196
Imaging across Synapses Specifying Synaptic Partners Volume 143 www.cell.com Number 3 October 29, 2010

Cell 101029

  • Upload
    prasath

  • View
    125

  • Download
    13

Embed Size (px)

DESCRIPTION

'CELL" issue 29/10/2010

Citation preview

Page 1: Cell 101029

Imaging acrossSynapses

Specifying SynapticPartners

Volum

e 143 Num

ber 3 Pages 327–486 O

ctober 29, 2010

Volume 143

www.cell.com

Number 3

October 29, 2010

INSERT ADVERT

cell143_3.c1.indd 1cell143_3.c1.indd 1 10/22/2010 2:03:06 PM10/22/2010 2:03:06 PM

Page 2: Cell 101029

World's most precise plate-based HRM system

HRM and Real-time PCR System

HRM Reagent Kits

Get There Faster

Salt Lake City, Utah, USA | 1-800-735-6544 | www.idahotech.com

Our LightScanner systems will take your lab to the next level of high-sensitive mutation screening and genotyping. As the pioneers of both rapid real-time PCR and Hi-Res Melting, Idaho Technology is the only company that offers a complete system capable of superior performance at an affordable price.

Proven technology and exceptional customer support from the inventors of rapid PCR, the LightCycler®, and Hi-Res Melting.

Reach your scientifi c destinations faster with the most accurate Hi-Res Melting® systems on the market.

LightScanner Express >>> Arrivals

Rapidly generate high quality gene expression data.Specialized for T/A homozygote small amplicon genotyping.Genotype samples with greater specificity than hydrolysis probe genotyping at a fraction of the cost.

Request a FREE sample of our Hi-Res Melting Master Mix by visiting www.idahotech.com

Visit us at

ASHG, Booth #923

Visit us at

ASHG, Booth #923

ASHG, Booth #923

Mutation Discovery | Genotyping | Gene Expression

Page 3: Cell 101029

H2O2 sterilization system acessory, plus BD labware consumables kit. Limited time upgrade offer with purchase of the Sterisonic™ GxP MCO 19AICUVH. ($900 Value)

BD stem cell starter kit with Sterisonic™ GxP quote. No purchase necessary! Act now. Supplies are limited. ($150 Value)

FREE! FREE! !FREE

OFFER[ DETAILS ONLINE ]

!FREEOFFER[ DETAILS ONLINE ]

NEW.

H2O2 sterilization system acessory, plus BD labware consumables kit. Limited time upgrade offer with purchase of the Sterisonic™ GxP MCO 19AICUVH. ($900 Value) FREE! !FREE

OFFER

NEW.

Spot on results.

Intelligent design

inCuSaFe™ copper enriched stainless steel interior

Single-beam, dual capture infrared CO2 sensor

SafeCell UV protection in situ

Hydrogen peroxide vapor sterilization in situ

Good laboratory technique Good laboratory technique Good laboratory technique1

vapor sterilization vapor sterilization vapor sterilization

Good laboratory technique Good laboratory technique Good laboratory technique Good laboratory technique

+

protection protection protection

Hydrogen peroxide Hydrogen peroxide Hydrogen peroxide Hydrogen peroxide +

+

Single-beam, dual capture infrared CO infrared CO infrared CO

SafeCell UV SafeCell UV SafeCell UV SafeCell UV SafeCell UV SafeCell UV +

+

inCuSaFe inCuSaFe inCuSaFe stainless steel interior stainless steel interior stainless steel interior

Single-beam, dual capture Single-beam, dual capture Single-beam, dual capture Single-beam, dual capture Single-beam, dual capture Single-beam, dual capture +

+

©2009 Sanyo Biomedical OWS 1015 05/09

Spot on results.The industry’s fi rst in situ H2O2 sterilization with the fastest turnaround.

For maximum productivity in clinical, general purpose or the most highly compliant GMP applications, the new SANYO Sterisonic™ GxP CO2 incubator offers an impressive return on investment. With multiple contamination control safeguards, exclusive on-board H2O2 sterilization, FDA-21CFR data capture and graphical LCD display, the Sterisonic™ GxP rewards good laboratory technique with performance you can trust. Learn more, visit www.sterisonic.com or call 800-858-8442.

pictured above: Sterisonic™ GxP MCO-19AICUVH with rapid H2O2 vapor sterilization system.

www.sterisonic.com

The rapid in situ H2O2 sequence returns the fully sterilized Sterisonic™ GxP to normal use quicker

than any competitive incubator in the world.

Sterisonic™ GxP Performance and Productivity Delivers Best Effi ciency Value:

2 Hours 14 Hours

SANYOSterisonic™

Brand X

H2O2 sterilization vs. high heat sterilization = Uptime (Hours) = Downtime (Hours)

Page 4: Cell 101029

Biopotential.Unlock extraordinary potential with stem cell technologies from Sigma®.

Stem cell biology offers astonishing research potential; Sigma® Life Science has the innovations you need to discover the promise it holds. Access a world of RNAi with the MISSION® RNAi Library, efficiently edit genes of interest using advanced CompoZr® ZFN technology, and characterize your stem cells with our Prestige Antibodies®, powered by Atlas Antibodies. The applications are endless—and so are the possibilities.

bioreprogramming

wherebiobegins.com/bioreprogramming

Sigma, MISSION, CompoZr and Prestige Antibodies are registered trademarks of Sigma-Aldrich and its affiliate Sigma-Aldrich Biotechnology, L.P.

Biopotential_locklady ad_Cell Press.indd 1 8/5/2010 3:10:38 PM

Page 5: Cell 101029

Editor

Emilie Marcus

Senior Deputy Editor

Elena Porro

Deputy Editors

Robert Kruger

Connie M. Lee

Scientific Editors

Karen Carniol

Michaeleen Doucleff

Fabiola Rivas

Niki Scaplehorn

Lara Szewczak

Senior Managing Editor

Meredith Adinolfi

Deputy Managing Editor

Andy Smith

Lead Illustrator

Andrew A. Tang

Illustrators

Yvonne Blanco

Kate Mahan

Production Staff

Reyna Clancy

Editorial Assistant

Mary Beth O’Leary

Editorial Board

Abul Abbas

C. David Allis

Genevieve Almouzni

Uri Alon

Angelika Amon

Johan Auwerx

Richard Axel

Cori Bargmann

Konrad Basler

Bonnie Bassler

David Baulcombe

Jeffrey Benovic

Carolyn Bertozzi

Wendy Bickmore

Elizabeth Blackburn

Joan Brugge

Lewis Cantley

Joanne Chory

David Clapham

Andrew Clark

Hans Clevers

Stephen Cohen

Pascale Cossart

Jeff Dangl

Ted Dawson

Pier Paolo di Fiore

Marileen Dogterom

Julian Downward

Bruce Edgar

Steve Elledge

Anne Ephrussi

Ronald Evans

Witold Filipowicz

Marco Foiani

Elaine Fuchs

Yukiko Goda

Stephen Goff

Joe Goldstein

Douglas Green

Leonard Guarente

Taekjip Ha

Daniel Haber

Ulrike Heberlein

Nobutaka Hirokawa

Mark Hochstrasser

Arthur Horwich

Tony Hunter

James Hurley

Richard Hynes

Thomas Jessell

Narry Kim

Mary-Claire King

David Kingsley

Frank Kirchhoff

Richard Kolodner

John Kuriyan

Robert Lamb

Mark Lemmon

Beth Levine

Wendell Lim

Jennifer Lippincott-Schwartz

Dan Littman

Richard Losick

Scott Lowe

Tom Maniatis

Matthias Mann

Kelsey Martin

Joan Massague

Iain Mattaj

Satyajit Mayor

Ruslan Medzhitov

Craig Mello

Tom Misteli

Tim Mitchison

Alex Mogilner

Paul Nurse

Roy Parker

Dana Pe’er

Kathrin Plath

Carol Prives

Klaus Rajewsky

Venki Ramakrishnan

Rama Ranganathan

Anne Ridley

James Roberts

Alexander Rudensky

Helen Saibil

Joshua Sanes

Randy Schekman

Ueli Schibler

Joseph Schlessinger

Hans Scholer

Trina Schroer

Geraldine Seydoux

Kevan Shokat

Pamela Sklar

Nahum Sonenberg

James Spudich

Paul Sternberg

Bruce Stillman

Azim Surani

Keiji Tanaka

Craig Thompson

Robert Tjian

Jurg Tschopp

Ulrich von Andrian

Gerhard Wagner

Detlef Weigel

Alan Weiner

Jonathan Weissman

Matthew Welch

Tian Xu

Shinya Yamanaka

Marino Zerial

Xiaowei Zhuang

Huda Zoghbi

Cell Office

Cell, Cell Press, 600 Technology Square, 5th Floor, Cambridge, Massachusetts 02139

Phone: (+1) 617 661 7057, Fax: (+1) 617 661 7061, E-mail: [email protected]

Online Publication: http://www.cell.com

Cell (ISSN0092-8674) is published biweeklybyCell Press, 600TechnologySquare, 5th Floor,Cambridge, Massachusetts02139.The institutional subscription rate for

2010 is $1,360 (US and Canada) or $1,532 (elsewhere). The individual subscription rate is $202 (US and Canada) or $305 (elsewhere). The individual copy price is $50.

Periodicals postage paid at Boston, Massachusetts and additional mailing offices. Postmaster: send address changes to Elsevier Customer Service Americas,

Cell Press Journals, 11830 Westline Industrial Drive, St. Louis, MO 63146, USA.

The paper used in this publication meets the requirments of ANSI/NISO Z39.48-1992 (Permanence of Paper). Printed by Dartmouth Printing Company, Hanover, NH.

Page 6: Cell 101029

to read the latest issue of any Cell Press journal.BE THE FIRST

Register for Cell Press Email Alerts and get the complete table of contents as soon as the issue publishes online — FREE!

Cell Press Email Alerts deliver the news, research, and commentaries featured in eachjournal’s latest issue, including the full title of every article, direct links to the articles, and the complete author list. Plus, to save you time, each research article has a brief summary highlighting its significant findings.

You don’t have to be a subscriber to sign up for Cell Press Email Alerts. While subscribers have instant access to the full text of all articles listed in the Email Alerts, non-subscribers can read the abstracts of all articles as well as the full text of the issue’s Featured Article.

www.cellpress.com

C

M

Y

CM

MY

CY

CMY

K

AD7.pdf 7/24/2008 12:07:29 PM

Page 7: Cell 101029

Cell Press

President & CEO

Lynne Herndon

Editor in Chief, Vice President of Content Development

Emilie Marcus

Vice President of Marketing and Publishing

Els Bosma

Vice President of Web Development and Operations

Keith Wollman

Director of Marketing

Jonathan Atkinson

Production Manager

Meredith Adinolfi

Press Officer

Cathleen Genova

Display Advertising

Northeast/Mid-Atlantic: Victoria Macomber, ph: 508 928

1255; fax: 508 928 1256; e-mail: [email protected]

Midwest/Southeast/Eastern Canada: Inez Herrero-Redman,

ph: 585 678 4395; fax: 585 678 4722; e-mail: i.herrero@elsevier.

com

Northwest/Southwest/Western Canada: Lori Young, ph: 646

370 6312; fax: 212 462 1915; e-mail: [email protected]

California: Elizabeth Loennborn, ph: 714 655 1877; fax: 214 452

9627; e-mail: [email protected]

UK/Europe: James Kenney, ph: +44 20 7424 4216; fax: +44 18

6585 3136; e-mail: [email protected]

Asia: Wendy Xie, ph: +86 10 8520 8827; e-mail: w.xie@

elsevier.com

Classified Advertising

United States and Canada:

Gordon Sheffield, Key Account Manager, ph: 617 386 2189; fax:

617 397 2805; e-mail: [email protected]

UK, Europe, and Asia:

Sabrina Dodge, Key Account Manager, ph: +44 20 7424 4997;

fax: +44 18 6585 3136; e-mail: [email protected]

ª2010 Elsevier Inc. All rights reserved.

This journal and the individual contributions contained in it are protected

under copyright by Elsevier Inc., and the following terms and conditions

apply to their use:

Photocopying:

Single photocopies of single articles may be made for personal use as al-

lowed by national copyright laws. Permission of the Publisher and payment

of a fee arerequired forall otherphotocopying, including multipleorsystem-

atic copying, copying for advertising or promotional purposes, resale, and

all forms of document delivery. Special rates are available for educational

institutions that wish to make photocopies for nonprofit educational class-

room use. For information on how to seek permission, visit www.elsevier.

com/permissions or call (+44) 1865 843830 (UK) / (+1) 215 239 3804 (US).

Permissions:

For information on how to seek permission, visit www.elsevier.com/

permissions or call (+44) 1865 843830 (UK) / (+1) 215 239 3804 (US).

Derivative Works:

Subscribers may reproduce tables of contents or prepare lists of articles

including summaries for internal circulation within their institutions.

Permission of the Publisher is required for resale or distribution outside

the institution. Permission of the Publisher is required for all other deriv-

ative works, including compilations and translations (please consult

www.elsevier.com/permissions).

Electronic Storage or Usage:

Permission of the Publisher is required to store or use electronically any

material contained in this journal, including any article or part of an article

(please consult www.elsevier.com/permissions). Except as outlined

above, no part of this publication may be reproduced, stored in a retrieval

system, or transmitted in any form or by any means, electronic, mechan-

ical, photocopying, recording, or otherwise, without prior written permis-

sion of the Publisher.

Notice:No responsibility is assumed by the Publisher for any injury and/or dam-

age to persons or property as a matter of products liability, negligence,

or otherwise, or from any use or operation of any methods, products,

instructions, or ideas contained in the material herein. Because of rapid

advances in the medical sciences, in particular, independent verification

of diagnoses and drug dosages should be made. Although all advertis-

ing material is expected to conform to ethical (medical) standards, inclu-

sion in this publication does not constitute a guarantee or endorsement

of the quality or value of such product or of the claims made of it by its

manufacturer.

Reprints:

Article reprints are available through Cell’s reprint service; for informa-

tion, contact Nicholas Pavlow (e-mail: [email protected]; ph: (+1)

212 633 3960).

Subscription Orders and Inquiries:

Mail, fax, or e-mail address changes to Elsevier Customer Service Amer-

icas, allowing 4–6 weeks for processing. Lost or damaged issues will be

replaced, subject to availability, if Cell Press is notified within the claim

period (US and airmail delivery: 3 months from issue date; surface deliv-

ery: 4 months from issue date). Periodical delivery in the US can take up

to 3 weeks. Airmail delivery can take 2–4 weeks.

The price of a single copy of Cell is $50 (excluding special issues).

All orders must be prepaid and in writing. Please include the volume

and issue number, payment (check or credit card, MasterCard, Visa,

or American Express only), and a delivery address. Allow 4–6 weeks

for delivery.

Mailing address: Elsevier Customer Service Americas, Cell Press

Journals, 11830 Westline Industrial Drive, St. Louis, MO 63146,

USA. Toll-free phone within USA/Canada: 866 314 2355; phone for

outside US/Canada: (+1) 314 453 7038; fax: (+1) 314 523 5170; e-mail:

[email protected]; internet: www.cellpress.com or <www.cell.com>.

Funding Body Agreements and Policies:

Elsevier has established agreements and developed policies to allow au-

thors whose articles appear in journals published by Elsevier to comply

with potential manuscript archiving requirements as specified as condi-

tions of their grant awards. To learn more about existing agreements and

policies, visit http://www.cell.com/cellpress/FundingBodyAgreements.

Guide for Authors:

For a full and complete guide for authors, please go to www.cell.com/

authors.

Page 8: Cell 101029

www.neb.com

CLONING & MAPPING DNA AMPLIFICATION& PCR RNA ANALYSIS PROTEIN EXPRESSION

& ANALYSISGENE EXPRESSION

& CELLULAR ANALYSIS

UNDERSTANDING CHANGE

New tools to advance epigenetics researchFor over 35 years, New England Biolabs has been committed to understanding the mechanisms of restriction and

methylation of DNA. This expertise in enzymology has recently led to the development of a suite of validated

products for epigenetics research. These unique solutions to study DNA methylation are designed to address some

of the challenges of the current methods. EpiMark™ validated reagents simplify epigenetics research and expand the

potential for biomarker discovery.

EpiMark™ validated products include:

• Newly discovered methylation-dependent restriction enzymes

• A novel kit for 5-hmC and 5-mC analysis and quantitation

• Methyltransferases

• Histones

• Genomic DNAs

To learn how these products can help you to better understand epigenetic changes, visit neb.com/epigenetics.

Simplify DNA methylation analysis with MspJI

MspJI recognizes methylated and hydroxymethylated DNA and cleaves out 32 bp fragments for downstream sequencing analysis. Overnight digestion of 1 µg of genomic DNA from various sources with or without MspJI is shown. Note: Yeast DNA does not contain methylated DNA, therefore no 32-mer is detected.

Plant Hela (Maize) Yeast – + – + – + MspJI

32 bp

Page 9: Cell 101029

Leading EdgeCell Volume 143 Number 3, October 29, 2010

IN THIS ISSUE

SELECT

331 Gut Microbes

PREVIEWS

335 ATRX: Put Me on Repeat I. Whitehouse and T. Owen-Hughes

337 Egg’s ZP3 Structure Speaks Volumes P.M. Wassarman and E.S. Litscher

339 Monocytes Join theDendritic Cell Family

F. Sallusto and A. Lanzavecchia

341 Ephecting Excitatory Synapse Development M.B. Dalva

REVIEW

343 Chemoaffinity Revisited: Dscams,Protocadherins, andNeural Circuit Assembly

S.L. Zipursky and J.R. Sanes

SNAPSHOT

486 Neural Crest T. Sauka-Spengler and M. Bronner-Fraser

Page 10: Cell 101029

Orders (toll-free) 1-877-616-2355 | Technical support (toll-free) 1-877-678-8324 [email protected] | Inquiries [email protected] | Environmental Commitment eco.cellsignal.com

www.cellsignal.comfor quality products you can trust...

© 2010 Cell Signaling Technology, Inc. Cell Signaling Technology

® and PathScan® are registered tradem

arks of Cell Signaling Technology, Inc. Alexa Fluor ® is a registered tradem

ark of Molecular Probes, Inc.

PathScan® Signaling Nodes

Multiplex IF Kitfrom Cell Signaling Technology

Immunofluorescent analysis of MCF7 (human breast adenocarcinoma) cells insulin-treated for 5 minutes, using PathScan® Signaling Nodes Multiplex IF Kit #8999.

PathScan® Signaling Nodes Multiplex IF Kit #8999 from

Cell Signaling Technology provides a novel multiplex assay to

simultaneously assess signaling through key pathway nodes

(activated-Akt, p44/42, and S6 Ribosomal Protein) using

automated imaging or laser scanning high content platforms,

or manual immunofluorescence microscopy. The kit provides

reagents necessary to perform 100 assays (based on 100 µl

assay volume).

:: The kit allows the analysis of multiple pathway endpoints within a single sample, saving time and reagents.

:: The kit is produced and optimized in-house with the highest quality antibodies, providing you with the greatest possible specificity and sensitivity.

:: Technical support is provided by our in-house IF group who developed the product and knows it best.

#8999 Kit Targets Detection Dye Ex(max) (nm) Em(max) (nm)

Phospho-Akt (Ser473) Alexa Fluor® 555 555 565

Phospho-p44/42 (Erk1/2) (Thr202/Tyr204) Alexa Fluor® 488 495 519

Phospho-S6 Ribosomal Protein (Ser235/236) Alexa Fluor® 647 650 665

The Alexa Fluor® dye conjugated secondary antibodies are sold under license from

Invitrogen, Inc., for research use only in im

munocytochem

istry, imm

unohistochemistry, high content screening (HCS) analysis, or flow cytom

etry applications.Antibodies and Related Reagents for Signal Transduction Research

Page 11: Cell 101029

ArticlesCell Volume 143 Number 3, October 29, 2010

355 DNA Damage-Mediated Inductionof a Chemoresistant Niche

L.A. Gilbert and M.T. Hemann

367 ATR-X Syndrome Protein Targets TandemRepeats and Influences Allele-SpecificExpression in a Size-Dependent Manner

M.J. Law, K.M. Lower, H.P.J. Voon, J.R. Hughes,D. Garrick, V. Viprakasit, M. Mitson, M. De Gobbi,M. Marra, A. Morris, A. Abbott, S.P. Wilder, S. Taylor,G.M. Santos, J. Cross, H. Ayyub, S. Jones, J. Ragoussis,D. Rhodes, I. Dunham, D.R. Higgs, and R.J. Gibbons

379 Upf1 Senses 30UTR Lengthto Potentiate mRNA Decay

J.R. Hogg and S.P. Goff

390 The Long Noncoding RNA, Jpx,Is a Molecular Switchfor X Chromosome Inactivation

D. Tian, S. Sun, and J.T. Lee

404 Insights into Egg Coat Assembly andEgg-Sperm Interaction from theX-Ray Structure of Full-Length ZP3

L. Han, M. Monne, H. Okumura, T. Schwend,A.L. Cherry, D. Flot, T. Matsuda, and L. Jovine

416 Microbial Stimulation FullyDifferentiates Monocytes to DC-SIGN/CD209+

Dendritic Cells for Immune T Cell Areas

C. Cheong, I. Matos, J.-H. Choi, D.B. Dandamudi,E. Shrestha, M.P. Longhi, K.L. Jeffrey, R.M. Anthony,C. Kluger, G. Nchinda, H. Koh, A. Rodriguez, J. Idoyaga,M. Pack, K. Velinzon, C.G. Park, and R.M. Steinman

430 Endophilin Functions as a Membrane-Bending Molecule and Is Delivered toEndocytic Zones by Exocytosis

J. Bai, Z. Hu, J.S. Dittman, E.C.G. Pym,and J.M. Kaplan

442 EphB-Mediated Degradation of the RhoAGEF Ephexin5 Relieves a DevelopmentalBrake on Excitatory Synapse Formation

S.S. Margolis, J. Salogiannis, D.M. Lipton,C. Mandel-Brehm, Z.P. Wills, A.R. Mardinly,L. Hu, P.L. Greer, J.B. Bikoff, H.-Y.H. Ho,M.J. Soskis, M. Sahin, and M.E. Greenberg

456 Imaging Activity-Dependent Regulationof Neurexin-Neuroligin Interactions Usingtrans-Synaptic Enzymatic Biotinylation

A. Thyagarajan and A.Y. Ting

(continued)

Page 12: Cell 101029

017.

A1.0

115.

A ©

201

0 E

pp

end

orf A

G

� Joystick provides intuitive control

� Patented axial injection movement of the capillary

� Semi-Automatic microinjection into adherent cells

� Pre-pull capillaries for reproducible injection of adherent cells

CB

A

search level

injection level

carrier

microinjection

InjectManNI 2

www.eppendorf.com • Email: [email protected]

In the U.S.: Eppendorf North America, Inc. 800-645-3050 • In Canada: Eppendorf Canada Ltd. 800-263-8715

Microinjection is one of the core methods to introduce foreign DNA and other non-permeable molecules into cells. Nuclear injection of plasmid DNA enables rapid expression of proteins in specific cells within a population.

The menu-controlled, programmable micromanipulator InjectMan NI 2 is ideally suited for microinjection of adherent cells. Connection with the FemtoJet and the axial mounting allows injections at 45˚ angle reducing cell damage during injection and increases cell viability. This guarantees a very rapid, safe and reproducible microinjection process.

Eppendorf InjectMan NI 2 microinjector has it all:� Motorized X-Y-Z movements provide precise movement� Pre-setting and storage of up to 2 locations in X-Y-Z,

saves time in returning to pre-set work locations� Automated Home function for rapid capillary exchange� Joystick-controlled provides overall ergonomic manipulator� Fine adjustment of work speed made easy with

positioning wheel� Can be adapted to all common microscopes

For more information visit www.eppendorf.com

Microinjection simplified!

Page 13: Cell 101029

470 Nucleosome-Interacting ProteinsRegulated by DNAand Histone Methylation

T. Bartke, M. Vermeulen, B. Xhemalce,S.C. Robson, M. Mann, and T. Kouzarides

RETRACTION

485 Retraction Notice to: Assembly ofEndogenous oskar mRNA Particles forMotor-Dependent Transport inthe Drosophila Oocyte

A. Trucco, I. Gaspar, and A. Ephrussi

ANNOUNCEMENTS

POSITIONS AVAILABLE

On the cover: Intercellular protein-protein interactions are integral for many biological pro-

cesses, including synapse formation and maturation. In this issue, Thyagarajan and Ting

(pp. 456–469) report a method, biotin labeling of intercellular contacts (BLINC), to image

the dynamics of the trans-synaptic neurexin-neuroligin complex. Synaptic activity causes

neurexin-neuroligin complexes to expand in size, which is important for recruitment of

AMPA receptors during synapse maturation. On the cover, connections between branching

arms are highlighted as intercellular connections are by BLINC. Artwork by Bang Wong,

Broad Institute.

Page 14: Cell 101029

Visit www.semrock.com for a complete list of exclusively hard-coated fi lters.

A Unit of Corporation

We invite you to experience the difference at no risk. All Semrock fi lters have a 30 day, no questions asked return policy. We’ll even help you make the evaluation. Give us a call at 866-736-7625 for details.

Controlling the Brightness and Contrast of Your Image with Filters

Image courtesy of Mike Davidson at Molecular Expressions

Most Semrock fl uorescence fi lter sets are designed to provide a good balance between high brightness and high contrast under standard imaging conditions. However, brightness and contrast may actually be mutually exclusive properties in many cases. When the fl uorescence signal level from a sample is low, wider passbands on the excitation and emission fi lters allow maximum collection effi ciency. Single-molecule imaging and other low fl uorophore concentration applications typically utilize wide passband fi lters, or even long-wave-pass emission fi lters. For these applications, careful sample preparation is required to minimize undesired sample autofl uorescence.

Filter sets with narrower passbands are better for imaging samples labeled with multiple, especially closely spaced, fl uorophores to minimize bleedthrough. Narrower passbands also minimize fl uorophore photo-bleaching and decrease relative noise from undesired autofl uorescence, especially in samples with inherently high background. For these reasons better contrast images are achieved with narrower-passband fi lters from samples with suffi cient fl uorescence signal.

Advances in thin-fi lm fi lter technology pioneered by Semrock and embodied in all BrightLine® fl uorescence fi lters permit the highest-performance fl uorescence imaging while resolving the longevity and handling issues that plague fi lters made with older, soft-coating technology.

400 450 500 550 6000

10

20

30

40

50

60

70

80

90

100

Tran

smis

sion

(%)

Wavelength (nm)

FITC-5050A “High Brightness” Filter Set

FITC-2024A “High Contrast” Filter Set

650

400 450 500 550 6000

10

20

30

40

50

60

70

80

90

100

Tran

smis

sion

(%)

Wavelength (nm)650

100%

145%

Brightness Contrast

71%

100%57%

Brightness Contrast

140%

Choice of fi lter sets for the popular fl uorophore FITC allows users to select the set that best suits their imaging conditions. Spectra shown are for “Highest Brightness” (top) and “Highest Contrast” performance. The bar-graph overlays show how brightness and contrast compare to the FITC-3540C set, which is designed to achieve the best balance.

Page 15: Cell 101029

Leading Edge

In This Issue

Thymus Harbors Fugitive Cancer CellsPAGE 355

The microenvironment around a tumor influences many aspects oftumorigenesis and pathogenicity. In this issue, Gilbert and Hemann findthat, when lymphomas are treated with chemotherapeutics, the endothelialcells surrounding them respond with a prosurvival program. In someorgans, such as the thymus, this program leads to the propagation ofanti-apoptotic signals to residual tumor cells, creating a chemo-resistantniche that can subsequently support tumor relapse.

Variable Repeats Underlie Variable PenetrancePAGE 367

Mutations in the chromatin-remodeling protein ATRX cause mentalretardation and the blood disease a thalassemia, but patients with identical

ATRX mutations exhibit a wide range of phenotypes. Now, Law et al. demonstrate that ATRX binds to G-rich tandemrepeats near disease-related genes, and the magnitude of the transcriptional effect in ATRX mutants correlates with thesize of the tandem repeats. These findings suggest that ATRX helps overcome the inhibitory effects of G quadruplexstructures and illustrate a mechanism for variable disease penetrance.

Sneaking a Peek at FertilizationPAGE 404

Fertilization begins with an encounter between a sperm and an egg.However, structural information about this interaction has been verydifficult to obtain. Now, Han et al. present the full-length crystal structureof ZP3, a component of the egg coat that binds sperm at fertilization.The structure provides insights into egg coat assembly and suggestshow sperm binding may be regulated by a hypervariable region of ZP3.These findings hold promise for the rational design of nonhormonalcontraceptives.

Upf1 Sizes Up the 30UTRPAGE 379

The nonsense-mediated decay (NMD) pathway is responsible for selectively degrading messenger RNAs withextended 30 untranslated regions (30UTRs). Here, Hogg and Goff provide a mechanism for how NMD measuresthe length of the 30UTR. They show that a key NMD factor, Upf1, associates with mRNAs in a 30UTR length-depen-dent manner and that a retroviral element can stimulate translational readthrough and disrupt NMD. These findingspoint to a two-step model for NMD in which 30UTR length surveillance by Upf1 is followed by initiation of RNAdecay.

Xist-ential DilemmaPAGE 390

X-inactivation creates equal sex chromosome dosage for mammalian males and females. Xist, a long noncoding RNA,coats the silenced chromosome while an antisense RNA, Tsix, blocks Xist on the active chromosome. Tian et al. nowidentify another regulator of Xist, the RNA Jpx, which activates Xist. Tsix and Jpx antagonize each other, and theirdynamic balance determines whether an X chromosome is inactivated. Thus, Xist is controlled by two RNA switches:Tsix for the active X and Jpx for the inactive X.

Cell 143, October 29, 2010 ª2010 Elsevier Inc. 327

Page 16: Cell 101029

R&D Systems, Inc. www.RnDSystems.com

R&D Systems Europe, Ltd. www.RnDSystems.co.uk

R&D Systems China Co., Ltd. www.RnDSystemsChina.com.cn

For research use only. Not for use in diagnostic procedures.

Cancer

Development

Endocrinology

Glycobiology

Immunology

Neuroscience

Proteases

Signal Transduction

Stem Cells

For more information visit our website at www.RnDSystems.com/go/Neuroscience

R&D Systems o� ers a wide range of high quality products for neuroscience research. In addition to high performance antibodies, we o� er the most referenced collection of premium quality proteins and ELISA kits in the industry. Our catalog also includes primary rat and mouse cortical stem cells, and kits for the expansion, di� erentiation, and identi� cation of neural stem cells.

Neuroscience ResearchR&D Systems Products for

R&D Systems Tools for Cell Biology Research™

Performance.Results.

Progress.

Plexin-B2 Notch-2 O4

RGM-B HSPH1

GFAP BSRP-A Vanilloid R1

AD111_CellPress.indd 1 8/31/10 4:53 PM

Page 17: Cell 101029

Mo-DCs, Less BacteriaPAGE 416

Dendritic cells (DCs), critical antigen-presenting cells for immune control,are normally derived from bone marrow precursors distinct from mono-cytes. Here, Cheong et al. uncover a rapid conversion of blood monocytesto fully differentiated DCs, called Mo-DCs, which are recruited from bloodmonocytes into lymph nodes by the lipopolysaccharide component ofbacteria cell walls. Mo-DCs are as active as classical DCs when testedfor antigen-presenting function, and they are more numerous than classicalDCs, making Mo-DCs the dominant antigen-presenting cell in response togram-negative bacteria.

Synaptic Supply and DemandPAGE 430

Synapses operate over an extremely broad range of action potential firingrates (from <1 to >50 Hz), which demands that processes underlying synaptic transmission are also stable over a cor-responding dynamic range. Here, Bai et al. show that the rate of vesicle exocytosis at synapses regulates the avail-ability of endophilin, a protein required for endocytosis at synapses. Linking the delivery of endophilin to exocytosisfunctionally couples the rates of synaptic vesicle exocytosis and endocytosis, providing a stabilizing mechanism forsynaptic transmission.

Synapse Maturation in High DefPAGE 456

The interaction between neurexin and neuroligin across a synapse isthought to play a role in synapse development, but direct functionalevidence is lacking. In this issue, Thyagarajan and Ting report a methodto label and image protein-protein interactions at cell junctions, such asneuronal synapses. They show that neurexin-neuroligin adhesioncomplexes expand in response to synaptic activity, and this expansionpromotes the recruitment of neurotransmitter receptors, which eventuallyleads to synapse maturation.

Curb Your Synaptic EnthusiasmPAGE 442

For synapses to form at the right place and time, the development ofexcitatory synapses must be limited. Margolis et al. now show that Ephexin5, a Rho guanine-nucleotide exchangefactor (GEF), controls the number of synapses formed by restricting the synaptogenic activity of Ephrin B2 (EphB2).Moreover, alleviating this brake on synapse development requires the coordinate function of both EphB and theAngelman syndrome E3 ubiquitin ligase, Ube3A, providing a link between EphB signaling and the pathophysiologyunderlying the neurogenetic disorder Angelman Syndrome.

Chromatin CryptographersPAGE 470

Histone and DNA modifications recruit proteins that regulate chromatin function. Using a proteomics approach incombination with recombinant nucleosomes methylated on both DNA and histone H3, Bartke et al. now identifychromatin-binding proteins, including origin recognition complex (ORC) and Fbxl11/KDM2A, which are modulatedby these two distinct classes of modifications. This study presents a new tool for studying the dynamics betweendifferent types of chromatin modifications and demonstrates that epigenetic readers can decode the landscape ofchromatin modifications.

Cell 143, October 29, 2010 ª2010 Elsevier Inc. 329

Page 18: Cell 101029

Expand your stem cell library and save today on the latest books on stem cells

and regenerative medicine

Cell Stem Cell subscribers save 25% on their book orderSecure ordering online at elsevierdirect.comEnter promo code 28024 at check outPrices and publication dates subject to change without notice.

Stem Cells Scientific Facts and FictionChristine Mummery, Ian Wilmut, Anja Van de Stolpe and Bernard RoelenNovember 2010 | 400 pages | Paperback | $79.95 | €57.95 | £48.99 | ISBN: 9780123815354

Principles of Regenerative Medicine, 2nd EditionAnthony Atala and Robert LanzaNovember 2010 | 1400 pages | Hardback | $199.95 | €143.00 | £125.00 | ISBN: 9780123814227

Heart Development and Regeneration, 2-Volume SetNadia Rosenthal and Richard P. HarveyJune 2010 | 1072 pp. | Hardback | $199.95 | €143.00 | £125.00 | AU$296.00 | ISBN: 9780123813329

Essentials of Stem Cell Biology, 2nd EditionRobert Lanza, Roger Pedersen, John Gearhart, E. Donnall Thomas, Brigid Hogan, James Thomson, Douglas Melton and Sir Ian WilmutJune 2009 | 600 pp. | Hardback | $199.95 | €134.00 | £125.00 | AU$302.00 | ISBN: 9780123747297

Foundations of Regenerative Medicine Clinical and Therapeutic ApplicationsAnthony Atala, Robert Lanza, James Thomson and Robert NeremSeptember 2009 | 750 pp. | Hardback | $99.95 | €66.95 | £60.99|AU$148.00 | ISBN: 9780123750853

Stem Cell Anthology From Stem Cell Biology, Tissue Engineering, Regenerative Medicine, Cloning and Stem Cell MethodsBruce M. CarlsonOctober 2009 | 450 pp. | Hardback | $150.00 | €100.00 | £95.00 |AU$222.00 | ISBN: 9780123756824

Essential Stem Cell Methods A Volume in the Reliable Lab Solutions SeriesRobert Lanza and Irina KlimanskayaApril 2009 | 628 pp. | Paperback | $75.00 | €50.95 | £45.99 |AU$111.00 | ISBN: 9780123750617

Tissue EngineeringClemens van Blitterswijk, Peter Thomsen, Jeffrey Hubbell, Ranieri Cancedda, Anders Lindahl Sahlgrenska,Jerome Sohier and David F. WilliamsMarch 2008 | 760 pp. | Hardback | $115.00 | €76.95 | £69.99 |AU$170.00 | ISBN: 9780123708694

Human Stem Cell Manual A Laboratory GuideJeanne F. Loring, Robin L. Wesselschmidt and Philip H. SchwartzJune 2007 | 488 pp. | Spiral bound | $88.95 | €59.95 | £53.99 |AU$132.00 | ISBN: 9780123704658

Handbook of Stem Cells 2-Volume Set with CD-ROM Vol. 1–2Vol. 1 – Embryonic Stem CellsVol. 2 – Adult & Fetal Stem CellsRobert Lanza, Roger Pedersen, Helen Blau, E. Donnall Thomas, John Gearhart, James Thomson, Brigid Hogan, Catherine Verfaillie, Douglas Melton, Irving Weissman, Malcolm Moore and Michael WestSeptember 2004 | 1,760 pp. | Hardback | $566.00 | €380.00 | £345.00 | AU$817.00 | ISBN: 9780124366435

Page 19: Cell 101029

Leading Edge

Select: Gut Microbes

Our intestines host trillions of bacteria, most of which are beneficial to our health most of the time. Occasionally,however, a change in conditions, or the entry of a pathogenic strain, leads to disease. Recent papers shed new lightonto the complex interactions that determine intestinal health and disease.

Preparing the Gut for Bacterial EncounterIntestinal cells first come into contact with bacteria after birth, as they transition fromthe sterile uterine environment to the outside microbe-filled world. Given that encoun-ters with bacteria normally trigger immune activation and inflammation, which maycause tissue damage, neonatal intestinal epithelial cells are programmed to undergoa period of tolerance, in which bacteria do not elicit an immune response. Chassinet al. (2010) now show in mouse that during tolerance the microRNA miR-146asuppresses the inflammatory pathway mediated by Toll-like receptors by repressingthe translation of the interleukin 1 receptor associated kinase 1 (IRAK1). Things, how-ever, are not as simple as they seem. In a surprising twist, the authors find that Toll-likereceptor 4 (TLR4) signaling in the neonate epithelium is required for the downregulationof IRAK1. Furthermore, both TLR4 and IRAK1 are required for maintaining elevatedmiR-146a levels, and both are also required for expression of genes that regulatecell survival, differentiation, and metabolism and hence promote cellular homeostasis.In other words, the immune pathway is not simply turned off at birth but is rather activelymodulated to accomplish tolerance and to allow intestinal cells to express the setof genes necessary for their maturation. Interestingly, the authors find that IRAK1expression reappears at weaning (21 days after birth in mice), when mice begin eating

solid food—the point at which mice may encounter pathogenic bacteria and need to mount an immune response. At thistime epithelial proliferation increases, ending the continuous TLR4/IRAK1 signaling that maintains tolerance andlowering miR-146a expression. The ultimate triggers that initiate and end tolerance remain unknown, however, and itwould be particularly interesting to study the regulation of the corresponding pathway in humans. Despite significantdifferences in the maturity of the neonate gut between mice and men, both have to cope with the sudden exposureto microbial stimuli after birth and to establish a life-long, stable host-microbe homeostasis.C. Chassin et al. (2010). Cell Host Microbe 8, 1–11.

Microbes Give Epithelial Proliferation a BoostRapid turnover of epithelial cells is a hallmark of healthy intestines. The rate of prolif-eration is regulated by both Wnt signaling and microbes, at least in adult tissue. Astudy from the Guillemin lab (Cheesman et al., 2010) now investigates the roles ofmicrobes and Wnt signaling during development in the intestines of zebrafish larvae.The larval period corresponds to the time zebrafish first encounter microbes, analo-gous to the neonatal period in humans and mice, and is the time when the epithelialproliferation rate is first established in the intestines. Cheesman et al. (2010) findthat microbes in the larval gut and Wnt signaling promote epithelial proliferation, asthey do in adults. They then ask whether microbes use the Wnt pathway to promoteproliferation and find that the answer is complex. They provide evidence that oneparticular resident bacterium, Aeromonas veronii, secretes a proliferation signal thatacts on intestinal cells to promote the accumulation of b-catenin, a key componentof the Wnt signaling pathway. A mutation in TCF4, a transcription factor downstreamof Wnt signaling, partially blocks the effect of A. veronii on cell proliferation. Theseresults suggest that resident microbes promote proliferation in part through effects on the Wnt pathway. However,the authors also show that axin, an upstream regulator of the Wnt pathway, does not affect the response to microbes,and that microbes act through the Myd88 protein, an adaptor downstream of Toll-like receptors (TLRs). Thus, it seemsthat the microbial pathway for regulating proliferation intersects the Wnt pathway but also acts independently, througha mechanism that will need to be explored in future studies.S. Cheesman et al. (2010). Proc. Natl. Acad. Sci. USA. Published online October 4, 2010. 10.1073/pnas.1000072107.

Intraepithelial lipopolysaccharide

(red) in a 6-day-old mouse gut

promotes epithelial tolerance.

Image courtesy of M. Hornef.

The epithelium (green) of the zebra-

fish larval intestine and lumenalbacteria (red). Image courtesy of

K. Guillemin.

Cell 143, October 29, 2010 ª2010 Elsevier Inc. 331

Page 20: Cell 101029

years of leadership in human genetics research,

education and service.

1948–2008www.ashg.org

60

Page 21: Cell 101029

Thriving in Inflammation’s WakeWinter et al. (2010) report an ingenious strategy used by the entericpathogen Salmonella enterica serotype Typhimurium to create a growthadvantage for itself in the gut. In so doing, the authors tie together twopreviously unrelated observations: (1) that S. Typhimurium causes acuteintestinal inflammation, which allows the bacterium to outcompete othermicrobes in the gut, and (2) that S. Typhimurium can use tetrathionate asan electron acceptor for respiration leading to enhanced growth, atleast in vitro. Winter et al. now show that inflammation caused by thepathogen leads to production of tetrathionate in the mouse intestine.The authors show that a compound produced in the cecum, thiosulfate(S2O3

2�), can be converted to tetrathionate by reactive oxygen species,which are produced by neutrophils during inflammation. Salmonellastrains lacking ttrA, a gene required for tetrathionate-dependentrespiration, do not grow as well as wild-type bacteria both in vitro andin vivo, and in vivo they cannot outcompete other microbes in the gut.The paper thus suggests that Salmonella has a good reason for inducinginflammation in the intestine: byproducts of inflammation, includingoxygen radicals and tetrathionate, allow it to thrive. The authors mentionthat another enteric pathogen, Yersinia enterocolitica, also harbors the

gene cluster that confers tetrathionate respiration ability, and future work will reveal whether other pathogenic bacteriause similar mechanisms for competing with host microbes. These results also raise the possibility of targeting thetetrathionate respiration pathway to specifically inhibit the growth of pathogens and not resident bacteria.S. Winter et al. (2010). Nature 467, 426–429.

Bacterial Toxins’ Multiple Choice: A or BClostridium difficile infections cause life-threatening diarrhea andinflammation, occurring most frequently when antibiotic treatmenteliminates other bacterial strains in the intestine. C. difficile producestwo toxins, toxin A and toxin B, that both target Rho GTPases, leadingto cytoskeletal disruption. Previous reports came to conflicting conclu-sions regarding the relative importance of each toxin to the virulenceof C. difficile. Some studies suggest that toxin A is sufficient for bacterialvirulence, and that toxin B alone cannot cause virulence, whereasanother study suggests the opposite, that toxin B was virulent on itsown, but not toxin A. Kuehne et al. (2010) re-examine this issue bygenerating mutant strains of C. difficile lacking toxin A, toxin B, orboth. They show that strains harboring just one toxin, either A or B,are virulent both in cultured cells and in a hamster model for the disease. Only when both toxins are knocked out(in a double-mutant strain, the first double mutant produced in C. difficile) does the bacteria become avirulent. Theauthors conclude that both toxins contribute to the disease and suggest that both need to be taken into account inthe design of treatments for C. difficile infections. Given that both toxins are glucosyltransferases targeting thesame GTPases, the results do raise the question of why the bacteria need two toxins, and what, if any, advantagehaving both toxins confers.S. Kuehne et al. (2010). Nature 467, 711–714.

Ilil Carmi

S. Typhimurium. Image by Rocky Mountain Labora-

tories, NIAID, NIH.

C. difficile bacteria. Image courtesy of S. Baban.

Cell 143, October 29, 2010 ª2010 Elsevier Inc. 333

Page 22: Cell 101029

XenoWorksTM Microinjection WorkstationSmooth, responsive, precise. Always.

ONE DIGITAL DRIVE, NOVATO, CA. 94949 PHONE: 415.883.0128 | FAX: 415.883.0572

EMAIL: [email protected] | WWW.SUTTER.COM

MicromanipulatorHighly ergonomic inverted joystickOne-touch coarse and fine controlSuperior mechanical stability

Digital MicroinjectorDual channel pneumatic microinjectorHolds, transfers and injects - all from a single remote keypad

Analog MicroinjectorUse with oil, water, or airInterchangeable syringes

F I N E S U R G I C A L

I N S T R U M E N T S

F O R R E S E A R C H ™

SHIPPING GLOBALLYSINCE 1974

Request a catalog at finescience.com or call 1-800-521-2109.

Naturally beautiful.

Page 23: Cell 101029

Leading Edge

Previews

ATRX: Put Me on RepeatIestyn Whitehouse1 and Tom Owen-Hughes2,*1Molecular Biology Program, Sloan-Kettering Institute, New York, NY 10065, USA2Wellcome Trust Centre for Gene Regulation and Expression, College of Life Sciences, University of Dundee, Dundee DD1 5EH, UK*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.10.021

Mutations in the chromatin-remodeling protein ATRX cause alpha thalassaemia and mental retar-dation, but the severity of the disorder is independent of the specific mutation. In this issue of Cell,Law et al. (2010) demonstrate that ATRX alters gene expression by binding to G-rich tandemrepeats, and the degree of transcriptional silencing caused by ATRX mutations correlates withthe number of repeats.

The alpha thalassaemia/mental retarda-

tion syndrome X-linked gene, or ATRX,

encodes a large helicase protein involved

in maintaining chromatin structure. Pa-

tients with mutations in the ATRX gene

typically exhibit severe mental retardation,

development defects, and a blood disease

called alpha thalassaemia, characterized

by a deficiency in the alpha globin protein.

Approximately 113 unique mutations in the

ATRX gene have been identified from >180

families, but exactly how these mutations

alter gene expression is not well under-

stood. Now Law et al. (2010) make signifi-

cant progress toward answering this

question by identifying where ATRX binds

in both the human and mouse genomes.

In addition, they provide an explanation

for why patients with identical mutations

in the ATRX gene display a broad variation

in phenotypes.

The ATRX gene encodes at least two

alternatively spliced transcripts that

give rise to slightly different proteins of

265 kDa and 280 kDa. The C-terminal

region contains a helicase/ATPase do-

main that shares sequence homology

with the Sucrose Non-Fermenting 2

(SNF2) family of chromatin-remodeling

enzymes (Gibbons et al., 1995). The

N-terminal region contains the ADD

(ATRX-DNMT3-DNMT3L) domain with a

plant homodomain zinc finger that may

interact with the tail of histone H3 (Argen-

taro et al., 2007). Mutations in ATRX that

cause alpha thalassaemia/mental retar-

dation syndrome X-linked or ATR-X

syndrome correlate with high sequence

conservation in these two domains, with

�30% and �50% of the mutations occur-

ring in the helicase/ATPase and ADD

domains, respectively.

ATRX is known to interact with the

death domain-associated protein DAXX

(Xue et al., 2003). More recently, re-

searchers demonstrated that DAXX is a

histone chaperone with specificity for

the histone H3 variant, H3.3 (Drane

et al., 2010; Goldberg et al., 2010; Lewis

et al., 2010; Wong et al., 2010). Although

both ATRX and DAXX are required for

H3.3 incorporation at telomeres, H3.3

incorporation in coding regions and near

binding sites of transcription factors

depends on a different histone chap-

erone, called Hira (Goldberg et al., 2010).

Thus, it is still unclear what factors deter-

mine where ATRX and DAXX incorporate

H3.3.

The new findings by Law et al. make

great strides toward answering this

question. Previous immunofluorescence

studies found that ATRX preferentially

interacts with a number of repetitive

DNA sequences, such as arrays of DNA

encoding ribosomal RNA (i.e., rDNA

arrays), a Y-specific satellite, and a repeat

sequence adjacent to a telomere (PMID:

10742099). This led Law et al. to investi-

gate whether ATRX binds to other repeti-

tive elements across the genome. Many

standard genome-wide protocols pre-

clude such analysis because repetitive

DNA elements give rise to spurious false

positive signals, and thus these se-

quences are routinely removed from

study. To overcome this technical hurdle,

Law et al. adapt a chromatin immunopre-

cipitation sequencing approach (ChIP-

Seq) by normalizing the signal intensity

to the size of the repeat. This provides a

relatively unbiased view of ATRX binding

across the genome and reveals �1000

stringent targets for ATRX.

A key finding of this study is that, in both

human and mouse cells, the targets of

ATRX include CpG islands (i.e., regions

of the genome with a high frequency of un-

methylated cytosine guanine dinucleo-

tides) and G-rich tandem repeats. Both

these DNA patterns are found at repeats

in telomeres, sequences adjacent to

telomeres (i.e., subtelomeric regions),

and rDNA. Moreover, Law and colleagues

show that ATRX predominantly binds to

G-rich tandem repeats in or near genes

that often display altered expression

patterns in patients with ATR-X syndrome.

Law and colleagues found that, in

erythroid cells, ATRX strongly localizes

�1 kb upstream of the alpha globin

gene HBM. This peak of ATRX binding

occurs within a tandem repeat, called

jz VNTR (CGCGGGGCGGGGG)n, where

the number of repeats (n) varies between

individuals. Interestingly, Law et al. find

that when ATRX is mutated, the most

downregulated genes in this gene cluster

are the alpha-like globin genes closest

to the jz VNTR repeat, and their down-

regulation scales with their proximity to

the ATRX binding site. The identification

of ATRX binding sequences within the

alpha globin gene cluster provides a direct

explanation as to why patients with

mutations in ATRX exhibit alpha thalas-

saemia.

Then Law and colleagues go a step

further and provide a molecular explana-

tion for how two individuals with the

same mutations in ATRX could have

different severities of alpha thalassaemia.

They demonstrate that patients with the

largest expansion of the jz VNTR repeat

have the greatest reduction in the expres-

sion of the alpha globin gene. At the

Cell 143, October 29, 2010 ª2010 Elsevier Inc. 335

Page 24: Cell 101029

extreme end of the spectrum, this ulti-

mately leads to a total silencing or ‘‘mono-

allelic expression’’ of the alpha globin

gene.

Some tandem repeat sequences that

are rich in guanine nucleotides assemble

into non-B-form structures called

G-quadruplexes or G4 DNA (Figure 1).

These structures form readily in vitro

and, once created, are very stable. Law

and colleagues note that �50% of the

ATRX target sequences are predicted to

likely adopt the G-quadruplex conforma-

tion, and they demonstrate that seven of

these target sequences do indeed form

G-quadruplex structures in isolation.

Moreover, ATRX preferentially binds to

the quadruplex structure over the B-form

DNA in vitro. These observations suggest

a model in which ATRX localizes to

specific regions of the genome through

its association to G-quadruplexes. Then

through its interaction with DAXX, ATRX

directs the incorporation of H3.3 into

that region of the genome (Figure 1).

However, this model is only a hypoth-

esis, and exactly what ATRX does at

these target sites is still a key question.

Several studies have shown that ATRX

can alter nucleosome structure (Lewis

et al., 2010; Xue et al., 2003), but this

does not rule out the possibility that the

primary substrate of the ATRX helicase

motor is nonchromatin DNA. For example,

if ATRX removes DNA quadruplexes, the

association with DAXX might be sufficient

to promote chromatin assembly, which

could then stabilize a B-DNA conforma-

tion (Figure 1). Consistent with this hy-

pothesis, a recent study found that

lowering the levels of ATRX compromises

telomere integrity (Wong et al., 2010).

Another important question is, what is

the function of H3.3 at these ATRX target

sequences? Assembled onto DNA

throughout the cell cycle, H3.3 is a non-

replicative histone that is often consid-

ered a hallmark of transcriptionally active

chromatin in genetic regions. However,

H3.3 is also known to mark many impor-

tant regulatory DNA sequences, including

both active gene promoters and DNA

elements with insulator activity (Jin et al.,

2009). Interestingly, the presence of

H3.3 at sites directed by ATRX is required

for repression of transcription at telomeric

repeats (Goldberg et al., 2010). Further-

more, ATRX has additional links to repres-

sive chromatin. For example, it associates

with the heterochromatin proteins HP1a

and HP1b (Berube et al., 2000), and the

presence of these proteins at telomeres

is dependent on ATRX (Wong et al., 2010).

Together these results suggest that

perhaps ATRX and H3.3 maintain the

boundary between regions of transcrip-

tionally active chromatin and inactive

chromatin (i.e., heterochromatin). Loss of

ATRX may then result in the spreading of

heterochromatin along DNA, resulting in

the progressive silencing of nearby genes

in cis, such as the alpha globin cluster.

Such a scenario would also deplete the

effective concentration of the protein

factors necessary for the formation of

heterochromatin, providing a plausible

explanation for the defects in telomeric

silencing seen in cells with ATRX and

DAXX mutations. Clearly, the new findings

by Law and colleagues demonstrate that

repetitive DNA may be ‘‘simple’’ in terms

of DNA sequence, but functionally they

are anything but.

REFERENCES

Argentaro, A., Yang, J.C., Chapman, L., Kowalc-

zyk, M.S., Gibbons, R.J., Higgs, D.R., Neuhaus,

D., and Rhodes, D. (2007). Proc. Natl. Acad. Sci.

USA 104, 11939–11944.

Berube, N.G., Smeenk, C.A., and Picketts, D.J.

(2000). Hum. Mol. Genet. 9, 539–547.

Drane, P., Ouararhni, K., Depaux, A., Shuaib, M.,

and Hamiche, A. (2010). Genes Dev. 24,

1253–1265.

Gibbons, R.J., Picketts, D.J., Villard, L., and Higgs,

D.R. (1995). Cell 80, 837–845.

Goldberg, A.D., Banaszynski, L.A., Noh, K.M.,

Lewis, P.W., Elsaesser, S.J., Stadler, S., Dewell,

S., Law, M., Guo, X.Y., Li, X., et al. (2010). Cell

140, 678–691.

Jin, C.Y., Zang, C.Z., Wei, G., Cui, K.R., Peng,

W.Q., Zhao, K.J., and Felsenfeld, G. (2009). Nat.

Genet. 41, 941–945.

Law, M.J., Lower, K.M., Voon, H.P.J., Hughes,

J.R., Garrick, D., Viprakasit, V., Mitson, M., De

Gobbi, M., Marra, M., Morris, A., et al. (2010). Cell

143, this issue, 367–378.

Lewis, P.W., Elsaesser, S.J., Noh, K.M., Stadler,

S.C., and Allis, C.D. (2010). Proc. Natl. Acad. Sci.

USA 107, 14075–14080.

Wong, L.H., McGhie, J.D., Sim, M., Anderson,

M.A., Ahn, S., Hannan, R.D., George, A.J., Morgan,

K.A., Mann, J.R., and Choo, K.H.A. (2010).

Genome Res. 20, 351–360.

Xue, Y.T., Gibbons, R., Yan, Z.J., Yang, D.F.,

McDowell, T.L., Sechi, S., Qin, J., Zhou, S.L.,

Higgs, D., and Wang, W.D. (2003). Proc. Natl.

Acad. Sci. USA 100, 10635–10640.

Figure 1. Could ATRX Help to Convert G-Quadruplex DNA to Duplex DNA?Some DNA sequences rich in guanine and cytosine nucleotides are capable of adopting a G4-quadruplexconfiguration instead of the standard B-form duplex (left). The chromatin-remodeling protein ATRX bindspreferentially to DNA sequences that have the potential to form G4-quadruplexes (Law et al., 2010). ATRXbelongs to a family of proteins that can translocate along duplex DNA, which may help to convertG4-quadruplexes to duplex DNA. ATRX also associates with a histone chaperone, DAXX, which can directthe assembly of nucleosomes containing the histone variant H3.3 (middle). Nucleosome assembly maythen further stabilize DNA in a duplex configuration (right).

336 Cell 143, October 29, 2010 ª2010 Elsevier Inc.

Page 25: Cell 101029

Leading Edge

Previews

Egg’s ZP3 Structure Speaks VolumesPaul M. Wassarman1,* and Eveline S. Litscher1

1Department of Developmental and Regenerative Biology, Mount Sinai School of Medicine, 1468 Madison Avenue, New York, NY 10029, USA*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.10.013

Binding of mammalian sperm to eggs depends in part on ZP3, a glycoprotein in the egg’s extracel-lular coat, the zona pellucida. In this issue, Han et al. (2010) describe the structure of an avian ZP3homolog, providing insights into ZP3 processing and polymerization and the roles of the ZP3polypeptide and its carbohydrate in sperm binding.

The plasma membrane of mammalian

eggs is surrounded by a relatively thick

extracellular coat called the zona pellu-

cida (ZP). It is composed of long intercon-

nected fibrils that consist of only a few

proteins held together by noncovalent in-

teractions. For example, the mouse egg’s

ZP consists of three glycosylated pro-

teins, called ZP1–3, that are synthesized,

secreted, and assembled by growing

oocytes (Wassarman, 2008). ZP proteins

have been conserved for more than 600

million years, and proteins closely related

to ZP1–3 are found in the ZP of all

mammalian eggs, including humans, as

well as in the extracellular coat (vitelline

envelope) of nonmammalian eggs. During

fertilization, sperm must bind to and then

penetrate the ZP in order to reach and

fuse with the egg’s plasma membrane to

produce a zygote. It has been known for

some time that sperm bind to the ZP of

unfertilized eggs but do not bind to the

ZP of fertilized eggs (Figure 1A) (Florman

and Ducibella, 2006). In this context,

a wide variety of evidence suggests that

ZP3 functions as a receptor during

binding of sperm to eggs (Wassarman

and Litscher, 2008). Both ZP3 polypep-

tide and its attached carbohydrate groups

have been implicated in binding of sperm

to the ZP, but it has not been possible to

reconcile the results of three decades of

experiments on ZP3 with a three-dimen-

sional structure for the protein. Now Han

et al. (2010) overcome the many problems

associated with crystallization of ZP3 and

determine the structure of full-length

chicken ZP3 (cZP3) at 2.0 A resolution

by X-ray crystallographic methods.

All ZP proteins are synthesized as pre-

cursor polypeptides possessing an N-ter-

minal signal sequence and a C-terminal

propeptide that contains a transmem-

brane domain, a protease cleavage site,

and a hydrophobic patch (external hydro-

phobic patch, EHP) (Figure 1B). The latter

is thought to interact with another hydro-

phobic patch (internal hydrophobic patch,

IHP) along the nascent polypeptide to

prevent premature polymerization of ZP

proteins. During secretion of ZP proteins

the propeptide, including the EHP, is

excised from nascent polypeptides,

thereby enabling them to polymerize.

Furthermore, ZP proteins are founding

members of a very large class of proteins

that have diverse functions and are found

in a variety of tissues in all multicellular

eukaryotes (Jovine et al., 2005). All of

these proteins possess a ZP domain that

consists of �260 amino acids and 8–12

conserved Cys residues present as disul-

fides. Each ZP domain has an N-terminal

(ZP-N) and C-terminal (ZP-C) subdomain

separated by a short linker region (Fig-

ure 1B). The structure of ZP-N represents

a new subtype of the immunoglobulin (Ig)-

like fold (Monne et al., 2008) and is

thought to be responsible for generating

polymers of ZP proteins. In this context,

it has been shown that mutations in ZP-

N can result in severe pathologies, such

as infertility, deafness, and cancer. It is

likely that polymers assembled by

different types of ZP domain proteins

share a similar structure.

The structure of cZP3 provides a wealth

of information about ZP proteins (Fig-

ure 1C). Han et al. (2010) show that

ZP-C adopts an Ig-like fold with the

same topology as ZP-N, suggesting that

ZP proteins may have arisen by duplica-

tion of a common Ig-like domain. Within

crystals, cZP3 forms antiparallel dimers

held together by interactions between

ZP-N and ZP-C of opposing molecules,

and Han et al. (2010) show that dimer

formation is essential for cZP3 secretion

from cells. These findings are consistent

with the propensity of purified ZP proteins

to polymerize in vitro and with the inability

of mouse oocytes lacking either ZP2 or

ZP3 to assemble a ZP in vivo (Wassar-

man, 2008). The latter has been attributed

to the failure to form intracellular ZP2-ZP3

dimers that can then polymerize in the

extracellular space into long fibrils. From

the structure of cZP3 it appears likely

that disulfides of the ZP-C subdomain

determine whether ZP proteins form

homo- or heteropolymers. However, ad-

ditional experiments that address this

issue, including the generation of mutant

ZP proteins, will be required to confirm

such a role for ZP-C disulfides.

The structure of cZP3 reveals that, as

previously proposed (Jovine et al., 2005),

the EHP present in the propeptide acts

as a ‘‘molecular glue’’ that maintains the

dimer in a conformation required for se-

cretion but that is incompatible with poly-

merization of the dimer into higher-order

structures. The structure of the cZP3

dimer also suggests that the transmem-

brane region of the propeptide may

specifically orient the precursor molecule

during proteolytic processing at the oo-

cyte membrane, thus enabling it to be

incorporated into the ZP. Indeed, this

conclusion is consistent with previous

findings (Jovine et al., 2005).

It has been proposed that the

C-terminal region of ZP3 lying just down-

stream of its ZP domain is, at least in

part, the binding site for sperm (Fig-

ure 1B) (Wassarman and Litscher, 2008).

Several studies have concluded that this

particular region of the polypeptide

Cell 143, October 29, 2010 ª2010 Elsevier Inc. 337

Page 26: Cell 101029

exhibits considerable inter-

specific sequence diversity

due to positive Darwinian

selection (Turner and Hoek-

stra, 2008) and could form

the basis of species-re-

stricted fertilization. On the

other hand, whether sperm

binding to ZP3 depends on

the protein’s polypeptide,

carbohydrate, or both is un-

clear. Although a role for

carbohydrate in many other

types of cell-cell adhesion is

well established (Varki et al.,

2009), its role in sperm-egg

interaction remains contro-

versial (Clark and Dell, 2006).

Han et al. (2010) address

the role of ZP3’s carbohydrate

in sperm binding directly

because their engineered

cZP3 possesses a single

O-glycan, probably Galb1-

3GalNAc, linked to threonine

168. The glycan is located on

thesurface ofcZP3 inaflexible

region of the polypeptide and

should be readily accessible

to sperm (Figure 1C). Thus,

the glycan, together with the

nearby cZP3 hypervariable

C-terminal polypeptide, could

form a docking platform for

sperm. Han et al. (2010)

analyze the binding of chicken

sperm to wild-type cZP3 and

to a mutant cZP3 in which

threonine 168 was converted

to alanine. They find that elim-

ination of the O-glycan causes

a large decrease (�80%) in

sperm binding to cZP3. This

result provides convincing evidence for a

role of this carbohydrate in sperm binding

to cZP3. It is of interest that this O-glycan

site, called site 1, is retained from cZP3 to

human ZP3. Another O-glycan site, called

site 2, lies very close to site 1 and may also

be involved in sperm binding.

In a recent report, Gahlay et al. (2010)

concluded that sperm fail to bind to the ZP

of fertilized eggs due to limited proteolysis

of ZP2 shortly after fertilization (Figure 1A).

Han et al. (2010) suggest that these findings

may be due to structural rearrangements

within the extracellular coat following fertil-

ization that result in shielding of the ZP3-

binding surface (i.e., its polypeptide and

O-glycan). However, this explanation does

not account for several observations. First,

ZP3purified fromunfertilizedeggZP inhibits

binding of sperm to eggs, but ZP3 purified

from fertilized egg ZP does not. Second,

solubilized ZP and purified ZP3 from unfer-

tilized eggs induce sperm to undergo

cellular exocytosis (i.e., the acrosome reac-

tion), but solubilized ZP and purified ZP3

from fertilized eggs do not. Rather, these

observations indicate that ZP3

is somehow modified shortly

after fertilization, possibly by

cortical granule enzymes, and

thereby inactivated as

a receptor for sperm. Further

structural studies are needed

to resolve this thorny issue.

In conclusion, the paper

by Han et al. (2010) is a major

breakthrough in the pur-

suit of mechanisms involved

in mammalian fertilization.

Comparable structural studies

on other ZP proteins, as well

as on other sperm and egg

proteins thought to participate

in fertilization, may lead to an

understanding of mutations

that cause infertility, the devel-

opment of new means of

contraception, and other ad-

vances inhuman reproduction.

REFERENCES

Clark, G.F., and Dell, A. (2006). J.

Biol. Chem. 281, 13853–13856.

Florman, H.M., and Ducibella, T.

(2006). Mammalian fertilization. In

Physiology of Reproduction, J.D.

Neill, ed. (New York: Academic

Press), pp. 55–112.

Gahlay, G., Gauthier, L., Baibakov,

B., Epifano, O., and Dean, J.

(2010). Science 329, 216–219.

Han, L., Monne, M., Okumura, H.,

Schwend, T., Cherry, A.L., Flot, D.,

Matsuda, T., and Jovine, L. (2010).

Cell 143, this issue, 404–415.

Jovine, L., Darie, C.C., Litscher,

E.S., and Wassarman, P.M. (2005).

Annu. Rev. Biochem. 74, 83–114.

Monne, M., Han, L., Schwend, T., Burendahl, S.,

and Jovine, L. (2008). Nature 456, 653–657.

Turner, L.M., and Hoekstra, H.E. (2008). Int. J. Dev.

Biol. 52, 769–780.

Varki, A., Cummings, R.D., Esko, J.D., Freeze,

H.H., Hart, G.W., and Etzler, M.E. (2009). Essen-

tials of Glycobiology, Second Edition (Cold Spring

Harbor, NY: Cold Spring Harbor Laboratory

Press)., pp 784.

Wassarman, P.M. (2008). J. Biol. Chem. 283,

24285–24289.

Wassarman, P.M., and Litscher, E.S. (2008). Int. J.

Dev. Biol. 52, 665–676.

Figure 1. Clues to Sperm-Egg Binding(A) A fully grown oocyte is surrounded by a thick extracellular coat, the zonapellucida (ZP), that is composed of glycoproteins. Sperm bind tightly to theZP of unfertilized eggs, but they are unable to bind to the ZP of fertilizedeggs because the ZP glycoproteins are modified following fertilization.(B) The glycoprotein ZP3 is a key component of the ZP of all mammalian eggsand apparently serves as a receptor for sperm binding. The mature ZP3 poly-peptide has an N-terminal signal sequence (red), a ZP domain that consists oftwo subdomains, ZP-N and ZP-C (blue), and a C-terminal region that hasa protease cleavage site (yellow), an external hydrophobic patch (EHP, green),and a transmembrane domain (gray).(C) In the X-ray crystallographic structure of an avian ZP3 homolog, the glyco-protein forms a dimer in which the ZP-N subdomain of one molecule interactswith the ZP-C subdomain of another molecule to hold the dimer together (Hanet al., 2010).

338 Cell 143, October 29, 2010 ª2010 Elsevier Inc.

Page 27: Cell 101029

Leading Edge

Previews

Monocytes Join theDendritic Cell FamilyFederica Sallusto1,* and Antonio Lanzavecchia1,*1Institute for Research in Biomedicine, CH-6500 Bellinzona, Switzerland*Correspondence: [email protected] (F.S.), [email protected] (A.L.)

DOI 10.1016/j.cell.2010.10.022

Dendritic cells are professional antigen-presenting cells that mediate immunity and tolerance.Cheong et al. (2010) uncover a new route for dendritic cell production in vivo. They show that inresponse to infection by gram-negative bacteria, monocytes are recruited to the lymph node wherethey rapidly differentiate into dendritic cells that present antigens to T cells.

Monocytes are circulating cells of the

mononuclear phagocyte system that have

been typically considered the precursors

of tissue macrophages. It therefore came

as a surprise that monocytes cultured

with the cytokines interleukin-4 (IL-4) and

granulocyte macrophage colony-stimu-

lating factor (GM-CSF) become dendritic

cells, the professional antigen-presenting

cells that initiate T cell responses in

lymphoid tissues (Sallusto and Lanzavec-

chia, 1994). These monocyte-derived

dendritic cells (Mo-DCs) capture soluble

antigens with high efficiency and respond

to microbial and inflammatory stimuli with

coordinated changes that enhance their

capacity for antigen presentation and

T cell stimulation. However, after more

than 15 years of study, the role of mono-

cytes and Mo-DCs in induction of T cell

responses in vivo remains unclear. In this

issue, Steinman and colleagues (Cheong

et al., 2010) show that in response to

injection of lipopolysaccharide (LPS) or

gram-negative bacteria, mouse mono-

cytes migrate to peripheral lymph nodes.

There they rapidly acquire the key proper-

ties of dendritic cells, such as a probing

morphology and the capacity to present

exogenous antigens to T cells that express

the cell surface markers CD4 (CD4+) and

CD8 (CD8+) (Cheong et al., 2010). These

data are compelling and the evidence

suggests that Mo-DCs have a prominent

role in initiating adaptive immunity to

gram-negative bacteria.

Two types of resident dendritic cells with

specialized functions are found in lymph

nodes and spleen (Figure 1): dendritic cells

that express CD8 and CD205 capture and

present cell-associated antigens to CD8+

T cells in association with major histocom-

patibility complex (MHC) class I mole-

cules, a mechanism known as cross-

presentation, whereas dendritic cells that

express CD11b, but not CD8 or DEC-

205, capture and present soluble antigens

to CD4+ T cells in association with MHC

class II molecules. These two subsets

develop under the influence of Flt3-L

(Fms-like tyrosine kinase 3 ligand) from

pre-dendritic cells, circulating precursors

that have lost the capacity to differentiate

along the monocyte/macrophage lineage

(Liu and Nussenzweig, 2010). Several

studies using cell transfer experiments or

reporter mice provide definitive evidence

that in the steady state monocytes do not

contribute significantly to the dendritic

cell population of lymphoid organs (Jakub-

zick et al., 2008; Naik et al., 2006).

To detect Mo-DCs, Cheong et al. used

an antibody to mouse DC-SIGN, a lectin

receptor expressed on human Mo-DCs

generated in vitro but not on classical

dendritic cells (Geijtenbeek et al., 2000).

The authors show that mouse DC-SIGN/

CD209 is expressed at low levels on fresh

monocytes and upregulated upon culture

with GM-CSF and IL-4, concomitant with

loss of the monocyte markers Ly6C and

c-fms/CD115. Using this antibody to stain

tissue sections, the authors find very few

DC-SIGN-positive cells in lymph nodes

in the steady state. Strikingly, however,

in mice challenged with LPS, large

numbers of cells expressing DC-SIGN

rapidly appear in the paracortical T cell

areas of lymph nodes (Figure 1). Direct

evidence that DC-SIGN-positive dendritic

cells are derived from monocytes comes

from experiments with mice that express

the diphtheria toxin receptor in cells

of the monocyte/macrophage lineage.

When mice are treated with diphtheria

toxin, DC-SIGN-positive cells fail to accu-

mulate in lymph nodes following LPS

challenge.

Using an ingenious in vivo labeling

approach to isolate DC-SIGN-expressing

cells from lymph nodes, the authors show

that the newly recruited Mo-DCs effi-

ciently present and cross-present to

CD4+ and CD8+ T cells soluble and cell-

associated antigens that have been taken

up in vivo. These cells are even more

potent than the two resident subsets of

dendritic cells. Interestingly, Mo-DCs

occupy a slightly different niche in the

T cell area as compared to resident

dendritic cells, suggesting that T cells

may be differentially exposed to either

cell type. Further studies using intravital

microscopy combined with interventions

to selectively deplete particular subsets

of dendritic cells will be required to define

the relative contributions of Mo-DCs to

the induction of T cell proliferation and

differentiation in vivo.

The usefulness of antibodies to surface

markers in these studies cannot be over-

emphasized, given the extensive hetero-

geneity and functional specialization of

dendritic cells. In addition to DC-SIGN/

CD209, Cheong et al. show that two other

markers can be used to identify mouse

Mo-DCs in vivo: the mannose receptor

(MMR/CD206), which is also upregulated

in human Mo-DCs, and CD14, the LPS

coreceptor, which is expressed on human

and mouse monocytes. These reagents

provide useful tools for future studies on

the role of Mo-DCs in immune response.

Cell 143, October 29, 2010 ª2010 Elsevier Inc. 339

Page 28: Cell 101029

How do monocytes and Mo-DCs reach

lymph nodes? The conventional view is

that monocytes first enter infected or in-

flamed nonlymphoid tissues where they

capture antigen,mature, and subsequently

migrate to the draining lymph nodes via the

afferent lymph (Leon et al., 2007; Randolph

et al., 1999). In contrast, Cheong et al.

report that monocytes migrate into lymph

nodes in a manner dependent on the cell

adhesion molecule CD62L and the chemo-

kine receptor CCR7, consistent with

a direct migration from blood through the

high endothelial venules (Figure 1). The

implication from these new findings is that

depending on the nature of the microbe

and its route of entry, monocytes can pref-

erentially use one or the other pathway of

migration and differentiation.

An intriguing finding of the study by

Cheong et al. is that monocyte migration

to lymph nodes and differentiation to

antigen-presenting Mo-DCs could be eli-

cited by administration of LPS or gram-

negative bacteria, but not by administra-

tion of other Toll-like receptor (TLR)

agonists or gram-positive bacteria. A

possible explanation may lie in a marked

upregulation of TLR4 and CD14 on mouse

Mo-DCs, but it is worth asking whether

this phenomenon is indeed limited to

systemic challenge with gram-negative

bacteria. Another intriguing finding is

that Mo-DCs are found only in peripheral

lymph nodes but not in spleen or mesen-

teric lymph nodes. Other studies have

shown that monocytes can be recruited

to the spleen by macrophages infected

with Listeria monoytogenes (Serbina

et al., 2003). There the monocytes

develop into inflammatory dendritic cells

that mediate protection from infection

through production of tumor necrosis

factor (TNF) and inducible nitric oxide syn-

thase (iNOS) but do not contribute to

induction of T cell responses. It is possible

that different gradients of chemokines

and cytokines elicited in different tissues

by microbial infection may determine the

recruitment of monocytes and their differ-

entiation into either inflammatory effector

cells or antigen-presenting dendritic cells.

It will be interesting to determine

whether and which cytokines trigger the

rapid differentiation of dendritic cells

from monocytes in the system used by

Cheong et al. GM-CSF is capable of

driving Mo-DC differentiation in vitro and

together with M-CSF and Flt3-L has been

shown to regulate the differentiation of

monocytes, macrophages, and dendritic

cells from bone marrow progenitors

in vivo (Schmid et al., 2010). A better

understanding of the role of GM-CSF and

other cytokines in the generation of Mo-

DCs may provide new insights into their

use as adjuvants for vaccination.

Dendritic cells are highly heterogeneous

with respect to their capacity to respond to

microbial and danger stimuli, to process

and present self and non-self antigens,

and to produce cytokines and costimula-

tory molecules that lead to different

immune responses, from Th1-, Th2-, and

Th17-mediated effector responses to

suppression and tolerance. The new study

by Cheong et al. provides conclusive

evidence that monocytes belong to the

extended dendritic cell family and intro-

duces useful tools to study the role of

Mo-DCs in these different types of immune

responses.

REFERENCES

Cheong, C., Matos, I., Choi, J.-H., Dandamudi,

D.B., Shrestha, E., Longhi, M.P., Jeffrey, K.L., An-

thony, R.M., Kluger, C., Nchinda, G., et al. (2010).

Cell 143, this issue, 416–429.

Geijtenbeek, T.B., Torensma, R., van Vliet, S.J.,

van Duijnhoven, G.C., Adema, G.J., van Kooyk,

Y., and Figdor, C.G. (2000). Cell 100, 575–585.

Jakubzick, C., Bogunovic, M., Bonito, A.J., Kuan,

E.L., Merad, M., and Randolph, G.J. (2008).

J. Exp. Med. 205, 2839–2850.

Leon, B., Lopez-Bravo, M., and Ardavin, C. (2007).

Immunity 26, 519–531.

Liu, K., and Nussenzweig, M.C. (2010). Immunol.

Rev. 234, 45–54.

Naik, S.H., Metcalf, D., van Nieuwenhuijze, A.,

Wicks, I., Wu, L., O’Keeffe, M., and Shortman, K.

(2006). Nat. Immunol. 7, 663–671.

Randolph, G.J., Inaba, K., Robbiani, D.F., Stein-

man, R.M., and Muller, W.A. (1999). Immunity 11,

753–761.

Sallusto, F., and Lanzavecchia, A. (1994). J. Exp.

Med. 179, 1109–1118.

Schmid, M.A., Kingston, D., Boddupalli, S., and

Manz, M.G. (2010). Immunol. Rev. 234, 32–44.

Serbina, N.V., Salazar-Mather, T.P., Biron, C.A.,

Kuziel, W.A., and Pamer, E.G. (2003). Immunity

19, 59–70.

Figure 1. Dendritic Cell Differentiation and Antigen Presentation in the Lymph NodeLymph node-resident dendritic cells comprise DEC-205+ and DEC-205� cells that are the progeny ofa circulating pre-dendritic cell precursor. The two types of resident dendritic cells are specialized forpresentation of antigen to CD8+ and CD4+ T cells, respectively. Upon systemic challenge with lipopolysac-charide (LPS) or gram-negative bacteria, blood monocytes enter the lymph node through the high endo-thelial venules and rapidly differentiate to dendritic cells that efficiently present antigen to CD4+ and CD8+

T cells. These cells can be identified according to the expression of DC-SIGN, MMR (macrophagemannose receptor), and CD14. Monocyte-derived dendritic cells can also migrate to the lymph nodefrom infected or inflamed tissues through the afferent lymph.

340 Cell 143, October 29, 2010 ª2010 Elsevier Inc.

Page 29: Cell 101029

Leading Edge

Previews

Ephecting Excitatory Synapse DevelopmentMatthew B. Dalva1,*1Department of Neuroscience, University of Pennsylvania School of Medicine, 421 Curie Boulevard, Philadelphia, PA 19104, USA*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.10.017

Alterations in synapse number and morphology are associated with devastating psychiatric andneurologic disorders. In this issue of Cell, Margolis et al. (2010) show that the RhoA-guanineexchange factor (GEF) Ephexin5 limits the numbers of excitatory synapses that neurons receive,thus identifying a new mechanism controlling synaptogenesis.

The anatomical and functional basis for

communication between neurons is the

synapse, a specialized site of cell-cell

contact. Synapses consist of a presyn-

aptic terminal, with neurotransmitter-filled

vesicles, and a postsynaptic terminal con-

taining receptors. Work during the past 10

years has demonstrated a significant role

for a number of trans-synaptic adhesion

proteins in the process of synapse forma-

tion (Dalva et al., 2007). Prominent among

these are the EphB family of receptor tyro-

sine kinases. EphBs are required for the

formation of normal numbers of excitatory

synapses, acting through control of filopo-

dia motility to mediate the formation of

these connections during specific devel-

opmental times (Dalva et al., 2007; Kayser

et al., 2008). Although a number of positive

regulators of synapse formation have

been described, we know less about the

factors that prevent neurons from gener-

ating too many contacts. In this issue of

Cell, an elegant and comprehensive paper

by Margolis et al. (2010) shows that the

RhoA-guanine exchange factor (GEF)

Ephexin5 (also called Vsm-Rho-GEF

[Ogita et al., 2003]) limits the synaptogenic

activity of EphB2, restricting synapse

formation. EphB2, in turn, limits Ephexin5

activity by promoting its degradation by

the E3 ligase Ube3A, relieving the restric-

tions on synapse formation. Of note,

Ube3A is the gene that is defective in the

neurogenetic cognitive disorder known

as Angelman syndrome, which strikes

about one in 10,000 live births (Dan, 2009).

