30.11.2012 Views

Gene Section - Atlas of Genetics and Cytogenetics

Gene Section - Atlas of Genetics and Cytogenetics

Gene Section - Atlas of Genetics and Cytogenetics

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.

Volume 16 1 - Number 2 1<br />

May February - September 2012<br />

1997


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

Scope<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

The <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology is a peer reviewed on-line journal in<br />

open access, devoted to genes, cytogenetics, <strong>and</strong> clinical entities in cancer, <strong>and</strong> cancer-prone diseases.<br />

It presents structured review articles ("cards") on genes, leukaemias, solid tumours, cancer-prone diseases, more<br />

traditional review articles on these <strong>and</strong> also on surrounding topics ("deep insights"), case reports in hematology,<br />

<strong>and</strong> educational items in the various related topics for students in Medicine <strong>and</strong> in Sciences.<br />

Editorial correspondance<br />

Jean-Loup Huret<br />

<strong>Gene</strong>tics, Department <strong>of</strong> Medical Information,<br />

University Hospital<br />

F-86021 Poitiers, France<br />

tel +33 5 49 44 45 46 or +33 5 49 45 47 67<br />

jlhuret@<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org or Editorial@<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org<br />

Staff<br />

Mohammad Ahmad, Mélanie Arsaban, Marie-Christine Jacquemot-Perbal, Maureen Labarussias, Vanessa Le<br />

Berre, Anne Malo, Catherine Morel-Pair, Laurent Rassinoux, Alain Zasadzinski.<br />

Philippe Dessen is the Database Director, <strong>and</strong> Alain Bernheim the Chairman <strong>of</strong> the on-line version (Gustave<br />

Roussy Institute – Villejuif – France).<br />

The <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology (ISSN 1768-3262) is published 12 times<br />

a year by ARMGHM, a non pr<strong>of</strong>it organisation, <strong>and</strong> by the INstitute for Scientific <strong>and</strong> Technical Information <strong>of</strong><br />

the French National Center for Scientific Research (INIST-CNRS) since 2008.<br />

The <strong>Atlas</strong> is hosted by INIST-CNRS (http://www.inist.fr)<br />

http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org<br />

© ATLAS - ISSN 1768-3262<br />

The PDF version <strong>of</strong> the <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology is a reissue <strong>of</strong> the original articles published in collaboration with<br />

the Institute for Scientific <strong>and</strong> Technical Information (INstitut de l’Information Scientifique et Technique - INIST) <strong>of</strong> the French National Center for Scientific<br />

Research (CNRS) on its electronic publishing platform I-Revues.<br />

Online <strong>and</strong> PDF versions <strong>of</strong> the <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology are hosted by INIST-CNRS.


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2)<br />

Editor<br />

Jean-Loup Huret<br />

(Poitiers, France)<br />

Editorial Board<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

Sreeparna Banerjee (Ankara, Turkey) Solid Tumours <strong>Section</strong><br />

Aless<strong>and</strong>ro Beghini (Milan, Italy) <strong>Gene</strong>s <strong>Section</strong><br />

Anne von Bergh (Rotterdam, The Netherl<strong>and</strong>s) <strong>Gene</strong>s / Leukaemia <strong>Section</strong>s<br />

Judith Bovée (Leiden, The Netherl<strong>and</strong>s) Solid Tumours <strong>Section</strong><br />

Vasantha Brito-Babapulle (London, UK) Leukaemia <strong>Section</strong><br />

Charles Buys (Groningen, The Netherl<strong>and</strong>s) Deep Insights <strong>Section</strong><br />

Anne Marie Capodano (Marseille, France) Solid Tumours <strong>Section</strong><br />

Fei Chen (Morgantown, West Virginia) <strong>Gene</strong>s / Deep Insights <strong>Section</strong>s<br />

Antonio Cuneo (Ferrara, Italy) Leukaemia <strong>Section</strong><br />

Paola Dal Cin (Boston, Massachussetts) <strong>Gene</strong>s / Solid Tumours <strong>Section</strong><br />

Louis Dallaire (Montreal, Canada) Education <strong>Section</strong><br />

Brigitte Debuire (Villejuif, France) Deep Insights <strong>Section</strong><br />

François Desangles (Paris, France) Leukaemia / Solid Tumours <strong>Section</strong>s<br />

Enric Domingo-Villanueva (London, UK) Solid Tumours <strong>Section</strong><br />

Ayse Erson (Ankara, Turkey) Solid Tumours <strong>Section</strong><br />

Richard Gatti (Los Angeles, California) Cancer-Prone Diseases / Deep Insights <strong>Section</strong>s<br />

Ad Geurts van Kessel (Nijmegen, The Netherl<strong>and</strong>s) Cancer-Prone Diseases <strong>Section</strong><br />

Oskar Haas (Vienna, Austria) <strong>Gene</strong>s / Leukaemia <strong>Section</strong>s<br />

Anne Hagemeijer (Leuven, Belgium) Deep Insights <strong>Section</strong><br />

Nyla Heerema (Colombus, Ohio) Leukaemia <strong>Section</strong><br />

Jim Heighway (Liverpool, UK) <strong>Gene</strong>s / Deep Insights <strong>Section</strong>s<br />

Sakari Knuutila (Helsinki, Finl<strong>and</strong>) Deep Insights <strong>Section</strong><br />

Lidia Larizza (Milano, Italy) Solid Tumours <strong>Section</strong><br />

Lisa Lee-Jones (Newcastle, UK) Solid Tumours <strong>Section</strong><br />

Edmond Ma (Hong Kong, China) Leukaemia <strong>Section</strong><br />

Roderick McLeod (Braunschweig, Germany) Deep Insights / Education <strong>Section</strong>s<br />

Cristina Mecucci (Perugia, Italy) <strong>Gene</strong>s / Leukaemia <strong>Section</strong>s<br />

Yasmin Mehraein (Homburg, Germany) Cancer-Prone Diseases <strong>Section</strong><br />

Fredrik Mertens (Lund, Sweden) Solid Tumours <strong>Section</strong><br />

Konstantin Miller (Hannover, Germany) Education <strong>Section</strong><br />

Felix Mitelman (Lund, Sweden) Deep Insights <strong>Section</strong><br />

Hossain Mossafa (Cergy Pontoise, France) Leukaemia <strong>Section</strong><br />

Stefan Nagel (Braunschweig, Germany) Deep Insights / Education <strong>Section</strong>s<br />

Florence Pedeutour (Nice, France) <strong>Gene</strong>s / Solid Tumours <strong>Section</strong>s<br />

Elizabeth Petty (Ann Harbor, Michigan) Deep Insights <strong>Section</strong><br />

Susana Raimondi (Memphis, Tennesse) <strong>Gene</strong>s / Leukaemia <strong>Section</strong><br />

Mariano Rocchi (Bari, Italy) <strong>Gene</strong>s <strong>Section</strong><br />

Alain Sarasin (Villejuif, France) Cancer-Prone Diseases <strong>Section</strong><br />

Albert Schinzel (Schwerzenbach, Switzerl<strong>and</strong>) Education <strong>Section</strong><br />

Clelia Storlazzi (Bari, Italy) <strong>Gene</strong>s <strong>Section</strong><br />

Sabine Strehl (Vienna, Austria) <strong>Gene</strong>s / Leukaemia <strong>Section</strong>s<br />

Nancy Uhrhammer (Clermont Ferr<strong>and</strong>, France) <strong>Gene</strong>s / Cancer-Prone Diseases <strong>Section</strong>s<br />

Dan Van Dyke (Rochester, Minnesota) Education <strong>Section</strong><br />

Roberta Vanni (Montserrato, Italy) Solid Tumours <strong>Section</strong><br />

Franck Viguié (Paris, France) Leukaemia <strong>Section</strong><br />

José Luis Vizmanos (Pamplona, Spain) Leukaemia <strong>Section</strong><br />

Thomas Wan (Hong Kong, China) <strong>Gene</strong>s / Leukaemia <strong>Section</strong>s


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

<strong>Gene</strong> <strong>Section</strong><br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2)<br />

Volume 16, Number 2, February 2012<br />

Table <strong>of</strong> contents<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

LMO1 (LIM domain only 1 (rhombotin 1)) 82<br />

Norihisa Saeki, Hiroki Sasaki<br />

MAP2 (microtubule-associated protein 2) 85<br />

Ashika Jayanthy, Vijayasaradhi Setaluri<br />

MIR200C (microRNA 200c) 90<br />

Sarah Jurmeister, Stefan Uhlmann, Özgür Sahin<br />

PXN (paxillin) 98<br />

Tiffany Pierson, Brendan C Stack Jr<br />

RPS27 (ribosomal protein S27) 101<br />

Tiffany Pierson, Brendan C Stack Jr<br />

STAT5B (signal transducer <strong>and</strong> activator <strong>of</strong> transcription 5B) 104<br />

Am<strong>and</strong>a M Del Rosario, Teresa M Bernaciak, Corinne M Silva<br />

AAMP (angio-associated, migratory cell protein) 109<br />

Marie E Beckner<br />

BCL2L15 (BCL2-like 15) 113<br />

Maria-Angeliki S Pavlou, Christos K Kontos<br />

DUSP6 (dual specificity phosphatase 6) 117<br />

Zhenfeng Zhang, Balazs Halmos<br />

GZMA (granzyme A (granzyme 1, cytotoxic T-lymphocyte-associated serine esterase 3)) 121<br />

Elena Catalan, Diego Sanchez-Martinez, Julián Pardo<br />

MIER1 (mesoderm induction early response 1 homolog (Xenopus laevis)) 125<br />

Laura L Gillespie, Gary D Paterno<br />

PA2G4 (proliferation-associated 2G4, 38kDa) 129<br />

Anne Hamburger, Arundhati Ghosh, Smita Awasthi<br />

PRDM1 (PR domain containing 1, with ZNF domain) 133<br />

Wayne Tam<br />

Leukaemia <strong>Section</strong><br />

del(11)(q23q23) MLL/ARHGEF12 139<br />

Jean-Loup Huret<br />

t(3;11)(q21;q23) 141<br />

Jean-Loup Huret


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

t(11;14)(q13;q32) in multiple myeloma Huret JL, Laï JL<br />

Solid Tumour <strong>Section</strong><br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2)<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

Head <strong>and</strong> Neck: Squamous cell carcinoma: an overview 143<br />

Audrey Rousseau, Cécile Badoual<br />

Cancer Prone Disease <strong>Section</strong><br />

Rombo syndrome 154<br />

Jean-Loup Huret<br />

Deep Insight <strong>Section</strong><br />

Cohesins <strong>and</strong> cohesin-regulators: Role in Chromosome Segregation/Repair<br />

<strong>and</strong> Potential in Tumorigenesis 155<br />

José L Barbero<br />

Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells 163<br />

Maryam Mehrpour, Joëlle Botti, Patrice Codogno<br />

Case Report <strong>Section</strong><br />

A new case <strong>of</strong> t(5;14)(q31;q32) in a pediatric acute lymphoblastic leukemia<br />

presenting with hypereosinophilia 181<br />

Marta Gallego, Mariela Coccé, María Felice, Jorge Rossi, Silvia E<strong>and</strong>i, Gabriela Sciuccati, Cristina Alonso<br />

Chromosomal translocation t(X;11)(q22;q23) involving the MLL gene 183<br />

Adriana Zamecnikova, Soad Al Bahar, Hassan A Al Jafar, Rames P<strong>and</strong>ita


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

<strong>Gene</strong> <strong>Section</strong><br />

Mini Review<br />

LMO1 (LIM domain only 1 (rhombotin 1))<br />

Norihisa Saeki, Hiroki Sasaki<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

<strong>Gene</strong>tics Division, National Cancer Center Research Institute, Tokyo 104-0045, Japan (NS, HS)<br />

Published in <strong>Atlas</strong> Database: August 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/<strong>Gene</strong>s/LMO1ID33ch11p15.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI LMO1ID33ch11p15.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Identity<br />

Other names: rhombotin 1, RBTN1, RHOM1,<br />

TTG1<br />

HGNC (Hugo): LMO1<br />

Location: 11p15.4<br />

Local order: Telomeric to STK33 gene;<br />

centromeric to RIC3 gene.<br />

DNA/RNA<br />

Description<br />

4,4 kb (from the beginning <strong>of</strong> the 1st exon to the<br />

end <strong>of</strong> the 5th exon) consisting <strong>of</strong> 5 exons.<br />

Transcription<br />

mRNA <strong>of</strong> approximately 700 - 1000 bp, depending<br />

on the splicing variants.<br />

Pseudogene<br />

Not reported.<br />

LMO1 gene is located between RIC3 <strong>and</strong> STK33 genes in Chromosome 11p15. The arrow indicates the orientation <strong>of</strong> the genes.<br />

Structure <strong>of</strong> LMO1 gene. The gene consists <strong>of</strong> 5 exons <strong>and</strong> several splicing variants were reported.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 82


LMO1 (LIM domain only 1 (rhombotin 1)) Saeki N, Sasaki H<br />

Protein<br />

Description<br />

The protein contains two highly conserved,<br />

cysteine-rich motifs known as LIM domains, which<br />

interact with other proteins. As LMO1 has no<br />

DNA-binding domain, its DNA binding ability is<br />

dependent on the other proteins with which it<br />

interacts. The proteins known for binding to LMO1<br />

include TAL1/SCL <strong>and</strong> LDB1, <strong>and</strong> the molecules<br />

have an oncogenic function in T cell acute leukemia<br />

with the chromosomal translocation<br />

t(11;14)(p15;q11).<br />

Expression<br />

LMO1 expression is prominent in the central<br />

nervous system in both human <strong>and</strong> mouse. In mice<br />

it was observed by an in situ hybridization<br />

technique in the cerebral cortex, diencephalon,<br />

mesencephalon, cerebellum, myeloencephalon <strong>and</strong><br />

spinal cord. Northern blot analysis revealed its<br />

expression in the thymus, kidney <strong>and</strong> placenta <strong>of</strong><br />

adult mice.<br />

Localisation<br />

Nucleus.<br />

Function<br />

Transcription factor.<br />

Mutations<br />

Note<br />

Not reported.<br />

Implicated in<br />

Neuroblastoma<br />

Disease<br />

Genome-wide association studies revealed a<br />

corelation between neuroblastoma <strong>and</strong> a single<br />

nucleotide polymorphism in LMO1 gene, rs110419<br />

(A/G), <strong>of</strong> which the A allele was shown to promote<br />

LMO1 expression <strong>and</strong> to corelate the cases. The<br />

DNA copy number gain <strong>of</strong> the LMO1 locus due to<br />

duplication was also demonstrated to associate with<br />

the cases. These findings suggest that the gene has a<br />

role in the development <strong>and</strong>/or progression <strong>of</strong><br />

neuroblastoma. It was also reported that LMO3 is a<br />

neuroblastoma oncogene.<br />

Apoptosis <strong>of</strong> the gastric epithelial<br />

cells<br />

Note<br />

In some gastric-cancer derived cell lines, LMO1 is<br />

upregulated by TGFbeta signalling <strong>and</strong> induces<br />

their apoptosis through enhancing GSDMA<br />

expression. This TGFbeta-LMO1-GSDMA cascade<br />

is considered a mechanism for apoptosis induction<br />

in the gastric epithelial cells, <strong>and</strong> has a role in<br />

maintaining their homeostasis.<br />

LMO1 induces apoptosis <strong>of</strong> the pit cells in TGFbeta<br />

signalling. This function is assumed to have a role in<br />

homeostasis <strong>of</strong> the gastric epithelium.<br />

T-cell acute lymphoblastic leukemia<br />

(T-ALL)<br />

Disease<br />

Originally, the LMO1 gene was identified at a<br />

break point <strong>of</strong> the translocation t(11;14)(p15;q11),<br />

which was frequently observed in T-ALL. Using<br />

LMO1 as a probe, LMO2 <strong>and</strong> LMO3 were<br />

identified. LMO1 is expressed in inmature T cells<br />

but suppressed in the mature cells, <strong>and</strong><br />

overexpression <strong>of</strong> the gene in thymocytes,<br />

hematopoeitic progeniter cells located in the<br />

thymus, resulted in developing T-ALL in mice.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 83


LMO1 (LIM domain only 1 (rhombotin 1)) Saeki N, Sasaki H<br />

LMO1 acts cooperatively with SCL (Stem cell<br />

leukemia) as a transcriptional activator or represser,<br />

dependent on the genes <strong>and</strong> the cells, <strong>and</strong> enforced<br />

expression <strong>of</strong> LMO1 <strong>and</strong> SCL in the thymocytes<br />

inhibits T-cell differentiation <strong>and</strong> causes T-ALL. It<br />

is known that the LMO1-SCL complex regulates<br />

several genes, including suppression <strong>of</strong> NFKB1<br />

(nuclear factor <strong>of</strong> kappa light polypeptide gene<br />

enhancer in B-cells 1) <strong>and</strong> PTCRA (pre T-cell<br />

antigen receptor alpha), activation <strong>of</strong> NKX3.1<br />

(NK3 homeobox 1).<br />

References<br />

McGuire EA, Hockett RD, Pollock KM, Bartholdi MF,<br />

O'Brien SJ, Korsmeyer SJ. The t(11;14)(p15;q11) in a Tcell<br />

acute lymphoblastic leukemia cell line activates<br />

multiple transcripts, including Ttg-1, a gene encoding a<br />

potential zinc finger protein. Mol Cell Biol. 1989<br />

May;9(5):2124-32<br />

Boehm T, Foroni L, Kaneko Y, Perutz MF, Rabbitts TH.<br />

The rhombotin family <strong>of</strong> cysteine-rich LIM-domain<br />

oncogenes: distinct members are involved in T-cell<br />

translocations to human chromosomes 11p15 <strong>and</strong> 11p13.<br />

Proc Natl Acad Sci U S A. 1991 May 15;88(10):4367-71<br />

Boehm T, Spillantini MG, S<strong>of</strong>roniew MV, Surani MA,<br />

Rabbitts TH. Developmentally regulated <strong>and</strong> tissue specific<br />

expression <strong>of</strong> mRNAs encoding the two alternative forms<br />

<strong>of</strong> the LIM domain oncogene rhombotin: evidence for<br />

thymus expression. Oncogene. 1991 May;6(5):695-703<br />

McGuire EA, Rintoul CE, Sclar GM, Korsmeyer SJ. Thymic<br />

overexpression <strong>of</strong> Ttg-1 in transgenic mice results in T-cell<br />

acute lymphoblastic leukemia/lymphoma. Mol Cell Biol.<br />

1992 Sep;12(9):4186-96<br />

Ono Y, Fukuhara N, Yoshie O. TAL1 <strong>and</strong> LIM-only proteins<br />

synergistically induce retinaldehyde dehydrogenase 2<br />

expression in T-cell acute lymphoblastic leukemia by<br />

acting as c<strong>of</strong>actors for GATA3. Mol Cell Biol. 1998<br />

Dec;18(12):6939-50<br />

Valge-Archer V, Forster A, Rabbitts TH. The LMO1 <strong>and</strong><br />

LDB1 proteins interact in human T cell acute leukaemia<br />

with the chromosomal translocation t(11;14)(p15;q11).<br />

Oncogene. 1998 Dec 17;17(24):3199-202<br />

Herblot S, Steff AM, Hugo P, Aplan PD, Hoang T. SCL <strong>and</strong><br />

LMO1 alter thymocyte differentiation: inhibition <strong>of</strong> E2A-<br />

HEB function <strong>and</strong> pre-T alpha chain expression. Nat<br />

Immunol. 2000 Aug;1(2):138-44<br />

Aoyama M, Ozaki T, Inuzuka H, Tomotsune D, Hirato J,<br />

Okamoto Y, Tokita H, Ohira M, Nakagawara A. LMO3<br />

interacts with neuronal transcription factor, HEN2, <strong>and</strong> acts<br />

as an oncogene in neuroblastoma. Cancer Res. 2005 Jun<br />

1;65(11):4587-97<br />

Chang PY, Draheim K, Kelliher MA, Miyamoto S. NFKB1 is<br />

a direct target <strong>of</strong> the TAL1 oncoprotein in human T<br />

leukemia cells. Cancer Res. 2006 Jun 15;66(12):6008-13<br />

Saeki N, Kim DH, Usui T, Aoyagi K, Tatsuta T, Aoki K,<br />

Yanagihara K, Tamura M, Mizushima H, Sakamoto H,<br />

Ogawa K, Ohki M, Shiroishi T, Yoshida T, Sasaki H.<br />

GASDERMIN, suppressed frequently in gastric cancer, is a<br />

target <strong>of</strong> LMO1 in TGF-beta-dependent apoptotic<br />

signalling. Oncogene. 2007 Oct 4;26(45):6488-98<br />

Kusy S, Gerby B, Goardon N, Gault N, Ferri F, Gérard D,<br />

Armstrong F, Ballerini P, Cayuela JM, Baruchel A, Pflumio<br />

F, Roméo PH. NKX3.1 is a direct TAL1 target gene that<br />

mediates proliferation <strong>of</strong> TAL1-expressing human T cell<br />

acute lymphoblastic leukemia. J Exp Med. 2010 Sep<br />

27;207(10):2141-56<br />

Wang K, Diskin SJ, Zhang H, Attiyeh EF, Winter C, Hou C,<br />

Schnepp RW, Diamond M, Bosse K, Mayes PA, Glessner<br />

J, Kim C, Frackelton E, Garris M, Wang Q, Glaberson W,<br />

Chiavacci R, Nguyen L, Jagannathan J, Saeki N, Sasaki<br />

H, Grant SF, Iolascon A, Mosse YP, Cole KA, Li H, Devoto<br />

M, McGrady PW, London WB, Capasso M, Rahman N,<br />

Hakonarson H, Maris JM. Integrative genomics identifies<br />

LMO1 as a neuroblastoma oncogene. Nature. 2011 Jan<br />

13;469(7329):216-20<br />

This article should be referenced as such:<br />

Saeki N, Sasaki H. LMO1 (LIM domain only 1 (rhombotin<br />

1)). <strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012;<br />

16(2):82-84.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 84


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

<strong>Gene</strong> <strong>Section</strong><br />

Mini Review<br />

MAP2 (microtubule-associated protein 2)<br />

Ashika Jayanthy, Vijayasaradhi Setaluri<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

Department <strong>of</strong> Dermatology, University <strong>of</strong> Wisconsin-Madison, Madison, Wisconsin 53706, USA.<br />

(AJ, VS)<br />

Published in <strong>Atlas</strong> Database: August 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/<strong>Gene</strong>s/MAP2ID44216ch2q34.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI MAP2ID44216ch2q34.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Identity<br />

Other names: DKFZp686I2148, MAP2A,<br />

MAP2B, MAP2C,<br />

HGNC (Hugo): MAP2<br />

Location: 2q34<br />

Note<br />

The protein encoded by this gene plays a role in<br />

dendrite morphogenesis in the vertebrate central<br />

nervous system. This function is accomplished by<br />

regulating microtubule stability <strong>and</strong> preventing the<br />

depolymerization <strong>of</strong> microtubules by stiffening<br />

them. It is known to have a number <strong>of</strong> distinctive<br />

is<strong>of</strong>orms.<br />

DNA/RNA<br />

Description<br />

The DNA consists <strong>of</strong> three or four t<strong>and</strong>em repeats<br />

that code for 31 amino acid long residues.<br />

The gene contains 15 exons <strong>and</strong> has a size <strong>of</strong><br />

310064 base pairs. Alternative splicing <strong>of</strong> this gene<br />

gives rise to a large variety <strong>of</strong> transcripts <strong>and</strong><br />

is<strong>of</strong>orms.<br />

Transcription<br />

Four transcription variants have been characterized.<br />

- MAP2 Is<strong>of</strong>orm 5 (NM_001039538.1 --><br />

NP_001034627.1): The difference in this is<strong>of</strong>orm is<br />

characterized by the 5' UTR which results in the<br />

production <strong>of</strong> a longer protein when compared to<br />

is<strong>of</strong>orm 2.<br />

- MAP2 Is<strong>of</strong>orm 1 (NM_002374.3 --><br />

NP_002365.3): This is<strong>of</strong>orm is thought to be the<br />

longest encoded <strong>and</strong> contains three alternative inframe<br />

exons when compared to is<strong>of</strong>orm 2. The<br />

tubulin binding <strong>and</strong> MAP2 projection domains are<br />

conserved.<br />

- MAP2 Is<strong>of</strong>orm 2 (NM_031845.2 --><br />

NP_114033.2): This is the shortest transcript.<br />

- MAP2 Is<strong>of</strong>orm 4 (NM_031847.2 --><br />

NP_114035.2): One alternate in-frame exon<br />

compared to is<strong>of</strong>orm 2.<br />

Pseudogene<br />

None.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 85


MAP2 (microtubule-associated protein 2) Jayanthy A, Setaluri V<br />

Variant 5 mRNA <strong>Gene</strong><br />

Exon Start End Length Start End Length<br />

1 1 230 230 1 230 230<br />

2 231 280 50 83546 83595 50<br />

3 281 345 65 155990 156054 65<br />

4 346 422 77 201007 201083 77<br />

5 423 713 291 229096 229386 291<br />

6 714 827 114 254526 254639 114<br />

7 828 905 78 256704 256781 78<br />

8 906 967 62 276231 276292 62<br />

9 968 1138 171 280404 280574 171<br />

10 1139 1286 148 281534 281681 148<br />

11 1287 1627 341 285868 286208 341<br />

12 1628 1720 93 299549 299641 93<br />

13 1721 1802 82* 301633 301744 112*<br />

14 1803 1915 113 305804 305916 113<br />

15 1916 5844 3929 306136 310064 3929<br />

TOTAL LENGTH : 5844*<br />

Variant 1 mRNA <strong>Gene</strong><br />

Exon Start End Length Start End Length<br />

1 1 77 77 155633 155709 77<br />

2 78 142 65 155990 156054 65<br />

3 143 219 77 201007 201083 77<br />

4 220 510 291 229096 229386 291<br />

5 511 624 114 254526 254639 114<br />

6 625 702 78 256704 256781 78<br />

7 703 4428 3726 268579 272304 3726<br />

8 4429 4635 207 272496 272702 207<br />

9 4636 4770 135 272871 273005 135<br />

10 4771 4832 62 276231 276292 62<br />

11 4833 4980 148 281534 281681 148<br />

12 4981 5321 341 285868 286208 341<br />

13 5322 5403 82 301663 301744 82<br />

14 5404 5516 113 305804 305916 113<br />

15 5517 9445 3929 306136 310064 3929<br />

TOTAL LENGTH : 9445<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 86


MAP2 (microtubule-associated protein 2) Jayanthy A, Setaluri V<br />

Variant 4 mRNA <strong>Gene</strong><br />

Exon Start End Length Start End Length<br />

1 1 77 77 155633 155709 77<br />

2 78 142 65 155990 156054 65<br />

3 143 219 77 201007 201083 77<br />

4 220 510 291 229096 229386 291<br />

5 511 624 114 254526 254639 114<br />

6 625 702 78 256704 256781 78<br />

7 703 764 62 276231 276292 62<br />

8 765 912 148 281534 281681 148<br />

9 913 1253 341 285868 286208 341<br />

10 1254 1346 93 299549 299641 93<br />

11 1347 1428 82 301663 301744 82<br />

12 1429 1541 113 305804 305916 113<br />

13 1542 5470 3929 306136 310064 3929<br />

TOTAL LENGTH : 5470<br />

Variant 2 mRNA <strong>Gene</strong><br />

Exon Star End Length Start End Length<br />

1 1 77 77 155633 155709 77<br />

2 78 142 65 155990 156054 65<br />

3 143 219 77 201007 201083 77<br />

4 220 510 291 229096 229386 291<br />

5 511 624 114 254526 254639 114<br />

6 625 702 78 256704 256781 78<br />

7 703 764 62 276231 276292 62<br />

8 765 912 148 281534 281681 148<br />

9 913 1253 341 285868 286208 341<br />

10 1254 1335 82 301663 301744 82<br />

11 1336 1448 113 305804 305916 113<br />

12 1449 5377 3929 306136 310064 3929<br />

TOTAL LENGTH : 5377<br />

*There is a discrepancy between base pair numbers recorded for the transcript <strong>and</strong> the gene for this exon.<br />

Protein<br />

Note<br />

MAP2 is an approximately 1827 amino acid long<br />

protein with an estimated molecular weight <strong>of</strong> 200<br />

kDa, with the exact molecular weight varying by<br />

is<strong>of</strong>orm. Four is<strong>of</strong>orms have been characterized, but<br />

additional ones are thought to exist. The protein<br />

undergoes post-translational phosphorylation upon<br />

DNA damage. The phosphorylation is hypothesized<br />

to be catalyzed by ATM or ATR. In the rat brain,<br />

the various is<strong>of</strong>orms were characterized at different<br />

stages <strong>of</strong> development - is<strong>of</strong>orm MAP2B is found<br />

throughout rat brain development, while MAP2A<br />

appears towards the end <strong>of</strong> the second week <strong>of</strong> postnatal<br />

life. MAP2C is found during the early<br />

development <strong>of</strong> the brain, but after maturation is<br />

only found in the neural cells <strong>of</strong> the retina, olfactory<br />

bulb <strong>and</strong> the cerebellum.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 87


MAP2 (microtubule-associated protein 2) Jayanthy A, Setaluri V<br />

Description<br />

MAP2 is a mostly unfolded protein that changes<br />

conformation upon binding to its target molecule. A<br />

domain near its carboxyl terminus enables MAP2<br />

protein to bind to the microtubules. A 31 amino<br />

acid long repeating motif is characteristic <strong>of</strong> this<br />

protein. However, it is found that this motif is not<br />

sufficient by itself to bind to microtubules. Two<br />

contiguous sequences on either end <strong>of</strong> this<br />

repeating structure on both the amino <strong>and</strong> carboxyl<br />

ends enable the binding <strong>of</strong> this protein to the<br />

microtubules. A proline rich domain on the amino<br />

end is thought to be especially crucial in this<br />

process. The protein is known to have three tubulin<br />

binding domains spanning residues 1160-1691;<br />

1692-1722; 1723-1754. The protein also has a<br />

projection domain which extends from residues<br />

377-1505. All is<strong>of</strong>orms have a conserved Cterminal<br />

domain which contain tubulin binding<br />

repeats <strong>and</strong> a conserved N-terminal projection<br />

domain. The projection domain varies in size across<br />

is<strong>of</strong>orms, has a net negative charge <strong>and</strong> exerts a<br />

long range repulsive force. This gives a potential<br />

mechanism that explains how MAP2 regulates the<br />

spacing between microtubules.<br />

Expression<br />

MAP2 plays a major role in dendrite<br />

morphogenesis <strong>and</strong> is normally expressed in<br />

neurons. It has also been reported to be ectopically<br />

expressed in several cancers including melanoma<br />

<strong>and</strong> breast cancer.<br />

Localisation<br />

MAP2 mRNAs were found in the avian neuronal<br />

cell body cytoplasm; however the protein was<br />

found localized to the dendrites in mammals<br />

(Crist<strong>of</strong>anilli et al., 2004). The same report shows<br />

that avian neuronal MAP2 mRNA lacks a dendritic<br />

targeting element in its 3' UTR. Local expression<br />

within dendrites is hypothesized to be more suited<br />

to regulate need based synthesis. Tubulin, a protein<br />

expressed in both axons <strong>and</strong> dendrites is known to<br />

be expressed in the cytoplasm <strong>of</strong> the cell body<br />

showing that location specific expression <strong>of</strong><br />

proteins is important to the maintenance <strong>of</strong> polarity<br />

<strong>of</strong> the neural cells.<br />

Function<br />

MAP2 stabilizes microtubule bundling <strong>and</strong><br />

stiffening through the interactions <strong>of</strong> several weak<br />

binding sites to the microtubules on the protein.<br />

The strength <strong>of</strong> bundling <strong>of</strong> microtubules is directly<br />

correlated to the strength <strong>of</strong> the binding to MAP2.<br />

This enables the microtubules to support outgrowth<br />

from the cells. When MAP2 was expressed by<br />

transfection in non neuronal cells, it induced the<br />

rearrangement <strong>of</strong> the microtubules into long<br />

bundles. These bundles enabled outgrowths from<br />

the non neuronal cells.<br />

Experiments done with knockout mice show that<br />

the role <strong>of</strong> MAP2 in neuronal morphogenesis may<br />

be redundant (Teng et al., 2001). Single knockouts<br />

<strong>of</strong> MAP2 did not show any severe phenotypes but<br />

simultaneous knockouts <strong>of</strong> MAP2 <strong>and</strong> MAP1B died<br />

in the prenatal stage (Teng et al., 2001). There are<br />

several reports <strong>of</strong> functional redundancy amongst<br />

the MAP proteins.<br />

Homology<br />

A microtubule associated protein with similar<br />

function to MAP2 is known to be expressed in the<br />

rat (Rattus norvegicus), chicken (Gallus gallus) <strong>and</strong><br />

lizard (Anolis carolinensis) with varying levels <strong>of</strong><br />

sequence homology. In the fruit fly, the tau gene<br />

seems to perform a similar function.<br />

Mutations<br />

Note<br />

35 SNPs associated with MAP2 have been<br />

identified. The CAGs, which are a set <strong>of</strong><br />

trinucleotide sequences starting at exon 1 <strong>of</strong> the<br />

MAP2 gene on the 5' UTR region, are conserved in<br />

the general population (Kalcheva et al., 1999).<br />

Implicated in<br />

Melanoma<br />

Note<br />

It has been found by Soltani et al. that primary<br />

melanomas that express the MAP2 gene have a<br />

lower rate <strong>of</strong> metastasis later on than primary<br />

melanomas that do not express the MAP2 gene. It<br />

has been proposed that MAP2 expression disrupts<br />

microtubule formation in cancer cells <strong>and</strong> interferes<br />

with cell cycle progression.<br />

Multiple sclerosis lesions<br />

Note<br />

Novel transcript <strong>of</strong> MAP2 that expresses exon 13 is<br />

shown to be up-regulated in multiple sclerosis<br />

lesions (Shafit-Zagardo et al., 1998). They<br />

proposed that this transcript is involved in<br />

remyelination <strong>of</strong> oligodendrocytes.<br />

Alzheimer's disease<br />

Note<br />

Abnormal hyperphosphorylation <strong>of</strong> several tau<br />

proteins, MAP1 <strong>and</strong> MAP2 have been implicated in<br />

leading to progressive degeneration <strong>and</strong> loss <strong>of</strong><br />

connectivity between neurons.<br />

Various diseases<br />

Note<br />

MAP2 expression is altered in response to various<br />

illnesses <strong>and</strong> thus is used as a marker in the<br />

diagnosis <strong>of</strong> many specific illnesses <strong>and</strong> especially<br />

as a marker <strong>of</strong> neuronal differentiation.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 88


MAP2 (microtubule-associated protein 2) Jayanthy A, Setaluri V<br />

References<br />

Shafit-Zagardo B, Kalcheva N. Making sense <strong>of</strong> the<br />

multiple MAP-2 transcripts <strong>and</strong> their role in the neuron. Mol<br />

Neurobiol. 1998 Apr;16(2):149-62<br />

Kalcheva N, Lachman HM, Shafit-Zagardo B. Survey for<br />

CAG repeat polymorphisms in the human MAP-2 gene.<br />

Psychiatr <strong>Gene</strong>t. 1999 Mar;9(1):43-6<br />

Shafit-Zagardo B, Kress Y, Zhao ML, Lee SC. A novel<br />

microtubule-associated protein-2 expressed in<br />

oligodendrocytes in multiple sclerosis lesions. J<br />

Neurochem. 1999 Dec;73(6):2531-7<br />

Teng J, Takei Y, Harada A, Nakata T, Chen J, Hirokawa N.<br />

Synergistic effects <strong>of</strong> MAP2 <strong>and</strong> MAP1B knockout in<br />

neuronal migration, dendritic outgrowth, <strong>and</strong> microtubule<br />

organization. J Cell Biol. 2001 Oct 1;155(1):65-76<br />

Crist<strong>of</strong>anilli M, Thanos S, Brosius J, Kindler S, Tiedge H.<br />

Neuronal MAP2 mRNA: species-dependent differential<br />

dendritic targeting competence. J Mol Biol. 2004 Aug<br />

20;341(4):927-34<br />

Dehmelt L, Halpain S. The MAP2/Tau family <strong>of</strong><br />

microtubule-associated proteins. Genome Biol.<br />

2005;6(1):204<br />

Soltani MH, Pichardo R, Song Z, Sangha N, Camacho F,<br />

Satyamoorthy K, Sangueza OP, Setaluri V. Microtubuleassociated<br />

protein 2, a marker <strong>of</strong> neuronal differentiation,<br />

induces mitotic defects, inhibits growth <strong>of</strong> melanoma cells,<br />

<strong>and</strong> predicts metastatic potential <strong>of</strong> cutaneous melanoma.<br />

Am J Pathol. 2005 Jun;166(6):1841-50<br />

Iqbal K, Liu F, Gong CX, Alonso Adel C, Grundke-Iqbal I.<br />

Mechanisms <strong>of</strong> tau-induced neurodegeneration. Acta<br />

Neuropathol. 2009 Jul;118(1):53-69<br />

This article should be referenced as such:<br />

Jayanthy A, Setaluri V. MAP2 (microtubule-associated<br />

protein 2). <strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012;<br />

16(2):85-89.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 89


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

<strong>Gene</strong> <strong>Section</strong><br />

Review<br />

MIR200C (microRNA 200c)<br />

Sarah Jurmeister, Stefan Uhlmann, Özgür Sahin<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

Division <strong>of</strong> Molecular Genome Analysis, German Cancer Research Center (DKFZ), Division <strong>of</strong><br />

Molecular Genome Analysis, Im Neuenheimer Feld 580, Heidelberg, Germany (SJ, SU, ÖS)<br />

Published in <strong>Atlas</strong> Database: August 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/<strong>Gene</strong>s/MIR200CID51054ch12p13.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI MIR200CID51054ch12p13.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Identity<br />

Other names: hsa-mir-200c, MIRN200C, mir-200c<br />

HGNC (Hugo): MIR200C<br />

Location: 12p13.31<br />

Local order: Based on Mapviewer <strong>Gene</strong>s on<br />

Sequence, genes flanking MIRN200C oriented<br />

from centromere to telomere on 12q13.31 are:<br />

- ATN1; Atrophin 1, 12q13.31<br />

- U7; U7 small nuclear 1, 12q13.31<br />

- C12orf57; Chromosome 12 open reading frame<br />

57, 12q13.31<br />

- PTPN6; Protein tyrosine phosphatase, nonreceptor<br />

type 6, 12q13.31<br />

- MIRN200C; microRNA 200c, 12q13.31<br />

- MIRN141; microRNA 141, 12q13.31<br />

- snoU89; small nucleolar RNA U89, 12q31.1<br />

- PHB2; Prohibitin 2, 12q31.1<br />

DNA/RNA<br />

Description<br />

miR-200c belongs to the miR-200 family, which<br />

consists <strong>of</strong> 5 members with two different<br />

chromosomal locations: miR-200c <strong>and</strong> miR-141 are<br />

A. Stem-loop structure <strong>of</strong> hsa-mir-200c (precursor miRNA).<br />

located on chromosome 12p13 <strong>and</strong> miR-200a, miR-<br />

200b <strong>and</strong> miR-429 are located on 1p36. This family<br />

is frequently downregulated upon the progression<br />

<strong>of</strong> tumors <strong>and</strong> maps to fragile chromosomal<br />

regions. Members <strong>of</strong> this family are important<br />

regulators <strong>of</strong> epithelial-to-mesenchymal transition<br />

(EMT) <strong>and</strong> metastasis.<br />

Transcription<br />

miRNAs are generally transcribed by RNA<br />

polymerase II.<br />

hsa-mir-200c (precursor miRNA)<br />

Accession: MI0000650<br />

Length: 68 bp<br />

Sequence:5'-CCCUCGUCUUACCCAGCAGUG<br />

UUUGGGUGCGGUUGGGAGUCUCUAAUACU<br />

GCCGGGUAAUGAUGGAGG-3'<br />

hsa-miR-200c (mature miRNA)<br />

Accession: MIMAT0000617<br />

Length: 23<br />

Sequence: 5'-UAAUACUGCCGGGUAAUGAU<br />

GGA-3'<br />

Pseudogene<br />

No reported pseudogenes.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 90


MIR200C (microRNA 200c) Jurmeister S, et al.<br />

B. The miR-200 family members. The human miR-200 family is located in two fragile chromosomal regions on 1p36.33 (200b,<br />

200a <strong>and</strong> 429) <strong>and</strong> 12p13.31 (200c <strong>and</strong> 141), respectively. It consists <strong>of</strong> two clusters based on seed sequence similarity: miR-<br />

200bc/429 (red) <strong>and</strong> 200a/141 (blue), distinguished by a single nucleotide change (U to C). (Source: Uhlmann et al., 2010,<br />

Oncogene).<br />

Protein<br />

Note<br />

microRNAs are not translated into amino acids.<br />

Mutations<br />

Note<br />

<strong>Gene</strong> mutations have not been described.<br />

Implicated in<br />

Bladder cancer<br />

Prognosis<br />

Loss <strong>of</strong> miR-200c expression was found to be<br />

associated with disease progression <strong>and</strong> poor<br />

outcome in 100 stage T1 bladder tumor patients<br />

(Wiklund et al., 2011).<br />

Oncogenesis<br />

Deep sequencing <strong>of</strong> nine bladder urothelial<br />

carcinomas <strong>and</strong> matched normal urothelium<br />

revealed that the miR-200c/141 cluster is<br />

upregulated in bladder cancer (Han et al., 2011).<br />

Consistently, a study comparing miRNA expression<br />

patterns by microarray in 27 invasive <strong>and</strong> 30<br />

superficial bladder tumors with 11 normal urothelia<br />

found that miR-200c was upregulated in bladder<br />

tumors compared to normal urothelium; however,<br />

expression <strong>of</strong> miR-200c was reduced in invasive<br />

compared to non-invasive tumors due to promoter<br />

hypermethylation (Wiklund et al., 2011).<br />

Furthermore, microarray miRNA analysis <strong>of</strong> 43<br />

primary tumors (10 colon, 10 bladder, 13 breast <strong>and</strong><br />

10 lung cancers) <strong>and</strong> matched lymph node<br />

metastases revealed that miR-200c <strong>and</strong> other miR-<br />

200 family members are downregulated in<br />

metastases compared to primary tumors (Baffa et<br />

al., 2009). This suggests that while miR-200c may<br />

have oncogenic function in bladder cancer, it<br />

interferes with invasion <strong>and</strong> metastasis.<br />

Mechanistically, miR-200c has been implicated in<br />

the regulation <strong>of</strong> epithelial-to-mesenchymal<br />

transition (EMT) in bladder cancer cells. A<br />

comparison <strong>of</strong> nine bladder cancer cell lines<br />

revealed a correlation between high expression <strong>of</strong><br />

miR-200c (<strong>and</strong> fellow miR-200 family member<br />

miR-200b) <strong>and</strong> epithelial phenotype (Adam et al.,<br />

2009). The same study also reported that miR-200c<br />

expression reverses resistance to anti-EGFR<br />

therapy in bladder cancer cell lines through<br />

targeting ERRFI-1.<br />

Breast cancer<br />

Oncogenesis<br />

A double-negative feedback loop between ZEB<br />

family transcription factors <strong>and</strong> the miR-200 family<br />

was shown to regulate EMT in different cell<br />

systems, including breast cancer cells (Burk et al.,<br />

2008). Moreover, expression <strong>of</strong> miR-200c was<br />

revealed to be activated by p53, resulting in<br />

induction <strong>of</strong> EMT in mammary epithelial cells upon<br />

loss <strong>of</strong> p53 (Chang et al., 2011). Loss <strong>of</strong> p53 was<br />

positively correlated with expression <strong>of</strong> ZEB1 <strong>and</strong><br />

negatively correlated with expression <strong>of</strong> miR-200c<br />

<strong>and</strong> E-Cadherin in 106 breast tumor specimens.<br />

miRNA microarray analysis <strong>of</strong> 43 primary tumors<br />

(10 colon, 10 bladder, 13 breast <strong>and</strong> 10 lung<br />

cancers) <strong>and</strong> matched lymph node metastases<br />

revealed that miR-200c <strong>and</strong> other miR-200 family<br />

members are downregulated in metastases<br />

compared to primary tumors (Baffa et al., 2009).<br />

Moreover, miR-200c <strong>and</strong> other miR-200 family<br />

members were shown to be underexpressed in the<br />

aggressive claudin-low subtype <strong>of</strong> breast cancer,<br />

which displays an EMT-like gene expression<br />

signature (Herschkowitz et al., 2011). In contrast,<br />

luminal breast cancers, which have a more<br />

epithelial-like phenotype <strong>and</strong> a better clinical<br />

prognosis, express high levels <strong>of</strong> miR-200c<br />

(Bockmeyer et al., 2011).<br />

Re-expression <strong>of</strong> the miR-200 family in aggressive<br />

breast cancer cells was shown to inhibit<br />

experimental lung metastasis (Ahmad et al., 2011).<br />

In contrast, another study reported that miR-200c<br />

promotes colonization <strong>of</strong> breast cancer cells<br />

(Dykxhoorn et al., 2009). In in vitro assays, miR-<br />

200c suppresses migration <strong>and</strong> invasion <strong>of</strong> breast<br />

cancer cells through various mechanisms, including<br />

targeting <strong>of</strong> ZEB1/ZEB2, PLCG1, moesin <strong>and</strong><br />

fibronectin (Korpal et al., 2008; Uhlmann et al.,<br />

2010; Howe et al., 2011).<br />

miR-200c also targets stem cell factors such as<br />

BMI1, <strong>and</strong> downregulation <strong>of</strong> miR-200c was shown<br />

to be characteristic <strong>of</strong> breast cancer stem cells<br />

(Shimono et al., 2009). Furthermore, miRNA<br />

microarray analysis revealed that miR-200c is<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 91


MIR200C (microRNA 200c) Jurmeister S, et al.<br />

downregulated in breast cancer cells with acquired<br />

resistance to cisplatin (Pogribny et al., 2010).<br />

Colorectal cancer<br />

Prognosis<br />

Kaplan-Meier survival analysis <strong>of</strong> 24 colorectal<br />

cancer patients suggested that high expression <strong>of</strong><br />

miR-200c was associated with decreased overall<br />

survival (Xi et al., 2006).<br />

Oncogenesis<br />

Analysis <strong>of</strong> miR-200c expression in 24 colorectal<br />

cancer biopsies <strong>and</strong> matched normal samples by<br />

qRT-PCR revealed that miR-200c is overexpressed<br />

in colorectal tumors compared to normal tissue (Xi<br />

et al., 2006). Furthermore, microarray miRNA<br />

analysis <strong>of</strong> 43 primary tumors (10 colon, 10<br />

bladder, 13 breast <strong>and</strong> 10 lung cancers) <strong>and</strong><br />

matched lymph node metastases revealed that miR-<br />

200c <strong>and</strong> other miR-200 family members are<br />

downregulated in metastases compared to primary<br />

tumors (Baffa et al., 2009).<br />

Endometrial cancer<br />

Disease<br />

Endometrial carcinoma; endometrial<br />

carcinosarcoma.<br />

Oncogenesis<br />

miRNA microarray analysis <strong>of</strong> four endometrial<br />

endometrioid carcinomas <strong>and</strong> four normal<br />

endometrial tissue samples showed that miR-200c<br />

<strong>and</strong> other miR-200 family members were<br />

overexpressed in cancerous compared to normal<br />

tissue (Lee et al., 2011). Inhibition <strong>of</strong> miR-200c<br />

decreased the growth <strong>of</strong> endometrial carcinoma<br />

cells (Lee et al., 2011). In contrast, an analysis <strong>of</strong><br />

miR-200c expression levels in five endometrial<br />

cancer <strong>and</strong> normal endometrial cell lines suggested<br />

that miR-200c is lower in cell lines derived from<br />

aggressive cancer compared to those derived from<br />

less aggressive cancer or normal endometrial<br />

epithelium (Cochrane et al., 2009). Restoration <strong>of</strong><br />

miR-200c expression in aggressive endometrial<br />

cancer cells reduced their migration <strong>and</strong> invasion<br />

<strong>and</strong> increased their sensitivity to microtubuletargeting<br />

chemotherapeutic agents, at least in part<br />

through targeting TUBB3 (Cochrane et al., 2009;<br />

Cochrane et al., 2010; Howe et al., 2011). In a panel<br />

<strong>of</strong> 23 endometrial carcinosarcomas, which are<br />

composed <strong>of</strong> mixed populations <strong>of</strong> epithelial-like<br />

<strong>and</strong> mesenchymal-like cells, miR-200c <strong>and</strong> other<br />

miR-200 family members were found to be<br />

downregulated in the mesenchymal components <strong>of</strong><br />

the tumors compared to the epithelial components<br />

(Castilla et al., 2011); this is consistent with the<br />

established role <strong>of</strong> the miR-200 family in<br />

suppression <strong>of</strong> epithelial-to-mesenchymal<br />

transition.<br />

Esophageal cancer<br />

Prognosis<br />

In a panel <strong>of</strong> 98 esophageal cancer patients treated<br />

with preoperative chemotherapy <strong>and</strong> surgery,<br />

expression <strong>of</strong> miR-200c was associated with<br />

shortened overall survival <strong>and</strong> poor response to<br />

chemotherapy, potentially through upregulation <strong>of</strong><br />

the Akt signaling pathway (Hamano et al., 2011).<br />

Oncogenesis<br />

qRT-PCR analysis <strong>of</strong> miR-200 expression levels in<br />

17 patients with Barrett's esophagus <strong>and</strong> 20 patients<br />

with esophageal adenocarcinoma indicated that<br />

miR-200c is downregulated during cancer<br />

progression from normal epithelium through<br />

Barrett's esophagus to esophageal adenocarcinoma<br />

(Smith et al., 2011). In contrast, another study on<br />

98 esophageal cancer patients treated with<br />

preoperative chemotherapy <strong>and</strong> surgery found that<br />

miR-200c was expressed at higher levels in the<br />

tumor than in normal tissue (Hamano et al., 2011).<br />

Germ cell tumors<br />

Disease<br />

Germinoma; yolk sac tumors.<br />

Oncogenesis<br />

Diagnosis. Microarray analysis <strong>of</strong> 25 germ cell<br />

tumors <strong>and</strong> subsequent validation by qRT-PCR in<br />

10 independent samples identified miR-200c as<br />

overexpressed in yolk sac tumors compared to<br />

germinoma (Murray et al., 2010).<br />

Head <strong>and</strong> neck cancer<br />

Disease<br />

Squamous cell carcinoma; spindle cell carcinoma.<br />

Oncogenesis<br />

miR-200c was significantly downregulated in a<br />

panel <strong>of</strong> 30 spindle cell carcinomas (which display<br />

a mesenchymal-like phenotype) compared to<br />

normal mucosa as determined by qRT-PC (Zidar et<br />

al., 2011). In contrast, expression levels <strong>of</strong> miR-<br />

200c in 30 squamous cell carcinomas were<br />

comparable to normal tissue.<br />

Liver cancer<br />

Oncogenesis<br />

Diagnosis. Due to its low expression in liver<br />

compared to other tissues, miR-200c has been<br />

suggested as a biomarker to distinguish<br />

hepatocellular carcinoma from liver metastases<br />

(Barshack et al., 2010).<br />

miRNA microarray analysis <strong>of</strong> 92 primary<br />

hepatocellular carcinomas <strong>and</strong> 9 hepatocellular<br />

carcinoma cell lines identified miR-200c as a<br />

microRNA that is upregulated by p53 (Kim et al.,<br />

2011). Increased expression <strong>of</strong> miR-200c results in<br />

downregulation <strong>of</strong> transcriptional repressors ZEB1<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 92


MIR200C (microRNA 200c) Jurmeister S, et al.<br />

<strong>and</strong> ZEB2, suggesting a role for p53-mediated<br />

regulation <strong>of</strong> miR-200c in suppression <strong>of</strong> EMT.<br />

miR-200c was reported to be underexpressed in<br />

benign liver tumors compared to hepatocellular<br />

carcinoma (Ladeiro et al., 2008); miR-200c levels<br />

were determined by qRT-PCR in two sets <strong>of</strong> tumors<br />

(first set: 18 benign tumors, 28 hepatocellular<br />

carcinomas; second set: 12 benign tumors, 22<br />

hepatocellular carcinomas).<br />

Lung cancer<br />

Prognosis<br />

qRT-PCR analysis <strong>of</strong> miR-200c expression levels<br />

in 70 non-small cell lung cancer (NSCLC) patients<br />

revealed that high expression <strong>of</strong> miR-200c was<br />

associated with reduced overall survival (Liu et al.,<br />

2011).<br />

Oncogenesis<br />

Treatment <strong>of</strong> immortalized human bronchial<br />

epithelial cells with tobacco carcinogens was shown<br />

to induce an EMT-like phenotype <strong>and</strong> stem-cell like<br />

properties (Tellez et al., 2011). Quantification <strong>of</strong><br />

miRNA levels by qRT-PCR in combination with<br />

bisulfite sequencing <strong>and</strong> chromatin<br />

immunoprecipitation revealed that these changes<br />

are accompanied by epigenetic silencing <strong>of</strong> miR-<br />

200c <strong>and</strong> other EMT-regulating microRNAs,<br />

suggesting that loss <strong>of</strong> miR-200c contributes to<br />

transformation <strong>of</strong> lung epithelial cells. In contrast,<br />

miRNA microarray analysis <strong>of</strong> six NSCLCs <strong>and</strong><br />

matched adjacent normal tissue revealed that miR-<br />

200c is upregulated in NSCLC compared to healthy<br />

tissue (Liu et al., 2011). This finding was further<br />

validated in 70 lung carcinomas <strong>and</strong> matched<br />

normal tissue by qRT-PCR.<br />

Several studies have reported that miR-200c can<br />

repress invasion <strong>and</strong> metastasis <strong>of</strong> lung cancer cells.<br />

Firstly, low expression <strong>of</strong> miR-200c <strong>and</strong> other miR-<br />

200 family members was associated with increased<br />

metastatic potential in a syngeneic mouse model <strong>of</strong><br />

lung adenocarcinoma, <strong>and</strong> re-expression <strong>of</strong> miR-<br />

200 family members in these cell lines prevented<br />

EMT <strong>and</strong> metastasis (Gibbons et al., 2009).<br />

Secondly, miR-200c was shown to be<br />

downregulated by promoter hypermethylation in<br />

invasive NSCLC cell lines, <strong>and</strong> re-expression <strong>of</strong><br />

miR-200c reduced the invasive potential <strong>of</strong> these<br />

cell lines (Ceppi et al., 2010). Furthermore,<br />

microarray miRNA analysis <strong>of</strong> 43 primary tumors<br />

(10 colon, 10 bladder, 13 breast <strong>and</strong> 10 lung<br />

cancers) <strong>and</strong> matched lymph node metastases<br />

revealed that miR-200c <strong>and</strong> other miR-200 family<br />

members are downregulated in metastases<br />

compared to primary tumors (Baffa et al., 2009).<br />

Finally, low expression <strong>of</strong> miR-200c in 69 primary<br />

lung tumors was correlated with lymph node<br />

metastases (Ceppi et al., 2010).<br />

Mechanistically, the Notch lig<strong>and</strong> Jagged2 was<br />

shown to suppress expression <strong>of</strong> miR-200 family<br />

members, resulting in induction <strong>of</strong> EMT <strong>and</strong><br />

increased metastatic potential (Yang et al., 2011).<br />

Moreover, miR-200c <strong>and</strong> fellow miR-200 family<br />

member miR-200b target VEGFR, which also<br />

contributes to invasion <strong>and</strong> metastasis (Roybal et<br />

al., 2011).<br />

Malignant pleural mesothelioma<br />

Oncogenesis<br />

Diagnosis. miR-200c has been proposed as a<br />

biomarker to distinguish malignant pleural<br />

mesothelioma from lung adenocarcinoma <strong>and</strong> lung<br />

metastases <strong>of</strong> other carcinomas. miRNA microarray<br />

expression pr<strong>of</strong>iling <strong>of</strong> 10 lung adenocarcinomas<br />

<strong>and</strong> 15 mesotheliomas revealed that miR-200c is<br />

reduced in mesothelioma (Gee et al., 2010). This<br />

result was further confirmed by qRT-PCR in a set<br />

<strong>of</strong> 100 mesotheliomas <strong>and</strong> 32 lung<br />

adenocarcinomas. Similarly, microRNA microarray<br />

analysis <strong>of</strong> 7 malignant pleural mesotheliomas <strong>and</strong><br />

97 carcinomas <strong>of</strong> various origins also identified<br />

miR-200c as underexpressed in mesotheliomas<br />

compared to the carcinoma samples, <strong>and</strong><br />

differential expression levels <strong>of</strong> miR-200c <strong>and</strong> two<br />

other microRNAs could successfully be used to<br />

distinguish between malignant pleural<br />

mesothelioma <strong>and</strong> other types <strong>of</strong> cancer (Benjamin<br />

et al., 2010).<br />

Melanoma<br />

Oncogenesis<br />

Analysis <strong>of</strong> miR-200c expression levels in a panel<br />

<strong>of</strong> 10 melanoma cell lines by qRT-PCR showed that<br />

miR-200c is overexpressed in many <strong>of</strong> these cell<br />

lines compared to normal melanocytes (Elson-<br />

Schwab et al., 2010). Overexpression <strong>of</strong> miR-200c<br />

in melanoma cell lines resulted in a shift towards<br />

amoeboid type <strong>of</strong> migration, possibly through<br />

targeting <strong>of</strong> MARCKS.<br />

Ovarian cancer<br />

Prognosis<br />

High expression <strong>of</strong> miR-200c was found to be<br />

correlated with decreased progression-free <strong>and</strong><br />

overall survival in a panel <strong>of</strong> 20 serous ovarian<br />

cancer patients (Nam et al., 2008). In contrast, a<br />

study investigating microRNA expression pr<strong>of</strong>iles<br />

in a total <strong>of</strong> 144 patients with epithelial ovarian<br />

cancer found that low expression <strong>of</strong> miR-200c was<br />

associated with increased progression-free <strong>and</strong><br />

overall survival (Marchini et al., 2011). Similarly,<br />

high expression <strong>of</strong> miR-200c was correlated with<br />

response to chemotherapy <strong>and</strong> decreased risk <strong>of</strong><br />

disease recurrence in a panel <strong>of</strong> 57 patients with<br />

serous ovarian carcinoma (Leskela et al., 2010).<br />

Oncogenesis<br />

miR-200c was found to be overexpressed in a panel<br />

<strong>of</strong> 20 serous ovarian carcinomas compared to 8<br />

normal ovarian tissues by miRNA microarray<br />

analysis (Nam et al., 2008). Similarly, increased<br />

expression <strong>of</strong> miR-200c compared to normal ovary<br />

(n=15) was reported for serous, endometroid <strong>and</strong><br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 93


MIR200C (microRNA 200c) Jurmeister S, et al.<br />

clear cell ovarian carcinoma in a series <strong>of</strong> 69 cancer<br />

specimens.<br />

Expression <strong>of</strong> miR-200c was correlated with E-<br />

Cadherin levels in 36 primary ovarian carcinomas<br />

(Park et al., 2008). The regulatory effect <strong>of</strong> miR-<br />

200c on EMT has been shown to be mediated<br />

through targeting <strong>of</strong> ZEB1 <strong>and</strong> ZEB2, which<br />

transcriptionally repress E-Cadherin (Gregory et al.,<br />

2008; Korpal et al., 2008; Park et al., 2008). Reexpression<br />

<strong>of</strong> miR-200c in aggressive ovarian<br />

cancer cell lines was shown to reduce their<br />

migratory capacity; however, this effect appears to<br />

be independent <strong>of</strong> E-Cadherin expression<br />

(Cochrane et al., 2010). Furthermore, forced<br />

expression <strong>of</strong> miR-200c has been reported to<br />

sensitize ovarian cancer cells to paclitaxel treatment<br />

due to downregulation <strong>of</strong> miR-200c target gene<br />

TUBB3 (Cochrane et al., 2009; Cochrane et al.,<br />

2010).<br />

miR-200c was also shown to be downregulated in a<br />

subpopulation <strong>of</strong> the ovarian cancer cell line<br />

OVCAR3 expressing the cancer stem cell marker<br />

CD133 (Guo et al., 2011).<br />

Pancreatic cancer<br />

Prognosis<br />

In a panel <strong>of</strong> 99 pancreatic cancer patients, high<br />

expression <strong>of</strong> miR-200c was associated with<br />

increased overall survival (Yu et al., 2010).<br />

Oncogenesis<br />

Downregulation <strong>of</strong> miR-200c <strong>and</strong> other miR-200<br />

family members has been observed in gemcitabineresistant<br />

pancreatic cancer cell lines (Li et al., 2009;<br />

Ali et al., 2010). miR-200c has also been suggested<br />

to have a stemness-inhibiting function in pancreatic<br />

cancer cells through targeting <strong>of</strong> stem cell factors<br />

such as Bmi1 (Wellner et al., 2009).<br />

A double-negative feedback loop between ZEB<br />

family transcription factors <strong>and</strong> the miR-200 family<br />

was shown to regulate EMT in different cell<br />

systems, including pancreatic cancer cells (Burk et<br />

al., 2008). Consistently, high expression <strong>of</strong> miR-<br />

200c was shown to be associated with decreased<br />

invasive behavior in a panel <strong>of</strong> six pancreatic<br />

cancer cell lines, <strong>and</strong> miR-200c expression was<br />

correlated with E-Cadherin levels in pancreatic<br />

cancer specimens <strong>and</strong> cell lines (Yu et al., 2010).<br />

Overexpression <strong>of</strong> miR-200c in pancreatic cancer<br />

cell lines resulted in upregulation <strong>of</strong> E-Cadherin<br />

expression <strong>and</strong> reduced invasion but stimulated<br />

proliferation.<br />

miRNA expression pr<strong>of</strong>iling <strong>of</strong> various stages in a<br />

mouse model <strong>of</strong> multistep tumorigenesis <strong>of</strong> the<br />

pancreas revealed that miR-200c is downregulated<br />

in metastases <strong>and</strong> metastasis-like tumors (Olson et<br />

al., 2009). Moreover, miR-200c also targets<br />

components <strong>of</strong> the Notch pathway, which is<br />

aberrantly activated in pancreatic cancer (Brabletz<br />

et al., 2011). Undifferentiated, aggressive<br />

pancreatic adenocarcinomas were shown to have<br />

higher expression <strong>of</strong> ZEB1 <strong>and</strong> Notch pathway<br />

components <strong>and</strong> lower expression <strong>of</strong> miR-200c<br />

compared to differentiated tumors.<br />

In contrast to the studies described above, which<br />

suggest a metastasis-suppressing function for miR-<br />

200c in pancreatic cancer, a comparison <strong>of</strong> 16<br />

pancreatic ductal adenocarcinoma cell lines found<br />

that miR-200c expression was upregulated in the<br />

highly metastatic cell lines (Mees et al., 2010).<br />

Prostate cancer<br />

Oncogenesis<br />

Prostate cancer cells with EMT phenotype were<br />

found to have stem-cell like properties <strong>and</strong> express<br />

low levels <strong>of</strong> miR-200 family members (Kong et<br />

al., 2010). Overexpression <strong>of</strong> miR-200c reversed<br />

EMT <strong>and</strong> stem-cell like properties, in part due to<br />

targeting <strong>of</strong> Notch-1. miR-200c was also shown to<br />

target the Notch lig<strong>and</strong> Jagged1, resulting in<br />

decreased proliferation <strong>of</strong> metastatic prostate cancer<br />

cells (Vallejo et al., 2011).<br />

Renal cancer<br />

Disease<br />

Clear cell carcinoma (CCC); Chromophobe renal<br />

cell carcinoma (ChCC).<br />

Oncogenesis<br />

Diagnosis. miR-200c has been found to be<br />

specifically expressed in chromophobe renal cell<br />

carcinoma <strong>and</strong> has been suggested as one <strong>of</strong> a set <strong>of</strong><br />

microRNAs that can be used to distinguish between<br />

renal cell carcinoma subtypes (Fridman et al.,<br />

2010).<br />

miR-200c was found to be significantly<br />

downregulated in clear cell carcinoma compared to<br />

normal kidney in a panel <strong>of</strong> 16 CCCs, 4 ChCCs <strong>and</strong><br />

6 normal kidneys both by microarray analysis <strong>and</strong><br />

by qRT-PCR (Nakada et al., 2008). Furthermore,<br />

miR-200c expression was inversely correlated with<br />

expression <strong>of</strong> its target gene ZEB1 in these<br />

specimens. The downregulation <strong>of</strong> miR-200c in<br />

CCC was also confirmed by a second study<br />

comparing a total <strong>of</strong> 25 clear cell carcinomas <strong>and</strong><br />

matched adjacent normal tissue (Liu et al., 2010).<br />

Thyroid carcinoma<br />

Oncogenesis<br />

The expression <strong>of</strong> miR-200 family members,<br />

including miR-200c, was found to be<br />

downregulated in undifferentiated, aggressive<br />

anaplastic thyroid carcinoma compared to both<br />

normal tissue <strong>and</strong> well-differentiated papillary <strong>and</strong><br />

follicular thyroid carcinomas (Braun et al., 2010).<br />

Overexpression <strong>of</strong> the miR-200 family induced<br />

mesenchymal-to-epithelial transition <strong>and</strong> reduced<br />

invasion <strong>of</strong> ATC cells.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 94


MIR200C (microRNA 200c) Jurmeister S, et al.<br />

Various cancers<br />

References<br />

miR-200c target genes regulate numerous processes involved in cancer development <strong>and</strong> progression.<br />

Xi Y, Formentini A, Chien M, Weir DB, Russo JJ, Ju J,<br />

Kornmann M, Ju J. Prognostic Values <strong>of</strong> microRNAs in<br />

Colorectal Cancer. Biomark Insights. 2006;2:113-121<br />

Burk U, Schubert J, Wellner U, Schmalh<strong>of</strong>er O, Vincan E,<br />

Spaderna S, Brabletz T. A reciprocal repression between<br />

ZEB1 <strong>and</strong> members <strong>of</strong> the miR-200 family promotes EMT<br />

<strong>and</strong> invasion in cancer cells. EMBO Rep. 2008<br />

Jun;9(6):582-9<br />

Gregory PA, Bert AG, Paterson EL, Barry SC, Tsykin A,<br />

Farshid G, Vadas MA, Khew-Goodall Y, Goodall GJ. The<br />

miR-200 family <strong>and</strong> miR-205 regulate epithelial to<br />

mesenchymal transition by targeting ZEB1 <strong>and</strong> SIP1. Nat<br />

Cell Biol. 2008 May;10(5):593-601<br />

Korpal M, Lee ES, Hu G, Kang Y. The miR-200 family<br />

inhibits epithelial-mesenchymal transition <strong>and</strong> cancer cell<br />

migration by direct targeting <strong>of</strong> E-cadherin transcriptional<br />

repressors ZEB1 <strong>and</strong> ZEB2. J Biol Chem. 2008 May<br />

30;283(22):14910-4<br />

Ladeiro Y, Couchy G, Balabaud C, Bioulac-Sage P,<br />

Pelletier L, Rebouissou S, Zucman-Rossi J. MicroRNA<br />

pr<strong>of</strong>iling in hepatocellular tumors is associated with clinical<br />

features <strong>and</strong> oncogene/tumor suppressor gene mutations.<br />

Hepatology. 2008 Jun;47(6):1955-63<br />

Nakada C, Matsuura K, Tsukamoto Y, Tanigawa M,<br />

Yoshimoto T, Narimatsu T, Nguyen LT, Hijiya N, Uchida T,<br />

Sato F, Mimata H, Seto M, Moriyama M. Genome-wide<br />

microRNA expression pr<strong>of</strong>iling in renal cell carcinoma:<br />

significant down-regulation <strong>of</strong> miR-141 <strong>and</strong> miR-200c. J<br />

Pathol. 2008 Dec;216(4):418-27<br />

Nam EJ, Yoon H, Kim SW, Kim H, Kim YT, Kim JH, Kim<br />

JW, Kim S. MicroRNA expression pr<strong>of</strong>iles in serous<br />

ovarian carcinoma. Clin Cancer Res. 2008 May<br />

1;14(9):2690-5<br />

Park SM, Gaur AB, Lengyel E, Peter ME. The miR-200<br />

family determines the epithelial phenotype <strong>of</strong> cancer cells<br />

by targeting the E-cadherin repressors ZEB1 <strong>and</strong> ZEB2.<br />

<strong>Gene</strong>s Dev. 2008 Apr 1;22(7):894-907<br />

Adam L, Zhong M, Choi W, Qi W, Nicoloso M, Arora A,<br />

Calin G, Wang H, Siefker-Radtke A, McConkey D, Bar-Eli<br />

M, Dinney C. miR-200 expression regulates epithelial-tomesenchymal<br />

transition in bladder cancer cells <strong>and</strong><br />

reverses resistance to epidermal growth factor receptor<br />

therapy. Clin Cancer Res. 2009 Aug 15;15(16):5060-72<br />

Baffa R, Fassan M, Volinia S, O'Hara B, Liu CG, Palazzo<br />

JP, Gardiman M, Rugge M, Gomella LG, Croce CM,<br />

Rosenberg A. MicroRNA expression pr<strong>of</strong>iling <strong>of</strong> human<br />

metastatic cancers identifies cancer gene targets. J Pathol.<br />

2009 Oct;219(2):214-21<br />

Cochrane DR, Spoelstra NS, Howe EN, Nordeen SK,<br />

Richer JK. MicroRNA-200c mitigates invasiveness <strong>and</strong><br />

restores sensitivity to microtubule-targeting<br />

chemotherapeutic agents. Mol Cancer Ther. 2009<br />

May;8(5):1055-66<br />

Dykxhoorn DM, Wu Y, Xie H, Yu F, Lal A, Petrocca F,<br />

Martinvalet D, Song E, Lim B, Lieberman J. miR-200<br />

enhances mouse breast cancer cell colonization to form<br />

distant metastases. PLoS One. 2009 Sep 29;4(9):e7181<br />

Gibbons DL, Lin W, Creighton CJ, Rizvi ZH, Gregory PA,<br />

Goodall GJ, Thilaganathan N, Du L, Zhang Y, Pertsemlidis<br />

A, Kurie JM. Contextual extracellular cues promote tumor<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 95


MIR200C (microRNA 200c) Jurmeister S, et al.<br />

cell EMT <strong>and</strong> metastasis by regulating miR-200 family<br />

expression. <strong>Gene</strong>s Dev. 2009 Sep 15;23(18):2140-51<br />

Li Y, V<strong>and</strong>enBoom TG 2nd, Kong D, Wang Z, Ali S, Philip<br />

PA, Sarkar FH. Up-regulation <strong>of</strong> miR-200 <strong>and</strong> let-7 by<br />

natural agents leads to the reversal <strong>of</strong> epithelial-tomesenchymal<br />

transition in gemcitabine-resistant<br />

pancreatic cancer cells. Cancer Res. 2009 Aug<br />

15;69(16):6704-12<br />

Olson P, Lu J, Zhang H, Shai A, Chun MG, Wang Y, Libutti<br />

SK, Nakakura EK, Golub TR, Hanahan D. MicroRNA<br />

dynamics in the stages <strong>of</strong> tumorigenesis correlate with<br />

hallmark capabilities <strong>of</strong> cancer. <strong>Gene</strong>s Dev. 2009 Sep<br />

15;23(18):2152-65<br />

Shimono Y, Zabala M, Cho RW, Lobo N, Dalerba P, Qian<br />

D, Diehn M, Liu H, Panula SP, Chiao E, Dirbas FM, Somlo<br />

G, Pera RA, Lao K, Clarke MF. Downregulation <strong>of</strong> miRNA-<br />

200c links breast cancer stem cells with normal stem cells.<br />

Cell. 2009 Aug 7;138(3):592-603<br />

Wellner U, Schubert J, Burk UC, Schmalh<strong>of</strong>er O, Zhu F,<br />

Sonntag A, Waldvogel B, Vannier C, Darling D, zur<br />

Hausen A, Brunton VG, Morton J, Sansom O, Schüler J,<br />

Stemmler MP, Herzberger C, Hopt U, Keck T, Brabletz S,<br />

Brabletz T. The EMT-activator ZEB1 promotes<br />

tumorigenicity by repressing stemness-inhibiting<br />

microRNAs. Nat Cell Biol. 2009 Dec;11(12):1487-95<br />

Ali S, Ahmad A, Banerjee S, Padhye S, Dominiak K,<br />

Schaffert JM, Wang Z, Philip PA, Sarkar FH. Gemcitabine<br />

sensitivity can be induced in pancreatic cancer cells<br />

through modulation <strong>of</strong> miR-200 <strong>and</strong> miR-21 expression by<br />

curcumin or its analogue CDF. Cancer Res. 2010 May<br />

1;70(9):3606-17<br />

Benjamin H, Lebanony D, Rosenwald S, Cohen L, Gibori<br />

H, Barabash N, Ashkenazi K, Goren E, Meiri E,<br />

Morgenstern S, Perelman M, Barshack I, Goren Y,<br />

Edmonston TB, Chajut A, Aharonov R, Bentwich Z,<br />

Rosenfeld N, Cohen D. A diagnostic assay based on<br />

microRNA expression accurately identifies malignant<br />

pleural mesothelioma. J Mol Diagn. 2010 Nov;12(6):771-9<br />

Barshack I, Meiri E, Rosenwald S, Lebanony D, Bronfeld<br />

M, Aviel-Ronen S, Rosenblatt K, Polak-Charcon S,<br />

Leizerman I, Ezagouri M, Zepeniuk M, Shabes N, Cohen<br />

L, Tabak S, Cohen D, Bentwich Z, Rosenfeld N.<br />

Differential diagnosis <strong>of</strong> hepatocellular carcinoma from<br />

metastatic tumors in the liver using microRNA expression.<br />

Int J Biochem Cell Biol. 2010 Aug;42(8):1355-62<br />

Braun J, Hoang-Vu C, Dralle H, Hüttelmaier S.<br />

Downregulation <strong>of</strong> microRNAs directs the EMT <strong>and</strong><br />

invasive potential <strong>of</strong> anaplastic thyroid carcinomas.<br />

Oncogene. 2010 Jul 22;29(29):4237-44<br />

Ceppi P, Mudduluru G, Kumarswamy R, Rapa I, Scagliotti<br />

GV, Papotti M, Allgayer H. Loss <strong>of</strong> miR-200c expression<br />

induces an aggressive, invasive, <strong>and</strong> chemoresistant<br />

phenotype in non-small cell lung cancer. Mol Cancer Res.<br />

2010 Sep;8(9):1207-16<br />

Cochrane DR, Howe EN, Spoelstra NS, Richer JK. Loss <strong>of</strong><br />

miR-200c: A Marker <strong>of</strong> Aggressiveness <strong>and</strong><br />

Chemoresistance in Female Reproductive Cancers. J<br />

Oncol. 2010;2010:821717<br />

Elson-Schwab I, Lorentzen A, Marshall CJ. MicroRNA-200<br />

family members differentially regulate morphological<br />

plasticity <strong>and</strong> mode <strong>of</strong> melanoma cell invasion. PLoS One.<br />

