baghai T.C., Pasini A.,<br />

Schwarz M. and R Rupprecht (Germany) Influence of mirtazapine on plasma<br />

concentrations of neuroactive steroids in major depression and on 3αhydroxysteroid<br />

dehydrogenase activity<br />

• Hill M., Cibula D., Včelaková H, Kancheva L. and Pařízek A. (Czech Republic)<br />

Pregnanolone isomers and their polar conjugates in late pregnancy: A longitudinal<br />

study<br />

• Kancheva L , Včelaková H, Hill M., Vrbíková J. and Stárka L (Czech Republic)<br />

Neuroactive steroids in adult men<br />

• Bo E., Casella D., Martini M., Viglietti-Panzica C., Deviche P., Panzica G.C.<br />

(Italy) Photoperiod influences aromatase expression in junco hyemalis<br />

prosencephalon<br />

• Pawluski J.L., Walker C.A., Galea L.A.M. (Canada) Adult hippocampal<br />

neurogenesis is altered with maternal experience<br />

• Rahman M, Lindblad C., Johansson I-M, Bäckström T and Wang M-D (Sweden)<br />

Neurosteroid modulation of recombinant rat α 5 β 2 γ 2l and α 1 β 2 γ 2L GABA A receptors<br />

in xenopus oocyte


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

DOES 5α-REDUCTASE STIMULATION IMPROVE COGNITIVE FUNCTIONS,<br />

WHILE INHIBITION IMPROVE IMMUNITY?<br />

A.G. Alias<br />

Fulton State Hospital, Fulton, MO 65251, USA. Fax: +01- 573-441-9666<br />

e-mail: aliasag@yahoo.com<br />

The two hypotheses presented here may seem too good to be true. And it may well be so. If<br />

not, these hypotheses have substantial utility. In any case, these simplified versions need to<br />

be reformulated as more research results are accumulated, or existing ones unbeknownst to<br />

the author are incorporated. These hypotheses have been generated by the author’s<br />

empirical research of over 30 years. Those findings are summarized in two papers [1,2].<br />

Steroid 5α-reductase (5αR) stimulates the conversion of steroids such as<br />

testosterone (T) and progesterone (P) to dihydrotestosterone (DHT) and<br />

dihydroprogesterone (DHP), respectively [11,13,23]. DHT and DHP are further reduced by<br />

3α-hydroxysteroid dehydrogenase (3αHSD) to 3α-androstanediol (A-diol), and 3α,5αtetrahydroprogesterone<br />

(THP/ “Allo”) [11,13]. Such metabolic activities take place in<br />

central and peripheral nervous system sites as well [11,13].<br />

Frye et al [12] demonstrated that “DHT has cognitive enhancing effects,<br />

independent of estradiol, which are attenuated by a [3αHSD] inhibitor, indomethacin [by<br />

inhibiting<br />

A-diol formation from DHT].” In yet another animal model [5], T, but not nonaromatizable<br />

DHT, improved working memory, however. And Brinton et al [6]<br />

demonstrated that THP promoted “neurogenesis in vitro and in vivo in transgenic mouse<br />

model of Alzheimer's disease [AD].” And THP is decreased in the prefrontal cortex of AD<br />

patients; the “levels are inversely correlated with neuropathological disease stage” [21]. P<br />

has neuroprotective properties in experimental models of neurodegeneration [12,17,18].<br />

“[P] increased the levels of DHP and THP in plasma and hippocampus and prevented<br />

kainic-acid-induced neuronal loss” [8]. By contrast, the synthetic progestin<br />

medroxyprogesterone acetate (Provera) failed to mimic P in this experiment [8]. 5αR<br />

inhibitor finasteride blocked the increase in DHP and THP in plasma and hippocampus,<br />

following P administration, and also abolished the neuroprotective effect of P [8]. Further,<br />

indomethacin blocked the neuroprotective effect of both DHP and THP [8]. Nevertheless,<br />

indomethacin, along with other non-steroidal antiinflammatory agents (NSAIAs), has been<br />

used in the treatment of AD [17], though many “long-term, placebo-controlled clinical<br />

trials … produced negative results” [17]. A mechanism for the beneficial effects of<br />

NSAIDs against AD is thought to be “an allosteric modulation of gamma-secretase<br />

activity, the enzyme responsible for the formation of amyloid-beta” which is considered to<br />

be “independent from the anti-cyclooxygenase activity [of NSAIDs] and is related to the<br />

chemical structure of the compounds, with some NSAIDs being active (ibuprofen,<br />

sulindac, flurbiprofen, indomethacin, diclofenac) and others not (naproxen, aspirin,<br />

celecoxib)” [17].<br />

It has been known that substantial differences exist in innate immune functions<br />

between the sexes [3,7]. These differences have been (largely) attributed to sex hormonal<br />

influences [3,7]. Though estrogens are believed to potentiate immune functions and<br />

androgens to suppress them, the influences of sex hormones on immune functions are more<br />

complex, as are on cognitive functions. A well-known example to psychiatrists is the<br />

beneficial effects of female gender, as well as of estradiol [19], in the course and severity<br />

of schizophrenia, whereas, the prevalence of excess anxiety is nearly twice as common in<br />

147


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

women of menstruating age. And a single sublingual dose of T inhibits fear induced startle<br />

response in women [16].<br />

Anabolic hormones have been used as immune enhancing agents. “Ghione [14] was<br />

the first to demonstrate an ‘anti-infective’ action of 4-chlorotestosterone in experimental<br />

infection by Nocardia asteroids in rabbits and mice, and in experimental staphylococcal<br />

infection of mice” [22]. Indeed, more recently, addition of an anabolic agent, in addition to<br />

good nutrition, has been advocated in wound healing [9]. This should follow that T, with<br />

its anabolic properties, should have immune enhancing properties. However, numerous<br />

studies have shown otherwise [3,7]. And yet, the same group [7] later identified, and<br />

confimed, DHT as the crucial immune depressing hormone after trauma-hemorrhage [7-<br />

see Fig. 3]. More specifically, DHT suppresses interleukin-4 and gamma-interferon [4].<br />

And, the powerful immunosuppressant, cyclosporin A, stimulates 5αR [10]. Furthermore,<br />

Gilliver et al [15] demonstrated that “systemic [5αR inhibition mimicked] the effects of<br />

castration in a rat model of cutaneous wound healing” [15].<br />

Thus it is reasonable to hypothesize that 5αR stimulation could enhance certain<br />

cognitive functions, and that.5αR inhibition could enhance certain immune functions.<br />

References list<br />

1. Alias AG. Schizotypy and leadership: a contrasting model for deficit symptoms, and a possible therapeutic<br />

role for sex hormones. Med Hypotheses 2000;54:537-552.<br />

2. Alias AG. A role for 5alpha-reductase activity in the development of male homosexuality? Ann N Y Acad<br />

Sci 2004;1032:237-44.<br />

3. Angele MK, Ayala A, Cioffi WG, Bland KI, Chaudry IH. Testosterone: the culprit for producing<br />

splenocyte immune depression after trauma-hemorrhage. Am J Physiol 1998;274:C1530-1536.<br />

4. Araneo BA, Dowell T, Diegel M, Daynes RA. [DHT] exerts a depressive influence on the production of<br />

interleukin-4 (IL-4), IL-5, and gamma-interferon, but not IL-2 by activated murine T cells. Blood<br />

1991;78:688-699.<br />

5. Bimonte-Nelson HA, Singleton RS, Nelson ME, et al. [T], but not nonaromatizable [DHT], improves<br />

working memory ... nerve growth factor levels in aged male rats.. Exp Neurol 2003;181:301-12.<br />

6. Brinton RD, Wang JM. ... therapeutic potential of allopregnanolone to promote neurogenesis in vitro and<br />

in vivo in transgenic mouse model of Alzheimer's disease. Curr Alzheimer Res 2006;3:11-17.<br />

7. Choudhry MA, Bland KI, Chaudry IH. Gender and susceptibility to sepsis following trauma. Endocr<br />

Metab Immune Disord Drug Targets 2006;6:127-<strong>35</strong>.<br />

8. Ciriza I, Carrero P, Frye CA, Garcia-Segura LM. J Neurobiol 2006;66:916-28.<br />

9. Collins, 2004. N. The right mix: using nutritional interventions and an anabolic agent to manage a stage IV<br />

ulcer. Adv Skin Wound Care 2004;17:36,38-39.<br />

10. Cutolo M, Giusti M, Villaggio B, et al. Testosterone metabolism and cyclosporin A treatment in<br />

rheumatoid arthritis. Br J Rheumatol 1997;36:433-439.<br />

11. Frye C.A. Some rewarding effects of androgens may be mediated by actions of its 5α-reduced metabolite<br />

3α-androstanediol. Pharm Biochem Behav (2006/07 – in press).<br />

12. Frye CA, Edi nger KL, Seliga AM, Wawrzycki JM. Psychoneuroendocrinology 2004;29:1019-1027.<br />

13. Garcia-Segura LM, Melcangi RC. Steroids and glial cell function. Glia 2006;54:485-498.<br />

14. Ghione M. Antinfective action of an anabolic steroid. Proc Soc Exp Biol Med 1958;97:773-775.<br />

15. Gilliver SC, Ashworth JJ, Mills SJ, Hardman MJ, Ashcroft GS. J Cell Science 2006;119:722-732.<br />

16. Hermans EJ, Putman P, Baas JM, et al. Biol Psychiatry 2006;59:872-874.<br />

17. Imbimbo BP. The potential role of non-steroidal anti-inflammatory drugs in treating Alzheimer's disease.<br />

Expert Opin Investig Drugs 2004;13:1469-8141.<br />

18. Koenig HL, Schumacher M, Ferzaz B, et al. Science 1995;268:1500-1503.<br />

19. Kulkarni J, Riedel A, de Castella AR, et al. Schizophr Research 2001;48:137-144.<br />

20. Marx CE, Trost WT, Shampine LJ, et al. The Neurosteroid Allopregnanolone Is Reduced in Prefrontal<br />

Cortex in Alzheimer’s Disease. Biol psychiatry 2006;60:1287-1294.<br />

21. Melcangi RC, Garcia-Segura LM. Therapeutic approaches to peripheral neuropathy based on neuroactive<br />

steroids. Expert Rev Neurotherapeutics 2006;6:1121-1125.<br />

22. Tolentino P. Androgens and antibody formation. Pharmacol and Therap 1975;1:209-216.<br />

23. Wilson JD, Griffin JE, Russell DW. Steroid 5α–reductase 2 deficiency. Endocr Rev 1993;14:577-593.<br />

148


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

INFLUENCE OF MIRTAZAPINE ON PLASMA CONCENTRATIONS OF<br />

NEUROACTIVE STEROIDS IN MAJOR DEPRESSION AND ON 3α-<br />

HYDROXYSTEROID DEHYDROGENASE ACTIVITY<br />

Eser D. 1 , Schüle C. 1 , Romeo E. 2 , Uzunov DP. 3 , di Michele F. 2 , Baghai T.C. 1 , Pasini A.<br />

2 , Schwarz M. 1 and R Rupprecht 1<br />

1 Department of Psychiatry and Psychotherapy, Ludwig-Maximilian-University,<br />

Nussbaumstr. 7, 80336 Munich, Germany<br />

2 IRCCS Santa Lucia, Tor Vergata University, Via Ardeatina 306, 00179 Rome, Italy<br />

3 Novartis Institutes for BioMedical Research, Neuroscience Research, Novartis Pharma<br />

AG, WSJ-386.3.26, CH-4002 Basel, Switzerland<br />

E-mail: Daniela.Eser@med.uni-muenchen.de, Tel.: ++49-89-5160-5876, Fax: ++49-89-<br />

5160-5391<br />

Concentrations of 3α-reduced neuroactive steroids are altered in depression and normalize<br />

after antidepressant pharmacotherapy with SSRIs. We investigated the impact of<br />

mirtazapine on the activity of a key neurosteroidogenic enzyme, the 3α-hydroxysteroid<br />

dehydrogenase (3α-HSD), and on the levels of neuroactive steroids in relation to clinical<br />

response. Twenty-three drug-free inpatients suffering from major depression (DSM-IV<br />

criteria) underwent 5-week treatment with mirtazapine (45 mg/day). Plasma samples were<br />

taken weekly at 8:00 AM and quantified for neuroactive steroids by means of combined<br />

gas chromatography/mass spectrometry analysis. Enzyme activity was determined by<br />

assessment of steroid conversion rates. Irrespectively of clinical outcome, there were<br />

significant increases in 3α-reduced neuroactive steroids after mirtazapine treatment,<br />

whereas 3β-reduced steroids were significantly decreased. In-vitro investigations<br />

demonstrated a dose-dependent inhibitory effect of mirtazapine on the activity of the<br />

microsomal 3α-HSD in the oxidative direction, which is compatible with an enhanced<br />

formation of 3α-reduced neuroactive steroids. However, the changes in neuroactive steroid<br />

concentrations more likely reflect direct pharmacological effects of this antidepressant<br />

rather than clinical improvement in general.<br />

149


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

PREGNANOLONE ISOMERS AND THEIR POLAR CONJUGATES IN LATE<br />

PREGNANCY: A LONGITUDINAL STUDY<br />

Hill M. 1 , Cibula D. 2 , Včelaková H, Kancheva L. 1 and Pařízek A. 2<br />

1 Institute of Endocrinology, Narodni triad 8, Prague 11694, Czech Republic<br />

2 Clinic of Gynecology and Obstetrics, General Faculty Hospital, Prague<br />

Pregnanolone isomers (PIs) are known as modulators of GABA A receptors (GABA A -r).<br />

The 3α-PIs operate as the positive modulators and the 3β-PIs are their inactive competitors<br />

for the receptors. The polar conjugates of PIs negatively modulate the GABA A -r and they<br />

are also active on NMDA-receptors depending on the position of C5 hydrogen. Free 5β-PIs<br />

may also operate on nuclear pregnane X receptors (PXr) influencing uterine contractility<br />

and on calcium channels of type T participating in pain transmission. This study addresses<br />

the question of whether changes in the biosynthesis and metabolism of neuroactive<br />

pregnanolone isomers (PIs) may be associated with the timing of human parturition. Using<br />

the GC-MS, the time pro<strong>file</strong>s of unconjugated pregnenolone isomers (PIs)<br />

allopregnanolone (3α-hydroxy-5α-pregnan-20-one, P3α5α), pregnanolone (3α-hydroxy-<br />

5β-pregnan-20-one, P3α5β), isopregnanolone (3β-hydroxy-5α-pregnan-20-one, P3β5α)<br />

and epipregnanolone (3β-hydroxy-5β-pregnan-20-one, P3β5β), pregnenolone, their polar<br />

conjugates, progesterone, 5α-dihydroprogesterone (P5α), and 5β-dihydroprogesterone<br />

(P5β) were monitored in the circulation of 30 healthy women to obtain data from the 10 th<br />

week before parturition (WBP) up to labor at one-week intervals (Fig. 1). Changes in the<br />

steroid levels were evaluated by two-way ANOVA with WBP and subject as independent<br />

factors. The mean concentrations of free PIs ranged from 2–50 nmol/L (Fig. 1) but their<br />

polar conjugates were in 40–140 fold excess over free PIs (Fig. 2). The decelerating<br />

biosynthesis of the P5β was found from the 7 th WBP till parturition (Fig. 3) but their<br />

escalating sulfation was found in all PIs from the 10 th or 9 th WPB (Fig. 4). The conjugation<br />

capacity for P3α5β in pregnancy is limited. As we recently reported, the ratio of<br />

conjugated to free steroid (C/F) for P3β5α was higher in NPW (150:1) that in the pregnant<br />

women (PW) (60:1). However, the C/F of P3α5β rise in the third trimester (Fig. 2). In<br />

contrast, the C/F was higher in PW for both P3α5α and P3β5α. The ratio for P3α5α was<br />

about 65 in PW but only about 15 in NPW and the ratio for P3β5α being about 80 for PW<br />

dropped to about half value in NPW. The changes in sulfation capacity probably diminish<br />

the difference between P3α5α and P3α5β levels in pregnancy. The accelerating ratio of<br />

conjugated to free P3β5α points to potential importance of the steroid for the timing of<br />

parturition. This conjugate exerts a reverse modulating effect on the GABA A -r than the<br />

unconjugated 3αPIs. The modulation efficiencies of both substances are comparable in<br />

absolute values, however the circulating polar conjugates of P3β5α are in a great excess<br />

over the P3α5α. In contrast to the expectedly inferior physiological role of P3α5β in<br />

NPW, the steroid may participate in sustaining of pregnancy and conversely, its reduced<br />

biosynthesis at accelerating conjugation may contribute to inducement of human<br />

parturition given its effectiveness in uterine relaxation via the PXr-dependent mechanism,<br />

its relative abundance and potency (like P3α5α) to suppress the activity of oxytocin<br />

producing cells via positive modulation of GABA A -r.<br />

The study was supported by grants 1A/8649-5, NR/8991-3, NR/9055-4, of the Internal<br />

Grant Agency of the Czech Ministry of Health and GAČR 303/06/1817.<br />

150


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

P3!5! (nmol/L)<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Week: F=8.58, p


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

NEUROACTIVE STEROIDS IN ADULT MEN<br />

Kancheva L , Včelaková H, Hill M., Vrbíková J. and Stárka L<br />

Institute of Endocrinology, Steroid Hormone Unit, Národní 8, Prague 11694, Czech<br />

Republic, fax +420224905325, email lkantcheva@endo.cz<br />

Pregnane steroids (PS), and androstane metabolites (AM) modulate ionotropic receptors.<br />

Women produce significant amounts of neuroactive progesterone metabolites. The steroid<br />

neuromodulators in men are mostly of testicular origin. The 3-oxo-4-ene androgens are<br />

converted to their 3α- and 3β-hydroxy-5α/5β-reduced metabolites (Fig. 1). The<br />

neuromodulating effects of AM and PS prompted us to follow their circulating levels to<br />

estimate metabolic pathways in periphery that may implicate brain concentrations of the<br />

steroids. Accordingly, the levels of 20 steroids and 16 steroid polar conjugates including<br />

17-oxo- and 17β-hydroxy-derivatives of 5α/β-androstane-3α/β-hydroxy-AM were<br />

quantified in 15 men (16-62 years) using GC-MS (Tab. 1).<br />

The levels of AM were in excess over the respective PS, which pointed to the potential<br />

physiological effect of AM in men. The levels of conjugated AM were by 2 or 3 orders of<br />

magnitude higher compared to free steroids (Tab. 1). The neuroinhibitory unconjugated<br />

3α-AM represent only a minute fragment from the pool of the circulating AM.<br />

OSO 2 O<br />

OSO 2 O<br />

OSO 2 O<br />

X<br />

X<br />

X<br />

OSO 2 O<br />

H<br />

H O<br />

H<br />

O<br />

H<br />

H O<br />

H<br />

OSO 2 O<br />

H<br />

X :...=O...Androsterone/3!-s ulf ate<br />

X:...-OH ...5!-Androstane-3!,17"-diol /3!-sulf ate<br />

17"-sulf ate /3!,17"-dis ulf ate<br />

X:...=O...5!-Androstane-3,17-dione<br />

X:...-OH...5!-D ihy drotestosterone/17"-s ulf ate<br />

X :...=O...Epiandrosterone/3"-sulf ate<br />

X:...-OH ...5!-Andros tane-3",17"-diol /3"-s ulf ate<br />

17"-s ulf ate /3",17"-disulf ate<br />

X<br />

OSO 2 O<br />

O<br />

X:...=O...Androstenedione<br />

X:...-OH ...Testosterone/17"-s ulf ate<br />

OSO 2 O OSO 2 O OSO 2 O<br />

X<br />

X<br />

X<br />

OSO 2 O<br />

H<br />

H O<br />

H<br />

O<br />

H<br />

H O<br />

H<br />

OSO 2 O<br />

H<br />

X :...=O...Etiocholanolone/3!-s ulf ate<br />

X:...-OH ...5"-Androstane-3!,17"-diol /3!-sulf ate<br />

17"-sulf ate /3!,17"-disulf ate<br />

X:...=O...5"-Androstane-3,17-dione<br />

X:...-OH...5"-D ihy drotestosterone/17"-s ulf ate<br />

X :...=O...Epietioc holanolone/3"-sulf ate<br />

X:...-OH ...5"-Andros tane-3",17"-diol /3"-s ulf ate<br />

17"-sulf ate /3",17"-disulf ate<br />

FIGURE 1. Biosynthesis of androstane metabolites<br />

152


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

Table 1. Levels of neuroactive androstane and pregnane steroids including their precursors and<br />

ratios of conjugates to free steroids in the serum of adult men (16-65 years old)<br />

Unconjugated<br />

steroids<br />

(n=15)<br />

[nmol/L]<br />

Steroid polar<br />

conjugates<br />

(n=10)<br />

[nmol/L]<br />

Conjugates/free<br />

steroids<br />

(n=10)<br />

[nmol/L]<br />

Median<br />

Lower<br />

quartile<br />

Upper<br />

quartile<br />

Median<br />

Lower<br />

quartile<br />

Upper<br />

quartile<br />

Median<br />

Lower<br />

quartile<br />

Upper<br />

quartile<br />

Steroid<br />

Pregnenolone 2.19 1.63 2.89 292 216 424 163 110 226<br />

17-Hydroxy-pregnenolone 2.98 2.54 8.74 --- --- --- --- --- ---<br />

Dehydroepiandrosterone 13.8 10.0 17.1 6848 3200 8981 593 280 1080<br />

5-Androstene-3 ! , 17 ! -diol 2.18 1.78 3.21 1840 1367 2300 1002 731 1722<br />

Testosterone 21.7 19.0 31.9 --- --- --- --- --- ---<br />

Androstenedione 5.43 4.32 7.74 --- --- --- --- --- ---<br />

5 " -Androstane-3, 17-dione (A5 " ) 0.366 0.195 0.700 --- --- --- --- --- ---<br />

Androsterone (A3 " 5 " ) 0.649 0.539 0.889 25<strong>35</strong> 1387 3293 4173 3185 6114<br />

Epiandrosterone (A3 ! 5 " ) 0.212 0.1<strong>35</strong> 0.313 510 3<strong>35</strong> 654 2941 2658 4198<br />

Etiocholanolone (A3 " 5 ! ) 0.510 0.189 0.971 166 76 242 346 158 475<br />

Epietiocholanolone (A3 ! 5 ! ) 0.017 0.011 0.067 65 45 140 3339 2383 5060<br />

5 " -Dihydrotestosterone 1.36 1.12 1.65 13.9 11.9 20.9 12.3 8.9 13.1<br />

5 " -Androstane-3 " , 17 ! -diol (A3 " 5 " 17 ! ) 0.475 0.<strong>35</strong>2 0.548 154 133 199 347 201 521<br />

5 " -Androstane-3 ! , 17 ! -diol (A3 ! 5 " 17 ! ) 0.149 0.057 0.251 246 149 346 2199 747 5723<br />

5 ! -Androstane-3 " , 17 ! -diol (A3 " 5 ! 17 ! ) 0.067 0.045 0.120 70.7 17.8 96.1 803 630 1434<br />

5 ! -Androstane-3 ! , 17 ! -diol (A3 ! 5 ! 17 ! ) 0.085 0.062 0.171 10.8 8.7 13.7 120 78 347<br />

Allopregnanolone (P3 " 5 " ) 0.341 0.154 0.399 9.0 6.8 14.4 50.4 20.6 93.4<br />

Isopregnanolone (P3 ! 5 " ) 0.233 0.177 0.367 15.0 12.7 26.6 76.0 44.6 124.4<br />

Pregnanolone (P3 " 5 ! ) --- --- --- 38.4 24.6 51.6 --- --- ---<br />

Epipregnanolone (P3 ! 5 ! ) --- --- --- 6.94 3.39 8.89 --- --- ---<br />

The ratios of conjugates to free steroids in AM were by 1-2 orders of magnitude higher<br />

compared to values found for the corresponding PS. The neuroinhibiting reduced 3α-5α/β-<br />

AM prevailed over the inactive 3β-derivatives. Strong correlations were detected between<br />

3α-, 3-oxo- and 3β- derivatives in both AM and PS. A5α correlated with A3α5α (r=0.604,<br />

p


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

PHOTOPERIOD INFLUENCES AROMATASE EXPRESSION IN JUNCO<br />

HYEMALIS PROSENCEPHALON<br />

Bo E., Casella D., Martini M., Viglietti-Panzica C., Deviche P.*, Panzica G.C.<br />

Dep.Anatomy, Pharmacology and Forensic Medicine, University of Turin, Italy.<br />

*School of Life Sciences, Arizona State University, Phoenix, AZ, USA<br />

Aromatase (ARO) is produced by the CYP19 gene and it’s responsible of estradiol (E2)<br />

biosynthesis from testosterone (T). In the avian brain, this enzyme is localized in regions<br />

controlling sexual behavior, as the POM (preoptic medial nucleus), the BnST (bed nucleus<br />

of the stria terminalis), the VMN (ventromedial nucleus) and the medial amygdala. In<br />

passerine birds, ARO was found also in hippocampus and in HVC (High Vocal Center).<br />

ARO expression in male galliforms is related to circulating levels of T, it almost<br />

disappears in castrated quails and its expression is restored to the level of intact males<br />

when castrated birds receive exogenous T [1,2]. Gonadal hormones’ levels, in particular T,<br />

fluctuate in wild birds when exposed to different seasonal conditions, in particular they<br />

may change according to the lenght of the photoperiod, paralleling changes of gonadal<br />

functions.<br />

Studies on the effects of seasonal changes on ARO system of wild birds are rare [3,5], and<br />

never performed on a migratory species. The aim of the present study was to investigate<br />

brain ARO immunoreactivity distribution in a migratory songbird, Junko hyemalis. In<br />

addition, we have also investigated changes in limbic and hypothalamic distribution of<br />

ARO according to the exposure of these birds to different photoperiods.<br />

Birds were collected from a local population in Fairbanks, Alaska (65°N, 148°W), in<br />

different periods and were divided in two groups:the first one (not photostimulated, no-<br />

PHOT) includes three animals captured in September at the age of three months and<br />

exposed to a short photoperiod (8h of light and 16h of dark) for 8 months. The second one<br />

(photostimulated, PHOT) includes three one year old animals captured in May (natural<br />

photoperiod of 20h of light and 4h of dark) and sacrificed immediately after the capture.<br />

The animals were perfused and processed for the immunocitochemistry using the Harada<br />

antibody to detect ARO immunoreactivity. To identificate brain nuclei we used the Serinus<br />

Canaria atlas with modifications for the Junco hyemalis [4]. The sections were analyzed to<br />

count the cell number in selected regions and the results were statistically analyzed.<br />

In PHOT birds, ARO positive neurons were observed in POM, BnST, VMN, medial<br />

amygdala, hippocampus, nidopallium, nidopallium caudalis, and reticular formation of the<br />

mesencephalon. We have quantitatively analyzed hypothalamic and limbic nuclei that are<br />

related to reproductive behavior in birds (i.e., POM, BnST, and VMN).<br />

In all these structures we observed a strongly significant decrease of the number of AROimmunoreactive<br />

(ir) elements in no-PHOT animals, comparable to the decrease observed<br />

in castrated male quail.<br />

It seems therefore that, as demonstrated in quail [1], in Junko there is a small population of<br />

ARO-ir cells that are insensitive to the fluctuations of circulating T levels. We can<br />

hypothesizes that, when photoperiod increases and circulating T levels raise, these<br />

elements start to locally convert T into E2 stimulating at the same time the surrounding<br />

cells to produce ARO.<br />

Acknowledgements. This work was supported by grants from PRIN, Università di Torino,<br />

Fondazione CRT, and Regione Piemonte<br />

154


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

Reference list<br />

[1] N. Aste, G.C. Panzica, P. Aimar, C. Viglietti-Panzica, N. Harada, A. Foidart and J.<br />

Balthazart, Morphometric studies demonstrate that aromatase-immunoreactive cells<br />

are the main target of androgens and estrogens in the quail medial preoptic nucleus,<br />

Exp. Brain Research 101 (1994) 241-252.<br />

[2] N. Aste, G.C. Panzica, C. Viglietti-Panzica, N. Harada and J. Balthazart,<br />

Distribution and effects of testosterone on aromatase mRNA in the quail forebrain:<br />

a non-radioactive in situ hybridization study, Journal of Chemical Neuroanatomy<br />

14 (1998) 103-115.<br />

[3] A. Foidart, B. Silverin, M. Baillen, N. Harada and J. Balthazart, Neuroanatomical<br />

distribution and seasonal variations of aromatase activity and aromataseimmunoreactive<br />

cells in the Pied Flycatcher (Ficedula hypoleuca), Hormones and<br />

Behavior 33 (1998) 180-196.<br />

[4] G.C. Panzica, L. Plumari, E. Garcia-Ojeda and P. Deviche, Central vasotocinimmunoreactive<br />

system in a male passerine bird (Junco hyemalis), Journal of<br />

Comparative Neurology 409 (1999) 105-117.<br />

[5] L.V. Riters, M. Eens, R. Pinxten, D.L. Duffy, J. Balthazart and G.F. Ball, Seasonal<br />

changes in courtship song and the medial preoptic area in male European starlings<br />

(Sturnus vulgaris), Hormones and Behavior 38 (2000) 250-261.<br />

155


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

ADULT HIPPOCAMPAL NEUROGENESIS IS ALTERED WITH MATERNAL<br />

EXPERIENCE<br />

Pawluski J.L., Walker C.A., Galea L.A.M.<br />

Program in Neuroscience, Department of Psychology and Brain Research Centre,<br />

University of British Columbia, 2136 West Mall, Vancouver B.C., V6T 1Z4, CANADA,.<br />

Fax: 001-604-822- 6923, email: jodi@psych.ubc.ca<br />

Adult neurogenesis in the dentate gyrus of the hippocampus is influenced by steroid<br />

hormones (estradiol and corticosterone), which fluctuate during the estrous cycle,<br />

pregnancy and lactation. Pregnancy and lactation have been demonstrated to be a time of<br />

maximal neural and behavioural plasticity as a host of learning and memory processes are<br />

activated in the mother to ensure the acquisition of adequate maternal care and<br />

reproductive success. Recent work has shown that motherhood differentially affects<br />

hippocampus-dependent learning and memory performance [1] and hippocampal<br />

morphology [2] and these effects differ with reproductive experience (number of times<br />

pregnant and given birth). Thus, it is possible that hippocampal neurogenesis may also be<br />

affected by reproductive experience. However, very little research has investigated the role<br />

of motherhood on hippocampal neurogenesis. The present study aimed to thoroughly<br />

investigate the role of motherhood and/or pup exposure on hippocampal neurogenesis via<br />

cell proliferation and cell survival. Four groups of female Sprague-Dawley rats were used;<br />

multiparous, primiparous, nulliparous, and sensitized (pup-exposed nulliparous females).<br />

All rats were injected with BrdU (200 mg/kg) 24 hours after birth/pup-exposure with agematched<br />

controls. Rats were perfused either 24 hours (cell proliferation) or 21 days (cell<br />

survival) after injection. Parous/sensitized rats remained with pups until perfusion. Results<br />

show there is a significant decrease in BrdU-labeled cells in the dentate gyrus surviving<br />

throughout lactation in primiparous dams compared to multiparous, nulliparous and<br />

sensitized rats. Interestingly this effect appears to be independent of pup-exposure. In<br />

addition, multiparous rats have a greater percentage of cells surviving throughout lactation<br />

compared to all other groups. Future research aims to determine the hormonal mechanisms<br />

mediating these changes.<br />

Reference list<br />

1. Pawluski, J.L., Walker, S.K., and Galea, L.A.M (2006). Reproductive experience differentially<br />

affects spatial reference and working memory performance in the mother. Hormones and Behavior.<br />

49(2):143-9.<br />

2. Pawluski, J.L. and Galea L.A.M. (2006). Hippocampal morphology is differentially affected by<br />

reproductive experience. Journal of Neurobiology. 66(1):71-81.<br />

156


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

PRIMARY CELL CULTURES FROM FETAL BOVINE BRAIN: AN IN VITRO<br />

MODEL TO STUDY NEUROACTIVE STEROIDS<br />

Peruffo A., Buson G. Cozzi B. and Ballarin C.<br />

Department of Experimental Veterinary Science, University of Padova,<br />

Viale dell’Università 16, <strong>35</strong>020 Legnaro – Agripolis (PD), ITALY<br />

Phone +39.049.8272626, Fax +39.049.8272669, e-mail: antonella.peruffo@unipd.it<br />

Primary cell cultures represent a simplified experimental model useful to control and<br />

modulate media’s composition and study the effects of neuroactive steroids on neural cells.<br />

We set up a procedure to obtain primary cell cultures from fetal bovine brain to study the<br />

mRNA expression and localization of P450 aromatase (P450 AROM ) at the cellular level in<br />

neurons and astrocytes. Aim of the study was to verify if primary cultures from fetal<br />

bovine brain may be a reliable model for the expression of P450 AROM . We chose the bovine<br />

as experimental species for studying the role of neural aromatization in the sexual<br />

differentiation of the central nervous system because of its large brain, extended duration<br />

of gestation and incidence of spontaneous intersex calves.<br />

Hypothalamus and frontal cortex areas were isolated from a series of bovine fetuses of<br />

different developmental stages. Previous published data from our laboratory [1]<br />

demonstrated that tissue fragments obtained from the bovine cerebral cortex and<br />

hypothalamus could be cryopreserved in liquid nitrogen and used for primary cell cultures.<br />

We compared mRNA expression of P450 AROM in both fetal brain tissue and primary cell<br />

cultures harvested from the same cerebral region. Furthermore, we detected the presence<br />

and localization of the enzyme by immunohistochemistry on fetal tissue and by<br />

immunocytochemistry on primary cell cultures.<br />

The mRNA expression of P450 AROM was confirmed using RT-PCR analysis.<br />

Immunohistochemistry performed with an anti P450 AROM antibody was used to identify<br />

immunoreactive neural cell in hypothalamic sections and to study the cellular localization<br />

of the enzyme in cultured neurons and astrocytes by confocal microscopy. Cellular lisates<br />

obtained from cell cultures were analysed by Western blot to detect the P450 AROM protein<br />

encoded by the aromatase transcripts.<br />

Neural cells from primary cultures were able to express P450 AROM mRNA. The enzyme<br />

encoded by transcripts was detected by Western blot and its localization in both neurons<br />

and astrocytes was confirmed by immunocytochemistry. The presence of P450 AROM in<br />

neuron and astrocytes was observed in rat cerebral cortex [2] and recently also in the<br />

human temporal cortex [3].<br />

We conclude that primary cultures from fetal bovine brain could represent a good in vitro<br />

model for future investigations on P450 AROM expression, activity and regulation.<br />

Reference List<br />

1. Peruffo, A., Massimino, M.L., Ballarin, C., Carmignoto, G., Rota, A., Cozzi, B., 2004. Primary cultures<br />

from fetal bovine brain. Neuroreport 15, 1719-1722.<br />

2. Zwain, I.H. and Yen, S.S.C. 1999. Neurosteroidogenesis in astrocytes, oligodendrocytes, and neurons of<br />

cerebral cortex of rat brain. Endocrinology 140, 3843-3852.<br />

3. Yague., J.G., Munoz, A., de Monasterio-Schrader, P., Defelipe, J., Garcia-Segura, L.M., Azcoitia, I.<br />

2006. Aromatase expression in the human temporal cortex. Neuroscience 138, 389-401.<br />

157


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

NEUROSTEROID MODULATION OF RECOMBINANT RAT α 5 β 2 γ 2L AND α 1 β 2 γ 2L<br />

GABA A RECEPTORS IN XENOPUS OOCYTE<br />

Rahman M., Lindblad C., Johansson I-M, Bäckström T. and Wang M-D<br />

Umeå Neurosteroid Research Center, Department of Clinical Science, Obstetrics and Gynecology, Umeå<br />

University, S-901 85 Umeå, Sweden. Phone: +46 90 785 3323 Fax: +46 90 77 60 06;<br />

E-mail: mingde.wang@obgyn.umu.se<br />

GABA A receptors containing α 5 -subunit have an important role in cognitive function. As<br />

the potentiating effect of 3α-hydroxy ring-A reduced steroids depends on subunit<br />

combinations of GABA A receptors, the antagonistic effect of 3β-hydroxypregnane steroids<br />

may vary between α 5 -subunit and α 1 -subunit containing receptors. We investigated the<br />

effect of steroid agonists and antagonists in recombinant α 1 β 2 γ 2L and α 5 β 2 γ 2L receptors<br />

expressed in Xenopus oocytes using a two electrodes voltage-clamp technique. We did not<br />

find any significant difference in potency and efficacy of GABA response between the<br />

α 1 β 2 γ 2L and α 5 β 2 γ 2L receptor. Compared to the α 1 β 2 γ 2L receptor, a significantly lower<br />

degree of desensitization was observed in the α 5 β 2 γ 2L receptor. In addition, the potency of<br />

3α-OH-5α-pregnan-20-one (3α5αP); 5α-pregnan-3α,21-diol-20-one (3α5αTHDOC) and<br />

5α-androstane-3α,17β-diol (3α5αADL) to enhance GABA response was significantly<br />

higher in the α 5 β 2 γ 2L receptor, whereas their efficacy remained unchanged between the<br />

receptors. In either receptor, the efficacy of the 3α5αTHDOC was significantly higher than<br />

that of 3α5αP and 3α5αADL. The efficacy of 5β-pregnan-3β,21-diol-20-one(UC1015)<br />

and 5α-pregnan-3β,20α-diol(UC1019) to inhibit GABA response and 5β-pregnan-3β, 20βdiol<br />

(UC1020) to inhibit 3α5αTHDOC enhanced GABA response were higher in the<br />

α 5 β 2 γ 2L receptor compared to α 1 β 2 γ 2L receptor. The potencies of 3β-hydroxypregnane<br />

steroids to inhibit GABA response or 3α5αTHDOC enhanced responses did not vary<br />

between the α 1 β 2 γ 2L and α 5 β 2 γ 2L receptors. Interestingly, the potencies and efficacies of<br />

3β-hydroxy-pregnane steroids to inhibit GABA response were positively correlated to<br />

potencies and efficacies inhibiting 3α5αTHDOC enhanced GABA response. Results from<br />

the current study revealed a different pattern of steroid modulation in the rat α 1 β 2 γ 2L and<br />

α 5 β 2 γ 2L receptors.<br />

Reference list<br />

Rahman, M., Lindblad, C., Johansson, I.M., Backstrom, T. and Wang, M.D., Neurosteroid modulation of<br />

recombinant rat alpha(5)beta(2)gamma(2L) and alpha(1)beta(2)gamma(2L) GABA(A) receptors in Xenopus<br />

oocyte, Eur J Pharmacol, 547 (2006) 37-44.<br />

158


Posters’ Exhibition:<br />

Neuroprotective effects<br />

• Barreto G., Veiga S., Azcoitia I., Garcia-Segura L.M., Garcia-Ovejero D. (Spain)<br />

Testosterone decreases reactive astroglia and reactive microglia after brain injury<br />

in male rats: role of its metabolites estradiol and dihydrotestosterone<br />

• Berumen L.C., Tecozautla A., Sánchez-Ramos M.A., García-Servín M., García-<br />

Alcocer G. (México) Steroid hormone effects on 5-HT 5A serotonin receptor-like<br />

immunolabelling in the rat hippocampus<br />

• F.Biamonte, G.Assenza, R.Marino, D.Caruso, S.Crotti, R.C.Melcangi, R.Cesa,<br />

P.Strata, F.Keller. (Italy) Interaction between estrogens and reelin in purkinje cell<br />

development<br />

• Carroll J.C., Emily R. Rosario, Lilly Chang, Frank Z. Stanczyk, Salvatore Oddo,<br />

Frank M. LaFerla, Christian J. Pike (USA) Progesterone blocks estrogen<br />

regulation of alzheimer-like neuropathology in female 3xTG-AD mice<br />

• Danza G., Cecchi C., Pensalfini A., Formigli L., Nosi D., Stefani M, Liguri G,<br />

Rosati F., Dichiara F., Morello M.,Pieraccini G., Serio M., Peri A. (Italy) The<br />

estrogen-regulated gene Seladin-1/DHCR24 exerts its neuroprotective effects<br />

through membrane cholesterol modulation<br />

• Fargo K.N., Sengelaub D.R. (USA) Androgenic, but not estrogenic, protection of<br />

motoneurons from somal and dendritic atrophy induced by the death of neighboring<br />

motoneurons<br />

• Forsberg M. K., Hallberg M., Nyberg F., Svensson A-L (Sweden) Neuronal<br />

protection of dehydroepiandrosterone on PC12 cells pretreated with β-amyloid<br />

• S. Giatti, I. Roglio, M. Pesaresi, R. Bianchi, G. Cavaletti, L.M. Garcia-Segura, G.<br />

Lauria, R.C. Melcangi (Italy) Progesterone and its derivatives as protective agents<br />

in experimental diabetic neuropathy<br />

• Jarrahi M, Vafaei AA, Rashidy-Pour A (Iran) An evaluation of the effect of<br />

dexamethasone in preventing Tourniquet neurepathy in rats<br />

• Kibaly C., Meyer L., Patte-Mensah C. and Mensah-Nyagan A.G. (France)<br />

Involvement of endogenous dehydroepiandrosterone in the modulation of spinal<br />

nociceptive mechanisms


• Lee J.E., Kim H.J., Kang H.S., Ahn H.S., and Gye M.C. (Korea) Postnatal<br />

changes in the expression of aquaporin 1 and effect of estrogen on the expression<br />

in ovariectomized mouse brain<br />

• I. Roglio, S. Giatti, M. Pesaresi, R. Bianchi, G. Cavaletti, D. Caruso, S, Scurati,<br />

L.M. Garcia-Segura, G. Lauria, R.C. Melcangi (Italy) Testosterone derivatives are<br />

neuroprotective agents in experimental diabetic neuropathy<br />

• Rosario, E.R., Carroll, J.C., Pike, C.J. (USA) Estrogen and androgens regulate<br />

Alzheimer-like neuropathology in male 3xTG-AD mice<br />

• Schaeffer V., Patte-Mensah C., Eckert A., Mensah-Nyagan A.G. (France) Effects<br />

of beta amyloid peptide 1-42 and oxidative stress on neurosteroid formation in<br />

human neuroblastoma cells<br />

• Szegő É.M., Kékesi K.A., Juhász G., Ábrahám I.M. (Hungary) Effect of estrogen<br />

treatment on protein expression pattern in famale mice brain using fluorescent<br />

differential 2-D gel electrophoresis (dige)<br />

• Tapia González S., Diz-Chaves Y., Pernía O., Carrero P., Garcia-Segura L.M.<br />

(Spain) Selective estrogen receptor modulators decrease microglia activation in<br />

the cerebellum of male rats


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

TESTOSTERONE DECREASES REACTIVE ASTROGLIA AND REACTIVE<br />

MICROGLIA AFTER BRAIN INJURY IN MALE RATS: ROLE OF ITS<br />

METABOLITES ESTRADIOL AND DIHYDROTESTOSTERONE<br />

Barreto G. 1* , Veiga S. 1 , Azcoitia I. 2 , Garcia-Segura L.M. 1 , Garcia-Ovejero D. 3<br />

(1) Instituto Cajal, C.S.I.C., Avenida Doctor Arce 37, E-28002 Madrid, Spain.* Email:<br />

gemilio.sampaio@cajal.csic.es FAX: +34-915854754<br />

(2) Departamento de Biología Celular, Facultad de Biología, Universidad Complutense,<br />

Madrid, Spain<br />

(3) Laboratorio de Neuroinflamación, Hospital Nacional de Parapléjicos, Toledo, Spain<br />

Stab wound injury in the brain elicits a complex cascade of events involving glial cells.<br />

Astrocytes and microglia actively react to brain damage, participating in the repair of the<br />

disrupted blood-brain barrier and in the reorganization of the injured neuronal circuits. The<br />

response of astrocytes and microglia to brain injury is modulated by local factors produced<br />

in the injured tissue and also by substances transported by the systemic circulation. Among<br />

other peripheral molecules, the hormones secreted by the gonads exert a modulation of<br />

reactive gliosis. In this study we have assessed the effect of testosterone therapy on gliosis<br />

after a stab wound injury affecting the hippocampal formation. In addition, to determine<br />

whether the effects of testosterone were mediated by its metabolites, some animals were<br />

treated with estradiol or dihydrotestosterone (DHT). Wistar male rats were bilaterally<br />

orchidectomized at the age of 2 months to reduce circulating levels of testicular secretions.<br />

A penetrating injury affecting the hippocampal formation was performed one month after<br />

orchidectomy. The effects of early and late treatments after injury with testosterone or its<br />

metabolites estradiol and DHT were assessed. Therefore, a group of animals received one<br />

subcutaneous injection of testosterone (5 mg/Kg), estradiol (1 mg/Kg) or DHT (5 mg/Kg)<br />

on days 0, 1 and 2 after injury (early steroid administration). A second group of animals<br />

were injected with the same steroids at the same doses on days 5, 6 and 7 after injury<br />

(delayed steroid administration). One week after lesion, animals were killed by fixative<br />

perfusion and the brains processed for immunohistochemistry for vimentin, a marker of<br />

reactive astroglia or MHC-II, a marker of reactive microglia. The number of vimentin<br />

immunoreactive astrocytes was assessed in the hippocampus using the optical disector<br />

method. The volume fraction of MHC-II immunoreactive microglia was estimated<br />

according to the point-counting method of Weibel. Both early and delayed administration<br />

of testosterone resulted in a significant decrease in the number of vimentinimmunoreactive<br />

astrocytes in the studied area (0-345 micrometers from the lateral border<br />

of the wound). Early and delayed treatments with estradiol also resulted in a decrease in<br />

the number of vimentin-immnoreactive astrocytes compared to control values. DHT<br />

administration, either early or delayed, did not affect the number of vimentin<br />

immunoreactive astrocytes. The volume fraction of MHC-II immunoreactive microglia<br />

showed a significant decrease in the animals that received testosterone or estradiol in both<br />

early and delayed treatments. In contrast, the volume fraction of MHC-II immunoreactive<br />

cells was not affected by the delayed administration of DHT. However, early<br />

administration of DHT significantly reduced the volume fraction of MHC-II<br />

immunoreactive cells. These findings indicate that both early and delayed testosterone<br />

administration after a stab wound injury reduce astroglia and microglia reactivity in male<br />

rats. Testosterone may exert its effects on reactive gliosis by acting on androgen receptors<br />

or after local conversion to estradiol. Previous studies have shown that a stab wound injury<br />

induces the expression of aromatase, the enzyme that converts testosterone in estradiol, in<br />

161


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

reactive astrocytes [2]. Furthermore, a stab wound injury induces the expression of<br />

estrogen receptors in reactive astroglia [1]. Therefore, testosterone may be converted into<br />

estradiol in reactive astrocytes and then estradiol may act by a paracrine or autocrine<br />

mechanism on reactive astrocytes expressing estrogen receptors. In addition, estradiol may<br />

indirectly reduce reactive astrogliosis by preventing neuronal death. In agreement with this<br />

possibility, we have detected in the present study that the early or late administration of<br />

estradiol after injury to adult orchidectomized male rats reduces reactive astrogliosis in the<br />

borders of the wound. Therefore, it may be postulated that at least part of the early and late<br />

effects of testosterone on reactive astrogliosis are mediated by its local conversion to<br />

estradiol. Local conversion to estradiol may also in part mediate the effects of testosterone<br />

on microglia. Previous studies have shown that estradiol may reduce microglia activation<br />

in female rodents [3] and our present findings indicate that early or late administration of<br />

estradiol may also reduce microglia activation in male rats. Testosterone may also be<br />

converted in the brain to its reduced metabolite DHT, by the enzyme 5alpha-reductase,<br />

which is also expressed in glial cells. DHT is a potent agonist of androgen receptors and<br />

many of the biological effects of testosterone are mediated by this metabolite. However,<br />

since DHT did not affect the number of vimentin immunoreactive astrocytes, we may<br />

conclude that the effect of testosterone on reactive astrogliosis is not mediated by the<br />

conversion in DHT and the subsequent activation of androgen receptors. Furthermore,<br />

androgen receptors have not been detected in reactive astrocytes in the rat brain. In<br />

contrast, early expression of androgen receptors has been detected in reactive microglia<br />

after a stab wound [1]. Our present findings, indicating that early administration of DHT<br />

reduces the volume fraction of MHC-II immunoreactive cells in the hippocampus in the<br />

proximity of the wound, suggest that part of the early effect of testosterone on reactive<br />

microglia may be mediated by its conversion to DHT and the consecutive action on<br />

androgen receptors. Therefore, we may conclude that early and late effects of testosterone<br />

on reactive astroglia and reactive microglia may be at least in part mediated by estradiol,<br />

while DHT may mediate part of the early effects of testosterone on reactive microglia.<br />

Therefore, our findings suggest that metabolism of testosterone is an essential mechanism<br />

involved in its effects on reactive astroglia and reactive microglia.<br />

Supported by Ministerio de Educación y Ciencia, Spain (SAF 2005-00272) and the<br />

European Union (EWA project: LSHM-CT-2005-518245).<br />

References list<br />

[1] D. Garcia-Ovejero, S. Veiga, L.M. Garcia-Segura and L.L. DonCarlos, Glial<br />

expression of estrogen and androgen receptors after rat brain injury, J Comp Neurol 450<br />

(2002) 256-271.<br />

[2] L.M. Garcia-Segura, S. Veiga, A. Sierra, R.C. Melcangi and I. Azcoitia, Aromatase: a<br />

neuroprotective enzyme, Prog Neurobiol 71 (2003) 31-41.<br />

[3] E. Vegeto, C. Bonincontro, G. Pollio, A. Sala , S. Viappiani, F. Nardi, A. Brusadelli, B.<br />