Only a small fraction of the contacts

between neuronal membranes yields

anatomically definable synaptic struc-

tures, suggesting that, in addition to

mechanisms that generate synapses,

neurons must have ways to restrict

synapse formation. Known negative regu-

lators of synapse formation act through

a variety of mechanisms. For instance,

increased neuronal activity, acting

through the transcription factor MEF2

(Flavell et al., 2006), and restricted delivery

of presynaptic proteins to synaptic sites

(Patel and Shen, 2009) can each limit

synapse development. Margolis et al.

now show that the guanine exchange

factor Ephexin5 constrains synapse

formation by restricting a specific inducer

of synapse formation, EphB2 (Figure 1).

Ephexins are a family of five GEFs, of

which only Ephexin1 and Ephexin5 are

highly expressed in the brain (Sahin

et al., 2005). GEFs control GTPase activa-

tion by catalyzing the exchange of GDP

for GTP. When phosphorylated by

EphA4, Ephexin1 has potent RhoA-acti-

vating characteristics, making these

GEFs likely mediators of RhoA-depen-

dent reorganization of the actin cytoskel-

eton in the nervous system. Ephexin1

mediates ephrin-A-dependent growth

cone collapse, and mice lacking Ephexin1

have muscle weakness and impaired

synaptic transmission at the neuromus-

cular junction, likely due to malformation

of the active zone (Shamah et al., 2001;

Shi et al., 2010). However, the function

of Ephexin5 has remained obscure.

To identify candidate molecules that

might constrain the number of synapses

formed downstream of EphB2, perhaps

by inhibiting cell motility, Margolis and

colleagues first examine the pattern of

expression of a number of candidate

RhoA GEFs, finding that expression of

Ephexin5 matches the pattern of EphB

expression. Moreover, in a well-controlled

series of experiments, the authors demon-

strate that, whereas Ephexin1 interacts

selectively with EphA4, Ephexin5 interacts

selectively with EphB2 in vitro and in vivo,

has RhoA activating ability that relies on its

Dbl-homology domain, and fails to acti-

vate either rac1 or CDC-42 GTPases.

The RhoA activity in Ephexin5 knockout

mice is reduced compared with controls,

suggesting that Ephexin5 is a major deter-

minant of RhoA levels in the brain.

The authors then use a comprehensive

approach to examine the role of Ephexin5

in the control of synapse number. They

use shRNA to knock out Ephexin5 in

cultured neurons and also test synapse

formation in neurons produced from

Ephexin5 knockout mice. In both cases,

neurons lacking Ephexin5 generate more

excitatory synapses compared to

controls. In contrast, overexpression of

Ephexin5 results in a marked decrease

in the number of synapses. Importantly,

these effects depend on the guanine

nucleotide exchange activity of Ephexin5.

Then, in a clever series of experiments

using brain slices from a conditional

Ephexin5 knockout mouse, the authors

show that Ephexin5 activity also restricts

synapse formation in intact neuronal

circuits. Thus, the Ephexin5 GEF limits

the number of excitatory synapses that

neurons make in vitro and in vivo.

Margolis et al. next show that the

effects of Ephexin5 are due to a restric-

tion of EphB2 function during synapse

development. Of interest, although the

effects of Ephexin5 on synapse density

depend on EphB2 kinase activity,

EphB2 activation actually inactivates

Ephexin5 by phosphorylation of a specific

tyrosine residue, and the inactivation of

Ehpexin5 is required for EphB-depen-

dent synapse formation. These results

suggest a negative feedback loop,

whereby Ephexin5 negatively regulates

EphB2, which in turn inhibits Ephexin5

via phosphorylation.

Cell 143, October 29, 2010 ª2010 Elsevier Inc. 341

Page 30: Cell 101029

In conducting these experiments, the

authors note that the expression level of

Ephexin5 is reduced in the presence

of EphB2, raising the possibility that

Ephexin5 is regulated by proteasomal

degradation. In fact, the authors demon-

strate that proteasomal destabilization of

the Ephexin5 protein is tightly regulated

by EphB2 in vitro and in vivo. In cell lines,

the expression of EphB2 promotes

a decrease in Ephexin5 levels, and this

effect requires phosphorylation of

Ephexin5. Furthermore, a blockade of

the proteasome prevents EphB2-depen-

dent degradation of Ephexin5. In vivo,

Ephexin5 protein levels are high during

times of low synapse formation (P0-P3)

and low during periods of rapid synapse

addition (P7-P21). However, mRNA levels

of Ephexin5 remain constant throughout,

consistent with the idea that phosphoryla-

tion of Ephexin5 by EphB2 leads to

Ephexin5 degradation. Of interest,

previous reports indicate that EphB2

controls synapse formation via regulation

of filopodial motility during a similar period

of development, suggesting that changes

in Ephexin5 protein levels are the likely

mechanism in initiating or limiting these

events (Kayser et al., 2008). Finally, the

authors show that Ephexin5 is ubiquiti-

nated in brain lysates and that it interacts

with the E3 ligase Ube3A, which is

required for Ephexin5 degradation.

The link to Ube3A is noteworthy because

this E3 ligase is defective in 90% of Angel-

man syndrome cases (reviewed in Dan,

2009). In the current study, the authors

link Ephexin5 to the etiology of Angelman

syndrome using a mouse model of the

disease in which the maternal inherited

copy of Ube3A is deleted (Ube3Am�/p+).

In brains of these mice, the levels of

Ephexin5 expression and the amount of

ubiquitinated Ephexin5 protein are in-

creased. Moreover, neurons cultured

from these mice are insensitive to ephrin-

B1 treatment. In these neurons, ephrin-B1

fails to induce reduced levels of Ephexin5

expression. These results lead the authors

to suggest that the cognitive defects in An-

gelman syndrome might result from

increased levels of Ephexin5 protein.

Margolis et al. have defined a mecha-

nism that restricts that activity of a specific

synaptogenic factor in vivo and in func-

tional neuronal circuits. EphB2 initiates

synapse development by interacting with

specific presynaptic ephrin-B proteins.

Ephexin5 suppresses this activity, and

EphB2 relieves this repression by phos-

phorylating and directing Ephexin5 for

degradation by the E3 ligase Ube3A

(Figure 1). These findings cement EphBs

as a key regulator of excitatory synapse

development and suggest the interesting

possibility that other known synaptogenic

factors will have similarly selective restric-

tive mechanisms. How Ephexin5 acts

to restrict EphB2-dependent synapse

formation remains unknown, but con-

sidering that RhoA activation typically

suppresses cell motility, these findings

suggest that Ephexin5 might limit EphB2

function during synapse formation by

downregulating the motility of dendritic

filopodia that EphB2 has previously been

shown to mediate. The authors suggest

that this may be the case by indicating

that Ephexin5 may limit filopodial motility

in preliminary unpublished work. Beyond

its impact on understanding synapse

development, the study provides a tanta-

lizing and exciting potential mechanism

to explain the cognitive and behavioral

defects in patients with Angelman

syndrome.

ACKNOWLEDGMENTS

NIDA, the NIMH, and the Dana Foundation support

M.B.D.’s work.

REFERENCES

Dalva, M.B., McClelland, A.C., and Kayser, M.S.

(2007). Nat. Rev. Neurosci. 8, 206–220.

Dan, B. (2009). Epilepsia 50, 2331–2339.

Flavell, S.W., Cowan, C.W., Kim, T.K., Greer, P.L.,

Lin, Y., Paradis, S., Griffith, E.C., Hu, L.S., Chen,

C., and Greenberg, M.E. (2006). Science 311,

1008–1012.

Kayser, M.S., Nolt, M.J., and Dalva, M.B. (2008).

Neuron 59, 56–69.

Margolis, S.S., Salogiannis, J., Lipton, D.M., Man-

del-Brehm, C., Wills, Z.P., Mardinly, A.R., Hu, L.,

Greer, P.L., Bikoff, J.B., Ho, H.-Y.H., et al. (2010).

Cell 143, this issue, 442–455.

Ogita, H., Kunimoto, S., Kamioka, Y., Sawa, H.,

Masuda, M., and Mochizuki, N. (2003). Circ. Res.

93, 23–31.

Patel, M.R., and Shen, K. (2009). Science 323,

1500–1503.

Sahin, M., Greer, P.L., Lin, M.Z., Poucher, H., Eber-

hart, J., Schmidt, S., Wright, T.M., Shamah, S.M.,

O’connell, S., Cowan, C.W., et al. (2005). Neuron

46, 191–204.

Shamah, S.M., Lin, M.Z., Goldberg, J.L., Estrach,

S., Sahin, M., Hu, L., Bazalakova, M., Neve, R.L.,

Corfas, G., Debant, A., and Greenberg, M.E.

(2001). Cell 105, 233–244.

Shi, L., Butt, B., Ip, F.C., Dai, Y., Jiang, L., Yung,

W.H., Greenberg, M.E., Fu, A.K., and Ip, N.Y.

(2010). Neuron 65, 204–216.

Figure 1. Ephexin5 Represses Synapse DevelopmentMargolis et al. (2010) show that the guanine exchange factor (GEF) Ephexin5 inhibits synapse formation byactivating RhoA prior to the activation of the EphB2 receptor by its ephrin-B ligands (left). Once engagedby ligand, EphB2 promotes Ephexin5 phosphorylation, leading to its ubiquitination and degradation by theE3 ubiquitin ligase Ube3A (center). EphB2 can then coordinate synapse maturation by interacting withpresynaptic ephrin-Bs, regulating the maturation of dendritic spines and recruiting glutamate receptors(AMPA receptor and NMDA receptor) to the synapse (right).

342 Cell 143, October 29, 2010 ª2010 Elsevier Inc.

Page 31: Cell 101029

Leading Edge

Review

Chemoaffinity Revisited: Dscams,Protocadherins, andNeural Circuit AssemblyS. Lawrence Zipursky1,* and Joshua R. Sanes2,*1Department of Biological Chemistry, Howard Hughes Medical Institute, David Geffen School of Medicine, University of California,Los Angeles, Los Angeles, CA 90095, USA2Center for Brain Science and Department of Molecular and Cellular Biology, Harvard University, Cambridge, MA 02138, USA

*Correspondence: [email protected] (S.L.Z.), [email protected] (J.R.S.)

DOI 10.1016/j.cell.2010.10.009

The chemoaffinity hypothesis for neural circuit assembly posits that axons and their targets bearmatching molecular labels that endow neurons with unique identities and specify synapsesbetween appropriate partners. Here, we focus on two intriguing candidates for fulfilling this role,Drosophila Dscams and vertebrate clustered protocadherins (Pcdhs). In each, a complex genomiclocus encodes large numbers of neuronal transmembrane proteins with homophilic binding spec-ificity, individual members of which are expressed combinatorially. Although these propertiessuggest that Dscams and Pcdhs could act as specificity molecules, they may do so in ways thatchallenge traditional views of how neural circuits assemble.

IntroductionTwo experiments have had a decisive influence on our ideas

about how neurons form the complex patterns of synaptic

connections that underlie mental activities. Both were performed

long ago, relied on simple behavioral assays, didn’t involve mole-

cules, and focused on regeneration following nerve injury in

adults rather than development.

In the first, John Langley (Langley, 1895) analyzed regenera-

tion in the autonomic nervous system of a cat. He had found

that axons from multiple levels of the spinal cord enter the supe-

rior cervical ganglion through a common nerve and connect with

neurons that then innervate distinct peripheral organs. For

example, sympathetic neurons innervated by axons from the first

thoracic segment controlled pupil dilation, those innervated from

the next segment controlled vasoconstriction of the ear, and so

on. Langley cut the nerve, awaited regeneration, and asked

whether ‘‘the fibres of each spinal nerve become connected

with only those nerve cells with which they are normally con-

nected, or will they run indiscriminately to such cells as may be

on their course.?’’ The answer was clearly the former: even

though axons entered the ganglion together and encountered

intermixed targets, they formed functionally appropriate connec-

tions.

In a second and more extensive series of experiments, Roger

Sperry (Sperry, 1943, 1944, 1963) cut the optic nerves of

amphibia (newts, toads, and frogs), then assessed the return of

visual function following regeneration. (Central axons regenerate

poorly in mammals but well in lower vertebrates.) The lens casts

an image of the world on the retina and this image is then pro-

cessed and transmitted through the optic nerve to form topo-

graphic maps in central nuclei. In fact, useful vision was restored,

implying that regenerated axons had formed proper connec-

tions. Most dramatically, when the eye was rotated, orderly but

counterproductive vision was restored: the animal behaved as

if it saw the world upside-down and reversed. The clear implica-

tion was that retinal axons had reconnected with their original

synaptic targets in the brain, not the targets that would now

make functional sense. Sperry went on to perform physiological

and anatomical experiments that provided definitive support for

this view (reviewed in Sperry, 1963).

Langley and Sperry drew similar conclusions. Langley (1895)

reasoned that there must be ‘‘some special chemical relation

between each class of nerve fibre and each class of nerve cell,

which induces each fibre to grow towards a cell of its own class

and there to form its terminal branches.’’ Sperry (1944) hypothe-

sized that ‘‘the ingrowing optic fibers must possess specific

properties of some sort by which they are differentially distin-

guished.[and]. neurons of the optic tectum are also biochemi-

cally dissimilar, possessing differential affinities for fibers arising

from different retinal quadrants.’’ Moreover, both realized that

the recognition was likely to involve interactions along the path

that axons follow as they grow toward their targets as well as

at the target itself—processes now called axon guidance and

target selection, respectively (Langley, 1895; Sperry, 1963).

In retrospect, we see that the power of these experiments

came from analyzing regeneration in adults rather than develop-

ment in embryos. The studies were initially criticized as having

limited relevance to how the nervous system wires up as it forms.

Yet, during regeneration, confounding factors associated with

normal developmental processes, such as precisely timed

generation of neurons, orderly arrival of axons, and limitations

of spatial access, were eliminated. Sperry’s eye rotation experi-

ment even eliminated activity and experience as instructive

factors. That is not to say that such factors have no role; indeed

Cell 143, October 29, 2010 ª2010 Elsevier Inc. 343

Page 32: Cell 101029

it is now clear that specificity arises from a combination of all of

these processes and more (Sanes and Yamagata, 2009; Sanes

and Zipursky, 2010). Nonetheless, the work of Langley and

Sperry led to a molecular view that remains largely unchallenged:

neurons must be chemically specified in ways that guide them to

and promote synapse formation with appropriate targets.

MoleculesFollowing further experiments, Sperry (1963) formalized the

chemoaffinity hypothesis, stating that neurons bear ‘‘individual

identification tags. [with] each axon linking only with certain

neurons to which it becomes selectively attached by specific

chemical affinities.’’ He believed this individualization could

require ‘‘millions, and possibly billions, of chemically differenti-

ated neuron types.’’ What sort of molecules might do the trick?

Three general possibilities have been suggested.

One is that the differences might be quantitative rather than

qualitative with neurons being specified by molecular gradients

of adhesive molecules encoding ‘‘matching values between

the retinal and tectal maps’’ (Sperry, 1963). Later, Gierer (1983)

formalized the model. Based on these ideas, intensive efforts

were made to isolate such ‘‘gradient molecules’’; eventually

Bonhoeffer and others showed that complementary gradients

of Eph kinases in retina and their ligands, ephrins, in tectum do

indeed play critical roles in establishment of the retinotectal

and other topographically organized maps (Drescher et al.,

1995; Cheng et al., 1995; McLaughlin and O’Leary, 2005). It is

less obvious, however, that gradients could endow axons with

the ability to distinguish among neuronal types that are physi-

cally intermingled rather than spatially arrayed. For example,

the specificity required to form microcircuits within the retina or

cortex, or connections within invertebrate ganglia, may require

qualitatively distinct molecular tags.

A second possibility is that diversity arises from the combined

action of many unrelated molecules that act in different ways.

Indeed, axons are guided to their targets by a combination of

short-range (contact-mediated) and long-range (diffusible)

cues that act as both attractants and repellents. Many such guid-

ance molecules and receptors have been identified—ephrins,

semaphorins, netrins, plexins, robos, slits, and so on (Dickson,

2002). Most of them turn out to be products of gene families of

small or moderate size (up to �20 for semaphorins). Studies of

synaptic specificity suggest that the same mechanisms and, in

some cases, the very same molecules act in this process. In

the few cases of target recognition that have yielded to analysis,

synaptic specificity results from soluble, membrane-bound and

matrix-associated proteins of multiple families that act on

multiple cell types as both attractants and repellents (Sanes

and Yamagata, 2009). This hybrid strategy may seem inelegant,

but that does not make it implausible. In fact Jacob (1977) and

others have argued that this is how evolution works—as

a tinkerer, cobbling together whatever variety of mechanisms

are already available as products of prior evolution, not as an

engineer, prospectively designing a maximally efficient solution.

Finally, a particularly attractive scenario is that multigene fami-

lies of adhesion molecules with distinct binding specificities are

differentially expressed among neurons of a population and

thereby stamp each individual with a distinct identity. This idea

was formalized under names such as ‘‘area code hypothesis’’

(Dreyer, 1998). During the 1990s, three families were proposed

to play this role: the classical and type II cadherins (�20 genes;

Takeichi, 2007), the neurexins and neuroligins (3–4 genes each,

but a far larger number of alternatively spliced isoforms; Sudhof,

2008), and the olfactory receptors, a group of �1000 G protein-

coupled receptors expressed by olfactory sensory neurons

(Buck and Axel, 1991).

Cadherins, neurexins, and neuroligins have turned out to be

critical players in neural development, but to date there is little

evidence that they act as determinants of synaptic specificity.

The olfactory receptors, in contrast, are clearly required for the

precise targeting of olfactory sensory axons to glomerular

targets in the olfactory bulb. Each neuron expresses just one

of the thousand receptors, and all neurons expressing the

same receptor send axons to a single pair of glomeruli in the

olfactory bulb. If a receptor is deleted, neurons that would

have expressed it innervate the bulb diffusely. When one

receptor is genetically replaced by another, the axon is retar-

geted to an ectopic location, which often corresponds to the

proper target of neurons that endogenously expressed that

receptor (Mombaerts et al., 1996; Mombaerts, 2006). These

‘‘receptor swap’’ experiments demonstrated an instructive role

for olfactory receptors in circuit assembly and led to the specu-

lation that they recognized complementary nonodorant ligands

expressed by targets in the bulb. It now seems likely, however,

that these receptors act not by interacting with targets directly

but rather by differentially modulating levels of intracellular

messengers in a ligand-independent fashion; the messengers,

in turn, regulate expression of more conventional axon guidance

molecules (Sakano, 2010). Thus, olfactory receptors are

determinants of specificity, but surprisingly, they act in a rather

indirect way.

Are there, then, large families of cell-cell recognition molecules

that specify assembly of neural circuits? Over the past few years,

two families, the Dscams in insects and the clustered protocad-

herins (Pcdhs) in vertebrates, have emerged as promising candi-

dates. In both cases, complex genomic loci encode large sets

of proteins that are expressed in combinatorial patterns by indi-

vidual neurons, mediate homophilic binding, and play critical

roles in neural development. In the next sections of this Review,

we summarize these recent findings. The results lead us to argue

that Dscams and Pcdhs are not transsynaptic ‘‘chemoaffinity’’

molecules in the sense that has generally been envisioned.

Instead, they contribute to neural specificity in unexpected

ways, suggesting a new view of how large families of cell-surface

molecules contribute to circuit assembly.

DscamsDscam proteins are highly conserved single-pass transmem-

brane domain proteins of the immunoglobulin (Ig) superfamily

(Hattori et al., 2008). Fly Dscam1 was identified through

a biochemical interaction of its C-terminal cytoplasmic domain

with an adaptor protein, Dock, previously shown to function in

growth cones during axon guidance. It comprises 10 Ig domains,

6 fibronectin type III repeats, a single transmembrane domain,

and a C-terminal cytoplasmic tail. Like the adaptor, Dscam1 is

widely expressed in the developing nervous system and

344 Cell 143, October 29, 2010 ª2010 Elsevier Inc.

Page 33: Cell 101029

essential for guidance of a subclass of axons (Schmucker et al.,

2000). Dscam1 thus joined a large group of immunoglobulin

superfamily molecules (Shapiro et al., 2007) known to function

as receptors for transmembrane and soluble molecules that

guide axons to their targets.

Sequence analysis of Dscam1 cDNAs and the genomic

locus revealed a feature that set it apart from other neuronal Ig

superfamily members, including Drosophila Dscam2–4 and the

vertebrate Dscams: its primary transcript is subject to massive

alternative splicing (Figure 1A). The Dscam1 gene in Drosophila

and other arthropods contains four blocks of tandemly arranged

alternative exons. In Drosophila, these encode 12, 48, 33, and 2

variations on Ig2, Ig3, Ig7, and the transmembrane domain,

respectively (although the same domains come in alternative

flavors in other species, the number of variants and their

sequences are highly variable). Any individual mature mRNA

contains just one exon from each block. As splicing at each block

is independent of the other three, the Dscam1 locus has the

potential to encode 19,008 ectodomains (12 3 48 3 33) tethered

to the membrane by one of two alternative transmembrane

segments.

A first clue to the mechanisms by which Dscam1 functions

came from studies of the mushroom body (MB), a central brain

structure involved in learning and memory. Each MB contains

some 2500 neurons; their axons bifurcate at a common branch

point, and the resulting sister branches then segregate to two

different pathways (Figure 2A). Removing Dscam1 from all MB

neurons led to massive disruption, but removing Dscam1 from

Figure 1. Dscam1 Gene and Proteins(A) The Drosophila Dscam1 gene contains groups of alternative exons that encode 12 different variants for the N-terminal half of Ig2 (purple), 48 different variantsfor the N-terminal half of Ig3 (orange), and 33 different variants for Ig7 (blue), as well as two different variants for the transmembrane domain (TM) (brown). Splicingleads to the incorporation of one alternative exon from each group, and as such, Dscam1 encodes 19,008 (i.e., 12 3 48 3 33) different ectodomains.(B) Results of adhesion assays in which Dscams with each of the 12 Ig2 variants were tested for binding to each other. The Ig3 and Ig7 variants were held constant.Each isoform binds to itself but rarely, if at all, to other isoforms. The numbers indicate the alternative Ig2 domain and are arranged as they would be in adendrogram, such that those closest to each on the grid are closest to each other in sequence. Inset shows binding represented as fold over background(BKGD) (Adapted from Wojtowicz et al., 2007).(C) Summary of results in (B). Homophilic binding occurs between identical isoforms that match at all three variable Ig domains. Isoform pairs that contain only twomatches and differ at the third variable domain bind poorly or not at all to one another. This summarizes the results for Ig2; the properties of the other variabledomains are analogous.(D) Dscam1 monomers have a rigid horseshoe-shaped amino terminus (Ig1–4) and a flexible tail. Dimerization leads to a large conformational change, resulting ina complex of two S-shaped monomers with direct contacts between opposing Ig2, Ig3, and Ig7 variable domains.

Cell 143, October 29, 2010 ª2010 Elsevier Inc. 345

Page 34: Cell 101029

single MB neurons in an otherwise wild-type background gave

a more interpretable result: the two sister branches of the mutant

neuron formed but frequently failed to segregate to the two

different pathways (Wang et al., 2002a). This finding, together

with biochemical and expression studies described below, led

to the notion that homophilic binding of Dscam1 proteins on

sister branches from the same cell promotes repulsive interac-

tions between them, thus ensuring that they diverge and grow

along separate pathways (Zhan et al., 2004; Wojtowicz et al.,

2004). Dscam1 proteins also promote repulsion between

dendrites of the same cell (Zhu et al., 2006; Matthews et al.,

2007; Hughes et al., 2007; Soba et al., 2007). This process is

best characterized in the dendrites of sensory ‘‘dendritic arbori-

zation’’ or da neurons (Figure 2B). The da dendrites elaborate

highly branched sensory endings in the body wall. As dendrites

arborize in a narrow plane, one might expect that dendrites

from the same cell would frequently encounter and cross over

one another, but such self-crossing seldom occurs. In the

absence of Dscam1, however, self-dendrites cross frequently.

Thus, as in MB neurons, Dscam1 promotes the repulsion of

processes of the same cell. This selective repulsion between

dendrites of the same cell promotes uniform coverage of a

receptive field while allowing processes of different neurons to

share the field.

Recently, Millard et al. (2010) found that Dscam-mediated

repulsive interactions among prospective postsynaptic

elements also contribute to synaptic specificity (Figure 2C). In

vertebrates, typical synapses comprise a presynaptic terminal

and a single postsynaptic element. In flies, however, the majority

of synapses are multiple contact synapses with a single presyn-

aptic site releasing neurotransmitter onto 2–5 postsynaptic

elements. The best characterized of these are so-called tetrad

synapses between presynaptic terminals of photoreceptor

neurons and postsynaptic elements of lamina neurons (Sanes

and Zipursky, 2010). Each photoreceptor axon makes some 50

tetrad synapses, with each tetrad containing two invariant

elements, one each from an L1 and an L2 neuron; all 50 tetrads

comprise postsynaptic elements from the same two cells.

Dscam1 acts in a redundant fashion with its paralog, Dscam2,

to control tetrad composition. In the absence of both Dscam1

and Dscam2, the invariant pairing breaks down with many

tetrads comprising two L1 or two L2 branches rather than one

of each. This phenotype led to a model in which Dscams provide

L1 and L2 neurites with the ability to distinguish between self and

non-self, thus preventing them from providing two elements to

a single tetrad.

Together, these results suggest a common theme to Dscam1

function in multiple aspects of neural circuit assembly: it medi-

ates self-recognition among neurites of a single cell followed

by their repulsion from each other. This process was originally

observed in leech neurons and termed self-avoidance (Kramer

and Kuwada, 1983; Kramer and Stent, 1985). Kramer and

colleagues emphasized that self-avoidance was important

because it could promote uniform coverage of receptive or

projective fields by individual neurons, while allowing multiple

neurons to share the same field. More recent studies of Dscams

show that self-avoidance can also affect axonal pathfinding

and synaptic connectivity. Nonetheless, the phenomenon of

self-avoidance was little-studied over the subsequent two

decades, perhaps because it was so difficult to envision molec-

ular mechanisms that could allow a neurite to distinguish other

neurites of the same cell from neurites of seemingly identical

cells—in other words, the problem of distinguishing self from

non-self. The chemoaffinity hypothesis provided a framework

for seeking molecules that mediate specific intercellular interac-

tions, but there was no corresponding framework for under-

standing selective interactions among neurites of a single cell.

Figure 2. Multiple Roles of Dscam1 and 2 in Neural Development(A) Dscam1 mediates self-avoidance in axons of mushroom body (MB)neurons. Each of the MB neurons (2 of 2500 are shown) extends a singleaxon that bifurcates and sends one branch medially and the other dorsally.Each MB neuron expresses a unique combination of isoforms. As a conse-quence, sister branches recognize each other through Dscam1 matching.This signals repulsion and subsequent segregation of axons to separate path-ways. When Dscam1 is removed from a single MB neuron, its branches oftenfail to segregate.(B) Dscam1 mediates self-avoidance in dendrites of da sensory neurons.Dendrites of each neuron are splayed out but can cross dendrites of otherda neurons. As the dendrites extend on a flat surface, crossing is associatedwith direct contact between arbors. Deletion of Dscam1 from a single daneuron leads to disordered arbors in which dendrites from the same cell some-times fasciculate or cross each other. Reducing Dscam1 diversity in all daneurons leads to segregation of their dendrites from each other.(C) Dscam1 and Dscam2 act redundantly to pattern synapses of photore-ceptor (R) axons on L1 and L2 dendrites in the lamina. In each cartridge, Raxons form tetrad synapses in which postsynaptic partners always includeone L1 and one L2; the other pair comprises combinations of elements fromother cell types. They lie above and below the L1/L2 pair (not shown). TheT-bar is a presynaptic structural specialization. In the absence of Dscam1and 2, the repulsion between prospective postsynaptic elements of L1s andbetween L2s is lost, so some tetrads include two elements from the sameL1 or same L2 cells.

346 Cell 143, October 29, 2010 ª2010 Elsevier Inc.

Page 35: Cell 101029

We will see below that analysis of Dscam1 has provided a way to

understand this process.

ProtocadherinsThere are two Dscam genes in vertebrates (Dscam and

DscamL). Early analysis indicates that they promote both

class-specific avoidance and transsynaptic target recognition

in the restricted subsets of retinal neurons that express them

(Fuerst et al., 2008, 2009; Yamagata and Sanes, 2008). However,

these are garden-variety genes with few alternatively spliced

isoforms, more like fly Dscam2–4 than Dscam1. So, they are

unlikely to promote diversity in the way that fly Dscam1 does.

However, another set of genes, the clustered protocadherins

(Pcdhs; Morishita and Yagi, 2007), show intriguing similarities

to fly Dscam1, raising the possibility that they play analogous

roles.

In 1998, T. Yagi and colleagues reported identification of a

group of eight homologous transmembrane proteins that they

called ‘‘cadherin-related neuronal receptors’’ or CNRs (Kohmura

et al., 1998). CNRs were fascinating for several reasons. First,

their ectodomains placed them squarely within the cadherin

superfamily, many other members of which had been implicated

in numerous developmental processes (Takeichi, 2007).

Second, their expression was largely restricted to the nervous

system. Third, immunohistochemical studies showed that they

were concentrated at synaptic sites. Finally, sequences of the

eight CNRs indicated that they had related ectodomains but

identical cytoplasmic domains, suggesting their coexistence

in a genomic cluster.

Shortly thereafter, Wu and Maniatis (1999) found that the

CNRs are derived from a large genomic locus that encodes

a total of >50 genes (58 in mice; Figure 3A) now called clustered

protocadherins. (Several other distantly related protocadherins

reside at other genomic loci; they are members of the cadherin

superfamily generally, but their expression and roles seem

quite distinct from those of the clustered protocadherins.) Exons

encoding complete extracellular and transmembrane domains

are arranged in three groups called Pcdh-a, Pcdh-b, and

Pcdh-g, with 14, 22, and 22 members in the mouse genome,

respectively. For the Pcdh-a and -g clusters, each ectodo-

main-encoding exon is joined to 3 invariant (constant) exons

that encode their common cytoplasmic domain. The Pcdh-b vari-

able exons, which have been less studied to date, appear to

encode complete proteins with short cytoplasmic domains;

this locus has no constant exons. The cytoplasmic domains of

the clustered Pcdhs differ from each other and all lack the canon-

ical catenin-binding domains present in classical cadherins. Like

the Pcdh-as, the Pcdh-b and -g genes are expressed primarily in

the nervous system, and their protein products are concentrated

at, but not restricted to, synaptic sites (Wang et al., 2002c;

Phillips et al., 2003; Junghans et al., 2008).

Kohmura et al. (1998) and Wu and Maniatis (1999) envisioned

several strategies by which Pcdh proteins could be generated

from Pcdh-a and -g genes: by genomic rearrangement as occurs

in the T cell receptor and immunoglobulin loci, by alternative

splicing of a large pre-mRNA, as occurs in Drosophila Dscam1,

or by alternative use of separate promoters upstream of each

first exon. The third alternative is now known to be the correct

one (Tasic et al., 2002; Wang et al., 2002b). Each exon is

preceded by a promoter and produces a transcript in which

the first exon is spliced to the common exons. Pcdh proteins

then interact with other products of the cluster to form hetero-

multimers (Murata et al., 2004; Schreiner and Weiner, 2010).

Thus, many Pcdh proteins, like Dscams, are generated from

a single genomic locus, though the methods of achieving this

diversity are fundamentally different.

Figure 3. The Protocadherin Gene Cluster

and Its Protein Products(A) The Pcdh gene cluster contains exons thatencode 58 extracellular and transmembranedomains—14 in the a group (purple) and 22 eachin the b (orange) and g (blue) groups. Each ectodo-main contains 6 cadherin repeats. Ectodcomainsare more related to others within a group than tothose in other groups with the exception of aC1,aC2, and gC3–5 domains (asterisks), which aremore closely related to each other than to neigh-boring members within their group. Each ectodo-main is preceded by a promoter. Alternativesplicing joins an a or g ectodomain/transmem-brane exon to three constant exons in the group.b exons encode complete proteins with shortintracellular domains.(B) Results of adhesion assays in which each ofseven Pcdh-gs was tested for binding to threeisoforms. Each isoform bound preferentially toitself (redrawn from data in Schreiner and Weiner,2010).(C) Cadherin domains EC2 and EC3 mediatethe specificity of homophilic binding between iso-forms (redrawn from data in Schreiner and Weiner,2010).(D) Crystal structures of the EC1 domain of Pcdh-aand immunoglobulin domain 7 of Dscam showingthe overall similarity of the b sandwich structureof Ig and cadherin repeats.

Cell 143, October 29, 2010 ª2010 Elsevier Inc. 347

Page 36: Cell 101029

Functions of the clustered a and g Pcdhs have been investi-

gated in targeted mouse mutants. Mice lacking Pcdh-as are

viable and fertile but display subtle neural defects. Perhaps

most interesting is a projection error of olfactory sensory neu-

rons. In wild-type animals, axons of olfactory sensory neurons

that express the same olfactory receptor converge to innervate

a few glomeruli, usually one on each side of the olfactory bulb.

In the Pcdh-a mutants, sorting is incomplete, and axons ex-

pressing the same receptor end up forming several small

supernumerary glomeruli (Hasegawa et al., 2008). Likewise,

serotonergic fibers are aberrantly distributed in the brains of

Pcdh-a mutants (Katori et al., 2009). Interestingly, the glomerular

defects in Pcdh-a mutants show parallels with those observed in

the olfactory system of fly Dscam1 mutants (Hummel et al.,

2003). These results suggest that Pcdhs play roles in axon guid-

ance or targeting.

Loss of Pcdh-gs, in contrast, leads to devastating neurological

defects and neonatal lethality (Wang et al., 2002c). At a cellular

level, the most striking phenotype is apoptosis of a substantial

fraction of many neuronal subtypes (Wang et al., 2002c; Prasad

et al., 2008; Lefebvre et al., 2008; Su et al., 2010). Death occurs

during the period of naturally occurring cell death (Prasad et al.,

2008; Lefebvre et al., 2008) and appears to be an accentuation of

this process. It is observed in many areas and neuronal popula-

tions but is not ubiquitous—for example, some neuronal

subtypes are spared in retina and spinal cord, and little loss is

seen in cerebral cortex, cerebellum, and hippocampus (Wang

et al., 2002c; Lefebvre et al., 2008).

The number of synapses is also decreased in Pcdh-gmutants,

but this could be an indirect consequence of decreased neuron

number. To test this possibility, the cell death phenotype was

largely eliminated by deleting the proapoptotic gene Bax. Effects

of this manipulation differed between spinal cord and retina:

synapse number remained depressed in the former but not in

the latter (Weiner et al., 2005; Lefebvre et al., 2008). Moreover,

Pcdh-gs appeared to be dispensable for synaptic function and

specificity in retina, as electrophysiological recordings indicate

that complex computation of visual features can occur in their

absence (Lefebvre et al., 2008). Thus, Pcdh-gs may be directly

required for synapse formation or maintenance in some but not

all regions of the nervous system.

In summary, molecular and genetic studies have revealed that

Dscams and Pcdhs are critical for assembly of neural circuits.

But do they endow individual neurons with unique identities

required to wire up correctly? For this hypothesis to be taken

seriously, one would need to demonstrate (1) that individual

neurons express distinct sets of Dscams and Pcdhs, (2) that

the proteins mediate highly specific intercellular interactions,

and (3) that their diversity is required for their function. Recent

evidence supports all three of these conditions for Dscams

and the first two for Pcdhs.

Combinatorial and Stochastic ExpressionDscam1

The Dscam1 gene encodes 19,008 different ectodomain iso-

forms. How many are actually expressed, and what cells express

them? Sequence analysis of cDNAs prepared from various

developmental stages and neuronal subpopulations revealed

that all but a single alternative exon were found in mRNA and

most were present in multiple populations. More recently,

high-throughput sequencing of some 3 million cDNAs from

whole animals indicated that more than 17,000 potential combi-

nations of isoforms are indeed expressed (B. Graveley, personal

communication).

To gain insight into patterns of isoform expression, Chess and

colleagues analyzed cDNAs prepared from purified neuronal

subtypes or from single neurons (Neves et al., 2004; Zhan

et al., 2004). Little specificity was found in the expression of

alternative exons encoding Ig2 and Ig3, although there were

cell type-specific biases in the utilization of exons encoding

Ig7. Experimental results and an independent statistical analysis

generated the estimate that a single neuron expresses 10–50

isoforms. Although it remains unknown whether all mRNAs are

translated into proteins, these studies provide strong evidence

that Dscam1 isoforms are expressed in a biased stochastic

fashion. Thus, as a consequence of alternative splicing and

combinatorial expression, Dscam1 appears to endow each

Drosophila neuron with a unique molecular identity.

Pcdhs

With few isoform-specific antibodies available, expression of

individual Pcdh isoforms has been analyzed primarily by in situ

hybridization and RT-PCR. Most isoforms are broadly expressed

throughout the developing and adult nervous systems, although

expression levels vary among isoforms and with age. Expression

patterns also vary among isoforms, and some exhibit interesting

concentrations in particular laminae or cell types, but the overall

impression is one of overlapping rather than mutually exclusive

expression at the regional level (Zou et al., 2007; Junghans

et al., 2008). Likewise, at the cellular level, double labeling for

any two isoforms shows partial overlap (Kohmura et al., 1998;

Wang et al., 2002c).

Single-cell RT-PCR analysis of Purkinje cells, chosen because

they are large and relatively uniform, provided strong evidence

for stochastic, combinatorial expression of Pcdhs in individual

cells (Esumi et al., 2005; Kaneko et al., 2006). Each Purkinje

neuron expressed 1–3 of the first 12 (that is, 50) Pcdh-a isoforms

and 1–3 of the first 19 Pcdh-g isoforms. In most cases, expres-

sion was monoallelic. There was no obvious relationship

between the Pcdh-a and Pcdh-g isoforms that a Purkinje cell

expressed. The 30 members of each cluster—the final 2 Pcdh-as

and the final 3 Pcdh-gs—exhibited a different pattern. It had

already been noted that these 5 isoforms were more closely

related by sequence to each other than to neighboring members

within their group, and they had been called ‘‘C’’ isoforms (Pcdh-

aC1-2 and Pcdh-gC3-5) in recognition of this relationship. All 5 C

isoforms were expressed biallelically by Purkinje neurons.

Although limited to Purkinje cells, these data allow estimation

of the number of distinct identities that Pcdh-a and -g expression

could confer on neurons. Assuming each cell has the potential to

express 1–3 isoforms each of Pcdh-a and -g, there are some

350,000 possible combinations. Expression of Pcdh-bs consid-

erably increases the number of combinations. These proteins

may function in complexes: Pcdh-as and Pcdh-gs form hetero-

multimers with no detectable isoform specificity, Pcdh-gs facili-

tate transport of Pcdh-as to the cell surface (Murata et al., 2004)

and Pcdh-bs associate with a and g Pcdhs (Han et al., 2010). It is

348 Cell 143, October 29, 2010 ª2010 Elsevier Inc.

Page 37: Cell 101029

interesting, though probably coincidental, that if Pcdhs form

complexes comprising one of each subfamily, the number of

possible combinations (14 3 22 3 22) is similar to that generated

by independent inclusion of alternative exons in Dscam1 (12 3

48 3 33).

Dscams and Pcdhs Exhibit Isoform-SpecificHomophilic BindingDscams

A comprehensive set of binding studies revealed that different

Dscam isoforms exhibit an unprecedented range of homophilic

adhesive specificities. Wojtowicz et al. (2004, 2007) assayed

recombinant proteins containing each of the 12 alternative Ig2s

in the context of constant Ig3 and 7 (Figure 1B), each of the 48

Ig3s in the context of constant Ig2 and 7, and each of the 33

Ig7s in the context of constant Ig2 and 3. In nearly all cases,

any individual Dscam isoform bound far better to other proteins

of the same isoform than to other isoforms, even when the differ-

ences between them were small. The very few cases of hetero-

philic interactions occurred between highly related isoforms.

Thus, Dscams show isoform-specific homophilic binding that

relies on the matching of all three variable Ig domains (Figure 1C).

Based on these studies, it was predicted that the Dscam locus

encodes some 18,024 isoforms with isoform-specific homophilic

binding (12 3 47 3 32, because one Ig3 variant is not expressed

and one Ig7 variant fails to bind).

Two X-ray structures of the Dscam1 Ig domains provided

insight into the structural basis for this remarkable binding

specificity (Meijers et al., 2007; Sawaya et al., 2008). The eight

N-terminal Ig domains form a two-fold symmetric double

S-shaped dimer (Figure 1D). The three variable domain inter-

faces comprise the majority of contacts between the two mole-

cules. Each interface is formed by pairing of a polypeptide strand

with the same strand in the opposing molecule in an antiparallel

fashion, with binding specificity being determined by the shape

and charge complementarity of the interface surfaces. The two

sharp turns within the double S-shaped structure, between Ig2

and Ig3 and between Ig5 and Ig6, facilitate the matching of the

variable domains in the two opposing molecules. The comple-

mentary surfaces of each variable domain fit together like

children’s blocks.

The structural analysis also provided a way to understand

why matching of all three variable domains is required for

binding. The Ig2 and Ig3 interfaces are intramolecularly con-

strained, so a mismatch in either one disrupts matching at

the other. Similarly, the four strands at the Ig7 interface are

internally constrained, so mismatching between any one

prevents the formation of the interface between the others.

An intramolecular interface between Ig5 and Ig6 is also crucial

for homophilic binding. This interface stabilizes the large con-

formational change that forms the double S shape on dimeriza-

tion, thereby bringing the Ig2-Ig3 and Ig7 interfaces together.

Thus, the combined interactions at four interfaces (Ig2-Ig2,

Ig3-Ig3, Ig7-Ig7, and Ig5-Ig6) lead to all-or-none binding spec-

ificity. The conformational change may also initiate the signal

transduction process that converts initial homophilic binding

into the repulsive response that mediates self-avoidance of

sister neurites.

Pcdhs

For many years, attempts to assay adhesive interactions among

Pcdh proteins gave equivocal results (Morishita and Yagi, 2007).

Very recently, however, Schreiner and Weiner (2010) showed

that Pcdh-gs exhibit isoform-specific homophilic binding. They

used a novel, quantitative cell adhesion assay to analyze 7 of

the 22 different Pcdh-g isoforms. Each isoform exhibited homo-

philic binding activity when transfected into cells devoid of

endogenous classical cadherins and protocadherins (Figure 3B).

Binding specificity was highly reminiscent of the strict isoform-

specific homophilic binding exhibited by Dscam1 isoforms.

To explore the molecular basis for this specificity, Schreiner

and Weiner asked which of the six Pcdh-g cadherin domains

(EC1–EC6) were required for homophilic binding (Figure 3C).

Mutations in EC1 domains disrupted homophilic binding, but

swapping EC1 domains between different isoforms did not alter

binding specificity. In this respect, protocadherins differ from

classical cadherins, in which EC1 is required not only for binding

per se but also for isoform specificity (Morishita et al., 2006;

Shapiro et al., 2007). Additional domain swaps revealed that

both EC2 and EC3 domains contain binding specificity determi-

nants (Figure 3C). Moreover, some chimeras unable to bind

either parent were able to bind homophilically; the generation

of novel specificities was also a feature of Dscam swaps (Wojto-

wicz et al., 2007). These findings establish that EC1 is required

for binding but not specificity, whereas EC2 and EC3 provide

the specificity determinants. Pairing of matched EC2 and EC3

domains from Pcdh molecules on opposing membranes might

occur by a strand-swap mechanism, as occurs in EC1 of clas-

sical cadherins (Shapiro et al., 2007), or by an antiparallel pairing

similar to that found for Ig2 and Ig3 in Dscam1.

Schreiner and Weiner (2010) also extended results of Kaneko

et al. (2006) by showing that Pcdh-gs form cis-tetramers in an

isoform-independent fashion and that this, in turn, expands the

binding specificity repertoire. They demonstrated that cells

expressing different ratios of Pcdh-gs exhibit selective binding

for cells expressing the same ratio. In another interesting exper-

iment, they tested the ability of cells expressing four isoforms to

bind to cells transected with the same or different four isoforms.

Cells sharing only one or two isoforms bound very poorly

whereas cells with three or four shared isoforms showed signif-

icant and similar levels of binding. Thus, cells expressing

different Pcdh-g combinations have distinct binding specific-

ities. If Pcdh-as and -bs contribute to the binding properties of

heteromultimeric Pcdh complexes, their combinatorial expres-

sion could greatly expand the repertoire of specificities.

Dscam Diversity Is Essential for PatterningNeural Circuit AssemblyAs described above, Dscams and Pcdhs are required for

numerous aspects of circuit assembly. But is there a special

role for their diversity? The question remains unanswered for

Pcdhs but has been addressed for Dscam1.

In a first test of whether Dscam1 diversity is required for neural

circuit assembly, the genomic region encoding the variable ecto-

domains was replaced with a cDNA fragment encoding only

a single isoform. Marked defects persisted within the peripheral

and central nervous systems, including in MB and da neurons,

Cell 143, October 29, 2010 ª2010 Elsevier Inc. 349

Page 38: Cell 101029

establishing that diversity is, indeed, essential (Hattori et al.,

2007).

To determine whether specific Dscam1 isoforms are required,

Wang et al. (2004) and Hattori et al. (2007) used a series of dele-

tions removing different sets of exons 4. No defects were seen

for either MB or da neurons, indicating that self-avoidance

does not rely upon any specific isoforms. Indeed, a single arbi-

trarily chosen isoform is sufficient for normal patterning of

a single da or MB neuron, as long as the surrounding neurons

express the wild-type gene and, thus, express different isoforms

(Figure 2B). This argues that self-avoidance relies solely on

differences between the isoforms expressed on neurons rather

than the particulars of their identity.

How much diversity is required? To address this question,

Hattori et al. (2009) constructed a series of knock-in mutants

through homologous recombination, generating animals

carrying 12, 24, 576, 1152, or 4752 isoforms. Both MB and da

neurons required between 1152 and 4752 isoforms for normal

patterning of axons and dendrites. Although extensive diversity

(thousands of isoforms) was not required for a neurite from

a single neuron to recognize and be repelled from a sister neurite,

it was essential to ensure that neurites did not inappropriately

recognize non-self as self. Thus, during neuronal differentiation

the biased stochastic expression of some 10–50 isoforms and

a large repertoire of isoforms from which to choose ensures

that each neuron is sufficiently different from its neighbors.

This allows them to distinguish between self and non-self with

high fidelity and this, in turn, ensures normal assembly of neural

circuits.

Conclusions and SpeculationsWe have emphasized striking molecular parallels between the

Dscams and Pcdhs (Table 1), all of which suggest that they

may play similar roles. Both are well suited by pedigree to

mediate intercellular interactions: they belong to the two largest

and best established families of cell adhesion molecules, the

immunoglobulin superfamily for Dscams and the cadherin super-

family for Pcdhs (Shapiro et al., 2007). Both are encoded by

complex genomic loci, with remarkable mechanisms to produce

many proteins from a single locus. For both, expression is gener-

ally stochastic and combinatorial rather than cell type specific,

endowing neurons with large numbers of individual identities.

Finally, both exhibit isoform-specific homophilic binding by a

mechanism involving interactions of multiple Ig (Dscam1) or

cadherin (Pcdh) domains.

Perhaps their most striking similarity, though, is their failure—

shared with olfactory receptors—to conform to a long-held

expectation of how synaptic connectivity is encoded. Sperry

(1963) hypothesized that neurons bear ‘‘individual identification

tags’’ that encode ‘‘specific chemical affinities.’’ But there is no

evidence to date that olfactory receptors, Dscams, or Pcdhs

act as transsynaptic ‘‘locks and keys’’ to match pre- and post-

synaptic partners. Instead, olfactory receptors regulate intracel-

lular messenger levels, Dscam1 mediates self-avoidance, and

the most striking role for Pcdhs identified so far is in neuronal

survival. Put bluntly, it is hard to imagine that families will be

found that are better suited than these to function as chemoaffin-

ity molecules. So, if they don’t serve this function, we need to

seriously consider the possibility that there is something wrong

with the conventional view. We argue that only limited diversity

is required for synaptic recognition and that the large-scale

diversity that does exist serves other purposes.

How many recognition molecules are required to form appro-

priate synaptic connections? In fact, in most regions of the

developing central nervous system, an ingrowing axon is faced

with the task of distinguishing among several to several dozen

cell types, not the ‘‘millions and perhaps billions’’ that Sperry

(1963) envisioned. The neuron’s birth will have placed it out of

reach of many of the neuronal types present in the nervous

system as a whole. Complex navigational machinery will have

guided the growth cone to a restricted target region, narrowing

the range of options still further. Within the target, the choice of

individual synaptic partners from a class of essentially equivalent

neurons may not matter much, and to the extent that it does

Table 1. Diversity, Expression, and Roles of Olfactory Receptors, Drosophila Dscam1, and Clustered Protocadherins

Feature Olfactory Receptors Dscam1 Clustered Pcdhs

# Genes 1000 1 58

Diversity mechanism Separate genes Alternative splicing Promoter choice and multimerization

Expression 1/cell (monoallelic) 10–50/cell �6/cella (largely monoallelic)

Number of protein products 1000 19,008 ectodomains 12,650 predicted Pcdh-g tetramersb

Ligands Odorants Self (homophilic binding)

and netrin

Self (homophilic binding of Pcdh-gs);

may also have other ligands

Require diversity? Yes Yes Unknown

Mechanism Modulate second messenger

levels intracellularly

Homophilic interactions between

processes of a cell leading to

repulsion

Unknown

Developmental phenotype Targeting of olfactory axons

to glomeruli in olfactory bulb

Axon and dendrite branching,

synaptic specificity

Axon targeting, synapse formation or

maintenance, neuronal survivala Assuming an average of two isoforms from each of the a, b, and g groups, not including aC and gC isoforms.b This estimate is based on the proposal that Pcdh-gs form tetramers (Schreiner and Weiner, 2010) and the assumptions that cells express four Pcdh-g

isoforms and that there is no isoform specificity to multimerization. Diversity could be lower if these assumptions are incorrect or greater in that Pcdh

oligomers can contain a and b as well as g isoforms (see text).

350 Cell 143, October 29, 2010 ª2010 Elsevier Inc.

Page 39: Cell 101029

matter, it may be regulated by quantitative topographic gradi-

ents of a few key molecules (Gierer, 1983). Thus, choices

required for synaptic specificity may be mediated in a fashion

analogous to those made by growth cones during axon guidance

based on the repeated reuse of a limited set of cell recognition

and secreted molecules.

If Dscam and Pcdh do not underlie transsynaptic recognition,

what is their purpose? Expression patterns provide a clue. For

these molecules to function in synaptic matching, the adhesive

repertoire of the two partners would need to be precisely

matched, requiring exquisite control of splicing or promoter

utilization. It is hard to imagine mechanisms capable of such

coordination. Indeed, although this type of regulation may occur

in some cases, expression appears largely stochastic rather than

determinative. Stochasticity is poorly suited to intercellular

recognition and synaptic selection but perfectly suited to self-

recognition and self-avoidance. The reason is that self-recogni-

tion requires each cell to distinguish itself from all of the other

cells it encounters (often many thousands), whereas synaptic

recognition, as we have argued, may require distinctions among

just a few dozen neuronal classes. When diversity is decreased,

for example genetically (Hattori et al., 2009), multiple neurons

would bear many of the same isoforms, and a neuron would

be likely to mistake a neighbor’s neurite for its own. The

combined features of vast diversity, isoform-specific binding,

and stochastic gene expression, which Dscam1 and Pcdhs

share, provide a simple and elegant way to provide all neurons

with an ability to distinguish between self and non-self. Indeed,

the functional experiments reviewed here show that Dscam1

proteins work precisely this way to promote self-avoidance.

Self-avoidance, in turn, allows patterning of dendritic arbors

and axon branches, thereby promoting uniform coverage of

receptor fields and branch segregation. Given their complemen-

tary phylogenetic distribution—Dscam diversity occurs in arthro-

pods but not vertebrates, whereas clustered Pcdhs occur only in

vertebrates—it is attractive to speculate that Pcdh diversity

plays roles in vertebrates similar to those played by Dscam1 in

flies. In principle, this idea can be tested genetically in mice, as

has been done in flies, by deletion and reintroduction of specific

Pcdhs isoforms.

In summary, then, recent studies of vertebrates and inverte-

brates argue that chemospecificity does exist on the scale

envisioned by Sperry but does not play the roles that have gener-

ally been envisioned for it. Although chemoaffinity seems likely to

underlie synaptic specificity, the number of tags required may be

limited. Conversely, cells do carry ‘‘identification tags’’ that

enable distinctions ‘‘almost at the level of the single neuron’’

(Sperry, 1963), but these tags act for self-recognition, an issue

unanticipated by Sperry and largely ignored by his successors.

So in the end, Sperry was right about the need for individual

cell identification tags, but we suspect that he would have

been surprised by the step in circuit assembly where they are

required.

ACKNOWLEDGMENTS

We thank Renate Hellmiss for drawing the figures, Dr. Michael Sawaya for the

structures shown in Figure 3D, and Daisuke Hattori, Julie Lefebvre, Kelsey

Martin, Woj Wojtowicz, and members of our labs for comments on the

manuscript. Work on protocadherins and Dscams in J.R.S.’ lab and on

Dscam1 in S.L.Z.’s lab is supported by the NIH. S.L.Z. is an investigator of

the Howard Hughes Medical Institute.

REFERENCES

Buck, L., and Axel, R. (1991). A novel multigene family may encode odorant

receptors: a molecular basis for odor recognition. Cell 65, 175–187.

Dickson, B.J. (2002). Molecular mechanisms of axon guidance. Science 298,

1959–1964.

Cheng, H.J., Nakamoto, M., Bergemann, A.D., and Flanagan, J.G. (1995).

Complementary gradients in expression and binding of ELF-1 and Mek4

in development of the topographic retinotectal projection map. Cell 82,

371–381.

Drescher, U., Kremoser, C., Handwerker, C., Loschinger, J., Noda, M., and

Bonhoeffer, F. (1995). In vitro guidance of retinal ganglion cell axons by

RAGS, a 25 kDa tectal protein related to ligands for Eph receptor tyrosine

kinases. Cell 82, 359–370.

Dreyer, W.J. (1998). The area code hypothesis revisited: olfactory receptors

and other related transmembrane receptors may function as the last digits in

a cell surface code for assembling embryos. Proc. Natl. Acad. Sci. USA 95,

9072–9077.

Esumi, S., Kakazu, N., Taguchi, Y., Hirayama, T., Sasaki, A., Hirabayashi, T.,

Koide, T., Kitsukawa, T., Hamada, S., and Yagi, T. (2005). Monoallelic yet

combinatorial expression of variable exons of the protocadherin-a gene

cluster in single neurons. Nat. Genet. 37, 171–176.

Fuerst, P.G., Koizumi, A., Masland, R.H., and Burgess, R.W. (2008). Neurite

arborization and mosaic spacing in the mouse retina require DSCAM. Nature

451, 470–474.

Fuerst, P.G., Bruce, F., Tian, M., Wei, W., Elstrott, J., Feller, M.B., Erskine, L.,

Singer, J.H., and Burgess, R.W. (2009). DSCAM and DSCAML1 function in

self-avoidance in multiple cell types in the developing mouse retina. Neuron

64, 484–497.

Gierer, A. (1983). Model for the retino-tectal projection. Proc. R. Soc. Lond. B

Biol. Sci. 218, 77–93.

Han, M.H., Lin, C., Meng, S., and Wang, X. (2010). Proteomics analysis reveals

overlapping functions of clustered protocadherins. Mol. Cell. Proteomics 9,

71–83.

Hasegawa, S., Hamada, S., Kumode, Y., Esumi, S., Katori, S., Fukuda, E.,

Uchiyama, Y., Hirabayashi, T., Mombaerts, P., and Yagi, T. (2008). The proto-

cadherin-a family is involved in axonal coalescence of olfactory sensory

neurons into glomeruli of the olfactory bulb in mouse. Mol. Cell. Neurosci.

38, 66–79.

Hattori, D., Demir, E., Kim, H.W., Viragh, E., Zipursky, S.L., and Dickson, B.J.

(2007). Dscam diversity is essential for neuronal wiring and self-recognition.

Nature 449, 223–227.

Hattori, D., Millard, S.S., Wojtowicz, W.M., and Zipursky, S.L. (2008).

Dscam-mediated cell recognition regulates neural circuit formation. Annu.

Rev. Cell Dev. Biol. 24, 597–620.

Hattori, D., Chen, Y., Matthews, B.J., Salwinski, L., Sabatti, C., Grueber, W.B.,

and Zipursky, S.L. (2009). Robust discrimination between self and non-self

neurites requires thousands of Dscam1 isoforms. Nature 461, 644–648.

Hughes, M.E., Bortnick, R., Tsubouchi, A., Baumer, P., Kondo, M., Uemura, T.,

and Schmucker, D. (2007). Homophilic Dscam interactions control complex

dendrite morphogenesis. Neuron 54, 417–427.

Hummel, T., Vasconcelos, M.L., Clemens, J.C., Fishilevich, Y., Vosshall, L.B.,

and Zipursky, S.L. (2003). Axonal targeting of olfactory receptor neurons in

Drosophila is controlled by Dscam. Neuron 37, 221–231.

Jacob, F. (1977). Evolution and tinkering. Science 196, 1161–1166.

Junghans, D., Heidenreich, M., Hack, I., Taylor, V., Frotscher, M., and Kemler,

R. (2008). Postsynaptic and differential localization to neuronal subtypes

Cell 143, October 29, 2010 ª2010 Elsevier Inc. 351

Page 40: Cell 101029

of protocadherin b16 in the mammalian central nervous system. Eur. J. Neuro-

sci. 27, 559–571.

Kaneko, R., Kato, H., Kawamura, Y., Esumi, S., Hirayama, T., Hirabayashi, T.,

and Yagi, T. (2006). Allelic gene regulation of Pcdh-a and Pcdh-g clusters

involving both monoallelic and biallelic expression in single Purkinje cells. J.

Biol. Chem. 281, 30551–30560.

Katori, S., Hamada, S., Noguchi, Y., Fukuda, E., Yamamoto, T., Yamamoto, H.,

Hasegawa, S., and Yagi, T. (2009). Protocadherin-a family is required for sero-

tonergic projections to appropriately innervate target brain areas. J. Neurosci.

29, 9137–9147.

Kohmura, N., Senzaki, K., Hamada, S., Kai, N., Yasuda, R., Watanabe, M.,

Ishii, H., Yasuda, M., Mishina, M., and Yagi, T. (1998). Diversity revealed by

a novel family of cadherins expressed in neurons at a synaptic complex.

Neuron 20, 1137–1151.

Kramer, A.P., and Kuwada, J.Y. (1983). Formation of the receptive fields of

leech mechanosensory neurons during embryonic development. J. Neurosci.

3, 2474–2486.

Kramer, A.P., and Stent, G.S. (1985). Developmental arborization of sensory

neurons in the leech Haementeria ghilianii. II. Experimentally induced varia-

tions in the branching pattern. J. Neurosci. 5, 768–775.

Langley, J.N. (1895). Note on regeneration of prae-ganglionic fibres of the

sympathetic. J. Physiol. 18, 280–284.

Lefebvre, J.L., Zhang, Y., Meister, M., Wang, X., and Sanes, J.R. (2008). g-Pro-

tocadherins regulate neuronal survival but are dispensable for circuit formation

in retina. Development 135, 4141–4151.

Matthews, B.J., Kim, M.E., Flanagan, J.J., Hattori, D., Clemens, J.C., Zipursky,

S.L., and Grueber, W.B. (2007). Dendrite self-avoidance is controlled by

Dscam. Cell 129, 593–604.

McLaughlin, T., and O’Leary, D.D. (2005). Molecular gradients and develop-

ment of retinotopic maps. Annu. Rev. Neurosci. 28, 327–355.

Meijers, R., Puettmann-Holgado, R., Skiniotis, G., Liu, J.H., Walz, T., Wang,

J.H., and Schmucker, D. (2007). Structural basis of Dscam isoform specificity.

Nature 449, 487–491.

Millard, S.S., Lu, Z., Zipursky, S.L., and Meinertzhagen, I.A. (2010). Drosophila

dscam proteins regulate postsynaptic specificity at multiple-contact

synapses. Neuron 67, 761–768.

Mombaerts, P. (2006). Axonal wiring in the mouse olfactory system. Annu. Rev.

Cell Dev. Biol. 22, 713–737.

Mombaerts, P., Wang, F., Dulac, C., Chao, S.K., Nemes, A., Mendelsohn, M.,

Edmondson, J., and Axel, R. (1996). Visualizing an olfactory sensory map. Cell

87, 675–686.

Morishita, H., and Yagi, T. (2007). Protocadherin family: diversity, structure,

and function. Curr. Opin. Cell Biol. 19, 584–592.

Morishita, H., Umitsu, M., Murata, Y., Shibata, N., Udaka, K., Higuchi, Y.,

Akutsu, H., Yamaguchi, T., Yagi, T., and Ikegami, T. (2006). Structure of the

cadherin-related neuronal receptor/protocadherin-a first extracellular cad-

herin domain reveals diversity across cadherin families. J. Biol. Chem. 281,

33650–33663.

Murata, Y., Hamada, S., Morishita, H., Mutoh, T., and Yagi, T. (2004).

Interaction with protocadherin-g regulates the cell surface expression of pro-

tocadherin-a. J. Biol. Chem. 279, 49508–49516.

Neves, G., Zucker, J., Daly, M., and Chess, A. (2004). Stochastic yet biased

expression of multiple Dscam splice variants by individual cells. Nat. Genet.

36, 240–246.

Phillips, G.R., Tanaka, H., Frank, M., Elste, A., Fidler, L., Benson, D.L., and

Colman, D.R. (2003). g-protocadherins are targeted to subsets of synapses

and intracellular organelles in neurons. J. Neurosci. 23, 5096–5104.

Prasad, T., Wang, X., Gray, P.A., and Weiner, J.A. (2008). A differential

developmental pattern of spinal interneuron apoptosis during synaptogenesis:

insights from genetic analyses of the protocadherin-g gene cluster. Develop-

ment 135, 4153–4164.

Sakano, H. (2010). Neural map formation in the mouse olfactory system.

Neuron 67, 530–542.

Sanes, J.R., and Yamagata, M. (2009). Many paths to synaptic specificity.

Annu. Rev. Cell Dev. Biol. 25, 161–195.

Sanes, J.R., and Zipursky, S.L. (2010). Design principles of insect and

vertebrate visual systems. Neuron 66, 15–36.

Sawaya, M.R., Wojtowicz, W.M., Andre, I., Qian, B., Wu, W., Baker, D.,

Eisenberg, D., and Zipursky, S.L. (2008). A double S shape provides the

structural basis for the extraordinary binding specificity of Dscam isoforms.

Cell 134, 1007–1018.

Schmucker, D., Clemens, J.C., Shu, H., Worby, C.A., Xiao, J., Muda, M.,

Dixon, J.E., and Zipursky, S.L. (2000). Drosophila Dscam is an axon guidance

receptor exhibiting extraordinary molecular diversity. Cell 101, 671–684.

Schreiner, D., and Weiner, J.A. (2010). Combinatorial homophilic interaction

between {g}-protocadherin multimers greatly expands the molecular diversity

of cell adhesion. Proc. Natl. Acad. Sci. USA 107, 14893–14898.

Shapiro, L., Love, J., and Colman, D.R. (2007). Adhesion molecules in the

nervous system: structural insights into function and diversity. Annu. Rev.

Neurosci. 30, 451–474.

Soba, P., Zhu, S., Emoto, K., Younger, S., Yang, S.J., Yu, H.H., Lee, T., Jan,

L.Y., and Jan, Y.N. (2007). Drosophila sensory neurons require Dscam for

dendritic self-avoidance and proper dendritic field organization. Neuron 54,

403–416.

Sperry, R.W. (1943). Visuomotor coordination in the newt (Triturus uiridescens)

after re-generation of the optic nerve. J. Comp. Neurol. 79, 33–55.

Sperry, R.W. (1944). Optic nerve regeneration with return of vision in anurans.

J. Neurophysiol. 7, 57–69.

Sperry, R.W. (1963). Chemoaffinity in the orderly growth of nerve fiber patterns

and connections. Proc. Natl. Acad. Sci. USA 50, 703–710.

Su, H., Marcheva, B., Meng, S., Liang, F.A., Kohsaka, A., Kobayashi, Y., Xu,

A.W., Bass, J., and Wang, X. (2010). g-protocadherins regulate the functional

integrity of hypothalamic feeding circuitry in mice. Dev. Biol. 339, 38–50.

Sudhof, T.C. (2008). Neuroligins and neurexins link synaptic function to

cognitive disease. Nature 455, 903–911.

Takeichi, M. (2007). The cadherin superfamily in neuronal connections and

interactions. Nat. Rev. Neurosci. 8, 11–20.

Tasic, B., Nabholz, C.E., Baldwin, K.K., Kim, Y., Rueckert, E.H., Ribich, S.A.,

Cramer, P., Wu, Q., Axel, R., and Maniatis, T. (2002). Promoter choice deter-

mines splice site selection in protocadherin a and g pre-mRNA splicing. Mol.

Cell 10, 21–33.

Wang, J., Zugates, C.T., Liang, I.H., Lee, C.H., and Lee, T. (2002a). Drosophila

Dscam is required for divergent segregation of sister branches and

suppresses ectopic bifurcation of axons. Neuron 33, 559–571.

Wang, J., Ma, X., Yang, J.S., Zheng, X., Zugates, C.T., Lee, C.H., and Lee, T.

(2004). Transmembrane/juxtamembrane domain-dependent Dscam distribu-

tion and function during mushroom body neuronal morphogenesis. Neuron

43, 663–672.

Wang, X., Su, H., and Bradley, A. (2002b). Molecular mechanisms governing

Pcdh-g gene expression: evidence for a multiple promoter and cis-alternative

splicing model. Genes Dev. 16, 1890–1905.

Wang, X., Weiner, J.A., Levi, S., Craig, A.M., Bradley, A., and Sanes, J.R.

(2002c). g protocadherins are required for survival of spinal interneurons.

Neuron 36, 843–854.

Weiner, J.A., Wang, X., Tapia, J.C., and Sanes, J.R. (2005). g protocadherins

are required for synaptic development in the spinal cord. Proc. Natl. Acad. Sci.

USA 102, 8–14.

Wojtowicz, W.M., Flanagan, J.J., Millard, S.S., Zipursky, S.L., and Clemens,

J.C. (2004). Alternative splicing of Drosophila Dscam generates axon

guidance receptors that exhibit isoform-specific homophilic binding. Cell

118, 619–633.

352 Cell 143, October 29, 2010 ª2010 Elsevier Inc.

Page 41: Cell 101029

Wojtowicz, W.M., Wu, W., Andre, I., Qian, B., Baker, D., and Zipursky, S.L.

(2007). A vast repertoire of Dscam binding specificities arises from modular

interactions of variable Ig domains. Cell 130, 1134–1145.

Wu, Q., and Maniatis, T. (1999). A striking organization of a large family of

human neural cadherin-like cell adhesion genes. Cell 97, 779–790.

Yamagata, M., and Sanes, J.R. (2008). Dscam and Sidekick proteins direct

lamina-specific synaptic connections in vertebrate retina. Nature 451, 465–469.

Zhan, X.L., Clemens, J.C., Neves, G., Hattori, D., Flanagan, J.J., Hummel, T.,

Vasconcelos, M.L., Chess, A., and Zipursky, S.L. (2004). Analysis of Dscam

diversity in regulating axon guidance in Drosophila mushroom bodies. Neuron

43, 673–686.

Zhu, H., Hummel, T., Clemens, J.C., Berdnik, D., Zipursky, S.L., and Luo, L.

(2006). Dendritic patterning by Dscam and synaptic partner matching in the

Drosophila antennal lobe. Nat. Neurosci. 9, 349–355.

Zou, C., Huang, W., Ying, G., and Wu, Q. (2007). Sequence analysis and

expression mapping of the rat clustered protocadherin gene repertoires.

Neuroscience 144, 579–603.

Cell 143, October 29, 2010 ª2010 Elsevier Inc. 353

Page 42: Cell 101029
Page 43: Cell 101029

DNA Damage-Mediated Inductionof a Chemoresistant NicheLuke A. Gilbert1 and Michael T. Hemann1,*1The Koch Institute for Integrative Cancer Research at MIT, Massachusetts Institute of Technology, Cambridge, MA 02139, USA

*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.09.043

SUMMARY

While numerous cell-intrinsic processes are knownto play decisive roles in chemotherapeutic response,relatively little is known about the impact of the tumormicroenvironment on therapeutic outcome. Here, weuse a well-established mouse model of Burkitt’slymphoma to show that paracrine factors in thetumor microenvironment modulate lymphoma cellsurvival following the administration of genotoxicchemotherapy. Specifically, IL-6 and Timp-1 arereleased in the thymus in response to DNA damage,creating a ‘‘chemo-resistant niche’’ that promotesthe survival of a minimal residual tumor burden andserves as a reservoir for eventual tumor relapse.Notably, IL-6 is released acutely from thymic endo-thelial cells in a p38-dependent manner followinggenotoxic stress, and this acute secretory responseprecedes the gradual induction of senescence intumor-associated stromal cells. Thus, conventionalchemotherapies can induce tumor regression whilesimultaneously eliciting stress responses that pro-tect subsets of tumor cells in select anatomical loca-tions from drug action.

INTRODUCTION

While significant progress has been made in the application of

chemotherapy over the past 40 years, most chemotherapeutic

regimens ultimately fail to cure cancer patients (Holen and Saltz,

2001). Even tumors that show dramatic initial responses to

therapy frequently relapse as chemoresistant malignancies.

This chemoresistance is thought to arise as a consequence of

cell intrinsic genetic changes including upregulation of drug

efflux pumps, activation of detoxifying enzymes or apoptotic

defects (Bleau et al., 2009). However, recent data suggests

that resistance to chemotherapy can also result from cell

extrinsic factors such as cytokines and growth factors (Eckstein

et al., 2009; Williams et al., 2007). Additionally, other studies have

suggested that rare cancer stem cells are the source of eventual

tumor relapse following therapy, as these cells are thought to be

drug resistant due to increased genomic stability, decreased

oxidative stress or the presence of multiple drug resistance

transporters (Visvader and Lindeman, 2008).

Modern combinatorial chemotherapeutic regimes can reduce

patient tumor burdens to undetectable levels, yet in many cases

these tumors will relapse (Corradini et al., 1999). Thus, even

when a patient is classified as being in complete remission,

surviving cancer cells can persist in particular anatomical

locations. This remnant population of cancer cells has been

described as minimal residual disease (MRD). MRD is generally

not macroscopic and may not be at the site of the primary tumor,

making this phenomenon difficult to dissect experimentally

(Ignatiadis et al., 2008). While MRD is a significant clinical

problem, few models exist to study residual tumor burden

following therapy. Thus, it remains unclear whether the cancer

cells that compose the MRD burden are surviving following

chemotherapy due to stochastic events, intrinsic drug resis-

tance, or microenvironmental cues.

Efforts to experimentally recapitulate the response of human

tumors in vivo to chemotherapy have generally relied upon xeno-

grafts of human tumors transplanted into immunodeficient mice

(Sharpless and Depinho, 2006). These models have proven inef-

fective in predicting drug efficacy, likely due to a failure to repro-

duce the complexity of a tumor with its associated complement

of stromal, immune and endothelial cells. This autochthonous

tumor microenvironment includes a complex mixture of pro-

and antineoplastic factors (Hideshima et al., 2007). Both malig-

nant and untransformed cells within a tumor influence the

balance of growth factors, chemokines and cytokines found in

the tumor microenvironment. These factors play key roles in

regulating tumor cell proliferation, and survival through the acti-

vation of diverse signaling pathways, including the Jak/Stat,

NFkB, Smad, and PI3K pathways (Nguyen et al., 2009).

While numerous studies have addressed the role of tumor-

proximal factors in tumor growth or metastasis, relatively few

have addressed the role of the tumor microenvironment in

chemotherapeutic outcome (Hanahan and Weinberg, 2000).

Here we show that two cytokines, IL-6 and Timp-1, protect

lymphoma cells from cell death induced by genotoxic chemo-

therapy, such that small molecule inhibition of cytokine-induced

signaling potentiates chemotherapeutic efficacy. We further

show that IL-6 release occurs as a result of p38 MAP Kinase

activation in tumor-associated endothelial cells acutely follow-

ing DNA damage. This acute cytokine release also occurs in

treated human endothelial and hepatocellular carcinoma

cells, suggesting that acute secretory responses may occur in

numerous contexts. In the thymus, rapid cytokine release

precedes the induction of senescence – a process recently

shown to promote sustained cytokine release in cultured cells

Cell 143, 355–366, October 29, 2010 ª2010 Elsevier Inc. 355

Page 44: Cell 101029

(Acosta et al., 2008; Coppe et al., 2008; Kuilman et al., 2008;

Wajapeyee et al., 2008). Thus, genotoxic drugs can, paradoxi-

cally, elicit prosurvival signaling in select anatomical sites,

providing a reservoir of minimal residual disease that subse-

quently fuels tumor relapse.

RESULTS

The Thymus Represents a ChemoprotectiveTumor MicroenvironmentTo investigate the dynamics of lymphoma response and relapse

following chemotherapy, we used a well-established preclinical

model of human Burkitt’s lymphoma—the Em-myc mouse

(Adams et al., 1985). Tumors from these mice can be trans-

planted into immunocompetent syngeneic recipient mice, and

the resulting tumors are pathologically indistinguishable from

autochthonous tumors (Burgess et al., 2008). Six to 8 week old

mice were tail vein injected with GFP-tagged Em-myc p19Arf�/�

B lymphoma cells. At tumor onset all mice displayed a character-

istic disseminated pattern of disease with lymphoma cells in the

peripheral lymph nodes, spleen and mediastinum. Mice were

treated with the maximum tolerated dose of the front-line

chemotherapeutic doxorubicin at the time of lymphoma mani-

festation. Three days after administration of doxorubicin, all

mice displayed tumor regression and peripheral tumor clear-

ance, measured by lymph node palpation. These mice were

sacrificed at four days posttreatment and sites of minimal

residual disease were identified by GFP imaging. Interestingly,

the majority of surviving lymphoma cells were in the mediastinal

cavity (Figure 1A), a central component of the thoracic cavity that

encapsulates the heart, esophagus, trachea and a large amount

of lymphatic tissue including the mediastinal lymph nodes and

the thymus.

To analyze the effect of drug treatment on specific tumor

niches, we harvested all primary lymphoid organs, including

peripheral lymph nodes, thymus, spleen and bone marrow,

following doxorubicin treatment. All tissues sampled showed

extensive lymphoma cell apoptosis and restoration of normal

organ architecture. Peripheral lymph nodes, spleen and bone

marrow exhibited nearly complete tumor clearance with rare

surviving lymphoma cells (Figure 1C and Figure S1A available

online). In contrast, many surviving B lymphoma cells could

CLymph Node

Thymus

0

1

2

3

4

5

6

7

8

Doxorubicin - +

Fo

ld

ch

an

ge (# L

ive C

ells)

Ratio

o

f T

hym

us/L

ym

ph

N

od

es

BA

D

Doxorubicin - + +

β-Tubulin

Lymph

Node Thymus

γ-H2AX

E

Overall S

urvival %

Days Following Treatment

20

40

60

80

100

5 10 15 20 25 30 35

0

C57BL/6 Rag1-/-

C57BL/6

Rag1+/+

0.0015

Untreated Doxorubicin

10mg/kg

Cervical Nodes

Thymus

p<0.0001

Axillary/Brachial

Nodes

Axillary/Brachial

Nodes

Figure 1. The Thymus Represents a Chemo-

protective Niche that Harbors Surviving

Lymphoma Cells Following Doxorubicin

Treatment

(A) Lymphoma-bearing mice were imaged for

whole body fluorescence prior to treatment and

4 days following a single dose of 10 mg/kg doxoru-

bicin. Representative mice are shown.

(B) Ratios of live GFP-tagged Em-myc p19Arf�/� B

lymphoma cells in the thymus versus peripheral

lymph nodes were quantified by flow cytometry,

before (n = 4 mice) and 48 hr after (n = 5 mice)

doxorubicin treatment. Average ratios are indi-

cated with a line.

(C) Hematoxylin and eosin (H&E) sections of lymph

node and thymus from a tumor-bearing bearing

mouse 48 hr after doxorubicin treatment. The

dotted line in the thymus demarcates a small

region of infiltrating lymphocytes neighboring

a larger region of surviving lymphoma cells.

Representative fields are shown at 403 magnifica-

tion.

(D) A western blot showing g-H2AX levels in FACS

sorted GFP-positive lymphoma cells from the

thymus and peripheral lymph nodes following

doxorubicin treatment. b-Tubulin serves as a

loading control. The untreated sample is a lysate

from cultured lymphoma cells.

(E) A Kaplan-Meier curve showing the overall

survival of tumor-bearing C57BL/6 (n = 8) or

C57BL/6 Rag1�/� (n = 5) mice following doxoru-

bicin treatment. The p value was calculated using

a log rank test.

See also Figure S1.

356 Cell 143, 355–366, October 29, 2010 ª2010 Elsevier Inc.

Page 45: Cell 101029

be seen in the thymus. To quantify this phenotype, cells were

harvested from peripheral lymph nodes and the thymus

following treatment, and the number of surviving GFP positive

lymphoma cells was assessed by flow cytometry. The number

of viable lymphoma cells in the thymus relative to the lymph

nodes increased 6.5-fold following doxorubicin treatment

(Figure 1B). Thus, the thymus represents a chemoprotective

niche that protects lymphoma cells from doxorubicin-induced

cell death.

To rule out the possibility that the selective survival of tumor

cells in the thymus was due to the specific exclusion of doxoru-

bicin from the mediastinum, we sorted live GFP-positive tumor

cells from the lymph nodes and thymus 12 hr after doxorubicin

treatment and blotted for g-H2AX, a marker of DNA damage

(Morrison and Shen, 2005). Western blot analysis showed that

cells in both anatomical locations undergo the same amount of

DNA damage (Figure 1D). Additionally, flow cytometry of medi-

astinal lymphoma cells failed to identify any sub-population of

lymphoma cells with decreased g-H2AX fluorescence

(Figure S1B). These data suggest that the thymus offers no phys-

ical barrier to drug delivery.

Minimal Residual Tumor Burden in the Thymus FuelsTumor Relapse Following ChemotherapyGiven the persistence of tumor cells in the thymus following

chemotherapy, we next sought to determine whether tumor cells

in the thymus contributed to lymphoma relapse. To this end, we

A

C

B

Fo

ld

C

ha

ng

e

(#

L

iv

e C

ells

)

Fo

ld

C

ha

ng

e

(#

L

iv

e C

ells

)

Lymph Node Thymus

No CM TCM BMM LNM No CM TCM LNM

0

0.5

1.0

1.5

2.0

2.5

3.0

No

rm

alize

d S

ig

na

l In

te

ns

ity

(x

10

6)

TARC

TIMP1

IL-6

KC

MCP-1

MIP-1ααG-CSF

sICAM

C5a

TNFαα

IL-16

IL-1αα

SDF-1

MCP-5

CXCL-1

IL-1ra

M-CSF

MIP-2

TREM-1

IL-2

CXCL-9

GM-CSF

Doxorubicin - ++ -

detaertnUMn02 niciburoxoD

0

1

2

3

4

5

6

7

8

9

0

0.25

0.5

0.75

1.0

1.25

p<0.0001

Figure 2. Thymic Conditioned Media Contains Soluble

Chemoprotective Factors

(A) A graph showing lymphoma cell survival in the presence of

doxorubicin alone or in the presence of conditioned media. The

data are represented as mean ± standard error of the mean

(SEM) (n = 3).

(B) A graph showing the growth of lymphoma cells cultured in

the absence or presence of conditioned media. The data are

represented as mean ± SEM (n = 3).

(C) Cytokine array analysis of conditioned media from

untreated and doxorubicin treated lymph nodes and thymus.