2010 Oct 4;5(10)<br />

Fridman E, Dotan Z, Barshack I, David MB, Dov A, Tabak<br />

S, Zion O, Benjamin S, Benjamin H, Kuker H, Avivi C,<br />

Rosenblatt K, Polak-Charcon S, Ramon J, Rosenfeld N,<br />

Spector Y. Accurate molecular classification <strong>of</strong> renal<br />

tumors using microRNA expression. J Mol Diagn. 2010<br />

Sep;12(5):687-96<br />

Gee GV, Koestler DC, Christensen BC, Sugarbaker DJ,<br />

Ugolini D, Ivaldi GP, Resnick MB, Houseman EA, Kelsey<br />

KT, Marsit CJ. Downregulated microRNAs in the<br />

differential diagnosis <strong>of</strong> malignant pleural mesothelioma.<br />

Int J Cancer. 2010 Dec 15;127(12):2859-69<br />

Kong D, Banerjee S, Ahmad A, Li Y, Wang Z, Sethi S,<br />

Sarkar FH. Epithelial to mesenchymal transition is<br />

mechanistically linked with stem cell signatures in prostate<br />

cancer cells. PLoS One. 2010 Aug 27;5(8):e12445<br />

Leskelä S, Le<strong>and</strong>ro-García LJ, Mendiola M, Barriuso J,<br />

Inglada-Pérez L, Muñoz I, Martínez-Delgado B, Redondo<br />

A, de Santiago J, Robledo M, Hardisson D, Rodríguez-<br />

Antona C. The miR-200 family controls beta-tubulin III<br />

expression <strong>and</strong> is associated with paclitaxel-based<br />

treatment response <strong>and</strong> progression-free survival in<br />

ovarian cancer patients. Endocr Relat Cancer.<br />

2011;18(1):85-95<br />

Liu H, Brannon AR, Reddy AR, Alexe G, Seiler MW,<br />

Arreola A, Oza JH, Yao M, Juan D, Liou LS, Ganesan S,<br />

Levine AJ, Rathmell WK, Bhanot GV. Identifying mRNA<br />

targets <strong>of</strong> microRNA dysregulated in cancer: with<br />

application to clear cell Renal Cell Carcinoma. BMC Syst<br />

Biol. 2010 Apr 27;4:51<br />

Mees ST, Mardin WA, Wendel C, Baeumer N, Willscher E,<br />

Senninger N, Schleicher C, Colombo-Benkmann M, Haier<br />

J. EP300--a miRNA-regulated metastasis suppressor gene<br />

in ductal adenocarcinomas <strong>of</strong> the pancreas. Int J Cancer.<br />

2010 Jan 1;126(1):114-24<br />

Murray MJ, Saini HK, van Dongen S, Palmer RD,<br />

Muralidhar B, Pett MR, Piipari M, Thornton CM, Nicholson<br />

JC, Enright AJ, Coleman N. The two most common<br />

histological subtypes <strong>of</strong> malignant germ cell tumour are<br />

distinguished by global microRNA pr<strong>of</strong>iles, associated with<br />

differential transcription factor expression. Mol Cancer.<br />

2010 Nov 8;9:290<br />

Pogribny IP, Filkowski JN, Tryndyak VP, Golubov A,<br />

Shpyleva SI, Kovalchuk O. Alterations <strong>of</strong> microRNAs <strong>and</strong><br />

their targets are associated with acquired resistance <strong>of</strong><br />

MCF-7 breast cancer cells to cisplatin. Int J Cancer. 2010<br />

Oct 15;127(8):1785-94<br />

Uhlmann S, Zhang JD, Schwäger A, Mannsperger H,<br />

Riazalhosseini Y, Burmester S, Ward A, Korf U, Wiemann<br />

S, Sahin O. miR-200bc/429 cluster targets PLCgamma1<br />

<strong>and</strong> differentially regulates proliferation <strong>and</strong> EGF-driven<br />

invasion than miR-200a/141 in breast cancer. Oncogene.<br />

2010 Jul 29;29(30):4297-306<br />

Yu J, Ohuchida K, Mizumoto K, Sato N, Kayashima T,<br />

Fujita H, Nakata K, Tanaka M. MicroRNA, hsa-miR-200c,<br />

is an independent prognostic factor in pancreatic cancer<br />

<strong>and</strong> its upregulation inhibits pancreatic cancer invasion but<br />

increases cell proliferation. Mol Cancer. 2010 Jun 28;9:169<br />

Ahmad A, Aboukameel A, Kong D, Wang Z, Sethi S, Chen<br />

W, Sarkar FH, Raz A. Phosphoglucose<br />

isomerase/autocrine motility factor mediates epithelialmesenchymal<br />

transition regulated by miR-200 in breast<br />

cancer cells. Cancer Res. 2011 May 1;71(9):3400-9<br />

Bockmeyer CL, Christgen M, Müller M, Fischer S, Ahrens<br />

P, Länger F, Kreipe H, Lehmann U. MicroRNA pr<strong>of</strong>iles <strong>of</strong><br />

healthy basal <strong>and</strong> luminal mammary epithelial cells are<br />

distinct <strong>and</strong> reflected in different breast cancer subtypes.<br />

Breast Cancer Res Treat. 2011 Mar 17;<br />

Brabletz S, Bajdak K, Meidh<strong>of</strong> S, Burk U, Niedermann G,<br />

Firat E, Wellner U, Dimmler A, Faller G, Schubert J,<br />

Brabletz T. The ZEB1/miR-200 feedback loop controls<br />

Notch signalling in cancer cells. EMBO J. 2011 Feb<br />

16;30(4):770-82<br />

Castilla MÁ, Moreno-Bueno G, Romero-Pérez L, Van De<br />

Vijver K, Biscuola M, López-García MÁ, Prat J, Matías-<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 96


MIR200C (microRNA 200c) Jurmeister S, et al.<br />

Guiu X, Cano A, Oliva E, Palacios J. Micro-RNA signature<br />

<strong>of</strong> the epithelial-mesenchymal transition in endometrial<br />

carcinosarcoma. J Pathol. 2011 Jan;223(1):72-80<br />

Chang CJ, Chao CH, Xia W, Yang JY, Xiong Y, Li CW, Yu<br />

WH, Rehman SK, Hsu JL, Lee HH, Liu M, Chen CT, Yu D,<br />

Hung MC. p53 regulates epithelial-mesenchymal transition<br />

<strong>and</strong> stem cell properties through modulating miRNAs. Nat<br />

Cell Biol. 2011 Mar;13(3):317-23<br />

Guo R, Wu Q, Liu F, Wang Y. Description <strong>of</strong> the CD133+<br />

subpopulation <strong>of</strong> the human ovarian cancer cell line<br />

OVCAR3. Oncol Rep. 2011 Jan;25(1):141-6<br />

Hamano R, Miyata H, Yamasaki M, Kurokawa Y, Hara J,<br />

Moon JH, Nakajima K, Takiguchi S, Fujiwara Y, Mori M,<br />

Doki Y. Overexpression <strong>of</strong> miR-200c induces<br />

chemoresistance in esophageal cancers mediated through<br />

activation <strong>of</strong> the Akt signaling pathway. Clin Cancer Res.<br />

2011 May 1;17(9):3029-38<br />

Han Y, Chen J, Zhao X, Liang C, Wang Y, Sun L, Jiang Z,<br />

Zhang Z, Yang R, Chen J, Li Z, Tang A, Li X, Ye J, Guan<br />

Z, Gui Y, Cai Z. MicroRNA expression signatures <strong>of</strong><br />

bladder cancer revealed by deep sequencing. PLoS One.<br />

2011 Mar 28;6(3):e18286<br />

Herschkowitz JI, Zhao W, Zhang M, Usary J, Murrow G,<br />

Edwards D, Knezevic J, Greene SB, Darr D, Troester MA,<br />

Hilsenbeck SG, Medina D, Perou CM, Rosen JM.<br />

Comparative oncogenomics identifies breast tumors<br />

enriched in functional tumor-initiating cells. Proc Natl Acad<br />

Sci U S A. 2011 Jun 1;<br />

Howe EN, Cochrane DR, Richer JK. Targets <strong>of</strong> miR-200c<br />

mediate suppression <strong>of</strong> cell motility <strong>and</strong> anoikis resistance.<br />

Breast Cancer Res. 2011 Apr 18;13(2):R45<br />

Kim T, Veronese A, Pichiorri F, Lee TJ, Jeon YJ, Volinia S,<br />

Pineau P, Marchio A, Palatini J, Suh SS, Alder H, Liu CG,<br />

Dejean A, Croce CM. p53 regulates epithelialmesenchymal<br />

transition through microRNAs targeting<br />

ZEB1 <strong>and</strong> ZEB2. J Exp Med. 2011 May 9;208(5):875-83<br />

Lee JW, Park YA, Choi JJ, Lee YY, Kim CJ, Choi C, Kim<br />

TJ, Lee NW, Kim BG, Bae DS. The expression <strong>of</strong> the<br />

miRNA-200 family in endometrial endometrioid carcinoma.<br />

Gynecol Oncol. 2011 Jan;120(1):56-62<br />

Liu XG, Zhu WY, Huang YY, Ma LN, Zhou SQ, Wang YK,<br />

Zeng F, Zhou JH, Zhang YK. High expression <strong>of</strong> serum<br />

miR-21 <strong>and</strong> tumor miR-200c associated with poor<br />

prognosis in patients with lung cancer. Med Oncol. 2011<br />

Apr 24;<br />

Marchini S, Cavalieri D, Fruscio R, Calura E, Garavaglia D,<br />

Nerini IF, Mangioni C, Cattoretti G, Clivio L, Beltrame L,<br />

Katsaros D, Scarampi L, Menato G, Perego P, Chiorino G,<br />

Buda A, Romualdi C, D'Incalci M. Association between<br />

miR-200c <strong>and</strong> the survival <strong>of</strong> patients with stage I epithelial<br />

ovarian cancer: a retrospective study <strong>of</strong> two independent<br />

tumour tissue collections. Lancet Oncol. 2011<br />

Mar;12(3):273-85<br />

Roybal JD, Zang Y, Ahn YH, Yang Y, Gibbons DL, Baird<br />

BN, Alvarez C, Thilaganathan N, Liu DD, Saintigny P,<br />

Heymach JV, Creighton CJ, Kurie JM. miR-200 Inhibits<br />

lung adenocarcinoma cell invasion <strong>and</strong> metastasis by<br />

targeting Flt1/VEGFR1. Mol Cancer Res. 2011<br />

Jan;9(1):25-35<br />

Smith CM, Watson DI, Leong MP, Mayne GC, Michael MZ,<br />

Wijnhoven BP, Hussey DJ. miR-200 family expression is<br />

downregulated upon neoplastic progression <strong>of</strong> Barrett's<br />

esophagus. World J Gastroenterol. 2011 Feb<br />

28;17(8):1036-44<br />

Tellez CS, Juri DE, Do K, Bernauer AM, Thomas CL,<br />

Damiani LA, Tessema M, Leng S, Belinsky SA. EMT <strong>and</strong><br />

stem cell-like properties associated with miR-205 <strong>and</strong> miR-<br />

200 epigenetic silencing are early manifestations during<br />

carcinogen-induced transformation <strong>of</strong> human lung<br />

epithelial cells. Cancer Res. 2011 Apr 15;71(8):3087-97<br />

Vallejo DM, Caparros E, Dominguez M. Targeting Notch<br />

signalling by the conserved miR-8/200 microRNA family in<br />

development <strong>and</strong> cancer cells. EMBO J. 2011 Feb<br />

16;30(4):756-69<br />

Wiklund ED, Bramsen JB, Hulf T, Dyrskjøt L, Ramanathan<br />

R, Hansen TB, Villadsen SB, Gao S, Ostenfeld MS, Borre<br />

M, Peter ME, Ørnt<strong>of</strong>t TF, Kjems J, Clark SJ. Coordinated<br />

epigenetic repression <strong>of</strong> the miR-200 family <strong>and</strong> miR-205<br />

in invasive bladder cancer. Int J Cancer. 2011 Mar<br />

15;128(6):1327-34<br />

Yang Y, Ahn YH, Gibbons DL, Zang Y, Lin W,<br />

Thilaganathan N, Alvarez CA, Moreira DC, Creighton CJ,<br />

Gregory PA, Goodall GJ, Kurie JM. The Notch lig<strong>and</strong><br />

Jagged2 promotes lung adenocarcinoma metastasis<br />

through a miR-200-dependent pathway in mice. J Clin<br />

Invest. 2011 Apr 1;121(4):1373-85<br />

Zidar N, Boštjančič E, Gale N, Kojc N, Poljak M, Glavač D,<br />

Cardesa A. Down-regulation <strong>of</strong> microRNAs <strong>of</strong> the miR-200<br />

family <strong>and</strong> miR-205, <strong>and</strong> an altered expression <strong>of</strong> classic<br />

<strong>and</strong> desmosomal cadherins in spindle cell carcinoma <strong>of</strong><br />

the head <strong>and</strong> neck--hallmark <strong>of</strong> epithelial-mesenchymal<br />

transition. Hum Pathol. 2011 Apr;42(4):482-8<br />

This article should be referenced as such:<br />

Jurmeister S, Uhlmann S, Sahin Ö. MIR200C (microRNA<br />

200c). <strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012;<br />

16(2):90-97.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 97


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

<strong>Gene</strong> <strong>Section</strong><br />

Mini Review<br />

PXN (paxillin)<br />

Tiffany Pierson, Brendan C Stack Jr<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

Department <strong>of</strong> Otolaryngology-Head <strong>and</strong> Neck Surgery, University <strong>of</strong> Arkansas for Medical Sciences,<br />

AR 72205, USA (TP, BCJrS)<br />

Published in <strong>Atlas</strong> Database: August 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/<strong>Gene</strong>s/PXNID41953ch12q24.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI PXNID41953ch12q24.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Identity<br />

Other names: FLJ16691<br />

HGNC (Hugo): PXN<br />

Location: 12q24.23<br />

Local order: Information about the local order <strong>of</strong><br />

PXN can be found at ensembl.org.<br />

DNA/RNA<br />

Description<br />

The PXN gene is 55.314 kb <strong>and</strong> consists <strong>of</strong> 12<br />

exons. This gene is a member <strong>of</strong> the Human CCDS<br />

set: CCDS44996, CCDS44997, CCDS44998.<br />

Transcription<br />

The transcript is 3788 base pairs long. 4 is<strong>of</strong>orms<br />

have been identified.<br />

Pseudogene<br />

Not known.<br />

Protein<br />

Description<br />

4 is<strong>of</strong>orms have been identified by alternative<br />

splicing. The 1st is<strong>of</strong>orm is the normal variant <strong>and</strong><br />

is comprised <strong>of</strong> 591 AA <strong>and</strong> weighs 68 kDa. The<br />

amino terminus region contains 5 LD-motifs, while<br />

the carboxy terminus contains 4 LIM-zinc binding<br />

domains. The protein also contains a proline rich<br />

region <strong>and</strong> several potential phosphorylation sites.<br />

Expression<br />

Epithelium.<br />

Localisation<br />

Found in the cytoplasm closely apposed to the<br />

plasma membrane at sites <strong>of</strong> focal adhesion to the<br />

extracellular matrix.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 98


PXN (paxillin) Pierson T, Stack BCJr<br />

Function<br />

Focal adhesion protein: This protein is a<br />

cytoskeletal component involved in focal actinmembrane<br />

attachments to the extracellular matrix.<br />

PXN can interact with multiple structural molecules<br />

<strong>and</strong> regulatory proteins to modulate adhesion,<br />

motility <strong>and</strong> survival <strong>of</strong> the cell by changing actin<br />

dynamics. Some PXN binding proteins have<br />

oncogenic equivalents, allowing cells to bypass<br />

normal adhesion <strong>and</strong> GF signaling cascades.<br />

Regulation: PXN activity is regulated by various<br />

kinases. Adhesion <strong>and</strong> GF's stimulate these kinases<br />

to phosphorylate LD motifs or LIM domains.<br />

Molecules such as Vinculin, FAK <strong>and</strong> SRC<br />

phosphorylate tyrosine residues <strong>of</strong> the N-terminal<br />

LD motifs. This results in recruitment <strong>of</strong><br />

downstream effectors (like CRK) to mediate<br />

changes in cell motility or in modulation <strong>of</strong> gene<br />

expression via MAPK pathways. N-terminal serine<br />

phosphorylation has also been identified.<br />

Phosphorylation <strong>of</strong> serine <strong>and</strong> threonine residues <strong>of</strong><br />

C-terminal LIM domains results in recruitment to<br />

focal adhesions. Identification <strong>of</strong> the C-terminal<br />

kinases is currently under investigation.<br />

Abbreviations: CRK (CT10 sarcoma oncogene<br />

cellular homolog ); FAK (focal adhesion kinase);<br />

GF (growth factor); MAPK (mitogen activated<br />

protein kinase); SRC (Rous sarcoma oncogene<br />

cellular homolog).<br />

Homology<br />

Member <strong>of</strong> the paxillin family, containing the 4<br />

LIM-zinc binding domains.<br />

Mutations<br />

Germinal<br />

Not known.<br />

Somatic<br />

Several single nucleotide polymorphisms have been<br />

identified. Point mutations between the LD1 <strong>and</strong><br />

LD2 motifs have been associated with lung cancer,<br />

the A127T mutation being the most frequent<br />

mutation (Jagadeeswaran, et al., 2008).<br />

Implicated in<br />

Head <strong>and</strong> neck cancers<br />

Note<br />

PXN overexpression has been reported in various<br />

head <strong>and</strong> neck cancers (Li et al., 2008; Dai et al.,<br />

2010; Shi et al., 2010). Metallopanstimulin-1<br />

expression has been associated with reduced PXN<br />

levels <strong>and</strong> tumor growth rate (Dai et al., 2010).<br />

Lung cancer<br />

Note<br />

A significant correlation was found between the<br />

presence <strong>of</strong> the A127T mutation between the LD1<br />

<strong>and</strong> LD2 regions <strong>of</strong> PXN with non small cell lung<br />

cancer (Jagadeeswaran et al., 2008). A possible<br />

mechanism is that mutations between the LD1 <strong>and</strong><br />

2 regions confer resistance to calpain mediated<br />

proteolysis <strong>of</strong> PXN (Cortesio et al., 2011).<br />

However, two later studies did not find this<br />

mutation to exist in lung cancer or any other solid<br />

tumor (Pallier et al., 2009; Kim et al., 2011).<br />

Overexpression <strong>of</strong> PXN in non small cell lung<br />

cancer has been reported with less controversy<br />

(Jagadeeswaran et al., 2008; Zhao et al., 2010;<br />

Mackinnon et al., 2011). The overexpression could<br />

possibly be due to rearrangements on chromosome<br />

12 (Wu et al., 2010).<br />

Breast cancer<br />

Note<br />

Metastatic potential was found to be directly related<br />

to PXN levels (Cai et al., 2010). The relationship<br />

between PXN <strong>and</strong> Her-2 expression is<br />

controversial. A study in 2007 found a direct<br />

relationship between the 2 markers (Short et al.,<br />

2007) while a 2011 study found no such link<br />

(Panousis et al., 2011).<br />

Prostate cancer<br />

Note<br />

PXN up regulation was found to promote adhesion<br />

<strong>and</strong> motility <strong>of</strong> prostate cancer cells (Bokobza et al.,<br />

2010).<br />

To be noted<br />

Note<br />

The link between PXN mutations <strong>and</strong> increased<br />

growth rate <strong>and</strong> invasion <strong>of</strong> cancer cells is<br />

controversial. On the contrary, amplification <strong>and</strong>/or<br />

overexpression <strong>of</strong> PXN has been consistently<br />

reported in the literature.<br />

References<br />

Sattler M, Pisick E, Morrison PT, Salgia R. Role <strong>of</strong> the<br />

cytoskeletal protein paxillin in oncogenesis. Crit Rev<br />

Oncog. 2000;11(1):63-76<br />

Brown MC, Turner CE. Paxillin: adapting to change.<br />

Physiol Rev. 2004 Oct;84(4):1315-39<br />

Conway WC, Van der Voort van Zyp J, Thamilselvan V,<br />

Walsh MF, Crowe DL, Basson MD. Paxillin modulates<br />

squamous cancer cell adhesion <strong>and</strong> is important in<br />

pressure-augmented adhesion. J Cell Biochem. 2006 Aug<br />

15;98(6):1507-16<br />

Short SM, Yoder BJ, Tarr SM, Prescott NL, Laniauskas S,<br />

Coleman KA, Downs-Kelly E, Pettay JD, Choueiri TK,<br />

Crowe JP, Tubbs RR, Budd TG, Hicks DG. The expression<br />

<strong>of</strong> the cytoskeletal focal adhesion protein paxillin in breast<br />

cancer correlates with HER2 overexpression <strong>and</strong> may help<br />

predict response to chemotherapy: a retrospective<br />

immunohistochemical study. Breast J. 2007 Mar-<br />

Apr;13(2):130-9<br />

Deakin NO, Turner CE. Paxillin comes <strong>of</strong> age. J Cell Sci.<br />

2008 Aug 1;121(Pt 15):2435-44<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 99


PXN (paxillin) Pierson T, Stack BCJr<br />

Jagadeeswaran R, Surawska H, Krishnaswamy S,<br />

Janamanchi V, Mackinnon AC, Seiwert TY, Loganathan S,<br />

Kanteti R, Reichman T, Nallasura V, Schwartz S, Faoro L,<br />

Wang YC, Girard L, Tretiakova MS, Ahmed S, Zumba O,<br />

Soulii L, Bindokas VP, Szeto LL, Gordon GJ, Bueno R,<br />

Sugarbaker D, Lingen MW, Sattler M, Krausz T,<br />

Vigneswaran W, Natarajan V, Minna J, Vokes EE,<br />

Ferguson MK, Husain AN, Salgia R. Paxillin is a target for<br />

somatic mutations in lung cancer: implications for cell<br />

growth <strong>and</strong> invasion. Cancer Res. 2008 Jan 1;68(1):132-<br />

42<br />

Li BZ, Lei W, Zhang CY, Zhou F, Li N, Shi SS, Feng XL,<br />

Chen ZL, Hang J, Qiu B, Wan JT, Shao K, Xing XZ, Tan<br />

XG, Wang Z, Xiong MH, He J. Increased expression <strong>of</strong><br />

paxillin is found in human oesophageal squamous cell<br />

carcinoma: a tissue microarray study. J Int Med Res. 2008<br />

Mar-Apr;36(2):273-8<br />

Sheibani N, Tang Y, Sorenson CM. Paxillin's LD4 motif<br />

interacts with bcl-2. J Cell Physiol. 2008 Mar;214(3):655-<br />

61<br />

Pallier K, Houllier AM, Le Corre D, Cazes A, Laurent-Puig<br />

P, Blons H. No somatic genetic change in the paxillin gene<br />

in nonsmall-cell lung cancer. Mol Carcinog. 2009<br />

Jul;48(7):581-5<br />

Bokobza SM, Ye L, Kynaston HG, Jiang WG. Growth <strong>and</strong><br />

differentiation factor-9 promotes adhesive <strong>and</strong> motile<br />

capacity <strong>of</strong> prostate cancer cells by up-regulating FAK <strong>and</strong><br />

Paxillin via Smad dependent pathway. Oncol Rep. 2010<br />

Dec;24(6):1653-9<br />

Dai Y, Pierson SE, Dudney WC, Stack BC Jr.<br />

Extraribosomal function <strong>of</strong> metallopanstimulin-1: reducing<br />

paxillin in head <strong>and</strong> neck squamous cell carcinoma <strong>and</strong><br />

inhibiting tumor growth. Int J Cancer. 2010 Feb<br />

1;126(3):611-9<br />

Cai H, Zhang T, Tang WX, Li SL. [Expression <strong>of</strong> paxillin in<br />

breast cancer cell with high <strong>and</strong> low metastatic<br />

potentiality]. Sichuan Da Xue Xue Bao Yi Xue Ban. 2010<br />

Jan;41(1):91-4<br />

Shi J, Wang S, Zhao E, Shi L, Xu X, Fang M. Paxillin<br />

expression levels are correlated with clinical stage <strong>and</strong><br />

metastasis in salivary adenoid cystic carcinoma. J Oral<br />

Pathol Med. 2010 Aug 1;39(7):548-51<br />

Wu DW, Cheng YW, Wang J, Chen CY, Lee H. Paxillin<br />

predicts survival <strong>and</strong> relapse in non-small cell lung cancer<br />

by microRNA-218 targeting. Cancer Res. 2010 Dec<br />

15;70(24):10392-401<br />

Zhao Y, Zhang X, Guda K, Lawrence E, Sun Q, Watanabe<br />

T, Iwakura Y, Asano M, Wei L, Yang Z, Zheng W, Dawson<br />

D, Willis J, Markowitz SD, Satake M, Wang Z. Identification<br />

<strong>and</strong> functional characterization <strong>of</strong> paxillin as a target <strong>of</strong><br />

protein tyrosine phosphatase receptor T. Proc Natl Acad<br />

Sci U S A. 2010 Feb 9;107(6):2592-7<br />

Cortesio CL, Boateng LR, Piazza TM, Bennin DA,<br />

Huttenlocher A. Calpain-mediated proteolysis <strong>of</strong> paxillin<br />

negatively regulates focal adhesion dynamics <strong>and</strong> cell<br />

migration. J Biol Chem. 2011 Mar 25;286(12):9998-10006<br />

Kim MS, Yoo NJ, Lee SH. Absence <strong>of</strong> paxillin gene<br />

mutation in lung cancer <strong>and</strong> other common solid cancers.<br />

Tumori. 2011 Mar-Apr;97(2):211-3<br />

Mackinnon AC, Tretiakova M, Henderson L, Mehta RG,<br />

Yan BC, Joseph L, Krausz T, Husain AN, Reid ME, Salgia<br />

R. Paxillin expression <strong>and</strong> amplification in early lung<br />

lesions <strong>of</strong> high-risk patients, lung adenocarcinoma <strong>and</strong><br />

metastatic disease. J Clin Pathol. 2011 Jan;64(1):16-24<br />

Panousis D, Patsouris E, Lagoudianakis E, Pappas A,<br />

Kyriakidou V, Voulgaris Z, Xepapadakis G, Manouras A,<br />

Athanassiadou AM, Athanassiadou P. The value <strong>of</strong><br />

TOP2A, EZH2 <strong>and</strong> paxillin expression as markers <strong>of</strong><br />

aggressive breast cancer: relationship with other<br />

prognostic factors. Eur J Gynaecol Oncol. 2011;32(2):156-<br />

9<br />

This article should be referenced as such:<br />

Pierson T, Stack BCJr. PXN (paxillin). <strong>Atlas</strong> <strong>Gene</strong>t<br />

Cytogenet Oncol Haematol. 2012; 16(2):98-100.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 100


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

<strong>Gene</strong> <strong>Section</strong><br />

Mini Review<br />

RPS27 (ribosomal protein S27)<br />

Tiffany Pierson, Brendan C Stack Jr<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

Department <strong>of</strong> Otolaryngology-Head <strong>and</strong> Neck Surgery, University <strong>of</strong> Arkansas for Medical Sciences,<br />

AR 72205, USA (TP, BCJrS)<br />

Published in <strong>Atlas</strong> Database: August 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/<strong>Gene</strong>s/RPS27ID45550ch1q21.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI RPS27ID45550ch1q21.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Identity<br />

Other names: MPS-1, MPS1, S27<br />

HGNC (Hugo): RPS27<br />

Location: 1q21.3<br />

Local order: Human RPS27 is found on<br />

chromosome 1: 153963235 - 153964626 bp from<br />

pter. Information about the local order for RPS27<br />

can be found at ensembl.org. Four transcripts have<br />

been identified, but only the first will be discussed<br />

below.<br />

DNA/RNA<br />

Description<br />

The RPS27 gene is comprised <strong>of</strong> 1.39 kb <strong>and</strong><br />

consists <strong>of</strong> 4 exons. This gene is a member <strong>of</strong> the<br />

Human CCDS set: CCDS1059.<br />

Transcription<br />

The transcript is 350 base pairs long.<br />

Pseudogene<br />

Multiple RPS27 pseudogenes are dispersed<br />

throughout the genome. The RPS27L pseudogene,<br />

located at 15q22.2, is known to encode a protein<br />

that shares 96% <strong>of</strong> its amino acid sequence with<br />

RPS27 (Balasubramanian et al., 2009).<br />

Protein<br />

Description<br />

RPS27 is a 9461 Da protein composed <strong>of</strong> 84 amino<br />

acids. The protein contains a C4 zinc finger<br />

domain, similar to steroid <strong>and</strong> thyroid hormones,<br />

which enables DNA binding. RPS27 is found in<br />

both the cytoplasm <strong>and</strong> the nucleus.<br />

Expression<br />

Ubiquitous expression. Expressed at high levels in<br />

actively dividing cells <strong>and</strong> in cancers <strong>of</strong> ectodermal<br />

origin, as well as in melanoma (Fern<strong>and</strong>ez-Pol et<br />

al., 1993). When overexpressed, it is secreted into<br />

serum (Lee et al., 2004).<br />

Function<br />

1. Component <strong>of</strong> the 40S ribosomal subunit in the<br />

cytoplasm: ribosomes carry out translation <strong>of</strong><br />

proteins. The eukaryotic ribosome is made up <strong>of</strong> a<br />

small 40S <strong>and</strong> a large 60S subunit. Together these<br />

subunits are comprised <strong>of</strong> 4 different rRNA species<br />

<strong>and</strong> almost 80 different RP's (ribosomal proteins).<br />

As a component <strong>of</strong> the 40S subunit, RPS27 is found<br />

near RPS18 <strong>and</strong> covalently bound to translation<br />

initiation factor eIF3.<br />

2. A mediator <strong>of</strong> cellular proliferation <strong>and</strong> survival:<br />

expression is induced by a variety <strong>of</strong> growth factors<br />

<strong>and</strong> other signaling molecules, including TGF-beta<br />

<strong>and</strong> cAMP; RPS27 can bind to cAMP response<br />

elements <strong>of</strong> DNA (Fern<strong>and</strong>ez-Pol et al., 1993).<br />

3. Oncogenesis (see below).<br />

Homology<br />

Member <strong>of</strong> the ribosomal protein S27e family.<br />

Mutations<br />

Note<br />

Single nucleotide polymorphisms have been<br />

identified, but have not been linked to disease.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 101


RPS27 (ribosomal protein S27) Pierson T, Stack BCJr<br />

Implicated in<br />

Various carcinomas <strong>and</strong> melanoma<br />

Note<br />

RPS27 overexpression has been reported in many<br />

cancers including prostate cancer (Fern<strong>and</strong>ez-Pol et<br />

al., 1997), colorectal cancer (Ganger et al., 1997),<br />

liver cancer (Ganger et al., 2001), breast cancer<br />

(Atsuta et al., 2002), head <strong>and</strong> neck squamous cell<br />

cancer (HNSCC) (Stack et al., 1999; Stack et al.,<br />

2004; Lee et al., 2004), gastric cancer (Wang, et al.,<br />

2006), as well as, melanoma (Santa Cruz et al.,<br />

1997).<br />

Since high serum levels <strong>of</strong> RPS27 have been found<br />

in cancer patients, especially in head <strong>and</strong> neck<br />

squamous cell carcinoma (HNSCC), the protein can<br />

be used as a tumor marker (Fern<strong>and</strong>ez-Pol et al.,<br />

1996; Lee et al., 2004; Stack et al. 2004).<br />

Prognosis<br />

It was reported that RPS27 levels correlate with<br />

tumor stage in patients with gastric cancer, thus<br />

high levels serve as a poor prognostic indicator<br />

(Wang et al., 2006).<br />

Oncogenesis<br />

The mechanism behind RPS27 overexpression is<br />

currently under investigation. One explanation<br />

recently <strong>of</strong>fered arises from the relationship<br />

between RPS27, MDM2 <strong>and</strong> p53: RPS27 is a p53<br />

repressible protein (He <strong>and</strong> Sun, 2007; Li et al.,<br />

2007). A 2011 study found that it competes with<br />

p53 for a central acidic binding domain on MDM2.<br />

Once bound, MDM2 is stimulated to ubiquinate <strong>and</strong><br />

degrade the RPS27 or p53, whichever it is bound<br />

to. When RPS27 levels are elevated, it can outcompete<br />

p53 for MDM2 binding <strong>and</strong> subsequent<br />

degradation, thus stabilizing p53 levels. This would<br />

be an appropriate cellular response to genotoxic<br />

stress. The same study also found that mutant p53<br />

cannot suppress RPS27, only the wild-type can.<br />

Since mutated p53 is found in almost 50% <strong>of</strong> all<br />

human cancers, RPS27 overexpression logically<br />

follows. Furthermore, stabilization <strong>of</strong> mutant p53<br />

levels associated with RPS27 abundance could<br />

provide malignant cells with a growth advantage<br />

(Xiong et al., 2011).<br />

RPS27 knockdown was found to enhance<br />

spontaneous apoptosis <strong>of</strong> tumor cells via caspase-3<br />

activation (Wang et al., 2006; Yang et al., 2011).<br />

HNSCC: some have questioned if RPS27<br />

overexpression is the cause or result <strong>of</strong> cancer. A<br />

2010 study overexpressed RPS27 in a line <strong>of</strong><br />

HNSCC cells to study the impact on tumor<br />

behavior. They found that RPS27 overexpression<br />

resulted in reduced cancer cell growth, proliferation<br />

rate <strong>and</strong> angiogenesis. RPS27 overexpression was<br />

also found to reduce the mRNA <strong>of</strong> Paxillin, a focal<br />

adhesion protein up regulated in HNSCC <strong>and</strong> many<br />

other cancer cells. RPS27 induced Paxillin<br />

repression <strong>of</strong>fers a possible explanation for the<br />

decreased HNSCC growth (Dai et al., 2010).<br />

References<br />

Fern<strong>and</strong>ez-Pol JA, Klos DJ, Hamilton PD. A growth factorinducible<br />

gene encodes a novel nuclear protein with zinc<br />

finger structure. J Biol Chem. 1993 Oct 5;268(28):21198-<br />

204<br />

Fern<strong>and</strong>ez-Pol JA, Fletcher JW, Hamilton PD, Klos DJ.<br />

Expression <strong>of</strong> metallopanstimulin <strong>and</strong> oncogenesis in<br />

human prostatic carcinoma. Anticancer Res. 1997 May-<br />

Jun;17(3A):1519-30<br />

Ganger DR, Hamilton PD, Fletcher JW, Fern<strong>and</strong>ez-Pol JA.<br />

Metallopanstimulin is overexpressed in a patient with<br />

colonic carcinoma. Anticancer Res. 1997 May-<br />

Jun;17(3C):1993-9<br />

Santa Cruz DJ, Hamilton PD, Klos DJ, Fern<strong>and</strong>ez-Pol JA.<br />

Differential expression <strong>of</strong> metallopanstimulin/S27<br />

ribosomal protein in melanocytic lesions <strong>of</strong> the skin. J<br />

Cutan Pathol. 1997 Oct;24(9):533-42<br />

Stack BC Jr, Dalsaso TA, Lee C Jr, Lowe VJ, Hamilton<br />

PD, Fletcher JW, Fern<strong>and</strong>ez-Pol JA. Overexpression <strong>of</strong><br />

MPS antigens by squamous cell carcinomas <strong>of</strong> the head<br />

<strong>and</strong> neck: immunohistochemical <strong>and</strong> serological<br />

correlation with FDG positron emission tomography.<br />

Anticancer Res. 1999 Nov-Dec;19(6C):5503-10<br />

Ganger DR, Hamilton PD, Klos DJ, Jakate S, McChesney<br />

L, Fern<strong>and</strong>ez-Pol JA. Differential expression <strong>of</strong><br />

metallopanstimulin/S27 ribosomal protein in hepatic<br />

regeneration <strong>and</strong> neoplasia. Cancer Detect Prev.<br />

2001;25(3):231-6<br />

Atsuta Y, Aoki N, Sato K, Oikawa K, Nochi H, Miyokawa N,<br />

Hirata S, Kimura S, Sasajima T, Katagiri M. Identification<br />

<strong>of</strong> metallopanstimulin-1 as a member <strong>of</strong> a tumor<br />

associated antigen in patients with breast cancer. Cancer<br />

Lett. 2002 Aug 8;182(1):101-7<br />

Lee WJ, Keefer K, Hollenbeak CS, Stack BC Jr. A new<br />

assay to screen for head <strong>and</strong> neck squamous cell<br />

carcinoma using the tumor marker metallopanstimulin.<br />

Otolaryngol Head Neck Surg. 2004 Oct;131(4):466-71<br />

Stack BC Jr, Hollenbeak CS, Lee CM, Dunphy FR, Lowe<br />

VJ, Hamilton PD. Metallopanstimulin as a marker for head<br />

<strong>and</strong> neck cancer. World J Surg Oncol. 2004 Dec 14;2:45<br />

Gilkes DM, Chen L, Chen J. MDMX regulation <strong>of</strong> p53<br />

response to ribosomal stress. EMBO J. 2006 Nov<br />

29;25(23):5614-25<br />

Wang YW, Qu Y, Li JF, Chen XH, Liu BY, Gu QL, Zhu ZG.<br />

In vitro <strong>and</strong> in vivo evidence <strong>of</strong> metallopanstimulin-1 in<br />

gastric cancer progression <strong>and</strong> tumorigenicity. Clin Cancer<br />

Res. 2006 Aug 15;12(16):4965-73<br />

He H, Sun Y. Ribosomal protein S27L is a direct p53 target<br />

that regulates apoptosis. Oncogene. 2007 Apr<br />

26;26(19):2707-16<br />

Li J, Tan J, Zhuang L, Banerjee B, Yang X, Chau JF, Lee<br />

PL, H<strong>and</strong>e MP, Li B, Yu Q. Ribosomal protein S27-like, a<br />

p53-inducible modulator <strong>of</strong> cell fate in response to<br />

genotoxic stress. Cancer Res. 2007 Dec 1;67(23):11317-<br />

26<br />

Balasubramanian S, Zheng D, Liu YJ, Fang G, Frankish A,<br />

Carriero N, Robilotto R, Cayting P, Gerstein M.<br />

Comparative analysis <strong>of</strong> processed ribosomal protein<br />

pseudogenes in four mammalian genomes. Genome Biol.<br />

2009;10(1):R2<br />

Dai Y, Pierson SE, Dudney WC, Stack BC Jr.<br />

Extraribosomal function <strong>of</strong> metallopanstimulin-1: reducing<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 102


RPS27 (ribosomal protein S27) Pierson T, Stack BCJr<br />

paxillin in head <strong>and</strong> neck squamous cell carcinoma <strong>and</strong><br />

inhibiting tumor growth. Int J Cancer. 2010 Feb<br />

1;126(3):611-9<br />

Xiong X, Zhao Y, He H, Sun Y. Ribosomal protein S27-like<br />

<strong>and</strong> S27 interplay with p53-MDM2 axis as a target, a<br />

substrate <strong>and</strong> a regulator. Oncogene. 2011 Apr<br />

14;30(15):1798-811<br />

Yang ZY, Qu Y, Zhang Q, Wei M, Liu CX, Chen XH, Yan<br />

M, Zhu ZG, Liu BY, Chen GQ, Wu YL, Gu QL. Knockdown<br />

<strong>of</strong> metallopanstimulin-1 inhibits NF-κB signaling at different<br />

levels: The role <strong>of</strong> apoptosis induction <strong>of</strong> gastric cancer<br />

cells. Int J Cancer. 2011 Jul 27;<br />

This article should be referenced as such:<br />

Pierson T, Stack BCJr. RPS27 (ribosomal protein S27).<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2):101-<br />

103.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 103


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

<strong>Gene</strong> <strong>Section</strong><br />

Review<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

STAT5B (signal transducer <strong>and</strong> activator <strong>of</strong><br />

transcription 5B)<br />

Am<strong>and</strong>a M Del Rosario, Teresa M Bernaciak, Corinne M Silva<br />

Department <strong>of</strong> Biological Engineering, Massachusetts Institute <strong>of</strong> Technology, 77 Massachusetts<br />

Avenue, Cambridge, MA 02139, USA (AMDR), NIAID/NIH/DHHS, 6610 Rockledge Drive,<br />

Bethesda, MD 20817, USA (TMB), DEM/NIDDK/NIH, 6707 Democracy Blvd, Bethesda, MD<br />

20892, USA (CMS)<br />

Published in <strong>Atlas</strong> Database: August 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/<strong>Gene</strong>s/STAT5BID217ch17q21.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI STAT5BID217ch17q21.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Identity<br />

Other names: STAT5<br />

HGNC (Hugo): STAT5B<br />

Location: 17q21.2<br />

DNA/RNA<br />

Description<br />

The STAT5b gene is composed <strong>of</strong> 77229 base pairs<br />

<strong>and</strong> contains 19 exons. As exon 1 contains only the<br />

5' UTR, there are 18 coding exons. The mRNA is<br />

composed <strong>of</strong> 5090 base pairs.<br />

Transcription<br />

There is one major transcript.<br />

The STAT5b gene. The STAT5b gene is composed <strong>of</strong> 19 exons that are transcribed as a 5090 nucleotide RNA transcript. The<br />

RNA is translated into the STAT5b protein containing 787 amino acids.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 104


STAT5B (signal transducer <strong>and</strong> activator <strong>of</strong> transcription 5B) Del Rosario AM, et al.<br />

Schematic <strong>of</strong> the STAT5b protein. The STAT5b protein contains several conserved domains: the coiled-coil domain, the DNA<br />

binding domain, the SH2 domain, <strong>and</strong> the carboxy-terminal transactivation domain. While phosphorylation <strong>of</strong> Y699 is required for<br />

transcriptional activity, there are multiple tyrosine <strong>and</strong> serine phosphorylation sites that have been identified under specific<br />

conditions <strong>and</strong> in certain cell types.<br />

Protein<br />

Note<br />

STAT5b is composed <strong>of</strong> 787 amino acids (92 kD).<br />

Description<br />

STAT5b is a member <strong>of</strong> the signal transducer <strong>and</strong><br />

activator <strong>of</strong> transcription (STAT) family. The<br />

STAT proteins contain several conserved domains:<br />

the coiled-coil domain, the DNA binding domain,<br />

the SH2 domain, <strong>and</strong> the carboxy-terminal<br />

transactivation domain. STATs remain latent in the<br />

cytoplasm until the binding <strong>of</strong> a cytokine or growth<br />

factor to its receptor, resulting in recruitment <strong>of</strong> the<br />

STAT to the lig<strong>and</strong> receptor complex (Levy <strong>and</strong><br />

Darnell, 2002; Herrington et al., 1999; Heim et al.,<br />

1995). The STAT protein is then phosphorylated by<br />

receptor tyrosine kinases or non-receptor tyrosine<br />

kinases, such as Janus kinases (JAKs) <strong>and</strong> Src<br />

family members. This phosphorylation results in<br />

SH2 domain mediated dimerization <strong>of</strong> STATs <strong>and</strong><br />

their translocation to the nucleus. In the nucleus,<br />

STAT dimers bind to consensus DNA sequences<br />

<strong>and</strong> recruit additional transcription machinery to<br />

initiate specific gene regulation. To date, seven<br />

members <strong>of</strong> the STAT family have been identified<br />

(STAT1, STAT2, STAT3, STAT4, STAT5a,<br />

STAT5b, <strong>and</strong> STAT6) (Silva, 2004; Calò et al.,<br />

2003; Moriggl et al., 1996; Liu et al., 1996; Liu et<br />

al., 1997).<br />

STAT5b is activated by a variety <strong>of</strong> stimuli,<br />

including interleukins, erythropoietin, growth<br />

hormone (GH), prolactin (Prl), <strong>and</strong> epidermal<br />

growth factor (EGF). Activation <strong>of</strong> STAT5b results<br />

in phosphorylation <strong>of</strong> tyrosine 699. Phosphorylation<br />

<strong>of</strong> this tyrosine is required for DNA binding <strong>and</strong><br />

transcriptional activity. Mutation <strong>of</strong> Y699 <strong>of</strong><br />

STAT5b inhibits stimulant-induced tyrosine<br />

phosphorylation, DNA binding, <strong>and</strong> transcriptional<br />

activity (Gebert et al., 1997; Gouilleux et al., 1994;<br />

Kloth et al., 2003). Additional tyrosine<br />

phosphorylation sites (Y679, Y725, Y740, <strong>and</strong><br />

Y743) <strong>and</strong> serine phosphorylation sites (S715,<br />

S731) have been shown to alter STAT5b<br />

transcriptional activity (Kloth et al., 2002; Weaver<br />

<strong>and</strong> Silva, 2006; Yamashita et al., 2001; Park et al.,<br />

2001; Decker <strong>and</strong> Kovarik, 2000).<br />

While no classic nuclear localization signal (NLS)<br />

composed <strong>of</strong> a cluster <strong>of</strong> basic amino acids has been<br />

reported for the STAT5b, the STAT5b dimer is<br />

actively translocated through the nuclear pore<br />

complex <strong>and</strong> accumulates in the nucleus upon<br />

phosphorylation (Xu <strong>and</strong> Massagué, 2004).<br />

STAT5b can be negatively regulated by<br />

phosphatase-mediated dephosphorylation,<br />

ubiquitination-promoting proteosome degradation,<br />

or by negative feedback loops.<br />

Expression<br />

Ubiquitous.<br />

Localisation<br />

STAT5b is localized in the cytoplasm <strong>and</strong><br />

translocates to the nucleus upon phosphorylation <strong>of</strong><br />

Y699. However, unphosphorylated STAT5b has<br />

also been reported to be found in the nucleus<br />

(Brown <strong>and</strong> Zeidler, 2008; Iyer <strong>and</strong> Reich, 2008;<br />

Zeng et al., 2002).<br />

Function<br />

Transcription factor. STAT5b mediates the<br />

transcription <strong>of</strong> numerous genes in various cell<br />

signaling pathways involved in cellular<br />

proliferation, differentiation, <strong>and</strong> cell survival. The<br />

STATs bind TTC(N3)GAA gamma-interferonactivating<br />

sequence (GAS) sites in the promoters <strong>of</strong><br />

target genes.<br />

Homology<br />

Shares homology with the other STAT family<br />

members (STAT1, 2, 3, 4, 5a, <strong>and</strong> 6). Additionally,<br />

STAT5a <strong>and</strong> STAT5b are 94% similar at the amino<br />

acid level, differing primarily at the C-terminus<br />

(Teglund et al., 1998; Silva et al., 1996; Lin et al.,<br />

1996; Liu et al., 1995).<br />

Mutations<br />

Note<br />

To date, there are 6 reported cases <strong>of</strong> humans<br />

having a mutant STAT5b, <strong>and</strong> these cases result<br />

from five different STAT5b mutations. The first<br />

STAT5b mutation in a human to be reported was<br />

the A630P STAT5b mutant. This single point<br />

mutation in the SH2 domain causes missfolding <strong>of</strong><br />

STAT5b. A nonsense mutation in the coiled-coil<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 105


STAT5B (signal transducer <strong>and</strong> activator <strong>of</strong> transcription 5B) Del Rosario AM, et al.<br />

domain (R152X) results in the absence <strong>of</strong><br />

detectable STAT5b protein. Insertion <strong>of</strong> a<br />

nucleotide in the DBD at position 1102<br />

(Q368fsX376) or 1191 (N398E) causes a frameshift<br />

mutation resulting in a non-functional truncated<br />

STAT5b. Likewise, a single nucleotide deletion in<br />

the linker domain at position 1680 (E561R) also<br />

results in a truncated STAT5b. In each <strong>of</strong> the 6<br />

cases, STAT5b protein is not detectable, but<br />

STAT5a protein levels are unchanged. These<br />

reports are from homozygous patients while the<br />

parents are heterozygous for the STAT5b mutation<br />

<strong>and</strong> display a normal phenotype. The phenotype <strong>of</strong><br />

each STAT5b mutant is similar: pronounced short<br />

stature, growth hormone insensitivity despite<br />

normal to high levels <strong>of</strong> GH in the serum, <strong>and</strong><br />

extremely low IGF-I <strong>and</strong> IGFBP-3 levels (Chia et<br />

al., 2006; Hwa et al., 2007; Nadeau et al., 2011).<br />

Implicated in<br />

Solid tumors<br />

Note<br />

STAT5b is implicated in prostate cancer (Koptyra<br />

et al., 2011; Clevenger, 2004), breast cancer<br />

(Bernaciak et al., 2009; Peck et al., 2011; Strauss et<br />

al., 2006; Sultan et al., 2005; Yamashita et al.,<br />

2003), lung cancer (Sánchez-Ceja et al., 2006),<br />

head <strong>and</strong> neck cancer (Koppikar et al., 2008),<br />

ovarian cancer (Chen et al., 2004), hepatocellular<br />

carcinoma (Lee et al., 2006), cervical cancer (Lopez<br />

et al., 2011), <strong>and</strong> colorectal cancer (Du et al., 2011).<br />

Leukemias <strong>and</strong> lymphomas<br />

Note<br />

STAT5b is involved in the proliferation <strong>of</strong> chronic<br />

myeloid leukemia (CML) <strong>and</strong> acute myeloid<br />

leukemia (AML) cells (Baśkiewicz-Masiuk <strong>and</strong><br />

Machalińkski, 2004; Sternberg <strong>and</strong> Gillil<strong>and</strong>, 2004;<br />

Hoover et al., 2001; de Groot et al., 1999).<br />

Additionally, STAT5b has been found to fuse with<br />

the retinoic acid receptor-alpha (RARalpha) gene in<br />

a subset <strong>of</strong> acute promyelocytic leukemias (APLL)<br />

(Arnould et al., 1999). Furthermore, STAT5b plays<br />

a role in the development <strong>of</strong> lymphoblastic<br />

lymphoma (Bessette et al., 2008; Nieborowska-<br />

Skorska et al., 2001).<br />

Laron type dwarfism II<br />

Note<br />

Laron type dwarfism II (LTD2) is mediated by<br />

defects in STAT5b (Nadeau et al., 2011; Freeth et<br />

al., 1998; Chia et al., 2006).<br />

Graft-versus-host disease<br />

Note<br />

Constitutively active STAT5b increases expansion<br />

<strong>of</strong> regulatory T cells (Treg), <strong>and</strong> these Tregs are<br />

more potent suppressors <strong>of</strong> graft-versus-host<br />

disease in vivo, compared to wild-type Tregs<br />

(Vogtenhuber et al., 2010).<br />

Crohn's disease/colitis<br />

Note<br />

Growth hormone reduces mucosal inflammation in<br />

colitis by activating STAT5b, such that STAT5b<br />

deficient mice demonstrated more severe colitis<br />

compared to wild-type mice (Han et al., 2006).<br />

Diabetes <strong>and</strong> metabolic disorder<br />

Note<br />

Upon leptin stimulation, the leptin receptor can<br />

mediate STAT5b tyrosine phosphorylation <strong>and</strong><br />

transcriptional activity in the liver, gastrointestinal<br />

tract, <strong>and</strong> brain (Mütze et al., 2007; Ghilardi et al.,<br />

1996; Gong et al., 2007).<br />

References<br />

Gouilleux F, Wakao H, Mundt M, Groner B. Prolactin<br />

induces phosphorylation <strong>of</strong> Tyr694 <strong>of</strong> Stat5 (MGF), a<br />

prerequisite for DNA binding <strong>and</strong> induction <strong>of</strong> transcription.<br />

EMBO J. 1994 Sep 15;13(18):4361-9<br />

Heim MH, Kerr IM, Stark GR, Darnell JE Jr. Contribution <strong>of</strong><br />

STAT SH2 groups to specific interferon signaling by the<br />

Jak-STAT pathway. Science. 1995 Mar 3;267(5202):1347-<br />

9<br />

Liu X, Robinson GW, Gouilleux F, Groner B,<br />

Hennighausen L. Cloning <strong>and</strong> expression <strong>of</strong> Stat5 <strong>and</strong> an<br />

additional homologue (Stat5b) involved in prolactin signal<br />

transduction in mouse mammary tissue. Proc Natl Acad<br />

Sci U S A. 1995 Sep 12;92(19):8831-5<br />

Ghilardi N, Ziegler S, Wiestner A, St<strong>of</strong>fel R, Heim MH,<br />

Skoda RC. Defective STAT signaling by the leptin receptor<br />

in diabetic mice. Proc Natl Acad Sci U S A. 1996 Jun<br />

25;93(13):6231-5<br />

Lin JX, Mietz J, Modi WS, John S, Leonard WJ. Cloning <strong>of</strong><br />

human Stat5B. Reconstitution <strong>of</strong> interleukin-2-induced<br />

Stat5A <strong>and</strong> Stat5B DNA binding activity in COS-7 cells. J<br />

Biol Chem. 1996 May 3;271(18):10738-44<br />

Liu X, Robinson GW, Hennighausen L. Activation <strong>of</strong> Stat5a<br />

<strong>and</strong> Stat5b by tyrosine phosphorylation is tightly linked to<br />

mammary gl<strong>and</strong> differentiation. Mol Endocrinol. 1996<br />

Dec;10(12):1496-506<br />

Moriggl R, Gouilleux-Gruart V, Jähne R, Berchtold S,<br />

Gartmann C, Liu X, Hennighausen L, Sotiropoulos A,<br />

Groner B, Gouilleux F. Deletion <strong>of</strong> the carboxyl-terminal<br />

transactivation domain <strong>of</strong> MGF-Stat5 results in sustained<br />

DNA binding <strong>and</strong> a dominant negative phenotype. Mol Cell<br />

Biol. 1996 Oct;16(10):5691-700<br />

Silva CM, Lu H, Day RN. Characterization <strong>and</strong> cloning <strong>of</strong><br />

STAT5 from IM-9 cells <strong>and</strong> its activation by growth<br />

hormone. Mol Endocrinol. 1996 May;10(5):508-18<br />

Gebert CA, Park SH, Waxman DJ. Regulation <strong>of</strong> signal<br />

transducer <strong>and</strong> activator <strong>of</strong> transcription (STAT) 5b<br />

activation by the temporal pattern <strong>of</strong> growth hormone<br />

stimulation. Mol Endocrinol. 1997 Apr;11(4):400-14<br />

Liu X, Robinson GW, Wagner KU, Garrett L, Wynshaw-<br />

Boris A, Hennighausen L. Stat5a is m<strong>and</strong>atory for adult<br />

mammary gl<strong>and</strong> development <strong>and</strong> lactogenesis. <strong>Gene</strong>s<br />

Dev. 1997 Jan 15;11(2):179-86<br />

Freeth JS, Silva CM, Whatmore AJ, Clayton PE. Activation<br />

<strong>of</strong> the signal transducers <strong>and</strong> activators <strong>of</strong> transcription<br />

signaling pathway by growth hormone (GH) in skin<br />

fibroblasts from normal <strong>and</strong> GH binding protein-positive<br />

Laron Syndrome children. Endocrinology. 1998<br />

Jan;139(1):20-8<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 106


STAT5B (signal transducer <strong>and</strong> activator <strong>of</strong> transcription 5B) Del Rosario AM, et al.<br />

Teglund S, McKay C, Schuetz E, van Deursen JM,<br />

Stravopodis D, Wang D, Brown M, Bodner S, Grosveld G,<br />

Ihle JN. Stat5a <strong>and</strong> Stat5b proteins have essential <strong>and</strong><br />

nonessential, or redundant, roles in cytokine responses.<br />

Cell. 1998 May 29;93(5):841-50<br />

de Groot RP, Raaijmakers JA, Lammers JW, Jove R,<br />

Koenderman L. STAT5 activation by BCR-Abl contributes<br />

to transformation <strong>of</strong> K562 leukemia cells. Blood. 1999 Aug<br />

1;94(3):1108-12<br />

Herrington J, Rui L, Luo G, Yu-Lee LY, Carter-Su C. A<br />

functional DNA binding domain is required for growth<br />

hormone-induced nuclear accumulation <strong>of</strong> Stat5B. J Biol<br />

Chem. 1999 Feb 19;274(8):5138-45<br />

Yamashita H, Nevalainen MT, Xu J, LeBaron MJ, Wagner<br />

KU, Erwin RA, Harmon JM, Hennighausen L, Kirken RA,<br />

Rui H. Role <strong>of</strong> serine phosphorylation <strong>of</strong> Stat5a in<br />

prolactin-stimulated beta-casein gene expression. Mol Cell<br />

Endocrinol. 2001 Oct 25;183(1-2):151-63<br />

Arnould C, Philippe C, Bourdon V, Gr goire MJ, Berger R,<br />

Jonveaux P. The signal transducer <strong>and</strong> activator <strong>of</strong><br />

transcription STAT5b gene is a new partner <strong>of</strong> retinoic acid<br />

receptor alpha in acute promyelocytic-like leukaemia. Hum<br />

Mol <strong>Gene</strong>t. 1999 Sep;8(9):1741-9<br />

Decker T, Kovarik P. Serine phosphorylation <strong>of</strong> STATs.<br />

Oncogene. 2000 May 15;19(21):2628-37<br />

Hoover RR, Gerlach MJ, Koh EY, Daley GQ. Cooperative<br />

<strong>and</strong> redundant effects <strong>of</strong> STAT5 <strong>and</strong> Ras signaling in<br />

BCR/ABL transformed hematopoietic cells. Oncogene.<br />

2001 Sep 13;20(41):5826-35<br />

Nieborowska-Skorska M, Slupianek A, Xue L, Zhang Q,<br />

Raghunath PN, Hoser G, Wasik MA, Morris SW, Skorski T.<br />

Role <strong>of</strong> signal transducer <strong>and</strong> activator <strong>of</strong> transcription 5 in<br />

nucleophosmin/ anaplastic lymphoma kinase-mediated<br />

malignant transformation <strong>of</strong> lymphoid cells. Cancer Res.<br />

2001 Sep 1;61(17):6517-23<br />

Park SH, Yamashita H, Rui H, Waxman DJ. Serine<br />

phosphorylation <strong>of</strong> GH-activated signal transducer <strong>and</strong><br />

activator <strong>of</strong> transcription 5a (STAT5a) <strong>and</strong> STAT5b: impact<br />

on STAT5 transcriptional activity. Mol Endocrinol. 2001<br />

Dec;15(12):2157-71<br />

Kloth MT, Catling AD, Silva CM. Novel activation <strong>of</strong><br />

STAT5b in response to epidermal growth factor. J Biol<br />

Chem. 2002 Mar 8;277(10):8693-701<br />

Levy DE, Darnell JE Jr. Stats: transcriptional control <strong>and</strong><br />

biological impact. Nat Rev Mol Cell Biol. 2002<br />

Sep;3(9):651-62<br />

Zeng R, Aoki Y, Yoshida M, Arai K, Watanabe S. Stat5B<br />

shuttles between cytoplasm <strong>and</strong> nucleus in a cytokinedependent<br />

<strong>and</strong> -independent manner. J Immunol. 2002<br />

May 1;168(9):4567-75<br />

Calò V, Migliavacca M, Bazan V, Macaluso M, Buscemi M,<br />

Gebbia N, Russo A. STAT proteins: from normal control <strong>of</strong><br />

cellular events to tumorigenesis. J Cell Physiol. 2003<br />

Nov;197(2):157-68<br />

Kloth MT, Laughlin KK, Biscardi JS, Boerner JL, Parsons<br />

SJ, Silva CM. STAT5b, a Mediator <strong>of</strong> Synergism between<br />

c-Src <strong>and</strong> the Epidermal Growth Factor Receptor. J Biol<br />

Chem. 2003 Jan 17;278(3):1671-9<br />

Yamashita H, Iwase H, Toyama T, Fujii Y. Naturally<br />

occurring dominant-negative Stat5 suppresses<br />

transcriptional activity <strong>of</strong> estrogen receptors <strong>and</strong> induces<br />

apoptosis in T47D breast cancer cells. Oncogene. 2003<br />

Mar 20;22(11):1638-52<br />

Baśkiewicz-Masiuk M, Machaliński B. The role <strong>of</strong> the<br />

STAT5 proteins in the proliferation <strong>and</strong> apoptosis <strong>of</strong> the<br />

CML <strong>and</strong> AML cells. Eur J Haematol. 2004 Jun;72(6):420-<br />

9<br />

Chen H, Ye D, Xie X, Chen B, Lu W. VEGF, VEGFRs<br />

expressions <strong>and</strong> activated STATs in ovarian epithelial<br />

carcinoma. Gynecol Oncol. 2004 Sep;94(3):630-5<br />

Clevenger CV. Roles <strong>and</strong> regulation <strong>of</strong> stat family<br />

transcription factors in human breast cancer. Am J Pathol.<br />

2004 Nov;165(5):1449-60<br />

Silva CM. Role <strong>of</strong> STATs as downstream signal<br />

transducers in Src family kinase-mediated tumorigenesis.<br />

Oncogene. 2004 Oct 18;23(48):8017-23<br />

Sternberg DW, Gillil<strong>and</strong> DG. The role <strong>of</strong> signal transducer<br />

<strong>and</strong> activator <strong>of</strong> transcription factors in leukemogenesis. J<br />

Clin Oncol. 2004 Jan 15;22(2):361-71<br />

Xu L, Massagué J. Nucleocytoplasmic shuttling <strong>of</strong> signal<br />

transducers. Nat Rev Mol Cell Biol. 2004 Mar;5(3):209-19<br />

Sultan AS, Xie J, LeBaron MJ, Ealley EL, Nevalainen MT,<br />

Rui H. Stat5 promotes homotypic adhesion <strong>and</strong> inhibits<br />

invasive characteristics <strong>of</strong> human breast cancer cells.<br />

Oncogene. 2005 Jan 27;24(5):746-60<br />

Chia DJ, Subbian E, Buck TM, Hwa V, Rosenfeld RG,<br />

Skach WR, Shinde U, Rotwein P. Aberrant folding <strong>of</strong> a<br />

mutant Stat5b causes growth hormone insensitivity <strong>and</strong><br />

proteasomal dysfunction. J Biol Chem. 2006 Mar<br />

10;281(10):6552-8<br />

Han X, Osuntokun B, Benight N, Loesch K, Frank SJ,<br />

Denson LA. Signal transducer <strong>and</strong> activator <strong>of</strong> transcription<br />

5b promotes mucosal tolerance in pediatric Crohn's<br />

disease <strong>and</strong> murine colitis. Am J Pathol. 2006<br />

Dec;169(6):1999-2013<br />

Lee TK, Man K, Poon RT, Lo CM, Yuen AP, Ng IO, Ng KT,<br />

Leonard W, Fan ST. Signal transducers <strong>and</strong> activators <strong>of</strong><br />

transcription 5b activation enhances hepatocellular<br />

carcinoma aggressiveness through induction <strong>of</strong> epithelialmesenchymal<br />

transition. Cancer Res. 2006 Oct<br />

15;66(20):9948-56<br />

Sánchez-Ceja SG, Reyes-Maldonado E, Vázquez-<br />

Manríquez ME, López-Luna JJ, Belmont A, Gutiérrez-<br />

Castellanos S. Differential expression <strong>of</strong> STAT5 <strong>and</strong> BclxL,<br />

<strong>and</strong> high expression <strong>of</strong> Neu <strong>and</strong> STAT3 in non-smallcell<br />

lung carcinoma. Lung Cancer. 2006 Nov;54(2):163-8<br />

Strauss BL, Bratthauer GL, Tavassoli FA. STAT 5a<br />

expression in the breast is maintained in secretory<br />

carcinoma, in contrast to other histologic types. Hum<br />

Pathol. 2006 May;37(5):586-92<br />

Weaver AM, Silva CM. Modulation <strong>of</strong> signal transducer<br />

<strong>and</strong> activator <strong>of</strong> transcription 5b activity in breast cancer<br />

cells by mutation <strong>of</strong> tyrosines within the transactivation<br />

domain. Mol Endocrinol. 2006 Oct;20(10):2392-405<br />

Gong Y, Ishida-Takahashi R, Villanueva EC, Fingar DC,<br />

Münzberg H, Myers MG Jr. The long form <strong>of</strong> the leptin<br />

receptor regulates STAT5 <strong>and</strong> ribosomal protein S6 via<br />

alternate mechanisms. J Biol Chem. 2007 Oct<br />

19;282(42):31019-27<br />

Hwa V, Camacho-Hübner C, Little BM, David A, Metherell<br />

LA, El-Khatib N, Savage MO, Rosenfeld RG. Growth<br />

hormone insensitivity <strong>and</strong> severe short stature in siblings:<br />

a novel mutation at the exon 13-intron 13 junction <strong>of</strong> the<br />

STAT5b gene. Horm Res. 2007;68(5):218-24<br />

Mütze J, Roth J, Gerstberger R, Hübschle T. Nuclear<br />

translocation <strong>of</strong> the transcription factor STAT5 in the rat<br />

brain after systemic leptin administration. Neurosci Lett.<br />

2007 May 7;417(3):286-91<br />

Bessette K, Lang ML, Fava RA, Grundy M, Heinen J,<br />

Horne L, Spolski R, Al-Shami A, Morse HC 3rd, Leonard<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 107


STAT5B (signal transducer <strong>and</strong> activator <strong>of</strong> transcription 5B) Del Rosario AM, et al.<br />

WJ, Kelly JA. A Stat5b transgene is capable <strong>of</strong> inducing<br />

CD8+ lymphoblastic lymphoma in the absence <strong>of</strong> normal<br />

TCR/MHC signaling. Blood. 2008 Jan 1;111(1):344-50<br />

Brown S, Zeidler MP. Unphosphorylated STATs go<br />

nuclear. Curr Opin <strong>Gene</strong>t Dev. 2008 Oct;18(5):455-60<br />

Iyer J, Reich NC. Constitutive nuclear import <strong>of</strong> latent <strong>and</strong><br />

activated STAT5a by its coiled coil domain. FASEB J.<br />

2008 Feb;22(2):391-400<br />

Koppikar P, Lui VW, Man D, Xi S, Chai RL, Nelson E,<br />

Tobey AB, Gr<strong>and</strong>is JR. Constitutive activation <strong>of</strong> signal<br />

transducer <strong>and</strong> activator <strong>of</strong> transcription 5 contributes to<br />

tumor growth, epithelial-mesenchymal transition, <strong>and</strong><br />

resistance to epidermal growth factor receptor targeting.<br />

Clin Cancer Res. 2008 Dec 1;14(23):7682-90<br />

Bernaciak TM, Zareno J, Parsons JT, Silva CM. A novel<br />

role for signal transducer <strong>and</strong> activator <strong>of</strong> transcription 5b<br />

(STAT5b) in beta1-integrin-mediated human breast cancer<br />

cell migration. Breast Cancer Res. 2009;11(4):R52<br />

Vogtenhuber C, Bucher C, Highfill SL, Koch LK, Goren E,<br />

Panoskaltsis-Mortari A, Taylor PA, Farrar MA, Blazar BR.<br />

Constitutively active Stat5b in CD4+ T cells inhibits graftversus-host<br />

disease lethality associated with increased<br />

regulatory T-cell potency <strong>and</strong> decreased T effector cell<br />

responses. Blood. 2010 Jul 22;116(3):466-74<br />

Du W, Wang YC, Hong J, Su WY, Lin YW, Lu R, Xiong H,<br />

Fang JY. STAT5 is<strong>of</strong>orms regulate colorectal cancer cell<br />

apoptosis via reduction <strong>of</strong> mitochondrial membrane<br />

potential <strong>and</strong> generation <strong>of</strong> reactive oxygen species. J Cell<br />

Physiol. 2011 Aug 8;<br />

Koptyra M, Gupta S, Talati P, Nevalainen MT. Signal<br />

transducer <strong>and</strong> activator <strong>of</strong> transcription 5a/b: biomarker<br />

<strong>and</strong> therapeutic target in prostate <strong>and</strong> breast cancer. Int J<br />

Biochem Cell Biol. 2011 Oct;43(10):1417-21<br />

Lopez TV, Lappin TR, Maxwell P, Shi Z, Lopez-Marure R,<br />

Aguilar C, Rocha-Zavaleta L. Autocrine/paracrine<br />

erythropoietin signalling promotes JAK/STAT-dependent<br />

proliferation <strong>of</strong> human cervical cancer cells. Int J Cancer.<br />

2011 Dec 1;129(11):2566-76<br />

Nadeau K, Hwa V, Rosenfeld RG. STAT5b deficiency: an<br />

unsuspected cause <strong>of</strong> growth failure, immunodeficiency,<br />

<strong>and</strong> severe pulmonary disease. J Pediatr. 2011<br />

May;158(5):701-8<br />

Peck AR, Witkiewicz AK, Liu C, Stringer GA, Klimowicz<br />

AC, Pequignot E, Freydin B, Tran TH, Yang N, Rosenberg<br />

AL, Hooke JA, Kovatich AJ, Nevalainen MT, Shriver CD,<br />

Hyslop T, Sauter G, Rimm DL, Magliocco AM, Rui H. Loss<br />

<strong>of</strong> nuclear localized <strong>and</strong> tyrosine phosphorylated Stat5 in<br />

breast cancer predicts poor clinical outcome <strong>and</strong> increased<br />

risk <strong>of</strong> antiestrogen therapy failure. J Clin Oncol. 2011 Jun<br />

20;29(18):2448-58<br />

This article should be referenced as such:<br />

Del Rosario AM, Bernaciak TM, Silva CM. STAT5B (signal<br />

transducer <strong>and</strong> activator <strong>of</strong> transcription 5B). <strong>Atlas</strong> <strong>Gene</strong>t<br />

Cytogenet Oncol Haematol. 2012; 16(2):104-108.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 108


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

<strong>Gene</strong> <strong>Section</strong><br />

Review<br />

AAMP (angio-associated, migratory cell<br />

protein)<br />

Marie E Beckner<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

Department <strong>of</strong> Pathology, Louisiana State University Health Sciences Center - Shreveport, USA<br />

(MEB)<br />

Published in <strong>Atlas</strong> Database: September 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/<strong>Gene</strong>s/AAMPID533ch2q35.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI AAMPID533ch2q35.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Identity<br />

HGNC (Hugo): AAMP<br />

Location: 2q35<br />

DNA/RNA<br />

Note<br />

The NCBI RefSeq (Aug-2011) consensus sequence<br />

for AAMP differs at the 5-prime end from the<br />

manually sequenced version published initially. A<br />

diagram <strong>of</strong> the NCBI ReqfSeq (Aug-2011) version<br />

for rAAMP is shown here.<br />

Description<br />

The AAMP gene (NCBI RefSeq, Aug-2011)<br />

encompasses 6042 bp; 11 exons.<br />

Transcription<br />

1859 bp mRNA (NCBI RefSeq, Aug-2011).<br />

Pseudogene<br />

None are known.<br />

rAAMP encoding a 434 aa protein in normal cells. The codon for AAMP's initiating methionine, the stop codon, <strong>and</strong> the polyadenylation<br />

sites are indicated at 85-87, 1387-1389, <strong>and</strong> 1774-1779, respectively. Untranslated sequence is indicated in the<br />

hatched regions at the 3' end <strong>of</strong> exon 1 <strong>and</strong> the 5' end <strong>of</strong> exon 11. The poly-adenosine tail, 1793-1859, is included.<br />

The bulk <strong>of</strong> AAMP's sequence is constituted by WD40 domains that are known to commonly fold to form a platform for active<br />

portions <strong>of</strong> a protein. Thus the stretch <strong>of</strong> contiguous glutamic acid residues should be available for interactions with other<br />

proteins. Also, binding <strong>of</strong> other proteins may occur to regions <strong>of</strong> these repeats. The WD40 repeats are known to mediate<br />

aggregation <strong>of</strong> subunits to form complex, multi-protein structures. Two immunoglobulin-type domains are designated by pairs <strong>of</strong><br />

cysteines, (C96 - C130) <strong>and</strong> (C216 - C265). AAMP may play a negative role in Nod2-mediated NF-kB activation via a physical<br />

interaction in the region indicated. A physical interaction between AAMP <strong>and</strong> thromboxane A2 receptors (TPalpha <strong>and</strong> TPbeta)<br />

has been suggested to occur via multiple sites with no particular domain identified in localization studies.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 109