Viviani, P. Ciana and A. Maggi, Estrogen prevents the lipopolysaccharide-induced<br />

inflammatory response in microglia, J Neurosci 21 (2001) 1809-1818.<br />

162


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

STEROID HORMONE EFFECTS ON 5-HT 5A SEROTONIN RECEPTOR-LIKE<br />

IMMUNOLABELLING IN THE RAT HIPPOCAMPUS<br />

Berumen L.C. 1 , Tecozautla A. 1 , Sánchez-Ramos M.A. 2 , García-Servín M. 3 , García-<br />

Alcocer G. 1<br />

1 Universidad Autónoma de Querétaro, Facultad de Química, Centro Universitario S/N,<br />

Cerro de las Campanas, 76010, Querétaro, México, Fax+52-442-1921302, email:<br />

berumen@uaq.mx. 2 Universidad Autónoma de Querétaro, Facultad de Ciencias Naturales,<br />

Qro. México. 3 Instituto de Neurobiología, Campus UNAM-UAQ, Juriquilla, Qro. México.<br />

The activity of different neurotransmitter systems can be modulated by steroid hormones.<br />

The effect of steroids over serotonergic pathway has been widely studied, by binding of<br />

radioligands, selective receptor antibodies and hibridization probes, and the use of agonist<br />

and antagonist drugs [1, 2]. The partial agonism of classical drugs for receptors, where 5-<br />

hydroxitriptamine (5-HT, serotonin) is their principal agonist, and the finding of different<br />

types of serotonin receptors led us to evaluate the response of 5-HT 5A receptors to steroid<br />

hormones, particularly estradiol and progesterone, using specific antibodies to label the<br />

presence of this protein in the hippocampal CA1 region.<br />

In the present study, thirty rats were ovariectomized and injected with 50µg/Kg 17-betaestradiol<br />

benzoate (E2), 7.5 mg/Kg progesterone (P) or the combination of both steroids;<br />

corn oil was used for ovariectomized controls. The 5-HT 5A –like immunosignal in the<br />

hippocampal CA1 region decreased in the group treated with E2 compared to the<br />

ovariectomized control group, and also decreased in the E2+P supplemented group. In<br />

contrast, the expression of 5-HT 5A receptor in the hippocampus of P-treated rats was not<br />

significantly different from ovariectomized controls. We also examined 5-HT 2C receptor<br />

immunolabelling, and found the contrary tendence, that is, decreased signal in<br />

hippocampus of P-supplemented rats compared to E2- or E2+P-supplemented ones.<br />

Studies have been made in order to understand the discrete effects found over serotonergic<br />

pathways, but the existence of several receptor proteins coupled to different transduction<br />

signals sometimes overlapping or cross-linked makes it complex to analyze [3]. Here we<br />

found that 5-HT 5A receptor responded similar to 5-HT 1 receptor [4], although the latter<br />

receptor is coupled to Gi-protein while 5-HT 5A receptor has been reported to stimulate<br />

adenylate cyclase activity, not yet defined [2]. In conclusion 5-HT 5A receptors participate<br />

in the modulation of serotonin signaling exerted by sexual hormones estradiol and<br />

progesterone.<br />

Acknowledgements<br />

The authors appreciate the work of Berenice Flores and Karina Hernández. This work was<br />

supported by PROMEP/103.5/05/1798.<br />

Reference list<br />

[1] Hoyer, D., Hannon, J.P., Martin, G.R., 2002. Molecular, pharmacological and functional diversity of 5-<br />

HT receptors. Pharm Biochem Beh. 71, 533-554.<br />

[2] Kroeze, W.K., Roth, B.L., 2006. Molecular biology and genomic organization of G protein-coupled<br />

serotonin receptors. In The Serotonin Receptors: from molecular pharmacology to human therapeutics (Roth<br />

BL, ED) Humana Press New Jersey. 1, 1-38.<br />

[3] Mouillet-Richard, S., Pietri, M., Schneider, B., Vidal, C., Mutel, V., Launay, J.M., Kellermann, O., 2005.<br />

Modulation of serotonergic receptor signaling and cross-talk by prion protein. J Biol Chem. 280(6), 4592-<br />

601<br />

[4] Biegon, A., Reches, A., Snyder, L., McEwen, B.S., 1983. Serotonergic and noradrenergic receptors in the<br />

rat brain: modulation by chronic exposure to ovarian hormones. Life Sci. 32(17), 2015-21.<br />

163


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

INTERACTION BETWEEN ESTROGENS AND REELIN IN PURKINJE CELL<br />

DEVELOPMENT<br />

F.Biamonte 1 , G.Assenza 1 , R.Marino 1 , D.Caruso 2 , S.Crotti 2 , R.C.Melcangi 3 , R.Cesa 4 ,<br />

P.Strata 4 , F.Keller 1<br />

1. Lab of Dev Neurosci, Univ Campus Bio-Medico, Roma, Italy. 2. Dept of Pharmacol Sci, Univ of<br />

Milan, Italy 3. Dpt of Endocrin and Ctr of Excellence on Neurodegen Dis, Univ of Milan, Italy. 4.<br />

Rita Levi Montalcini Ctr for Brain Res, Univ of Turin, Italy<br />

Mariani et al. have demonstrated that mutations like reeler and staggerer in the<br />

heterozygous state, leading to haploinsufficiency of the gene product, cause a loss of<br />

Purkinje cells (PC) in the adult mouse cerebellum. Furthermore, the loss of PC is more<br />

prominent in male than female heterozygous mice. In order to test whether the PC loss<br />

may start at earlier stages, we have assessed PC numbers in neonatal cerebella of<br />

heterozygous malea (rl/+ M ) and female (rl/+ F ) vs. wild-type (wt) littermates using<br />

stereological counting methods. At ages between P10 and P18, PC numbers are decreased<br />

in male rl/+ M . Of all cell populations of the cerebellar cortex and their input and output<br />

nuclei, PC appear to be selectively affected, since the total number of neurons in deep<br />

cerebellar nuclei and in the inferior olivary nucleus, as well as the volume of the cerebellar<br />

cortex, are not reduced rl/+ M . PC are well known to be sensitive to gonadal sex steroids,<br />

particularly estrogens, and express the aromatase enzyme converting androgens into<br />

estrogens. In principle, it could be that estrogens protect rl/+ F from reelin<br />

haploinsufficiency or, alternatively, that androgens increase the effect of reelin<br />

haploinsufficiency in rl/+ M . To investigate the interaction between estrogens and reelin in<br />

our animal model, we injected 4-OHtamoxifen (TMX), an estrogen receptor antagonist,<br />

into the cisterna magna of P4 mice of either sex and genotype, and then assessed PC<br />

numbers again at P10-P18. We found that TMX selectively reduces PC numbers in rl/+ F<br />

and wt F , but has no effect in rl/+ M or wt M . To confirm the interaction between reelin and<br />

estrogens, we did similar experiments with 17-β-Estradiol (17βΕ2): 17βΕ2 was found to<br />

increase the number of PC in rl/+ M but had no effect in rl/+ F and wt F . Furthermore, we<br />

started to assess tissue levels of gonadal steroids and their precursors in cerebella of either<br />

sex and genotype at various developmental times, using mass spectrometry. The<br />

preliminary results indicate decreased 17βΕ2 levels and increased testosterone levels in<br />

cerebella of rl/+ M , a finding that is consistent with an aromatase deficit in rl/+ M .<br />

Experiments aimed at testing the effect of androgens are in progress. Our current results<br />

converge toward a model where the same genetic mutation is more penetrant in males than<br />

in females because of the protective effect of estrogens. Our animal model could be<br />

relevant for genetically complex neurodevelopmental disorders, such as autism, where<br />

genetic and hormonal factors have been recently postulated to interact (see e.g.<br />

Knickmeyer et al., Horm Behav 49:282-292, 2006).<br />

This work was supported by grants from the Fondation Jerome Lejeune, from<br />

NAAR-AUTISM SPEAKS (Grant number 1391)<br />

164


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

PROGESTERONE BLOCKS ESTROGEN REGULATION OF ALZHEIMER-<br />

LIKE NEUROPATHOLOGY IN FEMALE 3XTG-AD MICE<br />

Carroll J.C. a,b , Rosario E.R. a,b , Chang L. c ,. Stanczyk F.Z. c , Oddo S. d , LaFerla F.M. d ,<br />

and Pike C.J. b<br />

a Neuroscience Graduate Program, University of Southern California, Los Angeles, CA,<br />

b Davis School of Gerontology, c Department of Obstetrics and Gynecology, d Department of<br />

Neurobiology and Behavior, University of California Irvine, Irvine, CA<br />

cjpike@usc.edu<br />

fax: 213-740-4787<br />

Estrogen depletion in post-menopausal women is a significant risk factor for the<br />

development of Alzheimer’s disease (AD) and estrogen-based hormone therapy may be<br />

capable of reducing this risk. However, the effects of progesterone both alone and in<br />

combination with estrogen on AD pathology remain unknown. In this study, we utilized<br />

the 3xTg-AD mouse model of AD to investigate the effects of estrogen and progesterone<br />

on beta-amyloid and tau pathology as well as hippocampal-dependent memory deficits.<br />

Female 3xTg-AD mice were ovariectomized at 3 mo of age and immediately replaced with<br />

a 90d, subcutaneous, slow-release pellet containing either 17-beta-estradiol, progesterone,<br />

placebo, or both 17-beta-estradiol and progesterone pellets. After 3 months, depletion of<br />

sex steroid hormones in female 3xTg-AD mice significantly increased beta-amyloid<br />

accumulation in the CA1 of hippocampus, subiculum, and frontal cortex and worsened<br />

memory performance. Continuous replacement with 17-beta-estradiol prevented these<br />

effects. Continuous progesterone replacement alone had no beneficial effect on betaamyloid<br />

pathology or cognitive deficits but did significantly decrease tau<br />

immunoreactivity. However, when both hormones were replaced in combination,<br />

progesterone attenuated the beneficial effect of 17-beta-estradiol on beta-amyloid<br />

pathology and memory deficits. These results support the hypothesis that estrogen<br />

treatment may be beneficial in reducing the risk of AD and suggest possible therapeutic<br />

strategies regarding estrogen and progesterone-based hormone therapy.<br />

This research was funded by NIH grant AG23739 (CJP) and AG026572 (R.Brinton/CJP).<br />

JCC was supported by NIH Grant AG00093 (C. Finch). ERR was supported by NIH Grant<br />

NS52143 (ERR).<br />

165


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

THE ESTROGEN-REGULATED GENE SELADIN-1/DHCR24 EXERTS ITS<br />

NEUROPROTECTIVE EFFECTS THROUGH MEMBRANE CHOLESTEROL<br />

MODULATION<br />

Danza G. 1 , Cecchi C. 2 , Pensalfini A. 2 , Formigli L. 3 , Nosi D. 3 , Stefani M 2 , Liguri G 2 ,<br />

Rosati F. 1 , Dichiara F. 1 , Morello M. 1 ,Pieraccini G. 4 , Serio M. 1 , Peri A. 1<br />

1 Endocrine Unit, Department of Clinical Physiopathology, University of Florence, Center<br />

for Research, Transfer and High Education on Chronic, Inflammatory, Degenerative and<br />

Neoplastic Disorders for the Development of Novel Therapies (DENOThe), Viale<br />

Pieraccini, 6, 50139 Florence, Italy (g.danza@dfc.unifi.it Fax n. +39 055 4271371)<br />

2 Department of Biochemical Sciences and Interuniversity Centre for the Study of the<br />

Molecular Basis of Neurodegenerative Diseases, University of Florence, viale Morgagni<br />

50, 50134 Florence, Italy<br />

3 Department of Anatomy, Histology and Forensic Medicine, University of Florence.<br />

4 Interdepartmental Mass Spectrometry Center, University of Florence.<br />

Alzheimer’s disease (AD) is a progressive, degenerative disorder of the brain characterized<br />

by loss of neurons and synapses in selective brain regions. Neuronal damage is apparently<br />

due to the altered production of intracellular neurofibrillary tangles and by extra-cellular<br />

and perivascular deposit of amyloid beta (Abeta) peptides. The reason for selective brain<br />

vulnerability has not been fully explained yet. However, a few years ago low levels of<br />

expression of a novel gene named seladin-1 (for Selective Alzheimer disease indicator 1)<br />

have been demonstrated in brain areas involved in AD [1]. Seladin-1 confers protection<br />

against Abeta mediated toxicity and from oxidative stress in vitro. In addition, it inhibits<br />

caspase 3 activity, a key mediator of apoptosis, thus protecting from apoptotic death. We<br />

have demonstrated previously that the expression of seladin-1 is up-regulated by estrogen<br />

and the selective estrogen receptor modulators (SERMs) tamoxifen and raloxifene, in a<br />

long-term cell culture of human foetal neuroblasts from human olfactory epithelium (FNC)<br />

that express both alpha and beta estrogen receptors. In these cells estrogen and SERMs<br />

significantly increased cell resistance to Abeta toxicity. Furthermore, upon seladin-1<br />

silencing, the neuroprotective effects of estrogen were abolished, indicating that this<br />

protein is a mediator of estrogen-related neuroprotection. Remarkably, seladin-1 has been<br />

found to be identical to DHCR24, the enzyme that catalyzes the last step of cholesterol<br />

biosynthesis by reducing the Delta 24 double bond of desmosterol [2]. Therefore, seladin-<br />

1/DHCR24 can be depicted as a multi-faced protein, which appears to have either antiapoptotic<br />

as well as enzymatic properties as a key enzyme of cholesterol biosynthesis.<br />

Neuronal cell death in AD is partly induced by the interaction of the Abeta peptides with<br />

the plasma membrane of target cells. It has been proposed that Abeta acts forming specific<br />

channels in the plasma membrane that allow a toxic flux of Ca2+ ions into the cell. In<br />

PC12 and in GT1-7 neurons it has been demonstrated that the enrichment of the plasma<br />

membrane with cholesterol, by modifying membrane fluidity, prevents the incorporation<br />

and pore formation of Abeta into cell membranes. Moreover we have previously shown<br />

that the variable susceptibility of different cell types to amyloid toxicity significantly<br />

correlates to membrane cholesterol content.<br />

Because the synthesis de novo of cholesterol is essential for cholesterol supply in the<br />

central nervous system, in this study we established the role of seladin-1/DHCR24 in the<br />

regulation of membrane cholesterol and in the control of the cytotoxicity of Abeta<br />

peptides.<br />

166


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

We modulated the enzymatic activity of seladin-1/DHCR24 by transient overexpression of<br />

the gene or using the specific inhibitor 5,22E-cholestadien-3β-ol (Δ22) in the human<br />

neuroblastoma cell line SH-SY5Y. These treatments determined a 30-40% increase or a<br />

15-25% decrease, respectively, of the membrane cholesterol (measured by GC/MS<br />

analysis). Cell membranes were also enriched or depauperated in cholesterol using specific<br />

reagents (PEG-cholesterol and methyl-beta-cyclodextrin, respectively) and the effect of all<br />

these treatments on Abeta toxicity was evaluated. We found that both seladin-1<br />

overexpressing cells with high membrane cholesterol and cells enriched in PEGcholesterol<br />

showed a significantly lower susceptibility to Aβ1-42 aggregates compared to<br />

control neuroblastoma cells. Confocal microscopic analysis, using monoclonal anti-Aβ<br />

antibodies or Fluo3-AM as fluorescent calcium indicator, revealed that the difference in<br />

cell viability was related to the ability of high membrane cholesterol content to: (i) reduce<br />

the interaction of amyloid assemblies with the plasma membrane; (ii) prevent intracellular<br />

Ca 2+ spikes induced by oligomeric inclusion. Conversely, membrane cholesterol loss in<br />

neuroblastoma cells treated with Δ22 or with methyl-beta-cyclodextrin triggered a quicker<br />

amyloid accumulation on cell surfaces and a higher cytosolic Ca 2+ increase resulting in a<br />

reduced cell survival. These data demonstrate that one of the mechanisms of<br />

seladin1/DHCR24 neuroprotection is dependent on the modulation of membrane<br />

cholesterol.<br />

Reference list<br />

1. Greeve, I., Hermans-Borgmeyer, I., Brellinger, C., Kasper, D., Gomez-Isla, T., Behl,<br />

C.,Levkau, B., Nitsch. RM., 2000. The human DIMINUTO/DWARF1 homolog<br />

seladin-1/DHCR24 confers resistance to Alzheimer’s disease-associated<br />

neurodegeneration and oxidative stress. J Neurosci. 20, 7345-52.<br />

2. Waterham, HR., Koster, J., Romeijn, GJ., Hennekam, RC., Vreken, P., Andersson,<br />

HC., FitzPatrick, DR., Kelley, RI., Wanders, R.J, 2001. Mutations in the 3betahydroxysterol<br />

Delta-reductase gene cause desmosterolosis, an autosomal recessive<br />

disorder of cholesterol biosynthesis. Am J Hum Genet. 69, 685-94.<br />

167


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

ANDROGENIC, BUT NOT ESTROGENIC, PROTECTION OF MOTONEURONS<br />

FROM SOMAL AND DENDRITIC ATROPHY INDUCED BY THE DEATH OF<br />

NEIGHBORING MOTONEURONS<br />

Fargo K.N. * , and Sengelaub D.R.<br />

Program in Neuroscience and Department of Psychological and Brain Sciences, Indiana<br />

University, 1101 East 10th Street, Bloomington, Indiana 47405, USA;<br />

sengelau@indiana.edu, Fax (812) 855-4691<br />

* This author now at Loyola University Chicago, Stritch School of Medicine;<br />

kfargo@lumc.edu<br />

Motoneuron loss is a significant medical problem, capable of causing severe movement<br />

disorders or even death. We have been investigating the effects of motoneuron loss on<br />

surviving motoneurons in a lumbar motor nucleus, the spinal nucleus of the<br />

bulbocavernosus (SNB). SNB motoneurons undergo marked dendritic and somal atrophy<br />

following the experimentally-induced death of other nearby SNB motoneurons. However,<br />

treatment with testosterone at the time of lesioning completely prevents this atrophy [1,<br />

2]. Because testosterone can be metabolized into the estrogen estradiol (as well as other<br />

physiologically active steroid hormones), it was unknown whether the protective effect of<br />

testosterone was an androgen effect, an estrogen effect, or both. In the present experiment,<br />

we used a retrogradely-transported neurotoxin to kill the majority of SNB motoneurons on<br />

one side of the spinal cord only, in adult male rats. Some animals were also treated with<br />

either testosterone, the androgen dihydrotestosterone (which cannot be converted into<br />

estradiol), or the estrogen estradiol. As seen previously, partial motoneuron loss led to<br />

reductions in soma area and in dendritic length and extent. Testosterone and<br />

dihydrotestosterone prevented these reductions, but estradiol had no protective effect.<br />

These results indicate that the neuroprotective effect of testosterone on the morphology of<br />

SNB motoneurons following partial motoneuron depletion is an androgen effect rather than<br />

an estrogen effect.<br />

This work supported by NIH-NINDS NS047264 (D.R.S.) and T32 DC 00012<br />

Reference list<br />

1. Fargo, K.N., Sengelaub, D.R., 2004. Testosterone manipulation protects motoneurons from dendritic<br />

atrophy after contralateral motoneuron depletion. J Comp Neurol. 469, 96-106.<br />

2. Fargo, K.N., Sengelaub, D.R., 2004. Exogenous testosterone prevents motoneuron atrophy induced by<br />

contralateral motoneuron depletion. J Neurobiol. 60, 348-<strong>35</strong>9.<br />

168


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

NEURONAL PROTECTION OF DEHYDROEPIANDROSTERONE ON PC12<br />

CELLS PRETREATED WITH β-AMYLOID<br />

Forsberg M. K 1 ., Hallberg M 2 ., Nyberg F 2 ., Svensson A-L 1<br />

Department of Pharmaceutical Biosciences, 1 Division of Pharmacology, 2 Division of<br />

Biological Research on Drug Dependence, Uppsala University, P.O. Box 591, SE-751 24<br />

Uppsala, Sweden, Fax +46-18501920, e-mail: marie.forsberg@farmbio.uu.se<br />

The neuroactive steroid dehydroepiandrosterone (DHEA) is the most abundant<br />

prohormone produced by the adrenal glands. It is synthesized from cholesterol and<br />

secreted to peripheral tissues where it can be converted into androgens and/or estrogens<br />

[Labrie et al., 2005]. It is an important source of sex steroids apart from the gonads, leaving<br />

the adrenals as the only source of testosterone and estrogens after menopause in women.<br />

Neurosteroids have been implicated in loss of memory and memory acquisition in rodents<br />

[Brown et al., 2000]. Some neurosteroids are able to enhance memory performance in rats<br />

[reviewed in Mellon and Griffin, 2002]. It has been suggested that DHEA and its sulfate ester<br />

may protect the brain from neurodegeneration [Compagnone and Mellon, 2000]. DHEA has<br />

been shown to increase neuronal survival and differentiation. Furthermore, DHEA has<br />

been reported to protect hippocampal neurons against NMDA-induced toxicity in vitro<br />

[Kimonides et al 1998, Cardounel et al., 1999]. Although DHEA have shown neuroprotective<br />

properties, the mechanism(s) for its neuroprotective effects is not known.<br />

A plausible link between DHEA, other neurosteroids and neurodegenerative disorders like<br />

Alzheimer´s disease (AD) has been discussed. In AD the levels of neurosteroids such as<br />

DHEA and allopregnanolone are reduced compared to controls [Weill-Engerer et al., 2002]. It<br />

has recently been suggested that beta-amyloid (Aβ) peptide can activate DHEA synthesis.<br />

DHEA are also suggested to feed back onto glial cells and protect them against Aβinduced<br />

toxicity [Brown et al. 2000]. Beta-amyloid (Aβ) is a peptide that is a byproduct of the<br />

transmembrane protein amyloid precursor protein (APP). AD are characterized<br />

pathologically by deposits of Aβ peptide in the brain. Aβ peptide has been implicated in<br />

cell death during the course of AD and exerts toxic effects on neurons both in vivo and in<br />

vitro [reviewed in Jellinger, 2006].<br />

In the present study, the effect of DHEA on Aβ-induced toxicity was investigated in rat<br />

pheochromocytoma PC12 cells. PC12 cells were exposed to Aβ alone and in the presence<br />

of DHEA for 24 hours. Untreated cells were used as controls. The number of necrotic<br />

cells were determined by the tryptane blue exclusion assay. The apoptosis was evaluated<br />

by measurement of caspase-3 activity using immunocytochemistry [Östergren et al., 2005].<br />

Treatment of PC12 cells with Aβ (10 -6 M) significantly increased the number of necrotic<br />

cells. When PC12 cells were treated with Aβ in the presence of DHEA (10 -9 and 10 -6 M)<br />

the number of necrotic cells were significantly reduced.<br />

Treatment of PC12 cells with Aβ (10 -6 M) caused a significant increased number of<br />

caspase-3-positive cells. When PC12 cells were treated with Aβ in the presence of DHEA<br />

(10 -9 and 10 -6 M) the number of caspase-3-positive cells were significantly reduced.<br />

These results demonstrated that DHEA can protect PC12 cells against Aβ-induced toxicity.<br />

The mechanism(s) for the effect of DHEA is not fully understood but some evidence<br />

points towards the sigma receptors as a possible site of action. To investigate whether the<br />

169


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

neuroprotective effect of DHEA was achived through sigma receptors, PC12 cells were<br />

treated with DHEA in the presence of BD-1047 (a sigma-1 antagonist and sigma-2 agonist<br />

[Matsumoto R.R., et al.1995]). BD-1047 (10 -6 M) was observed to partly attenuate the toxicity<br />

induced by Aβ (10 -6 M) in PC12 cells. However, BD-1047 (10 -6 M) was not able to<br />

prevent the neuroprotective effect of DHEA against Aβ-induced toxicity.<br />

These data demonstrate that DHEA exerts neuroprotective properties, but the mechanism<br />

behind is still unclear and needs to be investigated. Knowledge of the role of neurosteroids<br />

on processes that are ongoing in Alzheimer brains might lead to clinical important<br />

applications.<br />

Reference list<br />

Brown R.C., Cascio C. and Papadopoulos V.,2000.<br />

Pathways of neurosteroid biosynthesis in cell lines from human brain: regulation of<br />

dihydroepiandrosterone formation by oxidative stress and β-amyloid peptide. J Neurochem, 74:847-859.<br />

Cardounel A., Regelson W. and Kalimi M., 1999.<br />

Dihydroepiandrosterone protects hippocampal neurons against neurotoxin-induced cell death: mechanism<br />

of action. Proc Soc Exp Biol Med, 222:145-149.<br />

Compagnone N.A. and Mellon S.H., 2000.<br />

Neurosteroids: Biosynthesis and function of these novel neuromodulators. Frontiers in<br />

Neuroendocrinology 21:1-56.<br />

Jellinger K.A., 2006.<br />

Alzheimer 100 – highlights in the history of Alzheimer research. J Neural Transm. 113:1603-1623.<br />

Kimonides VG, Khatibi NH, Svendsen CN, Sofroniew MV, Herbert J., 1998.<br />

Dehydroepiandrosterone (DHEA) and DHEA-sulfate (DHEAS) protect hippocampal neurons against<br />

excitatory amino acid-induced neurotoxicity. Neurobiology 95:1852-7.<br />

Labrie F, Luu-The V, Bélanger A, Lin X-S, Simard J, Pelletier G., 2005.<br />

Is dehydroepiandrosterone a hormone? Journal of Endocrinology 187:169-96.<br />

Matsumoto R.R., Bowen W.D., Tomm. A., Vo V.N., Truong D.D., De Costa B.R., 1995.<br />

Characterization of two novel σ receptor ligands: Antidystonic effects in rats suggest σ receptor<br />

antagonism.<br />

European Journal of Pharmacology 280:301-310.<br />

Mellon S.H. and Griffin L.D., 2002.<br />

Neurosteroids: biochemistry and clinical significance. Trends Endocrinology Metabolism 13:<strong>35</strong>-43.<br />

Weill-Engerer, S., David, J. P., Sazdovitch, V., Liere, P., Eychenne, B., Pianos, A., Schumacher, M.,<br />

Delacourte, A., Baulieu, E. E., Akwa, Y., 2002.<br />

Neurosteroid quantification in human brain regions: comparison between Alzheimer's and nondemented<br />

patients. J. Clin. Endocrinol. Metab. 87:5138-43.<br />

Östergren, A., Svensson, A-L., Lindquist, N. G., & Brittebo, E. B., 2005.<br />

Dopamine melanin-loaded PC12 cells: a model for studies on pigmented neurons. Pigment Cell<br />

Research 18:306-314.<br />

170


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

PROGESTERONE AND ITS DERIVATIVES AS PROTECTIVE AGENTS IN<br />

EXPERIMENTAL DIABETIC NEUROPATHY<br />

S. Giatti #1 , I. Roglio 1 , M. Pesaresi 1 , R. Bianchi 2 , G. Cavaletti 3 , L.M. Garcia-Segura 4 ,<br />

G. Lauria 5 , R.C. Melcangi 1<br />

1 Dept. of Endocrinology and Center of Excellence on Neurodegenerative Diseases, University of Milan,<br />

Milano, Italy;<br />

2 Dept. of Molecular Biochemistry and Pharmacology, "Mario Negri" Institute for<br />

Pharmacological Research, Milano, Italy; 3 Dept. of Neurosciences and Biomedical Technologies, University<br />

of Milan"Bicocca", Monza, Italy; 4 Instituto Cajal, C.S.I.C., Madrid, Spain; 5 Neuromuscular Diseases Unit,<br />

National Neurological Institute “Carlo Besta”, Milano, Italy.<br />

# Presenting author: Dept. of Endocrinology and Center of Excellence on Neurodegenerative Diseases,<br />

University of Milan, via Balzaretti 9, 20133, Milano Italy.<br />

silvia.giatti@guest.unimi.it<br />

Deleterious effects of diabetes on the nervous system are responsible for several disorders<br />

including damage to the peripheral nervous system. However, in spite of the number of<br />

studies on human and experimental diabetic neuropathy, the current therapeutic arsenal is<br />

meagre. Consequently, the search for substances to protect the nervous system from the<br />

degenerative effects of diabetes has high priority in biomedical research.<br />

Because of the importance of neuroactive steroids in the control of the nervous system<br />

functions and of their neuroprotective effects in several experimental models of<br />

neurodegenerative diseases, we have assessed whether chronic treatment with progesterone<br />

(P), and its derivatives, dihydroprogesterone (DHP) and tetrahydroprogesterone (THP),<br />

had protective effects against (STZ)-induced peripheral neuropathy at the<br />

neurophysiological, functional, biochemical and neuropathological levels. Data obtained<br />

have indicated that chronic treatment for 1 month with P, or with its derivatives, DHP and<br />

THP, counteracted the impairment of nerve conduction velocity (NCV) and thermal<br />

threshold, restored skin innervation density, and improved Na + ,K + -ATPase activity and<br />

mRNA levels of myelin proteins, such as glycoprotein zero and peripheral myelin protein<br />

22. Moreover, protective effects of these neuroactive steroids in STZ-rat were also evident<br />

on morphological degeneration of the sciatic nerve. Indeed, treatment with P or DHP<br />

induced a significant reduction in the number of fibers with myelin infoldings. Altogether<br />

these observations suggest that these neuroactive steroids might be useful protective agents<br />

in diabetic neuropathy. Interestingly, different receptors seem to be involved in these<br />

effects. Thus, while morphological integrity of myelin, the expression of myelin proteins<br />

and Na + ,K + -ATPase activity are only influenced by P and DHP, (i.e., two neuroactive<br />

steroids interacting with P receptor, PR), NCV, thermal nociceptive threshold and intraepidermal<br />

nerve fiber density are also affected by THP, which interacts with GABA-A<br />

receptor. Because, a therapeutic approach with specific synthetic receptor ligands could<br />

avoid the typical side effects of steroids, future experiments will be devoid to evaluate the<br />

role of PR and GABA-A receptor in these protective effects.<br />

(PRIN-2005060584_004 and FIRST from University of Milan).<br />

171


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

AN EVALUATION OF THE EFFECT OF DEXAMETHASONE IN PREVENTING<br />

TOURNIQUET NEUREPATHY IN RATS<br />

Jarrahi M, Vafaei AA, Rashidy-Pour A<br />

Physiology Research center, Semnan University of Medical Sciences, Semnan, Iran<br />

E-mail: jarrahi44@yahoo.com<br />

INTRODUCTION<br />

Paralysis after the use of a tourniquet is a well recognized clinical phenomenon. Although ischemia<br />

to nerve causes fiber degeneration [1] Inflammation following IR injury has been shown to lead to<br />

the induction of the ca2+-independent (inducible) form of NOS (iNOS) [2]. Dexamethasone is a<br />

glucocorticoid that has been shown to inhibit iNOS production by repressing gene expression,<br />

inhibiting protein synthesis and decreasing iNOS protein stability [3,4]. The present experiment<br />

was designed to evaluate the effects of dexamethasone during post tourniquet IR injury.<br />

MATERIAL AND METHODS:<br />

36 male Wistar rats (200-250 gr) were chosen and divided randomly into 6 groups as control,<br />

vehicle, tourniquet-vehicle, Dex1, Dex 2 and Dex 3. Tourniquet was applied to the right hind limb<br />

of all Animals except of control and vehicle groups for 3 hours at the beginning of experiments.<br />

Animals of tourniquet-vehicle, Dex1, Dex 2 and Dex 3 groups were injected 30 min before<br />

reperfusion by 1 cc Normal saline containing 4% ethanol, 1mg/kg, 2mg/kg and 3mg/kg<br />

Dexamethasone dissolved in saline and Alcohol respectively and motor nerve conduction velocity<br />

(MNCV) was measured one week after releasing the tourniquet. MNCV of tibial nerve in<br />

response to nerve stimulation were measured at 1 week post-tourniquet release.<br />

The procedure was done under ketamine-xylocin (90mg/kg-100mg/kg,Ip) anesthesia. Ischemia of<br />

the lower hind limb was produced by placing a pneumatic tourniquet around the right tigh and<br />

inflating the cuff to115 to125 mmHg. After one week for evaluation of MNCV the animals were<br />

acclimated for at leas thirty minutes in a temperature controlled room maintained at 25± 0.5 Cº<br />

before measurements were reperoformed. The procedure was done under urethane anesthesia (1.5<br />

g/kg), with insitu preparation in a pool of liquid paraffin previously saturated with saline to avoid<br />

nerve dehydration. Rectal temperature was monitored throughout anesthesia for this procedure (2 –<br />

4 min). No hypothermia was observed and all rats remaining between 37 and 38 C. The right<br />

sciatic nerve was stimulated first at the sciatic notch and then at the Achilles tendon. Stimulation<br />

comprises single 0.1 ms pulses of amplitude 1– 4 V were delivered via fine bipolar needle<br />

electrodes. Consequent each stimulation an electromyogram (EMG) was recorded, again using fine<br />

needle electrodes, from the gastrocinemus muscle via a 250X gain AC preamplifier on a single<br />

beam storage oscilloscope. The temporal separation of the peaks of the EMGs, was induced by<br />

stimulation the two sites, was measured using dividers. The mean of six measurements was taken<br />

on each occasion. Nerve length, separating the two points (sciatic notch and Achilles tendon) was<br />

measured using vernier caliper. The MNCV was calculated by the difference between the two<br />

stimulating electrodes.<br />

RESULTS:<br />

Application of the tourniquet for 3 h decreased significantly (p


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

90<br />

*<br />

MNCV (M/S)<br />

60<br />

30<br />

*<br />

0<br />

Control Vehicle TOUR DEX1 DEX2 DEX3<br />

Experimental groups<br />

Fig 1. Effect of Matricaria chamomilla extract dissolved in olive oil with comparison of olive oil<br />

on the percentage of wound healing in days after beginning of experiments. *P


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

INVOLVEMENT OF ENDOGENOUS DEHYDROEPIANDROSTERONE IN THE<br />

MODULATION OF SPINAL NOCICEPTIVE MECHANISMS<br />

Kibaly C., Meyer L., Patte-Mensah C. and Mensah-Nyagan A.G.<br />

Institut des Neurosciences Cellulaires et Intégratives, UMR 7168/LC2-CNRS, Université Louis<br />

Pasteur, Equipe Stéroïdes et Système nociceptif, 21 rue René Descartes, 67084 Strasbourg Cedex,<br />

France. Fax: +33 388 613 347; e-mail: gmensah@neurochem.u-strasbg.fr<br />

The excessive advertisement of dehydroepiandrosterone (DHEA) as a miraculous<br />

anti-aging drug resulted in its widespread self-administration in Europe and the USA<br />

where DHEA is available without medical prescription. However, basic data supporting<br />

beneficial effects of DHEA are rare and the risks related to long-term supplementation of<br />

DHEA are almost completely unknown. Here, we used a multidisciplinary approach to<br />

show that endogenous DHEA locally synthesized in sensory networks of the spinal cord<br />

(SC) is a pro-nociceptive molecule the production of which is down-regulated by<br />

neuropathic rats to cope with their chronic pain state. Real-time polymerase chain reaction<br />

after reverse transcription revealed a down-regulation of the gene encoding cytochrome<br />

P450c17 (P450c17), the key DHEA-synthesizing enzyme, in the SC of rats submitted to<br />

neuropathic pain generated by sciatic nerve ligature. Pulse-chase experiments combined<br />

with high-performance liquid chromatography and flow scintillation detection showed<br />

decreased P450c17 enzymatic activity in the SC of neuropathic-pain rats.<br />

Radioimmunoassays demonstrated that, in vivo, the chronic pain dramatically reduced<br />

DHEA concentration in the SC. Moreover, in vivo blockade of DHEA production in the<br />

SC by intrathecal administration of ketoconazole, a pharmacological inhibitor of P450c17,<br />

induced analgesia in neuropathic-pain rats. Unlike ketoconazole, DHEA potentiated both<br />

thermal hyperalgesia and mechanical allodynia characterizing the neuropathic pain state.<br />

The results draw attention on the potential risks linked to abusive use of DHEA,<br />

particularly in victims of neuropathic pain. Perspectives are opened for analgesic strategies<br />

based on the selective modulation of DHEA or neurosteroid biosynthetic pathways in<br />

neural centers controlling pain sensation.<br />

174


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

POSTNATAL CHANGES IN THE EXPRESSION OF AQUAPORIN 1 AND EFFECT<br />

OF ESTROGEN ON THE EXPRESSION IN OVARIECTOMIZED MOUSE BRAIN<br />

Lee J.E., Kim H.J., Kang H.S., Ahn H.S., and Gye M.C.*<br />

Department of Life Science, Hanyang University, Seoul 133-791, Korea<br />

Fax: +82-2-2298-8646<br />

e-mail: mcgye@hanyang.ac.kr<br />

1. Introduction<br />

Brain aquaporins (AQP) play important roles in the dynamic regulation of brain water<br />

homeostasis and the production of cerebrospinal fluid (CSF) under normal, as well as<br />

pathological, conditions. In females, endogenous estrogen can affect cognitive functioning<br />

during the menstrual cycle [1], preserve certain aspects of cognitive function in<br />

postmenopausal women [2], act as a neuroprotective and neuroregenerative agent in stroke<br />

and traumatic brain injuries, and reduce the risk of developing Alzheimer's disease [3]. To<br />

elucidate the role estrogen in the maintenance of brain water homeostasis, changes in the<br />

expression of AQP1, an important structural element of choroid plexus following ovariectomy<br />

(OVX), as well as the effect of estrogen (17ß estradiol) on the expression of AQP1 in OVX<br />

brain, was examined in mice.<br />

2. Materials and Methods<br />

Normal cyclic female mice were subjected to oveaiectomy (OVX) under ketamine anesthesia<br />

(100 mg/kg); a sham operation was performed as a control. After recovery, to allow clearance<br />

of circulating estrogen following OVX or sham, mice were rested for two weeks, and then<br />

subjected to estrogen treatment. Additionally, some OVX mice received intraperitoneal<br />

injections of 17ß estradiol (E 2 ) at 20 µg/head (single injection) or 1 µg/head (for 7 days)<br />

dissolved in sesame oil (4 heads for each treatment). As a control, sesame oil was injected as a<br />

vehicleOVX mice were sacrificed 24 h after final dosing of E2. Sham operated mice were<br />

sacrificed at the estrous stage. Hippocampus were dissected and subjected to protein and PCR<br />

analysis. Primers for AQP1 were designated 5’-AGTATGACCTGG ATGCTGAC-3’<br />

(forward) and 5'-ACTCCTCCATGATGTCAAAG-3' (reverse) according to the mouse AQP1<br />

cDNA sequence (GenBank Acc, NM_007472, product size: 360bp). To confirm the estrogen<br />

response in brain expression of transthyretin (TTR) synthesized in choroids plexus was<br />

verified (Tang et al., 2004). Primers for TTR were 5’-aga cgt ggc tgt aaa agt gt-3’ (forward)<br />

and 5'-ctg tag gag tat ggg ctg ag-3' (reverse) according to the mouse TTR cDNA sequence<br />

(GenBank Acc, NM_013697, product size: 273bp). GAPDH mRNA was amplified as an<br />

internal control. The primers for GAPDH were 5'-AGTGGAGATTGTTGCCATCAACGAC-<br />

3' (forward) and 5'-GGGAGTTGCTGTTGAAGTCGCAGGA-3' (reverse) (GenBank Acc,<br />

NM001001303, product size: 360bp). PCR products were analyzed on 2% agarose gels<br />

containing ethidium bromide. The relative abundance of AQP1 mRNA versus GAPDH was<br />

determined by performing densitometric analysis of the amplicons. For immunohistochemical<br />

analysis of AQP1, cryocuts (5 µm) of brain at estrous stage were made and processed for<br />

immunohistochemical analysis using AQP1 antibody (rabbit polyclonal, SantaCruz, CA,<br />

1:200 dilution) and HRP conjugated secondary antibody. As a negative control, primary<br />

antibody was omitted. Following counterstaining with Mayer hematoxylin, slides were<br />

mounted permanently. Observation and photography were performed with a microscope<br />

equipped digital camera (DFC320, Leica, Germany).<br />

175


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

3. Results<br />

Throughout brain development, the immunoreactivity of AQP1 was found in the choroid<br />

plexus. Ependyma, pia, and veins were also positive for AQP1 immunoreactivity. AQP1<br />

mRNA level showed gradual decrease until postnatal day 8 but reincreased at adulthood. Two<br />

forms of AQP1 polypeptides with Mr. <strong>35</strong> and 28 kDa in brain. Both forms were much higher<br />

in the fetal brain but only <strong>35</strong>kDa form was detected in adult brain. AQP1 was significantly<br />

down regulated together with transthyretin in OVX brain compared with the sham control at<br />

the estrous stage. In OVX female, repeated dosage with E 2 (1 or 10 µg/head) for 7 days<br />

significantly augmented AQP1 mRNA level together with TTR mRNA in OVX female<br />

brains; however, single high dose of E 2 (20 µg/head) resulted in no significant changes in<br />

AQP1 expression in OVX brains.<br />

4. Conclusion<br />

Together, these results suggest that expression of AQP1 in female brain tissue is tightly<br />

regulated by estrogen. Alteration of AQP1 expression in brain should be considered as a<br />

likely candidate mechanism for the pathological changes in the CNS in estrogen deficiency.<br />

Reference list<br />

[1] E. Hogervorst, J. Williams, M. Budge, W. Riedel, and J. Jolles, The nature of the effect of<br />

female gonadal hormone replacement therapy on cognitive function in postmenopausal<br />

women: a meta-analysis, Neuroscience 101, (2000), pp. 485-512.<br />

[2] M.H. Birkhauser, J. Strnad, C. Kampf and M. Bahro, Oestrogens and Alzheimer's disease.<br />

Int. J. Geriatr. Psychiatry 15 (2000), pp. 600-609.<br />

[3] J.I. Koenig, Estrogen and brain function. Trends Endocrinol. Metab. 12 (2001), pp. 4-6.<br />

176


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

TESTOSTERONE DERIVATIVES ARE NEUROPROTECTIVE AGENTS IN<br />

EXPERIMENTAL DIABETIC NEUROPATHY<br />

I. Roglio #1 , S. Giatti 1 , M. Pesaresi 1 , R. Bianchi 2 , G. Cavaletti 3 , D. Caruso 4 , S, Scurati 4 ,<br />

L.M. Garcia-Segura 5 , G. Lauria 6 , R.C. Melcangi 1<br />

1 Dept. of Endocrinology and Center of Excellence on Neurodegenerative Diseases, University of Milan,<br />

Milano, Italy;<br />

2 Dept. of Molecular Biochemistry and Pharmacology, "Mario Negri" Institute for<br />

Pharmacological Research, Milano, Italy; 3 Dept. of Neurosciences and Biomedical Technologies, University<br />

of Milan"Bicocca", Monza, Italy; 4 Dept. of Pharmacological Sciences, University of Milan, Milano, Italy;<br />

5 Instituto Cajal, C.S.I.C., Madrid, Spain; 6 Neuromuscular Diseases Unit, National Neurological Institute<br />

“Carlo Besta”, Milano, Italy.<br />

# Presenting author: Dept. of Endocrinology and Center of Excellence on Neurodegenerative Diseases,<br />

University of Milan, via Balzaretti 9, 20133, Milano Italy.<br />

ilaria.roglio@unimi.it<br />

Our recent studies have shown that progesterone and its metabolites exert important<br />

protective effects on peripheral nerves against pathological alterations induced by diabetes.<br />

In the present study we have assessed whether other members of the neuroactive steroid<br />

family, such as testosterone (T) and its derivatives, dihydrotestosterone (DHT) and 5alpha<br />

-androstan-3alpha, 17beta-diol (3alpha -diol) also exert protection against diabetic<br />

neuropathy. Diabetes was induced in adult male rats by the injection of streptozotocin.<br />

After 3 months of diabetes, T plasma levels, evaluated by liquid chromatography/mass<br />

spectrometry, showed a decrease associated to an increase of 3alpha-diol levels. In<br />

contrast, in sciatic nerve of diabetic rats low levels of DHT were observed and the<br />

expression of the enzyme converting T into DHT (i.e., the 5alpha-reductase), evaluated by<br />

real time PCR, was also reduced. Chronic treatment for 1 month with DHT or 3alpha-diol<br />

increased tail nerve conduction velocity and partially counteracted the increase of thermal<br />

threshold induced by diabetes. Treatment with DHT increased tibial Na + ,K + -ATPase<br />

activity and the gene expression of myelin protein P0 (i.e., a protein which plays a crucial<br />

role in maintaining the multilamellar structure of peripheral nerve myelin). Not only DHT<br />

and 3alpha-diol, but also T treatment, were able to totally restore the reduction of<br />

intraepidermal nerve fiber density induced by diabetes. Altogether these observations<br />

indicate that T derivatives can reverse behavioral, neurophysiological, morphological and<br />

biochemical alterations induced by peripheral diabetic neuropathy. These findings further<br />

strengthen the evidence that different members of the neuroactive steroid family exert<br />

relevant protective effects at the level of peripheral nerves<br />

(PRIN-2005060584_004 and FIRST from University of Milan).<br />

177


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

ESTROGEN AND ANDROGENS REGULATE ALZHEIMER-LIKE<br />

NEUROPATHOLOGY IN MALE 3XTG-AD MICE<br />

Rosario E.R., Carroll J.C., and Pike C.J.<br />

Davis School of Gerontology, University of Southern California 3715 McClintock Avenue<br />

Los Angeles, CA 90089-0191 USA Tel: 213-740-4205 Fax: 213-740-4787<br />

Email: cjpike@usc.edu<br />

Normal, age-related, androgen depletion in men is a recently identified risk factor<br />

for the development of Alzheimer’s disease (AD). Recently, using a triple-transgenic<br />

mouse model of AD (3xTg-AD), we found that androgen depletion in males results in<br />

robust increases in accumulation of β-amyloid (Aβ), the protein implicated as the primary<br />

causal factor in AD pathogenesis. Androgen treatment, in the form of the nonaromatizable<br />

androgen dihydrotestosterone (DHT), prevented increased neural<br />

accumulation of Aβ; however, the mechanism of androgen regulation of Aβ remains<br />

unclear. Although several androgen actions in the brain are mediated through androgen<br />

pathways involving androgen receptors (AR), androgen actions may also result from<br />

estrogen-mediated pathways. To examine whether androgen actions on Aβ accumulation<br />

involve estrogen and / or androgen pathways we examined the effects of testosterone (T),<br />

dihydrotestosterone (DHT), and 17β-estradiol (E2) in androgen depleted male mice. Male<br />

3xTg-AD mice were gonadectomized (GDX) to deplete endogenous levels of androgens at<br />

3 months of age, and exposed to subcutaneous, slow-release drug delivery pellets<br />

containing either vehicle, 10 mg DHT, 10mg T, 0.025mg E2, or 0.01mg E2. After 4<br />

months of hormone treatment (7 months of age), mice were evaluated for severity of ADlike<br />

neuropathology. Similar to our previous results, we observed a significant increase in<br />

Aβ pathology in GDX mice in the subiculum and CA1 of hippocampus and amygdala, an<br />

effect prevented by DHT treatment. Treatment with testosterone also prevented increased<br />

accumulation of Aβ pathology in both hippocampus and amygdala. Interestingly, estrogen<br />

prevented Aβ accumulation in the subiculum and CA1 of hippocampus but had no effect in<br />

amygdala. These findings suggest that androgen regulation of Aβ is mediated through<br />

both androgen and estrogen pathways depending on the brain region.<br />

Supported by the Alzheimer’s Association (IIRG-04-1274; CJP) and NIH grants AG23739<br />

(CJP), NS52143 (ERR), AG00093 (JCC).<br />

178


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

EFFECTS OF BETA AMYLOID PEPTIDE 1-42 AND OXIDATIVE STRESS ON<br />

NEUROSTEROID FORMATION IN HUMAN NEUROBLASTOMA CELLS<br />

Schaeffer V. + , Patte-Mensah C. + , Eckert A.°, Mensah-Nyagan A.G. +#<br />

+<br />

Institut des Neurosciences Cellulaires et Intégratives, Equipe Stéroïdes et Système<br />

Nociceptif, UMR 7168/LC2, CNRS, Université Louis Pasteur, 21 rue René Descartes,<br />

67 084 Strasbourg cedex, France. Fax: +33 (0)3 88 61 33 47<br />

#<br />

e-mail:<br />

gmensah@neurochem.u-strasbg.fr.<br />

° Neurobiology Research Laboratory, Psychiatric University Clinic, Wilhelm Klein-Strasse<br />

27, CH-4025 Basel, Switzerland.<br />

Pharmacological and behavioral studies performed in animals suggest neurosteroid<br />

involvement in neuroprotection. However, the contribution of neurosteroidogenesis in the<br />

protection against degenerative processes in humans remains to be determined. In a recent<br />

work, we showed that the overexpression of key Alzheimer’s disease (AD) proteins (native<br />

tau, mutant P301L tau and amyloid peptide precursor) interferes with the process of<br />

neurosteroidogenesis in human neuroblastoma SH-SY5Y cells. In addition, we observed<br />

that extracellular treatment of SH-SY5Y cells with aggregated synthetic Beta-amyloid<br />

fragment 25-<strong>35</strong> (AB 25-<strong>35</strong> ) stimulated progesterone synthesis and had no effect on estradiol<br />

synthesis. Because the activity of AB 25-<strong>35</strong> can be different from that of the full-length AB 1-<br />