The data is represented graphically as normalized signal inten-

sity. Conditioned media was pooled from 3 or 4 mice for each

array.

examined therapeutic response in genetically

and surgically athymic mice. We injected control

or Rag1 deficient mice, which have severely atro-

phic thymuses, with lymphoma cells and then

treated tumor-bearing recipient animals with

doxorubicin. Overall survival and tumor free

survival were significantly extended in tumor-

bearing Rag1 deficient mice, relative to control

animals, suggesting that the presence of a func-

tional thymus promotes relapse and disease

progression (Figure 1E and Figure S1C). Similarly,

surgically thymectomized tumor-bearing mice also

showed extended tumor-free and overall survival

following therapy relative to control animals

(Figure S1D and data not shown). Notably, overall

survival in untreated tumor-bearing Rag1 deficient or thymec-

tomized mice was indistinguishable from that in control

animals (Figure S1E). Thus, the thymus harbors minimal residual

disease that contributes to tumor relapse following therapy in

this model.

Cultured Thymuses Secrete Prosurvival Factors In VitroPreferential lymphoma cell survival in the thymus following doxo-

rubicin treatment suggests that specific anatomical microenvi-

ronments may contain prosurvival factors absent in other

lymphoid organs. There is precedence for this phenomenon in

multiple myeloma, where the bone marrow microenvironment

promotes myeloma cell survival (Hideshima et al., 2007). To

address this possibility, we derived conditioned media from

the thymus (TM, for thymic media), bone marrow (BMM) and

lymph nodes (LNM) of mice treated with doxorubicin. Cultured

lymphoma cells were then treated with doxorubicin, along with

TM, LNM or BMM. Addition of TM provided a significant survival

advantage, with 10-fold more cells surviving 48 hr following

treatment (Figure 2A). This effect was specifically prosurvival,

as opposed to proproliferative, as these same conditioned

medias had little effect on lymphoma cell growth (Figure 2B). In

contrast, conditioned media derived from peripheral lymph

nodes had only a minimal effect on lymphoma cell survival (Fig-

ure 2A). Thus, soluble prosurvival factor(s) present in the thymic

microenvironment protect tumor cells from genotoxic chemo-

therapy.

Cell 143, 355–366, October 29, 2010 ª2010 Elsevier Inc. 357

Page 46: Cell 101029

Cytokine Levels Vary between Tumor-BearingAnatomical LocationsTo identify the factor(s) contributing to lymphoma cell survival in

the thymus, we performed cytokine arrays analyzing the abun-

dance of 40 cytokines and chemokines in conditioned media

from doxorubicin treated and untreated tumors (Figure 2C and

data not shown). Analysis of cytokine expression showed

significant differences between the thymic and lymph node

tumor microenvironments. Multiple factors related to cell

migration and cell cycle control were acutely upregulated

in thymic lymphomas, but not in peripheral lymphomas or

cultured lymphoma cells, following doxorubicin treatment.

These included the cytokines G-CSF, IL-1a, IL-1ra, IL-6, IL-16,

and the chemokines and growth factors KC, MCP-1, MCP-5,

MIP-2, and Timp-1 (Figure 2C).

Each of the upregulated factors was tested in vitro for the

ability to promote doxorubicin resistance. Of the 10 recombinant

proteins examined, only two produced a significant effect

on lymphoma cell survival following doxorubicin treatment

in vitro. Recombinant Interleukin-6 (IL-6), as a single agent,

was able to promote a 2.8-fold increase in the number of

surviving lymphoma cells 72 hr following doxorubicin treatment

(Figure 3A and Figure S2A). Similarly, addition of Tissue inhibitor

of metalloproteases 1 (Timp-1) resulted in a 3-fold increase in

surviving lymphoma cells following doxorubicin treatment

(Figure 3B). These factors had a combinatorial effect (Figure 3B),

as addition of both recombinant IL-6 and Timp-1 resulted in

a 4.5-fold increase lymphoma cell number following treatment

(Figure 3B). Importantly, neither factor alone or in combination

affected lymphoma cell growth, suggesting that this increase in

cell number was not due to enhanced cell proliferation

(Figure S2B). Additionally, recombinant IL-6 had no effect on

lymphoma cell motility in this setting (Figure S2C).

While these data show that both Timp-1 and IL-6 promote

chemoresistance in the thymus, we decided to focus our

efforts on the contribution of IL-6 to therapeutic response.

To determine whether IL-6 was acting to promote cell survival

in an autocrine fashion following release from lymphoma cells

or a paracrine fashion following release from surrounding

thymic cells, we performed lymphoma transplant experiments

in the presence or absence of IL-6. Specifically, IL-6+/+

lymphomas were transplanted into IL-6+/+ or IL-6�/� mice

(Figure 3C and Figure S2D). Tumor-bearing recipient mice

were then treated with the maximum tolerated dose of doxo-

rubicin and monitored for tumor free survival and overall

survival. Notably, while IL-6�/� and IL-6+/+ recipient mice

developed pathologically indistinguishable tumors, IL-6�/�

recipients displayed significantly longer tumor free survival

and overall survival following treatment than their IL-6+/+ coun-

terparts (Figure 3D and Figures S2F and S2G). Additionally,

histological analysis confirmed the lack of surviving lymphoma

cells in the thymus of IL-6�/� mice following treatment

(Figure 3E). Thus, IL-6 release from the tumor microenviron-

ment, rather than from the tumor itself, promotes tumor cell

survival.

To further interrogate the source of thymic IL-6, IL-6 levels

were examined in thymic lymphomas from doxorubicin-treated

IL-6+/+ and IL-6�/� recipient mice, as well as doxorubicin treated

Doxorubicin

20nM

Doxorubicin 20nM

+ IL-6 10ng/mL

Doxorubicin +

+

+++

IL-6

++

+-

-

-

Timp-1 -

+

+

+++

++

+-

-

-

-

+

+

+++

++

+-

-

-

-

24 hours 72 hours48 hours

A B

C

E

Overall S

urvival %

Days Following Treatment

5 10 15 20

20

40

60

80

100

0

Fo

ld

C

han

ge

(# L

ive C

ells)

Fo

ld

C

han

ge

(# L

ive C

ells)

0.0012

C57BL/6

C57BL/6 IL-6-/-

25

D

C57BL/6 IL-6-/-

C57BL/6

Eµ-myc p19Arf-/-

IL-6+/+

Lymph NodeThymus Lymph NodeThymus

C57BL/6 IL-6 -/-

C57BL/6

0

1

2

3

4

0

1

2

3

4

5p<0.0001Figure 3. IL-6 and Timp-1 Are Chemopro-

tective In Vitro and In Vivo

(A) A graph showing the fold change in lymphoma

cell number 72 hr after treatment with doxorubicin

as a single agent or doxorubicin plus recombinant

IL-6. The data are represented as mean ± SEM

(n = 4 independent experiments).

(B) A graph showing the relative survival of cultured

lymphoma cells at 24 hr intervals following treat-

ment with doxorubicin alone, doxorubicin plus

recombinant IL-6 or Timp-1, or doxorubicin plus

both IL-6 and Timp-1. The data are represented

as mean ± SEM (n = 3 independent experiments).

(C) A schematic diagram of the lymphoma trans-

plant experiment, showing injection of IL-6+/+

lymphoma cells into both IL-6+/+ and IL-6�/� recip-

ients.

(D) A Kaplan-Meier curve showing post-treatment

survival of IL-6+/+ (n = 17) or IL-6�/� (n = 5) mice

bearing IL-6+/+ lymphomas. All mice were treated

with a single dose of 10mg/kg doxorubicin. The

p value was calculated using a log rank test.

(E) H&E stained sections of lymphomas 72 hr

following doxorubicin treatment. The black dotted

line shown in the thymus from the IL-6+/+ recipient

mouse demarcates a zone of surviving lymphoma

cells that is absent in the other sections. Repre-

sentative fields are shown at 203 magnification.

See also Figure S2.

358 Cell 143, 355–366, October 29, 2010 ª2010 Elsevier Inc.

Page 47: Cell 101029

lymphoma cells in vitro (Figure 3C). No IL-6 was detected in

either IL-6+/+ lymphoma cells in vitro or in thymic or peripheral

tumors from IL-6�/� lymphoma-bearing mice (data not shown).

This result strongly suggests that the IL-6 present in the tumor

microenvironment following genotoxic stress is secreted in

a paracrine manner from resident thymic cells.

Genotoxic Chemotherapy Induces the Releaseof Prosurvival Cytokines In VivoRecent work has shown that DNA damage can induce a secre-

tory phenotype in cultured cells (Rodier et al., 2009). To deter-

mine whether IL-6 is similarly induced as a consequence of

genotoxic chemotherapy in vivo, we treated mice lacking

tumors with the maximally tolerated dose of doxorubicin.

18 hr later we assayed IL-6 levels by ELISA in untreated and

doxorubicin treated mice. As suggested by the cytokine array

E

Lymph

Node

Thymus

Fo

ld

C

ha

ng

e

(#

L

iv

e C

ells

)

C

Doxorubic

in

Doxorubic

in

+ T

CM

Doxorubic

in

+ T

CM

+ J

aki

Tu

mo

r F

re

e S

urv

iv

al %

Days Following Treatment

20

40

60

80

100

5 10 15 200

D

0.004

Doxorubicin

Doxorubicin +

Ag490

Untreated Doxorubicin Doxorubicin + Ag490

0

1

2

3

4

5

6

7

8

9

11

10

BA

pg

/m

g o

f IL

6

Thymic

Lymphoma

Peripheral

Lymphoma

Doxorubicin - + - +

pg

/m

g o

f IL

6

Thymus Peripheral

Lymph Node

Doxorubicin - + - +

0

5

10

15

20

25

30

35

40

45

55

50

0

10

20

30

40

50

60

70

80

90

p<0.01 p<0.05 Figure 4. Doxorubicin Induces the Release

of IL-6, and Inhibition of This Cytokine

Signaling Sensitizes Tumor Cells to Chemo-

therapy

(A) Quantification of IL-6 levels in conditioned

media from the thymus or lymph nodes of

untreated mice (n R 10) or mice treated for 18 hr

with 10mg/kg doxorubicin (n R 7). Values were

normalized by tissue weight. The data are repre-

sented as mean ± SEM.

(B) Quantification of IL-6 levels in conditioned

media derived from tumor-bearing thymuses or

lymph nodes of doxorubicin treated (n = 3) or

untreated (n = 3) mice. The data are represented

as mean ± SEM.

(C) A bar graph showing the fold change in number

of live cells at 48 hr following treatment with doxo-

rubicin alone or in combination with conditioned

media plus or minus a Jak2 inhibitor. The data

are represented as mean ± SEM (n = 3).

(D) A Kaplan-Meier curve showing tumor free

survival of lymphoma-bearing mice treated with

doxorubicin (n = 9) or doxorubicin plus two doses

of 50mg/kg AG-490 m-CF3 (n = 4). The p value

was calculated using a log rank test.

(E) H&E sections of lymphomas 72 hr after treat-

ment with doxorubicin or doxorubicin plus AG-

490 m-CF3. Black dotted lines distinguish surviving

lymphoma cells, which are largely absent in the

presence of AG-490 m-CF3, from infiltrating

immune cells. Representative fields are shown at

203 magnification.

See also Figure S3.

data, IL-6 was present at a constitutively

higher level in the thymus versus the

lymph nodes of mice (Figure 4A). Addi-

tionally, doxorubicin treatment signifi-

cantly increased the amount of IL-6 in

the thymus but not in peripheral lymph

nodes or the spleen (Figure 4A and

Figure S3A). Thus, genotoxic chemo-

therapy induces a stress response in

the thymus that includes the release of

IL-6, a prosurvival cytokine. Notably, IL-6 induction in the

thymus occurred within 18 hr of treatment, much more acutely

than has been reported for secretory phenotypes in cultured

cells (Rodier et al., 2009).

To confirm that this acute DNA damage-induced secretory

response also occurs in thymuses with substantial lymphoma

infiltration, we examined IL-6 levels in tumor-bearing mice.

Again, IL-6 was constitutively present at a higher level in the

thymic tumor microenvironment versus the peripheral tumor

microenvironment, and treatment with doxorubicin resulted in

a rapid and significant increase in IL-6 levels in thymic

lymphomas but not in peripheral lymphomas (Figure 4B). As

chemotherapy rapidly induces apoptosis in lymphoid malignan-

cies, these data are consistent with the idea that acute cytokine

release following DNA-damage may directly impact therapeutic

response.

Cell 143, 355–366, October 29, 2010 ª2010 Elsevier Inc. 359

Page 48: Cell 101029

Jak2 Signaling Is Required for Lymphoma Cell SurvivalFollowing Doxorubicin Treatment In Vivo and In VitroBoth IL-6 and Timp-1 have been shown to signal through

Jak2 and Stat3 (Heinrich et al., 1998; Lambert et al., 2003), sug-

gesting that doxorubicin efficacy could be potentiated if Jak

signaling were chemically inhibited. We tested this hypothesis

by treating lymphoma cells with doxorubicin and TM or doxoru-

bicin and TM plus a Jak2/Jak3 inhibitor. Addition of the Jak

inhibitor completely ablated the protective effect of TM (Fig-

ure 4C). Importantly, the Jak inhibitor had a minimal effect on

lymphoma cell growth in the presence of TM (data not shown).

Thus, Jak2/Jak3 signaling promotes the chemoprotective effect

of TM in vitro.

To determine whether this effect could be recapitulated

in vivo, we treated lymphoma-bearing mice with either doxoru-

bicin, Ag490 (a Jak2/3 inhibitor previously used in murine

in vivo studies) (Gu et al., 2005), or with a combination of both

doxorubicin and Ag490. Mice treated with doxorubicin and

Ag490 showed significantly longer tumor-free survival and over-

all survival than mice treated with doxorubicin alone (Figure 4D

and Figures S3B and S3C). Histological analysis of the thymus

in mice treated with both drugs showed few surviving lymphoma

cells, in sharp contrast with those treated with doxorubicin alone

(Figure 4E). Importantly, this was not due to a simple additive

effect of doxorubicin and Ag490-induced cytotoxicity, as mice

treated with Ag490 alone exhibited no tumor free survival or

extended overall survival when compared to mice treated with

a vehicle control (Figures S3D and data not shown). Thus,

Jak2/Jak3 inhibition can eliminate prosurvival signaling in the

thymic niche and potentiate doxorubicin cytotoxicity.

IL-6 Is Released from Thymic Endothelial CellsTo identify the cell type(s) responsible for IL-6 release from the

thymic stroma, we disassociated thymuses from untreated

mice and sorted known resident cells by characteristic surface

markers. Sorted cells were then plated in normal growth media,

and IL-6 levels were assessed after 48 hr. Notably, the vast

majority of IL-6 secreted following thymic disassociation was

released from thymic endothelial cells, while B, T, dendritic,

and thymic epithelial cells failed to produce any IL-6 levels above

background (Figure 5A). Resident macrophages produced trace

levels of IL-6 (Figure 5A and Figure S4). However, they did so at

a level that was more than ten-fold less than endothelial cells.

Similar results were seen for Timp-1, which was released almost

exclusively from thymic endothelial cells (Figure 5B).

Importantly, in mice treated with doxorubicin, IL-6 levels were

significantly induced in thymic endothelial cells (Figure 5F). While

IL-6 levels were also elevated in treated macrophages (Figures

S4A), this increase was not significant and represented less

than one tenth of the amount released from treated endothelial

cells - even when adjusted for total cell number. Additionally, no

significant increase in infiltrating macrophages or dendritic cells

was seen acutely following treatment (data not shown). Consis-

tent with the central role of endothelial cells in this secretory

response, pretreatment of mice with an inhibitor of VEGFR1/2 –

receptors necessary for endothelial cell proliferation - partially

inhibited doxorubicin-induced IL-6 release (Figure S4B). These

data indicate that resident endothelial cells are largely responsible

for the accumulation of prosurvival factors following chemo-

therapy in thismodel. To directly assess the relevanceof endothe-

lial cells to tumor cell survival, we co-cultured purified endothelial

cells and lymphoma cells in the presence of doxorubicin (Fig-

ure 5C). The presence of endothelial cells dramatically increased

lymphoma cell survival following treatment, with a 15-fold

increase in lymphoma cell number in co-cultured populations

relative to lymphoma cell-only populations 72 hr posttreatment.

Several studies have indicated that cytokines, including IL-6,

may exert a prosurvival benefit in target cells through induction

of antiapoptotic Bcl2 family members, including Bcl2, Bcl-XL

and Mcl-1 (Jourdan et al., 2000). Thus, we examined the protein

levels of Bcl2 family members in lymphoma cells treated with

thymic conditioned media. While Bcl2 and Mcl-1 levels were

unaffected (data not shown), Bcl-XL was consistently induced

2- to 4-fold (Figure 5D). To further examine whether Bcl-XL

contributes to cell survival in this context, we treated cells ex-

pressing a Bcl-XL shRNA with doxorubicin alone or doxorubicin

plus IL-6 (Figure 5E). Suppression of Bcl-XL blocked the ability of

IL-6 to promote doxorubicin resistance, suggesting that IL-6

mediated induction of Bcl-XL may be necessary for its role in

cell survival. This does not, however, preclude that other factors

may contribute to cell survival following exposure to IL-6, as

cytokines are known to activate numerous prosurvival pathways.

IL-6 Release from Endothelial Cells Is Dependentupon p38 MAP Kinase ActivityThe p38 MAP Kinase (p38) is known to be a key regulator of the

expression of inflammatory cytokines, including IL-6 (Medzhitov

and Horng, 2009). To determine whether p38 is required for DNA

damage-induced IL-6 release, treated and untreated thymic

endothelial cells were purified and probed by immunofluores-

cence for the presence of activated p38. Notably, treated endo-

thelial cells showed significantly higher phospho-p38 levels than

their untreated counterparts (Figure S4C). To examine the func-

tional relevance of this p38 activation, we plated thymic endothe-

lial cells from mice treated with doxorubicin in the presence or

absence of a p38 inhibitor (Figure 5F). Strikingly, the addition

of a p38 inhibitor not only prevented IL-6 induction, but actually

reduced the level of secreted IL-6 to below the level in untreated

cells. To investigate whether this DNA damage-induced IL-6

release is a conserved characteristic of endothelial cells, we per-

formed similar experiments in human vascular endothelial cells

(HUVECs). Cultured HUVECs were treated with doxorubicin

and conditioned media was collected 24 hr after treatment.

Here, doxorubicin elicited a threefold increase in the amount of

secreted IL-6 (Figure 5G). This process was also dependent

upon p38 activity, as concurrent treatment of HUVECs with

doxorubicin and a p38 inhibitor blocked IL-6 induction (Figures

5F and 5H).

Cell-based studies have implicated the ATM checkpoint

kinase in senescence-associated secretory phenotypes (SASP)

(Rodier et al., 2009). To examine the relevance of ATM to endo-

thelial IL-6 release, we treated HUVECs with doxorubicin and an

ATM inhibitor (Figure 5H). Surprisingly, as opposed to blocking

cytokine secretion, ATM inhibition significantly increased the

level of endothelial IL-6 release. These data suggest that the

biology of acute cytokine release may be distinct from SASP.

360 Cell 143, 355–366, October 29, 2010 ª2010 Elsevier Inc.

Page 49: Cell 101029

Cytotoxic Chemotherapy Induces Senescencein Thymic Stromal Cells In VivoRecent studies have shown that an autocrine IL-6 signaling

loop is induced upon oncogene activation and that this auto-

crine loop reinforces oncogene induced senescence (OIS)

(Coppe et al., 2008; Kuilman et al., 2008). This observation

led us to investigate whether IL-6 secretion in the thymus is

accompanied by drug-induced senescence. To determine

whether doxorubicin induces senescence in vivo, we harvested

the thymus and lymph nodes from mice 6 days following

treatment with doxorubicin. Tissues were frozen, sectioned

T c

ells

Epith

elia

l

Macrophage

Dendritic

B c

ells

Endoth

elia

l

pg

/m

L/1

06 C

ells

o

f IL

-6

1000

1500

2000

2500

3000

500

0

3500

Untreated Doxorubicin

pg

/m

L o

f IL

-6

pg

/m

L/1

06 C

ells

o

f IL

-6

2000

3000

4000

5000

1000

0

Doxorubic

in

Untreate

d

Doxorubic

in

+ S

B203580

10

15

25

5

0

50

75

25

0

20

pg

/m

L o

f IL

-6

Untreated

SB203580

KU55933

DoxorubicinUntreated

-+

-

---

--

-

+

+

+

+

+

-

-

--

BA

DC

FE

HG

T c

ells

Epith

elia

l

Macrophage

Dendritic

B c

ells

Endoth

elia

l

pg

/m

L/1

05 C

ells

o

f T

im

p-1

2000

3000

4000

5000

1000

0

10

15

5

0

20

Fo

ld

C

ha

ng

e

(#

L

iv

e C

ells

)

72 hours48 hours

Lymphoma

Endothelial -+

-+

+

+

+

+

Vinculin

Bcl-XL

TCMUntreated

Vector shBcl-XLshBcl-XL

Doxorubicin

IL-6 -+

-+

+

+

0.5

0.75

0.25

0

1.0

Fo

ld

C

ha

ng

e

p<0.0001

p<0.0001p<0.0001

Figure 5. Endothelial Cells Secrete IL-6 and

Timp-1 in Response to DNA Damage in

a p38 MAP Kinase-Dependent Manner

(A and B) (A) IL-6 and (B) Timp-1 levels were quan-

tified by ELISA in conditioned media derived from

sorted thymic cell populations The data are repre-

sented as mean ± SEM (n R 3 independent exper-

iments). Values were normalized to the number of

cells sorted.

(C) A graph showing lymphoma cell survival in

response to 20nM doxorubicin, with or without

endothelial cell co-culture. Fold change in cell

number was assessed at 48 and 72 hr posttreat-

ment. The data are represented as mean ± SEM

(n = 6 independent experiments).

(D) A western blot for Bcl-XL levels in lymphoma

cells in the presence or absence of TCM for

24 hr. The blot is representative of three indepen-

dent experiments.

(E) A graph showing the results of a GFP competi-

tion assay in cells partially transduced with

a Bcl-XL shRNA or a control vector. Fold change

in GFP percentage was assessed 48 hr following

treatment with 20nM doxorubicin. The data are

represented as mean ± SEM (n = 3).

(F) A bar graph showing the amount of IL-6 in

conditioned media from endothelial cells sorted

from the thymus of untreated mice (n = 5), mice

treated with doxorubicin (n = 8) or mice treated

with doxorubicin plus 10 mm SB203580 (n = 4).

Values were normalized to cell number. The data

are represented as mean ± SEM.

(G) A graph showing the amount of IL-6 present in

conditioned media from untreated and treated

human vascular endothelial cells (HUVECs). The

data are represented as mean ± SEM (n = 3).

(H) A graph showing the amount of IL-6 present in

conditioned media from HUVECs 48 hr after treat-

ment with doxorubicin alone or doxorubicin plus

either 10 mm SB203580 or 10 mm KU55933. The

data are represented as mean ± SEM (n = 3).

See also Figure S4.

and stained for b-Galactosidase activity –

a marker of cellular senescence (Dimri

et al., 1995). Tissues from untreated

mice showed no senescent cells (Fig-

ure 6A). In sharp contrast, b-Galactosi-

dase-positive cells were abundant in

the thymus, but not the lymph nodes,

of doxorubicin-treated mice (Figure S5A). Notably, this ‘‘senes-

cent’’ state was transient, as b-Galactosidase positive cells

were no longer present at twelve days following treatment

(Figure S5B). While the mechanism underlying the transient

presence of senescent cells in this context is unclear, these

data are consistent with the recognition and removal of senes-

cent cells by the innate immune system (Krizhanovsky et al.,

2008; Xue et al., 2007). Thus, doxorubicin can elicit the acute

release of prosurvival cytokines from non-tumor cells in the

thymus, coincident with a more gradual induction of senes-

cence.

Cell 143, 355–366, October 29, 2010 ª2010 Elsevier Inc. 361

Page 50: Cell 101029

To confirm that a similar ‘‘senescent’’ state also occurs in the

thymic tumor microenvironment, lymphoma-bearing mice were

treated with doxorubicin. Six days following treatment, mice

were sacrificed and tumors were harvested. Again, thymic tumor

sites showed the presence of disseminated senescent cells,

while the lymph nodes lacked any b-galactosidase positivity

(Figure 6A and Figure S5A). At least a component of this treat-

ment-induced senescent population was comprised of endothe-

lial cells, as purified CD31+/CD34+ cells showed senescent

phenotypes – including b–galactosidase activity (Figure S5C).

Notably, tumor-bearing thymuses and lymph nodes showed

similar numbers of macrophages and dendritic cells, suggesting

that the b-galactosidase positive cells in the thymus were

resident stromal cells as opposed to infiltrating immune cells

(Figure S5D).

IL-6 Modulates a General Response to DNA Damagein the ThymusThe presence of a prosurvival secretory response in the thymus

following chemotherapy led us to investigate whether IL-6 is

involved more generally in stress-induced thymic homeostasis.

Whole body irradiation, such as that occurring prior to bone

marrow transplantation, induces thymocyte cell death, periph-

eral leucopenia and thymic involution. This acute wave of thymo-

cyte death is followed by an acute regrowth of the thymus,

termed ‘‘thymic rebound’’ (Delrez et al., 1978). To assess

whether IL-6 induced by DNA damage is similarly cytoprotective

A Untreated

Doxorubicin

10mg/kg

Thymic

Lymphoma

Thymus

B

Day 5 Day 12

Mouse

genotype

IL-6+/+

IL-6+/+

IL-6-/-

IL-6-/-

p<0.001

p<0.01

Re

la

tiv

e W

eig

ht

0

0.5

0.75

0.25

p=0.32

p=0.69

0

0.5

0.75

0.25

1.00

Day 5 Day 12

Mouse

genotype

IL-6+/+

IL-6+/+

IL-6-/-

IL-6-/-

neelpSsumyhT

Re

la

tiv

e W

eig

ht

Figure 6. Genotoxic Damage Promotes Cellular

Senescence in Thymic Stromal Cells and Subsequent

IL-6-Mediated Thymic Rebound

(A) b-galactosidase staining of normal and tumor-bearing

thymuses and lymph nodes in the presence or absence of

doxorubicin-induced DNA damage. Representative fields are

shown at 203 magnification.

(B) A graph showing relative thymic and splenic weight follow-

ing genotoxic damage in the presence and absence of IL-6.

Organ weights are shown as the ratio of individual irradiated

thymus or spleen weights relative to the average unirradiated

thymicor spleen weight for each genotype. Each dot represents

an individual mouse, with a line demarcating the mean for each

cohort. The data are represented as mean ± SEM.

See also Figure S5.

in this setting, we irradiated wild-type and IL-6�/�

mice. 5 or 12 days later, all mice were sacrificed

and the spleen and thymus were harvested and

weighed. Thymic regrowth in IL-6�/� mice was

significantly reduced when compared to wild-

type control mice (Figure 6B), while no difference

was seen in the spleen. Therefore, IL-6 secretion in

the thymus may be critical for thymic growth and

repopulation following diverse genotoxic stresses.

Genotoxic Damage Promotes Acute IL-6Release and Chemoprotection in HumanLiver Cancer CellsPrevious descriptions of secretory phenotypes

have reported a gradual induction of cytokine

release following the onset of stress-induced cellular senes-

cence. This suggests that the release of prosurvival cytokines

may not occur rapidly enough to impact chemotherapeutic

response. The finding that doxorubicin can induce an acute

secretory response led us to investigate whether IL-6 induction

might be relevant to therapeutic response in contexts other

than the thymic microenvironment. Recent reports have impli-

cated IL-6 as a major contributor to the pathogenesis of hepato-

cellular carcinoma (HCC) (Naugler et al., 2007; Wong et al.,

2009). In humans, activating mutations in gp130, the obligate

signal transducing subunit of the IL-6 receptor, have been

recently identified (Rebouissou et al., 2009). Additionally recent

expression profiling identified the presence of an IL-6 induced

transcriptional signature in the tumor stroma that is associated

with poor prognosis in hepatocellular carcinoma (Hoshida

et al., 2008).

We treated an HCC cell line, Focus cells, with doxorubicin -

a front-line therapy for HCC - and measured the levels of

secreted IL-6 after 24 hr. Consistent with our endothelial cell

data, IL-6 secretion was increased over 3-fold acutely following

treatment (Figure 7A). Notably, treated cells lacked any markers

of senescence at this time point, indicating that senescence is

not required for acute cytokine release. In contrast with endothe-

lial cells, IL-6 secretion could be partially inhibited by either a p38

or an ATM inhibitor (Figure 7B). Thus, pathway requirements for

acute secretory phenotypes may be somewhat variable between

cell types.

362 Cell 143, 355–366, October 29, 2010 ª2010 Elsevier Inc.

Page 51: Cell 101029

We then investigated whether inhibition of IL-6 signaling could

enhance doxorubicin-induced cell death in HCC cells. We

treated Focus cells with doxorubicin alone, Ag490 alone or

doxorubicin in combination with Ag490. Treatment with both

doxorubicin and Ag490 resulted in more apoptosis and fewer

surviving cells in acute survival assays than either single agent

alone (Figure 7C). The combination treatment of doxorubicin

and Ag490 was also more effective than single agent therapy

when measured in a colony formation assay (Figure 7D). These

data suggest that acute drug-induced IL-6 release is chemopro-

tective in HCC and may contribute to the intrinsic chemoresist-

ance of these tumors.

DISCUSSION

The persistence of minimal residual disease following anticancer

therapy is strongly correlated with decreased survival in patients

(Xenidis et al., 2009). However, the mechanisms by which cells

survive in select contexts following chemotherapy are unclear.

In a mouse model of Burkitt’s lymphoma, we show that a specific

anatomical location, the thymus, confers a potent cytoprotective

benefit to lymphoma cells treated with genotoxic chemotherapy.

This surviving cell population is functionally relevant to disease

progression, as ablation of the thymus prolongs both tumor

free survival and overall survival following treatment with chemo-

therapy. While the importance of the thymic microenvironment to

tumor cell survival in human malignancy remains unclear, we

expect that factors contributing to drug resistance at this site

may also underlie MRD persistence at analogous locations in

human cancers.

C

Doxorubicin

Untreated Ag490

Ag490 +

Doxorubicin

D

Fo

ld

C

han

ge

(# L

ive C

ells)

No doxorubicin

Doxorubicin

Ag 490

8 µM

Ag 490

6 µM

Ag 490

4 µM

Ag 490

0 µM

3

10

11

12

13

1

0

2

Untreated Doxorubicin

pg

/m

L o

f IL

6

BA

pg

/m

L o

f IL

6

0

250

500

750

1000

50

75

100

125

150

25

0

175

Doxorubic

in

Doxorubic

in

+ S

B203580

Doxorubic

in

+ K

U55933

Untreate

d

p<0.0001p<0.01

p<0.01 Figure 7. DNA Damage Acutely Induces IL-

6 in Human Hepatocellular Carcinoma,

Promoting Both Cellular Survival and

Senescence

(A) IL-6 levels were quantified in conditioned

media derived from Focus cells treated with 200

nM doxorubicin for 24 hr. The data are repre-

sented as mean ± SEM (n = 3).

(B) A graph showing the amount of IL-6 present

in Focus cells 48 hr following treatment with either

SB203580 or KU55933, in the presence or

absence of doxorubicin. The data are represented

as mean ± SEM (n = 3).

(C) A graph showing the results of an acute cell

survival assay in which Focus cells were treated

with doxorubicin and increasing doses of Ag490,

as indicated, for 4 days. The data are represented

as mean ± SEM (n = 3 independent experiments).

(D) A colony formation assay showing Focus cells

that were treated with doxorubicin, Ag490 or both

for 24 hr before replating. Results are representa-

tive of three independent experiments.

The establishment of the thymic pro-

survival microenvironment occurs, para-

doxically, as a response to genotoxic

chemotherapy. Specifically, prosurvival

chemokines and cytokines are acutely

released following DNA damage. While

the complete signaling network leading from a DNA damage

response to cytokine release remains unclear, it involves the

activation of stress responsive kinases – most notably the MAP

kinase p38. Thus, cells exposed to genotoxic damage in vivo

can engage well-described cell cycle arrest and apoptotic

programs, as well as a physiological stress response pathway

leading to survival signaling. Importantly, the resulting secretory

response occurs not in the tumor cells themselves, but in prox-

imal endothelial cells. Drug treated endothelial cells release

IL-6 and Timp-1, which promote the induction of Bcl-XL in prox-

imal lymphoma cells. As a result, proapoptotic signaling induced

by the direct action of chemotherapy on tumor cells is countered

by antiapoptotic signaling emanating from the treated vascular

compartment in the tumor microenvironment.

Recent literature has shown that the induction of oncogene-

induced cellular senescence elicits a secretory response (Coppe

et al., 2008; Kuilman et al., 2008). Here we find that IL-6 is

induced acutely following DNA damage, prior to the onset of

senescence. This difference between a prosurvival response

that occurs within one day of treatment and a SASP that is

detectable only after 3-4 days is critical. Chemotherapy-induced

cell death generally occurs with 48 hr of treatment. Thus, a SASP

simply cannot effectively alter treatment response, as it occurs

well after tumor cell death decisions are made. However, our

data does not preclude a role for senescence in overall tumor

survival following therapy. Given that significant levels of senes-

cence occur in the thymus days after doxorubicin treatment, it is

possible that elevated IL-6 levels are maintained through the

establishment of a SASP. Thus, acute cytokine release and

subsequent senescence-related secretory phenotypes may

Cell 143, 355–366, October 29, 2010 ª2010 Elsevier Inc. 363

Page 52: Cell 101029

represent a general strategy to promote paracrine cell survival in

response to genotoxic stress in select microenvironments.

Secretory and inflammatory processes have been shown to be

critical for tissue repair and regeneration (Grivennikov et al.,

2009; Krizhanovsky et al., 2008). Thus, chemotherapy in the

thymic setting may activate physiological mechanisms of tissue

homeostasis. Consistent with this idea, the thymus is known to

engage prosurvival and growth mechanisms in response to other

cellular stresses. Following radiotherapy and subsequent thymic

atrophy (Muller-Hermelink et al., 1987), the thymus regrows and

replenishes the peripheral T cell population, leading to significant

thymic hyperplasia – even in adults with limited remaining thymic

tissue (Sfikakis et al., 2005). Factors that govern this process

have not been previously identified. Here we show that IL-6

modulates thymic recovery in response to DNA damage. This

suggests that lymphomas, and perhaps other malignancies,

can co-opt organ-specific prosurvival mechanisms.

Interestingly, serum IL-6 levels are elevated in many types of

cancer (Trikha et al., 2003), and high IL-6 levels are strongly

correlated with poor overall survival and accelerated disease

progression in a variety of cancers, including lymphomas

(Seymour et al., 1995). Furthermore, IL-6 levels are greatly

increased in metastatic disease versus non-metastatic disease

(Salgado et al., 2003). Consequently, stromal or tumor upregula-

tion of IL-6 may contribute to the intrinsic chemoresistance

commonly found in both primary and metastatic malignancies.

Additionally, tumor-directed inflammatory responses that result

in IL-6 release may similarly limit the efficacy of genotoxic

agents. These data demonstrate how both intrinsic genetic alter-

ations as well as chemoprotective microenvironments can play

decisive roles in the cellular response to genotoxic insults.

Thus, improved chemotherapeutic regimes may require a combi-

nation of cytotoxic agents, which target tumor cells, and tar-

geted therapeutics that inhibit prosurvival signaling from the

tumor-adjacent cells.

EXPERIMENTAL PROCEDURES

Cell Culture and Chemicals

Em-Myc;p19Arf�/� mouse B cell lymphomas were cultured in B cell medium

(45% DMEM/45% IMDM/10% FBS, supplemented with 2 mM L-glutamine

and 5mM b-mercaptoethanol). g-irradiated NIH 3T3 cells were used as feeder

cells. Focus cells were cultured in MEM with 10% FBS. HUVEC cells were

cultured in Endothelial Cell Growth Medium 2 (Lonza). Doxorubicin, Jak Inhib-

itor 1 and Ag490 mCF3 were purchased from Calbiochem. SB203580 and

KU55933 were purchased from Tocris Bioscience. Gefitinib was purchased

from LC labs. For in vivo studies, Ag490 m-CF3 was dissolved in DMSO and

then diluted 3:2 in DMEM plus 10% FBS.

Conditioned Media

Conditioned media was made from mouse tissues 18 hr after doxorubicin

treatment. Conditioned media for viability assays and cytokine arrays was

derived from organs from 3–4 pooled mice, while media for ELISAs was gener-

ated from individual mice. All tissues were dissociated manually in B cell

media. Thymic, bone marrow and lymph node conditioned media were condi-

tioned for 6 hr at 37�C. To isolate single-cell types from the thymus, tissue was

manual dissociated and washed two times in serum free DMEM, followed by

incubation for 1 hr at 37�C with Liberase (Roche, 1.3 Wunsch units/mL) and

Dnase I (0.15 mg/mL). To aid in dissociation, samples were manually pipetted

at 15 min intervals. Single-cell populations were sorted using FITC conjugated

antibodies to the following cell surface markers: CD45, CD19, CD11b, CD11c,

MHC II, CD31/CD34 for T cells, B cells, macrophages, dendritic cells, epithelial

and endothelial cells, respectively. Cells were plated and allowed to condition

media for 48 hr at 37�C and 5% CO2. All conditioned medias were cleared of

tissue and cells by centrifugation. All values shown for viability assays, ELISAs

and cytokine arrays are normalized to the weight of the dissected tissue or the

number of sorted cells. For the viability assays, conditioned medias were

diluted one to three. IL-6 ELISA kits were purchased from eBioscience. The

Timp-1 ELISA kit and mouse cytokine arrays were purchased from R&D

Biosystems.

In Vitro Viability, Competition, and Cell Growth Assays

For viability, competition and growth assays Em-Myc;p19Arf�/� lymphoma

cells were split into replicate wells of z500,000 cells in 24-well plates

or z125,000 cells in a 48-well plate. Every 24 hr, cultured cells were resus-

pended by pipeting and half of the culture was replaced with fresh medium.

Viability and cell number were determined by propidium iodide exclusion.

For the competition assay, lymphoma cells were partially infected with the indi-

cated retroviruses. The fold change for the competition assay is calculated by

dividing the percentage of GFP positive lymphoma cells in the treated popula-

tion by the percentage in untreated populations. Murine Timp-1 was

purchased from R&D Biosystems and used at 100ng/mL. All other cytokines

were purchased from Peprotech and used at 10ng/mL. Jak Inhibitor 1 was

used at a final concentration of 500nM, and Gefitinib was used at a final

concentration of 3 mM.

In Vivo Response to Chemotherapy

All mice were purchased from Jackson Laboratory. For survival assays, 1 3

106 Em-Myc;p19Arf�/� mouse lymphoma cells were injected by tail-vein injec-

tion into syngenic C57BL/6J, C57BL/6J IL-6�/� or C57BL/6J Rag1�/� mice.

Lymphoma burden was monitored by palpation of the axillary and brachial

lymph nodes. At the presentation of a substantial tumor burden (12–13 days

after injection), mice were treated with doxorubicin and/or Ag490 m-CF3.

Tumor free survival was monitored by palpation and in vivo GFP imaging using

a NightOwl imaging system (Berthold).

Thymic Rebound in Response to Radiation

Untreated 6 to 8 week old C57BL/6J or C57BL/6J IL-6�/� mice were sacrificed

to establish basal spleen and thymic weight. 6 to 8 week old C57BL/6J or

C57BL/6J IL-6�/� mice were irradiated with 4 or 5 Gray. 5 or 12 days later

all mice were sacrificed and the spleen and thymus were weighed.

Statistical Analysis

Statistical analysis was performed using GraphPad Prism4 software. Two-

tailed Student’s t tests were used, as indicated. Error bars represent mean ±

SEM. For comparison of survival curves, a Kaplan-Meier test was used.

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures and

five figures and can be found with this article online at doi:10.1016/j.cell.

2010.09.043.

ACKNOWLEDGMENTS

We thank Holly Criscione and Tyler Miller for their experimental assistance. We

thank Corbin Meacham for assistance with the cell migration assay. We would

also like to acknowledge Eliza Vasile in the Koch Institute Microscopy Core

Facility and Glen Paradis in the Koch Institute Flow Cytometry Core Facility

for advice and services. Roderick Bronson provided expert pathology anal-

ysis, and Justin Pritchard performed bioinformatic analysis of cytokine arrays.

We are grateful to Corbin Meacham, David Feldser, and Ross Dickins for crit-

ically reading the manuscript and the entire Hemann lab for helpful discus-

sions. M.T.H. is a Rita Allen Fellow and the Latham Family Career Development

Assistant Professor of Biology and is supported by NIH RO1 CA128803 and

Ludwig Center for Molecular Oncology at MIT. L.A.G. is supported by the

MIT Herman Eisen fellowship.

364 Cell 143, 355–366, October 29, 2010 ª2010 Elsevier Inc.

Page 53: Cell 101029

Received: November 25, 2009

Revised: March 31, 2010

Accepted: September 24, 2010

Published: October 28, 2010

REFERENCES

Acosta, J.C., O’Loghlen, A., Banito, A., Guijarro, M.V., Augert, A., Raguz, S.,

Fumagalli, M., Da Costa, M., Brown, C., Popov, N., et al. (2008). Chemokine

signaling via the CXCR2 receptor reinforces senescence. Cell 133, 1006–1018.

Adams, J.M., Harris, A.W., Pinkert, C.A., Corcoran, L.M., Alexander, W.S.,

Cory, S., Palmiter, R.D., and Brinster, R.L. (1985). The c-myc oncogene driven

by immunoglobulin enhancers induces lymphoid malignancy in transgenic

mice. Nature 318, 533–538.

Bleau, A.M., Hambardzumyan, D., Ozawa, T., Fomchenko, E.I., Huse, J.T.,

Brennan, C.W., and Holland, E.C. (2009). PTEN/PI3K/Akt pathway regulates

the side population phenotype and ABCG2 activity in glioma tumor stem-like

cells. Cell Stem Cell 4, 226–235.

Burgess, D.J., Doles, J., Zender, L., Xue, W., Ma, B., McCombie, W.R.,

Hannon, G.J., Lowe, S.W., and Hemann, M.T. (2008). Topoisomerase levels

determine chemotherapy response in vitro and in vivo. Proc. Natl. Acad. Sci.

USA 105, 9053–9058.

Coppe, J.P., Patil, C.K., Rodier, F., Sun, Y., Munoz, D.P., Goldstein, J., Nelson,

P.S., Desprez, P.Y., and Campisi, J. (2008). Senescence-associated secretory

phenotypes reveal cell-nonautonomous functions of oncogenic RAS and the

p53 tumor suppressor. PLoS Biol. 6, 2853–2868.

Corradini, P., Ladetto, M., Pileri, A., and Tarella, C. (1999). Clinical relevance of

minimal residual disease monitoring in non-Hodgkin’s lymphomas: a critical

reappraisal of molecular strategies. Leukemia 13, 1691–1695.

Delrez, M., Ikeh, V., Maisin, J.R., Mattelin, G., Haot, J., and Betz, E.H. (1978).

Influence of a mixture of chemical protectors on the lymphoid regeneration of

bone marrow and thymus in irradiated mice. Experientia 34, 1221–1222.

Dimri, G.P., Lee, X., Basile, G., Acosta, M., Scott, G., Roskelley, C., Medrano,

E.E., Linskens, M., Rubelj, I., Pereira-Smith, O., et al. (1995). A biomarker that

identifies senescent human cells in culture and in aging skin in vivo. Proc. Natl.

Acad. Sci. USA 92, 9363–9367.

Eckstein, N., Servan, K., Hildebrandt, B., Politz, A., von Jonquieres, G., Wolf-

Kummeth, S., Napierski, I., Hamacher, A., Kassack, M.U., Budczies, J., et al.

(2009). Hyperactivation of the insulin-like growth factor receptor I signaling

pathway is an essential event for cisplatin resistance of ovarian cancer cells.

Cancer Res. 69, 2996–3003.

Grivennikov, S., Karin, E., Terzic, J., Mucida, D., Yu, G.Y., Vallabhapurapu, S.,

Scheller, J., Rose-John, S., Cheroutre, H., Eckmann, L., and Karin, M. (2009).

IL-6 and Stat3 are required for survival of intestinal epithelial cells and

development of colitis-associated cancer. Cancer Cell 15, 103–113.

Gu, L., Zhuang, H., Safina, B., Xiao, X.Y., Bradford, W.W., and Rich, B.E.

(2005). Combinatorial approach to identification of tyrphostin inhibitors of

cytokine signaling. Bioorg. Med. Chem. 13, 4269–4278.

Hanahan, D., and Weinberg, R.A. (2000). The hallmarks of cancer. Cell 100,

57–70.

Heinrich, P.C., Behrmann, I., Muller-Newen, G., Schaper, F., and Graeve, L.

(1998). Interleukin-6-type cytokine signalling through the gp130/Jak/STAT

pathway. Biochem. J. 334, 297–314.

Hideshima, T., Mitsiades, C., Tonon, G., Richardson, P.G., and Anderson, K.C.

(2007). Understanding multiple myeloma pathogenesis in the bone marrow to

identify new therapeutic targets. Nat. Rev. Cancer 7, 585–598.

Holen, K.D., and Saltz, L.B. (2001). New therapies, new directions: advances

in the systemic treatment of metastatic colorectal cancer. Lancet Oncol. 2,

290–297.

Hoshida, Y., Villanueva, A., Kobayashi, M., Peix, J., Chiang, D.Y., Camargo, A.,

Gupta, S., Moore, J., Wrobel, M.J., Lerner, J., et al. (2008). Gene expression in

fixed tissues and outcome in hepatocellular carcinoma. N. Engl. J. Med. 359,

1995–2004.

Ignatiadis, M., Georgoulias, V., and Mavroudis, D. (2008). Micrometastatic

disease in breast cancer: clinical implications. Eur. J. Cancer 44, 2726–2736.

Jourdan, M., De Vos, J., Mechti, N., and Klein, B. (2000). Regulation of Bcl-

2-family proteins in myeloma cells by three myeloma survival factors: inter-

leukin-6, interferon-alpha and insulin-like growth factor 1. Cell Death Differ.

7, 1244–1252.

Krizhanovsky, V., Yon, M., Dickins, R.A., Hearn, S., Simon, J., Miething, C.,

Yee, H., Zender, L., and Lowe, S.W. (2008). Senescence of activated stellate

cells limits liver fibrosis. Cell 134, 657–667.

Kuilman, T., Michaloglou, C., Vredeveld, L.C., Douma, S., van Doorn, R.,

Desmet, C.J., Aarden, L.A., Mooi, W.J., and Peeper, D.S. (2008). Oncogene-

induced senescence relayed by an interleukin-dependent inflammatory

network. Cell 133, 1019–1031.

Lambert, E., Boudot, C., Kadri, Z., Soula-Rothhut, M., Sowa, M.L., Mayeux, P.,

Hornebeck, W., Haye, B., and Petitfrere, E. (2003). Tissue inhibitor of metallo-

proteinases-1 signalling pathway leading to erythroid cell survival. Biochem. J.

372, 767–774.

Medzhitov, R., and Horng, T. (2009). Transcriptional control of the inflamma-

tory response. Nat. Rev. Immunol. 9, 692–703.

Morrison, A.J., and Shen, X. (2005). DNA repair in the context of chromatin.

Cell Cycle 4, 568–571.

Muller-Hermelink, H.K., Sale, G.E., Borisch, B., and Storb, R. (1987). Pathology

of the thymus after allogeneic bone marrow transplantation in man. A histo-

logic immunohistochemical study of 36 patients. Am. J. Pathol. 129, 242–256.

Naugler, W.E., Sakurai, T., Kim, S., Maeda, S., Kim, K., Elsharkawy, A.M., and

Karin, M. (2007). Gender disparity in liver cancer due to sex differences in

MyD88-dependent IL-6 production. Science 317, 121–124.

Nguyen, D.X., Bos, P.D., and Massague, J. (2009). Metastasis: from dissemi-

nation to organ-specific colonization. Nat. Rev. Cancer 9, 274–284.

Rebouissou, S., Amessou, M., Couchy, G., Poussin, K., Imbeaud, S., Pilati, C.,

Izard, T., Balabaud, C., Bioulac-Sage, P., and Zucman-Rossi, J. (2009).

Frequent in-frame somatic deletions activate gp130 in inflammatory hepato-

cellular tumours. Nature 457, 200–204.

Rodier, F., Coppe, J.P., Patil, C.K., Hoeijmakers, W.A., Munoz, D.P., Raza,

S.R., Freund, A., Campeau, E., Davalos, A.R., and Campisi, J. (2009). Persis-

tent DNA damage signalling triggers senescence-associated inflammatory

cytokine secretion. Nat. Cell Biol. 11, 973–979.

Salgado, R., Junius, S., Benoy, I., Van Dam, P., Vermeulen, P., Van Marck, E.,

Huget, P., and Dirix, L.Y. (2003). Circulating interleukin-6 predicts survival in

patients with metastatic breast cancer. Int. J. Cancer 103, 642–646.

Seymour, J.F., Talpaz, M., Cabanillas, F., Wetzler, M., and Kurzrock, R. (1995).

Serum interleukin-6 levels correlate with prognosis in diffuse large-cell

lymphoma. J. Clin. Oncol. 13, 575–582.

Sfikakis, P.P., Gourgoulis, G.M., Moulopoulos, L.A., Kouvatseas, G., Theofilo-

poulos, A.N., and Dimopoulos, M.A. (2005). Age-related thymic activity in

adults following chemotherapy-induced lymphopenia. Eur. J. Clin. Invest.

35, 380–387.

Sharpless, N.E., and Depinho, R.A. (2006). The mighty mouse: genetically

engineered mouse models in cancer drug development. Nat. Rev. Drug

Discov. 5, 741–754.

Trikha, M., Corringham, R., Klein, B., and Rossi, J.F. (2003). Targeted anti-

interleukin-6 monoclonal antibody therapy for cancer: a review of the rationale

and clinical evidence. Clin. Cancer Res. 9, 4653–4665.

Visvader, J.E., and Lindeman, G.J. (2008). Cancer stem cells in solid tumours:

accumulating evidence and unresolved questions. Nat. Rev. Cancer 8,

755–768.

Wajapeyee, N., Serra, R.W., Zhu, X., Mahalingam, M., and Green, M.R. (2008).

Oncogenic BRAF induces senescence and apoptosis through pathways medi-

ated by the secreted protein IGFBP7. Cell 132, 363–374.

Williams, R.T., den Besten, W., and Sherr, C.J. (2007). Cytokine-dependent

imatinib resistance in mouse BCR-ABL+, Arf-null lymphoblastic leukemia.

Genes Dev. 21, 2283–2287.

Cell 143, 355–366, October 29, 2010 ª2010 Elsevier Inc. 365

Page 54: Cell 101029

Wong, V.W., Yu, J., Cheng, A.S., Wong, G.L., Chan, H.Y., Chu, E.S., Ng, E.K.,

Chan, F.K., Sung, J.J., and Chan, H.L. (2009). High serum interleukin-6 level

predicts future hepatocellular carcinoma development in patients with chronic

hepatitis B. Int. J. Cancer 124, 2766–2770.

Xenidis, N., Ignatiadis, M., Apostolaki, S., Perraki, M., Kalbakis, K., Agelaki, S.,

Stathopoulos, E.N., Chlouverakis, G., Lianidou, E., Kakolyris, S., et al. (2009).

Cytokeratin-19 mRNA-positive circulating tumor cells after adjuvant chemo-

therapy in patients with early breast cancer. J. Clin. Oncol. 27, 2177–2184.

Xue, W., Zender, L., Miething, C., Dickins, R.A., Hernando, E., Krizhanovsky,

V., Cordon-Cardo, C., and Lowe, S.W. (2007). Senescence and tumour clear-

ance is triggered by p53 restoration in murine liver carcinomas. Nature 445,

656–660.

366 Cell 143, 355–366, October 29, 2010 ª2010 Elsevier Inc.

Page 55: Cell 101029

ATR-X Syndrome Protein Targets TandemRepeats and Influences Allele-SpecificExpression in a Size-Dependent MannerMartin J. Law,1,8 Karen M. Lower,1,8 Hsiao P.J. Voon,1 Jim R. Hughes,1 David Garrick,1 Vip Viprakasit,3

Matthew Mitson,1 Marco De Gobbi,1 Marco Marra,7 Andrew Morris,4 Aaron Abbott,4 Steven P. Wilder,5

Stephen Taylor,2 Guilherme M. Santos,6 Joe Cross,1 Helena Ayyub,1 Steven Jones,7 Jiannis Ragoussis,4

Daniela Rhodes,6 Ian Dunham,5 Douglas R. Higgs,1 and Richard J. Gibbons1,*1Medical Research Council Molecular Haematology Unit2Computational Biology Research Group

Weatherall Institute of Molecular Medicine, John Radcliffe Hospital, Headington, Oxford OX3 9DS, UK3Department of Paediatrics, Faculty of Medicine, Siriaj Hospital, Mahidol University, Bangkok 10700, Thailand4The Wellcome Trust Centre for Human Genetics, Roosevelt Drive, Oxford OX3 7BN, UK5European Bioinformatics Institute, Wellcome Trust Genome Campus, Hinxton, Cambridge CB10 1SD, UK6Structural Studies Division, Medical Research Council, Laboratory of Molecular Biology, Cambridge CB2 0QH, UK7BCCA Genome Sciences Centre, University of British Columbia, Vancouver, British Columbia V6T 1Z4, Canada8These authors contributed equally to this work

*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.09.023

SUMMARY

ATRX is an X-linked gene of the SWI/SNF family,mutations in which cause syndromal mental retarda-tion and downregulation of a-globin expression.Here we show that ATRX binds to tandem repeat(TR) sequences in both telomeres and euchromatin.Genes associated with these TRs can be dysregu-lated when ATRX is mutated, and the change inexpression is determined by the size of the TR, pro-ducing skewed allelic expression. This reveals thecharacteristics of the affected genes, explains thevariable phenotypes seen with identical ATRX muta-tions, and illustrates a new mechanism underlyingvariable penetrance. Many of the TRs are G richand predicted to form non-B DNA structures (in-cluding G-quadruplex) in vivo. We show that ATRXbinds G-quadruplex structures in vitro, suggestinga mechanism by which ATRX may play a role invarious nuclear processes and how this is perturbedwhen ATRX is mutated.

INTRODUCTION

Although it is known that proteins of the Swi/Snf family are

required to facilitate a wide range of nuclear processes (e.g.,

replication, recombination, repair, transcription), the mecha-

nisms by which they operate in vivo are poorly understood (Flaus

et al., 2006). One such widely expressed protein (ATRX) was first

identified when it was shown that mutations in the X-linked gene

(ATRX) caused a form of syndromal mental retardation, with

multiple developmental abnormalities characteristically associ-

ated with a thalassaemia (ATR-X syndrome) (Gibbons et al.,

1995). To date 127 disease-causing mutations have been found,

most of which are located in two highly conserved domains of

the ATRX protein (Gibbons et al., 2008). At the N terminus these

lie within a globular domain (similar to that found in DNMT3 and

DNMT3L, the so-called ADD domain) including a plant homeo-

domain (PHD), which most probably binds the N-terminal tails

of histone H3 (Argentaro et al., 2007). At the C terminus there

are seven helicase subdomains that identify ATRX as a member

of the SNF2 family of chromatin-associated proteins (Figure 1A).

Although many of these proteins have been shown to remodel,

remove, or slide nucleosomes using in vitro assays, ATRX is

most closely related to a subgroup (including RAD54 and

ARIP4) that, despite acting as ATP-driven molecular motors,

perform poorly in such canonical assays, suggesting that they

have related but different chromatin-associated functions (Xue

et al., 2003 and unpublished data).

Some clues to the role of ATRX in vivo have come from

studying its distribution in the nucleus, the proteins with which

it interacts, and the effects of mutations. Using indirect immuno-

fluorescence, ATRX is found at heterochromatic repeats, at

rDNA repeats, at telomeric repeats, and within PML bodies,

which themselves are often associated with heterochromatic

structures including telomeres (Gibbons et al., 2000; McDowell

et al., 1999; Xue et al., 2003). Two robust protein-protein interac-

tions have been described. The first occurs with DAXX (Xue et al.,

2003) (a protein that is also found in PML bodies), which has

been implicated in both pro- and antiapoptotic pathways. The

second interaction occurs with HP1a and HP1b, proteins that

are widely associated with heterochromatin, including the telo-

mere (Berube et al., 2000). It has also been shown that mutations

in ATRX are consistently associated with alterations in the

Cell 143, 367–378, October 29, 2010 ª2010 Elsevier Inc. 367

Page 56: Cell 101029

pattern of DNA methylation at such repeat sequences (rDNA,

interstitial heterochromatic repeats, and subtelomeric repeats)

(Gibbons et al., 2000).

Recently, an important link has been established between

these observations and more functional studies. First, it has

been shown that ATRX and HP1 localize to the telomeres of

chromosomes in mouse embryonic stem cells (ESCs) (Wong

et al., 2010). Second, it has been shown that ATRX localizes to

telomeres in synchrony with the histone variant H3.3. Using

immunoprecipitation it was shown that ATRX and its partner

DAXX specifically interact with H3.3, which is found to be asso-

ciated with both active and inactive genes, regulatory elements,

and telomeres (Goldberg et al., 2010). It has recently been shown

that DAXX is an H3.3-specific chaperone (Drane et al., 2010;

Lewis et al., 2010), and in the absence of ATRX, H3.3 is no longer

recruited to telomeres whereas recruitment to the interstitial sites

that were analyzed appeared to be unaffected (Goldberg et al.,

2010). These observations suggest that ATRX plays an important

role in establishing or maintaining the chromatin environment of

telomeres and subtelomeric regions where it facilitates histone

A B

C D

E

Figure 1. Validation of ATRX ChIP Protocol

(A) Immunoblots of protein extracts from ATR-X patient and normal control lymphoblastoid cell lines using ATRX N- and C-terminal antibodies. The ATR-X patient

harbors an ATRX C-terminal deletion mutation affecting the C-terminal antibody epitope. Schematic diagram of ATRX shows protein isoforms, antibody epitope

regions, and conserved domains.

(B) The ATRX C-terminal antibody crosslinked to protein A-Sepharose was used to immunopurify ATRX from EBV cells. Eluted protein was analyzed by western

blot probed with the N-terminal mouse monoclonal ATRX antibody, 39f. The mock control lane contains sample immunopurified using normal rabbit IgG.

(C) Q-PCR analysis of ATRX ChIP at the major ribosomal RNA gene locus in erythroblast (n = 4) and Hep3B (n = 3). Error bars show standard deviations. Diagram

of the ribosomal RNA gene locus shows positions of rRNAs (red boxes), the promoter (arrow), and the Q-PCR primers (boxes above line).

(D) Direct mapping of human ATRX, SCL, and YY1 ChIP-seq reads to simple and interspersed repeats. Selected representative data are shown. For the complete

dataset, see Table S1.

(E) Direct mapping of ATRX ChIP-seq sequence reads to mouse simple and interspersed repeats.

See Figure S1 for further validation of the specificity of the ATRX ChIP.

368 Cell 143, 367–378, October 29, 2010 ª2010 Elsevier Inc.

Page 57: Cell 101029

replacement with the H3.3 variant (Drane et al., 2010; Lewis

et al., 2010).

Although these observations have provided new insight into

the potential role of ATRX at heterochromatic regions of the

genome, they have not identified the euchromatic targets of

ATRX and have not addressed the role of ATRX in regulating

gene expression. To date the only human genes whose expres-

sion is known to be affected by ATRX mutations lie in the

a-globin gene cluster (Gibbons et al., 1991). Although clearly

related to the b-globin cluster throughout evolution, ATRX muta-

tions do not affect b-globin expression. It has been noted that the

structure (e.g., GC content, repeat density, gene density),

nuclear organization (e.g., nuclear position, relationship to

chromosome territory, relationship to heterochromatin), and

epigenetic environment (e.g., timing of replication, chromatin

modification, DNA methylation) associated with these two

clusters are radically different (Higgs et al., 1998). Most notably

the human a-globin cluster lies very close to the telomere of

chromosome 16. It has previously been suggested that ATRX

is targeted to specific regions of the genome defined by their

genomic organization and/or chromatin structure. Thus muta-

tions in ATRX may affect one type of chromosomal region

(e.g., containing the a-globin genes) but not another (e.g., con-

taining the b-globin genes).

Here we have established the genome-wide distribution of the

ATRX protein in both mouse and human cells. We have confirmed

that ATRX binds directly to mouse telomeres and also shown that

ATRX is enriched at the telomeres and subtelomeric regions of

human chromosomes. Chromatin immunoprecipitation (ChIP)

sequencing identified 917 targets in primary human erythroid

cells (in which the globin genes are expressed) and 1305 targets

in mouse ESCs. The most prominent feature of the targets in both

human and mouse is the presence of variable number tandem

repeats (VNTRs), which in many (but not all) cases are G and C

rich and contain a high proportion of CpG dinucleotides. Of

particular interest we show that, when ATRX function is compro-

mised in ATR-X syndrome, the degree of perturbation in gene

expression is related to the size of the TR, and this may lead to

monoallelic expression. These findings explain the variable

phenotypes seen in patients with identical ATRX mutations and

provide a new mechanism underlying variable penetrance. A

common theme shared by telomeres and many of the subtelo-

meric targets of ATRX is their potential to form G-quadruplex

(G4) DNA structures. Here we show that ATRX binds G4 DNA

in vitro, suggesting a common mechanism by which ATRX may

influence a wide range of nuclear processes in the telomeric, sub-

telomeric, and interstitial regions of mammalian chromosomes.

RESULTS

Validation of an ATRX ChIP Protocolwith rDNA as a TargetDomain structure, interaction partners, and biochemical activity

currently implicate ATRX in the regulation of transcription via a

physical interaction with chromatin. To date, ATRX has been

implicated in histone H3.3 deposition at telomeres, but little is

known about ATRX function away from telomeres because no

direct ATRX target genes have been described. To address

this, an ATRX ChIP assay was developed using the ribosomal

gene loci (rDNA) as the first candidate targets. The rDNA loci

were chosen because immunofluorescence studies have previ-

ously shown that, in mitotic cells, ATRX is consistently found

on the short arms of the acrocentric chromosomes in human

colocalizing with the rDNA loci (McDowell et al., 1999); rDNA

also becomes hypomethylated at CpG dinucleotides in primary

peripheral blood mononuclear cells (PBMCs) from patients

with ATR-X syndrome (Gibbons et al., 2000).

ChIP analysis was performed with an ATRX antibody that

recognizes a C-terminal epitope only present in the full-length

ATRX isoform (Figure 1A). Western blot was used to confirm

that this antibody immunoprecipitates ATRX with detection

using an independent antibody (Figure 1B). ATRX ChIP enrich-

ment at rDNA was measured in primary erythroblasts and

Hep3B cells (Figure 1C). Consistent with its ubiquitous expres-

sion profile, ATRX binds rDNA in both cell types tested. It was

of interest that the maximal binding of ATRX occurs at the

transcribed region of the locus that is very rich in G and CpG

nucleotides. These observations confirm the specificity of the

ATRX C-terminal antibody, validate the ChIP assay, and identify

the ribosomal genes as direct ATRX targets.

ATRX Binds G-Rich Telomeric and SubtelomericRepetitive DNAHaving validated the ATRX ChIP protocol, we next addressed

whether, in addition to rDNA, other putative targets (heterochro-

matic repeats) identified by indirect immunofluorescence were

similarly bound by ATRX. To accomplish this, we took a ChIP-seq

approach using Illumina high-throughput, short read sequencing

to analyze primary human erythroid cells and mouse ESCs.

ATRX ChIP DNA from human primary erythroid cells was se-

quenced alongside sonicated input DNA as a control. The short

read mapping protocol used for ChIP sequencing (see below)

routinely discards nonunique genomic matches, precluding

analysis of direct binding to repeat sequences. To overcome

this, we interrogated the ATRX ChIP read library for perfect se-

quence matches to a variety of tandem and interspersed repeat

sequences. As a negative control, we used ChIP-seq data for

YY1 and SCL, transcription factors that have no known role at

heterochromatic repeats. YY1 and SCL ChIP DNA both showed

low enrichment of telomeric and nontelomeric satellite se-

quences (Figure 1D and Table S1 available online). ATRX ChIP

DNA showed striking enrichments for the G-rich telomeric

(TTAGGG)n repeats (�16-fold relative to input DNA) and telo-

mere-associated repeats (�10-fold relative to input) (Figure 1D

and Table S1). Similar results were obtained from the analysis

of ChIP-seq data from mouse ESCs (Figure 1E and Table S1).

Further confirmation of the specificity of the ATRX ChIP was

demonstrated by showing that ATRX enrichment was abolished

when ChIP was performed in mouse ESCs in which full-length

ATRX was knocked out (Figure S1).

These data therefore show that previously described immuno-

fluorescence studies reflect the binding of ATRX to telomeric

and subtelomeric repeat sequences. The presence of ATRX at

the subtelomeric TAR1 repeats is consistent with previous

observations that DNA methylation at subtelomeric repeats is

altered in patients with ATR-X syndrome (Gibbons et al., 2000).

Cell 143, 367–378, October 29, 2010 ª2010 Elsevier Inc. 369

Page 58: Cell 101029

Genome-wide Targets of ATRX Include CpG Islandsand G-Rich Tandem RepeatsHaving established that ATRX binds G-rich repetitive elements

associated with rDNA, telomeres, and subtelomeric repeats,

ATRX ChIP and input sequence reads were aligned to the

genome if five or fewer matches were detected (allowing for

three base-pair mismatches). Peak calling was performed on

the ATRX ChIP-seq alignments using an input correction penalty

to deplete peaks overlying enrichments of input reads. The input

correction penalty effectively eradicated many peaks overlying

DNA where there were differences in copy number between

the reference genome and the sequenced genome (e.g., at

pericentromeric satellite DNA).

Using these criteria in primary human erythroid cells we

identified 917 ATRX-binding sites genome-wide. The ChIP

enrichment at 14 sites (chosen to represent the different classes

of targets discussed below) was validated using Q-PCR. ATRX

binding at most of these sites was enriched above background

levels 10/14 (false discovery rate 4/14; Figure S2A). Of the 917

ATRX peaks called, approximately a third (324) were intergenic,

a third were present at promoter regions (326), and a third were in

the bodies of genes (267) (Figure 2A). All peaks were then

Promoter (326)

Intergenic (324)

Gene Body (267)

Total 917

Promoter (78)

Intergenic (771)

Gene Body (456)

Total 1305

Human Mouse

0

10

20

30

40

50

60

70

80

Human Mouse

Tandem Repeat Peaks

CpGi Peaks

% o

f P

ea

ks

A

B

qtel

Relative distribution across chromosome arms

C

1 10.8 0.6 0.2 0.6

Centromere Telomere

Relative distribution across chromosome arm

0 10.2 0.4 0.6 0.8

Mouse

ptel

Human

0.4 0.2 0 0.4 0.8

25

30

35

40

45

50

55

Me

an

T

an

de

m R

ep

ea

t

%G

C C

on

te

nt

Nu

mb

er o

f P

ea

ks

0.8 0.85 0.9 0.95 1

0

5

10

15

20

25

Nu

mb

er o

f P

ea

ks

25

30

35

40

45

50

55

0.8 0.85 0.9 0.95 1

Me

an

T

an

de

m R

ep

ea

t

%G

C C

on

te

nt

0

5

10

15

20

25

Figure 2. Genome-wide Comparison of Human and Mouse ATRX-Binding Site Characteristics

(A) Pie charts show the location of human and mouse ATRX-binding sites relative to genes.

(B) The proportion of human and mouse ATRX peaks overlapping with the two most common classes of human ATRX-binding sites, TRs and CpG islands (CpGi).

See also Figure S2C for genomic features associated with peaks.

(C) Ideograms showing the relative distribution of ATRX-binding sites across all human and mouse chromosomes. Each column represents the total number of

ATRX peaks within nonoverlapping 1/500 divisions of all chromosome arms. The zoomed panels show the telomeric region, overlayed with the mean %G+C

content of all tandem repeats throughout the same regions, for the respective human and mouse chromosomes. The sharp peak of subtelomeric targets in mouse

represents clusters of (TTAGGG)n adjacent to the telomeres of a subset of mouse chromosomes. See also Figure S2E for the distribution of TRs and Refseq genes

near telomeres.

See also Figure S2A for validation of targets by Q-PCR, Figure S2B for examples of ATRX-binding sites, Figure S2D for trinucleotide content of DNA sequence

underlying peaks, Figure S3A for histone modifications associated with peaks, and Figure S3B for histone H3.3 distribution associated with ATRX peaks.

370 Cell 143, 367–378, October 29, 2010 ª2010 Elsevier Inc.

Page 59: Cell 101029

examined for overlap with annotated genomic sequence

features (Figures S2B and S2C). Two striking observations arise

from this analysis: first, irrespective of location relative to genes,

human ATRX-binding sites commonly coincide with CpG islands

(Figure 2B); second, the predominant sequence feature that

ATRX binds in gene bodies and intergenic regions is tandem

repetitive DNA (Figure 2B and Figure S2C). Analysis of ATRX

binding in mouse ESCs (Figure 2A) identified a larger number

of ATRX targets (1305) and showed a similar enrichment at

TRs but less so at CpG islands (Figure 2B) (which occur much

less frequently in the mouse genome) (Waterston et al., 2002).

As the tandem repetitive ATRX targets at rDNA and telomeres

are G rich, we reasoned that this might be a common property of

other ATRX-bound TRs. To test this we calculated the tri-nucle-

otide sequence content of ATRX-bound tandem repetitive

targets. ATRX-bound TRs in both mouse and human are signifi-

cantly enriched for G and C and CpG, and they are depleted in

A- and T-containing trinucleotides relative to randomly selected

control repeats (Figure S2D and data not shown).

These findings are consistent with the observation that in

human cells, many ATRX-bound promoters are associated

with CpG islands. Genome-wide analysis (in human) showed

that there are no chromatin modifications consistently associ-

ated with binding of ATRX. Chromatin marks found at the

promoter and intragenic and intergenic binding sites show the

characteristic chromatin modifications associated with such

features (Figure S3A). Together the data suggest that ATRX

interacts predominantly with G and C and CpG-rich sequences

contained within TRs and promoters.

The Distribution of ATRX-Binding Sites Differsbetween Human and Mouse, Reflecting the DifferentDistributions of G-Rich Tandem RepeatsInitial analysis of the human ChIP-seq data suggested that ATRX

targets may be clustered at subtelomeric regions of the genome

(Figure 1D). This was confirmed when the proportions of ATRX-

binding sites were plotted as a function of their distance from the

nearest telomere (pooling data for all telomeres) (Figure 2C).

However, it has previously been shown that in humans, GC

content, CpG density, G-rich minisatellites, and gene density

are all increased in subtelomeric regions of the genome, and

this was confirmed here (Figure 2C and Figure S2F). In fact,

the distribution of ATRX targets in humans appears largely to

reflect the increase in GC content and G-rich TRs observed

toward telomeres rather than increased gene or general TR

density (Figure 2C and Figure S2E).

To explore this further, we compared the data from human

with those from mouse, a species with less extremes of GC

content and a different distribution of G-rich repeats (Waterston

et al., 2002). In mouse, the GC content of TRs is not increased

toward telomeres but is more evenly distributed across each

chromosome (Figure 2C). Although the majority of mouse targets

are associated with CpG islands or TRs (as in human), the mouse

ATRX targets are less concentrated at telomeres (Figure 2C).

This more even distribution of ATRX targets in mouse is consis-

tent with the more even distribution of GC content and G-rich

repeats in mouse compared to human (Figure 2C). These find-

ings focus attention on the fact that ATRX appears to bind

many G-rich TRs in different chromosomal environments rather

than genes within subtelomeric regions per se.

Analysis of H3.3 Distribution in the Absence of ATRXTelomeres are a site of rapid nucleosomal turnover as demon-

strated by the incorporation of histone H3.3 (Goldberg et al.,

2010). Furthermore, it has recently been shown that ATRX

recruits the histone H3.3-specific chaperone DAXX and facili-

tates H3.3 deposition at telomeres and pericentric DNA (Drane

et al., 2010; Lewis et al., 2010). In order to see whether H3.3 co-

localized with ATRX at its target sequences (predominantly TRs,

Figure 2B), data for the H3.3 distribution in mouse ESCs (Gold-

berg et al., 2010) were reanalyzed to determine the distribution

of H3.3 at ATRX-binding sites (Figure S3B). Peaks of H3.3 are

observed at genic and intergenic ATRX sites. ATRX has previ-

ously been shown to be required for H3.3 deposition at telo-

meres but not at promoters and transcription factor-binding sites

(Goldberg et al., 2010). In order to see if the H3.3 distribution

at these sites is dependent on ATRX, the patterns of H3.3 for

Atrxflox and Atrxnull mouse ESCs were compared. The distribu-

tion of H3.3 is only subtly perturbed at ATRX-binding sites in

gene bodies and intergenic sites (Figure S3B) with a slight dimi-

nution of the peak and increased signal in the adjacent

sequence. If ATRX is required for H3.3 incorporation it may be

only at a subset of these targets.

Analysis of Expression of ATRX Targets when ATRXIs MutatedAlthough we initially identified the human ATRX targets in

erythroid cells, because many of the affected genes are widely

expressed, we compared their expression in Epstein-Barr virus

(EBV)-transformed lymphocytes from normal individuals (n = 19)

with expression in EBV cells from individuals harboring natural

mutations in the ATRX gene (n = 23). Twenty ATRX targets

(expressed in EBV-transformed lymphocytes) were chosen for

analysis, including 9 ATRX promoter-binding targets and 11

tandem repetitive gene body targets. Four ATRX targets were

significantly altered in expression in ATR-X patients relative to

normal controls: NME4, SLC7A5, and RASA3 were downregu-

lated, whereas GAS8 was upregulated (Figure 3). Interestingly

all four novel targets contained tandem repetitive ATRX-binding

sites, whereas none of the nonrepetitive, promoter-binding site

target genes was affected. These data suggest that when

ATRX alters gene expression, this involves an interaction with

TRs associated with its target genes.

ATRX Exerts an Effect on Target Gene Expression viaan Interaction with G-Rich RepeatsTo examine the role of ATRX in regulating gene expression in

detail, we analyzed the subtelomeric region of chromosome 16

(16p13.3), which contains two ATRX targets (a-globin and

NME4), both of which are downregulated in ATR-X syndrome.

ChIP-seq analysis of this area was confirmed by ChIP-chip anal-

ysis (Figures 4A and 4B and Figure S4A). With this approach,

three consistent peaks of ATRX binding were seen in primary

erythroid cells. A small but reproducible enrichment was seen

at the probe closest to the telomere (telomeric repeats were

not included on the array). In erythroid cells, a broad region of

Cell 143, 367–378, October 29, 2010 ª2010 Elsevier Inc. 371

Page 60: Cell 101029

enrichment was seen across all the a-like globin genes with

maximum binding just upstream of the HBM globin gene. A third

peak was seen at the gene encoding a nucleoside kinase, NME4

(Figure 4A and Figure S4A). When we used Q-PCR (Figures S4B

and S4C), we noted that all peaks of ATRX binding localized at or

very close to regions of G-rich tandemly repetitive DNA. The sub-

telomeric peak shows an enrichment lying �150 bp from the

start of the telomeric satellite repeats (TTAGGG)n (Figure S4B).

The maximum peak of binding in the a-globin locus lies within

a VNTR (CGCGGGGCGGGGG)n 1 kb upstream from the HBM

promoter, called jz VNTR (Figures S4B and S4C). The peak at

NME4 is centered on an imperfect VNTR (CCCGG

CCCCCCCA)n within the first intron of the gene (Figures S4B

and S4C).

It has been previously shown that expression of RNA from the

HBA1 and HBA2 globin genes is downregulated in patients with

the ATR-X syndrome (Wilkie et al., 1990). However, maximal

ATRX binding occurs not at the HBA genes but in close proximity

to the HBM and NME4 genes. We therefore took an unbiased

approach using RT-PCR to measure expression of all 16 genes

in the 500 kb region in normal individuals (n = 19) and those

proven to have ATR-X syndrome (n = 20) (Figure 4C). Globin

gene expression was analyzed using cDNA derived from

erythroid cells, and other genes were analyzed using cDNA

from EBV cell lines (nonglobin mRNAs are of very low abundance

in erythrocytes). The most consistently and severely downregu-

lated genes (HBM and NME4) were those associated with the

greatest peaks of ATRX enrichment (Figures 4B and 4C). It was

of interest that other significantly downregulated genes (HBA2,

HBA1, HBQ, and DECR2) lie adjacent to these severely affected

genes. Furthermore, the degree of downregulation of each a-like

globin (HBM > HBA2 = HBA1 > HBQ) gene is related to its

proximity to the major peak of ATRX binding 1 kb upstream

from the HBM gene.

This observation explains the a thalassaemia seen in ATR-X

syndrome and why a-globin and not b-globin expression is

perturbed, as only the former locus is associated with G-rich

VNTRs (Higgs et al., 1998).

The Perturbation in Gene Expression Is Relatedto the Size of the Associated Tandem RepeatIn ATR-X syndrome, a-globin RNA expression is often downre-

gulated, but affected individuals show different degrees of

repression (Figure 4C). This gives rise to different degrees of

a thalassaemia and is reflected by varying proportions of red

cells containing HbH inclusions, ranging from 0%–30%

(Gibbons et al., 2008). Importantly, such variation is seen

between individuals with the same ATRX mutation (Figure S5A)

and occurs both within and between affected families. How-

ever, for any individual, the level of HbH is relatively constant

throughout life. If the downregulation of a-globin expression in

ATR-X syndrome resulted from a negative effect due to a TR

then one might predict that the effect would be more extreme

when the repeat is increased in size. The jz VNTR is highly poly-

morphic. The size of the TR alleles was measured in 43 ATR-X

individuals, and the average size in an individual was plotted

against the level of HbH inclusions observed. A significant

correlation (r value = 0.58; p = 0.0002) was seen between the

level of inclusions (reflecting the degree of a thalassaemia)

and the size of the TR (Figure 5A and Figure S5B). jz VNTR

lies within a block of linkage disequilibrium (Figure S5C and

Table S3); polymorphisms within this block also show a correla-

tion with the number of cells containing HbH inclusions. In

contrast with jz VNTR, another VNTR within this block,

30HVR, showed a low correlation between size and the severity

of a thalassaemia (Figure S5B). Given the rapid evolution of

VNTRs relative to the background haplotype, the strong corre-

lation associated with the jz VNTR strongly supports the

Figure 3. Dysregulation of ATRX Targets Genes

Q-PCR analysis of gene expression of ATRX ChIP target genes in ATR-X patient (n = 21) and normal control (n = 19) lymphoblastoid cDNAs. Gray dots represent

control samples. Black dots represent genes unaffected in patient samples. Data are normalized to the mean values of the control samples. Black bars represent

mean values. Red dots show genes affected in patient samples. p values are for a two-tailed Student’s t test. The presence of a TR or CpG island underling the

ATRX-binding sites is indicated.

372 Cell 143, 367–378, October 29, 2010 ª2010 Elsevier Inc.

Page 61: Cell 101029

A

B

C

Figure 4. ATRX Interacts with the a-Globin Locus and Influences Gene Expression

(A) Microarray analysis (black bars) of ATRX ChIP DNA enrichment across the 500 kb terminal region of chromosome 16p containing the a-globin genes

and surrounding ubiquitously expressed genes. ATRX ChIP DNA from erythroblasts (n = 4), fibroblasts (n = 1), and Hep3B (n = 2) cells were analyzed as well

as erythroblasts immunoprecipitated with control IgG (n = 2). Representative data are shown. See Figure S4A for full dataset for erythroblasts.

(B) ChIP-seq analysis of erythroblast ATRX ChIP and input DNA using Illumina short-read sequencing. Graphs are a 50 bp sliding window of mapped reads.

Black bars show peak calls.

(C) Q-PCR analysis of gene expression across the a-globin gene locus in ATR-X patient (n = 20) and normal control (n = 19) cDNAs (from erythroid cells for the

globin genes or lymphoblastoid cells for other genes). Expression was measured relative to GAPDH and the mean expression values for the normal controls were

set to 100%. Red bars represent means of ATR-X patient expression and p values are for a two-tailed t test. See Figures S4B and S4C for validation of targets by

Q-PCR and mapping of peaks to G-rich VNTRs.