AAMP (angio-associated, migratory cell protein) Beckner ME<br />

Protein<br />

Note<br />

The NCBI RefSeq (Aug-2011) consensus sequence<br />

for AAMP (434 aa) is shorter than initially reported<br />

(452 aa, 52 kDa).<br />

Description<br />

434 amino acids, 49 kDa protein (NCBI RefSeq,<br />

Aug-2011).<br />

Expression<br />

AAMP is widely expressed among many types <strong>of</strong><br />

mammalian cells <strong>and</strong> has been conserved in<br />

evolution.<br />

Localisation<br />

AAMP was initially detected in the cytoplasm <strong>of</strong><br />

many types <strong>of</strong> nucleated mammalian cells <strong>and</strong> was<br />

strongly expressed in endothelial cells,<br />

cytotrophoblasts, <strong>and</strong> poorly differentiated colon<br />

adenocarcinoma cells in lymphatics. AAMP has<br />

been observed at the luminal edges <strong>of</strong> endometrial<br />

cells <strong>and</strong> has been found in the extracellular<br />

environment <strong>of</strong> vascular-associated mesenchymal<br />

cells.<br />

Function<br />

Functional studies <strong>and</strong> associations suggest roles<br />

for AAMP in angiogenesis, including endothelial<br />

tube formation, migration <strong>of</strong> endothelial <strong>and</strong><br />

smooth muscle cells, neointima formation, <strong>and</strong><br />

thromboxane A receptor interactions. In immune<br />

cells AAMP may be involved in the regulation <strong>of</strong><br />

NF-kappaB activation mediated by Nod2. The<br />

common occurrence <strong>of</strong> WD40 domains in signaling<br />

proteins supports a signaling function as a<br />

possibility for AAMP. The epitope, ESESES, that<br />

AAMP shares with alpha-actinin <strong>and</strong> a smaller<br />

protein specific for fast skeletal muscle, suggests<br />

that it may have cytoskeletal interactions. Although<br />

AAMP was described as having heparin-binding<br />

capacity in melanoma cells, the NCBI RefSeq<br />

(Aug-2011) version <strong>of</strong> AAMP does not include the<br />

heparin-binding sequence (encodes RRLRR) in the<br />

coding sequence. The RefSeq version <strong>of</strong> AAMP<br />

identifies the initiating methionine (85-87) as being<br />

3' to the sequence encoding RRLRR (70-84).<br />

Homology<br />

AAMP shares homology with the other members <strong>of</strong><br />

the WD40 repeat superfamily. Several <strong>of</strong> the WD40<br />

repeat proteins also contain an amino terminal run<br />

<strong>of</strong> glutamic acid residues outside <strong>of</strong> their WD<br />

repeats. The two immunoglobulin-type domains in<br />

AAMP resemble those <strong>of</strong> the immunoglobulin<br />

superfamily members, including domains <strong>of</strong><br />

NCAM, DCC, NgCAM, etc.<br />

Mutations<br />

Note<br />

AAMP was initially cloned <strong>and</strong> sequenced as a<br />

transcript for a 452 aa, 52 kDa protein. It was<br />

obtained from an expression library derived from<br />

melanoma cells <strong>of</strong> a brain metastasis. However the<br />

amino terminus was missing from the library clone.<br />

AAMP's amino terminus was determined by 5'<br />

RACE from human brain mRNA. Two versions<br />

were obtained <strong>and</strong> both differed from the NCBI<br />

RefSeq (Aug-2011) version <strong>of</strong> AAMP. The one<br />

shown contains coding sequence upstream from the<br />

initiating methionine in the RefSeq version. This<br />

alternative form may represent a fusion protein due<br />

to an in-frame insertion or a chromosomal<br />

rearrangement. AAMP potentially gains heparin<br />

binding functionality when sequence upstream from<br />

AUG at nucleotides 85-87 is preceded by an<br />

alternative initiating methionine codon that enables<br />

translation <strong>of</strong> the codons for RRLRR.<br />

Germinal<br />

None are known.<br />

Somatic<br />

Chromosomal rearrangements or insertions that<br />

place an initiating methionine codon further<br />

upstream than the methionine at 85-87 permit<br />

AAMP to gain heparin binding function when<br />

nucleotides 70-84 are translated.<br />

rAAMP encoding a 452 aa protein (alternative form). An alternative AUG was detected when the amino terminus was<br />

manually sequenced with 5' RACE using brain mRNA. It is shown in red. Its upstream location permits 17 codons to encode<br />

amino acids, including 5 codons for RRLRR, a heparin-binding site. Sequences <strong>of</strong> the alternative forms described for AAMP are<br />

the same as the NCBI RefSeq (Aug-2011) version <strong>of</strong> AAMP for nucleotides 34-1792 <strong>and</strong> the first 25 adenosines in the poly-A<br />

tail.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 110


AAMP (angio-associated, migratory cell protein) Beckner ME<br />

Implicated in<br />

Gastrointestinal stromal tumor (GIST)<br />

Note<br />

GISTs are the most common mesenchymal tumors<br />

<strong>of</strong> the digestive tract <strong>and</strong> are believed to arise from<br />

the interstitial cells <strong>of</strong> Cajal. They respond to<br />

imitanib, a tyrosine kinase inhibitor, which is also<br />

used to treat myeloid leukemia. Expression <strong>of</strong><br />

AAMP was found to be increased in GISTs with<br />

mutated KIT. Expression <strong>of</strong> AAMP among various<br />

s<strong>of</strong>t tissue sarcomas <strong>and</strong> normal tissues was highest<br />

in the GISTs. AAMP ranks high among the<br />

upregulated genes in GISTs.<br />

Myeloid leukemia (chronic (CML) <strong>and</strong><br />

acute (AML) forms)<br />

Note<br />

AAMP is expressed in myeloid leukemia cell lines<br />

<strong>and</strong> its expression is repressed by imitanib<br />

(mainline treatment drug for chronic myeloid<br />

leukemia), deferasirox (iron chelator that decreases<br />

cell proliferation), <strong>and</strong> anisomycin, an inducer <strong>of</strong><br />

apoptosis.<br />

Lymphoma (B <strong>and</strong> T cell origins, non-<br />

Hodgkins <strong>and</strong> Hodgkins types)<br />

Note<br />

AAMP is expressed in activated T lymphocytes,<br />

monocytes, <strong>and</strong> lymphoma cells. Compared to<br />

normal B cells, AAMP is increased in non-<br />

Hodgkins lymphomas <strong>and</strong> in classical Hodgkins<br />

lymphoma.<br />

Melanoma (melanocyte origin,<br />

usually skin)<br />

Note<br />

In lysates <strong>of</strong> a melanoma cell line obtained from a<br />

brain metastasis, the size <strong>of</strong> the protein that reacted<br />

with anti-AAMP was 52 kDa, thus corresponding to<br />

the size predicted by an alternative transcript. The<br />

alternative version <strong>of</strong> AAMP could have resulted<br />

from a chromosomal rearrangement due to 2q35<br />

breakage in the amino terminal region <strong>of</strong> AAMP.<br />

Placement <strong>of</strong> AUG at the 5' end <strong>of</strong> nucleotide 34 to<br />

serve as an alternative initiating methionine codon<br />

permits expression <strong>of</strong> the heparin binding site,<br />

RRLRR (nucleotides 70-84 in the NCBI ReqSeq<br />

(Aug-2011) version <strong>of</strong> AAMP) as a gain <strong>of</strong><br />

function.<br />

Breast cancer (ductal cell origin most<br />

common)<br />

Note<br />

Expression pr<strong>of</strong>iling <strong>of</strong> human breast cancer cells<br />

versus mammary epithelial cells revealed higher<br />

expression <strong>of</strong> AAMP in the tumor cells. AAMP<br />

expression is higher in ductal carcinoma in situ<br />

(DCIS) with necrosis compared to DCIS without<br />

necrosis. However, expression pr<strong>of</strong>iling <strong>of</strong> DCIS<br />

<strong>and</strong> invasive ductal carcinoma (IDC) paired<br />

specimens from the same patients revealed<br />

decreased AAMP in IDC. AAMP expression may<br />

be relevant for the development <strong>of</strong> DCIS.<br />

Glial brain tumors (neuroectodermal<br />

cell origin)<br />

Note<br />

When compared with normal, glioblastomas (Grade<br />

IV astrocytomas) <strong>and</strong> oligodendrogliomas (Grades<br />

II - III) demonstrated slightly elevated levels <strong>of</strong><br />

AAMP expression. Expression <strong>of</strong> AAMP was<br />

increased in drug resistant glioblastomas, primary<br />

<strong>and</strong> recurrent.<br />

Colon neoplasia (benign adenomas<br />

<strong>and</strong> carcinoma arise from gl<strong>and</strong>ular<br />

cells <strong>of</strong> the mucosa in the<br />

gastrointestinal tract)<br />

Note<br />

Expression pr<strong>of</strong>iling <strong>of</strong> normal mucosa <strong>and</strong><br />

colorectal adenomas from the same patients showed<br />

that AAMP was slightly increased in the adenomas.<br />

However, a small dataset <strong>of</strong> colon tubular<br />

adenomas harboring focal adenocarcinomas, with<br />

microdissections <strong>of</strong> paired samples from each,<br />

showed that AAMP may become decreased in the<br />

incipient carcinomas. Although AAMP may play a<br />

role in the development <strong>of</strong> colonic neoplasia, it has<br />

not been shown to be positively involved in<br />

progression <strong>of</strong> adenomas to malignancy.<br />

Epidermoid carcinoma (keratinocyte<br />

origin if arises in skin)<br />

Note<br />

AAMP is expressed in squamous carcinoma cell<br />

lines. In one cell line studied with PTEN knocked<br />

down, the expression <strong>of</strong> AAMP also fell<br />

significantly. Also, for tumors from a radiosensitive<br />

cell line, AAMP expression fell significantly in<br />

radioresistant tumors derived from the sensitive cell<br />

line.<br />

Cervical cancer (usually arises from<br />

squamous type cells)<br />

Note<br />

In a study <strong>of</strong> cervical carcinoma HeLa cells treated<br />

with epidermal growth factor in a time course,<br />

expression <strong>of</strong> AAMP was increased at 2-8 hours.<br />

Ovarian cancer (<strong>of</strong>ten arises from<br />

serous cells)<br />

Note<br />

Cancer cells prepared from primary cultures <strong>of</strong><br />

ovarian papillary serous adenocarcinomas revealed<br />

increased AAMP expression in 2 <strong>of</strong> 3 carboplatin<br />

resistant tumors whereas none <strong>of</strong> the tumors from 3<br />

patients with sensitivity to carboplatin<br />

demonstrated comparable levels.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 111


AAMP (angio-associated, migratory cell protein) Beckner ME<br />

Papillary thyroid carcinoma (arises<br />

from cells in thyroid follicles)<br />

Note<br />

Analysis <strong>of</strong> papillary thyroid carcinoma tumors<br />

matched with normal thyroid from nine patients<br />

found that AAMP expression fell slightly in the<br />

tumors.<br />

Pulmonary carcinoma (multiple types<br />

<strong>of</strong> cells can be the origin)<br />

Note<br />

AAMP's expression in a lung alveolar<br />

adenocarcinoma cell line was down-regulated by<br />

TGF-beta that was added to induce an epithelialmesenchymal<br />

transition.<br />

Chromosomal rearrangements at<br />

2q35<br />

Note<br />

The location <strong>of</strong> AAMP at 2q35 imparts<br />

susceptibility to chromosomal rearrangements<br />

involving the terminal region <strong>of</strong> chromosome 2's<br />

long arm, q. Terminal regions are more susceptible<br />

to rearrangements than the midregions <strong>of</strong><br />

chromosomes. Fusion transcripts <strong>and</strong> proteins result<br />

from breakage <strong>and</strong> rearrangements that occur in<br />

unstable genomes. Several fusion proteins resulting<br />

from rearrangements involving other genes in this<br />

region have been reported in association with<br />

malignancy. Rearrangements that place in-frame<br />

coding sequence upstream from nucleotides 70-84<br />

in the NCBI ReqSeq (Aug-2011) version <strong>of</strong> AAMP<br />

permit a gain <strong>of</strong> function, i.e. heparin-binding, to<br />

occur.<br />

Breakpoints<br />

Note<br />

Clones obtained in 5' RACE studies <strong>of</strong> AAMP's<br />

amino terminus revealed a breakpoint between<br />

nucleotides 30 <strong>and</strong> 31 with inclusion <strong>of</strong> upstream<br />

sequence that did not match the NCBI RefSeq<br />

(Aug-2011) form <strong>of</strong> AAMP <strong>and</strong> also another<br />

version with a breakpoint between nucleotides 33<br />

<strong>and</strong> 34 <strong>and</strong> introduction <strong>of</strong> an alternative AUG. A<br />

schematic <strong>of</strong> the latter alternative form was shown<br />

earlier. The location <strong>of</strong> AAMP at 2q35 places it in a<br />

region near the end region <strong>of</strong> the chromosomal arm<br />

where breakpoints are more likely compared to the<br />

centromeric region. Importantly, in the sequences<br />

<strong>of</strong> the consensus <strong>and</strong> variant forms, in-frame<br />

codons are present for a heparin binding region that<br />

can be translated if an alternative initiating<br />

methionine is introduced upstream.<br />

References<br />

Beckner ME, Krutzsch HC, Stracke ML, Williams ST,<br />

Gallardo JA, Liotta LA. Identification <strong>of</strong> a new<br />

immunoglobulin superfamily protein expressed in blood<br />

vessels with a heparin-binding consensus sequence.<br />

Cancer Res. 1995 May 15;55(10):2140-9<br />

Beckner ME, Krutzsch HC, Klipstein S, Williams ST,<br />

Maguire JE, Doval M, Liotta LA. AAMP, a newly identified<br />

protein, shares a common epitope with alpha-actinin <strong>and</strong> a<br />

fast skeletal muscle fiber protein. Exp Cell Res. 1996 Jun<br />

15;225(2):306-14<br />

Beckner ME, Liotta LA. AAMP, a conserved protein with<br />

immunoglobulin <strong>and</strong> WD40 domains, regulates endothelial<br />

tube formation in vitro. Lab Invest. 1996 Jul;75(1):97-107<br />

Beckner ME, Krutzsch HC, Tsokos M, Moul DE, Liotta LA.<br />

The aggregated form <strong>of</strong> an AAMP derived peptide behaves<br />

as a heparin sensitive cell binding agent. Biotechnol<br />

Bioeng. 1997 May 20;54(4):365-72<br />

Beckner ME, Peterson VA, Moul DE. Angio-associated<br />

migratory cell protein is expressed as an extracellular<br />

protein by blood-vessel-associated mesenchymal cells.<br />

Microvasc Res. 1999 May;57(3):347-52<br />

All<strong>and</strong>er SV, Nupponen NN, Ringnér M, Hostetter G,<br />

Maher GW, Goldberger N, Chen Y, Carpten J, Elkahloun<br />

AG, Meltzer PS. Gastrointestinal stromal tumors with KIT<br />

mutations exhibit a remarkably homogeneous gene<br />

expression pr<strong>of</strong>ile. Cancer Res. 2001 Dec 15;61(24):8624-<br />

8<br />

Adeyinka A, Emberley E, Niu Y, Snell L, Murphy LC,<br />

Sowter H, Wyk<strong>of</strong>f CC, Harris AL, Watson PH. Analysis <strong>of</strong><br />

gene expression in ductal carcinoma in situ <strong>of</strong> the breast.<br />

Clin Cancer Res. 2002 Dec;8(12):3788-95<br />

Beckner ME, Jagannathan S, Peterson VA. Extracellular<br />

angio-associated migratory cell protein plays a positive<br />

role in angiogenesis <strong>and</strong> is regulated by astrocytes in<br />

coculture. Microvasc Res. 2002 May;63(3):259-69<br />

Blindt R, Vogt F, Lamby D, Zeiffer U, Krott N, Hilger-<br />

Eversheim K, Hanrath P, vom Dahl J, Bosserh<strong>of</strong>f AK.<br />

Characterization <strong>of</strong> differential gene expression in<br />

quiescent <strong>and</strong> invasive human arterial smooth muscle<br />

cells. J Vasc Res. 2002 Jul-Aug;39(4):340-52<br />

Francis P, Namløs HM, Müller C, Edén P, Fernebro J, et<br />

al. Diagnostic <strong>and</strong> prognostic gene expression signatures<br />

in 177 s<strong>of</strong>t tissue sarcomas: hypoxia-induced transcription<br />

pr<strong>of</strong>ile signifies metastatic potential. BMC Genomics. 2007<br />

Mar 14;8:73<br />

Holvoet P, Sinnaeve P. Angio-associated migratory cell<br />

protein <strong>and</strong> smooth muscle cell migration in development<br />

<strong>of</strong> restenosis <strong>and</strong> atherosclerosis. J Am Coll Cardiol. 2008<br />

Jul 22;52(4):312-4<br />

Vogt F, Zernecke A, Beckner M, Krott N, Bosserh<strong>of</strong>f AK,<br />

H<strong>of</strong>fmann R, Z<strong>and</strong>voort MA, Jahnke T, Kelm M, Weber C,<br />

Blindt R. Blockade <strong>of</strong> angio-associated migratory cell<br />

protein inhibits smooth muscle cell migration <strong>and</strong><br />

neointima formation in accelerated atherosclerosis. J Am<br />

Coll Cardiol. 2008 Jul 22;52(4):302-11<br />

Bielig H, Zurek B, Kutsch A, Menning M, Philpott DJ,<br />

Sansonetti PJ, Kufer TA. A function for AAMP in Nod2mediated<br />

NF-kappaB activation. Mol Immunol. 2009<br />

Aug;46(13):2647-54<br />

Reid HM, Wikström K, Kavanagh DJ, Mulvaney EP,<br />

Kinsella BT. Interaction <strong>of</strong> angio-associated migratory cell<br />

protein with the TPα <strong>and</strong> TPβ is<strong>of</strong>orms <strong>of</strong> the human<br />

thromboxane A₂ receptor. Cell Signal. 2011<br />

Apr;23(4):700-17<br />

This article should be referenced as such:<br />

Beckner ME. AAMP (angio-associated, migratory cell<br />

protein). <strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012;<br />

16(2):109-112.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 112


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

<strong>Gene</strong> <strong>Section</strong><br />

Mini Review<br />

BCL2L15 (BCL2-like 15)<br />

Maria-Angeliki S Pavlou, Christos K Kontos<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

Westfalische Wilhelms-Universitat Munster, ZMBE, Institute <strong>of</strong> Cell Biology, Stem Cell Biology <strong>and</strong><br />

Regeneration Group, Von-Esmarch-Str 56, 48149 Munster, Germany (MASP), Department <strong>of</strong><br />

Biochemistry <strong>and</strong> Molecular Biology, Faculty <strong>of</strong> Biology, University <strong>of</strong> Athens, 15701,<br />

Panepistimiopolis, Athens, Greece (CKK)<br />

Published in <strong>Atlas</strong> Database: September 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/<strong>Gene</strong>s/BCL2L15ID46259ch1p13.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI BCL2L15ID46259ch1p13.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Identity<br />

Other names: Bfk, C1orf178, FLJ22588<br />

HGNC (Hugo): BCL2L15<br />

Location: 1p13.2<br />

Local order: Centromere to telomere.<br />

DNA/RNA<br />

Description<br />

The BCl2L15 gene has a total length <strong>of</strong> 10734 nt<br />

<strong>and</strong> consists <strong>of</strong> 4 exons <strong>and</strong> 3 intervening introns<br />

(Coultas et al., 2003). The organization <strong>of</strong> the<br />

BCL2L15 gene, with the BH3 domain located on a<br />

single exon (exon 2) <strong>and</strong> the BH2 domain split<br />

between two exons (exons 3 <strong>and</strong> 4), is similar to<br />

that <strong>of</strong> other BCL2 family members, including<br />

BCL2, BCL2L1 (BCLX), BAX, <strong>and</strong> BAK1 (BAK)<br />

(Petros et al., 2004).<br />

Figure 1. Schematic representation <strong>of</strong> the BCL2L15 gene. Exons are shown as boxes <strong>and</strong> introns as connecting lines. The<br />

coding sequences are highlighted as red, while 5' <strong>and</strong> 3' untranslated regions (UTRs) are shown in white. Numbers inside or<br />

outside boxes indicate lengths (nt) <strong>of</strong> exons <strong>and</strong> introns, respectively, while numbers in parentheses indicate lengths (aa) <strong>of</strong><br />

protein is<strong>of</strong>orms. Arrows (↓) show the position <strong>of</strong> the start codons (ATG) <strong>and</strong> asterisks (*) denote the position <strong>of</strong> the stop codons<br />

(TGA). Question marks (?) indicate that the full-length sequence was not determined. Roman numerals indicate intron phases.<br />

The intron phase refers to the location <strong>of</strong> the intron within the codon; I denotes that the intron occurs after the first nucleotide <strong>of</strong><br />

the codon, II denotes that the intron occurs after the second nucleotide, <strong>and</strong> 0 means that the intron occurs between distinct<br />

codons. The figure is drawn to scale, except for the introns containing the (//) symbol.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 113


BCL2L15 (BCL2-like 15) Pavlou MAS, Kontos CK<br />

Transcription<br />

The BCL2L15 gene is subjected to alternative<br />

splicing, generating four splice variants, three <strong>of</strong><br />

which are considered as coding transcripts. Each<br />

coding splice variant consists <strong>of</strong> a distinctive exon<br />

combination <strong>and</strong> encodes a different protein<br />

is<strong>of</strong>orm. The predominant transcript, consisting <strong>of</strong><br />

4973 nt, includes all 4 exons <strong>and</strong> encodes is<strong>of</strong>orm<br />

a. The second transcript, predicted to encode<br />

is<strong>of</strong>orm b, contains exons 1, 2 <strong>and</strong> 4. The deletion<br />

<strong>of</strong> exon 3 does not result in frameshifting. The third<br />

transcript, putatively encoding is<strong>of</strong>orm d, consists<br />

<strong>of</strong> exons 1 <strong>and</strong> 4. The lack <strong>of</strong> exons 2 <strong>and</strong> 3 shifts<br />

the open reading frame.<br />

As aforementioned, except for alternatively spliced<br />

BCL2L15 coding variants, another noncoding<br />

transcript has also been identified. This one is<br />

composed <strong>of</strong> exons 1, 3 <strong>and</strong> 4, <strong>and</strong> was initially<br />

considered to encode is<strong>of</strong>orm c. Nonetheless, this<br />

transcript is a nonsense-mediated mRNA decay<br />

(NMD) c<strong>and</strong>idate, since deletion <strong>of</strong> exon 2<br />

generates a premature translation termination codon<br />

in exon 3.<br />

Interestingly, transcription <strong>of</strong> all BCL2L15<br />

alternatively spliced variants was noticed only in<br />

colon, while the full-length transcript was also<br />

detected in stomach, rectum, small intestine,<br />

cerebellum, testis <strong>and</strong> uterus (Dempsey et al.,<br />

2005). Moreover, despite the fact that a p53<br />

consensus binding site was identified upstream <strong>of</strong><br />

the transcription initiation site <strong>of</strong> BCL2L15, this<br />

gene does not constitute a transcriptional target <strong>of</strong><br />

p53 (TP53) (Ozören et al., 2009).<br />

Pseudogene<br />

Not identified so far.<br />

Protein<br />

Description<br />

The full-length BCL2L15 is<strong>of</strong>orm (is<strong>of</strong>orm a) is the<br />

predominant one. It consists <strong>of</strong> 163 amino acid<br />

residues <strong>and</strong> has a molecular mass <strong>of</strong> 17.7 kDa.<br />

BCL2L15 is<strong>of</strong>orm a contains a BH3 <strong>and</strong> a BH2<br />

domain, but no BH1, BH4 or hydrophobic tail<br />

(Coultas et al., 2003). Is<strong>of</strong>orm a is the predominant<br />

BCL2L15 is<strong>of</strong>orm <strong>and</strong> the only one that has been in<br />

vivo detected so far.<br />

The amino acid sequences <strong>of</strong> is<strong>of</strong>orms b <strong>and</strong> c are<br />

deduced from the mRNA sequences <strong>of</strong> the<br />

BCL2L15 alternatively spliced variants, <strong>and</strong> remain<br />

to be experimentally validated <strong>and</strong> in vivo detected.<br />

Is<strong>of</strong>orm b is a putative BH3-only protein <strong>of</strong> 88<br />

amino acid residues, with a calculated molecular<br />

mass <strong>of</strong> 9.6 kDa. This is<strong>of</strong>orm shares the same<br />

termini with BCL2L15 is<strong>of</strong>orm; still, it bears no<br />

BH2 domain. Finally, is<strong>of</strong>orm d is the smallest<br />

predicted BCL2L15 is<strong>of</strong>orm. This protein <strong>of</strong> 56<br />

amino acid residues, with a molecular mass <strong>of</strong> 6.3<br />

kDa, possesses no BCL2-homology (BH) domains<br />

(Dempsey et al., 2005).<br />

The N-terminal region <strong>of</strong> all BCL2L15 is<strong>of</strong>orms<br />

shares partial homology (ECIxNxLxxxFL peptide)<br />

with BID (Dempsey et al., 2005), a BH3-only<br />

proapoptotic member <strong>of</strong> the BCL2 family<br />

(Lomonosova <strong>and</strong> Chinnadurai, 2008). Moreover,<br />

all BCL2L15 is<strong>of</strong>orms contain a caspase-3/caspase-<br />

7 cleavage site (DEVD peptide) (Dempsey et al.,<br />

2005). This tetrapeptide, corresponding to amino<br />

acid residues 38-41, is responsible for the removal<br />

<strong>of</strong> an N-terminal peptide fragment <strong>and</strong> the<br />

subsequent activation <strong>of</strong> the predominant BCL2L15<br />

is<strong>of</strong>orm, at least during DNA damage-induced<br />

apoptosis (Dempsey et al., 2005; Ozören et al.,<br />

2009).<br />

Figure 2. Alignment <strong>of</strong> amino acid sequences <strong>and</strong> structural motifs <strong>of</strong> the BC2L15 protein is<strong>of</strong>orms. Light blue <strong>and</strong> pink<br />

denote the BH2 <strong>and</strong> BH3 domains, respectively, while the amino acid residues constituting the consensus sequence <strong>of</strong> each<br />

BCL2 homology domain are shown in dark blue <strong>and</strong> red color. Yellow highlights the site <strong>of</strong> caspase-3/7 cleavage (DEVD<br />

tetrapeptide), which is considered to be critical for the activation <strong>of</strong> the proapoptotic action <strong>of</strong> BCL2L15, at least in certain cell<br />

types <strong>and</strong>/or after certain stimuli, including DNA damage-induced apoptosis. Finally, light green highlights the ECIxNxLxxxFL<br />

peptide, which BCL2L15 is<strong>of</strong>orms share with BID; its conserved amino acid residues are shown in dark green.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 114


BCL2L15 (BCL2-like 15) Pavlou MAS, Kontos CK<br />

Expression<br />

The BCL2L15 protein is mainly expressed in<br />

tissues <strong>of</strong> the gastrointestinal tract, including the<br />

stomach, small intestine, colon <strong>and</strong> rectum<br />

(Dempsey et al., 2005; Ozören et al., 2009). The<br />

full-length is<strong>of</strong>orm has also been detected in several<br />

colorectal cancer cell lines, such as SW480, HT-29<br />

<strong>and</strong> HCT116 (Ozören et al., 2009).<br />

Localisation<br />

The BCL2L15 protein is localized to the cytoplasm<br />

<strong>of</strong> intestinal epithelial cells (Ozören et al., 2009). It<br />

does not possess any signal peptide or C-terminal<br />

membrane anchor <strong>and</strong>, consequently, it is not<br />

associated with any cellular organelles (Coultas et<br />

al., 2003; Ozören et al., 2009), unlike other<br />

members <strong>of</strong> the BCL2 family (Thomadaki <strong>and</strong><br />

Scorilas, 2006). The localization <strong>of</strong> the cleaved<br />

BCL2L15 has not been elucidated yet.<br />

Function<br />

BCL2L15 is a weakly proapoptotic member <strong>of</strong> the<br />

BCL2 family (Coultas et al., 2003; Dempsey et al.,<br />

2005; Pujianto et al., 2007). When overexpressed,<br />

the full-length BCL2L15 is<strong>of</strong>orm interacts with<br />

BCL2L1 long is<strong>of</strong>orm (BCLXL) <strong>and</strong> BCL2L2<br />

(BCLW), but not with BCL2 or BAD, as revealed<br />

by co-immunoprecipitation experiments (Ozören et<br />

al., 2009). Furthermore, it has been speculated that<br />

BCL2L15 acts most probably as an amplifier <strong>of</strong> the<br />

apoptotic signal rather than a trigger <strong>of</strong><br />

programmed cell death (Pujianto et al., 2007;<br />

Ozören et al., 2009).<br />

Given the weak proapoptotic activity <strong>of</strong> BCL2L15,<br />

it was initially suggested that the full-length<br />

BCL2L15 could represent the latent form <strong>of</strong> a<br />

potent BH3-only protein exerting its proapoptotic<br />

action once activated through proteolytic cleavage<br />

(Coultas et al., 2003), like caspase-8 cleavage <strong>of</strong><br />

BID (Li et al., 1998; Luo et al., 1998), at least in<br />

certain cell types or after certain stimuli. In support<br />

<strong>of</strong> this notion, it was shown that BCL2L15 becomes<br />

cleaved in a caspase-dependent manner during<br />

DNA damage-induced apoptosis <strong>and</strong> that truncated<br />

BCL2L15 (~13 kDa), corresponding to the part <strong>of</strong><br />

protein downstream <strong>of</strong> the caspase-3/7 cleavage<br />

site, is capable <strong>of</strong> inducing apoptosis in HCT116<br />

cells, in contrast to the full-length BCL2L15<br />

is<strong>of</strong>orm, which seems to be incapable <strong>of</strong> inducing<br />

apoptosis in HCT116 or SW480 colorectal cancer<br />

cells. Interestingly, the ability <strong>of</strong> the cleaved form<br />

<strong>of</strong> the BCL2L15 protein to induce apoptosis is<br />

dependent on the presence <strong>of</strong> the BAX or BAK1<br />

(BAK). Furthermore, co-expression <strong>of</strong> the<br />

antiapoptotic BCL2L1 long is<strong>of</strong>orm (BCLXL) or<br />

BCL2L2 (BCLW) antagonizes efficiently the<br />

killing activity <strong>of</strong> truncated BCL2L15 (Ozören et<br />

al., 2009).<br />

On the other h<strong>and</strong>, it has been proposed that the<br />

proapoptotic role <strong>of</strong> BCL2L15 may resemble more<br />

that <strong>of</strong> BAX <strong>and</strong> BAK1 (BAK) than that <strong>of</strong> BH3only<br />

proteins, since it most probably has a structure<br />

similar to that <strong>of</strong> BCL2 <strong>and</strong> BAX. In fact, the<br />

position <strong>of</strong> BH3 <strong>and</strong> BH2 domains in the BCL2L15<br />

protein is conserved relative to BAX <strong>and</strong> BCL2<br />

(Coultas et al., 2003).<br />

Potential phosphorylation at Ser-96 <strong>and</strong>/or Ser-42<br />

as well as other post-translational modifications <strong>of</strong><br />

BCL2L15 might change its subcellular localization<br />

<strong>and</strong> further regulate its physiological function<br />

(Dempsey et al., 2005; Pujianto et al., 2007).<br />

Homology<br />

Human BCL2L15 shares 71% amino acid identity<br />

<strong>and</strong> 80% similarity with the mouse ortholog.<br />

BCL2L15 bears the same combination <strong>of</strong> BCL2homology<br />

domains (BH2 <strong>and</strong> BH3) as the<br />

BCL2L14 long is<strong>of</strong>orm (BCLGL) <strong>and</strong> BCL2L12<br />

full-length is<strong>of</strong>orm, thus lacking other domains that<br />

are common among BCL2 family members (BH1<br />

<strong>and</strong> BH4) or a hydrophobic tail (Youle <strong>and</strong><br />

Strasser, 2008).<br />

Mutations<br />

Note<br />

A single nucleotide polymorphism (SNP) has been<br />

detected in the coding sequence (GAC→AAC) <strong>of</strong><br />

the BCL2L15 gene, which results in the substitution<br />

<strong>of</strong> an amino acid residue bearing a negatively<br />

charged side chain by an amino acid with a polar<br />

uncharged side chain (D→N).<br />

Implicated in<br />

Gastrointestinal cancer, particularly<br />

colorectal carcinoma<br />

Prognosis<br />

BCL2L15 mRNA expression is clearly reduced in a<br />

wide range <strong>of</strong> gastrointestinal malignancies.<br />

BCL2L15 mRNA levels are lower in colon tumors,<br />

compared to levels detected in matched normal<br />

colon tissue. Moreover, BCL2L15 mRNA<br />

expression is significantly downregulated in tumors<br />

<strong>of</strong> the small intestine, stomach <strong>and</strong> rectum. This<br />

reduction <strong>of</strong> BCL2L15 mRNA levels in<br />

gastrointestinal neoplasms implies that BCL2L15<br />

may contribute to the protective effect <strong>of</strong><br />

proapoptotic BCL2 family proteins against<br />

malignant transformation <strong>of</strong> the gastrointestinal<br />

tract (Dempsey et al., 2005).<br />

References<br />

Li H, Zhu H, Xu CJ, Yuan J. Cleavage <strong>of</strong> BID by caspase 8<br />

mediates the mitochondrial damage in the Fas pathway <strong>of</strong><br />

apoptosis. Cell. 1998 Aug 21;94(4):491-501<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 115


BCL2L15 (BCL2-like 15) Pavlou MAS, Kontos CK<br />

Luo X, Budihardjo I, Zou H, Slaughter C, Wang X. Bid, a<br />

Bcl2 interacting protein, mediates cytochrome c release<br />

from mitochondria in response to activation <strong>of</strong> cell surface<br />

death receptors. Cell. 1998 Aug 21;94(4):481-90<br />

Coultas L, Pellegrini M, Visvader JE, Lindeman GJ, Chen<br />

L, Adams JM, Huang DC, Strasser A. Bfk: a novel weakly<br />

proapoptotic member <strong>of</strong> the Bcl-2 protein family with a BH3<br />

<strong>and</strong> a BH2 region. Cell Death Differ. 2003 Feb;10(2):185-<br />

92<br />

Petros AM, Olejniczak ET, Fesik SW. Structural biology <strong>of</strong><br />

the Bcl-2 family <strong>of</strong> proteins. Biochim Biophys Acta. 2004<br />

Mar 1;1644(2-3):83-94<br />

Dempsey CE, Dive C, Fletcher DJ, Barnes FA, Lobo A,<br />

Bingle CD, Whyte MK, Renshaw SA. Expression <strong>of</strong> proapoptotic<br />

Bfk is<strong>of</strong>orms reduces during malignant<br />

transformation in the human gastrointestinal tract. FEBS<br />

Lett. 2005 Jul 4;579(17):3646-50<br />

Thomadaki H, Scorilas A. BCL2 family <strong>of</strong> apoptosis-related<br />

genes: functions <strong>and</strong> clinical implications in cancer. Crit<br />

Rev Clin Lab Sci. 2006 Jan;43(1):1-67<br />

Pujianto DA, Damdimopoulos AE, Sipilä P, Jalkanen J,<br />

Huhtaniemi I, Poutanen M. Bfk, a novel member <strong>of</strong> the<br />

bcl2 gene family, is highly expressed in principal cells <strong>of</strong><br />

the mouse epididymis <strong>and</strong> demonstrates a predominant<br />

nuclear localization. Endocrinology. 2007 Jul;148(7):3196-<br />

204<br />

Lomonosova E, Chinnadurai G. BH3-only proteins in<br />

apoptosis <strong>and</strong> beyond: an overview. Oncogene. 2008<br />

Dec;27 Suppl 1:S2-19<br />

Youle RJ, Strasser A. The BCL-2 protein family: opposing<br />

activities that mediate cell death. Nat Rev Mol Cell Biol.<br />

2008 Jan;9(1):47-59<br />

Ozören N, Inohara N, Núñez G. A putative role for human<br />

BFK in DNA damage-induced apoptosis. Biotechnol J.<br />

2009 Jul;4(7):1046-54<br />

This article should be referenced as such:<br />

Pavlou MAS, Kontos CK. BCL2L15 (BCL2-like 15). <strong>Atlas</strong><br />

<strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2):113-116.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 116


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

<strong>Gene</strong> <strong>Section</strong><br />

Review<br />

DUSP6 (dual specificity phosphatase 6)<br />

Zhenfeng Zhang, Balazs Halmos<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

Division <strong>of</strong> Hematology/Oncology, Herbert Irving Comprehensive Cancer Center, New York<br />

Presbyterian Hospital-Columbia University Medical Center, New York, NY, USA (ZZ, BH)<br />

Published in <strong>Atlas</strong> Database: September 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/<strong>Gene</strong>s/DUSP6ID46105ch12q21.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI DUSP6ID46105ch12q21.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Identity<br />

Other names: MKP-3, MKP3, PYST1, rVH6<br />

HGNC (Hugo): DUSP6<br />

Location: 12q21.33<br />

DNA/RNA<br />

Description<br />

The human DUSP6 gene is located on chromosome<br />

12q21.33 <strong>and</strong> consists <strong>of</strong> 3 exons. The full-length<br />

coding sequence <strong>of</strong> DUSP6 contains 1146<br />

nucleotides. The functional phosphatase domain <strong>of</strong><br />

DUSP6 is encoded by half <strong>of</strong> exon 2 <strong>and</strong> almost the<br />

entire sequence <strong>of</strong> exon 3.<br />

Transcription<br />

DUSP6 gene transcription can start from either the<br />

first ATG or alternatively the second ATG (Met14),<br />

<strong>and</strong> therefore two protein products are generated<br />

which usually demonstrate a double-b<strong>and</strong><br />

appearance in regular immunoblotting assays<br />

(Dowd et al., 1998; Zhang et al., 2010).<br />

Protein<br />

Description<br />

The full-length DUSP6 protein contains 381 amino<br />

acids <strong>and</strong> has a molecular weight <strong>of</strong> 44 kDa.<br />

DUSPs are characterized by a common structure<br />

comprising a C-terminal phosphatase domain that<br />

are defined by the active-site signature motif<br />

HCXXXXXR. The structure <strong>of</strong> DUSP proteins<br />

confers phosphatase activity for both<br />

phosphoserine/threonine <strong>and</strong> phosphotyrosine<br />

residues. An enzyme-dead DUSP6 expression<br />

construct can be generated via a 293 Cysteine to<br />

Serine/Glycine (C293S/G) point mutation (Wishart<br />

et al., 1995; Zhang et al., 2010; Zhou et al., 2006).<br />

Expression<br />

DUSP6 is expressed usually at low level in resting,<br />

nonstimulated cells in a variety <strong>of</strong> tissues <strong>and</strong> is<br />

induced as an early response gene after activation<br />

<strong>of</strong> the ERK-MAPK signaling pathway.<br />

The diagram depicts the structure <strong>of</strong> the DUSP6 gene (bottom) roughly aligned with its corresponding functional protein domains<br />

(middle <strong>and</strong> top). DUSP6 comprises a C-terminal catalytic domain <strong>and</strong> an N-terminal non-catalytic domain (middle). The 3 exons<br />

<strong>of</strong> DUSP6 (rectangles) are connected with lines representing introns.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 117


DUSP6 (dual specificity phosphatase 6) Zhang Z, Halmos B<br />

The diagram depicts the structural features <strong>of</strong> DUSP6. The highly conserved C-terminal domain <strong>of</strong> DUSP6 contains the<br />

canonical tyrosine/threonine-specific phosphatase signature sequence HCXXXXXR at the active site, where the cysteine acts as<br />

the essential enzymatic nucleophile <strong>and</strong> arginine interacts directly with the phosphate group on phosphotyrosine or<br />

phosphothreonine (Farooq et al., 2001). The amino-terminal domain <strong>of</strong> DUSP6 contains a specific arginine-rich kinase<br />

interaction motif (KIM) (Tárrega et al., 2005) <strong>and</strong> a leucine-rich nuclear export signal (NES) necessary <strong>and</strong> sufficient for nuclear<br />

export <strong>of</strong> the phosphatase (Karlsson et al., 2004).<br />

Localisation<br />

DUSP6 is a cytoplasmic dual specificity protein<br />

phosphatase.<br />

Function<br />

Mitogen-activated protein kinases (MAPK)<br />

constitute a highly conserved family <strong>of</strong> kinases that<br />

relay information from extracellular signals to<br />

downstream effectors that control diverse cellular<br />

processes such as proliferation, differentiation,<br />

migration, survival <strong>and</strong> apoptosis (Wada <strong>and</strong><br />

Penninger, 2004). A balance between the activities<br />

<strong>of</strong> upstream activators <strong>and</strong> various negative<br />

regulatory mechanisms <strong>of</strong> MAPK signaling, which<br />

terminate its activation, determines its biological<br />

outcomes. DUSP6 is a prototypical member <strong>of</strong> a<br />

subfamily <strong>of</strong> cytoplasmic MKPs, which includes<br />

DUSP7 <strong>and</strong> DUSP9 as well. These enzymes all<br />

display a high degree <strong>of</strong> substrate selectivity for<br />

ERK1 <strong>and</strong> ERK2 (Keyse, 2008). DUSP6 has been<br />

shown to act as a central feedback regulator<br />

attenuating ERK levels in developmental programs<br />

(Echevarria et al., 2005; Li et al., 2007). The<br />

cytoplasmic localization <strong>of</strong> DUSP6 is mediated by<br />

a chromosome region maintenance-1-dependent<br />

nuclear export pathway. DUSP6 appears to play a<br />

role in determining the subcellular localization <strong>of</strong><br />

ERK by serving as a cytoplasmic anchor for ERK,<br />

thereby mediating a spatio-temporal mechanism <strong>of</strong><br />

ERK signaling regulation. Cytoplasmic retention <strong>of</strong><br />

ERK requires both a functional kinase interaction<br />

motif <strong>and</strong> nuclear export site. Defects <strong>of</strong> these<br />

feedback regulation steps are thought to contribute<br />

to ERK-MAPK related oncogenesis. An in vivo<br />

study has identified DUSP6 as a negative feedback<br />

regulator <strong>of</strong> fibroblast growth factor-stimulated<br />

ERK signaling during murine development (Li et<br />

al., 2007). Several in vitro studies have<br />

demonstrated that DUSP6 acts as a negative<br />

regulator <strong>of</strong> fibroblast growth factor receptor<br />

signaling <strong>and</strong> endothelial cell platelet-derived<br />

growth factor receptor signaling via termination <strong>of</strong><br />

ERK activation (Ekerot et al., 2008; Jurek et al.,<br />

2009).<br />

Homology<br />

DUSP6 belongs to a subfamily <strong>of</strong> ten more closely<br />

related dual-specificity MAPK phosphatases<br />

(MKPs) within the larger cysteine-dependent dual<br />

specificity phosphatase (DUSP) family (Keyse,<br />

2008). While DUSP1 (MKP-1), DUSP4 (MKP-2),<br />

<strong>and</strong> DUSP9 (MKP4) dephosphorylate both ERKs,<br />

p38 <strong>and</strong> JNK, the phosphatases DUSP5 (Hvh-3),<br />

DUSP6 (MKP-3), <strong>and</strong> DUSP7 (MKP-X)<br />

exclusively target ERK1/2 MAPKs (Keyse, 2008).<br />

The N-terminal domain <strong>of</strong> all DUSPs has two<br />

regions <strong>of</strong> homology with the Cdc25 cell cycle<br />

regulatory phosphatase. The more conserved<br />

catalytic domain within DUSPs contains an active<br />

site sequence related to the prototypic VH-1<br />

phosphatase encoded by the vaccinia virus.<br />

Specificity <strong>of</strong> MKPs toward MAPKs relies on the<br />

KIM domain. Although each MKP targets different<br />

subsets <strong>of</strong> MAPKs, there is an overlap between<br />

their specificities (Bermudez et al., 2010).<br />

Mutations<br />

Note<br />

Although DUSP6 has been implicated as a<br />

c<strong>and</strong>idate tumor suppressor in several cancer<br />

setting, no mutations in the gene have been<br />

identified so far.<br />

Implicated in<br />

Various cancers<br />

Note<br />

DUSP6 null mice demonstrate enhanced ERK1/2<br />

phosphorylation leading to increased myocyte<br />

proliferation <strong>and</strong> cardiac hypercellularity (Maillet et<br />

al., 2008). DUSP6 has been identified as a potential<br />

novel tumor suppressor gene in pancreatic cancer<br />

since loss <strong>of</strong> DUSP6 expression might synergize<br />

with activating-mutated k-Ras resulting in<br />

increased activation <strong>of</strong> ERK1/2 MAP kinase <strong>and</strong><br />

thus contribute to the development <strong>of</strong> the malignant<br />

<strong>and</strong> invasive phenotype in pancreatic cancer<br />

(Furukawa et al., 2003). Loss <strong>of</strong> DUSP6 expression<br />

caused by oxidative stress-mediated degradation<br />

was also noted in ovarian cancer <strong>and</strong> correlated<br />

with high ERK1/2 activity (Chan et al., 2008).<br />

DUSP6 has also been identified as one <strong>of</strong> only three<br />

genes which are uniquely expressed in myeloma<br />

cells harboring a constitutively active mutant N-ras<br />

gene <strong>and</strong> is also overexpressed in human melanoma<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 118


DUSP6 (dual specificity phosphatase 6) Zhang Z, Halmos B<br />

cell lines with potent activating mutations in B-raf<br />

<strong>and</strong> in breast epithelial cells stably expressing H-<br />

Ras (Bloethner et al., 2005; Croonquist et al., 2003;<br />

Warmka et al., 2004), suggesting that the overexpression<br />

<strong>of</strong> DUSP6 seen in response to<br />

activating-mutated Ras or Raf might represent a<br />

compensatory increase in the negative feedback<br />

control <strong>of</strong> the ERK1/2 MAPK pathway, which lies<br />

downstream <strong>of</strong> these activated oncogenes. In<br />

support <strong>of</strong> this, the tetracycline-induced expression<br />

<strong>of</strong> a functional fusion protein between DUSP6 <strong>and</strong><br />

green fluorescent protein in H-ras transformed<br />

fibroblasts following injection into nude mice<br />

resulted in a large delay in tumor emergence <strong>and</strong><br />

growth as compared to the untreated control group<br />

(Marchetti et al., 2004 ). DUSP6 has been reported<br />

to be one <strong>of</strong> the most highly regulated genes in<br />

chronic myeloid leukemia cells upon imatinib<br />

treatment (Hakansson et al., 2008) <strong>and</strong> similarly<br />

DUSP6 is overexpressed upon inducible expression<br />

<strong>of</strong> the EGFRvIII oncogene in glioblastoma cells<br />

(Ramnarain et al., 2006). DUSP6 has also been<br />

demonstrated to be positively correlated with the<br />

activity <strong>of</strong> the oncogenic ERK pathway in nonsmall<br />

cell lung cancer tissue <strong>and</strong> is an ETSregulated<br />

negative feedback mediator <strong>of</strong> ERK<br />

signaling in lung cancer cells (Zhang et al., 2010).<br />

Prognosis<br />

Elevated DUSP6 RNA expression was reported to<br />

be a major negative predictor <strong>of</strong> survival in patients<br />

with resected non-small cell lung cancer as part <strong>of</strong> a<br />

five-gene signature model (Chen et al., 2007).<br />

References<br />

Wishart MJ, Denu JM, Williams JA, Dixon JE. A single<br />

mutation converts a novel phosphotyrosine binding domain<br />

into a dual-specificity phosphatase. J Biol Chem. 1995 Nov<br />

10;270(45):26782-5<br />

Dowd S, Sneddon AA, Keyse SM. Isolation <strong>of</strong> the human<br />

genes encoding the pyst1 <strong>and</strong> Pyst2 phosphatases:<br />

characterisation <strong>of</strong> Pyst2 as a cytosolic dual-specificity<br />

MAP kinase phosphatase <strong>and</strong> its catalytic activation by<br />

both MAP <strong>and</strong> SAP kinases. J Cell Sci. 1998 Nov;111 ( Pt<br />

22):3389-99<br />

Farooq A, Chaturvedi G, Mujtaba S, Plotnikova O, Zeng L,<br />

Dhalluin C, Ashton R, Zhou MM. Solution structure <strong>of</strong><br />

ERK2 binding domain <strong>of</strong> MAPK phosphatase MKP-3:<br />

structural insights into MKP-3 activation by ERK2. Mol<br />

Cell. 2001 Feb;7(2):387-99<br />

Croonquist PA, Linden MA, Zhao F, Van Ness BG. <strong>Gene</strong><br />

pr<strong>of</strong>iling <strong>of</strong> a myeloma cell line reveals similarities <strong>and</strong><br />

unique signatures among IL-6 response, N-ras-activating<br />

mutations, <strong>and</strong> coculture with bone marrow stromal cells.<br />

Blood. 2003 Oct 1;102(7):2581-92<br />

Furukawa T, Sunamura M, Motoi F, Matsuno S, Horii A.<br />

Potential tumor suppressive pathway involving<br />

DUSP6/MKP-3 in pancreatic cancer. Am J Pathol. 2003<br />

Jun;162(6):1807-15<br />

Alonso A, Sasin J, Bottini N, Friedberg I, Friedberg I,<br />

Osterman A, Godzik A, Hunter T, Dixon J, Mustelin T.<br />

Protein tyrosine phosphatases in the human genome. Cell.<br />

2004 Jun 11;117(6):699-711<br />

Karlsson M, Mathers J, Dickinson RJ, M<strong>and</strong>l M, Keyse SM.<br />

Both nuclear-cytoplasmic shuttling <strong>of</strong> the dual specificity<br />

phosphatase MKP-3 <strong>and</strong> its ability to anchor MAP kinase<br />

in the cytoplasm are mediated by a conserved nuclear<br />

export signal. J Biol Chem. 2004 Oct 1;279(40):41882-91<br />

Marchetti S, Gimond C, Roux D, Gothié E, Pouysségur J,<br />

Pagès G. Inducible expression <strong>of</strong> a MAP kinase<br />

phosphatase-3-GFP chimera specifically blunts fibroblast<br />

growth <strong>and</strong> ras-dependent tumor formation in nude mice. J<br />

Cell Physiol. 2004 Jun;199(3):441-50<br />

Wada T, Penninger JM. Mitogen-activated protein kinases<br />

in apoptosis regulation. Oncogene. 2004 Apr<br />

12;23(16):2838-49<br />

Warmka JK, Mauro LJ, Wattenberg EV. Mitogen-activated<br />

protein kinase phosphatase-3 is a tumor promoter target in<br />

initiated cells that express oncogenic Ras. J Biol Chem.<br />

2004 Aug 6;279(32):33085-92<br />

Bloethner S, Chen B, Hemminki K, Müller-Berghaus J,<br />

Ugurel S, Schadendorf D, Kumar R. Effect <strong>of</strong> common B-<br />

RAF <strong>and</strong> N-RAS mutations on global gene expression in<br />

melanoma cell lines. Carcinogenesis. 2005 Jul;26(7):1224-<br />

32<br />

Echevarria D, Martinez S, Marques S, Lucas-Teixeira V,<br />

Belo JA. Mkp3 is a negative feedback modulator <strong>of</strong> Fgf8<br />

signaling in the mammalian isthmic organizer. Dev Biol.<br />

2005 Jan 1;277(1):114-28<br />

Tárrega C, Ríos P, Cejudo-Marín R, Blanco-Aparicio C,<br />

van den Berk L, Schepens J, Hendriks W, Tabernero L,<br />

Pulido R. ERK2 shows a restrictive <strong>and</strong> locally selective<br />

mechanism <strong>of</strong> recognition by its tyrosine phosphatase<br />

inactivators not shared by its activator MEK1. J Biol Chem.<br />

2005 Nov 11;280(45):37885-94<br />

Ramnarain DB, Park S, Lee DY, Hatanpaa KJ, Scoggin<br />

SO, Otu H, Libermann TA, Raisanen JM, Ashfaq R, Wong<br />

ET, Wu J, Elliott R, Habib AA. Differential gene expression<br />

analysis reveals generation <strong>of</strong> an autocrine loop by a<br />

mutant epidermal growth factor receptor in glioma cells.<br />

Cancer Res. 2006 Jan 15;66(2):867-74<br />

Zhou B, Zhang J, Liu S, Reddy S, Wang F, Zhang ZY.<br />

Mapping ERK2-MKP3 binding interfaces by<br />

hydrogen/deuterium exchange mass spectrometry. J Biol<br />

Chem. 2006 Dec 15;281(50):38834-44<br />

Chen HY, Yu SL, Chen CH, Chang GC, Chen CY, Yuan A,<br />

Cheng CL, Wang CH, Terng HJ, Kao SF, Chan WK, Li HN,<br />

Liu CC, Singh S, Chen WJ, Chen JJ, Yang PC. A fivegene<br />

signature <strong>and</strong> clinical outcome in non-small-cell lung<br />

cancer. N Engl J Med. 2007 Jan 4;356(1):11-20<br />

Li C, Scott DA, Hatch E, Tian X, Mansour SL. Dusp6<br />

(Mkp3) is a negative feedback regulator <strong>of</strong> FGF-stimulated<br />

ERK signaling during mouse development. Development.<br />

2007 Jan;134(1):167-76<br />

Chan DW, Liu VW, Tsao GS, Yao KM, Furukawa T, Chan<br />

KK, Ngan HY. Loss <strong>of</strong> MKP3 mediated by oxidative stress<br />

enhances tumorigenicity <strong>and</strong> chemoresistance <strong>of</strong> ovarian<br />

cancer cells. Carcinogenesis. 2008 Sep;29(9):1742-50<br />

Ekerot M, Stavridis MP, Delavaine L, Mitchell MP, Staples<br />

C, Owens DM, Keenan ID, Dickinson RJ, Storey KG,<br />

Keyse SM. Negative-feedback regulation <strong>of</strong> FGF signalling<br />

by DUSP6/MKP-3 is driven by ERK1/2 <strong>and</strong> mediated by<br />

Ets factor binding to a conserved site within the<br />

DUSP6/MKP-3 gene promoter. Biochem J. 2008 Jun<br />

1;412(2):287-98<br />

Håkansson P, Nilsson B, Andersson A, Lassen C, Gullberg<br />

U, Fioretos T. <strong>Gene</strong> expression analysis <strong>of</strong> BCR/ABL1dependent<br />

transcriptional response reveals enrichment for<br />

genes involved in negative feedback regulation. <strong>Gene</strong>s<br />

Chromosomes Cancer. 2008 Apr;47(4):267-75<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 119


DUSP6 (dual specificity phosphatase 6) Zhang Z, Halmos B<br />

Keyse SM. Dual-specificity MAP kinase phosphatases<br />

(MKPs) <strong>and</strong> cancer. Cancer Metastasis Rev. 2008<br />

Jun;27(2):253-61<br />

Maillet M, Purcell NH, Sargent MA, York AJ, Bueno OF,<br />

Molkentin JD. DUSP6 (MKP3) null mice show enhanced<br />

ERK1/2 phosphorylation at baseline <strong>and</strong> increased<br />

myocyte proliferation in the heart affecting disease<br />

susceptibility. J Biol Chem. 2008 Nov 7;283(45):31246-55<br />

Jurek A, Amagasaki K, Gembarska A, Heldin CH,<br />

Lennartsson J. Negative <strong>and</strong> positive regulation <strong>of</strong> MAPK<br />

phosphatase 3 controls platelet-derived growth factorinduced<br />

Erk activation. J Biol Chem. 2009 Feb<br />

13;284(7):4626-34<br />

Bermudez O, Pagès G, Gimond C. The dual-specificity<br />

MAP kinase phosphatases: critical roles in development<br />

<strong>and</strong> cancer. Am J Physiol Cell Physiol. 2010<br />

Aug;299(2):C189-202<br />

Zhang Z, Kobayashi S, Borczuk AC, Leidner RS,<br />

Laframboise T, Levine AD, Halmos B. Dual specificity<br />

phosphatase 6 (DUSP6) is an ETS-regulated negative<br />

feedback mediator <strong>of</strong> oncogenic ERK signaling in lung<br />

cancer cells. Carcinogenesis. 2010 Apr;31(4):577-86<br />

This article should be referenced as such:<br />

Zhang Z, Halmos B. DUSP6 (dual specificity phosphatase<br />

6). <strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012;<br />

16(2):117-120.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 120


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

<strong>Gene</strong> <strong>Section</strong><br />

Review<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

GZMA (granzyme A (granzyme 1, cytotoxic Tlymphocyte-associated<br />

serine esterase 3))<br />

Elena Catalan, Diego Sanchez-Martinez, Julián Pardo<br />

Dpto Bioquimica y Biologia Molecular y Celular, Fac Ciencias, Univ Zaragoza, Spain (EC, DSM),<br />

Dpto Bioquimica y Biologia Molecular y Celular, Fac Ciencias, Univ Zaragoza, Spain; Fundacion<br />

Aragon I+D (ARAID), Zaragoza, Spain (JP)<br />

Published in <strong>Atlas</strong> Database: September 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/<strong>Gene</strong>s/GZMAID51130ch5q11.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI GZMAID51130ch5q11.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Identity<br />

Other names: CTLA3, HFSP<br />

HGNC (Hugo): GZMA<br />

Location: 5q11.2<br />

Local order: Size: 7607 bases. Coordinates:<br />

54398473.<br />

DNA/RNA<br />

Description<br />

The GZMA gene, with 7607 bases in length,<br />

consists <strong>of</strong> 5 exons <strong>and</strong> 4 introns. GZMA gene is<br />

located in a gene cluster together with granzyme K<br />

(figure 1) (Grossman et al., 2003).<br />

Transcription<br />

There are at least two transcripts <strong>of</strong> human GZMA<br />

whose expression is differentially regulated by<br />

glucocorticoid (Ruike et al., 2007). These<br />

transcripts generate two is<strong>of</strong>orms, GZMAα <strong>and</strong><br />

GZMAβ, which have respective first exons: exon<br />

1a <strong>and</strong> exon 1b (figure 1):<br />

GZMAα (exon 1a): canonical sequence,<br />

GZMAβ (exon 1b): lack aa 1-17; aa 18-23 LLLIPE<br />

--> MTKGLR.<br />

Figure 1. Genomic organization <strong>of</strong> human GZMA. A, human GZMA cluster. Arrow indicate the direction <strong>of</strong> transcription. B,<br />

representation <strong>of</strong> the GZMA genetic locus. White: untranslated regions; Blue: leader sequence; Green: mature enzyme. Solid<br />

lanes: splicing between the first <strong>and</strong> second exons. gre: glucocorticoid response element (adapted from Ruike et al., 2007).<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 121


GZMA (granzyme A (granzyme 1, cytotoxic T-lymphocyteassociated<br />

serine esterase 3))<br />

Catalan E, et al.<br />

Figure 2. Diagram <strong>of</strong> the crystal structure <strong>of</strong> human granzyme A dimer (Bell et al., 2003; Hink-Schauer et al., 2003). The<br />

cystein groups involved in disulphide bond-mediated dimer (green) <strong>and</strong> the three aminoacids forming the catalytic triad (red, blue<br />

<strong>and</strong> yellow) are shown. Representation from PDB (accession code 1OP8) deposited by Hink-Schauer C, Estébanez-Perpiñá E,<br />

Kurschus FC, Bode W, Jenne DE. Nat Struct Biol. 2003 Jul;10(7):535-40.<br />

Protein<br />

Description<br />

Granzyme A is a tryptase (cleave proteins after Lys<br />

or Arg residues) expressed mainly in cytotoxic cells<br />

(cytotoxic T <strong>and</strong> Natural Killer cells) (Masson et<br />

al., 1986; Simon et al., 1986; Young et al., 1986).<br />

Protein is expressed as a preproenzyme (Jenne et<br />

al., 1988) containing a signal sequence that<br />

mediates targeting <strong>of</strong> the nascent enzyme to the ER.<br />

Cleavage <strong>of</strong> the signal peptide produces an inactive<br />

proenzyme that contains an N-terminal dipeptide<br />

that needs to be cleaved to produce an active<br />

protease. In the Golgi, a mannose-6-phosphate tag<br />

is added for transporting the proenzyme to<br />

cytotoxic granules. Within the cytotoxic granule,<br />

the N-terminal dipeptide is removed by cathepsin C<br />

(dipeptidyl peptidase I) (Pham et al., 1999),<br />

producing the active enzyme that is kept inactive at<br />

low pH. Native granzyme A is expressed as a dimer<br />

(Bell et al., 2003; Hink-Schauer et al., 2003).<br />

Expression<br />

Cytotoxic CD8+ T cells, Natural Killer cells, CD4+<br />

T cells, gamma-delta T cells, type II pneumocytes,<br />

alveolar macrophages, bronchiolar epithelial cells.<br />

Localisation<br />

Cytotoxic granules.<br />

Function<br />

Granzyme A is delivered from CTL or NK<br />

cytotoxic granules to the cytoplasm <strong>of</strong> target cell by<br />

a mechanism dependent on perforin (Baran et al.,<br />

2009; Praper et al., 2011; Thiery et al., 2011).<br />

There are some controversial findings about the<br />

physiological function <strong>of</strong> gzmA.<br />

It has been reported that human GzmA induces<br />

perforin-mediated caspase-independent cell death in<br />

some tumors cell lines (Hayes et al., 1989; Shi et<br />

al., 1992; Beresford et al., 1999; Shresta et al.,<br />

1999; Pardo et al., 2004). GzmA translocates to the<br />

nucleus <strong>and</strong> mitochondria where key substrates<br />

such as mitochondrial complex I protein, NADH<br />

dehydrogenase Fe-S protein 3 (NDUFS3) is<br />

cleaved, inducing the production <strong>of</strong> Radical<br />

Oxygen Species (ROS). ROS production induces<br />

the activation <strong>of</strong> the SET complex that translocates<br />

into the nucleus in order to repair DNA damage<br />

induced by ROS. Once there, granzyme A cleaves<br />

components <strong>of</strong> the endoplasmic reticulumassociated<br />

SET complex, releasing the<br />

endonuclease NM23H1 that induces single str<strong>and</strong><br />

nicks in the DNA <strong>and</strong> ultimately cell death<br />

(Lieberman, 2011).<br />

Other authors have reported that the cytotoxic<br />

potential <strong>of</strong> granzyme A is low, but induce<br />

expression <strong>of</strong> pro-inflammatory cytokines in<br />

monocytes-like cells by a caspase-1 dependent<br />

mechanism (Metkar et al., 2008).<br />

Granzyme A is able to cleave several extracellular<br />

substrates like thrombin receptor, fibronectin,<br />

collagen IV, proteinase-activated receptor-2, Prourokinase<br />

plasminogen activator <strong>and</strong> myelin basic<br />

protein (Kramer et al., 1987; Buzza et al., 2006;<br />

Hendel et al., 2011).<br />

Granzyme <strong>and</strong> granzyme B double deficient mice<br />

are more susceptible than granzyme B deficient<br />

mice to transplanted tumors suggesting a<br />

contribution <strong>of</strong> granzyme A to tumor control in<br />

vivo (Pardo et al., 2002; Cao et al., 2007).<br />

Homology<br />

Mouse granzyme A; Rat granzyme A; Chicken<br />

granzyme A; Fish granzyme A (Common Carp,<br />

Atlantic cod, Channel catfish) (Praveen et al., 2006;<br />

Praveen et al., 2006; Wernersson et al., 2006).<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 122


GZMA (granzyme A (granzyme 1, cytotoxic T-lymphocyteassociated<br />

serine esterase 3))<br />

Mutations<br />

Note: Not known.<br />

Implicated in<br />

Sepsis (Froelich et al., 2009; Hendel et al., 2011)<br />

Disease<br />

Several findings suggest that gzmA contributes to<br />

septic shock. Native <strong>and</strong> recombinant human<br />

granzyme A as well as a human NK cell line<br />

expressing gzmA induces human adherent<br />

peripheral blood mononuclear cells to express<br />

proinflammatory cytokines including interleukin-<br />

1beta interleukin-6, inteleukin-8 <strong>and</strong> TNF-alpha<br />

(Sower et al., 1996; Metkar et al., 2008). Granzyme<br />

A deficient mice are more resistant than wild type<br />

mice to septic shock induced by LPS (Metkar et al.,<br />

2008).<br />

Rheumatoid arthritis<br />

Prognosis<br />

Granzyme A levels are higher in serum <strong>and</strong><br />

synovial fluid <strong>of</strong> patients with rheumatoid arthritis<br />

(Griffiths et al., 1992; Nordstrom et al., 1992;<br />

Kummer et al., 1994; Tak et al., 1994; Muller-<br />

Ladner et al., 1995; Spaeny-Dekking et al., 1998;<br />

Tak et al., 1999).<br />

Chronic obstructive pulmonary<br />

disease<br />

Prognosis<br />

Granzyme A is expressed in type II pneumocytes <strong>of</strong><br />

patients with severe chronic obstructive pulmonary<br />

disease (Vernooy et al., 2007).<br />

Hypersensitivity pneumonitis<br />

Prognosis<br />

Granzyme A is elevated in bronchoalveolar lavage<br />

fluid from patients with hypersensitivity<br />

pneumonitis (Tremblay et al., 2000).<br />

Sjögren's syndrome<br />

Prognosis<br />

Granzyme A is expressed in salivary gl<strong>and</strong>s from<br />

patients with Sjögren's syndrome (Alpert et al.,<br />

1994).<br />

Poxvirus infection<br />

Disease<br />

Granzyme A deficient mice are more susceptible<br />

than wild type mice to mousepox virus (ectromelia)<br />

(Mullbacher et al., 1996).<br />

Herpes virus infection<br />

Disease<br />

Granzyme A deficient mice are more susceptible<br />

than wild type mice to herpes simplex virus type 1<br />

(HSV-1) (Pereira et al., 2000) <strong>and</strong> mouse<br />

cytomegalovirus (CMV) infection (Riera et al.,<br />

2000).<br />

References<br />

Catalan E, et al.<br />

Masson D, Zamai M, Tschopp J. Identification <strong>of</strong> granzyme<br />

A isolated from cytotoxic T-lymphocyte-granules as one <strong>of</strong><br />

the proteases encoded by CTL-specific genes. FEBS Lett.<br />

1986 Nov 10;208(1):84-8<br />

Simon MM, Hoschützky H, Fruth U, Simon HG, Kramer<br />

MD. Purification <strong>and</strong> characterization <strong>of</strong> a T cell specific<br />

serine proteinase (TSP-1) from cloned cytolytic T<br />

lymphocytes. EMBO J. 1986 Dec 1;5(12):3267-74<br />

Young JD, Hengartner H, Podack ER, Cohn ZA.<br />

Purification <strong>and</strong> characterization <strong>of</strong> a cytolytic pore-forming<br />

protein from granules <strong>of</strong> cloned lymphocytes with natural<br />

killer activity. Cell. 1986 Mar 28;44(6):849-59<br />

Kramer MD, Simon MM. Are Proteinases Functional<br />

Molecules <strong>of</strong> T Lymphocytes? Immunol Today.<br />

1987;8:140-2. (REVIEW)<br />

Jenne DE, Tschopp J.. Granzymes, a family <strong>of</strong> serine<br />

proteases released from granules <strong>of</strong> cytolytic T<br />

lymphocytes upon T cell receptor stimulation. Immunol<br />

Rev. 1988 Mar;103:53-71. (REVIEW)<br />

Hayes MP, Berrebi GA, Henkart PA.. Induction <strong>of</strong> target<br />

cell DNA release by the cytotoxic T lymphocyte granule<br />

protease granzyme A. J Exp Med. 1989 Sep 1;170(3):933-<br />

46.<br />

Griffiths GM, Alpert S, Lambert E, McGuire J, Weissman<br />

IL.. Perforin <strong>and</strong> granzyme A expression identifying<br />

cytolytic lymphocytes in rheumatoid arthritis. Proc Natl<br />

Acad Sci U S A. 1992 Jan 15;89(2):549-53.<br />

Nordstrom DC, Konttinen YT, Sorsa T, Nykanen P,<br />

Pettersson T, Santavirta S, Tschopp J.. Granzyme Aimmunoreactive<br />

cells in synovial fluid in reactive <strong>and</strong><br />

rheumatoid arthritis. Clin Rheumatol. 1992 Dec;11(4):529-<br />

32.<br />

Shi L, Kraut RP, Aebersold R, Greenberg AH.. A natural<br />

killer cell granule protein that induces DNA fragmentation<br />

<strong>and</strong> apoptosis. J Exp Med. 1992 Feb 1;175(2):553-66.<br />

Alpert S, Kang HI, Weissman I, Fox RI.. Expression <strong>of</strong><br />

granzyme A in salivary gl<strong>and</strong> biopsies from patients with<br />

primary Sjogren's syndrome. Arthritis Rheum. 1994<br />

Jul;37(7):1046-54.<br />

Kummer JA, Tak PP, Brinkman BM, van Tilborg AA, Kamp<br />

AM, Verweij CL, Daha MR, Meinders AE, Hack CE,<br />

Breedveld FC.. Expression <strong>of</strong> granzymes A <strong>and</strong> B in<br />

synovial tissue from patients with rheumatoid arthritis <strong>and</strong><br />

osteoarthritis. Clin Immunol Immunopathol. 1994<br />

Oct;73(1):88-95.<br />

Tak PP, Kummer JA, Hack CE, Daha MR, Smeets TJ,<br />

Erkelens GW, Meinders AE, Kluin PM, Breedveld FC..<br />

Granzyme-positive cytotoxic cells are specifically<br />

increased in early rheumatoid synovial tissue. Arthritis<br />

Rheum. 1994 Dec;37(12):1735-43.<br />

Muller-Ladner U, Kriegsmann J, Tschopp J, Gay RE, Gay<br />

S.. Demonstration <strong>of</strong> granzyme A <strong>and</strong> perforin messenger<br />

RNA in the synovium <strong>of</strong> patients with rheumatoid arthritis.<br />

Arthritis Rheum. 1995 Apr;38(4):477-84.<br />

Mullbacher A, Ebnet K, Bl<strong>and</strong>en RV, Hla RT, Stehle T,<br />

Museteanu C, Simon MM.. Granzyme A is critical for<br />

recovery <strong>of</strong> mice from infection with the natural cytopathic<br />

viral pathogen, ectromelia. Proc Natl Acad Sci U S A. 1996<br />

Jun 11;93(12):5783-7.<br />

Sower LE, Klimpel GR, Hanna W, Froelich CJ..<br />

Extracellular activities <strong>of</strong> human granzymes. I. Granzyme<br />

A induces IL6 <strong>and</strong> IL8 production in fibroblast <strong>and</strong><br />

epithelial cell lines. Cell Immunol. 1996 Jul 10;171(1):159-<br />

63.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 123


GZMA (granzyme A (granzyme 1, cytotoxic T-lymphocyteassociated<br />

serine esterase 3))<br />

Spaeny-Dekking EH, Hanna WL, Wolbink AM, Wever PC,<br />

Kummer JA, Swaak AJ, Middeldorp JM, Huisman HG,<br />

Froelich CJ, Hack CE.. Extracellular granzymes A <strong>and</strong> B in<br />

humans: detection <strong>of</strong> native species during CTL responses<br />

in vitro <strong>and</strong> in vivo. J Immunol. 1998 Apr 1;160(7):3610-6.<br />

Beresford PJ, Xia Z, Greenberg AH, Lieberman J..<br />

Granzyme A loading induces rapid cytolysis <strong>and</strong> a novel<br />

form <strong>of</strong> DNA damage independently <strong>of</strong> caspase activation.<br />

Immunity. 1999 May;10(5):585-94.<br />

Pham CT, Ley TJ.. Dipeptidyl peptidase I is required for<br />

the processing <strong>and</strong> activation <strong>of</strong> granzymes A <strong>and</strong> B in<br />

vivo. Proc Natl Acad Sci U S A. 1999 Jul 20;96(15):8627-<br />

32.<br />

Shresta S, Graubert TA, Thomas DA, Raptis SZ, Ley TJ..<br />

Granzyme A initiates an alternative pathway for granulemediated<br />

apoptosis. Immunity. 1999 May;10(5):595-605.<br />

Tak PP, Spaeny-Dekking L, Kraan MC, Breedveld FC,<br />

Froelich CJ, Hack CE.. The levels <strong>of</strong> soluble granzyme A<br />

<strong>and</strong> B are elevated in plasma <strong>and</strong> synovial fluid <strong>of</strong> patients<br />

with rheumatoid arthritis (RA). Clin Exp Immunol. 1999<br />

May;116(2):366-70.<br />

Pereira RA, Simon MM, Simmons A.. Granzyme A, a<br />

noncytolytic component <strong>of</strong> CD8(+) cell granules, restricts<br />

the spread <strong>of</strong> herpes simplex virus in the peripheral<br />

nervous systems <strong>of</strong> experimentally infected mice. J Virol.<br />

2000 Jan;74(2):1029-32.<br />

Riera L, Gariglio M, Valente G, Mullbacher A, Museteanu<br />

C, L<strong>and</strong>olfo S, Simon MM.. Murine cytomegalovirus<br />

replication in salivary gl<strong>and</strong>s is controlled by both perforin<br />

<strong>and</strong> granzymes during acute infection. Eur J Immunol.<br />

2000 May;30(5):1350-5.<br />

Tremblay GM, Wolbink AM, Cormier Y, Hack CE..<br />

Granzyme activity in the inflamed lung is not controlled by<br />

endogenous serine proteinase inhibitors. J Immunol. 2000<br />

Oct 1;165(7):3966-9.<br />

Pardo J, Balkow S, Anel A, Simon MM.. Granzymes are<br />

essential for natural killer cell-mediated <strong>and</strong> perf-facilitated<br />

tumor control. Eur J Immunol. 2002 Oct;32(10):2881-7.<br />

Bell JK, Goetz DH, Mahrus S, Harris JL, Fletterick RJ,<br />

Craik CS.. The oligomeric structure <strong>of</strong> human granzyme A<br />

is a determinant <strong>of</strong> its extended substrate specificity. Nat<br />

Struct Biol. 2003 Jul;10(7):527-34.<br />

Grossman WJ, Revell PA, Lu ZH, Johnson H, Bredemeyer<br />

AJ, Ley TJ.. The orphan granzymes <strong>of</strong> humans <strong>and</strong> mice.<br />

Curr Opin Immunol. 2003 Oct;15(5):544-52.<br />

Hink-Schauer C, Estebanez-Perpina E, Kurschus FC,<br />

Bode W, Jenne DE.. Crystal structure <strong>of</strong> the apoptosisinducing<br />

human granzyme A dimer. Nat Struct Biol. 2003<br />

Jul;10(7):535-40.<br />

Pardo J, Bosque A, Brehm R, Wallich R, Naval J,<br />

Mullbacher A, Anel A, Simon MM.. Apoptotic pathways are<br />

selectively activated by granzyme A <strong>and</strong>/or granzyme B in<br />

CTL-mediated target cell lysis. J Cell Biol. 2004 Nov<br />

8;167(3):457-68.<br />

Buzza MS, Bird PI.. Extracellular granzymes: current<br />

perspectives. Biol Chem. 2006 Jul;387(7):827-37.<br />

(REVIEW)<br />

Praveen K, Leary JH 3rd, Evans DL, Jaso-Friedmann L..<br />

Molecular characterization <strong>and</strong> expression <strong>of</strong> a granzyme<br />

<strong>of</strong> an ectothermic vertebrate with chymase-like activity<br />

expressed in the cytotoxic cells <strong>of</strong> Nile tilapia (Oreochromis<br />

niloticus). Immunogenetics. 2006 Feb;58(1):41-55. Epub<br />

2006 Feb 9.<br />

Catalan E, et al.<br />

Praveen K, Leary JH 3rd, Evans DL, Jaso-Friedmann L..<br />

Nonspecific cytotoxic cells <strong>of</strong> teleosts are armed with<br />

multiple granzymes <strong>and</strong> other components <strong>of</strong> the granule<br />

exocytosis pathway. Mol Immunol. 2006 Mar;43(8):1152-<br />

62. Epub 2005 Aug 30.<br />

Wernersson S, Reimer JM, Poorafshar M, Karlson U,<br />

Wermenstam N, Bengten E, Wilson M, Pilstrom L, Hellman<br />

L.. Granzyme-like sequences in bony fish shed light on the<br />

emergence <strong>of</strong> hematopoietic serine proteases during<br />

vertebrate evolution. Dev Comp Immunol.<br />

2006;30(10):901-18. Epub 2005 Dec 27.<br />

Cao X, Cai SF, Fehniger TA, Song J, Collins LI, Piwnica-<br />

Worms DR, Ley TJ.. Granzyme B <strong>and</strong> perforin are<br />

important for regulatory T cell-mediated suppression <strong>of</strong><br />

tumor clearance. Immunity. 2007 Oct;27(4):635-46. Epub<br />

2007 Oct 4.<br />

Ruike Y, Katsuma S, Hirasawa A, Tsujimoto G..<br />

Glucocorticoid-induced alternative promoter usage for a<br />

novel 5' variant <strong>of</strong> granzyme A. J Hum <strong>Gene</strong>t.<br />