42 which is the real pathogenic peptide involved in AD, we have now combined pulsechase<br />

experiments with HPLC and continuous flow scintillation detection to study the<br />

effect of AB 1-42 on neurosteroid production in SH-SY5Y cells. AB 1-42 mimicked the action<br />

of AB 25-<strong>35</strong> on progesterone formation. In contrast, the biosynthesis of estradiol, which was<br />

totally unaffected by AB 25-<strong>35</strong>, was stimulated in SH-SY5Y cells by AB 1-42 . In addition to<br />

the comparative analysis of the effects of AB 1-42 and AB 25-<strong>35</strong> , we have also investigated the<br />

action of oxidative stress (another pathological factor leading to AD) on neurosteroid<br />

synthesis in human neuroblastoma cells. A significant SH-SY5Y cell death was observed<br />

24h and 48h after treatment with the oxidative stressor H 2 O 2 . A decreased estradiol<br />

production was detected in SH-SY5Y cells 12h, 24h and 48h after treatment with H 2 O 2 .<br />

Our results show that the cellular components responsible for endogenous formation of<br />

estradiol in human neuroblastoma cells are selectively modified by various factors<br />

involved in the etiology of AD. Consequently, it appears that molecular and cellular<br />

elements contributing to estradiol production in nerve cells may potentially be interesting<br />

to decipher functional interactions between the process of neurosteroidogenesis and<br />

neurodegenerative or neuroprotective pathways. Elucidation of such interactions will<br />

certainly open new perspectives for the development of efficient therapies against<br />

neurodegenerative disorders.<br />

179


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

EFFECT OF ESTROGEN TREATMENT ON PROTEIN EXPRESSION PATTERN IN<br />

FAMALE MICE BRAIN USING FLUORESCENT DIFFERENTIAL 2-D GEL<br />

ELECTROPHORESIS (DIGE)<br />

Szegő É.M., Kékesi K.A., Juhász G., Ábrahám I.M.<br />

Research Group of Neurobiology at Eötvös Loránd University - Hungarian Academy of Sciences, Pázmány<br />

P. Stny 1/c, Budapest, Hungary<br />

e-mail: eva.szego@freemail.hu<br />

Steroid hormons can alter neuronal functions such as learning, behavior and they can<br />

effect the cell-viability as well. The main female gonadal hormone estrogen (E2) could<br />

affect neurons on two different ways: via direct DNA-binding and transcriptional activity<br />

of liganded ERs (classical effect); or via intracellular signaling system (non-classical<br />

effect) such as protein kinase A, calcium-calmodulin dependent protein kinase IV leading<br />

to activation of many transcription factors and consequent changes in gene transcription<br />

and protein expression in the brain. In this study we characterized the impact of estrogen<br />

on the brain proteome using proteomic approaches. Differential two-dimensional gel<br />

electrophoresis (DIGE) is one of the key tools for comparative proteomic research. This<br />

method enable us to separate complex protein mixtures with high resolution, DIGE is a<br />

technique commonly employed for protein profiling studies. In our experiemnts total<br />

protein was extracted from the brain of ovariectomized mice at 24 hrs following E2 or<br />

vehicle injection. Equal amounts of protein from individual animals were homogenized,<br />

labeled with Cy3 (absorbtion max: 553 nm, emission max: 572 nm) or Cy5 (A max: 648<br />

nm, E max: 669 nm) and were focused isoelectrically and run on the same analytical gel. A<br />

pool composed of equal aliquot of each sample in the study was labeled with Cy2 dye (A<br />

max: 491 nm, E max: 506 nm) and loaded on each gel as a between-gel reference to<br />

minimize the gel-to-gel variation. 800 microgram protein was loaded on preparative gels<br />

for MS analysis. Gels were scanned with Typhoon Trio+ confocal laser scanner and<br />

analyzed using DeCyder (spot detection, noise filtering) and BVA (gel matching, statistica)<br />

softwares (GE Healthcare). We identified approximately 3000 protein spots in our 2D-gels.<br />

After trypsin-digestion, proteins were analysed with ESI LC MS/MS method. Estradiol<br />

altered the abundance of 76 spots (p≤0.05), and 48 of these proteins were identified with<br />

MS. Proteins can be clustered into 7 groups according to their biological functions. Our<br />

result suggest that estrogen may alter the rate of protein synthesis in the brain. Following<br />

estrogen treatment the amount of enzymes of the TCA cycle and glycolisis decreased<br />

(cytrate synthase; glycerol phosphate dehydrogenase, phosphoglycerate kinase 1 etc), still<br />

the quantity of proteasomal subunit, calpain, aspartyl aminopeptidase,<br />

phosphoglucomutase increased. Moreover, the EGFR-binding protein, peroxiredoxin,<br />

glutathione-S-transferase also increased. The enhanced protein turnover and increased<br />

antioxidant mechanisms provide a new aspect to the known neuroprotective mechanism of<br />

estrogen in the brain.<br />

Using bioinformatics we also studied the promoter regions of the corresponding genes, and<br />

only three of them contain ERE, but all have at least 7 half EREs. Based on these<br />

experiments, estrogen may effect the protein expression pattern acting on half ERE and/or<br />

AP-1 sites rather than ERE. These data may reformulate our view about the function of<br />

classical estrogen effects and estrogen receptor–ERE interaction. Our results also indicate<br />

that we can sucsessfully utilize the DIGE methodology to evaluate the effect of estrogen<br />

and ovariectomy on the brain proteome. We suppose that the enhanced rate of protein<br />

turnover and antioxidant mechanisms can be an important aspect of estrogen induced<br />

neuroprotection.<br />

Research supported by: RET, MEDICHEM, OTKA, ETT<br />

180


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

SELECTIVE ESTROGEN RECEPTOR MODULATORS DECREASE<br />

MICROGLIA ACTIVATION IN THE CEREBELLUM OF MALE RATS<br />

Tapia González S.*, Diz-Chaves Y., Pernía O., Carrero P., Garcia-Segura L.M.<br />

Instituto Cajal, C.S.I.C., Avenida Doctor Arce, 37, E-28002, Madrid, Spain.<br />

*E-mail: stapia75@cajal.csic.es FAX: +34-915854754<br />

Estradiol prevents neuronal loss in diverse experimental models of<br />

neurodegenerative diseases and enhances cognitive skills in animals and humans. Antiinflammatory<br />

actions of the hormone may represent an important component of its<br />

neuroprotective effects and previous studies have shown that estrogen therapy reduces the<br />

activation of microglia after acute brain lesions in female rodents [3,4]. Selective estrogen<br />

receptor modulators (SERMs) may represent an alternative to estrogen therapy for the<br />

treatment or the prevention of neurodegenerative disorders in humans. In the present study<br />

we have assessed the neuroprotective potency of several SERMs in male rats injected with<br />

lipopolysaccharide (LPS). Adult (3 months old) Wistar male rats were given two<br />

intraperitoneal injections of LPS, separated by an interval of 3 days [2]. One hour before<br />

the administration of LPS the animals were injected with vehicle or estrogenic compounds.<br />

Animals were killed 7 days after the first injection of LPS and a morphometric analysis of<br />

OX42 and major histocompatibility complex class II (MHCII) immunoreactive microglia<br />

in the central white matter of the cerebellum was performed using the optical disector<br />

method. OX42 was used as a marker of total microglia and MHC-II as a marker of reactive<br />

microglia. The estrogenic compounds assessed were: 17 beta-estradiol (50, 250, 500 and<br />

700 micrograms/Kg bw), tamoxifen (0.6, 1 and 2 mg/Kg bw), raloxifene (1 mg/Kg bw),<br />

lasofoxifene (1 mg/Kg bw) and bazedoxifene (2 mg/Kg bw). The selected doses of<br />

estrogenic compounds were based on previous studies that analyzed their neuroprotective<br />

properties [1]. LPS induced a significant increase in the number of MHC-II<br />

immunoreactive cells in the white matter of the cerebellum. The estrogenic compounds did<br />

not affect the basal number of MHC-II immunoreactive cells. However, estradiol,<br />

tamoxifen, raloxifene and bazedoxifene significantly reduced the number of MHC-II<br />

immunoreactive cells in LPS injected animals. The number of OX42 immunoreactive cells<br />

was neither affected by LPS administration nor by treatment with estrogenic compounds,<br />

suggesting that changes in MHC-II reflect differences in microglia activation and not<br />

differences in proliferation or survival. In summary, our data suggest that some SERMs<br />

may exert anti-inflammatory effects in the brain, reducing reactive microglia.<br />

References list<br />

[1] I. Ciriza, P. Carrero, I. Azcoitia, S.G. Lundeen and L.M. Garcia-Segura, Selective estrogen receptor<br />

modulators protect hippocampal neurons from kainic acid excitotoxicity: differences with the effect of<br />

estradiol, J Neurobiol 61 (2004) 209-221.<br />

[2] Y.K. Ng and E.A. Ling, Induction of major histocompatibility class II antigen on microglial cells in<br />

postnatal and adult rats following intraperitoneal injections of lipopolysaccharide, Neurosci Res 28<br />

(1997) 111-118.<br />

[3] E. Vegeto, S. Belcredito, S. Etteri, S. Ghisletti, A. Brusadelli, C. Meda, A. Krust, S. Dupont, P. Ciana, P.<br />

Chambon and A. Maggi, Estrogen receptor-alpha mediates the brain antiinflammatory activity of<br />

estradiol, Proc Natl Acad Sci U S A 100 (2003) 9614-9619.<br />

[4] E. Vegeto, S. Belcredito, S. Ghisletti, C. Meda, S. Etteri and A. Maggi, The endogenous estrogen status<br />

regulates microglia reactivity in animal models of neuroinflammation, Endocrinology 147 (2006) 2263-<br />

2272.<br />

181


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

ALLOPREGNANOLONE (ALLO) EXERTS A PROTECTIVE EFFECT AGAINST<br />

OXIDATIVE STRESS IN NIEMANN PICK C CELLS<br />

Zampieri S * , Mellon SH # , Pittis MG * , Nevyjel M*, Bembi B*, Dardis A * .<br />

* Unita di Malattie Metaboliche, IRCCS Burlo Garofolo, Via dell’Istria 65/1, 34137,<br />

Trieste, Italy. FAX 390403785500, email: andrea.dardis@area.trieste.it<br />

# Department of Obstetrics, Gynecology and Reproductive Sciences, The Center for<br />

Reproductive Sciences, University of California, San Francisco, USA.<br />

Niemann Pick C disease (NPC) is an autosomal recessive neurodegenerative disorder<br />

caused by different mutations in the NPC1 (95% of cases) and NPC2 (5% of cases) genes.<br />

The abnormal function of either of these proteins leads to an accumulation of unesterified<br />

cholesterol and other lipids, such as sphingolipids and glycosphingolipids (GSLs) in the<br />

lysosomal compartment of the cell [1]. The molecular mechanisms underlying the<br />

pathophysiology in NPC disease are not clear. However, oxidative damage has been<br />

implicated in the pathophysiology of different neurological disorders and<br />

neurodegenerative diseases [2], and the effect of GSL accumulation on the intracellular<br />

redox state has been well documented [3], suggesting a possible role of oxidative stress in<br />

the pathophysiology of this disease. The synthesis of neurosteroids is altered in a time- and<br />

region-specific fashion in the BALB\c Niemann Pick type C mouse, and neurons and<br />

neuroglia expressing the steroidogenic enzymes are lost in the NPC mouse. In particular,<br />

the synthesis of ALLO is substantially diminished at birth, and decreases further over time.<br />

These data suggest a pivotal role of neurosteroidogenesis abnormalities in the phenotypic<br />

expression of the disease. In addition, the treatment of NP-C mice with ALLO increased<br />

their lifespan and delayed the onset of neurological impairment [4]. However, the<br />

molecular mechanism by which ALLO exerts these neuroprotective effects is not fully<br />

understood..<br />

In this study we determined whether the intracellular redox state might contribute to the<br />

pathophysiology of NPC disease, and analyzed the possible effects of (ALLO) on the<br />

oxidative damage in human NPC cells.<br />

We first analyzed the levels of reactive oxygen species (ROS) in cultured fibroblasts from<br />

NPC patients and healthy controls. ROS levels were approximately three times higher in<br />

NPC than in normal fibroblasts (p0.01). Higher<br />

concentration did not show a further benefitial effect. In addition, pretreatment of cells for<br />

24 h with 50 nM ALLO almost complete reverted peroxide-induced apoptosis (p


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

due, at least in part, to its antioxidant properties. In addition these results demonstrated that<br />

the effects of ALLO are quite pleiotropic and that probably other brain diseases, including,<br />

but not limited to congenital storage diseases, may benefit from similar treatments with<br />

neuro-active steroids.<br />

Reference list<br />

[1] Patterson, M.C., Vanier, M.T., Suzuki, K., Morris, J.A., Carstea, E.D., Neufeld, E..B., Blanchette-<br />

Mackie, E.J. and Pentchev, P.G., 2001, Niemann–Pick disease type C: a lipid trafficking disorder. In:<br />

C.R. Scriver, A.L. Beaudet, W.S. Sly, D. Valle, B. Childs, K.W. Kinzler and B. Vogelstein, Editors<br />

(eighth ed.), The Metabolic and Molecular Bases of Inherited Disease, McGraw Hill, New York, pp.<br />

3611–3634.<br />

[2] Culmsee, C., Landshamer, S., 2006. Molecular insights into mechanisms of the cell death program: role<br />

in the progression of neurodegenerative disorders. Curr Alzheimer Res. 3, 269-83<br />

[3] Garcia-Ruiz, C., Colell, A., Paris, R., and Fernandez-Checa J.C., 2000. Direct interaction of GD3<br />

ganglioside with mitochondria generates reactive oxygen species followed by mitochondrial permeability<br />

transition, cytochrome c release, and caspase activation. FASEB J. 14,847-858.<br />

[4] Griffin, L.D., Gong, W., Verot, L., Mellon, S.H., 2004. Niemann-Pick type C disease involves disrupted<br />

neurosteroidogenesis and responds to allopregnanolone. Nat Med. 10, 704-11.<br />

183


Posters’ Exhibition:<br />

Xenoestrogens and brain circuitries<br />

• Martini M., Miceli D., Palanza P., Viglietti-Panzica C., Panzica G.C. (Italy)<br />

Effects of Bisphenol A on the hypothalamic nitrinergic system of CD1 mouse<br />

• Sica M., Håkansson H., Halldin K. (Sweden) Effects on estrogenic activity in<br />

pituitary gland and hypothalamus of male ERE reporter mice, after single oral<br />

TCDD exposure


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

EFFECTS OF BISPHENOL A ON THE HYPOTHALAMIC NITRINERGIC<br />

SYSTEM OF CD1 MOUSE<br />

Martini M.*, Miceli D.*, Palanza P.°, Viglietti-Panzica C.*, Panzica G.C.*<br />

*Laboratory of Neuroendocrinology, Dept Anatomy, Pharmacology and Forensic<br />

Medicine, C.so M. D’Azeglio, 52. 10126 Torino (Italy)<br />

e-mail: mariaangela.martini@unito.it<br />

°Dept Evolutive And Functional Biology, Parma (Italy)<br />

Bisphenol A (BPA) is a well-known pollutant derived from plastic used for aliments and is<br />

characterized by the ability of binding to estrogen receptors [1,8]. It is therefore considered<br />

an endocrine disrupting chemical of the category of xenoestrogens [3]. Several studies<br />

have been performed to elucidate its toxicological properties and its impact on different<br />

behaviors in rodents [4,9,10]. However, studies on the effects of BPA on neural circuits are<br />

at the moment limited to the catecholaminergic system [2,5-7]. In the present study we<br />

have investigated the effects of early exposure to BPA on the nitrinergic system of adult<br />

mice of both sexes.<br />

Pregnant mice (11th gestational day) were daily treated with different doses of BPA up to<br />

8 days after delivery. In this way, the pups of both sexes were exposed for 10 prenatal and<br />

8 postnatal days at BPA. Mice treated in this way were sacrificed at the age of 2 months by<br />

intracardiac perfusion and brains were dissected, frozen and sectioned. Serial sections<br />

taken each 100 µm were then treated for the immunohistochemical detection of nNOS<br />

(antibody from Diasorin, at a dilution of 1:12,000). Quantitative analysis of the number of<br />

nNOS-ir cell bodies was performed for 6 nuclei (medial preoptic nucleus, POM; nucleus of<br />

the stria terminalis, BST; paraventricular nucleus; ventromedial nucleus; arcuate nucleus,<br />

and caudate-putamen). Significant effects of BPA treatments were observed only in the<br />

POM and in the ventral subdivision of BST (BSTv), in a sex-oriented way. In the POM,<br />

where the nNOS-ir population is sexually dimorphic in controls, the higher doses of BPA<br />

(40 µg/kg/day) induced a decrease in the cell number in males, whereas the intermediate<br />

doses (20 µg/kg/day) induced an increase in the females. In the BSTv, where no sex<br />

dimorphism was observed in controls, the intermediate doses of BPA induced a significant<br />

decrease of the cell number only in males.<br />

These results indicate that BPA has a powerful effect on specific portions of the nNOS-ir<br />

system (i.e. POM and BSTv) that are particularly important for the control of male sexual<br />

behavior. In addition, they confirm that precocious exposure to xenoestrogens, in particular<br />

to BPA, may have a high impact on the organization of specific neural pathways that can<br />

later affect complex behaviors or functions as those related to reproduction.<br />

Reference list<br />

[1] Brotons, J.A., Olea-Serrano, M.F., Villalobos, M., Pedraza, V. and Olea, N., Xenoestrogens released<br />

from lacquer coatings in food cans, Environ Health Perspect, 103 (1995) 608-12.<br />

[2] Chu, H.P. and Etgen, A.M., A Potential Role of Cyclic GMP in the Regulation of Lordosis Behavior<br />

of Female Rats, Hormones and Behavior, 32 (1997) 125–132.<br />

[3] Fujimoto, T., Kubo, K. and Aou, S., Prenatal exposure to bisphenol A impairs sexual differentiation<br />

of exploratory behavior and increases depression-like behavior in rats, Brain Res, 1068 (2006) 49-<br />

55.<br />

[4] Henley, D.V. and Korach, K.S., Endocrine-disrupting chemicals use distinct mechanisms of action<br />

to modulate endocrine system function, Endocrinology, 147 (2006) S25-32.<br />

[5] Hull, E.M., Du, J., Lorrain, D.S. and Matuszewich, L., Testosterone, preoptic dopamine, and<br />

copulation in male rats. In: Hormones, Brain, and Behavior (G.C.Panzica and J.Balthazart eds),<br />

Brain Research Bullettin, 44 (1997) 327-333.<br />

185


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

[6] Hull, E.M., Lorrain, D.S., Du, J., Matuszewich, L., Lumley, L.A., Putnam, S.K. and Moses, J.,<br />

Hormone-neurotransmitter interactions in the control of sexual behavior, Behavioural Brain<br />

Research, 105 (1999) 105-116.<br />

[7] Hull, K.L. and Harvey, S., GH as a co-gonadotropin: the relevance of correlative changes in GH<br />

secretion and reproductive state, Journal of Endocrinology, 172 (2002) 1-19.<br />

[8] Olea, N., Pulgar, R., Perez, P., Olea-Serrano, F., Rivas, A., Novillo-Fertrell, A., Pedraza, V., Soto,<br />

A.M. and Sonnenschein, C., Estrogenicity of resin-based composites and sealants used in dentistry,<br />

Environ Health Perspect, 104 (1996) 298-305.<br />

[9] vom Saal, F.S., Timms, B.G., Montano, M.M., Palanza, P., Thayer, K.A., S.C., N., Dhar, M.D.,<br />

Parmigiani, S. and Welshons, W.V., Prostate enlargement in mice due to fetal exposure to low doses<br />

of estradiol or diethylstilbestrol and opposite effects at high doses, Proceedings of the National<br />

Academy of Sciences of the United States of America, 94 (1997) 2056–2061.<br />

[10] Welshons, W.V., Thayer, K.A., Judy, B.M., Taylor, J.A., Curran, E.M. and vom Saal, F.S., Large<br />

effects from small exposures. I. Mechanisms for endocrine-disrupting chemicals with estrogenic<br />

activity, Environ Health Perspect, 111 (2003) 994-1006.<br />

186


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

EFFECTS ON ESTROGENIC ACTIVITY IN PITUITARY GLAND AND<br />

HYPOTHALAMUS OF MALE ERE REPORTER MICE, AFTER SINGLE ORAL<br />

TCDD EXPOSURE<br />

Sica M., Håkansson H., Halldin K.<br />

Institute of Environmental Medicine, Karolinska Institute, nobels väg 13, 171 77 Stockholm,<br />

Sweden<br />

Many chemicals are known to interfere with endocrine system. 2,3,7,8-tetrachlorodibenzo-pdioxin<br />

(TCDD) is a potent endocrine disrupter that provokes disturbances in male [1,2,3,4]<br />

and female reproduction [5,6], and in the central nervous system (CNS).<br />

Neurodevelopmental delays and neurobehavioral effects on children [6, 7], impairment of<br />

seratonine metabolism [8] and alteration in brain sexual dimorphism [9, 10], have been<br />

observed.<br />

The toxic effects of TCDD are due to binding to a soluble, ligand –activated transcription<br />

factor (AhR). During the last few years, there has been growing evidence for molecular<br />

interactions between estrogen receptor (ER) and AhR signaling [11, 12].<br />

The present study was designed to investigate rapid effects of TCDD on estrogen signaling<br />

in the pituitary gland and hypothalamus, with major focus on nuclei that control reproductive<br />

behavior. The main objective of our studies is to clarify whether disturbed estrogen<br />

signaling is involved in the observed effects of dioxin-like contaminants on reproductive<br />

function.<br />

For this purpose an in vivo model that detects activation of estrogen receptors (ERs) was<br />

used. 66 transgenic mice carrying a luciferase reporter gene under control of three ERE<br />

sequence (3X ERE reporter mice; 13) were dosed with 200µg/kg bw of TCDD or vehicle<br />

(corn oil) and sacrificed at several time points (6h, 24h, 3 days, 7 days, 14 days) after<br />

exposure. Brains and pituitary were removed from the skull and the pituitary were snap<br />

frozen in liquid nitrogen and used for luciferase assay. Brains were immersed in acrolein<br />

10%, frozen in isopentane, sectioned at cryostat and processed for anti-luciferase<br />

immunohistochemistry.<br />

Luciferase activity in brains has been determined by means anti-luciferase<br />

immunohistochemistry, in order to give a more detailed picture of the rapid effects of TCDD<br />

on estrogenic activity in the different nuclei of male mouse hypothalamus.<br />

Luciferase assay performed on pituitary did not show any clear change regarding luciferase<br />

activity in all the time point analyzed.<br />

Preliminary results on antiluciferase immunostaining show a wide distribution of luciferase<br />

immunoreactive cells on hypothalamus and limbic system. Luciferase immunoreactive<br />

elements was observed in lateral septum, medial preoptic area, amygdala, ventromedial<br />

nucleus and arcuate nucleus. The distribution of luciferase immunoreactive distribution is<br />

comparable to estrogen receptors alpha and beta distribution.<br />

At 7 days after TCDD exposure an increase in the number of luciferase immunoreactive cells<br />

was noticed in the medial preoptic nucleus and currently other time points and other nuclei<br />

of hypothalamus are under investigation.<br />

Our data show that TCDD may interfere on estrogen receptor signaling in a tissue-specific<br />

manner.<br />

187


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

Acknowledgments The authors wish to thank, Louise Lyrenäs, Christina Trossvik, Daniel Borg,<br />

Maria Herlin for excellent technical support. This work is supported from the EU-research<br />

projects BoneTox (QLRT-2002-02528) and CASCADE (FOOD-CT-2004-506319). Monica Sica is<br />

recipient of a grant from Blanceflor Boncompagni-Ludovisi nee Bildt Foundation.<br />

References list<br />

[1] Sommer R.J., Ippolito D.L., Peterson R.E., 1996. In utero and lactational exposure of the male<br />

Holtzman rat to 2,3,7,8-tetrachlorodibenzo-p-dioxin: decreased epididymal and ejaculated sperm<br />

numbers without alterations in sperm transit rate. Toxicol Appl Pharmacol. 140,146-539.<br />

[2] Bjerke D.L., Peterson R.E. 1994. Reproductive toxicity of 2,3,7,8-tetrachlorodibenzo-p-dioxin<br />

inmale rats: different effects of in utero versus lactational exposure. Toxicology and Applied<br />

Pharmacology 127,241-9.<br />

[3] Kakeyama M., Sone H., Miyabara Y., Tohyama C., 2003. Perinatal exposure to 2,3,7,8,-<br />

tetrachlorodibenzo-p-dioxin alters activity-dependent expression of BDNF mRNA in the neocortex<br />

and male rat sexual behavior in adulthood. Neurotoxicology 24, 207-217.<br />

[4] Ikeda M., Tamura M., Yamashita J., Suzuki C., Tomita T.,2005. Repeated in utero and lactational<br />

2,3,7,8-tetrachlorodibenzo-p-dioxin exposure affects male gonads in offspring, leading to sex ratio<br />

changes in F2 progeny. Toxicology and Applied Pharmacology 133, 285-294<br />

[5] Gray L.E and Ostby J.S., 1995. In utero 2,3,7,8-Tetraclhorodibenzo-p-dioxin (TCDD) alters<br />

reproductive morphology and function in female rat offspring. Toxicology and Applied<br />

Pharmacology 133,285-294.<br />

[6] Vorderstrasse B.A., Fenton S.E., Bohn A.A., Cundiff J.A., Lawrence B.P., 2004. A novel effect of<br />

dioxin: exposure during pregnancy severely impairs mammary gland differentiation. Toxicology<br />

Science. 78, 248-57<br />

[7] Patandin S., Lanting C.I., Mulder P.G., Boersma E.R., Sauer P.J., Weisglas-Kuperus N., 1999.<br />

Effects of environmental exposure to polychlorinated biphenyls and dioxins on cognitive abilities in<br />

Dutch children at 42 months of age. J Pediatr. 134,33-41.<br />

[8] Kuchiiwa S., Cheng S.B., Nagatomo I., Akasaki Y., Uchida M., Tominaga M., Hashiguchi W.,<br />

Kuchiiwa T., 2002. In utero and lactational exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin<br />

decreases serotonin immunoreactive neurons in raphe nuclei of male mouse offspring. Neurosci<br />

Lett. 317, 73-6<br />

[9] Zareba G., Hojo R., Zareba K.M., Watanabe C., Markowski V.P., Baggs R.B, Weiss B., 2002.<br />

Sexually dimorphic alterations of brain cortical dominance in rats prenatally exposed to TCDD.<br />

Journal of Applied Toxicology 129-37<br />

[10] Hojo R, Stern S., Zareba G, Markowski VP, Cox C., Weiss B. 2002. Sexually dimorphic behavioral<br />

response to prenatal dioxin exposure. Environmental Health Perspective 247-254.<br />

[11] Brunnberg S., Pettersson K., Rydin E., Matthews J., Hanberg A., Pongratz I.,2003. The basic helixloop-helix-PAS<br />

protein ARNT functions as a potent coactivator of estrogen receptor-dependent<br />

transcription. Proc Natl Acad Sci U S A. 27, 6517-22.<br />

[12] Chaffin C.L., Trewin A.L., Hutz R.J. 2000. Estrous cycle-dependent changes in the expression of<br />

aromatic hydrocarbon receptor (AHR) and AHR-nuclear translocator (ARNT) mRNAs in the rat<br />

ovary and liver Chem Biol Interact. 124, 205-1623<br />

[13] Lemmesn J.G., Arends R.J., Ven Boxtel A.L., Van der Saag P.T., Van Der Burg B., 2004. Tissue<br />

and time-ependent estrogen receptor activation in estrogen reporter mice Journal of Molecular<br />

Endocrinology 32,689-701.<br />

188


Posters’ Exhibition:<br />

Effects mediated by membrane receptors<br />

• Dieni C.V., Tobin V., Menzies JRW., Dutia M.B. (Italy) Effects of THDOC and<br />

allopregnanolone on the gabaergic current evoked by muscimol in the neurons of<br />

the rat medial vestibular nuclei<br />

• Frondaroli A., Grassi S., Pettorossi V.E. (Italy) Long-term effects of THDOC on<br />

the neuronal activity of the rat medial vestibular nuclei<br />

• Hariri O.R., Micevych P.E. (USA) The nongenomic effects of estrogen regulates<br />

the effects of oxytocin in the hypothalamic astrocytes<br />

• Lindblad C., Bierzniece V., Turkmen S., Bäckström T. and Johansson I-M.<br />

(Sweden) Metabolism prevent the UC1010 antagonistic effect to allopregnanolone<br />

in the morris water maze<br />

• G. Ragagnin, M. Rahman, E. Zingmark, J. Strömberg, P. Lundgren, M. Wang, T.<br />

Bäckström (Sweden) Structure-activity relationship of GABA A -steroids antagonists


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

EFFECTS OF THDOC AND ALLOPREGNANOLONE ON THE GABAERGIC<br />

CURRENT EVOKED BY MUSCIMOL IN THE NEURONS OF THE RAT<br />

MEDIAL VESTIBULAR NUCLEI<br />

Dieni C.V.*, Tobin V.°, Menzies JRW.°, and Dutia M.B.°<br />

* Department of Internal Medicine, Section of Human Physiology, University of Perugia,<br />

Via del Giochetto, 06100, Perugia, Italy, Fax: +39-0755857371<br />

email: cristinavd@libero.it<br />

° Centre for Integrative Physiology, Biomedical Sciences, Hugh Robson Building, George<br />

Square, Edinburgh, EH8, 9XD, UK, email: m.b.dutia@ed.ac.uk<br />

Introduction: Plasticity in the medial vestibular nuclei (MVN) after unilateral<br />

labyrinthectomy and the resultant vestibular compensation involve time-dependent<br />

changes in MVN neuron intrinsic excitability and a re-balancing of the neuronal activity of<br />

the bilateral MVN [3], in part through changes in GABA receptor function [2, 5].<br />

Neurosteroids are known to modulate GABA receptor function in many brain areas [1, 6]<br />

by primarily but not exclusively increasing the sensitivity of the receptor and thus<br />

increasing inhibitory drive. The enzymes responsible for the de novo synthesis of<br />

neurosteroids are present in the MVN [4] and previous work in our lab has shown that<br />

neurosteroid exposure can increase the inhibitory effect of the GABA A receptor agonist<br />

Muscimol on the spontaneous firing rate of MVN neurons (unpublished data). Thus, we<br />

hypothesised that the Muscimol-induced Cl - current in MVN neurons may be augmented<br />

by exposure to the neurosteroids Allopregnanolone (ALLO) and 3α, 5αtetrahydrodeoxycorticosterone<br />

(THDOC).<br />

Methods: Coronal brainstem slices including the rostral half of the MVN (200-250 µm)<br />

were prepared from Lister Hooded rats (P14-P21) and incubated (60 min, <strong>35</strong> o C) in<br />

oxygenated artificial cerebrospinal fluid (ACSF). The slices were then transferred to a<br />

recording chamber (<strong>35</strong> o C) with constant oxygenated perfusion of ACSF or drugs dissolved<br />

in ACSF (2 ml/min). Neurons were identified using IR-DIC optics and generally had ovoid<br />

soma with 15 µm diameter with at least 2 processes visible. Patch pipettes (7-8 MΩ) were<br />

filled with (in mM): CsMeSO 4 145, HEPES 5, EGTA 0.1, MgATP 5, pH 7.2, 290-295<br />

mOsm). Clamping the membrane potential at -30 mV and dialysis of the cells with cesium<br />

blocked sodium and potassium currents respectively, allowing us to record the whole cell<br />

Cl - current elicited by bath application of Muscimol (3 µM, 1 min). Currents were acquired<br />

using Axon 200B amplifier and Clampex v9 software and filtered at 2 kHz. Both<br />

amplitude and area of current recording (integral of evoked current minus holding current)<br />

were measured using Clampfit v9. These were measured before (control) and 5, 15 and 25<br />

min after the start of neurosteroid application (ALLO 3 µM or THDOC 3 µM, 5 min) and<br />

compared using either paired t-test or one-way ANOVAs with Holm-Sidak post-hoc<br />

analysis. P values of less than 0.05 were accepted as significantly different.<br />

Results: In 20 cells, the amplitude of the GABA current immediately after ALLO was<br />

significantly increased (124.8 ± 23.9 pA vs. 90.13 ± 13.8 pA, P


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

the same intervals as above but in the absence of neurosteroid were unchanged compared<br />

to the first response.<br />

Conclusions: These results indicate that ALLO and to a lesser extent THDOC augment<br />

Muscimol-induced GABAergic currents in MVN neurons, consistent with previous studies<br />

in other brain regions that have shown neurosteroids can enhance the sensitivity of the<br />

GABA A receptor to its ligand. An interesting finding is the persistence of this enhancement<br />

after wash out of the neurosteroid. This may reflect either binding of neurosteroid to the<br />

GABA A receptor or a persistent change in the receptor function. Further studies are<br />

ongoing to clarify the biochemical mechanisms utilised by the neurosteroids. However, our<br />

results demonstrate that neurosteroids can modify the neuronal activity of MVN neurons<br />

by facilitating GABA currents and therefore inhibitory drive. As vestibular compensation<br />

is dependent on stress hormones [3], and stimulation of the stress axis has been shown to<br />

increase production of neurosteroids in different brain areas [6], we suggest that post-UL<br />

compensatory changes may involve the actions of neurosteroids in the vestibular nuclei. A<br />

possible mechanism may be an enhancement of GABAergic drive to induce long lasting<br />

depression of discharge of the MVN neurons in the intact side, while GABAergic receptor<br />

sensitivity of the MVN neurons on the lesioned side is down-regulated [5].<br />

Reference list<br />

1. E.E. Baulieu, Neurosteroids: a novel function of the brain, Psychoneuroendocrinol. 23 (1998) 963-987.<br />

1998.<br />

2. S.A. Cameron, M.B. Dutia, Cellular basis of vestibular compensation: changes in intrinsic excitability of<br />

MVN neurons, Neuroreport 8 (1997) 2595-2599.<br />

3. S.A. Cameron, M.B. Dutia, Lesion-induced plasticity in rat vestibular nucleus neurones dependent on<br />

glucocorticoid receptor activation, J. Physiol. 518 (1999) 151-158.<br />

4. E. Dupont, J. Simard, V. Luu The, F. Labrie, G. Pelletier, Localization of 3 beta-hydroxysteroid<br />

dehydrogenase in rat brain as studied by in situ hybridization. Mol. Cell. Neurosci. 5 (1994) 119-123.<br />

5. T. Yamanaka, A. Him, S.A. Cameron, M.B. Dutia, Rapid compensatory changes in GABA receptor<br />

efficacy in rat vestibular neurones after unilateral labyrinthectomy, J. Physiol. 523 (2000) 413-424.<br />

6. R.H. Purdy, A.L. Morrow, P.H. Jr. Moore, S.M. Paul, Stress-induced elevations of γ-aminobutyric acid<br />

type A receptors-active steroids on the rat brain. Proc. Natl. Acad. Sci. USA 88 (1991) 4553-4557.<br />

191


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

LONG-TERM EFFECTS OF THDOC ON THE NEURONAL ACTIVITY OF THE<br />

RAT MEDIAL VESTIBULAR NUCLEI<br />

Frondaroli A., Grassi S., Pettorossi V.E.<br />

Department of Internal Medicine, Section of Human Physiology, University of Perugia, Via del Giochetto,<br />

06100, Perugia, Italy, Fax: +39-0755857371 email: sgrassi@unipg.it<br />

Aim: Stressfull events induce synthesis of neurosteroids (NS) in the brain, which rapidly<br />

influence the neuronal excitability by acting on different neurotransmitter receptors,<br />

including GABA A and ionotropic glutamate (NMDA and AMPA) receptors [1].<br />

There is evidence that the medial vestibular nuclei (MVN) are susceptible to stress<br />

hormones (glucocorticoids, GC), as activation of GC receptors plays a role in vestibular<br />

compensation, which leads to the re-balance of neuronal activity in the medial vestibular<br />

nuclei (MVN) after unilateral labyrinthectomy (UL) [2]. Besides GC, it is also possible<br />

that NS are involved in the compensation, since NS could be locally synthesized in the<br />

MVN [3]. To evidence the role of NS in the vestibular plasticity phenomena, it is firstly<br />

necessary to demonstrate their influence on the spontaneous and synaptically driven<br />

activities of the MVN neurons and, secondly, to evidence possible long term effects. The<br />

NS used in this study was the tetrahydrodeoxycorticosterone (THDOC), since its secretion<br />

is correlated with that of GC [5].<br />

Methods: The experiments were performed in rat transverse brainstem slices (<strong>35</strong>0 µm),<br />

containing the medial vestibular nuclei (MVN), which were perfused with oxygenated<br />

artificial cerebrospinal fluid (2 ml/min) and maintained at 30-31 °C. The field potentials<br />

were recorded in the ventral part (Vp) of the MVN by stimulating (40-100 µA intensity,<br />

0.07 ms duration and 0.06 Hz frequency) the ipsilateral vestibular afferents at the point<br />

where they enter the MVN. Single unit potentials were also extracellularly recorded. The<br />

effect of THDOC (3 µM, applied for 15 min), was assayed on the amplitude of the field<br />

potential N1 wave (expressed as a percentage of the baseline) and on spontaneous firing<br />

rate of single neurons (spikes/sec). To recognise the receptors which could be involved in<br />

mediating the THDOC effects, we used the antagonists for the following receptors:<br />

GABA A (Bicuculline, 30 µM), NMDA (DL-AP5, 100 µM), group I metabotropic<br />

glutamate (mGlu-I) (AIDA, 200 µM) and AMPA/kainate (NBQX, 10µM).<br />

Results: THDOC induced long-lasting changes of the field potential amplitude consisting<br />

in both depressions and potentiations (Fig. 1A). In addition, THDOC induced two<br />

temporally distinct effects, which were not necessary combined: the early one, developing<br />

at 7-12 min after the start of drug application, and the late one at 20 - 30 min. Under block<br />

of GABA A receptors (Bicuculline), field potential depressions were annulled, while the<br />

occurrence of potentiation remained unchanged (Fig. 1A). Under this condition, the block<br />

of NMDA (AP-5) and mGlu-I (AIDA) receptors did not significantly change the<br />

occurrence of early and late potentiations (Fig. 1A and B), contrary to that normally<br />

observed in the vestibular synaptic long-term potentiation [4]. The same result was<br />

obtained by analysing the effect of THDOC on the neuron firing rate in normal condition,<br />

under Bicuculline and under block of NMDA and mGlu-I receptors (Fig. 1C and D). By<br />

contrast, the block of AMPA/kainate (NBQX) receptors significantly modified the<br />

occurrences of both early and late potentiations, which were respectively significantly<br />

reduced and increased, compared with those observed under Bicuculline alone (Fig. 1D).<br />

Conclusions: These results show that THDOC changes synaptic and spontaneous activity<br />

of the MVN neurons by inducing long-lasting effects. These effects consisted in depression<br />

and potentiation showing short or long induction time (early and late effects). The analysis<br />

of the possible action mechanism of THDOC indicates that 1) early and late depressions<br />

192


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

depend on a facilitatory influence of THDOC on GABA A receptors, which mediate the<br />

inhibitory transmission in the MVN; 2) the early THDOC potentiation depends on a<br />

facilitation of AMPA/kainate mediated glutamate transmission, which might be due to the<br />

enhancement of receptor sensitivity to glutamate and/or to the increase of glutamate<br />

release; 3) the late THDOC potentiation could depend on the enhancement of intrinsic<br />

neuronal excitability, probably through an effect on ionic channels, since it is not annulled<br />

by the whole glutamate receptor block. In conclusion, THDOC modifies the activity of<br />

MVN neuronal circuitry rapidly, by increasing the inhibitory and excitatory transmission,<br />

through potentiation of GABA A and AMPA/kainite receptor activation, and changes the<br />

neuronal intrinsic excitability, later. The presence of different influences of THDOC, on<br />

vestibular activity, in terms of sign and timing of effects, may explain opposite behaviour<br />

of vestibular neurons during UL compensation, which requires an increase of activity in<br />

the ipsilesional side and decrease in the contralateral side.<br />

Reference list<br />

1. E.E. Baulieu, Neurosteroids: a novel function of the brain, Psychoneuroendocrinol. 23 (1998) 963-987.<br />

1998.<br />

2. S.A. Cameron, M.B. Dutia, Lesion-induced plasticity in rat vestibular nucleus neurones dependent on<br />

glucocorticoid receptor activation, J. Physiol. 518 (1999) 151-158.<br />

3. E. Dupont, J. Simard, V. Luu The, F. Labrie, G. Pelletier, Localization of 3 beta-hydroxysteroid<br />

dehydrogenase in rat brain as studied by in situ hybridization. Mol. Cell. Neurosci. 5 (1994) 119-123.<br />

4. S. Grassi, V.E. Pettorossi, Synaptic plasticity in the medial vestibular nuclei: role of glutamate<br />

receptors and retrograde messengers in rat brainstem slices. Prog. Neurobiol. 64/6 (2001) 527-553.<br />

5. R.H. Purdy, A.L. Morrow, P.H. Jr. Moore, S.M. Paul, Stress-induced elevations of γ-aminobutyric acid<br />

type A receptors-active steroids on the rat brain. Proc. Natl. Acad. Sci. USA 88 (1991) 4553-4557.<br />

Event occurrence (%)<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

A<br />

*<br />

*<br />

P otentiation (early + l ate)<br />

Depression<br />

Null effect<br />

P otentiation occurrence (%)<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

B<br />

*<br />

Early potentiation<br />

Late potentiation<br />

0<br />

0<br />

Event occurrence (%)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

C<br />

*<br />

*<br />

*<br />

Potentiation (early + late)<br />

Depression<br />

Null effect<br />

P otentiation occurrence (%)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

D<br />

*<br />

*<br />

*<br />

*<br />

Early potentiation<br />

Late potenziation<br />

0<br />

0<br />

Fig. 1 Occurrence percentage (mean ± SD,* p < 0.05) of the effects induced by THDOC in the MVN on the<br />

field potential amplitude (A and B) and on neuronal firing rate (C and D) under normal condition (control),<br />

under block of GABA A receptors (Bicuculline) and under block of NMDA (AP-5), mGlu-I (AIDA) and/or<br />

AMPA (NBQX) glutamate receptors.<br />

193


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

THE NONGENOMIC EFFECTS OF ESTROGEN REGULATES THE EFFECTS OF<br />

OXYTOCIN IN THE HYPOTHALAMIC ASTROCYTES<br />

Hariri O.R. * , Micevych P.E. *<br />

* University of California, Los Angeles, Le Conte Ave, Los Angeles, CA, USA. Fax +310-825-<br />

2224 Email: pmicevych@mednet.ucla.edu<br />

Astrocytes have important roles in the development and physiology of the central nervous<br />

system. In the hypothalamus, astrocytes have been implicated in sexual differentiation, synthesis of<br />

neurosteroids and synaptic function. Astrocytes can regulate neuronal activity through physical<br />

integration of synaptic signals, regulate interneuronal contacts and have been recently implicated in<br />

estrogen positive feedback responses. It is clear that astrocyte activity has implications for the<br />

functional understanding of a variety of different circuits systems. Astrocyte activity can be<br />

tracked by observing variations in intracellular free cytoplasmic calcium levels ([Ca 2+ ] i ). Previous<br />

studies have demonstrated that hypothalamic astrocytes respond to both estradiol and oxytocin<br />

(OT). Studies have shown that a high proportion of ventromedial hypothalamic neurons express<br />

estrogen receptors (ER) mRNAs and oxytocin receptor (OTR) mRNA. Although estradiol<br />

influences the expression of oxytocin receptors (OTR) in neurons, there is a paucity of studies that<br />

examined the interaction ER and OTR inastrocytes. In non-neuronal cells of the corpus luteum,<br />

chronic estradiol treatment caused the down-regulation of OTR. Because both estradiol and OT<br />

affect astrocytes and because their interaction may have important consequences for astrocyte<br />

function, we tested whether estradiol could modulate oxytocin signaling in hypothalamic<br />

astrocytes. Separate studies have shown that both rapid actions of estradiol and oxytocin increase<br />

free cytoplasmic calcium levels [Ca 2+ ] i in astrocytes, but the proximal steps of their signaling<br />

mechanisms have not been examined. The present study examines the interaction of E2 and OT on<br />

cell signaling by monitoring ([Ca 2+ ] i transients in post-pubertal astrocytes.<br />

We used digital fluorescence microscopy with Fura-4 as the Ca 2+ indicator and tested the<br />

response of post-pubertal hypothalamic astrocytes to both OT and estradiol. Application of<br />

estradiol (4nM) to the astrocyte cultures rapidly increased the [Ca+2]i by 570.2 ± 70 (relative<br />

units). Also, application of OT (400nM) increased the [Ca 2+ ] i 285.2 ± 42. In another experiment,<br />

astrocytes were treated with 400nM OT inducing a significant increase in [Ca 2+ ] i levels . After a<br />

washout of three minutes, addition of 2nM estradiol produced a significant increase in [Ca 2+ ] i . But<br />

when astrocytes were treated with estradiol (2nM) first and then OT (400nM), OT did not<br />

significantly increase the [Ca 2+ ] i levels, demonstrating a estradiol mediated attenuation of the OTinduced<br />

[Ca 2+ ] i response. In another experiment, astrocytes were treated chronically with 10nM<br />

estradiol for 22 hours. Astrocytes did not have a [Ca 2+ ] i response to OT challenge after the<br />

estradiol.<br />

Previous studies demonstrate that estradiol and OT rapidly activate phospholipase C (PLC)<br />

to induce the release of intracellular stores of calcium through the inositol trisphosphate (IP3)<br />

receptor, suggesting that OT and estradiol signaling converge onto a common intracellular<br />

pathway. Rapid, estradiol signaling in neurons appears to require an interaction of membrane ER<br />

with metabotropic glutamate receptor (mGluR), and the activation of [Ca 2+ ] i flux requires the type<br />

1a receptor (mGluR1a). To test whether the same mechanism was present in astrocytes, 20nM of<br />

LY367385, an mGluR1a antagonist, was added to the media. The estradiol-induced [Ca 2+ ] i<br />

transient was significantly attenuated. Interestingly, the OT-induced [Ca 2+ ] i transient was also<br />

prevented by LY367385, indicating that the convergence of E2 and OT signaling occurs at the<br />

mGluR1a, which activates a G-protein to activate the PLC-IP3 pathway.<br />

Although estradiol or OT separately rapidly increases [Ca 2+ ] i , flux in hypothalamic<br />

astrocytes, an order of treatment effect was also observed: pre-incubation/treatment with estradiol<br />

prevents OT signaling in hypothalamic astrocytes. These results demonstrate a negative estradiol<br />

regulation of OT signaling. We hypothesize that the association of the ER with mGluR1a prevents<br />

interaction of the OTR with the ER/mGluR complex, and therefore preventing OT signaling<br />

through the PLC-IP3 pathway. These results suggest regulatory mechanism through which<br />

estradiol modulates OT responses during pregnancy and lactation.<br />

194


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

METABOLISM PREVENT THE UC1010 ANTAGONISTIC EFFECT TO<br />

ALLOPREGNANOLONE IN THE MORRIS WATER MAZE<br />

Lindblad C., Bierzniece V., Turkmen S., Bäckström T. and Johansson I-M.<br />

Department of Clinical Sciences, Obstetrics and Gynecology, Umea University Hospital, S-901 85,<br />

Umea, Sweden. Fax: +46-90-776006, e-mail: charlotte.lindblad@obgyn.umu.se<br />

Background: The progesterone metabolite allopregnanolone (allo, 3α-OH-5α-pregnan-<br />

20-one) is a neurosteroid which act as a positive modulator on the GABA-A receptor and<br />

enhance the GABA activity. It is produced at many different sites and occasions. In<br />

humans, allo from corpus luteum is fluctuating during the menstrual cycle with its peak<br />

during the luteal phase. During pregnancy placenta produces allopregnanolone which is<br />

rising to high levels. Allopregnanolone is also increased during stress, but in this case the<br />

source is the adrenal glands. It can also be synthesized in the brain.<br />

Allopregnanolone has been shown to have negative effects on learning and memory in<br />

rat [2]. A key area for learning and memory is the hippocampus and it has been shown that<br />

the Morris water maze is a good model for testing spatial learning in rat [5]. In humans it<br />

has been shown that short-term spatial working memory is affected by menstrual cycle<br />

changes, with more errors in the luteal phase [4], when allo is accumulated in the brain [1].<br />

Since allo has these negative effects on cognition, mood and learning, it would be<br />

helpful to be able to block these negative effects. As an antagonist we have used UC1010<br />

(3β-OH-5α-pregnan-20-one) a 3β-isomer to allopregnanolone. UC1010 has been shown to<br />

work as an antagonist with several different in vitro systems. Wang et al. in 2000 showed<br />

that UC1010 antagonize the inhibitory effect of allopregnanolone on population spikes in<br />

the CA1 region of the rat hippocampus dose-dependently [6]. Lundgren et al. (2003)<br />

showed that allo-induced GABA-mediated chloride ion influx into rat cortical microsacs is<br />

inhibited by UC1010 [3].<br />

Aim: To block the negative allopregnanolone effects on learning in the Morris Water<br />

Maze with the antagonist UC1010.<br />

Material and Methods: Adult male Wistar rats were injected i.v. daily with allo (2 mg/kg,<br />

n=7), UC1010 (32 mg/kg, n=5), allo:UC1010 (2:32 mg/kg, n=5) and vehicle (10,7 ml/kg,<br />

n=4) respectively. Rats started a trial session for place navigation in the Morris Water<br />

maze 8 minutes after injection. Each session consisted of 4 searches for the platform.<br />

Animals were decapitated eight minutes after injection at day 8 (allo and vehicle) or 7<br />