Cell 143, 367–378, October 29, 2010 ª2010 Elsevier Inc. 373

Page 62: Cell 101029

proposal that it is directly responsible for the variability seen in

the level of HbH inclusions.

The effect of TR size was further examined at the NME4 locus.

Again the TR is highly polymorphic; in this case the presence

of an expressed A/G single-nucleotide polymorphism (SNP) al-

lowed us to determine the effect of the TR size on allele-specific

expression. In ATR-X cases informative for the expressed SNP,

the most downregulated allele is always in cis with the larger TR

(Figures 5B and 5C). In some cases the expression was monoal-

lelic (Figure 5D).

ATRX Targets Have the Potential to Form G4 DNA,and ATRX Binds to G4 DNA StructuresTandem repetitive sequences can take up a range of non-B DNA

conformations (reviewed in Bacolla and Wells, 2009). G-rich se-

quences such as telomeres, rDNA, G-rich TRs, as well as CpG

islands can form abnormal DNA structures in vitro referred to

as G-quadruplex (G4) under physiological conditions (reviewed

in Lipps and Rhodes, 2009). These structures form in G-rich

sequences that contain four tracts of at least three guanines

separated by other bases and are stabilized by G-quartets that

form between four DNA strands held together by Hoogsteen

hydrogen bonds. Such structures are particularly likely to form

when DNA becomes single stranded, for example during replica-

tion and transcription, and may interfere with these nuclear

processes.

To explore the possibility that ATRX targets might form G4

structures in vivo, a genome-wide bioinformatic analysis using

Quadparser was performed to identify regions that have the

potential to form G4 DNA (Huppert and Balasubramanian,

2005). Fifty percent of ATRX peaks were found to overlap with

putative quadruplex sequences (PQSs) (Figure 6A). Given the

difficulty sequencing G-rich repeats and their consequent

contraction in the reference genome, it is possible that PQSs

are under-called in this analysis.

The potential for an ATRX-binding site to form G4 was further

examined using circular dichroism (Paramasivan et al., 2007).

The NME4 TR is predicted to form G4. A 31 bp oligonucleotide

0.1

1

10

genomic cDNA genomic cDNA

controls ATR-X

Ratio

A:G

allele

A

C

D

B

500bp

XhoI

genomic

cDNA

control ATR-X ATR-X

A alleleG allele

1000bp800bp

500bp

geno

mic

cDNA

geno

mic

cDNA

geno

mic

cDNA

VNTR

330

350

370

390

410

430

450

470

0.0

06-0

.8

0.8

-1.8

2.0

-5.0

5.0

-6.8

7.1

-15.5

16-4

7

% Haemoglobin H +ve cells

VN

TR

le

ng

th

/ b

as

e p

airs

A/G

Figure 5. ATRX-Binding Variable Number

Tandem Repeats Act as Length-Dependent

Negative Regulators of Gene Expression

When ATR-X Is Mutated

(A) jz VNTR length was measured in ATR-X

patients with a thalassaemia (n = 42) using PCR

and agarose gel electrophoresis and plotted

against the degree of a thalassaemia as measured

by % red cells showing Haemoglobin H inclusions.

See Figure S5B to compare correlation of VNTR

size and % red cells showing Haemoglobin H

inclusions for jz VNTR and 30HVR. Spearman

ranked correlation r value = 0.58, p value =

0.0002. See also Figure S5A for variable severity

of a thalassaemia in ATR-X syndrome, see Figur-

e S5C and Table S3 for a-globin locus haplotype

and linkage analysis.

(B) Q-PCR-based allelic discrimination assay was

used to determine the ratios of each NME4 allele

present in both genomic DNA and cDNA from

controls and ATR-X patients. The y axis is the ratio

of A:G allele (SNP rs14293), shown on a logarithmic

scale. For control cDNA samples, the ratio of A:G

allele expression is 0.70 to 1.30, mean = 1.0,

n = 13. For ATR-X cDNA samples, the ratio of

A:G allele expression is 0.24 to 2.37,

mean = 0.84, n = 17. F-test p value = 5.74 3

10�5. For the green datapoint, the larger VNTR is

linked to the G allele; for the red datapoints, the

larger VNTR is linked to the A allele. For the blue

datapoint, alleles could not be discriminated

based on VNTR size.

(C) Schematic representation of the exon/intron

structure of NME4. White boxes represent exons.

PCR amplicons are shown as generated from

genomic DNA and cDNA. The presence of a poly-

morphic XhoI site generated by SNP rs14293 in

NME4 exon 4 is shown, which allows allelic discrim-

ination by PCR amplification followed by an XhoI

restriction digest assay, the restriction site being

present in the G allele and abolished in the A allele.

(D) Results show monoallelic expression of NME4

in two individuals with ATR-X syndrome.

374 Cell 143, 367–378, October 29, 2010 ª2010 Elsevier Inc.

Page 63: Cell 101029

representing the repeat unit of the NME4 TR was incubated in

conditions that favor G4 DNA formation. The circular dichroism

spectrum was obtained (Figure 6B). A positive ellipticity

maximum was observed at 260 nm and a negative ellipticity

minimum at 240 nm, consistent with a parallel G4 form. Another

smaller ellipticity maximum at 295 nm suggested the coexis-

tence of an antiparallel G4 form. A further six ATRX TR target

sequences were analyzed by circular dichroism (CD); the spec-

trographs were consistent with the formation of G4 including

one sequence that was not predicted by Quadparser to form

G4 (Figure S6A and Table S4).

Finally, we used a gel-shift assay to test whether ATRX could

interact with G4 DNA in vitro. A G-rich oligonucleotide was

preformed into G4 DNA, labeled, and incubated with full-length

recombinant ATRX (Figure 6C). ATRX specifically bound to the

G4 structure and no shift was observed when the structure

was denatured by boiling before adding to the binding reaction.

Further, binding to the formed G4 structure can be competed by

a molar excess of unlabeled formed G4 but is less effectively

competed by the denatured G4 oligonucleotide or another

structured nucleic acid (Holliday junction) (Figure 6C), indicating

that ATRX binds the G4 structure rather than the sequence per

se. These data indicate that ATRX may be recruited to telomeres,

other G-rich TR, and G-rich nonrepetitive DNA and interact with

G-quadruplex DNA.

DISCUSSION

Genome-wide analysis has shown that in euchromatin the

predominant targets of ATRX are sequences containing VNTRs.

Many of these are G and C rich with a high proportion of CpG

dinucleotides. These observations explain why ATRX mutations

affect the a-globin cluster but not the b-globin cluster and cause

a thalassaemia. The a cluster lies in a GC-rich subtelomeric

region containing a high density of CpG islands and G-rich TRs

that we have now shown are targeted by ATRX. The b-globin

cluster has none of these features. It may also explain why in

mouse there are a number of imprinted genes (that are also asso-

ciated with tandemly repeated sequences) whose expression is

affected by downregulation of ATRX (Kernohan et al., 2010).

The relationship between ATRX, VNTRs, and gene expression

is clearly illustrated by the fact that of the targets whose expres-

sion was analyzed, all affected genes were associated with TRs.

Furthermore, at some target genes, the degree by which gene

expression is altered is directly related to the size of the VNTR,

and in the case of one gene examined in detail (NME4), this

can result in monoallelic expression. This provides an explana-

tion for a long-standing question of why individuals with identical

ATRX mutations have variable degrees of a thalassaemia. As

they all have the same mutation and apparently wild-type

a-globin gene clusters, one would have predicted that they

would downregulate the a-globin genes to the same extent.

The highly significant relationship between the effect of the

ATRX deficiency and the natural variation in the VNTR specifi-

cally explains the variable penetrance of ATR-X syndrome but

more importantly identifies a new mechanism that might underlie

many other genetic traits with similar variable penetrance.

A clearly demonstrated but unexplained phenomenon is that,

in the absence of ATRX, expression of the target gene lying

closest to an ATRX peak is the most severely perturbed. How-

ever, adjacent cis-linked genes (up to 10 kb downstream of the

peak) are also affected. For example, although there is enrich-

ment of ATRX across the entire a-globin gene cluster, the main

peak lies close to HBM and is associated with the G-rich TR in

Unlabeled competitor(4 X)

rATRX

Shifted G4

G4 Probe

Linear Probe

A

C

B

Mol

.Ellip

.

-200

600

0

200

400

220 320240 260 280 300

Promoters Gene Body Intergenic All

% o

f p

eaks o

verlap

pin

g p

red

icted

G

4

ATRX Peaks

0

10

20

30

40

50

60

70

80

G4 HJD

Radiolabeled probe D D D G4 G4 G4 G4 G4 G4

Figure 6. ATRX Interaction with G-Quadruplex DNA

(A) The proportion of human ATRX ChIP-seq peak coordinates overlapping

with predicted G-quadruplex (G4) forming sequence.

(B) Circular dichroism. The presence of a positive ellipticity maximum at

260 nm and a negative ellipticity minimum at 240 nm suggests a predominantly

parallel G4 form. The small positive ellipticity maximum at 295 nm is suggestive

of the minor presence of an antiparallel G4 form. See Figure S6A for further

examples of ATRX target sequences forming G4 structures.

(C) Gel-shift assay with recombinant full-length ATRX protein and a [g-32P]ATP

end-labeled G-rich oligonucleotide either preformed into a G4 structure (G4) or

boiled and denatured (D). Reactions contained either 0, 2, or 4 nM rATRX.

Cold competition was performed with a 4-fold molar excess of either unlabeled

G4 formed oligo (G4), denatured oligo (D), or a Holliday junction (HJ).

Cell 143, 367–378, October 29, 2010 ª2010 Elsevier Inc. 375

Page 64: Cell 101029

the HBZ pseudogene. HBM is severely downregulated, but

HBA1 and HBA2 are also downregulated to a lesser degree.

Similarly, at NME4, although this gene is severely downregu-

lated, the adjacent gene (DECR2) is also affected but to a lesser

degree. It appears that ATRX normally binds to these G-rich TRs;

in the absence of ATRX, the repeats at these loci now exert a

repressive influence on transcription that spreads for some

distance from the repeat.

At present it is not clear how ATRX might recognize such

repeat sequences, but one possibility is that they form unusual,

non-B DNA structures in vivo, and in the case of the G-rich

repeats these may take the form of G-quadruplex structures.

Such structures have been demonstrated in vitro using repeats

from telomeres, rDNA, G-rich minisatellites, and CpG-rich

promoters (all ATRX targets), and half of the ATRX targets iden-

tified here are predicted to form G4 DNA. In keeping with the

observations described above, the longer the repeat the more

likely it is to form G4 DNA (Ribeyre et al., 2009). Such structures

have been notoriously difficult to identify in vivo, but the stron-

gest evidence for their existence is at telomeres where it has

been suggested that G4 structures may form during DNA repli-

cation and transcription (Lipps and Rhodes, 2009). It is therefore

of interest that ATRX is recruited to telomeres during replication

and that downregulaton of ATRX by RNAi provokes a DNA-

damage response (marked by gamma-H2AX) at telomeres

during S phase (Wong et al., 2010). Downregulation of ATRX

expression is also associated with an altered expression of

telomere-associated RNA (Goldberg et al., 2010). Both of these

observations would be consistent with ATRX playing a role in

recognizing and/or modifying G4 structures at telomeres and

by implication at other G-rich TRs in vivo. Nevertheless this is

not the only factor determining the localization of ATRX, as at

A/T-rich pericentric heterochromatin, the recruitment of ATRX

depends on the presence of H3K9me3 (Kourmouli et al., 2005).

A role for ATRX at G-rich repeats may also be linked to the

recent observation that ATRX is required for the incorporation

of the histone variant H3.3 at telomeric repeats (Drane et al.,

2010; Goldberg et al., 2010; Lewis et al., 2010). H3.3 may be

incorporated into chromatin in a replication-independent or

replication-dependent manner and has typically been found at

actively transcribed regions of the genome and regions of

inherent nucleosome instability where there is a rapid turnover

of histones during interphase (Schneiderman et al., 2009). TRs

with a propensity to form abnormal DNA structures are likely to

be regions of rapid nucleosome turnover. An appealing hypoth-

esis, therefore, is that ATRX influences gene expression by

recognizing unusual DNA configurations at TRs and converting

them to regular forms in part by facilitating incorporation of

H3.3. Consistent with this we find that H3.3 is found at genic

and intergenic ATRX-binding sites, the majority of which are

TRs. However the distribution of H3.3 is only subtly perturbed

at these sites when ATRX is disrupted. One possibility is that

there is a critical requirement for ATRX at a subset of TRs

(such as telomeres), whereas at other sites, other proteins can

intervene. Future studies will focus on determining the role of

ATRX in H3.3 deposition at specific sites.

The role of ATRX may be to recognize unusual forms of DNA

and facilitate their resolution in several contexts. In the absence

of ATRX, G4 forms may persist and affect many nuclear pro-

cesses including replication, transcription recombination, and

repair.

EXPERIMENTAL PROCEDURES

Western Blotting

For ATRX western blotting, the mouse monoclonal 39c (McDowell et al., 1999)

and rabbit polyclonal H-300 (Insight Biotechnology sc-1540) were used at 1:10

and 1:1000 dilutions, respectively. 23c and 39f recognize an epitope within

ATRX and ATRXt N-terminal to the ADD domain, and H-300 recognizes a

C-terminal epitope within 2193–2492 of full-length ATRX only.

Immunopurification

Nuclear extracts were prepared from wild-type lymphoblastoid cells as previ-

ously described (Dignam, 1990) and incubated overnight at 4�C with H-300

antibody crosslinked to protein A-Sepharose. The beads were washed four

times with 20 mM HEPES (pH 7.9), 0.5M KCl, 0.2 mM EDTA, 0.1% Tween,

0.5 mM DTT and immunoprecipitated protein eluted with 0.1 M glycine

(pH 2.5), then neutralized with 1 M KHPO4. A mock immunopurification was

performed as a control in the same way using normal rabbit IgG (Santa Cruz

sc-2027) crosslinked to protein A.

Chromatin Immunoprecipitation

ATRX chromatin immunoprecipitation was performed according to a published

method (Lee et al., 2006) with the following modifications. Cells were fixed

with 2 mM EGS (Pierce 26103) for 45 min at room temperature in PBS. Form-

aldehyde was then added to 1% for 20 min and quenched with 125 mM

glycine. Chromatin was sonicated to under 500 bp and lysates were immuno-

precipitated with 40 mg ATRX H300 (Insight Biotechnology sc-15408) antibody

or rabbit IgG control (Dako X0903). DNA was precipitated with 20 mg of carrier

glycogen and quantitated using a Qubit fluorimeter (Invitrogen).

Real-Time Q-PCR

Real-time Q-PCR validation of ChIP-seq peaks was performed using SYBR

green mastermix (Applied biosystems 4309155) or using Taqman probes

with a 23 Taq mastermix (Applied Biosystems 4304437). SYBR green primers

(Table S2) were designed using Macvector software and tested by running

a five point, 8-fold serial dilution of genomic DNA to obtain a standard curve

with r2 > 0.99. PCR products were analyzed by melting curve and 3% agarose

gel electrophoresis. Taqman probes were designed using Primer Express

(Applied Biosystems). ChIP enrichments were determined relative to a 3 point

dilution series of input DNA and normalized relative to GAPDH enrichment.

Cell Culture

Human primary erythroblast cultures were prepared using a two-phase liquid

culture system according to a published protocol (Fibach et al., 1991). HbH

inclusions were detected in peripheral blood from ATR-X patients as previ-

ously described (Gibbons et al., 1992). Consent was obtained according to

standard ethics approval guidelines.

Microarray

Fluorescently labeled ChIP and input DNA was analyzed with a custom tiled

microarray covering the subtelomeric region of human chromosome 16p as

previously describe (De Gobbi et al., 2007).

Gene Expression Analysis

RNA was extracted using Tri-reagent (Sigma) and quality checked by micro-

fluidics separation using a 2100 Bioanalyser with an RNA 6000 nano kit

(Agilent 5067-1511). One microgram was reverse transcribed with Superscript

III (Invitrogen). Real-time RT-PCR was performed using commercial Taqman

assays and custom assays. Primer sequences and product codes are listed

in the Extended Experimental Procedures.

High-Throughput Sequencing and Peak Analysis

See Extended Experimental Procedures.

376 Cell 143, 367–378, October 29, 2010 ª2010 Elsevier Inc.

Page 65: Cell 101029

Allelic Discrimination

The ratio of allele-specific transcripts was ascertained with real-time tech-

nology, using an assay designed by Applied Biosystems (Table S2). In brief,

a single amplicon was used, which in combination with two probes, each

specific for one nucleotide of the polymorphism and labeled with a different

fluorophore, allowed quantitation of each species. A standard curve with

known ratios of A:G alleles was used to ensure specificity and quantitativeness

of the assay, and results were confirmed with pyrosequencing (data not

shown). Monoallelic expression is demonstrated with a restriction enzyme

digest assay. The genomic PCR product is 846 bp, of which the G allele

generates fragments of 581 bp and 265 bp when digested with XhoI. The

cDNA PCR product is 854 bp, of which the G allele generates fragments of

563 bp and 291 bp when digested with XhoI. The A allele is undigested by

XhoI in both cases.

VNTR Size Measurement

jz VNTR allele lengths were measured in 43 ATR-X patients with a thalas-

saemia by PCR and agarose gel electrophoresis. PCR was performed in

16.6 mM (NH4)2SO4, 67 mM Tris-HCl (pH 8.8), 10% DMSO, 10 mM Beta mer-

captoethanol, 125 mM dNTP, 0.83 mM MgCl2, 0.7 M Betaine, 0.3 ml platinum

Taq (Invitrogen), 250 nM primers 154505F/155293R (Table S2), and 100 to

400 ng genomic DNA in a 60 ml reaction volume. 30 HVR allele sizes were

measured by radio-labeled Southern blotting using AluI digested genomic

DNA and a probe from pa30HVR.64 derived from genomic fragment

Chr16:175999-177279. VNTR sizes were determined with a Typhoon 9400

Variable Mode Imager and ImageQuant TLv2005 software.

Circular Dichroism Analysis

An oligonucleotide containing the repeat found within the VNTR of intron 1 of

NME4 (CCGGGGTGGGGGTGGGGGTGTGGGGGGGTGA) was diluted to

2 mM in 20 mM Tris HCl (pH 8) and 5 mM NaCl, heated to 95�C for 10 min

then slowly cooled. CD analysis was performed as previously described using

a Jasco 810 CD spectrometer (Giraldo et al., 1994).

G4 Gel Shifts

G4 DNA was formed using oligonucleotide OX1-T (containing the Oxytrichia

telomeric repeat sequence) and its structure confirmed as previously

described (Sun et al., 1998). A Holliday junction structure was formed as

previously described (Bachrati and Hickson, 2006). All DNA substrates were

gel-purified prior to use. G4 DNA was labeled with [g-32P]ATP using T4

polynucleotide kinase, and unincorporated nucleotides were removed using

a Sephadex G50 column. Where indicated the G4 probe was boiled for

10 min and quenched on ice to denature the G4 structure. Binding reactions

(10 ml volume) contained 2 fmol of 32P-labeled G4 DNA, full-length rATRX

protein as indicated (0, 20, or 40 fmol), 6 fmol T25 oligonucleotide to minimize

nonspecific binding in a buffer containing 33 mM Tris acetate (pH 7.9), 66 mM

Na acetate, 1 mM MgCl2, 100 mg/ml BSA, and 1 mM DTT. Where indicated,

unlabeled competitor DNA (G4, denatured G4, or Holliday junction) was added

to the reaction at 4-fold molar excess. Reactions were incubated on ice for

30 min. To each reaction 1 ml of 50% glycerol was added and samples were

loaded onto a 5% acrylamide gel and electrophoresed in 0.5 3 TBE at

5 V/cm for 4 hr at 4�C. The gel was dried on Whatman filter paper and

visualized by autoradiography.

ACCESSION NUMBERS

Our ChIP-seq and microarray datasets have been deposited in the GEO

database with accession number GSE22162.

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures,

six figures, and four tables and can be found with this article online at

doi:10.1016/j.cell.2010.09.023.

ACKNOWLEDGMENTS

We thank the members of the families studied for their participation; S. Butler

and J. Sloane-Stanley for tissue culture; C. Wippo, J. Huddleston, and

L. Marcelline for genotyping; P. Vathesatogkit and R. Totong for SNP valida-

tion; A. Goriely for assistance with pyrosequencing; C. Bachrati for gel-shift

probes; and W. Wood for critical reading of the manuscript. In addition to

others, the work was supported by the Medical Research Council and the

National Institute of Health Biomedical Research Centre Programme. V.V. is

supported by Thailand Research Fund (TRF) and BIOTEC, Thailand. K.M.L.

was supported by an Oxford Nuffield Medical Fellowship, Oxford University.

Received: April 29, 2010

Revised: August 3, 2010

Accepted: September 13, 2010

Published: October 28, 2010

REFERENCES

Argentaro, A., Yang, J.C., Chapman, L., Kowalczyk, M.S., Gibbons, R.J.,

Higgs, D.R., Neuhaus, D., and Rhodes, D. (2007). Structural consequences

of disease-causing mutations in the ATRX-DNMT3-DNMT3L (ADD) domain

of the chromatin-associated protein ATRX. Proc. Natl. Acad. Sci. USA 104,

11939–11944.

Bachrati, C.Z., and Hickson, I.D. (2006). Analysis of the DNA unwinding activity

of RecQ family helicases. Methods Enzymol. 409, 86–100.

Bacolla, A., and Wells, R.D. (2009). Non-B DNA conformations as determi-

nants of mutagenesis and human disease. Mol. Carcinog. 48, 273–285.

Berube, N.G., Smeenk, C.A., and Picketts, D.J. (2000). Cell cycle-dependent

phosphorylation of the ATRX protein correlates with changes in nuclear matrix

and chromatin association. Hum. Mol. Genet. 9, 539–547.

Cui, K., Zang, C., Roh, T.Y., Schones, D.E., Childs, R.W., Peng, W., and Zhao,

K. (2009). Chromatin signatures in multipotent human hematopoietic stem

cells indicate the fate of bivalent genes during differentiation. Cell Stem Cell

4, 80–93.

De Gobbi, M., Anguita, E., Hughes, J., Sloane-Stanley, J.A., Sharpe, J.A.,

Koch, C.M., Dunham, I., Gibbons, R.J., Wood, W.G., and Higgs, D.R. (2007).

Tissue-specific histone modification and transcription factor binding in alpha

globin gene expression. Blood 110, 4503–4510.

Dignam, J.D. (1990). Preparation of extracts from higher eukaryotes. Methods

Enzymol. 182, 194–203.

Drane, P., Ouararhni, K., Depaux, A., Shuaib, M., and Hamiche, A. (2010). The

death-associated protein DAXX is a novel histone chaperone involved in the

replication-independent deposition of H3.3. Genes Dev. 24, 1253–1265.

Fibach, E., Manor, D., Treves, A., and Rachmilewitz, E.A. (1991). Growth of

human normal erythroid progenitors in liquid culture: a comparison with colony

growth in semisolid culture. Int. J. Cell Cloning 9, 57–64.

Flaus, A., Martin, D.M., Barton, G.J., and Owen-Hughes, T. (2006). Identifica-

tion of multiple distinct Snf2 subfamilies with conserved structural motifs.

Nucleic Acids Res. 34, 2887–2905.

Gibbons, R.J., Wilkie, A.O.M., Weatherall, D.J., and Higgs, D.R. (1991).

A newly defined X linked mental retardation syndrome associated with a

thalassaemia. J. Med. Genet. 28, 729–733.

Gibbons, R.J., Suthers, G.K., Wilkie, A.O.M., Buckle, V.J., and Higgs, D.R.

(1992). X-linked a thalassemia/mental retardation (ATR-X) syndrome: Localisa-

tion to Xq12-21.31 by X-inactivation and linkage analysis. Am. J. Hum. Genet.

51, 1136–1149.

Gibbons, R.J., Picketts, D.J., Villard, L., and Higgs, D.R. (1995). Mutations in

a putative global transcriptional regulator cause X-linked mental retardation

with a-thalassemia (ATR-X syndrome). Cell 80, 837–845.

Gibbons, R.J., McDowell, T.L., Raman, S., O’Rourke, D.M., Garrick, D., Ayyub,

H., and Higgs, D.R. (2000). Mutations in ATRX, encoding a SWI/SNF-like

protein, cause diverse changes in the pattern of DNA methylation. Nat. Genet.

24, 368–371.

Cell 143, 367–378, October 29, 2010 ª2010 Elsevier Inc. 377

Page 66: Cell 101029

Gibbons, R.J., Wada, T., Fisher, C., Malik, N., Mitson, M., Steensma, D.,

Goudie, D., Fryer, A., Krantz, I., and Traeger-Synodinos, J. (2008). Mutations

in the chromatin associated protein ATRX. Hum. Mutat. 29, 796–802.

Giraldo, R., Suzuki, M., Chapman, L., and Rhodes, D. (1994). Promotion of

parallel DNA quadruplexes by a yeast telomere binding protein: a circular

dichroism study. Proc. Natl. Acad. Sci. USA 91, 7658–7662.

Goldberg, A.D., Banaszynski, L.A., Noh, K.M., Lewis, P.W., Elsaesser, S.J.,

Stadler, S., Dewell, S., Law, M., Guo, X., Li, X., et al. (2010). Distinct factors

control histone variant H3.3 localization at specific genomic regions. Cell

140, 678–691.

Higgs, D.R., Sharpe, J.A., and Wood, W.G. (1998). Understanding alpha globin

gene expression: a step towards effective gene therapy. Semin. Hematol. 35,

93–104.

Huppert, J.L., and Balasubramanian, S. (2005). Prevalence of quadruplexes in

the human genome. Nucleic Acids Res. 33, 2908–2916.

Jurka, J., Kapitonov, V.V., Pavlicek, A., Klonowski, P., Kohany, O., and

Walichiewicz, J. (2005). Repbase Update, a database of eukaryotic repetitive

elements. Cytogenet. Genome Res. 110, 462–467.

Kernohan, K.D., Jiang, Y., Tremblay, D.C., Bonvissuto, A.C., Eubanks, J.H.,

Mann, M.R., and Berube, N.G. (2010). ATRX partners with cohesin and

MeCP2 and contributes to developmental silencing of imprinted genes in the

brain. Dev. Cell 18, 191–202.

Kourmouli, N., Sun, Y., van der Sar, S., Singh, P.B., and Brown, J.P. (2005).

Epigenetic regulation of mammalian pericentric heterochromatin in vivo by

HP1. Biochem. Biophys. Res. Commun. 337, 901–907.

Langmead, B., Trapnell, C., Pop, M., and Salzberg, S.L. (2009). Ultrafast and

memory-efficient alignment of short DNA sequences to the human genome.

Genome Biol. 10, R25.

Lee, T.I., Johnstone, S.E., and Young, R.A. (2006). Chromatin immunoprecip-

itation and microarray-based analysis of protein location. Nat. Protoc. 1,

729–748.

Lewis, P.W., Elsaesser, S.J., Noh, K.M., Stadler, S.C., and Allis, C.D. (2010).

Daxx is an H3.3-specific histone chaperone and cooperates with ATRX in

replication-independent chromatin assembly at telomeres. Proc. Natl. Acad.

Sci. USA 107, 14075–14080.

Lipps, H.J., and Rhodes, D. (2009). G-quadruplex structures: in vivo evidence

and function. Trends Cell Biol. 19, 414–422.

McDowell, T.L., Gibbons, R.J., Sutherland, H., O’Rourke, D.M., Bickmore,

W.A., Pombo, A., Turley, H., Gatter, K., Picketts, D.J., Buckle, V.J., et al.

(1999). Localization of a putative transcriptional regulator (ATRX) at pericentro-

meric heterochromatin and the short arms of acrocentric chromosomes. Proc.

Natl. Acad. Sci. USA 96, 13983–13988.

Paramasivan, S., Rujan, I., and Bolton, P.H. (2007). Circular dichroism of

quadruplex DNAs: applications to structure, cation effects and ligand binding.

Methods 43, 324–331.

Rhead, B., Karolchik, D., Kuhn, R.M., Hinrichs, A.S., Zweig, A.S., Fujita, P.A.,

Diekhans, M., Smith, K.E., Rosenbloom, K.R., Raney, B.J., et al. (2009). The

UCSC Genome Browser database: update 2010. Nucleic Acids Res. 38,

D613–D619.

Ribeyre, C., Lopes, J., Boule, J.B., Piazza, A., Guedin, A., Zakian, V.A.,

Mergny, J.L., and Nicolas, A. (2009). The yeast Pif1 helicase prevents genomic

instability caused by G-quadruplex-forming CEB1 sequences in vivo. PLoS

Genet. 5, e1000475.

Robertson, G., Hirst, M., Bainbridge, M., Bilenky, M., Zhao, Y., Zeng, T.,

Euskirchen, G., Bernier, B., Varhol, R., Delaney, A., et al. (2007). Genome-

wide profiles of STAT1 DNA association using chromatin immunoprecipitation

and massively parallel sequencing. Nat. Methods 4, 651–657.

Schneiderman, J.I., Sakai, A., Goldstein, S., and Ahmad, K. (2009). The XNP

remodeler targets dynamic chromatin in Drosophila. Proc. Natl. Acad. Sci.

USA 106, 14472–14477.

Sun, H., Karow, J.K., Hickson, I.D., and Maizels, N. (1998). The Bloom’s

syndrome helicase unwinds G4 DNA. J. Biol. Chem. 273, 27587–27592.

Waterston, R.H., Lindblad-Toh, K., Birney, E., Rogers, J., Abril, J.F., Agarwal,

P., Agarwala, R., Ainscough, R., Alexandersson, M., An, P., et al. (2002).

Initial sequencing and comparative analysis of the mouse genome. Nature

420, 520–562.

Wilkie, A.O.M., Zeitlin, H.C., Lindenbaum, R.H., Buckle, V.J., Fischel-

Ghodsian, N., Chui, D.H.K., Gardner-Medwin, D., MacGillivray, M.H.,

Weatherall, D.J., and Higgs, D.R. (1990). Clinical features and molecular anal-

ysis of the a thalassemia/mental retardation syndromes. II. Cases without

detectable abnormality of the a globin complex. Am. J. Hum. Genet. 46,

1127–1140.

Wong, L.H., McGhie, J.D., Sim, M., Anderson, M.A., Ahn, S., Hannan, R.D.,

George, A.J., Morgan, K.A., Mann, J.R., and Choo, K.H. (2010). ATRX interacts

with H3.3 in maintaining telomere structural integrity in pluripotent embryonic

stem cells. Genome Res. 20, 351–360.

Xue, Y., Gibbons, R., Yan, Z., Yang, D., McDowell, T.L., Sechi, S., Qin, J.,

Zhou, S., Higgs, D., and Wang, W. (2003). The ATRX syndrome protein forms

a chromatin-remodeling complex with Daxx and localizes in promyelocytic

leukemia nuclear bodies. Proc. Natl. Acad. Sci. USA 100, 10635–10640.

378 Cell 143, 367–378, October 29, 2010 ª2010 Elsevier Inc.

Page 67: Cell 101029

Upf1 Senses 30UTR Lengthto Potentiate mRNA DecayJ. Robert Hogg1,3,* and Stephen P. Goff1,2,3,*1Department of Biochemistry and Molecular Biophysics2Department of Microbiology and Immunology3Howard Hughes Medical Institute

College of Physicians and Surgeons, Columbia University, New York, NY 10032, USA*Correspondence: [email protected] (J.R.H.), [email protected] (S.P.G.)

DOI 10.1016/j.cell.2010.10.005

SUMMARY

The selective degradation of mRNAs by thenonsense-mediated decay pathway is a qualitycontrol process with important consequences forhuman disease. From initial studies using RNAhairpin-tagged mRNAs for purification of messengerribonucleoproteins assembled on transcripts withHIV-1 30 untranslated region (30UTR) sequences, weuncover a two-step mechanism for Upf1-dependentdegradation of mRNAs with long 30UTRs. We demon-strate that Upf1 associates with mRNAs in a 30UTRlength-dependent manner and is highly enrichedon transcripts containing 30UTRs known to elicitNMD. Surprisingly, Upf1 recruitment and subsequentRNA decay can be antagonized by retroviral RNAelements that promote translational readthrough.By modulating the efficiency of translation termina-tion, recognition of long 30UTRs by Upf1 is uncoupledfrom the initiation of decay. We propose a model for30UTR length surveillance in which equilibriumbinding of Upf1 to mRNAs precedes a kineticallydistinct commitment to RNA decay.

INTRODUCTION

The nonsense-mediated decay (NMD) machinery executes

important regulatory and quality control functions by targeting

specific classes of messenger RNAs (mRNAs) for degradation

(Chang et al., 2007). In addition to degrading transcripts contain-

ing premature termination codons (PTCs) resulting from mutation

or rearrangement of genomic DNA or defects in mRNA biogen-

esis, the pathway is also responsible for regulating between 1%

and 10% of all genes in diverse eukaryotes (He et al., 2003;

Mendell et al., 2004; Rehwinkel et al., 2005; Wittmann et al.,

2006; Weischenfeldt et al., 2008). Transcripts preferentially tar-

geted by NMD include those with PTCs encoded by alternatively

spliced exons, introns downstream of the termination codon (TC),

long 30 untranslated regions (30UTRs), or upstream open reading

frames (uORFs; reviewed in Nicholson et al., 2010; Rebbapra-

gada and Lykke-Andersen, 2009). A characteristic that is

common to many NMD decay substrates is an extended distance

from the terminating ribosome to the mRNA 30 end (i.e., 30UTR

length). Degradation of aberrant mRNAs by NMD can affect the

progression of many human genetic disorders, an estimated

one-third of which derive from PTCs (Kuzmiak and Maquat,

2006). In addition, shortening of 30UTRs has been proposed to

relax regulation of mRNA stability and translation, promoting

cellular transformation (Sandberg et al., 2008; Wang et al.,

2008; Mayr and Bartel, 2009). These findings underscore the

importance of understanding the mechanisms by which 30UTR

length is sensed in the process of mRNA quality control.

The well-conserved superfamily I RNA helicase Upf1 is

a crucial component of the core NMD machinery. Like other

RNA helicases, Upf1 exhibits nonspecific but robust RNA

binding activity modulated by ATP binding and hydrolysis

(Weng et al., 1998; Bhattacharya et al., 2000). Though the func-

tional roles of Upf1’s ATPase and helicase activities are unclear,

mutations that abolish its ATPase activity prevent NMD (Weng

et al., 1996a, 1996b; Sun et al., 1998). In addition, Upf1 partici-

pates in a network of interactions with additional factors

proposed to mediate its association with mRNA targets and

regulate a cycle of Upf1 phosphorylation and dephosphorylation

required for establishment of translational repression and

recruitment of RNA decay enzymes (reviewed in Nicholson

et al., 2010; see below).

Within the context of a long 30UTR, additional mRNA features

and protein components of mRNPs can promote or inhibit

decay. For example, the exon-junction complex (EJC), a multi-

protein assembly deposited at exon-exon junctions in the

process of splicing, acts through Upf1 to strongly activate decay

(Le Hir et al., 2000, 2001; Kim et al., 2001; Lykke-Andersen et al.,

2001). The competition between Upf1 and cytoplasmic poly(A)-

binding protein 1 (PABPC1) for binding to the translation release

factors eRF1 and eRF3 has been proposed to be a crucial factor

in the decision to decay diverse transcripts (Ivanov et al., 2008;

Singh et al., 2008). Upf1 binding to release factors at the

terminating ribosome stimulates phosphorylation of Upf1 by

the SMG-1 kinase, translational repression, and recruitment of

decay factors (Kashima et al., 2006; Isken et al., 2008; Cho

et al., 2009). Conversely, binding of PABPC1 to release factors

is proposed to preserve transcript stability and translational

competence. In support of this model, artificial tethering

approaches and alterations in 30UTR structure designed to

Cell 143, 379–389, October 29, 2010 ª2010 Elsevier Inc. 379

Page 68: Cell 101029

mimic 30UTR shortening by bringing PABPC1 in proximity to the

termination codon can suppress Upf1-dependent decay (Amrani

et al., 2004; Behm-Ansmant et al., 2007; Eberle et al., 2008; Iva-

nov et al., 2008; Silva et al., 2008).

Here, we use affinity purification of hairpin-tagged mRNAs to

isolate and characterize endogenously assembled mRNP

complexes. With this approach, we show that Upf1 assembles

into mRNPs in a 30UTR length-dependent manner. Upf1 copuri-

fies to some extent with all transcripts tested but is highly en-

riched on mRNAs containing 30UTRs derived from known NMD

targets. The preferential association of Upf1 with mRNAs con-

taining NMD-sensitive 30UTRs is not affected by inhibition of

translation and NMD. Together with our finding that the effi-

ciency of Upf1 coimmunoprecipitation with 30UTR-derived

RNase H cleavage products correlates with fragment length,

these observations suggest a direct role for Upf1 in 30UTR length

sensing. To further investigate the in vivo dynamics of 30UTR

length surveillance and decay, we use retroviral elements to

induce translational readthrough of NMD-triggering termination

codons. Surprisingly, periodic readthrough events can reduce

steady-state Upf1 association with transcripts containing long

30UTRs and robustly inhibit NMD. Moreover, we show that rare

readthrough events permit steady-state Upf1 accumulation in

mRNPs but prevent initiation of mRNA decay.

Our data inform a model in which equilibrium binding of Upf1

senses 30UTR length and establishes an RNP state primed for

decay. The identification of potential decay targets by Upf1 is

coupled to a subsequent commitment to decay, the rate of which

is dependent on other aspects of mRNP structure and composi-

tion. Furthermore, our data indicate that the decision to decay

takes place over a kinetic interval corresponding to many trans-

lation termination events. This separation between 30UTR length

sensing and initiation of decay provides a mechanism to prevent

aberrant degradation of normal RNAs and presents an opportu-

nity for transcripts to evade cellular mRNA surveillance. Retrovi-

ruses may exploit this opportunity by inducing translational read-

through or frameshifting to periodically disrupt the recognition of

viral mRNAs as potential decay substrates.

RESULTS

RNA-Based Affinity Purification Identifies ProteinComponents of Messenger RNPsTo better understand cellular mRNA biogenesis and decay, we

have developed a generalizable technique for purification and

characterization of endogenously assembled mRNP complexes.

In this approach, we singly tag mRNAs with the naturally

occurring Pseudomonas phage 7 coat protein (PP7CP) binding

site, a 25 nucleotide (nt) stably folding hairpin (Figure 1A; Lim

and Peabody, 2002). Tagged RNAs are transiently or stably ex-

pressed in appropriate mammalian cell lines, allowing progres-

sion through endogenous RNA processing pathways. RNPs

assembled on the tagged RNAs are then purified from extracts

using a version of the PP7CP tagged with tandem Staphylo-

coccus aureus protein A domains. Previously, a similar method

was used to isolate complexes associated with several noncod-

ing RNAs (Hogg and Collins, 2007a, 2007b). In the process of

adapting this methodology to the purification of mRNPs, we

found that the use of traditional agarose-based resins afforded

inefficient purification of tagged mRNP complexes. In contrast,

nonporous magnetic resins allowed purification of tagged

mRNAs to near homogeneity following a single step of purifica-

tion (Figures 1A and 1C; additional data not shown).

Recent work in our laboratory has shown that HIV-1 30LTR

sequences play a crucial role in the regulation of viral mRNA

biogenesis (Valente and Goff, 2006; Valente et al., 2009).

To identify proteins specifically associated with HIV 30LTR

sequences, we constructed a series of PP7-tagged RNAs con-

taining the GFP open reading frame and alternative 30UTRs

(Figure 1B; see below). In our initial experiments, we used

a version of the HIV 30LTR containing a deletion in the U3 region

(DU3 LTR). The bovine growth hormone polyadenylation (bGH

pA) element of the pcDNA3.1 vector was used as a control, aid-

ing discrimination of proteins specifically bound to HIV 30LTR

sequence-containing RNPs. Silver staining of complexes puri-

fied from whole-cell extracts of transiently transfected 293T cells

revealed that, as expected, each tagged RNA associates with

A C

B

Figure 1. RNA Hairpin-Based Affinity Purifi-

cation of PP7-Tagged mRNAs

(A) (Left) Predicted secondary structure of the

PP7CP RNA-binding site (Lim and Peabody,

2002). (Right) Scheme for purification and analysis

of mRNPs containing specific mRNAs.

(B) Tagged mRNAs used for RNA-based affinity

purification. RNAs containing the GFP ORF, a

single copy of the PP7 RNA hairpin, and the bovine

growth hormone polyadenylation element (bGH,

top), an HIV 30LTR variant containing a deletion

in the U3 region (DU3 LTR, middle), or the full-

length HIV 30LTR (bottom). TC positions are indi-

cated by octagons.

(C) Purification of tagged mRNPs. Proteins copur-

ifying with bGH- or DU3 LTR-containing RNAs

or present in mock purifications from extracts

lacking tagged RNA were separated by SDS-

PAGE and detected by silver staining. (Inset)

Magnification of the band corresponding to Upf1

(see also Table S1).

380 Cell 143, 379–389, October 29, 2010 ª2010 Elsevier Inc.

Page 69: Cell 101029

a large number of proteins (Figure 1C and data not shown). Mock

purifications from extracts lacking tagged RNAs exhibited very

few contaminating proteins, indicating that the vast majority

of the proteins visible by silver staining were isolated via their

association with tagged mRNPs. Many components of the

purified mRNPs are found in all complexes purified, including

general translation factors, ribosomes, hnRNP proteins, and

other proteins that associate with common mRNA features

(Figure 1C and data not shown).

Tandem mass spectrometry of gel slices excised from HIV

DU3 LTR and bGH copurifying material identified four peptides

derived from the Upf1 protein in the DU3 LTR sample but none

in the bGH control sample (Table S1 available online). Immuno-

blotting of PP7-purified RNPs confirmed that Upf1 was enriched

on transcripts containing HIV 30LTR sequences, using immuno-

blotting for PABPC1 as a control for RNP recovery (Figure 2A).

We detected Upf1 in association with RNAs containing the

bGH pA element, but at much lower levels than those copurifying

A

B

C

E

D

F

Figure 2. 30UTR Length-Dependent Interaction of Upf1 with mRNAs

(A) Enrichment of Upf1 on RNAs containing LTR sequence. Proteins in whole-cell extracts of parental 293T cells (mock) or cells transiently transfected with the

indicated PP7-tagged RNAs (extract, left) or copurifying with tagged RNAs (RNP, right) were detected by immunoblotting with antibodies against endogenous

Upf1 (top) and PABPC1 (bottom).

(B) Sequence-independent assembly of Upf1 in mRNPs. RNPs containing the bGH pA element or full-length or DU3 LTRs in the sense (FLTR and DU3 LTR) or

antisense (DU3 AS and FLTR AS) orientations were subjected to purification and immunoblotting as in (A). (Bottom) Small fractions of input extract and purified

material were analyzed by northern blotting to detect tagged RNAs.

(C) Upf1 association depends on 30UTR length. (Top) Constructs encoding RNAs in which the HIV 30LTR was fused to the GFP ORF, placing the HIV nef ORF in

frame. RNAs contained the standard GFP TC (0) or a CAA codon in place of the GFP TC in tandem with artificially introduced TCs at 100 nt intervals (100, 200, 300,

and 400). (Bottom) RNPs were purified and analyzed by immunoblotting and northern blotting as in (A) and (B).

(D) Quantification of data in (C). Upf1 signal was normalized to RNA signal from northern blotting of a fraction of the purified material and arbitrarily set to 1 for the

construct containing the standard GFP TC. Error bars indicate ± SEM; n = 2.

(E) Coimmunoprecipitation of Upf1 with 30UTR-derived RNase H cleavage products. RNase H cleavage of PP7-GFP-FLTR mRNAs was directed using oligonu-

cleotides hybridizing to FLTR sequences 7, 211, or 305 nt downstream of the GFP TC. Extracts were subjected to immunoprecipitation with an anti-Upf1 antibody

or nonspecific IgG, and recovery of uncut mRNAs (top) and 30UTR fragments (bottom) was monitored by northern blotting using a probe against the indicated

portion of the HIV 30LTR sequence. See also Figure S1A.

(F) Quantification of the data in (E). Recovery of 30UTR fragments was normalized to the abundance of the fragments in extracts, using the uncut mRNAs as

internal controls. The recovery efficiency of the longest 30UTR fragment (7) was arbitrarily set to 1. Error bars indicate ± SEM; n = 2.

Cell 143, 379–389, October 29, 2010 ª2010 Elsevier Inc. 381

Page 70: Cell 101029

with RNAs containing the DU3 LTR sequence. Still higher levels

of Upf1 were isolated using a full-length LTR (FLTR) comprising

intact U3, R, and U5 LTR segments from the pNL4.3 reference

HIV genome (Figure 1B and Figure 2A). In agreement with obser-

vations of human Upf1 cosedimentation with bulk polysomes

and coimmunoprecipitation with diverse mRNAs (Pal et al.,

2001; Hosoda et al., 2005), we find that Upf1 associates to

some degree with all RNAs tested (Figure 2A, Figure 3A and

data not shown). Importantly, we additionally observe substan-

tial transcript-specific enrichment of Upf1 in mRNPs (see below).

Upf1 Associates with Transcripts in a 30UTRLength-Dependent MannerTo address the specificity of Upf1 association with RNPs con-

taining HIV 30LTR sequences, we first created tagged RNA

constructs in which the DU3 or full-length LTR elements were

cloned in the antisense orientation, with the bGH pA element

provided downstream to ensure proper 30 end maturation. The

requirement for an additional 30 end-processing element caused

the antisense 30UTRs to be �200 nt longer than their sense

equivalents (Figure 2B, see northern blot). As above, we

observed increasing Upf1 copurification with the bGH, DU3,

and FLTR RNAs, respectively (Figure 2B). Surprisingly, the levels

of Upf1 associated with the antisense LTR-containing RNAs

were slightly higher than with the corresponding sense 30UTRs.

Thus, the observed recruitment of Upf1 to LTR-containing

RNAs was not dependent on primary sequence or structural

features. Instead, our data suggested that Upf1 accumulation

in mRNPs might be dictated by 30UTR length.

Current models suggest that 30UTR length is a crucial determi-

nant of NMD susceptibility (Muhlemann, 2008; Rebbapragada

and Lykke-Andersen, 2009), but the mechanism by which

30UTR length is sensed remains unclear. To test the hypothesis

that Upf1 associates with transcripts in a 30UTR length-depen-

dent manner, we generated a series of 50 PP7-tagged RNAs

consisting of the GFP ORF fused to the HIV FLTR, such that

the fragment of the HIV nef ORF contained in the LTR was in

frame with the GFP ORF (Figure 2C). This series of constructs

contains single termination codons at �100 nt intervals, starting

with the original GFP termination codon and ending with the nef

termination codon �400 nt downstream. In this way, we varied

30UTR length by making only one (ablation of the GFP TC) or

two (ablation of the GFP TC combined with introduction of a

new in-frame TC) point mutations to the RNA primary sequence.

Using these constructs, we found that Upf1 copurification with

tagged mRNAs increased with 30UTR length (Figure 2C). The

relationship between Upf1 copurification and 30UTR length was

strikingly linear, consistent with sequence-nonspecific recogni-

tion of long 30UTRs by Upf1 (Figure 2D).

Our observations suggested that Upf1 might accomplish

30UTR length sensing by associating with 30UTRs. To better

understand the basis for 30UTR length-dependent accumulation

of Upf1 in mRNPs, we used RNase H and a series of oligonucle-

otides directed against HIV 30LTR sequence to site-specifically

cleave 50-tagged GFP-FLTR mRNAs at sites �7, �211, and

�305 nucleotides downstream of the GFP TC. Following RNase

H digestion, we immunoprecipitated endogenous Upf1 and as-

sayed mRNA recovery by northern blotting using a probe against

HIV 30LTR sequence. The RNase H cleavage conditions were

designed to leave a substantial fraction of the mRNAs intact,

allowing the use of full-length mRNAs as recovery controls.

FLTR-containing mRNAs were recovered with an antibody

against Upf1, but not nonspecific control goat IgG (Figure 2E

and Figure S1A). Consistent with our observations above, the

efficiency of 30UTR fragment coimmunoprecipitation increased

with RNA length (Figures 2E and 2F). These data suggest that

Upf1 association along the length of 30UTRs accounts for the

observed 30UTR length-dependent accumulation in mRNPs.

Upf1 Preferentially Associates with TranscriptsContaining NMD-Sensitive 30UTRsOur observation that Upf1 association correlates with 30UTR

length mirrors prior findings that 30UTR extension causes

progressive transcript destabilization in mammalian cells (Buhler

et al., 2006; Eberle et al., 2008; Singh et al., 2008). To assess the

functional significance of the enrichment of Upf1 on specific

A

B

Figure 3. Upf1 Preferentially Associates with Transcripts Containing

30UTRs Known to Trigger NMD

(A) PP7-tagged GFP mRNAs containing the indicated 30UTRs were transiently

expressed in 293T cells and subjected to affinity purification. Proteins present

in whole-cell extracts and purified RNPs were detected by immunoblotting

with antibodies against endogenous Upf1, PABPC1, SMG-1, and Upf2.

(Bottom) RNA was isolated from small fractions of extracts and purified mate-

rial and analyzed by northern blotting. See also Figures S1A and S1B.

(B) Upf1 recruitment is insensitive to cycloheximide treatment. 293T cells tran-

siently transfected with PP7-tagged GFP mRNAs containing the TRAM1 or

SMG5 30UTRs were treated (+) or not treated (�) with cycloheximide for 4 hr

prior to cell harvest and throughout extract preparation and affinity purification.

Immunoblotting and northern blotting were performed as in (A). Inhibition of

NMD by cycloheximide and persistence of Upf1 recruitment under conditions

of translation inhibition by puromycin and 50-proximal hairpins are illustrated in

Figures S1C–S1F.

382 Cell 143, 379–389, October 29, 2010 ª2010 Elsevier Inc.

Page 71: Cell 101029

mRNAs, we used a series of long 30UTRs shown by Singh and

colleagues (2008) to either promote or evade decay. As repre-

sentative NMD-insensitive long 30UTRs, we used the human

CRIPT1 (1515 nt) and TRAM1 (1494 nt) 30UTRs. To model targets

of 30UTR length-dependent NMD, we used the human SMG5

30UTR (1342 nt) and an artificial 30UTR comprising a portion of

the GAPDH ORF and the GAPDH 30UTR (GAP; 846 nt).

As above, we transiently transfected 293T cells with tagged

RNA constructs containing model 30UTRs and isolated mRNPs

from whole-cell extracts with PP7CP. Immunoblotting of purified

RNPs revealed that Upf1 association strongly correlated with

NMD sensitivity (Figure 3A). Very low levels of Upf1 copurified

with transcripts containing the NMD-insensitive CRIPT1 and

TRAM1 30UTRs. In contrast, transcripts containing the intronless

GAP and SMG5 30UTRs copurified high levels of Upf1, with the

SMG5 30UTR-containing mRNAs showing the greatest Upf1

recruitment. Likewise, antibodies against Upf1 coimmunopreci-

pitated mRNAs containing the GAP and SMG5 30UTRs at higher

efficiencies than mRNAs containing the CRIPT and TRAM

30UTRs (Figure S1A). We did not observe the NMD factors

SMG-1 or Upf2 in PP7-purified mRNPs, despite robust detection

of the proteins in whole-cell extracts used for purification

(Figure 3A). In similar experiments, comparable levels of Upf1

copurified with mRNAs containing the intronless GAP 30UTR

and a version of the GAP 30UTR containing the adenovirus

major-late intron (GAP AdML; Figure S1B). This observation

suggests that 30UTR length is a more significant determinant of

Upf1 association than the presence of a spliced intron down-

stream of the TC. Together, these findings indicate that the

extent of Upf1 association with a transcript is diagnostic of its

NMD susceptibility, consistent with previous experiments in

yeast, C. elegans, and human cells (Johansson et al., 2007;

Johns et al., 2007; Silva et al., 2008 ; Hwang et al., 2010). In addi-

tion, they raise the intriguing possibility that endogenous mRNAs

with long 30UTRs, such as the CRIPT1 and TRAM1 mRNAs,

evade NMD by preventing steady-state incorporation of Upf1

into mRNPs.

Preferential Accumulation of Upf1 on TranscriptsContaining NMD-Sensitive 30UTRs Is Independentof Ongoing NMDWe hypothesized that the enrichment of Upf1 on long 30UTR-

containing transcripts could reflect a direct role for the protein

in 30UTR-length sensing prior to the initiation of decay. To

address this possibility, we tested the effect of suppressing

NMD on Upf1 recruitment by treating cells with the translation

elongation inhibitor cycloheximide (Figure 3B). Because initiation

of NMD requires translation termination events, cycloheximide

potently inhibits NMD (Figure S1C). Following cell growth and

extract preparation in cycloheximide, we purified tagged RNAs

containing the TRAM1 and SMG5 30UTRs and analyzed Upf1

association by immunoblotting. Both in the presence and

absence of cycloheximide, SMG5 30UTR-containing RNAs

exhibited enhanced copurification of Upf1 relative to TRAM1

30UTR-containing control RNAs (Figure 3B). Identical results

were obtained using the CRIPT1 and GAP 30UTRs (data not

shown). Moreover, the same pattern of Upf1 accumulation in

mRNPs was observed upon inhibition of translation elongation

with puromycin or translation initiation with a cap-proximal

stable hairpin (Figures S1D–S1F). These data demonstrate that

Upf1 recruitment is independent of ongoing translation termina-

tion and NMD and is therefore well positioned to act as a key

determinant of 30UTR length sensing.

Translational Readthrough Events Reduce Upf1Association with mRNAs Containing Long 30UTRsTo probe the in vivo dynamics of 30UTR length recognition by

Upf1, we used retroviral elements to modulate the efficiency of

translation termination upstream of an NMD-inducing 30UTR.

Retroviruses control the relative production of Gag and Gag-

Pol precursor proteins using RNA motifs that induce regulated

readthrough or �1 frameshifting to bypass the gag termination

codon (Bolinger and Boris-Lawrie, 2009). The Moloney murine

leukemia virus pseudoknot (MLVPK), a well-characterized

example of the former class, causes misincorporation of an

amino acid at the gag termination codon with �4% frequency

(Figure S2A) (Wills et al., 1991). Because ribosomes transiting

through 30UTRs are presumably capable of inducing dramatic

remodeling of RNP structure, we reasoned that retroviral read-

through-promoting elements might disrupt the recognition of

potential NMD substrates.

To determine the effects of translational readthrough on Upf1

accumulation and NMD, we inserted the readthrough-promoting

MLVPK sequence in place of the standard GFP termination

codon in PP7-tagged RNA constructs, upstream of the artificial

GAP 30UTR. This model NMD-triggering 30UTR comprises 489

nt of the GAPDH open reading frame and 357 nt of the GAPDH

30UTR. Readthrough events thus result in termination at the

downstream GAPDH TC, a position that does not elicit NMD in

reporter transcripts (Figure 4A, top) (Singh et al., 2008; see

below). As controls, we inserted an additional termination codon

immediately downstream of the MLVPK to prevent readthrough

into the GAPDH ORF (MLVPK�C) (Figure 4A, middle) or mutated

the upstream termination codon to CAG to allow constitutive

translation of the GFP-GAPDH fusion protein (MLVPK CAG)

(Figure 4A, bottom). In addition, we used three MLVPK variants

with reduced readthrough efficiency to assess the competition

between readthrough and Upf1 association: mutation of the

wild-type UAG termination codon to UAA (�1.5% readthrough;

see Figure S2A for bicistronic dual-luciferase readthrough effi-

ciency assays; Feng et al., 1990), G11C (numbering from termi-

nation codon; < 1% readthrough; Felsenstein and Goff, 1992),

and A17G (�2% readthrough; Wills et al., 1994).

As expected, immunoblotting of purified RNPs revealed that

control mRNAs containing an additional termination codon

downstream of the MLVPK efficiently recruited Upf1 (�C)

(Figure 4B). In contrast, mRNAs in which the wild-type MLVPK

sequence (UAG termination codon) directed intermittent transla-

tion of the GAPDH ORF copurified significantly reduced amounts

of Upf1. In fact, mRNAs containing the wild-type MLVPK

sequence copurified levels of Upf1 similar to those associated

with mRNAs lacking an upstream termination codon (CAG)

(Figures 4B and 4C). Of interest, the extent of Upf1 association

was dependent on the efficiency of readthrough, as the three

MLVPK mutants exhibiting reduced readthrough activity copuri-

fied high levels of Upf1, similar to the no-readthrough control

Cell 143, 379–389, October 29, 2010 ª2010 Elsevier Inc. 383

Page 72: Cell 101029

(Figures 4B and 4C). As a control, we treated cells with puro-

mycin prior to cell extract preparation to confirm that the modu-

lation of Upf1 recruitment by MLVPK was dependent on transla-

tion. Indeed, mRNAs containing the wild-type MLVPK and the

no-readthrough and constitutive-readthrough controls all cop-

urified similar levels of Upf1 under conditions of puromycin treat-

ment (Figure S2C). In addition, the Mouse mammary tumor virus

(MMTV) �1 frameshifting element also inhibited steady-state

Upf1 association, demonstrating that the reduction in Upf1

recruitment was independent of the mechanism of termination

codon evasion (Figure S2B). These data suggest that the peri-

odic transit of ribosomes through the 30UTR during readthrough

events remodels the mRNP, displacing Upf1. Readthrough

events caused by the wild-type MLVPK were sufficiently

frequent to repress steady-state Upf1 accumulation, whereas

less active MLVPK variants permitted recovery of Upf1 binding

to mRNPs. Modulation of readthrough efficiency thus uncovers

an equilibrium of Upf1 binding and displacement that marks

long 30UTRs of potential decay targets.

Rare Readthrough Events Protect Transcripts fromDecay Independently of Disruption of Steady-State Upf1AssociationThe ability of readthrough-promoting elements to reduce

steady-state Upf1 association with mRNPs raised the possibility

that such elements could also stabilize targets of NMD in

mammalian cells. To assess RNA decay, we introduced MLVPK

variants into tetracycline (tet)-regulated b-globin reporter

RNAs containing the GAP artificial 30UTR (Figure 5A) (Singh

et al., 2008). The indicated tet-regulated constructs were

A

B

C

Figure 4. Effects of Translational Readthrough on Upf1 Association

(A) PP7-tagged RNAs containing the GFP ORF in frame with a fragment of the GAPDH ORF were modified to contain the wild-type pseudoknot (top, MLVPK), the

MLVPK with an additional in-frame TC downstream (middle, MLVPK –C), or an MLVPK sequence lacking a TC (bottom, MLVPK CAG). Positions of in-frame TCs

are indicated. TCs subject to suppression by the MLVPK are indicated by gray octagons; normal TCs are indicated by black octagons.

(B) Tagged mRNAs containing the indicated MLVPK variants were transiently expressed in 293T cells and used for RNA affinity purification. Proteins present in

whole-cell extracts and purified material were analyzed by immunoblotting for endogenous Upf1 and PABPC1, and tagged mRNAs were detected by northern

blotting. See also Figure S2 for determination of approximate readthrough efficiencies using a bicistronic luciferase assay, disruption of Upf1 binding to mRNAs

containing the MMTV �1 frameshifting element, and the effects of puromycin on Upf1 recruitment to MLVPK-containing mRNAs.

(C) Quantification of Upf1 copurification normalized to PABPC1 copurification, arbitrarily set to 1 for the CAG no-TC control. Error bars indicate ± SEM; n = 3.

384 Cell 143, 379–389, October 29, 2010 ª2010 Elsevier Inc.

Page 73: Cell 101029

cotransfected with CMV promoter-driven b-globin control RNAs

into HeLa Tet-off cells. After an interval of transcription induction,

we monitored the decay of tet-regulated RNAs for a 9 hr time

course by northern blotting (Figures 5B and 5C). As expected,

mRNAs lacking an upstream TC (MLVPK CAG) were much

more stable than RNAs containing the wild-type MLVPK

sequence followed immediately by an additional termination

codon (MLVPK –C). Remarkably, all of the MLVPK variants

tested, including the minimally active G11C mutant, increased

RNA stability to levels indistinguishable from control RNAs lack-

ing an upstream TC (Figures 5B and 5C; additional data not

shown). Using the MMTV �1 frameshift element to allow transla-

tion of the GAPDH ORF also rescued transcript stability, sug-

gesting that readthrough events stabilize NMD targets indepen-

dently of the mechanism of translational recoding (Figure S3).

Impaired MLVPK variants allow steady-state Upf1 accumula-

tion in mRNPs but robustly inhibit long 30UTR-mediated mRNA

decay. These findings imply that Upf1 recognition of long 30UTRs

is coupled to a kinetically deferred commitment to decay that is

subject to inhibition by rare readthrough events. The presence

of an EJC downstream of a TC substantially enhances mRNA

decay, potentially by activating Upf1 recruited by long 30UTRs.

We therefore reasoned that the EJC might overcome the effects

of readthrough by accelerating the initiation of decay. As a model

for EJC-stimulated decay, we used the GAP AdML intron-con-

taining 30UTR previously shown to direct efficient degradation

of reporter transcripts (Singh et al., 2008). Assays of GAP

AdML 30UTR-containing mRNA accumulation revealed that the

ability of MLVPK variants to promote RNA stability correlated

with readthrough efficiency (Figures 6A and 6B and Figure S4).

The wild-type MLVPK sequence permitted mRNA accumulation

to levels indistinguishable from those of mRNAs lacking an

upstream TC (MLVPK CAG), whereas RNAs containing the mildly

impaired UAA and A17G MLVPK variants accumulated to

slightly lower levels. The more severe G11C and A39U mutations

further reduced RNA accumulation but nevertheless allowed

�4-fold higher RNA levels than the no-readthrough control

(MLVPK �C). Decay assay time courses confirmed that the

MLVPK-containing RNAs were indeed stabilized relative to the

no-readthrough control transcripts (Figure S4). These data

show that even efficient EJC-stimulated NMD can be signifi-

cantly impaired by instances of readthrough occurring at less

than 1% of all possible termination events. In addition, the differ-

ential stability of transcripts undergoing readthrough of varying

efficiency points to the existence of a rate-limiting step down-

stream of Upf1 association that can be accelerated by the EJC

(see Discussion).

DISCUSSION

The identification of proteins copurifying with PP7-tagged RNAs

by mass spectrometry allows unbiased analysis of the effects of

RNA sequence, structure, and biogenesis pathway on mRNP

composition. Using this system, we show that the extent of

Upf1 association with specific transcripts strongly correlates

with NMD susceptibility, as Upf1 is highly enriched on transcripts

containing 30UTRs derived from known NMD targets. Analysis of

mRNPs containing HIV 30LTR-derived and other model 30UTR

sequences revealed that Upf1 recruitment increases with

30UTR length but is independent of ongoing translation and

NMD. Moreover, we show that Upf1 coimmunoprecipitates

30UTR-derived RNase H cleavage products in a fragment

length-dependent manner, suggesting that Upf1 is associated

along the length of 30UTRs. The enrichment of Upf1 on long

A

C

B

Figure 5. Rare Readthrough Events Stabilize Targets of NMD

(A) Schematic of tet-regulated b-globin reporter mRNA constructs used in RNA decay assays. Constructs contained the b-globin ORF and the GAPDH ORF frag-

ment in frame, with intervening MLVPK sequence to regulate readthrough. Positions of in-frame TCs are indicated. TCs subject to suppression by the MLVPK are

indicated by gray octagons; normal TCs are indicated by black octagons.

(B) Decay assays of reporter mRNAs containing MLVPK variants. Constructs encoding the tet-regulated transcripts described in (A) (pcTET2 bwt MLVPK GAP;

top bands) were cotransfected with the constitutively expressed wild-type b-globin reporter (pcbwtb; bottom bands) in HeLa Tet-off cells. RNA was harvested

30 min after transcription was halted by addition of doxycycline and at 3 hr intervals thereafter. See also Figure S3 for decay assays of mRNAs containing the

MMTV �1FS element.

(C) Quantification of decay assays. Levels of tet-regulated reporter mRNAs were normalized to levels of the wild-type b-globin transfection control. Error bars

indicate ± SEM; n = 3.

Cell 143, 379–389, October 29, 2010 ª2010 Elsevier Inc. 385

Page 74: Cell 101029

30UTR-containing transcripts may increase the probability that

Upf1 will outcompete PABPC1 for release factor binding and

trigger NMD, providing a potential mechanism for the correlation

between 30UTR length and transcript stability in human cells

(Buhler et al., 2006; Eberle et al., 2008; Singh et al., 2008).

With the expectation that readthrough events would cause

elongating ribosomes to periodically strip Upf1 and other

proteins from the mRNA downstream of the bypassed termina-

tion codon, we modulated readthrough efficiency to probe the

dynamics of Upf1 association. We show that the activity of

wild-type MLVPK, which causes �4% readthrough, suppresses

steady-state Upf1 recruitment to the artificial GAP 30UTR. Less

frequent readthrough events, in contrast, allow complete

recovery of Upf1 binding to mRNPs. Together, our observations

suggest that Upf1 accumulation on transcripts is not simply

dependent on the site of the vast majority of termination events

but is instead determined by sequence-nonspecific association

of Upf1 with 30UTRs. Differential disruption of Upf1 binding by

readthrough events of varying efficiency therefore reveals an

equilibrium of Upf1 association that can serve as a mechanism

for sensing 30UTR length.

Equilibrium binding of Upf1 to mRNPs may be influenced by

several factors, including Upf1 RNA binding affinity, interactions

with additional mRNP components, and disruption by elongating

ribosomes. The ATP binding and hydrolysis cycle of Upf1 modu-

lates the protein’s sequence-nonspecific RNA binding activity,

providing a potential mechanism to regulate Upf1 association

with mRNPs (Weng et al., 1998; Bhattacharya et al., 2000). It is

possible that Upf1 recruitment to long 30UTRs is mediated by

protein-protein interactions, but scrutiny of silver-stained gels

of purified RNPs and immunoblotting for known NMD factors

did not reveal additional proteins that showed patterns of copur-

ification similar to Upf1 (1C, Figure 3A, and data not shown). In

addition, Upf1 association was maintained despite treatment

with EDTA and inhibition of translation by multiple means, indi-

cating that an interaction with intact ribosomes is not responsible

for the observed accumulation of Upf1 in mRNPs (Figure 3B and

Figures S1C–S1G).

Based on our findings that rare readthrough events permit

Upf1 30UTR length-dependent accumulation in mRNPs but

inhibit decay, we propose a two-step model in which Upf1

senses 30UTR length to potentiate decay (Figure 7). In this model,

length-dependent equilibrium binding of Upf1 marks 30UTRs of

potential decay substrates and increases the probability of

Upf1 binding to release factors. Upf1 accumulation in mRNPs

is necessary, but not sufficient, to initiate mRNA degradation

and is instead followed by a kinetically distinct commitment to

decay. The decision to decay is determined by the activity of

Upf1 and additional mRNP factors, including the EJC and

PABPC1. Here, we provide evidence that 30UTR recognition

and decay commitment steps can be separated by modulating

readthrough efficiency: frequent readthrough events disrupt

30UTR length sensing by displacing Upf1 from 30UTR sequence,

and rare readthrough events allow Upf1 association but prevent

one or more rate-limiting steps required for initiation of decay.

Potential rate-limiting steps subject to disruption by infrequent

translational readthrough include ATP binding and hydrolysis

by Upf1, the formation of the Upf (Upf1-Upf2-Upf3b) or SURF

(SMG-1-Upf1-eRF1-eRF3) complexes, Upf1 phosphorylation,

and the recruitment and/or activity of the RNA degradation

machinery. We find that the presence of a spliced intron down-

stream of the TC partially restores decay in the context of rare

readthrough events, indicating that one key event may be stim-

ulated by the EJC, such as Upf1 ATPase activity or phosphory-

lation (Kashima et al., 2006; Wittmann et al., 2006; Chamieh

et al., 2008).

An important consequence of the delay between length-

dependent Upf1 accumulation in mRNPs and initiation of decay

may be improved quality control fidelity. Based on a release

factor-dependent nonsense error rate of 1 in 105 codons (Jør-

gensen et al., 1993), aberrant termination events are predicted

to occur on �50% of transcripts encoding 50 kDa proteins

during the course of 100 translation events. Integrating the deci-

sion to decay over several termination events provides a mecha-

nism to avoid degradation in response to translational errors

while preserving the ability to recognize DNA- or RNA-encoded

PTCs. This mechanism may also prevent immediate decay

upon binding of Upf1 to release factors at a normal TC, which

may occur at a significant frequency as suggested by studies

of Upf1 and PABP release factor association (Ivanov et al.,

2008; Singh et al., 2008).

In mammals, NMD has been proposed to exclusively target

newly exported mRNA during a pioneer round or rounds of trans-

lation that are biochemically distinguished by the presence of the

nuclear cap-binding proteins CBP80/20 in the mRNP (Maquat

A

B

Figure 6. Readthrough Inhibits EJC-Stimulated NMD

(A) Northern blot of RNA accumulation in HeLa Tet-off cells cotransfected with

constitutively expressed wild-type b-globin transcripts (pcbwtb; bottom

bands) and tet-regulated b-globin transcripts containing the GAP AdML

30UTR and the indicated MLVPK variants (pcTET2 bwt MLVPK GAP AdML;

top bands). See Figure S4 for decay assays.

(B) Quantification of RNA accumulation assays. Levels of tet-regulated

reporter RNAs were normalized to levels of the wild-type b-globin transfection

control. Error bars indicate ± SEM; n = 3.

386 Cell 143, 379–389, October 29, 2010 ª2010 Elsevier Inc.

Page 75: Cell 101029

et al., 2010). Replacement of CBP80/20 with translation initiation

factor eIF4E marks entry of an mRNP into bulk translation and is

thought to confer resistance to NMD. We observe substantial

stabilization of transcripts containing elements that direct read-

through at less than 1% efficiency, suggesting that the decision

to decay spans an interval corresponding to many translation

termination events. These findings mirror the effects of altering

termination efficiency in yeast, in which NMD surveillance is

conducted throughout the lifetime of a transcript (Zhang et al.,

1997; Maderazo et al., 2003; Keeling et al., 2004). The ability

of rare readthrough events to inhibit NMD in mammalian cells

is supported by multiple reports of readthrough-promoting

drugs or inefficient selenocysteine incorporation at UGA codons

inducing accumulation of decay targets (Bedwell et al., 1997;

Moriarty et al., 1998; Weiss and Sunde, 1998; Mehta et al.,

2004; Allamand et al., 2008; Salvatori et al., 2009). Therefore,

we hypothesize that mammalian NMD is able to degrade

mRNAs that have proceeded beyond the pioneer round(s) of

translation.

EXPERIMENTAL PROCEDURES

Constructs

For details of plasmid construction, see Extended Experimental Procedures.

Cell Culture and Extracts

293T cells were maintained, transfected, and used for cell extract preparation

essentially as described (Hogg and Collins, 2007b). For details, see Extended

Experimental Procedures.

RNA-Based Affinity Purification of mRNPs and Immunoprecipitation

All mRNP purification steps were performed at 4�C. Whole-cell extracts

prepared in 20 mM HEPES (pH 7.6), 150 mM NaCl, 2 mM MgCl2, 10% glycerol,

1 mM DTT, and protease inhibitors (HLB150) were supplemented with 0.1%

NP-40 and �1 ug/ml ZZ-tev-PP7CP expressed and purified in E. coli as

described (Hogg and Collins, 2007b). After rotating for 60 min, 4.5 mg of

M-270 epoxy Dynabeads (Invitrogen) conjugated with rabbit IgG (Oeffinger

et al., 2007) were added per ml of extract, followed by an additional 60 min

incubation. Beads were collected using a magnetic particle concentrator (Invi-

trogen) and extensively washed in HLB150 + 0.1% NP-40. RNP proteins were

eluted from beads using LDS buffer (Invitrogen) for immunoblotting or 1:50

diluted RNase A/T1 mix (Ambion) in 100 mM ammonium bicarbonate for silver

A

B

Figure 7. Model for 30UTR Length Surveillance by Upf1

(A) Equilibrium length-dependent binding of Upf1 marks long 30UTRs as potential decay targets. Other aspects of RNP structure and composition, such as the

EJC and PABPC1, can stimulate or repress decay potentiated by Upf1 association.

(B) Inhibition of Upf1-dependent decay by translational readthrough. (Top) 30UTR length-dependent accumulation of Upf1 increases the likelihood of Upf1 binding

to release factors and precedes the initiation of decay. (Middle) Readthrough induced by the wild-type MLVPK disrupts steady-state accumulation of Upf1 on

mRNAs containing long 30UTRs and inhibits decay. (Bottom) MLVPK variants causing low levels of readthrough allow recovery of Upf1 equilibrium binding but

inhibit a kinetically distinct decay commitment step. Candidate rate-limiting steps required for decay initiation are discussed in the main text.

Cell 143, 379–389, October 29, 2010 ª2010 Elsevier Inc. 387

Page 76: Cell 101029

staining and mass spectrometry. For details of mass spectrometry and RNase

H cleavage and immunoprecipitation experiments, see Extended Experi-

mental Procedures.

RNA Decay and Accumulation Assays

RNA decay assays were performed as described, with modifications (Singh

et al., 2008). For details, see Extended Experimental Procedures.

Detection of RNA and Protein

Northern blots were imaged on a Typhoon Trio and quantified using Image-

Quant software (G.E.). Immunoblots were imaged and quantified on the

Odyssey Infrared Imaging System (LI-COR). For information on probes and

antibodies used, see Extended Experimental Procedures.

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures, four

figures, and one table and can be found with this article online at doi:

10.1016/j.cell.2010.10.005.

ACKNOWLEDGMENTS

We thank Jens Lykke-Andersen and Brian Houck-Loomis for generously

providing reagents and Kathleen Collins, Lisa Postow, and Jason Rodriguez

for critical reading of the manuscript. Mass spectrometry was performed by

Mary Ann Gawinowicz in the Columbia University Medical Center protein

core facility. J.R.H. is supported by NRSA postdoctoral fellowship

1F32GM087737. S.P.G. is an investigator of the Howard Hughes Medical

Institute.

Received: April 16, 2010

Revised: August 3, 2010

Accepted: October 1, 2010

Published: October 28, 2010

REFERENCES

Allamand, V., Bidou, L., Arakawa, M., Floquet, C., Shiozuka, M., Paturneau-

Jouas, M., Gartioux, C., Butler-Browne, G.S., Mouly, V., Rousset, J.-P.,

et al. (2008). Drug-induced readthrough of premature stop codons leads to

the stabilization of laminin a2 chain mRNA in CMD myotubes. J. Gene Med.

10, 217–224.

Amrani, N., Ganesan, R., Kervestin, S., Mangus, D.A., Ghosh, S., and Jacob-

son, A. (2004). A faux 30-UTR promotes aberrant termination and triggers

nonsense-mediated mRNA decay. Nature 432, 112–118.

Bedwell, D.M., Kaenjak, A., Benos, D.J., Bebok, Z., Bubien, J.K., Hong, J.,

Tousson, A., Clancy, J.P., and Sorscher, E.J. (1997). Suppression of a CFTR

premature stop mutation in a bronchial epithelial cell line. Nat. Med. 3,

1280–1284.

Behm-Ansmant, I., Gatfield, D., Rehwinkel, J., Hilgers, V., and Izaurralde, E.

(2007). A conserved role for cytoplasmic poly(A)-binding protein 1 (PABPC1)

in nonsense-mediated mRNA decay. EMBO J. 26, 1591–1601.

Bhattacharya, A., Czaplinski, K., Trifillis, P., He, F., Jacobson, A., and Peltz,

S.W. (2000). Characterization of the biochemical properties of the human

Upf1 gene product that is involved in nonsense-mediated mRNA decay.

RNA 6, 1226–1235.

Bolinger, C., and Boris-Lawrie, K. (2009). Mechanisms employed by retrovi-

ruses to exploit host factors for translational control of a complicated pro-

teome. Retrovirology 6, 8.

Buhler, M., Steiner, S., Mohn, F., Paillusson, A., and Muhlemann, O. (2006).

EJC-independent degradation of nonsense immunoglobulin-m mRNA

depends on 30 UTR length. Nat. Struct. Mol. Biol. 13, 462–464.

Chamieh, H., Ballut, L., Bonneau, F., and Le Hir, H. (2008). NMD factors UPF2

and UPF3 bridge UPF1 to the exon junction complex and stimulate its RNA

helicase activity. Nat. Struct. Mol. Biol. 15, 85–93.

Chang, Y.-F., Imam, J.S., and Wilkinson, M.F. (2007). The nonsense-mediated

decay RNA surveillance pathway. Annu. Rev. Biochem. 76, 51–74.

Cho, H., Kim, K.M., and Kim, Y.K. (2009). Human proline-rich nuclear receptor

coregulatory protein 2 mediates an interaction between mRNA surveillance

machinery and decapping complex. Mol. Cell 33, 75–86.

Eberle, A.B., Stalder, L., Mathys, H., Orozco, R.Z., and Muhlemann, O. (2008).

Posttranscriptional gene regulation by spatial rearrangement of the 30 untrans-

lated region. PLoS Biol. 6, e92.

Felsenstein, K.M., and Goff, S.P. (1992). Mutational analysis of the gag-pol

junction of Moloney murine leukemia virus: requirements for expression of

the gag-pol fusion protein. J. Virol. 66, 6601–6608.

Feng, Y.X., Copeland, T.D., Oroszlan, S., Rein, A., and Levin, J.G. (1990).

Identification of amino acids inserted during suppression of UAA and UGA

termination codons at the gag-pol junction of Moloney murine leukemia virus.

Proc. Natl. Acad. Sci. USA 87, 8860–8863.

He, F., Li, X., Spatrick, P., Casillo, R., Dong, S., and Jacobson, A. (2003).

Genome-wide analysis of mRNAs regulated by the nonsense-mediated and

50 to 30 mRNA decay pathways in yeast. Mol. Cell 12, 1439–1452.

Hogg, J.R., and Collins, K. (2007a). Human Y5 RNA specializes a Ro ribonu-

cleoprotein for 5S ribosomal RNA quality control. Genes Dev. 21, 3067–3072.

Hogg, J.R., and Collins, K. (2007b). RNA-based affinity purification reveals

7SK RNPs with distinct composition and regulation. RNA 13, 868–880.

Hosoda, N., Kim, Y.K., Lejeune, F., and Maquat, L.E. (2005). CBP80 promotes

interaction of Upf1 with Upf2 during nonsense-mediated mRNA decay in

mammalian cells. Nat. Struct. Mol. Biol. 12, 893–901.

Hwang, J., Sato, H., Tang, Y., Matsuda, D., and Maquat, L.E. (2010). UPF1

association with the cap-binding protein, CBP80, promotes nonsense-medi-

ated mRNA decay at two distinct steps. Mol. Cell 39, 396–409.

Isken, O., Kim, Y.K., Hosoda, N., Mayeur, G.L., Hershey, J.W.B., and Maquat,

L.E. (2008). Upf1 phosphorylation triggers translational repression during

nonsense-mediated mRNA decay. Cell 133, 314–327.

Ivanov, P.V., Gehring, N.H., Kunz, J.B., Hentze, M.W., and Kulozik, A.E. (2008).

Interactions between UPF1, eRFs, PABP and the exon junction complex

suggest an integrated model for mammalian NMD pathways. EMBO J. 27,

736–747.

Johansson, M.J.O., He, F., Spatrick, P., Li, C., and Jacobson, A. (2007). Asso-

ciation of yeast Upf1p with direct substrates of the NMD pathway. Proc. Natl.

Acad. Sci. USA 104, 20872–20877.

Johns, L., Grimson, A., Kuchma, S.L., Newman, C.L., and Anderson, P. (2007).

Caenorhabditis elegans SMG-2 selectively marks mRNAs containing prema-

ture translation termination codons. Mol. Cell. Biol. 27, 5630–5638.

Jørgensen, F., Adamski, F.M., Tate, W.P., and Kurland, C.G. (1993). Release

factor-dependent false stops are infrequent in Escherichia coli. J. Mol. Biol.

230, 41–50.