2007;52(2):172-8. Epub 2006 Dec 19.<br />

Vernooy JH, Moller GM, van Suylen RJ, van Spijk MP,<br />

Cloots RH, Hoet PH, Pennings HJ, Wouters EF..<br />

Increased granzyme A expression in type II pneumocytes<br />

<strong>of</strong> patients with severe chronic obstructive pulmonary<br />

disease. Am J Respir Crit Care Med. 2007 Mar<br />

1;175(5):464-72. Epub 2006 Nov 30.<br />

Metkar SS, Menaa C, Pardo J, Wang B, Wallich R,<br />

Freudenberg M, Kim S, Raja SM, Shi L, Simon MM,<br />

Froelich CJ.. Human <strong>and</strong> mouse granzyme A induce a<br />

proinflammatory cytokine response. Immunity. 2008 Nov<br />

14;29(5):720-33. Epub 2008 Oct 23.<br />

Baran K, Dunstone M, Chia J, Ciccone A, Browne KA,<br />

Clarke CJ, Lukoyanova N, Saibil H, Whisstock JC,<br />

Voskoboinik I, Trapani JA.. The molecular basis for<br />

perforin oligomerization <strong>and</strong> transmembrane pore<br />

assembly. Immunity. 2009 May;30(5):684-95. Epub 2009<br />

May 14.<br />

Froelich CJ, Pardo J, Simon MM.. Granule-associated<br />

serine proteases: granzymes might not just be killer<br />

proteases. Trends Immunol. 2009 Mar;30(3):117-23. Epub<br />

2009 Feb 13. (REVIEW)<br />

Hendel A, Hiebert PR, Boivin WA, Williams SJ, Granville<br />

DJ.. Granzymes in age-related cardiovascular <strong>and</strong><br />

pulmonary diseases. Cell Death Differ. 2010<br />

Apr;17(4):596-606. Epub 2010 Feb 5. (REVIEW)<br />

Lieberman J.. Granzyme A activates another way to die.<br />

Immunol Rev. 2010 May;235(1):93-104. (REVIEW)<br />

Praper T, Sonnen A, Viero G, Kladnik A, Froelich CJ,<br />

Anderluh G, Dalla Serra M, Gilbert RJ.. Human perforin<br />

employs different avenues to damage membranes. J Biol<br />

Chem. 2011 Jan 28;286(4):2946-55. Epub 2010 Oct 2.<br />

Thiery J, Keefe D, Boulant S, Boucrot E, Walch M,<br />

Martinvalet D, Goping IS, Bleackley RC, Kirchhausen T,<br />

Lieberman J.. Perforin pores in the endosomal membrane<br />

trigger the release <strong>of</strong> endocytosed granzyme B into the<br />

cytosol <strong>of</strong> target cells. Nat Immunol. 2011 Jun<br />

19;12(8):770-7. doi: 10.1038/ni.2050.<br />

This article should be referenced as such:<br />

Catalan E, Sanchez-Martinez D, Pardo J. GZMA<br />

(granzyme A (granzyme 1, cytotoxic T-lymphocyteassociated<br />

serine esterase 3)). <strong>Atlas</strong> <strong>Gene</strong>t Cytogenet<br />

Oncol Haematol. 2012; 16(2):121-124.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 124


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

<strong>Gene</strong> <strong>Section</strong><br />

Mini Review<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

MIER1 (mesoderm induction early response 1<br />

homolog (Xenopus laevis))<br />

Laura L Gillespie, Gary D Paterno<br />

Terry Fox Cancer Research Labs, Division <strong>of</strong> Biomedical Sciences, Faculty <strong>of</strong> Medicine, Memorial<br />

University, St John's, NL, Canada (LLG, GDP)<br />

Published in <strong>Atlas</strong> Database: September 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/<strong>Gene</strong>s/MIER1ID50389ch1p31.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI MIER1ID50389ch1p31.txt<br />

This article is an update <strong>of</strong> :<br />

Gillespie LL, Paterno GD. MIER1 (mesoderm induction early response 1 homolog (Xenopus laevis)). <strong>Atlas</strong> <strong>Gene</strong>t Cytogenet<br />

Oncol Haematol 2010;14(10)<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Identity<br />

Other names: ER1, Er1, KIAA1610, MGC150641,<br />

MGC131940, MGC150640, MI-ER1, hMI-ER1,<br />

RP5-944N15.1, DKFZp781G0451<br />

HGNC (Hugo): MIER1<br />

Location: 1p31.3<br />

Note<br />

MIER1 was identified by differential display as an<br />

immediate-early gene activated during fibroblast<br />

growth factor (FGF) induction <strong>of</strong> mesoderm<br />

differentiation in Xenopus laevis.<br />

DNA/RNA<br />

A. Schematic illustrating the exon-intron organization <strong>of</strong> the human MIER1 gene. Exons are shown as red bars/vertical<br />

lines <strong>and</strong> introns as horizontal lines; exon numbers are indicated below each schematic. The light red bar indicates the<br />

facultative intron 16 <strong>and</strong> the position <strong>of</strong> the alpha <strong>and</strong> beta carboxy-terminal coding regions are indicated. Note that the beta<br />

coding region is located within the facultative intron. The three alternate starts <strong>of</strong> translation, ML-, MF- <strong>and</strong> MAE- are indicated as<br />

are the three polyadenylation signals (PAS): i, ii <strong>and</strong> iii.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 125


MIER1 (mesoderm induction early response 1 homolog<br />

(Xenopus laevis))<br />

Gillespie LL, Paterno GD<br />

B. Schematic illustrating the variant 5' <strong>and</strong> 3' ends <strong>of</strong> human MIER1 transcripts. Alternate 5' ends are generated from<br />

differential promoter usage (P1 or P2) or alternate inclusion <strong>of</strong> exon 3A. This leads to three alternate starts <strong>of</strong> translation,<br />

indicated as ML-, MF- <strong>and</strong> MAE-, <strong>and</strong> produces three distinct amino termini. The four variant 3' ends, a, bi, bii <strong>and</strong> biii, produced<br />

by alternative splicing or alternate PAS usage, result in transcripts readily distinguished by size (1.7 kb, 2.5 kb, 3.4 kb <strong>and</strong> 4.8 kb,<br />

respectively) on a Northern blot. It should be noted that three <strong>of</strong> the variant 3' ends, bi, bii <strong>and</strong> biii encode the same protein<br />

sequence <strong>and</strong> differ only in their untranslated region. * indicates beta encoding transcript that contains the alpha exon in its<br />

3'UTR. The locations <strong>of</strong> the alpha <strong>and</strong> beta carboxy-terminal coding regions <strong>and</strong> PAS i, ii <strong>and</strong> iii are indicated. The combination<br />

<strong>of</strong> three possible 5' ends with four possible 3' ends gives rise to 12 distinct transcripts, but only 6 distinct protein is<strong>of</strong>orms. In<br />

most adult tissues, the most abundant transcript is 4.8 kb. Additional transcripts have been reported in Ensembl.<br />

Description<br />

63 kb gene; 2 promoters controlling 2 distinct<br />

transcriptional start sites; 17 exons; intron 16 is<br />

facultative; 3 polyadenylation sites.<br />

Protein<br />

Description<br />

The six human MIER1 is<strong>of</strong>orms: M-3A-alpha (457<br />

aa), M-3A-beta (536 aa), ML-alpha (432 aa), MLbeta<br />

(511 aa), MAE-alpha (433 aa), <strong>and</strong> MAE-beta<br />

(512 aa), range in predicted molecular size from<br />

47.5 kDa-59 kDa; however all is<strong>of</strong>orms migrate<br />

slower than predicted on SDS-PAGE, with<br />

calculated molecular sizes ranging 78 kDa-90 kDa.<br />

Expression<br />

MIER1beta protein is expressed ubiquitously, while<br />

MIER1alpha protein is expressed mainly in a subset<br />

<strong>of</strong> endocrine organs <strong>and</strong> endocrine responsive<br />

tissues, including the pancreatic islets, adrenal<br />

gl<strong>and</strong>s, testis, ovary, hypothalamus, pituitary,<br />

parafollicular cells <strong>of</strong> the thyroid <strong>and</strong> mammary<br />

ductal epithelium.<br />

Localisation<br />

MIER1beta is nuclear in all adult cell types but is<br />

retained in the cytoplasm <strong>of</strong> the pre-gastrula<br />

Xenopus embryo. MIER1alpha is cytoplasmic in<br />

most cell types, but localized in the nucleus in<br />

normal mammary ductal epithelium. During<br />

progression to invasive breast carcinoma, its<br />

subcellular localization shifts from nuclear to<br />

exclusively cytoplasmic.<br />

Function<br />

MIER1alpha <strong>and</strong> beta function in transcriptional<br />

repression by at least two distinct mechanisms:<br />

recruitment <strong>and</strong> regulation <strong>of</strong> chromatin modifying<br />

enzymes, including HDAC1, HDAC2, CBP <strong>and</strong><br />

G9a; interaction with transcription factors, such as<br />

Sp1 <strong>and</strong> ERalpha, to repress transcription <strong>of</strong> their<br />

respective target genes. MIER1alpha inhibits<br />

estrogen-stimulated anchorage-independent growth<br />

<strong>of</strong> breast carcinoma cells.<br />

Homology<br />

The MIER1 gene family contains two other<br />

members, MIER2 <strong>and</strong> MIER3. The MIER1 gene is<br />

conserved in chimpanzee, dog, cow, mouse, rat,<br />

chicken, frog, zebrafish, fruit fly, <strong>and</strong> C. elegans.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 126


MIER1 (mesoderm induction early response 1 homolog<br />

(Xenopus laevis))<br />

Gillespie LL, Paterno GD<br />

Schematic illustrating the common internal domains <strong>of</strong> the MIER1 is<strong>of</strong>orms <strong>and</strong> the variant amino- (N-) <strong>and</strong> carboxy- (C-<br />

) termini. Transcription from the P1 promoter produces proteins that either begin with M-L- or with the sequence encoded by<br />

exon 3A (MFMFNWFTDCLWTLFLSNYQ). Transcription from the P2 promoter produces a protein that begins with M-A-E-. The<br />

variant N-termini <strong>of</strong> the MIER1 is<strong>of</strong>orms are followed by common internal sequence containing several distinct domains: acidic,<br />

which function in transcriptional activation (Paterno et al., 1997); ELM2, responsible for recruitment <strong>of</strong> HDAC1 (Ding et al., 2003);<br />

SANT, which interacts with Sp1 (Ding et al., 2004) <strong>and</strong> PSPPP, which is required for MIER1 activity in the Xenopus embryo<br />

(Teplitsky et al., 2003). The two alternate C-termini, alpha <strong>and</strong> beta, result from removal or inclusion <strong>and</strong> read-through <strong>of</strong> intron<br />

16, respectively. The alpha C-terminus contains a classic LXXLL motif for interaction with nuclear receptors; the beta C-terminus<br />

contains a nuclear localization signal (NLS).<br />

Implicated in<br />

Breast cancer<br />

Note<br />

Initial studies showed that total MIER1 mRNA<br />

levels were increased in breast carcinoma cell lines<br />

<strong>and</strong> tumour samples (Paterno et al., 1998); in a<br />

more recent study, no consistent difference in<br />

MIER1alpha protein expression levels between<br />

normal breast <strong>and</strong> tumour samples was detected<br />

(McCarthy et al., 2008). Immunohistochemical<br />

analysis <strong>of</strong> patient biopsies revealed that<br />

MIER1alpha protein is expressed primarily in<br />

ductal epithelial cells in normal breast tissue, with<br />

little or no expression in the surrounding stroma; in<br />

breast carcinoma samples, its expression is<br />

restricted to tumour cells. While there is no<br />

difference in expression levels, the subcellular<br />

localization <strong>of</strong> MIER1alpha changes dramatically<br />

during tumour progression: MIER1alpha is nuclear<br />

in 75% <strong>of</strong> normal breast samples <strong>and</strong> in 77% <strong>of</strong><br />

hyperplasia, but in breast carcinoma, only 51% <strong>of</strong><br />

ductal carcinoma in situ, 25% <strong>of</strong> invasive lobular<br />

carcinoma <strong>and</strong> 4% <strong>of</strong> invasive ductal carcinoma<br />

contained nuclear MIER1alpha (McCarthy et al.,<br />

2008). This shift from nuclear to cytoplasmic<br />

localization <strong>of</strong> MIER1alpha during breast cancer<br />

progression suggests that loss <strong>of</strong> nuclear<br />

MIER1alpha contributes to the development <strong>of</strong><br />

invasive breast carcinoma. MIER1alpha inhibits<br />

ERalpha transcriptional activity <strong>and</strong> overexpression<br />

<strong>of</strong> MIER1alpha in breast carcinoma cells inhibits<br />

estrogen-stimulated anchorage-independent growth<br />

(McCarthy et al., 2008).<br />

References<br />

Paterno GD, Li Y, Luchman HA, Ryan PJ, Gillespie LL.<br />

cDNA cloning <strong>of</strong> a novel, developmentally regulated<br />

immediate early gene activated by fibroblast growth factor<br />

<strong>and</strong> encoding a nuclear protein. J Biol Chem. 1997 Oct<br />

10;272(41):25591-5<br />

Paterno GD, Mercer FC, Chayter JJ, Yang X, Robb JD,<br />

Gillespie LL. Molecular cloning <strong>of</strong> human er1 cDNA <strong>and</strong> its<br />

differential expression in breast tumours <strong>and</strong> tumourderived<br />

cell lines. <strong>Gene</strong>. 1998 Nov 5;222(1):77-82<br />

Luchman HA, Paterno GD, Kao KR, Gillespie LL.<br />

Differential nuclear localization <strong>of</strong> ER1 protein during<br />

embryonic development in Xenopus laevis. Mech Dev.<br />

1999 Jan;80(1):111-4<br />

Post JN, Gillespie LL, Paterno GD. Nuclear localization<br />

signals in the Xenopus FGF embryonic early response 1<br />

protein. FEBS Lett. 2001 Jul 27;502(1-2):41-5<br />

Paterno GD, Ding Z, Lew YY, Nash GW, Mercer FC,<br />

Gillespie LL. Genomic organization <strong>of</strong> the human mi-er1<br />

gene <strong>and</strong> characterization <strong>of</strong> alternatively spliced is<strong>of</strong>orms:<br />

regulated use <strong>of</strong> a facultative intron determines subcellular<br />

localization. <strong>Gene</strong>. 2002 Jul 24;295(1):79-88<br />

Ding Z, Gillespie LL, Paterno GD. Human MI-ER1 alpha<br />

<strong>and</strong> beta function as transcriptional repressors by<br />

recruitment <strong>of</strong> histone deacetylase 1 to their conserved<br />

ELM2 domain. Mol Cell Biol. 2003 Jan;23(1):250-8<br />

Teplitsky Y, Paterno GD, Gillespie LL. Proline365 is a<br />

critical residue for the activity <strong>of</strong> XMI-ER1 in Xenopus<br />

embryonic development. Biochem Biophys Res Commun.<br />

2003 Sep 5;308(4):679-83<br />

Beausoleil SA, Jedrychowski M, Schwartz D, Elias JE,<br />

Villén J, Li J, Cohn MA, Cantley LC, Gygi SP. Large-scale<br />

characterization <strong>of</strong> HeLa cell nuclear phosphoproteins.<br />

Proc Natl Acad Sci U S A. 2004 Aug 17;101(33):12130-5<br />

Ding Z, Gillespie LL, Mercer FC, Paterno GD. The SANT<br />

domain <strong>of</strong> human MI-ER1 interacts with Sp1 to interfere<br />

with GC box recognition <strong>and</strong> repress transcription from its<br />

own promoter. J Biol Chem. 2004 Jul 2;279(27):28009-16<br />

Post JN, Luchman HA, Mercer FC, Paterno GD, Gillespie<br />

LL. Developmentally regulated cytoplasmic retention <strong>of</strong> the<br />

transcription factor XMI-ER1 requires sequence in the<br />

acidic activation domain. Int J Biochem Cell Biol. 2005<br />

Feb;37(2):463-77<br />

Thorne LB, Grant AL, Paterno GD, Gillespie LL. Cloning<br />

<strong>and</strong> characterization <strong>of</strong> the mouse ortholog <strong>of</strong> mi-er1. DNA<br />

Seq. 2005 Jun;16(3):237-40<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 127


MIER1 (mesoderm induction early response 1 homolog<br />

(Xenopus laevis))<br />

Nousiainen M, Silljé HH, Sauer G, Nigg EA, Körner R.<br />

Phosphoproteome analysis <strong>of</strong> the human mitotic spindle.<br />

Proc Natl Acad Sci U S A. 2006 Apr 4;103(14):5391-6<br />

Blackmore TM, Mercer CF, Paterno GD, Gillespie LL. The<br />

transcriptional c<strong>of</strong>actor MIER1-beta negatively regulates<br />

histone acetyltransferase activity <strong>of</strong> the CREB-binding<br />

protein. BMC Res Notes. 2008 Aug 22;1:68<br />

McCarthy PL, Mercer FC, Savicky MW, Carter BA, Paterno<br />

GD, Gillespie LL. Changes in subcellular localisation <strong>of</strong> MI-<br />

ER1 alpha, a novel oestrogen receptor-alpha interacting<br />

protein, is associated with breast cancer progression. Br J<br />

Cancer. 2008 Aug 19;99(4):639-46<br />

Gillespie LL, Paterno GD<br />

Thorne LB, McCarthy PL, Paterno GD, Gillespie LL.<br />

Protein expression <strong>of</strong> the transcriptional regulator MI-ER1<br />

alpha in adult mouse tissues. J Mol Histol. 2008<br />

Feb;39(1):15-24<br />

Wang L, Charroux B, Kerridge S, Tsai CC. Atrophin<br />

recruits HDAC1/2 <strong>and</strong> G9a to modify histone H3K9 <strong>and</strong> to<br />

determine cell fates. EMBO Rep. 2008 Jun;9(6):555-62<br />

This article should be referenced as such:<br />

Gillespie LL, Paterno GD. MIER1 (mesoderm induction<br />

early response 1 homolog (Xenopus laevis)). <strong>Atlas</strong> <strong>Gene</strong>t<br />

Cytogenet Oncol Haematol. 2012; 16(2):125-128.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 128


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

<strong>Gene</strong> <strong>Section</strong><br />

Review<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

PA2G4 (proliferation-associated 2G4, 38kDa)<br />

Anne Hamburger, Arundhati Ghosh, Smita Awasthi<br />

University <strong>of</strong> Maryl<strong>and</strong> School <strong>of</strong> Medicine, Department <strong>of</strong> Pathology <strong>and</strong> University <strong>of</strong> Maryl<strong>and</strong><br />

Greenebaum Cancer Center, Baltimore, USA (AH, AG, SA)<br />

Published in <strong>Atlas</strong> Database: September 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/<strong>Gene</strong>s/PA2G4ID41628ch12q13.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI PA2G4ID41628ch12q13.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Identity<br />

Other names: EBP1, HG4-1, ITAF45, p38-2G4<br />

HGNC (Hugo): PA2G4<br />

Location: 12q13.2<br />

Note<br />

PA2G4 encodes a cell-cycle regulated protein<br />

capable <strong>of</strong> interacting with DNA, RNA <strong>and</strong> protein.<br />

The gene was isolated as a DNA binding protein<br />

(p38-2g4) (Radomski <strong>and</strong> Jost, 1995) <strong>and</strong> also as an<br />

ErbB3-interacting protein (EBP1) (Yoo et al.,<br />

2000). Two different is<strong>of</strong>orms <strong>of</strong> EBP1 play a role<br />

in cell survival, cell cycle arrest <strong>and</strong> differentiation.<br />

The long form may have an oncogenic function<br />

when overexpressed, <strong>and</strong> the short form acts as a<br />

tumor suppressor (Liu et al., 2006). EPB1 also<br />

functions as a transcriptional repressor <strong>of</strong> E2F1regulated<br />

genes (Zhang et al., 2004) <strong>and</strong> the<br />

<strong>and</strong>rogen receptor (AR) (Zhang et al., 2005)<br />

through its interactions with histone deacetylases<br />

<strong>and</strong> Sin3A.<br />

DNA/RNA<br />

The alignment <strong>of</strong> PA2G4 mRNA to its genomic sequence.<br />

Description<br />

The PA2G4 gene contains 13 exons. The sizes <strong>of</strong><br />

the exons 1-13 are 88, 128, 105, 69, 92, 63, 78, 78,<br />

133, 94, 127, 53 <strong>and</strong> 65 bp (to the stop codon).<br />

Exon 1 contains the translation initiation ATG.<br />

Exon 13 contains the stop codon.<br />

Transcription<br />

The human PA2G4 promoter contains several<br />

putative transcription factor binding sites. The<br />

major transcript length is 2643 nt. Two proteins are<br />

translated due to alternative splicing (Liu et al.,<br />

2006). An alternatively spliced version missing 29<br />

NT between the first <strong>and</strong> third ATGs has been<br />

observed.<br />

The PA2G4 promoter contains two t<strong>and</strong>em DNA<br />

elements that bind E2F1. E2F1 increases<br />

endogenous EBP1 mRNA levels in cancer cells, but<br />

decreases EBP1 mRNA abundance in non<br />

transformed cells (Judah et al., 2010).<br />

Pseudogene<br />

Six pseudogenes, located on chromosomes 3, 6, 9,<br />

18, 20 <strong>and</strong> X, have been identified.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 129


PA2G4 (proliferation-associated 2G4, 38kDa) Hamburger A, et al.<br />

The linear schematic <strong>of</strong> EBP1. Functional domains, including Nucleolar Localization Signal (NuLS), σ70 RNA binding region<br />

(σ70), amphipathic helical domain (AHD), LXXLL nuclear receptor binding motif (LX) <strong>and</strong> demonstrated in vivo phosphorylation<br />

sites (*).<br />

Protein<br />

Description<br />

p38-2G4 was initially isolated as a DNA binding<br />

protein from mouse Ehrlich ascites cells (Radomski<br />

<strong>and</strong> Jost, 1995). The MW <strong>of</strong> this protein is<br />

predicted to be 38058 D, consisting <strong>of</strong> 340 amino<br />

acids. The human orthologue EBP1 was later<br />

identified as an ErbB3 binding protein <strong>of</strong> the same<br />

MW as the mouse protein (Yoo et al., 2000). This<br />

form migrates at approximately 42 kD in SDS-<br />

PAGE gels. Later, a larger 394 amino acid form<br />

(predicted MW 43787 D, migrating at 48 kD) was<br />

observed in mammalian cells (Xia et al., 2001). The<br />

two forms have been demonstrated to be the result<br />

<strong>of</strong> alternative splicing (Liu et al., 2006) or usage <strong>of</strong><br />

alternative translation initiation sites (Xia et al.,<br />

2001). Amino acids 1-48 are required for nucleolar<br />

localization <strong>and</strong> the C terminal domain (aa 364-<br />

394) is required for interactions with nucleic acids<br />

(Moonie et al., 2007) <strong>and</strong> protein (Zhang et al.,<br />

2002).<br />

EBP1 is post translationally modified at several<br />

phosphorylation sites (Ser 360 (Ahn et al., 2006),<br />

Ser 363 (Akinmade et al., 2007) <strong>and</strong> Thr 261<br />

(Akinmade et al., 2008)) in vivo. The protein<br />

stability <strong>of</strong> the short form is regulated by<br />

ubiquitination (Liu et al., 2009). The short form is<br />

also sumoylated by the TLF/FUS E3 ligase <strong>and</strong> this<br />

sumoylation is required for the anti-proliferative<br />

effects <strong>of</strong> EBP1 (Oh et al., 2010).<br />

The crystal structure <strong>of</strong> both murine (Monie et al.,<br />

2007) <strong>and</strong> human (Kowalinski et al., 2007) EBP1<br />

has been solved. There is a core domain that is<br />

homologous to methionine aminopeptidases,<br />

although no enzymatic activity has been reported.<br />

The C terminal domain containing a Lys-rich<br />

nuclear hormone receptor binding motif (LKALL)<br />

was reported to mediate RNA binding (Monie et al.,<br />

2007).<br />

Expression<br />

EBP1 has been found to be ubiquitously expressed<br />

with high expression levels in skeletal muscle (Yoo<br />

et al., 2000).<br />

Localisation<br />

Under logarithmically growing conditions in cell<br />

culture, EBP1 localizes to the nucleolus <strong>and</strong> the<br />

cytoplasm (Xia et al., 2001; Squatrito et al., 2004).<br />

Upon stimulation with the ErbB3 lig<strong>and</strong> heregulin,<br />

the short form <strong>of</strong> EBP1 is recruited to the nucleus in<br />

AU565 breast cancer cells (Yoo et al., 2000).<br />

Sumoylation is required for nuclear translocation<br />

(Oh et al., 2010). In primary normal epithelial cells,<br />

EBP1 is confined to the cytoplasm (Zhang et al.,<br />

2008b).<br />

Function<br />

EBP1 was initially isolated as a cell cycle-regulated<br />

DNA binding protein (Radomski <strong>and</strong> Jost, 1995)<br />

<strong>and</strong> has been shown to induce cell cycle arrest in<br />

the G2/M phase <strong>of</strong> the cell cycle (Zhang et al.,<br />

2005). EBP1 acts as a corepressor for several<br />

proliferation-associated genes including Cyclin D1,<br />

E2F1 (Zhang <strong>and</strong> Hamburger, 2004) <strong>and</strong> the<br />

<strong>and</strong>rogen receptor (Zhang et al., 2005). EBP1<br />

inhibits transcription <strong>of</strong> these genes by recruiting<br />

HDAC2 via Sin3A to the E2F1 <strong>and</strong> AR-regulated<br />

promoters (Zhang et al., 2005). EBP1 interacts with<br />

RB1 <strong>and</strong> the interaction is enhanced upon EBP1<br />

dephosphorylation (Xia et al., 2001).<br />

EBP1 was isolated as an ErbB3 binding protein<br />

using a yeast-two hybrid screen (Yoo et al., 2000).<br />

The interactions <strong>of</strong> EBP1 with ErbB3 is disrupted<br />

by the ErbB3 lig<strong>and</strong> heregulin, leading to EBP1<br />

nuclear translocation. This leads to the eventual<br />

inhibition <strong>of</strong> heregulin-stimulated proliferation,<br />

presumably due to the repression <strong>of</strong> proliferation<br />

associated genes (Zhang et al., 2008a).<br />

EBP1 also binds RNA <strong>and</strong> associates with 28S, 18S<br />

<strong>and</strong> 5.8S mature rRNAs, several rRNA precursors<br />

<strong>and</strong> probably U3 small nucleolar RNA. It has been<br />

implicated in the regulation <strong>of</strong> intermediate <strong>and</strong> late<br />

steps <strong>of</strong> rRNA processing (Squatrito et al., 2004;<br />

Squatrito et al., 2006). EBP1 also mediates capindependent<br />

translation <strong>of</strong> specific viral IRES<br />

(internal ribosome entry site) (Pilipenko et al.,<br />

2000). EBP1 regulates translation <strong>of</strong> AR mRNA<br />

(Zhou et al., 2010).<br />

EBP1 has also been implicated in protein stability<br />

via its interaction with the proteasome.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 130


PA2G4 (proliferation-associated 2G4, 38kDa) Hamburger A, et al.<br />

Overexpression <strong>of</strong> EBP1 results in decreased<br />

stability <strong>of</strong> ErbB2 protein in breast cancer cells via<br />

a proteasome-mediated pathway (Lu et al., 2011).<br />

The long form <strong>of</strong> EBP1 binds to the p53 E3 ligase<br />

HDM2, enhancing HDM2-p53 interactions <strong>and</strong><br />

promoting p53 degradation (Kim et al., 2011).<br />

The long (p48) <strong>and</strong> short (p42) forms <strong>of</strong> EBP1 have<br />

opposing biological effects, with the longer form<br />

inducing cell survival <strong>and</strong> the shorter form<br />

inhibiting cell growth (Liu et al., 2006). The long<br />

form binds HDM2, promoting degradation <strong>of</strong> p53<br />

(Kim et al., 2010).<br />

Homology<br />

Similar (30% identity) to the 42 kDA DNA binding<br />

protein SF00553 in S. pombe yeast (Yamada et al.,<br />

1994) <strong>and</strong> StEBP1 in potato (Horvath et al., 2006).<br />

The orthologue in potatoes (StEBP1) has 69%<br />

sequence similarity to human EBP1 <strong>and</strong> can inhibit<br />

growth <strong>of</strong> human breast cancer cell lines <strong>and</strong> E2F1<br />

expression in these cells.<br />

Mutations<br />

Germinal<br />

No mutations in the PA2G4 gene have been<br />

reported.<br />

Somatic<br />

None reported.<br />

Implicated in<br />

Prostate cancer<br />

Prognosis<br />

Decreased expression <strong>of</strong> EBP1 is associated with<br />

higher tumor grade <strong>and</strong> metastasis in prostate<br />

cancer (Zhang et al., 2008b). However, another<br />

study indicated EBP1 expression increased with<br />

disease progression (Gannon et al., 2008).<br />

Breast cancer<br />

Prognosis<br />

Deletion <strong>of</strong> EBP1 results in tamoxifen resistance in<br />

breast cancer (Zhang et al., 2008a). However,<br />

patients with a high level <strong>of</strong> EBP1 protein have a<br />

poor clinical outcome (Ou et al., 2006).<br />

Glioblastoma<br />

Prognosis<br />

Glioblastoma patients expressing a high level <strong>of</strong><br />

p48 EBP1 have a worse prognosis than those<br />

expressing lower levels <strong>of</strong> the protein (Kim et al.,<br />

2010; Kwon <strong>and</strong> Ahn, 2011).<br />

References<br />

Yamada H, Mori H, Momoi H, Nakagawa Y, Ueguchi C,<br />

Mizuno T. A fission yeast gene encoding a protein that<br />

preferentially associates with curved DNA. Yeast. 1994<br />

Jul;10(7):883-94<br />

Radomski N, Jost E. Molecular cloning <strong>of</strong> a murine cDNA<br />

encoding a novel protein, p38-2G4, which varies with the<br />

cell cycle. Exp Cell Res. 1995 Oct;220(2):434-45<br />

Pilipenko EV, Pestova TV, Kolupaeva VG, Khitrina EV,<br />

Poperechnaya AN, Agol VI, Hellen CU. A cell cycledependent<br />

protein serves as a template-specific translation<br />

initiation factor. <strong>Gene</strong>s Dev. 2000 Aug 15;14(16):2028-45<br />

Yoo JY, Wang XW, Rishi AK, Lessor T, Xia XM, Gustafson<br />

TA, Hamburger AW. Interaction <strong>of</strong> the PA2G4 (EBP1)<br />

protein with ErbB-3 <strong>and</strong> regulation <strong>of</strong> this binding by<br />

heregulin. Br J Cancer. 2000 Feb;82(3):683-90<br />

Xia X, Cheng A, Lessor T, Zhang Y, Hamburger AW.<br />

Ebp1, an ErbB-3 binding protein, interacts with Rb <strong>and</strong><br />

affects Rb transcriptional regulation. J Cell Physiol. 2001<br />

May;187(2):209-17<br />

Squatrito M, Mancino M, Donzelli M, Areces LB, Draetta<br />

GF. EBP1 is a nucleolar growth-regulating protein that is<br />

part <strong>of</strong> pre-ribosomal ribonucleoprotein complexes.<br />

Oncogene. 2004 May 27;23(25):4454-65<br />

Zhang Y, Hamburger AW. Heregulin regulates the ability <strong>of</strong><br />

the ErbB3-binding protein Ebp1 to bind E2F promoter<br />

elements <strong>and</strong> repress E2F-mediated transcription. J Biol<br />

Chem. 2004 Jun 18;279(25):26126-33<br />

Zhang Y, Akinmade D, Hamburger AW. The ErbB3 binding<br />

protein Ebp1 interacts with Sin3A to repress E2F1 <strong>and</strong> ARmediated<br />

transcription. Nucleic Acids Res.<br />

2005;33(18):6024-33<br />

Ahn JY, Liu X, Liu Z, Pereira L, Cheng D, Peng J, Wade<br />

PA, Hamburger AW, Ye K. Nuclear Akt associates with<br />

PKC-phosphorylated Ebp1, preventing DNA fragmentation<br />

by inhibition <strong>of</strong> caspase-activated DNase. EMBO J. 2006<br />

May 17;25(10):2083-95<br />

Horváth BM, Magyar Z, Zhang Y, Hamburger AW, Bakó L,<br />

Visser RG, Bachem CW, Bögre L. EBP1 regulates organ<br />

size through cell growth <strong>and</strong> proliferation in plants. EMBO<br />

J. 2006 Oct 18;25(20):4909-20<br />

Liu Z, Ahn JY, Liu X, Ye K. Ebp1 is<strong>of</strong>orms distinctively<br />

regulate cell survival <strong>and</strong> differentiation. Proc Natl Acad<br />

Sci U S A. 2006 Jul 18;103(29):10917-22<br />

Ou K, Kesuma D, Ganesan K, Yu K, Soon SY, Lee SY,<br />

Goh XP, Hooi M, Chen W, Jikuya H, Ichikawa T, Kuyama<br />

H, Matsuo E, Nishimura O, Tan P. Quantitative pr<strong>of</strong>iling <strong>of</strong><br />

drug-associated proteomic alterations by combined 2nitrobenzenesulfenyl<br />

chloride (NBS) isotope labeling <strong>and</strong><br />

2DE/MS identification. J Proteome Res. 2006<br />

Sep;5(9):2194-206<br />

Squatrito M, Mancino M, Sala L, Draetta GF. Ebp1 is a<br />

dsRNA-binding protein associated with ribosomes that<br />

modulates eIF2alpha phosphorylation. Biochem Biophys<br />

Res Commun. 2006 Jun 9;344(3):859-68<br />

Akinmade D, Lee M, Zhang Y, Hamburger AW. Ebp1mediated<br />

inhibition <strong>of</strong> cell growth requires serine 363<br />

phosphorylation. Int J Oncol. 2007 Oct;31(4):851-8<br />

Kowalinski E, Bange G, Bradatsch B, Hurt E, Wild K,<br />

Sinning I. The crystal structure <strong>of</strong> Ebp1 reveals a<br />

methionine aminopeptidase fold as binding platform for<br />

multiple interactions. FEBS Lett. 2007 Sep<br />

18;581(23):4450-4<br />

Monie TP, Perrin AJ, Birtley JR, Sweeney TR,<br />

Karakasiliotis I, Chaudhry Y, Roberts LO, Matthews S,<br />

Goodfellow IG, Curry S. Structural insights into the<br />

transcriptional <strong>and</strong> translational roles <strong>of</strong> Ebp1. EMBO J.<br />

2007 Sep 5;26(17):3936-44<br />

Akinmade D, Talukder AH, Zhang Y, Luo WM, Kumar R,<br />

Hamburger AW. Phosphorylation <strong>of</strong> the ErbB3 binding<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 131


PA2G4 (proliferation-associated 2G4, 38kDa) Hamburger A, et al.<br />

protein Ebp1 by p21-activated kinase 1 in breast cancer<br />

cells. Br J Cancer. 2008 Mar 25;98(6):1132-40<br />

Gannon PO, Koumakpayi IH, Le Page C, Karakiewicz PI,<br />

Mes-Masson AM, Saad F. Ebp1 expression in benign <strong>and</strong><br />

malignant prostate. Cancer Cell Int. 2008 Nov 24;8:18<br />

Zhang Y, Akinmade D, Hamburger AW. Inhibition <strong>of</strong><br />

heregulin mediated MCF-7 breast cancer cell growth by<br />

the ErbB3 binding protein EBP1. Cancer Lett. 2008a Jul<br />

8;265(2):298-306<br />

Zhang Y, Linn D, Liu Z, Melamed J, Tavora F, Young CY,<br />

Burger AM, Hamburger AW. EBP1, an ErbB3-binding<br />

protein, is decreased in prostate cancer <strong>and</strong> implicated in<br />

hormone resistance. Mol Cancer Ther. 2008b<br />

Oct;7(10):3176-86<br />

Liu Z, Oh SM, Okada M, Liu X, Cheng D, Peng J, Brat DJ,<br />

Sun SY, Zhou W, Gu W, Ye K. Human BRE1 is an E3<br />

ubiquitin ligase for Ebp1 tumor suppressor. Mol Biol Cell.<br />

2009 Feb;20(3):757-68<br />

Judah D, Chang WY, Dagnino L. EBP1 is a novel E2F<br />

target gene regulated by transforming growth factor-β.<br />

PLoS One. 2010 Nov 10;5(11):e13941<br />

Kim CK, Nguyen TL, Joo KM, Nam DH, Park J, Lee KH,<br />

Cho SW, Ahn JY. Negative regulation <strong>of</strong> p53 by the long<br />

is<strong>of</strong>orm <strong>of</strong> ErbB3 binding protein Ebp1 in brain tumors.<br />

Cancer Res. 2010 Dec 1;70(23):9730-41<br />

Oh SM, Liu Z, Okada M, Jang SW, Liu X, Chan CB, Luo H,<br />

Ye K. Ebp1 sumoylation, regulated by TLS/FUS E3 ligase,<br />

is required for its anti-proliferative activity. Oncogene. 2010<br />

Feb 18;29(7):1017-30<br />

Zhou H, Mazan-Mamczarz K, Martindale JL, Barker A, Liu<br />

Z, Gorospe M, Leedman PJ, Gartenhaus RB, Hamburger<br />

AW, Zhang Y. Post-transcriptional regulation <strong>of</strong> <strong>and</strong>rogen<br />

receptor mRNA by an ErbB3 binding protein 1 in prostate<br />

cancer. Nucleic Acids Res. 2010 Jun;38(11):3619-31<br />

Kwon IS, Ahn JY. P48 Ebp1 acts as a downstream<br />

mediator <strong>of</strong> Trk signaling in neurons, contributing neuronal<br />

differentiation. Neurochem Int. 2011 Feb;58(2):215-23<br />

Lu Y, Zhou H, Chen W, Zhang Y, Hamburger AW. The<br />

ErbB3 binding protein EBP1 regulates ErbB2 protein levels<br />

<strong>and</strong> tamoxifen sensitivity in breast cancer cells. Breast<br />

Cancer Res Treat. 2011 Feb;126(1):27-36<br />

This article should be referenced as such:<br />

Hamburger A, Ghosh A, Awasthi S. PA2G4 (proliferationassociated<br />

2G4, 38kDa). <strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol<br />

Haematol. 2012; 16(2):129-132.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 132


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

<strong>Gene</strong> <strong>Section</strong><br />

Review<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

PRDM1 (PR domain containing 1, with ZNF<br />

domain)<br />

Wayne Tam<br />

Department <strong>of</strong> Pathology <strong>and</strong> Laboratory Medicine, Weill Cornell Medical College, New York, NY,<br />

USA (WT)<br />

Published in <strong>Atlas</strong> Database: September 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/<strong>Gene</strong>s/PRDM1ID41831ch6q21.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI PRDM1ID41831ch6q21.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Identity<br />

Other names: BLIMP-1, BLIMP1, PRDI-BF1<br />

HGNC (Hugo): PRDM1<br />

Location: 6q21<br />

DNA/RNA<br />

Description<br />

The gene encompasses 23,6 kb DNA in humans,<br />

from 106534195 to 106557814 (hg19-Feb, 2009) in<br />

the long arm <strong>of</strong> chromosome 6. It encodes 7 exons.<br />

The open reading frame spans all 7 exons.<br />

Figure 1.<br />

Transcription<br />

The mRNAs encoded by PRDM1 have two<br />

transcript is<strong>of</strong>orms, PRDM1alpha <strong>and</strong> PRDM1beta,<br />

which are 5164 <strong>and</strong> 4675 bp long, respectively. The<br />

shorter is<strong>of</strong>orm is generated by usage <strong>of</strong> the<br />

alternative promoter located in intron 3 <strong>and</strong><br />

contains a different 5' untranslated region. It lacks<br />

the 5' inframe portion <strong>of</strong> the coding region present<br />

in PRDM1alpha. A deletion exon 6 (exon 7 in<br />

mice) variant has also been described (Schmidt et<br />

al., 2008).<br />

Pseudogene<br />

None.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 133


PRDM1 (PR domain containing 1, with ZNF domain) Tam W<br />

Protein<br />

Description<br />

PRDM1alpha is the larger is<strong>of</strong>orm <strong>and</strong> contains 825<br />

amino acids, with a predicted molecular weight <strong>of</strong><br />

92 kD. It has a PR domain at the N-terminal portion<br />

(86-207 aa) <strong>of</strong> the protein, which is related to the<br />

SET domain (SM00317, SMART) found in many<br />

histone methyltransferases. In contrast to bona fide<br />

SET-domain proteins, the PR domain in PRDM1<br />

does not possess intrinsic histone methyltransferase<br />

activity. Five C2H2-type zinc fingers (SM00355,<br />

SMART), which represent the DNA binding<br />

domain, are present at the C-terminal portion <strong>of</strong> the<br />

protein (575-707 aa). The middle part <strong>of</strong> PRDM1<br />

(about 300-400 aa) is rich in proline <strong>and</strong> serine.<br />

PRDM1beta lacks the N-terminal 101 amino acids<br />

<strong>of</strong> the PRDM1alpha, <strong>and</strong> has a truncated PR<br />

domain. PRDM1beta has been shown to be<br />

functionally impaired in its transcriptional<br />

repression activity (Gyory et al., 2003). The<br />

proximal 3 zinc fingers in PRDM1/Blimp-1 delta<br />

exon 6 variant are disrupted (Schmidt et al., 2008).<br />

This protein variant has auto-regulatory potential<br />

through negative regulation <strong>of</strong> PRDM1/Blimp-1<br />

expression <strong>and</strong> functions.<br />

Expression<br />

PRDM1/Blimp-1 is expressed in several tissues <strong>and</strong><br />

organs, including hematopoietic tissues, skin,<br />

central nervous system, testis <strong>and</strong> gut. Within the<br />

hematopoietic system, PRDM1/Blimp-1 is<br />

expressed in plasma cells, B cells, T cells, natural<br />

killer cells, monocytes, granulocytes <strong>and</strong> dendritic<br />

cells. In the B cell population, PRDM1/Blimp-1<br />

starts to be expressed when they are committed to<br />

undergo terminal differentiation (Cattoretti et al.,<br />

2005). Consistent with its expression in normal<br />

lymphoid cells, PRDM1 is also expressed in 100%<br />

<strong>of</strong> multiple myeloma, <strong>and</strong> in various frequency in B<br />

<strong>and</strong> T lymphomas (Garcia et al., 2006; D'Costa et<br />

al., 2009). PRDM1/Blimp-1 expression is regulated<br />

by a number <strong>of</strong> transcription activators <strong>and</strong><br />

repressors (Martins <strong>and</strong> Calame, 2008).<br />

MicroRNAs have also been implicated in the downregulation<br />

<strong>of</strong> PRDM1 in both normal germinal<br />

center B cells (Malumbres et al., 2009; Zhang et al.,<br />

2009; Gururajan et al., 2010) <strong>and</strong> in neoplastic<br />

lymphoid cells (Nie et al., 2008; Nie et al., 2010).<br />

Among PRDM1 is<strong>of</strong>orms, PRDM1alpha is the<br />

Figure 2.<br />

most abundantly expressed, although relative<br />

abundance between PRDM1alpha <strong>and</strong> PRDM1beta<br />

may vary between cell types (Garcia et al., 2006).<br />

PRDM1/Blimp-1 delta 6 is<strong>of</strong>orm is preferentially<br />

expressed in resting B cells (Schmidt et al., 2008).<br />

Localisation<br />

Nuclear.<br />

Function<br />

Mechanisms <strong>of</strong> action<br />

PRDM1/Blimp-1 is a transcription repressor that<br />

binds to specific DNA sequences through its zinc<br />

fingers, <strong>and</strong> functions as a scaffold for recruiting<br />

co-repressors that catalyze histone modifications. It<br />

has no known intrinsic histone methyltransferase<br />

activity. PRDM1/blimp-1 has been shown to<br />

mediate transcriptional silencing via interactions<br />

with H3 lysine methyltransferase G9a (Gyory et al.,<br />

2004), histone deacetylases HDAC1 <strong>and</strong> HDAC2<br />

(Yu et al., 2000), <strong>and</strong> H3 lysine demethylase LSD1<br />

(Su et al., 2009). PRDM1 can also tether Groucho<br />

family <strong>of</strong> transcription factors to mediate repression<br />

<strong>of</strong> gene transcription (Ren et al., 1999). Interaction<br />

<strong>of</strong> PRDM1/Blimp-1 with these co-repressors is<br />

mediated through the proline/serine rich domain<br />

<strong>and</strong>/or the zinc fingers. PRDM1 exerts its biological<br />

functions by repressing expressions <strong>of</strong> various<br />

target genes, which can be cell-type specific. For<br />

example, PRDM1/Blimp-1 mediates cell-cycle<br />

arrest in B cells by repressing c-myc, <strong>and</strong> in CD4<br />

positive helper T cells by inhibiting expressions <strong>of</strong><br />

IL-2 <strong>and</strong> Fos, an IL-2 activator (Martins <strong>and</strong><br />

Calame, 2008).<br />

Biologic functions<br />

PRDM1 was originally identified as a novel<br />

repressor <strong>of</strong> human beta-interferon gene expression<br />

called PRDI-BF1 (positive regulatory domain I<br />

binding factor 1) (Keller <strong>and</strong> Maniatis, 1991). Later<br />

Blimp-1 (B lymphocyte-induced maturation<br />

protein), which represents the murine homolog <strong>of</strong><br />

PRDI-BF1(Huang, 1994), was discovered as a<br />

protein the enforced expression <strong>of</strong> which was<br />

sufficient to drive plasma cell differentiation <strong>of</strong> a<br />

mouse lymphoma cell line with an activated B cell<br />

phenotype (Turner et al., 1994). Studies since then<br />

have revealed a role <strong>of</strong> PRDM1/Blimp-1 in<br />

multiple cell types.<br />

PRDM1/Blimp-1 is the master regulator <strong>of</strong> plasma<br />

cell differentiation, critical for the generation <strong>of</strong><br />

immunoglobulin-secreting plasma cells <strong>and</strong><br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 134


PRDM1 (PR domain containing 1, with ZNF domain) Tam W<br />

maintenance <strong>of</strong> long-lived plasma cells (Shapiro-<br />

Shelef et al., 2003; Shapiro-Shelef et al., 2005;<br />

Martins <strong>and</strong> Calame, 2008). It mediates terminal B<br />

cell differentiation by extinguishing gene programs<br />

related to B cell signaling <strong>and</strong> proliferation (Shaffer<br />

et al., 2002), <strong>and</strong> allowing induction <strong>of</strong> XBP1<br />

which coordinates phenotypic changes, including<br />

expansion <strong>of</strong> endoplasmic reticulum <strong>and</strong> increased<br />

protein synthesis, that define the plasma cell<br />

phenotype (Shaffer et al., 2004).<br />

PRDM1/Blimp-1 is also required for T cell<br />

homeostasis (Kallies et al., 2006; Martins et al.,<br />

2006; Calame, 2010). It attenuates proliferation <strong>and</strong><br />

survival <strong>of</strong> CD4 positive helper T cells by<br />

repressing IL-2-dependent activation; inhibits Th1<br />

differentiation by down-regulating IFNgamma,<br />

Tbx21 <strong>and</strong> BCL6; <strong>and</strong> antagonizes BCL6dependent<br />

follicular T cell differentiation.<br />

PRDM1/Blimp-1 also plays critical role in the<br />

differentiation <strong>of</strong> CD8-positive T cells.<br />

Besides its critical functions in B <strong>and</strong> T cells,<br />

PRDM1/Blimp-1 promotes differentiation <strong>of</strong><br />

macrophages (Chang et al., 2000). It is also<br />

important for homeostatic development <strong>of</strong> dendritic<br />

cells as well as dendritic cell maturation (Chan et<br />

al., 2009). PRDM1/Blimp-1 negatively regulates<br />

NK cell activation by suppressing effector cytokine<br />

production (Smith et al., 2010).<br />

In the non-hematopoietic cell types,<br />

PRDM1/Blimp-1 plays an important role in the<br />

terminal differentiation <strong>of</strong> keratinocytes<br />

(Magnusdottir et al., 2007) <strong>and</strong> sebaceous cells<br />

(Sellheyer <strong>and</strong> Krahl, 2010). It also regulates<br />

postnatal reprogramming <strong>of</strong> intestinal enterocytes<br />

(Harper et al., 2011).<br />

Furthermore, PRDM1/Blimp-1 is required in<br />

embryonic developmental cell fate specification <strong>and</strong><br />

organ morphogenesis, as demonstrated by mouse<br />

<strong>and</strong> zebrafish model systems (Bik<strong>of</strong>f et al., 2009).<br />

Blimp-1 specifies cell fate <strong>of</strong> primordial germ cells<br />

(Vincent et al., 2005), neural crests <strong>and</strong> neuron<br />

progenitors (Roy <strong>and</strong> Ng, 2004), <strong>and</strong> also muscle<br />

fiber identity (Baxendale et al., 2004; H<strong>of</strong>sten et al.,<br />

2008). Blimp-1 is also critical for the normal<br />

development <strong>of</strong> the cardiovascular system,<br />

forelimbs <strong>and</strong> vibrissae (Robertson et al., 2007).<br />

Homology<br />

PRDM1 is conserved in chimpanzee, dog, cow,<br />

mouse, rat, chicken <strong>and</strong> zebrafish.<br />

Mutations<br />

Somatic<br />

Somatic mutations have been identified in diffuse<br />

large B cell lymphomas (DLBCL) at a frequency <strong>of</strong><br />

about 8 to 23% (Tam et al., 2006; Pasqualucci et<br />

al., 2006; M<strong>and</strong>elbaum et al., 2010). One group did<br />

not detect mutations in 82 cases <strong>of</strong> DLBCLs in the<br />

Chinese population, suggesting geographic<br />

differences in PRDM1 mutations in DLBCL (Liu et<br />

al., 2007). PRDM1 mutations are exclusively<br />

detected in about 24 to 35% <strong>of</strong> the activated B<br />

cell(ABC)/non-germinal center B cell (non-GCB)<br />

subtype <strong>of</strong> DLBCL, <strong>and</strong> have not been identified in<br />

GCB-like DLBCL. In addition, about 20% <strong>of</strong><br />

primary DLBCL <strong>of</strong> the central nervous system, a<br />

subtype <strong>of</strong> DLBCL, harbor PRDM1 mutations<br />

(Courts et al., 2008). PRDM1 mutation was also<br />

found in one case <strong>of</strong> primary effusion lymphoma<br />

out <strong>of</strong> 2 cases analyzed (Tate et al., 2007). The<br />

types <strong>of</strong> somatic mutations seen in PRDM1 include<br />

splice site mutations, frameshift insertion/deletion<br />

<strong>and</strong> less frequently, nonsense <strong>and</strong> missense<br />

mutations. The vast majority <strong>of</strong> these mutations<br />

result in inactivating truncations lacking one or<br />

more <strong>of</strong> the critical domains including PR, proline<br />

rich region, <strong>and</strong> the zinc fingers. Missense<br />

mutations have been demonstrated to affect<br />

PRDM1 functions (M<strong>and</strong>elbaum et al., 2010). Most<br />

<strong>of</strong> the somatic mutations (about 90%) are<br />

associated with inactivation <strong>of</strong> the other allele by<br />

deletion, consistent with biallelic inactivation<br />

characteristic <strong>of</strong> a tumor suppressor gene.<br />

Mutations in acute leukemias <strong>and</strong> solid tumors have<br />

been not identified (Hangaishi <strong>and</strong> Kurokawa,<br />

2010).<br />

Implicated in<br />

Diffuse large B cell lymphomas<br />

Prognosis<br />

Diffuse large B cell lymphoma (DLBCL) with<br />

partial plasmablastic differentiation, as indicated by<br />

PRDM1 expression, has a worse overall <strong>and</strong> failure<br />

free survival compared to conventional DLBLC<br />

without PRDM1 expression, suggesting that<br />

PRDM1 may be useful as a prognostic marker in<br />

DLBCL (Montes-Moreno et al., 2010).<br />

Expression <strong>of</strong> PRDM1beta has been implicated as a<br />

poor prognostic marker in DBLCL treated with<br />

CHOP (but not R-CHOP) <strong>and</strong> in T cell lymphomas.<br />

Resistance to chemotherapeutic agents mediated by<br />

PRDM1beta has been proposed as a potential<br />

mechanism (Liu et al., 2007; Zhao et al., 2008).<br />

Oncogenesis<br />

Inactivation or down-regulation <strong>of</strong> PRDM1 appears<br />

to be an important event in the pathogenesis <strong>of</strong><br />

DLBCLs <strong>of</strong> the ABC/non-GCB subtype. In about<br />

50% <strong>of</strong> DLBCLs <strong>of</strong> this subtype, PRDM1 is<br />

inactivated by truncating or missense mutations,<br />

biallele gene deletions or transcription repression<br />

by a constitutively active, translocated BCL6<br />

(M<strong>and</strong>elbaum et al., 2010). In non-GCB/ABC-like<br />

DLBCL without these genetic lesions,<br />

asynchronous PRDM1 mRNA <strong>and</strong> protein<br />

expressions have been observed, suggesting a posttranscriptional<br />

down-regulation (M<strong>and</strong>elbaum et<br />

al., 2010; Nie et al., 2010). MicroRNAs have been<br />

postulated as a potential mechanism <strong>of</strong> downregulation<br />

(Nie et al., 2010). All these findings are<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 135


PRDM1 (PR domain containing 1, with ZNF domain) Tam W<br />

consistent with a tumor suppressor role <strong>of</strong> PRDM1<br />

in DLBCL. In addition, PRDM1/Blimp-1 has been<br />

directly demonstrated in mouse models to function<br />

as a tumor suppressor gene (Calado et al., 2010;<br />

M<strong>and</strong>elbaum et al., 2010). Mice lacking<br />

PRDM1/Blimp-1 are predisposed to develop<br />

lymphoproliferative disorders resembling human<br />

ABC-like DLBCL. One <strong>of</strong> these mouse models also<br />

demonstrated a cooperative pathogenic effect<br />

between PRDM1 inactivation <strong>and</strong> constitutive NFkB<br />

activation. These findings suggest that PRDM1<br />

inactivation contributes to the pathogenesis <strong>of</strong><br />

ABC-like DLBCL by inhibiting terminal B cell<br />

differentiation induced by constitutive canonical<br />

NF-kB activation characteristically present in this<br />

lymphoma type.<br />

Natural killer cell neoplasms<br />

Oncogenesis<br />

PRDM1 is frequently (about 40%) deleted as part<br />

<strong>of</strong> a minimal deleted region in aggressive natural<br />

killer (NK) cell neoplasms such as extranodal NK/T<br />

cell lymphomas, nasal type, <strong>and</strong> aggressive NK cell<br />

leukemias (Iqbal et al., 2009; Karube et al., 2011).<br />

This deletion is generally associated with a downregulation<br />

<strong>of</strong> PRDM1 expression, associated with<br />

focal hypermethylation <strong>of</strong> the 5' region <strong>of</strong> the other<br />

PRDM1 allele. Enforced expression <strong>of</strong> PRDM1 in<br />

NK lymphoma cell lines results in cell cycle arrest<br />

<strong>and</strong> apoptosis (Karube et al., 2011). Inactivating<br />

nonsense mutations have also been found in two<br />

NK cell lines <strong>and</strong> one clinical NK neoplasm (Iqbal<br />

et al., 2009; Karube et al., 2011). These findings<br />

indicate a tumor suppressor role <strong>of</strong> PRDM1 in NK<br />

cell neoplasms.<br />

Radiation-therapy induced second<br />

malignant neoplasms<br />

Prognosis<br />

Pediatric Hodgkin lymphoma patients treated by<br />

radiation are at risk <strong>of</strong> developing second<br />

malignancies later in life. Two SNP variants were<br />

identified at 6q21 (intergenic between ATG5 <strong>and</strong><br />

PRDM1) in a genome wide association study that<br />

are associated with increased risk <strong>of</strong> second<br />

neoplasms for pediatric patients with Hodgkin<br />

lymphoma who received radiation therapy. These<br />

variants correlate with decreased basal PRDM1<br />

expression <strong>and</strong> reduced PRDM1 induction after<br />

radiation therapy, implicating PRDM1 in the<br />

etiology <strong>of</strong> radiation-therapy induced second<br />

malignancies (Best et al., 2011).<br />

Plasma cell myeloma<br />

Oncogenesis<br />

PRDM1 is consistently expressed in plasma cell<br />

myeloma. A plasmacytoma transgenic mouse<br />

model demonstrates that PRDM1/Blimp-1 is not a<br />

tumor suppressor gene in myeloma. Instead, it<br />

appears to be a limiting factor in plasma cell<br />

transformation (D'Costa et al., 2009). Reducing<br />

PRDM1/Blimp-1 dosage decreases incidence <strong>of</strong><br />

plasmacytoma in mice but does not cause reduction<br />

<strong>of</strong> normal plasma cells in nontransgenic mice. Loss<br />

<strong>of</strong> PRDM1/Blimp-1 eliminates plasmacytoma<br />

development.<br />

References<br />

Keller AD, Maniatis T. Identification <strong>and</strong> characterization <strong>of</strong><br />

a novel repressor <strong>of</strong> beta-interferon gene expression.<br />

<strong>Gene</strong>s Dev. 1991 May;5(5):868-79<br />

Huang S. Blimp-1 is the murine homolog <strong>of</strong> the human<br />

transcriptional repressor PRDI-BF1. Cell. 1994 Jul<br />

15;78(1):9<br />

Turner CA Jr, Mack DH, Davis MM. Blimp-1, a novel zinc<br />

finger-containing protein that can drive the maturation <strong>of</strong> B<br />

lymphocytes into immunoglobulin-secreting cells. Cell.<br />

1994 Apr 22;77(2):297-306<br />

Ren B, Chee KJ, Kim TH, Maniatis T. PRDI-BF1/Blimp-1<br />

repression is mediated by corepressors <strong>of</strong> the Groucho<br />

family <strong>of</strong> proteins. <strong>Gene</strong>s Dev. 1999 Jan 1;13(1):125-37<br />

Chang DH, Angelin-Duclos C, Calame K. BLIMP-1: trigger<br />

for differentiation <strong>of</strong> myeloid lineage. Nat Immunol. 2000<br />

Aug;1(2):169-76<br />

Yu J, Angelin-Duclos C, Greenwood J, Liao J, Calame K.<br />

Transcriptional repression by blimp-1 (PRDI-BF1) involves<br />

recruitment <strong>of</strong> histone deacetylase. Mol Cell Biol. 2000<br />

Apr;20(7):2592-603<br />

Shaffer AL, Lin KI, Kuo TC, Yu X, Hurt EM, Rosenwald A,<br />

Giltnane JM, Yang L, Zhao H, Calame K, Staudt LM.<br />

Blimp-1 orchestrates plasma cell differentiation by<br />

extinguishing the mature B cell gene expression program.<br />

Immunity. 2002 Jul;17(1):51-62<br />

Györy I, Fejér G, Ghosh N, Seto E, Wright KL.<br />

Identification <strong>of</strong> a functionally impaired positive regulatory<br />

domain I binding factor 1 transcription repressor in<br />

myeloma cell lines. J Immunol. 2003 Mar 15;170(6):3125-<br />

33<br />

Shapiro-Shelef M, Lin KI, McHeyzer-Williams LJ, Liao J,<br />

McHeyzer-Williams MG, Calame K. Blimp-1 is required for<br />

the formation <strong>of</strong> immunoglobulin secreting plasma cells<br />

<strong>and</strong> pre-plasma memory B cells. Immunity. 2003<br />

Oct;19(4):607-20<br />

Baxendale S, Davison C, Muxworthy C, Wolff C, Ingham<br />

PW, Roy S. The B-cell maturation factor Blimp-1 specifies<br />

vertebrate slow-twitch muscle fiber identity in response to<br />

Hedgehog signaling. Nat <strong>Gene</strong>t. 2004 Jan;36(1):88-93<br />

Gyory I, Wu J, Fejér G, Seto E, Wright KL. PRDI-BF1<br />

recruits the histone H3 methyltransferase G9a in<br />

transcriptional silencing. Nat Immunol. 2004 Mar;5(3):299-<br />

308<br />

Roy S, Ng T. Blimp-1 specifies neural crest <strong>and</strong> sensory<br />

neuron progenitors in the zebrafish embryo. Curr Biol.<br />

2004 Oct 5;14(19):1772-7<br />

Shaffer AL, Shapiro-Shelef M, Iwakoshi NN, Lee AH, Qian<br />

SB, Zhao H, Yu X, Yang L, Tan BK, Rosenwald A, Hurt<br />

EM, Petroulakis E, Sonenberg N, Yewdell JW, Calame K,<br />

Glimcher LH, Staudt LM. XBP1, downstream <strong>of</strong> Blimp-1,<br />

exp<strong>and</strong>s the secretory apparatus <strong>and</strong> other organelles,<br />

<strong>and</strong> increases protein synthesis in plasma cell<br />

differentiation. Immunity. 2004 Jul;21(1):81-93<br />

Cattoretti G, Angelin-Duclos C, Shaknovich R, Zhou H,<br />

Wang D, Alobeid B. PRDM1/Blimp-1 is expressed in<br />

human B-lymphocytes committed to the plasma cell<br />

lineage. J Pathol. 2005 May;206(1):76-86<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 136


PRDM1 (PR domain containing 1, with ZNF domain) Tam W<br />

Shapiro-Shelef M, Lin KI, Savitsky D, Liao J, Calame K.<br />

Blimp-1 is required for maintenance <strong>of</strong> long-lived plasma<br />

cells in the bone marrow. J Exp Med. 2005 Dec<br />

5;202(11):1471-6<br />

Vincent SD, Dunn NR, Sciammas R, Shapiro-Shalef M,<br />

Davis MM, Calame K, Bik<strong>of</strong>f EK, Robertson EJ. The zinc<br />

finger transcriptional repressor Blimp1/Prdm1 is<br />

dispensable for early axis formation but is required for<br />

specification <strong>of</strong> primordial germ cells in the mouse.<br />

Development. 2005 Mar;132(6):1315-25<br />

Garcia JF, Roncador G, García JF, Sánz AI, Maestre L,<br />

Lucas E, Montes-Moreno S, Fern<strong>and</strong>ez Victoria R,<br />

Martinez-Torrecuadrara JL, Marafioti T, Mason DY, Piris<br />

MA. PRDM1/BLIMP-1 expression in multiple B <strong>and</strong> T-cell<br />

lymphoma. Haematologica. 2006 Apr;91(4):467-74<br />

Kallies A, Hawkins ED, Belz GT, Metcalf D, Hommel M,<br />

Corcoran LM, Hodgkin PD, Nutt SL. Transcriptional<br />

repressor Blimp-1 is essential for T cell homeostasis <strong>and</strong><br />

self-tolerance. Nat Immunol. 2006 May;7(5):466-74<br />

Martins GA, Cimmino L, Shapiro-Shelef M, Szabolcs M,<br />

Herron A, Magnusdottir E, Calame K. Transcriptional<br />

repressor Blimp-1 regulates T cell homeostasis <strong>and</strong><br />

function. Nat Immunol. 2006 May;7(5):457-65<br />

Pasqualucci L, Compagno M, Houldsworth J, Monti S,<br />

Grunn A, N<strong>and</strong>ula SV, Aster JC, Murty VV, Shipp MA,<br />

Dalla-Favera R. Inactivation <strong>of</strong> the PRDM1/BLIMP1 gene<br />

in diffuse large B cell lymphoma. J Exp Med. 2006 Feb<br />

20;203(2):311-7<br />

Tam W, Gomez M, Chadburn A, Lee JW, Chan WC,<br />

Knowles DM. Mutational analysis <strong>of</strong> PRDM1 indicates a<br />

tumor-suppressor role in diffuse large B-cell lymphomas.<br />

Blood. 2006 May 15;107(10):4090-100<br />

Liu YY, Leboeuf C, Shi JY, Li JM, Wang L, Shen Y, Garcia<br />

JF, Shen ZX, Chen Z, Janin A, Chen SJ, Zhao WL.<br />

Rituximab plus CHOP (R-CHOP) overcomes PRDM1associated<br />

resistance to chemotherapy in patients with<br />

diffuse large B-cell lymphoma. Blood. 2007 Jul<br />

1;110(1):339-44<br />

Magnúsdóttir E, Kalachikov S, Mizukoshi K, Savitsky D,<br />

Ishida-Yamamoto A, Panteleyev AA, Calame K. Epidermal<br />

terminal differentiation depends on B lymphocyte-induced<br />

maturation protein-1. Proc Natl Acad Sci U S A. 2007 Sep<br />

18;104(38):14988-93<br />

Robertson EJ, Charatsi I, Joyner CJ, Koonce CH, Morgan<br />

M, Islam A, Paterson C, Lejsek E, Arnold SJ, Kallies A,<br />

Nutt SL, Bik<strong>of</strong>f EK. Blimp1 regulates development <strong>of</strong> the<br />

posterior forelimb, caudal pharyngeal arches, heart <strong>and</strong><br />

sensory vibrissae in mice. Development. 2007<br />

Dec;134(24):4335-45<br />

Tate G, Hirayama-Ohashi Y, Kishimoto K, Mitsuya T.<br />

Novel BLIMP1/PRDM1 gene mutations in B-cell<br />

lymphoma. Cancer <strong>Gene</strong>t Cytogenet. 2007 Jan<br />

15;172(2):151-3<br />

Courts C, Montesinos-Rongen M, Brunn A, Bug S, Siemer<br />

D, Hans V, Blümcke I, Klapper W, Schaller C, Wiestler OD,<br />

Küppers R, Siebert R, Deckert M. Recurrent inactivation <strong>of</strong><br />

the PRDM1 gene in primary central nervous system<br />

lymphoma. J Neuropathol Exp Neurol. 2008 Jul;67(7):720-<br />

7<br />

Martins G, Calame K. Regulation <strong>and</strong> functions <strong>of</strong> Blimp-1<br />

in T <strong>and</strong> B lymphocytes. Annu Rev Immunol. 2008;26:133-<br />

69<br />

Nie K, Gomez M, L<strong>and</strong>graf P, Garcia JF, Liu Y, Tan LH,<br />

Chadburn A, Tuschl T, Knowles DM, Tam W. MicroRNAmediated<br />

down-regulation <strong>of</strong> PRDM1/Blimp-1 in<br />

Hodgkin/Reed-Sternberg cells: a potential pathogenetic<br />

lesion in Hodgkin lymphomas. Am J Pathol. 2008<br />

Jul;173(1):242-52<br />

Schmidt D, Nayak A, Schumann JE, Schimpl A, Berberich<br />

I, Berberich-Siebelt F. Blimp-1Deltaexon7: a naturally<br />

occurring Blimp-1 deletion mutant with auto-regulatory<br />

potential. Exp Cell Res. 2008 Dec 10;314(20):3614-27<br />

von H<strong>of</strong>sten J, Elworthy S, Gilchrist MJ, Smith JC, Wardle<br />

FC, Ingham PW. Prdm1- <strong>and</strong> Sox6-mediated<br />

transcriptional repression specifies muscle fibre type in the<br />

zebrafish embryo. EMBO Rep. 2008 Jul;9(7):683-9<br />

Zhao WL, Liu YY, Zhang QL, Wang L, Leboeuf C, Zhang<br />

YW, Ma J, Garcia JF, Song YP, Li JM, Shen ZX, Chen Z,<br />

Janin A, Chen SJ. PRDM1 is involved in chemoresistance<br />

<strong>of</strong> T-cell lymphoma <strong>and</strong> down-regulated by the proteasome<br />

inhibitor. Blood. 2008 Apr 1;111(7):3867-71<br />

Bik<strong>of</strong>f EK, Morgan MA, Robertson EJ. An exp<strong>and</strong>ing job<br />

description for Blimp-1/PRDM1. Curr Opin <strong>Gene</strong>t Dev.<br />

2009 Aug;19(4):379-85<br />

Chan YH, Chiang MF, Tsai YC, Su ST, Chen MH, Hou<br />

MS, Lin KI. Absence <strong>of</strong> the transcriptional repressor Blimp-<br />

1 in hematopoietic lineages reveals its role in dendritic cell<br />

homeostatic development <strong>and</strong> function. J Immunol. 2009<br />

Dec 1;183(11):7039-46<br />

D'Costa K, Emslie D, Metcalf D, Smyth GK, Karnowski A,<br />

Kallies A, Nutt SL, Corcoran LM. Blimp1 is limiting for<br />

transformation in a mouse plasmacytoma model. Blood.<br />

2009 Jun 4;113(23):5911-9<br />

Iqbal J, Kucuk C, Deleeuw RJ, Srivastava G, Tam W,<br />

Geng H, Klinkebiel D, Christman JK, Patel K, Cao K, Shen<br />

L, Dybkaer K, Tsui IF, Ali H, Shimizu N, Au WY, Lam WL,<br />

Chan WC. Genomic analyses reveal global functional<br />

alterations that promote tumor growth <strong>and</strong> novel tumor<br />

suppressor genes in natural killer-cell malignancies.<br />

Leukemia. 2009 Jun;23(6):1139-51<br />

Malumbres R, Sarosiek KA, Cubedo E, Ruiz JW, Jiang X,<br />

Gascoyne RD, Tibshirani R, Lossos IS. Differentiation<br />

stage-specific expression <strong>of</strong> microRNAs in B lymphocytes<br />

<strong>and</strong> diffuse large B-cell lymphomas. Blood. 2009 Apr<br />

16;113(16):3754-64<br />

Su ST, Ying HY, Chiu YK, Lin FR, Chen MY, Lin KI.<br />

Involvement <strong>of</strong> histone demethylase LSD1 in Blimp-1mediated<br />

gene repression during plasma cell<br />

differentiation. Mol Cell Biol. 2009 Mar;29(6):1421-31<br />

Zhang J, Jima DD, Jacobs C, Fischer R, Gottwein E,<br />

Huang G, Lugar PL, Lagoo AS, Rizzieri DA, Friedman DR,<br />

Weinberg JB, Lipsky PE, Dave SS. Patterns <strong>of</strong> microRNA<br />

expression characterize stages <strong>of</strong> human B-cell<br />

differentiation. Blood. 2009 May 7;113(19):4586-94<br />

Calado DP, Zhang B, Srinivasan L, Sasaki Y, Seagal J,<br />

Unitt C, Rodig S, Kutok J, Tarakhovsky A, Schmidt-<br />

Supprian M, Rajewsky K. Constitutive canonical NF-κB<br />

activation cooperates with disruption <strong>of</strong> BLIMP1 in the<br />

pathogenesis <strong>of</strong> activated B cell-like diffuse large cell<br />

lymphoma. Cancer Cell. 2010 Dec 14;18(6):580-9<br />

Calame K. Blimp-1's maiden flight. J Immunol. 2010 Jul<br />

1;185(1):3-4<br />

Gururajan M, Haga CL, Das S, Leu CM, Hodson D, Josson<br />

S, Turner M, Cooper MD. MicroRNA 125b inhibition <strong>of</strong> B<br />

cell differentiation in germinal centers. Int Immunol. 2010<br />

Jul;22(7):583-92<br />

Hangaishi A, Kurokawa M. Blimp-1 is a tumor suppressor<br />

gene in lymphoid malignancies. Int J Hematol. 2010<br />

Jan;91(1):46-53<br />

M<strong>and</strong>elbaum J, Bhagat G, Tang H, Mo T, Brahmachary M,<br />

Shen Q, Chadburn A, Rajewsky K, Tarakhovsky A,<br />

Pasqualucci L, Dalla-Favera R. BLIMP1 is a tumor<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 137


PRDM1 (PR domain containing 1, with ZNF domain) Tam W<br />

suppressor gene frequently disrupted in activated B celllike<br />

diffuse large B cell lymphoma. Cancer Cell. 2010 Dec<br />

14;18(6):568-79<br />

Montes-Moreno S, Gonzalez-Medina AR, Rodriguez-Pinilla<br />

SM, Maestre L, Sanchez-Verde L, Roncador G, Mollejo M,<br />

García JF, Menarguez J, Montalbán C, Ruiz-Marcellan<br />

MC, Conde E, Piris MA. Aggressive large B-cell lymphoma<br />

with plasma cell differentiation: immunohistochemical<br />

characterization <strong>of</strong> plasmablastic lymphoma <strong>and</strong> diffuse<br />

large B-cell lymphoma with partial plasmablastic<br />

phenotype. Haematologica. 2010 Aug;95(8):1342-9<br />

Nie K, Zhang T, Allawi H, Gomez M, Liu Y, Chadburn A,<br />

Wang YL, Knowles DM, Tam W. Epigenetic downregulation<br />

<strong>of</strong> the tumor suppressor gene PRDM1/Blimp-1<br />

in diffuse large B cell lymphomas: a potential role <strong>of</strong> the<br />

microRNA let-7. Am J Pathol. 2010 Sep;177(3):1470-9<br />

Sellheyer K, Krahl D. Blimp-1: a marker <strong>of</strong> terminal<br />

differentiation but not <strong>of</strong> sebocytic progenitor cells. J Cutan<br />

Pathol. 2010 Mar;37(3):362-70<br />

Smith MA, Maurin M, Cho HI, Becknell B, Freud AG, Yu J,<br />

Wei S, Djeu J, Celis E, Caligiuri MA, Wright KL.<br />

PRDM1/Blimp-1 controls effector cytokine production in<br />

human NK cells. J Immunol. 2010 Nov 15;185(10):6058-67<br />

Best T, Li D, Skol AD, Kirchh<strong>of</strong>f T, Jackson SA, Yasui Y,<br />

Bhatia S, Strong LC, Domchek SM, Nathanson KL,<br />

Olopade OI, Huang RS, Mack TM, Conti DV, Offit K,<br />

Cozen W, Robison LL, Onel K. Variants at 6q21 implicate<br />

PRDM1 in the etiology <strong>of</strong> therapy-induced second<br />

malignancies after Hodgkin's lymphoma. Nat Med. 2011<br />

Jul 24;17(8):941-3<br />

Harper J, Mould A, Andrews RM, Bik<strong>of</strong>f EK, Robertson EJ.<br />

The transcriptional repressor Blimp1/Prdm1 regulates<br />

postnatal reprogramming <strong>of</strong> intestinal enterocytes. Proc<br />

Natl Acad Sci U S A. 2011 Jun 28;108(26):10585-90<br />

Karube K, Nakagawa M, Tsuzuki S, Takeuchi I, Honma K,<br />

Nakashima Y, Shimizu N, Ko YH, Morishima Y, Ohshima<br />

K, Nakamura S, Seto M. Identification <strong>of</strong> FOXO3 <strong>and</strong><br />

PRDM1 as tumor-suppressor gene c<strong>and</strong>idates in NK-cell<br />

neoplasms by genomic <strong>and</strong> functional analyses. Blood.<br />

2011 Sep 22;118(12):3195-204<br />

This article should be referenced as such:<br />

Tam W. PRDM1 (PR domain containing 1, with ZNF<br />

domain). <strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012;<br />

16(2):133-138.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 138


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

Leukaemia <strong>Section</strong><br />

Mini Review<br />

del(11)(q23q23) MLL/ARHGEF12<br />

Jean-Loup Huret<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

<strong>Gene</strong>tics, Dept Medical Information, University <strong>of</strong> Poitiers, CHU Poitiers Hospital, F-86021 Poitiers,<br />

France (JLH)<br />

Published in <strong>Atlas</strong> Database: September 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/Anomalies/del11q23q23ID1420.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI del11q23q23ID1420.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Clinics <strong>and</strong> pathology<br />

Disease<br />

Acute myeloid leukemia (AML)<br />

Epidemiology<br />

Two cases so far: a 38-year-old male patient with a<br />

history <strong>of</strong> occupational exposure to herbicides <strong>and</strong> a<br />