(UC1010 and allo:UC1010). Trunk blood and selected brain areas were taken care of for<br />

further analyses.<br />

Results: The negative effects of allopregnanolone on learning in the Morris water maze<br />

were not blocked by UC1010 (fig. 1A, filled squares). Instead UC1010 enhanced the<br />

negative allo effect, i.e. rats injected with both allo and UC1010 didn’t find the platform at<br />

all. Allopregnanolone (filled circles) inhibited learning as shown earlier. Animals injected<br />

with only UC1010 (open squares) had an allo-like learning curve, with slightly, but not<br />

significant lower latency to find the platform compared to allo group. Vehicle group (open<br />

circles) learned to find the platform already at day two of the experiment.<br />

The lack of learning in the treated groups was not due to loss of swimming skills or<br />

reduced motor activity. The only group that differed in swim speed was the UC1010<br />

group, and those animals had a higher speed compared to the vehicle and allo groups (fig.<br />

1B).<br />

195


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

We see is a changed search pattern in all treated groups compared to the vehicle group<br />

(fig. 1C). Allo and UC1010 groups decreased their time spent close to the wall<br />

(thigmotaxis) throughout the experiment. But in the group that received both allo and<br />

UC1010 there was instead a slight increase in thigmotaxis time.<br />

Preliminary data of allopregnanolone concentrations in hippocampus shows higher<br />

levels of allo in the group that was injected with only UC1010 (3300 nmol/kg) than in<br />

animals injected with allo (2100 nmol/kg). The highest levels were found in animals<br />

injected with both allo and UC1010 (4600 nmol/kg). All treated groups had allo levels<br />

)<br />

several thousand times higher than the vehicle group (4 nmol/kg).<br />

)<br />

s<br />

(<br />

y<br />

c<br />

n<br />

e<br />

t<br />

a<br />

L<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Latency<br />

0<br />

0 1 2 3 4 5<br />

Day<br />

)<br />

s<br />

/<br />

m<br />

c<br />

(<br />

A B C<br />

)<br />

s<br />

/<br />

m<br />

c<br />

(<br />

d<br />

e<br />

e<br />

p<br />

s<br />

Figure 1. Morris water maze results<br />

m<br />

i<br />

40<br />

30<br />

20<br />

Swim speed<br />

40<br />

dVehicle<br />

e<br />

Allo (2 mg/kg)<br />

e 30<br />

pUC1010 (32 mg/kg)<br />

sAllo:UC1010 (2:32 mg/kg)<br />

20<br />

m<br />

i<br />

10<br />

w<br />

S<br />

0<br />

0 1 2 3 4 5<br />

Swim speed<br />

Day<br />

Thigmotaxis time (%)<br />

s 100<br />

i Vehicle<br />

x Allo 80(2 mg/kg)<br />

a UC1010 (32 mg/kg)<br />

t 60<br />

o<br />

Allo:UC1010 (2:32 mg/kg)<br />

m<br />

g<br />

40<br />

i<br />

h<br />

20<br />

T<br />

0<br />

0 1 2 3 4 5<br />

Discussion: UC1010 didn’t 10 succeed to block the negative effect of allo on learning in the<br />

w<br />

Morris water maze even though S it had been shown to work as an antagonist to allo in vitro<br />

0<br />

[2,6]. This taken together with the allo like effects we had seen in the group injected with<br />

0 1 2 3 4 5<br />

UC1010 and the high concentration of allo Day in the UC1010 group that had not been injected<br />

with allo points towards a possible metabolism of UC1010 to allopregnanolone within the<br />

brain.<br />

%<br />

(<br />

e<br />

m<br />

i<br />

t<br />

Day<br />

Vehicle<br />

Allo (2 mg/kg)<br />

UC1010 (32 mg/kg)<br />

Allo:UC1010 (2:32 mg/kg)<br />

Vehicle<br />

Allo (2 mg/kg)<br />

UC1010 (32 mg/kg)<br />

Allo:UC1010 (2:32 mg/kg)<br />

Reference list<br />

[1] M. Bixo, A. Andersson, B. Winblad, R.H. Purdy, T. Bäckström, Progesterone, 5α-pregnane-3,20-dione<br />

and 3α-hydroxy-5α-pregnane-20-one in specific regions of the human female brain in different endocrine states,<br />

Brain Res. 764 (1997) 173-178<br />

[2] I-M. Johansson, V. Birzniece, C. Lindblad, T. Olsson, T. Bäckström, Allopregnanolone inhibits learning<br />

in the Morris water maze, Brain Res. 934 (2002) 125-131<br />

[3] P. Lundgren, J. Strömberg, T. Bäckström, M. Wang, Allopregnanolone-stimulated GABA-mediated<br />

chloride ion flux is inhibited by 3β-hydroxy-5α-pregnan-20-one (isoallopregnanolone), Brain Res. 982 (2003)<br />

45-53<br />

[4] M.S. Man, I. MacMillan, J. Scott, A.H. Young, Mood, neuropsychological function and cognitions in<br />

premenstrual dysphoric disorder, Psychol. Med. 29 (1999) 727-733<br />

[5] G. Riedel, J. Micheau, A.G. Lam, E. Roloff, S.J. Martin, H. Bridge, L. Hoz, B. Poeschel, J. McCulloch,<br />

R.G. Morris, Reversible neural inactivation reveals hippocampal participation in several memory processes,<br />

Nat. Neurosci. 2 (1999) 898-905<br />

[6] M.D. Wang, T. Bäckström, S. Landgren, The inhibitory effects of allopregnanolone and pregnanolone on<br />

the population spike, evoked in the rat hippocampal CA1 stratum pyramidale In vitro, can be blocked<br />

selectively by epiallopregnanolone, Acta Physiol. Scand. 169 (2000) 333-341<br />

196


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

STRUCTURE-ACTIVITY RELATIONSHIP OF GABA A -STEROIDS<br />

ANTAGONISTS<br />

G. Ragagnin*, M. Rahman, E. Zingmark, J. Strömberg, P. Lundgren, M. Wang, T.<br />

Bäckström<br />

*Umeå Neurosteroid Research Center, Department of Clinical science, Obstetrics and<br />

Gynecology, Norrland University Hospital by5B, tr5, 901 85 Umeå, Sweden<br />

Fax +46 90 776006; e-mail: gianna.ragagnin@obgyn.umu.se<br />

Neuroactive steroids have effects on serotonin, glutamate and GABA systems.<br />

Their importance is enormous since they are involved in the regulation of mood, memory,<br />

Alzheimer’s disease, stress-induced depression and burn-out syndrome, anxiety disorders,<br />

postnatal and major depression as well as premenstrual syndrome (PMS), premenstrual<br />

dysforic disorder (PMDD), and mood changes related to hormone replacement therapy,.<br />

They are 3α-hydroxy-5α/β metabolites of progesterone (allopregnanolone and<br />

pregnanolone), testosterone (3α-hydroxy-5α-androstan-diol), and deoxycorticosterone<br />

(tetra-hydro-deoxycorticosterone).<br />

Allopregnanolone acts via the GABA A receptor by changing receptor’s expression or<br />

sensitivity. It is involved in premenstrual mood changes and induces cognitive deficits,<br />

such as spatial-learning impairment.<br />

Me<br />

O<br />

Me<br />

HO<br />

H<br />

Allopregnanolone<br />

Our aim is to find steroids than could antagonize the negative effects of allopregnanolone.<br />

A X-ray analysis of the GABA A receptor is required in order to design an optimal structure<br />

of a possible agonist/inverse agonist/antagonist by simple computer-aided modelling.<br />

Such an analysis is still not available yet, because of the receptor’s complexity. It must also<br />

be noticed that there are at least 20 different subunit compositions in the brain and there<br />

are evidences that more than one active site for steroids are present in a single receptor.<br />

We have synthesized, in some steps followed by chromatographic purification, a number<br />

of new steroids with potential activity as partial inverse agonists against allopregnanolone<br />

and neutral antagonists against GABA. The compounds have been tested in vitro by<br />

voltage-clamp and patch-clamp methods.<br />

The systematic investigation of several isomer combinations of GABA A -active<br />

neurosteroid derivatives, led us to the finding that, as a general rule, the following features<br />

are of importance for an effective drug:<br />

- Geometry between rings A/B is crucial<br />

- Hydrogen-bond donator in position 3<br />

- Hydrogen-bond acceptor in position 20<br />

- Flexible bond at position 17<br />

In addition, the title compounds fitfull the Lipinski’s rule of five for effective oral<br />

administration, and display acceptable values for logBBB.<br />

197


Posters’ Exhibition:<br />

Corticosteroid effects and stress<br />

• Berry A., Giorgio M., Martin-Padura I.; Pelicci P.G., de Kloet E.R., Alleva E.,<br />

Minghetti L. and Cirulli F. (Italy) Mice carrying a deletion of the P66 Shc gene show<br />

reduced behavioural and neuroendocrine responses to stressful stimuli<br />

• Hirst JJ , Palliser HK, Yates DM, Walker DW (Australia) Role of 5αReductase<br />

enzymes in the fetal brain and placenta in neurosteroid production following<br />

intrauterine growth restriction<br />

• Macrì S., Cirulli F. , Pasquali P. , Bonsignore L.T. , Laviola G. (Italy) Neonatal<br />

exposure to low doses of corticosterone increases novelty seeking and resistance to<br />

bacteria infection in adult male mice<br />

• Maggio N. and Segal M (Israel) Corticosteroids modulation of long term<br />

potentiation along the septo_temporal axis of the hippocampus<br />

• Morley-Fletcher S, Mairesse J, Daszuta A, Soumier A, Banasr M, Zuena AR,<br />

Mocaer E, Matteucci P, Casolini P, Catalani A, Maccari S. (France)<br />

Neuroplasticity in the prenatal stress rat model of depression: effects of<br />

agomelatine treatment<br />

• Ognibene E., Adriani W., Macrì S., Laviola G (Italy) Neurobehavioural disorders<br />

in the infant reeler mouse model: interaction of genetic vulnerability and<br />

consequences of maternal separation<br />

• Vafaei AA, Taherian AA, Jarrahi M (Iran) Assessment of modulatory effects of<br />

corticosterone on anxiety related behavior in mice<br />

• Velickovic N.A., Djordjevic A.D., Horvat A.I., Demajo M.A. (Serbia) Late effects<br />

of ionizing irradiation on corticosteroid receptor expression in rat hippocampus:<br />

the role of hypothalamus-pituitary-adrenal axis


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

MICE CARRYING A DELETION OF THE P66 Shc GENE SHOW REDUCED<br />

BEHAVIOURAL AND NEUROENDOCRINE RESPONSES TO STRESSFUL<br />

STIMULI<br />

Berry A. (1, 2); Giorgio M. (3); Martin-Padura I. (3); Pelicci P. G. (3); de Kloet E.R. (2);<br />

Alleva E. (1); Minghetti L. (4), and Cirulli F. (1)<br />

(1) Section of Behavioral Neuroscience, Dept. BCN, ISS, Rome, Italy; (2) LACDR, Div. of Medical<br />

Pharmacology, Gorlaeus Labs., Leiden University, Leiden, The Netherlands, (3) Department of<br />

Experimental Oncology, IEO, Milan, Italy; (4) Section of Degenerative and Inflammatory<br />

Neurological Diseases, Dept. BCN, ISS, Rome, Italy.<br />

Targeted mutation of the p66 Shc gene in the 129Sv/Ev mouse strain has been shown to increase<br />

both resistance to oxidative stress (OS) and life span. A growing body of evidence suggests<br />

that there might be an interaction between stressful stimuli and OS. P66 mutant mice represent<br />

an ideal animal model to study such an interaction. P66-/- (KO) and p66+/+ (WT) mice (4-<br />

and 11-months-old subjects) were tested in the open field (OF) and in the elevated plus maze<br />

(EPM) to assess emotional reactivity. They also underwent a chronic restraint stress paradigm<br />

as well as an acute systemic LPS challenge. Data from the OF and the EPM give an overall<br />

indication that KO mice were characterized by a less anxious phenotype. Results from the<br />

chronic restraint did not reveal any difference in the HPA axis responsiveness. By contrast,<br />

results from the LPS challenge indicate a reduced HPA axis activation in adult KO subjects.<br />

These results indicate that behavioral responses to arousing stimuli, as well as neuroendocrine<br />

responses to an immunogenic stimulus, are reduced in KO mice in an age-dependent fashion.<br />

Supported by: Italian Ministry of Health (grant ex art. 56 to F. C. and L. M.) and by Marie<br />

Curie fellowship (The Genetic Basis of Disease) granted to Alessandra Berry.<br />

199


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

EFFECTS OF PRENATAL DEXAMETHASONE TREATMENT ON MOTOR<br />

DEXTERITY IN THE JUVENILE MARMOSET MONKEY<br />

Hauser J.*, Zuercher N.*, Feldon J.*, Diaz-Heijtz R.#, Dettling-Artho R.*, Knapman<br />

A.*, Pryce C.R.$<br />

* Behavioural Neurobiology Laboratory, Swiss Federal Institute of Technology – Zurich,<br />

Schoerenstrasse 16, CH- 8603 Schwerzenbach, Switzerland, Fax: +41 (0)44 655 72 02, e-mail:<br />

jonas.hauser@behav.biol.ethz.ch<br />

# Behavioural Neuroscience Laboratory, Karolinska Institute, Stockholm, Sweden<br />

$ Novartis Institutes for BioMedical Research Basel, Novartis Pharma AG, Basel, Switzerland<br />

There is considerable evidence that prenatal stress has deleterious effects on development<br />

of the central nervous system (CNS) and those functions that it controls, including endocrine<br />

homeostasis and behaviour [7]. Prenatal stress results in increased maternal hypothalamicpituitary-adrenal<br />

(HPA) axis activity and thereby in higher plasma corticosteroids levels in<br />

maternal blood as well as, to some extent, fetoplacental unit. Corticosteroid binds specifically<br />

to mineralocorticoid receptor and glucocorticoid receptor (GR), this last being recruited<br />

primarily under stressful condition. GR are expressed in many organs, including brain, where<br />

they can, as transcription factors, regulate genes expression. GR undergo fetal programming:<br />

the setting up of adult expression level of a receptor by the presence of its ligand during early<br />

development, a mechanism hypothesized to mediate long term effects of prenatal stress. In<br />

rats, prenatal stress has been associated with delayed early life motor development [6] and<br />

increased HPA axis activity [4], this was also observed with the specific GR agonist<br />

dexamethasone (DEX) [3,11]. Similarly in rhesus monkey, prenatal stress or ACTH exposure<br />

resulted in impairement in an adaptation of the human Brazelton Newborn Behavioural<br />

Assessment Scale [8,9]. These associations between both prenatal stress and prenatal DEX<br />

exposure and impaired early motor development clearly point to the importance of increased<br />

understanding of any causal relationships between prenatal exposure to synthetic GC and<br />

development of motor dexterity beyond infancy. This long-term effect are furthermore<br />

important to study as synthetic specific GR agonists, such as DEX, are commonly prescribed<br />

in obstetric medicine for the prophylactic treatment of morbid symptoms associated with<br />

preterm birth, most importantly intra-ventricular haemorrhage and respiratory distress<br />

syndrome [5]. Such treatment is endorsed by the US National Institutes of Health (NIH) [1],<br />

with emphasis on the importance to study potentially harmful long-term effect especially in<br />

the case of repeated exposure [2].<br />

In this study, we administered 5mg/kg/d (per os) DEX to pregnant marmoset monkeys<br />

(Callithrix jacchus; gestation length 144 days) for 7 days either during early gestation (day 42-<br />

48, putative stage of maximal neurogenesis) or late gestation (day 90-96, equivalent time of<br />

clinical treatment in pregnant women). We assessed postnatal developmental effects relative to<br />

controls in offspring between 3 and 7 months of age. Physical and endocrine developments<br />

were monitored monthly by measuring body weight, knee-heel length and collecting a blood<br />

sample for radioimmuno assay analyses of basal ACTH and cortisol plasma titres. Home-cage<br />

behaviours were scored one hour weekly during months 3, 5 and 7, using an ethogram based<br />

on one already published for the marmoset [10], with focus on the following relationships<br />

(behaviour elements in parentheses): infant–parent or infant-infant (social grooming, social<br />

play) and infant alone (mobility, scratch with hand or foot, exploration, self-grooming, eat<br />

200


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

solid food, solitary play, tail hair piloerection). Skilled motor dexterity – reach-to-grasp<br />

movement – was assessed using a conditioned test that was adapted from the elegantlydesigned<br />

and well-validated skilled reaching task for rodents [12], consisting in training and<br />

testing subjects to reach through a small opening for a reward, and measuring the frequency of<br />

specific reaching and grasping behaviours.<br />

In the EDEX subjects there was a female-specific increase in body weight and in body<br />

weight:knee-heel ratio. There were no differences between any of the treatment groups or<br />

between sex in terms of basal plasma ACTH and cortisol titres. When assessing motor<br />

dexterity in the skilled reaching task, the LDEX treatment led to a deficit, which was not<br />

improved by training, whereas the EDEX treatment resulted in a transient impairment, which<br />

was overcome by the end of the testing. This study extends the aforementioned reports of<br />

neonatal delay in motor reflexes development following prenatal GR activation, and provides<br />

novel evidence that prenatal DEX exposure impairs skilled motor function in juvenile<br />

primates, which is important to our understanding of the role of glucocorticoids in the<br />

ontogeny of motor behaviour and of the negative side effects of clinical prenatal<br />

glucocorticoid treatment.<br />

References list<br />

[1] Consensus Development Panel on the Effect of Corticosteroids for Fetal Maturation on Perinatal<br />

Outcomes: Effect of corticosteroids for fetal maturation on perinatal outcomes., JAMA, 273 (1995) 413-<br />

8.<br />

[2] Consensus Development Conference Statement: Antenatal corticosteroids revisited: repeat courses,<br />

Obstet Gynecol, 98 (2001) 144-50.<br />

[3] Burlet, G., Fernette, B., Blanchard, S., Angel, E., Tankosic, P., Maccari, S. and Burlet, A., Antenatal<br />

glucocorticoids blunt the functioning of the hypothalamic-pituitary-adrenal axis of neonates and disturb<br />

some behaviors in juveniles, Neuroscience, 133 (2005) 221-30.<br />

[4] Fride, E., Dan, Y., Feldon, J., Halevy, G. and Weinstock, M., Effects of prenatal stress on vulnerability<br />

to stress in prepubertal and adult rats, Physiol Behav, 37 (1986) 681-7.<br />

[5] Jobe, A.H. and Soll, R.F., Choice and dose of corticosteroid for antenatal treatments, Am J Obstet<br />

Gynecol, 190 (2004) 878-81.<br />

[6] Patin, V., Vincent, A., Lordi, B. and Caston, J., Does prenatal stress affect the motoric development of<br />

rat pups? Brain Res Dev Brain Res, 149 (2004) 85-92.<br />

[7] Ruiz, R.J. and Avant, K.C., Effects of maternal prenatal stress on infant outcomes: a synthesis of the<br />

literature, ANS Adv Nurs Sci, 28 (2005) 345-55.<br />

[8] Schneider, M.L. and Coe, C.L., Repeated social stress during pregnancy impairs neuromotor<br />

development of the primate infant, J Dev Behav Pediatr, 14 (1993) 81-7.<br />

[9] Schneider, M.L., Coe, C.L. and Lubach, G.R., Endocrine activation mimics the adverse effects of<br />

prenatal stress on the neuromotor development of the infant primate, Dev Psychobiol, 25 (1992) 427-39.<br />

[10] Stevenson, M.F. and Poole, T.B., An ethogram of the common marmoset (Calithrix jacchus jacchus):<br />

general behavioural repertoire, Anim Behav, 24 (1976) 428-51.<br />

[11] Welberg, L.A., Seckl, J.R. and Holmes, M.C., Prenatal glucocorticoid programming of brain<br />

corticosteroid receptors and corticotrophin-releasing hormone: possible implications for behaviour,<br />

Neuroscience, 104 (2001) 71-9.<br />

[12] Whishaw, I.Q., O'Connor, W.T. and Dunnett, S.B., The contributions of motor cortex, nigrostriatal<br />

dopamine and caudate-putamen to skilled forelimb use in the rat, Brain, 109 (Pt 5) (1986) 805-43.<br />

201


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

ROLE OF 5ΑREDUCTASE ENZYMES IN THE FETAL BRAIN AND PLACENTA IN<br />

NEUROSTEROID PRODUCTION FOLLOWING INTRAUTERINE GROWTH<br />

RESTRICTIONHirst JJ 1,2 , Palliser HK 1 , Yates DM 1 , Walker DW 3<br />

1 Mothers and Babies Research Centre & 2 School of Biomedical Sciences, University of Newcastle,<br />

Callaghan NSW 2308, Australia;<br />

3 Department of Physiology, Monash University, Melbourne Vic<br />

3800, Australia<br />

Email: Jon.Hirst@newcastle.edu.au<br />

Allopregnanolone concentrations throughout the brain are remarkable high during fetal life and fall<br />

dramatically after birth [1]. This decline may result for the loss of a supply of 5α-reduced precursors<br />

by the placenta. We have shown that acute episodes of fetal hypoxia-ischemia, induced by occlusion<br />

of the umbilical cord during late gestation, lead to a rapid and marked rise in both allopregnanolone<br />

concentrations and 5α-reductase expression in the in the fetal brain [2]. In recent studies we have<br />

shown that the suppression of this neurosteroid response by finasteride treatment, potentiates hypoxiaischemia<br />

induced brain injury [3]. This suggests that the stress induced rise in allopregnanolone<br />

concentrations play an important neuroprotective role in the fetal brain and may be responsible to the<br />

relative resistance of the fetal brain to hypoxic ischemic insults. Placental insufficiency reduces<br />

progesterone production and may limit the supply of precursors for these stress-induced responses.<br />

Alternatively, the intrauterine growth restriction (IUGR) caused by placental insufficiency has been<br />

shown to alter steroidogenic pathways in the fetus. This may increase the neuroprotective steroid<br />

concentrations and/or the expression of their synthetic enzymes. These observations further suggest<br />

that neurosteroidogenic enzyme expression in the brain and placenta may act together to protect the<br />

fetal brain.<br />

The aim of this study was to compare the expression of 5α-reductase enzymes in the brain and placenta<br />

of fetuses from normal pregnancies and IUGR fetuses.<br />

A model of IUGR was produced in guinea pigs by the surgical ablation of a proportion of the branches<br />

of the uterine artery supplying each placenta at 31-<strong>35</strong> days of gestation. Fetal brains and placentas were<br />

collected at term (65 days) and prepared for immunoblotting. 5α-Reductase type 1 and 5α-reductase<br />

type 2 expression in the cortex, hippocampal region and placentas of the IUGR fetuses (n=4) were<br />

compared to sham operated controls (n=4). Denistometry of the specific bands was normalized against<br />

actin.<br />

Birth weights of IUGR fetuses were significantly lower than those of the sham operated controls (44.9<br />

6.9g and 80.2 6.0g respectively; P=0.001). Fetal brain to body weight ratio was higher in the IUGR<br />

fetuses (170%; P=0.001) indicating brain sparing occurred in response to the chronic restriction of the<br />

placental blood supply. There was no difference in the expression of 5α-reductase 1 in the cortex,<br />

hippocampal region or placenta between sham and IUGR fetuses. However, the placenta was found to<br />

have the highest level of expression of 5α-reductase 1, followed by the cortex and then the<br />

hippocampal region (P


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

following preterm birth may raise the vulnerability of the preterm neonate to brain injury if<br />

neurosteroidogenic pathways in the brain are not adequately mature to take over the role of the<br />

placenta.<br />

Reference list<br />

1. Nguyen, P. N., Billiards, S.B., Walker, D. W., Hirst, J.J. 2003. Changes in 5α-pregnane steroids and<br />

neurosteroidogenic enzyme expression in the perinatal brain. Pediatr. Res. 53, 956-964<br />

2. Nguyen, P. N., Yan, E. B., Castillo-melendez, M., Walker, D. W., Hirst, J. J. 2004. Increased<br />

allopregnanolone concentrations in the fetal brain following umbilical cord occlusion. J. Physiol. 560, , 593-<br />

6021<br />

3. Yawno, T., Yan, E.B., Walker D.W., Hirst, J.J. 2006. Inhibition of neurosteroid synthesis increases asphyxiainduced<br />

brain injury in the late gestation fetal sheep. Soc Gynecol Invest 53, Abstr31.<br />

203


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

NEONATAL EXPOSURE TO LOW DOSES OF CORTICOSTERONE INCREASES<br />

NOVELTY SEEKING AND RESISTANCE TO BACTERIA INFECTION IN ADULT<br />

MALE MICE<br />

Macrì S.* ,1 , Cirulli F. 1 , Pasquali P. 2 , Bonsignore L.T. 1 , Laviola G 1 .<br />

1 Section of Behavioural Neuroscience, Department of Cell Biology and Neuroscience,<br />

2 Department of Food Safety and Veterinary Public Health<br />

Istituto Superiore di Sanità, Viale Regina Elena 299, 00161 Roma, Italy. Fax +39-06-4957821<br />

email: simone.macri@iss.it; laviola@iss.it<br />

Neonatal maternal separation studies in laboratory rodents suggest that variable levels of<br />

environmental demands may modify mother-offspring interaction and, in turn, regulate the<br />

development of individual stress and fear responses in the adult progeny [1]. However, there is<br />

no agreement on the view that maternal care is the unique mediator of this form of adaptive<br />

plasticity [1,2,3]. In contrast with this hypothesis, we recently demonstrated that long periods<br />

of maternal separation increase levels of maternal care compared to no separation, and that<br />

brief separations increase active maternal care and reduce HPA activation in adult offspring<br />

[2]. It has therefore been proposed that, among other mediators (e.g. maternal care, food<br />

intake) neonatal manipulations may modify circulating levels of corticosterone in dams and/or<br />

pups, and, in turn, offspring development [4]. Additionally, we proposed that the effects of<br />

neonatal corticosteroids may follow a U-shaped curve whereby, compared to intermediate<br />

concentrations, both low and high circulating levels of maternal corticosterone would upregulate<br />

offspring stress and fear responses. In the present study we investigated this<br />

hypothesis through supplementing mouse dams with corticosterone in the drinking water<br />

during the first week of lactation [5]. Here we used this experimental procedure to mimic the<br />

physiological HPA activation consequent to variable levels of environmental demands in<br />

mouse dams.<br />

Methods: lactating CD-1 mouse dams were given ad libitum access to water (vehicle) or water<br />

supplemented with corticosterone (low, L-CORT, 33µg/ml or high, H-CORT, 100µg/ml)<br />

during the first lactation week. In order to control for the effects of corticosterone treatment on<br />

dam-pup interaction, maternal behaviour was scored according to a detailed ethogram<br />

originally developed for rats [6,7]. Naïve male adult offspring were tested for levels of novelty<br />

seeking (PND > 65), and anxiety-related behaviours in the elevated plus maze. To address the<br />

development of the immune response, a separate group of adult mice were infected with<br />

Brucella, and residual colony forming units in the spleen were measured two weeks following<br />

infection. Finally, corticosterone plasma levels were addressed in an additional group of male<br />

mice under basal conditions, and 0, <strong>35</strong> and 95 min after the termination of 25-min restraint<br />

stress.<br />

Results: Compared to control and L-CORT, H-CORT dams showed a significant reduction in<br />

levels of active maternal care (p


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

(p


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

CORTICOSTEROIDS MODULATION OF LONG TERM POTENTIATION ALONG<br />

THE SEPTO_TEMPORAL AXIS OF THE HIPPOCAMPUS<br />

Maggio N. and Segal M.<br />

Department of Neurobiology, The Weizmann Institute of Science, Israel<br />

Behavioral stress is correlated with an increased release of corticosteroids. Within the<br />

hippocampus, a structure that is associated with memory processes, corticosteroids have been<br />

shown to modulate synaptic plasticity. However, in recent years, evidence has accumulated to<br />

indicate that spatial memory is associated primarily with the dorsal (septal) sector of the<br />

hippocampus, while the role of its ventral/temporal sector is not yet clearly defined. In<br />

addition, slices taken from the ventral hippocampus exhibit a much lower ability to produce<br />

LTP compared to those taken from the septal sector.<br />

We explored the ability to produce LTP in dorsal and ventral hippocampal slices taken from<br />

the acutely stressed rats. Following stress, LTP was impaired in dorsal hippocampus, as seen<br />

before. Surprisingly, LTP was enhanced in ventral hippocampal slices. Bath applied 1µM<br />

Corticosterone in control slices mimicked this effect. Furthermore, the glucocorticoid agonist<br />

(dexamethazone) also induced an LTP impairment in dorsal hippocampal slices while it did<br />

not have any effect in the ventral hippocampus. On the other hand, aldosterone, a selective<br />

mineralocorticosterone receptor (MR) agonist, selectively enhanced LTP in the ventral<br />

hippocampus. This effect was blocked by a calcium channel antagonist nifedipine, but not by<br />

APV, indicating an involvement of voltage gated calcium channels in the facilitating effect of<br />

steroids on LTP in the ventral hippocampus. We conclude that a differential distribution of<br />

MR and GR receptors along the septo-temporal axis of the hippocampus could account for this<br />

effect.<br />

206


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

NEUROPLASTICITY IN THE PRENATAL STRESS RAT MODEL OF<br />

DEPRESSION: EFFECTS OF AGOMELATINE TREATMENT<br />

1 Morley-Fletcher S, 1 Mairesse J, 2 Daszuta A, 2 Soumier A, 2 Banasr M, 3 Zuena AR,<br />

4 Mocaer E, 3 Matteucci P, 3 Casolini P, 3 Catalani A, Maccari 1,3 S.<br />

1 Lab. of Perinatal Stress, University of Lille 1 USTL, Villeneuve d'Ascq, FR sara.morleyfletcher@univ-lille1.fr,<br />

fax +33.320.434602 ; 2 IC2N, CNRS, Marseille, FR ; 3 Univ of Rome<br />

La Sapienza, IT; 4 I.R.I.S, Courbevoie, FR<br />

Hippocampal neurogenesis can be influenced by several environmental factors and stressful<br />

experiences diminish the formation of the hippocampal granule cells in a number of<br />

mammalian species [4]. In rats, chronic antidepressants treatment can oppose the action of<br />

stress on the morphology and proliferation of hippocampal neurons, inducing changes in<br />

neuroplasticity and increasing adult neurogenesis in animals [9]. Thus, behavioural effects of<br />

antidepressant may be mediated by the stimulation of neurogenesis in the hippocampus [12].<br />

Among the stress-induced animal models of depression that have been developed, the prenatal<br />

restraint stress (PS) in the rat represents a promising model of early stress with high face and<br />

predictive validity [7,10,11]. PS rats present a life span reduction of hippocampal<br />

neurogenesis [6], an impairment of hypothalamus-pituitary adrenal axis feedback inhibition<br />

[8], a generalized disorganization of circadian rhythms [3] and increased anxiety [13]. We<br />

evaluated the effect of a chronic treatment (6 weeks, 40 mg/kg i.p.) with the new<br />

antidepressant agomelatine, a melatonin agonist with 5-HT 2C antagonist properties [1], on<br />

hippocampal neurogenesis and on PSA-NCAM and BDNF expression, markers of<br />

neuroplasticity, in PS male adult rats. To evidence neurogenesis and cell survival, the<br />

thymidine-analogue bromodeoxyuridine (BrdU, 75 mg/kg i.p. twice daily for 4 days) was<br />

injected after 3 weeks of the agomelatine treatment which was then continued for additional 3<br />

weeks. Prenatal stress reduced hippocampal neurogenesis whereas it increased PSA-NCAM<br />

and BDNF. The effects of PS were reversed by the chronic agomelatine treatment.<br />

Interestingly, agomelatine’s effect on neurons’ survival was selectively observed in the ventral<br />

part of the dentate gyrus, a brain region specifically involved in anxiety [5]. Also, based on the<br />

potential involvement of glutamatergic system in neurogenesis and anxiety [2], we studied the<br />

long-term effects of PS and antidepressant on the expression of hippocampal mGluR5<br />

subtype. Prenatal stress reduced mGluR5 expression, whereas chronic agomelatine increased<br />

it. Finally, to investigate the functional behavioural impact of neurogenesis, we tested animals<br />

in the elevated-plus maze test to assess their anxiety-like response. PS rats treated with<br />

agomelatine spent more time on the open arms of the elevated-plus maze, suggesting a<br />

possible causal link between increased hippocampal neurogenesis and glutamatergic<br />

transmission and, the attenuated anxiety-like behaviour. The results obtained with agomelatine<br />

provide further evidence of neuroplasticity as one of the targets of antidepressants and, further<br />

reinforce the high predictive validity of the PS rat as animal model of depression.<br />

References list<br />

1. Banasr M, et al. (2006). Agomelatine, a new antidepressant, induces regional changes in hippocampal<br />

neurogenesis. Biol Psychiatry 2006 1:59(11):1087-96.<br />

2. Di Giorgi Gerevini V. et al. (2004). The mGlu5 metabotropic glutamate receptor is expressed in zones of<br />

active neurogenesis of the embryonic and postnatal brain. Brain Res.Dev.Brain Res. 150: 17-22<br />

207


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

3. Dugovic C. et al. (1999). High corticosterone levels in prenatally stressed rats predict persistent paradoxical<br />

sleep alterations. J.Neurosci. 19: 8656-8664<br />

4. Kempermann G, et al. (1997). More hippocampal neurons in adult mice living in an enriched environment.<br />

Nature 386:493-495.<br />

5. Kjelstrup KG, et al. (2002). Reduced fear expression after lesions of the ventral hippocampus. Proc Natl<br />

Acad Sci U S A 99:10825-10830.<br />

6. Lemaire V, et al. (2000). Prenatal stress produces learning deficits associated with an inhibition of<br />

neurogenesis in the hippocampus. Proc Natl Acad Sci U S A 97:11032-11037.<br />

7. Maccari S, et al. (2003). Prenatal stress and long-term consequences: implications of glucocorticoid<br />

hormones. Neurosci Biobehav Rev 27:119-127.<br />

8. Maccari S, et al. (1995). Adoption reverses the long-term impairment in glucocorticoid feedback induced by<br />

prenatal stress. J Neurosci 15:110-116.<br />

9. Malberg JE, et al. (2000). Chronic antidepressant treatment increases neurogenesis in adult rat<br />

hippocampus. J Neurosci 20:9104-9110.<br />

10. Morley-Fletcher S, et al. (2003). Prenatal stress in rats predicts immobility behavior in the forced swim test.<br />

Effects of a chronic treatment with tianeptine. Brain Res 989:246-251.<br />

11. Morley-Fletcher S, et al. (2004). Chronic treatment with imipramine reverses immobility behaviour,<br />

hippocampal corticosteroid receptors and cortical 5-HT(1A) receptor mRNA in prenatally stressed rats.<br />

Neuropharmacology 47:841-847.<br />

12. Santarelli L, et al. (2003). Requirement of hippocampal neurogenesis for the behavioral effects of<br />

antidepressants. Science 301:805-809.<br />

13. Vallee M, Mayo W, Dellu F, Le MM, Simon H, Maccari S (1997). Prenatal stress induces high anxiety and<br />

postnatal handling induces low anxiety in adult offspring: correlation with stress-induced corticosterone<br />

secretion. J Neurosci 17:2626-2636.<br />

208


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

NEUROBEHAVIOURAL DISORDERS IN THE INFANT REELER MOUSE MODEL:<br />

INTERACTION OF GENETIC VULNERABILITY AND CONSEQUENCES OF<br />

MATERNAL SEPARATION<br />

Ognibene E. 1 , Adriani W. 1 , Macrì S. 1 , Laviola G 1, *.<br />

1 Section of Behavioural Neuroscience, Department of Cell Biology and Neuroscience,<br />

Istituto Superiore di Sanità, Viale Regina Elena 299, 00161 Rome, Italy. Fax +39-06-4957821 email:<br />

laviola@iss.it<br />

Studies on heterozygous (HZ) reeler mice suggest a relationship between reelin (a protein of extra<br />

cellular matrix) haploinsufficiency and the presence of altered neural networks and behaviour.<br />

Neonatal adverse and/or stimulating experiences might interfere with the emergence of this geneticdependent<br />

phenotype. Repeated episodes of maternal separation early in ontogeny result in enduring<br />

neuroendocrine, neurochemical and behavioural alterations in the offspring. Therefore, in order to<br />

investigate whether developmental indexes of neurobehavioural disorders can be studied in the infant<br />

reeler mouse model, and whether ontogenetic adverse experiences may question or improve its<br />

suitability, homozygous reeler (RL), heterozygous (HZ) and wild-type (WT) mouse pups underwent<br />

maternal separation (SEP, 5 h/day) or handling (H, 3 min/day) on PND 2–6. As expected, a sex<br />

difference appeared, for measure of emotional and communicative behaviour in infant mice. In<br />

particular, male H pups emitted a higher number of ultrasonic vocalizations (USV) compared to H<br />

females (p < 0,05). Interestingly, such sex difference was not observed in the SEP group. On PND 7,<br />

compared to other genotypes, RL mouse pups from the H control group, showed reduced levels USV<br />

production and of locomotion. Surprisingly, this deficit in RL mice was fully reverted by maternal<br />

separation. Maternal separation per se reduced social motivation in the homing test at PND 9 in WT<br />

mice, with no effects on HZ and RL ones. Additionally, female pups emitted much lower levels of<br />

ultrasound production than males within the H control group. The deficit in both emotional and<br />

communicative capabilities, however, apparently disappeared when reeler pups were faced, early in<br />

development, with the SEP condition. In fact, repeated episodes of maternal separation exerted a strong<br />

contrasting effect on homozygous RL mice, dramatically increasing, and perhaps activating their<br />

masked USV capacity production. This behavioural activation rendered these pups similarly able to<br />

respond to a stress condition, such as the repeated maternal separation, as the other two genotypes.<br />

Although hypothetical, it is tenable that the “beneficial” effects of repeated maternal separation,<br />

observed in RL pups, reflect a compensatory process of neural plasticity involving the activation of<br />

hormonal steroid pathways. The repeated 5-h long episodes of maternal separation are frequently<br />

associated with a substantial corticosterone release in newborn pups [1]. In turn, such a corticosteroid<br />

release might exert a stimulatory action in RL pups from SEP dam condition, early in development,<br />

compared to those from H-control group. In line with this interpretation, it has been recently reported<br />

that i.c.v. steroid hormones administration on PND 4 compensates reelin haploinsufficiency in terms of<br />

Purkinje cell loss during development [2].<br />

The present results provide evidence that unusual stress and related hormonal stimulation early in<br />

development may (i) independently shape individual behavioural phenotype and (ii) interact with a<br />

genetic make-up to substantially modify its “natural” developmental trajectories.<br />

Reference list<br />

[1] M.V. Schmidt, S. Levine, S. Alam, D. Harbich, V. Sterlemann, K. Ganea, E.R. de Kloet, F. Holsboer, M.B. Muller.<br />

Metabolic signals modulate hypothalamic–pituitary–adrenal axis activation during maternal separation of the neonatal<br />

mouse. J. Neuroendocrinol. 18 (2006) 865–74.<br />

[2] G. Assenza, F. Biamonte, R. Cesa, P. Strata, F. Keller. Interaction between reelin and estrogens on Purkinje cells during<br />

development: a model of cerebellar pathology in autism and related disorders. In: Abstract national congress of the Italian<br />

Society for Neuroscience and joint Italian-Swedish neuroscience meeting Regina Isabella Congress Center Lacco Ameno.<br />

2006.<br />

209


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

ASSESSMENT OF MODULATORY EFFECTS OF CORTICOSTERONE ON ANXIETY<br />

RELATED BEHAVIOR IN MICE<br />

Vafaei AA, Taherian AA, Jarrahi M<br />

Physiology Research Center, Semnan University of Medical Sciences, Semnan, Iran<br />

E-mail: aavaf43@yahoo.com<br />

Previous studies indicated that glucocorticoid receptors probably involve on anxiety reactions. This<br />

study was designed to evaluate modulatory effects of corticosterone on anxiety in mice. In this study,<br />

seventy male albino mice (25 – 30 gr) were used. Also we used of Elevated plus Maze (EPM) model<br />

for assessment of anxiety. Corticosterone (CORT) as a glucocorticoid receptor agonist (0.1, 0.5, 1, 3<br />

and 10 mg/kg) or vehicle were injected IP, 30 min before of test. At the first time for increasing<br />

activity animals have put inside the black wall box for 5 min. Then animal transfer to the EPM and<br />

evaluation their anxiety reaction that including of number entrances and time spent in open arm.<br />

Results indicated that injection of CORT in doses of 1 and 3 mg/kg reduced of reaction anxiety and<br />

with compare to sham and control groups in the test group animals have more number of entrances and<br />

spent more time in open arm (P


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

Time spent in open arm (Sec)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Sham Cont CORT<br />

0.1<br />

CORT<br />

0.5<br />

*<br />

CORT<br />

1<br />

*<br />

CORT<br />

3<br />

CORT<br />

10<br />

Number of entrance to open<br />

arm<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Sham Cont CORT<br />

0.1<br />

CORT<br />

0.5<br />

*<br />

CORT<br />

1<br />

*<br />

CORT<br />

3<br />

CORT<br />

10<br />

Fig 1 A: The effect of CORT (0.1, 0.5, 1, 3<br />

and 10 mg/kg, SC) on anxiety related<br />

behavior in mice. (Time spent in open arm)<br />

*P


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

LATE EFFECTS OF IONIZING IRRADIATION ON CORTICOSTEROID<br />

RECEPTOR EXPRESSION IN RAT HIPPOCAMPUS: THE ROLE OF<br />

HYPOTHALAMUS-PITUITARY-ADRENAL AXIS<br />

Velickovic N. A., Djordjevic A. D., Horvat A. I., Demajo M. A.<br />

VINCA Institute for Nuclear Sciences, Laboratory for Molecular Biology and<br />

Endocrinology, PO Box 522, Belgrade 11000, Serbia, Fax: +381112455561<br />

e-mail: natasaxx@vin.bg.ac.yu<br />

Radiotherapy continues to be the primary treatment modality for malignant brain<br />

tumors. This treatment, however, inevitably involves the inclusion of normal CNS tissue<br />

in the radiation field. Experimental studies have shown that radiation exposure induces<br />

both short and long-term deleterious effects on the functional capacity of the<br />

hypothalamus-pituitary-adrenal (HPA) axis [2]. The primary element in HPA axis<br />

feedback regulatory mechanisms considered to be the hippocampus. The hippocampus is<br />

highly sensitive to corticosteroids (CS), exemplified by its enriched complement of CS<br />

receptors. Two types of CS receptors are found in the brain: mineralocorticoid (MR) and<br />

glucocorticoid (GR) receptors. Differential activation of this dual receptor system may<br />

account for the opposing actions of corticosteroids on neuronal proliferation, survival,<br />

and death in hippocampus [1].<br />

The aim of the present study was to estimate the effect of ionizing irradiation on<br />

corticosteroid receptor expression in the rat hippocampus and to compare the expression<br />

of GR protein with changes in HPA axis functioning during the late radiation response<br />

phase. Infantile (18 days old) male Wistar rats received a single dose of 10 Gy (the dose<br />

was selected to be clinically relevant [3]) in the head region from a Co 60 -source. Late<br />

effects of gamma-irradiation on HPA axis have been studied at the age of 42 days in<br />

basal and after restrain stress, followed by assessment of GR and MR mRNA levels and<br />

protein expression by RT-PCR and Western blot, respectively.<br />

In our study, radiation exposure lead to a decrease of stress-induced but not basal<br />

plasma corticosterone levels (Fig. 1, C-3 vs. IR-3, p< 0.05). Treatment with<br />

dexamethasone (DEX) revealed the nonsuppression of stress-induced HPA axis function<br />

in irradiated rats, thus suggesting the dampened sensitivity of negative feedback of<br />

glucocorticoids as a long-term consequence of irradiation (Fig. 1, IR-4 vs. C-4, p< 0.01).<br />

Corticosterone (ng/ml)<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

Basal<br />

Stress<br />

*<br />

**<br />

C-1 IR-1 C-2 IR-2 C-3 IR-3 C-4 IR-4<br />

DEX: - - + + - - + +<br />

Groups<br />

Figure 1. Plasma corticosterone<br />

levels in control (C) and irradiated<br />

(IR) animals under different<br />

treatment. Groups 1 and 3<br />

received vehicle injection, groups<br />

2 and 4 received DEX injection<br />

(30 µg/kg). Groups 3 and 4 were<br />

subjected to acute stress, groups 1<br />

and 2 were not subjected to this<br />

stress.<br />

Values are means ± SEM (n=7-9);<br />

* p< 0.05, ** p< 0.01 vs. controls.<br />

212


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

Since GR (coordinately with MR) mediates the corticosteroid control of the<br />

hippocampus on the HPA axis, we assessed the level of GR and MR mRNA and protein.<br />

Western blot analysis showed a decreased translocation of the GR from the cytoplasm to<br />

the nucleus in the presence of DEX, without affecting the level of GR protein in whole<br />

cell extracts. Nuclear trafficking of the GR is regulated by the macromolecular hsp90<br />

heterocomplex. Additional Western analyses were performed to determine whether<br />

deficits in GR translocation were associated with decreased chaperone expression or<br />

trafficking. Expression of hsp90 and hsp70 was not affected by irradiation in both nuclear<br />

and cytoplasmic extracts, thus indicating to some other mechanism lying in this<br />

phenomenon. In contrast, no change in nuclear (or cytosolic) MR protein was observed in<br />

rat hippocampus after irradiation, which may denote that the corticosteroid receptor<br />

translocation deficit was limited to the GR. The obtained results are confirmed by RT-<br />

PCR, since GR down-regulation is accompanied by reduced GR mRNA, whereas the MR<br />

mRNA level did not show statistically significant differences.<br />

In conclusion, our data suggest that irradiation of the cranial region may attenuate<br />

the translocation and activation of the GR by circulating hormones, thereby diminishing<br />

negative feedback on the HPA axis solely by GR, and not by MR, in the hippocampus.<br />

This study was supported by Project grant No. 143044, the Ministry of Science and<br />

Environmental Protection, Republic of Serbia.<br />

Reference list<br />

[1] Almeida, O.F.X., Conde, G. L., Crochemore, C., Demeneix, B. A., Ficher, D., Hassan, A. H. S., Meyer,<br />

M., Holsboer, F., Michaelidis, T. M., 2000. Subtle shifts in the ratio between pro- and antiapoptotic<br />

molecules after activation of corticosteroid receptors decide neuronal fate. FASEB J 14, 779-790.<br />

[2] der Meeren, A. V., Monti, P., Lebaron-Jacobs, L., Marquette, C., Gourmelon, P., 2001. Characterization<br />

of the acute inflammatory response after radiation exposure in mice and its regulation by Il4. Radiat.<br />

Res. 155, 858-865.<br />

[3] Schunior, A., Mullenix, P. J., Zengel, A. E., Landy, H., Howes, A., Tarbell, N. J., 1994. Radiation<br />

effects on growth are altered in rats by prednisone and methotrexate. Pediatr. Res. <strong>35</strong>, 416-423.<br />

213


Posters’ Exhibition:<br />

Others<br />

• Allieri F., Hardin-Pouzet H. (Italy) Monoaminergic regulation of AVP system in<br />

limbic nuclei during estrous cycle, possible implication in axiety and depression<br />

• Amini H., Salimpour S., Mirzaei M., Sabetkasaei M., Ahmadiani A. (Iran)<br />

Involvement of the enzyme 5alpha-reductase in morphine-induced dopamine<br />

release in the nucleus accumbens: a microdialysis study in rats<br />

• Andrade T.G.C.S., Sergio T.O., Broiz A.C.G., Avanzi V. (Brazil) The effect of<br />

oestradiol benzoate in the median raphe nucleus on the exploratory behaviour in<br />

forced swimming test<br />

• Andrade T.G.C.S. , Almada R.F., Nakamuta J.S., Avanzi V. . (Brazil) Effect of<br />

oestrogenic action in the median raphe nucleus on the exploratory behaviour of<br />

previously immobilized female rats, in the elevated plus maze<br />

• Aste N., Shimada K., Watanabe Y. and Saito, N. (Japan) Neurosteroidogensis in<br />

the quail brain<br />

• Atwood C.S., Wilson A.C., Bowen R.L., Vadakkadath Meethal S. and Liu T.<br />

(USA) The potential role of a neuronal autocrine/paracrine mechanism in the<br />

regulation of neurosteroid production: luteinizing hormone receptor mediates<br />

neurosteroid production via upregulation of steroidogenic acute regulatory protein<br />

expression<br />

• Bramanti V., Bronzi D., Raciti G., Avitabile M., Avola R. (Italy) Neurosteroidsgrowth<br />

factors interaction induces up and down regulation of GFAP and vimentin<br />

expression in astroglial cells maintained under serum-free stressed culture<br />

conditions<br />

• Balog J., Szegő É.M., Erdei F., Szabó G., Juhász G. and Ábrahám I.M. (Hungary)<br />

Sex differences in rapid estrogen action on GABAergic neurons in vivo<br />

• Campbell B.C. (USA) DHEAS and the hominoid brain<br />

• Carrillo B, Pinos H., Guillamón A., Panzica G.C., Pérez-Izquierdo M.A., Collado<br />

P. (Spain) Nitric oxide effects on sexual and maternal behavior in the female rat<br />

• Ceccarelli I., De Padova A.M., Fiorenzani P., Massafra C. and Aloisi A.M. (Italy)<br />

Single opioid administration modifies gonadal steroids in both the CNS and plasma<br />

of male rats<br />

• Chalbot S., Lecanu L., Greeson J. and Papadopoulos V. (USA) Oxidationdependent<br />

plasma DHEA formation as a diagnostic tool for Alzheimer’s disease<br />

pathology: results from a trial<br />

• Csakvari E., Hoyk S., Szajli A., Kurunczi A., Gyenes A., Berger A., and Parducz<br />

A. (Hungary) Lesion-induced glial reaction in the rat olfactory bulb: effect of<br />

DHEA and DHEA derivatives


• Diz-Chaves Y., Pernía O., Carrero P., Garcia-Segura L.M. (Spain) Dose-response<br />

study of antidepressant effects of estrogenic compounds in ovariectomized mice in<br />

the forced swim test<br />

• Do Rego J.L., Tremblay Y., Luu-The V., Acharjee S., Repetto E., Galas L., Castel<br />

H., Vallarino M., Kwon H.B., Bélanger A., Seong J.Y., Pelletier G., Tonon M.C.,<br />

Vaudry H. (France) Neuroanatomical and biochemical evidence for the<br />

occurrence of cytochrome P450 C17 in the frog brain. Regulation by vasotocin and<br />

mesotocin<br />

• Fester L., Zhou L., Bütow A., Huber C., von Lossow R., Jarry H., M. Rune G.M.<br />

(Germany) Synaptogenesis: promoted by cholesterol or estradiol?<br />

• Hiroi R., Neumaier J.F. (USA) Estrogen selectively increases tryptophan<br />

hydroxylase-2 and decreases 5HT1 B mRNA expressions in distinct subregions of<br />

rat dorsal raphe nucleus: association between gene expression and anxiety<br />

behavior in the open field<br />

• Kanematsu T. and Hirata M. (Japan) PRIP, a phospholipase C-related inactive<br />

protein, regulates GABA A receptor endocytosis<br />

• Löfgren M., Johansson I-, Meyerson B. and Bäckström T. (Sweden) Progesterone<br />

withdrawal sensitivity in female rats relates to differences in baseline behavior of<br />

risk taking and exploration<br />

• Longo D., Baldelli E., Zini I., Zoli M., Avoli M., Biagini G. (Italy) P450scc is<br />

induced in neuronal and glial cells after status epilepticus: modulatory effects of<br />

neurosteroids on epileptogenesis<br />

• Martín-García E., Darbra S., Pallarés M. (Spain) Alterations of neonatal levels of<br />

allopregnanolone and the novelty-directed behavioural response to<br />

intrahippocampal administration of allopregnanolone in adulthood<br />

• Milani P., Ginanneschi F., Biasella A., Bonifazi M., Rossi A., Mazzocchio R.<br />

(Italy) Heightened seizure susceptibility following the administration of human<br />

chorionic gonadotropin<br />

• Mizokami A., Kanematsu T. and Hirata M. (Japan) Roles of PRIP in trafficking of<br />

gamma2 subunit containing GABA A receptor<br />

• Nasir RH, Chen C, Bellinger D, Korrick SA (USA) Prenatal estrogens and the<br />

development of memory and learning<br />

• Nobahar M, Vafaei AA (Iran) Assessment of interaction between sex hormones<br />

and incidence of epilepsy crisis in female<br />

• Ohya T., Kodama M., Hayashi S. (Japan) Vasotocin/isotocin neurons are<br />

decreased after spawning in the female medaka fish (Oryzias latipes) brain:<br />

localization of aromatase and estrogen receptor homologue


• Parkash J. and Kaur G. (India) GnRH-Astrocytes interactions involved in GnRH<br />

neurosecretion: role of PSA-NCAM through changing activity and expression<br />

levels of polsialyltrasferase.<br />

• Prange-Kiel J, Jarry H, Kohlmann P, Schön M, Lohse C, Rune GM (Germany) Is<br />

there a link between the hypothalamo-pituitary-gonad axis and the hippocampus?<br />

• Romanò N., Jasoni C.L. and Herbison A.E. (New Zealand) Rapid actions of<br />

estrogen on adult GnRH neurons<br />

• Scurati S., Maschi O., Crotti S., De Angelis L., Melcangi R.C. and Caruso D.<br />

(Italy) Assessment of neuroactive steroid levels in plasma and nervous sistem by<br />

liquid chromatography-mass spectrometry<br />

• Timby E., Bäckström T., Nyberg S., Wihlbäck A.-C.N., Bixo M. (Sweden)<br />

Administration of allopregnanolone decreases secretion of gonadotropins in<br />

healthy women of fertile age<br />

• Venard C., Boujedaini N., Belon P., Mensah-Nyagan A.G. and Patte-Mensah C.<br />

(France) Pharmacological modulators of the glycinergic system regulate<br />

allopregnanolone biosynthesis in the rat spinal cord<br />

• Vlad A.G (Romania) It is possible that the unspecific steroids for gonadotropin<br />

system modulate neuronal pulsatility activity from POA–SCH of the hypothalamus<br />

for regulating LH–RH relase and ovulation trigger?