Kashima, I., Yamashita, A., Izumi, N., Kataoka, N., Morishita, R., Hoshino, S.,

Ohno, M., Dreyfuss, G., and Ohno, S. (2006). Binding of a novel SMG-1-Upf1-

eRF1-eRF3 complex (SURF) to the exon junction complex triggers Upf1 phos-

phorylation and nonsense-mediated mRNA decay. Genes Dev. 20, 355–367.

Keeling, K.M., Lanier, J., Du, M., Salas-Marco, J., Gao, L., Kaenjak-Angeletti,

A., and Bedwell, D.M. (2004). Leaky termination at premature stop codons

antagonizes nonsense-mediated mRNA decay in S. cerevisiae. RNA 10,

691–703.

Kim, V.N., Kataoka, N., and Dreyfuss, G. (2001). Role of the nonsense-

mediated decay factor hUpf3 in the splicing-dependent exon-exon junction

complex. Science 293, 1832–1836.

Kuzmiak, H.A., and Maquat, L.E. (2006). Applying nonsense-mediated mRNA

decay research to the clinic: progress and challenges. Trends Mol. Med. 12,

306–316.

Le Hir, H., Gatfield, D., Izaurralde, E., and Moore, M.J. (2001). The exon-exon

junction complex provides a binding platform for factors involved in mRNA

export and nonsense-mediated mRNA decay. EMBO J. 20, 4987–4997.

388 Cell 143, 379–389, October 29, 2010 ª2010 Elsevier Inc.

Page 77: Cell 101029

Le Hir, H., Izaurralde, E., Maquat, L.E., and Moore, M.J. (2000). The spliceo-

some deposits multiple proteins 20-24 nucleotides upstream of mRNA

exon-exon junctions. EMBO J. 19, 6860–6869.

Lim, F., and Peabody, D.S. (2002). RNA recognition site of PP7 coat protein.

Nucleic Acids Res. 30, 4138–4144.

Lykke-Andersen, J., Shu, M.D., and Steitz, J.A. (2001). Communication of the

position of exon-exon junctions to the mRNA surveillance machinery by the

protein RNPS1. Science 293, 1836–1839.

Maderazo, A.B., Belk, J.P., He, F., and Jacobson, A. (2003). Nonsense-

containing mRNAs that accumulate in the absence of a functional nonsense-

mediated mRNA decay pathway are destabilized rapidly upon its restitution.

Mol. Cell. Biol. 23, 842–851.

Maquat, L.E., Tarn, W.Y., and Isken, O. (2010). The pioneer round of transla-

tion: features and functions. Cell 142, 368–374.

Mayr, C., and Bartel, D.P. (2009). Widespread shortening of 3’UTRs by alterna-

tive cleavage and polyadenylation activates oncogenes in cancer cells. Cell

138, 673–684.

Mehta, A., Rebsch, C.M., Kinzy, S.A., Fletcher, J.E., and Copeland, P.R.

(2004). Efficiency of mammalian selenocysteine incorporation. J. Biol. Chem.

279, 37852–37859.

Mendell, J.T., Sharifi, N.A., Meyers, J.L., Martinez-Murillo, F., and Dietz, H.C.

(2004). Nonsense surveillance regulates expression of diverse classes of

mammalian transcripts and mutes genomic noise. Nat. Genet. 36, 1073–1078.

Moriarty, P.M., Reddy, C.C., and Maquat, L.E. (1998). Selenium deficiency

reduces the abundance of mRNA for Se-dependent glutathione peroxidase

1 by a UGA-dependent mechanism likely to be nonsense codon-mediated

decay of cytoplasmic mRNA. Mol. Cell. Biol. 18, 2932–2939.

Muhlemann, O. (2008). Recognition of nonsense mRNA: towards a unified

model. Biochem. Soc. Trans. 36, 497–501.

Nicholson, P., Yepiskoposyan, H., Metze, S., Zamudio Orozco, R., Kleinsch-

midt, N., and Muhlemann, O. (2010). Nonsense-mediated mRNA decay in

human cells: mechanistic insights, functions beyond quality control and the

double-life of NMD factors. Cell. Mol. Life Sci. 67, 677–700.

Oeffinger, M., Wei, K.E., Rogers, R., DeGrasse, J.A., Chait, B.T., Aitchison,

J.D., and Rout, M.P. (2007). Comprehensive analysis of diverse ribonucleopro-

tein complexes. Nat. Methods 4, 951–956.

Pal, M., Ishigaki, Y., Nagy, E., and Maquat, L.E. (2001). Evidence that phos-

phorylation of human Upfl protein varies with intracellular location and is medi-

ated by a wortmannin-sensitive and rapamycin-sensitive PI 3-kinase-related

kinase signaling pathway. RNA 7, 5–15.

Rebbapragada, I., and Lykke-Andersen, J. (2009). Execution of nonsense-

mediated mRNA decay: what defines a substrate? Curr. Opin. Cell Biol. 21,

394–402.

Rehwinkel, J., Letunic, I., Raes, J., Bork, P., and Izaurralde, E. (2005).

Nonsense-mediated mRNA decay factors act in concert to regulate common

mRNA targets. RNA 11, 1530–1544.

Salvatori, F., Breveglieri, G., Zuccato, C., Finotti, A., Bianchi, N., Borgatti, M.,

Feriotto, G., Destro, F., Canella, A., Brognara, E., et al. (2009). Production of

b-globin and adult hemoglobin following G418 treatment of erythroid

precursor cells from homozygous b039 thalassemia patients. Am. J. Hematol.

84, 720–728.

Sandberg, R., Neilson, J.R., Sarma, A., Sharp, P.A., and Burge, C.B. (2008).

Proliferating cells express mRNAs with shortened 30 untranslated regions

and fewer microRNA target sites. Science 320, 1643–1647.

Silva, A.L., Ribeiro, P., Inacio, A., Liebhaber, S.A., and Romao, L. (2008). Prox-

imity of the poly(A)-binding protein to a premature termination codon inhibits

mammalian nonsense-mediated mRNA decay. RNA 14, 563–576.

Singh, G., Rebbapragada, I., and Lykke-Andersen, J. (2008). A competition

between stimulators and antagonists of Upf complex recruitment governs

human nonsense-mediated mRNA decay. PLoS Biol. 6, e111.

Sun, X., Perlick, H.A., Dietz, H.C., and Maquat, L.E. (1998). A mutated human

homologue to yeast Upf1 protein has a dominant-negative effect on the decay

of nonsense-containing mRNAs in mammalian cells. Proc. Natl. Acad. Sci.

USA 95, 10009–10014.

Valente, S.T., Gilmartin, G.M., Mott, C., Falkard, B., and Goff, S.P. (2009). Inhi-

bition of HIV-1 replication by eIF3f. Proc. Natl. Acad. Sci. USA 106, 4071–4078.

Valente, S.T., and Goff, S.P. (2006). Inhibition of HIV-1 gene expression by

a fragment of hnRNP U. Mol. Cell 23, 597–605.

Wang, E.T., Sandberg, R., Luo, S., Khrebtukova, I., Zhang, L., Mayr, C.,

Kingsmore, S.F., Schroth, G.P., and Burge, C.B. (2008). Alternative isoform

regulation in human tissue transcriptomes. Nature 456, 470–476.

Weischenfeldt, J., Damgaard, I., Bryder, D., Theilgaard-Monch, K., Thoren,

L.A., Nielsen, F.C., Jacobsen, S.E.W., Nerlov, C., and Porse, B.T. (2008).

NMD is essential for hematopoietic stem and progenitor cells and for elimi-

nating by-products of programmed DNA rearrangements. Genes Dev. 22,

1381–1396.

Weiss, S.L., and Sunde, R.A. (1998). Cis-acting elements are required for sele-

nium regulation of glutathione peroxidase-1 mRNA levels. RNA 4, 816–827.

Weng, Y., Czaplinski, K., and Peltz, S.W. (1996a). Genetic and biochemical

characterization of mutations in the ATPase and helicase regions of the Upf1

protein. Mol. Cell. Biol. 16, 5477–5490.

Weng, Y., Czaplinski, K., and Peltz, S.W. (1996b). Identification and character-

ization of mutations in the UPF1 gene that affect nonsense suppression and

the formation of the Upf protein complex but not mRNA turnover. Mol. Cell.

Biol. 16, 5491–5506.

Weng, Y., Czaplinski, K., and Peltz, S.W. (1998). ATP is a cofactor of the Upf1

protein that modulates its translation termination and RNA binding activities.

RNA 4, 205–214.

Wills, N.M., Gesteland, R.F., and Atkins, J.F. (1991). Evidence that a down-

stream pseudoknot is required for translational read-through of the Moloney

murine leukemia virus gag stop codon. Proc. Natl. Acad. Sci. USA 88,

6991–6995.

Wills, N.M., Gesteland, R.F., and Atkins, J.F. (1994). Pseudoknot-dependent

read-through of retroviral gag termination codons: importance of sequences

in the spacer and loop 2. EMBO J. 13, 4137–4144.

Wittmann, J., Hol, E.M., and Jack, H.-M. (2006). hUPF2 silencing identifies

physiologic substrates of mammalian nonsense-mediated mRNA decay.

Mol. Cell. Biol. 26, 1272–1287.

Zhang, S., Welch, E.M., Hogan, K., Brown, A.H., Peltz, S.W., and Jacobson, A.

(1997). Polysome-associated mRNAs are substrates for the nonsense-

mediated mRNA decay pathway in Saccharomyces cerevisiae. RNA 3,

234–244.

Cell 143, 379–389, October 29, 2010 ª2010 Elsevier Inc. 389

Page 78: Cell 101029

The Long Noncoding RNA, Jpx,Is a Molecular Switchfor X Chromosome InactivationDi Tian,1,2,3 Sha Sun,1,2 and Jeannie T. Lee1,2,3,*1Howard Hughes Medical Institute2Department of Molecular Biology

Massachusetts General Hospital and Department of Genetics, Harvard Medical School, Boston, MA 02115, USA3Department of Pathology, Massachusetts General Hospital, Boston, MA 02114, USA*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.09.049

SUMMARY

Once protein-coding, the X-inactivation center (Xic) isnow dominated by large noncoding RNAs (ncRNA). Xchromosome inactivation (XCI) equalizes geneexpression between mammalian males and femalesby inactivating one X in female cells. XCI requiresXist, an ncRNA that coats the X and recruits Poly-comb proteins. How Xist is controlled remainsunclear but likely involves negative and positive regu-lators. For the active X, the antisense Tsix RNA is anestablished Xist repressor. For the inactive X, here,we identify Xic-encoded Jpx as an Xist activator.Jpx is developmentally regulated and accumulatesduring XCI. Deleting Jpx blocks XCI and is femalelethal. Posttranscriptional Jpx knockdown recapitu-lates the knockout, and supplying Jpx in transrescues lethality. Thus, Jpx is trans-acting and func-tions as ncRNA. Furthermore, DJpx is rescued bytruncating Tsix, indicating an antagonistic relation-ship between the ncRNAs. We conclude that Xist iscontrolled by two RNA-based switches: Tsix for Xaand Jpx for Xi.

INTRODUCTION

In the mammal, X chromosome inactivation (XCI) achieves

dosage balance between the sexes by transcriptionally silencing

one X chromosome in the female (Lyon, 1961; Lucchesi et al.,

2005; Wutz and Gribnau, 2007; Payer and Lee, 2008; Starmer

and Magnuson, 2009). During XCI, �1000 genes on the X are

subject to repression by the X-inactivation center (Xic) (Brown

et al., 1991). Multiple noncoding genes have been identified

within this 100–500 kb domain that, until �150 million years

ago, was dominated by protein-coding genes. The rise of Euthe-

rian mammals and the transition from imprinted to random XCI

led to region-wide ‘‘pseudogenization’’ (Duret et al., 2006; Da-

vidow et al., 2007; Hore et al., 2007; Shevchenko et al., 2007).

To date, four Xic-encoded noncoding genes have been ascribed

function in XCI, including Xist, Tsix, Xite, and RepA (Brockdorff

et al., 1992; Brown et al., 1992; Lee and Lu, 1999; Ogawa and

Lee, 2003; Zhao et al., 2008) (Figure 1A). The dominance of

ncRNAs brought early suspicion that long transcripts are favored

by allelic regulation during XCI and imprinting (for review, see

Wan and Bartolomei, 2008; Koerner et al., 2009; Lee, 2009;

Mercer et al., 2009). Indeed, the Xic region harbors many other

ncRNA (Simmler et al., 1996; Chureau et al., 2002), but many

have yet to be characterized.

One key player is Xist, a 17 kb ncRNA that initiates XCI as it

spreads along the X in cis (Brockdorff et al., 1992; Brown et al.,

1992; Penny et al., 1996; Marahrens et al., 1997; Wutz et al.,

2002) and recruits Polycomb repressive complexes to the X

(Plath et al., 2003; Silva et al., 2003; Schoeftner et al., 2006;

Zhao et al., 2008). In embryonic stem (ES) cell models that reca-

pitulate XCI during differentiation ex vivo, Xist expression is

subject to a counting mechanism that ensures repression in XY

cells and monoallelic upregulation in XX cells. Prior to differenti-

ation, Xist is expressed at a low basal level but is poised for

activation in the presence of supernumerary X chromosomes

(XX state). In the presence of only one X (XY), Xist becomes

stably silenced.

It has been proposed that Xist is under both positive and nega-

tive control (Lee and Lu, 1999; Lee, 2005; Monkhorst et al.,

2008). Negative regulation is achieved by the antisense gene,

Tsix. When Tsix is deleted or truncated, the Xist allele in cis is

derepressed (Lee and Lu, 1999; Lee, 2000; Luikenhuis et al.,

2001; Sado et al., 2001; Stavropoulos et al., 2001; Morey et al.,

2004; Vigneau et al., 2006; Ohhata et al., 2008). Tsix represses

Xist induction by several means, including altering the chromatin

state of Xist (Navarro et al., 2005; Sado et al., 2005; Sun et al.,

2006; Ohhata et al., 2008), deploying Dnmt3a’s DNA methyl-

transferase activity (Sado et al., 2005; Sun et al., 2006), recruiting

the RNAi machinery (Ogawa et al., 2008), and interfering with the

ability of Xist and RepA RNA to engage Polycomb proteins (Zhao

et al., 2008). In turn, Tsix is regulated by Xite, a proximal noncod-

ing element that interacts with Tsix’s promoter (Tsai et al., 2008)

and sustains Tsix expression on the future Xa (active X) (Ogawa

and Lee, 2003).

Significantly, whereas a Tsix deletion has major effects on Xist

in XX cells, it has little consequence in XY cells (Lee and Lu, 1999;

390 Cell 143, 390–403, October 29, 2010 ª2010 Elsevier Inc.

Page 79: Cell 101029

Ohhata et al., 2006). This difference led to the idea that Xist is not

only negatively regulated on Xa but also positively controlled on

Xi (inactive X) by factors that activate Xist (Lee and Lu, 1999).

Positive regulation finds support in that RepA—a short RNA

embedded within Xist—recruits Polycomb proteins to facilitate

Xist upregulation (Zhao et al., 2008; Hoki et al., 2009). Activators

outside of the Xist-Tsix-Xite region must also occur, as an 80 kb

transgene carrying only these genes cannot induce XCI (Lee

et al., 1999b). Furthermore, female cells carrying a heterozygous

deletion of Xist-Tsix-Xite still undergo XCI, indicating female cells

with only one copy of Xist, Tsix, and Xite still count two X chromo-

somes (Monkhorst et al., 2008). One such activator has been

proposed to be the E3 ubiquitin ligase, Rnf12, whose gene

resides �500 kb away from Xist (Jonkers et al., 2009). Overex-

pression of Rnf12 ectopically induces Xist expression in XY cells,

but Rnf12 is not required for Xist activation in XX cells, as its

knockout delays but does not abrogate expression. This implies

that essential Xist activator(s) must reside elsewhere.

Here, we seek to identify that essential factor. We draw hints

from an older study demonstrating that, while transgenes

carrying only Xist-Tsix-Xite cannot activate Xist, inclusion of

sequences upstream of Xist restores Xist upregulation (Lee

et al., 1999b). The Eutherian-specific noncoding gene, Jpx/

Enox (Chureau et al., 2002; Johnston et al., 2002; Chow et al.,

2003), lies �10 kb upstream of Xist, is transcribed in the opposite

orientation (Figure 1A), but remains largely uncharacterized. Jpx

lacks open reading frames but is relatively conserved in its 50

exons. Initial reports indicate that Jpx is neither developmentally

regulated nor sex specific and is therefore unlikely to regulate

XCI (Chureau et al., 2002; Johnston et al., 2002; Chow et al.,

2003). Although they imply a pseudogene status, chromosome

conformation capture (3C) suggests that Jpx resides within Xist’s

chromatin hub (Tsai et al., 2008). We herein study Jpx and

uncover a crucial role as ncRNA in the positive arm of Xist

regulation.

RESULTS

Jpx Escapes XCI and Is Upregulated during ES CellDifferentiationWe first analyzed Jpx expression patterns in ES cells, as an Xist

inducer might be expected to display developmental specificity

correlating with the kinetics of XCI. Time-course measurements

of Jpx and Xist during ES differentiation into embryoid bodies

(EB) showed that Jpx RNA levels increased 10- to 20-fold

between d0 and d12 and remained elevated in somatic cells

(Figure 1B and data not shown). Upregulation occurred in both

XX and XY cells. However, whereas Xist induction paralleled

Jpx upregulation in female cells, Xist remained suppressed in

male cells (Figure 1C). To determine whether Jpx originated

from Xa or Xi, we carried out allele-specific analysis in TsixTST/+

female cells, which are genetically marked by a Tsix mutation

that invariably inactivates the mutated X of 129 origin (X129)

instead of the wild-type Mus castaneus X (Xcas) (Ogawa et al.,

2008). On the basis of a Nla-III polymorphism, RT-PCR demon-

strated that both alleles of Jpx could be detected from d0 to d12,

indicating that Jpx escapes XCI (Figure 1D). On d0, there was

μ

0

5

10

15

20

25

30

Nor

mal

ized

Jp

x le

vel

0 4 8 12 MEF

Day of ES differentiation

B

WT

A

0

5

10

15

20

25

0 4 8 12Day of ES differentiation

Nor

mal

ized

Jp

x le

vel WT

C

0

0.5

1.0

1.5

2.0

2.5

0 4 8 12Day of ES differentiation

Nor

mal

ized

Xis

t le

vel

WT

D

0

100

200

300

400

500

600

Nor

mal

ized

Xis

t le

vel

0 4 8 12Day of ES differentiation

WT

WT

Tsix

TST

237(cas)

142(129)

(bp)WT

day 0 day 12

Tsix

TST

E

60%, n=61

Figure 1. Jpx Expression Increases 10- to 20-Fold during ES Cell

Differentiation

(A) The Xic and its noncoding genes. Rnf12 is coding and lies 500 kb away.

(B) Time-course analyses of Jpx expression by qRT-PCR in differentiating

female and male ES cells. Averages and standard error (SE) from three (female)

or four (male) independent differentiation experiments are plotted. Values are

normalized to Gapdh RNA and d0 Jpx levels are set to 1.0.

(C) Time-course analyses of Xist expression by qRT-PCR in differentiating

male and female ES cells. Averages and SE from six (male) and three (female)

independent differentiation experiments are plotted. All values are normalized

to Gadph RNA and d0 Xist is set to 1.0.

(D) Allele-specific RT-PCR analysis of Jpx in wild-type and TsixTST/+ female ES

cells on d0 and d12 of differentiation.

(E) RNA FISH indicates that Jpx escapes inactivation in 60% of d16 female

cells. N = 61. Xist clouds are present in 98% of cells. Xist RNA, green. Jpx

RNA, red.

Cell 143, 390–403, October 29, 2010 ª2010 Elsevier Inc. 391

Page 80: Cell 101029

Day of differentiation

Cel

l dea

th (%

)

20

401F31F8

WT

WT

Δ Jp

x/+

d0 d4 d8 d12

250μ 250μ 550μ 550μ

Time window for XCI

WT7B7 n

eo+

1F8 n

eo-

7H6 n

eo+

1F3 n

eo-

(kb)13.2 9.6 8.6

WTΔJpx Neo+ΔJpx Neo-

clone 1 clone 2

237 142 95/83

M 129 Cas 7H6 7B7 (bp)

ΔJpx/+ (1F8)

Xist Jpx

B

E

C

F

DAPI

Xist

Jpx

20

60

100

Day of differentiation

% C

ells

with

Xis

t RN

A fo

ciW

TΔJ

px/+

d0 d4 d8 d12

0 4 8 12

0

40

80

1F31F8

WT

G H

5 Xist1234

Bgl I

Sac I

Neo

BstZ17

I

Avr II

Pst I

3

5 34 Neo

LoxP LoxP

5 3

LoxP

Wildtype Sp

e I

4

Spe I

Mfe I

4

DT

(CpG)n

Targeting vector

Cre

ASa

c ISa

c I

ΔJpx Neo+

ΔJpx Neo-

Sac ISp

e I

Southern probes 1 2 3

Xist

Xist

Jpx exons

homologous targeting

0 4 6 10 12 16 20 24 288

05101520253035

Day of differentiation0 4 8 12 16

WT

Nor

mal

ized

Jp

x le

vel

D

ΔJpx/+

Figure 2. DJpx Causes Loss of XCI and Massive Cell Death in Female ES Cells

(A) The Jpx gene, targeting vector, and products of homologous targeting before and after Cre-mediated excision of the Neo positive-selection marker. DT,

diphtheria toxin for negative selection. (CpG)n, CpG island. Numbered boxes represent five Jpx exons.

(B) Top panel: Southern analysis of SacI-digested genomic DNA from DJpx/+ and WT female ES cells using probe 1. The Neo- female clones, 1F3 and 1F8, were

derived from the Neo+ 6H7 and 7B7 clones, respectively. Bottom panel: Allele-specific PCR analysis showed that the 129 allele was preferentially targeted over

the M. castaneus (cas) allele. The analysis for Neo+ 6H7 and 7B7 clones are shown. M, 100 bp markers.

(C) DNA FISH of DJpx/+ female ES cells. Xist probe (pSx9), FITC-labeled. The Jpx probe (Cy3-labeled, red) is located in the region of deletion.

392 Cell 143, 390–403, October 29, 2010 ª2010 Elsevier Inc.

Page 81: Cell 101029

nearly equal expression from both alleles; between d12 and d16,

expression from Xi accounted for 10%–35% of total Jpx.

RNA fluorescence in situ hybridization (FISH) showed that

98% of cells expressed Xist clouds, and Jpx RNA was present

on Xi in 60% (Figure 1E, n = 61). In such cells, Jpx RNA was

seen on both Xa and Xi. On Xi, Jpx RNA was always adjacent

to, not in, the Xist cloud, a juxtaposition characteristic of genes

that escape XCI (Clemson et al., 2006; Namekawa et al., 2010).

Thus, consistent with previous analysis (Chureau et al., 2002;

Johnston et al., 2002; Chow et al., 2003), our results indicate

ubiquitous, non-sex-specific Jpx expression. However, our

data demonstrate that Jpx upregulation is developmentally

regulated to correlate with Xist upregulation, and that Jpx

significantly escapes XCI.

Deleting Jpx Has No Effect on Male Cellsbut Is Female LethalTo test Jpx function, we knocked out a 5.17 kb region at the 50

end of Jpx that includes its major promoter, CpG island, and first

two exons (DJpx) (Figure 2A and Figure S1A available online). We

isolated four independently derived male ES clones and

confirmed homologous targeting by Southern analysis using

external and internal probes (Figure S1B and data not shown).

The Neo selectable marker was thereafter removed by Cre-

mediated excision. Following DNA FISH to verify the deletion

(Figure S1C), we analyzed two independent Neo- clones for

each. Because 1C4 and 1D4 male clones behaved identically,

we present data for 1C4 below.

DJpx/Y ES cells displayed no obvious phenotype when differ-

entiated into EB to induce XCI. Differentiation in suspension

culture from d0 to d4 (day 0 to 4) revealed no morphological

anomalies, and adherent outgrowth on gelatin-coated plates

after d4 yielded robust growth (Figure S1D). Consistent with

this, no elevation of cell death was detected (Figure S1E).

RT-PCR analysis showed that Xist was appropriately sup-

pressed during differentiation (Figure S1F), RNA FISH confirmed

that basal Xist expression became repressed (Figure S1G).

Furthermore, the X-linked genes, Pgk1, Mecp2, and Hprt, were

all expressed appropriately (Figures S1F and S1G). Strand-

specific qRT-PCR showed that Xist and Tsix levels in mutants

were not significantly different from those of wild-type cells at

any time (Figure S1H). We conclude that deleting Jpx has no

functional consequence for XY cells.

We also deleted Jpx in a hybrid female ES line (16.7) carrying X

chromosomes of different strain origin (X129/Xcas) (Lee and Lu,

1999). We isolated five independent female clones, verified

homologous targeting by Southern analysis using external and

internal probes (Figures 2A and 2B and data not shown), and

then removed the Neo marker by Cre-mediated excision.

Allele-specific analysis showed that, in all five cases, X129 was

targeted (Figure 2B), consistent with the targeting vector’s 129

origin. Following DNA FISH to confirm the deletion (Figure 2C),

we analyzed two independent Neo- clones, 1F3 and 1F8. RNA/

DNA FISH showed that >95% of mutant cells are XX throughout

differentiation. The two female clones behaved similarly.

To quantitate residual Jpx levels inDJpx/+ cells, we performed

qRT-PCR and found less RNA than expected (Figure 2D). On d0,

targeting of a single allele resulted in loss of approximately half of

Jpx RNA, as expected. However, during differentiation, Jpx

levels from the wild-type castaneus allele did not increase to

the extent anticipated. Between d8 and d16, Jpx was expressed

at only 10%–20% of wild-type levels (50% expected). This

disparity could not be explained by strain-specific differences,

as allele-specific analysis of wild-type cells demonstrated similar

allelic levels between d0 and d12 (Figure 1D). Deleting one Jpx

allele therefore resulted in effects on the homologous allele, sug-

gesting an expression feedback loop. Thus, a heterozygous

deletion severely compromises overall Jpx expression and

approximates a homozygous deletion.

To investigate effects on XCI, we differentiated ES cells into EB

to induce XCI. Although DJpx/+ and wild-type cells were indistin-

guishable on d0, differentiation uncovered profound effects.

Wild-type EB typically showed smooth and radiant borders

between d2 and d4 when grown in suspension, but mutant EB

exhibited necrotic centers, irregular edges, and disaggregation

(Figure 2E, arrows). The difference became more obvious during

the adherent phase (post-d4). Whereas wild-type EB adhered to

plates and displayed exuberant cellular outgrowth, mutant EB

attached poorly and showed scant outgrowth. The difference

was not due to Jpx effects on cell differentiation per se, as immu-

nostaining of stem cell markers showed that mutant EB appro-

priately downregulated Oct4 and Nanog upon differentiation

(Figure S2). Thus, whereas DJpx had little effect in males,

deleting one Jpx allele in females caused severe abnormalities

during differentiation.

The female-specific nature suggested a link to XCI, a process

tightly coupled to cell differentiation (Monk and Harper, 1979;

Navarro et al., 2008; Donohoe et al., 2009). To test this possi-

bility, we performed a time-course analysis of Xist expression

by RNA FISH (Figures 2F and 2G). In wild-type cells, XCI was

largely established by d8–d12, with 75.0% ± 4.8% (mean ±

SE) of female cells displaying large Xist clusters by d8 and

89.1% ± 3.4% by d12. However, in DJpx/+ cells, Xist upregula-

tion was severely compromised, with only 6.35% ± 1.77%

displaying Xist foci on d8 and no major increase on d12.

Strand-specific RNA FISH confirmed that large RNA clouds

(D) Time-course analyses of Jpx expression by qRT-PCR in differentiating WT and DJpx/+ female ES cells. Averages and standard errors (SE) from three

independent differentiation experiments are plotted, with values normalized first to Gapdh and then d0 WT Jpx levels are set to 1.0.

(E) Brightfield photographs of WT and DJpx/+ female ES cells from d0 to d12 of differentiation. Arrows point to disintegrating, necrotic EBs present in mutant

cultures.

(F) RNA FISH to examine the time course of Xist upregulation. Xist probe, Cy3-labeled pSx9.

(G) Plotted time course of Xist upregulation in WT and two DJpx/+ mutants, 1F3 and 1F8. Averages ± SE from three independent differentiation experiments are

shown. Sample sizes (n): d0, 595–621; d4, 922–1163; d8, 3013–4370; d12, 3272–4794.

(H) Massive cell death in mutant female cells. The trypan blue staining results of three independent differentiation experiments were averaged and plotted with SE

d0, n = 150–800 cells for d0; d4, n = 200–500 cells; all other time points, n = 500–2000 cells.

See also Figures S1– S3.

Cell 143, 390–403, October 29, 2010 ª2010 Elsevier Inc. 393

Page 82: Cell 101029

during differentiation were of Xist origin and residual pinpoint

signals were of Tsix (Figure S3). The Xist deficiency mirrored

poor EB growth and massive cell death over the same time

course (Figures 2E and 2G). The disparity was greatest between

d4 and d12, when mutant cell death approached ten times that

of wild-type cells (Figure 2H). Between d4 and d12, at least 85%

of mutant cells were lost. Because dead cells detached from

culture, the actual percentage of Xist+ cells was probably even

lower (< < 6%) than measurable by collecting attached cells

for RNA FISH.

Our data argue that Jpx is an activator of Xist. DJpx differs

from DRnf12, which merely delays Xist induction by two days

and does not prevent XCI (Jonkers et al., 2009). We believe

that DJpx blocks XCI rather than delays it, because Xist clouds

were rare up to d16. Whereas DRnf12/+ cells are fully capable

of expressing Xist, DJpx/+ cells have severely compromised

Xist expression at all time points. Moreover, whereas DRnf12/+

cells are viable, DJpx/+ cells undergo massive cell death during

differentiation. Therefore, Jpx serves an essential function and

precludes Xist induction when deficient.

Jpx Acts in trans

Interestingly, DJpx’s influence on Xist was not restricted in cis to

X129 but also blocked Xist upregulation on Xcas, implying that,

unlike other Xic-encoded factors, Jpx may be trans-acting. If so,

expressing Jpx from an autosomal transgene might rescue

DJpx/+ cells. To test this, we introduced a 90 kb BAC carrying

full-length Jpx (and no other intact gene) (Figure 3A) into DJpx/+

cells (1F8) and characterized two independent clones, Jpx+/�;

TgB2and Jpx+/�;TgB3. Both clones carried autosomal insertions,

and qPCR using primer pairs at different transgene positions indi-

cated that each clone carried one to two copies of the full-length

transgene (Figure 3B and data not shown). In both clones, Jpx

levels were restored between d0 and d12 (Figure 3C).

Significantly, both clones behaved differently from DJpx/+

cells and were more similar to wild-type cells. Whereas DJpx/+

cells differentiated poorly and displayed elevated cell death,

Jpx+/�;TgB2 and Jpx+/�;TgB3 cells differentiated well and

were fully viable (Figures 3D and 3E). Moreover, Xist expression

was fully restored in Jpx+/�;TgB2 and Jpx+/�;TgB3 cells, both

in steady-state levels and in the number of cells with Xist clouds

(Figures 3F–3H). We conclude that an autosomal Jpx transgene

rescues the X-linked Jpx deletion and that Jpx must therefore be

able to act in trans.

Jpx Acts as a Long ncRNAIn principle, Jpx could function as a positive regulator in several

ways. Jpx could operate as enhancer, given 3C analysis showing

interaction between Jpx and Xist within a defined chromatin hub

(Tsai et al., 2008). However, a luciferase reporter assay in stably

transfected female ES cells uncovered no obvious enhancer

within the deleted Jpx region (Figure S4). In this assay, Jpx not

only failed to enhance luciferase expression but actually

depressed it in some cases. A relative increase in expression

occurred between d0 and d2, but activation never exceeded

that of the Xist-only construct. While we cannot exclude an

enhancer, enhancer function would be difficult to reconcile

with Jpx’s trans effects.

Jpx’s trans-acting property might be better explained by a

diffusible ncRNA. To distinguish RNA-based mechanisms from

those of DNA, chromatin, and/or transcriptional activity, we

used shRNA to deplete Jpx RNA after it is transcribed and to

knock down both Jpx alleles. We generated clones of wild-

type female ES cells carrying one of three Jpx-specific shRNAs

directed against nonpolymorphic regions of exon 1 (Figure 4A:

shRNA-A, -B, -C) and analyzed two to three independent clones

with good knockdown efficiency for each (e.g., shRNA-A1, -A2,

-A3). Controls carrying scrambled shRNA (Scr) were generated

and analyzed in parallel. Using qRT-PCR with primer pairs

positioned in exon 1, we observed 70%–90% depletion of Jpx

RNA (Figure 4B). Allele-specific RT-PCR showed that 129 and

castaneus alleles were symmetrically targeted (Figure 4C).

Because all clones behaved similarly, results are shown for

representative clones.

Phenotype analysis indicated that all knockdown clones reca-

pitulated DJpx. Knockdown clones grew indistinguishably from

wild-type on d0 and only lost viability upon differentiation

(Figures 4D and 4E). Between d0 and d4, EB formed by shRNA

clones were inferior in size and quality to those of wild-type

and Scr control (Figure 4E). Between d4 and d12, knockdown

EB showed poor outgrowth and underwent massive cell

death at magnitudes comparable to those for DJpx/+ cells (Fig-

ures 4D and 4E). Xist RNA FISH indicated a deficiency of Xist+

cells in differentiating knockdown clones (Figures 4F and 4G).

Similarly, qRT-PCR demonstrated significantly lower Xist levels

when Jpx RNA was knocked down by Jpx-specific shRNAs

(Figure 4H). These data showed that targeting both Jpx alleles

for posttranscriptional RNA degradation recapitulates the

heterozygous deletion.

In DJpx/+ cells, only 10%–20% of Jpx RNA remained, though

the castaneus allele was not deleted. To determine the conse-

quences of further Jpx deletion, we introduced shRNA-C into

the heterozygous cells (1F8) and depleted Jpx RNA by another

�50% (Figure 4I). Further depletion did not worsen the already

severe phenotype, as Xist upregulation remained similarly com-

promised and EB viability remained poor (Figure 4I), possibly

because Jpx was already largely abrogated. Thus, posttran-

scriptional depletion of Jpx RNA achieves the equivalent of the

Jpx�/�state (�10% residual RNA) and argues that Jpx acts as

a long ncRNA.

Jpx Has a Mild cis PreferenceWhile DJpx eliminated almost all female cells during differentia-

tion, a very small subset persisted past d20 and continued to

proliferate, indicating that rare cells might bypass DJpx. To

investigate the XCI status of surviving cells, we expanded

survivors to d28, performed Xist RNA FISH, and found that Xist

induction occurred in almost all survivors (Figure 5A). To ask

which of two Xist alleles was upregulated, we performed allele-

specific RNA-DNA FISH and observed that Xist was induced

monoallelically from X129 or Xcas (Figure 5B; RNA/DNA FISH

showed that >95% of mutant cells are XX; only XX cells were

counted). However, Xcas was favored by a ratio of 65:35 in d28

survivors (Figure 5C), indicating that DJpx is a disadvantage for

the Xist allele linked to it. Allele-specific RT-PCR of Xist, Pgk1,

Mecp2, and Hprt ratios confirmed these findings (Figure 5D).

394 Cell 143, 390–403, October 29, 2010 ª2010 Elsevier Inc.

Page 83: Cell 101029

AXist

Tg

Tg

Xist

WT

Jp

x+

/-

+T

gB

2

Jp

x+

/-

+T

gB

3

DAPI

d8

D

d4d0

250μ 500μ75μ

d12

500μ

WT

Jp

x+

/-

+T

gB

2

Jp

x+

/-

(1F

8)

0 4 8 12Day of differentiation

G

0

10

20

30

40

E

0 4 8 12

50 WTJpx+/-Jpx+/-; TgB2Jpx+/-; TgB3

Day of differentiation

Cel

l dea

th (%

)

WTJpx+/-Jpx+/-; TgB2Jpx+/-; TgB3

% N

ucle

i with

Xis

t foc

i

15μ

d12d8d0

C

F

H

WT

Jp

x+

/-

+T

gB

2

Jp

x+

/-

(1F

8)

0 4 8 12Day of differentiation

0

5

10

15

20

25

30

0 4 8 12Day of differentiation

Xist

RN

A le

vels

WTJpx+/-Jpx+/-; TgB2Jpx+/-; TgB3

WTJpx+/-Jpx+/-; TgB2Jpx+/-; TgB3

B

Jpx Xist

Tsix

Xite Tsx

10 Kb

Cnbp2

Jpx BAC Tg

Ftx

RepA

20

40

60

80

100

0

20

40

60

80

100

0

Jpx

RN

A le

vels

Figure 3. Transgenic Jpx Rescues DJpx in trans

(A) Map of the Xic and 90 kb Jpx transgene.

(B) Multiprobe DNA FISH to localize Xist (pSx9, red) and Jpx (BAC, green) in two independent transgenic clones, TgB2 and TgB3. Arrows, Jpx transgene.

(C) Time-course analyses of Jpx expression by qRT-PCR in differentiating cells of indicated genotype. Averages ± SE from three independent differentiation

experiments are plotted. Values are normalized to Gapdh RNA and WT d0 Jpx level is set to 1.0.

(D) Brightfield photographs of WT and transgenic EB from d0 to d12.

(E) Cell death analysis of WT, knockout, and transgenic EB, performed as above.

(F) RNA FISH to examine the time course of Xist upregulation. Xist probe, Cy3-labeled pSx9.

(G) Quantitation of WT, knockout, and transgenic EB with Xist RNA foci (RNA FISH) from d0 to d12.

(H) qRT-PCR of steady-state Xist levels in WT, knockout, and transgenic EB from d0 to d12.

Cell 143, 390–403, October 29, 2010 ª2010 Elsevier Inc. 395

Page 84: Cell 101029

d0

d4

d8

d12

200μ

500μ

500μ

100μ

0

10

20

30

A E

B

D

G

Cel

l dea

th (

%)

0 4 8 12Day of differentiation

H

shRNA-C1shRNA-Scr

0

20

40

60

80

100

0100200300400500600

700

% N

ucle

i with

Xis

t RN

A fo

ci

Nor

mal

ized

Xist le

vels

0 4 8 12Day of differentiation

0 4 8 12Day of differentiation

Exon 1

Avr II(CpG)n

Sac I

Sac I

shRNAqPCR primers e1-R e1-F

Xist123

Bgl I

BstZ17

I

Avr II

Pst I

Spe I

Spe I

Mfe I

(CpG)n

Sac I

Sac I

Jpx

Jpx

ACB

0 4 8 12Day of differentiation

00.20.40.60.81.01.21.4

0 4 8 12

shRNA-A

shRNA-C

shRNA-B

WTScrA1A2A3

00.20.40.60.81.01.21.4

Day of differentiation0 4 8 12

Day of differentiation

WTScrB1B2B3

WTScrC1C2

WTScrC1C2

Nor

mal

ized

Jp

x le

vels

d0 d8

10μ

shRNA-C1shRNA-Scr

10μ

d0 d8F

Day of differentiation0 4 8 12

Day of differentiation0 4 8 12

0100200300400500600700

0

100200300400500600700 shRNA-A

WTScrA1A2A3

shRNA-BWTScrB1B2B3

shRNA-CWTScrC1C2

shRNA-CWTScrC1C2

I

Day of differentiation0 4 8 120 4 8 12

Day of differentiation

Cel

l dea

th (%

)

% N

ucle

i with

Xis

t RN

A fo

ci

WT1F81F8-C51F8-C7

Nor

mal

ized

Jp

x le

vels

0 4 8 12Day of differentiation

02468

101214161820

200

0

300400500600700800

Nor

mal

ized

Xis

t le

vels

0 4 8 12Day of differentiation

C

(bp)

237 (cas)142 (129)

WT C1A1 B1

shRNA

55 53 52 54 %129

00.20.40.60.81.01.21.4

100

100

80

60

40

20

0

60

40

20

0

Nor

mal

ized

Jp

x le

vels

Nor

mal

ized

Jp

x le

vels

Figure 4. Jpx Functions as a Long ncRNA

(A) A map of the 50 end of Jpx showing its exons (purple), shRNA locations, and qPCR primer positions.

(B) Significant knockdown of Jpx RNA in two to three independent clones for each Jpx-specific shRNA, but not in the scrambled shRNA clone (Scr). Jpx RNA

levels are normalized to WT levels for each day of differentiation. A1–A3 are clones for shRNA-A; B1–B3 for shRNA-B; and C1, C2 for shRNA-C.

(C) Residual Jpx RNA was extracted from d8 shRNA clones, A1, B1, and C1, and subjected to allele-specific RT-PCR (Nla-III polymorphism). The gel was blotted

and hybridized to an end-labeled oligo. Allelic fractionation shows similar ratios of 129:castaneus bands in WT and knockdown clones, suggesting that the

396 Cell 143, 390–403, October 29, 2010 ª2010 Elsevier Inc.

Page 85: Cell 101029

RNA FISH also demonstrated that Xist upregulation led to

silencing of genes in cis (Figure 5E), demonstrating that Jpx

does not affect gene silencing per se. The observed allelic biases

were the opposite of wild-type, which ordinarily favors inactivat-

ing X129 due to the strain-specific Xce modifier (Cattanach and

Isaacson, 1967). Thus, although trans-acting, Jpx has a measur-

able cis preference that is uncovered only in rare female survi-

vors (Figure 5F).

Antagonism between Tsix and Jpx in the Control of XistSeveral models for Xist regulation postulate a balancing act

between positive and negative factors (Lee and Lu, 1999; Lee,

2005; Monkhorst et al., 2008; Navarro et al., 2008; Donohoe

et al., 2009; Starmer and Magnuson, 2009; Ahn and Lee,

2010). Xite and Tsix clearly reside in the repressive regulatory

arm (Lee and Lu, 1999; Sado et al., 2001; Ogawa and Lee,

2003). DJpx’s phenotype suggests that Jpx may reside in

a parallel, opposing arm. To test the idea of Jpx and Tsix antag-

onism, we targeted the TsixTST mutation (Ogawa et al., 2008) into

DJpx/+ cells to truncate Tsix RNA on the chromosome bearing

DJpx (Figure 6A). Targeting was confirmed by Southern blot

analysis and allele-specific genotyping (Figure 6B and data not

shown). Intriguingly, truncating Tsix almost completely restored

viability and differentiation of DJpx/+ cells. Cell death analysis

showed that two independently derived double mutants, 1F8-

S1 and 1F8-S2, have reduced cell death between d6 and d12

when compared to the single mutant (Figure 6C). Cell death

was comparable to that of wild-type EB, though significantly

higher between d4 and d6. Furthermore, unlike single mutants,

double mutants exhibited normal EB morphology and outgrowth

(Figure 6D) and RNA FISH showed restoration of Xist upregula-

tion and kinetics (Figures 6E and 6F). These results demonstrate

that TsixTST suppresses DJpx.

We next asked how allelic choice was further affected in Jpx-

Tsix double mutants. Single mutations both skew XCI ratios, but

the polarity is opposite: TsixTST/+ cells exclusively inactivate X129

(Ogawa et al., 2008), whereas DJpx/+ survivors preferentially

inactivate Xcas (Figure 5). In the double mutant, allele-specific

RT-PCR for Xist, Pgk1, and Mecp2 expression revealed Tsix’s

dominance over Jpx (Figure 6G). Abrogating Tsix RNA not only

overcame the block to transactivate Xist, but also skewed choice

to favor X129. Therefore, when Tsix RNA is eliminated, the linked

Xist allele is induced despite a Jpx deficiency. To determine

whether further reduction of Jpx by shRNA knockdown affected

the rescue, we introduced shRNA-C into the double mutant but

did not observe additional effects on Xist expression or cell

viability (Figures 6H and 6I).

In principle, the rescue of DJpx by TsixTST could be interpreted

in two ways. One idea is that Tsix and Jpx reside a single genetic

pathway in which Jpx occurs upstream of Tsix and controls Xist

expression by suppressing Tsix’s repressive effect on Xist. We

do not favor this idea, given that deleting Jpx did not affect

Tsix levels in male cells (Figure S1H). Moreover, the Tsix-Jpx

double mutant was not identical in phenotype to TsixTST, as the

double mutant still demonstrated elevated cell death at early

time points in spite of rescuing Xist expression (Figure 6C).

Thus, we believe that the data collectively argue for parallel

pathways in which Tsix and Jpx independently control Xist

transcription. In this scenario, how can Xist be induced in double

mutants? One possibility is that residual Jpx levels from Xcas

were sufficient to activate Xist in trans. This alone cannot explain

the rescue, however, as residual Jpx from Xcas could not upregu-

late Xist at all in DJpx/+ cells (Figures 2D, 2F, and 2G). We

propose that eliminating the negative arm of regulation (via

TsixTST) created a hyper-permissive state for Xist upregulation

in which even very low Jpx expression might be sufficient to

induce Xist expression.

DISCUSSION

Our work demonstrates that Xist is controlled by two parallel

RNA switches—Tsix for Xa and Jpx for Xi. Whereas Tsix

represses Xist on Xa, Jpx activates Xist on Xi. How Jpx RNA

transactivates Xist is yet to be determined, but it is intriguing

that expression of one long ncRNA would be controlled by

another. Recapitulation of the knockout by posttranscriptional

knockdown of Jpx implies that the activator acts as an RNA.

Unlike other ncRNAs of the Xic, Jpx is trans-acting and diffusible.

Indeed, autosomally expressed Jpx RNA can rescue the

X-linked DJpx defect. We cannot exclude the possibility that

Jpx also acts as an enhancer, though our reporter assay did

not uncover such a property (Figure S4). Interestingly, 3C anal-

ysis previously revealed close chromatin contact between the

50 ends of Jpx and Xist in cis (Tsai et al., 2008). Their physical

proximity may underlie Jpx’s preference for the linked Xist allele

(Figure 5), as a diffusion-limited Jpx RNA would be expected to

preferentially bind the Xist allele closer to it.

Our findings place Jpx’s function in an epistatic context

(Figure 7A). Prior work has proposed that Xite and Tsix reside

at the top of the repressive pathway, controlling XCI counting

shRNAs affected both Jpx alleles. Only 10%–30% of Jpx RNA was left in the knockdowns and therefore the PCR was overcycled to visualize the low residual

levels of Jpx in the knockdown cells.

(D) Cell death assay shows that loss of Jpx RNA reduces cell viability during differentiation. Clones shRNA-C1 and -C2 are shown, but shRNA-A and -B clones

also show increased cell death.

(E) Brightfield images show poor EB formation and outgrowth in knockdowns but not Scr control.

(F) Xist RNA FISH shows loss of Xist upregulation when Jpx is knocked down using shRNA-C.

(G)Quantitationof the numberofcells withXistRNAclusters fromthree independentdifferentiationexperimentsofcontrolandknockdownclones.Average±SEshown.

(H) Quantitation of Xist RNA levels in control and knockdown clones from three independent differentiation experiments. RNA levels are normalized to d0 WT

values. Average ± SE shown. Differentiation of shRNA-A and -B knockdown clones were performed at the same time; therefore, WT and Scr values for

shRNA-A and shRNA-B are the same.

(I) Jpx knockdown inDJpx/+ cells (1F8) using shRNA-C. Independent clones, C5 and C7, behaved similarly to each other and also to their parent, 1F8, in all assays

shown. Average ± SE shown. All values are normalized to d0 WT.

See also Figure S4.

Cell 143, 390–403, October 29, 2010 ª2010 Elsevier Inc. 397

Page 86: Cell 101029

and choice by inducing homologous chromosome pairing

through Oct4 (Bacher et al., 2006; Xu et al., 2006; Donohoe

et al., 2009). X-X pairing would play an essential role in breaking

epigenetic symmetry by shifting the binding of Tsix- and Xite-

associated transcription factors from both X’s to the future Xa

(Xu et al., 2006; Nicodemi and Prisco, 2007; Donohoe et al.,

2009; Lee, 2009). Retained transcription factors would then

sustain Xite and Tsix expression and block Xist activation on

5um5um

A

C E EB d28

d28 EB

DAPI Xist Jpx Merge

Xi = XWT

Xi = XΔ

WT

ΔJp

x/+

DAPI Xist Pgk1 Merge

EB d28

Xist

Tsix

Pgk1

Mecp2

Hprt

+ 0

129

day 12 8 4

0.70

0.44 0.430.38 0.46 0.45 0.52 0.430.52

0.40 0.460.40 0.45 0.47 0.47 0.390.51

129

fraction

cas

129cas

0.75 0.75 0.71 0.73 0.74 0.76 0.78

0.67 0.60 0.60 0.52 0.510.650.64

129cas

129cas

129cas

0.52 0.53 0.47 0.41 0.45 0.44

0.54 0.49 0.53 0.60 0.64 0.650.530.52

Δ16 282420

0.41 0.41 0.46 0.39 0.36 0.43 0.38 0.40

0.42 0.41 0.40 0.59 0.59 0.620.480.46

0.62

0.45 0.650.45 0.44 0.44 0.720.67

0.48 0.450.45 0.44 0.40 0.51 0.400.47

0.46

0.44 0.40

Day of differentiation

20

80

100

0

60

40

41.2%+/-1.9%

35% +/- 2.7%

63.3%+/-3.0%

+ Δ + Δ + Δ + Δ + Δ + Δ + Δ

% n

ucle

i whe

re X

i = X

Δ ΔJpx/+

10μ

8 16 28

n=48

n=230

n=946

WT ΔJpx/+

97%, n > 2000 92% , n > 3000

% Xist+

ΔJpx/Y

ΔJpx/+

Massive cell death due to failed upregulationof either Xist allele(Jpx’s trans effect)

Normal:XCI not necessary

Raresurvivor

Allelic skewingof Xist

(Jpx’s cis effect)

F

D

B

129

fraction

129

fraction

129

fraction

129

fraction

Figure 5. Jpx’s Mild cis Preference Revealed in DJpx/+ Survivors

(A) Xist RNA FISH on d28 EB. Xist probe, Cy3-labeled pSx9. WT, 97% Xist+ cells (n > 2000). DJpx/+, 92% Xist+ cells (n > 3000).

(B) Allele-specific RNA/DNA FISH determines which X is Xi. FITC-labeled pSx9 probe detects Xist RNA and the Xist locus from both Xs, whereas the Cy3-labeled

Jpx probe detects only the wild-type X (the probe resides in the deleted region).

(C) Percentage of 1F8 mutant female cells where Xi = XD (i.e., X129). Averages ±SE from three independent differentiation experiments.

(D) Allele-specific RT-PCR of indicated transcripts from d0 to d28. The percentage of transcripts from the 129 allele (%129) is determined by phosphorimaging. +,

WT. D, 1F8 mutant. Values for lanes that are not visible are obtained after a longer exposure.

(E) Two-color RNA FISH for Xist and Pgk1 transcripts in d28 cells.

(F) Summary of DJpx effects on male and female ES cells.

398 Cell 143, 390–403, October 29, 2010 ª2010 Elsevier Inc.

Page 87: Cell 101029

Xa (Stavropoulos et al., 2001; Ogawa and Lee, 2003), in part by

interfering with the action of RepA RNA and Polycomb proteins

(Sun et al., 2006; Zhao et al., 2008).

Work from the current study supports the existence of a

parallel, but activating pathway for establishment of Xi. Jpx

resides in this pathway (Figure 7A). The RNA is upregulated

10- to 20-fold during ES differentiation and leads to monoallelic

Xist induction in female cells. The collective evidence suggests

that Jpx and RepA RNA collaborate to transcriptionally activate

Xist. In this model, loss of Tsix expression on the future Xi would

enable the RepA-Polycomb complex to load onto the Xist

chromatin and trimethylate H3-K27 on the Xist promoter (Zhao

et al., 2008), creating a permissive state in which Jpx RNA could

transactivate Xist.

In male cells, Jpx upregulation does not result in Xist induction

on the single X—similar to the Xa of female cells. As would be the

case for the female Xa, persistence of Tsix in male cells overrides

Jpx by recruiting silencers to the Xist promoter (Navarro et al.,

2005; Sado et al., 2005; Sun et al., 2006). In the context of Tsix

regulation, DNA methylation and RNAi have been invoked in

Xist silencing (Norris et al., 1994; Ariel et al., 1995; Zuccotti and

Monk, 1995; Sado et al., 2005; Sun et al., 2006; Ogawa et al.,

2008). By this model, female cells deficient for Jpx would be

unaffected on d0 because Jpx is normally not induced until cell

differentiation and the onset of XCI. Once induced, Jpx RNA

remains at high levels in somatic cells (Figure 1), implying that

continued presence of the activator may be necessary for lifelong

Xist expression in the female. Jpx may also play other roles during

development, given that the Tsix-Jpx double mutant rescues Xist

expression but does not fully rescue cell death (Figure 6).

In conclusion, our study identifies Jpx as an RNA-based

activator of Xist and supports a dynamic balance of activators

and repressors for XCI control. The fate of Xist appears to be

determined by a series of Xic-encoded RNA switches, reinforc-

ing the idea that long ncRNAs may be ideally suited to epigenetic

regulation involving allelic and locus-specific control (Lee, 2009).

Future work will help elucidate why the Xic, once protein-coding,

was replaced in recent evolutionary history by noncoding genes.

EXPERIMENTAL PROCEDURES

Construction of DJpx Cell Lines

Male (J1) and female (16.7) ES cell lines, culture condition, and cell differenti-

ation protocols have been described (Lee and Lu, 1999). To generated DJpx,

a 50 homology arm (6.5 kb BstZ17I-BglI of Jpx) was cloned into the NotI site of

vector, PgkNeo2LoxDTA. To the resulting vector was cloned the 30 arm

(6.19 kb AvrII-PstI fragment) into the NheI-SalI site, yielding a 5.17 kb deletion.

The targeting construct was linearized with XhoI and electroporated. For

screening, �2000 male and �2500 female G418-resistant clones were picked,

and 4 male and 5 female independent knockout clones were analyzed. To

excise the Neo selection marker, a Cre plasmid (pMC-CreN) was introduced

and G418-sensitive colonies identified. Homologous targeting was confirmed

by genomic Southern blots using 50 and 30 external probes, as well as internal

probes to rule out ectopic integrations. The templates for 50 and 30 external

probes were PCR products generated using primers: GAGCTCTGAGACA

CAGCGCAA and GCCAAAGGGGTTGTCATCTATG for the 50 probe (nt

84779–85380 of GenBank sequence AJ421479); and GCCCAGGAACTGA

GTTTTAGCACA and TGCTTATGGACGATCAAAGTGCC for the 30 probe

(nt 104761–105450 of AJ421479). To determine which allele was targeted in

females, we carried out allele-specific PCR analysis based on a Nla-III poly-

morphic site at nt 95,738 (GenBank sequence AJ421479) within Jpx (CATG

for the 129; CAAG for castaneus). Genomic DNA was amplified by primer pairs,

JpxUp (cggcgtccacatgtatacgtcc) and JpxLo (taggaatgagcctccccagcct) (Chur-

eau et al., 2002), to generate a 329 bp product (nt 95598–95926 of GenBank

AJ421479), which was then digested with Nla-III to yield 142, 95, 83, and

9 bp fragments for 129 and 237, 83, and 9 bp fragments for castaneus. All

female clones showed targeting of the 129 allele.

Generation of Transgenic Jpx Cell Lines

A 90 kb BAC transgene containing full-length Jpx (and no other known tran-

scribed sequences) and a Neo resistance marker was made by ET-cloning

(Yang and Seed, 2003) from BAC clone 399K20 (Invitrogen). Ten million 1F8

cells (DJpx/+) were electroporated with 20 mg linearized BAC DNA and

cultured under G418 selection (250 mg/ml). Two G418-resistant clones

(TgB2 and TgB3) were picked on d8 and expanded for analysis.

Generation of TsixTST DJpx/ ++ ES Cells

The TsixTST truncation vector has been described (Ogawa et al., 2008). The

DJpx/+ female line, 1F8, was electroporated with the TsixTST vector, 96 clones

were picked after puromycin selection, and targeting into X129 was determined

by Southern blot analysis and allele-specific PCR, as described (Ogawa et al.,

2008). Two independent clones, 1F8-S1 and 1F8-S2, were analyzed in parallel.

Generation of Jpx Knockdown Clones

To generate three shRNA knockdown plasmids, three nonpolymorphic

sequences from Jpx exon 1 were inserted into the EcoRI and NheI site of

pLKO1 (Addgene): shRNA-A, 50-CCGGcaccaggcttctgtaacttatCTCGAGataa

gttacagaagcctggtgTTTTTG-30; shRNA-B, 50-CCGGtagaggatgacttaataagga

CTCGAGtccttattaagtcatcctctaTTTTTG-30; and shRNA-C: 50-CCGGGGCGT

CCACATGTATACGTCCCTCGAGGGACGTATACATGTCGACGCCTTTTTG-30.16.7 cells were electroporated with either Jpx-specific or a scrambled (Scr)

shRNA vector and selected with puromycin for stable integration. Multiple

independent clones were picked (24 for shRNA-A, 24 for shRNA-B, and 48

for shRNA-C) and tested for Jpx knockdown efficiency by qRT-PCR (see

Quantitative RT-PCR). We analyzed two to three independent clones for each.

RNA and DNA FISH

FISH protocols and probes (Xist, Pgk1) have been described (Lee and Lu,

1999; Stavropoulos et al., 2001). The Jpx probe is a 3.7 kb fragment

(nt 93362–97039 of GenBank AJ421479) within the deleted region that was

cloned into pCR-Blunt II-Topo vector (Invitrogen) for Nick translation. For

two-color strand-specific RNA FISH, an FITC-labeled Xist riboprobe cocktail

was generated by in vitro transcription (MAXIscript kit, Ambion) to detect the

Xist strand, and Tsix was detected by Cy3-labeled pCC3, a 50 Tsix probe

that does not overlap Xist (Lee et al., 1999a; Ogawa and Lee, 2003).

Quantitative RT-PCR

Real-time PCR for Xist, Tsix, and Jpx was performed under the following

conditions: 95�C 3 min; 95�C 10 s, 60�C 20 s, 72�C 20 s, for a total of 40 cycles,

and 72�C 5 min. Melting curves for primer pairs were determined by increasing

temperatures from 60�C to 95�C at 0.5�C interval for 5 s. Primers for Xist

qRT-PCR were NS66 and NS33, and for Tsix NS18 and NS19 (Stavropoulos

et al., 2001). Primers for Jpx were e1-F, GCACCACCAGGCTTCTGTAAC,

and e1-R, GGGCATGTTCATTAATTGGCCAG.

Allele-Specific RT-PCR

Allele-specific RT-PCR was performed as described (Stavropoulos et al.,

2001; Ogawa and Lee, 2003). Total RNA was extracted by Trizol (Invitrogen)

and DNA was removed with DNase I treatment (Ambion). Reverse transcription

was then performed with SuperScript III First Strand Synthesis System (Invitro-

gen). Allele-specific primers were: NS66 and NS33 for Xist (Stavropoulos et al.,

2001), NS18 and NS19 for Tsix (Stavropoulos et al., 2001), NS43 and NS44 for

Mecp2 (Ogawa and Lee, 2003), KH106 and KH107 for Pgk1 (Huynh and Lee,

2003), and NS41 and NS70 for Hprt (Stavropoulos et al., 2001). Southern

blot was carried out using nested primers as probes as referenced above:

XSP1 for Xist, NS19 for Tsix, NS65 for Mecp2, KH106 for Pgk1, and NS59

for Hprt. For Jpx allele-specific RT-PCR, Jpx cDNA was amplified with JpxLow

Cell 143, 390–403, October 29, 2010 ª2010 Elsevier Inc. 399

Page 88: Cell 101029

BA

C

0 4 6 8 12

20

40

80

100

10

20

0

30

40

50

WT m

ale

WT f

emale

1F8

1F8-S

21F

8-S1

WT 6.8kb (129) WT 6.5kb (cas) TsixTST 5.6kb (129)

0

60

60

120 4 6 8

FE

Day of differentiation

Cel

l dea

th (%

)

Day of differentiation

% C

ells

with

Xis

t RN

A fo

ci

WT1F8-S11F8-S21F8

WT1F8-S11F8-S21F8

Tsix

EcoR

VSa

cII

SphI

Hind

III

SalI

BsrF

1Ec

oRV

EcoR

V

Tsix truncation vectorSouthern probe

ΔJpx

Homologous targeting

EcoR

VSa

cII

EcoR

V

SalI

BsrF

1

EcoR

V

DXPas34

LoxP trpA Puro IRES

LoxP trpA Puro IRESDXPas34

ΔJpx; TsixTST

ΔJpx

ΔJpx

Xist

Pgk1

Mecp2

129cas

129cas

129cas

0day 12 8 4

0.62 0.97

0.65 0.98

0.53 0.99

0.57 0.99

0.67 0.99

0.69 0.990.66 0.99

0.50 0.99

0.49 0.47

0.49 0.47

0.45 0.42

0.46 0.43

0.46 0.26

0.46 0.230.33 0.16

0.46 0.15

0.46 0.47

0.45 0.49

0.45 0.34

0.37 0.35

0.4 0.20

0.38 0.200.27 0.07

0.4 0.12

1F8WT S1 S2 1F8WT S1 S2 1F8WT S1 S2 1F8WT S1 S2

Tsix

d0

d8

10μ

WT (1F8-S1)(1F8)ΔJpx/+ ΔJpx;TsixTST

G

d0 d4 d8

WT

ΔJp

x/+

ΔJp

x;Tsix

TS

T

200μ 250μ 550μ

D

ΔJpx;TsixTST

ΔJpx;TsixTST

+C1

H

275μ0 4 8Day of differentiation

1000

1500

2000

2500

500

0Nor

mal

ized

Xis

t R

NA

leve

ls

WTΔJpx;Tsix

TST

ΔJpx;TsixTST

+C1

ΔJpx;TsixTST

+C2

I

129

fraction

129

fraction

129

fraction

Figure 6. A Tsix RNA Truncation Suppresses DJpx

(A) Targeting the Tsix truncation mutation (TsixTST) (Ogawa et al., 2008) to the DJpx chromosome in 1F8 female ES cells. TsixTST prematurely terminates Tsix at the

targeted triple polyA site (trpA) 1 kb downstream of the major Tsix promoter. 1F8-S1 and 1F8-S2 are two independently generated double mutant clones. IRES,

internal ribosome entry site. Puro, puromycin selection marker.

(B) Southern analysis using EcoRV digestion to confirm targeting. The X129 and Xcas alleles have an �300 bp DXPas34 length polymorphism. The X129 allele was

targeted in both 1F8-S1 and 1F8-S2.

(C) Cell death analysis shows that TsixTST partially rescues viability of DJpx/+ ES cells.

400 Cell 143, 390–403, October 29, 2010 ª2010 Elsevier Inc.

Page 89: Cell 101029

and JpxUp to generate a 329 bp product, which was digested with NlaIII.

End-labeled oligonucleotide, Jpx-P1, GGTGATGTGGGCACTGATCACTCATC,

was used as southern probe to recognize both castaneus specific 237 bp and

129 specific 142 bp band. All allelic signals were then quantitated by phos-

phorimaging.

Luciferase Assay

Promoterless pGL4.19 (Promega) was used to construct luciferase vectors.

To generate Jpx-pGL4, a 5.29 kb fragment (nt 92711–98009 of AJ421479),

corresponding to the knockout region (promoter, CpG island, and exons 1–2),

was cloned into the multiple cloning site. To construct Xist-pGL4, a 4.43 kb

fragment (nt 104971–109403 of AJ421479), containing the 500 bp region

upstream of Xist’s start site and the proximal 4 kb of exon 1, was cloned simi-

larly. Jpx-Xist-pGL4 was constructed by inserting the 5.29 kb Jpx fragment

upstream of Xist in Xist-pGL4 vector. Vectors were individually electroporated

into female ES cells, and 200–300 stably transfected clones from each vector

were pooled and subjected to luciferase assay at different differentiation time

points. qRT-PCR for luciferase was performed using primers, Luc-F1,

CAGCGCCATTCTACCCACTCG, and Luc-R1, GCTTCTGCCAGCCGAACGC.

Beta-actin was amplified as the internal control.

Cell Death Analysis

Cell death assays were performed as described (Stavropoulos et al., 2001). In

brief, on d0, both supernatant and attached ES cells were collected and

stained with trypan blue dye (Sigma). On other time points, both supernatant

and floating embryoid bodies (EBs on d4) or attached EBs (d6 and onward)

were collected and stained with trypan blue. The ratios of dead cells (blue)

to total cells (i.e., blue dead cells + clear viable cells) were calculated

and plotted as a function of time. Each sample was counted in duplicate or

triplicate.

Xist

TsixJpx/Enox

Pre-XCI(undifferentiated ES) XCI window XCI establishment

Xa

Xi

Xa

Xist

TsixJpx/Enox

Tsix blocks Xist induction and Jpxis expressed at low levels: Xistat basal levels.

Tsix blocks Xist induction and Jpxis expressed at low levels: Xistat basal levels.

Xist promoter is permanentlyrepressed and methylated. Jpx remains highly expressed but cannot induce Xist.

Jpx is highly expressed.

Xa: Xist promoter is methylated andpermanently repressed.

Xi: Jpx transactivates Xist. Initiationof chromosome-wide silencing.

Jpx is induced 10-20X.

Xa: Persistent Tsix blocks actionof Jpx and Xist induction. Tsix recruits silencers to the Xist promoter. Silencers =

Xi: Downregulation of Tsix rendersXist susceptible to activation by Jpx.Trans-acting Jpx =

Jpx is induced 10-20X.Persistent Tsix blocks Xist inductionand inactivates the Xist promoter byrecruiting silencers. Silencers =

Tsix

Xite

Xist

RepA

Jpx/Enox

XCI

Rnf12

A B

Figure 7. Model and Summary

(A) Proposed epistasis model: Xist is under positive-negative regulation by noncoding genes. Xite and Tsix repress Xist, whereas Jpx and RepA activate Xist.

Arrows, positive relationship. Blunt arrows, negative relationship. Rnf12 is a coding gene.

(B) Proposed events in male and female ES cells. Xist silencers (orange hexagons) include Dnmt3a and other chromatin modifications. Jpx (purple oval) is

depicted as a diffusible trans-acting RNA. Open lollipops, unmethylated Xist promoter. Filled lollipops, methylated Xist promoter.

(D) Brightfield photographs of wild-type, single, and double mutant female ES cells during differentiation.

(E) RNA FISH indicating that Xist upregulation (large red clouds) is rescued in double mutants.

(F) TsixTST restores Xist induction in DJpx/+ cells. Averages ± SD shown for three independent differentiation experiments.

(G) The pattern of allelic skewing is reversed in DJpx; TsixTST/+ cells.

(H and I) Further depletion of Jpx RNA by shRNA-C knockdown inDJpx; TsixTST/+ cells did not alter the phenotype of the double mutant, as shown by qRT-PCR of

Xist expression (H) and by EB outgrowth to d8 (I). DJpx; TsixTST/+, 1F8-S2. Two shRNA-C clones derived from 1F8-S2 were examined (C1, C2).

Cell 143, 390–403, October 29, 2010 ª2010 Elsevier Inc. 401

Page 90: Cell 101029

SUPPLEMENTAL INFORMATION

Supplemental Information includes four figures and can be found with this

article online at doi:10.1016/j.cell.2010.09.049.

ACKNOWLEDGMENTS

We thank Y. Jeon for providing Xist riboprobes, and A. Chess, D. Lessing, B.

Payer, and S. Pinter for critical reading of the manuscript and all members of

the laboratory for their helpful input throughout this project. This work was

funded by NIH grants K08-HD053824 to D.T., RO1-GM58839 to J.T.L., and

a Pathology training grant T32-CA009216. J.T.L. is also an Investigator of

the Howard Hughes Medical Institute.

Received: March 19, 2010

Revised: August 6, 2010

Accepted: September 17, 2010

Published: October 28, 2010

REFERENCES

Ahn, J.Y., and Lee, J.T. (2010). Retinoic acid accelerates downregulation of the

Xist repressor, Oct4, and increases the likelihood of Xist activation when Tsix is

deficient. BMC Dev. Biol. 10, 90.

Ariel, M., Robinson, E., McCarrey, J.R., and Cedar, H. (1995). Gamete-specific

methylation correlates with imprinting of the murine Xist gene. Nat. Genet. 9,

312–315.

Bacher, C.P., Guggiari, M., Brors, B., Augui, S., Clerc, P., Avner, P., Eils, R.,

and Heard, E. (2006). Transient colocalization of X-inactivation centres

accompanies the initiation of X inactivation. Nat. Cell Biol. 8, 293–299.

Brockdorff, N., Ashworth, A., Kay, G.F., McCabe, V.M., Norris, D.P., Cooper,

P.J., Swift, S., and Rastan, S. (1992). The product of the mouse Xist gene is

a 15 kb inactive X-specific transcript containing no conserved ORF and

located in the nucleus. Cell 71, 515–526.

Brown, C.J., Lafreniere, R.G., Powers, V.E., Sebastio, G., Ballabio, A.,

Pettigrew, A.L., Ledbetter, D.H., Levy, E., Craig, I.W., and Willard, H.F.

(1991). Localization of the X inactivation centre on the human X chromosome

in Xq13. Nature 349, 82–84.

Brown, C.J., Hendrich, B.D., Rupert, J.L., Lafreniere, R.G., Xing, Y., Lawrence,

J., and Willard, H.F. (1992). The human XIST gene: analysis of a 17 kb inactive

X-specific RNA that contains conserved repeats and is highly localized within

the nucleus. Cell 71, 527–542.

Cattanach, B.M., and Isaacson, J.H. (1967). Controlling elements in the mouse

X chromosome. Genetics 57, 331–346.

Chow, J.C., Hall, L.L., Clemson, C.M., Lawrence, J.B., and Brown, C.J. (2003).

Characterization of expression at the human XIST locus in somatic, embryonal

carcinoma, and transgenic cell lines. Genomics 82, 309–322.

Chureau, C., Prissette, M., Bourdet, A., Barbe, V., Cattolico, L., Jones, L.,

Eggen, A., Avner, P., and Duret, L. (2002). Comparative sequence analysis

of the X-inactivation center region in mouse, human, and bovine. Genome

Res. 12, 894–908.

Clemson, C.M., Hall, L.L., Byron, M., McNeil, J., and Lawrence, J.B. (2006).

The X chromosome is organized into a gene-rich outer rim and an internal

core containing silenced nongenic sequences. Proc. Natl. Acad. Sci. USA

103, 7688–7693.

Davidow, L.S., Breen, M., Duke, S.E., Samollow, P.B., McCarrey, J.R., and

Lee, J.T. (2007). The search for a marsupial XIC reveals a break with vertebrate

synteny. Chromosome Res. 15, 137–146.

Donohoe, M.E., Silva, S.S., Pinter, S.F., Xu, N., and Lee, J.T. (2009). The

pluripotency factor Oct4 interacts with Ctcf and also controls X-chromosome

pairing and counting. Nature 460, 128–132.

Duret, L., Chureau, C., Samain, S., Weissenbach, J., and Avner, P. (2006). The

Xist RNA gene evolved in eutherians by pseudogenization of a protein-coding

gene. Science 312, 1653–1655.

Hoki, Y., Kimura, N., Kanbayashi, M., Amakawa, Y., Ohhata, T., Sasaki, H., and

Sado, T. (2009). A proximal conserved repeat in the Xist gene is essential as

a genomic element for X-inactivation in mouse. Development 136, 139–146.

Hore, T.A., Koina, E., Wakefield, M.J., and Marshall Graves, J.A. (2007). The

region homologous to the X-chromosome inactivation centre has been

disrupted in marsupial and monotreme mammals. Chromosome Res. 15,

147–161.

Huynh, K.D., and Lee, J.T. (2003). Inheritance of a pre-inactivated paternal X

chromosome in early mouse embryos. Nature 426, 857–862.

Johnston, C.M., Newall, A.E., Brockdorff, N., and Nesterova, T.B. (2002).

Enox, a novel gene that maps 10 kb upstream of Xist and partially escapes

X inactivation. Genomics 80, 236–244.

Jonkers, I., Barakat, T.S., Achame, E.M., Monkhorst, K., Kenter, A.,

Rentmeester, E., Grosveld, F., Grootegoed, J.A., and Gribnau, J. (2009).

RNF12 is an X-encoded dose-dependent activator of X chromosome inactiva-

tion. Cell 139, 999–1011.

Koerner, M.V., Pauler, F.M., Huang, R., and Barlow, D.P. (2009). The function

of non-coding RNAs in genomic imprinting. Development 136, 1771–1783.

Lee, J.T. (2000). Disruption of imprinted X inactivation by parent-of-origin

effects at Tsix. Cell 103, 17–27.

Lee, J.T. (2005). Regulation of X-chromosome counting by Tsix and Xite

sequences. Science 309, 768–771.

Lee, J.T. (2009). Lessons from X-chromosome inactivation: long ncRNA as

guides and tethers to the epigenome. Genes Dev. 23, 1831–1842.

Lee, J.T., and Lu, N. (1999). Targeted mutagenesis of Tsix leads to nonrandom

X inactivation. Cell 99, 47–57.

Lee, J.T., Davidow, L.S., and Warshawsky, D. (1999a). Tsix, a gene antisense

to Xist at the X-inactivation centre. Nat. Genet. 21, 400–404.

Lee, J.T., Lu, N., and Han, Y. (1999b). Genetic analysis of the mouse X inacti-

vation center defines an 80-kb multifunction domain. Proc. Natl. Acad. Sci.

USA 96, 3836–3841.

Lucchesi, J.C., Kelly, W.G., and Panning, B. (2005). Chromatin remodeling in

dosage compensation. Annu. Rev. Genet. 39, 615–651.

Luikenhuis, S., Wutz, A., and Jaenisch, R. (2001). Antisense transcription

through the Xist locus mediates Tsix function in embryonic stem cells. Mol.

Cell. Biol. 21, 8512–8520.

Lyon, M.F. (1961). Gene action in the X-chromosome of the mouse (Mus

musculus L.). Nature 190, 372–373.

Marahrens, Y., Panning, B., Dausman, J., Strauss, W., and Jaenisch, R. (1997).

Xist-deficient mice are defective in dosage compensation but not spermato-

genesis. Genes Dev. 11, 156–166.

Mercer, T.R., Dinger, M.E., and Mattick, J.S. (2009). Long non-coding RNAs:

insights into functions. Nat. Rev. Genet. 10, 155–159.

Monk, M., and Harper, M.I. (1979). Sequential X chromosome inactivation

coupled with cellular differentiation in early mouse embryos. Nature 281,

311–313.

Monkhorst, K., Jonkers, I., Rentmeester, E., Grosveld, F., and Gribnau, J.

(2008). X inactivation counting and choice is a stochastic process: evidence

for involvement of an x-linked activator. Cell 132, 410–421.

Morey, C., Navarro, P., Debrand, E., Avner, P., Rougeulle, C., and Clerc, P.

(2004). The region 30 to Xist mediates X chromosome counting and H3 Lys-4

dimethylation within the Xist gene. EMBO J. 23, 594–604.

Namekawa, S.H., Payer, B., Huynh, K.D., Jaenisch, R., and Lee, J.T. (2010).

Two-step imprinted X inactivation: repeat versus genic silencing in the mouse.

Mol. Cell. Biol. 30, 3187–3205.

Navarro, P., Pichard, S., Ciaudo, C., Avner, P., and Rougeulle, C. (2005).

Tsix transcription across the Xist gene alters chromatin conformation without

affecting Xist transcription: implications for X-chromosome inactivation.

Genes Dev. 19, 1474–1484.

Navarro, P., Chambers, I., Karwacki-Neisius, V., Chureau, C., Morey, C.,

Rougeulle, C., and Avner, P. (2008). Molecular coupling of Xist regulation

and pluripotency. Science 321, 1693–1695.

402 Cell 143, 390–403, October 29, 2010 ª2010 Elsevier Inc.

Page 91: Cell 101029

Nicodemi, M., and Prisco, A. (2007). Symmetry-breaking model for X-chromo-

some inactivation. Phys. Rev. Lett. 98, 108104.

Norris, D.P., Patel, D., Kay, G.F., Penny, G.D., Brockdorff, N., Sheardown,

S.A., and Rastan, S. (1994). Evidence that random and imprinted Xist expres-

sion is controlled by preemptive methylation. Cell 77, 41–51.

Ogawa, Y., and Lee, J.T. (2003). Xite, X-inactivation intergenic transcription

elements that regulate the probability of choice. Mol. Cell 11, 731–743.

Ogawa, Y., Sun, B.K., and Lee, J.T. (2008). Intersection of the RNA Interference

and X-Inactivation Pathways. Science 320, 1336–1341.

Ohhata, T., Hoki, Y., Sasaki, H., and Sado, T. (2006). Tsix-deficient X chromo-

some does not undergo inactivation in the embryonic lineage in males:

implications for Tsix-independent silencing of Xist. Cytogenet. Genome Res.

113, 345–349.

Ohhata, T., Hoki, Y., Sasaki, H., and Sado, T. (2008). Crucial role of antisense

transcription across the Xist promoter in Tsix-mediated Xist chromatin

modification. Development 135, 227–235.

Payer, B., and Lee, J.T. (2008). X Chromosome Dosage Compensation: How

Mammals Keep the Balance. Annu. Rev. Genet. 42, 733–772.

Penny, G.D., Kay, G.F., Sheardown, S.A., Rastan, S., and Brockdorff, N.

(1996). Requirement for Xist in X chromosome inactivation. Nature 379,

131–137.

Plath, K., Fang, J., Mlynarczyk-Evans, S.K., Cao, R., Worringer, K.A., Wang,

H., de la Cruz, C.C., Otte, A.P., Panning, B., and Zhang, Y. (2003). Role of

histone H3 lysine 27 methylation in X inactivation. Science 300, 131–135.

Sado, T., Wang, Z., Sasaki, H., and Li, E. (2001). Regulation of imprinted

X-chromosome inactivation in mice by Tsix. Development 128, 1275–1286.

Sado, T., Hoki, Y., and Sasaki, H. (2005). Tsix silences Xist through modifica-

tion of chromatin structure. Dev. Cell 9, 159–165.

Schoeftner, S., Sengupta, A.K., Kubicek, S., Mechtler, K., Spahn, L., Koseki,

H., Jenuwein, T., and Wutz, A. (2006). Recruitment of PRC1 function at the

initiation of X inactivation independent of PRC2 and silencing. EMBO J. 25,

3110–3122.

Shevchenko, A.I., Zakharova, I.S., Elisaphenko, E.A., Kolesnikov, N.N.,

Whitehead, S., Bird, C., Ross, M., Weidman, J.R., Jirtle, R.L., Karamysheva,

T.V., et al. (2007). Genes flanking Xist in mouse and human are separated on

the X chromosome in American marsupials. Chromosome Res. 15, 127–136.

Silva, J., Mak, W., Zvetkova, I., Appanah, R., Nesterova, T.B., Webster, Z.,

Peters, A.H., Jenuwein, T., Otte, A.P., and Brockdorff, N. (2003). Establishment

of histone h3 methylation on the inactive X chromosome requires transient

recruitment of Eed-Enx1 polycomb group complexes. Dev. Cell 4, 481–495.

Simmler, M.C., Cunningham, D.B., Clerc, P., Vermat, T., Caudron, B., Cruaud,

C., Pawlak, A., Szpirer, C., Weissenbach, J., Claverie, J.M., et al. (1996). A 94

kb genomic sequence 30 to the murine Xist gene reveals an AT rich region

containing a new testis specific gene Tsx. Hum. Mol. Genet. 5, 1713–1726.

Starmer, J., and Magnuson, T. (2009). A new model for random X chromosome

inactivation. Development 136, 1–10.

Stavropoulos, N., Lu, N., and Lee, J.T. (2001). A functional role for Tsix

transcription in blocking Xist RNA accumulation but not in X-chromosome

choice. Proc. Natl. Acad. Sci. USA 98, 10232–10237.

Sun, B.K., Deaton, A.M., and Lee, J.T. (2006). A transient heterochromatic

state in Xist preempts X inactivation choice without RNA stabilization. Mol.

Cell 21, 617–628.