M4-AML, <strong>and</strong> a 77-year-old female patient with a<br />

M5a-AML (Kourlas et al., 2000; Shih el al., 2006).<br />

Evolution<br />

The M4-AML patient underwent complete<br />

remission, but died 6 months later from an<br />

unrelated cause.<br />

<strong>Cytogenetics</strong><br />

<strong>Cytogenetics</strong> morphological<br />

The del(11)(q23q23) has been missed by<br />

cytogenetic analysis; both cases were hyperploid<br />

with +8 <strong>and</strong> other abnormalities.<br />

<strong>Gene</strong>s involved <strong>and</strong><br />

proteins<br />

MLL<br />

Location<br />

11q23.3<br />

Protein<br />

A major transcript <strong>of</strong> 14982 bp produces a 3969<br />

amino acids protein from 36 <strong>of</strong> the 37 exons.<br />

Contains from N-term to C-term a binding site for<br />

MEN1, 3 AT hooks (binds to the minor grove <strong>of</strong><br />

DNA); 2 speckled nuclear localisation signals; 2<br />

repression domains RD1 <strong>and</strong> RD2: RD1 or CXXC:<br />

cystein methyl transferase, binds CpG rich DNA,<br />

has a transcriptional repression activity; RD2<br />

recruits histone desacetylases HDAC1 <strong>and</strong><br />

HDAC2; 3 plant homeodomains (cystein rich zinc<br />

finger domains, with homodimerization properties),<br />

1 bromodomain (may bind acetylated histones), <strong>and</strong><br />

1 plant homeodomain; these domains may be<br />

involved in protein-protein interaction; a FYRN <strong>and</strong><br />

a FRYC domain; a transactivation domain which<br />

binds CBP; may acetylates H3 <strong>and</strong> H4 in the HOX<br />

area; a SET domain: methyltransferase; methylates<br />

H3, including histones in the HOX area for<br />

allowing chromatin to be open to transcription.<br />

MLL is cleaved by taspase 1 into 2 proteins before<br />

entering the nucleus: a p300/320 N-term protein<br />

called MLL-N, <strong>and</strong> a p180 C-term protein, called<br />

MLL-C. The FYRN <strong>and</strong> a FRYC domains <strong>of</strong> native<br />

MLL associate MLL-N <strong>and</strong> MLL-C in a stable<br />

complex; they form a multiprotein complex with<br />

transcription factor TFIID. <strong>Gene</strong>ral transcription<br />

factor; maintains HOX genes expression in<br />

undifferentiated cells. Major regulator <strong>of</strong><br />

hematopoiesis <strong>and</strong> embryonic development; role in<br />

cell cycle regulation.<br />

ARHGEF12<br />

Location<br />

11q23.3<br />

Protein<br />

Better known as LARG, ARHGEF12 contains a<br />

PDZ (postsynaptic density protein, Drosophila disc<br />

large tumor suppressor, <strong>and</strong> zonula occludens-1<br />

protein) domain, which localize ARHGEF12 to the<br />

membrane, a regulator <strong>of</strong> G protein signalling-like<br />

domain (RGSL or RH), which binds to activated<br />

heterotrimeric G protein alpha12/13 subunits, a Dbl<br />

homology (DH) domain, responsible for exchange<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 139


del(11)(q23q23) MLL/ARHGEF12 Huret JL<br />

activity, <strong>and</strong> a pleckstrin homology (PH) domain,<br />

involved in the regulation <strong>of</strong> the process.<br />

Regulatory protein involved in the GDP/GTP<br />

exchange reaction <strong>of</strong> the Rho proteins; activates a<br />

Rho-GTPase-dependent signaling pathway;<br />

activated by FYN.<br />

Result <strong>of</strong> the chromosomal<br />

anomaly<br />

Hybrid gene<br />

Description<br />

5' MLL - 3' ARHGEF12<br />

Fusion protein<br />

Description<br />

Joins amino acid 1362 from MLL to amino acid<br />

309 from ARHGEF12. The fusion protein<br />

comprises the Ala/Gly/Ser-rich region, poly-Gly<br />

stretch, three AT hooks domains, poly-Pro<br />

stretches, <strong>and</strong> Zinc finger CXXC-type domain from<br />

MLL fused to the RGSL, DH, PH domains <strong>of</strong><br />

ARHGEF12.<br />

References<br />

Kourlas PJ, Strout MP, Becknell B, Veronese ML, Croce<br />

CM, Theil KS, Krahe R, Ruutu T, Knuutila S, Bloomfield<br />

CD, Caligiuri MA. Identification <strong>of</strong> a gene at 11q23<br />

encoding a guanine nucleotide exchange factor: evidence<br />

for its fusion with MLL in acute myeloid leukemia. Proc Natl<br />

Acad Sci U S A. 2000 Feb 29;97(5):2145-50<br />

Shih LY, Liang DC, Fu JF, Wu JH, Wang PN, Lin TL, Dunn<br />

P, Kuo MC, Tang TC, Lin TH, Lai CL. Characterization <strong>of</strong><br />

fusion partner genes in 114 patients with de novo acute<br />

myeloid leukemia <strong>and</strong> MLL rearrangement. Leukemia.<br />

2006 Feb;20(2):218-23<br />

This article should be referenced as such:<br />

Huret JL. del(11)(q23q23) MLL/ARHGEF12. <strong>Atlas</strong> <strong>Gene</strong>t<br />

Cytogenet Oncol Haematol. 2012; 16(2):139-140.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 140


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

Leukaemia <strong>Section</strong><br />

Mini Review<br />

t(3;11)(q21;q23)<br />

Jean-Loup Huret<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

<strong>Gene</strong>tics, Dept Medical Information, University <strong>of</strong> Poitiers, CHU Poitiers Hospital, F-86021 Poitiers,<br />

France (JLH)<br />

Published in <strong>Atlas</strong> Database: September 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/Anomalies/t0311q21q23ID1407.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI t0311q21q23ID1407.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Clinics <strong>and</strong> pathology<br />

Disease<br />

Acute leukemia<br />

Epidemiology<br />

Only two cases <strong>of</strong> acute leukemia to date (2 female<br />

patients); in one case, the phenotype was described:<br />

it was a case <strong>of</strong> biphenotypic acute leukemia<br />

(BAL), a B-ALL expressing CD13 (Hanson et al.,<br />

1993).<br />

Clinics<br />

No data.<br />

<strong>Cytogenetics</strong><br />

<strong>Cytogenetics</strong> morphological<br />

A complex karyotype was found in both cases. In<br />

one case, a t(9;22)(q34;q11) was also present.<br />

<strong>Gene</strong>s involved <strong>and</strong><br />

proteins<br />

Note<br />

The genes involved in the translocation were<br />

determined in one case (Meyer et al., 2005).<br />

EEFSEC<br />

Protein<br />

EEFSEC, <strong>of</strong>ten called SELB, is the elongation<br />

factor for delivery <strong>of</strong> selenocysteinyl-tRNA to the<br />

ribosome. Selenocysteine (Sec) is found in the<br />

active sites <strong>of</strong> enzymes, most <strong>of</strong> which are involved<br />

in redox reactions (Paleskava et al., 2010).<br />

MLL<br />

Location<br />

11q23<br />

Protein<br />

A major transcript <strong>of</strong> 14982 bp produces a 3969<br />

amino acids protein from 36 <strong>of</strong> the 37 exons.<br />

Contains from N-term to C-term a binding site for<br />

MEN1, 3 AT hooks (binds to the minor grove <strong>of</strong><br />

DNA); 2 speckled nuclear localisation signals; 2<br />

repression domains RD1 <strong>and</strong> RD2: RD1 or CXXC:<br />

cystein methyl transferase, binds CpG rich DNA,<br />

has a transcriptional repression activity; RD2<br />

recruits histone desacetylases HDAC1 <strong>and</strong><br />

HDAC2; 3 plant homeodomains (cystein rich zinc<br />

finger domains, with homodimerization properties),<br />

1 bromodomain (may bind acetylated histones), <strong>and</strong><br />

1 plant homeodomain; these domains may be<br />

involved in protein-protein interaction; a FYRN <strong>and</strong><br />

a FRYC domain; a transactivation domain which<br />

binds CBP; may acetylates H3 <strong>and</strong> H4 in the HOX<br />

area; a SET domain: methyltransferase; methylates<br />

H3, including histones in the HOX area for<br />

allowing chromatin to be open to transcription.<br />

MLL is cleaved by taspase 1 into 2 proteins before<br />

entering the nucleus: a p300/320 N-term protein<br />

called MLL-N, <strong>and</strong> a p180 C-term protein, called<br />

MLL-C. The FYRN <strong>and</strong> a FRYC domains <strong>of</strong> native<br />

MLL associate MLL-N <strong>and</strong> MLL-C in a stable<br />

complex; they form a multiprotein complex with<br />

transcription factor TFIID. <strong>Gene</strong>ral transcription<br />

factor; maintains HOX genes expression in<br />

undifferentiated cells. Major regulator <strong>of</strong><br />

hematopoiesis <strong>and</strong> embryonic development; role in<br />

cell cycle regulation.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 141


t(3;11)(q21;q23) Huret JL<br />

Result <strong>of</strong> the chromosomal<br />

anomaly<br />

Hybrid gene<br />

Description<br />

Fusion between MLL intron 9 <strong>and</strong> EEFSEC intron<br />

1, but with no maintenance <strong>of</strong> an open reading<br />

frame. This translocation seems to create<br />

nonfunctional fusion genes.<br />

References<br />

Hanson CA, Abaza M, Sheldon S, Ross CW, Schnitzer B,<br />

Stoolman LM.. Acute biphenotypic leukaemia:<br />

immunophenotypic <strong>and</strong> cytogenetic analysis. Br J<br />

Haematol. 1993 May;84(1):49-60.<br />

Meyer C, Schneider B, Reichel M, Angermueller S, Strehl<br />

S, Schnittger S, Schoch C, Jansen MW, van Dongen JJ,<br />

Pieters R, Haas OA, Dingermann T, Klingebiel T,<br />

Marschalek R.. Diagnostic tool for the identification <strong>of</strong> MLL<br />

rearrangements including unknown partner genes. Proc<br />

Natl Acad Sci U S A. 2005 Jan 11;102(2):449-54. Epub<br />

2004 Dec 30.<br />

Paleskava A, Konevega AL, Rodnina MV..<br />

Thermodynamic <strong>and</strong> kinetic framework <strong>of</strong> selenocysteyltRNASec<br />

recognition by elongation factor SelB. J Biol<br />

Chem. 2010 Jan 29;285(5):3014-20. Epub 2009 Nov 23.<br />

This article should be referenced as such:<br />

Huret JL. t(3;11)(q21;q23). <strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol<br />

Haematol. 2012; 16(2):141-142.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 142


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

Solid Tumour <strong>Section</strong><br />

Review<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

Head <strong>and</strong> Neck: Squamous cell carcinoma: an<br />

overview<br />

Audrey Rousseau, Cécile Badoual<br />

Universite d'Angers, Departement de Pathologie Cellulaire et Tissulaire, CHU Angers, 4 rue Larrey,<br />

49100 Angers, France (AR), Universite Rene Descartes Paris 5, Service d'Anatomie Pathologique,<br />

Hopital Europeen Georges Pompidou, 20 rue Leblanc, 75015 Paris, France (CB)<br />

Published in <strong>Atlas</strong> Database: September 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/Tumors/HeadNeckSCCID5090.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI HeadNeckSCCID5090.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Identity<br />

Note<br />

Head <strong>and</strong> neck squamous cell carcinoma (HNSCC)<br />

develops from the mucosal linings <strong>of</strong> the upper<br />

aerodigestive tract, comprising 1) the nasal cavity<br />

<strong>and</strong> paranasal sinuses, 2) the nasopharynx, 3) the<br />

hypopharynx, larynx, <strong>and</strong> trachea, <strong>and</strong> 4) the oral<br />

cavity <strong>and</strong> oropharynx. Squamous cell carcinoma<br />

(SCC) is the most frequent malignant tumor <strong>of</strong> the<br />

head <strong>and</strong> neck region. HNSCC is the sixth leading<br />

cancer by incidence worldwide. There are 500000<br />

new cases a year worldwide. Two thirds occur in<br />

industrialized nations. HNSCC usually develops in<br />

males in the 6 th <strong>and</strong> 7 th decade. It is caused by<br />

tobacco <strong>and</strong> alcohol consumption <strong>and</strong> infection<br />

with high-risk types <strong>of</strong> human papillomavirus<br />

(HPV). SCC <strong>of</strong>ten develops from preexisting<br />

dysplastic lesions. The five-year survival rate <strong>of</strong><br />

patients with HNSCC is about 40-50%.<br />

Classification<br />

SCC can occur either in 1) the nasal cavity <strong>and</strong><br />

paranasal sinuses, 2) the nasopharynx, 3) the<br />

hypopharynx, larynx, <strong>and</strong> trachea, or 4) the oral<br />

cavity <strong>and</strong> oropharynx. The 2005 World Health<br />

Organization (WHO) classification <strong>of</strong> Head <strong>and</strong><br />

Neck Tumors (Barnes et al., 2005) distinguishes<br />

different types <strong>of</strong> SCC:<br />

- Conventional<br />

- Verrucous<br />

- Basaloid<br />

- Papillary<br />

- Spindle cell (sarcomatoid)<br />

- Acantholytic<br />

- Adenosquamous<br />

- Cuniculatum<br />

Each variant can arise in any one <strong>of</strong> the 4 above<br />

mentioned head <strong>and</strong> neck regions, except for the<br />

cuniculatum type which only develops from the<br />

oral mucosa.<br />

SCC can be well-, moderately- or poorlydifferentiated,<br />

<strong>and</strong> either keratinizing or nonkeratinizing.<br />

Most cases are moderately to poorlydifferentiated.<br />

Precursor lesions (dysplasia) can be (arbitrarily)<br />

separated into mild, moderate, or severe (carcinoma<br />

in situ) (see below).<br />

Clinics <strong>and</strong> pathology<br />

Etiology<br />

The most important risk factors for developing<br />

HNSCC are tobacco smoking <strong>and</strong> alcohol<br />

consumption, which have a synergistic effect.<br />

Smoking habits that increase the risk <strong>of</strong> developing<br />

HNSCC are smoking black tobacco (compared to<br />

blond tobacco), smoking at a young age, long<br />

duration, high number <strong>of</strong> cigarettes per day, <strong>and</strong><br />

deep smoke inhalation (Benhamou et al., 1992).<br />

Avoiding cigarettes <strong>and</strong> alcohol could prevent<br />

about 90% <strong>of</strong> HNSCCs, especially laryngeal <strong>and</strong><br />

hypopharyngeal tumors. Tobacco chewing is a<br />

major cause <strong>of</strong> oral <strong>and</strong> oropharyngeal SCC in the<br />

Indian subcontinent, parts <strong>of</strong> South-East Asia,<br />

China <strong>and</strong> Taiwan, especially when consumed in<br />

betel quids containing areca nut (Znaor et al.,<br />

2003). In India, chewing accounts for nearly 50%<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 143


Head <strong>and</strong> Neck: Squamous cell carcinoma: an overview Rousseau A, Badoual C<br />

<strong>of</strong> oral <strong>and</strong> oropharyngeal tumors in men <strong>and</strong> over<br />

90% in women (Barnes et al., 2005). Significant<br />

risk increases <strong>of</strong> developing HNSCC have also<br />

been reported among non-drinking smokers <strong>and</strong>, to<br />

a lesser extent, non-smoking heavy drinkers. Heavy<br />

consumption <strong>of</strong> all types <strong>of</strong> alcoholic beverages<br />

(wine, beer, hard liquors) confers an increased risk<br />

(La Vecchia et al., 1999). Conversely, protective<br />

effects <strong>of</strong> diets rich in fresh fruits <strong>and</strong> vegetables<br />

have been described (Pelucchi et al., 2003).<br />

Some occupational exposures may be associated<br />

with a higher risk <strong>of</strong> developing HNSCC, especially<br />

<strong>of</strong> the larynx: polycyclic aromatic hydrocarbons,<br />

metal dust, cement dust, varnish, lacquer, etc...<br />

Significant associations were also found with<br />

ionizing radiation, diesel exhausts, sulphuric acid<br />

mists, <strong>and</strong> mustard gas (Barnes et al., 2005).<br />

The incidence <strong>of</strong> HNSCC in specific sites has been<br />

slowly declining during the past decade, due to a<br />

decrease in the prevalence <strong>of</strong> the more traditional<br />

risk factors, most notably smoking. However, in the<br />

Western World, HNSCCs, notably <strong>of</strong> the oral<br />

cavity <strong>and</strong> oropharynx, are becoming more<br />

prevalent, which may be related to an increase in<br />

oral <strong>and</strong> oropharyngeal HPV infections (Leeman et<br />

al., 2011). Indeed, recent studies have shown that<br />

infection with high-risk types <strong>of</strong> HPV (e.g. HPV-16<br />

<strong>and</strong> -18) is responsible for a subgroup <strong>of</strong> HNSCCs.<br />

HPV-positive tumors represent a different<br />

clinicopathological <strong>and</strong> molecular entity compared<br />

to HPV-negative cases (see below). HPV infection<br />

is now recognized as one <strong>of</strong> the primary causes <strong>of</strong><br />

oropharyngeal SCC (especially SCC <strong>of</strong> the tonsils<br />

<strong>and</strong> the base <strong>of</strong> the tongue). In the USA, about 40-<br />

80% <strong>of</strong> oropharyngeal cancers are caused by HPV,<br />

whereas in Europe the proportion varies from<br />

around 90% in Sweden to less than 20% in<br />

communities with the highest tobacco use (Marur et<br />

al., 2010). Patients tend to be younger, with no<br />

prior history <strong>of</strong> tobacco <strong>and</strong>/or heavy alcohol<br />

consumption. There is evidence that HPV-positive<br />

HNSCC is a sexually transmitted disease. A strong<br />

association between sexual behavior (oral sex) <strong>and</strong><br />

risk <strong>of</strong> oropharyngeal cancer as well as HPV-16positive<br />

HNSCC has been demonstrated (Smith et<br />

al., 2004; Gillison et al., 2008).<br />

Finally, certain inherited disorders, such as Fanconi<br />

anemia or Bloom syndrome, predispose to HNSCC<br />

(Kutler et al., 2003; Barnes et al., 2005).<br />

Epidemiology<br />

SCC is the most frequent malignant tumor <strong>of</strong> the<br />

head <strong>and</strong> neck region. HNSCC represents the sixth<br />

leading cancer by incidence <strong>and</strong> there are 500000<br />

new cases a year worldwide (Kamangar et al.,<br />

2006). Two thirds occur in industrialized nations.<br />

Most HNSCCs arise in the hypopharynx, larynx,<br />

<strong>and</strong> trachea, <strong>and</strong> in the oral cavity <strong>and</strong> oropharynx.<br />

The majority <strong>of</strong> laryngeal SCCs originate from the<br />

supraglottic <strong>and</strong> glottic regions. Tracheal SCCs are<br />

rare compared to laryngeal ones. The most common<br />

oropharyngeal site <strong>of</strong> involvement is the base <strong>of</strong> the<br />

tongue. Within the oral cavity, most tumors arise<br />

from the floor <strong>of</strong> the mouth, the ventrolateral<br />

tongue or the s<strong>of</strong>t palate complex.<br />

HNSCCs occur most frequently in the sixth <strong>and</strong><br />

seventh decades. They typically develop in men<br />

though women are more <strong>and</strong> more affected because<br />

<strong>of</strong> increased prevalence <strong>of</strong> smoking over the last<br />

two decades (Barnes et al., 2005). For laryngeal,<br />

hypopharyngeal <strong>and</strong> tracheal SCCs, the incidence<br />

in men is high in Southern <strong>and</strong> Central Europe,<br />

some parts <strong>of</strong> South America, <strong>and</strong> among Blacks in<br />

the United States. The lowest rates are recorded in<br />

South-East Asia <strong>and</strong> Central Africa. The disease is<br />

slightly more common in urban than in rural areas.<br />

For oral <strong>and</strong> oropharyngeal SCCs, the disease<br />

usually affects adults in the 5 th <strong>and</strong> 6 th decades <strong>of</strong><br />

life. Extremely elevated rates are observed in<br />

France, parts <strong>of</strong> Switzerl<strong>and</strong>, Northern Italy, Central<br />

<strong>and</strong> Eastern Europe, <strong>and</strong> parts <strong>of</strong> Latin America.<br />

Rates are high among both men <strong>and</strong> women<br />

throughout South Asia. In the US, incidence rates<br />

are two-fold higher in Blacks compared to Whites<br />

(Barnes et al., 2005).<br />

Clinics<br />

Clinical features <strong>of</strong> HNSCC depend on the<br />

localization <strong>of</strong> the tumor.<br />

Nasal <strong>and</strong> paranasal sinuses<br />

Patients with SCC arising in the nasal or paranasal<br />

sinuses may complain <strong>of</strong> nasal fullness, stuffiness,<br />

or obstruction, but also <strong>of</strong> epistaxis, rhinorrhea,<br />

pain, paraesthesia, swelling <strong>of</strong> the nose <strong>and</strong> cheek<br />

or <strong>of</strong> a palatal bulge. Some may present with a<br />

persistent or non-healing nasal sore or ulcer, a nasal<br />

mass, or in advanced cases, proptosis, diplopia, or<br />

lacrimation (Barnes et al., 2005; Thompson, 2006).<br />

Nasopharynx<br />

Most patients with nasopharyngeal carcinoma<br />

present with painless enlargement <strong>of</strong> upper cervical<br />

lymph nodes. Nasal symptoms, particularly bloodstained<br />

post-nasal drip are reported in half the<br />

cases. Serous otitis media following Eustachian<br />

tube obstruction is also common. Headaches <strong>and</strong><br />

cranial nerve involvement indicate more advanced<br />

disease. However, 10% <strong>of</strong> the patients are<br />

asymptomatic (Barnes et al., 2005; Thompson,<br />

2006).<br />

Hypopharynx, larynx, <strong>and</strong> trachea<br />

Hypopharyngeal <strong>and</strong> supraglottic tumors may be<br />

responsible <strong>of</strong> dysphagia, change in quality <strong>of</strong><br />

voice, foreign body sensation in the throat,<br />

haemoptysis, <strong>and</strong> odynophagia. Glottic SCC most<br />

commonly presents with hoarseness (Fig. 1). In<br />

case <strong>of</strong> subglottic tumor, dyspnea <strong>and</strong> stridor are<br />

frequent clinical features. SCC arising in the<br />

trachea may cause dyspnea, wheezing or stridor,<br />

acute respiratory failure, cough, haemoptysis, <strong>and</strong><br />

hoarseness (Barnes et al., 2005; Thompson, 2006).<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 144


Head <strong>and</strong> Neck: Squamous cell carcinoma: an overview Rousseau A, Badoual C<br />

Figure 1: Endoscopy. Exophytic <strong>and</strong> ulcerative SCC <strong>of</strong> the left vocal cord. Figure 2: Large exophytic <strong>and</strong> ulcerative SCC <strong>of</strong> the<br />

left amygdala. Fig. 1-2 were kindly provided by Dr S. Hans (Georges Pompidou European Hospital, Paris, France).<br />

Oral cavity <strong>and</strong> oropharynx<br />

Most patients display at the time <strong>of</strong> diagnosis signs<br />

<strong>and</strong> symptoms <strong>of</strong> locally advanced disease. Clinical<br />

features vary according to the exact site <strong>of</strong> the<br />

lesion. The most common presenting features are<br />

ulceration, pain, referred pain to the ear, difficulty<br />

with speaking, opening the mouth or chewing,<br />

difficulty <strong>and</strong> pain with swallowing, bleeding,<br />

weight loss, <strong>and</strong> neck swelling (Fig. 2). Cancer <strong>of</strong><br />

the buccal mucosa may present as an ulcer with<br />

indurated raised margins or as an exophytic growth.<br />

SCC <strong>of</strong> the floor <strong>of</strong> the mouth may arise as a red or<br />

ulcerated lesion or as a papillary growth. Cancer <strong>of</strong><br />

the gingiva usually presents as an<br />

ulceroproliferative growth. Cancer <strong>of</strong> the tongue<br />

may appear as an ulcer infiltrating deeply <strong>and</strong><br />

reducing the mobility <strong>of</strong> the tongue. SCC <strong>of</strong> the<br />

base <strong>of</strong> the tongue usually presents at a locally<br />

advanced stage as an ulcerated, painful, indurated<br />

growth. Cancer <strong>of</strong> the hard palate <strong>of</strong>ten presents as<br />

a papillary or exophytic growth rather than a flat or<br />

ulcerated lesion. Cancer <strong>of</strong> s<strong>of</strong>t palate <strong>and</strong> uvula<br />

<strong>of</strong>ten appears as an ulcerative lesion with raised<br />

margins or as a fungating mass. Occasionally,<br />

patients harbor enlarged cervical lymph nodes with<br />

no identifiable oral or oropharyngeal lesion. In very<br />

advanced disease, patients may present an<br />

ulceroproliferative lesion with areas <strong>of</strong> necrosis <strong>and</strong><br />

extension to surrounding structures, such bone,<br />

muscle <strong>and</strong> skin (Barnes et al., 2005; Thompson,<br />

2006).<br />

Pathology<br />

Precursor lesions<br />

The 2005 WHO classification <strong>of</strong> precursor lesions<br />

(Barnes et al., 2005) is as follows:<br />

- Squamous cell hyperplasia<br />

Hyperplasia describes increased cell numbers. This<br />

may be in the spinous layer (acanthosis) <strong>and</strong>/or in<br />

the basal/parabasal cell layers (progenitor<br />

compartment), termed basal cell hyperplasia. The<br />

architecture shows regular stratification without<br />

cellular atypia.<br />

- Mild dysplasia (Squamous Intraepithelial<br />

Neoplasia (SIN) 1)<br />

- Moderate dysplasia (SIN 2)<br />

- Severe dysplasia (SIN 3)<br />

- Carcinoma in situ (SIN 3)<br />

The Ljubljana classification <strong>of</strong> squamous<br />

intraepithelial lesions has also been proposed (see<br />

below) (Barnes et al., 2005).<br />

2005<br />

WHO<br />

Classification<br />

Squamous<br />

cell<br />

hyperplasia<br />

Mild<br />

dysplasia<br />

Moderate<br />

dysplasia<br />

Severe<br />

dysplasia<br />

Carcinoma in<br />

situ<br />

Squamous<br />

Intraepithelial<br />

Neoplasia<br />

(SIN)<br />

SIN 1<br />

SIN 2<br />

SIN 3<br />

Ljubljana<br />

Classification<br />

Squamous<br />

Intraepithelial<br />

Lesions (SIL)<br />

Squamous cell<br />

(simple)<br />

hyperplasia<br />

Basal/Parabasal<br />

cell hyperplasia<br />

Atypical<br />

hyperplasia<br />

Atypical<br />

hyperplasia<br />

SIN 3 Carcinoma in situ<br />

Invasive HNSCCs arise in most cases from<br />

preneoplastic lesions grouped under the term<br />

"dysplasia". Dysplastic lesions present with an<br />

increased likelihood <strong>of</strong> progressing to SCC. The<br />

altered epithelium displays architectural <strong>and</strong><br />

cytological changes that range from mild to severe<br />

(see below). Precursor lesions are mostly seen in<br />

the adult population <strong>and</strong> affect men more <strong>of</strong>ten than<br />

women. They are strongly associated with tobacco<br />

smoking <strong>and</strong> alcohol consumption, <strong>and</strong> especially a<br />

combination <strong>of</strong> the two. The duration <strong>of</strong> smoking,<br />

the type <strong>of</strong> tobacco <strong>and</strong> the practice <strong>of</strong> deep<br />

inhalation play a role in the development <strong>of</strong><br />

precursor lesions. Other etiological factors have<br />

been reported such as industrial pollution, specific<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 145


Head <strong>and</strong> Neck: Squamous cell carcinoma: an overview Rousseau A, Badoual C<br />

occupational exposures, <strong>and</strong> nutritional deficiency.<br />

Mean age <strong>of</strong> patients with first diagnosis <strong>of</strong><br />

precursor lesions is 48-56.5 years. Rarely,<br />

malignant transformation may develop from<br />

morphologically normal epithelium (Barnes et al.,<br />

2005; Thompson, 2006).<br />

The clinical picture <strong>of</strong> precursor lesions depends on<br />

the location <strong>and</strong> severity <strong>of</strong> the disease. In case <strong>of</strong><br />

dysplastic lesions <strong>of</strong> the hypopharynx, larynx or<br />

trachea, patients may present with fluctuating<br />

hoarseness, sore throat, <strong>and</strong>/or chronic cough <strong>of</strong> a<br />

few months' duration. However, some individuals<br />

may be asymptomatic.<br />

Endoscopically, the lesions may be discrete or<br />

diffuse, smooth or irregular, flat or exophytic.<br />

Precursor lesions may present as a small flat patch<br />

or as a large warty plaque. The surface may be<br />

brown to red (erythroplakia) or present with<br />

circumscribed whitish plaques (leukoplakia). White<br />

patches may be ulcerated. Leukoplakia, in contrast<br />

to erythroplakia, tends to be well demarcated <strong>and</strong><br />

seems to have a lower risk <strong>of</strong> malignant<br />

transformation. The lesions are commonly diffuse,<br />

with a thickened appearance. However, in a<br />

minority <strong>of</strong> cases, patchy atrophy may be present.<br />

In the larynx, the precursor lesions appear mainly<br />

along the anterior true vocal cords. Two thirds <strong>of</strong><br />

vocal cord lesions are bilateral. Occasionally,<br />

precursor lesions may present macroscopically as<br />

normal mucosa (Barnes et al., 2005; Thompson,<br />

2006).<br />

Microscopically, dysplasia is defined as<br />

architectural <strong>and</strong> cytological changes <strong>of</strong> the<br />

epithelium, without evidence <strong>of</strong> invasion. The<br />

diagnostic features <strong>of</strong> dysplasia are not uniformly<br />

accepted or interpreted. In dysplastic lesions, the<br />

epithelium presents with irregular stratification, loss<br />

<strong>of</strong> polarity <strong>of</strong> basal cells, drop-shaped rete ridges,<br />

increased number <strong>of</strong> mitotic figures, abnormal<br />

superficial mitoses, premature keratinization in<br />

single cells (dyskeratosis) <strong>and</strong> keratin pearls within<br />

rete pegs. Cytological changes include abnormal<br />

variation in nuclear or cell size <strong>and</strong> shape, increased<br />

nuclear to cytoplasmic ratio, increased nuclear size,<br />

atypical mitotic figures, increased number <strong>and</strong> size<br />

<strong>of</strong> nucleoli, <strong>and</strong> hyperchromasia. The spectrum <strong>of</strong><br />

dysplasia is divided for practical reasons into mild,<br />

moderate <strong>and</strong> severe. In mild dysplasia,<br />

architectural disturbances <strong>and</strong> cytological atypia are<br />

limited to the lower third <strong>of</strong> the epithelium. In<br />

moderate dysplasia, architectural <strong>and</strong> cytological<br />

changes extend into the middle third <strong>of</strong> the<br />

epithelium. Up-grading from moderate to severe<br />

dysplasia may be considered when there is marked<br />

cytological atypia. Severe dysplasia displays greater<br />

than two thirds altered epithelium. Carcinoma in<br />

situ presents with full thickness or almost full<br />

thickness architectural abnormalities accompanied<br />

by cytological atypia (Fig.3). Superficial <strong>and</strong><br />

atypical mitotic figures are commonly seen. By<br />

definition, invasion has not yet occurred.<br />

Differential diagnosis <strong>of</strong> dysplastic lesions includes<br />

reactive or regenerative squamous epithelium<br />

(Barnes et al., 2005; Thompson, 2006).<br />

Conventional type squamous cell carcinoma<br />

SCC is characterized by squamous differentiation<br />

(<strong>of</strong>ten seen as keratinization, sometimes with<br />

keratin pearl formation) <strong>and</strong> invasive growth with<br />

disruption <strong>of</strong> the basement membrane. Extension<br />

into the underlying tissue is <strong>of</strong>ten accompanied by a<br />

desmoplastic stromal reaction <strong>and</strong> a dense<br />

inflammatory infiltrate, mainly comprised <strong>of</strong><br />

lymphocytes <strong>and</strong> plasma cells. Angiolymphatic <strong>and</strong><br />

perineural invasion may be seen. SCC is graded<br />

into well-, moderately-, <strong>and</strong> poorly-differentiated.<br />

Well-differentiated SCC closely resembles normal<br />

squamous mucosa whereas moderatelydifferentiated<br />

SCC displays nuclear pleomorphism,<br />

mitoses (including atypical forms), <strong>and</strong> usually less<br />

keratinization (Fig. 4-6). In poorly-differentiated<br />

SCC, immature cells predominate, with numerous<br />

typical <strong>and</strong> atypical mitoses, minimal<br />

keratinization, <strong>and</strong> sometimes necrosis. Most SCCs<br />

are moderately-differentiated (Barnes et al., 2005;<br />

Thompson, 2006).<br />

HNSCCs express epithelial markers such as<br />

cytokeratins. In well-differentiated tumors, no<br />

additional stains are usually needed. In poorlydifferentiated<br />

lesions, immunohistochemistry may<br />

be useful. HNSCCs are immunopositive for<br />

cytokeratin cocktails, AE1/AE3 <strong>and</strong> pancytokeratin.<br />

CK5/CK6 <strong>and</strong> p63 are also excellent markers to<br />

detect squamous differentiation (Dabbs, 2006).<br />

Verrucous carcinoma<br />

Verrucous carcinoma (VC) is a non-metastasizing<br />

variant <strong>of</strong> well-differentiated SCC characterized by<br />

an exophytic, warty, slowly-growing tumor with<br />

pushing rather than infiltrative margins (Barnes et<br />

al., 2005). The larynx is the second most common<br />

site <strong>of</strong> VC in the head <strong>and</strong> neck region after the oral<br />

cavity. The gross appearance is <strong>of</strong> a broad-based,<br />

exophytic, warty, firm to hard, white mass.<br />

Microscopically, the thickened, club-shaped,<br />

projections are lined by well-differentiated<br />

squamous epithelium devoid <strong>of</strong> the malignant<br />

features commonly seen in SCC. Mitotic figures are<br />

rare <strong>and</strong> not atypical. The advancing margins are<br />

usually broad with a pushing appearance. A dense<br />

inflammatory response is <strong>of</strong>ten present in the<br />

underlying tissue. There is abundant surface<br />

keratosis ("church-spire" keratosis) (Fig. 7).<br />

Differential diagnosis includes verrucous<br />

hyperplasia <strong>and</strong> very well-differentiated SCC.<br />

Distinguishing these entities from verrucous<br />

carcinoma can be delicate. Analysis <strong>of</strong> a sample <strong>of</strong><br />

sufficient size which has been accurately oriented is<br />

necessary before rendering a definitive diagnosis.<br />

The separation <strong>of</strong> verrucous hyperplasia from<br />

verrucous carcinoma is <strong>of</strong>ten difficult, requiring<br />

clinical-pathological confrontation. Pure VC does<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 146


Head <strong>and</strong> Neck: Squamous cell carcinoma: an overview Rousseau A, Badoual C<br />

not metastasize <strong>and</strong> has an excellent prognosis.<br />

However, hybrid VC (displaying a conventional<br />

SCC component) has the potential to metastasize<br />

<strong>and</strong> should be managed as similarly staged SCC<br />

(Barnes et al., 2005; Thompson, 2006).<br />

Basaloid squamous cell carcinoma<br />

Basaloid squamous cell carcinoma is a high-grade<br />

variant <strong>of</strong> SCC composed <strong>of</strong> both basaloid <strong>and</strong><br />

squamous components (Barnes et al., 2005). It is an<br />

aggressive, rapidly growing tumor characterized by<br />

an advanced stage at the time <strong>of</strong> diagnosis (cervical<br />

lymph node metastases) <strong>and</strong> a poor prognosis. The<br />

basaloid component is comprised <strong>of</strong> small packed<br />

cells displaying hyperchromatic nuclei without<br />

nucleoli, <strong>and</strong> scant cytoplasm. The tumor grows in<br />

a solid pattern with a lobular configuration, <strong>and</strong><br />

sometimes a prominent peripheral palisading.<br />

Comedo-type necrosis is frequently seen (Fig. 8).<br />

Small cystic spaces containing PAS- <strong>and</strong> Alcian<br />

Blue-positive material <strong>and</strong> stromal hyalinization<br />

may be noticed. BSCC is always associated with a<br />

SCC component, usually located superficially. The<br />

SCC component may also present as focal<br />

squamous differentiation within the basaloid<br />

lobules. The junction between the two components<br />

may be abrupt. The differential diagnosis includes<br />

neuroendocrine carcinoma, adenoid cystic<br />

carcinoma, <strong>and</strong> adenosquamous carcinoma. BSCC<br />

requires aggressive multimodality treatment,<br />

including radical surgery (including neck<br />

dissection), radiotherapy, <strong>and</strong> chemotherapy<br />

(especially for metastatic disease). Survival rate is<br />

only 40% (Thompson, 2006).<br />

Papillary squamous cell carcinoma<br />

Papillary squamous cell carcinoma (PSCC) is a<br />

distinct variant <strong>of</strong> SCC characterized by an<br />

exophytic, papillary growth, <strong>and</strong> a favorable<br />

prognosis. PSCC presents as a s<strong>of</strong>t, friable,<br />

polypoid, exophytic, papillary tumor. It frequently<br />

arises from a thin stalk, but broad-based lesions<br />

have also been described. The tumor is<br />

characterized by a predominant papillary growth<br />

pattern. By definition, the lesion must demonstrate<br />

a dominant (> 70%) exophytic or papillary<br />

architectural growth pattern with unequivocal<br />

cytological evidence <strong>of</strong> malignancy. The papillary<br />

pattern consists <strong>of</strong> multiple, thin, delicate, fingerlike<br />

papillary projections. These papillae have thin<br />

fibrovascular cores covered by neoplastic,<br />

immature basaloid cells or more pleomorphic cells.<br />

Commonly, there is minimal keratosis. Foci <strong>of</strong><br />

necrosis <strong>and</strong> hemorrhage are common. Invasion<br />

may be difficult to define, especially in superficial<br />

biopsies. Stroma invasion consists <strong>of</strong> a single or<br />

multiple nests <strong>of</strong> tumor cells with dense<br />

lymphoplasmacytic inflammation at the tumorstroma<br />

interface. Differential diagnosis includes<br />

squamous papilloma, verrucous carcinoma, <strong>and</strong><br />

exophytic SCC. Though squamous papilloma <strong>and</strong><br />

verrucous carcinoma share similar architectural<br />

features with PSCC, the latter is easily recognized<br />

by atypia <strong>of</strong> the squamous epithelium. Patients with<br />

PSCC tend to have a better prognosis compared to<br />

those with site- <strong>and</strong> stage-matched conventional<br />

SCC. This is probably related to limited invasion in<br />

PSCC. Approximately, one third <strong>of</strong> patients<br />

develop recurrence, frequently more than once<br />

(Barnes et al., 2005; Thompson, 2006).<br />

Spindle cell carcinoma<br />

Spindle cell carcinoma is a biphasic tumor<br />

composed <strong>of</strong> a squamous cell carcinoma, either in<br />

situ <strong>and</strong>/or invasive, <strong>and</strong> a malignant spindle cell<br />

component with a mesenchymal appearance, but <strong>of</strong><br />

epithelial origin (Barnes et al., 2005). Spindle cell<br />

carcinoma most <strong>of</strong>ten occurs in males. It usually<br />

exhibits a polypoid appearance with a mean size <strong>of</strong><br />

2 cm. The surface is frequently ulcerated. The<br />

spindle cell component usually forms the bulk <strong>of</strong><br />

the tumor. It can be arranged in a diverse array <strong>of</strong><br />

appearances, including storiform, interlacing<br />

bundles or fascicles, <strong>and</strong> herringbone. The two<br />

components can abut directly against one another<br />

with areas <strong>of</strong> blending <strong>and</strong> continuity between<br />

them. Hypocellular areas with dense collagen<br />

deposition can be seen. Pleomorphism is <strong>of</strong>ten mild<br />

to moderate, without a severe degree <strong>of</strong> anaplasia.<br />

The tumor cells are plump fusiform cells, although<br />

they can be rounded <strong>and</strong> epithelioid. Rarely,<br />

metaplastic or frankly neoplastic cartilage or bone<br />

can be seen. Resemblance to fibrosarcoma or<br />

malignant fibrous histiocytoma is most common.<br />

Evidence for squamous epithelial derivation can be<br />

seen as either in situ carcinoma or as invasive SCC.<br />

The SCC component is usually minor to<br />

inconspicuous with the sarcomatoid part<br />

dominating. Carcinoma in situ can be obscured by<br />

extensive ulceration. Infiltrating SCC may be focal,<br />

requiring multiple sections for demonstration.<br />

Sometimes, only spindle cells are present; in such<br />

cases, SPCC can be mistaken for a true sarcoma.<br />

Metastases usually contain SCC alone or both SCC<br />

<strong>and</strong> the spindle cell component, <strong>and</strong> rarely, only the<br />

spindle cell component. SPCC can also be confused<br />

with reactive or benign spindle cell proliferation<br />

(such as nodular fasciitis), inflammatory<br />

my<strong>of</strong>ibroblastic sarcoma, low-grade<br />

my<strong>of</strong>ibroblastic sarcoma, <strong>and</strong> myoepithelial<br />

carcinoma. There is mounting molecular evidence<br />

that SPCC is a monoclonal epithelial neoplasm with<br />

a divergent (mesenchymal) differentiation, rather<br />

than a collision tumor. This is the one SCC variant<br />

in which immunohistochemistry may be <strong>of</strong> value.<br />

The individual neoplastic spindle cells react<br />

variably with keratin (AE1/AE3), EMA, <strong>and</strong> CK18,<br />

even though only 70% <strong>of</strong> cases will yield any<br />

epithelial immunoreactivity. A nonreactive or<br />

negative result should not dissuade the pathologist<br />

from the diagnosis, especially in the right setting.<br />

Spindle cell carcinoma metastasizes to the regional<br />

lymph nodes in up to 25% <strong>of</strong> cases, but distant<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 147


Head <strong>and</strong> Neck: Squamous cell carcinoma: an overview Rousseau A, Badoual C<br />

dissemination is less common (5-15%). The<br />

reported 5-year survival rate is between 65% <strong>and</strong><br />

95% (Barnes et al., 2005; Thompson, 2006).<br />

Acantholytic squamous cell carcinoma<br />

This is an uncommon histopathologic variant <strong>of</strong><br />

squamous cell carcinoma, characterized by<br />

acantholysis <strong>of</strong> the tumor cells, creating<br />

pseudolumina <strong>and</strong> false appearance <strong>of</strong> gl<strong>and</strong>ular<br />

differentiation. No special etiologic factor has been<br />

discovered for the mucosal acantholytic SCC. It is<br />

most frequent in sun-exposed areas <strong>of</strong> the head <strong>and</strong><br />

neck. The tumor is composed <strong>of</strong> SCC, but with foci<br />

<strong>of</strong> acantholysis in tumor nests, creating the<br />

appearance <strong>of</strong> gl<strong>and</strong>ular differentiation. The<br />

pseudolumina usually contain acantholytic <strong>and</strong><br />

dyskeratotic cells, or cellular debris, but they may<br />

be empty. They are more frequent in the deeper<br />

portions <strong>of</strong> the tumor. There is no evidence <strong>of</strong> true<br />

gl<strong>and</strong>ular differentiation or mucin production. The<br />

SCC component predominates, <strong>and</strong> is usually<br />

moderately-differentiated. The stroma is usually<br />

desmoplastic, with a lymphoplasmacytic response.<br />

The acantholysis may also form anastomosing<br />

spaces <strong>and</strong> channels mimicking angiosarcoma.<br />

Acantholytic SCC must be differentiated from<br />

adenosquamous carcinoma, adenoid cystic<br />

carcinoma, <strong>and</strong> mucoepidermoid carcinoma.<br />

Prognosis is similar to that <strong>of</strong> SCC. However, some<br />

reports suggest a more aggressive behavior (Barnes<br />

et al., 2005; Thompson, 2006).<br />

Adenosquamous carcinoma<br />

This rare aggressive neoplasm originates from the<br />

surface epithelium <strong>and</strong> is characterized by both<br />

squamous cell carcinoma <strong>and</strong> true adenocarcinoma.<br />

The larynx is the most frequent site <strong>of</strong> occurrence.<br />

Most patients (65%) present with lymph node<br />

metastases. Adenosquamous carcinoma occurs<br />

throughout the upper aerodigestive tract, <strong>of</strong>ten as an<br />

indurated submucosal nodule usually less than 1 cm<br />

in diameter. It can present as an exophytic or<br />

polypoid mass, or as poorly defined mucosal<br />

induration, frequently with ulceration. The main<br />

feature is both true adenocarcinoma <strong>and</strong> SCC. The<br />

two components occur in close proximity, but they<br />

tend to be distinct <strong>and</strong> separate, not intermingled as<br />

in mucoepidermoid carcinoma. The SCC<br />

component can present either as in situ or as an<br />

invasive SCC. The adenocarcinomatous component<br />

tends to occur in the deeper parts <strong>of</strong> the tumor. It<br />

consists <strong>of</strong> tubular structures that give rise to<br />

"gl<strong>and</strong>s within gl<strong>and</strong>s". The adenocarcinoma<br />

component can be tubular, alveolar, <strong>and</strong> gl<strong>and</strong>ular,<br />

although mucus-cell differentiation is not essential<br />

for the diagnosis. Mucin production is typically<br />

present, either intraluminal or intracellular, <strong>and</strong> can<br />

appear as signet ring cells. There is typically a<br />

sparse inflammatory cell infiltrate at the tumorstroma<br />

interface. Differential diagnosis includes<br />

mucoepidermoid carcinoma, acantholytic SCC, <strong>and</strong><br />

SCC invading seromucinous gl<strong>and</strong>s, <strong>and</strong><br />

necrotizing sialometaplasia. The most important<br />

differential diagnosis is from mucoepidermoid<br />

carcinoma as adenosquamous carcinoma has a<br />

poorer prognosis. Aggressive surgery with neck<br />

dissection yields an approximately 55% 2-year<br />

survival rate (Barnes et al., 2005; Thompson,<br />

2006).<br />

Carcinoma cuniculatum is a rare variant <strong>of</strong> oral<br />

cancer displaying similarities with lesions more<br />

commonly described in the foot in which the tumor<br />

infiltrates deeply into the bone. There is<br />

proliferation <strong>of</strong> stratified squamous epithelium in<br />

broad processes with keratin cores <strong>and</strong> keratinfilled<br />

crypts which seem to burrow into bone tissue,<br />

but lack obvious cytological features <strong>of</strong><br />

malignancy. Clinical-pathological correlation is<br />

<strong>of</strong>ten needed to make the diagnosis (Barnes et al.,<br />

2005).<br />

Nasopharyngeal carcinoma <strong>and</strong> lymphoepithelial<br />

carcinoma are rare entities distinct from<br />

conventional squamous cell carcinomas.<br />

Lymphoepithelial carcinoma (LEC) may develop in<br />

the nasal cavity <strong>and</strong> paranasal sinuses, the<br />

hypopharynx, larynx <strong>and</strong> trachea, <strong>and</strong> in the oral<br />

cavity <strong>and</strong> oropharynx. It is a poorly differentiated<br />

squamous cell carcinoma or histologically<br />

undifferentiated carcinoma accompanied by a<br />

prominent reactive lymphoplasmacytic infiltrate,<br />

morphologically similar to nasopharyngeal<br />

carcinoma. Most sinonasal LECs are associated<br />

with Epstein-Barr virus (EBV) infection (Barnes et<br />

al., 2005).<br />

Nasopharyngeal carcinoma (NPC) is a carcinoma<br />

arising in the nasopharynx that shows light<br />

microscopic or ultrastructural evidence <strong>of</strong><br />

squamous differentiation. It encompasses squamous<br />

cell carcinoma, non-keratinizing carcinoma<br />

(differentiated or undifferentiated), <strong>and</strong> basaloid<br />

squamous cell carcinoma. Keratinizing squamous<br />

cell carcinoma <strong>of</strong> the nasopharynx is<br />

morphologically similar to keratinizing squamous<br />

cell carcinomas occurring in other head <strong>and</strong> neck<br />

sites. NPC incidence is considerably higher in<br />

Chinese, Southeast Asians, North Africans, <strong>and</strong><br />

native people from the Arctic region. There is a<br />

near constant association <strong>of</strong> NPC with EBV,<br />

suggesting an oncogenic role <strong>of</strong> the virus. NPC<br />

harbors a highly malignant behavior with extensive<br />

loco-regional infiltration, early lymphatic spread,<br />

<strong>and</strong> hematogenous dissemination (Barnes et al.,<br />

2005).<br />

Histogenesis<br />

SCC originates from the squamous mucosa or from<br />

ciliated respiratory epithelium that has undergone<br />

squamous metaplasia (Barnes et al., 2005).<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 148


Head <strong>and</strong> Neck: Squamous cell carcinoma: an overview Rousseau A, Badoual C<br />

Figure 3: Carcinoma in situ. Full thickness architectural abnormalities <strong>and</strong> cytological atypia Hematoxylin <strong>and</strong> eosin staining<br />

(original magnification x200). Figure 4: Invasive SCC. Invasive growth with disruption <strong>of</strong> the basement membrane <strong>and</strong> extension<br />

into the underlying tissue. Hematoxylin <strong>and</strong> eosin staining (original magnification Fig. 4A: x20, Fig. 4B: x100). Figure 5: Invasive<br />

SCC. Mitoses (at 9 o' clock) <strong>and</strong> nuclear atypia. Hematoxylin <strong>and</strong> eosin staining (original magnification x400). Figure 6: Invasive<br />

SCC. Focal keratinization (left h<strong>and</strong> side). Hematoxylin <strong>and</strong> eosin staining (original magnification x400). Figure 7: Verrucous<br />

carcinoma. Thickened, club-shaped, projections <strong>of</strong> well-differentiated squamous epithelium <strong>and</strong> abundant surface keratosis<br />

("church-spire" keratosis). Hematoxylin <strong>and</strong> eosin staining (original magnification x20). Figure 8: Basaloid SCC. Solid pattern <strong>of</strong><br />

growth <strong>and</strong> central comedo-type necrosis. Hematoxylin <strong>and</strong> eosin staining (original magnification x100). Figure 9: p16<br />

immunostaining in basaloid SCC (original magnification x200).<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 149


Head <strong>and</strong> Neck: Squamous cell carcinoma: an overview Rousseau A, Badoual C<br />

<strong>Cytogenetics</strong><br />

Precursor lesions<br />

Malignant transformation <strong>of</strong> the mucosal lining is a<br />

genetic process resulting from accumulation <strong>of</strong><br />

multiple genetic alterations that dictates the<br />

frequency <strong>and</strong> pace <strong>of</strong> progression to invasive<br />

carcinoma. LOH studies indicate that the earliest<br />

alterations appear to target specific genes located on<br />

chromosomes 3p, 9p21 (CDKN2A), <strong>and</strong> 17p13<br />

(TP53). Alterations that tend to occur in association<br />

with higher grades <strong>of</strong> dysplasia <strong>and</strong> SCC include<br />

cyclin D1 amplification, PTEN inactivation, <strong>and</strong><br />

LOH at 13q21, 14q32, 6p, 8, 4q27, <strong>and</strong> 10q23<br />

(Barnes et al., 2005).<br />

There are no individual markers that reliably predict<br />

malignant transformation <strong>of</strong> dysplastic lesions.<br />

Ploidy studies <strong>of</strong> dysplastic leukoplakias showed<br />

that the great majority <strong>of</strong> aneuploid lesions<br />

developed SCC in the follow-up period, by contrast<br />

with 60% <strong>of</strong> tetraploid lesions <strong>and</strong> only about 3%<br />

<strong>of</strong> diploid lesions (Sudbo et al., 2001). Similar<br />

studies on erythroplakias confirmed the higher<br />

predictive potential <strong>of</strong> aneuploidy in identifying<br />

cases which progressed to SCC (Sudbo et al.,<br />

2002).<br />

Invasive squamous cell carcinoma<br />

HNSCC is a heterogeneous disease, comprising at<br />

least two distinct genetic subclasses: tumors that are<br />

caused by infection with high-risk types <strong>of</strong> HPV,<br />

<strong>and</strong> those that do not contain HPV. Approximately<br />

20% <strong>of</strong> HNSCCs contain transcriptionally active<br />

HPV whereas 60% harbor a TP53 mutation. In the<br />

remaining 20%, other genes encoding proteins in<br />

the p53 pathway may be targeted or these tumors<br />

may undergo p53-independent malignant<br />

progression.<br />

- HPV-negative squamous cell carcinoma<br />

Tobacco <strong>and</strong> alcohol-induced HNSCCs are<br />

characterized by TP53 mutation. The excess <strong>of</strong> G to<br />

T transversions <strong>and</strong> the codons more frequently<br />

affected were attributed to the carcinogenic effect<br />

<strong>of</strong> tobacco smoking (Barnes et al., 2005). Other<br />

genes are involved in the pathogenesis <strong>of</strong> HPVnegative<br />

tumors. CCND1, which encodes cyclin<br />

D1, is amplified or gained in more than 80% <strong>of</strong><br />

HPV-negative HNSCCs. CDKN2A (encoding p16)<br />

can be inactivated by mutation, homozygous<br />

deletion, or promoter hypermethylation (Barnes et<br />

al., 2005).<br />

EGFR (Epidermal Growth Factor Receptor) is<br />

overexpressed in most HNSCCs (Hama et al.,<br />

2009). The EGFR is a receptor tyrosine kinase<br />

belonging to the erbB family <strong>of</strong> cell surface<br />

receptors. Once phosphorylated, it can signal<br />

through MAPK, Akt, ERK, <strong>and</strong> Jak/STAT<br />

pathways. These pathways are related to cellular<br />

proliferation, apoptosis, invasion, angiogenesis, <strong>and</strong><br />

metastasis. Dysfunction <strong>of</strong> the receptor <strong>and</strong> its<br />

associated pathways occurs in 80-90% <strong>of</strong> HNSCCs<br />

(Kalyankrishna et al., 2006). Mutations <strong>and</strong><br />

amplifications <strong>of</strong> EGFR have been reported, albeit<br />

at relatively low frequencies. EGFR amplification<br />

has been detected in 10-30% <strong>of</strong> cases (Temam et<br />

al., 2007; Sheu et al., 2009). Even though few<br />

activating mutations have been found, the mutant<br />

form EGFRvIII has been detected in 42% <strong>of</strong><br />

HNSCCs (Sok et al., 2006). Interestingly, it has<br />

been shown that microscopically normal mucosa<br />

adjacent to invasive SCC displayed a high degree <strong>of</strong><br />

overexpression <strong>and</strong> that the upregulation <strong>of</strong> EGFR<br />

occurs in the transition from dysplasia to cancer<br />

(Gr<strong>and</strong>is et al., 1993; Shin et al., 1994). Elevated<br />

levels <strong>of</strong> EGFR expression have been associated to<br />

a poor clinical outcome (Chung et al., 2004;<br />

Temam et al., 2007). High copy number<br />

amplification has also been shown to portend a<br />

dismal prognosis in HNSCCs (Chung et al., 2006).<br />

However, overexpression <strong>of</strong> EGFR may be a<br />

biomarker for an improved response to therapy <strong>and</strong><br />

could serve as a predictive marker (Bentzen et al.,<br />

2005). The EGFR pathway can be targeted through<br />

the use <strong>of</strong> specific tyrosine kinase inhibitors (TKIs),<br />

monoclonal antibodies blocking receptor<br />

dimerization, <strong>and</strong> anti-sense oligodeoxynucleotides<br />

or siRNA blocking mRNA expression (Glazer et<br />

al., 2009).<br />

Both mutations <strong>and</strong> gene amplifications <strong>of</strong> MET<br />

have been described in HNSCCs. MET, the<br />

receptor for Hepatocyte Growth Factor (HGF) is a<br />

tyrosine kinase encoded by MET on chromosome<br />

7q31. It activates the AKT <strong>and</strong> Ras pathways <strong>and</strong><br />

influences growth, motility <strong>and</strong> angiogenesis in<br />

HNSCCs (Leeman et al., 2011).<br />

Finally, the PI3K-PTEN-AKT pathway is<br />

frequently activated in HNSCCs (Leeman et al.,<br />

2011).<br />

- HPV-induced squamous cell carcinoma<br />

HPVs are DNA viruses that show a tropism for<br />

squamous epithelium. HPV is a strictly<br />

epitheliotropic, circular double-str<strong>and</strong>ed DNA<br />

virus. There are more than 100 subtypes <strong>of</strong> HPV,<br />

some <strong>of</strong> which are involved in cervical<br />

carcinogenesis <strong>and</strong> have been designated as highrisk<br />

HPVs (e.g. HPV-16 <strong>and</strong> -18) (zur Hausen,<br />

2002; Moody et al., 2010). HPV-positive HNSCCs<br />

present with distinct molecular pr<strong>of</strong>iles compared to<br />

HPV-negative tumors whereas they harbor<br />

similarities with HPV-positive cervical SCCs. Most<br />

HPV-induced HNSCCs are caused by one subtype,<br />

HPV-16. HPV infection is an early, <strong>and</strong> probably<br />

initiating, oncogenic event in HNSCCs. High-risk<br />

oncogenic HPV subtypes have been shown to be<br />

capable <strong>of</strong> transforming oral epithelial cells through<br />

the viral oncoproteins E6 <strong>and</strong> E7. The E6 protein<br />

induces degradation <strong>of</strong> p53 through ubiquitinmediated<br />

proteolysis, leading to substantial loss <strong>of</strong><br />

p53 activity. The usual function <strong>of</strong> p53 is to arrest<br />

cells in G1 or induce apoptosis to allow host DNA<br />

to be repaired. E6-expressing cells are not capable<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 150


Head <strong>and</strong> Neck: Squamous cell carcinoma: an overview Rousseau A, Badoual C<br />

<strong>of</strong> this p53-mediated response to DNA damage <strong>and</strong>,<br />

hence, are susceptible to genomic instability. The<br />

E7 protein binds <strong>and</strong> inactivates the retinoblastoma<br />

tumor suppressor gene product pRB, causing the<br />

cell to enter S-phase, leading to cell cycle<br />

disruption. This functional inactivation <strong>of</strong> pRB also<br />

results in a reciprocal overexpression <strong>of</strong> p16<br />

protein. By immunohistochemistry, most HPVpositive<br />

HNSCCs show p16 overexpression (Marur<br />

et al., 2010) (Fig. 9). The combination <strong>of</strong> low<br />

EGFR <strong>and</strong> high p16 expression has been shown to<br />

highly correlate with better clinical outcome<br />

compared with high EGFR expression <strong>and</strong> low<br />

HPV titer or high EGFR <strong>and</strong> low p16 expression<br />

(Kumar et al., 2008). P16 expression in<br />

oropharyngeal SCCs has also been associated with<br />

longer survival times regardless <strong>of</strong> HPV status<br />

(Lewis et al., 2010).<br />

The best method for HPV detection is still<br />

controversial. PCR-based detection <strong>of</strong> HPV E6<br />

oncogene expression in frozen samples is generally<br />

regarded as the gold st<strong>and</strong>ard but in situ<br />

hybridization is also commonly used. P16<br />

immunohistochemistry could serve as a potential<br />

surrogate marker (Marur et al., 2010).<br />

As mentioned above, HPV-positive HNSCCs are<br />

typically TP53 wild-type. There also seems to be an<br />

inverse relationship between EGFR expression <strong>and</strong><br />

HPV status.<br />

Prognosis<br />

Precursor lesions<br />

Some precursor lesions are self-limiting <strong>and</strong><br />

reversible (particularly if apparent etiologic factors<br />

are removed), others persist <strong>and</strong> some progress to<br />

SCC. The likelihood <strong>of</strong> malignant change directly<br />

relates to the severity <strong>of</strong> dysplasia. However, it is<br />

clear that malignancy can develop from any grade<br />

<strong>of</strong> dysplasia or even from morphologically normal<br />

epithelium. Dysplastic lesions classified as<br />

moderate to severe have an 11% rate <strong>of</strong> malignant<br />

transformation. Diagnosis <strong>of</strong> precursor lesions<br />

implies a need for close follow-up <strong>and</strong> complete<br />

excision. Patients with carcinoma in situ require<br />

more extensive management (Thompson, 2006).<br />

Dysplastic lesions are frequently found in the<br />

surgical margins <strong>of</strong> invasive SCC, meaning such<br />

lesions can remain in the patient. These unresected<br />

fields act as an important source <strong>of</strong> local<br />

recurrences <strong>and</strong> second primary tumors that <strong>of</strong>ten<br />

occur in patients treated for HNSCC.<br />

Invasive squamous cell carcinoma<br />

The prognosis for patients with HNSCC is<br />

determined by the stage at presentation, established<br />

based on the extent <strong>of</strong> the tumor, as well as the<br />

presence <strong>of</strong> lymph-node metastases <strong>and</strong> distant<br />

metastases. About one third <strong>of</strong> patients presents<br />

with early-stage disease, whereas two thirds present<br />

with advanced cancer with lymph node metastases<br />

(Jemal et al., 2007). Early-stage tumors are treated<br />

with surgery or radiotherapy <strong>and</strong> have a favorable<br />

prognosis. The st<strong>and</strong>ard <strong>of</strong> care for advanced<br />

tumors is surgery combined with adjuvant radiation<br />

therapy <strong>and</strong>/or chemotherapy. Survival outcomes<br />

are poor (40-50% five-year survival rates) <strong>and</strong> the<br />

treatment is uniformly morbid. Organ-preservation<br />

protocols, with combined chemotherapy/radiation<br />

therapy <strong>and</strong> surgery for salvage, are increasingly<br />

performed. These protocols are particularly<br />

effective for young patients with a good<br />

performance status presenting with moderatelyadvanced<br />

laryngeal or pharyngeal SCC. Several<br />

characteristics <strong>of</strong> patients with HNSCC have been<br />

linked to favorable prognosis, including nonsmoker,<br />

minimum exposure to alcohol, good<br />

performance status, <strong>and</strong> absence <strong>of</strong> co-morbid<br />

disorders (Marur et al., 2010). Thirty-five to 55% <strong>of</strong><br />

patients with advanced-stage HNSCC remain<br />

disease-free 3 years after st<strong>and</strong>ard treatment.<br />

However, locoregional recurrence develops in 30%<br />

to 40% <strong>of</strong> patients <strong>and</strong> distant metastases occur in<br />

20% to 30% <strong>of</strong> HNSCCs (Forastiere et al., 2003).<br />

Locoregional recurrences <strong>of</strong>ten require a<br />

combination <strong>of</strong> surgery, radiation therapy, <strong>and</strong>/or<br />

chemotherapy, <strong>and</strong> metastatic disease is treated<br />

with chemotherapy.<br />

Recently, the use <strong>of</strong> targeted drugs has entered the<br />

field. Cetuximab is one <strong>of</strong> the most well studied<br />

monoclonal antibodies directed against EGFR.<br />

Binding <strong>of</strong> the antibody to EGFR prevents<br />

activation <strong>of</strong> the receptor by endogenous lig<strong>and</strong>s.<br />

An overall survival benefit <strong>and</strong> an increased<br />

duration <strong>of</strong> locoregional control have been observed<br />

in advanced HNSCCs treated with a combination <strong>of</strong><br />

radiation therapy <strong>and</strong> cetuximab, compared to<br />

radiation therapy alone (Bonner et al., 2006).<br />

It has been demonstrated that the presence <strong>and</strong> type<br />

<strong>of</strong> TP53 mutation is also <strong>of</strong> prognostic relevance.<br />

Several studies have shown a correlation between<br />

p53 mutation <strong>and</strong> lower response rates to<br />

chemotherapy <strong>and</strong> shorter overall survival times<br />

(Erber et al., 1998; Cabelguenne et al., 2000;<br />

Temam et al., 2000; Poeta et al., 2007).<br />

On the whole, survival has not markedly improved<br />

in recent decades because patients still frequently<br />

develop locoregional recurrences, distant<br />

metastases, <strong>and</strong> second primary tumors. Primary<br />

prevention could be achieved by cessation <strong>of</strong><br />

smoking <strong>and</strong> reduction <strong>of</strong> alcohol consumption.<br />

Patients with HPV-positive HNSCC tend to be<br />

younger <strong>and</strong> have a lower tobacco <strong>and</strong> alcohol<br />

consumption. They <strong>of</strong>ten present at a late stage with<br />

large metastatic cervical lymph nodes.<br />

Histopathologically, the tumor is <strong>of</strong>ten moderately<br />

to poorly-differentiated with basaloid features<br />

(Gillison et al., 2000). However, HPV-positive<br />

HNSCCs are associated with a more favorable<br />

clinical outcome regardless <strong>of</strong> treatment modalities,<br />

<strong>and</strong> this may be related to immune surveillance to<br />

viral antigens (Leemans et al., 2011). Prognosis is<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 151


Head <strong>and</strong> Neck: Squamous cell carcinoma: an overview Rousseau A, Badoual C<br />

better not only for patients treated with radiation<br />

therapy or concomitant chemotherapy/radiation<br />

therapy but also for patients treated with surgery<br />

alone (Lassen et al., 2009; Fischer et al., 2010). In<br />

patients with oropharyngeal SCC treated with<br />

surgery, the 5-year survival rates for p16-negative<br />

<strong>and</strong> p16-positive patients were 26.8% <strong>and</strong> 57.1%,<br />

respectively (Lassen et al., 2009). In another study,<br />

patients with HPV-positive oropharyngeal SCC had<br />

a 58% reduction in the risk <strong>of</strong> death (Ang et al.,<br />

2010). The better prognosis associated with HPVstatus<br />

has also been observed in high-grade<br />

basaloid SCCs <strong>of</strong> the oropharynx (Thariat et al.,<br />

2010). Most studies confirm that HPV is one <strong>of</strong> the<br />

most important independent prognostic factors in<br />

HNSCC. However, only the rate <strong>of</strong> locoregional<br />

recurrence, but not that <strong>of</strong> distant disease, is<br />

diminished in patients with HPV-positive SCC.<br />

Increased sensitivity to chemotherapy <strong>and</strong><br />

radiotherapy in HPV-positive oropharyngeal cancer<br />

may be related to absence <strong>of</strong> exposure to tobacco<br />

<strong>and</strong> presence <strong>of</strong> functional p53 protein. Increased<br />

survival <strong>of</strong> patients with HPV-positive SCC may be<br />

in part attributable to absence <strong>of</strong> dysplastic fields<br />

related to tobacco <strong>and</strong> alcohol exposure. So, HPVstatus<br />

<strong>of</strong> HNSCC is a prognostic factor for<br />

progression-free <strong>and</strong> overall survival <strong>and</strong> might also<br />

be a predictive factor. Use <strong>of</strong> HPV vaccines against<br />

infection <strong>and</strong> therapeutic vaccines in the adjuvant<br />

setting for locoregional recurrence <strong>and</strong> distant<br />

disease should be assessed in this form <strong>of</strong> HNSCC.<br />

References<br />

Benhamou CA, Laraqui N, Touhami M, Chekkoury A,<br />

Benchakroun Y, Samlali R, Kahlain A. [Tobacco <strong>and</strong><br />

cancer <strong>of</strong> the larynx: a prospective survey <strong>of</strong> 58 patients].<br />

Rev Laryngol Otol Rhinol (Bord). 1992;113(4):285-8<br />

Gr<strong>and</strong>is JR, Tweardy DJ. Elevated levels <strong>of</strong> transforming<br />

growth factor alpha <strong>and</strong> epidermal growth factor receptor<br />

messenger RNA are early markers <strong>of</strong> carcinogenesis in<br />

head <strong>and</strong> neck cancer. Cancer Res. 1993 Aug<br />

1;53(15):3579-84<br />

Shin DM, Ro JY, Hong WK, Hittelman WN. Dysregulation<br />

<strong>of</strong> epidermal growth factor receptor expression in<br />

premalignant lesions during head <strong>and</strong> neck tumorigenesis.<br />

Cancer Res. 1994 Jun 15;54(12):3153-9<br />

Erber R, Conradt C, Homann N, Enders C, Finckh M, Dietz<br />

A, Weidauer H, Bosch FX. TP53 DNA contact mutations<br />

are selectively associated with allelic loss <strong>and</strong> have a<br />

strong clinical impact in head <strong>and</strong> neck cancer. Oncogene.<br />

1998 Apr 2;16(13):1671-9<br />

La Vecchia C, Franceschi S, Favero A, Talamini R, Negri<br />

E. Alcohol intake <strong>and</strong> cancer <strong>of</strong> the upper digestive tract.<br />

Pattern <strong>of</strong> risk in Italy is different from that in Denmark.<br />

BMJ. 1999 May 8;318(7193):1289-90; author reply 1291<br />

Cabelguenne A, Blons H, de Waziers I, Carnot F, Houllier<br />

AM, Soussi T, Brasnu D, Beaune P, Laccourreye O,<br />

Laurent-Puig P. p53 alterations predict tumor response to<br />

neoadjuvant chemotherapy in head <strong>and</strong> neck squamous<br />

cell carcinoma: a prospective series. J Clin Oncol. 2000<br />

Apr;18(7):1465-73<br />

Gillison ML, Koch WM, Capone RB, Spafford M, Westra<br />

WH, Wu L, Zahurak ML, Daniel RW, Viglione M, Symer<br />

DE, Shah KV, Sidransky D. Evidence for a causal<br />

association between human papillomavirus <strong>and</strong> a subset<br />

<strong>of</strong> head <strong>and</strong> neck cancers. J Natl Cancer Inst. 2000 May<br />

3;92(9):709-20<br />

Temam S, Flahault A, Périé S, Monceaux G, Coulet F,<br />

Callard P, Bernaudin JF, St Guily JL, Fouret P. p53 gene<br />

status as a predictor <strong>of</strong> tumor response to induction<br />

chemotherapy <strong>of</strong> patients with locoregionally advanced<br />

squamous cell carcinomas <strong>of</strong> the head <strong>and</strong> neck. J Clin<br />

Oncol. 2000 Jan;18(2):385-94<br />

Sudbø J, Bryne M, Johannessen AC, Kildal W, Danielsen<br />

HE, Reith A. Comparison <strong>of</strong> histological grading <strong>and</strong> largescale<br />

genomic status (DNA ploidy) as prognostic tools in<br />

oral dysplasia. J Pathol. 2001 Jul;194(3):303-10<br />

Sudbø J, Kildal W, Johannessen AC, Koppang HS, Sudbø<br />

A, Danielsen HE, Risberg B, Reith A. Gross genomic<br />

aberrations in precancers: clinical implications <strong>of</strong> a longterm<br />

follow-up study in oral erythroplakias. J Clin Oncol.<br />

2002 Jan 15;20(2):456-62<br />

zur Hausen H. Papillomaviruses <strong>and</strong> cancer: from basic<br />

studies to clinical application. Nat Rev Cancer. 2002<br />

May;2(5):342-50<br />

Forastiere AA, Goepfert H, Maor M, Pajak TF, Weber R,<br />

Morrison W, Glisson B, Trotti A, Ridge JA, Chao C, Peters<br />

G, Lee DJ, Leaf A, Ensley J, Cooper J. Concurrent<br />

chemotherapy <strong>and</strong> radiotherapy for organ preservation in<br />

advanced laryngeal cancer. N Engl J Med. 2003 Nov<br />

27;349(22):2091-8<br />

Kutler DI, Auerbach AD, Satagopan J, Giampietro PF,<br />

Batish SD, Huvos AG, Goberdhan A, Shah JP, Singh B.<br />

High incidence <strong>of</strong> head <strong>and</strong> neck squamous cell carcinoma<br />

in patients with Fanconi anemia. Arch Otolaryngol Head<br />

Neck Surg. 2003 Jan;129(1):106-12<br />

Pelucchi C, Talamini R, Levi F, Bosetti C, La Vecchia C,<br />

Negri E, Parpinel M, Franceschi S. Fibre intake <strong>and</strong><br />

laryngeal cancer risk. Ann Oncol. 2003 Jan;14(1):162-7<br />

Znaor A, Brennan P, Gajalakshmi V, Mathew A, Shanta V,<br />

Varghese C, B<strong>of</strong>fetta P. Independent <strong>and</strong> combined effects<br />

<strong>of</strong> tobacco smoking, chewing <strong>and</strong> alcohol drinking on the<br />

risk <strong>of</strong> oral, pharyngeal <strong>and</strong> esophageal cancers in Indian<br />

men. Int J Cancer. 2003 Jul 10;105(5):681-6<br />

Chung CH, Parker JS, Karaca G, Wu J, Funkhouser WK,<br />

Moore D, Butterfoss D, Xiang D, Zanation A, Yin X,<br />

Shockley WW, Weissler MC, Dressler LG, Shores CG,<br />

Yarbrough WG, Perou CM. Molecular classification <strong>of</strong><br />

head <strong>and</strong> neck squamous cell carcinomas using patterns<br />

<strong>of</strong> gene expression. Cancer Cell. 2004 May;5(5):489-500<br />

Smith EM, Ritchie JM, Summersgill KF, Klussmann JP,<br />

Lee JH, Wang D, Haugen TH, Turek LP. Age, sexual<br />

behavior <strong>and</strong> human papillomavirus infection in oral cavity<br />

<strong>and</strong> oropharyngeal cancers. Int J Cancer. 2004 Feb<br />

20;108(5):766-72<br />

Barnes L, Eveson JW, Reichart P, Sidransky D.. Pathology<br />

<strong>and</strong> <strong>Gene</strong>tics <strong>of</strong> Head <strong>and</strong> Neck Tumours. World Health<br />

Organization Classification <strong>of</strong> Tumours. IARC Press, Lyon.<br />

2005.<br />

Bentzen SM, Atasoy BM, Daley FM, Dische S, Richman<br />

PI, Saunders MI, Trott KR, Wilson GD.. Epidermal growth<br />

factor receptor expression in pretreatment biopsies from<br />

head <strong>and</strong> neck squamous cell carcinoma as a predictive<br />

factor for a benefit from accelerated radiation therapy in a<br />

r<strong>and</strong>omized controlled trial. J Clin Oncol. 2005 Aug<br />

20;23(24):5560-7.<br />

Bonner JA, Harari PM, Giralt J, Azarnia N, Shin DM,<br />

Cohen RB, Jones CU, Sur R, Raben D, Jassem J, Ove R,<br />

Kies MS, Baselga J, Youssoufian H, Amellal N, Rowinsky<br />

EK, Ang KK.. Radiotherapy plus cetuximab for squamous-<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 152


Head <strong>and</strong> Neck: Squamous cell carcinoma: an overview Rousseau A, Badoual C<br />

cell carcinoma <strong>of</strong> the head <strong>and</strong> neck. N Engl J Med. 2006<br />

Feb 9;354(6):567-78.<br />

Chung CH, Ely K, McGavran L, Varella-Garcia M, Parker<br />

J, Parker N, Jarrett C, Carter J, Murphy BA, Netterville J,<br />

Burkey BB, Sinard R, Cmelak A, Levy S, Yarbrough WG,<br />

Slebos RJ, Hirsch FR.. Increased epidermal growth factor<br />

receptor gene copy number is associated with poor<br />

prognosis in head <strong>and</strong> neck squamous cell carcinomas. J<br />

Clin Oncol. 2006 Sep 1;24(25):4170-6.<br />

Dabbs D.. Diagnostic immunohistochemistry. Churchill<br />

Livingstone Elsevier, Philadelphia. 2006.<br />

Kalyankrishna S, Gr<strong>and</strong>is JR.. Epidermal growth factor<br />

receptor biology in head <strong>and</strong> neck cancer. J Clin Oncol.<br />

2006 Jun 10;24(17):2666-72. (REVIEW)<br />

Kamangar F, Dores GM, Anderson WF.. Patterns <strong>of</strong><br />

cancer incidence, mortality, <strong>and</strong> prevalence across five<br />

continents: defining priorities to reduce cancer disparities<br />

in different geographic regions <strong>of</strong> the world. J Clin Oncol.<br />