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

MONOAMINERGIC REGULATION OF AVP SYSTEM IN LIMBIC NUCLEI<br />

DURING ESTROUS CYCLE, POSSIBLE IMPLICATION IN AXIETY AND<br />

DEPRESSION<br />

F.Allieri1,2, H. Hardin-Pouzet2<br />

1 Lab: Neuroendocrinologia, Dip.Anatomia, Farmacologia e Med. Legale Univ Torino,<br />

Italy<br />

2 Lab. Neurobiologie des Signaux Intracellulaires, CNRS UMR 7101, Université Pierre et<br />

Marie Curie Paris, France<br />

Estrogen appear to influence the intensity of several psychiartric and neurological<br />

disorders, including depression. In particular, depression and anxiety are common health<br />

problems affecting women, particularly during reproductive years. The cerebral structures<br />

implicated in such mood disorders are principally located in the limbic areas, namely in the<br />

bed nucleus of the stria terminalis (BST), the anterior part of medial amygdala (MeA) and<br />

the posterior part of the medial amygdala (MeP). These nuclei contain parvocellular<br />

arginine-vasopressinergic (AVP) neurons, which are known to be sensitive to gonadal<br />

steroids and receive a dense innervation from the serotoninergic and dopaminergic<br />

systems. As well know serotoninergic and dopaminergic systems are strongly implicated in<br />

the modulation of some behavioural disorders, like depression and anxiety. To understand<br />

the possible role of estrogen on the monoamines system and the consequent influence of<br />

both of theme on the AVP parvocellular neurons present in the BST, the MeA and the<br />

MeP. Particularly we focused on the possible interaction of monoamines during the estrous<br />

cycle. We treated male and female, taken in the different phases of the estrous cycle, mice<br />

C 3 H with para-clorophenyl alanine (pCpA) and alpha-metylparatyrosine (α-MPT); these<br />

two pharmaco can inhibit whole patway of serotonin and dopamine. We observe the a<br />

prominent decrease of AVP protein quantity in the MeP alpha-MPT treated male mice and<br />

an increase of AVP protein quantity in the MeA and MeP of pCpA treated male mice,<br />

comparing to control animals. Normally in the female mice AVP circulating levels<br />

detected in blood change during the different phases of the estrous cycle; we observe that<br />

treatment with α-MPT and pCpA can affect the quantity of AVP protein detected in the<br />

BST, in the MeA and in the MeP, in comparison with control female and male mice. We<br />

detect that MeA and MeP respond in a different manner comparing the differnt phases of<br />

estrous cycle. These data suggest that the estrogen can lead monoamines regulation of<br />

AVP neurons. So we suggest a double control of estrogen and monoamines on the<br />

AVPergic system present in nuclei, as the BST, the MeA and the MeP, strongly related<br />

with the control of behaviour. Moreover this doble action can partecipate to the beginning<br />

of behavioural disorders like anxiety and depression.<br />

217


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

INVOLVEMENT OF THE ENZYME 5ALPHA-REDUCTASE IN MORPHINE-<br />

INDUCED DOPAMINE RELEASE IN THE NUCLEUS ACCUMBENS: A<br />

MICRODIALYSIS STUDY IN RATS<br />

Amini H.*, Salimpour S., Mirzaei M., Sabetkasaei M., Ahmadiani A.<br />

Department of Pharmacology, Neuroscience Research Center, Shaheed Beheshti Medical<br />

University, P.O. Box 198<strong>35</strong>-<strong>35</strong>5, Tehran, Iran. E-mail: hamini@sbmu.ac.ir<br />

The enzyme 5alpha-reductase (5alpha-R) is one of the key enzymes in the biosynthesis of<br />

neurosteroids. We have already reported that morphine increases the 5alpha-R activity in<br />

the rat CNS following acute and chronic administration, but the question remains about the<br />

significance of this finding in morphine effects. It is well known that morphine and other<br />

addictive drugs increase dopamine release in the nucleus accumbens as well as dopamine<br />

metabolites, 3, 4-dihydroxyphenylacetic acid (DOPAC) and homovanillic acid (HVA). In<br />

the present study, the effects of concomitant administration of morphine and finasteride (an<br />

inhibitor of 5alpha-R) on the extracellular levels of DOPAC and HVA in the nucleus<br />

accumbens of rats were studied using in vivo microdialysis and high performance liquid<br />

chromatography with electrochemical detection. Acute morphine (7 mg/kg, i.p.)<br />

treatment increased the levels of DOPAC and HVA in the nucleus accumbens to<br />

approximately two-fold of basal levels. Pretreatment with finasteride (5 mg/kg, i.p.), 2 h<br />

before morphine injection (7 mg/kg. i.p.) significantly changed the effects of morphine on<br />

the levels of DOPAC and HVA in the nucleus accumbens. These results suggest that the<br />

enzyme 5alpha-R involves in the mesolimbic dopaminergic pathway.<br />

218


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

THE EFFECT OF OESTRADIOL BENZOATE IN THE MEDIAN RAPHE<br />

NUCLEUS ON THE EXPLORATORY BEHAVIOUR IN FORCED SWIMMING<br />

TEST<br />

Andrade, T.G.C.S. 1 , Sergio, T.O., Broiz, A.C.G., Avanzi,V 1 .<br />

Laboratory of Physiology – University UNESP- Assis – São Paulo - Brazil 1 - Fax: 55 18<br />

3302-5849 –<br />

E-mail: raica@assis.unesp.br<br />

The median raphe nucleus (MRN) is located in the brain stem and constitutes one of the main<br />

systems of ascending serotonergic innervations. Based on clinical and experimental evidence,<br />

Deakin & Graeff [1] have hypothesized that the 5-HT MRN-hippocampal pathway underlies<br />

adaptation to chronic stress and the failure of this mechanism would impair the putative 5-HT 1A<br />

“resilience” system, slowing adaptation to stress and predisposing animals to depressive symptoms<br />

and humans to depression. Recently, oestrogenic receptors have been identified in the MRN [2]. In<br />

addition to that, oestrogen deficiencies have been related to the manifestation of anxiety and<br />

depression. Women in periods of low estrogenic concentration tend to present high level of<br />

anxiety, which might lead to occurrence of depression. It is known that oestrogen reposition<br />

increases the serotonergic neurotransmission with a direct influence on the sensibility of the 5-HT<br />

receptors in different areas of the brain, especially in the hippocampus. The main objective of the<br />

study presented here was to investigate the action of Oestradiol Benzoate (OB), microinjected in<br />

the MRN, on the behavioural responses in Forced Swimming Test (FST), an animal depression<br />

model [3]. With this objective we analysed ovariectomized female Wistar rats, 200g medium<br />

weight at the beginning of the experimental sessions, previously exposed to FST 24 hours before<br />

the cannulation. Seven days after the stereotaxic surgery for the insertion of the guiding cannula of<br />

access to the MRN, the animals were microinjected with OB, in doses of 600ng and 1200ng, in<br />

volume of 0,2ul, and were placed in FST. The control group received the same volume of saline or<br />

oil (diluent of OB). The results showed that the microinjection of OB in the MRN, in higher dose,<br />

led to an increase in the mobility in the test [F(3,38)=7.78; p=0.000] without causing an increase in<br />

motor activity in the arena [F(3,38)=0.26; p=0.851], similar to the effect of anti-depressive drugs in<br />

this test. These results are in accordance with other studies carried out in our laboratory in males<br />

and females with lesions in NMR or direct microinjection of 8-OH-DPAT, an agonistic of 5-HT<br />

receptors, in this structure. As conclusion, the beneficial effect of Oestradiol Benzoate may be<br />

measured by specific receptors in the MRN. Neuron located in this structure would also be<br />

modulated by endogen oestrogen, confirming studies, which point out that there are significant<br />

betterments in emotional symptoms related to depression, in periods of higher level of this<br />

hormone or due to oestrogen reposition.<br />

Support: FAPESP- Brazil<br />

Reference list<br />

1. DEAKIN, J.F.W., GRAEFF, F.G. 5-HT and mechanisms of defense. Journal of<br />

Psychopharmacology, 5(4):305-315, 1991.<br />

2. LERANTH, C. SHANABROUGH, M., UPHOUSEL, L. Oestrogen receptor-α in the<br />

serotonergic and supramammillary area calretinin-containing neurons of the female rat.<br />

Experimental Brain Research, 128:417-420, 1999.<br />

3. PORSOLT, R.D. et al. Behavioral despair in rats: a new model sensitive to antidepressant<br />

treatments. European Journal Pharmacology, 47: 379-391, 1978.<br />

219


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

EFFECT OF OESTROGENIC ACTION IN THE MEDIAN RAPHE NUCLEUS ON<br />

THE EXPLORATORY BEHAVIOUR OF PREVIOUSLY IMMOBILIZED<br />

FEMALE RATS, IN THE ELEVATED PLUS MAZE<br />

Andrade T.G.C.S. 1 , Almada 1 R.F., Nakamuta 1 J.S., Avanzi V 1 .<br />

Laboratory of Physiology – University UNESP- Assis – São Paulo - Brazil 1 - Fax: 55 18<br />

3302-5849 – E-mail: raica@assis.unesp.br<br />

Serotonin has been implicated in the aetiology of anxiety disorders. Gender differences in<br />

serotonergic functions have been observed in animal models of anxiety. Several studies<br />

have shown changes in the serotonergic activity correlated to the phases of the female<br />

hormonal cycle, and that ovarian hormones act on the synthesis, liberation, reuptake and<br />

catabolism of 5-HT. Clinical studies have reported marked emotional alterations with<br />

periods of increased anxiety, during low-oestrogen phases of the hormonal cycle.<br />

Oestrogen replacement improves those symptoms and it also increases serotonergic<br />

neurotransmission. In the same direction, experimental studies in rodents using animal<br />

models of anxiety have shown anxiety reduction during oestrogenic phases of the<br />

hormonal cycle as well as after oestrogen administration. It seems that oestrogen<br />

replacement increase serotonergic neurotransmission, influencing the sensitivity of 5-HT 1A<br />

receptors in several brain areas, especially in the hippocampus and raphe nuclei. The<br />

existence of oestrogen receptors (ER) in cellular bodies of the median raphe neurones has<br />

been demonstrated [1]. The median raphe nucleus (MRN) is located in the brain stem and<br />

constitutes one of the main systems of ascending serotonergic innervations and has been<br />

related to behavioural inhibitions and stress resistance mechanisms. The inefficiency of<br />

this structure could lead to the occurrence of emotional disturbances, such as anxiety and<br />

depression. Injection of oestradiol into the MRN impairs the 5-HT innervation of the<br />

hippocampus [2], which is critical for the process of behavioural inhibition that is<br />

supposed to underlie anxiety. Recently, the microinjection of the oestradiol benzoate into<br />

MRN occasioned anxiolytic effect in the elevated plus-maze [3]. This effect was abolished<br />

by microinjection into MRN of the Way1006<strong>35</strong>, a 5-HT 1A receptors antagonist [3,4,5].<br />

Like this, the study presented here had as its main objective to evaluate the effect of an<br />

acute stressor, immobilization, on behavioural responses of ovariectomized female rats in<br />

the elevated plus maze (EPM), an animal anxiety model [6] and to verify if the direct<br />

microinjection of Oestradiol Benzoate (OB) in the MRN would alter the anxiogenic effect<br />

caused by the previous exposition the acute stressor. Events, like restraint and social<br />

separation, have been described like powerfull stressors, unchained of inhibitory response<br />

in animal tests of anxiety [7]. Females Wistar rats, 200g medium weight at the beginning<br />

of the experimental sessions, kept in groups of five per box, under controlled temperature<br />

and illumination (12/12 hours light-dark cycles, 50 lux and 21º C ±1º) receiving food and<br />

water ad libitum were submitted to a bilateral ovariectomy and after fourteen days were<br />

submitted to a stereotaxic surgery for the implantation of an access cannula to the NMR.<br />

Seven days after, the animals were microinjected in the MRN with saline, Sesame Oil or<br />

OB, in doses 600 or 1200 ng (0,2 µ/min) and evaluated immediately after the<br />

administration of the drug in the EPM, for the period between 14 and 17 hours.<br />

Previously, on the sixth day were immobilized for two hours in a metal box and kept in<br />

individual box for 24 hours till the behavioural evaluation (social separation). The results<br />

showed that the exposition to previous stress (immobilization) led to an anxiogenic effect<br />

in the elevated plus maze (EPM). The ovariectomy produced effect only when the animals<br />

were previously immobilized. We verified that the microinjection of EB in the MRN<br />

220


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

neutralized the stressing effect of the immobilization in this animal model of anxiety,<br />

leading to pattern of exploratory behavioural responses similar to the non immobilized<br />

females: that is, a clear anxiolytic effect of this substance microinjected in the MRN; % of<br />

entrances in open arms/total [F(3,79)=14,<strong>35</strong>; p=0.000]; % of time in open arms/total<br />

[F(3,79)=11.90; p=0.000]. As conclusion, it seems that the anxiogenic effect due to the<br />

decreasing in the oestrogen levels may be evidenced when the females are exposed to<br />

previous stressors. Women in puerperium, in climacterium, or in premenstrual period, may<br />

present emotional symptoms when exposed to aversive events during these situations, as<br />

there is a low concentration of oestrogen. The direct microinjection of Oestradiol<br />

Benzoate in the MRN has led to behavioural desinhibition of the females. This may<br />

explain the salutary effect of endogen oestrogen and of the oestrogenic reposition in the<br />

prevention and/or treatment of anxiety disorders and affective disturbances.<br />

Support: FAPESP- Brazil<br />

Reference list<br />

1. LERANTH, C. SHANABROUGH, M., UPHOUSEL, L. Oestrogen receptor-α in the<br />

serotonergic and supramammillary area calretinin-containing neurons of the female rat.<br />

Experimental Brain Research, 128:417-420, 1999.<br />

2. PRANGE KIEL, J., RUNE, G., M., LERANTH, C., Median raphe mediates estrogenic<br />

effects to the hippocampus in female rats. European Journal Neuroscience, 249:341-<br />

<strong>35</strong>1, 2004.<br />

3. ANDRADE, T.G.C.S.; NAKAMUTA, J.S., AVANZI, V., GRAEFF, F.G. Anxiolytic<br />

effect of oestradiol in the median raphe nucleus mediated by 5-HT 1A. Behavioural Brain<br />

Research,. 163: 18-25, 2005.<br />

4. ANDRADE, T.G.C.S., AVANZI, V. Effect of oestradiol benzoate microinjected into<br />

median raphe nucleus on anxiety in the conditioned fear test. Frontiers in<br />

Neuroendocrinology, 27: 134-138.<br />

5. ANDRADE, T.G.C.S., VIANA, R.A., NAKAMUTA, J.S., AVANZI, V. Different<br />

effects of the oestradiol benzoate microinjected into raphe nuclei on anxiety. Frontiers<br />

in Neuroendocrinology, 27: 134-138, 2006.<br />

6. PELLOW, S., CHOPIN, P., FILE, S. E., BRILEY, M., Validation of open-closed arm<br />

entries in an elevated plus-maze as a measure of anxiety in the rat. Journal<br />

Neuroscience Methods, 14: 149-167, 1985.<br />

7. KENETT, G.A., DOURISH, C.T., CURZON, G. Antidepressant-like action of 5-HT1A<br />

agonist and conventional antidepressants in an animal model of depression. European<br />

Journal Pharmacology, 134: 265-274, 1987.<br />

221


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

NEUROSTEROIDOGENSIS IN THE QUAIL BRAIN<br />

Aste N., Shimada K., Watanabe Y. and Saito N.<br />

Laboratory of Animal Physiology, Graduate School of Bioagricultural Sciences, Nagoya<br />

University,Chikusa, Nagoya 464-8601 Japan, Tel & Fax: +81-52-789-4067.<br />

aste@agr.nagoya-u.ac.jp<br />

The brain, besides being a target for gonadal and adrenal steroids, is also a site of steroids<br />

synthesis. Steroids of brain origin (neurosteroids), may work locally by modulating<br />

centrally controlled functions related to reproduction, learning and memory formation,<br />

which show sexual differences [1, 4]. Furthermore in situ formation of estradiol from<br />

cholesterol may play a role in brain sexual differentiation [3]. It is therefore of importance<br />

to understand whether the synthesis and the location of neurosteroidogenic enzymes show<br />

sexual dimorphism either in the adult or in the developing animals.<br />

Among birds, Japanese quail has been long employed as a model for studying the role of<br />

gonadal steroids on the activation of copulatory behavior and on brain sexual<br />

differentiation [2]. The activity and the presence of mRNA of cytochrome P450 side-chain<br />

cleavage enzyme (P450scc), 3β-hydroxysteroid dehydrogenase/Δ5-Δ4-isomerase (3β-<br />

HSD), cytochrome P450 17α-hydroxylase/c17, 20-lyase (P450c17), which are required to<br />

synthesize testosterone from cholesterol, have been demonstrated in quail brain [5].<br />

However very little attention has been given to sexual differences in their content and<br />

distribution.<br />

In this study we assessed for the first time the presence of a sexual dimorphism for<br />

P450scc, 3β-HSD and P450c17 mRNA level in discrete regions of adult Japanese quail.<br />

Moreover the sexually dimorphic expression of these enzymes and of aromatase (which<br />

conerts testosterone into estradiol) was investigated in the prosencephalon of embryos at<br />

several ages spanning the critical period for brain sexual differentiation (E7, E9, E11, E15)<br />

using real-time PCR as a method.<br />

Expression of P450scc, 3β-HSD and P450c17 mRNAs was detected in all the investigated<br />

regions (telencephalon, diencephalon, optic lobes, cerebellum, brainstem) and showed<br />

regional differences. P450scc mRNA level did not show sexual differences. It was higher<br />

in the diencephalon and very low in the cerebellum, the other regions displaying<br />

intermediate values. The level of 3β-HSD and P450c17 mRNAs showed sexual<br />

dimorphism in the optic lobes, where higher mRNA level was detected in males than in<br />

females.<br />

P450c17 mRNA level showed large region-related variations in males being relatively<br />

higher in the optic lobes and in the brainstem than in the cerebellum and in the<br />

telencephalon. In females the region-related variations were comparatively of modest<br />

amplitude.<br />

3β-HSD mRNA level showed a more uniform distribution. This enzyme was more<br />

expressed in the diencephalon of females and in the optic lobes of males.<br />

In quail embryos 3β-HSD mRNA level showed a strong sexual dimorphism at E7, E9 and<br />

E15. At E7, females showed higher 3β-HSD mRNA level than males whereas males had<br />

more 3β-HSD mRNA than females at E9 and E15. In females, 3β-HSD mRNA was at its<br />

highest level at E7. In males, the relative level of 3β-HSD mRNA was significantly higher<br />

at E9 and E15 than at E7 and at E11.<br />

222


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

Conversely, the level of P450scc and of P450c17 mRNAs did not show sexual dimorphism<br />

nor development-dependent regulation. The expression of aromatase varied as a function<br />

of the embryonic age being significantly elevated at E9 and at E15 when compared to E7<br />

and E11 in both the sexes.<br />

In conclusion our study showed for the first time a sexual dimorphism in mRNA level of<br />

neurosteroidogenic enzymes in the adult and in the embryonic quail brain; provided the<br />

first description for P450scc gene expression in the optic lobes and confirmed the<br />

widespread capacity of quail brain to synthesize steroids starting from cholesterol.<br />

Further investigation is needed to understand the exact anatomical location of the<br />

described dimorphisms.<br />

Reference list<br />

1. Ball GF, and Balthazart J. Hormonal regulation of brain circuits mediating male sexual behavior in<br />

birds. Physiol Behav. 83 (2004) 329-46.<br />

2. Balthazart J., Baillien M., Cornil CA., and Ball GF. Preoptic aromatase modulates male sexual<br />

behavior: slow and fast mechanisms of action. Physiol Behav. 83 (2004) 247-70.<br />

3. Holloway CC, and Clayton DF. Estrogen synthesis in the male brain triggers development of the avian<br />

song control pathway in vitro. Nat Neurosci. 4 (2001) 170-5.<br />

4. Kawato, S., Yamada, M. and Kimoto, T. Brain neurosteroids are 4 th generation neuromessengers in the<br />

brain: Cell biophysical analysis of steroid signal transduction. Adv. in Biophys. (2003) 1-48.<br />

5 Tsutsui, K., Matsunaga, M., Miysbara, H., and Ukena, K. Neurosteroids biosynthesis in the quail brain: a<br />

review.Journal of Exp. Zool. 305A (2006) 733-742.<br />

223


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

THE POTENTIAL ROLE OF A NEURONAL AUTOCRINE/PARACRINE<br />

MECHANISM IN THE REGULATION OF NEUROSTEROID PRODUCTION:<br />

LUTEINIZING HORMONE RECEPTOR MEDIATES NEUROSTEROID<br />

PRODUCTION VIA UPREGULATION OF STEROIDOGENIC ACUTE<br />

REGULATORY PROTEIN EXPRESSION<br />

Atwood C.S.*, Wilson A.C.*, Bowen R.L.°, Vadakkadath Meethal S.* and Liu T.*<br />

*Department of Medicine, University of Wisconsin and the Geriatrics Research, Education<br />

and Clinical Center, VA Hospital, 2500 Overlook Terrace., Madison, WI 53705, U.S.A.<br />

Fax +608-2807291 e-mail: csa@medicine.wisc.edu<br />

°OTB Research, Raleigh, North Carolina 27615, U.S.A.<br />

Neurosteroids are synthesized by both neurons and glia, however the hormonal<br />

regulation of neurosteroid synthesis is unknown. In the gonads, steroid production is<br />

induced by luteinizing hormone (LH) and its signaling is mediated via luteinizing<br />

hormone/human chorionic gonadotropin (LH/hCG) receptors. Importantly, LH/hCG<br />

receptors have been shown to be expressed by neuronal cells throughout the brain<br />

(reviewed in [1, 2]). Since LH/hCG can cross the blood-brain barrier [3]; are present in<br />

cerebrospinal fluid [4]; and are expressed by neuronal cells [5, 6], we tested whether LH<br />

also might signal via neuronal LH/hCG receptors to modulate neurosteroid synthesis.<br />

Treatment of differentiated rat primary hippocampal neurons and human M17<br />

neuroblastoma cells with LH (100 mIU/ml) resulted in a 2-fold increase in pregnenolone<br />

secretion in both cell types, suggesting an increase in P450scc mediated cleavage of<br />

cholesterol to pregnenolone and its secretion from neurons [7]. To explore how LH might<br />

regulate the synthesis of pregnenolone, the precursor for steroid synthesis, we treated rat<br />

primary hippocampal neurons with LH (0, 10 and 100 mIU/ml) and measured changes in<br />

the expression of LH receptor and steroidogenic acute regulatory protein (StAR). LH<br />

induced a rapid (within 30 min.) increase in the expression of StAR, but induced a dosedependent<br />

decrease in LH receptor expression. Consistent with these results, the<br />

suppression of serum LH in young rats treated with leuprolide acetate for 4 months<br />

downregulated StAR expression but increased LH receptor expression in the brain. Taken<br />

together, these results indicated that LH induces neuronal pregnenolone production by<br />

modulating the expression of the LH receptor, increasing mitochondrial cholesterol<br />

transport and increasing P450scc mediated cleavage of cholesterol for pregnenolone<br />

synthesis and secretion.<br />

The source of LH (and hCG in humans) to signal for neurosteroid production is unclear;<br />

neuronal LH may be from pituitary LH secreted into the blood stream that crosses the<br />

blood-brain barrier [3]. Alternatively, LH has been localized to the cytoplasm of neurons<br />

in the cerebral cortex and hippocampus of human brain (e.g. [5]), raising the possibility<br />

that neurons synthesize LH de novo. Since gonadotropin-releasing hormone receptor 1<br />

(GnRHR1) has been localized to the limbic system of the brain, we tested whether<br />

GnRH1-induced neuronal LH expression by treating cultured human M17 neuroblastoma<br />

cells with GnRH1 for 6 h. M17 neuroblastoma cells expressed LHbeta mRNA while<br />

immunoblot analyses indicated the presence of 3 LH variants (~30, 47 and 60 kDa) that<br />

were upregulated by low concentrations of GnRH1, but downregulated at higher GnRH1<br />

concentrations [6]. LH expression also increased in differentiating embryonic rat primary<br />

cortical neurons. Our results demonstrate that neurons expressing GnRHR1 respond to<br />

GnRH1 by upregulating LH production. Post-reproductive surges in GnRH1 secretion<br />

may explain the accumulation of LH in pyramidal neurons of the aged human.<br />

224


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

The contribution of steroids produced in the brain, versus those produced in the<br />

periphery, to normal neuronal function is unknown. Given the importance of steroids to<br />

brain function, it is possible that neurosteroid production may be a mechanism to fine-tune<br />

the level of sex steroids in the brain. The neuronal production of sex steroids may be more<br />

crucial to brain function when gonadal sex steroid production decreases with menopause<br />

and andropause. This is supported by a recent study in mice showing that despite the<br />

decrease in serum sex steroids following ovariectomy, brain estrogen levels remain as high<br />

as control mice [8], possibly due to an increased production of neurosteroids as a result of<br />

the ovariectomy-induced increases in serum LH [9]. In humans however, brain<br />

neurosteroid levels have been shown to decrease when serum LH levels are increasing with<br />

age [10]. Neurosteroid concentrations also are decreased in individuals with AD versus<br />

age-matched controls even though serum 17β-estradiol levels are unchanged [8] and serum<br />

LH levels are elevated [11]. Further studies are required to determine the contribution of<br />

neuronal versus peripheral LH in mediating neurosteroid production.<br />

Taken together, our results suggest that the production of neurosteroids may be<br />

regulated in an autocrine/paracrine manner, whereby GnRH neurons secrete GnRH1 which<br />

binds to GnRHR1 in the limbic system for the local production of LH. Secreted neuronal<br />

LH then binds to LH receptors to elicit neurosteroid synthesis and secretion. It might be<br />

predicted that like the hypothalamic-pituitary-gonadal (HPG) axis, neurosteroids<br />

negatively feedback on GnRH neurons to regulate neuronal hormone synthesis. Thus, we<br />

propose that neurosteroid levels may be modulated by neuronal autocrine/paracrine<br />

mechanisms in addition to peripherally produced steroids.<br />

Reference list<br />

1. Lei Z.M., Rao C.V., Neural actions of luteinizing hormone and human chorionic gonadotropin. Semin.<br />

Reprod. Med. 19 (2001) 103-109.<br />

2. Vadakkadath Meethal S., Atwood C.S., The role of hypothalamic-pituitary-gonadal hormones in the<br />

normal structure and functioning of the brain. Cell Mol. Life Sci. 62 (2005) 257-270.<br />

3. Lukacs H., Hiatt E.S., Lei Z.M., Rao C.V., Peripheral and intracerebroventricular administration of<br />

human chorionic gonadotropin alters several hippocampus-associated behaviors in cycling female rats.<br />

Horm. Behav. 29 (1995) 42-58.<br />

4. Bagshawe K.D., Orr A.H. & Rushworth A.G., Relationship between concentrations of human<br />

chorionic gonadotrophin in plasma and cerebrospinal fluid. Nature 217 (1968) 950-951.<br />

5. Bowen R.L., Smith M.A., Harris P.L., Kubat Z., Martins R.N., Castellani R.J., Perry G., Atwood C.S.,<br />

Elevated luteinizing hormone expression colocalizes with neurons vulnerable to Alzheimer's disease<br />

pathology. J. Neurosci. Res. 70 (2002) 514-518.<br />

6. Wilson, A.C., Salamat, M.S., Haasl, R.J., Roche, K.M., Karande, A., Vadakkadath Meethal, S.,<br />

Terasawa, E., Bowen, R.L. and Atwood, C.S., Human neurons express Type I GnRH receptor and<br />

respond to GnRH I by increasing luteinizing hormone expression. J Endocrinol. 191 (2006) in press.<br />

7. Liu T., Wimalasena J., Bowen R.L. & Atwood C.S., Luteinizing hormone receptor mediates neuronal<br />

pregnenolone production via upregulation of steroidogenic acute regulatory protein expression. J.<br />

Neurochem. in press.<br />

8. Yue X., Lu M., Lancaster T., Cao P., Honda S., Staufenbiel M., Harada N., Zhong Z., Shen Y., Li R.,<br />

Brain estrogen deficiency accelerates Abeta plaque formation in an Alzheimer's disease animal model.<br />

Proc. Natl. Acad. Sci. U. S. A. 102 (2005) 19 198-19 203.<br />

9. Parlow A. F., Effect of ovariectomy on pituitary and serum gonadotrophins in the mouse.<br />

Endocrinology. 74 (1964) 102-107.<br />

10. Rosario E.R., Chang L., Stanczyk F.Z., Pike C.J., Age-related testosterone depletion and the<br />

development of Alzheimer disease. JAMA. 292 (2004) 1431-1432.<br />

11. Short R.A., Bowen R.L., O'Brien P.C. & Graff-Radford N.R., Elevated gonadotropin levels in patients<br />

with Alzheimer disease. Mayo Clin. Proc. 76 (2001) 906-909<br />

225


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

NEUROSTEROIDS-GROWTH FACTORS INTERACTION INDUCES UP AND<br />

DOWN REGULATION OF GFAP AND VIMENTIN EXPRESSION IN<br />

ASTROGLIAL CELLS MAINTAINED UNDER SERUM-FREE STRESSED<br />

CULTURE CONDITIONS<br />

V. Bramanti 1 , D. Bronzi 2 , G. Raciti 3 , M. Avitabile 3 , R. Avola 1<br />

1<br />

Department of Chemical Sciences, Section of Biochemistry and Molecular Biology, via<br />

A. Doria 6, 95125 University of Catania Italy Fax 0039-0957384220 – e-mail:<br />

ravola@unict.it<br />

2 Department of Physiological Sciences, University of Catania, Italy.<br />

3<br />

Department of Biological Chemistry, Medical Chemistry and Molecular Biology,<br />

University of Catania, Italy.<br />

The glucocorticoid dexametasone (DEX) influences astroglial cells regulating glial<br />

fibrillary acidic protein (GFAP) expression (1, 2) and play also a pivotal role in neuronal<br />

differentiation in culture.<br />

Interesting evidences (3) demonstrated that DEX can differentially regulate the expression<br />

of growth factors [bFGF, Nerve Growth Factor (NGF) and S-100 beta], in hippocampal<br />

astrocytes in vitro, suggesting that one of the mechanisms through which glucocorticoids<br />

(GCs) affect hippocampal functions may be regulated by the expression of astrocytederived<br />

growth factors.<br />

Neuron – glia cross-talk is modulated by mitogenic growth factors (GFs) EGF, IGF-I,<br />

Insulin (INS) and bFGF able to induce astroglial and neuronal cell proliferation and<br />

differentiation under different culture conditions.<br />

Serum deprivation is one exogenous stimulus, like glucocorticoids and cAMP, capable to<br />

enhance spontaneous or growth factor-induced astroglial differentiation (4).<br />

The aim of present investigation was to study the interactions between GFs and DEX on<br />

cytoskeletal proteins GFAP and vimentin (VIM) expression under different experimental<br />

conditions.<br />

Condition 1: 24h pre-treatment with bFGF, subsequent 72h switching in serum-free<br />

medium (SFM) and final addition of GFs alone or by two in the last 24h, after a prolonged<br />

(60h) DEX treatment.<br />

Condition 2: 36h pre-treatment with DEX and bFGF in the last 24h followed by SFM for<br />

60h and final GFs addition for 24h alone or by two.<br />

Western blot analysis data showed a marked GFAP expression in cultures submitted to<br />

condition 1 comparing results to untreated or treated controls. In particular, the maximum<br />

level of GFAP expression was observed when EGF or INS or both together were added in<br />

a prolonged 60h DEX treatment, comparing the results to both untreated and pretreated<br />

control cultures. This finding well correlates with differentiative role played by<br />

glucocorticoids interacting with the “competence” factor bFGF and demonstrates as<br />

increased GFAP expression mostly depends on maturation, rather than proliferating status<br />

of astroglial cells in culture.<br />

Under the same culture conditions (condition 1), VIM expression was instead significantly<br />

reduced after GFs addition in the last 24h of 60h DEX treatment, respect to control DEXpretreated<br />

ones. On the contrary, referring data to untreated controls, VIM expression was<br />

significantly enhanced after GFs addition.<br />

GFAP showed also a significant increase in EGF or INS or IGF-I or both EGF+INS<br />

treated astrocytes submitted to condition 2 respect to both control untreated or treated<br />

cultures.<br />

226


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

VIM expression was up an down regulated under condition 2. In particular, the expression<br />

of vimentin was significantly increased in cultures pretreated for 36h with DEX and bFGF<br />

in the last 24h, switching under serum free conditions (SFC) for 60h and final treatment for<br />

24h with INS or EGF+IGF-I or EGF+INS, comparing the results to both untreated and<br />

pretreated control ones.<br />

On the contrary, vimentin expression was markedly decreased after IGF-I addition, when<br />

data were compared to both untreated and pretreated controls.<br />

Our data demonstrate that the pre-treatment with “competence” factor bFGF, the<br />

subsequent switching for a long period under SFC and final treatment with DEX for 60h<br />

induces an up and down regulation of cytoskeletal protein expression depending on<br />

mitogenic synergistic effect evoked by some “progression” growth factors , like EGF o<br />

INS or both together.<br />

Collectively, our results indicate that progression growth factors addition can regulate<br />

GFAP and VIM expression, depending on pre-treatment with DEX and/or bFGF in<br />

cultures switched in SFC. This suggests an interactive dialogue between these two class of<br />

neuroactive molecules and confirm the complex role played by glucocorticoids and both<br />

‘‘competence’’ and ‘‘progression’’ growth factors, regulating astrocytic cytosckeletal<br />

network under stressed and adversed environmental conditions.<br />

Reference list:<br />

1. Avola R, et al., Clin Exp. Hypertension 2004 May;26(4):323-33<br />

2. Avola R, et al. Clin Exp. Hypertension 2002 Oct-Nov;24(7-8):753-67<br />

3. Niu H, et al., Brain Res Mol Brain Res 1997 51: 97-105.<br />

4. Loo T. et al., J Neuroscience Research 1995, 42(2): 184-91.<br />

227


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

SEX DIFFERENCES IN RAPID ESTROGEN ACTION ON GABAERGIC<br />

NEURONS IN VIVO<br />

Balog J. 1 , Szegő É.M. 1 , Erdei F. 2 , Szabó G. 2 , Juhász G. 1 and Ábrahám I.M. 1<br />

1<br />

Research Group of Neurobiology of Hungarian Academy of Sciences, Eötvös Loránd<br />

University, Pázmány Péter sétány 1/c, Budapest, H-1117, Hungary<br />

2<br />

Laboratory of Molecular Biology and Genetics, Institute of Experimental Medicine of<br />

Hungarian Academy of Sciences, Szigony utca 43, Budapest, H-1083, Hungary<br />

email: turboka@inf.elte.hu<br />

GABA is the most important inhibitory transmitter in the brain. GABAergic neurons<br />

play pivotal role in all brain function including cognition, perception, regulation of stress<br />

response or fertility. Previous findings showed that there is clear sex difference in function<br />

of GABAergic neurons such as the GABA release or rate limiting enzyme of GABA<br />

production. The mechanism of this sexually dimorphic GABAergic functions is not clear.<br />

One possibility that estrogen induces sexually dimorphic actions on these neurons.<br />

Besides the well-established estrogen receptor (ER)-mediated direct genomic effect,<br />

estrogen also exerts rapid nonclassical actions via different signaling pathways. Previously<br />

we have demonstrated that estrogen can induce a rapid activation of signaling pathways<br />

regulating transcription factors such as cAMP responsive element binding protein (CREB)<br />

via ER in different brain regions and in cholinergic and gonadotropin relasing hormon<br />

(GnRH) neurons in vivo. In our present study, using the CREB phosphorylation as an<br />

indicator of rapid estrogen action, we examined the estrogen’s effect on CREB<br />

phosphorylation in the GABAergic neurons. In addition we also determined the ER<br />

expression of GABAergic neurons in different brain areas.<br />

In our experiments we used GAD65-GFP transgenic mice expressing green<br />

fluorescent protein (GFP) under the control of the GAD65 gene’s regulatory domain. 99%<br />

of all GABAergic neurons exhibited GFP fluorescence and there were no GFP<br />

fluorescence without GAD expression in the GAD-GFP transgenic mice. In order to<br />

eliminate endogenous estrogens, all experiments were performed on gonadectomized adult<br />

female or male mice to which exogenous estrogen was then administered.<br />

In our first experiment we have examined the rapid effect of estrogen on CREB<br />

phosphorylation in GAD-GFP neurons in different brain areas by means of quantitative<br />

fluorescent immunohistochemistry. Whereas estradiol had no effect on the numbers of<br />

GAD-GFP neurons expressing CREB, an increase in pCREB expression was detected in<br />

GAD-GFP neurons in medial preoptic area (mPOA), median preoptic area (MnPO) and<br />

subparaventricular area (SpA) in female mice 15 min following estrogen administration.<br />

pCREB in females<br />

pCREB in males<br />

100<br />

100<br />

% in GAD-GFP neurons<br />

80<br />

60<br />

40<br />

20<br />

*<br />

* *<br />

% in GAD-GFP neurons<br />

80<br />

60<br />

40<br />

20<br />

0<br />

LS MnPO BNST mPOA SI SpA mAMY STR<br />

0<br />

LS MnPO mPOA BNST SI SpA mAMY STR<br />

Fig. 1. Effect of estrogen on CREB phosphorylation in GAD-GFP neurons in specific brain areas in<br />

gonadectomized GAD-GFP female and male transgenic mice. *p


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

By contrast, the estrogen did not alter the pCREB expression of GAD-GFP neurons<br />

in any of the brain regions analyzed in males (fig 1).<br />

In the next experiment ER-expression of GAD-GFP neurons was determined in the<br />

brain using quantitative fluorescent immunohistochemistry. Our results showed that GAD-<br />

GFP neurons exclusively expressed ERα in lateral septum (LS), MnPO while GAD-GFP<br />

positive neurons were labelled for ERα and ERβ in bed nucleus of stria terminalis (BNST),<br />

mPOA, substantia innominata (SI), SpA, medial amygdala (mAMY). By contrast, GAD-<br />

GFP neurons do not express ERs in striatum (STR). There were significant difference in<br />

ERα levels of GAD-GFP neurons of female and male mice in mAMY, SI and SpA.<br />

Regarding the ERβ expression of GAD-GFP neurons, there were significant differences<br />

between sexes in BNST, and mAMY.<br />

In summary, the present study reveals that clear sex differences exists in the ability of<br />

estrogen to phosphorylate CREB within GABAergic neurons and the ERs expression in<br />

specific brain regions in vivo. Our findings also demonstrate that the anatomical difference<br />

of estrogen-induced CREB phosphorylation in GABAergic neurons in female and male<br />

mice does not correspond to the sexdimorphic expression of ERs in GABAergic neurons.<br />

229


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

DHEAS AND THE HOMINOID BRAIN<br />

Campbell B.C.<br />

Anthropology Department, Harvard Unversity, 11 Divinity Ave., Cambridge MA 02138<br />

USA Fax: 617-496-8041 email: bccampb@fas.harvard.edu<br />

Dehydroepiandrosterone (DHEA) and its sulfate (DHEAS), the most common steroids in<br />

the human body, have has been demonstrated to have a wide variety of physiological<br />

effects on the skeletal, immunological, vascular, and nervous systems. Furthermore,<br />

adrenarche, the prepubertal rise in adrenal production of DHEAS, is a distinctive feature of<br />

hominoid life history, documented for humans and the African apes, but lacking among<br />

other primates. Yet, the evolution of adrenarche remain a mystery. [1].<br />

This lack of understanding may reflect three major sources of confusion in the current<br />

literature: 1) the wide variety of effects exhibited by DHEA makes it difficult to establish<br />

its main function; 2) DHEA can be converted into a variety of other steroids, including<br />

estrogen and testosterone making it unclear if DHEA itself has specific effects or acts<br />

primarily as a source of estrogen and testosterone; 3) evidence for a well-defined receptor<br />

for DHEA, physiologist’s hallmark of functionality, has been lacking, further obscuring<br />

attempts to delineate the specific actions (if any) of DHEAS.<br />

However, the characterization of DHEA as a neurosteroid [2], and recent evidence of a<br />

specific DHEA receptor [3] provide new insights into the role of adrenarche. As a<br />

neurosteroid, the original function of DHEAS may have been neurological, focusing<br />

attention on a possible role in brain expansion during human evolution. The existence of a<br />

DHEA receptor would suggest that DHEA represents a meaningful signal that can be<br />

altered by selection. The fact that DHEA represents one step in the biosynthetic steroid<br />

pathway makes DHEA a likely candidate for an evolvable system. Furthermore, the fact<br />

that increases in DHEA continue into the early 20s mean that DHEA production could<br />

coordinate life history changes in both brain and body throughout development.<br />

I argue that adrenarche emerged in the great apes to support the development of continued<br />

brain maturation associated with a long life span, extended juvenile period and a fissionfusion<br />

social organization. I base this argument on two lines of evidence; 1) the role of<br />

DHEA in primate fetal brain development; 2) anatomical and genetic features specific to<br />

the hominoid brain which suggest increasing complexity relative to other primates.<br />

Together these lines of evidence suggest that adrenarche may be a marker of an extended<br />

period of juvenile development that leads to slow somatic growth while promoting<br />

increased neuroplasticity and enhanced social learning.<br />

Among primates, DHEA appears to be particular important during prenatal development<br />

[4]. DHEAS is produced by the fetal adrenal gland in larger amounts as precursor to<br />

placental estradiol production. Both DHEA and DHEAS have been shown to effect the<br />

development of fetal neural cells [5], suggesting that the two hormones may play a special<br />

role in promoting primate fetal brain development. Thus the development of a distinct zona<br />

reticularis and DHEA production by the adrenal gland starting around the age of three and<br />

continuing through adolescence [6] suggests a similar neurological function throughout<br />

childhood development.<br />

230


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

While brain maturation undoubtedly involves many cellular processes DHEA may play a<br />

particular role in shaping the development of neural conntections. DHEAS is known to<br />

have a variety of neuronal effects, including acting at both the GABA [7], and NMDA [8]<br />

receptor, as well as promoting glutamate release [9]. The effect of DHEA on glutamate<br />

receptors is of special interest since humans and great apes share a hominoid specific gene<br />

thought to promote glutamate turnover [10]. Thus increasing levels of DHEA at adrenarche<br />

would potentiate the excitotoxic effects of glutamate in shaping neural connections.<br />

NDMA receptors are present at high density in the striatum, amygdala and prefrontal<br />

cortex, parts of the mesolimbic and corticolimbic systems important in emotion and the<br />

control of behavior.<br />

Taken together the available literature suggests that adrenarche may represent an<br />

adaptation among humans and the great apes to increase levels of DHEA in support of<br />

brain development in response to an extend period of development and complex<br />

environment.<br />

Reference list<br />

1. Campbell, B.C. 2006. Adrenarche and the evolution of human life history. Am J Hum Biol. 18,569-589<br />

2. Baulieu, E.E. 1998. Neurosteroids: a novel function of the brain. Pyschoneuroendocrinol 23,963-987.<br />

3. Chen, F., Knecht, K., Birzin, E., Fisher, J., Wilkinson, H., Mojena, M., Moreno, C.T., Schimdt, A.,<br />

Harada, S., Freedman, L.P., Reszka, A.A. 2005. Direct agonist/antagonist functions of<br />

dehydroepiandrosterone. Endocrinology 4568-4576.<br />

4. Mesiano, S., Jaffe, R.B. 1997. Developmental and functional biology of the primate fetal<br />

adrenal cortex. Endocrine Rev. 18,378-403.<br />

5. Compagnone, N.A., Mellon, S.H. 1998. Dehydroepiandrosterone: a potential signaling molecule for<br />

neocortical organization during development. PNAS (USA). 95,4678-4683.<br />

6. Remer, T., Boye, K.R., Hartmann, M.F., Wudy, S.A. 2005. Urinary markers of adrenarche: reference<br />

values in health subjects, aged 3-18 years. J Clin Endocrinol Metabol. 90,2015-2021.<br />

7. Majewska, M., Demigoren, S., Spivak, C.E., London, E.D. 1990. The neurosteroid<br />

dehydroepiandrosterone sulfate is an allosteric antagonist of the GABAa receptor. Brain Res.<br />

526,143-146.<br />

8. Monnet, F.P., Maurice, T. 2006. The sigma1 protein as a target for the non-genomic effects of neuro<br />

(active) steroids: molecular, physiological, and behavioral aspects. J Pharmacol Sci. 100,93-118.<br />

9. Lhullier, F.L.R., Nicolaidis, R., Riera, N.G., Ciprirna, F., Junqueira, D., Dahm, K.C.S., Brusque, A.M.,<br />

Souza, D.O. 2004. Dehydroepiandrosterone increases synaptosomal glutamate release and improves<br />

the performance in inhibitory avoidance task. Pharmacol Biochem Behav. 77,606-610.<br />