Tsai, C.L., Rowntree, R.K., Cohen, D.E., and Lee, J.T. (2008). Higher order

chromatin structure at the X-inactivation center via looping DNA. Dev. Biol.

319, 416–425.

Vigneau, S., Augui, S., Navarro, P., Avner, P., and Clerc, P. (2006). An essential

role for the DXPas34 tandem repeat and Tsix transcription in the counting

process of X chromosome inactivation. Proc. Natl. Acad. Sci. USA 103,

7390–7395.

Wan, L.B., and Bartolomei, M.S. (2008). Regulation of imprinting in clusters:

noncoding RNAs versus insulators. Adv. Genet. 61, 207–223.

Wutz, A., and Gribnau, J. (2007). X inactivation Xplained. Curr. Opin. Genet.

Dev. 17, 387–393.

Wutz, A., Rasmussen, T.P., and Jaenisch, R. (2002). Chromosomal silencing

and localization are mediated by different domains of Xist RNA. Nat. Genet.

30, 167–174.

Xu, N., Tsai, C.L., and Lee, J.T. (2006). Transient homologous chromosome

pairing marks the onset of X inactivation. Science 311, 1149–1152.

Yang, Y., and Seed, B. (2003). Site-specific gene targeting in mouse embry-

onic stem cells with intact bacterial artificial chromosomes. Nat. Biotechnol.

21, 447–451.

Zhao, J., Sun, B.K., Erwin, J.A., Song, J.J., and Lee, J.T. (2008). Polycomb

proteins targeted by a short repeat RNA to the mouse X chromosome. Science

322, 750–756.

Zuccotti, M., and Monk, M. (1995). Methylation of the mouse Xist gene in

sperm and eggs correlates with imprinted Xist expression and paternal X-inac-

tivation. Nat. Genet. 9, 316–320.

Cell 143, 390–403, October 29, 2010 ª2010 Elsevier Inc. 403

Page 92: Cell 101029

Insights into Egg Coat Assembly andEgg-Sperm Interaction from theX-Ray Structure of Full-Length ZP3Ling Han,1,5 Magnus Monne,1,5,6 Hiroki Okumura,1,2,5 Thomas Schwend,1,7 Amy L. Cherry,1 David Flot,3,8

Tsukasa Matsuda,4 and Luca Jovine1,*1Department of Biosciences and Nutrition and Center for Biosciences, Karolinska Institutet, Halsovagen 7, Huddinge SE-141 83, Sweden2Department of Applied Biological Chemistry, Faculty of Agriculture, Meijo University, 1-501 Shiogamaguchi, Tempaku-ku,

Nagoya 468-8502, Japan3EMBL Grenoble, 6 Rue Jules Horowitz, BP 181, 38042 Grenoble Cedex 9, France4Department of Applied Molecular Biosciences, Graduate School of Bioagricultural Sciences, Nagoya University, Furo-cho,

Chikusa-ku, Nagoya 464-8601, Japan5These authors contributed equally to this work6Present address: Department of Pharmaco-Biology, University of Bari, Via E. Orabona 4, Bari I-70125, Italy7Present address: Biomolecular Mass Spectrometry and Proteomics Group, Utrecht University, Padualaan 8, 3584 CH, Utrecht,

The Netherlands8Present address: ESRF, 6 Rue Jules Horowitz, BP 220, 38043 Grenoble Cedex 9, France

*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.09.041

SUMMARY

ZP3, a major component of the zona pellucida (ZP)matrix coating mammalian eggs, is essential forfertilization by acting as sperm receptor. By retaininga propeptide that contains a polymerization-blockingexternal hydrophobic patch (EHP), we determinedthe crystal structure of an avian homolog of ZP3 at2.0 A resolution. The structure unveils the fold of acomplete ZP domain module in a homodimericarrangement required for secretion and reveals howEHP prevents premature incorporation of ZP3 intothe ZP. This suggests mechanisms underlying poly-merization and how local structural differences, re-flected by alternative disulfide patterns, control thespecificity of ZP subunit interaction. Close relativepositioning of a conserved O-glycan important forsperm binding and the hypervariable, positivelyselected C-terminal region of ZP3 suggests a con-certed role in the regulation of species-restrictedgamete recognition. Alternative conformations ofthe area around the O-glycan indicate how spermbinding could trigger downstream events via intra-molecular signaling.

INTRODUCTION

The first fundamental step of animal fertilization is binding

between egg and sperm, whose fusion generates a zygote that

will develop into a new individual. A specialized extracellular

matrix of the egg, called zona pellucida (ZP) in mammals and

vitelline envelope (VE) in nonmammals, is crucial for this process

by directly mediating species-restricted recognition between

gametes (Wassarman and Litscher, 2008). In the mouse, the ZP

consists of glycoproteins ZP1 (100 kDa), ZP2 (120 kDa), and

ZP3 (83 kDa). These components are coordinately secreted by

growing oocytes and polymerize into mm-long filaments with

a structural repeat of 14 nm. Pairs of filaments are then cross-

linked by homodimers of the less abundant ZP1 subunit, giving

rise to the three-dimensional (3D), 6.5mm thick ZP matrix. In other

mammals, the egg coat also contains a fourth subunit (ZP4) that is

�30% identical to ZP1; moreover, proteins homologous to

mammalian ZP1–4 constitute the VE of other vertebrates, and

highly related molecules comprise the egg coat of species evolu-

tionarily very distant from mammals, like molluscs and ascidians.

The basic structure of the ZP/VE has thus been conserved over

more than 600 million years of evolution (Monne et al., 2006).

As indicated by in vitro sperm binding experiments (Bleil and

Wassarman, 1980) and exemplified by the phenotype of ZP3

null mice, which produce eggs that lack a ZP and are completely

infertile (Liu et al., 1996; Rankin et al., 1996), mouse ZP3 (mZP3)

is essential for fertilization in vivo by acting as receptor for sperm

(Wassarman and Litscher, 2008). This is supported by numerous

studies in different mammalian species, including human (Barratt

et al., 1993), as well as in other vertebrates such as chicken

(Bausek et al., 2004) and Xenopus (Vo and Hedrick, 2000).

However, the specific ZP3 determinants recognized by sperm

are highly controversial, and the molecular basis of gamete

interaction remains elusive (Gahlay et al., 2010; Wassarman

and Litscher, 2008; Shur, 2008).

The domain structure of ZP3 reflects its dual biological func-

tion. Most of the protein consists of a polymerization module of

260 residues, the so-called ZP domain (Bork and Sander,

1992), followed by a C-terminal region of 40 amino acids that

is specific to ZP3 and has been implicated in interaction with

sperm (Wassarman and Litscher, 2008). The ZP module is not

404 Cell 143, 404–415, October 29, 2010 ª2010 Elsevier Inc.

Page 93: Cell 101029

only conserved in egg coat components but is also found in

many other secreted eukaryotic proteins with variable architec-

ture and biological function (Jovine et al., 2005; Bork and Sander,

1992). It is responsible for the incorporation of ZP3 and other

subunits into the ZP (Jovine et al., 2002) and consists of two

domains, ZP-N and ZP-C, that are separated by a protease-

sensitive linker (Jovine et al., 2004). Whereas ZP-N is thought

to constitute a basic building block of ZP filaments (Monne

et al., 2008), ZP-C may mediate the specificity of interaction

between subunits (Kanai et al., 2008; Sasanami et al., 2006).

These processes are controlled by an external hydrophobic

patch (EHP) contained within the C-terminal propeptide of ZP

component precursors and an internal hydrophobic patch (IHP)

inside the ZP module (Jovine et al., 2004).

A recent crystal structure of the ZP-N domain of mZP3 had

important implications for the architecture of animal egg coats

(Monne et al., 2008). However, it could not address the function

of ZP3 as a sperm receptor, and, apart from a cryo-electron

microscopy study of glycoprotein endoglin at 25 A resolution

(Llorca et al., 2007), no structural information is available on

the complete ZP module and the regulation of its biological func-

tion. Here we present the high-resolution structure of full-length

ZP3, providing crucial insights into both the mechanism of ZP

module-mediated polymerization and the sperm binding activity

of this key reproductive protein.

RESULTS

Protein Engineering and Structure DeterminationBiogenesis of ZP3 requires processing of an N-terminal signal

peptide, formation of six intramolecular disulfide bonds, and

loss of a C-terminal propeptide that contains a polymerization-

blocking EHP and a single-spanning transmembrane domain

(TM). The latter event depends on cleavage of the protein

precursor at a consensus furin-cleavage site (CFCS) located

between the ZP-C domain and the EHP (Figure S1A and

Figure S2A available online; Wassarman and Litscher, 2008). As

a result of this complex maturation pathway, correctly folded

recombinant ZP3 can only be efficiently expressed in mammalian

cells. However, due to its heavy and heterogeneous glycosylation

(accounting for �50% of the total apparent mass of the mouse

protein), as well as its tendency to aggregate when concentrated

or enzymatically deglycosylated (Zhao et al., 2004; E. Litscher

and P. Wassarman, personal communication), full-length ZP3

has eluded attempts at structure determination for over 25 years.

To overcome this impasse, we focused on chicken ZP3

(cZP3), a naturally hypoglycosylated homolog that contains

a single N-glycosylation site and is 53% identical to human

ZP3 (Takeuchi et al., 1999; Waclawek et al., 1998). A series of

progressively modified, C-terminally histidine-tagged constructs

(Figure S1A) were expressed in Chinese hamster ovary (CHO)

cells (Figure S1B), which were previously shown to produce a re-

combinant avian ZP3 protein that is indistinguishable from its

native counterpart (Sasanami et al., 2003). Deletion of the TM

and inactivation of the N-glycosylation site and the CFCS re-

sulted in construct cZP3-3 (Figure S1A), which was secreted

from cells as a single homogeneous species of 41 kDa (Figur-

e S1B, lane 9). Because it retained the EHP at its C terminus,

this protein did not aggregate and could be purified by immobi-

lized metal affinity chromatography (IMAC), followed by size-

exclusion chromatography (SEC). The latter suggested that

cZP3 exists as a dimer (Figure S1C), in agreement with crosslink-

ing experiments (Figure S1D) and sedimentation equilibrium

studies of human ZP3 (Zhao et al., 2004).

cZP3-3 had relatively low solubility and yielded only weakly dif-

fracting crystals. However, its solubility could be significantly

improved by limited trypsinization, which resulted in loss of an

N-terminal fragment (residues Y21–R46; Figure S3) that is not

conserved among ZP3s and is missing in the mature avian protein

(Pan et al., 2000; Waclawek et al., 1998). Further mass spectro-

metric (MS) analysis of trypsinized forms of cZP3-3 and cZP3-4,

a better expressed construct carrying a deletion of P23-H52

(Figures S1A and S1B, lane 11), revealed that the improvement

in solubility was in fact due to proteolysis of a second fragment

(R348–R358) immediately preceding the inactivated CFCS

(Figure S3 and Figure S4). Trypsinized cZP3-3 and cZP3-4

(cZP3-3T/4T) produced tetragonal crystals that diffracted to high

resolution despite 71% solvent content (Figures S5A and S5B).

The structure of cZP3-4T was solved by molecular replacement

using the ZP-N domain of mZP3 (Monne et al., 2008) as search

model and refined against both a dataset at 2.0 A resolution and

an earlier 2.6 A dataset that better resolved a functionally impor-

tant O-linked carbohydrate (Table S1 and Figures S5C–S5F).

Overall Architecture of the ZP3 HomodimerIn the asymmetric unit, two molecules of ZP3 embrace each

other in antiparallel orientation to form a flat, Yin-Yang-shaped

homodimer (Figures 1A and 1B). Although part of the linker

between ZP-N and ZP-C (E158–R166) is disordered in the elec-

tron density map, the connectivity between the two domains is

unequivocally determined by their relative positions in the crystal.

In this arrangement, the two ZP modules of the dimer are held

together by interactions between ZP-N and ZP-C domains that

belong to opposite subunits. On the other hand, no ZP-N/ZP-N

or ZP-C/ZP-C contacts are observed within the dimer (Figure 1A).

Interaction with ZP-C Induces Local Rearrangementsof Two Conserved ZP-N Domain RegionsThe structure of a maltose-binding protein-mZP3 ZP-N fusion

revealed that the ZP-N domain belongs to a distinct immuno-

globulin (Ig) superfamily subtype, characterized by an E' strand

and two invariant disulfides that link the first four Cys of the ZP

module with C1-C4, C2-C3 connectivity (Monne et al., 2008).

Consistent with the fact that the model of mZP3 ZP-N was suffi-

cient to phase the structure of cZP3-4T despite representing

only 28% of the scattering mass in the asymmetric unit, the

secondary structure elements of cZP3 and mZP3 ZP-Ns can

be superimposed with a Ca root-mean-square distance of 0.9 A

(Figure S6A). However, as further discussed below, contacts

with ZP-C cause significant local differences in a conserved

region within the long FG loop of the ZP-N domain, as well as

around its invariant C2-C3 disulfide.

The ZP Module Is Internally SymmetricAs hinted by initial molecular replacement solutions that placed

additional copies of ZP-N at the position of ZP-C, the latter

Cell 143, 404–415, October 29, 2010 ª2010 Elsevier Inc. 405

Page 94: Cell 101029

domain also adopts an Ig-like fold, so that 50% of the residues

of cZP3-4T are involved in b strands (Figure 1 and Figure S2A).

ZP-N and ZP-C display no significant sequence similarity and

have different disulfide connectivity (Boja et al., 2003). Neverthe-

less, despite replacement of the C and E' strands of ZP-N by

a single C strand in ZP-C (which also contains additional A',

A", and C' strands), the b sandwiches of the two domains share

a common topology (Figure 1C). As a consequence, each ZP

module has internal symmetry (Figures S6C and S6D).

The EHP Is Coupled to an Invariant ZP3 Disulfideat the Core of the ZP-C DomainAnalysis of purified cZP3-3T and -4T suggested that the

C-terminal tryptic peptide produced by ZP3 cleavage at R358

remained noncovalently associated with the rest of the protein

(Figure S3 and Figure S4). Surprisingly, the electron density

A B

C

Figure 1. Overall Structure and Topology of

cZP3

(A) Cartoon diagram of the cZP3 homodimer struc-

ture, formed by two ZP modules each consisting of

a ZP-N and a ZP-C domain. In the upper molecule,

b sheets and disulfides are colored according to

the topology scheme in (C), except for ZP-C

strands A (IHP; orange) and G (EHP; dark cyan).

Dashed lines represent disordered loops. The

lower ZP module is colored by secondary struc-

ture, with the IHP and EHP depicted as above

and disulfides in magenta.

(B) Side view of the cZP3 homodimer with ZP-N

and ZP-C domains in gray and black, respectively.

The IHP and EHP lie at the domain interface. The

C-terminal linkage from the EHP to the TM is indi-

cated by a dark cyan dashed line.

(C) Topology scheme with secondary structure

and disulfide connectivity.

See also Figure S1, Figure S2, Figure S5,

Figure S6, and Table S1.

map reveals that the EHP sequence con-

tained in this peptide constitutes the G

strand of the ZP-C domain and is thus

an integral part of the ZP3 fold (Figure 1

and Figure 2A). Immediately next to the

EHP, a C5-C7 disulfide staples the F

strand of ZP-C to the neighboring C

strand. This linkage is conserved in all

ZP3 homologs (Kanai et al., 2008) and

forms a short right-handed hook that is

preceded by a b bulge in the C strand

and protrudes toward the center of the

ZP-C hydrophobic core (Figure 2A). To

gain insights into the functional role of

C5-C7 and other ZP3 disulfides, we indi-

vidually mutated all Cys pairs of cZP3-4

(Figure 2B). As shown in lane 6, C5-C7 is

the only disulfide whose mutation does

not completely abolish secretion of ZP3.

This result suggests that the invariant

C5-C7 pair of ZP3 is involved in other

functions besides protein folding, consistent with absence of

both of these Cys in a subset of Drosophila ZP module proteins

with a different biological function (Fernandes et al., 2010).

Cysteine Clustering in a Structurally VariableZP3-Specific SubdomainInsertions within the C'D and FG loops of ZP-C give rise to a

C-terminal ZP-C subdomain (Figures 1A and 1C) that is con-

served in the type I ZP module of ZP3 homologs but is not found

in either the type II ZP module of other ZP subunits or unrelated

Ig-like domains. The ZP-C subdomain has a remote similarity

to EGF domains based on secondary structure and consists

of a short 310 helix C"D and a three-stranded b sheet that is

connected to a longer, mixed 310/a helix F"G through C6-C11

and C8-C9 disulfides (Figure 2C and Figure S5F). This connec-

tivity was confirmed by the anomalous signal of sulfur and is

406 Cell 143, 404–415, October 29, 2010 ª2010 Elsevier Inc.

Page 95: Cell 101029

consistent with partial disulfide bond assignments of pig ZP3

(C8-C9; C6-C10/C11; C12-C11/C10; Kanai et al., 2008). On the other

hand, it differs from the disulfide pattern of fish, mouse, rat, and

human ZP3, where the same Cys residues form a C6-C8 and,

presumably, a C9-C11 bridge (Figure S2B; Kanai et al., 2008; Da-

rie et al., 2004; Boja et al., 2003). The structure reveals that, even

though these Cys are spaced in sequence, they are closely clus-

tered in space on top of a platform created by invariant W322

(Figure 2C). This 3D arrangement immediately suggests how

the alternative C6-C8, C9-C11 connectivity could be accommo-

dated in the ZP-C subdomain. At the same time, the structure

explains why cZP3 adopts the C6-C11, C8-C9 pattern, as the

helical conformation of residues R329–T337 would not be

compatible with a C9-C11 disulfide.

Consistent with the latter observation, attempts to force partial

formation of the alternative connectivity in cZP3 by mutating

either C6 and C8 or C9 and C11 resulted in nonsecreted protein

products (Figure 2D, lanes 1–4). This suggests that formation

of helix F"G is an early event in cZP3 folding that commits the

C-terminal disulfides to the C6-C11, C8-C9 connectivity.

Conversely, the same region of the protein probably adopts

a different conformation in order to form the C9-C11 disulfide

observed in other homologs of ZP3. In support of this conclu-

sion, mutations that either interfere with disulfide-mediated teth-

ering of helix F"G to the rest of the subdomain (Figure 2B, lanes

7–10; Figure 2D, lanes 5–6) or delete the residues between loop

F'F" and the CFCS (Figure 2E, top panel, lanes 3–6) are not toler-

ated by cZP3, whereas the corresponding amino acids are not

required for secretion of mZP3 constructs when the TM is

present (Figure 2E, bottom panel, lane 4).

ZP-N/ZP-C Contacts at the Homodimer InterfaceAre Essential for ZP3 BiogenesisElectrostatic complementarity between the ZP-N and ZP-C

domains of opposite ZP modules plays a major role at the inter-

face of the homodimer, which buries 2450 A2 of surface area.

The main interaction involves a positively charged protrusion

formed by the long FG loop of the ZP-N domain of one mole-

cule and a negatively charged cleft between ZP-C and the

C-terminal subdomain of the other (Figure 3A). The tip of the

ZP-N FG loop, which was loosely packed against maltose-

binding protein in the ZP-N fusion crystals (Monne et al.,

2008), forms a short F' b strand (Figure S6A) that generates

an intermolecular antiparallel b sheet with the E' strand of

ZP-C (Figure 3B). This involves a highly conserved FXF motif

and is strengthened by hydrophobic contacts between the

side chains of the F' strand and surrounding residues L204,

Y243, and cis-P241. Additionally, conserved R142 forms a

salt bridge with invariant D254 and an hydrogen bond with

Y243. Deletion of the ZP-N F' strand or mutation of the neigh-

boring R142 in ZP-C almost completely inhibits secretion (Fig-

ure 3C), indicating that dimer formation is a prerequisite for

the biogenesis of ZP3.

Intramolecular Interaction between ZP-N and ZP-CIs Hydrophobically Mediated by the EHPAs described above, the protein used for crystallization re-

tained a noncovalently bound C-terminal proteolytic fragment,

whose EHP sequence forms the G strand of the ZP-C b sand-

wich (Figure 1 and Figure 2A). Notably, this positions the EHP

not only next to the aforementioned invariant C5-C7 disulfide

A

C D

E

B Figure 2. ZP-C Disulfide Connectivity

(A) 2Fobs-Fcalc map of the region around invariant

disulfide C5-C7 and the EHP G strand, contoured

at 1 s. Dashed lines indicate hydrogen bonds.

(B) All disulfide-forming Cys pairs were individually

substituted by Ala. The constructs were expressed

and cell lysate (L) and conditioned medium (M;

concentrated 10 times, unless otherwise indi-

cated) were analyzed by immunoblot.

(C) C-terminal subdomain disulfide arrangement,

showing the close proximity of C6, C8, C9, and

C11. Black mesh is a 3.7 A resolution phased

anomalous difference map, calculated using

diffraction data collected at 7.75 keV and con-

toured at 4 s.

(D) Cys mutations preventing the native disulfide

connectivity of the ZP-C subdomain abolish pro-

tein secretion. Medium was concentrated 5 times.

(E) Removal of C-terminal residues W322–R358 in-

hibited secretion of cZP3 whether the TM was

present (DSCS) or not (DC-term). In corresponding

mZP3 mutants lacking S309-K346 (Jovine et al.,

2002), the TM rescued protein secretion.

See also Figure S2, Figure S3, and Figure S4.

Cell 143, 404–415, October 29, 2010 ª2010 Elsevier Inc. 407

Page 96: Cell 101029

staple (Figure 2A) but also close to the E'-F-G extension of

ZP-N, as well as the IHP sequence that constitutes the A strand

of ZP-C (Figure 1, Figure 4A, and Figure S2A). Both of these

elements have been implicated in ZP module-mediated poly-

merization (Schaeffer et al., 2009; Monne et al., 2008; Jovine

et al., 2004).

Analysis of the 635/610 A2 interface between the adjacent

ZP-N and ZP-C domains of the ZP module reveals a central

role of hydrophobic interactions around absolutely conserved

P376 in the EHP. This residue, which is in cis conformation

and forms a b bulge together with invariant G375, is flanked

by L290 and forms a stack of rings with highly conserved

ZP-C amino acids Y292 and P235 (Figure 4A). The resulting

surface interacts with V114, L127, V147, and P149 on the

outside of the E'-F-G sheet of ZP-N, as well as with P87. This

stretches the CD loop of ZP-N, causing the C2-C3 disulfide to

adopt an unusual left-handed conformation and, in turn, to pull

the underlying EE' region, which does not form the a helix

observed in isolated mZP3 ZP-N (Figure S6B). Moreover, a

Q116-E196 hydrogen bond and an R125-E196 ionic interaction

are observed at one end of the interface, whereas variable

contacts involving R288 are found at the other (Figure 4A).

However, analysis of mutants shows that the hydrophobic

contacts play a much more important role than these other kinds

of interactions. In agreement with complementary mutational

studies of invariant EHP residues (Schaeffer et al., 2009; Jovine

et al., 2004), mutation of Y292 and P235 severely inhibits ZP3

secretion (Figure 4B, lanes 1–6), whereas an E196A mutant is

secreted at levels comparable to the wild-type protein (Fig-

ure 4B, lane 8).

ZP3 Cleavage Causes Slow Spontaneous Dissociationof the EHP at Physiological TemperatureApart from being involved in the ring stack and hydrogen-

bonding to the neighboring F and A" b strands, the EHP makes

many other interactions with the ZP-C domain. These include

hydrophobic contacts with residues of the A, B, and F strands

as well as F199 and conserved P171, F172, and F202 and a

salt bridge between D371 and H296 on the F strand. Consistent

with this array of interactions, our biochemical analysis of

trypsinized cZP3 shows that the EHP is tightly bound to the

core of the protein and is not removed by SEC or IMAC, even

upon extensive washing. This raises the issue of whether the

EHP can dissociate spontaneously, or if this is dependent on

interaction between cZP3 and other ZP subunits. To answer

this question, we incubated cZP3-4T at 39�C (the body temper-

ature of the chicken) for 30 hr. As shown in Figure 4C (lanes 1–3),

this resulted in loss of approximately 40% of the EHP from the

sample, a reasonable proportion considering that avian VE

assembly requires several weeks. SEC analysis (Figure 4D)

revealed that—like mature native cZP3 (Bausek et al., 2004)—

much of the remaining protein had formed different oligomeric

states and large-molecular-weight species (gray profile) in

comparison with an identical sample incubated at 4�C (violet

profile), or with uncut protein incubated for the same time at

39�C (red profile). Consistent with the fact that this experiment

was performed in the absence of other egg coat subunits, elec-

tron microscopy indicated that the material in the void volume

peak of Figure 4D consisted of amorphous aggregates rather

than polymers (data not shown).

An Evolutionarily Conserved O-Glycan Plays a MajorRole in Sperm BindingIn the structure of cZP3-4T, one molecule in the homodimer has

visible density for part of the ZP-N/ZP-C linker region, which can

be modeled from residue P167 onward (Figure 1). Additional

electron density was found next to T168, which belongs to

A

B

C

Figure 3. The Homodimer Interface

(A) Complementary electrostatic surface potential of the ZP-N FG loop and the

cleft between ZP-C and the C-terminal subdomain.

(B) An intermolecular antiparallel b sheet is formed by the ZP-N F' strand of one

monomer and ZP-C E' strand of the other.

(C) Whereas mutation of cis-P241 does not affect secretion, R142A and

mutants lacking the ZP-N F' strand (D139–141 and D139–142) disrupt dimer

formation and are essentially not secreted.

See also Figures S1C and S1D.

408 Cell 143, 404–415, October 29, 2010 ª2010 Elsevier Inc.

Page 97: Cell 101029

a highly conserved PTWXPF ZP3 motif (Figure S2A). MS analysis

identified a E158–R181 peptide containing a 365 Da Hex-

HexNAc modification (Figure S7A) that, based on carbohydrate

composition analysis of cZP3 (Takeuchi et al., 1999) and lectin-

binding experiments (Figure S7B, lanes 2 and 5), was interpreted

as Galb1-3GalNAc (T antigen). This disaccharide could be fitted

into the electron density map of the 2.6 A structure (Figure 5A),

whereas density for the second carbohydrate residue was not

as well defined in the 2.0 A crystal.

Considering the evolutionary conservation of this site, which

has been denominated ‘‘site 1’’ and is also modified with core

1-related glycans in native mZP3, native rat ZP3, and human

ZP3 expressed in transgenic mice or CHO-Lec3.2.8.1 cells (Cha-

labi et al., 2006; Boja et al., 2005; Zhao et al., 2004; Boja et al.,

2003), T168 was mutated to Ala in order to assess the carbohy-

drate function. The mutant protein was expressed and secreted

as efficiently as the wild-type, excluding a role for the T168

O-glycan in ZP3 biogenesis (Figure 5B). This is consistent with

the observation that the Thr is substituted by other amino acids

in a subset of ZP3 sequences from fish, where the protein has an

equivalent structural, but not receptorial, role. As expected from

the lack of a single O-linked sugar chain, a small change was

observed in the migration of the mutant protein (Figure 5B),

which no longer bound to either jacalin or peanut agglutinin

(Figures S7B and S7C). This hinted at the lack of additional

O-glycans, which was confirmed by both inspection of electron

density maps and extensive MS analysis of both cZP3-4 and

native cZP3. The fact that this protein carries a single O-linked

carbohydrate at T168 allowed us to conclusively evaluate the

role of this particular sugar chain in sperm binding, in the

absence of possible compensatory effects from other glycans.

Quantification of protein binding to the tip of chicken sperm

head (Figure 5C) showed that the T168A mutation caused a

decrease of �80% in binding relative to wild-type cZP3-4 (Fig-

ure 5D), indicating an important role of the conserved O-glycan

in avian gamete interaction.

DISCUSSION

Thirty years after ZP3 was identified (Bleil and Wassarman,

1980), this work yields structural information on an egg protein

region directly recognized by sperm at the beginning of fertiliza-

tion. Combined with mutational and in vitro binding studies,

the structure provides insights into many aspects of ZP3 biology,

ranging from secretion and polymerization to interaction with

sperm. Moreover, it has important implications for human

reproductive medicine.

Evolution of the ZP and Role of the ZP Module DimerInterfaceOur previous crystal structure of the ZP-N domain of mZP3

(Monne et al., 2008) strongly supported the suggestion that

additional copies of ZP-N are found within N-terminal exten-

sions of ZP1, ZP2, and ZP4 (Callebaut et al., 2007). By showing

A D

B

C

Figure 4. The Domain Interface and EHP

(A) Interface between ZP-N and ZP-C domains of

the same monomer, with the ZP-C G strand

(EHP) in the center and ZP-C A strand (IHP) in

the background. Note the close position of Y124,

an invariant residue in the E'-F-G extension of

ZP-N that was suggested to be important for poly-

merization (Fernandes et al., 2010; Monne et al.,

2008; Legan et al., 2005). Pink mesh is an aver-

aged kick omit map of the EHP contoured at 1 s.

The set of interactions involving N129, D131, and

R288 is observed in chain A of the 2.0 A resolution

structure.

(B) Mutation of Y292 and P235, which stack with

P376 in the EHP, severely inhibits secretion,

whereas mutation of E196 has no effect. Medium

was concentrated 5 times.

(C and D) Analysis of EHP dissociation. Purified

cZP3-4/4T proteins were incubated either at

4�C or at 39�C for 30 hr and molecules with

and without EHP/6His-tag were separated by

IMAC. SDS-PAGE of cZP3-4T samples incubated

at 39�C (C) shows the EHP/6His-tag peptide in

the IMAC-bound sample (lane 2, red arrow).

Whereas 40% of cZP3-4T incubated at 39�Cwas found in the flow-through (FT; compare

lane 3 to lane 1), cZP3-4T incubated at 4�C and

cZP3-4 incubated at 39�C remained bound to

the column (data not shown). Lanes 4–7, analysis

of fractions from the SEC peaks numbered in (D).

(D) SEC analyses of eluted cZP3-4T incubated at

4�C and cZP3-4 incubated at 39�C are shown in violet and red, respectively (left-hand scale), and that of FT of cZP3-4T incubated at 39�C is shown in gray

with four distinct peaks corresponding to different oligomeric forms (right-hand scale).

See also Figure S1C, Figure S3, Figure S4, and Figure S6B.

Cell 143, 404–415, October 29, 2010 ª2010 Elsevier Inc. 409

Page 98: Cell 101029

that the ZP-C domain adopts an Ig-like fold with the same

topology as ZP-N (Figures 1A and 1C), the X-ray map of full-

length ZP3 reveals that the ZP module contains internal

symmetry (Figures S6C and S6D). Considering that the ZP-C

domain has so far been found only within the context of

a complete ZP module, this observation raises the possibility

that ZP-N and ZP-C—and thus essentially the whole mammalian

egg coat—originated by duplication of a common ancestral

Ig-like domain. Moreover, conservation of ZP-N and ZP-C resi-

dues that mediate formation of the antiparallel ZP module

homodimer (Figure 1A and Figure 3B), which is essential for

ZP3 secretion (Figure 3C), suggests that this quaternary struc-

ture is also important for the function of other ZP subunits and

unrelated ZP module proteins. In agreement with this conclu-

sion, a C582-C582 interchain disulfide that characterizes human

endoglin (Llorca et al., 2007) can be readily modeled on the basis

of the ZP module arrangement observed in the ZP3 crystal. The

biological importance of the dimer interface is further highlighted

by a recent study of fish embryo hatching, identifying R167 of

medaka ZP3 as a target cleavage site of hatching enzymes

(Yasumasu et al., 2010). Because this residue corresponds

to cZP3 R142 (Figure S2A), which plays an essential role at

the interface (Figure 3B and Figure 3C, lane 2), the structure

immediately suggests how hatching enzymes could solubilize

egg coat filaments by disrupting the stability of ZP module

dimers. Considering that a mammalian homolog of fish hatching

enzymes is expressed in unfertilized oocytes and preimplanta-

tion embryos (Quesada et al., 2004), conservation of the RjTcleavage site in human ZP3 (Figure S2A) might indicate that a

similar mechanism is involved in human embryo hatching and

implantation.

A D B

C

Figure 5. T168 Carries an O-Glycan Important for

Sperm Binding

(A) Averaged kick omit map (0.8 s; green mesh) and

composite omit map (0.9s; red mesh) of the Galb1-3GalNAc

chain attached to T168.

(B) cZP3-4 T168A mutant protein shows a migration shift

relative to the wild-type during SDS-PAGE.

(C) Chicken sperm were incubated with cZP3-4 and its

mutant T168A and bound protein were detected by

immunofluorescence (green). Corner inserts show

TOTO3-stained (red) sperm.

(D) Statistically highly significant difference in the sperm-

binding activity of cZP3-4 and T168A. Data are represented

as mean ± standard error of the mean (SEM).

See also Figure S7.

Mechanism of Protein PolymerizationInhibition by the EHP and Implicationsfor ZP AssemblyPrevious mutational studies suggested that

cleavage of the membrane-bound precursors

of ZP module proteins around the CFCS

releases a block to polymerization by causing

dissociation of the EHP (Schaeffer et al., 2009;

Jovine et al., 2004). However, how the EHP

inhibits polymer assembly at the molecular level,

and what is its relationship with other elements

involved in polymerization such as the IHP (Schaeffer et al.,

2009; Jovine et al., 2004) and the ZP-N E'-F-G extension (Fer-

nandes et al., 2010; Monne et al., 2008; Legan et al., 2005),

was unknown.

The structure of ZP3 reveals that, rather than simply shielding

a surface-exposed polymerization interface, the EHP penetrates

through the core of the molecule by constituting b strand G of

the ZP-C domain (Figure 1). This strand directly faces the IHP

(ZP-C b strand A) and makes contacts with the E'-F-G face

of ZP-N (Figure 4A). Although stable within the context of the

uncleaved protein precursor, the resulting ZP-N/ZP-C interface

is dominated by hydrophobic contacts involving the EHP. This

suggests that the two domains must undergo significant rear-

rangements upon cleavage of ZP3 at the CFCS and dissociation

of the C-terminal propeptide. Thus, the EHP blocks premature

protein polymerization by acting as a ‘‘molecular glue’’ that

keeps the ZP module in a conformation that is essential for

secretion (Figure 4B) but not compatible with formation of

higher-order structures.

In agreement with studies on soluble fish egg coat protein

precursors secreted by the liver (Sugiyama et al., 1999), our

in vitro analysis of EHP ejection shows that, even in constructs

lacking a TM, the propeptide containing the EHP must be phys-

ically cleaved before the latter is released from the protein

(Figures 4C and 4D). This implies that, regardless of the presence

of C-terminal membrane-anchoring elements, the patch can only

be ejected from the side of the homodimer opposite to where the

CFCS lies; this is where the C-terminal ends of the two ZP3

subunits come almost in contact with each other (Figure 1B).

Coupling of this structural constraint, probably deriving from

the sharp kink made by the invariant GP sequence of the EHP

410 Cell 143, 404–415, October 29, 2010 ª2010 Elsevier Inc.

Page 99: Cell 101029

(Figure 4A), with membrane anchoring may play an important

role in ZP assembly by orienting the ZP3 precursor so that it

can properly interact with other subunits upon cleavage at the

CFCS (Figure 1B). This would explain why, although the TM is

not required for secretion, it is essential for incorporation of

ZP3 into the mouse ZP (Jovine et al., 2002).

Following cleavage, dissociation of the EHP must cause

exposure of a large hydrophobic region on ZP-C (Figure 4A), trig-

gering interaction with its cognate ZP-N or another ZP module.

This might depend on strand- or domain-swapping events

involving the exposed IHP and the E'-F-G extension of ZP-N,

which in the structure does not interact with other parts of the

homodimer (Figure 1A and Figure 4A). Moreover, because of

the direct structural relationship between the EHP and the F

strand of ZP-C (Figure 2A), rearrangements connected with ZP

assembly may also involve the C5-C7 disulfide staple, which is

conserved in ZP3 homologs from fish to human despite not

being essential for secretion (Figure 2B). Notably, a similar disul-

fide has been found in CD4 and implicated in domain swapping,

CD4 dimerization, and entry of HIV-1 into CD4+ cells (Sane-

jouand, 2004).

A Structural Basis for the Specificity of Egg CoatSubunit InteractionEven though several ZP module-containing proteins can

homopolymerize, formation of egg coat filaments requires ZP3

(a type I subunit) and at least one type II (ZP1/ZP2/ZP4-like)

component (Jovine et al., 2005; Boja et al., 2003). Furthermore,

in spite of very high sequence identity, only certain combinations

of heterologous ZP subunits can productively interact to form a

ZP (Hasegawa et al., 2006). How is the specificity of ZP assembly

regulated at the molecular level?

Our crystallographic and mutational analysis indicates that,

although clustering of conserved Cys within the ZP-C subdo-

main of ZP3 (Figure 2C) can account for the two disulfide

connectivities observed in different ZP3 homologs (Figure S2B),

these alternative patterns must be accommodated by local

differences in the surrounding structural elements (Figures 2C–

2E). Because the ZP-C domain mediates interaction between

type I and type II ZP subunits (Okumura et al., 2007; Sasanami

et al., 2006), and because different ZP3 disulfide connectivities

are reflected by changes in the disulfide patterns of cognate

type II proteins (Kanai et al., 2008), this suggests that the tertiary

structure of the ZP-C subdomain of ZP3 determines the speci-

ficity of egg coat assembly. Considering that pig and mouse

ZP3 adopt different disulfide patterns (Kanai et al., 2008), this

conclusion explains why pig ZP2 does not incorporate into the

mouse ZP when secreted by transgenic animals (Hasegawa

et al., 2006).

Sperm Binding and Modulation of the Specificityof Gamete InteractionCarbohydrates of ZP3 have been repeatedly implicated in

binding to sperm, but there is highly conflicting evidence about

the chemical nature and location of the bioactive glycans, as

well as about their functional importance relative to the polypep-

tide moiety of the protein (Wassarman and Litscher, 2008; Shur,

2008). Nevertheless, many studies from different laboratories

support the idea that initial species-restricted binding between

mammalian gametes is mediated by ZP3 O-glycans (Florman

and Wassarman, 1985) and involves a C-terminal region of the

molecule that, in the mouse, is encoded by exon 7 of the Zp3

gene (Figure S2A; Wassarman and Litscher, 2008; Kinloch

et al., 1995). This region varies between species as a result of

positive Darwinian selection (Swann et al., 2007; Turner and

Hoekstra, 2006; Jansa et al., 2003; Swanson et al., 2001) and,

based on mZP3 mutants expressed in embryonal carcinoma

cells, was suggested to contain a sperm-combining site (SCS;

Figure S2A) carrying active O-glycans at S332 and S334 (Chen

et al., 1998). This hypothesis was challenged by MS analysis of

purified ZP material, which indicated that the same sites are

not glycosylated in native mZP3 (Boja et al., 2003). A suggestion

was thus made that the functional O-glycans of the native protein

are instead located at site 1 and/or a downstream Ser/Thr-rich

region called ‘‘site 2’’ (Figure S2A; Chalabi et al., 2006). More

recently, the biological importance of S332 and S334 in vivo

was excluded based on the fertility of ZP3�/� mice expressing

a ZP3 transgene where these residues are mutated, although

alternative binding sites were not identified (Gahlay et al.,

2010). How can these results be reconciled with the strong

evidence for a role of O-linked carbohydrates in binding to sperm

(Florman and Wassarman, 1985)?

The data presented in Figure 5 provide direct evidence in favor

of the importance of ZP3 site 1 O-glycans in gamete interaction.

At the same time, they allow evaluation of the relationship

between the various ZP3 sites that have been implicated in

sperm binding, by projecting them on top of the structure of

cZP3. As shown in Figures 1A and 1B, the interdomain loop

carrying T168 folds back onto itself, positioning site 2 next to

site 1 on top of ZP-C (Figure 6A). On the other side of the b sand-

wich, disulfide C10-C12 in the ZP-C subdomain, which partly

overlaps with the exon 7/SCS region (Figure S2A), fastens the

C-terminal region of mature ZP3 to helix F"G (Figure 2C) so

that it bends toward the interdomain loop (Figures 1A and 1B).

The resulting �120� inversion in chain direction is necessary

for inserting the EHP at the core of the ZP module, explaining

why the C10-C12 connectivity is invariant between ZP3 homologs

(Figure S2B). At the same time, this has the effect of positioning

the C-terminal half of the SCS on the same surface of the mole-

cule as sites 1 and 2 (Figure 6A). Although this region and the

CFCS that follows it are disordered in the electron density, the

approximate positions of mZP3 S332 and S334 can be easily

inferred because these residues would immediately follow

P343, the last visible SCS residue in the cZP3 map. By revealing

that site 1, site 2, and the SCS are all exposed within a restricted

area on the same surface of ZP3, our structure suggests that any

of them could in principle contribute to carbohydrate-mediated

sperm binding, as long as it is modified with the correct type of

sugar chain in either native ZP3 (sites 1 and/or 2) or recombinant

ZP3 produced in embryonal carcinoma cells (SCS). As shown in

Figures 6B and 6C, spatial clustering of the sites also immedi-

ately suggests how—regardless of glycosylation—the hypervari-

able SCS and very C-terminal part of mature ZP3 could affect the

specificity of gamete interaction by modulating the recognition of

sites 1 and 2. At the same time, the conformational flexibility of

the C-terminal region of ZP3, which could be amplified by the

Cell 143, 404–415, October 29, 2010 ª2010 Elsevier Inc. 411

Page 100: Cell 101029

aforementioned species-dependent variations in the local

structure of the ZP-C subdomain, could clearly provide opportu-

nities for protein-based recognition. This is consistent with the

observation that sperm binding is highly reduced but not com-

pletely abolished in the T168A mutant (Figure 5D) and agrees

with the growing evidence that gamete interaction probably

relies on multiple distinct binding events (Shur, 2008). With rela-

tion to this point, it is interesting to notice how the conserved

glycosylation sites of ZP3 are located within a linker region

whose flexibility is probably important for ZP-N/ZP-C domain

rearrangements during polymerization, and the orientation of

the hypervariable C-terminal region results from the position of

the EHP within the protein precursor. These structural consider-

ations suggest how the sperm recognition function of ZP3 might

have arisen during evolution as a specialization of its polymeriza-

tion activity.

Regardless of which exact ZP3 epitope(s) are recognized by

sperm in the mouse, the data by Gahlay et al. (2010) suggest

that lack of sperm binding to the murine ZP following fertilization

is not the result of ZP3 carbohydrate cleavage or modification

but rather depends on proteolytic processing of ZP2. These

results are still compatible with an important role of ZP3 carbo-

hydrates in sperm binding, as ZP2 cleavage could act indirectly

by causing structural rearrangements that ultimately shield the

ZP3-binding surface identified by our structural and functional

studies.

Downstream Transmission of Sperm BindingInformationOrdering of the O-glycosylated interdomain linker region, which

is not involved in crystal contacts with symmetry-related

molecules, has remarkable effects on underlying ZP-C domain

residues (Figure 7A). In the ZP3 monomer where T168 is ordered,

the conserved neighboring residue W169 stacks against the side

chain of E180 and forms a short b sheet by inducing the forma-

tion of a B' b strand within the ZP-C BC loop. Consequently, an

invariant residue in this loop, H219, flips inwards making

hydrogen bonds with main chain carbonyl oxygens of S215 in

strand B and V220 in strand B'. The presence of two different

conformations within the crystal allows us to hypothesize how

information about sperm binding might be transmitted through

ZP3. It is possible that in the unbound protein the linker region

around T168 is highly flexible. However, upon sperm binding

this zone assumes a more ordered conformation that is stable

(Figure 7B) and transmits a signal through the molecule as a

result of H219 flipping. This may lead to stimulation of the

acrosome reaction, a process that depends on the polypeptide

moiety of ZP3 (Wassarman and Litscher, 2008; Shur, 2008).

Alternatively, the conformational switch could be part of the

structural changes of the ZP that take place during the block

to polyspermy, and regulate the accessibility of the O-glycan

before and after fertilization.

Relevance for Human Reproductive MedicineAntibodies against ZP proteins, and in particular the C-terminal

region of ZP3, have been shown to be powerful tools for inhibit-

ing fertilization of domestic animals and wildlife, including

primates (Kaul et al., 2001; Millar et al., 1989). However, variable

efficiency and safety concerns suggest that immunocontracep-

tion is unlikely to become a feasible option for humans. At the

same time, no completely novel contraceptive method has

been introduced in the last 50 years to address the continuous

growth of the world population (McLaughlin and Aitken, 2010).

By allowing the development of small-molecule compounds

that specifically target the sperm binding surface shown in

Figure 6A, the structure of ZP3 could pave the way to the rational

design of nonhormonal contraceptives. Moreover, structural

information on the molecule will be essential for understanding

ZP mutations linked to human infertility at the molecular level.

site 1

conservedO-glycosylation

domain

hZP3 S331mZP3 S332

A

cZP3 T168hZP3 T156mZP3 T155

cZP3 S346hZP3 S333mZP3 S334

ZP-C1

site 2hZP3 T163mZP3 T162

cZP3 S177mZP3 S164

hZP3 S166mZP3 S165

cZP3 P343

GalNac

Gal

cZP3 S174hZP3 T162

sperm-combining

site

B

C

site 1

site 1

T 168

P167

N163

R166

V220

N173

P301 G305 S328 N320

E336

N333F360

Q356

M352

E353

S357R354

L349R348

R347 L345

R329

N333

E336

T337

N3 39

L345R347

L349

M352

E353

S357

variable conserved

positively selected

correlated change

C

P308

N350

V323

P342P343

site 2

site 2

Figure 6. Conserved O-Glycosylation Sites are Clustered on the

Same Protein Surface as Hypervariable, Positively Selected Regions

of ZP3

(A) Conserved O-glycosylation sites 1 and 2 and the SCS are exposed on the

same surface of ZP3.

(B) Top view of the ZP-C domain, colored according to amino acid conserva-

tion of ZP3 homologs from amphibian to human. Approximately 70% of the

most variable residues in ZP-C are located in the depicted area. The figure

includes a model of disordered C-terminal residues L345–F360 (black outline),

which were added to the crystal structure and relaxed by molecular dynamics.

Statistically significant variable residues, as well as invariant P167 and T168,

are marked. cZP3 sites 1 and 2 are indicated, with the conserved site 1

O-glycan shown in stick representation.

(C) Mapping of positively selected sites (red) onto the model of mature cleaved

ZP3, oriented as in (B). Two sites showing correlated changes are colored

in violet.

412 Cell 143, 404–415, October 29, 2010 ª2010 Elsevier Inc.

Page 101: Cell 101029

EXPERIMENTAL PROCEDURES

Protein Expression and Purification

Protocols used for DNA construct generation, protein expression in CHO cells,

and protein purification are outlined in the Extended Experimental Procedures.

Protein and Carbohydrate Analysis

Methods used for immunoblot analysis, oligomeric state determination, cross-

linking in solution, mass spectrometry, and lectin binding are described in the

Extended Experimental Procedures.

Crystallization and Data Collection

Crystals of cZP3-4T (25 mg/ml) were grown in 0.1 M Na citrate (pH 5.0), 10 mM

Tris-HCl (pH 8.0), 3%–13% PEG 6000, 50 mM NaCl (Figure S5A). They

appeared in 1–5 days at 4�C and were cryoprotected by stepwise addition

of PEG 6000 and PEG MME 550 to a final solution of 0.1 M Na citrate

(pH 5.0), 10 mM Tris-HCl (pH 8.0), 6% PEG 6000, 30% PEG MME 550,

50 mM NaCl, after which they were flash frozen in liquid nitrogen. Datasets

were collected at the European Synchrotron Radiation Facility (ESRF), Greno-

ble (Table S1). Details of structure determination and refinement, as well as

structure analysis and molecular dynamics simulation, are provided in the

Extended Experimental Procedures.

Sperm Binding Assays

Semen collected from 15 White Leghorn cocks was frozen in liquid nitrogen as

described (Japanese patent No. 2942822). Sperm (1.5 3 104/ml) were

incubated with protein (5 ng/ml = 134 nM) in 20 mM Na-HEPES (pH 7.4),

150 mM NaCl at 37�C for 15 min. They were then fixed onto a glass slide

with 3% paraformaldehyde, blocked with 2% BSA, and incubated with anti-

5His (QIAGEN; 1:1,000), followed by Alexa Fluor-488 goat anti-mouse IgG

(Invitrogen, 1:300). Imaging was performed on an Axioplan2 microscope

equipped with LSM5 PASCAL laser scanning confocal optics (Zeiss) in

multitrack mode. 488 nm excitation and 505–530 nm band-pass emission

filters were used for imaging Alexa-Fluor 488. Stacks of 7–11 images taken

at 0.5 mm intervals along the Z axis were merged, and signal intensities of

the tip region of sperm heads were measured. Differential interference contrast

images were taken by the same system. Analysis was performed with ImageJ

(http://imagej.nih.gov/ij/), using a negative control-based integrated density

cutoff of 10,000. t test statistical analysis was performed with InStat (Graph-

Pad Software, Inc.). Animal procedures were approved by the Nagoya Univer-

sity Institutional Animal Care and Use Committee.

ACCESSION NUMBERS

Atomic coordinates and structure factors are deposited in the Protein

Data Bank with accession codes 3NK3 (2.6 A resolution) and 3NK4 (2.0 A

resolution).

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures,

seven figures, and one table and can be found with this article online at

doi:10.1016/j.cell.2010.09.041.

ACKNOWLEDGMENTS

We thank CMC Biologics for expression plasmids pDEF38 and pNEF38; the

ESRF for provision of synchrotron radiation facilities and Joanne McCarthy

for assistance; Pavel Afonine, Ralf Grosse-Kunstleve, and Tom Terwilliger

for help with PHENIX; Elmar Krieger and Alessandra Villa for help with molec-

ular dynamics simulations; Hans Hebert for electron microscopy; Hisako

Watanabe for help with sperm preparation; Franco Cotelli, Eveline Litscher,

Rune Toftgard, and Paul Wassarman for discussions and comments. This

work was supported by the Center for Biosciences; the Swedish Research

Council (grants 2005-5102 and 2007-6068); the European Community (Marie

Curie ERG 31055); the Scandinavia-Japan Sasakawa Foundation; Grant-in-

aids from the Japan Society for the Promotion of Science and MEXT; and an

EMBO Young Investigator award to L.J. Author Contributions: L.H. expressed

proteins and analyzed mutants; M.M. generated constructs, purified and crys-

tallized proteins, carried out model building, and refined the structures; H.O.

generated and characterized constructs, expressed proteins, and performed

and analyzed sperm binding assays; T.S. performed mass spectrometric anal-

ysis; A.L.C. analyzed crystallographic data; D.F. assisted data collection at

ESRF; T.M. performed and analyzed sperm binding assays; L.J. directed the

research, solved the structure, took part in structure refinement, ran molecular

dynamics simulations, and wrote the manuscript with contributions from all

other authors. L.J. dedicates this work to Marta, Smilla, and Sofia.

Received: June 23, 2010

Revised: August 11, 2010

Accepted: August 24, 2010

Published online: October 21, 2010

REFERENCES

Barratt, C.L.R., Andrews, P.A., McCann, C.T., Hornby, D.P., and Cooke, I.D.

(1993). Recombinant human ZP3 expressed in Chinese hamster ovary cells

(CHO) is a potent inducer of the acrosome reaction. Hum. Reprod. (8 Suppl.),

Abstr. no. 407.

Bausek, N., Ruckenbauer, H.H., Pfeifer, S., Schneider, W.J., and Wohlrab, F.

(2004). Interaction of sperm with purified native chicken ZP1 and ZPC proteins.

Biol. Reprod. 71, 684–690.

Bleil, J.D., and Wassarman, P.M. (1980). Mammalian sperm-egg interaction:

Identification of a glycoprotein in mouse egg zonae pellucidae possessing

receptor activity for sperm. Cell 20, 873–882.

H219E180

A B C

Monomer 2

H219

W169

E180

GalNAcT168

C

B'

A B

G'

Monomer 1

Gal

A

V220

S215

L176

N218

B W169 β secondary structure

N218 β secondary structure

V220 β secondary structure

H219 - S215 hydrogen bonding

H219 - V220 hydrogen bonding

E180 - L176 hydrogen bonding

0 1 2 3 4 5 6 7 8 9 10simulation time (ns)

Figure 7. Alternative Conformations of the Conserved O-Linked Site

Region

(A) The interdomain loop containing O-glycosylated T168 is disordered in

monomer 2 (left) but adopts an ordered structure in monomer 1 by interacting

with the BC loop of ZP-C (right).

(B) Key elements of the ordered loop conformation are stable during the course

of independent 10 ns molecular dynamics simulations.

Cell 143, 404–415, October 29, 2010 ª2010 Elsevier Inc. 413

Page 102: Cell 101029

Boja, E.S., Hoodbhoy, T., Fales, H.M., and Dean, J. (2003). Structural charac-

terization of native mouse zona pellucida proteins using mass spectrometry.

J. Biol. Chem. 278, 34189–34202.

Boja, E.S., Hoodbhoy, T., Garfield, M., and Fales, H.M. (2005). Structural

conservation of mouse and rat zona pellucida glycoproteins. Probing the

native rat zona pellucida proteome by mass spectrometry. Biochemistry 44,

16445–16460.

Bork, P., and Sander, C. (1992). A large domain common to sperm receptors

(Zp2 and Zp3) and TGF-b type III receptor. FEBS Lett. 300, 237–240.

Callebaut, I., Mornon, J.P., and Monget, P. (2007). Isolated ZP-N domains

constitute the N-terminal extensions of Zona Pellucida proteins. Bioinfor-

matics 23, 1871–1874.

Chalabi, S., Panico, M., Sutton-Smith, M., Haslam, S.M., Patankar, M.S.,

Lattanzio, F.A., Morris, H.R., Clark, G.F., and Dell, A. (2006). Differential

O-glycosylation of a conserved domain expressed in murine and human

ZP3. Biochemistry 45, 637–647.

Chen, J., Litscher, E.S., and Wassarman, P.M. (1998). Inactivation of the

mouse sperm receptor, mZP3, by site-directed mutagenesis of individual

serine residues located at the combining site for sperm. Proc. Natl. Acad.

Sci. USA 95, 6193–6197.

Darie, C.C., Biniossek, M.L., Jovine, L., Litscher, E.S., and Wassarman, P.M.

(2004). Structural characterization of fish egg vitelline envelope proteins by

mass spectrometry. Biochemistry 43, 7459–7478.

Fernandes, I., Chanut-Delalande, H., Ferrer, P., Latapie, Y., Waltzer, L.,

Affolter, M., Payre, F., and Plaza, S. (2010). Zona pellucida domain proteins

remodel the apical compartment for localized cell shape changes. Dev. Cell

18, 64–76.

Florman, H.M., and Wassarman, P.M. (1985). O-linked oligosaccharides of

mouse egg ZP3 account for its sperm receptor activity. Cell 41, 313–324.

Gahlay, G., Gauthier, L., Baibakov, B., Epifano, O., and Dean, J. (2010).

Gamete recognition in mice depends on the cleavage status of an egg’s

zona pellucida protein. Science 329, 216–219.

Hasegawa, A., Kanazawa, N., Sawai, H., Komori, S., and Koyama, K. (2006).

Pig zona pellucida 2 (pZP2) protein does not participate in zona pellucida

formation in transgenic mice. Reproduction 132, 455–464.

Jansa, S.A., Lundrigan, B.L., and Tucker, P.K. (2003). Tests for positive

selection on immune and reproductive genes in closely related species of

the murine genus Mus. J. Mol. Evol. 56, 294–307.

Jovine, L., Qi, H., Williams, Z., Litscher, E., and Wassarman, P.M. (2002).

The ZP domain is a conserved module for polymerization of extracellular

proteins. Nat. Cell Biol. 4, 457–461.

Jovine, L., Qi, H., Williams, Z., Litscher, E.S., and Wassarman, P.M. (2004).

A duplicated motif controls assembly of zona pellucida domain proteins.

Proc. Natl. Acad. Sci. USA 101, 5922–5927.

Jovine, L., Darie, C.C., Litscher, E.S., and Wassarman, P.M. (2005). Zona

pellucida domain proteins. Annu. Rev. Biochem. 74, 83–114.

Kanai, S., Kitayama, T., Yonezawa, N., Sawano, Y., Tanokura, M., and Nakano,

M. (2008). Disulfide linkage patterns of pig zona pellucida glycoproteins ZP3

and ZP4. Mol. Reprod. Dev. 75, 847–856.

Kaul, R., Sivapurapu, N., Afzalpurkar, A., Srikanth, V., Govind, C.K., and

Gupta, S.K. (2001). Immunocontraceptive potential of recombinant bonnet

monkey (Macaca radiata) zona pellucida glycoprotein-C expressed in

Escherichia coli and its corresponding synthetic peptide. Reprod. Biomed.

Online 2, 33–39.

Kinloch, R.A., Sakai, Y., and Wassarman, P.M. (1995). Mapping the mouse

ZP3 combining site for sperm by exon swapping and site-directed mutagen-

esis. Proc. Natl. Acad. Sci. USA 92, 263–267.

Legan, P.K., Lukashkina, V.A., Goodyear, R.J., Lukashkin, A.N., Verhoeven,

K., Van Camp, G., Russell, I.J., and Richardson, G.P. (2005). A deafness

mutation isolates a second role for the tectorial membrane in hearing. Nat.

Neurosci. 8, 1035–1042.

Liu, C., Litscher, E.S., Mortillo, S., Sakai, Y., Kinloch, R.A., Stewart, C.L., and

Wassarman, P.M. (1996). Targeted disruption of the mZP3 gene results in

production of eggs lacking a zona pellucida and infertility in female mice.

Proc. Natl. Acad. Sci. USA 93, 5431–5436.

Llorca, O., Trujillo, A., Blanco, F.J., and Bernabeu, C. (2007). Structural model

of human endoglin, a transmembrane receptor responsible for hereditary

hemorrhagic telangiectasia. J. Mol. Biol. 365, 694–705.

McLaughlin, E.A., and Aitken, R.J. (2010). Is there a role for immunocontracep-

tion? Mol. Cell. Endocrinol. Published online April 20, 2010. 10.1016/j.mce.

2010.04.004.

Millar, S.E., Chamow, S.M., Baur, A.W., Oliver, C., Robey, F., and Dean, J.

(1989). Vaccination with a synthetic zona pellucida peptide produces long-

term contraception in female mice. Science 246, 935–938.

Monne, M., Han, L., and Jovine, L. (2006). Tracking down the ZP domain: From

the mammalian zona pellucida to the molluscan vitelline envelope. Semin.

Reprod. Med. 24, 204–216.

Monne, M., Han, L., Schwend, T., Burendahl, S., and Jovine, L. (2008). Crystal

structure of the ZP-N domain of ZP3 reveals the core fold of animal egg coats.

Nature 456, 653–657.

Okumura, H., Aoki, N., Sato, C., Nadano, D., and Matsuda, T. (2007).

Heterocomplex formation and cell-surface accumulation of hen’s serum

zona pellucida B1 (ZPB1) with ZPC expressed by a mammalian cell line

(COS-7): a possible initiating step of egg-envelope matrix construction. Biol.

Reprod. 76, 9–18.

Pan, J., Sasanami, T., Nakajima, S., Kido, S., Doi, Y., and Mori, M. (2000).

Characterization of progressive changes in ZPC of the vitelline membrane of

quail oocyte following oviductal transport. Mol. Reprod. Dev. 55, 175–181.

Quesada, V., Sanchez, L.M., Alvarez, J., and Lopez-Otin, C. (2004). Identifica-

tion and characterization of human and mouse ovastacin: a novel metallopro-

teinase similar to hatching enzymes from arthropods, birds, amphibians, and

fish. J. Biol. Chem. 279, 26627–26634.

Rankin, T., Familari, M., Lee, E., Ginsberg, A., Dwyer, N., Blanchette-Mackie,

J., Drago, J., Westphal, H., and Dean, J. (1996). Mice homozygous for an

insertional mutation in the Zp3 gene lack a zona pellucida and are infertile.

Development 122, 2903–2910.

Sanejouand, Y.H. (2004). Domain swapping of CD4 upon dimerization.

Proteins 57, 205–212.

Sasanami, T., Ohtsuki, M., Ishiguro, T., Matsushima, K., Hiyama, G., Kansaku,

N., Doi, Y., and Mori, M. (2006). Zona Pellucida Domain of ZPB1 controls

specific binding of ZPB1 and ZPC in Japanese quail (Coturnix japonica). Cells

Tissues Organs 183, 41–52.

Sasanami, T., Toriyama, M., and Mori, M. (2003). Carboxy-terminal proteolytic

processing at a consensus furin cleavage site is a prerequisite event for quail

ZPC secretion. Biol. Reprod. 68, 1613–1619.

Schaeffer, C., Santambrogio, S., Perucca, S., Casari, G., and Rampoldi, L.

(2009). Analysis of uromodulin polymerization provides new insights into the

mechanisms regulating ZP domain-mediated protein assembly. Mol. Biol.

Cell 20, 589–599.

Shur, B.D. (2008). Reassessing the role of protein-carbohydrate complemen-

tarity during sperm-egg interactions in the mouse. Int. J. Dev. Biol. 52,

703–715.

Sugiyama, H., Murata, K., Iuchi, I., Nomura, K., and Yamagami, K. (1999).

Formation of mature egg envelope subunit proteins from their precursors

(choriogenins) in the fish, Oryzias latipes: loss of partial C-terminal sequences

of the choriogenins. J. Biochem. 125, 469–475.

Swann, C.A., Cooper, S.J., and Breed, W.G. (2007). Molecular evolution of the

carboxy terminal region of the zona pellucida 3 glycoprotein in murine rodents.

Reproduction 133, 697–708.

Swanson, W.J., Yang, Z., Wolfner, M.F., and Aquadro, C.F. (2001). Positive

Darwinian selection drives the evolution of several female reproductive

proteins in mammals. Proc. Natl. Acad. Sci. USA 98, 2509–2514.

Takeuchi, Y., Nishimura, K., Aoki, N., Adachi, T., Sato, C., Kitajima, K., and

Matsuda, T. (1999). A 42-kDa glycoprotein from chicken egg-envelope, an

avian homolog of the ZPC family glycoproteins in mammalian zona pellucida.

414 Cell 143, 404–415, October 29, 2010 ª2010 Elsevier Inc.

Page 103: Cell 101029

Its first identification, cDNA cloning and granulosa cell-specific expression.

Eur. J. Biochem. 260, 736–742.

Turner, L.M., and Hoekstra, H.E. (2006). Adaptive evolution of fertilization

proteins within a genus: variation in ZP2 and ZP3 in deer mice (Peromyscus).

Mol. Biol. Evol. 23, 1656–1669.

Vo, L.H., and Hedrick, J.L. (2000). Independent and hetero-oligomeric-

dependent sperm binding to egg envelope glycoprotein ZPC in Xenopus

laevis. Biol. Reprod. 62, 766–774.

Waclawek, M., Foisner, R., Nimpf, J., and Schneider, W.J. (1998). The chicken

homologue of zona pellucida protein-3 is synthesized by granulosa cells. Biol.

Reprod. 59, 1230–1239.

Wassarman, P.M., and Litscher, E.S. (2008). Mammalian fertilization: the egg’s

multifunctional zona pellucida. Int. J. Dev. Biol. 52, 665–676.

Yasumasu, S., Kawaguchi, M., Ouchi, S., Sano, K., Murata, K., Sugiyama, H.,

Akama, T., and Iuchi, I. (2010). Mechanism of egg envelope digestion by

hatching enzymes, HCE and LCE in medaka, Oryzias latipes. J. Biochem.

148, 439–448.

Zhao, M., Boja, E.S., Hoodbhoy, T., Nawrocki, J., Kaufman, J.B., Kresge, N.,

Ghirlando, R., Shiloach, J., Pannell, L., Levine, R.L., et al. (2004). Mass

spectrometry analysis of recombinant human ZP3 expressed in glycosyla-

tion-deficient CHO cells. Biochemistry 43, 12090–12104.

Cell 143, 404–415, October 29, 2010 ª2010 Elsevier Inc. 415

Page 104: Cell 101029

Microbial Stimulation FullyDifferentiatesMonocytestoDC-SIGN/CD209+

Dendritic Cells for Immune T Cell AreasCheolho Cheong,1,5,* Ines Matos,1,5 Jae-Hoon Choi,1 Durga Bhavani Dandamudi,1 Elina Shrestha,1 M. Paula Longhi,1

Kate L. Jeffrey,2 Robert M. Anthony,3 Courtney Kluger,1 Godwin Nchinda,1 Hyein Koh,1 Anthony Rodriguez,1

Juliana Idoyaga,1 Maggi Pack,1 Klara Velinzon,4 Chae Gyu Park,1,* and Ralph M. Steinman1,*1Laboratory of Cellular Physiology and Immunology and Chris Browne Center for Immunology and Immune Diseases2Laboratory of Lymphocyte Signaling3Laboratory of Molecular Genetics and Immunology4Laboratory of Molecular Immunology, Howard Hughes Medical Institute

The Rockefeller University, 1230 York Avenue, New York, NY 10065, USA5These authors contributed equally to this work

*Correspondence: [email protected] (C.C.), [email protected] (C.G.P.), [email protected] (R.M.S.)

DOI 10.1016/j.cell.2010.09.039

SUMMARY

Dendritic cells (DCs), critical antigen-presenting cellsfor immune control, normally derive from bonemarrow precursors distinct from monocytes. It isnot yet established if the large reservoir of mono-cytes can develop into cells with critical features ofDCs in vivo. We now show that fully differentiatedmonocyte-derived DCs (Mo-DCs) develop in miceand DC-SIGN/CD209a marks the cells. Mo-DCs arerecruited from blood monocytes into lymph nodesby lipopolysaccharide and live or dead gram-nega-tive bacteria. Mobilization requires TLR4 and itsCD14 coreceptor and Trif. When tested for antigen-presenting function, Mo-DCs are as active as clas-sical DCs, including cross-presentation of proteinsand live gram-negative bacteria on MHC I in vivo.Fully differentiated Mo-DCs acquire DC morphologyand localize to T cell areas via L-selectin and CCR7.Thus the blood monocyte reservoir becomes thedominant presenting cell in response to selectmicrobes, yielding DC-SIGN+ cells with critical func-tions of DCs.

INTRODUCTION

Recent advances have clarified the origin of dendritic cells (DCs),

a hematopoietic lineage specialized to present antigens and

both initiate and control immunity (Heath and Carbone, 2009;

Melief, 2008). In the bone marrow, a common monocyte-DC

precursor (Fogg et al., 2006) gives rise to monocytes and other

precursors termed common DC precursors (Naik et al., 2007;

Onai et al., 2007) and pre-cDCs (Liu et al., 2009). The latter

express intermediate levels of CD11c integrin and begin to

synthesize MHC II products. Pre-cDCs move into the blood to

seed both lymphoid and nonlymphoid tissues forming CD11chi,

MHC IIhi DCs (Liu et al., 2009; Ginhoux et al., 2009). DCs in the

steady state are dependent upon the hematopoietin, Flt3-L

(D’Amico and Wu, 2003), whereas monocytes require macro-

phage colony-stimulating factor (M-CSF) (Geissmann et al.,

2010). Flt3-L�/� mice have a severe deficit of DCs (Naik et al.,

2007; Onai et al., 2007; Liu et al., 2009; Waskow et al., 2008),

whereas monocytes are missing in mice lacking M-CSF receptor

(c-fms or CD115) (Heard et al., 1987; Ginhoux et al., 2006). Thus,

most DCs in the steady state are independent of monocytes.

Nevertheless, monocytes also can differentiate into DCs.

Although first studied as macrophage precursors, mainly in vitro

(de Villiers et al., 1994; Johnson et al., 1977), monocytes were

later recognized to have an added potential to develop into

DCs (monocyte-derived DCs [Mo-DCs]). This too has been

studied primarily in cultures of human blood monocytes (Romani

et al., 1994; Sallusto and Lanzavecchia, 1994). Monocytes, upon

culture for several days in GM-CSF and IL-4, acquire a typical

probing or dendritic morphology, lose the capacity to phagocy-

tose, and adhere to various tissue culture surfaces but acquire

strong capacities to initiate immunity. Mo-DCs can immunize

humans (Dhodapkar et al., 1999; Schuler-Thurner et al., 2000)

and home to the T cell areas of lymph nodes (LNs) (De Vries

et al., 2003). Monocytes are �20 times more abundant than

DCs in blood and marrow, so the mobilization of this monocyte

reservoir in vivo to generate potent antigen-presenting DCs

needs to be elucidated.

Several reports have begun to document in mice the differen-

tiation of CD11c� and MHC II� blood monocytes into large

numbers of CD11c+ MHC II+ Mo-DCs during different models,

e.g., Leishmania major infection via the skin (Leon et al., 2007),

intravenous infection with Listeria monocytogenes (Serbina

et al., 2003), influenza virus infection via the airway (Nakano

et al., 2009), Aspergillus fumigatus in the lung (Hohl et al.,

2009), T cell-mediated colitis (Siddiqui et al., 2010), and injection

416 Cell 143, 416–429, October 29, 2010 ª2010 Elsevier Inc.

Page 105: Cell 101029

of the adjuvant, alum (Kool et al., 2008). These Mo-DCs pre-

sented protein antigens to TCR transgenic CD4+ T cells and

are distinguished from classical DCs by expression of the Gr-

1/Ly6C monocyte markers. However, many classical functional

features of DCs have not been assessed, including a peculiar

probing morphology, localization to T cell areas of lymphoid

organs in a position to find and activate rare clones of specific

T cells, and efficient antigen capture and processing.

The latter includes the capacity for cross-presentation. This is

the processing of captured proteins onto MHC I without the need

for synthesis in antigen-presenting cells (APCs) (Heath and Car-

bone, 2001). Through cross-presentation to CD8+ T cells, DCs

present nonreplicating antigens, e.g., from dying cells (Liu

et al., 2002; Luckashenak et al., 2008), noninfectious microbes

(Moron et al., 2003), and immune complexes (Regnault et al.,

1999). The CD8+ subset of classical DCs are specialized for

cross-presentation (den Haan et al., 2000; Schnorrer et al.,

2006; Dudziak et al., 2007; Sancho et al., 2009), but Mo-DCs

have not been assessed in vivo.

To address these gaps, markers are required to identify

Mo-DCs. Here we describe a unique approach using recently

isolated monoclonal anti-DC-SIGN/CD209a antibodies (Cheong

et al., 2010). We had previously defined in mice the DC-SIGN or

CD209a gene syntenic with human DC-SIGN/CD209 (Park et al.,

2001). DC-SIGN is a hallmark of human Mo-DCs in culture (Geij-

tenbeek et al., 2000b) but is not detected on the rich network of

presumably monocyte-independent DCs in human LNs in the

steady state (Granelli-Piperno et al., 2005). We now find that

anti-mouse DC-SIGN/CD209a monoclonal antibodies (mAbs)

distinguish Mo-DCs from classical DCs in cell suspensions

and tissue sections. We will report that the full differentiation of

monocytes to DC-SIGN/CD209a+ Mo-DCs does occur in vivo

and can be initiated by lipopolysaccharide (LPS) or LPS-ex-

pressing bacteria. In contrast to prior reports on inflammatory

monocytes, these Mo-DCs rapidly lose expression of monocyte

markers Gr-1/Ly6C and CD115/c-fms, markedly upregulate

expression of TLR4 and CD14, acquire the probing morphology

of DCs, localize to the T cell areas, and through Trif signal-

ing become powerful antigen-capturing and -presenting cells,

including cross-presentation of gram-negative bacteria.

RESULTS

DC-SIGN/CD209a Marks Mouse Mo-DCs with StrongAntigen-Presenting ActivityTo determine if new mAbs to mouse DC-SIGN/CD209a can

identify Mo-DCs, as occurs with cultured human Mo-DCs (Geij-

tenbeek et al., 2000b), we cultured bone marrow monocytes

(SSClo cells with high Ly6C and CD11b; Figure S1A available

online; Naik et al., 2006) with two cytokines, GM-CSF and IL-4,

as described for blood monocytes (Schreurs et al., 1999). After

4–7 days, we recovered �80% of the plated cells. Most had con-

verted to large nonadherent cells that extended and retracted

sheet-like processes in several directions from the cell body

(Figure 1A, left), which is the hallmark, probing morphology of

DCs (Steinman and Cohn, 1973; Lindquist et al., 2004). A poly-

clonal Ab to mouse DC-SIGN detected low levels of the

30 kDa protein in fresh monocytes, but within 2 days of culture,

DC-SIGN and MHC II were upregulated markedly (Figure 1A,

right), particularly with IL-4 and GM-CSF in combination,

whereas no DC-SIGN was expressed by marrow granulocytes

similarly cultured (Figure S1B).

To establish differentiation to DCs, we confirmed that fresh

marrow and blood monocytes did not react with mAbs to DC-

SIGN, MHC II, or CD11c (Figure S1B), but when cultured in

GM-CSF and IL-4, strong reactivity developed (Figure 1B, top).

The combination of GM-CSF and IL-4, but not single cytokines

or other hematopoietins like Flt3-L and M-CSF, allowed

monocytes to express MHC II and CD11c and develop a DC

morphology. When we compared marrow monocytes before

and after culture in GM-CSF and IL-4 (Figure 1C, left, days

0 and 4) to spleen monocytes and classical DCs (Figure 1C, right

panels), we found that Mo-DCs like spleen DCs lacked M-CSF

receptor or CD115, a key receptor for monocyte development,

whereas both marrow and splenic monocytes expressed

CD115 (Figure 1C). Splenic but not Mo-DCs expressed Flt3 or

CD135 (Figure 1C), the receptor for Flt3-L, a major hematopoie-

tin for DCs derived from nonmonocytic precursors.

During differentiation, Mo-DCs also lost the Gr-1 and Ly6C

markers of monocytes and reduced their levels of F4/80 but

retained high expression of CD11b and CD172a found on

both monocytes and DEC-205� CD8� monocyte-independent,

spleen DCs (Figure 1C). Monocytes and Mo-DCs lacked CD8aa,

expressed by the DEC-205+ CD8+ subset of splenic DCs, but

Mo-DCs expressed high levels of CD24, like DEC-205+ CD8+

splenic DCs (not shown). The data in Figures 1B and 1C indicate

that monocytes acquire many surface features of splenic DCs

except that Mo-DCs express DC-SIGN and lack Flt3 or CD135.

To test if DC-SIGN+ Mo-DCs shared functions with splenic

DCs, we used the mixed leukocyte reaction (MLR), an example

of the immune-initiating function of DCs (Steinman and Witmer,

1978). In these and all T cell studies, we used CSFE-labeled

T cells and monitored the expansion of dividing or CFSElo cells,

as in Figures S1C and S1D. Mo-DCs induced with GM-CSF and

IL-4 stimulated a strong MLR, whereas monocytes cultured

under other conditions were weak (GM-CSF) or inactive (IL-4,

M-CSF, Flt3-L) (Figure S1C).

To evaluate presentation of protein antigens, we used TCR

transgenic T cells as responders and compared Mo-DCs to

two subsets of classical splenic DCs (DEC-205+ and DEC-

205�, corresponding to CD8+ and CD8� DCs). We used 40mg/ml,

a limiting concentration malarial circumsporozoite protein (CSP,

expressed in bacteria), and Ovalbumin (OVA). The Mo-DCs were

superior APCs when using graded doses of each type of DC

(Figure 1D, green).

To compare Mo-DCs with classical DCs that had also been

derived from marrow cultures, we used a Flt3-L culture system

as described by Naik et al. (2005) (Figure S1E). Over a range of

protein concentrations and cell doses, Mo-DCs were superior

cross-presenting cells relative to Flt3-L expanded, CD8+, and

CD8� DC equivalents (Figure S1F). The Mo-DCs also were supe-

rior to CD8+ DCs when irradiated, stably expressing OVA-CHO

cells were used as the antigen (Figure 1E). Thus in vitro derived

Mo-DCs are marked by DC-SIGN and are functionally strong

APCs, including cross-presentation.

Cell 143, 416–429, October 29, 2010 ª2010 Elsevier Inc. 417

Page 106: Cell 101029

TLR4 Agonists Rapidly Recruit DC-SIGN+ Cellsto the T Cell Area of Lymph NodesTo find out if comparable Mo-DCs develop in vivo in response

to microbial stimuli, we treated mice intravenously (i.v.) with

agonists for individual Toll-like receptors (TLRs) and looked

for DC-SIGN/CD209a+ cells in LNs 12–24 hr later. We also

assessed mannose receptor/CD206 because both CD206

(Sallusto et al., 1995) and DC-SIGN/CD209 (Geijtenbeek et al.,

2000b) are induced when cultured human monocytes become

Mo-DCs. Using LPS, we observed a 10-fold increase in

A B

C

D E

Figure 1. DCs Derived from Marrow Monocytes Express DC-SIGN and Are Potent APCs

(A) Marrow monocytes (Figure S1) were cultured in GM-CSF and IL-4 for 4–7 days. (Left) DIC image with typical dendritic morphology. (Right) Western blot with

rabbit polyclonal aDC-SIGN and mAb KL295 aMHC II.

(B) As in (A), showing MHC II, CD11c, and DC-SIGN Alexa 647-MMD3 (or isotype control, middle panel) on Mo-DCs.

(C) Surface markers on freshly isolated monocytes, GM-CSF/IL-4-induced Mo-DCs, and fresh spleen populations.

(D) Presentation of CSP or OVA, 40 mg/ml, to TCR transgenic T cells by graded doses of Mo-DCs or CD11chi DEC-205+ and DEC-205� DCs from spleen. Gating

strategy for CFSElo T cells is in Figure S1D.

(E) Presentation of stably transduced, irradiated CHO-OVA cells by graded doses of different populations of DCs cultured from bone marrow (DC:T cell ratio on

the x axis), including the equivalents of CD8+ and CD8� classical DCs from Flt3-L expanded marrow cultures (Figure S1E). Representative of 2–3 experiments in

triplicate or quadruplicate cultures.

Error bars = standard deviation (SD) (D and E).

418 Cell 143, 416–429, October 29, 2010 ª2010 Elsevier Inc.

Page 107: Cell 101029

DC-SIGN/CD209a+ CD206+ cells in skin-draining nodes 12–

24 hr later (Figure 2A) but not in spleen or mesenteric nodes

(not shown). Expansion took place in C3H/HeN but not C3H/

HeJ TLR4 mutant mice, indicating a need for TLR4 (Figure 2A,

compare top and bottom right). However, DC-SIGN/CD209a+

cells did not expand to other TLR agonists like Pam3CSK4,

poly(I:C), Flagellin, R848, and CpG, for TLR2, 3, 5, 7/8, and 9,

respectively (Figure 2B).

To determine whether Mo-DCs localize to T cell areas like

authentic DCs, we used new anti-DC-SIGN mAbs (Cheong

et al., 2010) to label lymph node sections. In PBS mice, there

were relatively few DC-SIGN+ cells, mainly in interfollicular

regions, beneath SIGN-R1/CD209b+ subcapsular macrophages

and between B220+ B cell follicles (Figure 2C, left and Fig-

ure S2A). However, 12 hr after LPS i.v., DC-SIGN+ cells were

abundant and localized to T cell areas, regions in which DCs

have been shown to present antigens to recirculating antigen-

specific T cells (Stoll et al., 2002; Mempel et al., 2004; Miller

et al., 2004; Shakhar et al., 2005) (Figure 2C). Likewise, DC-

SIGN+ cells accumulated in the T cell areas when we injected

LPS-bearing, heat-killed E. coli i.v. and subcutaneously (s.c.)

A B

C

D

E

Figure 2. Mobilization of DC-SIGN+ Mo-DCs

to the T Cell Areas of Lymph Nodes

(A) TLR4-competent (C3H/HeN) or TLR4 mutant

(C3H/HeJ) mice were injected with 5 mg of LPS

i.v. After 24 hr, lymph node cells were stained

intracellularly with Alexa 647 MMD3 a-DC-SIGN

and Alexa 488 a-MMR/CD206 mAbs.

(B) Mice were injected i.v. with 10 mg of a-DC-

SIGN-Alexa 647 mAb and 5 mg of TLR agonist.