2006 May 10;24(14):2137-50.<br />

Sok JC, Coppelli FM, Thomas SM, Lango MN, Xi S, Hunt<br />

JL, Freilino ML, Graner MW, Wikstr<strong>and</strong> CJ, Bigner DD,<br />

Gooding WE, Furnari FB, Gr<strong>and</strong>is JR.. Mutant epidermal<br />

growth factor receptor (EGFRvIII) contributes to head <strong>and</strong><br />

neck cancer growth <strong>and</strong> resistance to EGFR targeting. Clin<br />

Cancer Res. 2006 Sep 1;12(17):5064-73.<br />

Thompson LDR.. Head <strong>and</strong> Neck Pathology. Foundations<br />

in Diagnostic Pathology Series. Churchill Livingstone,<br />

Elsevier, Philadelphia. 2006.<br />

Jemal A, Siegel R, Ward E, Murray T, Xu J, Thun MJ..<br />

Cancer statistics, 2007. CA Cancer J Clin. 2007 Jan-<br />

Feb;57(1):43-66.<br />

Poeta ML, Manola J, Goldwasser MA, Forastiere A, Benoit<br />

N, Califano JA, Ridge JA, Goodwin J, Kenady D, Saunders<br />

J, Westra W, Sidransky D, Koch WM.. TP53 mutations <strong>and</strong><br />

survival in squamous-cell carcinoma <strong>of</strong> the head <strong>and</strong> neck.<br />

N Engl J Med. 2007 Dec 20;357(25):2552-61.<br />

Temam S, Kawaguchi H, El-Naggar AK, Jelinek J, Tang H,<br />

Liu DD, Lang W, Issa JP, Lee JJ, Mao L.. Epidermal<br />

growth factor receptor copy number alterations correlate<br />

with poor clinical outcome in patients with head <strong>and</strong> neck<br />

squamous cancer. J Clin Oncol. 2007 Jun 1;25(16):2164-<br />

70.<br />

Gillison ML, D'Souza G, Westra W, Sugar E, Xiao W,<br />

Begum S, Viscidi R.. Distinct risk factor pr<strong>of</strong>iles for human<br />

papillomavirus type 16-positive <strong>and</strong> human papillomavirus<br />

type 16-negative head <strong>and</strong> neck cancers. J Natl Cancer<br />

Inst. 2008 Mar 19;100(6):407-20. Epub 2008 Mar 11.<br />

Kumar B, Cordell KG, Lee JS, Worden FP, Prince ME,<br />

Tran HH, Wolf GT, Urba SG, Chepeha DB, Teknos TN,<br />

Eisbruch A, Tsien CI, Taylor JM, D'Silva NJ, Yang K,<br />

Kurnit DM, Bauer JA, Bradford CR, Carey TE.. EGFR, p16,<br />

HPV Titer, Bcl-xL <strong>and</strong> p53, sex, <strong>and</strong> smoking as indicators<br />

<strong>of</strong> response to therapy <strong>and</strong> survival in oropharyngeal<br />

cancer. J Clin Oncol. 2008 Jul 1;26(19):3128-37. Epub<br />

2008 May 12.<br />

Glazer CA, Chang SS, Ha PK, Califano JA.. Applying the<br />

molecular biology <strong>and</strong> epigenetics <strong>of</strong> head <strong>and</strong> neck<br />

cancer in everyday clinical practice. Oral Oncol. 2009 Apr-<br />

May;45(4-5):440-6. Epub 2008 Jul 31. (REVIEW)<br />

Hama T, Yuza Y, Saito Y, O-uchi J, Kondo S, Okabe M,<br />

Yamada H, Kato T, Moriyama H, Kurihara S, Urashima M..<br />

Prognostic significance <strong>of</strong> epidermal growth factor receptor<br />

phosphorylation <strong>and</strong> mutation in head <strong>and</strong> neck squamous<br />

cell carcinoma. Oncologist. 2009 Sep;14(9):900-8. Epub<br />

2009 Sep 2.<br />

Lassen P, Eriksen JG, Hamilton-Dutoit S, Tramm T, Alsner<br />

J, Overgaard J.. Effect <strong>of</strong> HPV-associated p16INK4A<br />

expression on response to radiotherapy <strong>and</strong> survival in<br />

squamous cell carcinoma <strong>of</strong> the head <strong>and</strong> neck. J Clin<br />

Oncol. 2009 Apr 20;27(12):1992-8. Epub 2009 Mar 16.<br />

Sheu JJ, Hua CH, Wan L, Lin YJ, Lai MT, Tseng HC,<br />

Jinawath N, Tsai MH, Chang NW, Lin CF, Lin CC, Hsieh<br />

LJ, Wang TL, Shih IeM, Tsai FJ.. Functional genomic<br />

analysis identified epidermal growth factor receptor<br />

activation as the most common genetic event in oral<br />

squamous cell carcinoma. Cancer Res. 2009 Mar<br />

15;69(6):2568-76. Epub 2009 Mar 10.<br />

Ang KK, Harris J, Wheeler R, Weber R, Rosenthal DI,<br />

Nguyen-Tan PF, Westra WH, Chung CH, Jordan RC, Lu<br />

C, Kim H, Axelrod R, Silverman CC, Redmond KP, Gillison<br />

ML.. Human papillomavirus <strong>and</strong> survival <strong>of</strong> patients with<br />

oropharyngeal cancer. N Engl J Med. 2010 Jul<br />

1;363(1):24-35. Epub 2010 Jun 7.<br />

Fischer CA, Zlobec I, Green E, Probst S, Storck C, Lugli A,<br />

Tornillo L, Wolfensberger M, Terracciano LM.. Is the<br />

improved prognosis <strong>of</strong> p16 positive oropharyngeal<br />

squamous cell carcinoma dependent <strong>of</strong> the treatment<br />

modality? Int J Cancer. 2010 Mar 1;126(5):1256-62.<br />

Lewis JS Jr, Thorstad WL, Chernock RD, Haughey BH,<br />

Yip JH, Zhang Q, El-M<strong>of</strong>ty SK.. p16 positive oropharyngeal<br />

squamous cell carcinoma:an entity with a favorable<br />

prognosis regardless <strong>of</strong> tumor HPV status. Am J Surg<br />

Pathol. 2010 Aug;34(8):1088-96.<br />

Marur S, D'Souza G, Westra WH, Forastiere AA.. HPVassociated<br />

head <strong>and</strong> neck cancer: a virus-related cancer<br />

epidemic. Lancet Oncol. 2010 Aug;11(8):781-9. Epub 2010<br />

May 5. (REVIEW)<br />

Moody CA, Laimins LA.. Human papillomavirus<br />

oncoproteins: pathways to transformation. Nat Rev<br />

Cancer. 2010 Aug;10(8):550-60. Epub 2010 Jul 1.<br />

(REVIEW)<br />

Thariat J, Badoual C, Faure C, Butori C, Marcy PY, Righini<br />

CA.. Basaloid squamous cell carcinoma <strong>of</strong> the head <strong>and</strong><br />

neck: role <strong>of</strong> HPV <strong>and</strong> implication in treatment <strong>and</strong><br />

prognosis. J Clin Pathol. 2010 Oct;63(10):857-66.<br />

(REVIEW)<br />

Leemans CR, Braakhuis BJ, Brakenh<strong>of</strong>f RH.. The<br />

molecular biology <strong>of</strong> head <strong>and</strong> neck cancer. Nat Rev<br />

Cancer. 2011 Jan;11(1):9-22. Epub 2010 Dec 16.<br />

(REVIEW)<br />

This article should be referenced as such:<br />

Rousseau A, Badoual C. Head <strong>and</strong> Neck: Squamous cell<br />

carcinoma: an overview. <strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol<br />

Haematol. 2012; 16(2):143-153.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 153


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

Cancer Prone Disease <strong>Section</strong><br />

Short Communication<br />

Rombo syndrome<br />

Jean-Loup Huret<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

<strong>Gene</strong>tics, Dept Medical Information, University <strong>of</strong> Poitiers, CHU Poitiers Hospital, F-86021 Poitiers,<br />

France (JLH)<br />

Published in <strong>Atlas</strong> Database: September 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/Kprones/RomboID10169.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI RomboID10169.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Identity<br />

Inheritance<br />

Rare disorder, with less than 10 cases described,<br />

with a probable autosomal dominant transmission,<br />

as suggested by the family tree <strong>of</strong> four generations<br />

in the princeps report (Michaëlsson et al., 1981).<br />

Clinics<br />

Phenotype <strong>and</strong> clinics<br />

Skin changes appear at the age <strong>of</strong> 6-10 years, with<br />

cyanotic redness, acral erythema, thin implantation<br />

<strong>of</strong> hair <strong>and</strong> absent eyelashes (hypotrichosis).<br />

Atrophoderma vermiculatum (severe skin atrophy)<br />

<strong>of</strong> the face <strong>and</strong> sun-exposed areas, telangiectasia<br />

<strong>and</strong> milia-like papules develop in adulthood.<br />

Histology <strong>of</strong> the skin shows highly irregular<br />

distribution <strong>of</strong> elastin in the upper dermis, with<br />

areas without elastin <strong>and</strong> others with clumps <strong>of</strong><br />

elastin, vascular proliferation <strong>and</strong> lymphocytes<br />

infiltration (Michaëlsson et al., 1981; Van Steensel<br />

et al., 2001).<br />

Differential Diagnosis<br />

Resembles Bazex-Dupré-Christol syndrome, which<br />

is a X-linked dominant disease.<br />

Neoplastic risk<br />

Basal cell carcinomas are a frequent complication.<br />

<strong>Gene</strong>s involved <strong>and</strong><br />

proteins<br />

Note<br />

The gene involved in this rare disease is unknown.<br />

References<br />

Michaëlsson G, Olsson E, Westermark P. The Rombo<br />

syndrome: a familial disorder with vermiculate<br />

atrophoderma, milia, hypotrichosis, trichoepitheliomas,<br />

basal cell carcinomas <strong>and</strong> peripheral vasodilation with<br />

cyanosis. Acta Derm Venereol. 1981;61(6):497-503<br />

van Steensel MA, Jaspers NG, Steijlen PM. A case <strong>of</strong><br />

Rombo syndrome. Br J Dermatol. 2001 Jun;144(6):1215-8<br />

This article should be referenced as such:<br />

Huret JL. Rombo syndrome. <strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol<br />

Haematol. 2012; 16(2):154.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 154


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

Deep Insight <strong>Section</strong><br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

Cohesins <strong>and</strong> cohesin-regulators: Role in<br />

Chromosome Segregation/Repair <strong>and</strong> Potential<br />

in Tumorigenesis<br />

José L Barbero<br />

Cell Proliferation <strong>and</strong> Development Program, Chromosome Dynamics in Meiosis Laboratory, Centro<br />

de Investigaciones Biologicas (CSIC), Ramiro de Maeztu 9, E-28040 Madrid, Spain (JLB)<br />

Published in <strong>Atlas</strong> Database: September 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/Deep/CohesinsID20100.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI CohesinsID20100.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Running title: Cohesins <strong>and</strong> tumorigenesis<br />

Keywords. Aneuploidy, cancer, cell cycle control, chromosome segregation, chromosome stability, cohesin,<br />

cohesin-regulators, DNA-repair, sister chromatid cohesion, tumorigenesis.<br />

Summary<br />

Cohesin is the name <strong>of</strong> a multifunctional protein complex, which was initially discovered <strong>and</strong> characterized by<br />

its role in maintain cohered sister chromatids during chromosome segregation in cell division. However, in the<br />

last years a large number <strong>of</strong> studies <strong>and</strong> results have evidenced the implication <strong>of</strong> cohesin complexes in different<br />

crucial processes in cell life such as DNA replication, control <strong>of</strong> gene expression, heterochromatin formation,<br />

<strong>and</strong> DNA-repair. The canonical cohesin complex consists in four subunits named SMC1, SMC3 (Structural<br />

Maintenance <strong>of</strong> Chromosomes) <strong>and</strong> SCC1, SCC3 (Sister Chromatid Cohesion). Two last subunits have also<br />

denominated RAD21 for SCC1 <strong>and</strong> STAG for SCC3 in mammals. These four subunits, named also cohesin<br />

individually, are able to form a ring-like structure (figure 1A), which could modulate different local chromatin<br />

conformations depending on the interactions with diverse cohesin-interacting proteins, which have been<br />

designated as cohesin c<strong>of</strong>actors <strong>and</strong>/or cohesin-regulators.<br />

Chromosomal instability is one <strong>of</strong> the hallmarks <strong>of</strong> cancer, generating chromosomic aberrations including<br />

aneuploidy, loss <strong>of</strong> heterozygosity, chromosomal translocations etc. Mutations in genes encoding for proteins<br />

that control cell cycle are potential c<strong>and</strong>idates in the generation <strong>of</strong> genome instability. Thus, cohesins <strong>and</strong><br />

cohesin-regulators, which are key players in chromosome segregation <strong>and</strong> in DNA-damage repair, are obviously<br />

aspirant molecules for this research.<br />

Basic concepts <strong>of</strong> cohesins in<br />

chromosome segregation <strong>and</strong><br />

DNA-damage repair<br />

Chromosome miss-segregation <strong>and</strong> aneuploidy are<br />

frequently observed in most <strong>of</strong> cancer cells. Perhaps<br />

the two cellular mechanisms more critical for<br />

chromosomal stability in which cohesins are<br />

involved are chromosome segregation during cell<br />

division <strong>and</strong> DNA-damage repair, therefore, in this<br />

paper, I will center on the cohesins <strong>and</strong> cohesininteracting<br />

proteins involved in these two cellular<br />

important processes <strong>and</strong> their links with tumor<br />

formation <strong>and</strong> development.<br />

Although the multiple roles <strong>of</strong> cohesins remodeling<br />

chromatin structure have been extensively <strong>and</strong><br />

detailed revised in the last time (for two example<br />

reviews see Barbero, 2009 <strong>and</strong> Nasmyth <strong>and</strong><br />

Haering, 2009), following, I describe here briefly<br />

the most relevant concepts <strong>of</strong> cohesin dynamic in<br />

order to underst<strong>and</strong> their potential involvement in<br />

tumorigenesis.<br />

Cohesin complexes preserve sister chromatid<br />

cohesion during cell division in mitosis <strong>and</strong> meiosis<br />

by binding along the arm <strong>and</strong> centromeres <strong>of</strong><br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 155


Cohesins <strong>and</strong> cohesin-regulators: Role in Chromosome<br />

Segregation/Repair <strong>and</strong> Potential in Tumorigenesis<br />

chromosomes. For this function, cohesin needs to<br />

the action <strong>of</strong> other proteins; among the best<br />

characterized <strong>of</strong> these c<strong>of</strong>actors are the followings:<br />

the adherin complex formed by SCC2 <strong>and</strong> SCC4<br />

proteins is required for the loading <strong>of</strong> cohesin<br />

complexes to chromosomes; the cohesion<br />

establishment/maintenance proteins Eco1<br />

acetyltransferase in yeast <strong>and</strong> the mammalian<br />

homologues ESCO1 <strong>and</strong> ESCO2, which, by<br />

acetylation <strong>of</strong> the SMC3 subunit <strong>of</strong> cohesin<br />

complex, stabilizes the binding <strong>of</strong> cohesin to<br />

chromatin; PDS5A <strong>and</strong> B (Precocious Dissociation<br />

<strong>of</strong> Sister), WAPL (Wings Apart-Like) <strong>and</strong><br />

SORORIN are proteins involved in the maintenance<br />

<strong>of</strong> cohesion through their interaction with cohesin<br />

complex subunits <strong>and</strong>/or with other cohesinregulators<br />

(figures 1B <strong>and</strong> 2A); Shugoshin ("the<br />

Barbero JL<br />

guardian <strong>of</strong> the spirit" in Japanese) protein family<br />

SGO1 <strong>and</strong> SGO2, which are essentially implicated<br />

in the protection <strong>of</strong> sister chromatid cohesion at the<br />

centromeric regions <strong>and</strong>, finally, the cohesion<br />

removal proteins PLK1 (Polo Like Kinase 1),<br />

Aurora B <strong>and</strong> securin/separase complex (figure<br />

2A). The two first molecules are protein-kinases<br />

that phosphorylate cohesin, essentially STAG<br />

subunit, triggering the removal <strong>of</strong> arm cohesins.<br />

Separase is a specific protease, which is inhibited<br />

by its c<strong>of</strong>actor securin; activation <strong>of</strong> the anaphase<br />

promoting complex/cyclosome (APC/C) leads to<br />

ubiquitination <strong>and</strong> degradation <strong>of</strong> securin, allowing<br />

cleavage <strong>of</strong> SCC1/RAD21 from centromeric<br />

cohesin complexes by separase <strong>and</strong> triggering the<br />

onset <strong>of</strong> anaphase.<br />

Figure 1. Cohesin complex <strong>and</strong> cohesin-regulators. A. Ring model <strong>of</strong> cohesin complex formed by four subunits SMC1,<br />

SMC3, RAD21/SCC1 <strong>and</strong> STAG/SCC3. Subunits SMC1, SMC3 <strong>and</strong> RAD21 conform the ring-like structure. STAG protein<br />

interacts with RAD21 to complete the cohesin complex. B. Examples <strong>of</strong> cohesin-regulators. PDS5 <strong>and</strong> WAPL form a protein<br />

complex, which is associated to the cohesin complex by STAG interaction. Sororin is other cohesin-regulator that is also involved<br />

in the control <strong>of</strong> cohesin ring dynamic.<br />

Figure 2. Scheme <strong>of</strong> cohesin metabolism during chromosome segregation <strong>and</strong> DNA-damage repair. A. Dynamic <strong>of</strong><br />

cohesin complex in chromosome segregation indicating the characterized cohesin-regulators involved in cohesin loading to<br />

chromosomes, sister chromatid cohesion establishment <strong>and</strong> maintenance, <strong>and</strong> cohesion dissolution during cell division. B.<br />

Involvement <strong>of</strong> cohesin complexes during DNA-damage repair. Cohesin are recruited to double str<strong>and</strong> break (DSB) areas to<br />

facilitate the repair process. Some cohesin-regulators, such us Scc2/Scc4 loading complex <strong>and</strong> Eco1 cohesion establishment<br />

factor are also required for new cohesin loading in these chromosome damaged regions. See text for more details.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 156


Cohesins <strong>and</strong> cohesin-regulators: Role in Chromosome<br />

Segregation/Repair <strong>and</strong> Potential in Tumorigenesis<br />

All these cohesin-regulators were firstly<br />

characterized studying its function in sister<br />

chromatid cohesion, but shortly after has been<br />

found evidence <strong>of</strong> its involvement in DNA repair<br />

<strong>and</strong> other cohesin tasks. The initial result on the<br />

participation <strong>of</strong> cohesins in theses mechanisms was<br />

already reported before cohesins were known to<br />

mediate sister chromatid cohesion; the Scc1<br />

ortholog from Schizosaccharomyces pombe was<br />

first identified as a protein whose mutation causes<br />

sensitivity to radiation because damaged DNA<br />

cannot be properly repaired <strong>and</strong>, thus, called Rad21<br />

(Radiation-sensitivity) (Birkenbihl <strong>and</strong> Subramani,<br />

1992). Following experiments, essentially in<br />

budding yeast, showed that cohesin mutants are<br />

defective in repair damaged DNA <strong>and</strong> provided<br />

evidence that DNA repair depends on the function<br />

<strong>of</strong> cohesin to mediate sister chromatid cohesion. In<br />

addition, studies <strong>of</strong> Rad21-depleted chicken cells<br />

have shown that vertebrate cohesin also functions<br />

in both segregation <strong>and</strong> repair (Sonoda et al., 2001).<br />

This cohesin requirement could be explained<br />

because DNA double str<strong>and</strong> breaks (DSB) are<br />

preferentially repaired by recombination between<br />

sister chromatids <strong>and</strong> the cohesion between them<br />

would facilitate this process. In this sense, cohesin<br />

complexes are recruited to sites <strong>of</strong> DSB to<br />

contribute to DNA repair <strong>of</strong> these damaged regions<br />

<strong>of</strong> chromosomes (figure 2B). This cohesin<br />

recruitment also required: from Eco1/ESCO<br />

acetyltransferase (Heidinger-Pauli et al., 2009) <strong>and</strong><br />

from a functional SCC2/SCC4 adherin complex<br />

(Ström et al., 2004). In addition to machinery <strong>of</strong><br />

chromosome cohesion, another specific DNAdamage<br />

repair proteins are necessary for the<br />

cohesin localization to DSB sites (figure 2B). A<br />

component <strong>of</strong> the DNA-damage sensing complex<br />

MRX (Mre11/Rad50/Xsr2) is required for cohesin<br />

assembly around the DSB. Phosphorylation <strong>of</strong><br />

histone H2AX by Mec1/Tel1 generates what is<br />

known as gH2AX. Yeast strains expressing a nonphosphorylatable<br />

H2AX fail to recruit cohesin, thus<br />

suggesting that gH2AX may act as a signal for<br />

cohesin assembly (Unal et al., 2004). Interestingly,<br />

some modifications by phosphorylation <strong>of</strong> SMC1<br />

<strong>and</strong> SMC3 cohesin subunits are carried out by<br />

specific kinases following IR <strong>and</strong> UV damage.<br />

Mutations in SMC1 or SMC3 that prevent<br />

phosphorylation result in abrogated DNA-damage<br />

responses (Kitagawa et al., 2004).<br />

Although, currently the participation <strong>of</strong> other<br />

chromosome cohesion cohesin-regulators in DNA<br />

repair is poorly understood, probably future<br />

research will incorporate also some <strong>of</strong> these<br />

molecules to control <strong>of</strong> DNA repair mechanisms<br />

(figure 2B).<br />

Barbero JL<br />

Cohesins: chromosome<br />

segregation, DNA-damage repair<br />

<strong>and</strong> cancer<br />

Obviously, the multiple roles <strong>of</strong> cohesins (figure 3)<br />

make it difficult to determine how their lack <strong>of</strong><br />

function contributes to the generation <strong>and</strong><br />

development <strong>of</strong> tumors <strong>and</strong> probably in many cases,<br />

were involved several processes. In this review, I<br />

focus on the two functions <strong>of</strong> cohesins most clearly<br />

related to genome stability, chromosome<br />

segregation <strong>and</strong> DNA-damage repair, <strong>and</strong> on the<br />

currently available data <strong>and</strong> results linking<br />

mutations on cohesin <strong>and</strong> cohesin-regulators genes<br />

<strong>and</strong> tumorogenesis.<br />

Figure 3. Cohesin functions. Scheme <strong>of</strong> the main cell life<br />

processes in which cohesins have been functionally<br />

characterized. The two cohesin tasks in red are the<br />

principal focus <strong>of</strong> this review.<br />

Cohesin complex subunits<br />

RAD21 cohesin subunit has long been linked with<br />

cancer chemotherapy; inhibition <strong>of</strong> RAD21<br />

expression by RNA interference in human breast<br />

cancer cells enhanced the cytotoxicity <strong>of</strong> etoposide<br />

<strong>and</strong> bleomycin in these cells (Atienza et al., 2005).<br />

In agreement with this result, Xu et al. (2011)<br />

reported that overexpression <strong>of</strong> RAD21 gives<br />

resistance to chemotherapy in high-grade luminal,<br />

basal <strong>and</strong> HER2 breast cancers. Mice lacking<br />

Rad21 gene function present early embryonal<br />

lethality, but heterozygous Rad21 +/- animals were<br />

obtained by breeding. Rad21 +/- mice were viable<br />

<strong>and</strong> developed to apparently normal adulthood<br />

without morphological defects. However, Rad21 +/-<br />

mice have enhanced sensitivity to whole body<br />

irradiation, indicating that Rad21 gene dosage is<br />

critical for ionizing radiation (IR) response (Xu et<br />

al., 2010).<br />

The study <strong>of</strong> 11 somatic mutations in 132 human<br />

colorectal cancers identified 6 <strong>of</strong> them mapping to 3<br />

cohesin, SMC1a, SMC3 <strong>and</strong> STAG3, genes <strong>and</strong> 4 to<br />

a cohesin-regulator SCC2 gene (Barber et al.,<br />

2008).<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 157


Cohesins <strong>and</strong> cohesin-regulators: Role in Chromosome<br />

Segregation/Repair <strong>and</strong> Potential in Tumorigenesis<br />

Chromosomal instability is a characteristic <strong>of</strong><br />

colorectal cancer cells, resulting in chromosome<br />

gain or loss. It is possible to argue that abnormal<br />

cohesin pathway activity leads to chromosome<br />

missegregation <strong>and</strong> chromosome instability. This<br />

hypothesis is supported by the observation that<br />

colorectal cancer cells exhibit up to 100-fold higher<br />

rates <strong>of</strong> missegregation than normal cells. In<br />

addition, using a microcell mediated chromosome<br />

transfer <strong>and</strong> expression microarray analysis,<br />

Notaridou et al. (2011) identified the cohesin<br />

subunit STAG3 gene as one <strong>of</strong> the nine genes<br />

associated with functional suppression <strong>of</strong><br />

tumorogenicity in ovarian cancer cell lines <strong>and</strong> as a<br />

c<strong>and</strong>idate gene associated with risk <strong>and</strong><br />

development <strong>of</strong> epithelial ovarian cancer. Kalejs et<br />

al. (2006) found aberrant expression <strong>of</strong> meioticspecific<br />

genes, including the meiotic specific<br />

cohesin genes REC8 <strong>and</strong> STAG3 in a lymphoma<br />

cell model.<br />

The other two members <strong>of</strong> the STAG cohesin<br />

family (STAG1 <strong>and</strong> STAG2) have been also<br />

implicated in cancer. The most frequent cause <strong>of</strong><br />

familial clear cell renal cell carcinoma (RCC) is<br />

von Hippel-Lindau disease <strong>and</strong> the VHL tumor<br />

suppressor gene (TSG) is inactivated in most<br />

sporadic clear cell RCC. To identify c<strong>and</strong>idate<br />

genes for renal tumorigenesis, Foster et al. (2007)<br />

characterized a translocation, t(3;6)(q22;q16.1)<br />

associated with multicentric RCC without evidence<br />

<strong>of</strong> VHL target gene dysregulation. The gene<br />

encoding for the human cohesin subunit STAG1<br />

map within close proximity to the breakpoints <strong>and</strong><br />

thus it is a c<strong>and</strong>idate gene involved in RCC. In<br />

array studies searching for genome alterations in a<br />

series <strong>of</strong> 167 malignant myeloid diseases, Rocquain<br />

et al. (2010) found recurrent deletions <strong>of</strong> RAD21<br />

<strong>and</strong> STAG2 genes, suggesting that cohesin<br />

components are new players in leukemogenesis.<br />

Chromosomal translocations are also frequently<br />

found in different cancer cells. The results studying<br />

a novel non-TCR chromosome translocations<br />

t(3;11)(q25;p13) <strong>and</strong> t(X;11)(q25;p13) activating<br />

LMO2 by juxtaposition with MBNL1 <strong>and</strong> STAG2<br />

are consistent with LMO2 upregulation via capture<br />

<strong>of</strong> MBNL1 or STAG2 regulatory elements effected<br />

by t(3;11) or t(X;11), respectively (Chen et al.,<br />

2011). With the aim to identify genomic alterations,<br />

associated with exposure to radiation, Hess et al.<br />

(2011) used array comparative genomic<br />

hybridization to analyze a main (n=52) <strong>and</strong> a<br />

validation cohort (n=28) <strong>of</strong> PTC from patients aged<br />


Cohesins <strong>and</strong> cohesin-regulators: Role in Chromosome<br />

Segregation/Repair <strong>and</strong> Potential in Tumorigenesis<br />

al., 2011), linking again lost <strong>of</strong> chromosome<br />

cohesion with genomic instability.<br />

WAPL was also identified as an oncogene in uterine<br />

cervical cancer <strong>and</strong> it is induced by human<br />

papillomavirus (HPV) E6 <strong>and</strong> E7 oncoproteins.<br />

WAPL overexpression induces apparition <strong>of</strong><br />

multinucleated cells <strong>and</strong> increases the number <strong>of</strong><br />

chromatid breaks in the cell, contributing to<br />

molecular mechanisms <strong>of</strong> tumor progression from<br />

HPV-infected cells to cervical carcinoma<br />

(Ohbayashi et al., 2007). Later, these authors<br />

reported that human WAPL gene encodes a large<br />

number <strong>of</strong> spliced variants <strong>and</strong> that the expression<br />

patterns <strong>of</strong> these variants could have diagnostic<br />

potential for cervical lesions (Oikawa et al., 2008).<br />

Sororin, also known as cell division cycle<br />

associated 5 (CDCA5) protein, has been recently<br />

identified as an up-regulated gene in mostly lung<br />

cancers using a cDNA array containing 27648<br />

genes or expressed sequence tags (Nguyen et al.,<br />

2010). Sororin is phosphorylated by extracellular<br />

signal-regulated kinase (ERK) at Ser79 <strong>and</strong> Ser209<br />

in vivo. The suppression <strong>of</strong> sororin expression by<br />

siRNAs or the inhibition <strong>of</strong> the interaction between<br />

sororin <strong>and</strong> ERK inhibited the growth <strong>of</strong> lung<br />

cancer cells indicating a functional role <strong>of</strong><br />

activation <strong>of</strong> CDCA5/sororin in lung cell cancer<br />

proliferation.<br />

To investigate the putative role <strong>of</strong> the centromere<br />

cohesion guardian shugoshin 1 (SGO1) in human<br />

colorectal cancer, Iwaizumi et al. (2009) performed<br />

SGO1 knockdown using shRNA expression vector.<br />

Human SGO1 knockdown cells proliferated slowly<br />

<strong>and</strong> presented marked <strong>of</strong> chromosomal instability<br />

(CIN) in the form <strong>of</strong> aneuploidy. Other<br />

characteristics <strong>of</strong> these transfected cells were<br />

increased centrosome amplification, the presence <strong>of</strong><br />

binucleated cells, <strong>and</strong> mitotic catastrophes. The<br />

results <strong>of</strong> this study showed that SGO1 downregulation<br />

leads CIN in human colorectal cancer<br />

cells <strong>and</strong> it could be a molecule involved in the CIN<br />

pathway found in colorectal cancer progression.<br />

Cohesin-regulators: cohesion<br />

dissolution<br />

The complex separase <strong>and</strong> its inhibitor securin are<br />

responsible for the total dissolution <strong>of</strong> sister<br />

chromatid cohesion in anaphase. Mammalian<br />

securin gene was originally identified in 1997 <strong>and</strong><br />

characterized as pituitary tumor-transforming gene<br />

(Pttg1), which encodes the PTTG protein, from rat<br />

pituitary tumor cells (Pei <strong>and</strong> Melmed, 1997).<br />

PTTG/securin is highly expressed in various tumors<br />

<strong>and</strong> it can induce human cellular transformation.<br />

PTTG/securin is associated with more aggressive<br />

tumor behavior <strong>and</strong> has been identified as one <strong>of</strong> 17<br />

key signature genes associated with metastatic<br />

disease (Ramaswamy et al., 2003). In addition, a<br />

PTTG binding factor (PBF) was identified through<br />

its interaction with PTTG <strong>and</strong> it was characterized<br />

Barbero JL<br />

as a proto-oncogene that is upregulated in several<br />

cancers (Smith et al., 2010). PTTG1/securin is also<br />

overexpressed in hepatocellular carcinoma. Chronic<br />

infection with hepatitis B virus (HBV) is the main<br />

causal factor for hepatocellular carcinoma <strong>and</strong> the<br />

viral protein HBx plays an essential role in the<br />

pathogenesis <strong>of</strong> hepatic tumors. To investigate the<br />

putative correlation between the abnormal<br />

expression <strong>of</strong> PTTG1 <strong>and</strong> the tumorigenic<br />

mechanism <strong>of</strong> HBx, Molina-Jiménez et al. (2010)<br />

analyzed the PTTG1 expression in biopsies from<br />

patients chronically infected with HBV in different<br />

disease stages <strong>and</strong> from HBx transgenic mouse<br />

model. These authors found that HBx viral protein<br />

promotes an accumulation <strong>of</strong> PTTG1 by inhibition<br />

<strong>of</strong> PTTG1 ubiquitination <strong>and</strong> degradation. The<br />

molecular mechanism/s by which HBx carried out<br />

this inhibition is currently under research.<br />

Separase is the endopeptidase that cleaves RAD21<br />

cohesin subunit during to metaphase/anaphase<br />

transition causing the removal <strong>of</strong> cohesins <strong>and</strong> the<br />

separation <strong>of</strong> sister chromatids. Overexpression <strong>of</strong><br />

separase induces premature separation <strong>of</strong><br />

chromatids, lagging chromosomes, <strong>and</strong> anaphase<br />

bridges. In a mouse mammary transplant model,<br />

induction <strong>of</strong> separase expression in the transplanted<br />

FSK3 cells for 3-4 weeks results in the formation <strong>of</strong><br />

aneuploid tumors in the mammary gl<strong>and</strong> (Zhang et<br />

al., 2008). In a later report, Meyer et al. (2009)<br />

showed that separase is significantly overexpressed<br />

in osteosarcoma, breast, <strong>and</strong> prostate tumor<br />

specimens. There is a strong correlation <strong>of</strong> tumor<br />

status with the localization <strong>of</strong> separase into the<br />

nucleus throughout all stages <strong>of</strong> the cell cycle. In<br />

addition, overexpression <strong>of</strong> separase transcript<br />

strongly correlates with high incidence <strong>of</strong> relapse,<br />

metastasis, <strong>and</strong> lower 5-year overall survival rate in<br />

breast <strong>and</strong> prostate cancer patients, suggesting that<br />

separase is an oncogene.<br />

Polo-like kinase 1 (PLK1) <strong>and</strong> Aurora B are two<br />

protein-kinases that have as substrates cohesins <strong>and</strong><br />

other proteins involved in chromosome segregation.<br />

PlK1 is overexpressed in various human cancers,<br />

<strong>and</strong> this is mostly associated with poor prognosis<br />

(Strebhardt <strong>and</strong> Ullrich, 2006). The first data to<br />

associate PLK1 with neoplastic growth were<br />

generated by studies showing that PLK1<br />

concentrations are also increased in primary cancer<br />

tissues (Holtrich et al., 1994). This prompted a<br />

number <strong>of</strong> studies that subsequently demonstrated<br />

that PLK1 is overexpressed in a broad spectrum <strong>of</strong><br />

human tumors compared with normal controls.<br />

Furthermore, some reports have indicated that<br />

PLK1 expression is a reliable marker for<br />

identifying a high risk <strong>of</strong> metastasis (Dai et al.,<br />

2000). More recently, Ito el al. (2010) described the<br />

post-transcriptional regulation <strong>of</strong> Plk1 expression<br />

by RNA interference mediates by miR-593* <strong>and</strong><br />

Plk1 downregulation in EC cells decreases cell<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 159


Cohesins <strong>and</strong> cohesin-regulators: Role in Chromosome<br />

Segregation/Repair <strong>and</strong> Potential in Tumorigenesis<br />

proliferation in vitro via G2/M cell cycle arrest, <strong>and</strong><br />

drastically suppresses tumor formation in vivo.<br />

Aurora B kinase is involved in different key<br />

functions during chromosome segregation to<br />

preserve genomic stability. Examples <strong>of</strong> these<br />

functions are: sister chromatid cohesion,<br />

chromosome condensation, mitotic spindle<br />

assembly, syntelic chromosome attachments <strong>and</strong><br />

spindle assembly checkpoint (for a review see<br />

Vader <strong>and</strong> Lens, 2008). Aurora B is overexpressed<br />

in cancer cells, <strong>and</strong> an increased level <strong>of</strong> Aurora B<br />

correlates with advanced stages <strong>of</strong> colorectal<br />

cancer. Overexpression <strong>of</strong> Aurora B results in<br />

multi-nucleation <strong>and</strong> polyploidy in human cells<br />

(Tatsuka et al., 1998) <strong>and</strong>, additionally, it has been<br />

reported that Aurora B overexpression induces<br />

chromosomes lagging in metaphase, chromosome<br />

segregation error, <strong>and</strong> errors in cytokinesis, <strong>and</strong><br />

thus suggesting a direct link between Aurora B <strong>and</strong><br />

carcinogenesis (Ota et al., 2002). These findings<br />

<strong>and</strong> the crucial roles <strong>of</strong> Aurora B <strong>and</strong> PLK1 in<br />

chromosome dynamics during cell cycle have led to<br />

consider these two kinases as important targets for<br />

cancer therapy (de Cárcer et al., 2007; Strebhardt,<br />

2010; Libertini et al., 2010).<br />

Concluding remarks<br />

Cohesin complex, initially characterized as a ring<br />

protein complex that maintains sister chromatids<br />

together during chromosome segregation, is now<br />

considered a real architect <strong>of</strong> chromatin structure<br />

during essential dynamic DNA processes. In many<br />

cases, these processes are designed to safeguard the<br />

stability <strong>of</strong> genetic material <strong>and</strong> its proper<br />

distribution to the daughter cells. Thus, it is not<br />

surprising that when there are errors/problems in<br />

the cohesin complex metabolism related with this<br />

guardian function, one <strong>of</strong> the likely results was the<br />

formation <strong>of</strong> a tumor. Although this review focuses<br />

essentially on the role <strong>of</strong> cohesins in chromosome<br />

segregation <strong>and</strong> DNA-damage repair <strong>and</strong> their<br />

connection with tumorigenesis, other functions <strong>of</strong><br />

cohesins are also possibly related with the<br />

development <strong>of</strong> human tumors. In this sense,<br />

recently Baysal et al. 2011, described that germ line<br />

mutations in SDHD, a mitochondrial complex II<br />

(succinate dehydrogenase) subunit gene at<br />

chromosome b<strong>and</strong> 11q23, cause highly penetrant<br />

paraganglioma tumors when transmitted through<br />

fathers. In contrast, maternal transmission rarely, if<br />

ever, leads to tumor development. They observed<br />

that hypermethylated adrenal tissues show<br />

increased binding <strong>of</strong> the chromatin-looping factor<br />

cohesin relative to the hypomethylated tissues,<br />

suggesting that this differential allelic interaction<br />

may result in maternal downregulation <strong>of</strong> SDHD<br />

<strong>and</strong> the parent-<strong>of</strong>-origin dependent tumor<br />

susceptibility.<br />

In recent years, an increasing number <strong>of</strong> scientific<br />

works showed that cohesin functions are mediated<br />

by the action <strong>of</strong> other proteins. These molecules can<br />

Barbero JL<br />

be subdivided into two kind <strong>of</strong> cohesin-interacting<br />

proteins: those that regulate different aspects <strong>of</strong> the<br />

cohesin metabolism <strong>and</strong> necessary for several<br />

functions (such as SCC2/SCC4, Eco1/ESCO) <strong>and</strong><br />

those that contribute to one cohesin specific role by<br />

a spatial-temporal interaction with cohesin complex<br />

(such as CCCTC-binding factor (CTCF) <strong>and</strong><br />

MEDIATOR in the control <strong>of</strong> gene expression<br />

(Wendt et al., 2008; Kagey et al., 2010)). In<br />

addition, post-translational modifications, such as<br />

acetylation <strong>and</strong> phosphorylation, in specific<br />

residues <strong>of</strong> cohesin subunits are also required for<br />

specific cohesin functions, suggesting the putative<br />

existence <strong>of</strong> a cohesin code similarly to well<br />

established histone code.<br />

All these findings point to the initial denomination<br />

<strong>of</strong> cohesin is currently very limited <strong>and</strong> therefore<br />

some authors are beginning to use other terms, such<br />

us chromatin-looping factor (Baysal et al., 2011)<br />

or, from my point <strong>of</strong> view, more convenient<br />

chromosome architectins, which, by interaction<br />

with specific regulator proteins, model precise<br />

tridimensional structures at local regions <strong>of</strong><br />

chromosomes to perform different <strong>and</strong> specific<br />

functions depending on the spatial-temporal<br />

requirements <strong>of</strong> cell life. The future research on the<br />

molecular mechanisms <strong>of</strong> both, the cohesininteracting<br />

proteins <strong>and</strong> the specific cohesin posttranslational<br />

modifications, <strong>and</strong> on the alterations in<br />

the cohesin network during pathological conditions<br />

is crucial in determining the relationships between<br />

this interesting ring protein complex <strong>and</strong> the<br />

formation/development <strong>of</strong> tumors in humans.<br />

Acknowledgments<br />

We thank Dr. Adela Calvente for her help in the<br />

figure design <strong>and</strong> critical comments. We apologize<br />

to all colleagues whose important contributions<br />

have not been referenced due to space restrictions.<br />

This work was supported by the Spanish Ministerio<br />

de Ciencia e Innovación (grant BFU2009-<br />

08975/BMC) <strong>and</strong> CSIC (grant PIE-201120E020).<br />

References<br />

Birkenbihl RP, Subramani S. Cloning <strong>and</strong> characterization<br />

<strong>of</strong> rad21 an essential gene <strong>of</strong> Schizosaccharomyces<br />

pombe involved in DNA double-str<strong>and</strong>-break repair.<br />

Nucleic Acids Res. 1992 Dec 25;20(24):6605-11<br />

Holtrich U, Wolf G, Bräuninger A, Karn T, Böhme B,<br />

Rübsamen-Waigmann H, Strebhardt K. Induction <strong>and</strong><br />

down-regulation <strong>of</strong> PLK, a human serine/threonine kinase<br />

expressed in proliferating cells <strong>and</strong> tumors. Proc Natl Acad<br />

Sci U S A. 1994 Mar 1;91(5):1736-40<br />

Pei L, Melmed S. Isolation <strong>and</strong> characterization <strong>of</strong> a<br />

pituitary tumor-transforming gene (PTTG). Mol Endocrinol.<br />

1997 Apr;11(4):433-41<br />

Tatsuka M, Katayama H, Ota T, Tanaka T, Odashima S,<br />

Suzuki F, Terada Y. Multinuclearity <strong>and</strong> increased ploidy<br />

caused by overexpression <strong>of</strong> the aurora- <strong>and</strong> Ipl1-like<br />

midbody-associated protein mitotic kinase in human<br />

cancer cells. Cancer Res. 1998 Nov 1;58(21):4811-6<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 160


Cohesins <strong>and</strong> cohesin-regulators: Role in Chromosome<br />

Segregation/Repair <strong>and</strong> Potential in Tumorigenesis<br />

Dai W, Li Y, Ouyang B, Pan H, Reissmann P, Li J, Wiest J,<br />

Stambrook P, Gluckman JL, N<strong>of</strong>fsinger A, Bejarano P.<br />

PRK, a cell cycle gene localized to 8p21, is downregulated<br />

in head <strong>and</strong> neck cancer. <strong>Gene</strong>s Chromosomes Cancer.<br />

2000 Mar;27(3):332-6<br />

Pezzi N, Prieto I, Kremer L, Pérez Jurado LA, Valero C,<br />

Del Mazo J, Martínez-A C, Barbero JL. STAG3, a novel<br />

gene encoding a protein involved in meiotic chromosome<br />

pairing <strong>and</strong> location <strong>of</strong> STAG3-related genes flanking the<br />

Williams-Beuren syndrome deletion. FASEB J. 2000<br />

Mar;14(3):581-92<br />

Sonoda E, Matsusaka T, Morrison C, Vagnarelli P, Hoshi<br />

O, Ushiki T, Nojima K, Fukagawa T, Waizenegger IC,<br />

Peters JM, Earnshaw WC, Takeda S. Scc1/Rad21/Mcd1 is<br />

required for sister chromatid cohesion <strong>and</strong> kinetochore<br />

function in vertebrate cells. Dev Cell. 2001 Dec;1(6):759-<br />

70<br />

Ota T, Suto S, Katayama H, Han ZB, Suzuki F, Maeda M,<br />

Tanino M, Terada Y, Tatsuka M. Increased mitotic<br />

phosphorylation <strong>of</strong> histone H3 attributable to AIM-1/Aurora-<br />

B overexpression contributes to chromosome number<br />

instability. Cancer Res. 2002 Sep 15;62(18):5168-77<br />

Ramaswamy S, Ross KN, L<strong>and</strong>er ES, Golub TR. A<br />

molecular signature <strong>of</strong> metastasis in primary solid tumors.<br />

Nat <strong>Gene</strong>t. 2003 Jan;33(1):49-54<br />

Kitagawa R, Bakkenist CJ, McKinnon PJ, Kastan MB.<br />

Phosphorylation <strong>of</strong> SMC1 is a critical downstream event in<br />

the ATM-NBS1-BRCA1 pathway. <strong>Gene</strong>s Dev. 2004 Jun<br />

15;18(12):1423-38<br />

Ström L, Lindroos HB, Shirahige K, Sjögren C.<br />

Postreplicative recruitment <strong>of</strong> cohesin to double-str<strong>and</strong><br />

breaks is required for DNA repair. Mol Cell. 2004 Dec<br />

22;16(6):1003-15<br />

Unal E, Arbel-Eden A, Sattler U, Shr<strong>of</strong>f R, Lichten M,<br />

Haber JE, Koshl<strong>and</strong> D. DNA damage response pathway<br />

uses histone modification to assemble a double-str<strong>and</strong><br />

break-specific cohesin domain. Mol Cell. 2004 Dec<br />

22;16(6):991-1002<br />

Atienza JM, Roth RB, Rosette C, Smylie KJ, Kammerer S,<br />

Rehbock J, Ekblom J, Denissenko MF. Suppression <strong>of</strong><br />

RAD21 gene expression decreases cell growth <strong>and</strong><br />

enhances cytotoxicity <strong>of</strong> etoposide <strong>and</strong> bleomycin in<br />

human breast cancer cells. Mol Cancer Ther. 2005<br />

Mar;4(3):361-8<br />

Kalejs M, Ivanov A, Plakhins G, Cragg MS, Emzinsh D,<br />

Illidge TM, Erenpreisa J. Upregulation <strong>of</strong> meiosis-specific<br />

genes in lymphoma cell lines following genotoxic insult <strong>and</strong><br />

induction <strong>of</strong> mitotic catastrophe. BMC Cancer. 2006 Jan<br />

9;6:6<br />

Strebhardt K, Ullrich A. Targeting polo-like kinase 1 for<br />

cancer therapy. Nat Rev Cancer. 2006 Apr;6(4):321-30<br />

de Cárcer G, Pérez de Castro I, Malumbres M. Targeting<br />

cell cycle kinases for cancer therapy. Curr Med Chem.<br />

2007;14(9):969-85<br />

Foster RE, Abdulrahman M, Morris MR, Prigmore E,<br />

Gribble S, Ng B, Gentle D, Ready S, Weston PM,<br />

Wiesener MS, Kishida T, Yao M, Davison V, Barbero JL,<br />

Chu C, Carter NP, Latif F, Maher ER. Characterization <strong>of</strong> a<br />

3;6 translocation associated with renal cell carcinoma.<br />

<strong>Gene</strong>s Chromosomes Cancer. 2007 Apr;46(4):311-7<br />

Ohbayashi T, Oikawa K, Yamada K, Nishida-Umehara C,<br />

Matsuda Y, Satoh H, Mukai H, Mukai K, Kuroda M.<br />

Unscheduled overexpression <strong>of</strong> human WAPL promotes<br />

chromosomal instability. Biochem Biophys Res Commun.<br />

2007 May 11;356(3):699-704<br />

Barbero JL<br />

Ryu B, Kim DS, Deluca AM, Alani RM. Comprehensive<br />

expression pr<strong>of</strong>iling <strong>of</strong> tumor cell lines identifies molecular<br />

signatures <strong>of</strong> melanoma progression. PLoS One. 2007 Jul<br />

4;2(7):e594<br />

Barber TD, McManus K, Yuen KW, Reis M, Parmigiani G,<br />

Shen D, Barrett I, Nouhi Y, Spencer F, Markowitz S,<br />

Velculescu VE, Kinzler KW, Vogelstein B, Lengauer C,<br />

Hieter P. Chromatid cohesion defects may underlie<br />

chromosome instability in human colorectal cancers. Proc<br />

Natl Acad Sci U S A. 2008 Mar 4;105(9):3443-8<br />

Oikawa K, Akiyoshi A, Tanaka M, Takanashi M, Nishi H,<br />

Isaka K, Kiseki H, Idei T, Tsukahara Y, Hashimura N,<br />

Mukai K, Kuroda M. Expression <strong>of</strong> various types <strong>of</strong><br />

alternatively spliced WAPL transcripts in human cervical<br />

epithelia. <strong>Gene</strong>. 2008 Oct 15;423(1):57-62<br />

Vader G, Lens SM. The Aurora kinase family in cell<br />

division <strong>and</strong> cancer. Biochim Biophys Acta. 2008<br />

Sep;1786(1):60-72<br />

Wendt KS, Yoshida K, Itoh T, B<strong>and</strong>o M, Koch B,<br />

Schirghuber E, Tsutsumi S, Nagae G, Ishihara K, Mishiro<br />

T, Yahata K, Imamoto F, Aburatani H, Nakao M, Imamoto<br />

N, Maeshima K, Shirahige K, Peters JM. Cohesin<br />

mediates transcriptional insulation by CCCTC-binding<br />

factor. Nature. 2008 Feb 14;451(7180):796-801<br />

Yamamoto F, Yamamoto M. Identification <strong>of</strong> genes that<br />

exhibit changes in expression on the 8p chromosomal arm<br />

by the Systematic Multiplex RT-PCR (SM RT-PCR) <strong>and</strong><br />

DNA microarray hybridization methods. <strong>Gene</strong> Expr.<br />

2008;14(4):217-27<br />

Zhang N, Ge G, Meyer R, Sethi S, Basu D, Pradhan S,<br />

Zhao YJ, Li XN, Cai WW, El-Naggar AK,<br />

Balad<strong>and</strong>ayuthapani V, Kittrell FS, Rao PH, Medina D, Pati<br />

D. Overexpression <strong>of</strong> Separase induces aneuploidy <strong>and</strong><br />

mammary tumorigenesis. Proc Natl Acad Sci U S A. 2008<br />

Sep 2;105(35):13033-8<br />

Barbero JL. Cohesins: chromatin architects in<br />

chromosome segregation, control <strong>of</strong> gene expression <strong>and</strong><br />

much more. Cell Mol Life Sci. 2009 Jul;66(13):2025-35<br />

Heidinger-Pauli JM, Unal E, Koshl<strong>and</strong> D. Distinct targets <strong>of</strong><br />

the Eco1 acetyltransferase modulate cohesion in S phase<br />

<strong>and</strong> in response to DNA damage. Mol Cell. 2009 May<br />

15;34(3):311-21<br />

Iwaizumi M, Shinmura K, Mori H, Yamada H, Suzuki M,<br />

Kitayama Y, Igarashi H, Nakamura T, Suzuki H, Watanabe<br />

Y, Hishida A, Ikuma M, Sugimura H. Human Sgo1<br />

downregulation leads to chromosomal instability in<br />

colorectal cancer. Gut. 2009 Feb;58(2):249-60<br />

Meyer R, F<strong>of</strong>anov V, Panigrahi A, Merchant F, Zhang N,<br />

Pati D. Overexpression <strong>and</strong> mislocalization <strong>of</strong> the<br />

chromosomal segregation protein separase in multiple<br />

human cancers. Clin Cancer Res. 2009 Apr 15;15(8):2703-<br />

10<br />

Nasmyth K, Haering CH. Cohesin: its roles <strong>and</strong><br />

mechanisms. Annu Rev <strong>Gene</strong>t. 2009;43:525-58<br />

Denes V, Pilichowska M, Makarovskiy A, Carpinito G,<br />

Geck P. Loss <strong>of</strong> a cohesin-linked suppressor APRIN<br />

(Pds5b) disrupts stem cell programs in embryonal<br />

carcinoma: an emerging cohesin role in tumor<br />

suppression. Oncogene. 2010 Jun 10;29(23):3446-52<br />

Ito T, Sato F, Kan T, Cheng Y, David S, Agarwal R, Paun<br />

BC, Jin Z, Olaru AV, Hamilton JP, Selaru FM, Yang J,<br />

Matsumura N, Shimizu K, Abraham JM, Shimada Y, Mori<br />

Y, Meltzer SJ. Polo-like kinase 1 regulates cell proliferation<br />

<strong>and</strong> is targeted by miR-593* in esophageal cancer. Int J<br />

Cancer. 2010 Dec 17;<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 161


Cohesins <strong>and</strong> cohesin-regulators: Role in Chromosome<br />

Segregation/Repair <strong>and</strong> Potential in Tumorigenesis<br />

Kagey MH, Newman JJ, Bilodeau S, Zhan Y, Orl<strong>and</strong>o DA,<br />

van Berkum NL, Ebmeier CC, Goossens J, Rahl PB,<br />

Levine SS, Taatjes DJ, Dekker J, Young RA. Mediator <strong>and</strong><br />

cohesin connect gene expression <strong>and</strong> chromatin<br />

architecture. Nature. 2010 Sep 23;467(7314):430-5<br />

Libertini S, Abagnale A, Passaro C, Botta G, Portella G.<br />

Aurora A <strong>and</strong> B kinases--targets <strong>of</strong> novel anticancer drugs.<br />

Recent Pat Anticancer Drug Discov. 2010 Nov;5(3):219-41<br />

Molina-Jiménez F, Benedicto I, Murata M, Martín-Vílchez<br />

S, Seki T, Antonio Pintor-Toro J, Tortolero M, Moreno-<br />

Otero R, Okazaki K, Koike K, Barbero JL, Matsuzaki K,<br />

Majano PL, López-Cabrera M. Expression <strong>of</strong> pituitary<br />

tumor-transforming gene 1 (PTTG1)/securin in hepatitis B<br />

virus (HBV)-associated liver diseases: evidence for an<br />

HBV X protein-mediated inhibition <strong>of</strong> PTTG1 ubiquitination<br />

<strong>and</strong> degradation. Hepatology. 2010 Mar;51(3):777-87<br />

Nguyen MH, Koinuma J, Ueda K, Ito T, Tsuchiya E,<br />

Nakamura Y, Daigo Y. Phosphorylation <strong>and</strong> activation <strong>of</strong><br />

cell division cycle associated 5 by mitogen-activated<br />

protein kinase play a crucial role in human lung<br />

carcinogenesis. Cancer Res. 2010 Jul 1;70(13):5337-47<br />

Rocquain J, Gelsi-Boyer V, Adélaïde J, Murati A,<br />

Carbuccia N, Vey N, Birnbaum D, Mozziconacci MJ,<br />

Chaffanet M. Alteration <strong>of</strong> cohesin genes in myeloid<br />

diseases. Am J Hematol. 2010 Sep;85(9):717-9<br />

Smith VE, Franklyn JA, McCabe CJ. Pituitary tumortransforming<br />

gene <strong>and</strong> its binding factor in endocrine<br />

cancer. Expert Rev Mol Med. 2010 Dec 3;12:e38<br />

Strebhardt K. Multifaceted polo-like kinases: drug targets<br />

<strong>and</strong> antitargets for cancer therapy. Nat Rev Drug Discov.<br />

2010 Aug;9(8):643-60<br />

Xu H, Balakrishnan K, Malaterre J, Beasley M, Yan Y,<br />

Essers J, Appeldoorn E, Tomaszewski JM, Vazquez M,<br />

Verschoor S, Lavin MF, Bertoncello I, Ramsay RG, McKay<br />

MJ. Rad21-cohesin haploinsufficiency impedes DNA repair<br />

<strong>and</strong> enhances gastrointestinal radiosensitivity in mice.<br />

PLoS One. 2010 Aug 12;5(8):e12112<br />

Baysal BE, McKay SE, Kim YJ, Zhang Z, Alila L, Willett-<br />

Brozick JE, Pacak K, Kim TH, Shadel GS. Genomic<br />

imprinting at a boundary element flanking the SDHD locus.<br />

Hum Mol <strong>Gene</strong>t. 2011 Nov 15;20(22):4452-61<br />

Chen S, Nagel S, Schneider B, Kaufmann M, Meyer C,<br />

Zaborski M, Kees UR, Drexler HG, MacLeod RA. Novel<br />

non-TCR chromosome translocations t(3;11)(q25;p13) <strong>and</strong><br />

t(X;11)(q25;p13) activating LMO2 by juxtaposition with<br />

MBNL1 <strong>and</strong> STAG2. Leukemia. 2011 Oct;25(10):1632-5<br />

Barbero JL<br />

Hagemann C, Weigelin B, Schommer S, Schulze M, Al-<br />

Jomah N, Anacker J, Gerngras S, Kühnel S, Kessler AF,<br />

Polat B, Ernestus RI, Patel R, Vince GH. The cohesininteracting<br />

protein, precocious dissociation <strong>of</strong> sisters<br />

5A/sister chromatid cohesion protein 112, is up-regulated<br />

in human astrocytic tumors. Int J Mol Med. 2011<br />

Jan;27(1):39-51<br />

Hess J, Thomas G, Braselmann H, Bauer V, Bogdanova T,<br />

Wienberg J, Zitzelsberger H, Unger K. Gain <strong>of</strong><br />

chromosome b<strong>and</strong> 7q11 in papillary thyroid carcinomas <strong>of</strong><br />

young patients is associated with exposure to low-dose<br />

irradiation. Proc Natl Acad Sci U S A. 2011 Jun<br />

7;108(23):9595-600<br />

Notaridou M, Quaye L, Dafou D, Jones C, Song H, Høgdall<br />

E, Kjaer SK, Christensen L, Høgdall C, Blaakaer J,<br />

McGuire V, Wu AH, Van Den Berg DJ, Pike MC, Gentry-<br />

Maharaj A, Wozniak E, Sher T, Jacobs IJ, Tyrer J,<br />

Schildkraut JM, Moorman PG, Iversen ES, Jakubowska A,<br />

Mędrek K, Lubiński J, Ness RB, Moysich KB, Lurie G,<br />

Wilkens LR, Carney ME, Wang-Gohrke S, Doherty JA,<br />

Rossing MA, Beckmann MW, Thiel FC, Ekici AB, Chen X,<br />

Beesley J, Gronwald J, Fasching PA, Chang-Claude J,<br />

Goodman MT, Chenevix-Trench G, Berchuck A, Pearce<br />

CL, Whittemore AS, Menon U, Pharoah PD, Gayther SA,<br />

Ramus SJ. Common alleles in c<strong>and</strong>idate susceptibility<br />

genes associated with risk <strong>and</strong> development <strong>of</strong> epithelial<br />

ovarian cancer. Int J Cancer. 2011 May 1;128(9):2063-74<br />

Solomon DA, Kim T, Diaz-Martinez LA, Fair J, Elkahloun<br />

AG, Harris BT, Toretsky JA, Rosenberg SA, Shukla N,<br />

Ladanyi M, Samuels Y, James CD, Yu H, Kim JS,<br />

Waldman T. Mutational inactivation <strong>of</strong> STAG2 causes<br />

aneuploidy in human cancer. Science. 2011 Aug<br />

19;333(6045):1039-43<br />

Xu H, Yan M, Patra J, Natrajan R, Yan Y, Swagemakers S,<br />

Tomaszewski JM, Verschoor S, Millar EK, van der Spek P,<br />

Reis-Filho JS, Ramsay RG, O'Toole SA, McNeil CM,<br />

Sutherl<strong>and</strong> RL, McKay MJ, Fox SB. Enhanced RAD21<br />

cohesin expression confers poor prognosis <strong>and</strong> resistance<br />

to chemotherapy in high grade luminal, basal <strong>and</strong> HER2<br />

breast cancers. Breast Cancer Res. 2011 Jan 21;13(1):R9<br />

This article should be referenced as such:<br />

Barbero JL. Cohesins <strong>and</strong> cohesin-regulators: Role in<br />

Chromosome Segregation/Repair <strong>and</strong> Potential in<br />

Tumorigenesis. <strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol.<br />

2012; 16(2):155-162.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 162


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

Deep Insight <strong>Section</strong><br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in<br />

mammalian cells<br />

Maryam Mehrpour, Joëlle Botti, Patrice Codogno<br />

INSERM U984, Faculty <strong>of</strong> Pharmacy, University Paris-Sud 11, 92296 Chatenay-Malabry Cedex,<br />

France (MM, JB, PC)<br />

Published in <strong>Atlas</strong> Database: September 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/Deep/AutophagyID20102.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI AutophagyID20102.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Introduction<br />

Cell homeostasis depends on the balance between<br />

the production <strong>and</strong> destruction <strong>of</strong> macromolecules<br />

<strong>and</strong> organelles. There are two major systems in<br />

eukaryotic cells that degrade cellular components:<br />

the ubiquitin proteasome system (UPS) <strong>and</strong> the<br />

lysosome. The UPS only degrades proteins, mainly<br />

short-lived proteins that have to be tagged by<br />

ubiquitin to be recognized by the proteasome<br />

(Ciechanover et al., 2000). The lysosomal system is<br />

responsible for degrading macromolecules,<br />

including proteins, <strong>and</strong> for the turnover <strong>of</strong><br />

organelles by autophagy (Mizushima et al., 2008).<br />

Recent evidence demonstrates that cross-talk <strong>and</strong><br />

cooperation exist between the UPS <strong>and</strong> autophagy<br />

(Korolchuk et al., 2009; Lamark <strong>and</strong> Johansen,<br />

2009; Nedelsky et al., 2008). The term "autophagy"<br />

was coined by Christian de Duve soon after his<br />

discovery <strong>of</strong> lysosomes (see reference Klionsky,<br />

2007 for an historical view <strong>of</strong> autophagy). The<br />

seminal discovery <strong>of</strong> ATG (AuTophaGy) genes,<br />

originally in yeast <strong>and</strong> subsequently in multicellular<br />

organisms, has provided an important breakthrough<br />

in the underst<strong>and</strong>ing <strong>of</strong> macroautophagy <strong>and</strong> <strong>of</strong> its<br />

functions in physiology <strong>and</strong> diseases (Klionsky et<br />

al., 2003; Nakatogawa et al., 2009). However the<br />

term "autophagy" also embraces microautophagy<br />

<strong>and</strong> chaperone-mediated autophagy (Klionsky,<br />

2007) that we will briefly describe here. In contrast<br />

to macroautophagy, which starts with the formation<br />

<strong>of</strong> a vacuole, known as the autophagosome, that<br />

sequesters cytoplasmic components,<br />

microautophagy consists <strong>of</strong> the direct uptake <strong>of</strong><br />

fractions <strong>of</strong> the cytoplasm by the lysosomal<br />

membrane. Macro- <strong>and</strong> microautophagy are<br />

conserved from yeast to humans. These processes<br />

were originally described as bulk degradation<br />

mechanisms. However, selective forms <strong>of</strong><br />

macroautophagy <strong>and</strong> microautophagy target<br />

organelles (mitophagy, pexophagy, ribophagy,<br />

ERphagy, piecemeal microautophagy <strong>of</strong> the<br />

nucleus), protein aggregates (aggrephagy), lipid<br />

droplets (lipophagy), glycogen <strong>and</strong> microorganisms<br />

that invade the intracellular milieu (xenophagy)<br />

(Beau et al., 2008; Kraft et al., 2009; van der Vaart<br />

et al., 2008). Microautophagy is dependent on GTP<br />

hydrolysis <strong>and</strong> on calcium (Uttenweiler <strong>and</strong> Mayer,<br />

2008). However the molecular regulation <strong>of</strong><br />

microautophagy remains to be unraveled. Bulk<br />

microautophagy does not seem to be dependent on<br />

Atg proteins, whereas selective forms <strong>of</strong><br />

microautophagy require different sets <strong>of</strong> Atg<br />

proteins (Beau et al., 2008; Kraft et al., 2009; van<br />

der Vaart et al., 2008). Chaperone-mediated<br />

autophagy (CMA) is a selective form <strong>of</strong> autophagy<br />

that has so far only been described in mammalian<br />

cells (Cuervo, 2009). Substrates for CMA contain a<br />

KFERQ-related motif in their amino acid sequence.<br />

This motif is recognized by the cytosolic<br />

constitutive chaperone hsc70 (heat shock cognate <strong>of</strong><br />

the Hsp70 family). This recognition allows the<br />

lysosomal delivery <strong>of</strong> CMA substrates to occur.<br />

The lysosomal membrane protein, LAMP-2A,<br />

serves as a receptor in the translocation <strong>of</strong> unfolded<br />

polypeptides across the lysosomal membrane.<br />

KFERQ-like motifs are found mainly in cytosolic<br />

proteins, <strong>and</strong> it is estimated that about 30% <strong>of</strong><br />

cytosolic proteins contain this motif. CMA<br />

performs several general functions, such as the<br />

elimination <strong>of</strong> oxidazed proteins <strong>and</strong> the removal <strong>of</strong><br />

misfolded proteins, <strong>and</strong> also provides amino acids<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 163


Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells Mehrpour M, et al.<br />

during prolonged periods <strong>of</strong> starvation. It is<br />

interesting to note that cross-talk occurs between<br />

macroautophagy <strong>and</strong> CMA during starvation<br />

(Kaushik et al., 2008; Massey et al., 2006). When<br />

CMA is stimulated, macroautophagy is first<br />

induced <strong>and</strong> then declines. The molecular basis for<br />

this switch remains to be identified. Prevention <strong>of</strong><br />

the age-related decline <strong>of</strong> CMA is beneficial for the<br />

homeostasis <strong>of</strong> organs <strong>and</strong> function (Zhang <strong>and</strong><br />

Cuervo, 2008). This observation is indicative <strong>of</strong> the<br />

potential importance <strong>of</strong> CMA <strong>and</strong> macroautophagy,<br />

as we discuss below, as possible anti-aging<br />

mechanisms. CMA is also involved in more<br />

specific functions, such as antigen presentation by<br />

MHC class II molecules, neurone survival, <strong>and</strong><br />

kidney growth (Cuervo, 2009).<br />

Molecular <strong>and</strong> cellular aspects <strong>of</strong><br />

macroautophagy<br />

Autophagosome formation. Autophagy is initiated<br />

by the formation <strong>of</strong> a double membrane-bound<br />

vacuole known as the autophagosome. The size <strong>of</strong><br />

this vacuole can range from 300 to 900 nm.<br />

Autophagosomes are non-degradative vacuoles that<br />

sequester cytoplasmic material. The boundary<br />

membrane arises from a single membrane, known<br />

as the phagophore or isolation membrane (reviewed<br />

in Yang <strong>and</strong> Klionsky, 2010) (Figure 1). Once<br />

completed, autophagosomes receive inputs from the<br />

endocytic pathway, <strong>and</strong> thus acquire acidic <strong>and</strong><br />

degradative capacities. These acidic <strong>and</strong><br />

degradative vacuoles form single-membrane<br />

compartments known as amphisomes (reviewed in<br />

Yang <strong>and</strong> Klionsky, 2010). The last stage <strong>of</strong><br />

autophagy is the fusion <strong>of</strong> the amphisomes with<br />

lysosomes to degrade their autophagic cargoes <strong>and</strong><br />

to recycle nutrients (to meet the metabolic dem<strong>and</strong>)<br />

<strong>and</strong> membranes (to permit ongoing lysosomal<br />

function). Whereas the mobility <strong>of</strong> autophagic<br />

vacuoles is dependent on microtubules, fusion<br />

events between autophagic vacuoles <strong>and</strong> lysosomes<br />

seem to be independent <strong>of</strong> cytoskeletal elements. A<br />

long-st<strong>and</strong>ing question regarding autophagy was to<br />

identify the origin <strong>of</strong> the phagophore <strong>and</strong> to<br />

decipher the molecular basis <strong>of</strong> the biogenesis <strong>of</strong><br />

autophagosomes. The discovery <strong>of</strong> ATG genes in<br />

yeast was a major milestone in our underst<strong>and</strong>ing <strong>of</strong><br />

autophagy (Nakatogawa et al., 2009). More than 15<br />

Atg proteins, plus class III PI3K or hVps34, are<br />

required to construct the autophagosome. These<br />

Atg proteins are hierarchically recruited at the PAS<br />

(Nakatogawa et al., 2009). The formation <strong>of</strong> the<br />

autophagosome is a multistep process that includes<br />

the biogenesis <strong>of</strong> the isolation membrane, followed<br />

by its elongation <strong>and</strong> closure. The process also<br />

requires the shuttling <strong>of</strong> Atg9, the only<br />

transmembrane Atg, between a peripheral site <strong>and</strong><br />

the isolation membrane (Nakatogawa et al., 2009)<br />

(Figure 2).<br />

Figure 1. Schematic view <strong>of</strong> the autophagic pathway. Autophagy is initiated by the nucleation <strong>of</strong> an isolation membrane or<br />

phagophore. Several different membrane pools contribute to the formation <strong>of</strong> the phagophore. This membrane then elongates<br />

<strong>and</strong> closes on itself to form an autophagosome. In most cases, once the autophagosome has been formed it receives input from<br />

the endocytic pathway (early <strong>and</strong> late endosomes <strong>and</strong> multivesicular bodies-MVB). These steps are collectively termed<br />

maturation. The amphisomes that result from the fusion <strong>of</strong> autophagosomes with late endosomes/MVB are acidic <strong>and</strong> hydrolytic<br />

vacuoles.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 164


Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells Mehrpour M, et al.<br />

Figure 2. Regulation <strong>of</strong> autophagy <strong>and</strong> its relationship with signaling molecules <strong>and</strong> apoptotic mediators. In the<br />

presence <strong>of</strong> amino acids, growth factors <strong>and</strong> energy, the mTOR complex 1 (mTORC1) represses autophagy by inhibiting the<br />

kinase activity <strong>of</strong> ULK1. In contrast, in the absence <strong>of</strong> amino acids <strong>and</strong> growth factors, or in response to an increase in the<br />

AMP/ATP ratio (via activation <strong>of</strong> AMPK), mTORC1 is inhibited <strong>and</strong> autophagy is initiated by the ULK1 complex. In this complex,<br />

Atg13 <strong>and</strong> FIP200 are substrates for ULK1 kinase activity. We do not know whether ULK1 has any other substrates. During<br />

starvation, c-JUN N-terminal kinase 1 (JNK1) is activated. By phosphorylating Bcl-2, JNK1abolishes its inhibitory effect on the<br />

activity <strong>of</strong> the Beclin 1:hVps34:Atg14 complex. The phosphorylation <strong>of</strong> Beclin 1 by Death-associated protein kinase (DAPK) also<br />

triggers the dissociation <strong>of</strong> Bcl-2 from Beclin 1. Not shown in the Figure, BH3-containing proteins can dissociate the Beclin 1:Bcl-<br />

2 interaction by competing with the Beclin 1 BH3 domain independently <strong>of</strong> the modification <strong>of</strong> the phosphorylation status <strong>of</strong> the<br />

proteins in the complex. The activity <strong>of</strong> the Beclin 1:hVps34:Atg14 complex is important for the nucleation <strong>of</strong> the autophagosomal<br />

membrane. The functional relationship between the ULK1:Atg13:FIP200 (initiation) <strong>and</strong> Beclin 1:hVps34:Atg14 (nucleation)<br />

complexes remains to be determined. Production <strong>of</strong> PtdIns3P by hVps34 in the Beclin 1:hVps34:Atg14 complex allows the<br />

recruitment <strong>of</strong> WIPI-1 <strong>and</strong> Atg2 to occur. The expansion <strong>and</strong> closure <strong>of</strong> the autophagosomal membrane are dependent on the<br />

Atg12 <strong>and</strong> LC3 conjugation systems. The Atg12-Atg5 conjugate associated with Atg16 contributes to the stimulation <strong>of</strong> the<br />

conjugation <strong>of</strong> LC3-I to phosphatidylethanolamine (PE) to generate LC3-II. Expansion <strong>of</strong> the autophagosomal membrane is<br />

probably dependent on the supply <strong>of</strong> lipids by Atg9 that cycles between a peripheral pool <strong>and</strong> the growing isolation membrane or<br />

phagophore. The anti-apoptotic protein FLIP inhibits autophagy by interacting with Atg3. In this Figure, protein kinases with<br />

substrates in the autophagic machinery are boxed in rectangles. Mediators <strong>of</strong> apoptosis are boxed in orange. These mediators<br />

regulate autophagy at the nucleation step (Bcl-2 <strong>and</strong> DAPK) <strong>and</strong> at the expansion/closure step (FLIP).<br />

Recently strong arguments have been advanced for<br />

the role <strong>of</strong> the endoplasmic reticulum (ER) in the<br />

initiation <strong>of</strong> autophagy (Axe et al., 2008; Hayashi-<br />

Nishino et al., 2009; Ylä-Anttila et al., 2009). The<br />

ULK1 complex (a complex composed <strong>of</strong> ULK1:<br />

the mammalian ortholog <strong>of</strong> yeast Atg1, FIP200: the<br />

mammalian ortholog <strong>of</strong> the yeast Atg13, Atg17,<br />

Atg101), <strong>and</strong> the PI3K complex (a complex<br />

composed <strong>of</strong> Beclin 1: the mammalian ortholog <strong>of</strong><br />

yeast Atg6, Atg14, hVps34, hVps15 <strong>and</strong> AMBRA<br />

1) congregate at the PAS to initiate autophagy in<br />

response to nutrient starvation. The kinase activity<br />

<strong>of</strong> ULK1 is controlled by the kinase mTOR in the<br />

mTORC1 complex sensitive to rapamycin<br />

(Hosokawa et al., 2009). The protein Atg14 plays a<br />

key role in the ER targeting <strong>of</strong> the PI3K complex.<br />

How the ULK1 <strong>and</strong> PI3K complexes are<br />

coordinately regulated remains to be elucidated.<br />

The production <strong>of</strong> PtdIns3P by hVps34 recruits<br />

WIPI-1/2 (Atg18) <strong>and</strong> DFPC1, which are both<br />

PtdIns3P binding proteins. DFCP1 is located at the<br />

Golgi in resting cells, but in response to autophagy<br />

stimulation it is recruited to an ER structure known<br />

as the omegasome (Axe et al., 2008). The<br />

omegasome serves as a PAS to accommodate the<br />

two ubiquitin-like conjugation systems (Atg12-<br />

Atg5, Atg16 <strong>and</strong> LC3-II, the<br />

phosphatidylethanolamine containing LC3, the<br />

mammalian ortholog <strong>of</strong> the yeast Atg8) that act<br />

sequentially to elongate the phagophore membrane<br />

<strong>and</strong> thus form the autophagosome (Nakatogawa et<br />

al., 2009). More recently it has been suggested that<br />

the mitochondrial outer membrane may be another<br />

source <strong>of</strong> the isolation membrane (Hailey et al.,<br />

2010). According to this scenario, the<br />

mitochondria-ER contact site provides the growing<br />

phagophore with lipids. The plasma membrane,<br />

through Atg16L1 decorated vesicles derived from<br />

coated pits, is also a source <strong>of</strong> membrane for the<br />

phagophore (Ravikumar et al., 2010a). Finally,<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 165


Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells Mehrpour M, et al.<br />