10. Birke, F., Kaessman, H. 2004 Birth and adaptive evolution of a hominoid gene that supports high<br />

neurotransmitter flux. Nature Genetics 36,1061-1063.<br />

231


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

NITRIC OXIDE EFFECTS ON SEXUAL AND MATERNAL BEHAVIOR IN THE<br />

FEMALE RAT<br />

Carrillo B, Pinos H., Guillamón A., Panzica G.C. 1 , Pérez-Izquierdo M.A., Collado P.<br />

Departamento de Psicobiología, Universidad Nacional de Educación a Distancia.<br />

C/ Juan del Rosal, 10. P.O Box 60148 28040-Madrid, Spain.<br />

1 Rita Levi Montalcini, Dipartimento di Anatomia, Farmacologia e Medicina Legale, Laboratorio di<br />

Neuroendocrinología, Università di Torino, Torino, Italy.<br />

Nitric oxide (NO) is a gas that represents a new form of neurotransmission. NO plays an<br />

important role in the regulation of reproductive behaviors such as maternal and sexual<br />

behavior [1-4] and its production is regulated by estradiol in several mammalian species<br />

[5-8]. Sexual behavior in female rats is complex and it is constituted by proceptive and<br />

receptive behaviors, all of them can be measured [9]. In previous studies, it has been<br />

shown that NO affects sexual behavior in female rats throughout its influence in hormonal<br />

secretion and in the activity of some cerebral structures related to this behavior [3].<br />

Although, the possible NO role in the maternal behavior is controversial [10, 11].<br />

In the present study, we investigate the role of NO in sexual and maternal behaviors in<br />

Wistar female rat. For this purpose, we have administered intraperitoneally (i.p.) a NO<br />

precursor (L- Arginina) and a NO inhibitor (L-NAME).<br />

Sexual behavior: In order to test sexual bahavior 22 female Wistar rats three month old<br />

were studied. Before the test Ss. were ovariectomized and primed with estradiol benzoate<br />

and progesterone to induce sexual behavior. Animals were randomly assigned to one of<br />

three groups: control (i.p. saline injected), precursor (i.p injected. with a dose of 25mg/kg<br />

of L-Arginina) and inhibitor (i.p. injected with a dose of 25 mg/kg of L-NAME). All<br />

experimental treatment was administered before testing sexual behavior. Data from this<br />

experiment have shown a significant difference between control and L-NAME groups, and<br />

between L-NAME and L-Arginina groups. It can be concluded that proceptive behaviors<br />

were not affected either by L-Arginina or L-NAME but lordotic behaviour decresed when<br />

NO inhibitor is inyected.<br />

Natural maternal behaviour: To test maternal behavior 28 primiparous Wistar female rats<br />

weere studied inmediately after delivery. The primiparous were distributed in three groups:<br />

control (i.p. saline injected), NO precursor (i.p. injected with a dose of 25mg/kg of L-<br />

Arginina) and NO inhibitor (i.p. injected with a dose of 25 mg/kg of L-NAME). Five<br />

components of maternal behavior were recorded, and only three were significantly affected<br />

by the treatment: nest building, grooming and licking behavior. In all of them results<br />

indicate that L-Arginina interferes with these maternal behavior patterns.<br />

In conclusion, these data suggest different NO effects depending on the reproductive<br />

behavior we are studing: NO inhibitor decreases female sexual behavior and NO precursor<br />

seems to deteriorate some aspects of maternal behavior.<br />

232


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

Reference list<br />

1. Hull, E.M., Lumley, L.A., Matuszewich, L., Dominguez, J., Moses, J., Lorrain,<br />

D.S., 1994. The roles of nitric oxide in sexual function of male rats.<br />

Neuropharmacology. Nov, 33 (11), 1499-504.<br />

2. Hull, E.M., Lorrain, D.S., Du, J.F., Matuszewich, L., Lumley, L.A., Putnam, S.K.,<br />

Moses, J., 1999. Hormone-neurotransmitter interactions in the control of sexual<br />

behavior. Behv. Brain Res., 105, 105-116.<br />

3. Mani, S.K., Allen, J.M.C, Rettori, V., McCann, S.M., O’Malley, B.W., Clark, J.H.,<br />

1994. Nitric oxide mediates sexual behavior in female rats. Proc. Natl. Acad. Sci.<br />

91, 6468-6472.<br />

4. Benelli, A., Bertolini, A., Poggioli, R., Cavazzuti, E., Calza, L., Giardino, L.,<br />

Arletti, R., 1995 Nitric oxide is involved in male sexual behaviour of rats. Eur. J.<br />

Pharmacol. 294, 505-510.<br />

5. Collado, P., Guillamón, A., Pinos, H., Pérez-Izquierdo, M.A., García-Falgueras, A.,<br />

Carrillo, B., Rodríguez, C., Panzica, G.C., 2003a. NADPH-diaphorase activity<br />

increases during estrous phase in the bed nucleus of the accessory olfactory tract in<br />

the female rat. Brain Res. 983, 223-229.<br />

6. Ceccatelli, S., Grandison, L., Scott, R.E.M., Pfaff, D.W., Kow, L.M., 1996.<br />

Estradiol regulation of nitric oxide synthase mRNAs in rat hypothalamus.<br />

Neuroendocrinology 64, <strong>35</strong>7-363.<br />

7. Okamura, H., Yokosuka, M., Hayashi, S., 1994. Estrogenic induction of NADPHdiaphorase<br />

activity in the preoptic neurons containing estrogen receptor<br />

immunoreactivity in the female rat. J. Neuroendocrinol. 6, 597-601.<br />

8. Rachman, I.M., Unnerstall, J.R., Pfaff, D.W., Cohen, R.S., 1998. Regulation of<br />

neuronal nitric oxide synthase mRNA in lordosis-relevant neurons of the<br />

ventromedial hypothalamus following short-term estrogen treatment. Mol. Brain<br />

Res. 59,105-108.<br />

9. Beach, F.A., 1976. Sexual attractivity, proceptivity, and receptivity in female<br />

mammals. Horm Behav. Mar, 7 (1), 105-38.<br />

10. Popeski, N., Woodside, B., 2004. Central nitric oxide synthase inhibition disrupts<br />

maternal behavior in the rat. Behav Neurosci. Dec, 118 (6), 1305-16.<br />

11. Numan, M., 2004. Maternal behaviors: central integration or independent parallel<br />

circuits? Theoretical comment on Popeski and Woodside (2004). Behav Neurosci.,<br />

Dec, 118 (6), 1469-72.<br />

Sources of support MCyT: BSO2003-02526 (Paloma Collado) and BSO2003-08962<br />

(Antonio Guillamón).<br />

233


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

SINGLE OPIOID ADMINISTRATION MODIFIES GONADAL STEROIDS IN<br />

BOTH THE CNS AND PLASMA OF MALE RATS<br />

Ceccarelli I., De Padova A.M., Fiorenzani P., Massafra C. and Aloisi A.M.<br />

Pain and Stress Neurophysiology Lab., Neuroscience and Applied Physiology Section,<br />

Department of Physiology, University of Siena, Via Aldo Moro, 2, 53100 Siena, Italy<br />

While morphine remains one of the most widely used opioid agonists for the treatment of<br />

painful conditions, other opioid agonists are also commonly employed. Because of the<br />

interactions between opioids and gonadal hormones, in particular the opioid-induced<br />

hypogonadism, this study investigated the effects of widely used opioids on plasma<br />

testosterone and estradiol levels and brain testosterone levels in male rats. Animals were<br />

subcutaneously injected with two concentrations of morphine (5 or 10 mg/kg), fentanyl<br />

(0.05 or 0.1 mg/kg), tramadol (10 or 40 mg/kg), buprenorphine (0.05 or 0.1 mg/kg) or<br />

saline (0.7 ml/kg). Four or 24 hours after treatment, the rats were deeply anesthetized to<br />

collect blood samples from the abdominal aorta and to perfuse the brains with saline.<br />

Plasma and brain hormone levels were measured by radioimmunoassay. In rats studied 4<br />

hours after treatment, all opioids, but tramadol 10 mg/kg, decreased plasma testosterone in<br />

comparison with saline administration. At the same time, plasma estradiol levels were<br />

lower than control in the groups treated with the low doses of morphine, tramadol and<br />

buprenorphine, while estradiol remained at control levels in the other groups. Twenty-four<br />

hours after treatment, plasma testosterone levels were different (higher) than control only<br />

in the animals treated with the low doses of morphine, fentanyl and buprenorphine.<br />

Estradiol was lower than control in the low dose groups, while the high doses did not<br />

produce any changes with respect to control. Four hours after treatment, brain testosterone<br />

was drastically decreased in all groups treated with the higher dose, except buprenorphine,<br />

in which it remained at control levels. All groups returned to control levels at 24 hours<br />

after treatment. The different magnitude and time-course of the effects of the different<br />

opiate agonists on testosterone and estradiol levels are likely due to their different<br />

mechanism of action.<br />

234


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

OXIDATION-DEPENDENT PLASMA DHEA FORMATION AS A DIAGNOSTIC<br />

TOOL FOR ALZHEIMER’S DISEASE PATHOLOGY: RESULTS FROM A<br />

TRIAL<br />

Chalbot S. *† , Lecanu L. *† , Greeson J. ‡ and Papadopoulos V *† .<br />

* Georgetown University Medical Center, Department of Biochemistry, Molecular and<br />

Cellular Biology, 3900 Reservoir Rd NW, Washington DC, 20057, USA. † Samaritan<br />

Pharmaceuticals Research Laboratories, Georgetown University Medical Center. ‡<br />

Samaritan Pharmaceuticals Inc., 101 Convention Drive Center, Las Vegas NV, 89109,<br />

USA<br />

Sonia Chalbot, snc9@georgetown.edu, Fax (202)687-2<strong>35</strong>4<br />

Alzheimer’s disease (AD) is a progressive, irreversible neurodegenerative disorder<br />

characterized by loss of memory and other cognitive functions leading to dementia.<br />

Although powerful neuropsychological tests are available for the diagnosis and monitoring<br />

of the evolution of the disease, conclusive diagnosis of AD may only be made after postmortem<br />

examination of brain tissue for the presence of large numbers of plaques and<br />

tangles. Thus, the need for an Alzheimer’s disease biochemical marker (biomarker). The<br />

ideal AD biomarker must reflect a fundamental feature of AD neuropathology and should<br />

be reliable, reproducible, and non-invasive. Accordingly, such biomarker may improve our<br />

understanding of AD origin and diagnosis, and help monitor AD progression and the<br />

efficacy of AD therapeutic interventions.<br />

In search of a biomarker directed at the fundamental CNS pathophysiology of AD,<br />

we proposed to apply our recent findings on the presence of an alternative, oxidative<br />

stress-mediated, pathway of neurosteroid biosynthesis in the brain [1]. Brain cells can<br />

convert cholesterol to pregnenolone, which is the precursor for a number of steroid<br />

modulators of neuronal functions, including DHEA. We identified a novel, brain- and cellspecific<br />

mechanism for DHEA biosynthesis present in the rat, bovine and human species.<br />

In this scheme, DHEA biosynthesis is mediated by a cytochrome P450 17α-hydroxylase<br />

(CYP17)-independent mechanism involving a yet unidentified hydroperoxide precursor.<br />

This alternative pathway is regulated by agents, such as Fe ++ and β-amyloid (Aβ) peptide,<br />

both pro-oxidant agents abundant in AD brain. Using brain tissue specimens from control<br />

and AD patients we subsequently provided evidence that DHEA levels are elevated in AD<br />

brain tissue specimens and DHEA is formed in the AD brain by the oxidative stressmediated<br />

metabolism of an hydroxyperoxy-steroid precursor, thus depleting the levels of<br />

the precursor present in plasma. In the present study, we proposed to test for the presence<br />

of this DHEA precursor in human plasma using a simple Fe ++ -based reaction and<br />

determine the amounts of DHEA formed.<br />

A total of 40 patients were included in this study, 12 age-matched control, 10 AD mild, 3<br />

AD moderate, 8 AD severe and 7 MCI (mild cognitive impairment). Blood samples were<br />

collected on heparin to prevent any interaction between EDTA and our experimental<br />

protocol. Deuterated DHEA as the internal standard was added to human plasma and<br />

20mM FeSO4 solution. Tubes were shaken for 60 minutes at 37 °C. The incubations were<br />

stopped with the addition of ethyl acetate. After decantation and centrifugation, the organic<br />

phase was extracted and dry under nitrogen. The extraction process was repeated three<br />

times. Then the dry residue was partitioned between methanol/water (8:2 (v/v) and<br />

petroleum ether to eliminate the lipids and sterols. After decantation the methanol/water<br />

phase containing steroids for analysis was evaporated under nitrogen. The dry residues<br />

2<strong>35</strong>


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

were derivatized with using BSTFA in pyridine, evaporated under a stream of nitrogen and<br />

then dissolved in toluene. Samples were injected into GC–MS (GC17A/QP5050A,<br />

Shimadzu). DHEA was detected in selected ion monitoring (SIM) mode using the<br />

fragments (m/z=304; deuterated-DHEA, m/z=272). Statistical analyses were performed the<br />

software JMP 5.1 (SAS Institute, Palo Alto CA). The test used included ANOVA,<br />

Dunnett’s method and correlation.<br />

Plasma oxidation induced an increase of DHEA levels measured in control patient sera<br />

whereas no dramatic change was observed in patients diagnosed with severe AD. This<br />

significant difference of the biochemical pattern between both these groups (pF = 0.0<strong>35</strong>). The merged group displayed significant<br />

difference compared to the control group (p=0.012) and a very important difference versus<br />

the AD mild group (p=0.066). These results suggest that comparing DHEA levels in<br />

plasma before and after oxidation by FeSO4 permits to differentiate between healthy and<br />

AD mild patients on one side, and AD moderate and AD severe patients on the other side.<br />

The % of DHEA variation after oxidation correlates with the Mini Mental State Evaluation<br />

(MMSE) of the patients. The lower the MMSE, the lower the DHEA increase (Correlation<br />

Coefficient = 0.44, R 2 =0.196). The slope of the linear regression curve was determined to<br />

not be due to a random effect but to a statistically significant percent Diff DHEA/MMSE<br />

relationship (F ratio = 7.54, Prob>F = 0.009). The changes seen were independent of age<br />

and sex of the patients.<br />

These preliminary results suggest that the comparison of DHEA levels in patient<br />

plasma before and after oxidation by FeSO4 could provide a useful tool to diagnose<br />

Alzheimer’s disease. These results also suggest that the proposed assay will not<br />

misdiagnose MCI patients as Alzheimer’s patients. In addition, since a significant<br />

correlation was observed between DHEA variation (%) and patient MMSE, the developed<br />

methodology might be useful in the prediction/diagnosis of severity of the disease as well<br />

as to follow up on effects of various therapies used. It is evident that the validity of the<br />

proposed methodology and interpretation of these data will be significantly refined with<br />

the inclusion of more patients to this trial, in particular in the AD moderate group, as well<br />

as non-AD demented patients like the ones with fronto-temporal dementia. The correlation<br />

of the data obtained with clinical information on treatments taken by the patients, drugs<br />

used, doses, and duration of treatments would further define the utility of this test for<br />

Alzheimer’s disease pathology.<br />

Reference list<br />

1. Brown R.C., Han Z., Cascio C. and Papadopoulos V. Oxidative stress-mediated<br />

DHEA formation in Alzheimer’s disease pathology. Neurobiol. Aging 2003, 24(1):<br />

57-65.<br />

236


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

LESION-INDUCED GLIAL REACTION IN THE RAT OLFACTORY BULB:<br />

EFFECT OF DHEA AND DHEA DERIVATIVES<br />

Csakvari E., Hoyk S., Szajli A. * , Kurunczi A., Gyenes A., Berger A., and Parducz A.<br />

Laboratory of Molecular Neurobiology, Institute of Biophysics, Biological Research<br />

Center, Temesvari krt. 62., H-6701 Szeged, Hungary,<br />

E-mail: parducz@brc.hu; Fax: +36-62-433-133<br />

* Department of Organic Chemistry, University of Szeged, Hungary<br />

The neuroactive steroid, dehydroepiandrosterone (DHEA) was shown to influence<br />

the glial reactions of the peripherally denervated olfactory bulb in adult male rats. GFAP<br />

and vimentin immunostaining revealed that the deafferentation-induced reactive gliosis in<br />

the glomerular layer of the olfactory bulb was significantly diminished by DHEA. Western<br />

blot experiments have also shown that both chronic and acute DHEA treatment resulted in<br />

a significant decrease of GFAP expression levels. These findings indicate that DHEA<br />

attenuates glial reaction to denervation and may regulate glial plasticity in the olfactory<br />

glomeruli.<br />

To reveal the possible molecular mechanism of DHEA effect we selected different<br />

DHEA derivatives and studied their effects on the glial reactions.<br />

Denervation was achieved by destroying the olfactory mucosa with ZnSO 4 (0.17<br />

M) irrigation of the nasal cavities. The neurosteroids were applied in different doses during<br />

7 days. Rats were killed on day 7 after chemical denervation and reactive gliosis was<br />

monitored in the olfactory bulb using GFAP and vimentin immunohistochemistry.<br />

Qualitative changes in GFAP expression were analyzed by western blot.<br />

This work was supported by grants OTKA T 043436, M36252 and RET 08/2004<br />

237


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

DOSE-RESPONSE STUDY OF ANTIDEPRESSANT EFFECTS OF ESTROGENIC<br />

COMPOUNDS IN OVARIECTOMIZED MICE IN THE FORCED SWIM TEST<br />

Diz-Chaves Y*., Pernía O., Carrero P., Garcia-Segura L.M.<br />

Instituto Cajal, C.S.I.C., Avenida Doctor Arce, 37, E-28002, Madrid, Spain.<br />

*E-mail: ydiz@cajal.csic.es. FAX: 34-915854754<br />

Estradiol (E2) may influence depressive symptoms of women and decrease<br />

depressive behavior among rodents. Removal of the primary source of E2 (ovariectomy,<br />

ovx) increases depressive behavior of female rats and mice and E2 replacement reverses<br />

this effect. Previous reports have shown the ability of E2 to reduce depressive behavior in<br />

rats and mice in the forced swim test (FST), a well-established behavioral paradigm<br />

typically used to test the efficacy of antidepressants. In mice, the doses of estradiol used<br />

were 100 or 200 micrograms/Kg, and it has been described that E2 decreased duration of<br />

immobility in the FST in a dose dependent manner in different ovx mouse strains, after a<br />

minimum of three consecutive daily doses. In addition, previous studies have shown that<br />

selective estrogen receptor modulators (SERMs) with actions at estrogen receptor (ER)<br />

beta reduced depressive behavior in rats, but little is know about mice. The aim of this<br />

study was to determine the effective dose of different estrogenic compounds: E2, the ER<br />

alpha-specific SERM, PPT and the ER beta-specifc SERM, DPN, to induce an<br />

antidepressant effect on ovariectomized female mice in the FST. Female C57BL6 mice<br />

(n=104) of two months of age, were ovariectomized bilaterally under Rompun (2%) and<br />

ketamin (50 mg/Kg) anesthesia. Mice were housed in groups of six and maintained on a<br />

12h light-dark cycle in a temperature-controlled room with free access to food and water.<br />

After two weeks, animals were randomly assigned to an experimental group and<br />

behavioral studies performed. Mice were administered sesame oil vehicle, 17 beta-E2<br />

(Sigma), PPT or DPN (Tocris) in single doses of 50, 100 or 200 micrograms/Kg. As a<br />

positive control, some mice were administered a single injection of a tricyclic<br />

antidepressant, desipramine hydrochloride (DMI; Sigma) in a dose of 10 mg/Kg, or saline<br />

vehicle. Mice were tested on two occasions. First, in order to discard a possible influence<br />

of drug treatments on locomotor activity, the effect of the estrogenic compounds and<br />

antidepressant drug test was tested in the VersaMax activity monitor (Accuscan<br />

Instruments). Mice were placed in one of the four squares of a cage of 42 cm that<br />

mechanically recorded the number of beam breaks that occurred during a 5 minutes period.<br />

Mice were habituated during three consecutive days during 1 hour period and immediately<br />

after the last day of habituation, mice were administered the vehicle, the estrogenic<br />

compounds or DMI, and different locomotor activities (horizontal and vertical activity,<br />

total distance, number of movements, stereotype and rearing activity) were recorded 24-<br />

hours later. After two weeks, mice were tested in the forced swim test, according to the<br />

method of Porsolt. Swimming sessions were conducted by placing mice into an individual<br />

Plexiglas cylinder (29 cm height x 12 cm diameter) filled with water at 25ºC. Two<br />

swimming sessions were conducted: an initial 5-minutes pretest, one hour after drugs<br />

administration, followed by a 6-minutes test, 24 hours later. Test sessions were run<br />

between 11:00 and 15:00 hours and videotaped for later scoring. Two distinct types of<br />

behavior were assessed: mobility and immobility. One-Way analyses of variance<br />

(ANOVA) were utilized to determine if there were differences among the effects of<br />

estrogenic compounds on the basal activity or the forced swim test. In the present study we<br />

observed that none of the doses of estrogenic compounds analyzed produced changes in<br />

the different locomotor activities studied: horizontal and vertical activity, total distance,<br />

238


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

and number of movements, stereotype and rearing behavior. By contrast, DMI reduced the<br />

horizontal locomotor activities studied. Furthermore, a single acute injection of the<br />

different studied doses of 17 beta-E2, were unable to decrease immobility in the FST.<br />

However, at the higher dose tested (200 micrograms/Kg) we observed a tendency to<br />

increased immobility, but statically differences were not found. The lack of effectiveness<br />

of E2 to reduce depressive behavior observed in this study may be due in part to the<br />

duration of treatment and the strain of mice studied. It has been described that baseline<br />

immobility and sensitivity to antidepressants in the FST are strain dependent in mice. Also,<br />

it is known that high supraphysiological doses of E2 do not decrease immobility in the FST<br />

test. As observed for E2, a single acute injection of the ER alpha-specific SERM, PPT, did<br />

not affect immobility at any of the doses studied. However, the ER beta-specific SERM,<br />

DPN, significantly increased the duration of immobility at the dose of 100 micrograms/Kg.<br />

The lower dose (50 micrograms/Kg), did not have a significant effect. This finding seems<br />

to be in apparent contradiction to previous reports suggesting that E2 antidepressive effects<br />

are mediated by ER beta. However, the data presented here did not rule out the importance<br />

of ER beta, as important differences in the methodology could explain the discrepant result<br />

obtained. In most of the published studies, the FST procedure was performed in a single<br />

day, where the aspect of novelty and stress had great importance. Studying the<br />

antidepressive behavior in a two swimming sessions, eliminate the stress induced by<br />

novelty and the increased immobility observed in mice treated with DPN at 100<br />

micrograms/Kg, may be due to a less anxious behavior. As was expected, DMI, used as a<br />

positive control drug, significantly reduced duration of immobility when compared to the<br />

respective saline control. The diminution in locomotor activity did not interfere with the<br />

expression of active behaviors indicating that the anti-immobility effect of DMI is specific<br />

in the FST. In summary, in contrast to what it has been observed in ovx rats, our findings<br />

indicate that the acute administration of the studied estrogenic compounds does not have<br />

an antidepressive action in ovx mice. Although this may represent a species difference,<br />

further studies should determine whether chronic treatments with estrogenic compounds<br />

might have antidepressive actions in mice.<br />

Supported by Ministerio de Educación y Ciencia, Spain (SAF 2005-00272) and the<br />

European Union (EWA project: LSHM-CT-2005-518245).<br />

239


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

NEUROANATOMICAL AND BIOCHEMICAL EVIDENCE FOR THE OCCURRENCE<br />

OF CYTOCHROME P450 C17 IN THE FROG BRAIN. REGULATION BY VASOTOCIN<br />

AND MESOTOCIN<br />

Do Rego J.L.*, Tremblay Y. † , Luu-The V. † , Acharjee S. ‡ , Repetto E. § , Galas L.*, Castel H.*,<br />

Vallarino M. § , Kwon H.B. ‡ , Bélanger A. † , Seong J.Y.°, Pelletier G. † , Tonon M.C.*, Vaudry<br />

H.*<br />

*INSERM U413, Lab. Cell. Mol. Neuroendocrinol., Eur. Inst. Pept. Res. (IFRMP23), Univ. Rouen, 76821<br />

Mont-Saint-Aignan, France. Fax +33 2<strong>35</strong> 14 6946. e-mail: jean-luc.do-rego@univ-rouen.fr<br />

† Lab. Ontog. Reprod., and MRC Group Mol. Endocrinol. Oncol., CHU Laval, Québec G1V 4G2, Canada<br />

‡ Horm. Res. Center, Chonnam National Univ., Gwangju 500-757, Korea<br />

§ Dept. Exp. Biol., Univ. Genova, 16132 Genova, Italy<br />

°Lab. G Prot. Coupled Recept., Korea Univ. Coll. Med., Seoul 136-705, Korea<br />

It is now clearly established that the brain has the capability of synthesizing various<br />

biologically active steroids including 17-hydroxypregnenolone (17OH-PREG),<br />

17-hydroxyprogesterone (17OH-P), dehydroepiandrosterone (DHEA) and androstenedione<br />

(AD) [5]. However, the presence, distribution and activity of cytochrome P450 17alphahydroxylase<br />

/ C17, 20-lyase (P450 C17 ), a key enzyme required for the biosynthesis of these<br />

steroids in the central nervous system (CNS), are poorly documented. We took advantage<br />

of the availability of an antiserum raised against bovine testicular P450 C17 to determine the<br />

distribution of P450 C17 in the frog CNS. Immunohistochemical studies showed that<br />

P450 C17 -like immunoreactivity is widely distributed in the frog brain and pituitary.<br />

Prominent populations of P450 C17 -containing cells were observed in a number of nuclei of<br />

the telencephalon, diencephalon, mesencephalon and metencephalon as well as in the pars<br />

distalis and pars intermedia of the pituitary. The expression of P450 C17 in the brain and<br />

pituitary was also confirmed by Western blot analysis. In the brain, P450 C17 -like<br />

immunoreactivity was predominantly located in neurons. However, a small proportion of<br />

P450 C17 -positive cells, notably large cells in the optic tectum, in the tectal lamina six and in<br />

the pretectal grey of the mesencephalon, were identified as glial cells [2]. Labeling of<br />

consecutive sections of frog diencephalon with the antiserum against P450 C17 and the<br />

antiserum against 3beta-hydroxysteroid dehydrogenase / delta 5 -delta 4 isomerase (3beta-<br />

HSD) revealed that, in several hypothalamic nuclei, P450 C17 -positive cell bodies also<br />

contain 3beta-HSD-like immunoreactivity [2]. Incubation of telencephalon, diencephalon,<br />

mesencephalon, metencephalon or pituitary explants with tritiated pregnenolone (Preg)<br />

resulted in the formation of several tritiated steroids including 17OH-PREG, 17OH-P,<br />

DHEA and AD. De novo synthesis of C 21 17-hydroxy-steroids and C 19 ketosteroids was<br />

reduced in a concentration-dependent manner by ketoconazole, a P450 C17 inhibitor,<br />

demonstrating the presence of authentic P450 C17 activity in the frog brain [2].<br />

The areas where P450 C17 - and 3beta-HSD-positive cell bodies are located are richely<br />

innervated by vatocin (VT)- and mesotocin (MT)-immunoreactive fibers [9]. In<br />

amphibians, VT and MT, that are orthologues of mammalian vasopressin (VP) and<br />

oxytocin (OT), respectively, play a crucial role in the control of sexual behaviors [4, 6, 10].<br />

Since several neurosteroids also regulate reproduction-related behaviors [7], we have<br />

investigated the possible effect of VT and MT in the control of neurosteroid production.<br />

Double immunohistochemical labeling of frog brain sections with polyclonal antibodies<br />

against 3beta-HSD or P450 C17 and a monoclonal antibody against VP/VT [8] revealed the<br />

presence of VT/MT-positive fibers in close proximity of neurons expressing the<br />

steroidogenic enzymes 3beta-HSD and P450 C17 [3]. High concentrations of VT and MT<br />

receptor mRNAs were observed in diencephalic nuclei containing the 3beta-HSD and<br />

P450 C17 neuronal cell bodies [1, 3]. Exposure of frog hypothalamic explants to graded<br />

240


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

concentrations of VT or MT produced a dose-dependent increase in the formation of<br />

progesterone (P), 17OH-PREG, 17OH-P and DHEA. The stimulatory effect of VT and MT<br />

on neurosteroid production was suppressed by the 3beta-HSD inhibitor trilostane and the<br />

P450 C17 inhibitor ketoconazole. Time-course experiments revealed that a 30-min<br />

incubation of hypothalamic explants with VT or MT was sufficient to induce a robust<br />

increase in neurosteroid production and that the maximum effect was observed after a 2-h<br />

exposure to VT or MT, suggesting that VT and MT activate steroidogenic enzymes at a<br />

posttranslational level [3]. The stimulatory effect of VT and MT on neurosteroid<br />

biosynthesis was mimicked by VP and OT, as well as by a selective V1b receptor agonist,<br />

while a V2 receptor agonist and an OT receptor agonist had no effect. VT-induced<br />

neurosteroid production was completely suppressed by selective V1a receptor antagonists,<br />

and was not affected by a V2 receptor antagonist or an OT receptor antagonist.<br />

Concurrently, the effect of MT on neurosteroidogenesis was markedly attenuated by a<br />

selective OT receptor antagonist and a V1a receptor antagonist but not by a V2 receptor<br />

antagonist [3].<br />

In conclusion, the present study provides the first detailed immunohistochemical<br />

mapping of the steroidogenic enzyme P450 C17 in the brain and pituitary of any vertebrate.<br />

These data provide additional evidence that CNS neurons and pituitary cells can synthesize<br />

androgens. Our data also demonstrate for the first time the existence of a regulatory effect<br />

of VT and MT on neurosteroid biosynthesis, suggesting that some of the behavioral effects<br />

of VT and MT may be mediated through modulation of the activity of P450 C17 - and 3beta-<br />

HSD-expressing neurons.<br />

Supported by INSERM (U413), a France-Québec exchange program (INSERM-FRSQ), France-Korean<br />

exchange programs (INSERM-KOSEF and STAR), the Regional Platform in Cell Imaging, and the Conseil<br />

Régional de Haute-Normandie.<br />

Reference list<br />

[1] Acharjee, S., Do Rego, J.L., Oh, D.Y., Ahn, R.S., Lee, K., Vaudry, H., Kwon, H.B., Seong, J.Y., 2004.<br />

Molecular cloning, pharmacological characterization, and histochemical distribution of frog vasotocin<br />

and mesotocin receptors. J. Mol. Endocrinol. 33, 293–313.<br />

[2] Do Rego, J.L., Tremblay, Y., Luu-The, V., Repetto, E., Castel, H., Vallarino, M., Bélanger, A., Pelletier,<br />

G., Vaudry, H., 2006. Immunohistochemical localization and biological activity of the steroidogenic<br />

enzyme cytochrome P450 17alpha-hydroxylase/C17, 20-lyase (P450C17) in the frog brain and<br />

pituitary. J. Neurochem. (in press).<br />

[3] Do Rego, J.L., Acharjee, S., Seong, J.Y., Galas, L., Alexandre, D., Bizet, P., Burlet, A., Kwon, H.B.,<br />

Luu-The, V., Pelletier, G., Vaudry, H., 2006. Vasotocin and mesotocin stimulate the biosynthesis of<br />

neurosteroids in the frog brain. J. Neurosci. 26, 67496760.<br />

[4] Iwata, T., Toyoda, F., Yamamoto, K., Kikuyama, S., 2000. Hormonal control of urodele reproductive<br />

behavior. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 126, 221–229.<br />

[5] Mellon, S., Vaudry, H., 2001. Biosynthesis of neurosteroids and regulation of their synthesis. Int. Rev.<br />

Neurobiol. 46, 33–78.<br />

[6] Moore, F.L., Miller, L.J., 1983. Arginine vasotocin induces sexual behavior of newts by acting on cells<br />

in the brain. Peptides 4, 97-102.<br />

[7] Moore, F.L., Boyd, S.K., Kelley, D.B., 2005. Historical perspective: hormonal regulation of behaviors<br />

in amphibians. Horm. Behav. 48, 373383.<br />

[8] Robert, F.R., Léon-Henri, B.P., Chapleur-Château, M.M., Girr, M.N., Burlet, A.J., 1985. Comparison of<br />

three immunoassays in the screening and characterization of monoclonal antibodies against argininevasopressin.<br />

J. Neuroimmunol. 9, 205-220.<br />

[9] Smeets, W.J.A.J., González, A., 2001. Vasotocin and mesotocin in the brains of amphibians: state of the<br />

art. Microsc. Res. Tech. 54, 125–136.<br />

[10] Woolley, S.C., Sakata, J.T., Crews, D., 2004. Evolutionary insights into the regulation of courtship<br />

behavior in male amphibians and reptiles. Physiol. Behav. 83, 347–360.<br />

241


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

SYNAPTOGENESIS: PROMOTED BY CHOLESTEROL OR ESTRADIOL?<br />

Fester L. 1 , Zhou L. 1 , Bütow A. 1 , Huber C. 1 , von Lossow R. 1 , Jarry H. 2 , Rune G.M. 1 *<br />

Institute of Anatomy I: Cellular Neurobiology, University Medical Center Hamburg<br />

Eppendorf, Martinistr. 52, 20246 Hamburg, Germany; Department of Experimental<br />

Endocrinology, University of Göttingen, Robert Koch-Str. 40, 30707 Göttingen, Germany<br />

Cholesterol of glial origin has been demonstrated to promote CNS synaptogenesis (1). As<br />

neuron-derived estradiol also regulates synapse density, we questioned whether cholesterol<br />

promotes synapse formation directly or indirectly by providing elevated substrate levels for<br />

neuronal estrogen synthesis. In this study, we provide evidence that cholesterol-induced<br />

synaptogenesis results from its metabolization to estradiol. Estradiol release from the<br />

cultures into the medium was 8-fold higher, stimulated by cholesterol. Cholesterolpromoted<br />

synaptogenesis, as demonstrated by spine synapse counting and by quantitative<br />

evaluation of pre- and postsynaptic protein expression, is abolished when cholesterol and<br />

letrozole, a potent aromatase inhibitor, are simultaneously applied to hippocampal cultures.<br />

Most importantly, downregulation of synapse formation after knock-down of StAR is only<br />

rescued by estradiol but not by cholesterol.<br />

Acknowledgement: This study was supported by the DFG (Ru 436/4-1)<br />

Reference list<br />

1. Mauch et al., Science 2001 Nov 9; 294(5545):1<strong>35</strong>4-7<br />

242


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

ESTROGEN SELECTIVELY INCREASES TRYPTOPHAN HYDROXYLASE-2<br />

AND DECREASES 5HT1 B mRNA EXPRESSIONS IN DISTINCT SUBREGIONS<br />

OF RAT DORSAL RAPHE NUCLEUS: ASSOCIATION BETWEEN GENE<br />

EXPRESSION AND ANXIETY BEHAVIOR IN THE OPEN FIELD<br />

Hiroi R. * , and Neumaier J.F. o<br />

* Department of Psychology, University of Washington, USA.<br />

o Department of Psychiatry and Behavioral Sciences, Neumaier Lab, University of<br />

Washington, Box <strong>35</strong>125, Seattle, Washington, USA. Fax +1-206-341-5804<br />

e-mail: neumaier@u.washington.edu<br />

Mounting evidence suggests that estrogen has anxiolytic effects [1,6,7]. We recently found<br />

that estrogen decreased anxiety behavior in rats in the open field test and progesterone<br />

reversed this effect [4]. Ovarian steroids also regulate the serotonergic neurons in the<br />

dorsal raphe nucleus (DRN), which are implicated in the etiology of affective disorders.<br />

The effects of estrogen on serotonin synthesis, release and reuptake may affect the overall<br />

availability of serotonin in the forebrain, and in turn affect behavior. Therefore, we<br />

examined the effects of ovarian steroids on mRNA expression of a brain specific isoform<br />

of tryptophan hydroxylase (TPH2), the rate-limiting enzyme for serotonin synthesis, and<br />

on inhibitory serotonin autoreceptors, 5-HT 1A and 5-HT 1B [2,5]. In addition, we examined<br />

whether these mRNA levels in discrete subregions of DRN correlated with anxiety<br />

behavior. Ovariectomized rats were treated for two weeks with placebo, estrogen, or<br />

estrogen plus progesterone, exposed to the open field test, and subsequently processed for<br />

TPH2, 5-HT 1A , and 5-HT 1B in situ hybridization histochemistry. Estrogen significantly and<br />

selectively increased TPH2 mRNA optical density in the mid ventromedial and caudal<br />

subregions of the DRN (by 31-41%); there were no changes in median raphe nucleus.<br />

Combined estrogen and progesterone treatment had no effect on the TPH2 mRNA in any<br />

of the DRN subregions, suggesting that progesterone reversed the effects of estrogen with<br />

no further effect on gene expression. Furthermore, TPH2 mRNA in caudal DRN was<br />

associated with lower anxiety-like behavior whereas TPH2 mRNA in rostral dorsomedial<br />

DRN was associated with increased anxiety-like behavior. On the other hand, estrogen had<br />

no effect on 5-HT 1A in any of the subregions of the DRN, while selectively decreased<br />

5-HT 1B mRNA in the mid-ventromedial subregion. This decrease in 5-HT 1B mRNA was<br />

associated with higher TPH2 mRNA and with higher anxiety-like behavior. These results<br />

suggest that estrogen may increase TPH2 synthesis and reduce 5-HT 1B autoreceptor in a<br />

coordinated fashion, thereby increasing the capacity for serotonin synthesis and release in<br />

distinct forebrain regions that modulate specific components of anxiety behavior. To test<br />

this idea, we are currently working on manipulating gene expression by knockdown and<br />

overexpression to block or mimic the effects of estrogen on anxiety behavior. Thus far, we<br />

have successfully achieved selective knockdown of TPH2 mRNA in the rat midrostral<br />

DRN, as measured by immunohistochemistry and western blot assays, with preliminary<br />

behavioral results that support our hypothesis [3].<br />

Reference list<br />

[1] Frye, C.A. and Walf, A.A., Estrogen and/or progesterone administered systemically<br />

or to the amygdala can have anxiety-, fear-, and pain-reducing effects in<br />

ovariectomized rats. Behav Neurosci 118 (2004) 306-13.<br />

[2] Hiroi, R., McDevitt, R.A. and Neumaier, J.F., Estrogen selectively increases<br />

tryptophan hydroxylase-2 mRNA expression in distinct subregions of rat dorsal<br />

243


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

raphe nucleus: association between gene expression and anxiety behavior in the<br />

open field, Biol Psychiatry, 60 (2006) 288-95.<br />

[3] Hiroi, R. and Neumaier, J.F., Morpholino antisense oligonucleotide-mediated<br />

knockdown of tryptophan hydroxylase-2 in a discrete subregion of rat dorsal raphe<br />

nucleus. Abstract of the 36th Annual Meeting of the Society for Neuroscience,<br />

p199.16.<br />

[4] Hiroi, R. and Neumaier, J.F., Differential effects of ovarian steroids on anxiety<br />

versus fear as measured by open field test and fear-potentiated startle, Behav Brain<br />

Res, 166 (2006) 93-100.<br />

[5] Hiroi, R. and Neumaier, J.F., Estrogen selectively decreases 5-HT 1B mRNA in<br />

distinct subregions of rat dorsal raphe nucleus: Inverse association between gene<br />

expression and anxiety behavior in the open field. Annual Northwest Chapter<br />

Meeting of the Society for Neuroscience, 2006.<br />

[6] Walf A.A. and Frye C.A., ERbeta-selective estrogen receptor modulators produce<br />

antianxiety behavior when administered systemically to ovariectomized rats.<br />

Neuropsychopharmacology, 30 (2005) 1598-609.<br />

[7] Walf A.A. and Frye C.A., A review and update of mechanisms of estrogen in the<br />

hippocampus and amygdala for anxiety and depression behavior.<br />

Neuropsychopharmacology, 31 (2006) 1097-111.<br />

244


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

PRIP, A PHOSPHOLIPASE C-RELATED INACTIVE PROTEIN, REGULATES<br />

GABA A RECEPTOR ENDOCYTOSIS<br />

Kanematsu T. and Hirata M.<br />

Laboratory of Molecular and Cellular Biochemistry, Faculty of Dental Science, Kyushu<br />

University, 3-1-1, Maidashi, Higashi-ku, 812-8582, Fukuoka, Japan.<br />

e-mail: inositol@dent.kyushu-u.ac.jp Fax: +81-92-642-6322.<br />

Efficacy of synaptic inhibition depends on number of cell surface expressed<br />

GABA A receptors. In spite of growing number of reports, the detailed molecular<br />

mechanisms involved in regulation of the receptor number still remain unclear. Our recent<br />

studies revealed that PRIP (phospholipase C-related but catalytically inactive protein)<br />

regulates GABA A receptor signaling by analyzing PRIP knockout (KO) mouse [1]. In the<br />

present study, we studied the involvement of PRIP in the modulation of postsynaptic<br />

GABA A receptor number by brain-derived neurotrophic factor (BDNF), which rapidly<br />

down-regulates GABA A receptor surface number. The exposure to BDNF reduced the<br />

GABA-evoked inhibitory current (I GABA ) in cultured hippocampal neuron of wild type<br />

mice, whereas a little potentiation was observed in the PRIP-KO mice, corresponding to<br />

the surface expression of GABA A receptor number. As PRIP bound to beta subunits of<br />

GABA A receptor, we mapped the region in PRIP responsible for the interaction with the<br />

beta-subunits, and the peptide mimicking that region blocked the attenuation of I GABA in<br />

wild type hippocanpal neurons in response to BDNF application [2]. GABA A receptor<br />

endocytosis is mediated by clathrin/AP2 protein complex. Since clathrin/AP2 protein<br />

complex was co-immunoprecipitated with PRIP, PRIP might be involved in the<br />

clathrin/AP2-mediated GABA A receptor endocytosis [3]. These results indicate that PRIP<br />

plays an importnat role in the process of GABA A receptors endocytosis by the direct<br />

interaction with GABA A receptor beta-subunits.<br />

Reference list<br />

1. Kanematsu T., Jang I. S., Yamaguchi T., Nagahama H., Yoshimura K., Hidaka K., Matsuda M., Takeuchi<br />

H., Misumi Y., Nakayama K., Yamamoto T., Akaike N., Hirata M. and Nakayama K. (2002) Role of the<br />

PLC-related, catalytically inactive protein p130 in GABA A receptor function. EMBO J. 21, 1004-1011.<br />

2. Kanematsu T., Yasunaga A., Mizoguchi Y., Kuratani A., Kittler J. T., Jovanovic J. N., Takenaka K.,<br />

Nakayama K. I., Fukami K., Takenawa T., Moss S. J., Nabekura J. and Hirata M. (2006) Modulation of<br />

GABA A receptor phosphorylation and membrane trafficking by phospholipase C-related inactive<br />

protein/protein phosphatase 1 and 2A signaling complex underlying BDNF-dependent regulation of<br />

GABAergic inhibition. J. Biol. Chem. 281, 22180-22189.<br />

3. Kanematsu T., Fujii M., Mizokami A., Kittler J. T., Nabekura J., Moss S. J. and Hirata M. Phospholipase<br />

C-related inactive protein is implicated in the constitutive internalization of GABA A receptors mediated by<br />

clathrin and AP2 adaptor complex. J. Neurochem. In press.<br />

245


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

PROGESTERONE WITHDRAWAL SENSITIVITY IN FEMALE RATS RELATES<br />

TO DIFFERENCES IN BASELINE BEHAVIOR OF RISK TAKING AND<br />

EXPLORATION<br />

Löfgren M., Johansson I-M, Meyerson B. a and Bäckström T.<br />

Department of Clinical Science, Obstetrics and Gynecology, Umeå Neurosteroid Research<br />

Center, Building 5B, 5 th floor, Umeå University Hospital, SE-901 85 Umeå, Sweden.<br />

a Department of Neuroscience, Division of Pharmacology. Box 593, BMC SE-751 24<br />

Uppsala, Sweden. Fax: +46-90-776006<br />

Email: Magnus.Lofgren@obgyn.umu.se<br />

Background: Progesterone effects and the subsequent withdrawal properties are<br />

similar to those of GABA A receptor acting drugs. The effects of progesterone are most<br />

likely caused by allopregnanolone (3alpha-hydroxy-5alpha-pregnane-20-one). This<br />

progesterone metabolite is produced during the luteal phase of the menstrual cycle, during<br />

pregnancy and by stressful events. In women elevated levels of allopregnanolone are<br />

correlated with negative mood symptoms during the luteal phase. Interestingly normal<br />

plasma concentrations of gonadal steroids trigger PMS symptoms more readily in<br />

susceptible women. In male rats the stable baseline behavior of risk taking and exploration<br />

has been shown to influence the severity of progesterone withdrawal (PWD).<br />

Method: 32 female Wistar rats were tested in their diestrus phase in the Open Field<br />

(OF) for baseline behavior of risk taking and exploration. From the OF data the rats were<br />

divided into high and low responders (HR/LR) and further assigned to either placebo or<br />

treatment. Vaginal lavage was performed daily. Injections were given i.p. twice daily for<br />

six days, either 5 mg/kg progesterone in conjunction with 10 µg/kg 17β estradiol, or vehicle<br />

(sesame oil). Blood samples for corticosterone (CORT) analysis were collected after the<br />

behavioral tests. At WD (24h) the animals were tested in the Elevated Plus Maze (EPM).<br />

Results: The high risk taking and exploring rats showed greater aversion of the<br />

open arms in the EPM and had lower CORT levels at PWD. The low risk taking rats did<br />

not show an adverse reaction at PWD. Within the control group the time in the inner parts<br />

of the baseline OF test correlated with time spent on the open arms of the EPM WD test.<br />

The CORT concentrations in plasma collected after the tests were correlated in the control<br />

group.<br />

Conclusions: Baseline exploration and risk taking behavior measured in the OF<br />

predicted the PWD reaction in female rats, with greater sensitivity found in high<br />

responders. The stability of the plasma CORT indicated a consistent individual stress<br />

response.<br />

246


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

P450scc IS INDUCED IN NEURONAL AND GLIAL CELLS AFTER STATUS<br />

EPILEPTICUS: MODULATORY EFFECTS OF NEUROSTEROIDS ON<br />

EPILEPTOGENESIS<br />

Longo D.*, Baldelli E.*, Zini I.*, Zoli M.*, Avoli M. # , Biagini G.*<br />

*Dipartimento di Scienze Biomediche, Università di Modena e Reggio Emilia, via Campi<br />

287, 41100 Modena; # Montreal Neurological Institute, McGill University, Montreal,<br />

Quebec, Canada.<br />

P. I.: Longo Daniela, Dipartimento di Scienze Biomediche, Sezione di Fisiologia,<br />

Università di Modena e Reggio Emilia, via Campi 287, 41100 Modena; phone +39 059<br />

2055<strong>35</strong>8; fax +39 059 2055363; e-mail: danylong@virgilio.it<br />

The conversion of cholesterol into pregnenolone by the rate-limiting enzyme<br />

cholesterol side-chain cleavage cytochrome P450 (P450scc) is a critical step in the<br />

synthesis of brain-derived steroids (neurosteroids). Pregnenolone is subsequently<br />

metabolized through various enzymatic steps into main final products such as<br />

allopregnanolone and allotetrahydrodeoxycorticosterone. Steroids may play several roles<br />

in the nervous systems, as they can regulate the axonal growth, synaptogenesis and, in<br />

general, neuronal trophism. In addition, steroids display modulatory properties on<br />

glutamate and gamma-amino butyric acid (GABA) receptors. It is noteworthy that<br />

allopregnanolone and allotetrahydrodeoxycorticosterone function as GABA A<br />

receptor<br />

agonists, so enhancing GABA inhibitory effects on neuronal targets that could be critical<br />

in modulating seizure susceptibility.<br />

Although neuronal cells possess the molecular machinery to produce neurosteroids,<br />

these molecules are mainly synthesized in glial cells, particularly in oligodendrocytes and<br />

astrocytes. This last glial cell type is highly activated by neuronal damage, but it is<br />

presently unclear whether this activation leads to a enhanced neurosteroid synthesis.<br />

Temporal lobe epilepsy (TLE) associated with hippocampal sclerosis is characterized by<br />

reactive gliosis. In animal models mimicking this human disease, astrocytes show<br />

hypertrophy and, consequently, increased staining for the marker glial fibrillary acidic<br />

protein (GFAP). This phenomenon, defined as “glial reactivity”, is particularly pronounced<br />

in the early period that follows status epilepticus (SE), but no information is still available<br />

on the possible changes in neurosteroid levels after SE or whether such a modification<br />

could affect the normal course of epileptogenesis.<br />

In this work, we induced SE by injecting pilocarpine (380 mg/kg i.p.) in Sprague-<br />

Dawley rats (270-300 gr body weight). Seizures were blocked after 3 hours from the<br />

beginning of SE and subsequently the animals were sacrificed at 5 different time intervals<br />

(1, 3, 7, 21 days after SE) and analyzed with immunohistochemical procedures. To this<br />

aim, we used polyclonal antibodies against GFAP (1:500, DAKO, Glostrup, Denmark) for<br />

astroglial and against heme oxygenase-1 (HO-1, 1:500, Stressgen, Victoria, BC, Canada)<br />

for microglial cells, while oligodendrocytes were identified with a monoclonal antibody<br />

against human 2',3'-cyclic nucleotide-3'-phosphodiesterase (CNPase, 1:100, Sigma-<br />

Aldrich, Milan, Italy). A polyclonal anti-P450scc antibody (1:200, Chemicon, Tamecula,<br />