(C) Labeling of frozen sections with the indicated

mAb 12–24 hr after PBS, 5 mg LPS i.v., or 5 3

106 heat-killed E. coli or B. subtilis i.v. Alexa 647

B220 mAb marks B cell areas (blue). 1003 magni-

fication.

(D and E) Lymph node sections from PBS- or LPS-

treated mice were stained with the indicated mAb.

4003 magnification.

but not LPS-lacking B. subtilis by these

routes (Figure 2C) or Listeria monocyto-

genes s.c. (not shown).

To determine if DC-SIGN+ cells were

distinct from DCs and macrophages

in the lymph node, we double-labeled

for DC-SIGN and several markers. In

PBS-injected mice, the few DC-SIGN+

cells were distinct from macrophages

in subcapsular and medullary regions

of lymph node, which in steady state

express CD206 (Figure 2D) and SIGN-

R1/CD209b (Figure S2). However, in

LPS-injected mice, there was a major

expansion of cells in the T cell area

expressing both CD206 and DC-SIGN/

CD209a (Figure 2D and Figure S2). The

DC-SIGN+ cells mobilized to the T cell

areas by LPS were clearly distinct from

other DCs, which expressed higher levels of CD11c, as well as

DEC-205/CD205 and Langerin/CD207 (Figure 2E and Fig-

ure S2). Also, DC-SIGN+ cells did not colabel with markers that

are abundant on lymph node macrophages, such as SIGN-R1/

CD209b and CD169 (Figure 2E, right panels and Figure S2)

and F4/80 (not shown). DC-SIGN/CD209a+ Mo-DCs also lacked

CD115 and Ly6C found on monocytes and inflammatory mono-

cytes (Geissmann et al., 2003) (not shown). Therefore, DC-SIGN

marks abundant cells in the T cell areas from LPS-treated mice,

which express molecules distinct from classical DCs, macro-

phages, and monocytes.

Mo-DCs Can Be Selectively Labeled with InjectedAnti-DC-SIGN/CD209a Antibody and Isolatedfrom Classical DCs in Lymph NodesTo compare the properties of LPS-mobilized DC-SIGN+ cells

to other DCs in LNs, we needed a strategy to separate the cell

types. However, the problem we faced was that most DC-

SIGN is inside the cell and not on the cell surface, preventing

the separation of cell-surface-labeled DC-SIGN+ cells. To

Cell 143, 416–429, October 29, 2010 ª2010 Elsevier Inc. 419

Page 108: Cell 101029

A

B

C D

E F

Figure 3. DC-SIGN+ Mo-DCs Are Induced upon Treatment with LPS or LPS-Bacteria

(A) Thirty micrograms Alexa 488 MMD3 a-DC-SIGN or control mAb were injected i.v. with LPS into WT or DC-SIGN�/� mice. Twelve hours later, lymph node

sections were fixed and stained with rabbit a-Alexa 488 to visualize the injected mAb in green. a-MMR/CD206 (red) identifies Mo-DCs, and A647 B220 mAb

(blue) B cells. 4003 magnification.

(B) Separation of three lymph node DC populations 12 hr after injecting i.v. 10 mg of Alexa 647 MMD3 a-DC-SIGN mAb plus 5 mg of LPS. Skin-draining lymph node

cells were stained for lymphocyte lineage markers (CD3, CD19, NK1-1 [or DX-5]), CD11c, and DEC-205. Live, lineage� CD11c+ cells were gated and three pop-

ulations defined (Pop#1, #2, #3). Isotypes for DC-SIGN and DEC-205 are mouse IgG2c and rat IgG2a, respectively.

420 Cell 143, 416–429, October 29, 2010 ª2010 Elsevier Inc.

Page 109: Cell 101029

overcome this obstacle, during injection of LPS (or PBS con-

trols), we also included 10 mg of Alexa dye-labeled MMD3 anti-

DC-SIGN mAb, or isotype-matched control mAb, to allow the

DC-SIGN+ cells to take up the fluorescent mAb. When we exam-

ined sections of the injected LNs (Figure 3A), we found that the

injected anti-DC-SIGN mAb labeled abundant dendritic profiles

in the T cell area, but only if the mice had received LPS. No

such profiles were seen if we injected isotype control mAb, or

if we injected Alexa 488-labeled MMD3 into DC-SIGN�/� mice

(Figure 3A), which did mobilize numerous macrophage man-

nose receptor (MMR)/CD206+ cells in response to LPS (Fig-

ure 3A and Figure S2E). The anti-DC-SIGN mAb-targeted cells

did not express detectable CD115, but this M-CSF receptor

strongly marked lymph node medullary macrophages, and had

low levels of CD11c but no DEC-205, which were expressed

by classical DCs in the lymph node (not shown).

Therefore to isolate Mo-DCs, we injected LPS together

with labeled MMD3 mAb (or isotype control mAb) and made

cell suspensions. To identify DCs, we gated on lymphocyte

lineage-negative, CD11c+ cells, and we surface labeled for

DEC-205 on cross-presenting classical DCs. In LNs from mice

injected with LPS plus-labeled MMD3 mAb, there was a specifi-

cally stained DC-SIGN+ population, as there was no staining if

isotype control mAb was injected (Figure 3B), or if we studied

DC-SIGN�/� mice (Figure S3A). Labeling with MMD3 was com-

parable in wild-type (WT) and Fc receptor g�/� mice, further indi-

cating that labeling required DC-SIGN and was not Fc mediated

(Figure S3B). The CD11c+ lymphocyte-negative cells also had

DEC-205+ and DEC-205� populations, both lacking DC-SIGN.

Thus LNs from LPS-treated mice have three populations:

population #1 corresponds to DC-SIGN/CD209a+ DCs, which

we will show derive from monocytes, whereas populations #2

and #3 correspond to DEC-205+ (including CD8+, Figure S3A)

and DEC-205� resident DCs (Vremec and Shortman, 1997)

(Figure 3B and Figures S3A and S3B).

When tested for surface markers following cell sorting, all three

populations of DCs from LPS-treated LNs expressed high levels

of MHC II, which is expected of DCs, and all expressed CD40, 80,

and 86 with the DC-SIGN+ and DEC-205+ subsets having the high-

est levels (Figure 3C). However, the DC-SIGN+ population had

lower levels of CD11c (not shown). We also verified that the sorted

DC-SIGN+ cells had the probing morphology of DCs (Figure 3D and

Movie S1 for video). All three DC populations likewise failed to stain

for CD115/c-fms, but DC-SIGN+ cells lacked CD135/Flt3, which

was expressed by lymph node resident DCs (Figure 3E). Like

DEC-205� classical DCs, DC-SIGN+ DCs were CD11b+ and

CD172a/SIRPahi, F4/80+, CD24lo, and CD8� (Figure 3E).

To test LPS-bearing bacteria, we injected the labeled MMD3

mAb together with either dead or live E. coli and, 12 hr later,

stained cells from draining LNs. Either dead or live E. coli, but

not dead or live B. subtilis that lacked LPS, mobilized DC-

SIGN+ cells and upregulated CD86 on splenic DCs if injected

i.v. (Figure 3F and Figure S3C). These data indicate that cells

with the morphology and markers of Mo-DCs accumulate

in vivo in response to LPS and LPS+ bacteria, and they resemble

CD8� DEC-205� resident DCs except for selective DC-SIGN/

CD209a and MMR/CD206 expression, two uptake receptors

abundant on human Mo-DCs ex vivo (Sallusto et al., 1995; Gra-

nelli-Piperno et al., 2005).

DC-SIGN+ MMR+ Mo-DCs in LPS-Stimulated LymphNodes Derive from MonocytesTo determine whether LPS mobilized DC-SIGN+ cells from mono-

cytes, we injected 2 3 106 marrow monocytes from CD45.2+ mice

i.v. into CD45.1+ hosts. Next day, the mice were injected i.v. with

labeled MMD3 mAb and 5 mg of LPS. Twenty-four hours later,

skin-drainingLNs were tested byflow cytometry forMo-DCrecruit-

ment. In three experiments, with three mice each, LPS induced

an increase in CD45.2+ donor-derived, DC-SIGN/CD209a+ and

MMR/CD206+ cells in all mice, whereas donor-derived cells were

absent in nodes of PBS-injected mice (Figure 4A).

To establish the monocyte origin of LPS-recruited Mo-DCs by

an alternative method, we focused on LysMcre 3 iDTR mice, in

which treatment with diphtheria toxin (DT) depletes monocytes

and macrophages (Goren et al., 2009). We confirmed that a

single dose of DT i.v. decreased >80% of blood monocytes

12 hr later (Figure 4B). DT-treated, LPS-injected WT mice gener-

ated CD11c+ DC-SIGN+ cells normally (Figure 4B, right, arrow),

but DT-treated, LPS-injected LysMcre 3 iDTR mice failed to

generate Mo-DCs, although the classical monocyte-indepen-

dent DC subsets were normally represented (Figure 4B, right).

Likewise in tissue sections, DC-SIGN+ DCs were not recruited

into the T cell areas of LNs of LPS-treated LysMcre 3 iDTR

mice upon DT treatment, but DEC-205+ DCs were abundant in

LPS- and DT-treated WT and LysMcre 3 iDTR mice (Figure 4C,

green versus red), again showing that Mo-DCs derived from

monocytes, whereas classical DCs did not.

To test whether the spleen was needed, a recently recognized

source of monocytes (Swirski et al., 2009), we studied sple-

nectomized mice. However after LPS injection, these mice

normally mobilized DC-SIGN/CD209a+ MMR/CD206+ Mo-DCs

(Figure S4A).

To selectively deplete classical DCs, we employed Flt3�/�

mice, which lack classical DCs because of a need for Flt3

signaling. We confirmed a loss of classical DCs in Flt3�/� mice

(Waskow et al., 2008), but in contrast, LPS comparably mobi-

lized Mo-DCs from Flt3�/� and WT mice using either DC-SIGN/

CD209a or MMR/CD206 as markers (Figure 4D and Fig-

ure S4B). To determine whether cell proliferation was involved,

we labeled mice with BrdU during the 12 hr treatment with

LPS, but no labeling was evident in contrast to the basal level

of BrdU labeling of classical DCs (Figure S4C). These results pro-

vide considerable evidence for the monocyte origin of DC-SIGN+

DCs in LNs from LPS-treated mice.

(C) Expression of maturation markers on three DC populations.

(D) Representative morphology (DIC images) of DC-SIGN+ cells sorted from LNs of LPS-treated mice as in (B). 6003 magnification.

(E) Three DC populations as in (B) were sorted and stained with PE-mAbs.

(F) As in (B), but fluorescence-activated cell sorting (FACS) analyses and total numbers of lineage� CD11c+ cells from mice 12 hr after i.v. injection of MMD3 a-DC-

SIGN mAb plus killed E. coli or B. subtilis (data with live organisms are in Figure S3C).

Cell 143, 416–429, October 29, 2010 ª2010 Elsevier Inc. 421

Page 110: Cell 101029

L-Selectin and CCR7 Are Required for LPS to GenerateMo-DCsTo begin to identify mechanisms of Mo-DC mobilization, we

evaluated the lymph node homing molecule used by lympho-

cytes, L-selectin/CD62L, which is also expressed on monocytes

prior to their becoming Mo-DCs (e.g., in Figure 1C). We treated

mice with isotype control or anti-CD62L (MEL-14) mAb and

1 hr later injected LPS. Anti-CD62L blocked Mo-DC formation

in LNs using immunolabeling of sections and cell suspensions

(Figures 5A and 5B).

To identify the required chemokine receptors, we tested

four chemokine receptor knockout mice. Accumulation of

DC-SIGN+ Mo-DCs was critically dependent on CCR7 (Fig-

ure 5C, right). Only a partial but statistically significant decrease

in Mo-DCs was noted in CCR2�/� mice (Figure S5A), whereas

CCR5 and CCR6 were not necessary (Figure 5C). Monocytes

disappeared normally from the blood in LPS-treated CCR7�/�

mice, and CCR7 was not required to generate Mo-DCs in vitro

(Figures S5B and S5C). In all these experiments, we verified

that spleen DCs in the knockout mice responded normally to

LPS by upregulating CD86 (Figure 5C, right). To establish that

the need for CCR7 was cell intrinsic, we made mixed bone

marrow chimeras with 50:50 mixes of WT and CCR7�/� donor

cells, each marked with CD45.1 and CD45.2 and injected into

CD45.1+ WT hosts. Six weeks later, we certified chimerism in

the blood (Figure 5D, left) and injected LPS to recruit DC-

SIGN/CD209a+ MMR/CD206+ DCs. LPS greatly reduced the

number of monocytes in the blood (Figure 5D, middle), but only

CD45.1+ WT cells and not CD45.2+ CCR7�/� cells formed

Mo-DCs (Figure 5D, right), indicating that the need for CCR7

by Mo-DCs is cell intrinsic.

Mo-DCs Efficiently Present Proteins and BacteriaCaptured In Vivo to T CellsTo test the antigen-presenting functions of Mo-DCs, we initially

sorted three populations of CD11chi DCs from inflamed LNs using

CD11c, DEC-205, and DC-SIGN as markers as in Figure 3B. All

three DC types from LPS-treated mice effectively stimulated allo-

geneic T cells in the MLR assay, with Mo-DCs being moderately

more active (Figure 6A, left and Figure S6). Surprisingly, Mo-DCs

were comparable or superior to classical DCs in presenting two

different proteins (OVA, which is glycosylated, and CSP, which

is nonglycosylated) to CD8+ and CD4+ TCR transgenic T cells

(Figure 6A). Thus just like the Mo-DCs that can be generated in

culture by adding GM-CSF and IL-4 to monocytes, LPS-mobi-

lized Mo-DCs in vivo are as good or better presenting cells than

classical DCs, including cross-presentation.

To consider antigen capture in vivo, we injected LPS, then

soluble CSP or OVA protein s.c. 10 hr later. At 12 hr, or 2 hr after

CSP/OVA injection, we isolated DC-SIGN+ Mo-DCs as well as

DEC-205+ and DEC-205�, DC-SIGN� classical DCs from the no-

des. When added in graded doses to TCR transgenic, CD4+ and

CD8+ T cells without further antigen, Mo-DCs were again

comparable or superior to classical DCs for both CSP and

OVA (Figure 6B and Figure S6), showing that these cells capture

and present on both MHC I and II in vivo.

A B

C D

Figure 4. Monocyte Origin of DC-SIGN+ Mo-DCs

(A) CD45.2+ marrow monocytes were transferred i.v. into CD45.1+ hosts. Twenty-four hours later, PBS or 5 mg of LPS was injected i.v. with 10 mg Alexa 647-MMD3

a-DC-SIGN, and 24 hr later, DC-SIGN+ CD206+ DCs of CD45.2 origin were enumerated. This is one of three similar experiments.

(B) WT and LysMCre3 iDTR mice were injected with DT, and 12 hr later, blood monocytes (Ly6G� CD115+ CD11b+ Ly6Chi/lo) were analyzed (left panels). Twenty-

four hours after DT, 5 mg of LPS plus 10 mg of MMD3-Alexa 647 mAb were given i.v., and 12 hr later, skin-draining lymph node cells were analyzed as CD19/CD3/

NK1.1� and CD11c high and segregated into three DC populations (right) to look for DC-SIGN+ Mo-DCs (arrow).

(C) Lymph node sections were stained for Mo-DCs with a-DC-SIGN (BMD10, green), resident DCs with a-DEC-205 (NLDC145, red), and B cells with a-B220 (blue)

at 1003 magnification.

(D) WT and Flt3�/� mice were injected with 5 mg of LPS and 10 mg of MMD3-Alexa 647 a-DC SIGN mAb to enumerate Mo-DCs expressing DC-SIGN (blue) or

MMR/CD206 (red) 24 hr later. Shown are cells/106 lymph node cells from two independent experiments with 2 mice/group.

422 Cell 143, 416–429, October 29, 2010 ª2010 Elsevier Inc.

Page 111: Cell 101029

To determine the type of T cell that was developing in

response to antigen-presenting Mo-DCs, we collected the

medium from 4 day cocultures of OT-II CD4+ TCR transgenic

T cells with Mo-DCs that had captured OVA in vivo. The Mo-DCs

induced strong production of IFN-g and IL-2 but not IL-4, IL-10,

or IL-17 (Figure 6C), suggesting Th1 differentiation.

To evaluate presentation of bacterial antigens, we injected

recombinant E. coli OVA (or E. coli control). Twelve hours later,

we isolated three populations of DCs from the LNs. Mo-DCs

were even more effective than DEC-205+ classical cross-pre-

senting DCs, whereas DEC-205� DCs did not cross-present

bacteria (Figure 6D).

Mo-DCs Selectively Express CD14, a NeededCoreceptor for Trif-Dependent LPS SignalingTo begin to understand why LPS and gram-negative bacteria

were superior agonists for mobilizing Mo-DCs, we first used

quantitative PCR to assess expression of several TLRs in

marrow monocytes and Mo-DCs. Both cells expressed several

TLRs, but TLR4 and its coreceptor CD14 were markedly upregu-

lated in Mo-DCs (Figure 7A).

To pursue the contribution of the LPS coreceptor, CD14, we

used monoclonal anti-CD14 to show that monocytes were selec-

tively CD14+ in blood (Figure S7A), whereas among CD11chi DCs

in the LNs from LPS-stimulated mice, only DC-SIGN+ Mo-DCs

were CD14+ (Figure 7B). When we studied CD14�/� mice, which

lacked CD14 on monocytes (Figure S7B), LPS injection failed to

mobilize Mo-DCs (Figure 7C). We then compared mice lacking

the MyD88 and Trif adaptors for TLR4 signaling, where CD14

is a known coreceptor for MyD88-independent, Trif-dependent

signaling (Jiang et al., 2005). Trif, not MyD88, was essential

for LPS to mobilize Mo-DCs (Figure 7D) and to upregulate

CD86 on splenic DCs (Figure S7C). CD14+ DCs accumulated

with identical kinetics to DC-SIGN+ Mo-DCs, peaking at 24 hr

and becoming the dominant DCs in LNs (Figure 7E and Fig-

ure S7D). Together, the data indicate that CD14, a coreceptor

for TLR4, is upregulated by LPS and is essential for Mo-DC

differentiation via Trif signaling.

A B

C

D

Figure 5. L-Selectin and CCR7-Dependent Trafficking of DC-SIGN+ Mo-DCs

(A and B) mAb to block L-selectin (MEL-14, 100 mg i.v.) was given 1 hr before injection of LPS and a-DC-SIGN mAb. After 24 hr, LNs were analyzed by staining at

1003 magnification (A) or FACS (B).

(C) Chemokine receptor KO mice were injected with LPS i.v., and 24 hr later, lymph node cells were stained for intracellular DC-SIGN and MMR/CD206. Systemic

injection of LPS was confirmed by CD86 upregulation on spleen DCs (right).

(D) Blood chimerism 6 weeks after lethal irradiation and reconstitution with CD45.1 (WT) and CD45.2 (CCR7�/�) in CD45.1 hosts (left). Twelve hours after LPS,

blood monocytes had largely disappeared (middle). LNs from these same animals were stained for CD45.1, CD45.2, DC-SIGN, and MHC II to show that Mo-DCs

were WT in origin.

Cell 143, 416–429, October 29, 2010 ª2010 Elsevier Inc. 423

Page 112: Cell 101029

To find out if selective expression of CD14 provided an inde-

pendent means to isolate Mo-DCs after injecting antigens

in vivo, we compared CD14 surface labeling to MMD3 in vivo

labeling. With either approach, Mo-DCs were similar and supe-

rior cross-presenting DCs (Figure 7E). Thus monocyte differenti-

ation to DCs in response to LPS requires CD14, which serves as

an alternative marker to identify and isolate Mo-DCs from clas-

sical DCs.

DISCUSSION

One can use the term ‘‘authentic’’ for the Mo-DCs described

here for several reasons, which have not previously been noted

for inflammatory monocytes. The Mo-DCs are dendritic cells in

terms of their motility because they are nonadherent cells that

continually form and retract processes in the living state, iden-

tical to the probing morphology of DCs in the T cell areas of living

LNs (Lindquist et al., 2004). These Mo-DCs also concentrate in

the T cell areas, again a classic feature of DCs and a location

that facilitates clonal selection of antigen-specific T cells from

the recirculating repertoire. The Mo-DCs are very similar in

phenotype to DCs in lymphoid tissues including the loss of

markers that were used previously to positively identify inflam-

matory monocytes in vivo, i.e., Ly6C and Gr-1 antigens and

CD115/c-fms receptor.

Importantly, when Mo-DCs are compared functionally to clas-

sical DCs from the same LNs, the former are not only active but

can be superior in stimulating the MLR and in presenting protein

antigens, administered in vitro and also in vivo prior to testing as

presenting cells. A large amount of previous emphasis has been

A

B

C D

Figure 6. Presentation of Malaria CS and OVA Proteins by Three Types of DCs

(A). C57BL/6 or B6 3 BALB/c F1 mice were injected i.v. with 5 mg LPS for 12 hr to isolate three DC fractions (>95% purity), as in Figure 3B. Graded doses were

added with 40 mg/ml protein to 50,000 CFSE-labeled T cells, and 3–4 days later, CFSElo T cells were counted. An MLR was also run to verify DC activity.

(B) As in (A), but mice received 5 mg of LPS i.v. for 12 hr, as well as 50 mg of CSP or OVA protein s.c. in each paw for 2 hr before DC and B cell isolation. Repre-

sentative data of two experiments in triplicate or quadruplicate cultures are shown. Error bars = SD.

(C) As in (B), but enzyme-linked immunosorbent assay (ELISA) was used to measure the indicated cytokines in the medium of cocultures in which different types of

lymph node DCs were used to present OVA to OT-II CD4+ T cells. Error bars = SD.

(D) 5–10 3106 live E. coli-OVA or control E. coli were injected s.c. with 10 mg MMD3 mAb. Twelve hours later, three populations of lymph node DCs were isolated

and used to stimulate OT-I CD8+ T cells. This is representative of two experiments in triplicate. Error bars = SD.

424 Cell 143, 416–429, October 29, 2010 ª2010 Elsevier Inc.

Page 113: Cell 101029

placed on the superior cross-presenting activity of the CD8+ or

DEC-205+ subset of DCs, but the Mo-DCs we describe can be

equally or more effective than CD8+ DCs, including for bacteria

injected in vivo. Thus Mo-DCs are equivalent in many functional

respects to DCs, except that they are monocyte dependent,

whereas numerous prior studies show that classical DCs are

monocyte independent (Naik et al., 2006; Varol et al., 2007)

and derive from a committed pre-cDC in the bone marrow (Liu

et al., 2009). None of these new functional features of Mo-DCs

have been described before for monocyte-derived cells in

various inflammatory conditions.

The finding that permitted our research was the derivation of

mAbs to DC-SIGN or CD209a that recognized this lectin in tissue

sections, much of which are intracellular in location (Cheong

et al., 2010). The new anti-DC-SIGN/CD209a mAbs allowed us

to visualize the LPS-induced mobilization of Mo-DCs in the

T cell areas and distinguish them from the resident DCs there.

Previously, a combination of CD11b and CD11c markers were

used to help identify inflammatory monocytes with some

features of DCs (Leon et al., 2007; Serbina et al., 2003; Nakano

et al., 2009; Hohl et al., 2009; Siddiqui et al., 2010; Kool et al.,

2008), but these integrins are not sufficient to permit localization

A B

C

D

E F

Figure 7. Mo-DCs Selectively Express CD14, a Required Coreceptor for Their Mobilization

(A) Quantitative PCR to assess expression of mRNA for several TLRs and CD14 in marrow monocytes and Mo-DCs. Error bars = SD.

(B) DC-SIGN+ Mo-DCs colabel for CD14 expression.

(C) CD14�/� mice fail to mobilize Mo-DCs in response to LPS.

(D) DC-SIGN+ Mo-DCs are mobilized in MyD88�/� but not MyD88�/� 3 Trif�/� mice. The numbers of DC-SIGN+ Mo-DCs per million lymph node cells are on the

panels.

(E) Kinetics of formation and disappearance of Mo-DCs in LNs from LPS-treated mice, monitored by in vivo labeling of lymphocyte-negative, CD11chi DCs with

MMD3 anti-DC-SIGN mAb or by ex vivo labeling for CD14. Mo-DC’s numbers are averages of two mice each per time point.

(F) Mice were injected with LPS for 12 hr, and 2 hr prior to isolation CSP was injected. Mo-DCs were labeled either with MMD3 mAb in vivo or with anti-CD14 ex

vivo and used to stimulate CSP-specific CD8+ T cells.

Error Bars = SD. Representative data of at least two independent experiments (A–F) are shown.

Cell 143, 416–429, October 29, 2010 ª2010 Elsevier Inc. 425

Page 114: Cell 101029

in situ, and the Mo-DCs actually have lower levels of CD11c than

classical DCs. Previous isolations also used antibodies to Ly6C

or Gr-1, but these markers are lost from the Mo-DCs described

here.

Although DC-SIGN/CD209a was critical for identifying

authentic Mo-DCs in vivo, functions for this lectin need research.

We showed, for example, that DC-SIGN�/� monocytes become

Mo-DCs (marked by MMR/CD206) in the T cell areas, just like WT

monocytes, when the mice are given LPS (Figure S2E). Therefore

DC-SIGN seems not to be involved in Mo-DC mobilization and

differentiation. Also Mo-DCs cultured from DC-SIGN�/� mice

still present antigens to OT-I and OT-II transgenic T cells com-

parably to WT (not shown). DC-SIGN/CD209 can play patho-

genic roles, either in transmitting infectious agents like HIV and

CMV in the case of cultured human Mo-DCs (Geijtenbeek

et al., 2000a; Halary et al., 2002) or in transducing inhibitory

signals as seen when human DC-SIGN/CD209 interacts with

mycobacteria (Geijtenbeek et al., 2003; Tailleux et al., 2003).

DC-SIGN/CD209 could also have protective functions for cap-

ture and presentation of glycan-modified antigens (Tacken

et al., 2005). Also the pathway described here to mobilize

DC-SIGN/CD209a+ DCs could generate new vaccination strate-

gies, given the powerful antigen presentation and immune stim-

ulatory consequences of this full DC differentiation pathway.

We have identified one molecular pathway to produce

Mo-DCs in vivo, which is rapid differentiation from blood mono-

cytes upon administration of TLR4 agonists to mice. The clas-

sical method to produce DC-SIGN/CD209+ MMR/CD206+

Mo-DCs from human (Romani et al., 1994; Sallusto and Lanza-

vecchia, 1994) and mouse (Schreurs et al., 1999; Agger et al.,

2000) blood monocytes takes several days of culture in GM-

CSF and IL-4, but here we show that LPS and LPS+ live and

dead bacteria act rapidly within hours. Blood monocytes

drop to 20% of their normal levels 6–12 hr after i.v. LPS, and

at the same time, cells move into LNs and differentiate into

DC-SIGN/CD209a+ MMR/CD206+ Mo-DCs. This influx requires

CCR7 and CD62L, both expressed by bone marrow and blood

monocytes. Among the agonists for Toll-like receptors that

we studied, only LPS via TLR4 had this capacity to induce

Mo-DCs. In spite of hundreds of studies of the response of

mice to LPS, this mobilization of antigen-presenting cells was

not previously appreciated.

To explain the peculiar role of TLR4 agonists, we first exam-

ined gene expression for several TLRs. Whereas monocytes

expressed many TLRs, only TLR4 increased markedly when

monocytes differentiated into Mo-DCs in culture. This was also

the case for the CD14 coreceptor for TLR4, which mediates

MyD88-independent and Trif-dependent TLR4 signaling (Jiang

et al., 2005). Xu et al. have shown previously that GM-CSF/IL-

4-derived DCs produce cytokines in response to several ago-

nists, e.g., Pam3Cys and ODN1826 (Xu et al., 2007), which we

found did not mobilize Mo-DCs from monocytes in vivo.

However, a key feature of the Mo-DCs that are mobilized by

LPS is that they express CD14, which not only proved to be an

independent marker for Mo-DCs but was also essential for their

generation.

We would like to propose that the mobilization of Mo-DCs

described here has two roles. One is part of the innate response

to gram-negative bacteria and other agents that contain agonists

for the TLR4-CD14 complex, although this will require additional

studies of the functional properties of Mo-DCs such as the

production of cytokines and chemokines. A second is as a segue

to the adaptive immune response. During the TLR4-based

response, Mo-DCs increase while classical DCs decrease, so

that Mo-DCs become the dominant cell for induction of effective

and combined CD4+ and CD8+ T cell immunity, with or without

the requirement for bacterial replication in this newly mobilized

DC reservoir.

EXPERIMENTAL PROCEDURES

Mice

DC-SIGN�/� mice were from the Consortium for Functional Glycomics

(Scripps Res. Inst., La Jolla, CA, USA). Flt3�/� (I.R. Lemischka, Mount Sinai

School of Medicine), GMCSF-R�/� (G. Begley, Amgen), MyD88�/�(S. Akira,

Univ. of Osaka), and MyD88�/� 3 Trif�/� (E. Pamer, Memorial Sloan-Kettering

Cancer Center) were provided by M. Nussenzweig (Rockefeller Univ.),

iDTR mice by A. Waisman (Univ. of Mainz), and FcR g�/� mice by J. Ravetch

(Rockefeller Univ.). C57BL/6 (CD45.1 or CD45.2), C3H/HeJ, chemokine

receptor (CCR2, CCR5, CCR6, and CCR7), Lysozyme-M Cre (LysMcre), and

CD14�/�mice were from Jackson Labs and C3H/HeN and splenectomized

mice from Taconic Farms. Mice in specific pathogen-free conditions were

studied at 6–10 weeks according to institutional guidelines of the Rockefeller

University.

Lipopolysaccharide and Bacteria

LPS from E. coli 055:B5 (Sigma) was given i.v., s.c., or intraperitoneally (i.p.) at

a dose of 5 mg to induce Mo-DCs. For optimal LPS activity, stocks had to be

dissolved at 10 mg/ml or higher. Other TLR agonists were purchased from Inviv-

ogen and injected i.v. at 5 mg/mouse. We also tested bacteria at a dose of 5 3

106 per mouse, both heat-killed and live bacteria (E. coli DH5a, B. subtilis).

To evaluate presentation of proteins from bacteria, recombinant E. coli ex-

pressing OVA was used.

Bone Marrow Monocytes and DCs

Monocytes were sorted on a FACSAria (BD Biosciences) as SSClo, CD11bhi,

Ly6Chi or as Ly6G�, CD11bhi, Ly6Chi cells, the latter ensuring higher yields.

To generate Mo-DCs, monocytes were cultured with cytokines (M-CSF,

GM-CSF, GM-CSF, IL-4; PeproTech) at 20 ng/ml or Flt3-L at 200 ng/ml in

RPMI with 5% FBS and antibiotic-antimycotic plus b-mercaptoethanol (Invi-

trogen). At 4–7 days, nonadherent cells were removed to test function, or for

M-CSF, adherent cells were recovered with Cellstripper nonenzymatic cell

dissociation solution (Mediatech). Alternatively, to generate DCs, total bone

marrow was cultured with Flt3-L (400 ng/ml) for 9 days as described (Naik

et al., 2005), and the equivalents of CD8+ and CD8� spleen DCs were sorted

as CD24hi CD11blo and CD24lo CD11bhi cells, respectively.

Monocyte and Bone Marrow Transfer

23 106 CD45.2+ marrow monocytes were transferred to 4- to 6-week CD45.1+

mice (>8 weeks gave poor results). For mixed marrow chimeras, 50:50

mixtures of knockout (KO) and WT marrow were injected i.v. into lethally irra-

diated (5.5 Gy twice, 3 hr apart) mice. To deplete monocytes, DT (Sigma) in

PBS (1 mg/ml, stored at �80�C) was injected i.v. to LysMcre 3 iDTR mice at

25 ng/g weight (�500 ng/mouse).

Antibodies, Flow Cytometry, and Microscopy

Rabbit polyclonal antibody to a 14 amino acid cytoplasmic domain pep-

tide of DC-SIGN and mAbs to DC-SIGN (BMD10, BMD30, and MMD3)

were described (Cheong et al., 2010). mAbs were conjugated with biotin

or Alexa 647 (Invitrogen) following manufacturer’s instructions. These

bound specifically to CHO cells stably expressing mouse DC-SIGN. 22D1

(a-SIGN-R1/CD209b), SER4 (a-CD169), L31 (a-CD207), NLDC145 (a-DEC-

205/CD205), N418 (a-CD11c), KL295 (a-MHC II I-Ab/d b), GL117 (rat IgG2a

426 Cell 143, 416–429, October 29, 2010 ª2010 Elsevier Inc.

Page 115: Cell 101029

control), and MEL-14 (a-CD62L) mAbs were purified from hybridoma superna-

tants or purchased from eBioscience, and they were tested to be endotoxin

free (QCL-1000 kit, BioWhittaker). We purchased mAbs conjugated to different

fluorochromes to CD19, CD3, NK1.1, DX-5, CD206, CD11b, I-A/I-E (MHC II),

CD135, CD172a, CD14, and Ly6G from BD Bioscience; MMR/CD206 from

Biolegend; PE-a-mouse CD115, CD8a, Gr-1, CD11b, CD40, CD24, Mac-3,

CD62L, and CD14 from eBioscience; F4/80 and Ly6C (PE or Alexa 647) from

AbD Serotec. For BrdU labeling, 200 ml of 10 mg/ml of BrdU was injected

i.p. for 12 hr; staining followed manufacturer’s instruction (FITC BrdU flow

kit, BD).

Lymph Node Cells and Sections

Skin-draining nodes were treated with collagenase D (400 U/ml) for 30 min at

37�C. Cells were preincubated 10 min with 2.4G2 mAb at 4�C to block Fc

receptors, stained with fluorescent mAbs, acquired on a BD-LSRII, and

analyzed using flowjo (Treestar). To label Mo-DCs in vivo, we injected 10 mg

of Alexa 647-MMD3 a-DC-SIGN or control mouse IgG2c mAb along with

LPS. Lymphocytes (CD3+, CD19+, DX5+, or NK1-1+) and B220+ plasmacytoid

DCs were excluded, and three populations of CD11chi cells were separated

as DC-SIGN+, DEC-205+ (Alexa 488-NLDC145 mAb) and DEC-205�

DC-SIGN� DCs. CD19+ cells were also sorted. 10 mm OCT-embedded lymph

node sections were acetone-fixed, stained with BMD10 or BMD30 CD209a

mAb for 1 hr at room temperature or 4�C overnight, followed by mouse

anti-rat IgG2a-HRP for 30 min and Tyramide-signal amplification (Invitrogen).

B220-Alexa 647 stained B cell areas in confocal microscopy (LSM510, Zeiss).

We also injected into live mice 30 mg Alexa 488 MMD3 anti-DC-SIGN or

isotype control mAb i.v. Tissues were fixed in 4% HCHO/PBS for 20 min,

then 0.5% Triton X-100 for 15 min, and stained with rabbit anti-Alexa

488 and anti-rabbit HRP to label using TSA Alexa 488. For live-cell DIC

imaging, Mo-DCs were seeded on glass bottom culture dishes (MatTek)

and examined in an Olympus LCV110U incubator fluorescence microscope.

Confocal and live-cell images were analyzed with MetaMorph software

(Universal Imaging).

Splenic Monocytes and DCs

These were sorted from collagenase-digested spleen as monocytes (CD19�

CD3� DX-5� CD11b+ CD11cdim Ly6G� Ly6C+) and two classical DC subsets

(CD19� CD3� DX-5� CD11chi and either DEC-205+ or DEC-205� cells).

Antigen Presentation

T cells specific for OVA (OT-I, OT-II) or malarial (P. yoeli) circumsporozoite

protein (CSP) were cultured with graded doses of DCs or B cells. OVA

(LPS-free, Seikagaku Corp.) or CSP (Choi et al., 2009) was added in graded

doses but usually at 40 mg/ml in vitro, or the proteins were injected for 2 hr

in vivo (50 mg/foot pad) during LPS mobilization of Mo-DCs. In some exper-

iments, we used irradiated CHO cells stably transduced with OVA as the

source of antigen. Splenic transgenic T cells were enriched after Fc block

by excluding B220+, F4/80+, NK1.1+, I-Ab+, and CD4+ or CD8+ T cells using

anti-rat IgG Dynabeads (Invitrogen), labeled with 5 mM CFSE (Invitrogen) and

added to round bottom microtest wells at 50,000/well. After 3 days for OT-I

or 4 days for OT-II and CS T cells, proliferation of live (Aqua dye negative,

Invitrogen) T cells was evaluated by CFSE dilution and staining with mAb

to Va2 for the OT-I or OT-II TCR and Vb8.1/8.2 for CSP. For the MLR,

DCs from C57BL/6 mice were added in graded doses to CFSE-labeled

BALB/c T cells (NK1.1, I-A, B220, F4/80 negative cells) and assayed at

day 4.

Quantitative PCR for TLR and CD14 Expression by Monocytes

and Mo-DCs

Taqman probes (AssayID) were used for TLR4 (Mm00445273_m1), TLR2

(Mm00442346_m1), TLR3 (Mm00628112_m1), TLR7(Mm00446590_m1),

TLR9 (Mm00446193_m1), and CD14(Mm00438094_g1) from Applied Biosys-

tems. The relative expression was normalized by TATA-box binding protein

(TBP) housekeeping gene expression. All qPCR experiments were performed

with LightCycler 480 Real-Time PCR System (Roche).

SUPPLEMENTAL INFORMATION

Supplemental Information includes seven figures and one movie and can be

found with this article online at doi:10.1016/j.cell.2010.09.039.

ACKNOWLEDGMENTS

We thank members of the Steinman lab for valuable discussion, J. Adams for

graphics, A.J. North and S.A. Galdeen for DIC imaging at the bioimaging

resource center, S. Mazel and C. Bare for flow cytometry at the resource

center of Rockefeller University, Y. Oh and I. Jang for CSP preparation, J.D.

Schauer for mAb purification, J. Gonzalez for ELISA assays (Rockefeller

University Center for Clinical and Translational Science, UL1RR024143 from

National Center for Research Resource). We thank the Consortium for Func-

tional Glycomics supported by NIGMS (GM62116) for DC-SIGN/CD209a�/�

mice. We were supported by grants from the NIH (AI40045 and AI40874),

the Bill and Melinda Gates Foundation (R.M.S.), New York Community Trust’s

Francis Florio funds for blood diseases (C.C.), and a Fundacao para a Ciencia e

Tecnologia PhD scholarship (I.M. SFRH/BD/41073/2007).

Received: May 11, 2010

Revised: August 10, 2010

Accepted: September 23, 2010

Published: October 28, 2010

REFERENCES

Agger, R., Petersen, M.S., Toldbod, H.E., Holtz, S., Dagnaes-Hansen, F.,

Johnsen, B.W., Bolund, L., and Hokland, M. (2000). Characterization of murine

dendritic cells derived from adherent blood mononuclear cells in vitro. Scand.

J. Immunol. 52, 138–147.

Cheong, C., Matos, I., Choi, J.H., Schauer, J.D., Dandamudi, D.B., Shrestha,

E., Makeyeva, J.A., Li, X., Li, P., Steinman, R.M., et al. (2010). New monoclonal

anti-mouse DC-SIGN antibodies reactive with acetone-fixed cells. J. Immunol.

Methods 360, 66–75.

Choi, J.H., Do, Y., Cheong, C., Koh, H., Boscardin, S.B., Oh, Y.S., Bozzacco,

L., Trumpfheller, C., Park, C.G., and Steinman, R.M. (2009). Identification of

antigen-presenting dendritic cells in mouse aorta and cardiac valves. J. Exp.

Med. 206, 497–505.

D’Amico, A., and Wu, L. (2003). The early progenitors of mouse dendritic cells

and plasmacytoid predendritic cells are within the bone marrow hemopoietic

precursors expressing Flt3. J. Exp. Med. 198, 293–303.

de Villiers, W.J.S., Fraser, I.P., Hughes, D.A., Doyle, A.G., and Gordon, S.

(1994). Macrophage-colony-stimulating factor selectively enhances macro-

phage scavenger receptor expression and function. J. Exp. Med. 180,

705–709.

De Vries, I.J., Krooshoop, D.J., Scharenborg, N.M., Lesterhuis, W.J., Diepstra,

J.H., Van Muijen, G.N., Strijk, S.P., Ruers, T.J., Boerman, O.C., Oyen, W.J.,

et al. (2003). Effective migration of antigen-pulsed dendritic cells to lymph no-

des in melanoma patients is determined by their maturation state. Cancer Res.

63, 12–17.

den Haan, J.M., Lehar, S.M., and Bevan, M.J. (2000). CD8+ but not CD8-

dendritic cells cross-prime cytotoxic T cells in vivo. J. Exp. Med. 192, 1685–

1696.

Dhodapkar, M., Steinman, R.M., Sapp, M., Desai, H., Fossella, C., Krasovsky,

J., Donahoe, S.M., Dunbar, P.R., Cerundolo, V., Nixon, D.F., et al. (1999).

Rapid generation of broad T-cell immunity in humans after single injection of

mature dendritic cells. J. Clin. Invest. 104, 173–180.

Dudziak, D., Kamphorst, A.O., Heidkamp, G.F., Buchholz, V., Trumpfheller, C.,

Yamazaki, S., Cheong, C., Liu, K., Lee, H.W., Park, C.G., et al. (2007). Differen-

tial antigen processing by dendritic cell subsets in vivo. Science 315, 107–111.

Fogg, D.K., Sibon, C., Miled, C., Jung, S., Aucouturier, P., Littman, D.R.,

Cumano, A., and Geissmann, F. (2006). A clonogenic bone marrow progenitor

specific for macrophages and dendritic cells. Science 311, 83–87.

Cell 143, 416–429, October 29, 2010 ª2010 Elsevier Inc. 427

Page 116: Cell 101029

Geijtenbeek, T.B.H., Kwon, D.S., Torensma, R., van Vliet, S.J., van Duijnhoven,

G.C.F., Middel, J., Cornelissen, I.L.M.H.A., Nottet, H.S.L.M., KewalRamani,

V.N., Littman, D.R., et al. (2000a). DC-SIGN, a dendritic cell specific HIV-1

binding protein that enhances trans-infection of T cells. Cell 100, 587–597.

Geijtenbeek, T.B.H., Torensma, R., van Vliet, S.J., van Duijnhoven, G.C.F.,

Adema, G.J., van Kooyk, Y., and Figdor, C.G. (2000b). Identification of DC-

SIGN, a novel dendritic cell-specific ICAM-3 receptor that supports primary

immune responses. Cell 100, 575–585.

Geijtenbeek, T.B., van Vliet, S.J., Koppel, E.A., Sanchez-Hernandez, M.,

Vandenbroucke-Grauls, C.M., Appelmelk, B., and van Kooyk, Y. (2003). Myco-

bacteria target DC-SIGN to suppress dendritic cell function. J. Exp. Med. 197,

7–17.

Geissmann, F., Jung, S., and Littman, D.R. (2003). Blood monocytes consist of

two principal subsets with distinct migratory properties. Immunity 19, 71–82.

Geissmann, F., Manz, M.G., Jung, S., Sieweke, M.H., Merad, M., and Ley, K.

(2010). Development of monocytes, macrophages, and dendritic cells.

Science 327, 656–661.

Ginhoux, F., Tacke, F., Angeli, V., Bogunovic, M., Loubeau, M., Dai, X.M.,

Stanley, E.R., Randolph, G.J., and Merad, M. (2006). Langerhans cells arise

from monocytes in vivo. Nat. Immunol. 7, 265–273.

Ginhoux, F., Liu, K., Helft, J., Bogunovic, M., Greter, M., Hashimoto, D., Price,

J., Yin, N., Bromberg, J., Lira, S.A., et al. (2009). The origin and development of

nonlymphoid tissue CD103+ DCs. J. Exp. Med. 206, 3115–3130.

Goren, I., Allmann, N., Yogev, N., Schurmann, C., Linke, A., Holdener, M.,

Waisman, A., Pfeilschifter, J., and Frank, S. (2009). A transgenic mouse model

of inducible macrophage depletion: effects of diphtheria toxin-driven lysozyme

M-specific cell lineage ablation on wound inflammatory, angiogenic, and

contractive processes. Am. J. Pathol. 175, 132–147.

Granelli-Piperno, A., Pritsker, A., Pack, M., Shimeliovich, I., Arrighi, J.-F., Park,

C.G., Trumpfheller, C., Piguet, V., Moran, T.M., and Steinman, R.M. (2005).

Dendritic cell-specific intercellular adhesion molecule 3-grabbing noninte-

grin/CD209 is abundant on macrophages in the normal human lymph node

and is not required for dendritic cell stimulation of the mixed leukocyte reac-

tion. J. Immunol. 175, 4265–4273.

Halary, F., Amara, A., Lortat-Jacob, H., Messerle, M., Delaunay, T., Houles, C.,

Fieschi, F., Arenzana-Seisdedos, F., Moreau, J.F., and Dechanet-Merville, J.

(2002). Human cytomegalovirus binding to DC-SIGN is required for dendritic

cell infection and target cell trans-infection. Immunity 17, 653–664.

Heard, J.M., Roussel, M.F., Rettenmier, C.W., and Sherr, C.J. (1987). Multiline-

age hematopoietic disorders induced by transplantation of bone marrow cells

expressing the v-fms oncogene. Cell 51, 663–673.

Heath, W.R., and Carbone, F.R. (2001). Cross-presentation, dendritic cells,

tolerance and immunity. Annu. Rev. Immunol. 19, 47–64.

Heath, W.R., and Carbone, F.R. (2009). Dendritic cell subsets in primary and

secondary T cell responses at body surfaces. Nat. Immunol. 10, 1237–1244.

Hohl, T.M., Rivera, A., Lipuma, L., Gallegos, A., Shi, C., Mack, M., and Pamer,

E.G. (2009). Inflammatory monocytes facilitate adaptive CD4 T cell responses

during respiratory fungal infection. Cell Host Microbe 6, 470–481.

Jiang, Z., Georgel, P., Du, X., Shamel, L., Sovath, S., Mudd, S., Huber, M.,

Kalis, C., Keck, S., Galanos, C., et al. (2005). CD14 is required for MyD88-inde-

pendent LPS signaling. Nat. Immunol. 6, 565–570.

Johnson, W.D., Jr., Mei, B., and Cohn, Z.A. (1977). The separation, long-term

cultivation, and maturation of the human monocyte. J. Exp. Med. 146, 1613–

1626.

Kool, M., Soullie, T., van Nimwegen, M., Willart, M.A., Muskens, F., Jung, S.,

Hoogsteden, H.C., Hammad, H., and Lambrecht, B.N. (2008). Alum adjuvant

boosts adaptive immunity by inducing uric acid and activating inflammatory

dendritic cells. J. Exp. Med. 205, 869–882.

Leon, B., Lopez-Bravo, M., and Ardavin, C. (2007). Monocyte-derived

dendritic cells formed at the infection site control the induction of protective

T helper 1 responses against Leishmania. Immunity 26, 519–531.

Lindquist, R.L., Shakhar, G., Dudziak, D., Wardemann, H., Eisenreich, T., Dus-

tin, M.L., and Nussenzweig, M.C. (2004). Visualizing dendritic cell networks

in vivo. Nat. Immunol. 5, 1243–1250.

Liu, K., Iyoda, T., Saternus, M., Kimura, Y., Inaba, K., and Steinman, R.M.

(2002). Immune tolerance after delivery of dying cells to dendritic cells

in situ. J. Exp. Med. 196, 1091–1097.

Liu, K., Victora, G.D., Schwickert, T.A., Guermonprez, P., Meredith, M.M., Yao,

K., Chu, F.F., Randolph, G.J., Rudensky, A.Y., and Nussenzweig, M.C. (2009).

In vivo analysis of dendritic cell development and homeostasis. Science 324,

392–397.

Luckashenak, N., Schroeder, S., Endt, K., Schmidt, D., Mahnke, K., Bach-

mann, M.F., Marconi, P., Deeg, C.A., and Brocker, T. (2008). Constitutive

crosspresentation of tissue antigens by dendritic cells controls CD8+ T cell

tolerance in vivo. Immunity 28, 521–532.

Melief, C.J. (2008). Cancer immunotherapy by dendritic cells. Immunity 29,

372–383.

Mempel, T.R., Henrickson, S.E., and Von Andrian, U.H. (2004). T-cell priming

by dendritic cells in lymph nodes occurs in three distinct phases. Nature

427, 154–159.

Miller, M.J., Safrina, O., Parker, I., and Cahalan, M.D. (2004). Imaging the single

cell dynamics of CD4+ T cell activation by dendritic cells in lymph nodes.

J. Exp. Med. 200, 847–856.

Moron, V.G., Rueda, P., Sedlik, C., and Leclerc, C. (2003). In vivo, dendritic

cells can cross-present virus-like particles using an endosome-to-cytosol

pathway. J. Immunol. 171, 2242–2250.

Naik, S.H., Proietto, A.I., Wilson, N.S., Dakic, A., Schnorrer, P., Fuchsberger,

M., Lahoud, M.H., O’Keeffe, M., Shao, Q.X., Chen, W.F., et al. (2005). Gener-

ation of splenic CD8+ and CD8- dendritic cell equivalents in Fms-like tyrosine

kinase 3 ligand bone marrow cultures. J. Immunol. 174, 6592–6597.

Naik, S.H., Metcalf, D., van Nieuwenhuijze, A., Wicks, I., Wu, L., O’Keeffe, M.,

and Shortman, K. (2006). Intrasplenic steady-state dendritic cell precursors

that are distinct from monocytes. Nat. Immunol. 7, 663–671.

Naik, S.H., Sathe, P., Park, H.Y., Metcalf, D., Proietto, A.I., Dakic, A., Carotta,

S., O’Keeffe, M., Bahlo, M., Papenfuss, A., et al. (2007). Development of plas-

macytoid and conventional dendritic cell subtypes from single precursor cells

derived in vitro and in vivo. Nat. Immunol. 8, 1217–1226.

Nakano, H., Lin, K.L., Yanagita, M., Charbonneau, C., Cook, D.N., Kakiuchi, T.,

and Gunn, M.D. (2009). Blood-derived inflammatory dendritic cells in lymph

nodes stimulate acute T helper type 1 immune responses. Nat. Immunol. 10,

394–402.

Onai, N., Obata-Onai, A., Schmid, M.A., Ohteki, T., Jarrossay, D., and Manz,

M.G. (2007). Identification of clonogenic common Flt3+M-CSFR+ plasmacy-

toid and conventional dendritic cell progenitors in mouse bone marrow.

J. Exp. Med. 193, 233–238.

Park, C.G., Takahara, K., Umemoto, E., Yashima, Y., Matsubara, K., Matsuda,

Y., Clausen, B.E., Inaba, K., and Steinman, R.M. (2001). Five mouse homo-

logues of the human dendritic cell C-type lectin, DC-SIGN. Int. Immunol. 13,

1283–1290.

Regnault, A., Lankar, D., Lacabanne, V., Rodriguez, A., Thery, C., Rescigno,

M., Saito, T., Verbeek, S., Bonnerot, C., Ricciardi-Castagnoli, P., et al.

(1999). Fcg receptor-mediated induction of dendritic cell maturation and major

histocompatibility complex class I-restricted antigen presentation after

immune complex internalization. J. Exp. Med. 189, 371–380.

Romani, N., Gruner, S., Brang, D., Kampgen, E., Lenz, A., Trockenbacher, B.,

Konwalinka, G., Fritsch, P.O., Steinman, R.M., and Schuler, G. (1994). Prolifer-

ating dendritic cell progenitors in human blood. J. Exp. Med. 180, 83–93.

Sallusto, F., and Lanzavecchia, A. (1994). Efficient presentation of soluble

antigen by cultured human dendritic cells is maintained by granulocyte/macro-

phage colony-stimulating factor plus interleukin 4 and downregulated by

tumor necrosis factor a. J. Exp. Med. 179, 1109–1118.

Sallusto, F., Cella, M., Danieli, C., and Lanzavecchia, A. (1995). Dendritic

cells use macropinocytosis and the mannose receptor to concentrate

428 Cell 143, 416–429, October 29, 2010 ª2010 Elsevier Inc.

Page 117: Cell 101029

macromolecules in the major histocompatibility class II compartment: down-

regulation by cytokines and bacterial products. J. Exp. Med. 182, 389–400.

Sancho, D., Joffre, O.P., Keller, A.M., Rogers, N.C., Martinez, D., Hernanz-

Falcon, P., Rosewell, I., and Reis e Sousa, C. (2009). Identification of a dendritic

cell receptor that couples sensing of necrosis to immunity. Nature 458,

899–903.

Schnorrer, P., Behrens, G.M., Wilson, N.S., Pooley, J.L., Smith, C.M., El-Suk-

kari, D., Davey, G., Kupresanin, F., Li, M., Maraskovsky, E., et al. (2006). The

dominant role of CD8+ dendritic cells in cross-presentation is not dictated

by antigen capture. Proc. Natl. Acad. Sci. USA 103, 10729–10734.

Schreurs, M.W.J., Eggert, A.A.O., de Boer, A.J., Figdor, C.G., and Adema, G.J.

(1999). Generation and functional characterization of mouse monocyte-

derived dendritic cells. Eur. J. Immunol. 29, 2835–2841.

Schuler-Thurner, B., Dieckmann, D., Keikavoussi, P., Bender, A., Maczek, C.,

Jonuleit, H., Roder, C., Haendle, I., Leisgang, W., Dunbar, R., et al. (2000).

Mage-3 and influenza-matrix peptide-specific cytotoxic T cells are inducible

in terminal stage HLA-A.2.1+ melanoma patients by mature monocyte-derived

dendritic cells. J. Immunol. 165, 3492–3496.

Serbina, N.V., Salazar-Mather, T.P., Biron, C.A., Kuziel, W.A., and Pamer, E.G.

(2003). TNF/iNOS-producing dendritic cells mediate innate immune defense

against bacterial infection. Immunity 19, 59–70.

Shakhar, G., Lindquist, R.L., Skokos, D., Dudziak, D., Huang, J.H., Nussenz-

weig, M.C., and Dustin, M.L. (2005). Stable T cell-dendritic cell interactions

precede the development of both tolerance and immunity in vivo. Nat. Immu-

nol. 6, 707–714.

Siddiqui, K.R., Laffont, S., and Powrie, F. (2010). E-cadherin marks a subset of

inflammatory dendritic cells that promote T cell-mediated colitis. Immunity 32,

557–567.

Steinman, R.M., and Cohn, Z.A. (1973). Identification of a novel cell type in

peripheral lymphoid organs of mice. I. Morphology, quantitation, tissue distri-

bution. J. Exp. Med. 137, 1142–1162.

Steinman, R.M., and Witmer, M.D. (1978). Lymphoid dendritic cells are potent

stimulators of the primary mixed leukocyte reaction in mice. Proc. Natl. Acad.

Sci. USA 75, 5132–5136.

Stoll, S., Delon, J., Brotz, T.M., and Germain, R.N. (2002). Dynamic imaging of

T cell-dendritic cell interactions in lymph nodes. Science 296, 1873–1876.

Swirski, F.K., Nahrendorf, M., Etzrodt, M., Wildgruber, M., Cortez-Retamozo,

V., Panizzi, P., Figueiredo, J.L., Kohler, R.H., Chudnovskiy, A., Waterman, P.,

et al. (2009). Identification of splenic reservoir monocytes and their deployment

to inflammatory sites. Science 325, 612–616.

Tacken, P.J., de Vries, I.J., Gijzen, K., Joosten, B., Wu, D., Rother, R.P., Faas,

S.J., Punt, C.J., Torensma, R., Adema, G.J., et al. (2005). Effective induction of

naive and recall T-cell responses by targeting antigen to human dendritic cells

via a humanized anti-DC-SIGN antibody. Blood 106, 1278–1285.

Tailleux, L., Schwartz, O., Herrmann, J.-L., Pivert, E., Jackson, M., Amara, A.,

Legres, L., Dreher, D., Nicod, L.P., Gluckman, C.J., et al. (2003). DC-SIGN is

the major Mycobacterium tuberculosis receptor on human dendritic cells.

J. Exp. Med. 197, 121–127.

Varol, C., Landsman, L., Fogg, D.K., Greenshtein, L., Gildor, B., Margalit, R.,

Kalchenko, V., Geissmann, F., and Jung, S. (2007). Monocytes give rise to

mucosal, but not splenic, conventional dendritic cells. J. Exp. Med. 204,

171–180.

Vremec, D., and Shortman, K. (1997). Dendritic cells subtypes in mouse

lymphoid organs. Cross-correlation of surface markers, changes with incuba-

tion, and differences among thymus, spleen, and lymph nodes. J. Immunol.

159, 565–573.

Waskow, C., Liu, K., Darrasse-Jeze, G., Guermonprez, P., Ginhoux, F., Merad,

M., Shengelia, T., Yao, K., and Nussenzweig, M. (2008). The receptor tyrosine

kinase Flt3 is required for dendritic cell development in peripheral lymphoid

tissues. Nat. Immunol. 9, 676–683.

Xu, Y., Zhan, Y., Lew, A.M., Naik, S.H., and Kershaw, M.H. (2007). Differential

development of murine dendritic cells by GM-CSF versus Flt3 ligand has impli-

cations for inflammation and trafficking. J. Immunol. 179, 7577–7584.

Cell 143, 416–429, October 29, 2010 ª2010 Elsevier Inc. 429

Page 118: Cell 101029

Endophilin Functions as a Membrane-Bending Molecule and Is Delivered toEndocytic Zones by ExocytosisJihong Bai,1,2 Zhitao Hu,1,2 Jeremy S. Dittman,3 Edward C.G. Pym,1,2 and Joshua M. Kaplan1,2,*1Department of Molecular Biology, Massachusetts General Hospital, Boston, MA 02114, USA2Department of Neurobiology, Harvard Medical School, Boston, MA 02115, USA3Department of Biochemistry, Weill Cornell Medical College, New York, NY 10065, USA

*Correspondence: [email protected] 10.1016/j.cell.2010.09.024

SUMMARY

Two models have been proposed for endophilinfunction in synaptic vesicle (SV) endocytosis. Thescaffolding model proposes that endophilin’s SH3domain recruits essential endocytic proteins,whereas the membrane-bending model proposesthat the BAR domain induces positively curvedmembranes. We show that mutations disrupting thescaffolding function do not impair endocytosis,whereas those disrupting membrane bending causesignificant defects. By anchoring endophilin tothe plasma membrane, we show that endophilinacts prior to scission to promote endocytosis. Des-pite acting at the plasma membrane, the majority ofendophilin is targeted to the SV pool. Photoactivationstudies suggest that the soluble pool of endophilin atsynapses is provided by unbinding from the adjacentSV pool and that the unbinding rate is regulated byexocytosis. Thus, endophilin participates in an asso-ciation-dissociation cycle with SVs that parallels thecycle of exo- and endocytosis. This endophilin cyclemay provide a mechanism for functionally couplingendocytosis and exocytosis.

INTRODUCTION

Neurotransmitter released at synapses is drawn from a pool of

recycling synaptic vesicles (SVs). SVs are consumed by exocy-

tosis and are recycled by endocytosis. To maintain a releasable

pool of SVs, the rates of exo- and endocytosis must remain in

balance. Stimuli that increase SV exocytosis rates produce

corresponding increases in endocytosis rates, whereas endocy-

tosis is arrested following blockade of exocytosis (Dittman and

Ryan, 2009). Relatively little is known about how endocytosis is

regulated or how the competing processes of SV exocytosis

and endocytosis are coordinately regulated.

To begin addressing these questions, we focused on the en-

docytic protein endophilin. Endophilin is a conserved protein

harboring two functional domains: an N-terminal BAR (Bin–am-

phiphysin–Rvs) domain and a C-terminal SH3 (Src homology 3)

domain. Inactivation of endophilin produces profound defects

in SV endocytosis (Schuske et al., 2003; Verstreken et al.,

2002); however, the mechanism by which endophilin promotes

endocytosis has remained controversial.

Several studies suggest that endophilin acts primarily as

a scaffold, recruiting other essential endocytic proteins via its

SH3 domain (Dickman et al., 2005; Gad et al., 2000; Ringstad

et al., 1999; Schuske et al., 2003; Verstreken et al., 2002,

2003). Endophilin’s SH3 domain robustly binds to proline-rich

domains (PRDs) in dynamin and synaptojanin. Antibodies or

peptides that interfere with endophilin’s SH3-mediated interac-

tions impair SV recycling and cause accumulation of clathrin-

coated vesicles at lamprey synapses. In flies and worms,

mutants lacking endophilin have decreased synaptic abundance

of synaptojanin (Schuske et al., 2003; Verstreken et al., 2003).

Based on these data, endophilin was proposed to primarily

function as a molecular scaffold.

Analysis of endophilin’s BAR domain suggests an alternative

model. Recombinant BAR domains bind liposomes and induce

positive curvature of their membranes, as evidenced by the

conversion of spherical liposomes into elongated tubules

(Farsad et al., 2001). The endophilin BAR domain also alters

membrane morphology in transfected cells (Itoh et al., 2005).

Based on these data, endophilin (and potentially all BAR

proteins) was proposed to function by bending membranes.

Crystallographic studies suggested a potential mechanism for

the endophilin membrane-bending activity (Gallop et al., 2006;

Masuda et al., 2006). Homodimers of the endophilin BAR domain

form a concave membrane-binding surface, and specific

hydrophobic residues in the BAR domain are proposed to insert

into the outer membrane leaflet. Both of these features are pre-

dicted to promote positive membrane curvature. Although these

studies clearly demonstrate endophilin’s membrane-bending

ability, whether this activity is required for its endocytic function

has not been tested. Although all BAR domains share these

in vitro membrane-bending activities, each BAR protein regu-

lates distinct steps in membrane trafficking. Relatively little is

known about how BAR proteins are specifically targeted to

distinct membrane-trafficking events.

Here we examine the functional importance of the scaffolding

and membrane-bending activities of endophilin. We show that

430 Cell 143, 430–441, October 29, 2010 ª2010 Elsevier Inc.

Page 119: Cell 101029

the membrane-bending activity is essential for endophilin’s func-

tion and that endophilin undergoes an association-dissociation

cycle with SVs that parallels the cycle of exo- and endocytosis.

We propose that this endophilin cycle provides an activity

dependent mechanism for delivering endophilin to endocytic

zones.

RESULTS

Endophilin Function Requires the BAR Domainbut Not the SH3 DomainDue to their endocytosis defects, unc-57 endophilin mutants

have a smaller pool of SVs and a corresponding decrease in

synaptic transmission (Schuske et al., 2003). We exploited three

unc-57 mutant phenotypes as in vivo assays of endophilin func-

tion. First, unc-57 mutants had decreased locomotion rates

(Figures 1A and 1B; wild-type [WT] 147 ± 9 mm/s, unc-57 33 ±

3 mm/s; p < 0.001). Second, unc-57 mutants had a decreased

rate of excitatory postsynaptic currents (EPSCs) at body muscle

NMJs (Figures 1C and 1D; EPSC rates: WT 38 ± 2.1 Hz, unc-57

12 ± 0.9 Hz; p < 0.001). The mean EPSC amplitude was not

significantly altered in unc-57 mutants (see Figure S1A available

online; EPSC amplitudes: WT 22.7 ± 1.4 pA, unc-57 20.7 ± 0.6

pA; p = 0.25). Third, when endocytosis rates are diminished,

the SV protein synaptobrevin becomes increasingly trapped in

the plasma membrane. We utilized SynaptopHluorin (SpH) to

E F

A B

C D

Figure 1. The UNC-57 BAR Domain

Promotes SV Endocytosis through Its

Membrane Interactions

The phenotypes of wild-type (WT), unc-57(e406)

endophilin mutants, and the indicated transgenic

strains were compared. Transgenes were mCherry

tagged UNC-57 variants, including full-length (FL;

residues 1–379), BAR domain (residues 1–283),

and DN (residues 27-379). Transgenes were ex-

pressed in all neurons, using the snb-1 promoter.

Expression levels of these transgenes are shown

in Figure S1.

(A) Representative 1 min locomotion trajectories

are shown (n = 20 animals for each genotype).

The starting points for each trajectory were aligned

for clarity. (B) Locomotion rates are compared for

the indicated genotypes. Representative traces

(C) and summary data for endogenous EPSC rates

(D) are shown. Representative images (E) and

summary data (F) for axonal SpH fluorescence in

the dorsal nerve cord are shown for the indicated

genotypes. The number of worms analyzed for

each genotype is indicated. **, p < 0.001

compared to WT controls. ##, p < 0.001 when

compared to unc-57 mutants. Error bars, standard

error of the mean (SEM). See also Figure S1 and

Figure S2.

measure changes in surface synaptobre-

vin (Dittman and Kaplan, 2006). SpH

consists of a pH-sensitive GFP tag fused

to the extracellular domain of synap-

tobrevin. In SVs, SpH fluorescence is

quenched by the acidic pH of the vesicle lumen. Following SV

fusion, SpH fluorescence on the plasma membrane is de-

quenched (Dittman and Ryan, 2009). Endophilin mutants had

an 83% increase in SpH axon fluorescence compared to wild-

type controls, consistent with a defect in SV endocytosis

(Figures 1E and 1F).

Using these assays, we tested the importance of the BAR and

SH3 domains for endophilin’s function. Full-length and truncated

UNC-57 proteins were expressed in unc-57 mutants. Each

construct was tagged with mCherry at the C terminus, to control

for differences in transgene expression (Figure S1B). Quantita-

tive RT-PCR analysis showed that unc-57 transgenes were

expressed at approximately twice the level of the endogenous

unc-57 mRNA (Figure S1C). Mutant UNC-57 proteins lacking

the SH3 domain fully rescued the locomotion, EPSC, and

SpH defects (Figure 1). By contrast, UNC-57 proteins containing

a BAR domain mutation that disrupts membrane binding (DN,

deletion of N-terminal 26 residues; Gallop et al., [2006]) lacked

rescuing activity in all three assays (Figure 1) . Thus, UNC-57 en-

docytic function requires the membrane-binding BAR domain

but does not require the SH3 domain.

Testing the Scaffolding ModelAlthough the SH3 domain was not required for rescuing activity

(Figure 1), it remained possible that UNC-57 primarily functions

as a scaffold molecule recruiting other endocytic proteins. We

did several experiments to further test the scaffolding model.

Cell 143, 430–441, October 29, 2010 ª2010 Elsevier Inc. 431

Page 120: Cell 101029

Consistent with prior studies (Schuske et al., 2003; Verstreken

et al., 2003), we found that the fluorescence intensity of

GFP::UNC-26 synaptojanin puncta was slightly reduced in the

unc-57 endophilin mutants (83% wild-type level; p < 0.01)

(Figures S2A and S2B). Photobleaching experiments demon-

strated that approximately 50% of GFP::UNC-26 was immobile

in wild-type animals and that this immobile fraction was unal-

tered in either unc-57 mutants or in mutants rescued with the

BAR domain (Figures S2D–S2F). Consequently, synaptojanin

must have additional binding partners beyond endophilin at

synapses. Expressing mutant UNC-57 proteins lacking the

SH3 domain rescued the unc-57 endocytic defects but failed

to rescue the UNC-26 synaptojanin localization defects. In fact

rescued animals had less UNC-26 puncta fluorescence than

was observed in unc-57 mutants (50% and 80% wild-type levels,

respectively, p < 0.001) (Figures S2A and S2B). Similarly,

expressing a mutant UNC-26 protein lacking the PRD rescued

the locomotion defects of unc-26 mutants (Figure S2C). These

results agree with a prior study showing that mutations prevent-

ing the interaction of mouse synaptojanin and endophilin caused

only modest endocytic defects (Mani et al., 2007). Collectively,

these results support the notion that interactions between endo-

philin and synaptojanin do not play an essential role in SV endo-

cytosis, although it remains possible that these interactions

regulate endocytosis in some manner. These experiments also

suggest that the modest changes in UNC-26 synaptojanin tar-

geting are unlikely to account for the unc-57 endocytic defect.

To further address the scaffolding model, we analyzed two

additional endocytic proteins. GFP-tagged dynamin (DYN-

1::GFP) and the AP2a-subunit (APT-4::GFP) were both localized

to diffraction-limited puncta adjacent to presynaptic elements

(labeled with mRFP::SNB-1), suggesting that these reporters

are localized to perisynaptic endocytic zones. DYN-1 and

APT-4 puncta intensities were significantly increased in unc-57

mutants (Figure S2), indicating increased synaptic abundance

when endophilin was absent. Mislocalization of DYN-1 in

unc-57 mutants could arise from the absence of DYN-1 interac-

tions with the UNC-57 SH3. Contrary to this idea, expression of

mutant UNC-57 proteins lacking the SH3 corrected the DYN-1

puncta defects, whereas those carrying mutations that prevent

membrane binding (DN) abolished rescuing activity. These

data suggest that the increased synaptic recruitment of DYN-1

and APT-4 observed in unc-57 mutants is a secondary conse-

quence of the endocytic defect and do not support a role for

endophilin as a molecular scaffold.

Testing the Membrane-Bending ModelTo test the membrane-bending model, we analyzed mutations

that disrupt various aspects of BAR domain function in vitro.

For these experiments we used the BAR domain derived from

rat endophilin A1 (rEndoA1) because the impact of these muta-

tions on BAR domain activity and structure has only been

analyzed for the mammalian proteins. Transgenes were

expressed at similar levels (Figure S3A). Expression of rEndoA1

BAR rescued the locomotion, SpH, and EPSC defects of unc-57

mutants (Figure 2 and Figure S3B). Mutations disrupting

membrane binding [rEndoA1 BAR(DN)] failed to rescue both

the locomotion and SpH defects of unc-57 mutants (data not

shown), consistent with the results we obtained with the UNC-

57(DN) mutant. These results indicate that the rEndoA1 BAR

domain retains endocytic function in C. elegans neurons.

Endophilin’s tubulation activity in vitro is diminished by muta-

tions that prevent dimerization of the BAR domain (DH1I) and by

mutations that replace hydrophobic residues in the H1 helix with

polar residues (M70S/I71S double mutant) (Gallop et al., 2006).

Conversely, membrane-bending activity is enhanced by a muta-

tion that increases hydrophobicity of the H1 helix (A66W)

(Masuda et al., 2006). Due to its increased membrane-bending

activity, the A66W protein also lacks tubulation activity and,

instead, promotes vesiculation of liposomes. None of these

mutations significantly alters the membrane-binding activity of

the BAR domain in vitro (Gallop et al., 2006; Masuda et al., 2006).

Transgenes encoding mutant rEndoA1 BAR domains were

expressed in unc-57 mutants. Both the dimerization mutant

(DH1I) and the tubulation defective mutant (M70S/I71S) had

significantly less rescuing activity for the unc-57 locomotion,

SpH, and EPSC rate defects compared to the wild-type rEndoA1

BAR domain (Figure 2 and Figure S3B). Interestingly, the A66W

mutant (which has enhanced membrane-bending activity) also

exhibited decreased rescuing activity in all three assays (Figure 2

and Figure S3B). None of these tubulation mutants significantly

altered endogenous EPSC amplitudes (Figures S3C and S3D).

These results indicate that endophilin mutations altering

membrane tubulation activity produce corresponding defects

in SV endocytosis in vivo, consistent with the membrane-

bending model. These results also suggest that decreased and

increased membrane-bending activities are both detrimental to

SV endocytosis.

Specificity of the BAR DomainMembrane association and in vitro tubulation activities are

common features of most if not all BAR domains; however,

only a few BAR domain proteins have been implicated in SV

endocytosis. Thus, BAR domains must have other features

that confer specificity for their corresponding membrane-traf-

ficking functions. To test this idea we analyzed BAR domains

derived from two other proteins. The endophilin B and amphi-

physin BAR domains both have in vitro tubulation activity (Farsad

et al., 2001; Peter et al., 2004). Nonetheless, neither BAR domain

was able to rescue the unc-57 locomotion defects (Figure 2D),

although both were well expressed and targeted to axons

(data not shown). By contrast, efficient rescue was observed

with transgenes expressing rat and lamprey endophilin A

proteins. These results suggest that only endophilin A BAR

domains can promote SV endocytosis.

To compare their functional properties, we expressed the BAR

domains derived from the three rat endophilin A proteins in

unc-57 mutants. The rEndoA1 and A2 BAR domains fully

rescued the unc-57 locomotion defect, whereas the A3 BAR

domain had significantly less rescuing activity (Figure 2F).

Comparing the H1 helix sequence of these isoforms suggested

an explanation for this discrepancy. The rEndoA3 H1 helix

contains a hydrophobic tyrosine residue at position 64, whereas

the corresponding residue in the A1 and A2 isoforms is serine

(Figure 2E). An rEndoA3(Y64S) transgene had significantly

improved rescuing activity for the unc-57 locomotion defect

432 Cell 143, 430–441, October 29, 2010 ª2010 Elsevier Inc.

Page 121: Cell 101029

(Figure 2F). These results suggest that sequence differences in

the H1 helix contribute to the functional specificity of BAR

domains.

Endophilin Is Targeted to the SV PoolTo further examine how endophilin functions in endocytosis, we

analyzed where endophilin is localized in presynaptic elements.

For these experiments we utilized an UNC-57 construct

(UNC-57::CpG) containing two fluorophores, mCherry and

photo-activatable GFP (PAGFP). Expressing UNC-57::CpG in

all neurons (with the snb-1 promoter) efficiently rescued the

unc-57 locomotion defect (data not shown), suggesting that

this chimeric protein was functional.

UNC-57::CpG was highly enriched at synapses (synapse/

axon ratio = 8.6 ± 0.6; n = 38; Figure 3 and Figures 4E and 4F).

UNC-57 fluorescence colocalized with two SV markers,

GFP::SNB-1 (synaptobrevin) and GFP::RAB-3 (Figure 3A and

Figure S4A). By contrast the majority of UNC-57 fluorescence

did not colocalize with the endocytic markers APT-4::GFP and

DYN-1::GFP (Figure 3B and data not shown). Thus, at steady

state the majority of UNC-57 was targeted to the SV pool. This

conclusion is consistent with prior studies suggesting that endo-

philin cofractionates with SVs in biochemical purifications and

that anti-endophilin antibodies labeled SVs in immunoelectron

micrographs (Fabian-Fine et al., 2003; Takamori et al., 2006).

To further investigate how UNC-57 associates with the SV

pool, we analyzed unc-104 KIF1A mutants. In unc-104 mutants,

anterograde transport of SV precursors is defective, resulting in

a dramatic decrease in the abundance of SVs at synapses, and

a corresponding increase in the abundance of SVs in neuronal

cell bodies (Hall and Hedgecock, 1991). We found a similar shift

in UNC-57 abundance from axons to cell bodies in unc-104

mutants (Figure 3C), consistent with prior studies (Schuske

et al., 2003). These results suggest that UNC-57 and SV

FD

A B C

E

Figure 2. The Membrane-Bending Activity

of Endophilin A BAR Domains Promotes SV

Endocytosis

(A–C) Transgenes encoding wild-type and mutant

BAR domains (1–247) from rEndoA1 BAR were

analyzed for their ability to rescue locomotion rate

(A), the surface Synaptobrevin (SpH) (Figure S3B),

and EPSC rate (B and C) defects of unc-57 mutants.

The DH1, A66W, and M70S,I71S mutations alter

membrane tubulation activity but have little or no

effect on membrane binding in vitro (Gallop et al.,

2006; Masuda et al., 2006). All transgenes were

tagged with mCherry at the C terminus to assess

differences in expression levels (Figure S3).

(D) Transgenes expressing BAR domains derived

from different proteins were compared for their

ability to rescue the locomotion rate defect of

unc-57 mutants. BAR domains are indicated as

follows: rat endophilin A (rEndo A1, A2, and A3; resi-

dues 1-247); lamprey endophilin A (LampEndo; resi-

dues 1-248); C. elegans endophilin B (CeEndo B;

residues 1-267); rat endophilin B (rEndo B; residues

1-247); and ratamphiphysin (ramphiphysin; residues

1-250).

(E) Alignment of the H1 helix sequence is shown for the indicated BAR domains. The A66 residue (green, arrow) is required for tubulation activity (Masuda et al.,

2006). rEndo A3 has a sequence polymorphism (S64Y) compared to the A1 and A2 isoforms.

(F) Rescuing activities of rEndo A1, A2, A3, and A3(Y64S) BAR domains for the unc-57 mutant locomotion defect are compared. All transgenes were expressed in

all neurons using the snb-1 promoter. The number of animals analyzed for each genotype is indicated. ** and *, significant differences compared to WT (p < 0.001

and p < 0.01, respectively). ##, p < 0.001 when compared to unc-57 mutants.

Error bars, SEM. See also Figure S3.

A

B

C

Figure 3. Endophilin Is Targeted to the SV Pool at Presynaptic

TerminalsFull-length unc-57 endophilin was tagged at the C terminus with mCherry and

photoactivatable GFP (designated as CpG) (schematic shown in Figure 4A).

(A and B) The distribution of UNC-57::CpG mcherry fluorescence in DA neuron

dorsal axons is compared with a coexpressed SV (GFP::SNB-1 [A]) or endo-

cytic marker (APT-4::GFP AP2a [B]).

(C) Targeting of UNC-57::CpG to presynaptic terminals was strongly reduced

in unc-104(e1265) KIF1A mutants.

See also Figure S4.

Cell 143, 430–441, October 29, 2010 ª2010 Elsevier Inc. 433

Page 122: Cell 101029

precursors are cotransported to synapses by UNC-104 KIF1A,

as would be expected if UNC-57 were associated with SV

precursors. The UNC-57 targeting defect in unc-104 mutants is

unlikely to be a secondary consequence of an underlying defect

in active zone assembly because targeting of several active zone

proteins was unaltered in the unc-104 mutants (Kohn et al., 2000;

Koushika et al., 2001). Taken together, these results support the

idea that the majority of UNC-57 is targeted to SVs, despite the

fact that endophilin functions at endocytic zones (which

are lateral to the SV pool).

Given the preceding results, we would expect that the SV pool

contains binding sites that retain UNC-57. To test this idea we

analyzed UNC-57::CpG dispersion following photoactivation at

individual synapses (Figures 4A–4D). Photoactivated synaptic

UNC-57 rapidly dispersed into the axon (t = 28.1 ± 3.3 s;

n = 22), whereas the mCherry signal was unaltered. Mobile pho-

toactivated UNC-57 was rapidly recaptured at adjacent

synapses (Figure S4B). Photoactivated UNC-57 was not

observed in axons between synapses, presumably because

our imaging rate (1 frame/s; Figure S4B) was not fast enough

to detect the diffusion of mobile UNC-57. Although we could

not directly measure its diffusion rate, these results suggest

that UNC-57 released from the SV pool diffuses as a soluble

cytoplasmic protein (which would have a predicted t �140 ms).

A prior study showed that a subpopulation of SVs is mobile

and can be shared between adjacent synapses (Darcy et al.,

2006). Three results suggest that dispersion of UNC-57::CpG

is unlikely to reflect mobility of SVs bound to UNC-57: (1) photo-

recovery of an SV marker (GFP::RAB-3) was much slower than

that of UNC-57::GFP (Figure S4C); (2) given the slow mobility

of SVs, if the mobile fraction of UNC-57 remained bound to

SVs, we should have detected dispersion of photoactivated

UNC-57 in axons between synapses; and (3) a small fraction of

SVs (2%–4%) are mobile in cultured neurons (Darcy et al.,

2006; Jordan et al., 2005), whereas 60% of UNC-57 was

exchanged in 25 s (as measured by both photoactivation and

FRAP) (Figure 4C and Figure S4D). These data indicate that SV

mobility cannot account for the dispersion of photoactivated

UNC-57 and instead support the idea that dispersion is medi-

ated by unbinding of UNC-57 from the SV pool.

Exocytosis Regulates Endophilin Binding to the SV PoolBecause endophilin is targeted to the SV pool, it is possible that

endophilin is delivered to endocytic zones by exocytosis. If this

were the case, we would expect that mutations altering exocy-

tosis rates would also alter UNC-57 recruitment to synapses.