Golgi apparatus <strong>and</strong> post-Golgi compartments<br />

containing Atg9 also contribute to the formation <strong>of</strong><br />

the autophagosome membrane (Mari et al., 2010;<br />

Ohashi <strong>and</strong> Munro, 2010). Whatever the origin <strong>of</strong><br />

the membrane, Atg proteins are retrieved from the<br />

autophagosome membrane after closure, with the<br />

exception <strong>of</strong> a fraction <strong>of</strong> LC3-II, which is<br />

transported into the lysosomal compartment (Yang<br />

<strong>and</strong> Klionsky, 2010). Following on from this<br />

discovery, several methods based on the analysis <strong>of</strong><br />

the LC3 protein have been developed to monitor<br />

autophagy (Mizushima et al., 2010).<br />

The role <strong>of</strong> LC3 <strong>and</strong> <strong>of</strong> other members <strong>of</strong> the<br />

mammalian Atg8 family (GABARAPs) remains to<br />

be fully elucidated. However, a recent study shows<br />

that LC3s <strong>and</strong> GABARAPs are involved in<br />

different steps <strong>of</strong> autophagosome biogenesis<br />

(Weidberg et al., 2010). LC3s mediate the<br />

elongation <strong>of</strong> the autophagic membrane, <strong>and</strong><br />

GABARAPs mediate a downstream event probably<br />

associated with the closure <strong>of</strong> the autophagosome<br />

membrane. Atg8 proteins induce membrane fusion,<br />

which is involved in autophagy (Nakatogawa et al.,<br />

2007; Weidberg et al., 2011a). However, the<br />

biogenesis <strong>of</strong> autophagosomes has also been shown<br />

to involve SNAREs in yeast <strong>and</strong> mammalian cells<br />

(Moreau et al., 2011; Nair et al., 2011). Atg8<br />

proteins can also serve as a scaffold for recruiting<br />

proteins that may regulate events upstream <strong>and</strong><br />

downstream <strong>of</strong> the formation <strong>of</strong> autophagosomes<br />

(Garcia-Marcos et al., 2011; Itoh et al., 2011;<br />

Mauvezin et al., 2010). Moreover, LC3 contributes<br />

to the selectivity shown by autophagy towards<br />

different cell structures, protein aggregates <strong>and</strong><br />

microorganisms via the recognition <strong>of</strong> an LIR (LC3<br />

Interacting Region) on target proteins such as P62<br />

<strong>and</strong> NBR1 (Noda et al., 2008; Kraft et al., 2009;<br />

Weidberg et al., 2011b).<br />

After their formation, autophagosomes can merge<br />

with endocytic compartments (early <strong>and</strong> late<br />

endosomes, multivesicular bodies can merge with<br />

autophagosomes) before fusing with the lysosomal<br />

compartment (Liou et al., 1997; Razi et al., 2009;<br />

Stromhaug <strong>and</strong> Seglen, 1993). The term<br />

"amphisome" (from the Greek roots, amphi: both<br />

<strong>and</strong> soma: body) has been coined by Per O. Seglen<br />

(reviewed in Fengsrud et al., 2004) for the vacuole<br />

that results from the fusion <strong>of</strong> an autophagosome<br />

with an endosome. The late stage <strong>of</strong> autophagy<br />

depends on molecules that regulate the maturation<br />

<strong>of</strong> autophagosomes, including their fusion with<br />

endosomes <strong>and</strong> lysosomes, as well as on the<br />

acidification <strong>of</strong> the autophagic compartments, <strong>and</strong><br />

the recycling <strong>of</strong> metabolites from the lysosomal<br />

compartment (Figure 1). These steps are <strong>of</strong> a<br />

fundamental importance for the flux (defined here<br />

as extending from the cargo sequestration step to<br />

that <strong>of</strong> its lysosomal degradation) <strong>of</strong> material<br />

through the autophagic pathway (Codogno <strong>and</strong><br />

Meijer, 2005). Any blockade in the maturation <strong>of</strong><br />

autophagosomes, fusion with the lysosomal<br />

compartment or impairment <strong>of</strong> the lysosomal<br />

function or biogenesis would result in an<br />

accumulation <strong>of</strong> autophagosomes that would<br />

inevitably slowdown or interrupt the autophagic<br />

flux (Eskelinen, 2005; Rubinsztein et al., 2009).<br />

Maturation <strong>and</strong> degradation <strong>of</strong><br />

autophagosomes<br />

Rubicon <strong>and</strong> UVRAG. Rubicon <strong>and</strong> UVRAG (UV<br />

irradiation resistance associated gene) are two<br />

Beclin 1-binding proteins that regulate the<br />

maturation <strong>of</strong> autophagosomes <strong>and</strong> endocytic<br />

trafficking (Liang et al., 2006; Matsunaga et al.,<br />

2009; Zhong et al., 2009). These findings suggest<br />

that the Beclin 1: hVps34: UVRAG: Rubicon<br />

complex down-regulates these trafficking events,<br />

whereas the Beclin 1: hVps34: UVRAG complex<br />

upregulates the maturation <strong>of</strong> autophagosomes <strong>and</strong><br />

the endocytic trafficking (Matsunaga et al., 2009;<br />

Zhong et al., 2009). Therefore, Beclin 1 regulates<br />

both the formation <strong>of</strong> autophagosomes (via its<br />

interaction with Atg14L) <strong>and</strong> the maturation <strong>of</strong><br />

autophagosomes (via its interaction with UVRAG<br />

<strong>and</strong> Rubicon).<br />

Rab proteins. Colombo <strong>and</strong> co-workers (Gutierrez<br />

et al., 2004), <strong>and</strong> Eskelinen <strong>and</strong> co-workers (Jager<br />

et al., 2004) have shown that Rab7 is required for<br />

autophagosome maturation. Autophagosome<br />

maturation is dependent on interactions with class<br />

C Vps proteins <strong>and</strong> UVRAG (Liang et al., 2008).<br />

This function <strong>of</strong> UVRAG is independent <strong>of</strong> its<br />

interaction with Beclin 1, <strong>and</strong> stimulates Rab7<br />

GTPase activity <strong>and</strong> the fusion <strong>of</strong> autophagosomes<br />

with late endosomes/lysosomes. Interestingly,<br />

Rab11 is required for the fusion <strong>of</strong> autophagosomes<br />

<strong>and</strong> multivesicular bodies (MVB) during starvationinduced<br />

autophagy in the erythroleukemic cell line<br />

K562 (Fader et al., 2008). These findings suggest<br />

that specific membrane-bound compartment fusion<br />

processes during the maturation <strong>of</strong> autophagosomes<br />

engage different sets <strong>of</strong> Rab proteins, <strong>and</strong> possibly<br />

associated cohort proteins. Other Rab proteins such<br />

as Rab22 <strong>and</strong> Rab24 have subcellular locations<br />

compatible with a role in autophagy (Egami et al.,<br />

2005; Mesa et al., 2001; Olkkonen et al., 1993).<br />

ESCRT <strong>and</strong> Hrs. The endosomal sorting complex<br />

required for transport (ESCRT) mediates the<br />

biogenesis <strong>of</strong> MVB <strong>and</strong> the sorting <strong>of</strong> proteins in<br />

the endocytic pathway (Raiborg <strong>and</strong> Stenmark,<br />

2009). It has recently been demonstrated that the<br />

multisubunit complex ESCRT III is needed for<br />

autophagosomes to fuse with MVB <strong>and</strong> lysosomes<br />

to generate amphisomes <strong>and</strong> autolysosomes,<br />

respectively (Filimonenko et al., 2007; Lee et al.,<br />

2007; Rusten et al., 2007). ESCRT III dysfunction<br />

associated with the autophagic pathway may have<br />

important implications in neurodegenerative<br />

diseases (such as frontotemporal dementia <strong>and</strong><br />

amyotrophic lateral sclerosis) (Filimonenko et al.,<br />

2007; Lee et al., 2007). The Hrs protein (hepatocyte<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 166


Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells Mehrpour M, et al.<br />

growth factor-regulated tyrosine kinase substrate)<br />

plays a major role in endosomal sorting upstream <strong>of</strong><br />

ESCRT complexes (Raiborg <strong>and</strong> Stenmark, 2009).<br />

Hrs contains a FYVE domain that binds specifically<br />

to PtdIns3P, <strong>and</strong> facilitates the maturation <strong>of</strong><br />

autophagosomes (Tamai et al., 2007). This raises<br />

the intriguing possibility that PtdIns3P may be<br />

required for the formation <strong>of</strong> the autophagosome<br />

<strong>and</strong> its maturation. However, the role <strong>of</strong> ESCRT<br />

proteins in autophagy remains to be elucidated. It is<br />

impossible to rule out the possibility that these<br />

proteins could be involved in the closing <strong>of</strong><br />

autophagosomes (reviewed in Rusten <strong>and</strong><br />

Stenmark, 2009).<br />

SNAREs. Soluble N-ethylmaleimide-sensitive<br />

factor attachment protein receptors (SNAREs) are<br />

basic elements in intracellular membrane fusion<br />

(Gurkan et al., 2007; Rothman <strong>and</strong> Wiel<strong>and</strong>, 1996).<br />

In yeast the vacuolar t-SNAREs Vam3 (Darsow et<br />

al., 1997) <strong>and</strong> Vti1 (Ishihara et al., 2001), are<br />

needed for complete fusion to occur between the<br />

autophagosome <strong>and</strong> the vacuole (the name given to<br />

the lysosome in yeast) in S. cerevisiae. The<br />

mammalian homologue <strong>of</strong> Vt1, Vti1b, may be<br />

involved in a late stage <strong>of</strong> autophagy, because the<br />

maturation <strong>of</strong> autophagic vacuoles is delayed in<br />

hepatocytes isolated from mice in which Vti1b has<br />

been deleted (<strong>Atlas</strong>hkin et al., 2003). More recently,<br />

Colombo <strong>and</strong> colleagues (Fader et al., 2009) have<br />

reported that VAMP3 <strong>and</strong> VAMP7, two v-<br />

SNAREs, control the fusion between<br />

autophagosomes <strong>and</strong> MVB <strong>and</strong> fusion <strong>of</strong><br />

amphisomes with lysosomes, respectively.<br />

Endo/lysosomal membrane proteins. LAMPs<br />

(Lysosomal associated membrane proteins) are a<br />

family <strong>of</strong> heavily-glycosylated, endo/lysosomal<br />

transmembrane proteins (Eskelinen et al., 2003).<br />

Autophagic degradation has been shown to be<br />

impaired in hepatocytes isolated from LAMP-2<br />

deficient mice (Tanaka et al., 2000). However, no<br />

defect in autophagy was observed in LAMP-2<br />

deficient mouse fibroblasts (Eskelinen et al., 2004).<br />

A blockade in the later stage <strong>of</strong> autophagy only<br />

occurs in fibroblasts that are deficient for both<br />

LAMP-1 <strong>and</strong> LAMP-2. The differences in the<br />

autophagic activity observed between hepatocytes<br />

<strong>and</strong> fibroblasts may be responsible for the cell typespecific<br />

effects <strong>of</strong> LAMP-1 <strong>and</strong> LAMP-2 depletion<br />

(Eskelinen, 2005).<br />

DRAM (Damage-regulated autophagy modulator)<br />

encodes a 238-amino acid protein which is<br />

conserved through evolution, but has no ortholog in<br />

yeast (Crighton et al., 2006). DRAM is a direct<br />

target <strong>of</strong> p53. The protein is a multispanning<br />

transmembrane protein present in the lysosome.<br />

DRAM may regulate late stages <strong>of</strong> autophagy, but<br />

surprisingly it also controls the formation <strong>of</strong><br />

autophagosomes (Crighton et al., 2006). This<br />

suggests the possibility <strong>of</strong> a new paradigm in which<br />

feedback signals from the lysosomes control the<br />

early stages <strong>of</strong> autophagy.<br />

Microtubules. The destabilization <strong>of</strong> microtubules<br />

by either vinblastin (Høyvik et al., 1991) or<br />

nocodazole (Aplin et al., 1992) blocks the<br />

maturation <strong>of</strong> autophagosomes, whereas their<br />

stabilization by taxol increases the fusion between<br />

autophagic vacuoles <strong>and</strong> lysosomes (Yu <strong>and</strong><br />

Marzella, 1986). More recent findings have<br />

confirmed the role <strong>of</strong> microtubules in the fusion<br />

with the acidic compartment (Jahreiss et al., 2008;<br />

Kochl et al., 2006; Webb et al., 2004).<br />

Autophagosomes move bidirectionally along<br />

microtubules. Their centripetal movement is<br />

dependent on the dynein motor (Kimura et al.,<br />

2008; Ravikumar et al., 2005). Two types <strong>of</strong> fusion<br />

have been documented (Jahreiss et al., 2008): 1.<br />

Complete fusion <strong>of</strong> the autophagosome with the<br />

lysosome, 2. Transfer <strong>of</strong> material from the<br />

autophagosome to the lysosomal compartment<br />

following a kiss-<strong>and</strong>-run fusion process in which<br />

two separate vesicles are maintained. However,<br />

fusion with lysosomes can be microtubuleindependent<br />

during starvation-induced autophagy<br />

when autophagosomes are formed in the vicinity <strong>of</strong><br />

lysosomes (Fass et al., 2006).<br />

Acidification <strong>and</strong> degradation<br />

ATPases. Vacuolar ATPases (v-ATPase) are<br />

ubiquitous, multi-subunit proteins located in the<br />

acidic compartment (Forgac, 2007). Inhibition <strong>of</strong><br />

the activity <strong>of</strong> v-ATPase by bafilomycin A1 or<br />

concanamycin A blocks the lysosomal pumping <strong>of</strong><br />

H+ <strong>and</strong> consequently inhibits lysosomal enzymes,<br />

which are active at low pH. It has been proposed<br />

that bafilomycin A1 blocks the late stages <strong>of</strong><br />

autophagy by interfering with the fusion <strong>of</strong><br />

autophagosomes with endosomes <strong>and</strong> lysosomes<br />

(Yamamoto et al., 1998) or preventing the<br />

lysosomal degradation <strong>of</strong> sequestered material<br />

(Fass et al., 2006; Mousavi et al., 2001). Overall,<br />

the resulting effect <strong>of</strong> the inhibition <strong>of</strong> v-ATPase is<br />

to interrupt the autophagic flux as determined by<br />

the lysosomal inhibition <strong>of</strong> autophagic cargo.<br />

Interestingly it has recently been demonstrated that<br />

a deficiency <strong>of</strong> vacuolar H+-ATPase homolog<br />

(VMA21), a chaperone that binds to the c" subunit<br />

<strong>of</strong> the v-ATPase in the ER <strong>and</strong> which is responsible<br />

for X-linked myopathy with excessive autophagy<br />

(XMEA), causes an accumulation <strong>of</strong> autophagic<br />

vacuoles <strong>and</strong> interrupts autophagy flux in striated<br />

muscle cells (Ramach<strong>and</strong>ran et al., 2009).<br />

ATPases associated with various cellular activity<br />

proteins (AAA ATPases) are a family <strong>of</strong> proteins<br />

broadly engaged in intracellular membrane fusion<br />

(White <strong>and</strong> Lauring, 2007). N-ethylmaleimide<br />

sensitive factor (NSF) is an AAA ATPase, which<br />

binds to SNARE complexes <strong>and</strong> utilizes ATP<br />

hydrolysis to disassemble them, thus facilitating<br />

SNARE recycling. In yeast mutants lacking sec18<br />

(the yeast homolog <strong>of</strong> NSF), autophagosomes are<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 167


Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells Mehrpour M, et al.<br />

formed, but dot not fuse with the vacuole (Ishihara<br />

et al., 2001). However, we do not know whether the<br />

ATPase activity <strong>of</strong> NSF plays a role in the later<br />

stages <strong>of</strong> autophagy in mammalian cells.<br />

Nevertheless the activity <strong>of</strong> NSF is attenuated<br />

during starvation, which provides a possible<br />

explanation for the slow fusion between<br />

autophagosomes <strong>and</strong> lysosomes observed when<br />

autophagy is induced by starvation (Fass et al.,<br />

2006). Suppressor <strong>of</strong> K+ transport growth defect 1<br />

(SKD1-Vps4), another AAA ATPase protein, is<br />

required for the maturation <strong>of</strong> autophagosomes<br />

(Nara et al., 2002) in mammalian cells. Vps4<br />

controls the assembly <strong>of</strong> ESCRT complexes on the<br />

multivesicular membrane, <strong>and</strong> is involved in<br />

autophagosome maturation (Rusten et al., 2007) in<br />

Drosophila, <strong>and</strong> autophagosome fusion with the<br />

vacuole in yeast (Shirahama et al., 1997).<br />

Degradation <strong>and</strong> lysosomal efflux. By virtue <strong>of</strong><br />

lysosomal degradation autophagy contributes to<br />

regulating the metabolism <strong>of</strong> carbohydrates, lipids<br />

<strong>and</strong> proteins (Kotoulas et al., 2006; Kovsan et al.,<br />

2009; Mortimore <strong>and</strong> Pösö, 1987). Like<br />

acidification defects in the endo/lysosomal<br />

compartment, defects in the transport or the<br />

expression <strong>of</strong> lysosomal enzymes induce a blockade<br />

<strong>of</strong> autophagy, which is characterized by an<br />

accumulation <strong>of</strong> autophagic vacuoles (Koike et al.,<br />

2005; Yogalingam <strong>and</strong> Pendergast, 2008). The final<br />

stage <strong>of</strong> autophagy is the efflux <strong>of</strong> metabolites<br />

generated by the lysosomal degradation <strong>of</strong><br />

macromolecules into the cytosol (reviewed in<br />

Lloyd, 1996). Atg22 has recently been identified as<br />

a permease that recycles amino acids from the<br />

vacuole in S. cerevisiae (Yang et al., 2006).<br />

Cytoplasmic <strong>and</strong> nuclear<br />

regulation <strong>of</strong> mammalian<br />

autophagy<br />

Several recent reviews have been dedicated to the<br />

regulation <strong>of</strong> autophagy by signaling pathways<br />

(Codogno <strong>and</strong> Meijer, 2005; He <strong>and</strong> Klionsky,<br />

2009; Meijer <strong>and</strong> Codogno, 2009). In this section<br />

we would like to focus on signaling pathways with<br />

identified targets in the molecular machinery <strong>of</strong><br />

autophagosome formation. We will also discuss the<br />

role <strong>of</strong> signaling pathways <strong>and</strong> transcription factors<br />

in the regulation <strong>of</strong> the expression <strong>of</strong> genes<br />

involved in controlling autophagy.<br />

Cytoplasmic regulation<br />

mTORC1 <strong>and</strong> mTORC2. Many signals, including<br />

growth factors, amino acids, glucose, <strong>and</strong> energy<br />

status, are integrated by the kinase mTOR (Kim et<br />

al., 2002). The induction <strong>of</strong> autophagy by the<br />

inhibition <strong>of</strong> TOR under conditions <strong>of</strong> starvation is<br />

conserved from yeast to mammals (Blommaart et<br />

al., 1995; Noda <strong>and</strong> Ohsumi, 1998). The mTOR<br />

pathway involves two functional complexes:<br />

mTORC1 <strong>and</strong> mTORC2. Both these complexes are<br />

involved in the regulation <strong>of</strong> autophagy (Fig. 2).<br />

mTORC1, the rapamycin-sensitive mTOR complex<br />

1, contains the mTOR catalytic subunit, raptor<br />

(regulatory associated protein <strong>of</strong> mTOR, a protein<br />

that acts as a scaffold for the mTOR-mediated<br />

phosphorylation <strong>of</strong> mTOR substrates), GbL <strong>and</strong><br />

PRAS40 (proline-rich Akt substrate <strong>of</strong> 40 kDa).<br />

The binding <strong>of</strong> FKBP12 to mTOR inhibits the<br />

mTOR-raptor interaction, suggesting a mechanism<br />

for rapamycin-specific inhibition <strong>of</strong> mTOR<br />

signaling (Oshiro et al., 2004). This mTOR-raptor<br />

interaction, <strong>and</strong> its regulation by nutrients <strong>and</strong>/or<br />

rapamycin are dependent on GbL (Kim et al.,<br />

2003). A major unanswered question about the<br />

stimulation <strong>of</strong> autophagy during starvation is how<br />

amino acids signal to mTOR (Meijer <strong>and</strong><br />

Dubbelhuis, 2004). It has been suggested that<br />

hVps34 may have a role in the amino acid signaling<br />

to mTORC1 (Byfield et al., 2005; Nobukuni et al.,<br />

2005). Thus it appears that hVps34 acts both as a<br />

down-regulator <strong>of</strong> autophagy (acting as an amino<br />

acid sensor) <strong>and</strong> as an up-regulator (because it is a<br />

component <strong>of</strong> Beclin 1 complexes) <strong>of</strong> autophagy.<br />

However, recent observations in Drosophila <strong>and</strong><br />

mammalian cells suggest that Rag GTPases (Rasrelated<br />

small GTPases) activate TORC1 in response<br />

to amino acids by promoting its redistribution to a<br />

specific subcellular compartment, which contains<br />

the TORC1 activator Rheb (Ras homolog enriched<br />

in brain, a GTP-binding protein) (Kim et al., 2008a;<br />

Sancak et al., 2008). Moreover, the rate-limiting<br />

factor that enables essential amino acids to inhibit<br />

mTORC1 has been identified as L-glutamine<br />

(Nicklin et al., 2009). L-glutamine uptake is<br />

regulated by solute carrier family 1, member 5<br />

(SLC1A5). Loss <strong>of</strong> SLC1A5 function activates<br />

autophagy, <strong>and</strong> inhibits cell growth. Thus, Lglutamine<br />

sensitivity is attributable to<br />

SLC7A5/SLC3A2, a bidirectional transporter that<br />

regulates the simultaneous efflux <strong>of</strong> L-glutamine<br />

out <strong>of</strong> cells <strong>and</strong> the transportation <strong>of</strong> Lleucine/essential<br />

amino acids into cells (Nicklin et<br />

al., 2009).<br />

The other mTOR complex, mTORC2, which is less<br />

sensitive to rapamycin, includes mTOR, rictor<br />

(rapamycin-insensitive companion <strong>of</strong> mTOR), GbL,<br />

SIN1 (SAPK-interacting protein 1) <strong>and</strong> PROTOR<br />

(protein observed with rictor) (Sarbassov et al.,<br />

2006). The mTORC2 complex phosphorylates the<br />

Ser 473 <strong>of</strong> Akt/PKB, thereby contributing to the<br />

activation <strong>of</strong> this important cell-survival kinase<br />

(Sarbassov et al., 2006). Phosphorylated Akt/PKB<br />

down-regulates the activity <strong>of</strong> the transcriptional<br />

factor Forkhead Box O3 (FoxO3). Interestingly,<br />

FoxO3 has been shown to stimulate autophagy in<br />

muscle cells by increasing the transcription <strong>of</strong><br />

several genes involved in autophagy (see below)<br />

(Mammucari et al., 2007).<br />

Signaling segments acting upstream <strong>of</strong> mTORC1<br />

<strong>and</strong> mTORC2 that regulate autophagy have been<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 168


Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells Mehrpour M, et al.<br />

discussed in recent reviews that the reader can<br />

consult for more information (Codogno <strong>and</strong> Meijer,<br />

2005; Meijer <strong>and</strong> Codogno, 2009).<br />

mTORC1 substrates <strong>and</strong> the regulation <strong>of</strong><br />

autophagy. As discussed above,ULK1, Atg13 <strong>and</strong><br />

FIP200 form a stable complex that signals to the<br />

autophagic machinery downstream <strong>of</strong> mTORC1.<br />

Importantly, mTORC1 is incorporated into the<br />

ULK1:Atg13:FIP200 complex via ULK1 in a<br />

nutrient-dependent manner. mTOR phosphorylates<br />

both ULK1 <strong>and</strong> Atg13. Under starvation conditions<br />

or in response to rapamycin treatment, mTORC1<br />

dissociates from the ULK1 complex, resulting in<br />

the activation <strong>of</strong> ULK1. Activated ULK1<br />

autophosphorylates, <strong>and</strong> also phosphorylates both<br />

Atg13 <strong>and</strong> FIP200 to initiate autophagy (Hosokawa<br />

et al., 2009). The location <strong>of</strong> phosphorylation sites,<br />

as well as the role played by other members <strong>of</strong> the<br />

ULK family, ULK2 <strong>and</strong> ULK3, remain to be<br />

determined. Once activated, mTORC1 favors cell<br />

growth by promoting translation via the<br />

phosphorylation <strong>of</strong> 70 kDa, polypeptide 1<br />

ribosomal protein S6 kinase-1 (p70S6K), <strong>and</strong><br />

phosphorylation <strong>of</strong> the inhibitor <strong>of</strong> translation<br />

initiation, 4E-BP1. Interestingly, Neufeld <strong>and</strong><br />

coworkers showed that p70S6K is up-regulated<br />

during starvation-induced autophagy in the<br />

Drosophila fat body (Scott et al., 2004).<br />

Mammalian cells probably have a regulatory<br />

feedback pathway involving S6K that regulates<br />

autophagy (Klionsky et al., 2005). p70S6K is<br />

known to phosphorylate <strong>and</strong> inhibit IRS1<br />

downstream <strong>of</strong> the insulin receptor (reviewed in<br />

Meijer <strong>and</strong> Codogno, 2009). This loop could<br />

provide a way to regulate the activity <strong>of</strong> mTORC1<br />

during starvation-induced autophagy.<br />

AMP-activated kinase (AMPK). Apart <strong>of</strong> being a<br />

sensor, mTOR can also sense changes in the<br />

cellular energy via AMPK. Activation <strong>of</strong> AMPK<br />

inhibits mTOR-dependent signaling, by interfering<br />

the activity <strong>of</strong> the GTPase Rheb, <strong>and</strong> protein<br />

synthesis (Meijer <strong>and</strong> Codogno, 2004). This is<br />

consistent with the switch <strong>of</strong>f <strong>of</strong> ATP-dependent<br />

processes (Hardie, 2004) during period <strong>of</strong> energy<br />

crisis. In starved cells when AMP/ATP ratio<br />

increases, the binding <strong>of</strong> AMP to AMPK favors its<br />

activation by the AMPK kinase LKB1 (Corradetti<br />

et al., 2004; Shaw et al., 2004). Moreover,<br />

Ca 2+ /calmodulin-dependent kinase kinase b<br />

(CaMKK-b) has been identified to be an AMPK<br />

kinase (Hawley et al., 2005; Woods et al., 2005).<br />

The activity <strong>of</strong> AMPK is required to induce<br />

autophagy in response to starvation in mammalian<br />

cells (Meley et al., 2006) <strong>and</strong> in yeast (Wang et al.,<br />

2001) in a TORC1-dependent manner. Moreover,<br />

induction autophagy is also dependent on the<br />

inhibition <strong>of</strong> mTORC1 by AMPK in non-starved<br />

cells in response to an increase in free cytosolic<br />

Ca 2+ (Høyer-Hansen et al., 2007). In this setting, the<br />

activation <strong>of</strong> AMPK <strong>and</strong> stimulation <strong>of</strong> autophagy<br />

are dependent on CaMKK-b. Induction <strong>of</strong><br />

autophagy through the activation <strong>of</strong> AMPK is<br />

probably extended to other settings such as hypoxia<br />

(Degenhardt et al., 2006; Laderoute et al., 2006).<br />

AMPK is probably a general regulator <strong>of</strong> autophagy<br />

upstream <strong>of</strong> mTOR (Høyer-Hansen <strong>and</strong> Jaattela,<br />

2007; Meijer <strong>and</strong> Codogno, 2007). Another<br />

potential c<strong>and</strong>idate <strong>of</strong> autophagy regulation<br />

downstream <strong>of</strong> AMPK is the Elongation factor-2<br />

kinase (eEF-2 kinase) that controls the rate <strong>of</strong><br />

peptide elongation (Hait et al., 2006). Activation <strong>of</strong><br />

eEF-2 kinase increases autophagy while slowing<br />

down protein translation (Wu et al., 2006). The<br />

activity <strong>of</strong> eEF-2 kinase is regulated by mTOR,<br />

S6K <strong>and</strong> AMPK (Browne et al., 2004; Browne <strong>and</strong><br />

Proud, 2002). During periods <strong>of</strong> ATP depletion,<br />

AMPK is activated <strong>and</strong> eEF-2 kinase is<br />

phosphorylated (Browne et al., 2004) making a<br />

balance between inhibition <strong>of</strong> peptide elongation<br />

<strong>and</strong> induction <strong>of</strong> autophagy. How eEF-2 kinase<br />

impinges on the molecular machinery <strong>of</strong> autophagy<br />

remains to be investigated. Moreover, autophagy is<br />

activated by AMPK in a p53-dependent manner<br />

(Feng et al., 2005). But the cytoplasmic form <strong>of</strong> p53<br />

has been shown to have an inhibitory effect on<br />

autophagy (Tasdemir et al., 2008).<br />

Finally, AMPK also triggers the initiation <strong>of</strong><br />

autophagosome formation by phosphorylating<br />

ULK1 (Egan et al., 2011; Kim et al., 2011) (Figure<br />

2).<br />

Beclin 1:hVps34 complexes. As discussed in the<br />

preceding sections, the trimer Beclin<br />

1:hVps34:hVps15 can interact with various<br />

different partners to control the formation <strong>and</strong> the<br />

maturation <strong>of</strong> autophagosomes. Recently, the antiapoptotic<br />

protein Bcl-2, <strong>and</strong> anti-apoptotic<br />

members <strong>of</strong> the Bcl-2 family such as Bcl-XL were<br />

shown to inhibit autophagy (Erlich et al., 2007;<br />

Maiuri et al., 2007; Pattingre et al., 2005). Bcl-<br />

2/Bcl-XL binds Beclin 1 through a BH3 domain that<br />

mediates the docking <strong>of</strong> the latter to the BH3binding<br />

groove. The constitutive Bcl-2/Bcl-<br />

XL:Beclin 1 interaction is disrupted by signals that<br />

promote autophagy. Importantly, c-Jun N-terminal<br />

kinase 1 (JNK-1) phosphorylates 3 amino acids in<br />

the N-terminal loop <strong>of</strong> Bcl-2 to trigger its release<br />

from Beclin 1 (Wei et al., 2008) in response to<br />

starvation or ceramide treatment (Pattingre et al.,<br />

2009; Wei et al., 2008). In a reciprocal manner, the<br />

BH3 domain <strong>of</strong> Beclin 1 can be phosphorylated by<br />

death-associated protein kinase (DAPk), which has<br />

the effect <strong>of</strong> reducing its affinity for Bcl-XL<br />

(Zalckvar et al., 2009). A second mechanism that<br />

leads to the dissociation <strong>of</strong> the complex involves<br />

the competitive displacement <strong>of</strong> BH3-domain<br />

Beclin 1 from Bcl-2/Bcl-XL by other BH3containing<br />

proteins with proapoptotic properties<br />

such BH3-only member <strong>of</strong> the Bcl-2 family (BAD),<br />

the pro-apoptotic member <strong>of</strong> the Bcl-2 family, Bax-<br />

, <strong>and</strong> BH3-mimetics (Luo <strong>and</strong> Rubinsztein, 2010;<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 169


Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells Mehrpour M, et al.<br />

Maiuri et al., 2007). The role <strong>of</strong> the hypoxiainducible<br />

BH3-only proteins BNIP3 <strong>and</strong> BNIP3L in<br />

the dissociation <strong>of</strong> the Beclin 1:Bcl-2 complex will<br />

be discussed in the next section. Overall these<br />

findings point to the central role <strong>of</strong> Beclin1:hVps34<br />

complexes in the cross-talk between autophagy <strong>and</strong><br />

apoptosis. Interestingly, a recent study reports that<br />

the anti-apoptotic protein FLIP, which blocks the<br />

activation <strong>of</strong> caspase 8 downstream <strong>of</strong> death<br />

receptors, is also an anti-autophagy molecule by<br />

blocking the formation <strong>of</strong> LC3-II via its interaction<br />

with Atg3 (Lee et al., 2009) (Figure 2).<br />

Inositol 1,4,5-trisphosphate (IP3) receptor.<br />

Autophagy can also be induced via an mTORindependent<br />

pathway by lowering myo-inositol<br />

1,4,5-trisphosphate (IP3) levels (Sarkar et al., 2005).<br />

This effect can be achieved pharmacologically with<br />

drugs such as lithium or L-690 330, which disrupt<br />

the metabolism <strong>of</strong> inositol by inhibiting inositol<br />

monophosphatase (IMP). Rubinsztein <strong>and</strong><br />

coworkers found that L-type Ca2+ channel<br />

antagonists, the K+ATP channel opener, <strong>and</strong> Gi<br />

signaling activators all induce autophagy (Williams<br />

et al., 2008). These drugs reveal a cyclical mTORindependent<br />

pathway regulating autophagy, in<br />

which cAMP regulates IP3 levels, influencing<br />

calpain activity, which completes the cycle by<br />

cleaving <strong>and</strong> activating Gs alpha, which regulates<br />

cAMP levels. These data also suggest that insults<br />

that elevate intracytosolic Ca2+ (such as<br />

excitotoxicity) inhibit autophagy, thus retarding the<br />

clearance <strong>of</strong> aggregate-prone proteins. Both genetic<br />

<strong>and</strong> pharmacological inhibition <strong>of</strong> the IP3 receptor<br />

(IP3R) strongly stimulate autophagy. Kroemer <strong>and</strong><br />

coworkers have shown that the IP3R antagonist<br />

xestospongin B induces autophagy by disrupting a<br />

molecular complex formed between the IP3R <strong>and</strong><br />

Beclin 1, an interaction that is regulated by Bcl-2<br />

(Vicencio et al., 2009). The IP3R is known to be<br />

located in the membranes <strong>of</strong> the ER as well as in<br />

ER-mitochondrial contact sites, <strong>and</strong> IP3R blockade<br />

triggers the autophagy <strong>of</strong> both ER <strong>and</strong><br />

mitochondria, as observed in starvation-induced<br />

autophagy. ER stressors, such as tunicamycin <strong>and</strong><br />

thapsigargin, also induce autophagy <strong>of</strong> the ER <strong>and</strong>,<br />

to a lesser extent, <strong>of</strong> mitochondria. Autophagy<br />

triggered by starvation or IP3R blockade is inhibited<br />

by Bcl-2, <strong>and</strong> Bcl-XL located at the ER, but not at<br />

the mitochondrial outer membrane (Pattingre et al.,<br />

2005; Vicencio et al., 2009). In contrast, ER stressinduced<br />

autophagy is not inhibited by Bcl-2 or Bcl-<br />

XL. Autophagy promoted by IP3R inhibition cannot<br />

be attributed to a modulation <strong>of</strong> steady-state Ca2+<br />

levels in the ER or in the cytosol (Vicencio et al.,<br />

2009).<br />

Other cytoplasmic autophagy regulation<br />

mechanisms. The function <strong>of</strong> Atg proteins in<br />

autophagy can be regulated not only by proteinprotein<br />

interactions <strong>and</strong> phosphorylation, but also<br />

by oxidation, acetylation, <strong>and</strong> proteolytic cleavage.<br />

Elazar <strong>and</strong> colleagues (Scherz-Shouval et al., 2007)<br />

have reported that the oxidation <strong>of</strong> a cysteine<br />

residue near the catalytic site <strong>of</strong> Atg4A <strong>and</strong> Atg4B<br />

is required during starvation-induced autophagy.<br />

During starvation, the deacetylation <strong>of</strong> Atg5, Atg7,<br />

LC3 <strong>and</strong> Atg12 is important to stimulate autophagy.<br />

The acetylation is dependent on the activity <strong>of</strong> p300<br />

(Lee <strong>and</strong> Finkel, 2009), <strong>and</strong> the deacetylation is<br />

probably under the control <strong>of</strong> the histone<br />

deacetylase sirtuin 1 (Lee et al., 2008). Atg5, Atg7<br />

<strong>and</strong> Beclin 1 are substrates for calpains (Kim et al.,<br />

2008b; Xia et al., 2010; Yousefi et al., 2006), <strong>and</strong><br />

Atg4D <strong>and</strong> Beclin 1 are substrates for caspases<br />

(Betin <strong>and</strong> Lane, 2009; Cho et al., 2009; Luo <strong>and</strong><br />

Rubinsztein, 2010). The cleavage <strong>of</strong> Beclin 1 by<br />

caspase 3, <strong>and</strong> that <strong>of</strong> Atg5 by calpain 1 inhibit<br />

autophagy (Luo <strong>and</strong> Rubinsztein, 2010; Xia et al.,<br />

2010). The cleavage <strong>of</strong> Atg proteins by caspases<br />

<strong>and</strong> calpains has been proposed as a possible<br />

additional mechanism modulating autophagy.<br />

Interestingly, the truncated form <strong>of</strong> Atg5 generated<br />

by calpain 1 cleavage is translocated into the<br />

mitochondria <strong>and</strong> induces apoptosis (Yousefi et al.,<br />

2006).<br />

Nuclear regulation <strong>of</strong> autophagy<br />

JNK/c-Jun. The sphingolipid ceramides have been<br />

shown to increase the expression <strong>of</strong> Beclin 1 in<br />

human cancer cell lines (Scarlatti et al., 2004). In<br />

cancer cell lines exposed to ceramide, Li et al. have<br />

observed activation <strong>of</strong> JNK <strong>and</strong> increased<br />

phosphorylation <strong>of</strong> c-Jun (Li et al., 2009). They<br />

also showed that c-Jun controls the transcription <strong>of</strong><br />

Becn1 (we have adopted this nomenclature for the<br />

gene encoding Beclin 1), <strong>and</strong> the induction <strong>of</strong><br />

autophagic cell death in response to ceramide.<br />

Activation <strong>of</strong> JNK, resulting in the upregulation <strong>of</strong><br />

Beclin 1 expression, has also been reported in the<br />

autophagic cell death <strong>of</strong> human colon cancer cells<br />

induced by the stimulation <strong>of</strong> the human death<br />

receptor 5 (Park et al., 2009).<br />

NF-κB. The NF-κB transcription factor, which<br />

plays a plethoric role in inflammation, immunity<br />

<strong>and</strong> cancer (Karin, 2006), has also been implicated<br />

in regulating autophagy. A conserved NF-κB<br />

binding site has recently been unveiled in the<br />

promoter <strong>of</strong> the murine <strong>and</strong> human gene that<br />

encodes Beclin 1 (Copetti et al., 2009). The authors<br />

have shown that p65/RelA, a member <strong>of</strong> the NF-κB<br />

family, upregulates the expression <strong>of</strong> Beclin 1 <strong>and</strong><br />

stimulates autophagy in several cellular systems.<br />

Autophagy stimulation has also been observed after<br />

the activation <strong>of</strong> NF-κB during the heat shock<br />

response (Nivon et al., 2009). However, in contrast<br />

to this stimulatory role <strong>of</strong> NF-κB in the regulation<br />

<strong>of</strong> autophagy, the inhibition <strong>of</strong> NF-κB favors<br />

TNFα-dependent <strong>and</strong> starvation-dependent<br />

autophagy (Djavaheri-Mergny et al., 2006; Fabre et<br />

al., 2007). Moreover, Schlottmann et al. reported<br />

that activation <strong>of</strong> NF-κB prevents autophagy in<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 170


Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells Mehrpour M, et al.<br />

macrophages by downregulating the expression <strong>of</strong><br />

Atg5 <strong>and</strong> Beclin 1 (Schlottmann et al., 2008).<br />

E2F1. E2F transcription factors are known to be<br />

involved in cellular proliferation, but also in DNA<br />

repair, differentiation <strong>and</strong> development (DeGregori<br />

<strong>and</strong> Johnson, 2006). E2F1 has been shown to bind<br />

to the promoter <strong>of</strong> Becn1, although an effect <strong>of</strong><br />

E2F1 on Beclin 1 expression remains to be<br />

demonstrated (Weinmann et al., 2001). More<br />

recently, the activation <strong>of</strong> E2F1 has been shown to<br />

induce autophagy by upregulating the expression <strong>of</strong><br />

the autophagy genes Map1lc3, Ulk1, Atg5 <strong>and</strong><br />

Dram (we have adopted the nomenclature Map1lc3<br />

for the gene encoding LC3) (Polager et al., 2008).<br />

The E2F1-mediated induction <strong>of</strong> Map1lc3, Ulk1<br />

<strong>and</strong> Dram is direct (interaction with the promoter),<br />

whereas the up-regulation <strong>of</strong> the expression <strong>of</strong> Atg5<br />

is indirect.<br />

HIF-1. HIF-1 (hypoxia-inducible factor-1) is a<br />

transcription factor, which regulates the<br />

transcription <strong>of</strong> hundred <strong>of</strong> genes in response to<br />

hypoxia (Manalo et al., 2005). Recently Zhang et al<br />

have demonstrated that hypoxia-induced<br />

mitochondrial autophagy via HIF-1 mediated<br />

induction <strong>of</strong> Bnip3 (Zhang et al., 2008). In a similar<br />

way, Bellot et al. have shown that hypoxia-induced<br />

autophagy is mediated by HIF-1, which induces the<br />

expression <strong>of</strong> BNIP3 <strong>and</strong> BNIP3L (Bellot et al.,<br />

2009). BNIP3 <strong>and</strong> BNIP3L play important roles in<br />

the induction <strong>of</strong> autophagy by disrupting the<br />

interaction <strong>of</strong> Beclin 1 with Bcl-2 via their BH3<br />

domain. HIF-1 could also regulate the expression <strong>of</strong><br />

Beclin 1 <strong>and</strong> Atg5, probably indirectly although<br />

according to a recent report the silencing <strong>of</strong> HIF-1<br />

in cultured chondrocytes was associated with a<br />

reduced level <strong>of</strong> Beclin 1 (Bohensky et al., 2007).<br />

FoxO proteins. Three members <strong>of</strong> the FoxO family<br />

<strong>of</strong> transcription factors, FoxO1, FoxO3, <strong>and</strong> FoxO4<br />

are regulated by Akt phosphorylation in response to<br />

growth factor <strong>and</strong> insulin stimulation. FoxO<br />

proteins are phosphorylated by Akt, which renders<br />

them inactive in the presence <strong>of</strong> growth factors.<br />

When Akt is repressed, FoxO proteins are<br />

translocated into the nucleus, bind to DNA, <strong>and</strong><br />

transactivate its target genes (Salih <strong>and</strong> Brunet,<br />

2008). Several studies <strong>of</strong> protein degradation during<br />

muscle atrophy show that FoxO3 can induce the<br />

expression <strong>of</strong> multiple autophagy genes, including<br />

Map1lc3, Atg12, Becn1, Atg4b, Ulk1, Pik3c3 (we<br />

have adopted the nomenclature Pik3c3 for the gene<br />

encoding hVps34), Bnip3/Bnip3l, <strong>and</strong> Gabarapl1<br />

(Mammucari et al., 2007; Zhao et al., 2007), <strong>and</strong><br />

then upregulates autophagy. FoxO3 can bind<br />

directly to the promoter region <strong>of</strong> some <strong>of</strong> these<br />

genes, such as Map1lc3, Atg12, Gabarapl1 <strong>and</strong><br />

Bnip3/Bnip3l. Expression <strong>of</strong> a constitutive form <strong>of</strong><br />

FoxO3 induces autophagy in adult mouse skeletal<br />

muscle. Recently another member <strong>of</strong> the FoxO<br />

protein family, FoxO1, has been shown to regulate<br />

the expression <strong>of</strong> key autophagy genes, Pik3c3,<br />

Atg12, <strong>and</strong> Gabarapl1 in hepatocytes in an insulindependent<br />

manner (Liu et al., 2009).<br />

p53. p53 is a pivotal factor involved in regulating<br />

cell death <strong>and</strong> survival, <strong>and</strong> in regulating<br />

metabolism (Vousden <strong>and</strong> Prives, 2009). When p53<br />

is activated by cellular stress, p53 accumulates in<br />

the cell nucleus, where it transactivates several<br />

autophagy-modulating genes including Dram<br />

(damage-regulated autophagy modulator) <strong>and</strong> Tigar<br />

(TP53-induced glycolysis <strong>and</strong> apoptosis regulator).<br />

DRAM stimulates the accumulation <strong>of</strong><br />

autophagosomes, probably by regulating<br />

autophagosome-lysosome fusion to generate<br />

autophagolysosomes (Crighton et al., 2006).<br />

TIGAR, through its fructose 2, 6-bisphosphatase<br />

function, can modulate the glycolytic pathway <strong>and</strong><br />

indirectly contribute to the fall in the intracellular<br />

level <strong>of</strong> ROS (Bensaad et al., 2006). Recently, the<br />

same group showed that TIGAR can also modulate<br />

the intracellular ROS level in response to nutrient<br />

starvation or metabolic stress, <strong>and</strong> consequently<br />

inhibit autophagy via an mTOR-independent<br />

pathway (Bensaad et al., 2009). So, while DRAM<br />

<strong>and</strong> TIGAR are both transactivated by p53, DRAM<br />

promotes autophagy whereas TIGAR inhibits<br />

autophagy. The complexity <strong>of</strong> the autophagic<br />

response to p53 is further increased by the ability <strong>of</strong><br />

cytoplasmic p53 to limit autophagy (Tasdemir et<br />

al., 2008).<br />

Transcription factor EB (TFEB). Recently the<br />

bHLH-leucine zipper transcription factor TFEB has<br />

been shown to control lysosomal biogenesis <strong>and</strong><br />

function. TFEB is a master gene in the gene<br />

regulatory network CLEAR (CLEAR: Coordinated<br />

Lysosomal Enhancement And Regulation) that<br />

binds to CLEAR target sites in the promoter <strong>of</strong><br />

lysosomal genes <strong>and</strong> increases lysosomal<br />

biogenesis (Sardiello et al., 2009). TFEB not only<br />

controls the expression <strong>of</strong> lysosomal proteins, it<br />

also regulates the autophagy transcription program<br />

during starvation (Settembre et al., 2011). TFEB is<br />

retained in the cytosol because it phosphorylates<br />

MAP kinase (ERK2). The reduction <strong>of</strong> TFEB<br />

phosphorylation during starvation triggers its<br />

nuclear transport where it activates a transcription<br />

program that activates the biogensesis <strong>of</strong> lysosomes<br />

<strong>and</strong> stimulates autophagy. TFEB has a broad range<br />

<strong>of</strong> activities, because it controls the activity <strong>of</strong><br />

genes involved in various different steps <strong>of</strong><br />

autophagy, including autophagosome formation<br />

(Map1lc3, Wipi), cargo recognition (Sqstm1,) <strong>and</strong><br />

autophagosome fusion with the lysosomal<br />

compartment (Uvrag, Vps11, Vps18). The fact that<br />

autophagy <strong>and</strong> lysosomal formation are both<br />

coordinated by TFEB <strong>of</strong>fers a possible therapeutic<br />

target that could be used to boost the autophagic<br />

pathway when appropriate.<br />

Other regulators <strong>of</strong> Atg genes expression.<br />

Recently the phosphorylation <strong>of</strong> eIF2α by PERK<br />

has been shown to be essential for the conversion <strong>of</strong><br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 171


Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells Mehrpour M, et al.<br />

LC3-I to LC3-II during ER-stress induced by the<br />

polyQ72 or dysferlin L1341P mutant (Fujita et al.,<br />

2007; Kouroku et al., 2007). In polyQ72 loaded<br />

mammalian cells, the phosphorylation <strong>of</strong> eIF2α<br />

upregulates the expression <strong>of</strong> Atg12 (Kouroku et<br />

al., 2007). During the unfolded protein response,<br />

triggered by hypoxia, the transcription factors<br />

ATF4 <strong>and</strong> CHOP, which are regulated by PERK,<br />

increase the expression Map1lc3b <strong>and</strong> Atg5,<br />

respectively (Rouschop et al., 2010).<br />

Glyceraldehyde-3-phosphate dehydrogenase<br />

(GAPDH), a multifunctional enzyme known to play<br />

a role in glycolysis as well as having other less<br />

well-understood roles such as transcriptional<br />

coactivation, has also been shown to upregulate the<br />

transcription <strong>of</strong> Atg12 to protect cells against<br />

caspase-independent cell death (Colell et al., 2007).<br />

Beclin 1 is one <strong>of</strong> the essential components<br />

involved in autophagosome formation, <strong>and</strong> its level<br />

usually increases during autophagy. For example, in<br />

HBV (hepatitis B virus)-infected hepatocytes, the<br />

HBV x protein increases autophagy by upregulating<br />

the expression <strong>of</strong> Beclin 1 (Tang et al., 2009). In<br />

human monocytes <strong>and</strong> human myeloid leukemia<br />

cells, vitamin D3 has been shown to induce<br />

autophagy by upregulating both Beclin 1 <strong>and</strong> Atg5<br />

(Wang et al., 2008; Yuk et al., 2009). The<br />

transcription factor implicated has not been<br />

identified, but it has been shown that in human<br />

monocytes the effect <strong>of</strong> Vitamin D3 is mediated via<br />

cathelicidin.<br />

Very recently, Zhu et al. (Zhu et al., 2009) observed<br />

for the first time the regulation <strong>of</strong> autophagy by<br />

miRNA. They showed that miR-30a targets Beclin<br />

1 mRNA, <strong>and</strong> down-regulates the expression <strong>of</strong><br />

Beclin 1.<br />

Autophagy in physiology<br />

Autophagosome formation occurs at a basal rate in<br />

most cells <strong>and</strong> controls the quality <strong>of</strong> the cytoplasm<br />

by initiating the elimination <strong>of</strong> protein aggregates<br />

<strong>and</strong> <strong>of</strong> damaged organelles (Ravikumar et al.,<br />

2010b) (Figure 3). This autophagy-dependent<br />

quality control is also important to limit the<br />

production <strong>of</strong> ROS.<br />

Stimulation <strong>of</strong> autophagy during periods <strong>of</strong><br />

starvation is an evolutionarily-conserved response<br />

to stress in eukaryotes (Yang <strong>and</strong> Klionsky, 2010).<br />

Under starvation conditions, the degradation <strong>of</strong><br />

proteins <strong>and</strong> lipids allows the cell to adapt its<br />

metabolism <strong>and</strong> meet its energy needs (Figure 3).<br />

Figure 3. Physiological <strong>and</strong> pathological roles <strong>of</strong> autophagy. The physiological role <strong>of</strong> basal autophagy is to clean the<br />

cytoplasm <strong>of</strong> damaged organelles <strong>and</strong> protein aggregates. This function is essential for cell fitness by limiting the accumulation <strong>of</strong><br />

ROS. The stimulation <strong>of</strong> autophagy during periods <strong>of</strong> starvation plays a major role in many tissues, but with some exceptions,<br />

such as the brain, in providing nutrients for cell metabolism, for the biosynthesis <strong>of</strong> macromolecules, <strong>and</strong> to maintain the level <strong>of</strong><br />

ATP. Autophagy is involved in an early stage <strong>of</strong> development (pre-implantation <strong>of</strong> the fertilized oocyte), <strong>and</strong> differentiation.<br />

Autophagy declines during aging, <strong>and</strong> the restoration <strong>of</strong> autophagy extends life span in various species. Autophagy contributes<br />

to both innate <strong>and</strong> adaptive immunity. Defective autophagy is observed in many human diseases, <strong>and</strong> its stimulation is beneficial<br />

in most cases. Autophagy plays a more complex role in cancer, because it can be a tumor suppressor mechanism, but can also<br />

become a cytoprotective mechanism in tumors, where it contributes to cell survival in a context <strong>of</strong> metabolic stress <strong>and</strong> in<br />

response to cancer treatments.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 172


Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells Mehrpour M, et al.<br />

The stimulation <strong>of</strong> autophagy plays a major role at<br />

birth to maintain energy levels in various tissues<br />

after the maternal nutrient supply via the placenta<br />

ceases (Kuma et al., 2004). Moreover, starvationinduced<br />

autophagy is cytoprotective by blocking<br />

the induction <strong>of</strong> apoptosis upstream <strong>of</strong><br />

mitochondrial events (Boya et al., 2005).<br />

Autophagy is essential during development <strong>and</strong><br />

differentiation. The pre-implantation period after<br />

oocyte fertilization is dependent on autophagic<br />

degradation <strong>of</strong> components <strong>of</strong> the oocyte cytoplasm<br />

(Tsukamoto et al., 2008). Autophagy remodeling <strong>of</strong><br />

the cytoplasm is involved during the differentiation<br />

<strong>of</strong> erythrocytes, lymphocytes <strong>and</strong> adipocytes<br />

(Ravikumar et al., 2010b). Autophagy is crucial for<br />

the homeostasis <strong>of</strong> immune cells, <strong>and</strong> contributes to<br />

the regulation <strong>of</strong> self-tolerance (Nedjic et al., 2008).<br />

The pioneering work <strong>of</strong> Bergamini <strong>and</strong> colleagues<br />

in the field <strong>of</strong> autophagy (Yang <strong>and</strong> Klionsky,<br />

2010) suggested that its stimulation during calorie<br />

restriction may contribute to extending lifespan in<br />

the rat. Recent data have shown that the induction<br />

<strong>of</strong> autophagy increases longevity in a large panel <strong>of</strong><br />

species (Eisenberg et al., 2009). The antiaging<br />

effect <strong>of</strong> autophagy probably depends, at least in<br />

part, on its quality control function that limits the<br />

accumulation <strong>of</strong> aggregation-prone protein <strong>and</strong><br />

damaged mitochondria. Calorie restriction<br />

stimulates autophagy via the activation <strong>of</strong> the<br />

deacetylase sirtuin-1 (Morselli et al., 2010). Targets<br />

<strong>of</strong> sirtuin-1 include Atg proteins 5, 7 <strong>and</strong> 8. In<br />

several cell lines, deacetylation <strong>of</strong> these Atg<br />

proteins is necessary for autophagy to be stimulated<br />

by nutrient deprivation (Lee et al., 2008). It would<br />

be interesting to investigate whether this regulation<br />

<strong>of</strong> autophagy is conserved in tissues that<br />

preferentially use fatty acids as oxidizable<br />

substrates during starvation, where a high level <strong>of</strong><br />

acetyl-CoA is to be expected.<br />

Autophagy <strong>and</strong> pathology<br />

As mentioned above, basal autophagy is important<br />

as a housekeeping process to prevent the deposition<br />

<strong>of</strong> aggregation-prone proteins in the cytoplasm<br />

(Figure 3). Several neurodegenerative diseases,<br />

including Huntington's, Alzheimer's <strong>and</strong><br />

Parkinson's diseases, are characterized by the<br />

accumulation <strong>of</strong> such protein aggregates in the<br />

brain (Ravikumar et al., 2010b). As a protective<br />

measure, the stimulation <strong>of</strong> autophagy limits the<br />

accumulation <strong>of</strong> toxic products <strong>and</strong> protects<br />

neurons against degeneration. The important<br />

pathological role <strong>of</strong> autophagy in aggregation-prone<br />

protein disease is strengthened by a recent study<br />

showing that a drug that enhances autophagy<br />

promotes the degradation <strong>of</strong> mutant, aggregationprone<br />

α1-antitrypsin in the liver, <strong>and</strong> consequently<br />

reduces hepatic fibrosis (Hidvegi et al., 2010).<br />

Autophagy is involved in the clearance <strong>of</strong><br />

aggregation-prone proteins in muscle diseases,<br />

including limb girdle muscular dystrophy type 2B,<br />

Miyoshi myopathy, <strong>and</strong> sporadic inclusion body<br />

myositis. Blockade <strong>of</strong> the autophagic pathway leads<br />

to the cardiomyopathy <strong>and</strong> myopathy <strong>of</strong> Danon<br />

disease (Ravikumar et al., 2010b). Autophagy is<br />

also involved in muscle atrophy but it is unclear<br />

whether autophagy has a beneficial effect by<br />

promoting survival during catabolic conditions, or a<br />

detrimental effect by causing atrophy. In the heart,<br />

basal autophagy is necessary to maintain cellular<br />

homeostasis <strong>and</strong> is upregulated in response to stress<br />

in hypertensive heart disease, heart failure, cardiac<br />

hypertrophy, <strong>and</strong> ischemia-reperfusion (Nakai et<br />

al., 2007).<br />

In the pancreas, autophagy is required to maintain<br />

the architecture <strong>and</strong> function <strong>of</strong> pancreatic β-cells<br />

(Ebato et al., 2008). Defective hepatic autophagy<br />

probably makes a major contribution to insulin<br />

resistance <strong>and</strong> to predisposition to type-2 diabetes<br />

<strong>and</strong> obesity (Yang et al., 2010).<br />

In infectious diseases, autophagy is involved in the<br />

elimination <strong>of</strong> intracellular pathogens (bacteria,<br />

viruses <strong>and</strong> parasites), <strong>and</strong> thus contributes to the<br />

innate immunity (Levine <strong>and</strong> Deretic, 2007; Virgin<br />

<strong>and</strong> Levine, 2009). Autophagy acts as an effector <strong>of</strong><br />

Toll-like receptor (TLR) signaling. TLR lig<strong>and</strong>s<br />

induce autophagy to promote the delivery <strong>of</strong><br />

infecting pathogens to the lysosomes (Levine <strong>and</strong><br />

Deretic, 2007). Autophagy contributes to adaptive<br />

immunity by generating antigenic peptides that are<br />

exposed on the cell surface in association with<br />

MHCAtg16L1 class II for presentation to CD4positive<br />

T cells, or by promoting the development<br />

<strong>and</strong> survival <strong>of</strong> B <strong>and</strong> T cells (Paludan et al., 2005;<br />

Pua et al., 2007). Recently polymorphisms <strong>of</strong> the<br />

genes that encode <strong>and</strong> IRGM, two autophagy genes<br />

essential for the elimination <strong>of</strong> intracellular<br />

pathogens, have been associated with Croh's<br />

disease, a chronic inflammatory bowel disease<br />

(Virgin <strong>and</strong> Levine, 2009).<br />

Cancer is frequently associated with defects in<br />

autophagy, but the role <strong>of</strong> autophagy in cancer is<br />

clearly complex, because autophagy is also required<br />

in the later stages <strong>of</strong> tumor progression to enable<br />

tumor cells to cope with metabolic stress (caused by<br />

limited supplies <strong>of</strong> oxygen <strong>and</strong> nutrients) (Levine,<br />

2007). The link between autophagy <strong>and</strong> cancer is<br />

further strengthened by the fact that several <strong>of</strong> the<br />

proteins involved in regulating autophagy are<br />

known to be tumor suppressor genes or<br />

oncoproteins (Morselli et al., 2009). Several <strong>of</strong> the<br />

functions <strong>of</strong> autophagy, such as the elimination <strong>of</strong><br />

defective organelles, which reduces oxidative stress<br />

<strong>and</strong> prevents DNA damage, also contribute to its<br />

tumor suppressor role (Mathew et al., 2009;<br />

Mathew et al., 2007). Remarkably, autophagy<br />

facilitates effective glucose uptake <strong>and</strong> glycolytic<br />

flux in Ras-transformed cells (Lock et al., 2011).<br />

Moreover, the loss <strong>of</strong> autophagy in Rastransformed<br />

cells is associated with reduced oxygen<br />

consumption <strong>and</strong> lower levels <strong>of</strong> the tricarboxylic<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 173


Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells Mehrpour M, et al.<br />

acid (TCA) intermediate (Guo et al., 2011). The<br />

high basal level <strong>of</strong> autophagy observed in tumors<br />

with Ras mutation is required for cancer cell<br />

survival (Yang et al., 2011). In these tumors,<br />

autophagy certainly constitutes an Achilles heel that<br />

could be used in the fight against cancer. More<br />

generally, inhibiting autophagy is a challenging<br />

concept; because in many tumors autophagy is a<br />

stress response to anti-cancer treatments (Kondo et<br />

al., 2005).<br />

References<br />

Yu QC, Marzella L. Modification <strong>of</strong> lysosomal proteolysis in<br />

mouse liver with taxol. Am J Pathol. 1986 Mar;122(3):553-<br />

61<br />

Mortimore GE, Pösö AR. Intracellular protein catabolism<br />

<strong>and</strong> its control during nutrient deprivation <strong>and</strong> supply. Annu<br />

Rev Nutr. 1987;7:539-64<br />

Høyvik H, Gordon PB, Berg TO, Strømhaug PE, Seglen<br />

PO. Inhibition <strong>of</strong> autophagic-lysosomal delivery <strong>and</strong><br />

autophagic lactolysis by asparagine. J Cell Biol. 1991<br />

Jun;113(6):1305-12<br />

Aplin A, Jasionowski T, Tuttle DL, Lenk SE, Dunn WA Jr.<br />

Cytoskeletal elements are required for the formation <strong>and</strong><br />

maturation <strong>of</strong> autophagic vacuoles. J Cell Physiol. 1992<br />

Sep;152(3):458-66<br />

Olkkonen VM, Dupree P, Killisch I, Lütcke A, Zerial M,<br />

Simons K. Molecular cloning <strong>and</strong> subcellular localization <strong>of</strong><br />

three GTP-binding proteins <strong>of</strong> the rab subfamily. J Cell Sci.<br />

1993 Dec;106 ( Pt 4):1249-61<br />

Strømhaug PE, Seglen PO. Evidence for acidity <strong>of</strong><br />

prelysosomal autophagic/endocytic vacuoles<br />

(amphisomes). Biochem J. 1993 Apr 1;291 ( Pt 1):115-21<br />

Blommaart EF, Luiken JJ, Blommaart PJ, van Woerkom<br />

GM, Meijer AJ. Phosphorylation <strong>of</strong> ribosomal protein S6 is<br />

inhibitory for autophagy in isolated rat hepatocytes. J Biol<br />

Chem. 1995 Feb 3;270(5):2320-6<br />

Lloyd JB. Metabolite efflux <strong>and</strong> influx across the lysosome<br />

membrane. Subcell Biochem. 1996;27:361-86<br />

Rothman JE, Wiel<strong>and</strong> FT. Protein sorting by transport<br />

vesicles. Science. 1996 Apr 12;272(5259):227-34<br />

Darsow T, Rieder SE, Emr SD. A multispecificity syntaxin<br />

homologue, Vam3p, essential for autophagic <strong>and</strong><br />

biosynthetic protein transport to the vacuole. J Cell Biol.<br />

1997 Aug 11;138(3):517-29<br />

Liou W, Geuze HJ, Geelen MJ, Slot JW. The autophagic<br />

<strong>and</strong> endocytic pathways converge at the nascent<br />

autophagic vacuoles. J Cell Biol. 1997 Jan 13;136(1):61-<br />

70<br />

Shirahama K, Noda T, Ohsumi Y. Mutational analysis <strong>of</strong><br />

Csc1/Vps4p: involvement <strong>of</strong> endosome in regulation <strong>of</strong><br />

autophagy in yeast. Cell Struct Funct. 1997 Oct;22(5):501-<br />

9<br />

Noda T, Ohsumi Y. Tor, a phosphatidylinositol kinase<br />

homologue, controls autophagy in yeast. J Biol Chem.<br />

1998 Feb 13;273(7):3963-6<br />

Yamamoto A, Tagawa Y, Yoshimori T, Moriyama Y,<br />

Masaki R, Tashiro Y. Bafilomycin A1 prevents maturation<br />

<strong>of</strong> autophagic vacuoles by inhibiting fusion between<br />

autophagosomes <strong>and</strong> lysosomes in rat hepatoma cell line,<br />

H-4-II-E cells. Cell Struct Funct. 1998 Feb;23(1):33-42<br />

Ciechanover A, Orian A, Schwartz AL. Ubiquitin-mediated<br />

proteolysis: biological regulation via destruction.<br />

Bioessays. 2000 May;22(5):442-51<br />

Tanaka Y, Guhde G, Suter A, Eskelinen EL, Hartmann D,<br />

Lüllmann-Rauch R, Janssen PM, Blanz J, von Figura K,<br />

Saftig P. Accumulation <strong>of</strong> autophagic vacuoles <strong>and</strong><br />

cardiomyopathy in LAMP-2-deficient mice. Nature. 2000<br />

Aug 24;406(6798):902-6<br />

Ishihara N, Hamasaki M, Yokota S, Suzuki K, Kamada Y,<br />

Kihara A, Yoshimori T, Noda T, Ohsumi Y.<br />

Autophagosome requires specific early Sec proteins for its<br />

formation <strong>and</strong> NSF/SNARE for vacuolar fusion. Mol Biol<br />

Cell. 2001 Nov;12(11):3690-702<br />

Mesa R, Salomón C, Roggero M, Stahl PD, Mayorga LS.<br />

Rab22a affects the morphology <strong>and</strong> function <strong>of</strong> the<br />

endocytic pathway. J Cell Sci. 2001 Nov;114(Pt 22):4041-9<br />

Mousavi SA, Kjeken R, Berg TO, Seglen PO, Berg T,<br />

Brech A. Effects <strong>of</strong> inhibitors <strong>of</strong> the vacuolar proton pump<br />

on hepatic heterophagy <strong>and</strong> autophagy. Biochim Biophys<br />

Acta. 2001 Feb 9;1510(1-2):243-57<br />

Wang Z, Wilson WA, Fujino MA, Roach PJ. Antagonistic<br />

controls <strong>of</strong> autophagy <strong>and</strong> glycogen accumulation by<br />

Snf1p, the yeast homolog <strong>of</strong> AMP-activated protein kinase,<br />

<strong>and</strong> the cyclin-dependent kinase Pho85p. Mol Cell Biol.<br />

2001 Sep;21(17):5742-52<br />

Weinmann AS, Bartley SM, Zhang T, Zhang MQ, Farnham<br />

PJ. Use <strong>of</strong> chromatin immunoprecipitation to clone novel<br />

E2F target promoters. Mol Cell Biol. 2001<br />

Oct;21(20):6820-32<br />

Browne GJ, Proud CG. Regulation <strong>of</strong> peptide-chain<br />

elongation in mammalian cells. Eur J Biochem. 2002<br />

Nov;269(22):5360-8<br />

Kim DH, Sarbassov DD, Ali SM, King JE, Latek RR,<br />

Erdjument-Bromage H, Tempst P, Sabatini DM. mTOR<br />

interacts with raptor to form a nutrient-sensitive complex<br />

that signals to the cell growth machinery. Cell. 2002 Jul<br />

26;110(2):163-75<br />

Nara A, Mizushima N, Yamamoto A, Kabeya Y, Ohsumi Y,<br />

Yoshimori T. SKD1 AAA ATPase-dependent endosomal<br />

transport is involved in autolysosome formation. Cell Struct<br />

Funct. 2002 Feb;27(1):29-37<br />

<strong>Atlas</strong>hkin V, Kreykenbohm V, Eskelinen EL, Wenzel D,<br />

Fayyazi A, Fischer von Mollard G. Deletion <strong>of</strong> the SNARE<br />

vti1b in mice results in the loss <strong>of</strong> a single SNARE partner,<br />

syntaxin 8. Mol Cell Biol. 2003 Aug;23(15):5198-207<br />

Eskelinen EL, Tanaka Y, Saftig P. At the acidic edge:<br />

emerging functions for lysosomal membrane proteins.<br />

Trends Cell Biol. 2003 Mar;13(3):137-45<br />

Kim DH, Sarbassov DD, Ali SM, Latek RR, Guntur KV,<br />

Erdjument-Bromage H, Tempst P, Sabatini DM. GbetaL, a<br />

positive regulator <strong>of</strong> the rapamycin-sensitive pathway<br />

required for the nutrient-sensitive interaction between<br />

raptor <strong>and</strong> mTOR. Mol Cell. 2003 Apr;11(4):895-904<br />

Klionsky DJ, Cregg JM, Dunn WA Jr, Emr SD, Sakai Y,<br />

S<strong>and</strong>oval IV, Sibirny A, Subramani S, Thumm M, Veenhuis<br />

M, Ohsumi Y. A unified nomenclature for yeast autophagyrelated<br />

genes. Dev Cell. 2003 Oct;5(4):539-45<br />

Browne GJ, Finn SG, Proud CG. Stimulation <strong>of</strong> the AMPactivated<br />

protein kinase leads to activation <strong>of</strong> eukaryotic<br />

elongation factor 2 kinase <strong>and</strong> to its phosphorylation at a<br />

novel site, serine 398. J Biol Chem. 2004 Mar<br />

26;279(13):12220-31<br />

Corradetti MN, Inoki K, Bardeesy N, DePinho RA, Guan<br />

KL. Regulation <strong>of</strong> the TSC pathway by LKB1: evidence <strong>of</strong><br />

a molecular link between tuberous sclerosis complex <strong>and</strong><br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 174


Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells Mehrpour M, et al.<br />

Peutz-Jeghers syndrome. <strong>Gene</strong>s Dev. 2004 Jul<br />

1;18(13):1533-8<br />

Eskelinen EL, Schmidt CK, Neu S, Willenborg M, Fuertes<br />

G, Salvador N, Tanaka Y, Lüllmann-Rauch R, Hartmann<br />

D, Heeren J, von Figura K, Knecht E, Saftig P. Disturbed<br />

cholesterol traffic but normal proteolytic function in LAMP-<br />

1/LAMP-2 double-deficient fibroblasts. Mol Biol Cell. 2004<br />

Jul;15(7):3132-45<br />

Fengsrud, M., Lunde Sneve, M., yverbye, A., <strong>and</strong> Seglen,<br />

P.O. Structural aspects <strong>of</strong> mammalian autophagy In<br />

Autophagy, D.J. Klionsky, ed. (Georgetown, TX, L<strong>and</strong>es<br />

Bioscience), 2004;11-25.<br />

Gutierrez MG, Munafo DB, Beron W, Colombo MI.. Rab7<br />

is required for the normal progression <strong>of</strong> the autophagic<br />

pathway in mammalian cells. J Cell Sci. 2004 Jun 1;117(Pt<br />

13):2687-97. Epub 2004 May 11.<br />

Hardie DG.. The AMP-activated protein kinase pathway-new<br />

players upstream <strong>and</strong> downstream. J Cell Sci. 2004<br />

Nov 1;117(Pt 23):5479-87. (REVIEW)<br />

Jager S, Bucci C, Tanida I, Ueno T, Kominami E, Saftig P,<br />

Eskelinen EL.. Role for Rab7 in maturation <strong>of</strong> late<br />

autophagic vacuoles. J Cell Sci. 2004 Sep 15;117(Pt<br />

20):4837-48. Epub 2004 Aug 31.<br />

Kuma A, Hatano M, Matsui M, Yamamoto A, Nakaya H,<br />

Yoshimori T, Ohsumi Y, Tokuhisa T, Mizushima N.. The<br />

role <strong>of</strong> autophagy during the early neonatal starvation<br />

period. Nature. 2004 Dec 23;432(7020):1032-6. Epub<br />

2004 Nov 3.<br />

Meijer AJ, Codogno P.. Regulation <strong>and</strong> role <strong>of</strong> autophagy<br />

in mammalian cells. Int J Biochem Cell Biol. 2004<br />

Dec;36(12):2445-62. (REVIEW)<br />

Meijer AJ, Dubbelhuis PF.. Amino acid signalling <strong>and</strong> the<br />

integration <strong>of</strong> metabolism. Biochem Biophys Res Commun.<br />

2004 Jan 9;313(2):397-403. (REVIEW)<br />

Oshiro N, Yoshino K, Hidayat S, Tokunaga C, Hara K,<br />

Eguchi S, Avruch J, Yonezawa K.. Dissociation <strong>of</strong> raptor<br />

from mTOR is a mechanism <strong>of</strong> rapamycin-induced<br />

inhibition <strong>of</strong> mTOR function. <strong>Gene</strong>s Cells. 2004<br />

Apr;9(4):359-66.<br />

Scarlatti F, Bauvy C, Ventruti A, Sala G, Cluzeaud F,<br />

V<strong>and</strong>ewalle A, Ghidoni R, Codogno P.. Ceramidemediated<br />

macroautophagy involves inhibition <strong>of</strong> protein<br />

kinase B <strong>and</strong> up-regulation <strong>of</strong> beclin 1. J Biol Chem. 2004<br />

Apr 30;279(18):18384-91. Epub 2004 Feb 17.<br />

Scott RC, Schuldiner O, Neufeld TP.. Role <strong>and</strong> regulation<br />

<strong>of</strong> starvation-induced autophagy in the Drosophila fat body.<br />

Dev Cell. 2004 Aug;7(2):167-78.<br />

Shaw RJ, Bardeesy N, Manning BD, Lopez L, Kosmatka<br />

M, DePinho RA, Cantley LC.. The LKB1 tumor suppressor<br />

negatively regulates mTOR signaling. Cancer Cell. 2004<br />

Jul;6(1):91-9.<br />

Webb JL, Ravikumar B, Rubinsztein DC.. Microtubule<br />

disruption inhibits autophagosome-lysosome fusion:<br />

implications for studying the roles <strong>of</strong> aggresomes in<br />

polyglutamine diseases. Int J Biochem Cell Biol. 2004<br />

Dec;36(12):2541-50.<br />

Boya P, Gonzalez-Polo RA, Casares N, Perfettini JL,<br />

Dessen P, Larochette N, Metivier D, Meley D, Souquere S,<br />

Yoshimori T, Pierron G, Codogno P, Kroemer G.. nhibition<br />

<strong>of</strong> macroautophagy triggers apoptosis. Mol Cell Biol. 2005<br />

Feb;25(3):1025-40.<br />

Byfield MP, Murray JT, Backer JM.. hVps34 is a nutrientregulated<br />

lipid kinase required for activation <strong>of</strong> p70 S6<br />

kinase. J Biol Chem. 2005 Sep 23;280(38):33076-82.<br />

Epub 2005 Jul 27.<br />

Codogno P, Meijer AJ.. Autophagy <strong>and</strong> signaling: their role<br />

in cell survival <strong>and</strong> cell death. Cell Death Differ. 2005<br />

Nov;12 Suppl 2:1509-18. (REVIEW)<br />

Egami Y, Kiryu-Seo S, Yoshimori T, Kiyama H.. Induced<br />

expressions <strong>of</strong> Rab24 GTPase <strong>and</strong> LC3 in nerve-injured<br />

motor neurons. Biochem Biophys Res Commun. 2005 Dec<br />

2;337(4):1206-13. Epub 2005 Oct 6.<br />

Eskelinen EL.. Maturation <strong>of</strong> autophagic vacuoles in<br />

Mammalian cells. Autophagy. 2005 Apr;1(1):1-10. Epub<br />

2005 Apr 28. (REVIEW)<br />

Feng Z, Zhang H, Levine AJ, Jin S.. The coordinate<br />

regulation <strong>of</strong> the p53 <strong>and</strong> mTOR pathways in cells. Proc<br />