CA, USA) was used to evaluate the putative source of neurosteroid production. Neuronal<br />

cell bodies were identified with a monoclonal antibody against NeuN (1:100; Chemicon).<br />

A triple immunolabelling was obtained by incubating first the neuronal marker along with<br />

an antibody for glial cells, then by developing the reaction for P450scc.<br />

In another set of experiments, the animals were treated with daily s.c. injections of<br />

100 mg/kg finasteride (Ivy Chiral Chemicals, NJ, USA) or with 30% hydroxypropyl-β-<br />

247


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

cyclodextrin in water (vehicle-treated, n = 3 for NECs and n = 13 for the pilocarpine<br />

group), starting 3 days after SE and continuing for 18 days, in order to block neurosteroid<br />

synthesis and to evaluate the latency period prior to the onset of spontaneous recurrent<br />

seizures. The animals were videorecorded 6 h/day 7 days/week to monitor time of<br />

appearance and frequency of spontaneous seizures. Only stage 5 (generalized tonic-clonic<br />

convulsions) seizures were scored. Results were analyzed with one-way analysis of<br />

variance followed by the Games–Howell test for multiple comparisons. The Kaplan–Meier<br />

method was used to estimate the rate of onset of stage 5 spontaneous seizures after SE.<br />

These curves were compared by the log rank test.<br />

We found that P450scc is upregulated in several areas of the hippocampal<br />

formation (CA1, CA3, dentate gyrus, subiculum, entorhinal cortex) as well as in<br />

extrahippocampal areas (amygdala, neocortex) few days after SE. In particular, we<br />

observed a remarkable increase of P450scc in the stratum lacunosum-moleculare as well as<br />

in the pyramidal cell layer of the CA3 hippocampal subfield. The triple immunolabelling<br />

evidenced that most P450scc-positive elements co-stained with the GFAP antibody. In<br />

addition, anti-CNPase and anti-P450scc co-stained cells were also identified. Interestingly,<br />

we found also that putative microglial cells stained with the anti-HO-1 antibody were<br />

positive to P450scc: some of them were clearly localized in the vessel wall of markedly<br />

dilated blood vessels. The CA3 pyramidal cell layer was also characterized by doublelabelled<br />

positive neurons.<br />

Then, we analyzed the intensity of immunolabelling with semiquantitative<br />

microdensitometric methods as function of the different time intervals considered. In<br />

particular, we observed that positive cell counts as well as the intensity of the<br />

immunostaining increased progressively from day 1 to 3, both in the pyramidal and<br />

lacunosum-molecular layer of CA3. However, in the pyramidal layer both values<br />

decreased from day 7 reaching the basal levels at day 14, while in the stratum lacunosummoleculare<br />

cell counts and the intensity of immunostaining were still significantly<br />

(p


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

ALTERATIONS OF NEONATAL LEVELS OF ALLOPREGNANOLONE AND<br />

THE NOVELTY-DIRECTED BEHAVIOURAL RESPONSE TO<br />

INTRAHIPPOCAMPAL ADMINISTRATION OF ALLOPREGNANOLONE IN<br />

ADULTHOOD<br />

Martín-García E., Darbra S., Pallarés M.<br />

Departament de Psicobiologia i Metodologia en Ciències de la Salut,<br />

Institut de Neurociències, Universitat Autònoma de Barcelona, 08193<br />

Bellaterra, Barcelona, Spain. Fax: +34 93 581 20 01.<br />

E-mail: marc.pallares@uab.es.<br />

Recent findings indicate that neurosteroids could act as important keys during the brain<br />

development. Neonatal allopregnanolone (AlloP) administration produces behavioural<br />

changes observable into adulthood. Moreover, fluctuations in neonatal AlloP could result<br />

in altered pharmacological properties of the GABA A receptor system in adulthood. The<br />

aim of the present work is to screen whether developmentally altered neurosteroid levels<br />

influence the behavioural response to intrahippocampal administration of AlloP, a GABA A<br />

positive modulating neurosteroid, in adulthood. For this purpose, pups received AlloP (10<br />

mg/kg, s.c.) or the 5alpha-reductase inhibitor (finasteride, 50 mg/kg, s.c.) or vehicle since<br />

the fifth to the tenth postnatal day. At maturity (i.e. 90 old-days) a bilateral cannulae was<br />

implanted into the hippocampus (AP, -3.6 mm; L, ±1.8 mm; V, 2.8 mm). After recovery<br />

from surgery, animals received an administration of AlloP (0.2 µg) or vehicle 5 min before<br />

they were tested in the open field test, a paradigm of novelty-directed exploration and<br />

neophobia. The evaluation of habituation in a new environment, a primitive form of nonassociative<br />

learning, was also evaluated. Results showed that the habituation of activity in<br />

an open field test in adulthood was affected by alterations of neonatal levels of AlloP.<br />

Furthermore, animals that received perinatal administration of finasteride showed<br />

anxiolityc-like behaviour in adulthood. Thus, fluctuations in neonatal AlloP affect the<br />

novelty-directed behaviour. This effect seems to be mediated by alterations of the mature<br />

functions of the hippocampus, possibly via the GABA A receptor.<br />

249


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

HEIGHTENED SEIZURE SUSCEPTIBILITY FOLLOWING THE<br />

ADMINISTRATION OF HUMAN CHORIONIC GONADOTROPIN<br />

Milani P., Ginanneschi F., Biasella A., Bonifazi* M., Rossi A., Mazzocchio R.<br />

Department of Neurological and Behavioral Sciences, Section of Clinical<br />

Neurophysiology, University of Siena, Italy.<br />

Fax +39057740327 e-mail: dr.milani@yahoo.com<br />

*Department of Physiology, University of Siena, Siena, Italy<br />

It is known that the intramuscular injection of hCG lowers the threshold for motor<br />

evoked responses (MEPs) in the first dorsal interosseous (FDI) muscle to transcranial<br />

magnetic stimulation (TMS) in humans [1]. We describe the case of a patient with a<br />

clinically silent left-sided nasofrontal dermoid cyst who, while being treated with hCG/LH<br />

for hypogonadotropic hypogonadism, presented with simple partial seizures, ipsilateral to<br />

the cyst, with secondary generalization. Motor cortex excitability was studied by single and<br />

paired TMS and MEPs were recorded from FDI. Resting motor threshold (RMT), active<br />

motor threshold (AMT), MEP size, intracortical inhibition (ICI) and intracortical<br />

facilitation (ICF) were tested during and after suspension of hormonal therapy. RMT and<br />

AMT were lower, MEP size was larger, ICI was decreased while ICF was unchanged<br />

during treatment. This indicated an increased intracortical excitability during hormonal<br />

therapy. It is concluded that treatment with hCG/LH may favour seizure onset in the<br />

presence of potentially epileptogenic lesions such as an intracranial dermoid cyst.<br />

Reference list<br />

1. Bonifazi M, Ginanneschi F, Della Volpe R, Rossi A. Effects of gonadal steroids on the input-output<br />

relationship of the corticospinal pathway in humans. Brain Res 2004;1011:187-94.<br />

250


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

ROLES OF PRIP IN TRAFFICKING OF GAMMA2 SUBUNIT CONTAINING<br />

GABA A RECEPTOR<br />

Mizokami A., Kanematsu T. and Hirata M.<br />

Laboratory of Molecular and Cellular Biochemistry, Faculty of Dental Science, Kyushu<br />

University, 3-1-1, Maidashi, Higashi-ku, 812-8582, Fukuoka, Japan.<br />

e-mail: akiko-k @dent.kyushu-u.ac.jp Fax: +81-92-642-6322.<br />

GABA A receptors are a family of ligand-gated ion channels that are pentamer<br />

composed predominantly of alpha, beta, and gamma subunits. They are the major target of<br />

the endogenous inhibitory neurotransmitter (GABA) and have been implicated in a variety<br />

of brain functions including sedation, hypnosis, anxiety, learning and memory. The<br />

heterologous subunit composition of the GABA A receptor is known to be associated with<br />

the distinct pharmacological and physiological properties. In the present study, we have<br />

elucidated that PRIP (phospholipase C-related, but catalytically inactive protein) regulates<br />

the GABA signaling via the receptors by analyzing PRIP knockout (KO) mice; the<br />

sensitivity to diazepam was reduced as assessed by biochemical, electrophysiological and<br />

behavioral analyses of PRIP KO mice, suggesting the dysfunction of the gamma2 subunitcontaining<br />

GABA A receptors, a target of diazepam, a typical benzodiazepine type drug.<br />

We then examined the mechanisms by which PRIP molecule regulates cell-surface<br />

expression of gamma2 subunit-containing GABA A receptor. Disruption of the direct<br />

interaction between PRIP and the beta subunit of GABA A receptors by PRIP-binding<br />

peptide inhibited cell-surface expression of gamma2 subunit-containing GABA A receptors,<br />

while the expression of alpha and beta subunits were not altered by the peptide in GH3 and<br />

HEK293 cells. Collectively, PRIP molecules are involved in trafficking of gamma2<br />

subunit-containing GABA A receptors to cell-surface membrane, probably by facilitating<br />

the function of GABARAP, GABA A receptor associated protein.<br />

251


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

PRENATAL ESTROGENS AND THE DEVELOPMENT OF MEMORY AND<br />

LEARNING<br />

Nasir RH 1 , Chen C 2 , Bellinger D 1,3,4 , Korrick SA 2,3,4 .<br />

1 Children’s Hospital, Boston, 2 Channing Laboratory, Brigham and Women’s Hospital, 3 Harvard Medical<br />

School, 4 Harvard School of Public Health.<br />

Developmental Medicine Center, Fegan 10, Children’s Hospital Boston,300 Longwood Ave, Boston MA<br />

02115, USA. Ramzi.nasir@childrens.harvard.edu. Fax: 1-617-738-0252.<br />

Background: In animal studies experimental alterations of estrogen levels early in<br />

development result in alteration of memory and learning patterns later in life [5]. It is not<br />

known what role prenatal estrogens play in the development of memory systems in humans<br />

or if variability in estrogen levels between individuals during fetal development contributes<br />

to differences in memory function later in life.<br />

Objective: Evaluate the relationship between umbilical cord serum estradiol (E2) level and<br />

memory and learning at age eight years.<br />

Design/Methods: This study utilizes data from an ongoing birth cohort study examining<br />

relations between in utero exposures to environmental pollutants and neurodevelopmental<br />

outcomes among 788 children residing near a polychlorinated biphenyl (PCB)<br />

contaminated Superfund site in New Bedford, MA. We examined data from 326 children<br />

enrolled in this parent study for whom data are available on umbilical cord serum E2 level<br />

and 8-year neurodevelopmental and behavioral assessments. Memory and learning were<br />

assessed using the Wide Range Assessment of Memory and Learning (WRAML) which is<br />

a standardized instrument used clinically to evaluate three distinct memory functions:<br />

visual memory, verbal memory, and learning. Scores are standardized to a mean (standard<br />

deviation) of 100 (15). In this preliminary analysis, non linearities were noted in the<br />

relationship between E2 and WRAML indices. Therefore, nonparametric smoothing was<br />

used to describe the dose-response relationship of E2 with each WRAML index (SAS<br />

PROC GAM)[1]. Models were adjusted for age at exam and the examiner who<br />

administered the test. Potential sex specific differences in the impact of E2 on memory and<br />

learning were assessed by stratifying by sex.<br />

Results: Selected population characteristics are presented in Table 1. Children were<br />

generally full term and healthy at birth. Mean E2 levels were higher in male infants than in<br />

females as has been described in other populations [4]<br />

Table 1: Selected population characteristics (n=326)<br />

Characteristic Mean (sd) Characteristic N(%)<br />

Estradiol (pg/ml) - Girls 162 (50)<br />

Male 8174.7 (5457) Low income 90 (32)<br />

household<br />

Female 6879.5 (4849.9) Child breastfed 107 (39)<br />

Birth Weight (g) 3383 (4<strong>35</strong>.9) Maternal smoking 93 (30)<br />

during pregnancy<br />

Gestational age 39.7 (1.3) Married parents 182 (62)<br />

(weeks)<br />

Age at exam (years) 7.9 (0.4) White mother 238 (81)<br />

The mean WRAML score (standard deviation) for visual memory was 90 (12), verbal<br />

memory 88 (13), and learning index 97 (13). Thus, values were slightly lower than the<br />

standardization population. Nonparametric smooths describing the relation between E2<br />

and WRAML verbal memory and learning indices are shown in Figure 1. In this figure<br />

there is a suggestion of a biphasic dose response: at moderate cord serum E2 levels,<br />

252


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

increasing E2 levels are generally associated with improved verbal memory and learning<br />

whereas at higher cord serum E2 levels, the inverse relationship holds. However, the<br />

paucity of data at higher E2 levels makes the latter association less certain. Additionally,<br />

the impact of E2 on WRAML indices appears to be different among boys and girls,<br />

particularly at low E2 levels where girls have better, and boys poorer, performance with<br />

increasing cord serum E2 levels.<br />

Figure 1<br />

Smoothing plot of the adjusted relationship between cord serum Estradiol and WRAML<br />

Verbal Memory Index (right), and Learning Index (left), adjusted for age at exam and<br />

examiner.<br />

Discussion/Conclusion: The key findings in this preliminary analysis are that prenatal<br />

estrogen exposures (measured as cord serum E2 levels) may impact subsequent memory<br />

and learning skills at age 8, and that this impact potentially differs by both E2 level and<br />

sex. Our observed non-linear dose-response relationship may be related to complex<br />

interactions with other environmental or biological covariates, including other sex steroids.<br />

The potential differential impact of E2 on memory and learning in boys compared with<br />

girls is intriguing but consistent with previously published evidence of sexual dimorphism<br />

of the brain and differential impact of neurosteroids according to sex [2,3]. The next steps<br />

in the analysis include further model development by adjustment for a wide range of<br />

covariates and potential confounders, using the GAM model findings to develop possible<br />

parametric representations of the dose-response relationship, assess the possible sex<br />

difference in E2 effects, and fuller assessment for interaction including hypothesized<br />

interactions between E2 levels and other environmental exposures with likely estrogenic<br />

activity such as PCBs.<br />

References list<br />

[1] SAS/STAT user’s guide. Version 9.1., SAS Institute, Inc Cary, NC, 2002-2003.<br />

[2] C.N. Jacklin, K.T. Wilcox and E.E. Maccoby, Neonatal sex-steroid hormones and cognitive abilities<br />

at six years, Dev Psychobiol 21 (1988) 567-574.<br />

[3] T.J. Shors and G. Miesegaes, Testosterone in utero and at birth dictates how stressful experience<br />

will affect learning in adulthood, Proc Natl Acad Sci U S A 99 (2002) 13955-13960.<br />

[4] R. Troisi, N. Potischman, J.M. Roberts, G. Harger, N. Markovic, B. Cole, D. Lykins, P. Siiteri and<br />

R.N. Hoover, Correlation of serum hormone concentrations in maternal and umbilical cord samples,<br />

Cancer Epidemiol Biomarkers Prev 12 (2003) 452-456.<br />

[5] C.L. Williams, A.M. Barnett and W.H. Meck, Organizational effects of early gonadal secretions on<br />

sexual differentiation in spatial memory, Behav Neurosci 104 (1990) 84-97.<br />

253


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

ASSESSMENT OF INTERACTION BETWEEN SEX HORMONES AND<br />

INCIDENCE OF EPILEPSY CRISIS IN FEMALE<br />

Nobahar M, Vafaei AA<br />

Faculty of Nursing and Paramedical, Semnan University of Medical Sciences, Semnan, Iran<br />

E-mail: Nobahar43@yahoo.com<br />

INTRODUCTION:<br />

Epilepsy is one of the disorders with chronic, recurrent and sudden changes in neurological function<br />

due to an electrical abnormality of cerebrum [1]. Previous studies have shown that around forty five<br />

million through out the world are suffering from epilepsy and is estimated that between 0.5 and 2<br />

percent of the population could acquire this disorder at any age [2]. Also the steroid hormones<br />

estradiol and progesterone not only regulate the reproductive system but have other central nervous<br />

system effects that can directly affect a variety of behaviors. Generally, estradiol has been shown to<br />

have activating effects, including the ability to increase seizure activity, while progesterone has<br />

been shown to have depressant effects, including anticonvulsant properties. Because levels of these<br />

hormones fluctuate across the menstrual cycle, it is important to understand how changes in these<br />

hormone levels may influence levels of excitability in the brain, especially in women who have<br />

seizure patterns that are related to their menstrual cycle, a phenomenon known as catamenial<br />

epilepsy. Seizures are generally random events and thus most women with epilepsy will have had a<br />

seizure near their menstrual cycle at some point in time [3]. Ovarian steroid hormones alter<br />

excitability of neurons of the central nervous system. Estrogen reduces inhibition at the GABA A<br />

receptor, enhances excitation at the glutamate receptor, and increases the number of excitatory<br />

neuronal synapses. Progesterone enhances GABA mediated inhibition, increases GABA synthesis,<br />

and increases the number of GABA A receptors. In animal models of epilepsy, estrogen increases and<br />

progesterone decreases the likelihood that a seizure will occur. Women with epilepsy may<br />

experience changes in seizures at puberty, during the menstrual cycle. These seizure patterns are<br />

believed to be associated with changes in estrogen and progesterone levels. Seizure control may also<br />

change during perimenopause because of fluctuations in estrogen and progesterone. In female<br />

patients with epilepsy manifestation of complex partial seizures and generalized tonic-clonic<br />

seizures may be influenced by sexual steroid hormones and progesterone. The term catamential<br />

seizure refers to a seizure manifestation in relation to the menstrual cycle during the few days before<br />

menstruation the first days of menstruation and near the middle of the cycle before ovulation [4].<br />

The aim of this study was evaluation of interaction between of changes of sex hormone level and<br />

incidence of epileptically crisis in female patients.<br />

METHODS: This study has been done as a clinical trial study that during one year we investigate<br />

of all female that conflicted of epilepsy. At the first time we record of demographic data include of<br />

age, sex and so on. Then we collected data regarding of situation of epilepsy crisis especially during<br />

of menstrual cycle by interview and questioner.<br />

RESULTS: The results indicated that mean of age was 17 years old, 27% of them had family<br />

history and also there is significantly correlation between of sex hormone and incidence of epilepsy<br />

crisis (P


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

VASOTOCIN/ISOTOCIN NEURONS ARE DECREASED AFTER SPAWNING IN<br />

THE FEMALE MEDAKA FISH (ORYZIAS LATIPES) BRAIN: LOCALIZATION<br />

OF AROMATASE AND ESTROGEN RECEPTOR HOMOLOGUE<br />

Ohya T., Kodama M., and Hayashi S.<br />

Laboratory of Endocrinology, Graduate School of Integrated Science, Yokohama City<br />

University, 22-2 Seto, Kanazawa-ku, Yokohama, 236-0027, Japan. Fax +81-45-787-2380<br />

e-mail: tamaki037@ybb.ne.jp.<br />

In teleosts, the distribution of neurons in the preoptic-hypothalamic region and their<br />

associated neurohypophysial hormones, such as vasotocin (VT), appears to be different<br />

among species. This differential distribution is thought to reflect the social and/or sexual<br />

status of individuals within a species. In the previous study, we analyzed the number, size<br />

and the distribution of vasotocin/isotocin (VT/IT) neurons in the brains of both male and<br />

female medaka using immunohistochemistry. VT/IT neurons were similarly located in an<br />

inverted L-shape in the nucleus preopticus (NPO) in both sexes, as has been already<br />

reported in salmonids [1]. However, computer-assisted image analysis revealed sexual<br />

dimorphism in the number of VT/IT-immunoreactive (ir) neurons, with greater numbers<br />

found in males as compared to females. Further, in the female brain, the number of VT/ITir<br />

neurons decreased significantly after spawning. In pre-spawning compared to postspawning<br />

females, the small-sized VT/IT-ir neurons dominated [2]. Sexual differentiation<br />

of the medaka is fully dependent upon the steroid status during the early developmental<br />

stages and steroids are also known to trigger sex-specific behavior in the adult medaka [3].<br />

In addition, we have already reported that aromatase-ir cells ware localized in the nucleus<br />

preopticus parvocelluralis posterioris [4], where the VT/IT-ir fibers which have originated<br />

from the magnocellulalis of the NPO extended facing to the third ventricle [2].<br />

Furthermore, mRNA of estrogen receptor homologue seems to be localized in the preoptic<br />

area of the medaka brain [5]. Although, roles of the activational effects and/or<br />

organizational effects of the steroids via aromatase and estrogen receptor to the VT/IT<br />

neurons is not clarified yet, our findings strongly suggest that VT and/or IT neurons may<br />

be functionally related to ovulation and/or the reproductive axes through connections to<br />

their steroidal status in the medaka.<br />

Reference list<br />

1. Ohya, T., Ando, H., Ueda, H., Urano, A., 1998. Subnuclei of the nucleus preopticus in sockeye salmon:<br />

presence of sexual dimorphism. Prog. Jpn. Soc. Comp. Endocrinol. 13, 16<br />

2. Ohya, T., Hayashi, S., 2006. Vasotocin/Isotocin-immunoreactive neurons in the Medaka fish brain are<br />

sexually dimorphic and their numbers decrease after spawning in the female. Zool. Sci. 23, 23-29<br />

3. Yamamoto, T., 1959 The effects of estrone dosage level upon the percentage of sex-reversals in genetic<br />

male (XY) of the medaka (Oryzias latipes). J. Exp. Zool. 141, 133-153<br />

4. Hayashi, S., Kodama, M., 2005 Two kinds of immunohistochemical distribution of aromatase in the<br />

medaka brain. Neurosci. Res. 52, 205.<br />

5. Kawahara, T., Okada, H., Yamashita, I., 2000. Cloning and expression of genomic and complementary<br />

DNAs encoding an estrogen receptor in the medaka fish, Oryzias latipes. Zool. Sci. 17, 643-649<br />

255


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

GNRH-ASTROCYTES INTERACTIONS INVOLVED IN GNRH<br />

NEUROSECRETION: ROLE OF PSA-NCAM THROUGH CHANGING<br />

ACTIVITY AND EXPRESSION LEVELS OF POLSIALYLTRASFERASE<br />

Parkash J. and Kaur G.<br />

Department of Biotechnology, Guru Nanak Dev University, Amritsar, Punjab, India.<br />

Abstract: In recent years compelling evidence has been provided that cell–cell interactions<br />

involving non-neuronal cells, such as glial and endothelial cells, are important in<br />

regulating the secretion of GnRH, the neuro-peptide that control the sexual development<br />

and adult reproductive function. We have shown previously that morphological changes<br />

occur in the external zone of the median eminence (ME) allowing certain GnRH nerve<br />

terminals to contact the perivascular space on the day of proestrous. Using dual label<br />

immunofluorescence in conjunction with confocal microscopy, we determined that<br />

terminals and perikarya of GnRH neurons in adult cycling female rats are intimately<br />

associated with polysialylated form of neural cell adhesion molecule (PSA-NCAM). In the<br />

preoptic area (POA) the intense PSA-NCAM immunoreactivity was evident around the<br />

periphery of GnRH cell bodies. This morphological remodeling includes a reduction in<br />

astrocytic coverage of GnRH neurons during proestrous phase of cycling rats. Due to this<br />

neuronal glial interaction GnRH neurons and glial cells continue to express PSA-NCAM in<br />

cyclic fashion indicating that PSA plays important role in the neurosecretory activity of the<br />

hypothalamus in adult brain. Our data addressed new communication pathway between<br />

glia cells and GnRH neurons in the POA as well as ME regions of the hypothalamus in the<br />

central of GnRH release. The second goal of study was to determine the functional<br />

significance of PSA-NCAM molecule to both structural remodeling and neurosecretory<br />

activity of hypothalamus in adult brain, we have studied the expression of PSA-NCAM on<br />

GnRH axon terminals and on the glial cells in the POA as well as ME-ARC regions of<br />

hypothalamus in the proestrous and diestrous phase of cycling rats by using confocal<br />

microscope. Both GnRH and PSA-NCAM immunostaining was much higher in the POA<br />

as well as in the ME regions from proestrous phase rats, whereas, in diestrous phase rats<br />

their expression significantly reduced. The co-localization of PSA-NCAM was studied<br />

with GFAP in the POA as well as in the ME-ARC regions of hypothalamus using dual<br />

immunohistofluorescent staining. Taken together, our observations add to the growing<br />

evidence that PSA-NCAM play permissive role for neuronal-glial remodeling and further<br />

suggest a functional role of PSA-NCAM in the GnRH release mechanism.<br />

256


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

IS THERE A LINK BETWEEN THE HYPOTHALAMO-PITUITARY-GONAD AXIS<br />

AND THE HIPPOCAMPUS?<br />

Prange-Kiel J 1 , Jarry H 2 , Kohlmann P 1 , Schön M 1 , Lohse C 1 , Rune GM 1<br />

1 Institute of Anatomy I: Cellular Neurobiology, Martinistrasse 52, University of Hamburg,<br />

20246 Hamburg, Germany; prange-kiel@uke.uni-hamburg.de, fax: ++49-40-42803-4966<br />

2 Experimental Endocrinology, University of Göttingen, Göttingen, Germany<br />

Spine density in the hippocampus varies with fluctuating estradiol serum<br />

concentrations during the estrous cycle in rats, although application of exogenous estradiol to<br />

hippocampal slice cultures does not increase spine number. Since we recently demonstrated<br />

that hippocampal neurons of the rat synthesize estradiol (E 2 ) de novo, which in turn regulates<br />

spine density, we are now looking for potential regulators of estrogen synthesis in the<br />

hippocampus.<br />

First, we showed by unbiased electron microscopic stereological calculation that cyclic<br />

changes in spine synapse density occur specifically in the hippocampus, but not in the neocortex.<br />

In accordance, we demonstrated by real-time RT-PCR gonadotrophin-releasing<br />

hormone receptor (GnRH-R)-mRNA expression to be specific in the hippocampus.<br />

Immunohistochemistry showed that a subpopulation of hippocampal cells expresses<br />

gonadotrophin-releasing hormone receptor (GnRH-R). This expression is differentially<br />

regulated by E 2 , as shown by treatment of hippocampal dispersion cultures with E 2 and the<br />

aromatase-inhibitor letrozole, immunohistochemistry and subsequent image analysis. GnRH<br />

treatment of hippocampal slice and dispersion cultures resulted in a dose-dependent regulation<br />

of hippocampal E 2 synthesis (measured by RIA) and synaptic proteins. GnRH effects could be<br />

abolished by using the GnRH antagonist antide. If GnRH-induced estrogen synthesis was<br />

suppressed with the specific aromatase inhibitor letrozole, the GnRH effect on synaptic<br />

plasticity was also abolished.<br />

We conclude that GnRH specifically mediates cyclic changes in hippocampal E 2 synthesis and<br />

via this mechanism synaptic plasticity. Therefore, GnRH may link the hypothalamo-pituitarygonad<br />

axis to the hippocampus.<br />

257


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

RAPID ACTIONS OF ESTROGEN ON ADULT GnRH NEURONS<br />

Romanò N.*, Jasoni C.L.* and Herbison A.E.*<br />

*Department of Physiology, University of Otago, New Zealand<br />

Gonadotropin-releasing hormone (GnRH) neurons are the principal regulators of<br />

reproductive function and fertility. Estrogen feedback on the GnRH neuronal network is<br />

essential for its correct functioning and regulation. In addition to the possibility of<br />

classical genomic actions of estrogen, recent studies have suggested that rapid estrogen<br />

actions also occur at the GnRH neurons [1,2]. To evaluate this, we have been<br />

undertaking experiments using a new transgenic mouse line, in which the geneticallyencoded<br />

calcium indicator ratiometric-pericam is expressed in GnRH neurons. We show<br />

that 17-β estradiol can rapidly influence intracellular calcium levels in GnRH neurons<br />

from adult female mice by either promoting or suppressing calcium transients in a dosedependent<br />

rapid manner. Whereas the inhibitory effect is reproducible using the<br />

membrane-insoluble estradiol conjugate E 2 -6-BSA, the stimulatory effect is not. This<br />

suggests the presence of at least two distinct pathways of rapid estrogen actions, one of<br />

which is mediated by a membrane receptor. As GnRH neurons have been reported to<br />

express estrogen receptor β (ER-β) [3], we repeated the experiments on a GnRHpericam/ER-β<br />

knock-out mouse line. Both the stimulatory and the inhibitory responses<br />

to estrogen were seen on GnRH neurons from these animals, suggesting that ER-β is not<br />

essential for the rapid effects on calcium levels. Similar results were obtained in the<br />

presence of TTX, suggesting a direct effect of estrogen on GnRH neurons. These<br />

studies identify multiple pathways and mechanisms of rapid estrogen action upon the<br />

GnRH neuronal phenotype.<br />

Reference list<br />

1. Temple JL, Laing E, Sunder A, Wray S – J Neurosci. 2004 24:6326-33<br />

2. Abraham IM, Han SK, Todman MG, Korach KS, Herbison AE – J Neurosci. 2003 23:5771-7<br />

3. Herbison AE, Pape JR – Front. Neuroendocrinol. 2001 22:292-308<br />

258


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

ASSESSMENT OF NEUROACTIVE STEROID LEVELS IN PLASMA AND NERVOUS<br />

SISTEM BY LIQUID CHROMATOGRAPHY-MASS SPECTROMETRY<br />

Scurati S.# 1 , Maschi O. 1 , Crotti S. 1 , De Angelis L. 1 , Melcangi R.C. 2 and Caruso D. 1<br />

1 Dept. of Pharmacological Sciences and 2 Dept. of Endocrinology, University of Milano, Milano, Italy.<br />

#Presenting author: Dept. of Pharmacological Sciences, University of Milano, via Balzaretti 9, 20133,<br />

Milano Italy, samuele.scurati@unimi.it<br />

Several methods for the quantification of steroids in plasma and tissue have been so far<br />

developed. However, even if these methods have allowed significant progress in endocrine<br />

research, they have manifest limits for analysis in the nervous system, where low levels of several<br />

steroids in small amount of tissue samples are present. For instance, immunometric techniques<br />

present low specificity due to cross reactivity [1], while gaschromatography connected to mass<br />

spectrometry, which has been used for its high specificity, requires a very critical derivatization<br />

steps to enhance sensitivity, but only few steroids in a single analysis are detected. Liquid<br />

chromatography coupled with mass spectrometry (LC-MS) might be considered as a good<br />

alternative. Indeed, this analytical technique shows high specificity and since generally it does not<br />

require any derivatization step, permits to detect different steroids in a single analysis.<br />

On this basis, the aim of our study was to develop a LC-tandem mass spectrometry (LC-MS/MS)<br />

method for the simultaneous quantification of the main neuroactive steroids present in plasma and<br />

in different nervous tissues, namely cerebral cortex, cerebellum, lombar portion of spinal cord,<br />

sciatic and brachial nerves. Plasma (0.5ml) and nervous tissues (100mg) were obtained from 5-<br />

month old male Sprague Dawley rats, extracted and purified according to Vallée et al. with minor<br />

modifications [1]. LC-MS/MS was performed with a linear ion trap-mass spectrometer (LTQ-<br />

ThermoElectron Co) in positive MS/MS mode using an atmospheric pressure chemical ionization<br />

(APCI) source. Each steroid was identified on the basis of the mass spectra of reference the<br />

compounds and the quantitative analysis was obtained by means of calibration curves using three<br />

deuterium labeled internal standards (D4-E, D4-PREG, D9-PROG). For each steroid we obtained a<br />

good linearity, as shown by linear regression coefficient (R 2 ≥ 0.99), and good reproducibility, as<br />

demonstarted by the analysis of five different calibration curves. This LC-MS/MS method allows,<br />

in the same analysis, the detection and quantification of eight different neuroactive steroids:<br />

pregnenolone (PREG), progesterone (PROG) and its derivatives, dihydroprogesterone (DHP) and<br />

tetrahydroprogesterone (THP), testosterone (T) and its derivative, dihydrotestosterone (DHT) and<br />

5alpha-androstan-3alpha, 17beta-diol (3alpha-diol), and 17beta-estradiol (17beta-E 2 ).<br />

Neuroactive steroid levels were assessed as pg/µl in plasma and as pg/mg in nervous structures.<br />

The results have indicated that PREG, the first steroid formed by cholesterol and precursor for the<br />

further steroids, is present not only in plasma but also in all the nervous tissues here considered.<br />

The same pattern is evident in case of PROG and T. However, some differences occur in case of<br />

their metabolites. Indeed, in case of PROG metabolites (i.e., DHP and THP), DHP is detected in<br />

plasma and in nervous structure, such as cerebellum and brachial nerve, but it is under the detection<br />

limit in cerebral cortex, spinal cord and sciatic nerve. In addition, the metabolite of DHP, i.e. THP,<br />

is present in cerebellum, spinal cord, sciatic and brachial nerves but it is under the detection limit in<br />

plasma and cerebral cortex. In case of T metabolites (DHT and 3alpha-diol), DHT is present in all<br />

nervous structures here considered, but under the detection limit in plasma while its metabolite, i.e.<br />

3alpha-diol, is detected both in plasma and nervous structures. Finally, 17beta-E 2 is identified in<br />

plasma and in all nervous structures tested with the exception of cerebral cortex and spinal cord.<br />

In conclusion, we here demonstrate that LC-MS/MS method allows the assessment of<br />

neuroactive steroids in plasma and in structures of central and peripheral nervous system with high<br />

sensitivity and specificity.<br />

References list<br />

1) Anal. Biochem. 2000; 287, 153-166<br />

259


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

ADMINISTRATION OF ALLOPREGNANOLONE DECREASES SECRETION OF<br />

GONADOTROPINS IN HEALTHY WOMEN OF FERTILE AGE<br />

Timby E., Bäckström T., Nyberg S., Wihlbäck A.-C.N., and Bixo M.<br />

Department of Clinical Science, Obstetrics and Gynecology, University Hospital of Umeå,<br />

Umeå University, S-901 85 Umeå, Sweden<br />

Fax +46-90137540, e-mail: erika.timby@vll.se<br />

OBJECTIVE<br />

Allopregnanolone is an endogenous neuroactive steroid secreted by the ovarian and<br />

adrenal glands. It is also synthesized de novo in the central nervous system.Through<br />

binding to the GABA-A receptor it enhances inhibitory neurotransmission. Altered levels<br />

of allopregnanolone have been reported in major depression and premenstrual dysphoric<br />

disorder, although results are contradictory. In animals, memory and learning are impaired<br />

by allopregnanolone.The concept of neuroactive steroids does not include endocrine action<br />

per se [1].<br />

The aim of this study was to determine any effect of allopregnanolone on part of the<br />

pituitary-ovarian axis in healthy volunteers.<br />

MATERIAL AND METHODS<br />

10 healthy women with regular menstrual cycles and without hormonal treatment were<br />

challenged with allopregnanolone administered intravenously in a cumulative dose of 0.09<br />

mg/kg. The doses were increasing over time starting with 0.015 mg/kg at 0 min and then<br />

0.03 and 0.045 mg/kg at 30 and 60 min respectively. The study was performed during the<br />

follicular phase of the menstrual cycle. All subjects had the allopregnanolone injections<br />

during the same time (morning.). Baseline levels of allopregnanolone (ALLO), folliclestimulating<br />

hormone (FSH), luteinizing hormone (LH), estradiol (E2) and progesterone<br />

(PG) were drawn from serum before the first allopregnanolone injection. Following<br />

allopregnanolone injections blood samples for ALLO, FSH, LH, E2 and PG were collected<br />

at 5,13, 21, <strong>35</strong>, 43, 61, 65, 73, 81, 95, 105, 115, 150, 330, 600, 780 min. Saccadic eye<br />

movement parameters and subjective ratings of sedation were also measured during the<br />

study day. Levels of FSH, LH, E2 and PG were measured with an Immulite system.<br />

Changes of serum levels of each hormone were analyzed by one-way ANOVA (analysis of<br />

variance) with repeated measures using time-point as within-subjects factor. The SPSS<br />

statistical package was used for the analyses. P values less than 0.05 were considered<br />

statistically significant.<br />

RESULTS<br />

Intravenously administered allopregnanolone in a cumulative dose of 0.09 mg/kg reduces<br />

SEV significantly and increases subjective report of sedation which were correlated to<br />

serum levels of allopregnanolone [2]. Exogenous allopregnanolone in a cumulative<br />

intravenous dose of 0.09 mg/kg significantly reduces serum levels of FSH<br />

(F(16,144)=2,184, p=0,008). LH was reduced significantly (F(16,144)=2,633, p=0,001).<br />

Allopregnanolone as an intravenous injection of a cumulative dose of 0.09 mg/kg does not<br />

affect the serum levels of estradiol. Serum levels of progesterone did not increase, but a<br />

significant lower level was seen at 5 min (p=0.033) and 780 min (p=0.023)<br />

(F(16,144)=6,153, p< 0,001).<br />

260


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

CONCLUSION<br />

Our results suggest that allopregnanolone affects the secretion of FSH and LH while E2<br />

and PG are not increased, thus the effect is not mediated by negative feed-back by the<br />

ovarian hormones. Our hypothesis is that this could be an action mediated by the GABA-<br />

A receptor.<br />

Reference list<br />

1. Baulieu EE. Neurosteroids: a new function in the brain. Biol Cell 1991; 71:3-10.<br />

2. Timby E, Balgård M, Nyberg S, Spigset O, Andersson A, Porankiewicz-Asplund J, Purdy R,<br />

Zhu D, Bäckström T, Sundström Poromaa I. Pharmacokinetic and behavioral effects of<br />

allopregnanolone in healthy women. Psychopharmacology (berl) 1-11, 2005.<br />

261


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

PHARMACOLOGICAL MODULATORS OF THE GLYCINERGIC SYSTEM<br />

REGULATE ALLOPREGNANOLONE BIOSYNTHESIS IN THE RAT SPINAL<br />

CORD<br />

Venard C.*, Boujedaini N. + , Belon P. + , Mensah-Nyagan A.G.* # and Patte-Mensah C.*<br />

*Institut des Neurosciences Cellulaires et Intégratives, Equipe Stéroïdes et Système Nociceptif, UMR<br />

7168/LC2 CNRS – 21, rue René Descartes – Université Louis Pasteur, 67084, Strasbourg, France. Fax: +33<br />

(0)3 88 61 33 47 # e-mail: gmensah@neurochem.u-strasbg.fr.<br />

+ Laboratoires BOIRON, Sainte-Foy-lès-Lyon, France.<br />

The neurosteroid allopregnanolone, also called 5alpha,3alpha-tetrahydroprogesterone<br />

(5a,3a-THP), controls several important neurobiological processes including anxiety, stress, pain,<br />

neuroprotection, depression and motor activity. It is well demonstrated that 5a,3a-THP regulates<br />

the nervous system activity through a potent allosteric stimulation of GABA A receptors. 5a,3a-THP<br />

is also a modulator of T-type calcium channels.<br />

The biosynthesis of 5a,3a-THP requires enzymatic activities of 5a-reductase (5a-R) and 3ahydroxysteroid<br />

oxidoreductase (3a-HSOR) which convert progesterone (PROG) into 5alphadihydroprogesterone<br />

(5a-DHP) and 5a,3a-THP, respectively. We have recently observed that the<br />

spinal cord (SC), which controls several neurophysiological mechanisms, contains active forms of<br />

5a-R and 3a-HSOR. Biochemical analyses revealed that SC slices are capable of converting<br />

[ 3 H]PROG into [ 3 H]5a-DHP and [ 3 H]5a,3a-THP indicating that the spinal tissue is an active<br />

producing site of 5a,3a-THP. Moreover, we demonstrated that, owing to its ability to potentiate the<br />

GABAergic inhibitory transmission, 5a,3a-THP synthesized in the SC modulates nociceptive<br />

mechanisms controlling pain sensitivity.<br />

Because the glycinergic system is also important for inhibitory neurotransmission, we<br />

combined pulse-chase experiments with HPLC analysis and flow scintillation characterization to<br />

investigate the effects of glycine and strychnine (a major antagonist of glycinergic receptors) on<br />

5a,3a-THP biosynthesis. Glycine at 10 -10 M strongly enhanced [ 3 H]PROG conversion into<br />

[ 3 H]5a,3a-THP in the SC. The amount of [ 3 H]5a,3a-THP newly synthesized from [ 3 H]PROG in the<br />

presence of glycine (10 -10 M) was 117% higher than the control. Strychnine (10 -5 M), which<br />

antagonizes glycinergic actions, did not modify by its own 5a,3a-THP synthesis in the SC.<br />

Strychnine is an alkaloid exhibiting structural analogy with gelsemine, the major active<br />

chemical agent of Gelsemium sempervirens (G. sempervirens) described as an anxiolytic and<br />

analgesic substance. Therefore, we have also assessed the actions of purified gelsemine and G.<br />

sempervirens on 5a,3a-THP formation in SC slices. Gelsemine (10 -10 M) and G. sempervirens (10 -<br />

10 M) were both capable of increasing the amount of [ 3 H]5a,3a-THP synthesized from [ 3 H]PROG.<br />

The stimulatory effects of gelsemine (10 -10 M) and G. sempervirens (10 -10 M) were respectively<br />

185% and 600% higher than the controls.<br />

Taken together, our results suggest that gelsemine, which exerts similar effects as glycine<br />

on neurosteroidogenesis, may be an activator of the central inhibition via the potentiation of<br />

glycinergic transmission. Since gelsemine enhances the production of 5a,3a-THP, a highly active<br />

stimulator of GABA A receptors, gelsemine may reinforce significantly inhibitory signalling in the<br />

central nervous system through simultaneous activation of both GABAergic and glycinergic<br />

systems. Consequently, anxiolytic and analgesic actions of G. sempervirens might be explained by<br />

the presence of gelsemine in this medical drug. However, as the stimulatory effect of G.<br />

sempervirens (10 -10 M) on 5a,3a-THP formation is higher than that of gelsemine (10 -10 M), the data<br />

suggest that additional ingredients present in G. sempervirens composition may potentiate the<br />

stimulatory action of gelsemine on neurosteroid biosynthesis.<br />

The work was supported by the contrat CIFRE n° 2005/659 between CNRS-ULP and Laboratories<br />

BOIRON.<br />

262


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

IT IS POSSIBLE THAT THE UNSPECIFIC STEROIDS FOR GONADOTROPIN<br />

SYSTEM MODULATE NEURONAL PULSATILITY ACTIVITY FROM POA–<br />

SCH OF THE HYPOTHALAMUS FOR REGULATING LH–RH RELASE AND<br />

OVULATION TRIGGER?<br />

Vlad A.G.<br />

Dept. of Endocrinology.Spitalul Clinic Judetean.1900- Timisoara. ROMANIA<br />

Email aurvlad@yahoo.com<br />

We describe in 1971 the pulsatile rhythmic activity of the neurons from POA of<br />

Hypothalamus also and rhythmic activity of some neurons from CA1, CA2 of<br />

Hippocampus. This fact was followed by revealance that LH-RH was pulsatile released by<br />

neuronal system from hypothalamus. We demonstrate that this pulsatile activity of the<br />

neuronal system behave as a, pacemaker, mechanism for ovulation triggering as a basic<br />

mechanism of the gonadotropin system which start during puberty by a genetic program.<br />

In same time by clinical study on the puberty troubles of the younger girls and mature<br />

woman with menstrual disorders, treatment with very small dose of some steroids can<br />

regulate these troubles. Our study was performed on a lot of girls and mature womans,<br />

2500 cases, with amenorheea, pspaniomenorheea with infertility. We assayed blood<br />

samples for LH, FSH, PROL, ESTRADIOL, PROGESTERON and TESTOSTERONE.<br />

We monitorized menstrual cycles during 1 – 3 years. We give some small quantity of<br />

chorionoc gonadotropin 500 ei weekly and small amount of Medoxyprogesteron,<br />

Orgametril and anabolic steroids Nandrolone. We do not give Estradiol or Estradiol<br />

combinations. From this kind of therapeutic strategy we was very surprised that menstrual<br />

troubles disappeared and obtain a regularized menses and establish the fertility probed by<br />

frequent pregnancy with normal baby. These effects put a very interesting problem about<br />

specificity of active site receptors from the neuronal membrane from POA – SCH neurons<br />

which behave as, pacemaker, triggering of LH- RH pulsatile release. In a previous paper<br />

recently we demonstrate that the pulsatile activity of neuronal units from POA – SCH from<br />

anterior hypothalamus start this automaticity by a genetic program and this kind of activity<br />

is modulated probably by neuroactive steroids and exogenous (adrenal and gonadal)<br />

steroids. Our data after applied unspecific steroids evoke a very interesting problem about<br />

functionality of neuronal units from gonadotropin system first is automaticity and second<br />

is possible that the functionality of this automatic mechanism can have some reflex<br />

troubles which can be regulated by modulatory effects of very different steroids.<br />

263


Posters’ Exhibition:<br />

RTI: Steroid hormones and sexually dimorphic brain circuits<br />

• Bao A-M, Swaab D. F. (Netherlands) Gender difference in age-related number of<br />

corticotropin-releasing hormone expressing neurons in the human hypothalamic<br />

paraventricular nuclears and the role of sex hormones<br />

• Cannizzaro C., Plescia F., Barrile V., Diliberto I., La Barbera M., Noto G.,<br />

Mantia G. (Italy) Neurosteroid pregs differently affects learning and memory<br />

performance by altering emotionality in a gender-related manner<br />

• Gotti S., Martini M., Pradotto M., Viglietti-Panzica C., Panzica G.C. (Italy) Sexual<br />

dimorphism and estrous cycle effects on nitrinergic system in mouse hippocampus<br />

• Maccari S, Mairesse J, Zuena AR, Morley-Fletcher S, Matteucci P, Cinque C,<br />

Catalani A, Nicoletti F, Casolini P. (Italy) Prenatal stress has long-term influence<br />

on neuroplasticity: sex differences<br />

• Mura E., Furnari P., Plumari L., Viglietti-Panzica C., Panzica G.C. (Italy) Role of<br />

apoptosis in the sexual differentiation of the bed nucleus of stria terminalis of<br />

japanese quail<br />

• Oboti L., Peretto P., Fasolo A., Panzica G.C. (Italy) The accessory olfactory bulb<br />

of the adult mouse: neurogenesis and morphological analysis in the two sexes<br />

• Pinos H, Carrillo B, Pérez-Izquierdo M, Ortega E., Collado P. (Spain)<br />

Undernurishment and food rehabilitation effects on plasma leptin levels in Wistar<br />

rat


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

GENDER DIFFERENCE IN AGE-RELATED NUMBER OF CORTICOTROPIN-<br />

RELEASING HORMONE EXPRESSING NEURONS IN THE HUMAN<br />

HYPOTHALAMIC PARAVENTRICULAR NUCLEARS AND THE ROLE OF SEX<br />

HORMONES<br />

Bao A-M*, Swaab D.F.*<br />

*Netherlands Institute for Neuroscience, Meibergdreef 47, 1105 BA Amsterdam, The<br />

Netherlands, fax: 0031-20-6961006; e-mail: a.m.bao@nin.knaw.nl<br />

Previous studies showed that the activity of the corticotropin-releasing hormone<br />

(CRH) neurons in the hypothalamic paraventricular nucleus (PVN) forms the basis of the<br />

activity of the hypothalamo-pituitary-adrenal (HPA)-axis, which is increased at times of<br />

stress and in specific neurological and psychiatric disorders [1, 2]. Changes in the total<br />

number of CRH expressing neurons in the PVN reflect the change in activity of CRH<br />

neurons [1, 3-5], which was confirmed by in situ hybridization of CRH-mRNA analyzed in<br />

the same patients [6-8]. The total number of CRH neurons in the human hypothalamic<br />

PVN increases with aging [3, 4]. To determine whether such an age-related change<br />

depends on gender and whether circulating sex hormones play a role, we analyzed the total<br />

number of CRH immunoreactive neurons by immunocytochemistry and image analysis in<br />

postmortem hypothalamic PVN of 22 control subjects (11 males and 11 females) between<br />

the ages of 22 and 89 year, and 10 subjects with abnormal sex hormone status. Our data<br />

shows that there are gender differences in the number of CRH neurons in the hypothalamic<br />

PVN, in that (i) there is a significant age-related increase of CRH neurons in men (p =<br />

0.032), but not in women (p = 0.733); and (ii) men have significantly more CRH neurons<br />

than women (p=0.004). Male subjects with low testosterone levels due to castration<br />

showed significantly fewer CRH neurons than well-matched intact males (p = 0.008),<br />

while castrated male-to-female transsexuals with estrogen replacement showed normal<br />

numbers of CRH neurons. One male case, which had high estrogen levels due to an<br />

estrogen-producing tumor, showed a large number of CRH neurons. Thus, although both<br />

circulating androgens and estrogens seem to play a stimulatory role with respect to CRH<br />

neurons, the age-dependent increase in the number of CRH neurons in the PVN of men,<br />

which has been interpreted as a sign of activation of the CRH neurons with age, seems to<br />

result from factors other than age-related changes of circulating sex hormone levels. The<br />

authors favor the idea that a number of factors seem to be involved in the changes of CRH<br />

neuron activity, alterations in circulating sex hormone levels being only one of them [9].<br />

Reference list<br />

1. Swaab DF, Bao AM, Lucassen PJ. The stress system in the human brain in depression<br />

and neurodegeneration. Ageing Res Rev 2005;4(2):141-194.<br />

2. Sapolsky RM, Finch CE. Alzheimer's disease and some speculations about the<br />

evolution of its modifiers. Ann N Y Acad Sci 2000;924:99-103.<br />

3. Raadsheer FC, Hoogendijk WJ, Stam FC, Tilders FJ, Swaab DF. Increased numbers of<br />

corticotropin-releasing hormone expressing neurons in the hypothalamic<br />

paraventricular nucleus of depressed patients. Neuroendocrinology 1994;60(4):436-<br />