Natl Acad Sci U S A. 2005 Jun 7;102(23):8204-9. Epub<br />

2005 May 31.<br />

Hawley SA, Pan DA, Mustard KJ, Ross L, Bain J, Edelman<br />

AM, Frenguelli BG, Hardie DG.. Calmodulin-dependent<br />

protein kinase kinase-beta is an alternative upstream<br />

kinase for AMP-activated protein kinase. Cell Metab. 2005<br />

Jul;2(1):9-19.<br />

Klionsky DJ, Meijer AJ, Codogno P.. Autophagy <strong>and</strong><br />

p70S6 kinase. Autophagy. 2005 Apr;1(1):59-60; discussion<br />

60-1. Epub 2005 Apr 5.<br />

Koike M, Shibata M, Waguri S, Yoshimura K, Tanida I,<br />

Kominami E, Gotow T, Peters C, von Figura K, Mizushima<br />

N, Saftig P, Uchiyama Y.. Participation <strong>of</strong> autophagy in<br />

storage <strong>of</strong> lysosomes in neurons from mouse models <strong>of</strong><br />

neuronal ceroid-lip<strong>of</strong>uscinoses (Batten disease). Am J<br />

Pathol. 2005 Dec;167(6):1713-28.<br />

Kondo Y, Kanzawa T, Sawaya R, Kondo S.. The role <strong>of</strong><br />

autophagy in cancer development <strong>and</strong> response to<br />

therapy. Nat Rev Cancer. 2005 Sep;5(9):726-34.<br />

(REVIEW)<br />

Manalo DJ, Rowan A, Lavoie T, Natarajan L, Kelly BD, Ye<br />

SQ, Garcia JG, Semenza GL.. Transcriptional regulation <strong>of</strong><br />

vascular endothelial cell responses to hypoxia by HIF-1.<br />

Blood. 2005 Jan 15;105(2):659-69. Epub 2004 Sep 16.<br />

Nobukuni T, Joaquin M, Roccio M, Dann SG, Kim SY,<br />

Gulati P, Byfield MP, Backer JM, Natt F, Bos JL,<br />

Zwartkruis FJ, Thomas G.. Amino acids mediate<br />

mTOR/raptor signaling through activation <strong>of</strong> class 3<br />

phosphatidylinositol 3OH-kinase. Proc Natl Acad Sci U S<br />

A. 2005 Oct 4;102(40):14238-43. Epub 2005 Sep 21.<br />

Paludan C, Schmid D, L<strong>and</strong>thaler M, Vockerodt M, Kube<br />

D, Tuschl T, Munz C.. Endogenous MHC class II<br />

processing <strong>of</strong> a viral nuclear antigen after autophagy.<br />

Science. 2005 Jan 28;307(5709):593-6. Epub 2004 Dec 9.<br />

Pattingre S, Tassa A, Qu X, Garuti R, Liang XH,<br />

Mizushima N, Packer M, Schneider MD, Levine B.. Bcl-2<br />

antiapoptotic proteins inhibit Beclin 1-dependent<br />

autophagy. Cell. 2005 Sep 23;122(6):927-39.<br />

Ravikumar B, Acevedo-Arozena A, Imarisio S, Berger Z,<br />

Vacher C, O'Kane CJ, Brown SD, Rubinsztein DC.. Dynein<br />

mutations impair autophagic clearance <strong>of</strong> aggregate-prone<br />

proteins. Nat <strong>Gene</strong>t. 2005 Jul;37(7):771-6. Epub 2005 Jun<br />

26.<br />

Sarkar S, Floto RA, Berger Z, Imarisio S, Cordenier A,<br />

Pasco M, Cook LJ, Rubinsztein DC.. Lithium induces<br />

autophagy by inhibiting inositol monophosphatase. J Cell<br />

Biol. 2005 Sep 26;170(7):1101-11.<br />

Woods A, Dickerson K, Heath R, Hong SP, Momcilovic M,<br />

Johnstone SR, Carlson M, Carling D.. Ca2+/calmodulindependent<br />

protein kinase kinase-beta acts upstream <strong>of</strong><br />

AMP-activated protein kinase in mammalian cells. Cell<br />

Metab. 2005 Jul;2(1):21-33.<br />

Bensaad K, Tsuruta A, Selak MA, Vidal MN, Nakano K,<br />

Bartrons R, Gottlieb E, Vousden KH.. TIGAR, a p53-<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 175


Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells Mehrpour M, et al.<br />

inducible regulator <strong>of</strong> glycolysis <strong>and</strong> apoptosis. Cell. 2006<br />

Jul 14;126(1):107-20.<br />

Crighton D, Wilkinson S, O'Prey J, Syed N, Smith P,<br />

Harrison PR, Gasco M, Garrone O, Crook T, Ryan KM..<br />

DRAM, a p53-induced modulator <strong>of</strong> autophagy, is critical<br />

for apoptosis. Cell. 2006 Jul 14;126(1):121-34.<br />

Degenhardt K, Mathew R, Beaudoin B, Bray K, Anderson<br />

D, Chen G, Mukherjee C, Shi Y, Gelinas C, Fan Y, Nelson<br />

DA, Jin S, White E.. Autophagy promotes tumor cell<br />

survival <strong>and</strong> restricts necrosis, inflammation, <strong>and</strong><br />

tumorigenesis. Cancer Cell. 2006 Jul;10(1):51-64.<br />

DeGregori J, Johnson DG.. Distinct <strong>and</strong> Overlapping Roles<br />

for E2F Family Members in Transcription, Proliferation <strong>and</strong><br />

Apoptosis. Curr Mol Med. 2006 Nov;6(7):739-48.<br />

(REVIEW)<br />

Djavaheri-Mergny M, Amelotti M, Mathieu J, Besan‡on F,<br />

Bauvy C, Souquere S, Pierron G, Codogno P.. NF-kappaB<br />

activation represses tumor necrosis factor-alpha-induced<br />

autophagy. J Biol Chem. 2006 Oct 13;281(41):30373-82.<br />

Epub 2006 Jul 20.<br />

Fass E, Shvets E, Degani I, Hirschberg K, Elazar Z..<br />

Microtubules support production <strong>of</strong> starvation-induced<br />

autophagosomes but not their targeting <strong>and</strong> fusion with<br />

lysosomes. J Biol Chem. 2006 Nov 24;281(47):36303-16.<br />

Epub 2006 Sep 8.<br />

Hait WN, Wu H, Jin S, Yang JM.. Elongation factor-2<br />

kinase: its role in protein synthesis <strong>and</strong> autophagy.<br />

Autophagy. 2006 Oct-Dec;2(4):294-6. Epub 2006 Oct 3.<br />

Karin M.. Nuclear factor-kappaB in cancer development<br />

<strong>and</strong> progression. Nature. 2006 May 25;441(7092):431-6.<br />

(REVIEW)<br />

Kochl R, Hu XW, Chan EY, Tooze SA.. Microtubules<br />

facilitate autophagosome formation <strong>and</strong> fusion <strong>of</strong><br />

autophagosomes with endosomes. Traffic. 2006<br />

Feb;7(2):129-45.<br />

Kotoulas OB, Kalamidas SA, Kondomerkos DJ.. Glycogen<br />

autophagy in glucose homeostasis. Pathol Res Pract.<br />

2006;202(9):631-8. Epub 2006 Jun 16. (REVIEW)<br />

Laderoute KR, Amin K, Calaoagan JM, Knapp M, Le T,<br />

Orduna J, Foretz M, Viollet B.. 5'-AMP-activated protein<br />

kinase (AMPK) is induced by low-oxygen <strong>and</strong> glucose<br />

deprivation conditions found in solid-tumor<br />

microenvironments. Mol Cell Biol. 2006 Jul;26(14):5336-<br />

47.<br />

Liang C, Feng P, Ku B, Dotan I, Canaani D, Oh BH, Jung<br />

JU.. Autophagic <strong>and</strong> tumour suppressor activity <strong>of</strong> a novel<br />

Beclin1-binding protein UVRAG. Nat Cell Biol. 2006<br />

Jul;8(7):688-99. Epub 2006 Jun 25.<br />

Massey AC, Kaushik S, Sovak G, Kiffin R, Cuervo AM..<br />

Consequences <strong>of</strong> the selective blockage <strong>of</strong> chaperonemediated<br />

autophagy. Proc Natl Acad Sci U S A. 2006 Apr<br />

11;103(15):5805-10. Epub 2006 Apr 3.<br />

Meley D, Bauvy C, Houben-Weerts JH, Dubbelhuis PF,<br />

Helmond MT, Codogno P, Meijer AJ.. AMP-activated<br />

protein kinase <strong>and</strong> the regulation <strong>of</strong> autophagic proteolysis.<br />

J Biol Chem. 2006 Nov 17;281(46):34870-9. Epub 2006<br />

Sep 21.<br />

Sarbassov DD, Ali SM, Sengupta S, Sheen JH, Hsu PP,<br />

Bagley AF, Markhard AL, Sabatini DM.. Prolonged<br />

rapamycin treatment inhibits mTORC2 assembly <strong>and</strong><br />

Akt/PKB. Mol Cell. 2006 Apr 21;22(2):159-68. Epub 2006<br />

Apr 6.<br />

Wu H, Yang JM, Jin S, Zhang H, Hait WN.. Elongation<br />

factor-2 kinase regulates autophagy in human<br />

glioblastoma cells. Cancer Res. 2006 Mar 15;66(6):3015-<br />

23.<br />

Yang Z, Huang J, Geng J, Nair U, Klionsky DJ.. Atg22<br />

recycles amino acids to link the degradative <strong>and</strong> recycling<br />

functions <strong>of</strong> autophagy. Mol Biol Cell. 2006<br />

Dec;17(12):5094-104. Epub 2006 Oct 4.<br />

Yousefi S, Perozzo R, Schmid I, Ziemiecki A, Schaffner T,<br />

Scapozza L, Brunner T, Simon HU.. Calpain-mediated<br />

cleavage <strong>of</strong> Atg5 switches autophagy to apoptosis. Nat<br />

Cell Biol. 2006 Oct;8(10):1124-32. Epub 2006 Sep 24.<br />

Bohensky J, Shapiro IM, Leshinsky S, Terkhorn SP,<br />

Adams CS, Srinivas V.. HIF-1 regulation <strong>of</strong> chondrocyte<br />

apoptosis: induction <strong>of</strong> the autophagic pathway.<br />

Autophagy. 2007 May-Jun;3(3):207-14. Epub 2007 May<br />

14.<br />

Colell A, Ricci JE, Tait S, Milasta S, Maurer U, Bouchier-<br />

Hayes L, Fitzgerald P, Guio-Carrion A, Waterhouse NJ, Li<br />

CW, Mari B, Barbry P, Newmeyer DD, Beere HM, Green<br />

DR.. GAPDH <strong>and</strong> autophagy preserve survival after<br />

apoptotic cytochrome c release in the absence <strong>of</strong> caspase<br />

activation. Cell. 2007 Jun 1;129(5):983-97.<br />

Erlich S, Mizrachy L, Segev O, Lindenboim L, Zmira O,<br />

Adi-Harel S, Hirsch JA, Stein R, Pinkas-Kramarski R..<br />

Differential interactions between Beclin 1 <strong>and</strong> Bcl-2 family<br />

members. Autophagy. 2007 Nov-Dec;3(6):561-8. Epub<br />

2007 Jul 8.<br />

Fabre C, Carvalho G, Tasdemir E, Braun T, Ades L,<br />

Grosjean J, Boehrer S, Metivier D, Souquere S, Pierron G,<br />

Fenaux P, Kroemer G.. NF-kappaB inhibition sensitizes to<br />

starvation-induced cell death in high-risk myelodysplastic<br />

syndrome <strong>and</strong> acute myeloid leukemia. Oncogene. 2007<br />

Jun 14;26(28):4071-83. Epub 2007 Jan 8.<br />

Filimonenko M, Stuffers S, Raiborg C, Yamamoto A,<br />

Maler›d L, Fisher EM, Isaacs A, Brech A, Stenmark H,<br />

Simonsen A.. Functional multivesicular bodies are required<br />

for autophagic clearance <strong>of</strong> protein aggregates associated<br />

with neurodegenerative disease. J Cell Biol. 2007 Nov<br />

5;179(3):485-500.<br />

Forgac M.. Vacuolar ATPases: rotary proton pumps in<br />

physiology <strong>and</strong> pathophysiology. Nat Rev Mol Cell Biol.<br />

2007 Nov;8(11):917-29. (REVIEW)<br />

Fujita E, Kouroku Y, Isoai A, Kumagai H, Misutani A,<br />

Matsuda C, Hayashi YK, Momoi T.. Two endoplasmic<br />

reticulum-associated degradation (ERAD) systems for the<br />

novel variant <strong>of</strong> the mutant dysferlin: ubiquitin/proteasome<br />

ERAD(I) <strong>and</strong> autophagy/lysosome ERAD(II). Hum Mol<br />

<strong>Gene</strong>t. 2007 Mar 15;16(6):618-29. Epub 2007 Mar 1.<br />

Gurkan C, Koulov AV, Balch WE.. An evolutionary<br />

perspective on eukaryotic membrane trafficking. Adv Exp<br />

Med Biol. 2007;607:73-83. (REVIEW)<br />

Hoyer-Hansen M, Bastholm L, Szyniarowski P,<br />

Campanella M, Szabadkai G, Farkas T, Bianchi K,<br />

Fehrenbacher N, Elling F, Rizzuto R, Mathiasen IS,<br />

Jaattela M.. Control <strong>of</strong> macroautophagy by calcium,<br />

calmodulin-dependent kinase kinase-beta, <strong>and</strong> Bcl-2. Mol<br />

Cell. 2007 Jan 26;25(2):193-205.<br />

Hoyer-Hansen M, Jaattela M.. AMP-activated protein<br />

kinase: a universal regulator <strong>of</strong> autophagy? Autophagy.<br />

2007 Jul-Aug;3(4):381-3. Epub 2007 Jul 5.<br />

Klionsky DJ.. Autophagy: from phenomenology to<br />

molecular underst<strong>and</strong>ing in less than a decade. Nat Rev<br />

Mol Cell Biol. 2007 Nov;8(11):931-7. (REVIEW)<br />

Kouroku Y, Fujita E, Tanida I, Ueno T, Isoai A, Kumagai H,<br />

Ogawa S, Kaufman RJ, Kominami E, Momoi T.. ER stress<br />

(PERK/eIF2alpha phosphorylation) mediates the<br />

polyglutamine-induced LC3 conversion, an essential step<br />

for autophagy formation. Cell Death Differ. 2007<br />

Feb;14(2):230-9. Epub 2006 Jun 23.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 176


Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells Mehrpour M, et al.<br />

Lee JA, Beigneux A, Ahmad ST, Young SG, Gao FB..<br />

ESCRT-III dysfunction causes autophagosome<br />

accumulation <strong>and</strong> neurodegeneration. Curr Biol. 2007 Sep<br />

18;17(18):1561-7. Epub 2007 Aug 2.<br />

Levine B.. Cell biology: autophagy <strong>and</strong> cancer. Nature.<br />

2007 Apr 12;446(7137):745-7.<br />

Levine B, Deretic V.. Unveiling the roles <strong>of</strong> autophagy in<br />

innate <strong>and</strong> adaptive immunity. Nat Rev Immunol. 2007<br />

Oct;7(10):767-77. (REVIEW)<br />

Maiuri MC, Le Toumelin G, Criollo A, Rain JC, Gautier F,<br />

Juin P, Tasdemir E, Pierron G, Troulinaki K, Tavernarakis<br />

N, Hickman JA, <strong>Gene</strong>ste O, Kroemer G.. Functional <strong>and</strong><br />

physical interaction between Bcl-X(L) <strong>and</strong> a BH3-like<br />

domain in Beclin-1. EMBO J. 2007 May 16;26(10):2527-<br />

39. Epub 2007 Apr 19.<br />

Mammucari C, Milan G, Romanello V, Masiero E, Rudolf<br />

R, Del Piccolo P, Burden SJ, Di Lisi R, S<strong>and</strong>ri C, Zhao J,<br />

Goldberg AL, Schiaffino S, S<strong>and</strong>ri M.. FoxO3 controls<br />

autophagy in skeletal muscle in vivo. Cell Metab. 2007<br />

Dec;6(6):458-71.<br />

Mathew R, Kongara S, Beaudoin B, Karp CM, Bray K,<br />

Degenhardt K, Chen G, Jin S, White E.. Autophagy<br />

suppresses tumor progression by limiting chromosomal<br />

instability. <strong>Gene</strong>s Dev. 2007 Jun 1;21(11):1367-81. Epub<br />

2007 May 17.<br />

Meijer AJ, Codogno P.. AMP-activated protein kinase <strong>and</strong><br />

autophagy. Autophagy. 2007 May-Jun;3(3):238-40. Epub<br />

2007 May 14.<br />

Nakai A, Yamaguchi O, Takeda T, Higuchi Y, Hikoso S,<br />

Taniike M, Omiya S, Mizote I, Matsumura Y, Asahi M,<br />

Nishida K, Hori M, Mizushima N, Otsu K.. The role <strong>of</strong><br />

autophagy in cardiomyocytes in the basal state <strong>and</strong> in<br />

response to hemodynamic stress. Nat Med. 2007<br />

May;13(5):619-24. Epub 2007 Apr 22.<br />

Nakatogawa H, Ichimura Y, Ohsumi Y.. Atg8, a ubiquitinlike<br />

protein required for autophagosome formation,<br />

mediates membrane tethering <strong>and</strong> hemifusion. Cell. 2007<br />

Jul 13;130(1):165-78.<br />

Pua HH, Dzhagalov I, Chuck M, Mizushima N, He YW.. A<br />

critical role for the autophagy gene Atg5 in T cell survival<br />

<strong>and</strong> proliferation. J Exp Med. 2007 Jan 22;204(1):25-31.<br />

Epub 2006 Dec 26.<br />

Rusten TE, Vaccari T, Lindmo K, Rodahl LM, Nezis IP,<br />

Sem-Jacobsen C, Wendler F, Vincent JP, Brech A, Bilder<br />

D, Stenmark H.. ESCRTs <strong>and</strong> Fab1 regulate distinct steps<br />

<strong>of</strong> autophagy. Curr Biol. 2007 Oct 23;17(20):1817-25.<br />

Epub 2007 Oct 11.<br />

Scherz-Shouval R, Shvets E, Fass E, Shorer H, Gil L,<br />

Elazar Z.. Reactive oxygen species are essential for<br />

autophagy <strong>and</strong> specifically regulate the activity <strong>of</strong> Atg4.<br />

EMBO J. 2007 Apr 4;26(7):1749-60. Epub 2007 Mar 8.<br />

Tamai K, Tanaka N, Nara A, Yamamoto A, Nakagawa I,<br />

Yoshimori T, Ueno Y, Shimosegawa T, Sugamura K.. Role<br />

<strong>of</strong> Hrs in maturation <strong>of</strong> autophagosomes in mammalian<br />

cells. Biochem Biophys Res Commun. 2007 Sep<br />

7;360(4):721-7. Epub 2007 Jul 5.<br />

White SR, Lauring B.. AAA+ ATPases: achieving diversity<br />

<strong>of</strong> function with conserved machinery. Traffic. 2007<br />

Dec;8(12):1657-67. Epub 2007 Sep 26. (REVIEW)<br />

Zhao J, Brault JJ, Schild A, Cao P, S<strong>and</strong>ri M, Schiaffino S,<br />

Lecker SH, Goldberg AL.. FoxO3 coordinately activates<br />

protein degradation by the autophagic/lysosomal <strong>and</strong><br />

proteasomal pathways in atrophying muscle cells. Cell<br />

Metab. 2007 Dec;6(6):472-83.<br />

Axe EL, Walker SA, Manifava M, Ch<strong>and</strong>ra P, Roderick HL,<br />

Habermann A, Griffiths G, Ktistakis NT.. Autophagosome<br />

formation from membrane compartments enriched in<br />

phosphatidylinositol 3-phosphate <strong>and</strong> dynamically<br />

connected to the endoplasmic reticulum. J Cell Biol. 2008<br />

Aug 25;182(4):685-701.<br />

Beau I, Esclatine A, Codogno P.. Lost to translation: when<br />

autophagy targets mature ribosomes. Trends Cell Biol.<br />

2008 Jul;18(7):311-4. Epub 2008 May 27. (REVIEW)<br />

Ebato C, Uchida T, Arakawa M, Komatsu M, Ueno T,<br />

Komiya K, Azuma K, Hirose T, Tanaka K, Kominami E,<br />

Kawamori R, Fujitani Y, Watada H.. Autophagy is<br />

important in islet homeostasis <strong>and</strong> compensatory increase<br />

<strong>of</strong> beta cell mass in response to high-fat diet. Cell Metab.<br />

2008 Oct;8(4):325-32.<br />

Fader CM, Sanchez D, Furlan M, Colombo MI.. Induction<br />

<strong>of</strong> autophagy promotes fusion <strong>of</strong> multivesicular bodies with<br />

autophagic vacuoles in k562 cells. Traffic. 2008<br />

Feb;9(2):230-50. Epub 2007 Dec 7.<br />

Jahreiss L, Menzies FM, Rubinsztein DC.. The itinerary <strong>of</strong><br />

autophagosomes: from peripheral formation to kiss-<strong>and</strong>run<br />

fusion with lysosomes. Traffic. 2008 Apr;9(4):574-87.<br />

Epub 2008 Jan 7.<br />

Kaushik S, Massey AC, Mizushima N, Cuervo AM..<br />

Constitutive activation <strong>of</strong> chaperone-mediated autophagy<br />

in cells with impaired macroautophagy. Mol Biol Cell. 2008<br />

May;19(5):2179-92. Epub 2008 Mar 12.<br />

Kim E, Goraksha-Hicks P, Li L, Neufeld TP, Guan KL..<br />

Regulation <strong>of</strong> TORC1 by Rag GTPases in nutrient<br />

response. Nat Cell Biol. 2008a Aug;10(8):935-45. Epub<br />

2008 Jul 6.<br />

Kim JS, Nitta T, Mohuczy D, O'Malley KA, Moldawer LL,<br />

Dunn WA Jr, Behrns KE.. Impaired autophagy: A<br />

mechanism <strong>of</strong> mitochondrial dysfunction in anoxic rat<br />

hepatocytes. Hepatology. 2008b May;47(5):1725-36.<br />

Kimura S, Noda T, Yoshimori T.. Dynein-dependent<br />

movement <strong>of</strong> autophagosomes mediates efficient<br />

encounters with lysosomes. Cell Struct Funct.<br />

2008;33(1):109-22. Epub 2008 Apr 4.<br />

Lee IH, Cao L, Mostoslavsky R, Lombard DB, Liu J, Bruns<br />

NE, Tsokos M, Alt FW, Finkel T.. A role for the NADdependent<br />

deacetylase Sirt1 in the regulation <strong>of</strong><br />

autophagy. Proc Natl Acad Sci U S A. 2008 Mar<br />

4;105(9):3374-9. Epub 2008 Feb 22.<br />

Liang C, Lee JS, Inn KS, Gack MU, Li Q, Roberts EA,<br />

Vergne I, Deretic V, Feng P, Akazawa C, Jung JU..<br />

Beclin1-binding UVRAG targets the class C Vps complex<br />

to coordinate autophagosome maturation <strong>and</strong> endocytic<br />

trafficking. Nat Cell Biol. 2008 Jul;10(7):776-87. Epub 2008<br />

Jun 15.<br />

Mizushima N, Levine B, Cuervo AM, Klionsky DJ..<br />

Autophagy fights disease through cellular self-digestion.<br />

Nature. 2008 Feb 28;451(7182):1069-75. (REVIEW)<br />

Nedelsky NB, Todd PK, Taylor JP.. Autophagy <strong>and</strong> the<br />

ubiquitin-proteasome system: collaborators in<br />

neuroprotection. Biochim Biophys Acta. 2008<br />

Dec;1782(12):691-9. Epub 2008 Oct 10. (REVIEW)<br />

Nedjic J, Aichinger M, Emmerich J, Mizushima N, Klein L..<br />

Autophagy in thymic epithelium shapes the T-cell<br />

repertoire <strong>and</strong> is essential for tolerance. Nature. 2008 Sep<br />

18;455(7211):396-400. Epub 2008 Aug 13.<br />

Noda NN, Kumeta H, Nakatogawa H, Satoo K, Adachi W,<br />

Ishii J, Fujioka Y, Ohsumi Y, Inagaki F.. Structural basis <strong>of</strong><br />

target recognition by Atg8/LC3 during selective autophagy.<br />

<strong>Gene</strong>s Cells. 2008 Dec;13(12):1211-8. Epub 2008 Oct 22.<br />

Polager S, Ofir M, Ginsberg D.. E2F1 regulates autophagy<br />

<strong>and</strong> the transcription <strong>of</strong> autophagy genes. Oncogene. 2008<br />

Aug 14;27(35):4860-4. Epub 2008 Apr 14.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 177


Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells Mehrpour M, et al.<br />

Salih DA, Brunet A.. FoxO transcription factors in the<br />

maintenance <strong>of</strong> cellular homeostasis during aging. Curr<br />

Opin Cell Biol. 2008 Apr;20(2):126-36. Epub 2008 Apr 3.<br />

Sancak Y, Peterson TR, Shaul YD, Lindquist RA, Thoreen<br />

CC, Bar-Peled L, Sabatini DM.. The Rag GTPases bind<br />

raptor <strong>and</strong> mediate amino acid signaling to mTORC1.<br />

Science. 2008 Jun 13;320(5882):1496-501. Epub 2008<br />

May 22.<br />

Schlottmann S, Buback F, Stahl B, Meierhenrich R, Walter<br />

P, Georgieff M, Senftleben U.. Prolonged classical NFkappaB<br />

activation prevents autophagy upon E. coli<br />

stimulation in vitro: a potential resolving mechanism <strong>of</strong><br />

inflammation. Mediators Inflamm. 2008;2008:725854.<br />

Tasdemir E, Maiuri MC, Galluzzi L, Vitale I, Djavaheri-<br />

Mergny M, D'Amelio M, Criollo A, Morselli E, Zhu C,<br />

Harper F, Nannmark U, Samara C, Pinton P, Vicencio JM,<br />

Carnuccio R, Moll UM, Madeo F, Paterlini-Brechot P,<br />

Rizzuto R, Szabadkai G, Pierron G, Blomgren K,<br />

Tavernarakis N, Codogno P, Cecconi F, Kroemer G..<br />

Regulation <strong>of</strong> autophagy by cytoplasmic p53. Nat Cell Biol.<br />

2008 Jun;10(6):676-87. Epub 2008 May 4.<br />

Tsukamoto S, Kuma A, Murakami M, Kishi C, Yamamoto<br />

A, Mizushima N.. Autophagy is essential for<br />

preimplantation development <strong>of</strong> mouse embryos. Science.<br />

2008 Jul 4;321(5885):117-20.<br />

Uttenweiler A, Mayer A.. Microautophagy in the yeast<br />

Saccharomyces cerevisiae. Methods Mol Biol.<br />

2008;445:245-59.<br />

van der Vaart A, Mari M, Reggiori F.. A picky eater:<br />

exploring the mechanisms <strong>of</strong> selective autophagy in<br />

human pathologies. Traffic. 2008 Mar;9(3):281-9. Epub<br />

2007 Nov 6. (REVIEW)<br />

Wang J, Lian H, Zhao Y, Kauss MA, Spindel S.. Vitamin<br />

D3 induces autophagy <strong>of</strong> human myeloid leukemia cells. J<br />

Biol Chem. 2008 Sep 12;283(37):25596-605. Epub 2008<br />

Jul 15.<br />

Wei Y, Pattingre S, Sinha S, Bassik M, Levine B.. JNK1mediated<br />

phosphorylation <strong>of</strong> Bcl-2 regulates starvationinduced<br />

autophagy. Mol Cell. 2008 Jun 20;30(6):678-88.<br />

Williams A, Sarkar S, Cuddon P, Tt<strong>of</strong>i EK, Saiki S, Siddiqi<br />

FH, Jahreiss L, Fleming A, Pask D, Goldsmith P, O'Kane<br />

CJ, Floto RA, Rubinsztein DC.. Novel targets for<br />

Huntington's disease in an mTOR-independent autophagy<br />

pathway. Nat Chem Biol. 2008 May;4(5):295-305.<br />

Yogalingam G, Pendergast AM.. Abl kinases regulate<br />

autophagy by promoting the trafficking <strong>and</strong> function <strong>of</strong><br />

lysosomal components. J Biol Chem. 2008 Dec<br />

19;283(51):35941-53. Epub 2008 Oct 21.<br />

Zhang C, Cuervo AM.. Restoration <strong>of</strong> chaperone-mediated<br />

autophagy in aging liver improves cellular maintenance<br />

<strong>and</strong> hepatic function. Nat Med. 2008 Sep;14(9):959-65.<br />

Zhang H, Bosch-Marce M, Shimoda LA, Tan YS, Baek JH,<br />

Wesley JB, Gonzalez FJ, Semenza GL.. Mitochondrial<br />

autophagy is an HIF-1-dependent adaptive metabolic<br />

response to hypoxia. J Biol Chem. 2008 Apr<br />

18;283(16):10892-903. Epub 2008 Feb 15.<br />

Bellot G, Garcia-Medina R, Gounon P, Chiche J, Roux D,<br />

Pouyssegur J, Mazure NM.. Hypoxia-induced autophagy is<br />

mediated through hypoxia-inducible factor induction <strong>of</strong><br />

BNIP3 <strong>and</strong> BNIP3L via their BH3 domains. Mol Cell Biol.<br />

2009 May;29(10):2570-81. Epub 2009 Mar 9.<br />

Bensaad K, Cheung EC, Vousden KH.. Modulation <strong>of</strong><br />

intracellular ROS levels by TIGAR controls autophagy.<br />

EMBO J. 2009 Oct 7;28(19):3015-26. Epub 2009 Aug 27.<br />

Betin VM, Lane JD.. Caspase cleavage <strong>of</strong> Atg4D<br />

stimulates GABARAP-L1 processing <strong>and</strong> triggers<br />

mitochondrial targeting <strong>and</strong> apoptosis. J Cell Sci. 2009 Jul<br />

15;122(Pt 14):2554-66. Epub 2009 Jun 23.<br />

Cho DH, Jo YK, Hwang JJ, Lee YM, Roh SA, Kim JC..<br />

Caspase-mediated cleavage <strong>of</strong> ATG6/Beclin-1 links<br />

apoptosis to autophagy in HeLa cells. Cancer Lett. 2009<br />

Feb 8;274(1):95-100. Epub 2008 Oct 7.<br />

Copetti T, Bertoli C, Dalla E, Demarchi F, Schneider C..<br />

p65/RelA modulates BECN1 transcription <strong>and</strong> autophagy.<br />

Mol Cell Biol. 2009 May;29(10):2594-608. Epub 2009 Mar<br />

16.<br />

Eisenberg T, Knauer H, Schauer A, Buttner S,<br />

Ruckenstuhl C, Carmona-Gutierrez D, Ring J, Schroeder<br />

S, Magnes C, Antonacci L, Fussi H, Deszcz L, Hartl R,<br />

Schraml E, Criollo A, Megalou E, Weiskopf D, Laun P,<br />

Heeren G, Breitenbach M, Grubeck-Loebenstein B, Herker<br />

E, Fahrenkrog B, Frohlich KU, Sinner F, Tavernarakis N,<br />

Minois N, Kroemer G, Madeo F.. Induction <strong>of</strong> autophagy by<br />

spermidine promotes longevity. Nat Cell Biol. 2009<br />

Nov;11(11):1305-14. Epub 2009 Oct 4.<br />

Fader CM, Sanchez DG, Mestre MB, Colombo MI.. TI-<br />

VAMP/VAMP7 <strong>and</strong> VAMP3/cellubrevin: two v-SNARE<br />

proteins involved in specific steps <strong>of</strong> the<br />

autophagy/multivesicular body pathways. Biochim Biophys<br />

Acta. 2009 Dec;1793(12):1901-16. Epub 2009 Sep 23.<br />

Hayashi-Nishino M, Fujita N, Noda T, Yamaguchi A,<br />

Yoshimori T, Yamamoto A.. A subdomain <strong>of</strong> the<br />

endoplasmic reticulum forms a cradle for autophagosome<br />

formation. Nat Cell Biol. 2009 Dec;11(12):1433-7. Epub<br />

2009 Nov 8.<br />

He C, Klionsky DJ.. Regulation mechanisms <strong>and</strong> signaling<br />

pathways <strong>of</strong> autophagy. Annu Rev <strong>Gene</strong>t. 2009;43:67-93.<br />

(REVIEW)<br />

Hosokawa N, Hara T, Kaizuka T, Kishi C, Takamura A,<br />

Miura Y, Iemura S, Natsume T, Takehana K, Yamada N,<br />

Guan JL, Oshiro N, Mizushima N.. Nutrient-dependent<br />

mTORC1 association with the ULK1-Atg13-FIP200<br />

complex required for autophagy. Mol Biol Cell. 2009<br />

Apr;20(7):1981-91. Epub 2009 Feb 11.<br />

Korolchuk VI, Mansilla A, Menzies FM, Rubinsztein DC..<br />

Autophagy inhibition compromises degradation <strong>of</strong><br />

ubiquitin-proteasome pathway substrates. Mol Cell. 2009<br />

Feb 27;33(4):517-27.<br />

Kraft C, Reggiori F, Peter M.. Selective types <strong>of</strong> autophagy<br />

in yeast. Biochim Biophys Acta. 2009 Sep;1793(9):1404-<br />

12. Epub 2009 Mar 2. (REVIEW)<br />

Lee IH, Finkel T.. Regulation <strong>of</strong> autophagy by the p300<br />

acetyltransferase. J Biol Chem. 2009 Mar 6;284(10):6322-<br />

8. Epub 2009 Jan 5.<br />

Lee JS, Li Q, Lee JY, Lee SH, Jeong JH, Lee HR, Chang<br />

H, Zhou FC, Gao SJ, Liang C, Jung JU.. FLIP-mediated<br />

autophagy regulation in cell death control. Nat Cell Biol.<br />

2009 Nov;11(11):1355-62. Epub 2009 Oct 18.<br />

Li DD, Wang LL, Deng R, Tang J, Shen Y, Guo JF, Wang<br />

Y, Xia LP, Feng GK, Liu QQ, Huang WL, Zeng YX, Zhu<br />

XF.. The pivotal role <strong>of</strong> c-Jun NH2-terminal kinasemediated<br />

Beclin 1 expression during anticancer agentsinduced<br />

autophagy in cancer cells. Oncogene. 2009 Feb<br />

12;28(6):886-98. Epub 2008 Dec 8.<br />

Liu HY, Han J, Cao SY, Hong T, Zhuo D, Shi J, Liu Z, Cao<br />

W.. Hepatic autophagy is suppressed in the presence <strong>of</strong><br />

insulin resistance <strong>and</strong> hyperinsulinemia: inhibition <strong>of</strong><br />

FoxO1-dependent expression <strong>of</strong> key autophagy genes by<br />

insulin. J Biol Chem. 2009 Nov 6;284(45):31484-92. Epub<br />

2009 Sep 16.<br />

Mathew R, Karp CM, Beaudoin B, Vuong N, Chen G, Chen<br />

HY, Bray K, Reddy A, Bhanot G, Gelinas C, Dipaola RS,<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 178


Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells Mehrpour M, et al.<br />

Karantza-Wadsworth V, White E.. Autophagy suppresses<br />

tumorigenesis through elimination <strong>of</strong> p62. Cell. 2009 Jun<br />

12;137(6):1062-75.<br />

Matsunaga K, Saitoh T, Tabata K, Omori H, Satoh T,<br />

Kurotori N, Maejima I, Shirahama-Noda K, Ichimura T,<br />

Isobe T, Akira S, Noda T, Yoshimori T.. Two Beclin 1binding<br />

proteins, Atg14L <strong>and</strong> Rubicon, reciprocally<br />

regulate autophagy at different stages. Nat Cell Biol. 2009<br />

Apr;11(4):385-96. Epub 2009 Mar 8<br />

Meijer AJ, Codogno P.. Autophagy: regulation <strong>and</strong> role in<br />

disease. Crit Rev Clin Lab Sci. 2009;46(4):210-40.<br />

(REVIEW)<br />

Morselli E, Galluzzi L, Kepp O, Vicencio JM, Criollo A,<br />

Maiuri MC, Kroemer G.. Anti- <strong>and</strong> pro-tumor functions <strong>of</strong><br />

autophagy. Biochim Biophys Acta. 2009<br />

Sep;1793(9):1524-32. Epub 2009 Jan 21. (REVIEW)<br />

Nakatogawa H, Suzuki K, Kamada Y, Ohsumi Y..<br />

Dynamics <strong>and</strong> diversity in autophagy mechanisms: lessons<br />

from yeast. Nat Rev Mol Cell Biol. 2009 Jul;10(7):458-67.<br />

Epub 2009 Jun 3. (REVIEW)<br />

Nicklin P, Bergman P, Zhang B, Triantafellow E, Wang H,<br />

Nyfeler B, Yang H, Hild M, Kung C, Wilson C, Myer VE,<br />

MacKeigan JP, Porter JA, Wang YK, Cantley LC, Finan<br />

PM, Murphy LO.. Bidirectional transport <strong>of</strong> amino acids<br />

regulates mTOR <strong>and</strong> autophagy. Cell. 2009 Feb<br />

6;136(3):521-34.<br />

Nivon M, Richet E, Codogno P, Arrigo AP, Kretz-Remy C..<br />

Autophagy activation by NFkappaB is essential for cell<br />

survival after heat shock. Autophagy. 2009 Aug;5(6):766-<br />

83. Epub 2009 Aug 20.<br />

Park KJ, Lee SH, Lee CH, Jang JY, Chung J, Kwon MH,<br />

Kim YS.. Upregulation <strong>of</strong> Beclin-1 expression <strong>and</strong><br />

phosphorylation <strong>of</strong> Bcl-2 <strong>and</strong> p53 are involved in the JNKmediated<br />

autophagic cell death. Biochem Biophys Res<br />

Commun. 2009 May 15;382(4):726-9. Epub 2009 Mar 24.<br />

Pattingre S, Bauvy C, Carpentier S, Levade T, Levine B,<br />

Codogno P.. Role <strong>of</strong> JNK1-dependent Bcl-2<br />

phosphorylation in ceramide-induced macroautophagy. J<br />

Biol Chem. 2009 Jan 30;284(5):2719-28. Epub 2008 Nov<br />

23.<br />

Raiborg C, Stenmark H.. The ESCRT machinery in<br />

endosomal sorting <strong>of</strong> ubiquitylated membrane proteins.<br />

Nature. 2009 Mar 26;458(7237):445-52. (REVIEW)<br />

Ramach<strong>and</strong>ran N, Munteanu I, Wang P, Aubourg P,<br />

Rilstone JJ, Israelian N, Naranian T, Paroutis P, Guo R,<br />

Ren ZP, Nishino I, Chabrol B, Pellissier JF, Minetti C, Udd<br />

B, Fardeau M, Tailor CS, Mahuran DJ, Kissel JT, Kalimo<br />

H, Levy N, Manolson MF, Ackerley CA, Minassian BA..<br />

VMA21 deficiency causes an autophagic myopathy by<br />

compromising V-ATPase activity <strong>and</strong> lysosomal<br />

acidification. Cell. 2009 Apr 17;137(2):235-46.<br />

Razi M, Chan EY, Tooze SA.. Early endosomes <strong>and</strong><br />

endosomal coatomer are required for autophagy. J Cell<br />

Biol. 2009 Apr 20;185(2):305-21. Epub 2009 Apr 13.<br />

Rubinsztein DC, Cuervo AM, Ravikumar B, Sarkar S,<br />

Korolchuk V, Kaushik S, Klionsky DJ.. In search <strong>of</strong> an<br />

"autophagomometer". Autophagy. 2009 Jul;5(5):585-9.<br />

Epub 2009 Jul 23.<br />

Rusten TE, Stenmark H.. How do ESCRT proteins control<br />

autophagy? J Cell Sci. 2009 Jul 1;122(Pt 13):2179-83.<br />

Sardiello M, Palmieri M, di Ronza A, Medina DL, Valenza<br />

M, Gennarino VA, Di Malta C, Donaudy F, Embrione V,<br />

Polishchuk RS, Banfi S, Parenti G, Cattaneo E, Ballabio<br />

A.. A gene network regulating lysosomal biogenesis <strong>and</strong><br />

function. Science. 2009 Jul 24;325(5939):473-7. Epub<br />

2009 Jun 25.<br />

Tang H, Da L, Mao Y, Li Y, Li D, Xu Z, Li F, Wang Y,<br />

Tiollais P, Li T, Zhao M.. Hepatitis B virus X protein<br />

sensitizes cells to starvation-induced autophagy via upregulation<br />

<strong>of</strong> beclin 1 expression. Hepatology. 2009<br />

Jan;49(1):60-71.<br />

Vicencio JM, Ortiz C, Criollo A, Jones AW, Kepp O,<br />

Galluzzi L, Joza N, Vitale I, Morselli E, Tailler M, Castedo<br />

M, Maiuri MC, Molg¢ J, Szabadkai G, Lav<strong>and</strong>ero S,<br />

Kroemer G.. The inositol 1,4,5-trisphosphate receptor<br />

regulates autophagy through its interaction with Beclin 1.<br />

Cell Death Differ. 2009 Jul;16(7):1006-17. Epub 2009 Mar<br />

27.<br />

Virgin HW, Levine B.. Autophagy genes in immunity. Nat<br />

Immunol. 2009 May;10(5):461-70. (REVIEW)<br />

Vousden KH, Prives C.. Blinded by the Light: The Growing<br />

Complexity <strong>of</strong> p53. Cell. 2009 May 1;137(3):413-31.<br />

(REVIEW)<br />

Yla-Anttila P, Vihinen H, Jokitalo E, Eskelinen EL.. 3D<br />

tomography reveals connections between the phagophore<br />

<strong>and</strong> endoplasmic reticulum. Autophagy. 2009<br />

Nov;5(8):1180-5. Epub 2009 Nov 8.<br />

Yuk JM, Shin DM, Lee HM, Yang CS, Jin HS, Kim KK, Lee<br />

ZW, Lee SH, Kim JM, Jo EK.. Vitamin D3 induces<br />

autophagy in human monocytes/macrophages via<br />

cathelicidin. Cell Host Microbe. 2009 Sep 17;6(3):231-43.<br />

Zalckvar E, Berissi H, Mizrachy L, Idelchuk Y, Koren I,<br />

Eisenstein M, Sabanay H, Pinkas-Kramarski R, Kimchi A..<br />

DAP-kinase-mediated phosphorylation on the BH3 domain<br />

<strong>of</strong> beclin 1 promotes dissociation <strong>of</strong> beclin 1 from Bcl-XL<br />

<strong>and</strong> induction <strong>of</strong> autophagy. EMBO Rep. 2009<br />

Mar;10(3):285-92. Epub 2009 Jan 30.<br />

Zhong Y, Wang QJ, Li X, Yan Y, Backer JM, Chait BT,<br />

Heintz N, Yue Z.. Distinct regulation <strong>of</strong> autophagic activity<br />

by Atg14L <strong>and</strong> Rubicon associated with Beclin 1phosphatidylinositol-3-kinase<br />

complex. Nat Cell Biol. 2009<br />

Apr;11(4):468-76. Epub 2009 Mar 8.<br />

Zhu H, Wu H, Liu X, Li B, Chen Y, Ren X, Liu CG, Yang<br />

JM.. Regulation <strong>of</strong> autophagy by a beclin 1-targeted<br />

microRNA, miR-30a, in cancer cells. Autophagy. 2009<br />

Aug;5(6):816-23. Epub 2009 Aug 20.<br />

Hailey DW, Rambold AS, Satpute-Krishnan P, Mitra K,<br />

Sougrat R, Kim PK, Lippincott-Schwartz J.. Mitochondria<br />

supply membranes for autophagosome biogenesis during<br />

starvation. Cell. 2010 May 14;141(4):656-67.<br />

Hidvegi T, Ewing M, Hale P, Dippold C, Beckett C, Kemp<br />

C, Maurice N, Mukherjee A, Goldbach C, Watkins S,<br />

Michalopoulos G, Perlmutter DH.. An autophagyenhancing<br />

drug promotes degradation <strong>of</strong> mutant alpha1antitrypsin<br />

Z <strong>and</strong> reduces hepatic fibrosis. Science. 2010<br />

Jul 9;329(5988):229-32. Epub 2010 Jun 3.<br />

Kovsan J, Bashan N, Greenberg AS, Rudich A.. Potential<br />

role <strong>of</strong> autophagy in modulation <strong>of</strong> lipid metabolism. Am J<br />

Physiol Endocrinol Metab. 2010 Jan;298(1):E1-7. Epub<br />

2009 Nov 3. (REVIEW)<br />

Lamark T, Johansen T.. Autophagy: links with the<br />

proteasome. Curr Opin Cell Biol. 2010 Apr;22(2):192-8.<br />

Epub 2009 Dec 2.<br />

Luo S, Rubinsztein DC.. Apoptosis blocks Beclin 1dependent<br />

autophagosome synthesis: an effect rescued<br />

by Bcl-xL. Cell Death Differ. 2010 Feb;17(2):268-77. Epub<br />

2009 Aug 28.<br />

Mari M, Griffith J, Rieter E, Krishnappa L, Klionsky DJ,<br />

Reggiori F.. An Atg9-containing compartment that<br />

functions in the early steps <strong>of</strong> autophagosome biogenesis.<br />

J Cell Biol. 2010 Sep 20;190(6):1005-22.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 179


Mechanisms <strong>and</strong> regulation <strong>of</strong> autophagy in mammalian cells Mehrpour M, et al.<br />

Mauvezin C, Orpinell M, Francis VA, Mansilla F, Duran J,<br />

Ribas V, Palac¡n M, Boya P, Teleman AA, Zorzano A.. The<br />

nuclear c<strong>of</strong>actor DOR regulates autophagy in mammalian<br />

<strong>and</strong> Drosophila cells. EMBO Rep. 2010 Jan;11(1):37-44.<br />

Epub 2009 Dec 4.<br />

Mizushima N, Yoshimori T, Levine B.. Methods in<br />

mammalian autophagy research. Cell. 2010 Feb<br />

5;140(3):313-26.<br />

Morselli E, Maiuri MC, Markaki M, Megalou E, Pasparaki<br />

A, Palikaras K, Criollo A, Galluzzi L, Malik SA, Vitale I,<br />

Michaud M, Madeo F, Tavernarakis N, Kroemer G.. The<br />

life span-prolonging effect <strong>of</strong> sirtuin-1 is mediated by<br />

autophagy. Autophagy. 2010 Jan;6(1):186-8. Epub 2010<br />

Jan 2. (REVIEW)<br />

Ohashi Y, Munro S.. Membrane delivery to the yeast<br />

autophagosome from the Golgi-endosomal system. Mol<br />

Biol Cell. 2010 Nov 15;21(22):3998-4008. Epub 2010 Sep<br />

22.<br />

Ravikumar B, Moreau K, Jahreiss L, Puri C, Rubinsztein<br />

DC.. Plasma membrane contributes to the formation <strong>of</strong><br />

pre-autophagosomal structures. Nat Cell Biol. 2010a<br />

Aug;12(8):747-57. Epub 2010 Jul 18.<br />

Ravikumar B, Sarkar S, Davies JE, Futter M, Garcia-<br />

Arencibia M, Green-Thompson ZW, Jimenez-Sanchez M,<br />

Korolchuk VI, Lichtenberg M, Luo S, Massey DC, Menzies<br />

FM, Moreau K, Narayanan U, Renna M, Siddiqi FH,<br />

Underwood BR, Winslow AR, Rubinsztein DC.. Regulation<br />

<strong>of</strong> mammalian autophagy in physiology <strong>and</strong><br />

pathophysiology. Physiol Rev. 2010b Oct;90(4):1383-435.<br />

(REVIEW)<br />

Rouschop KM, van den Beucken T, Dubois L, Niessen H,<br />

Bussink J, Savelkouls K, Keulers T, Mujcic H, L<strong>and</strong>uyt W,<br />

Voncken JW, Lambin P, van der Kogel AJ, Koritzinsky M,<br />

Wouters BG.. The unfolded protein response protects<br />

human tumor cells during hypoxia through regulation <strong>of</strong> the<br />

autophagy genes MAP1LC3B <strong>and</strong> ATG5. J Clin Invest.<br />

2010 Jan 4;120(1):127-41. doi: 10.1172/JCI40027. Epub<br />

2009 Dec 14.<br />

Weidberg H, Shvets E, Shpilka T, Shimron F, Shinder V,<br />

Elazar Z.. LC3 <strong>and</strong> GATE-16/GABARAP subfamilies are<br />

both essential yet act differently in autophagosome<br />

biogenesis. EMBO J. 2010 Jun 2;29(11):1792-802. Epub<br />

2010 Apr 23.<br />

Xia HG, Zhang L, Chen G, Zhang T, Liu J, Jin M, Ma X,<br />

Ma D, Yuan J.. Control <strong>of</strong> basal autophagy by calpain1<br />

mediated cleavage <strong>of</strong> ATG5. Autophagy. 2010<br />

Jan;6(1):61-6. Epub 2010 Jan 13.<br />

Yang L, Li P, Fu S, Calay ES, Hotamisligil GS.. Defective<br />

hepatic autophagy in obesity promotes ER stress <strong>and</strong><br />

causes insulin resistance. Cell Metab. 2010 Jun<br />

9;11(6):467-78.<br />

Yang Z, Klionsky DJ.. Eaten alive: a history <strong>of</strong><br />

macroautophagy. Nat Cell Biol. 2010 Sep;12(9):814-22.<br />

(REVIEW)<br />

Egan DF, Shackelford DB, Mihaylova MM, Gelino S,<br />

Kohnz RA, Mair W, Vasquez DS, Joshi A, Gwinn DM,<br />

Taylor R, Asara JM, Fitzpatrick J, Dillin A, Viollet B, Kundu<br />

M, Hansen M, Shaw RJ.. Phosphorylation <strong>of</strong> ULK1<br />

(hATG1) by AMP-activated protein kinase connects energy<br />

sensing to mitophagy. Science. 2011 Jan<br />

28;331(6016):456-61. Epub 2010 Dec 23.<br />

Garcia-Marcos M, Ear J, Farquhar MG, Ghosh P.. A GDI<br />

(AGS3) <strong>and</strong> a GEF (GIV) regulate autophagy by balancing<br />

G protein activity <strong>and</strong> growth factor signals. Mol Biol Cell.<br />

2011 Mar 1;22(5):673-86. Epub 2011 Jan 5.<br />

Guo JY, Chen HY, Mathew R, Fan J, Strohecker AM,<br />

Karsli-Uzunbas G, Kamphorst JJ, Chen G, Lemons JM,<br />

Karantza V, Coller HA, Dipaola RS, Gelinas C, Rabinowitz<br />

JD, White E. Activated Ras requires autophagy to maintain<br />

oxidative metabolism <strong>and</strong> tumorigenesis. <strong>Gene</strong>s Dev. 2011<br />

Mar 1;25(5):460-70. Epub 2011 Feb 11.<br />

Itoh T, Kanno E, Uemura T, Waguri S, Fukuda M.. OATL1,<br />

a novel autophagosome-resident Rab33B-GAP, regulates<br />

autophagosomal maturation. J Cell Biol. 2011 Mar<br />

7;192(5):839-53.<br />

Kim J, Kundu M, Viollet B, Guan KL.. AMPK <strong>and</strong> mTOR<br />

regulate autophagy through direct phosphorylation <strong>of</strong> Ulk1.<br />

Nat Cell Biol. 2011 Feb;13(2):132-41. Epub 2011 Jan 23.<br />

Lock R, Roy S, Kenific CM, Su JS, Salas E, Ronen SM,<br />

Debnath J.. Autophagy facilitates glycolysis during Rasmediated<br />

oncogenic transformation. Mol Biol Cell. 2011<br />

Jan 15;22(2):165-78. Epub 2010 Nov 30.<br />

Moreau K, Ravikumar B, Renna M, Puri C, Rubinsztein<br />

DC.. Autophagosome precursor maturation requires<br />

homotypic fusion. Cell. 2011 Jul 22;146(2):303-17.<br />

Nair U, Jotwani A, Geng J, Gammoh N, Richerson D, Yen<br />

WL, Griffith J, Nag S, Wang K, Moss T, Baba M, McNew<br />

JA, Jiang X, Reggiori F, Melia TJ, Klionsky DJ.. SNARE<br />

proteins are required for macroautophagy. Cell. 2011 Jul<br />

22;146(2):290-302.<br />

Settembre C, Di Malta C, Polito VA, Garcia Arencibia M,<br />

Vetrini F, Erdin S, Erdin SU, Huynh T, Medina D, Colella P,<br />

Sardiello M, Rubinsztein DC, Ballabio A.. TFEB links<br />

autophagy to lysosomal biogenesis. Science. 2011 Jun<br />

17;332(6036):1429-33. Epub 2011 May 26.<br />

Weidberg H, Shpilka T, Shvets E, Abada A, Shimron F,<br />

Elazar Z.. LC3 <strong>and</strong> GATE-16 N termini mediate membrane<br />

fusion processes required for autophagosome biogenesis.<br />

Dev Cell. 2011a Apr 19;20(4):444-54.<br />

Weidberg H, Shvets E, Elazar Z.. Biogenesis <strong>and</strong> cargo<br />

selectivity <strong>of</strong> autophagosomes. Annu Rev Biochem. 2011b<br />

Jun 7;80:125-56.<br />

Yang S, Wang X, Contino G, Liesa M, Sahin E, Ying H,<br />

Bause A, Li Y, Stommel JM, Dell'antonio G, Mautner J,<br />

Tonon G, Haigis M, Shirihai OS, Doglioni C, Bardeesy N,<br />

Kimmelman AC.. Pancreatic cancers require autophagy for<br />

tumor growth. <strong>Gene</strong>s Dev. 2011 Apr 1;25(7):717-29. Epub<br />

2011 Mar 15.<br />

This article should be referenced as such:<br />

Mehrpour M, Botti J, Codogno P. Mechanisms <strong>and</strong><br />

regulation <strong>of</strong> autophagy in mammalian cells. <strong>Atlas</strong> <strong>Gene</strong>t<br />

Cytogenet Oncol Haematol. 2012; 16(2):163-180.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 180


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

Case Report <strong>Section</strong><br />

Paper co-edited with the European LeukemiaNet<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

A new case <strong>of</strong> t(5;14)(q31;q32) in a pediatric<br />

acute lymphoblastic leukemia presenting with<br />

hypereosinophilia<br />

Marta Gallego, Mariela Coccé, María Felice, Jorge Rossi, Silvia E<strong>and</strong>i, Gabriela Sciuccati,<br />

Cristina Alonso<br />

Laboratorio de Citogenetica - Servicio de <strong>Gene</strong>tica - Servicio de Inmunologia y Reumatologia -<br />

Servicio de Hemato-Oncologia, Hospital de Pediatria "Pr<strong>of</strong>. Dr. J. P. Garrahan", Buenos Aires,<br />

Argentina (MG, MC, MF, JR, SE, GS, CA)<br />

Published in <strong>Atlas</strong> Database: September 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/Reports/t514q31q32GallegoID100055.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI t514q31q32GallegoID100055.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Clinics<br />

Age <strong>and</strong> sex<br />

11 years old male patient.<br />

Previous history<br />

Preleukemia. The patient presented with a chronic<br />

eosinophilic leukemia 3 months before developing<br />

ALL. No previous malignancy. No inborn condition<br />

<strong>of</strong> note<br />

Organomegaly<br />

Hepatomegaly (4 cm from below costal rib),<br />

splenomegaly (3 cm from below costal rib), no<br />

enlarged lymph nodes , no central nervous system<br />

involvement.<br />

Blood<br />

WBC : 48 with 62% <strong>of</strong> eosinophils X 10 9 /l<br />

HB : 7.0g/dl<br />

Platelets : 79 X 10 9 /l<br />

Blasts : 15%<br />

Bone marrow : Normal cellularity was replaced by<br />

60% <strong>of</strong> lymphoblasts FAB L1 morphology%.<br />

Cyto-Pathology<br />

Classification<br />

Immunophenotype<br />

Pre-B ALL (EGIL classification B III).<br />

The blasts expressed CD45, CD19, CD10, CD34,<br />

HLA-DR, cCD79a, cCD22, Tdt <strong>and</strong> cytoplasmic<br />

micro chain, partial CD20 <strong>and</strong> CD33 <strong>and</strong> were<br />

negative for CD2, CD7, CD13, CD15, CD117 <strong>and</strong><br />

CD3.<br />

Diagnosis<br />

Acute lymphoblastic leukemia following a chronic<br />

eosinophilic leukemia.<br />

Survival<br />

Date <strong>of</strong> diagnosis: 03-2008<br />

Treatment: Chemotherapy for ALL (12-ALLIC 02<br />

protocol)<br />

Complete remission was obtained.<br />

Treatment related death : no<br />

Relapse : yes<br />

Phenotype at relapse<br />

During continuation phase hypereosinophilia was<br />

observed in peripheral blood, but low percentage <strong>of</strong><br />

lymphoblasts was detected during 2-3 weeks before<br />

relapse. After this finding, the patient presented<br />

CNS infiltration by eosinophils (70% <strong>of</strong> WBC<br />

detected in CSF). He presented a bone marrow<br />

infiltration by dysplastic eosinophils <strong>and</strong> less than<br />

5% <strong>of</strong> lymphoblasts after 18 months from achieving<br />

CR <strong>and</strong> a hematological relapse was diagnosed.<br />

Status: Death 06-2010<br />

Survival: 21 months<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 181


A new case <strong>of</strong> t(5;14)(q31;q32) in a pediatric acute<br />

lymphoblastic leukemia presenting with hypereosinophilia<br />

Karyotype<br />

Sample: Bone marrow<br />

Culture time: 24h<br />

B<strong>and</strong>ing: G b<strong>and</strong>ing<br />

Results<br />

Karyotype at time <strong>of</strong> diagnosis <strong>of</strong> ALL:<br />

46,XY,t(5;14)(q31;q32)[4]/46,XY[12]<br />

Karyotype at Relapse<br />

46,XY,t(3;8)(p21;q24),t(5;14)(q31;q32)[2]/46,XY[<br />

18]<br />

Other Molecular Studies<br />

Technics:<br />

RT-PCR non evaluable, due to control gene non<br />

amplifiable.<br />

Partial GTG b<strong>and</strong>ed karyotype showing t(5;14)(q31;q32).<br />

Comments<br />

To our knowledge nine cases (8M/1F) <strong>of</strong> ALL with<br />

eosinophilia <strong>and</strong> t(5;14)(q31;q32) have been<br />

reported in the literature. Five <strong>of</strong> them were<br />

described in childhood ALL. The prognosis <strong>of</strong><br />

t(5;14)(q31;q32) seems to be very poor. Our patient<br />

relapsed <strong>and</strong> died 21 months after diagnoses.<br />

References<br />

Tono-oka T, Sato Y, Matsumoto T, Ueno N, Ohkawa M,<br />

Shikano T, Takeda T. Hypereosinophilic syndrome in<br />

acute lymphoblastic leukemia with a chromosome<br />

translocation [t(5q;14q)]. Med Pediatr Oncol.<br />

1984;12(1):33-7<br />

Gallego M, et al.<br />

Hogan TF, Koss W, Murgo AJ, Amato RS, Fontana JA,<br />

VanScoy FL. Acute lymphoblastic leukemia with<br />

chromosomal 5;14 translocation <strong>and</strong> hypereosinophilia:<br />

case report <strong>and</strong> literature review. J Clin Oncol. 1987<br />

Mar;5(3):382-90<br />

McConnell TS, Foucar K, Hardy WR, Saiki J. Three-way<br />

reciprocal chromosomal translocation in an acute<br />

lymphoblastic leukemia patient with hypereosinophilia<br />

syndrome. J Clin Oncol. 1987 Dec;5(12):2042-4<br />

Baumgarten E, Wegner RD, Fengler R, Ludwig WD,<br />

Schulte-Overberg U, Domeyer C, Schüürmann J, Henze<br />

G. Calla-positive acute leukaemia with t(5q;14q)<br />

translocation <strong>and</strong> hypereosinophilia--a unique entity? Acta<br />

Haematol. 1989;82(2):85-90<br />

Grimaldi JC, Meeker TC. The t(5;14) chromosomal<br />

translocation in a case <strong>of</strong> acute lymphocytic leukemia joins<br />

the interleukin-3 gene to the immunoglobulin heavy chain<br />

gene. Blood. 1989 Jun;73(8):2081-5<br />

Fishel RS, Farnen JP, Hanson CA, Silver SM, Emerson<br />

SG. Acute lymphoblastic leukemia with eosinophilia.<br />

Medicine (Baltimore). 1990 Jul;69(4):232-43<br />

Meeker TC, Hardy D, Willman C, Hogan T, Abrams J.<br />

Activation <strong>of</strong> the interleukin-3 gene by chromosome<br />

translocation in acute lymphocytic leukemia with<br />

eosinophilia. Blood. 1990 Jul 15;76(2):285-9<br />

Chen Z, Morgan R, S<strong>and</strong>berg AA. Non-r<strong>and</strong>om<br />

involvement <strong>of</strong> chromosome 5 in ALL. Cancer <strong>Gene</strong>t<br />

Cytogenet. 1992 Jul 1;61(1):106-7<br />

Heerema NA, Palmer CG, Weetman R, Bertolone S.<br />

Cytogenetic analysis in relapsed childhood acute<br />

lymphoblastic leukemia. Leukemia. 1992 Mar;6(3):185-92<br />

Knuutila S, Alitalo R, Ruutu T. Power <strong>of</strong> the MAC<br />

(morphology-antibody-chromosomes) method in<br />

distinguishing reactive <strong>and</strong> clonal cells: report <strong>of</strong> a patient<br />

with acute lymphatic leukemia, eosinophilia, <strong>and</strong> t(5;14).<br />

<strong>Gene</strong>s Chromosomes Cancer. 1993 Dec;8(4):219-23<br />

Gmidène A, Sennana H, Elghezal H, Ziraoui S, Youssef<br />

YB, Elloumi M, Issaoui L, Harrabi I, Raynaud S, Saad A.<br />

Cytogenetic analysis <strong>of</strong> 298 newly diagnosed cases <strong>of</strong><br />

acute lymphoblastic leukaemia in Tunisia. Hematol Oncol.<br />

2008 Jun;26(2):91-7<br />

This article should be referenced as such:<br />

Gallego M, Coccé M, Felice M, Rossi J, E<strong>and</strong>i S, Sciuccati<br />

G, Alonso C. A new case <strong>of</strong> t(5;14)(q31;q32) in a pediatric<br />

acute lymphoblastic leukemia presenting with<br />

hypereosinophilia. <strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol.<br />

2012; 16(2):181-182.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 182


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

Case Report <strong>Section</strong><br />

Paper co-edited with the European LeukemiaNet<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

Chromosomal translocation t(X;11)(q22;q23)<br />

involving the MLL gene<br />

Adriana Zamecnikova, Soad Al Bahar, Hassan A Al Jafar, Rames P<strong>and</strong>ita<br />

Kuwait Cancer Control Center, Dep <strong>of</strong> Hematology, Laboratory <strong>of</strong> Cancer <strong>Gene</strong>tics, Kuwait (AZ, SA,<br />

RP), Dep <strong>of</strong> Hematology, Amiri Hospital, Kuwait (HAAJ)<br />

Published in <strong>Atlas</strong> Database: September 2011<br />

Online updated version : http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/Reports/tX11q22q23ZamecID100056.html<br />

Printable original version : http://documents.irevues.inist.fr/bitstream/DOI tX11q22q23ZamecID100056.txt<br />

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 2.0 France Licence.<br />

© 2012 <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology<br />

Clinics<br />

Age <strong>and</strong> sex<br />

15 months old male patient.<br />

Previous history<br />

No preleukemia. No previous malignancy. No<br />

inborn condition <strong>of</strong> note.<br />

Organomegaly<br />

No hepatomegaly, no splenomegaly, no enlarged<br />

lymph nodes, no central nervous system<br />

involvement.<br />

Blood<br />

WBC : 50.2 (neutrophils 2%, lymphocytes 68%,<br />

probably blasts, monocytes 16%, atypical<br />

lymphocytes 14% undifferenciated cells) X 10 9 /l<br />

HB : 10.2g/dl<br />

Platelets : 191 X 10 9 /l<br />

Blasts : 14%<br />

Bone marrow : The bone marrow was<br />

hypercellular with more than 50% blasts that were<br />

positive for CD45, CD33, CD15, CD13, CD14 <strong>and</strong><br />

HLA-DR.<br />

Cyto-Pathology<br />

Classification<br />

Cytology: AML M4<br />

Immunophenotype: M4. CD13+,CD14+, CD15+,<br />

CD33+, CD45+, CD64+ <strong>and</strong> MPO.<br />

Diagnosis: Acute myelomonocytic leukemia.<br />

Survival<br />

Date <strong>of</strong> diagnosis: 02-2010<br />

Treatment: Chemotherapy (ADE)<br />

Complete remission was obtained.<br />

Treatment related death : no<br />

Relapse : no<br />

Status: Alive. Last follow up: 02-2011<br />

Survival: 12 months<br />

Karyotype<br />

Sample: BM<br />

Culture time: 24h<br />

B<strong>and</strong>ing: G-b<strong>and</strong><br />

Karyotype at Relapse<br />

46,XY [5] / 46,Y,t(X;11)(q22;q23) [25]<br />

Other molecular cytogenetics technics<br />

Fluorescence in situ hybridization.<br />

Other molecular cytogenetics results<br />

MLL rearrangement was identified using the LSI<br />

MLL (11q23) Dual Color Break Apart<br />

Rearrangement probe (Abbott Molecular) revealing<br />

80% <strong>of</strong> cells with MLL rearrangement. The<br />

rearrangement was confirmed in metaphases<br />

demonstrating that the distal part <strong>of</strong> the MLL gene<br />

was juxtaposed to the der(X) chromosome.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 183


Chromosomal translocation t(X;11)(q22;q23) involving the<br />

MLL gene<br />

Zamecnikova A, et al.<br />

Top: partial karyotype <strong>of</strong> the patient. Bottom: fluorescence in situ studies were performed successively on G-b<strong>and</strong>ed slides<br />

prepared for chromosome analysis using a locus-specific, break-apart probe for MLL (green <strong>and</strong> red signals) <strong>and</strong> with<br />

centromeric X/Y probe (Vysis) (red <strong>and</strong> green signal). Hybridization on metaphase cells with the MLL probe detected one MLL<br />

signal on the normal chromosome 11 (red-green fusion signal) <strong>and</strong> signals on the der(11) (green signal) <strong>and</strong> the der(X),<br />

confirming MLL disruption. The red signal from MLL has moved to the derivative chromosome X, indicating that the breakpoint is<br />

in the 5' <strong>of</strong> the MLL gene. Arrows indicate derivative chromosomes, arrow heads are pointing to derivative chromosomes X <strong>and</strong><br />

11.<br />

Comments<br />

A previously healthy, 15 months-old boy presented<br />

with fever, <strong>and</strong> chest infection was diagnosed with<br />

acute myeloid leukemia FAB M4 in February 2010.<br />

His biochemistry was significant for GGT (8 IU/L;<br />

normal 9-40) <strong>and</strong> LDH 350 (normal 90-225).<br />

Chromosomal studies performed at diagnosis<br />

revealed the karyotype 46,Y,t(X;11)(q22;q23) in 25<br />

out <strong>of</strong> 30 metaphases. Fluorescence in situ<br />

hybridization study showed rearrangement <strong>of</strong> the<br />

MLL gene in interphase <strong>and</strong> metaphase cells<br />

revealing that the break-apart 5'-MLL segment is<br />

translocated to the derivative X chromosome. The<br />

patient achieved a complete hematological<br />

remission with chemotherapy one month later.<br />

Chromosomal <strong>and</strong> FISH studies performed in April,<br />

June, August <strong>and</strong> December confirmed the<br />

complete cytogenetic response without<br />

rearrangment <strong>of</strong> the MLL gene. He remains disease<br />

free 1 year from diagnosis. Our case together with<br />

the few reported similar cases suggest that<br />

chromosomal b<strong>and</strong> Xq22 is a recurring 11q23<br />

chromosomal partner in a subgroup <strong>of</strong> infant<br />

leukemia patients with AML.<br />

As the gene in Xq22 is yet unknown, it is therefore<br />

uncertain whether this translocation involve a new<br />

MLL partner. Due to the similar clinical features<br />

with patients with t(X;11)(q13;q23) involving the<br />

FOXO4/MLL genes, (such as occurrence in infants<br />

<strong>and</strong> young children diagnosed with acute<br />

myelomonocytic leukemia), the possibility <strong>of</strong><br />

involvement <strong>of</strong> FOXO4 or FOXO related gene in<br />

our patient cannot be excluded. In addition as MLL<br />

rearrangements are frequently confirmed in cases<br />

with highly complex changes, complex <strong>and</strong>/or<br />

cryptic changes cannot be excluded.<br />

References<br />

Harrison CJ, Cuneo A, Clark R, Johansson B, Lafage-<br />

Pochital<strong>of</strong>f M, Mugneret F, Moorman AV, Secker-Walker<br />

LM. Ten novel 11q23 chromosomal partner sites.<br />

European 11q23 Workshop participants. Leukemia. 1998<br />

May;12(5):811-22<br />

Ribeiro RC, Oliveira MS, Fairclough D, Hurwitz C, Mirro J,<br />

Behm FG, Head D, Silva ML, Raimondi SC, Crist WM.<br />

Acute megakaryoblastic leukemia in children <strong>and</strong><br />

adolescents: a retrospective analysis <strong>of</strong> 24 cases. Leuk<br />

Lymphoma. 1993 Jul;10(4-5):299-306<br />

Soszynska K, Mucha B, Debski R, Skonieczka K,<br />

Duszenko E, Koltan A, Wysocki M, Haus O. The<br />

application <strong>of</strong> conventional cytogenetics, FISH, <strong>and</strong> RT-<br />

PCR to detect genetic changes in 70 children with ALL.<br />

Ann Hematol. 2008 Dec;87(12):991-1002<br />

This article should be referenced as such:<br />

Zamecnikova A, Al Bahar S, Al Jafar HA, P<strong>and</strong>ita R.<br />

Chromosomal translocation t(X;11)(q22;q23) involving the<br />

MLL gene. <strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012;<br />

16(2):183-184.<br />

<strong>Atlas</strong> <strong>Gene</strong>t Cytogenet Oncol Haematol. 2012; 16(2) 184


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

OPEN ACCESS JOURNAL AT INIST-CNRS<br />

Instructions to Authors<br />

Manuscripts submitted to the <strong>Atlas</strong> must be submitted solely to the <strong>Atlas</strong>.<br />

Iconography is most welcome: there is no space restriction.<br />

The <strong>Atlas</strong> publishes "cards", "deep insights", "case reports", <strong>and</strong> "educational items".<br />

Cards are structured review articles. Detailed instructions for these structured reviews can be found at:<br />

http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/Forms/<strong>Gene</strong>_Form.html for reviews on genes,<br />

http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/Forms/Leukaemia_Form.html for reviews on leukaemias,<br />

http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/Forms/SolidTumour_Form.html for reviews on solid tumours,<br />

http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/Forms/CancerProne_Form.html for reviews on cancer-prone diseases.<br />

According to the length <strong>of</strong> the paper, cards are divided, into "reviews" (texts exceeding 2000 words), "mini reviews"<br />

(between), <strong>and</strong> "short communications" (texts below 400 words). The latter category may not be accepted for indexing by<br />

bibliographic databases.<br />

Deep Insights are written as traditional papers, made <strong>of</strong> paragraphs with headings, at the author's convenience. No length<br />

restriction.<br />

Case Reports in haematological malignancies are dedicated to recurrent -but rare- chromosomes abnormalities in<br />

leukaemias/lymphomas. Cases <strong>of</strong> interest shall be: 1- recurrent (i.e. the chromosome anomaly has already been described in<br />

at least 1 case), 2- rare (previously described in less than 20 cases), 3- with well documented clinics <strong>and</strong> laboratory findings,<br />

<strong>and</strong> 4- with iconography <strong>of</strong> chromosomes.<br />

It is m<strong>and</strong>atory to use the specific "Submission form for Case reports":<br />

see http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org/Reports/Case_Report_Submission.html.<br />

Educational Items must be didactic, give full information <strong>and</strong> be accompanied with iconography. Translations into French,<br />

German, Italian, <strong>and</strong> Spanish are welcome.<br />

Subscription: The <strong>Atlas</strong> is FREE!<br />

Corporate patronage, sponsorship <strong>and</strong> advertising<br />

Enquiries should be addressed to Editorial@<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org.<br />

Rules, Copyright Notice <strong>and</strong> Disclaimer<br />

Conflicts <strong>of</strong> Interest: Authors must state explicitly whether potential conflicts do or do not exist. Reviewers must disclose to<br />

editors any conflicts <strong>of</strong> interest that could bias their opinions <strong>of</strong> the manuscript. The editor <strong>and</strong> the editorial board members<br />

must disclose any potential conflict.<br />

Privacy <strong>and</strong> Confidentiality – Iconography: Patients have a right to privacy. Identifying details should be omitted. If<br />

complete anonymity is difficult to achieve, informed consent should be obtained.<br />

Property: As "cards" are to evolve with further improvements <strong>and</strong> updates from various contributors, the property <strong>of</strong> the<br />

cards belongs to the editor, <strong>and</strong> modifications will be made without authorization from the previous contributor (who may,<br />

nonetheless, be asked for refereeing); contributors are listed in an edit history manner. Authors keep the rights to use further<br />

the content <strong>of</strong> their papers published in the <strong>Atlas</strong>, provided that the source is cited.<br />

Copyright: The information in the <strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong> in Oncology <strong>and</strong> Haematology is issued for general<br />

distribution. All rights are reserved. The information presented is protected under international conventions <strong>and</strong> under<br />

national laws on copyright <strong>and</strong> neighbouring rights. Commercial use is totally forbidden. Information extracted from the<br />

<strong>Atlas</strong> may be reviewed, reproduced or translated for research or private study but not for sale or for use in conjunction with<br />

commercial purposes. Any use <strong>of</strong> information from the <strong>Atlas</strong> should be accompanied by an acknowledgment <strong>of</strong> the <strong>Atlas</strong> as<br />

the source, citing the uniform resource locator (URL) <strong>of</strong> the article <strong>and</strong>/or the article reference, according to the Vancouver<br />

convention. Reference to any specific commercial products, process, or service by trade name, trademark, manufacturer, or<br />

otherwise, does not necessarily constitute or imply its endorsement, recommendation, or favouring. The views <strong>and</strong> opinions<br />

<strong>of</strong> contributors <strong>and</strong> authors expressed herein do not necessarily state or reflect those <strong>of</strong> the <strong>Atlas</strong> editorial staff or <strong>of</strong> the web<br />

site holder, <strong>and</strong> shall not be used for advertising or product endorsement purposes. The <strong>Atlas</strong> does not make any warranty,<br />

express or implied, including the warranties <strong>of</strong> merchantability <strong>and</strong> fitness for a particular purpose, or assumes any legal<br />

liability or responsibility for the accuracy, completeness, or usefulness <strong>of</strong> any information, <strong>and</strong> shall not be liable whatsoever<br />

for any damages incurred as a result <strong>of</strong> its use. In particular, information presented in the <strong>Atlas</strong> is only for research purpose,<br />

<strong>and</strong> shall not be used for diagnosis or treatment purposes. No responsibility is assumed for any injury <strong>and</strong>/or damage to<br />

persons or property for any use or operation <strong>of</strong> any methods products, instructions or ideas contained in the material herein.<br />

See also: "Uniform Requirements for Manuscripts Submitted to Biomedical Journals: Writing <strong>and</strong> Editing for Biomedical<br />

Publication - Updated October 2004": http://www.icmje.org.<br />

http://<strong>Atlas</strong><strong>Gene</strong>ticsOncology.org<br />

© ATLAS - ISSN 1768-3262


<strong>Atlas</strong> <strong>of</strong> <strong>Gene</strong>tics <strong>and</strong> <strong>Cytogenetics</strong><br />

in Oncology <strong>and</strong> Haematology<br />

OPEN ACCESS JOURNAL AT INIST-CNRS

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!