444.<br />

4. Raadsheer FC, Oorschot DE, Verwer RW, Tilders FJ, Swaab DF. Age-related increase<br />

in the total number of corticotropin-releasing hormone neurons in the human<br />

265


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

paraventricular nucleus in controls and Alzheimer's disease: comparison of the<br />

disector with an unfolding method. J Comp Neurol 1994;339(3):447-457.<br />

5. Bao AM, Hestiantoro A, Van Someren EJ, Swaab DF, Zhou JN. Colocalization of<br />

corticotropin-releasing hormone and oestrogen receptor-alpha in the paraventricular<br />

nucleus of the hypothalamus in mood disorders. Brain 2005;128(Pt 6):1301-1313.<br />

6. Huitinga I, Erkut ZA, van Beurden D, Swaab DF. Impaired hypothalamus-pituitaryadrenal<br />

axis activity and more severe multiple sclerosis with hypothalamic lesions.<br />

Ann Neurol 2004;55(1):37-45.<br />

7. Goncharuk VD, Van Heerikhuize J, Swaab DF, Buijs RM. Paraventricular nucleus of<br />

the human hypothalamus in primary hypertension: activation of corticotropinreleasing<br />

hormone neurons. J Comp Neurol 2002;443(4):321-331.<br />

8. Raadsheer FC, van Heerikhuize JJ, Lucassen PJ, Hoogendijk WJ, Tilders FJ, Swaab<br />

DF. Corticotropin-releasing hormone mRNA levels in the paraventricular nucleus of<br />

patients with Alzheimer's disease and depression. Am J Psychiatry 1995;152(9):1372-<br />

1376.<br />

9. Bao AM, Fischer DF, Wu YH, Hol EM, Balesar R, Unmehopa UA, et al. A direct<br />

androgenic involvement in the expression of human corticotropin-releasing hormone.<br />

Mol Psychiatry 2006;11(6):567-576.<br />

266


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

NEUROSTEROID PREGS DIFFERENTLY AFFECTS LEARNING AND<br />

MEMORY PERFORMANCE BY ALTERING EMOTIONALITY IN A GENDER-<br />

RELATED MANNER<br />

Cannizzaro C., Plescia F., Barrile V., Diliberto I., La Barbera M., Noto G., Mantia G.<br />

Dept. of Pharmacological Sciences, Laboratory of Neuropsychopharmacology, University<br />

of Palermo Via del Vespro 129, 90127 Palermo, Italy, Fax +39 0916553212<br />

carla.cannizzaro@katamail.com<br />

Neurosteroids exert an important role as modulators of inhibitory and excitatory<br />

neurotransmission by interacting with different receptors [1] .In particular, pregnenolone<br />

sulphate (PREGS) rapidly alters neuronal excitability via non-genomic mechanism [4].<br />

This neurosteroid is a negative modulator of GABA A receptor , positive modulator of<br />

NMDA receptor [3] and agonists at sigma 1 receptor, influencing cognition as well as<br />

emotionality [2]. Indeed altered levels of neurosteroids together with alterations in<br />

acetylcholine transmission have been reported in human neurodegenerative pathologies<br />

like Alzheimer’s disease [6].<br />

In this study we investigated, in adult male and female rats, the effects of PREGS<br />

(10mg/kg s.c.) administrated 60 min before the test sessions, on: learning and memory<br />

performance, using a motivated, non-aversive, spatial and tactile/visual learning task, the<br />

“Can test”; emotionality, using the Elevated Plus Maze test (EPM). The Can Test protocol,<br />

following an habituation period, consisted in two separate parts: the baseline training (BT),<br />

i.e. four-day sessions consisting in ten trials each, and the longitudinal evaluation (LE), i.e<br />

four ten-trial sessions every two weeks, according to Popovich [5]. Rats, following their<br />

performance in the baseline training, were divided in non active (NA), and active (A)<br />

animals, according to the quality of their performance. Our results showed that in A-male<br />

rats PREGS induced an increase in the number of correct responses (p


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

Reference list<br />

1) Baulieu E.E., Robel P., Neurosteroids: a new brain function? J. Steroid Biochem. Mol. Biol. 37<br />

(1990) 395-403<br />

2) Darnaudéry M., Koehl M., Piazza P.V., Le Moal M., Mayo W., Pregnenolone sulfate increase<br />

hippocampal acetylcholine release and spatial recognition, Brain Res 852 (2000) 173-179<br />

3) Irwin R.P., Maragaskis N.J., Rogawski M.A., Purdy R.H., Farb D.H., Paul S.M., Pregnonolone<br />

sulphate augments NMDA receptor mediated increases in intracellular Ca 2+ in cultured rat<br />

hippocampal neurons, Neurosci. Lett. 141 (1992) 30-34<br />

4) Paul S.M., Purdy R.H., Neuroactive steroids, FASEB J. 6 (1992) 2311-2322<br />

5) Popoviç M, Biessels G-J, Isaacson RL, Gispen WH. Learning and memory in a streptozotocininduced<br />

diabetic rats in a novel spatial/object discrimination task. Behav Brain Res, 2001; 122: 201-<br />

207<br />

6) Vallée M., Mayo W., Darnaudéry M., Corpéchot C., Young J., Koel M., M. Moal Le, Baulieu E.E.,<br />

Robel P., Simon H., Neurosteroids: deficient cognitive performance in aged rats depends on low<br />

pregnenolone sulphate levels in the hippocampus, Proc. Natl. Acad. Sci. USA 94 (1997) 14865-<br />

14870<br />

268


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

SEXUAL DIMORPHISM AND ESTROUS CYCLE EFFECTS ON NITRINERGIC<br />

SYSTEM IN MOUSE HIPPOCAMPUS<br />

Gotti S., Martini M., Pradotto M., Viglietti-Panzica C., Panzica G.C.<br />

Laboratory of Neuroendocrinology, Dept Anatomy, Pharmacology and Forensic Medicine,<br />

C.so M. D’Azeglio, 52. 10126 Torino (Italy)<br />

e-mail: stefano.gotti@unito.it<br />

Fluctuating levels of estradiol and progesterone during the estrous cycle may<br />

induce structural changes in several brain nuclei including the hippocampus, where some<br />

neurons express estrogens receptors [1-3]. Nitric oxide plays a wide range of functions in<br />

the nervous system generally by acting as a neurotransmitter-like molecule that can<br />

contribute to both anterograde and retrograde signalling at the synapse. It has been<br />

demonstrated that long-term treatments with estradiol in ovariectomized females and with<br />

testosterone in castrated males induce neuronal nitric oxide synthase (nNOS) expression in<br />

rat hypothalamus [4-5], whereas changes in nNOS immunoreactivity or in associated<br />

NADPH-diaphorase activity were observed both in hypothalamus [6] and in the amygdala<br />

[7-8] during different phases of estrous cycle [for a review see 9]. Estradiol could induce<br />

nNOS expression in several brain regions in rodents. Therefore, to clarify if the<br />

hippocampal NO producing system is a target for gonadal hormones in physiological<br />

conditions, we have performed a comparison between two months old intact female and<br />

male mice in the expression on nNOS in the hippocampus. Moreover, we have investigated<br />

the effects of estrous cycle in the expression of nNOS immunoreactivity on two months<br />

old intact female mice.<br />

Immunoreactive cells were observed in all the hippocampal subregions: the higher number<br />

was detected in the pyramidal layer of CA1 region and in polymorph layer of dentate<br />

gyrus. Comparing proestrus female mice with male mice we observed a significantly<br />

dimorphism in the number of nNOS positive cells: female mice show a higher number of<br />

nNOS-immunoreactive elements in all hippocampal subregions. In the female, the number<br />

of nNOS positive neurons fluctuates during the estrous cycle, reaching its peak during<br />

proestrus and metaestrus, but these variations were statistically significant only in CA2<br />

region, that is a hippocampal region with fewer cells.<br />

In conclusions, our data demonstrate the presence of a sexually dimorphic population of<br />

nNOS positive neurons all over the hippocampus. This population is larger in females than<br />

in males, and the estrous cycle is playing a non-significant role for the expression of nNOS<br />

in the hippocampus. Sexually dimorphic structures or systems that are larger in females<br />

than in males are a minority within the rodent brain (i.e. locus coeruleus, AVPV) and the<br />

development of these structures is probably under the control of androgens and androgen<br />

receptors. Further studies should clarify the mechanisms that are influencing this sex<br />

dimorphism in the mouse hippocampus.<br />

Acknowledgements. This work was supported by Regione Piemone, University of Torino,<br />

Fondazione CRT, PRIN.<br />

269


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

Reference list<br />

[1] R.C. Melcangi and G.C. Panzica, Neuroactive steroids: old players in a new game,<br />

Neuroscience 138 (2006) 733-739.<br />

[2] E. Gould, C.S. Wooley, M. Frankfurt and B.S. McEwen, Gonadal steroids regulate<br />

dendritic spine density in hippocampal pyramidal cells in adulthood, Neuroscience<br />

10 (1990) 1286-1291.<br />

[3] C.S. Woolley and B.S. McEwen, Estradiol mediates fluctuation in hippocampal<br />

synapse density during the estrous cycle in the adult rat, J Neurosci 12 (1992)<br />

2549-2554.<br />

[4] S. Ceccatelli, L. Grandison, R.E. Scott, D.W. Pfaff and L.M. Kow, Estradiol<br />

regulation of nitric oxide synthase mRNAs in rat hypothalamus,<br />

Neuroendocrinology 64 (1996) <strong>35</strong>7-363.<br />

[5] J. Du and E.M. Hull, Effects of testosterone on neuronal nitric oxide synthase and<br />

tyrosine hydroxylase, Brain Research 836 (1999) 90-98.<br />

[6] M. Martini, M. Sica, C. Eva, C. Viglietti Panzica and G.C. Panzica, Dimorphism<br />

and effects of estrous cycle on the nitrinergic system in mouse hypothalamus,<br />

Hormones and Behavior 46 (2004) 95-96.<br />

[7] P. Collado, A. Guillamon, H. Pinos, M.A. Perez-Izquierdo, A. Garcıa-Falgueras, B.<br />

Carrillo, C. Rodrıguez and G.C. Panzica, NADPH-diaphorase activity increases<br />

during estrous phase in the bed nucleus of the accessory olfactory tract in the<br />

female rat, Brain Research 983 (2003) 223–229.<br />

[8] B. Carrillo, H. Pinos, G.C. Panzica, A. Guillamon and P. Collado, Nitrergic<br />

expression during estrous phase and morphologic sexual dimorphism in the medial<br />

amygdala anteroventral subdivision in rat, Hormones and Behavior 46 (2004) 85-<br />

86.<br />

[9] G.C. Panzica, C. Viglietti-Panzica, M. Sica, S. Gotti, M. Martini, H. Pinos, B.<br />

Carrillo and P. Collado, Effects of gonadal hormones on central nitric oxide<br />

producing systems, Neuroscience 138 (2006) 987-995.<br />

270


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

PRENATAL STRESS HAS LONG-TERM INFLUENCE ON<br />

NEUROPLASTICITY: SEX DIFFERENCES<br />

1,2 Maccari S, 1,2 Mairesse J, 2 Zuena AR, 1 Morley-Fletcher S, 2 Matteucci P, 2 Cinque C,<br />

2 Catalani A, 2,3 Nicoletti F, 2 Casolini P.<br />

1 Lab. of Perinatal Stress, University of Lille 1 USTL, Villeneuve d'Ascq, FR<br />

maccari@univ-lille1.fr, fax +33.320.434602 or +39.0649912524; 2 University of Rome La<br />

Sapienza, IT; 3 Neuromed, Pozzilli, IT.<br />

Prenatal Stress (PS) in rats is a well-documented model of early stress known to<br />

induce long-lasting neurobiological and behavioural alterations (i.e., altered circadian<br />

rhythms, high levels of “anxiety”, altered feedback mechanisms of the HPA axis,<br />

decreased hippocampal glucocorticoid receptors, increased hippocampal BDNF and basal<br />

Fos expression) (2,4,6). These results together suggest that PS affects the ability to cope<br />

with environmental challenges and alters neuroplasticity. In particular, we have found that<br />

PS rats prefer an active coping strategy when exposed to an anxiogenic environment,<br />

whereas they show less interest towards a more reassuring environment (7). This is<br />

reminiscent of what occurs in depression and anxiety, in which allostatic changes in<br />

neuroplasticity in limbic regions hamper the emotional response to the environment. (5).<br />

In the last two decades, however, effects of PS on neuroplasticity and on response to stress<br />

has been extensively studied, but most studies focus on male rats. One example is the<br />

effect of PS on hippocampal neurogenesis in male rats (1). Among the different factors that<br />

can be involved in the regulation of neurogenesis, metabotropic glutamate receptors<br />

(mGluR) seem to play a significant role. Indeed, we have recently shown that in male rats<br />

PS induced a decrease in neurogenesis, an attenuated activity of hippocampal group-I<br />

mGlu receptors associated with increased anxiety-like behaviors. Interestingly, in female<br />

rats, PS induced opposite long-term effects on group-I mGlu receptors and anxiety (2).<br />

Here we studied, in male and female rats, the long-lasting effects of PS on neurogenesis in<br />

two different hippocampal regions, the ventral part, mainly involved in the modulation of<br />

anxiety-like behaviours, and the dorsal part, more implicated in learning processes. In<br />

order to evidence cell survival, rats were injected with the thymidine-analog<br />

bromodeoxyuridine (BrdU 75 mg/kg i.p. twice daily for 4 days) 3 weeks before<br />

immunohistochemistry analysis. Two important results came out from this study: first, in<br />

female rats a higher percentage of cells differentiated in astroglia compared to males, even<br />

if the percentage of hippocampal neuronal differentiation was similar in the two sexes;<br />

second, PS differentially affect neurogenesis in male and female rats, in fact, in terms of<br />

cell survival, PS reduced neurogenesis in male rats both in the dorsal and in the ventral part<br />

of the hippocampus, but no effect was observed in females. These results show that the<br />

hippocampus of female rats seems to be less vulnerable to PS compared to males. This<br />

decreased vulnerability could be explained, at least in part, by the protective effect yielded<br />

by the increased activity of group-I mGlu receptors present in females, while the activity of<br />

the same receptors was decreased in the more vulnerable male rats. Furthermore, an higher<br />

percentage of astroglia in females could also play a role.<br />

271


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

References list<br />

1 Lemaire V, Koehl M, Le Moal M, Abrous DN. Prenatal stress produces learning deficits<br />

associated with an inhibition of neurogenesis in the hippocampus. Proc Natl Acad Sci U S<br />

A. 2000 Sep 26;97(20):11032-7<br />

2 Maccari S, et al. (2003). Prenatal stress and long-term consequences: implications of<br />

glucocorticoid hormones. Neurosci Biobehav Rev 27:119-127.<br />

3 Maccari S, Zuena AR, Cinque C, Morley-Fletcher S, Catalani A, Nicoletti F, Casolini P.<br />

Long term consequences of a restraint prenatal stress and hippocampal metabotropic<br />

receptors. Program No. 662.15.- (2004) San Diego, USA : Society for Neuroscience.<br />

4 Maccari S. Morley-Fletcher S, Mairesse J., Viltart O., Daszuta A., Soumier A., Hery M.,<br />

Gabriel C, Mocaer E, Zuena A., Matteucci P., Cinque C. Catalani A. Casolini P. Chronic<br />

treatment with agomelatine reversed the decrease in hippocampal cells neurogenesis and<br />

survival in prenatally stressed adult rats Program no. 566.8. (2005) Washington, DC:<br />

Society for Neuroscience, 2005.<br />

5 McEwen BS, Wingfield JC (2003) The concept of allostasis in biology and biomedicine.<br />

Horm Behav. 43(1):2-15Viltart O, Mairesse J, et al. (2006). Prenatal stress alters Fos<br />

protein expression in hippocampus and locus coeruleus stress-related brain structures.<br />

Psychoneuroendocrinology 31:769-780.<br />

6 Viltart O, Mairesse J, Darnaudery M, Louvart H, Vanbesien-Mailliot C, Catalani A,<br />

Maccari S. Prenatal stress alters Fos protein expression in hippocampus and locus<br />

coeruleus stress-related brain structures. Psychoneuroendocrinology. 2006 Jul;31(6):769-80<br />

7 Mairesse et al., Prenatal stress disrupts the negative correlation between neuronal activation<br />

in limbic regions and behavioral responses in rats exposed to high and low anxiogenic<br />

environments. Submitted.<br />

272


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

ROLE OF APOPTOSIS IN THE SEXUAL DIFFERENTIATION OF THE BED<br />

NUCLEUS OF STRIA TERMINALIS OF JAPANESE QUAIL<br />

Mura E., Furnari P., Plumari L., Viglietti-Panzica C., Panzica G.C.<br />

Laboratory of Neuroendocrinology, Dept Anatomy, Pharmacology and Forensic Medicine, C.so M.<br />

D’Azeglio, 52. 10126 Torino (Italy)<br />

e-mail: elena.mura@unito.it<br />

Programmed cell death (or apoptosis) is an essential process during brain<br />

development that plays an important role in the differentiation of areas involved in<br />

sexually dimorphic functions. In fact, in rodents a different incidence of apoptosis between<br />

males and females during the first weeks of life establishes the sexual dimorphism of some<br />

brain nuclei [2,5,10].<br />

In galliform birds, the parvocellular vasotocin (VT) system of the limbic-preoptic region<br />

shows a strong dimorphism [4,6] probably dependent by the exposure to estradiol during<br />

embryonic development [8]. In adult males VT-immunoreactive (VT-ir) cells and fibers<br />

are largely present in the bed nucleus of the stria terminalis (BST), the medial preoptic<br />

nucleus (POM), and the lateral septum (SL). In females very low levels of VT-ir fibers are<br />

present, whereas cell bodies are totally absent. This strong dimorphism is also confirmed<br />

by the detection of mRNA for VT by means of in situ hybridization [4,7]. The BST<br />

represents the major source of VT-ir fibers for the innervation of both POM and SL [1,9].<br />

In the present study we investigated the possible sexually dimorphic incidence of apoptosis<br />

during the first ten postnatal days within the Bed Nucleus of Stria Terminalis (BST) of the<br />

Japanese quail. Quail BST shows, in fact, a strong sexual dimorphism of the parvocellular<br />

vasotocin (VT) system in adult, while at birth males and females have a comparable<br />

amount of vasotocinergic elements [3].<br />

Male and female quail chicks were sacrificed by intracardiac perfusion and according to<br />

their age they were divided in the following groups: P1, P2, P3, P4, P5, P8 and P10. Brains<br />

were taken out of the skull, frozen with dry ice and cut with a cryostat. Coronal sections<br />

were stained with cresyl violet or immunostained for VT detection. Sections stained with<br />

cresyl violet were analyzed with a camera lucida in order to distinguish and count<br />

apoptotic cells (cells with picnotic nucleus and cells with fragmented nuclear material).<br />

Adjacent sections immunostained for VT were semiquantitative analyzed to estimate the<br />

extension of the VT system.<br />

The semiquantitative analysis of VT-immunoreactivity has confirmed the presence of the<br />

VT-ir system and birth in both sexes with a similar extension. VT-ir cells diminish in a<br />

progressive way during postnatal development up to day 10, whereas fibers have a<br />

moderate immunoreactivity in both sexes.<br />

The two-way ANOVA (age and sex as indipendent variables) for the total number of<br />

apoptotic cells demonstrated a strong effect for both factors. Two by two comparisons with<br />

student t-test demonstrated a statistically significant difference at day P3, with females<br />

showing a higher number of apoptotic cells than males.<br />

We can therefore conclude that, in Japanese quail, apoptosis in the BST is a sexually<br />

differentiated process that could be one of the mechanisms inducing the sexual<br />

dimorphism of VT-ir system of the BST.<br />

273


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

Reference list<br />

1. P. Absil, M. Papello, C. Viglietti Panzica, J. Balthazart and G.C. Panzica, The medial<br />

preoptic nucleus receives vasotocinergic inputs in male quail: a tract-tracing and<br />

immunocytochemical study, J Chem Neuroanat 24 (2002) 27-39.<br />

2. Y. Arai, Y. Sekine and S. Murakami, Estrogen and Apoptosis in the Developing<br />

Sexually Dimorphic Preoptic Area in Female Rats, Neurosci Res 25 (1996) 403-407.<br />

3. N. Aste, G. Baiamonte, R. Grossmann and G.C. Panzica, Postnatal development and<br />

sexual differentiation of quail vasotocin system, Italian J Anat Embriol 102, Suppl.<br />

(1997) 82.<br />

4. N. Aste, J. Balthazart, P. Absil, R. Grossmann, E. Mühlbauer, C. Viglietti-Panzica and<br />

G.C. Panzica, Anatomical and neurochemical definition of the nucleus of the stria<br />

terminalis in Japanese quail (Coturnix japonica), J Comp Neurol 396 (1998) 141-157.<br />

5. E.C. Davis, P. Popper and R.A. Gorski, The Role of Apoptosis in Sexual<br />

Differentiation of the Rat Sexually Dimorphic Nucleus of the Preoptic Area, Brain<br />

Res. 734 (1996) 10-18.<br />

6. A. Jurkevich, S.W. Barth, N. Aste, G.C. Panzica and R. Grossmann, Intracerebral sex<br />

differences in the vasotocin system in birds: possible implication on behavioral and<br />

autonomic functions, Horm Behav 30 (1996) 673-681.<br />

7. A. Jurkevich, S.W. Barth and R. Grossmann, Sexual dimorphism of arg-vasotocin<br />

gene expressing neurons in the telencephalon and dorsal diencephalon of the domestic<br />

fowl. An immunocytochemical and in situ hybridization study, Cell Tissue Res 287<br />

(1997) 69-77.<br />

8. G.C. Panzica, C. Castagna, C. Viglietti-Panzica, C. Russo, O. Tlemçani and J.<br />

Balthazart, Organizational effects of estrogens on brain vasotocin and sexual behavior<br />

in quail, J Neurobiol 37 (1998) 684-699.<br />

9. G.C. Panzica, E. Mura, M. Pessatti and C. Viglietti Panzica, Early embryonic<br />

administration of xenoestrogens alters vasotocin system and male sexual behavior of<br />

the Japanese quail, Dom An Endo 29 (2005) 436-445.<br />

10. M. Yoshida, K. Yuri, Z. Kizaki, T. Sawada and M. Kawata, The distributions of<br />

apoptotic cells in the medial preoptic areas of male and female neonatal rats, Neuros<br />

Res 36 (2000) 1-7.<br />

274


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

THE ACCESSORY OLFACTORY BULB OF THE ADULT MOUSE:<br />

NEUROGENESIS AND MORPHOLOGICAL ANALYSIS IN THE TWO SEXES<br />

Oboti L.*°, Peretto P.*, Fasolo A.*, Panzica G.C.°<br />

*Department of Animal and Human Biology, University of Turin, via Accademia<br />

Albertina 13, 10123 Turin, Italy<br />

e-mail: livio.oboti@unito.it<br />

Fax: +39-011-6704692<br />

°Department of Anatomy, Pharmacology and Forensic Medicine, University of Turin, Via<br />

Accademia Albertina 13, 10123, Italy<br />

The accessory olfactory bulb (AOB) is a sexually dimorphic structure [1] of the<br />

vomeronasal system whose development and function is regulated by steroid hormones [2].<br />

The AOB is part of the accessory olfactory pathway that is devoted to the processing of<br />

information concerned with reproduction and social behaviours. Evidences in rats indicate<br />

that the persistence of neurogenesis occurring in the adult subventricular zone (SVZ) also<br />

involves the AOB [3]. In the rat the amount of the SVZ newborn neurons is sexually<br />

dimorphic in the anterior part of the granular layer (GrL) of the AOB [4]. Recent<br />

comparative analyses have shown significant differences in the organization and activity of<br />

the adult neurogenic regions in several mammalian species [5]. Specific aim of our work<br />

was to analyze the neurogenic activity in the adult mouse AOB. In particular we have<br />

studied the distribution of newborn cells in the AOB different layers and functional<br />

subdivisions (anterior/posterior) of both sexes in the CD1 mice strain. Systemic injections<br />

of the exogenous marker of cell proliferation BrdU, and markers of immature (DCX) and<br />

mature (Neu-N, CR, GABA, NOS) neurons were used to evaluate neurogenic activity 1<br />

month after the BrdU treatment. Quantitative analyses were used to establish number,<br />

distribution, and differentiating phenotypes of newly formed neurons in the AOB of both<br />

sexes. Moreover, by using G0α immunoreactivity we studied the AOB volumes<br />

considering each single layer in its own functional subdivision in males and females.<br />

Our results indicate the occurrence of newly formed neurons in the AOB of adult CD1<br />

mice in both sexes. No significant differences were found between sexes in the distribution<br />

and phenotype of newly formed neurons. Conversely, in both males and females we have<br />

found that newly formed neurons are preferentially located (p=0,0002) in the anterior part<br />

of the AOB. Finally, the volumetric study did not show any sexual difference.<br />

These overall results validate the hypothesis that the occurrence of such neurogenic<br />

phenomenon along the accessory olfactory pathway can be similar in phylogenetically<br />

related species. Nevertheless, the striking result concerning the absence of a morphological<br />

sexual dimorphism in the CD1mice AOB reinforce the importance of detailed comparative<br />

studies to understand the basis of sexual dimorphism in the mammalian brain.<br />

This work is supported by Compagnia di San Paolo (Neurotransplant Project 2004.2019)<br />

Reference list<br />

1. Guillamon A. & Segovia S., 1997. “Sex differences in the vomeronasal system”. In:<br />

Hormones, Brain, and Behavior (G.C.Panzica and J.Balthazart eds). Brain Res Bull<br />

44: 377-382.<br />

2. Segovia S. & Guillamon A., 1984. “Effects of sex steroids on the development of the<br />

accessory olfactory bulb in the rat: a volumetric study”. Dev. Brain Res. 16:312-314.<br />

275


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

3. Bonfanti L., Peretto P., Merighi A., Fasolo A., 1997. “Newly generated cells from the<br />

rostral migratory stream in the accessory olfactory bulb of the adult rat”. Neuroscience<br />

81:489-502.<br />

4. Peretto P., Giachino C., Panzica G.C., Fasolo A., 2001. ”Sexually dimorphic<br />

neurogenesis is topographically matched with the anterior accessory olfactory bulb of<br />

the adult rat”. Cell Tissue Res. 306:385-389.<br />

5. Rakic P. 2004. ”Neuroscience: immigration denied”. Nature 427:685-686.<br />

276


4 th International Meeting STEROIDS AND NERVOUS SYSTEM<br />

Villa Gualino, TORINO, Italy. February 17-21 2007<br />

UNDERNURISHMENT AND FOOD REHABILITATION EFFECTS ON PLASMA<br />

LEPTIN LEVELS IN WISTAR RAT<br />

Pinos H, Carrillo B, Pérez-Izquierdo M, Ortega E (1), Collado P.<br />

Dpto Psicobiología, Fac. Psicología UNED, C/ Juan del Rosal nº 10, 28040 Madrid, Spain,<br />

pcollado@psi.uned.es (1) Dpto de Bioquímica y Biología Molecular, Fac Medicina,<br />

Universidad de Granada, Spain.<br />

Leptin is a peptide hormone secreted by adipose tissue which is crucial in the regulation of<br />

feeding behaviour and has an important role in the regulation of metabolim, energy<br />

balance and reproduction (Campfield et al, 1995; Pelleymounter et al, 1995; Ahima and<br />

Flier, 2000; Ahima et al, 2000; Ahima and Osei, 2004).. Recent studies have shown that<br />

during the development, this hormone can play a role like neurotrophic factor (Bouret and<br />

Simerly, 2004; Bouret et al, 2004; Steppan and Swick, 1999). In well nourished rats,<br />

plasma leptin levels are different between males and females, showing males higher levels<br />

than females of the same age.<br />

In a previous study, we have studied the effect of undernutrition from gestational period to<br />

weaning day, in the postnatal day 21, and food rehabilitation, on the morphology of the<br />

locus coeuruleus (LC) in male and female Wistar rats. Data showed that pre and postnatal<br />

food deprovation until 60-days of age results in a significant decrease in the number of<br />

neurons in the LC in rats of both sexes. However, food rehabilitation induces some<br />

recovery in the LC neuronal population in males, but not in females.<br />

Data from the present study has shown that a food restriction period from embryonic day<br />

6 until postnatal day 60, has a consequence a significant decrease of serum leptin levels in<br />

both, male and female rats during development at postnatal day 12 but only seems to have<br />

a permanent effect in males when the food restricted diet extents until adulthood. When a<br />

food rehabilitation (water and food ad libitum) is implemented early enough in males,<br />

serum leptin levels recover in some extent levels showed by adult control males.<br />

Undernutrition also has a long term effect on body weight in both sexes, but nutritional<br />

rehabilitation leads to some degree of body weight recovery depending on sex and age at<br />

which that rehabilitation was implemented. It could be suggested that undernutrition and a<br />

posterior nutritional rehabilitation during the first stages of life is affecting different<br />

developmental processes in male and female rats and therefore the response of body<br />

weight and serum leptin levels in each sex to these different nutritional conditions varies,<br />

being males more vulnerable than females.<br />

References list<br />

1. Ahima RS, Flier JS (2000) Leptin. Annu. Rev. Physiol. 62:413-437<br />

2. Ahima RS, Saper CB, Flier JS et al (2000) Leptin regulation of neuroendocrine<br />

systems. Front. Neuroendocrinol. 21:263-307.<br />

3. Ahima RS, Osei YO (2004) Leptin signalling. Physiol. Behav. 81:223-241.<br />

4. Bouret SG, Simerly RB (2004) Minireview: Leptin and development of<br />

hypothalamic feeding circuits. Endocrinology. 145:2621-2626.<br />

5. Bouret SG, Draper SJ, Simerly RB (2004) Trophic action of leptin on hypothalamic<br />

neurons that regulate feeding. Science. 304:108-110.<br />

277


Trabajos del Instituto Cajal. Tomo LXXXI, 2007<br />

6. Campfield LA, Smith FJ, Guisez Y et al (1995) Recombinant mouse OB protein:<br />

evidence for a peripheral signal linking adiposity and central neural networks.<br />

Science. 269:546-549.<br />

7. Pelleymounter MA, Cullen MJ, Baker MB et al (1995) Effects of the obese gene<br />

producto on body weight regulation in ob/ob mice. Science. 269:540-543.<br />

8. Pinos H, Collado P, Salas M et al (2004) Undernutrition and food rehabilitation<br />

effects on the locus coeruleus in the rat. Neuroreport. 15:1417-1420.<br />

9. Steppan CM, Swick AG (1999) A role for leptin in brain development. Biochem.<br />

Biophys. Res. Comm. 256:600-602.<br />

Sources od support:MCYT grant BSO2003-02526 (PC)<br />

278


AUTHOR INDEX<br />

A<br />

Abdelnabi M.......................89<br />

Ábrahám I.M. ......... 180, 228<br />

Abrous D.N.........................66<br />

Acharjee S........................ 240<br />

Adriani W. ....................... 209<br />

Ahmadiani A. .................. 218<br />

Ahn H.S............................ 175<br />

Alias A.G.......................... 147<br />

Alleva E...................... 97, 199<br />

Allieri F. ........................... 217<br />

Almada R.F. .................... 220<br />

Aloisi A.M........................ 234<br />

Amini H............................ 218<br />

Andrade T.G.C.S.... 219, 220<br />

Arcieri P..............................13<br />

Assenza G ........................ 164<br />

Aste N. .............................. 222<br />

Atwood C.S...................... 224<br />

Avanzi V................... 219, 220<br />

Avitabile M ...................... 226<br />

Avola R............................. 226<br />

Avoli M............................. 247<br />

Azcoitia I.......................... 161<br />

B<br />

Bäckström T...121, 158, 195,<br />

197, 246, 260,<br />

Baghai T.C....................... 149<br />

Bakker J..................... 57, 104<br />

Baldelli E.......................... 247<br />

Ballarin C. ....................... 157<br />

Balog J.............................. 228<br />

Balthazart J............... 59, 104<br />

Banasr M ......................... 207<br />

Bao A-M................... 125, 265<br />

Barha C ...............................63<br />

Barker J.M. ........................63<br />

Barreto G......................... 161<br />

Barrile V. ......................... 267<br />

Barton M.............................89<br />

Bass A.H................. 39, 56, 99<br />

Baulieu E.E.........................19<br />

Bélanger A. ...................... 240<br />

Belcher S.M. .................... 113<br />

Bellinger D....................... 252<br />

Belloni V..............................97<br />

Belon P. ............................ 262<br />

Bembi B............................ 182<br />

Benedusi V............... 138, 142<br />

Berger A........................... 237<br />

Bernardi G..........................28<br />

Berry A............................. 199<br />

Berumen L.C................... 163<br />

Biagini G...........................247<br />

Biamonte F .......................164<br />

Bianchi R ................. 171, 177<br />

Biasella A. .........................250<br />

Bierzniece V. ....................195<br />

Biggio G. ...........................118<br />

Bixo M...............................260<br />

Bo E ...................................154<br />

Bodo C.................................42<br />

Bonifazi M. .......................250<br />

Bonsignore L.T. ...............204<br />

Boujedaini N.....................262<br />

Bourque M............................8<br />

Bowen R.L. .......................224<br />

Bramanti V.......................226<br />

Brinton R.D. ..................7, 65<br />

Broiz A.C.G......................219<br />

Bronzi D ............................226<br />

Brunström B.....................139<br />

Buson G.............................157<br />

Bütow A. ...........................242<br />

C<br />

Cambiasso M.J. ...............120<br />

Campbell B.C...................230<br />

Campbell R.E.....................33<br />

Caniglia S............................13<br />

Cannizzaro C. ..................267<br />

Carrero P...10, 143, 181, 238<br />

Carrillo B................. 232, 277<br />

Carroll J.C.........78, 165, 178<br />

Carta M.................... 116, 118<br />

Caruso D....25, 164, 177, 259<br />

Casella D...........................154<br />

Casolini P................. 207, 271<br />

Castel H.............................240<br />

Catalani A................ 207, 271<br />

Cavaletti G............... 171, 177<br />

Ceccarelli I........................234<br />

Cecchi C............................166<br />

Cesa R ...............................164<br />

Chalbot S. .........................2<strong>35</strong><br />

Chang L. ...........................165<br />

Chen C...............................252<br />

Chen S. ...............................65<br />

Chung E. ............................65<br />

Cibula D............................150<br />

Cinque C ...........................271<br />

Ciriza I. ...............................10<br />

Cirulli F.................... 199, 204<br />

Clarkson J. .........................33<br />

Collado P.................. 232, 277<br />

Corrieri L. ..........................93<br />

Cozzi B. .............................157<br />

Crotti S...............25, 164, 259<br />

Csakvari E........................237<br />

D<br />

Danza G. ...........................166<br />

Darbra S. ..........................249<br />

Dardis A ............................182<br />

Darnaudéry M. ................134<br />

Daszuta A..........................207<br />

De Angelis L .....................259<br />

de Kloet E.R. ....................199<br />

De Nicola A.F. ........... 19, 115<br />

De Padova A.M................234<br />

Delitala G............................13<br />

Della Seta D. .......................93<br />

Demajo M.A. ....................212<br />

Desole M.S. .........................13<br />

Dessì-Fulgheri F.......... 93, 97<br />

Dettling-Artho R..............200<br />

Deviche P. .........................154<br />

Dewing P. ............................55<br />

Di Michele F. ....................149<br />

Di Paolo T. ............................8<br />

Diaz-Heijtz R....................200<br />

Dichiara F.........................166<br />

Dieni C.V. .........................190<br />

Diliberto I. ........................267<br />

Diz-Chaves Y.....10, 181, 238<br />

Djordjevic A.D.,...............212<br />

Dluzen D.E. ...........................8<br />

Do Rego J.L......................240<br />

Doodipala S.R.....................27<br />

Dutia M.B. ........................190<br />

Dykens J.A..........................77<br />

E<br />

Eckert A............................179<br />

Erdei F. .............................228<br />

Eser D................................149<br />

F<br />

Farabollini F.......................93<br />

Fargo K.N.........................168<br />

Fasolo A. ...........................275<br />

Feldon J.............................200<br />

Fester L. ............................242<br />

Fiorenzani P. ....................234<br />

Forlano P.M. ......................99<br />

Formigli L.........................166<br />

Forsberg M. K..................169<br />

Fowler C.D. ........................64<br />

Frondaroli A.....................192<br />

Frye C.A.............11, 106, 144


Furnari P. .........................273<br />

Fusani L. .............................93<br />

G<br />

Galas L. .............................240<br />

Galea L.A.M.............. 63, 156<br />

García-Alcocer G. ...........163<br />

Garcia-Ovejero D. ...........161<br />

Garcia-Segura L.M. 10, 143,<br />

161, 171, 177, 181, 238<br />

García-Servín M..............163<br />

Gass P................................132<br />

Gennuso F...........................13<br />

Giaquinta G........................13<br />

Giatti S ..................... 171, 177<br />

Ginanneschi F. .................250<br />

Giorgio M. ........................199<br />

Gong W ...............................73<br />

Gonzalez Deniselle M.C. 115<br />

Gonzalez S.L. ...................115<br />

Gorosito S.V. ....................120<br />

Gotti S. ..............................269<br />

Grassi S. ............................192<br />

Greeson J. .................. 68, 2<strong>35</strong><br />

Griffiths W.J. .....................21<br />

Grossmann R. ....................45<br />

Guennoun R. ............. 19, 115<br />

Guillamon A.............. 51, 232<br />

Gye M.C............................175<br />

Gyenes A. ..........................237<br />

H<br />

Håkansson H. ...................187<br />

Hallberg M .......................169<br />

Halldin K. ................ 139, 187<br />

Handa R.J...........................41<br />

Hardin-Pouzet H .............217<br />

Hariri O. .................. 111, 194<br />

Hauser J............................200<br />

Hausmann O. .....................84<br />

Hayashi S. .........................255<br />

Herbert J.............................67<br />

Herbison A.E............. 33, 258<br />

Higashi T.............................23<br />

Hill M. ...................... 150, 152<br />

Hirata M. ................. 245, 251<br />

Hiroi R. .............................243<br />

Hirst JJ..............................202<br />

Horvat A.I.........................212<br />

Hoyk S...............................237<br />

Huber C. ...........................242<br />

I<br />

Ibrahim A. ..........................68<br />

Imaizumi K.........................86<br />

Irwin R................................65<br />

Ishunina T.A. .....................43<br />

J<br />

Jarrahi M................. 172, 210<br />

Jarry H..................... 242, 257<br />

Jasoni C.L.........................258<br />

Johansson I-M 158, 195, 246<br />

Juhász G. ................. 180, 228<br />

K<br />

Kancheva L. ............ 150, 152<br />

Kanematsu T........... 245, 251<br />

Kang H.S...........................175<br />

Kaur G. .............................256<br />

Kékesi K.A........................180<br />

Keller F .............................164<br />

Keller M............................104<br />

Kibaly C..................... 81, 174<br />

Kim H.J.............................175<br />

Klein S.................................45<br />

Knapman A. .....................200<br />

Kodama M........................255<br />

Kohlmann P .....................257<br />

Kondo S...............................86<br />

Korrick SA .......................252<br />

Kurunczi A.......................237<br />

Kwon H.B .........................240<br />

L<br />

L’Episcopo F......................13<br />

La Barbera M. .................267<br />

Labombarda F.......... 19, 115<br />

LaFerla F.M .............. 78, 165<br />

Lauria G .................. 171, 177<br />

Laviola G ................. 204, 209<br />

Lavoie E. .............................89<br />

Le H.H...............................113<br />

Lecanu L.................... 68, 2<strong>35</strong><br />

Lee J.E. .............................175<br />

Liere P.................................19<br />

Liguri G ............................166<br />

Lindblad C. ............. 158, 195<br />

Liu B. .....................................8<br />

Liu T..................................224<br />

Löfgren M.........................246<br />

Lohse C .............................257<br />

Longo D.............................247<br />

Longone P...........................28<br />

Lundgren P.......................197<br />

Luu-The V. .......................240<br />

M<br />

Maccari S.........134, 207, 271<br />

Macrì S..................... 204, 209<br />

Maggi A.............. 91, 138, 142<br />

Maggio N. .........................206<br />

Mairesse J................ 207, 271<br />

Mameli M. ........................116<br />

Manieri G. ..........................28<br />

Mantia G...........................267<br />

Marchetti B. .......................13<br />

Marino R ..........................164<br />

Martín-García E..............249<br />

Martini M. .47, 154, 185, 269<br />

Martin-Padura I..............199<br />

Maruniak J.........................94<br />

Maschi O.................... 25, 259<br />

Massafra C. ......................234<br />

Materazzi P. .......................93<br />

Matteucci P.............. 207, 271<br />

Matthews S.M. .................133<br />

Mattsson A. ......................139<br />

Mayoral S.R. ....................141<br />

Mazzocchio R...................250<br />

McCourty A. ......................68<br />

McEwen B.S. ....................102<br />

Meffre D............................115<br />

Melcangi R.C. .... 25, 83, 164,<br />

171, 177, 259<br />

Mellon S.H................. 73, 182<br />

Mensah-Nyagan A.G. ......81,<br />

101, 174, 179, 262<br />

Menzies JRW...................190<br />

Meyer L.............. 81, 101, 174<br />

Meyerson B.......................246<br />

Miceli D.............................185<br />

Micevych P. ....... 55, 111, 194<br />

Miele E. ...............................13<br />

Milani P.............................250<br />

Minghetti L.......................199<br />

Mirzaei M. ........................218<br />

Mizokami A......................251<br />

Mocaer E ..........................207<br />

Morale M.C........................13<br />

Morello M.........................166<br />

Morissette M. .......................8<br />

Morley-Fletcher S. 134, 207,<br />

271<br />

Mostallino M.C................118<br />

Mura E..................... 139, 273<br />

N<br />

Nagahama A.......................23<br />

Nakamuta J.S...................220<br />

Nasir RH...........................252<br />

Neumaier J.F....................243<br />

Nevyjel M..........................182<br />

Nicoletti F .........................271<br />

Ninomiya Y.........................23<br />

Nobahar M .......................254<br />

Nosi D................................166<br />

Noto G. ..............................267<br />

Nyberg F ...........................169<br />

Nyberg S. ................. 121, 260


O<br />

Oboti L. ............................ 275<br />

Oddo S........................ 78, 165<br />

Ognibene E. ..................... 209<br />

Ohya T.............................. 255<br />

Ortega E........................... 277<br />

Ottinger M.A......................89<br />

P<br />

Palanza P. .................. 94, 185<br />

Pallarés M........................ 249<br />

Palliser HK ...................... 202<br />

Panzica G.C......47, 139, 154,<br />

185, 232, 269, 273, 275<br />

Papadopoulos V. ....... 68, 2<strong>35</strong><br />

Parducz A. ....................... 237<br />

Paris, J.J........................... 106<br />

Pařízek A. ........................ 150<br />

Parkash J ......................... 256<br />

Parmigiani S.......................94<br />

Pasini A. ..................... 28, 149<br />

Pasquali P. ....................... 204<br />

Patisaul H.B................. 54, 90<br />

Patte-Mensah C. ...... 81, 101,<br />

174, 179, 262<br />

Pawluski J.L.............. 63, 156<br />

Pelicci P. G....................... 199<br />

Pelletier G. ....................... 240<br />

Penn A.A.......................... 141<br />

Pensalfini A. .................... 166<br />

Peretto P........................... 275<br />

Pérez-Izquierdo M.A .... 232,<br />

277<br />

Peri A................................ 166<br />

Pernía O. ....10, 143, 181, 238<br />

Peruffo A.......................... 157<br />

Pesaresi M................ 171, 177<br />

Pettorossi V.E.................. 192<br />

Pieraccini G. .................... 166<br />

Pike C.J. .............78, 165, 178<br />

Pinos H. .................... 232, 277<br />

Pisu M.G. ......................... 118<br />

Pittis MG.......................... 182<br />

Plescia F. .......................... 267<br />

Plumari L......................... 273<br />

Polston E.K.........................90<br />

Ponzi D. ...............................94<br />

Porteous R. .........................33<br />

Pozzi S. ..................... 138, 142<br />

Pradotto M. ..................... 269<br />

Prange-Kiel J................... 257<br />

Pryce C.R................. 130, 200<br />

Q<br />

Quinn M. Jr........................89<br />

R<br />

Raciti C .............................226<br />

Ragagnin G.......................197<br />

Rahman M............... 158, 197<br />

Rashidy-Pour A ...............172<br />

Repetto E. .........................240<br />

Rhodes, M.E.....................106<br />

Rissman E.F. ......................42<br />

Ritz M.-F.............................84<br />

Roglio I..................... 171, 177<br />

Romano N...........................33<br />

Romanò N.........................258<br />

Romeo E..................... 28, 149<br />

Romeo R.D. ......................102<br />

Rosario E.R. ......78, 165, 178<br />

Rosati F. ............................166<br />

Rossi A. .............................250<br />

Rune G.M. ............... 242, 257<br />

Rupprecht R.............. 28, 149<br />

S<br />

Sabetkasaei M. .................218<br />

Saito N...............................222<br />

Salimpour S......................218<br />

Sánchez-Ramos M.A.......163<br />

Sanna E. ............................118<br />

Santucci D...........................97<br />

Sanz A. ..............................143<br />

Schaeffer V. ............... 81, 179<br />

Schön M ............................257<br />

Schonemann M. .................73<br />

Schüle C. ...........................149<br />

Schumacher M.......... 19, 115<br />

Schwarz M........................149<br />

Scurati S.............25, 177, 259<br />

Segal M..............................206<br />

Segovia S. ............................51<br />

Sengelaub D.R.................168<br />

Seong J.Y. .........................240<br />

Sergio T.O.........................219<br />

Serio M..............................166<br />

Serra M. ............................118<br />

Serra P.-A. ..........................13<br />

Shimada K. ................ 23, 222<br />

Sica M......................... 47, 187<br />

Simpkins J.W. ....................77<br />

Sinchak K. ........................111<br />

Smith J.T. ...........................37<br />

Soma K..............................111<br />

Soumier A .........................207<br />

Sousa N..............................129<br />

Spritzer M.D. .....................63<br />

Spurling L.........................113<br />

Stanczyk F.Z. ...................165<br />

Stárka L ............................152<br />

Stefani M...........................166<br />

Stein D.G.................... 79, 115<br />

Strata P .............................164<br />

Strömberg J......................197<br />

Sundström Poromaa I. ...121<br />

Svensson A-L....................169<br />

Swaab D.F....43, 53, 125, 265<br />

Szabó G. ............................228<br />

Szajli A..............................237<br />

Szegő É.M................. 180, 228<br />

T<br />

Taherian AA.....................210<br />

Talani G. ...........................118<br />

Tapia González S...... 10, 181<br />

Taziaux M.........................104<br />

Tecozautla A. ...................163<br />

Tena-Sempere M. ..............15<br />

Testa N. ...............................13<br />

Thompson N.......................89<br />

Timby E. ...........................260<br />

Tirolo C...............................13<br />

Tobin V. ............................190<br />

Tonon M.C. ......................240<br />

Tremblay Y. .....................240<br />

Turkmen S........................195<br />

U<br />

Uzunov DP........................149<br />

V<br />

Vadakkadath Meethal S.224<br />

Vafaei AA ........172, 210, 254<br />

Valenzuela C.F.................116<br />

Vallarino M. .....................240<br />

Vaudry H. .........................240<br />

Včelaková H ............ 150, 152<br />

Vegeto E................... 138, 142<br />

Veiga S. .............................161<br />

Velickovic N.A .................212<br />

Venard C...........................262<br />

Verzè L................................47<br />

Viglietti-Panzica C. . 47, 154,<br />

185, 269, 273<br />

Vlad A.G. ..........................263<br />

Vom Saal F.S......................94<br />

von Lossow R. ..................242<br />

Vrbíková J........................152<br />

W<br />

Walf A.A. ................... 11, 144<br />

Walker C.A. .....................156<br />

Walker DW ......................202<br />

Wang J.M...........................65<br />

Wang M-D ............... 158, 197<br />

Wang Y. ..............................21<br />

Wang Z.X. ..........................64<br />

Watanabe Y......................222


Whitehouse K.....................89<br />

Wihlbäck A.-C.N. ............260<br />

Wilson A.C. ......................224<br />

Y<br />

Yao W..................................68<br />

Yates DM ..........................202<br />

Z<br />

Zaccaroni M.......................97<br />

Zampieri S ........................182<br />

Zamudio, P.A. ..................116<br />

Zhao L...................................7<br />

Zhou L...............................242<br />

Zingmark E. ............ 121, 197<br />

Zini I..................................247<br />

Zoli M................................247<br />

Zsarnovszky A. ................113<br />

Zuena AR................. 207, 271<br />

Zuercher N. ......................200


Conference organized with the support of<br />

Università degli Studi di Torino<br />

Università degli Studi di Milano<br />

Facoltà di Scienze MFN, Torino<br />

Dipartimento di Anatomia, Farmacologia e Medicina Legale<br />

Fondazione Oasi, Troina, Italy<br />

Centro Rita Levi Montalcini, Torino<br />

Center of Excellence on Neurodegenerative diseases, Milano<br />

International Brain Research Organization (IBRO)<br />

Società Italiana di Neuroscienze<br />

National Science Foundation<br />

Regione Piemonte<br />

Provincia di Torino<br />

Comune di Torino<br />

Applied Biosystems<br />

Thermofisher<br />

BIORAD<br />

CELBIO<br />

DBA, Italy<br />

Nikon, Italy<br />

Vinci Biochem, Italy<br />

Elsevier Publisher<br />

Karger Publisher

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!