Consistent with this idea, UNC-57 puncta fluorescence was

significantly increased in both unc-18 Munc18 and unc-13

Munc13 mutants (12% and 1% wild-type EPSC rates,

respectively) (Madison et al., 2005; Weimer et al., 2003) (Figures

4E and 4F). Thus, decreased SV exocytosis was accompanied

by increased UNC-57 synaptic abundance. By contrast the

tom-1 Tomosyn mutation increases SV exocytosis (McEwen

et al., 2006) and caused a parallel decrease in UNC-57 puncta

fluorescence (data not shown). Double mutants lacking both

UNC-13 and TOM-1 had intermediate SV fusion rates (McEwen

et al., 2006) and UNC-57 synaptic abundance values that were

intermediate to those observed in either single mutant

(Figure 4F). These results show that bidirectional changes in

exocytosis rate produce opposite changes in UNC-57 synaptic

enrichment.

If exocytosis regulates UNC-57 targeting by altering binding to

the SV pool, exocytosis mutants should also alter the kinetics of

UNC-57 dispersion following photoactivation. Consistent with

this idea, dispersion rates were significantly reduced in unc-13

(t = 117.2 ± 13.6 s; n = 20; p < 0.001) and unc-18 (t = 136.2 ±

26.4 s; n = 12; p < 0.001) mutants compared to wild-type controls

(t = 28.1 ± 3.3 s; n = 22) (Figures 4C and 4D). An intermediate

dispersion rate was observed in tom-1 unc-13 double mutants

(t = 68.2 ± 6.5 s, n = 18) (Figure 4D). These results suggest

E F

B C

D

A Figure 4. Exocytosis Promotes Dissocia-

tion of Endophilin from the SV Pool

(A) Photoactivation of UNC-57::CpG at a single

synapse is shown schematically (above) and in

representative images (below).

(B and C) Representative images and traces of

photoactivated UNC-57::CpG green fluorescence

decay in wild-type (WT) and unc-13(s69) mutants.

The mCherry fluorescence was captured to control

for motion artifacts.

(D) Dispersion rates of photoactivated UNC-

57::CpG were quantified in the indicated geno-

types. Decay constants (t) are 28.1 ± 3.3 s for

WT; 117.2 ± 13.6 s for unc-13 (s69); 136.2 ±

26.4 s for unc-18 (e81); and 68.2 ± 6.5 s for

tom-1(nu468)unc-13(s69).

Representative images (E) and summary data (F)

for steady-state UNC-57::CpG mCherry fluores-

cence in the dorsal nerve cord axons were

compared for the indicated genotypes. (F)

Synaptic enrichment of UNC-57::CpG was calcu-

lated as follows: DF/F = (Fpeak � Faxon)/Faxon. The

number of animals analyzed for each genotype is

indicated. **, p < 0.001 compared to WT controls.

Error bars, SEM. See also Figure S5.

434 Cell 143, 430–441, October 29, 2010 ª2010 Elsevier Inc.

Page 123: Cell 101029

that exocytosis rates regulate UNC-57 dissociation from the SV

pool, thereby altering steady-state UNC-57 synaptic abun-

dance.

The exocytosis mutants utilized for these experiments

produce global changes in synaptic transmission at all

synapses. Consequently, changes in UNC-57 targeting at one

synapse may be caused by altered secretion at other synapses.

To address this possibility we inhibited exocytosis in a single

class of cholinergic neurons (the DA neurons) by expressing

a dominant-negative syntaxin mutant. Prior studies showed

that increasing the length of the linker between the transmem-

brane domain and the SNARE helix of Syntaxin inhibits

SNARE-mediated liposome fusion, presumably because the

longer juxtamembrane domain prevents close approximation

of the donor and target membranes (McNew et al., 1999). Trans-

genes expressing Tall Syntaxin in the DA neurons significantly

increased synaptic UNC-57 abundance (16.4 ± 1.0; p < 0.001)

and decreased the UNC-57 dispersion rate (t = 49.1 ± 3.3 s;

p < 0.001) compared to wild-type controls (synapse/axon ratio =

8.6 ± 0.6 and t = 28.1 ± 3.3 s, respectively) (Figure S5). These

results indicate that changes in exocytosis rates regulate

synaptic recruitment of UNC-57 in a cell autonomous manner,

as would be expected if exocytosis regulates UNC-57 binding

to the SV pool.

Structural Requirements for UNC-57 Regulationby ExocytosisWe did several experiments to determine how UNC-57

senses changes in exocytosis. A mutant UNC-57 protein

lacking the SH3 domain (BAR::CpG) rescues the unc-57

endocytic defects (Figure 1) and was properly targeted to the

SV pool (Figure S6A). In unc-13 mutants the BAR dispersion

rate was significantly decreased (WT t = 32.3 ± 2.8 s, unc-13

t = 122.6 ± 16.7; p < 0.001) (Figure 5A). By contrast

a mutation disrupting membrane binding, UNC-57(DN), elimi-

nated the effect of unc-13 mutations on dispersion rates

BA

C D

Figure 5. Structural Requirements for UNC-

57 Regulation by Exocytosis

Representative traces and summary data are

shown comparing the dispersion of mutant UNC-

57 proteins. Mutant proteins analyzed are: (A) WT

BAR domain lacking the SH3 (BAR reporter), and

full-length UNC-57 proteins containing the DN

(membrane-binding deficient) (B); A66W (tubula-

tion deficient) (C); and DH1I (dimerization deficient)

(D) mutations. Each mutant protein was tagged

with CpG, expressed in DA neurons, and their

dispersion rates compared following photoactiva-

tion in wild-type and unc-13 mutants. **, p < 0.001

compared to WT controls. n.s., nonsignificant.

Error bars, SEM. See also Figure S6.

(WT t = 16.2 ± 3.8 s, unc-13 t = 19.1 ±

1.7 s; p = 0.49) (Figure 5B) and sig-

nificantly reduced UNC-57 synaptic

enrichment (Figure S6). These results

demonstrate that the membrane-bindingactivity of the BAR domain is required for UNC-57 regulation

by exocytosis.

We next asked if membrane-bending activity of the BAR

domain is required for regulation by exocytosis. The UNC-57

(A66W) mutant had increased membrane-bending activity

in vitro and decreased rescuing ability in vivo. Nonetheless, the

dispersion rates A66W and wild-type UNC-57 were indistin-

guishable and were slowed to the same extent in unc-13 mutants

(Figure 5C). Similarly, the dimerization defective UNC-57(DH1I)

mutant was localized to presynaptic elements (Figure S6), and

its dispersion rate was significantly reduced in unc-13 mutants

(WT t = 20.5 ± 2.5 s, unc-13 t = 48.5 ± 5.2 s; p < 0.001)

(Figure 5D). These data suggest that endophilin monomers

bind to SVs and that exocytosis stimulates unbinding of

monomers from SVs. Thus, membrane-bending activity is not

required for UNC-57 binding to the SV pool or for its regulation

by exocytosis.

Although monomeric UNC-57 retained the ability to sense

changes in exocytosis, theDH1I dispersion rate was significantly

faster than that observed for wild-type UNC-57 (Figure 5D), and

the DH1I synaptic enrichment was also reduced (Figure S6B).

These results suggest that monomeric UNC-57 binds to SVs

with lower affinity than UNC-57 dimers.

RAB-3 Promotes UNC-57 Targeting to the SV PoolIf UNC-57 binds directly to SVs, we would expect that a protein

associated with SVs would promote its synaptic targeting.

Several results suggest that the RAB-3 GTP-binding protein

enhances UNC-57 recruitment to the SV pool. To test the role

of RAB-3, we analyzed aex-3 mutants. The aex-3 gene encodes

the GDP/GTP exchange factor for RAB-3 and AEX-6 Rab27

(AEX-3 Rab3GEF). Mutants lacking AEX-3 have an SV

exocytosis defect that is very similar to the defect observed in

rab-3; aex-6 Rab27 double mutants (Mahoney et al., 2006). As

in other exocytosis mutants, photoactivated UNC-57 dispersed

more slowly in aex-3 mutants (t = 45.5 ± 5.1 s; p < 0.01;

Cell 143, 430–441, October 29, 2010 ª2010 Elsevier Inc. 435

Page 124: Cell 101029

Figure 6C). Given this reduced UNC-57 dispersion rate, we

would expect that aex-3 mutants would have increased synaptic

enrichment of UNC-57. Surprisingly, UNC-57 synaptic enrich-

ment was significantly reduced (28% decrease) in aex-3 mutants

(WT 8.6 ± 0.6, aex-3 6.2 ± 0.6, p < 0.01; Figures 6A and 6B),

unlike the increased enrichment observed in other exocytosis

mutants (e.g., unc-13 32.5 ± 2.1). This result is not due to

a generalized decrease in the abundance of SV proteins because

aex-3 mutants had increased SNB-1 synaptobrevin accumula-

tion (38% increase, p < 0.001) (Ch’ng et al., 2008). These

results suggest that inactivating the AEX-3 Rab3GEF reduced

UNC-57 recruitment to the SV pool but did not prevent

exocytosis-dependent regulation of UNC-57 unbinding from

the SV pool.

Because aex-3 mutants also have an exocytosis defect (and

consequently a decreased UNC-57 dispersion rate), it is likely

that this experiment underestimates the magnitude of the aex-

3 defect in UNC-57 synaptic recruitment. To more accurately

assess the role of AEX-3, we analyzed unc-13; aex-3 double

mutants, in which SV exocytosis is nearly completely blocked.

UNC-57 synaptic enrichment was significantly reduced in

unc-13; aex-3 double mutants (39% decrease), when compared

CB

A

D E

Figure 6. RAB-3 and the Rab3 GEF (AEX-3) Regu-

late Endophilin Targeting to SVs

Representative images (A) and quantification (B) of UNC-

57::CpG synaptic enrichment in WT, aex-3, unc-13, and

unc-13;aex-3 double mutants are shown (Synaptic enrich-

ment: WT 8.6 ± 0.6; aex-3 6.2 ± 0.6; unc-13; aex-3 19.6 ±

1.2, unc-13 32.1 ± 2.2-fold). Dispersion rates of UNC-

57::CpG in WT (t = 28.1 ± 3.3 s) and aex-3 mutant (t =

45.5 ± 5.1 s) animals were compared in (C). (D and E)

UNC-57::CpG distribution in transgenic unc-13 mutant

animals with overexpressed RAB-3 (Q81L) or (T36N) was

studied. Overexpression of RAB-3 (Q81L), but not RAB-3

(T36N), significantly reduced UNC-57::CpG synaptic

enrichment in unc-13 mutants. **, p < 0.001 and *, p <

0.01, compared to WT controls. Error bars, SEM.

to unc-13 single mutants (unc-13; aex-3 19.6 ±

1.2, unc-13 32.1 ± 2.2 fold; p < 0.001; Figures

6A and 6B). Thus, changes in exocytosis cannot

explain the aex-3 mutant defect in UNC-57

synaptic recruitment. Instead, these results

support the idea that AEX-3 promotes UNC-57

recruitment to the SV pool.

We next asked if the AEX-3 substrate RAB-3

regulates UNC-57 targeting. In aex-3 mutants,

RAB-3 is absent from axons and accumulates

in neuronal cell bodies (Mahoney et al., 2006).

Therefore, defects in UNC-57 targeting could

arise from either lack of axonal RAB-3 or from

mis-regulation of RAB-3 GTP/GDP cycle.

Expression of a GTP-locked (Q81L) form of

RAB-3 significantly reduced UNC-57::CpG

synaptic accumulation in unc-13 mutants (Fig-

ures 6D and 6E). In contrast, the GDP-locked

(T36N) form of RAB-3 had no effect on UNC-

57 enrichment. Taken together, these data

suggest that AEX-3 and RAB-3,GTP regulate UNC-57 targeting

to the SV pool, even when exocytosis is blocked.

A Plasma Membrane-Anchored Endophilin Is Targetedto Endocytic Zonesunc-57 mutants accumulate coated membranes and invagi-

nated coated pits (Schuske et al., 2003). Based on these studies,

endophilin has been variously proposed to act before scission or

to promote uncoating of endocytic vesicles after scission. Our

preceding results suggest a third possibility. Endophilin may

also act prior to fusion, i.e., bound to SVs. To distinguish

between these possibilities, we designed a mutant form of endo-

philin that is constitutively bound to the plasma membrane

[UNC-57(PM)]. UNC-57(PM) contains full-length UNC-57 and

GFP fused to the N-terminus of the plasma membrane protein

UNC-64 Syntaxin1A (Figure S7A). To control for the impact of

Syntaxin’s cytoplasmic domains on UNC-57(PM), we also

analyzed a deletion mutant lacking the Syntaxin membrane-

spanning domain, termed UNC-57(Cyto).

We analyzed the subcellular distribution of UNC-57 when it is

constitutively anchored to the plasma membrane. Unlike UNC-

64 Syntaxin1A, which has a diffuse distribution on plasma

436 Cell 143, 430–441, October 29, 2010 ª2010 Elsevier Inc.

Page 125: Cell 101029

membranes (Figure S7B), UNC-57(PM) was highly enriched in

synaptic puncta (Figure 7). Although UNC-57(PM) and wild-

type UNC-57 were both punctate, their properties differed in

several respects. UNC-57(PM) puncta were significantly smaller

than UNC-57 puncta (Figure 7A; puncta width: UNC-57(PM):

0.51 ± 0.02 mm, n = 28; UNC-57: 0.75 ± 0.02 mm, n = 38;

p < 0.001). Furthermore, a majority of the UNC-57(PM) fluores-

cence was colocalized with the endocytic marker APT-4::GFP,

whereas far less colocalization was observed with SV pool,

labeled with either wild-type UNC-57::CpG (Figure 7) or mcher-

ry::RAB-3 (data not shown). By contrast wild-type UNC-57 had

the converse pattern, exhibiting greater colocalization with the

SV pool than with endocytic zones (Figure 3 and Figure S4A).

Because UNC-57(PM) and APT-4 puncta are both diffraction

limited, it remained possible that these proteins are localized to

distinct presynaptic subdomains that cannot be resolved by

conventional confocal microscopy. We did several experiments

to control for this possibility. First, UNC-57(PM) is unlikely to be

targeted to active zones because it failed to colocalize with the

active zone marker ELKS-1 (Figure 7A). Second, if UNC-57

(PM) is targeted to endocytic zones, then it should behave like

other endocytic zone proteins. We previously showed that

unc-13 mutations have opposite effects on the synaptic

abundance of SV proteins (increasing SNB-1 and RAB-3) versus

endocytic proteins (decreasing APT-4 AP2a) (Ch’ng et al., 2008;

Dittman and Kaplan, 2006). The reduced targeting of APT-4 to

endocytic zones is presumably caused by the decreased abun-

dance of SV cargo in the plasma membrane when exocytosis is

blocked (Dittman and Kaplan, 2006). UNC-57(PM) puncta fluo-

rescence was significantly decreased (42% ± 3% reduction; p

< 0.001) in unc-13 mutants (Figure 7B), which is similar to the

behavior of APT-4 (21% decrease; p < 0.001), and opposite to

the behavior of RAB-3 (26% increase; p < 0.01) (Ch’ng et al.,

2008). Thus, when exocytosis is blocked, UNC-57(PM) behaves

like an endocytic protein, and not like an SV-associated protein.

In contrast a membrane-anchored BAR domain [BAR(PM)], lack-

ing the SH3 domain, had a diffuse axonal distribution similar to

UNC-64 Syntaxin (Figure 7C). UNC-57(Cyto), which lacks the

Syntaxin transmembrane domain, behaved similarly to wild-

type UNC-57 and other SV proteins, i.e., its synaptic abundance

was increased in unc-13 mutants (Figure S7C). Thus, the

membrane-spanning domain anchors UNC-57(PM) to the

plasma membrane, preventing its association with the SV pool.

Once anchored in the plasma membrane, the SH3 domain

promotes UNC-57 targeting to endocytic zones.

UNC-57(PM) Rescues the Endocytic Defects of unc-57MutantsTo determine if UNC-57 functions on the plasma membrane, we

assayed the ability of UNC-57(PM) rescue the synaptic defects

of unc-57 mutants. The UNC-57(PM) transgene rescued the

unc-57 mutant locomotion and endogenous EPSC rate defects

(Figures 7D and 7E). Thus, a plasma membrane-anchored form

of UNC-57 retains the ability to promote endocytosis. These

results suggest that UNC-57 promotes endocytosis by

regulating a step that occurs prior to both scission and uncoating

of endocytic vesicles. Interestingly, the membrane-anchored

UNC-57(PM) protein had significantly less rescuing activity

than the UNC-57(Cyto) construct, which lacks the Syntaxin

membrane-spanning domain (Figures S7D–S7F). This discrep-

ancy suggests that the membrane-tethered protein cannot fully

reconstitute UNC-57’s endocytic function. For example,

A

D E

C

B Figure 7. Analysis of a Membrane-

Anchored UNC-57 Protein

(A) The distribution of UNC-57(PM) in DA neuron

axons was compared with coexpressed UNC-

57::CpG (upper panels), active zone [AZ] mark-

er ELKS-1::mcherry (middle panels), or endocytic

zone [EZ] marker APT-4::mcherry (AP2a, lower

panels). UNC-57(PM) comprises full-length

UNC-57 and GFP fused to the N-terminus of

UNC-64 Syntaxin 1A (schematic shown in

Figure S7A).

(B) GFP fluorescence of UNC-57(PM) in WT and

unc-13(s69) mutant animals were quantified.

UNC-57(PM) was expressed in all neurons with

the snb-1 promoter.

(C) Representative images are shown of wild-type

and mutant UNC-57(PM) proteins in dorsal cord

axons. The BAR(PM) protein corresponds to

UNC-57(PM) lacking the SH3 domain. The DN

(PM) protein lacks the N-terminal 26 residues of

UNC-57, which prevents membrane binding.

(D) Representative traces of endogenous EPSC

from WT, unc-57(e406) mutants, and transgenic

unc-57 animals carrying wild-type and mutant

UNC-57(PM) constructs.

Endogenous EPSC rates (left panel) and ampli-

tudes (right panel) are shown in (E). Significant differences (p < 0.001 by Student’s t test) are indicated as: **, compared to WT; and ##, compared to unc-57

mutants. n.s., nonsignificant.

Error bars, SEM. See also Figure S7.

Cell 143, 430–441, October 29, 2010 ª2010 Elsevier Inc. 437

Page 126: Cell 101029

UNC-57’s endocytic activity may be more potent when it is deliv-

ered via association with SVs. Alternatively, UNC-57 endophilin

may have additional functions that occur after scission.

Does membrane anchoring of UNC-57 bypass the require-

ment for direct interactions between membranes and the BAR

domain? Contrary to this idea, an UNC-57(PM) transgene

containing the DN mutation failed to rescue the unc-57 mutant

synaptic defects (Figures 7D and 7E), although this mutant

protein was efficiently targeted to synaptic puncta (Figure 7C).

The BAR(PM) protein, which lacks the SH3 domain, had a diffuse

axonal distribution (Figure 7C) yet rescued the EPSC defect to an

equivalent level as the UNC-57(PM) protein (Figures 7D and 7E).

Thus, a diffusely distributed membrane-anchored BAR domain

was sufficient to support SV endocytosis (Figures 7D and 7E).

Interestingly, endogenous EPSC amplitudes were significantly

larger in animals expressing BAR(PM) compared to those

observed in animals expressing UNC-57(PM) (p = 0.018; Figures

7D and 7E), suggesting that SV recycling had been subtly altered

by removing the SH3 domain.

DISCUSSION

Our results lead to six primary conclusions. First, endophilin

promotes SV endocytosis by acting as a membrane-bending

molecule, not as a molecular scaffold. Second, endophilin

functions on the plasma membrane, promoting an early step in

endocytosis (prior to scission of endocytic vesicles). Third,

endophilin A BAR domains are specialized to promote SV

endocytosis. Fourth, endophilin is targeted to synapses by its

association with the SV pool. Fifth, RAB-3 promotes endophilin

association with the SV pool. And sixth, endophilin dissociation

from the SV pool is regulated by exocytosis. Collectively, these

results argue that endophilin undergoes a membrane associa-

tion/dissociation cycle that parallels the SV cycle. Below we

discuss the implications of these results for understanding SV

endocytosis.

Endophilin Functions as a Molecular ScaffoldPrior studies proposed that endophilin primarily functions as

a scaffolding molecule, recruiting other endocytic proteins via

its SH3 domain (Dickman et al., 2005; Gad et al., 2000; Ringstad

et al., 1999; Schuske et al., 2003; Verstreken et al., 2002, 2003).

Consistent with these studies, we find that synaptojanin abun-

dance at synapses was modestly reduced, whereas DYN-1

and APT-4 AP2a abundance were increased in unc-57 mutants.

Several results argue against the idea that this putative

scaffolding function constitutes endophilin’s major role in endo-

cytosis. Deleting the SH3 domain did not impair the endocytic

function of UNC-57. Similarly, deleting the PRD did not impair

synaptojanin’s endocytic function, which agrees with analogous

experiments analyzing mouse synaptojanin (Mani et al., 2007).

Finally, changes in synaptojanin and dynamin targeting did not

correlate with rescue of the unc-57 endocytic defects. Thus,

altered recruitment of endocytic molecules is unlikely to account

for the endocytic defects of unc-57 mutants. Instead, these

localization defects are more likely a secondary consequence

of the endocytic defects.

What Is Endophilin’s Function in Endocytosis?Beyond scaffolding, several other mechanisms have been

proposed for endophilin’s endocytic function, including

promoting early steps (prior to scission) and later steps (e.g.,

uncoating of endocytosed vesicles). Our results indicate that

endophilin acts at the plasma membrane and, consequently,

must function prior to scission. An endophilin mutant that is

permanently anchored to the plasma membrane [UNC-57(PM)]

reconstitutes SV endocytosis when expressed in unc-57

mutants. UNC-57(PM) remains in the plasma membrane and

does not equilibrate into the recycled SV pool. Thus, at least

one aspect of endophilin function can be executed at the plasma

membrane. Our results do not exclude the possibility that endo-

philin also has a later function.

Our analysis suggests that the BAR domain, and its

membrane-bending activity, plays the primary and essential

function of endophilin in SV endocytosis. The curvature-inducing

activity of endophilin could promote internalization of cargo from

the plasma membrane. Consistent with this idea, the membrane-

anchored UNC-57(PM) protein was highly enriched at endocytic

zones. A prior study showed that endophilin accumulates along

the neck of plasma membrane invaginations following inactiva-

tion of dynamin, also consistent with endophilin acting prior to

scission (Ferguson et al., 2009). Alternatively, the membrane-

bending function of the BAR domain could act following scission,

perhaps by accelerating vesicle uncoating.

The SH3 domain is conserved in all endophilin proteins,

implying that it plays an important role. Although not essential

for endocytosis, several results indicate that the SH3 domain

regulates endophilin’s activity in certain contexts. Once

anchored to the plasma membrane, the SH3 domain targeted

UNC-57 to endocytic zones, presumably via interactions with

dynamin or synaptojanin. Although membrane-anchored con-

structs containing and lacking the SH3 domain [UNC-57(PM)

and BAR(PM)] rescued the unc-57 endocytic defects equally

well, EPSC amplitudes (a measure of quantal size) were signifi-

cantly increased by the BAR(PM) transgene. In principle, an

increased quantal size could be caused by delayed scission,

which would produce larger recycled SVs. Alternatively, this

defect could arise from faster refilling of recycled SVs with neuro-

transmitter (e.g., by increased recycling of VAChT transporters).

Whatever the mechanism involved, our results suggest that

endophilin alters quantal size only in specific circumstances

because EPSC amplitudes were not altered in unc-57 null

mutants. Similarly, at the Drosophila larval NMJ, Endophilin’s

effect on quantal size varied depending on the stimulus

rate (Dickman et al., 2005). Collectively, these results are most

consistent with the idea that endophilin has multiple functions

at the plasma membrane, perhaps including both internalization

of endocytic cargo and adjusting the timing of membrane

scission.

BAR Domain SpecificityMembrane-bending activity is a shared feature of most (perhaps

all) BAR proteins (Peter et al., 2004); however, only two BAR

proteins (endophilin and amphiphysin) have been implicated in

SV endocytosis. This suggests that BAR domains contain other

determinants that confer specificity for distinct membranes and

438 Cell 143, 430–441, October 29, 2010 ª2010 Elsevier Inc.

Page 127: Cell 101029

trafficking functions. In support of this idea, BAR domains

derived from endophilin B and amphiphysin did not rescue

unc-57 endocytic defects, whereas those derived from several

endophilin A proteins did rescue. The endocytic function of

rEndoA1 and A3 BAR domains differed significantly due to

a sequence difference in the H1 helix. Thus, the H1 helix may

confer functional specificity to BAR domains.

Endophilin Is Targeted to the SV PoolAlthough endophilin functions at endocytic zones, our results

suggest that that 90% of endophilin at presynaptic sites is bound

to the SV pool, whereas the remainder has a diffuse axonal distri-

bution. We propose that UNC-57 association with SVs is medi-

ated by at least two factors: direct binding of the BAR domain

to the SV membrane (disrupted by the DN mutant) and a second

RAB-3 dependent mode of SV binding (disrupted in aex-3 Rab3

GEF mutants). The RAB-3 effect is likely mediated by the GTP-

bound form of RAB-3 and is independent of RAB-3’s effect on

SV exocytosis. Further study is needed to determine if this is

mediated by direct binding of RAB-3 to UNC-57. Prior studies

also support endophilin’s association with the SV pool (Fabian-

Fine et al., 2003; Takamori et al., 2006).

SV Exocytosis Provides Soluble Endophilin at SynapsesOur results suggest that endophilin undergoes an association/

dissociation cycle with SVs and that dissociation from SVs is

stimulated by exocytosis. By analyzing a panel of mutants with

a range of exocytosis rates, we observed that the rate of

UNC-57 dispersion (or unbinding from the SV pool) was

positively correlated with the exocytosis rate. A mutant

UNC-57 lacking membrane-binding activity (DN) was not regu-

lated by the exocytosis rate, suggesting that binding of

UNC-57 to SVs is required to sense exocytosis. By contrast

neither tubulation defective mutants nor dimerization mutants

prevented UNC-57 regulation by exocytosis. Thus, distinct

biochemical properties of endophilin are required for binding to

SVs, sensing exocytosis, and promoting endocytosis. A conse-

quence of this mechanism for regulating endophilin availability

is that proteins previously thought to act solely during SV exocy-

tosis (e.g., RAB-3 and AEX-3 Rab3 GEF) also have the potential

to regulate endocytosis.

Implications for Regulating SV EndocytosisSV endocytosis is tightly coupled to exocytosis, which allows

neurotransmission to be sustained and presynaptic membrane

turnover to remain balanced. To date, the mechanism underlying

coupling of SV exo- and endocytosis is not well understood. Two

general models have been proposed. First, changes in presyn-

aptic calcium could potentially produce coordinated changes

in exo- and endocytosis because calcium potently regulates

both processes (Dittman and Ryan, 2009). A recent publication

proposed a second model, whereby rate-limiting endocytic

proteins are delivered to endocytic zones by associating with

SVs (Shupliakov, 2009). For example the endocytic proteins’ in-

tersectin and EPS15 were previously shown to associate with the

SV pool in resting synapses, but both are dynamically recruited

to endocytic zones following depolarization (Shupliakov, 2009).

Consistent with the latter model, we propose that wild-type

UNC-57 is delivered to synapses via its association with SVs,

that the endocytic pool of UNC-57 is provided by unbinding

from the adjacent SV pool, and that UNC-57 delivery to

endocytic zones is stimulated by exocytosis. The requirement

for SV-mediated delivery can be bypassed by artificially

anchoring UNC-57 to the plasma membrane. However, the

membrane-anchored protein had diminished rescuing activity,

implying that UNC-57’s endocytic activity is more potent when

delivered via association with SVs.

Several results support this model. Endophilin binds to the SV

pool, and dissociation from SV’s is stimulated by exocytosis. The

SV-bound pool of UNC-57 is likely to be inactive for several

reasons. First, SV binding sequesters endophilin away from en-

docytic zones. And second, our results suggest that endophilin

bound to SVs remains in an inactive, monomeric conformation.

Upon release from SVs, soluble endophilin monomers would be

free to form active dimers and to subsequently promote

membrane bending at endocytic zones. Because soluble

UNC-57 diffuses into the cytosol, we propose that exocytosis

would provide a pulse of active endophilin, thereby promoting

endocytosis at the adjacent endocytic zone. It is worth noting

that such an increase in endophilin concentration at endocytic

zones is transient, i.e., soluble endophilin concentration rapidly

decreases with time and distance, providing a tight temporal

and spatial control on exocytosis-endocytosis coupling. Calcium

regulation is unlikely to explain our results because presynaptic

Ca2+ currents were unaltered in Munc13-1/2 double knockout

neurons (Varoqueaux et al., 2002), yet unc-13 mutations potently

regulated UNC-57 unbinding from SVs. It is also possible that

both mechanisms act in concert to couple exo- and endocytosis.

Our results also predict that distinct endocytic mechanisms

may be employed during stimulus trains, versus those utilized

following stimulation. During a stimulus, soluble endophilin will be

continuously provided by ongoing SV exocytosis. By contrast,

following a stimulus, exocytosis rates decline, and the concen-

tration of soluble endophilin will drop dramatically. Thus, we

predict that endophilin does not play an important role in

compensatory endocytosis. Indeed, a slow form of SV endocy-

tosis persists in mutant flies lacking endophilin (Dickman et al.,

2005). Prior studies of dynamin-1 knockouts also support the

idea that distinct modes of endocytosis occur during versus after

stimulus trains (Ferguson et al., 2007). We speculate that delivery

of key endocytic proteins by SV exocytosis provides a potential

mechanism to explain the different modes of endocytosis that

occur at synapses. Because endophilin potentially functions at

multiple steps of the recycling pathway, these modes of endocy-

tosis may differ in several ways (e.g., endocytosis rate, quantal

size, and the rate at which recycled SVs become available for re-

release).

EXPERIMENTAL PROCEDURES

Strains

A full list of strains is provided in the Extended Experimental Procedures.

Transgenic animals were prepared by microinjection, and integrated trans-

genes were isolated following UV irradiation, as described (Dittman and

Kaplan, 2006).

Cell 143, 430–441, October 29, 2010 ª2010 Elsevier Inc. 439

Page 128: Cell 101029

Constructs

cDNAs of unc-57 and erp-1 were amplified from total mRNA extracted from

wild-type worms. cDNAs of rEndoA1, A2, A3, endophilin B1, and amphiphysin

were amplified from a cDNA library from Clontech (Mountain View, CA, USA).

cDNA of lamprey endophilin was synthesized by Genscript (Piscataway, NJ,

USA). All constructs were generated using modified pPD49.26 vectors.

A more detailed description of all constructs is provided in the Extended

Experimental Procedures.

In Vivo Microscopy and Image Analysis

Animals were immobilized with 2,3-Butanedione monoxamine (30 mg/ml;

Sigma-Aldrich), and images were collected with an Olympus FV-1000 confocal

microscope with an Olympus PlanApo 60 3 Oil 1.45 NA objective at 53 zoom,

a 488 nm Argon laser (GFP), a 559 nm diode laser (mCherry), and a 405 nm

diode laser (photoactivation). Detailed descriptions of the photoactivation

protocol and image analysis are provided in the Extended Experimental

Procedures.

Worm Tracking and Locomotion Analysis

Locomotion behavior of young adult animals (room temperature, off food) was

recorded on a Zeiss Discovery Stereomicroscope using Axiovision software.

The center of mass was recorded for each animal on each video frame using

object-tracking software in Axiovision. Imaging began 1 hr after worms were

removed from food.

Electrophysiology

Strains for electrophysiology were maintained at 20�C on plates seeded with

HB101. Adult worms were immobilized on Sylgard-coated coverslips with

cyanoacrylate glue. Dissections and whole-cell recordings were carried out

as previously described (Madison et al., 2005; Richmond and Jorgensen,

1999). Statistical significance was determined on a worm-by-worm basis

using the Mann-Whitney test or Student’s t test for comparison of mean

frequency and amplitude for endogenous EPSCs.

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures and

seven figures and can be found with this article online at doi:10.1016/j.cell.

2010.09.024.

ACKNOWLEDGMENTS

We thank the following for strains and reagents: Roger Tsien, Jennifer Lippin-

cott-Schwartz, and the C. elegans Genetics Stock Center. We thank members

of the Kaplan lab and Susana Garcia for critical comments. This work was

supported by a Jane Coffin Childs postdoctoral fellowship (J.B.) and by NIH

grants GM54728 (J.M.K.) and K99MH085039 (J.B.).

Received: February 11, 2010

Revised: June 2, 2010

Accepted: September 7, 2010

Published: October 28, 2010

REFERENCES

Ch’ng, Q., Sieburth, D., and Kaplan, J.M. (2008). Profiling synaptic proteins

identifies regulators of insulin secretion and lifespan. PLoS Genet. 4,

e1000283.

Darcy, K.J., Staras, K., Collinson, L.M., and Goda, Y. (2006). Constitutive

sharing of recycling synaptic vesicles between presynaptic boutons. Nat. Neu-

rosci. 9, 315–321.

Dickman, D.K., Horne, J.A., Meinertzhagen, I.A., and Schwarz, T.L. (2005).

A slowed classical pathway rather than kiss-and-run mediates endocytosis

at synapses lacking synaptojanin and endophilin. Cell 123, 521–533.

Dittman, J., and Ryan, T.A. (2009). Molecular circuitry of endocytosis at nerve

terminals. Annu. Rev. Cell Dev. Biol. 25, 133–160.

Dittman, J.S., and Kaplan, J.M. (2006). Factors regulating the abundance and

localization of synaptobrevin in the plasma membrane. Proc. Natl. Acad. Sci.

USA 103, 11399–11404.

Fabian-Fine, R., Verstreken, P., Hiesinger, P.R., Horne, J.A., Kostyleva, R.,

Zhou, Y., Bellen, H.J., and Meinertzhagen, I.A. (2003). Endophilin promotes

a late step in endocytosis at glial invaginations in Drosophila photoreceptor

terminals. J. Neurosci. 23, 10732–10744.

Farsad, K., Ringstad, N., Takei, K., Floyd, S.R., Rose, K., and De Camilli, P.

(2001). Generation of high curvature membranes mediated by direct endophi-

lin bilayer interactions. J. Cell Biol. 155, 193–200.

Ferguson, S.M., Brasnjo, G., Hayashi, M., Wolfel, M., Collesi, C., Giovedi, S.,

Raimondi, A., Gong, L.W., Ariel, P., Paradise, S., et al. (2007). A selective

activity-dependent requirement for dynamin 1 in synaptic vesicle endocytosis.

Science 316, 570–574.

Ferguson, S.M., Raimondi, A., Paradise, S., Shen, H., Mesaki, K., Ferguson, A.,

Destaing, O., Ko, G., Takasaki, J., Cremona, O., et al. (2009). Coordinated

actions of actin and BAR proteins upstream of dynamin at endocytic cla-

thrin-coated pits. Dev. Cell 17, 811–822.

Gad, H., Ringstad, N., Low, P., Kjaerulff, O., Gustafsson, J., Wenk, M., Di

Paolo, G., Nemoto, Y., Crun, J., Ellisman, M.H., et al. (2000). Fission and

uncoating of synaptic clathrin-coated vesicles are perturbed by disruption of

interactions with the SH3 domain of endophilin. Neuron 27, 301–312.

Gallop, J.L., Jao, C.C., Kent, H.M., Butler, P.J., Evans, P.R., Langen, R., and

McMahon, H.T. (2006). Mechanism of endophilin N-BAR domain-mediated

membrane curvature. EMBO J. 25, 2898–2910.

Hall, D.H., and Hedgecock, E.M. (1991). Kinesin-related gene unc-104 is

required for axonal transport of synaptic vesicles in C. elegans. Cell 65,

837–847.

Itoh, T., Erdmann, K.S., Roux, A., Habermann, B., Werner, H., and De Camilli,

P. (2005). Dynamin and the actin cytoskeleton cooperatively regulate plasma

membrane invagination by BAR and F-BAR proteins. Dev. Cell 9, 791–804.

Jordan, R., Lemke, E.A., and Klingauf, J. (2005). Visualization of synaptic

vesicle movement in intact synaptic boutons using fluorescence fluctuation

spectroscopy. Biophys. J. 89, 2091–2102.

Kohn, R.E., Duerr, J.S., McManus, J.R., Duke, A., Rakow, T.L., Maruyama, H.,

Moulder, G., Maruyama, I.N., Barstead, R.J., and Rand, J.B. (2000). Expres-

sion of multiple UNC-13 proteins in the Caenorhabditis elegans nervous

system. Mol. Biol. Cell 11, 3441–3452.

Koushika, S.P., Richmond, J.E., Hadwiger, G., Weimer, R.M., Jorgensen,

E.M., and Nonet, M.L. (2001). A post-docking role for active zone protein

Rim. Nat. Neurosci. 4, 997–1005.

Madison, J.M., Nurrish, S., and Kaplan, J.M. (2005). UNC-13 interaction with

syntaxin is required for synaptic transmission. Curr. Biol. 15, 2236–2242.

Mahoney, T.R., Liu, Q., Itoh, T., Luo, S., Hadwiger, G., Vincent, R., Wang, Z.W.,

Fukuda, M., and Nonet, M.L. (2006). Regulation of synaptic transmission by

RAB-3 and RAB-27 in Caenorhabditis elegans. Mol. Biol. Cell 17, 2617–2625.

Mani, M., Lee, S.Y., Lucast, L., Cremona, O., Di Paolo, G., De Camilli, P., and

Ryan, T.A. (2007). The dual phosphatase activity of synaptojanin1 is required

for both efficient synaptic vesicle endocytosis and reavailability at nerve termi-

nals. Neuron 56, 1004–1018.

Masuda, M., Takeda, S., Sone, M., Ohki, T., Mori, H., Kamioka, Y., and Mochi-

zuki, N. (2006). Endophilin BAR domain drives membrane curvature by two

newly identified structure-based mechanisms. EMBO J. 25, 2889–2897.

McEwen, J.M., Madison, J.M., Dybbs, M., and Kaplan, J.M. (2006). Antago-

nistic regulation of synaptic vesicle priming by Tomosyn and UNC-13. Neuron

51, 303–315.

McNew, J.A., Weber, T., Engelman, D.M., Sollner, T.H., and Rothman, J.E.

(1999). The length of the flexible SNAREpin juxtamembrane region is a critical

determinant of SNARE-dependent fusion. Mol. Cell 4, 415–421.

Patterson, G.H., and Lippincott-Schwartz, J. (2002). A photoactivatable GFP

for selective photolabeling of proteins and cells. Science 297, 1873–1877.

440 Cell 143, 430–441, October 29, 2010 ª2010 Elsevier Inc.

Page 129: Cell 101029

Peter, B.J., Kent, H.M., Mills, I.G., Vallis, Y., Butler, P.J., Evans, P.R., and

McMahon, H.T. (2004). BAR domains as sensors of membrane curvature:

the amphiphysin BAR structure. Science 303, 495–499.

Richmond, J.E., and Jorgensen, E.M. (1999). One GABA and two acetylcholine

receptors function at the C. elegans neuromuscular junction. Nat. Neurosci. 2,

791–797.

Ringstad, N., Gad, H., Low, P., Di Paolo, G., Brodin, L., Shupliakov, O., and De

Camilli, P. (1999). Endophilin/SH3p4 is required for the transition from early to

late stages in clathrin-mediated synaptic vesicle endocytosis. Neuron 24,

143–154.

Schuske, K.R., Richmond, J.E., Matthies, D.S., Davis, W.S., Runz, S., Rube,

D.A., van der Bliek, A.M., and Jorgensen, E.M. (2003). Endophilin is required

for synaptic vesicle endocytosis by localizing synaptojanin. Neuron 40,

749–762.

Shupliakov, O. (2009). The synaptic vesicle cluster: a source of endocytic

proteins during neurotransmitter release. Neuroscience 158, 204–210.

Takamori, S., Holt, M., Stenius, K., Lemke, E.A., Gronborg, M., Riedel, D., Ur-

laub, H., Schenck, S., Brugger, B., Ringler, P., et al. (2006). Molecular anatomy

of a trafficking organelle. Cell 127, 831–846.

Varoqueaux, F., Sigler, A., Rhee, J.S., Brose, N., Enk, C., Reim, K., and Rose-

nmund, C. (2002). Total arrest of spontaneous and evoked synaptic transmis-

sion but normal synaptogenesis in the absence of Munc13-mediated vesicle

priming. Proc. Natl. Acad. Sci. USA 99, 9037–9042.

Verstreken, P., Kjaerulff, O., Lloyd, T.E., Atkinson, R., Zhou, Y., Meinertzhagen,

I.A., and Bellen, H.J. (2002). Endophilin mutations block clathrin-mediated

endocytosis but not neurotransmitter release. Cell 109, 101–112.

Verstreken, P., Koh, T.W., Schulze, K.L., Zhai, R.G., Hiesinger, P.R., Zhou, Y.,

Mehta, S.Q., Cao, Y., Roos, J., and Bellen, H.J. (2003). Synaptojanin is re-

cruited by endophilin to promote synaptic vesicle uncoating. Neuron 40,

733–748.

Weimer, R.M., Richmond, J.E., Davis, W.S., Hadwiger, G., Nonet, M.L., and

Jorgensen, E.M. (2003). Defects in synaptic vesicle docking in unc-18

mutants. Nat. Neurosci. 6, 1023–1030.

Cell 143, 430–441, October 29, 2010 ª2010 Elsevier Inc. 441

Page 130: Cell 101029

EphB-Mediated Degradation of the RhoAGEF Ephexin5 Relieves a DevelopmentalBrake on Excitatory Synapse FormationSeth S. Margolis,1,3 John Salogiannis,1,3 David M. Lipton,1 Caleigh Mandel-Brehm,1 Zachary P. Wills,1 Alan R. Mardinly,1

Linda Hu,1 Paul L. Greer,1 Jay B. Bikoff,1 Hsin-Yi Henry Ho,1 Michael J. Soskis,1 Mustafa Sahin,2

and Michael E. Greenberg1,*1Department of Neurobiology, Harvard Medical School, 220 Longwood Avenue, Boston, MA 02115, USA2F.M. Kirby Neurobiology Center, Departments of Neurology, Children’s Hospital Boston, Harvard Medical School, Boston, MA 02115, USA3These authors contributed equally to this work

*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.09.038

SUMMARY

The mechanisms that promote excitatory synapseformation and maturation have been extensivelystudied. However, the molecular events that limitexcitatory synapse development so that synapsesform at the right time and place and in the correctnumbers are less well understood. We have identi-fied a RhoA guanine nucleotide exchange factor,Ephexin5, which negatively regulates excitatorysynapse development until EphrinB binding to theEphB receptor tyrosine kinase triggers Ephexin5phosphorylation, ubiquitination, and degradation.The degradation of Ephexin5 promotes EphB-dependent excitatory synapse development and ismediated by Ube3A, a ubiquitin ligase that is mutatedin the human cognitive disorder Angelman syndromeand duplicated in some forms of Autism SpectrumDisorders (ASDs). These findings suggest thataberrant EphB/Ephexin5 signaling during the devel-opment of synapses may contribute to the abnormalcognitive function that occurs in Angelman syn-drome and, possibly, ASDs.

INTRODUCTION

A crucial early step in the formation of excitatory synapses is the

physical interaction between the developing presynaptic

specialization and the postsynaptic dendrite (Jontes et al.,

2000; Ziv and Smith, 1996). This step in excitatory synapse

development is thought to be mediated by cell surface mem-

brane proteins expressed by the developing axon and dendrite

and appears to be independent of the release of the excitatory

neurotransmitter glutamate (reviewed in Dalva et al., 2007).

Several recent studies have revealed an important role for Ephrin

cell surface-associated ligands and Eph receptor tyrosine

kinases in this early cell-cell contact phase that is critical for

excitatory synapse formation (Dalva et al., 2000; Ethell et al.,

2001; Henkemeyer et al., 2003; Kayser et al., 2006; Kayser

et al., 2008; Lai and Ip, 2009; Murai et al., 2003). Ephs can be

divided into two classes, EphA and EphB, based on their ability

to bind the ligands EphrinA and EphrinB, respectively (reviewed

in Flanagan and Vanderhaeghen, 1998). EphBs are expressed

postsynaptically on the surface of developing dendrites, while

their cognate ligands, the EphrinBs, are expressed on both the

developing axon and dendrite (Grunwald et al., 2004; Grunwald

et al., 2001; Lim et al., 2008). When an EphrinB encounters an

EphB on the developing dendrite, EphB becomes autophos-

phorylated, thus increasing its catalytic kinase activity (reviewed

in Flanagan and Vanderhaeghen, 1998). This leads to a cascade

of signaling events including the activation of guanine nucleotide

exchange factors (GEFs) Tiam, Kalirin, and Intersectin, culmi-

nating in actin cytoskeleton remodeling that is critical for excit-

atory synapse development (reviewed in Klein, 2009). Consistent

with a role for EphBs in excitatory synapse development, EphB1/

EphB2/EphB3 triple knockout mice have fewer mature excit-

atory synapses in vivo in the cortex, and hippocampus (Henke-

meyer et al., 2003; Kayser et al., 2006). In addition, the disruption

of EphB function postsynaptically in dissociated hippocampal

neurons leads to defects in spine morphogenesis and a decrease

in excitatory synapse number (Ethell et al., 2001; Kayser et al.,

2006). Conversely, activation of EphBs in hippocampal neurons

leads to an increase in the number of dendritic spines and

functional excitatory synapses (Henkemeyer et al., 2003; Penzes

et al., 2003).These findings indicate that EphBs are positive

regulators of excitatory synapse development.

While there has been considerable progress in characterizing

the mechanisms by which EphBs promote excitatory synapse

development, it is not known if there are EphB-associated

factors that restrict the timing and extent of excitatory synapse

development. We hypothesized that neurons might have evolved

mechanisms which act as checkpoints to restrict EphB-medi-

ated synapse formation, and that the release from such synapse

formation checkpoints might be required if synapses are to form

at the correct time and place and in appropriate numbers.

We considered the possibility that likely candidates to mediate

the EphB-dependent restriction of excitatory synapse formation

might be regulators of RhoA, a small G protein that functions to

442 Cell 143, 442–455, October 29, 2010 ª2010 Elsevier Inc.

Page 131: Cell 101029

antagonize the effects of Rac (Tashiro et al., 2000). In previous

studies we identified a RhoA GEF, Ephexin1 (E1), which interacts

with EphA4 (Fu et al., 2007; Sahin et al., 2005; Shamah et al.,

2001). E1 is phosphorylated by EphA4 and is required for the

EphrinA-dependent retraction of axonal growth cones and

dendritic spines (Fu et al., 2007; Sahin et al., 2005). While E1

does not appear to interact with EphB, E1 is a member of a family

of five closely related GEFs. Of these GEFs, Ephexin5 (E5) (in

addition to E1) is highly expressed in the nervous system. There-

fore, we hypothesized that E5 might function to restrict the

EphB-dependent development of excitatory synapses by

activating RhoA.

In this study we report that EphB interacts with E5, that E5

suppresses excitatory synapse development by activating

RhoA, and that this suppression is relieved by EphrinB activation

of EphB during synapse development. Upon binding EphrinB,

EphB catalyzes the tyrosine phosphorylation of E5 which trig-

gers E5 degradation. We identify Ube3A as the ubiquitin ligase

that mediates E5 degradation, thus allowing synapse formation

to proceed. As Ube3A is mutated in Angelman syndrome and

duplicated in some forms of Autism Spectrum Disorders

(ASDs), these findings suggest a possible mechanism by which

the mutation of Ube3A might lead to cognitive dysfunction (Jiang

et al., 1998; Kishino et al., 1997). Specifically, we provide

evidence that in the absence of Ube3A, the level of E5 is elevated

and propose that this may lead to the enhanced suppression of

EphB-mediated excitatory synapse formation, thereby contrib-

uting to Angelman syndrome.

RESULTS

Ephexin5 Interacts with EphB2To identify mechanisms that restrict the ability of EphBs to

promote an increase in excitatory synapse number, we searched

for guanine nucleotide exchange factors (GEFs) that specifically

activate RhoA signaling, are expressed in the same population of

neurons that express EphB, are expressed at the same time

during development as EphB, and interact with EphB. Struc-

ture-function studies of GEFs identified amino acid residues in

the activation domain of Rho family GEFs that specifically iden-

tify the GEFs as activators of RhoA rather than Rac or Cdc42.

Applying this criterion, fourteen GEFs were identified that specif-

ically activate RhoA (Rossman et al., 2005). Of these GEFs we

found by in situ hybridization that E5 has a similar expression

pattern to EphB in the hippocampus (Figure 1A). These findings

raised the possibility that E5 might mediate the effect of EphB on

developing synapses.

We asked if E5 interacts physically with EphB. We transfected

HEK293T (293) cells with plasmids encoding Myc-tagged E5, E1,

or a vector control together with Flag-tagged EphB2 or EphA4

and asked if these proteins coimmunoprecipitate. Extracts

were prepared from the transfected 293 cells and EphA4 or

EphB2 immunoprecipitated with Flag antibodies. The immuno-

precipitates were subjected to SDS polyacrylamide gel electro-

phoresis (SDS-PAGE) and blotted with anti-Myc antibody

(a-Myc). We found that E5 coimmunoprecipitates with EphB2

but not with EphA4 (Figure 1B). The relatively weak E5 interaction

with EphA4 is consistent with published experiments (Ogita

E5-MycE1-MycFlag-EphA4Flag-EphB2

++ - -- +

-

+

++

++

- - --- -

- -+ +

- -

Input

A

IP: α-Flag

IB: α-Myc E5-Myc

E1-Myc

IB: α-FlagFlag-Eph

α-Myc

α-Myc

E5-Myc

E1-Myc

IB: α-Myc

Flag-Ephα-Flag

α-C-E5 InputIgG

150 kD

100 kD

150 kD

100 kD

B

IB:α-EphB2

IB:α-N-E5

IP:

EphB2

E5

α-EphB2 / / α-E5

EphB2

Anti-sense Sense Anti-sense Sense

DAPI DAPI DAPI DAPI

E5

C

% E

ph

B2

ove

rlap

wit

h E

505

1015

2025

30354045

DIV 2 DIV 4 DIV 8

*n.s.

E

D

CA1

DG

CA1

DG

CA1

DG

CA1

DG

Input

WT

IB:α-EphB2

α-C-E5 IP: α-C-E5

IB:α-EphB2

E5-/-

Figure 1. Ephexin5 Interacts with EphB2

(A) E5 and EphB2 are expressed in the CA1 region and dentate gyrus (DG) of

the hippocampus at P12. Adjacent 14 mm mouse brain sections were stained

for E5 or EphB2 using digoxigenin-labeled RNA probes to the antisense strand

or sense strand as a control (top). Lower panels show nuclear staining with

DAPI.

(B) Immunoprecipitation with a-Flag from 293 cell lysates previously trans-

fected with various combinations of overexpressing plasmids containing

E1-Myc, E5-Myc, Flag-EphB2, and/or Flag-EphA4, followed by immunoblot-

ting with a-Myc or a-Flag. Input protein levels are shown (bottom).

(C) Immunoprecipitation of mouse cortical lysates with IgG or a-C-E5, followed

by immunoblotting with a-EphB2 or a-N-E5 (left). Input protein levels are

shown (right).

(D) Immunoprecipitation of WT or E5�/� mouse cortical culture lysates with

a-C-E5 followed by immunoblotting with a-EphB2. Input EphB2 levels are

shown (bottom).

(E) Dissociated rat hippocampal neurons were stained using a-N-E5 (Blue) and

a-EphB2 (Red). A representative image of overlapped EphB2 and E5 is shown

(left). White rectangle outlines magnified dendritic region (right) showing

examples of EphB2/E5 colocalization (arrows). In three independent experi-

ments, quantification of overlapped EphB2/E5 puncta was determined at

DIV2, DIV4 and DIV8 and is represented as percent of EphB2 overlapped

with E5 (right). Error bars ± SEM; *p < 0.05, nonsignificant (n.s.).

See also Figure S1.

Cell 143, 442–455, October 29, 2010 ª2010 Elsevier Inc. 443

Page 132: Cell 101029

et al., 2003). By contrast, E1 is coimmunoprecipitated by EphA4

but not EphB2 (Shamah et al., 2001). These findings suggest that

E5 interacts preferentially with EphB2.

To extend this analysis we investigated whether EphB2

interacts with E5 in neurons. Neurons from embryonic day 16

(E16) mouse brains were lysed in RIPA buffer and the lysates

incubated with affinity purified anti-C-terminal E5 (a-C-E5) or

control (IgG) antibodies. The immunoprecipitates were then

resolved by SDS-PAGE and immunoblotted with affinity purified

anti-N-terminal E5 (a-N-E5) or EphB2 (a-EphB2) antibodies

(Figure 1C). This analysis revealed that endogenous, neuronal

EphB2 is immunoprecipitated by a-C-E5 but not IgG. Moreover,

using lysates from cortical cultures of wild-type or E5 knockout

mice (E5�/�, see Figure S1 available online), we find that

a-C-E5 immunoprecipitates EphB2 only from lysates when E5

is present (Figure 1D). Taken together, these findings suggest

that EphB interacts with Ephexin5 in neurons.

As an independent means of assessing if EphB and E5 interact

with one another, we used immunofluorescence microscopy to

determine if these two proteins colocalize in neurons. Cultured

mouse hippocampal neurons were transfected with a plasmid

expressing green fluorescent protein (GFP). The GFP-express-

ing neurons were imaged and quantified for the colocalization

of EphB2 and E5 puncta by staining with a-C-E5 and a-EphB2.

This analysis revealed that EphB2 and E5 colocalize along

dendrites (Figure 1E). We find that 40% of EphB staining over-

laps with a-C-E5 staining early during the development of

excitatory synapses. After eight days in vitro (DIV) the overlap

of EphB with E5 within neuronal dendrites decreases to below

the level that would be detected by random chance. This change

suggests that EphB interacts with E5 early during development

but that these two proteins may not interact later in development.

Ephexin5 Is a Guanine Nucleotide Exchange Factorthat Activates RhoATo determine if E5 activates RhoA, we transfected 293 cells with

a control plasmid or a plasmid that drives the expression of

Myc-tagged mouse E5. We prepared extracts from the trans-

fected cells and incubated the extracts with a GST-fusion protein

that includes the Rhotekin-Binding Domain (GST-RBD), a protein

domain that selectively interacts with active (GTP-bound) but not

inactive (GDP-bound) RhoA. Following SDS-PAGE of the

proteins in the extract that bind to GST-RBD, RhoA binding to

GST-RBD was measured by immunoblotting with a-RhoA anti-

bodies. We found that cells expressing E5 exhibited higher levels

of activated RhoA compared to cells transfected with a control

plasmid, indicating that E5 activates RhoA (Figure 2A).

When a similar series of experiments were performed using

a GST-fusion Pak-Binding Domain (GST-PBD) which specifically

interacts with active forms of two other Rho GTPases, Rac1 and

Cdc42, we found that E5 does not induce the binding of

GST-PBD to Rac1 or Cdc42. In contrast, E1-expressing cells

displayed enhanced binding of Rac1 and Cdc42 to GST-PBD.

We conclude that E5 activates RhoA but not Rac1 or Cdc42

(Figure S2A).

To determine whether E5 activation of RhoA requires the GEF

activity of E5, we generated a mutant form of E5 in which its GEF

activity is impaired. To identify the residues required for

Ephexin5 guanine nucleotide exchange activity we compared

its Dbl-homology (DH) domain to the DH domain of other

RhoA-specific GEFs (Snyder et al., 2002). We identified within

the a5 helix of E5’s DH domain three amino acids that are

conserved in other GEFs that, like E5, activate RhoA but not

Rac1 and Cdc42 (Figure S2B). To generate a form of E5 pre-

dicted to be inactive as a GEF, we mutated these three

conserved amino acids (L562, Q566, and R567) to alanine

(E5-LQR). Using the GST-RBD pull-down assay we found that

although E5-WT and E5-LQR are expressed at similar levels,

the E5-LQR mutant is significantly impaired relative to WT in its

ability to activate RhoA (Figure 2B). As a control, we mutated

other conserved residues within the a5 DH region to alanine

(Q547, S548, R555, and L556). When we tested this mutant we

observed no defect in RhoA activation, suggesting that the

E5-LQR mutation specifically disrupts the GEF activity of E5

and that the inability of the LQR mutant to activate RhoA is not

a general consequence of disrupting the a5 region of Ephexin5

(Figure S2C). Taken together, these findings indicate that E5

requires an intact conserved GEF domain to promote RhoA

activity in 293 cells, suggesting that E5 functions as a RhoA GEF.

We next asked if E5 expression affects RhoA activity in the

brain. We lysed P3 whole brains from wild-type or E5�/� mice

and performed a GST-RBD pull-down assay. This analysis re-

vealed a significant decrease in RhoA activation in brain extracts

from E5�/� mice compared to wild-type mice, suggesting that E5

is required to maintain wild-type levels of RhoA activity in the

brain (Figure 2C).

Ephexin5 Negatively Regulates ExcitatorySynapse NumberOur findings indicate that E5 interacts with EphB, a key regulator

of excitatory synapse development. Thus, we asked whether E5

plays a role in the development of excitatory synapses. We

generated two short hairpin RNA constructs that each knocks

down E5 protein levels when expressed in 293 cells or cultured

hippocampal neurons (Figures S3A–S3B). These shRNAs were

introduced into cultured hippocampal neurons together with a

plasmid that drives expression of green fluorescent protein

(GFP) to allow detection of the transfected cells. We found by

staining with a-N-E5 antibodies that the E5 shRNAs (E5-shRNA),

but not scrambled hairpin control shRNAs (ctrl-shRNA), effi-

ciently knocked down E5 expression in the transfected neurons

(Figure S3C).

By staining with antibodies that recognize pre- and postsyn-

aptic proteins or by visualizing dendritic spines in GFP trans-

fected neurons we observed a significant increase in the number

of excitatory synapses and dendritic spines that are present on

the E5-shRNA-expressing neurons compared to neurons ex-

pressing ctrl-shRNAs (Figures 3A and 3B). By contrast, we failed

to detect a significant change in dendritic spine length or width

under these conditions (Figure S3D). These findings suggest

that E5 functions to restrict spine/excitatory synapse number

but has no significant effect on spine morphology. Consistent

with these conclusions, we found that overexpression of E5 in

hippocampal neurons leads to a decrease in the number of excit-

atory synapses that are present on the E5-overexpressing

neurons (Figure 3C). This ability of E5 to negatively regulate

444 Cell 143, 442–455, October 29, 2010 ª2010 Elsevier Inc.

Page 133: Cell 101029

excitatory synapse number requires its RhoA GEF activity, as

overexpression of E5-LQR had no effect on synapse number

(Figure 3D).

To assess the effect of reducing E5 levels on the functional

properties of excitatory synapses, we recorded miniature excit-

atory postsynaptic currents (mEPSCs) from cultured hippo-

campal neurons transfected with E5-shRNA or ctrl-shRNA. We

observed an increase in the frequency and amplitude of mEPSCs

on neurons expressing E5-shRNA compared to ctrl-shRNA

(Figure 3E). This suggests that E5 acts postsynaptically to restrict

excitatory synapse function. The increase in mEPSC frequency

could be due to an increase in presynaptic vesicle release onto

the transfected neuron or an increase in the number of excitatory

synapses that are present on the transfected neuron. We favor

the latter possibility since our transfection protocol selectively

reduces E5 levels postsynaptically and also because an increase

in synapse number would be most consistent with the increase in

costaining of pre- and postsynaptic markers that we observe

when the level of E5 is reduced. The possibility that E5 functions

postsynaptically is further supported by immunofluorescence

staining experiments demonstrating that E5 is enriched in

dendrites relative to axons (Figure S1F).

As an independent means of assessing the importance of E5 in

the control of excitatory synapse number, we cultured hippo-

campal neurons from E5�/� mice or their wild-type littermates

for 10 days in vitro and then, following transfection of a GFP-ex-

pressing plasmid into these neurons, quantified the number of

excitatory synapses present on the transfected neuron at

DIV14. We observed a three-fold increase in the number of

synapses that are present on E5�/� neurons compared to E5+/�

neurons (Figure 4A). Taken together with the E5-shRNA knock-

down and E5 overexpression analyses, these findings suggest

that E5 acts postsynaptically to reduce excitatory synapse

number.

We next asked if E5 regulates synapse number in the context

of an intact developing neuronal circuit using conditional E5

(E5fl/fl) animals (see Figure S1). Upon introduction of Cre recom-

binase into E5fl/fl cells, exons 4–8 of the E5 gene are excised

resulting in a cell that no longer produces E5 protein (data not

A

α-Myc

α-Actin

RBD pull-down

GTPγS

α-RhoA

α-RhoA

Input

E5-Myc

WT WT

α-N-E5

α-Actin

α-RhoA

RBD pull-down

100 kD

25 kD

25 kD

α-RhoA

Input

KO KO KO

E5

Ephexin5 whole brain (P3)

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Rho

A s

ign

al (D

ensi

tom

etry

un

its)

*

C

B

RBD pull-down

Ctrl WT LQR

Input

α-RhoA

E5-Mycα-Myc

α-Actin

Ctrl WT

WT KO

Figure 2. Ephexin5 Is a GEF that Activates RhoA

(A) Lysates from 293 cells transfected with empty vector (Ctrl) or E5-Myc

overexpressing vector (WT) were assayed for endogenous RhoA activity using

the RBD pull-down assay and analyzed by immunoblotting with an antibody to

RhoA (top). GTPgS lane is a positive control for inducing RhoA activity.

Increased endogenous RhoA activity is demonstrated by presence of

a-RhoA signal in RBD pull-down lanes. Input protein levels and a-Actin loading

control are shown (bottom).

(B) Lysates from 293 cells transfected with empty vector (Ctrl), E5-Myc (WT) or

LQR mutant of E5-Myc (LQR) were assessed for RhoA activity as measured by

RBD assay described in (A). Input protein levels and a-Actin loading control are

shown (bottom).

(C) Presence of E5 is critical for wild-type levels of endogenous RhoA signaling

in vivo. P3 mouse whole brain lysates from WT or E5�/� (KO) littermates were

subjected to RBD pull-down assays as described in (A). A representative

immunoblot is shown (top). From three experiments, blinded to condition,

the quantification of a-RhoA signal in the RBD pull-down assay was normal-

ized to input RhoA signal (bottom). Error bars indicate ± SEM; *p < 0.05.

See also Figure S2.

Cell 143, 442–455, October 29, 2010 ª2010 Elsevier Inc. 445

Page 134: Cell 101029

A B

E

E5-shRNA

Ctrl-shRNA

GFP Synapses

GFP Synapses

0

0.5

1

1.5

2

No

rmal

ized

Exc

itat

ory

Syn

apse

Den

sity

10ng 20ng

*****

******

Ctrl-shRNAE5-shRNA 1E5-shRNA 2

N=

55

N=

56

N=

52

N=

52

N=

52

N=

48

D

μg of E5 transfected

0

0.2

0.4

0.6

0.81.0

1.2

1.4

1.6

1.8

0.125 0.5 0.125 0.5

E5-WTE5-LQR

**

**

No

rmal

ized

Exc

itat

ory

Syn

apse

Den

sity

N=

45

N=

40

N=

48

N=

45

Cu

mu

lati

ve F

ract

ion 1.0

0.8

0.6

0.4

0.2

0.0

1.0

0.8

0.6

0.4

0.2

0.01000050000

Inter-event Interval (ms)

Cu

mu

lati

ve F

ract

ion 1.0

0.8

0.6

0.4

0.2

0.0

1.0

0.8

0.6

0.4

0.2

0.06040200

Amplitude (pA)

Ctrl-shRNA

E5-shRNA

GFP

GFPSp

ine

Den

sity

(#Sp

ines

/10

μm)

0

1

2

3

4

5

6

7

8 **

N=

35

N=

38

10ng

E5-shRNACtrl-shRNA

C

0

50

100

150

200

250

300

00 0.05 0.1 0.25 0.5 0.75

μg of E5-Myc transfected

1.00

1.25

0.50

0.75

0.25

1.50

E5 in

tens

ity

(fol

d in

crea

se)

No

rmal

ized

Exc

itat

ory

Syn

apse

Den

sity **

**GFP

GFP/E5

0.25 μg of E5

Inte

r-ev

ent

inte

rval

(ms)

Am

plit

ud

e (p

A)

0100200

300400500600

700800900 ***

0

5

10

15

20

25

30 *

E5-shRNACtrl-shRNA N=12

N=14

1 sec10 pA

80 100

Figure 3. Ephexin5 Negatively Regulates Excitatory Synapse Number

(A) 10 ng of E5-shRNA or Ctrl-shRNA was cotransfected with GFP into rat hippocampal neurons at DIV14. At DIV18 dendritic spines were measured as described

in methods. Representative image illustrates dendritic spines. N indicates number of neurons assessed. Error bars indicate ± SEM; **p < 0.01, ANOVA.

(B) 10 ng or 20 ng of two different E5-shRNA or Ctrl-shRNA constructs were cotransfected with GFP into rat hippocampal neurons at DIV10. At DIV14 excitatory

synapses were measured as described in methods. Representative image illustrates quantified synapse puncta (white). Error bars indicate ± SEM; **p < 0.01,

***p < 0.005, ANOVA.

(C) DIV10 rat hippocampal neurons were cotransfected with GFP and increasing concentrations of E5-Myc or control plasmid. At DIV 14 excitatory synapses

(gray bars) and exogenous E5 expression (blue bars) were measured as described in methods. Representative image illustrates localization of E5-Myc on

transfected neuron (red). Error bars indicate ± SEM; **p < 0.01, ANOVA.

(D) Neurons were transfected with E5-Myc (E5-WT) or E5-LQR-Myc (E5-LQR) and quantified as in (C). Error bars indicate ± SEM; **p < 0.01, ANOVA.

(E) Quantification of mEPSC inter-event interval and amplitude from hippocampal neurons transfected as in (B) with 20 ng of shRNA. Cumulative distribution plots,

bar graphs and representative traces are shown. Error bars represent the standard deviation of the mean, ***p < 0.005, *p < 0.05.

See also Figure S3 and Figure S1.

446 Cell 143, 442–455, October 29, 2010 ª2010 Elsevier Inc.

Page 135: Cell 101029

shown). Organotypic slices were prepared from the hippo-

campus of the E5fl/fl mice or their wild-type littermates. Using

the biolistic transfection method, a plasmid expressing Cre re-

combinase was introduced into a low percentage of neurons in

the slices. We found that introduction of a Cre-expressing

plasmid into E5fl/fl neurons in the hippocampal slice led to a

significant increase in the density of dendritic spines present

on the Cre-expressing neurons relative to wild-type hippo-

campal slices transfected with Cre (Figure 4B). The length and

width of dendritic spines analyzed in these experiments showed

no significant difference between wild-type and E5�/� neurons

(Figure S4). Thus, elimination of E5 expression in neurons in

the context of an intact neuronal circuit leads to an increase in

the number of dendritic spines.

To assess the role of E5 in hippocampal circuit development

in vivo, we performed acute slice physiology experiments in

the CA1 region of the hippocampus from wild-type or E5�/�

mice. We find that relative to wild-type neurons, in E5�/� CA1

pyramidal neurons there are more frequent excitatory events

that have larger amplitude (Figure 4C). A possible explanation

for these findings is that when E5 function is disrupted during

in vivo development more excitatory synapses form resulting in

more excitatory postsynaptic events. To test this possibility,

we used array tomography to quantify the number of excitatory

synapses that form in the CA1 stratum radiatum of wild-type

and E5�/� mice. We observed a �2-fold increase in the number

of excitatory synapses within the CA1 region of the E5�/� hippo-

campus compared to wild-type mice (Figure 4D). Specifically,

the number of juxtaposed synapsin and PSD-95 puncta was

quantified and considered a measurement of the number of

excitatory synapses that form within the CA1 region of the hippo-

campus in vivo. This analysis revealed a significant increase in

the number of PSD-95 puncta but no change in the number of

synapsin puncta density (Figure 4D). This suggests that the

increase in excitatory synapse number in the stratum radiatum

of E5�/� mice is likely due to the absence of E5 postsynaptically

and that when E5 is present within dendrites it functions to nega-

tively regulate synapse number in vivo. On the basis of these

results, we conclude that a key function of E5 is to restrict excit-

atory synapse number during the development of neuronal

circuits.

Ephexin5 Restricts EphB2 Control of ExcitatorySynapse FormationWe next considered the possibility that the ability of E5 to

restrict excitatory synapse number might be controlled by

EphB2 signaling. To test this idea, we asked whether reducing

EphB2 signaling eliminates the increase in excitatory synapse

number detected when E5 levels are knocked down by expres-

sion of E5-shRNA. To block EphB2 activation, we introduced

into neurons a kinase dead version of EphB2 (EphB2-KD) which

has been previously shown to block EphB2 signaling (Dalva

et al., 2000). As described above, expression of E5-shRNA in

neurons leads to a significant increase in the number of

synapses that are present on the E5-shRNA-expressing neuron.

However, this increase was reversed if the E5-shRNA was

cotransfected with a plasmid that drives expression of EphB2-

KD, but was not affected by cotransfection of a control plasmid

(Figure 4E). These findings suggest that the increase in excit-

atory synapse number that occurs when E5 levels are reduced

requires EphB signaling. Consistent with this conclusion, we

find that if we overexpress wild-type EphB2 in neurons more

synapses are present on the EphB-expressing neuron.

However, this effect is reduced if E5 is overexpressed in

neurons together with EphB (Figure 4F). It is possible that the

ability of overexpressed E5 to suppress the synapse-promoting

effect of EphB2 reflects independent actions of these two

signaling molecules. However, given that EphB2 and E5 interact

with one another in neurons, the most likely interpretation of

these results is that E5 functions directly to restrict the

synapse-promoting effects of EphB2. If this were the case, we

would predict that for EphB2 to positively regulate excitatory

synapse development it would be necessary to inactivate and/

or degrade E5.

EphB Mediates Phosphorylation of Ephexin5at Tyrosine-361We considered the possibility that since EphB2 is a tyrosine

kinase it might inhibit the GEF activity or expression of the E5

protein by catalyzing the tyrosine phosphorylation of E5. In sup-

port of this possibility, stimulation of dissociated mouse hippo-

campal neurons with EphrinB1 (EB1) for 15 min led to an

increase in the level of E5 tyrosine phosphorylation as detected

by probing immunoprecipitated E5 with the pan-anti-phospho-

tyrosine antibody, 4G10 (Figure 5A).

We have previously shown that EphrinA1 stimulation of

cultured neurons leads to the tyrosine phosphorylation of E1 at

tyrosine 87 (Sahin et al., 2005). On the basis of this finding we

hypothesized that exposure of neurons to EB1 might promote

the phosphorylation of the analogous tyrosine residue (Y361)

on E5 (Figure 5B) and that phosphorylation at this site might

lead to E5 inactivation. To address this possibility, we overex-

pressed EphB2 in 293 cells together with wild-type E5 or

a mutant form of E5 in which Y361 is converted to a phenylalanine

(E5-Y361F). Lysates were prepared from the transfected cells

and after SDS-PAGE were immunoblotted with 4G10 (Figure 5C).

We found that in the presence of EphB2, E5-WT, but not

E5-Y361F, becomes tyrosine phosphorylated. These findings

suggest that EphB2 catalyzes the tyrosine phosphorylation of

E5 primarily at Y361.

To show definitively that E5 Y361 is tyrosine phosphorylated,

we generated an E5 phospho-Y361 antibody (a-pY361). To

demonstrate that this antibody specifically recognizes the

Y361-phosphorylated form of E5, we immunoblotted cell lysates

prepared from 293 cells that express EphB2 and either E5-WT or

E5-Y361F with a-pY361. This analysis demonstrated that

a-pY361 bind to wild-type E5 but not E5-Y361F (Figure 5C).

Furthermore, using a-pY361 we found that when wild-type

EphB2, but not a kinase dead or cytoplasmic truncated version

of EphB2, is expressed in 293 cells together with E5, E5

becomes phosphorylated at Y361 (Figure S5A). In contrast,

when EphA4 or EphA2 were expressed in 293 cells we detected

little to no phosphorylation of E5 at Y361 (Figure S5B). These

findings suggest that EphB2, but not EphAs, promote E5 Y361

phosphorylation (pY361).

Cell 143, 442–455, October 29, 2010 ª2010 Elsevier Inc. 447

Page 136: Cell 101029

D

n.s.

PSD-95Synapsin1

No

rmal

ized

Pu

nct

a D

ensi

ty

A

0

1

2

3

4

5

6

7

Inte

r-ev

ent

Inte

rval

(s)

Am

plit

ud

e (p

A)

0

2

4

6

8

10

12

14

16

N=

12

N=

12

*** *

CWTE5 -/-

B

0

0.5

1.0

1.5

2.0

2.5

3.0

3.5GFP

GFP

No

rmal

ized

Exc

itat

ory

Syn

apse

Den

sity

N=

55

N=

55

***

E5E5

+/-

-/-WT + CreE5 + Crefl/fl

0

20

40

60

80

100

120

140

Spin

e D

ensi

ty (#

Spin

es/1

00 μ

m) ***

N>

3000

sp

ines

/co

nd

itio

n

WT E5 -/-

F

*****

**

N=

30

N=

30

N=

30

N=

30

EphB2 - + - +0

0.20.40.60.81.01.21.41.61.82.0

E5-MycCtrl

E

No

rmal

ized

Ex

cita

tory

Syn

apse

Den

sity **

*****

N=

30

N=

30

N=

30

N=

300.20.40.60.81.01.21.41.61.8

EphB2KD - + - +0

E5-shRNACtrl-shRNA

WTE5 -/-

CA1 CA1

PSD95/SynapsinPSD95/Synapsin

Synapsin1Synapsin1

PSD95PSD95

0

0.5

1

1.5

2

2.5

0

0.2

0.4

0.6

0.8

1

1.2

0

0.5

1

1.5

2

2.5

3

3.5Synapses

No

rmal

ized

Pu

nct

a D

ensi

ty

No

rmal

ized

Pu

nct

a D

ensi

ty

**

No

rmal

ized

Ex

cita

tory

Syn

apse

Den

sity

WTE5 -/-

WTE5 -/-

E5 + Crefl/fl

WT + Cre

Figure 4. Ephexin5 Restricts EphB2 Control of Excitatory Synapse Formation

(A) E16 hippocampi from E5+/� or E5�/� mice were dissected and dissociated for culture. At DIV10 dissociated neurons were transfected with GFP. At DIV14

neurons were fixed, stained, and excitatory synapses were measured as described in methods. Error bars indicate ± SEM; ***p < 0.005, ANOVA.

(B) Organotypic slices from WT or E5fl/fl mice were biolistically transfected with Cre-recombinase (Cre) and dendritic spines were quantified as described in

methods. Representative images are shown (left). Error bars indicate ± SEM; ***p < 0.005, KS test.

(C) Quantification of mEPSC inter-event interval and amplitude from acute hippocampal brain slices prepared from P12-P14 WT or E5�/� mice. Error bars repre-

sent the standard deviation of the mean; ***p < 0.005, *p < 0.05.

(D) Hippocampi from three independent littermate pairs consisting of P12 WT and E5�/� mice were prepared as described in methods for quantification of

synapses, Synapsin1 and PSD-95 using array tomography. Error bars ± SEM; *p < 0.05, Mann-Whitney U-Test.

(E) Increase in excitatory synapse number following loss of E5 requires EphB2 signaling. At DIV10, control plasmid (�) or EphB2KD plasmid (+) were coexpressed

in dissociated mouse hippocampal neurons with GFP and either Ctrl-shRNA or E5-shRNA. At DIV14 excitatory synapses were measured as described in

methods. Error bars indicate ± SEM; **p < 0.01, ***p < 0.005, ANOVA.

448 Cell 143, 442–455, October 29, 2010 ª2010 Elsevier Inc.

Page 137: Cell 101029

We also found by immunoblotting with a-pY361 that E5

is phosphorylated at Y361 in the hippocampus of wild-type but

not E5�/� mice (Figure S5C), and that EB1 stimulation of cultured

hippocampal neurons leads to E5 Y361 phosphorylation (Fig-

ure 5D). By immunofluorescence microscopy we detect

punctate a-pY361 staining along the dendrites of EB1-treated

wild-type neurons, but less staining in untreated neurons (Fig-

ure 5E). This result suggests that E5 becomes newly phosphor-

ylated at Y361 upon exposure of hippocampal neurons to EB1.

EphB2-Mediated Degradation of Ephexin5Is Kinase- and Proteasome DependentWe asked if EB1 stimulation of E5 Y361 phosphorylation leads

to a change in E5 activity or expression. To investigate this

possibility we asked if EphB suppresses E5-dependent RhoA

activation in a phosphorylation-dependent manner. We trans-

fected 293 cells with E5 in the presence or absence of EphB2

and measured RhoA activity using the RBD pull-down assay

(Figure 5F). We found that E5-dependent RhoA activation was

reduced in 293 cells expressing EphB2 and E5 compared to cells

expressing E5 alone. These findings are consistent with the

possibility that EphB2-mediated tyrosine phosphorylation of E5

either leads to a suppression of E5’s ability to activate RhoA,

or alternatively might trigger a decrease in E5 protein expression

resulting in a decrease in RhoA activation. We found this latter

possibility to be the case (Figure 5F, E5 loading control). Further-

more, when we compared lysates from the brains of wild-type or

EphB2�/� mice, we observed that E5 phosphorylation at Y361 is

decreased while the levels of E5 expression are increased in the

lysates from EphB2�/� mice (Figure 5G). These data suggest that

EphB2 functions to phosphorylate and degrade E5.

Consistent with the idea that E5 expression is destabilized in

the presence of EphB, we observed that in the dendrites of

cultured hippocampal neurons overexpressing EphB2, endoge-

nous E5 expression levels are reduced compared to control

transfected neurons or neurons transfected with a kinase dead

version of EphB2 (Figures S6A and S6B). When neurons were

exposed to EB1 compared to EA1 for 60 min, we found by immu-

noblotting of neuronal extracts, or immunofluorescence staining

with a-N-E5, that exposure to EB1 leads to a decrease in E5

expression (Figure 6A). The lack of complete loss of E5 expres-

sion by Western blot may be due to the fact that EB1 stimulation

leads to dendritic and not somatic loss of E5 expression. More-

over, immunofluorescence staining revealed a loss of E5 puncta

specifically within the dendrites of EB1-stimulated neurons,

consistent with the possibility that EB1/EphB-mediated degra-

dation of E5 relieves an inhibitory constraint that suppresses

excitatory synapse formation on dendrites (Figure 6A). In support

of this idea, we find by immunoblotting of extracts from mouse

hippocampi that endogenous E5 protein levels are highest at

postnatal day 3 prior to the time of maximal synapse formation

and then decrease as synapse formation peaks in the postnatal

period (Figure S6C). Northern blotting revealed that this

decrease in E5 protein is not due to a change in the level of E5

mRNA expression (Figure S6C). Given that E5 protein levels

decrease dramatically during the time period P7-P21 when

synapse formation is maximal, these findings suggest that E5

may need to be degraded prior to synapse formation.

We asked whether EphB-mediated degradation of E5 could be

reconstituted in heterologous cells. When EphB and Myc-tagged

E5 were coexpressed in 293 cells we observed a significant

decrease in E5 protein expression in the presence of EphB2

(Figure 6B). The presence of EphB2 had no effect on the level of

expression of a related GEF, E1 (Figure 6B). We asked whether

EphB-mediated degradation of E5 depends upon Y361 phos-

phorylation. We found that in 293 cells overexpressing Myc-

tagged E5, the coexpression of EphB2, but not EphB2-KD,

resulted in a significant decrease in E5 levels (Figure 6C). This

suggests that EphB tyrosine kinase activity is required for E5

degradation. The EphB-mediated reduction in E5 levels is depen-

dent on Y361 phosphorylation, as EphB2 expression had no effect

on the level of E5 Y361F expression (Figure 6D). This suggests that

the phosphorylation of E5 at Y361 triggers E5 degradation.

We considered the possibility that the Y361 phosphorylation-

dependent decrease in E5 protein levels might be due to

EphB-dependent stimulation of E5 proteasomal degradation.

Consistent with this possibility we found that addition of the

proteasome inhibitor lactacystin to 293 cells leads to a reversal

of the EphB-dependent decrease in E5 protein levels, as

measured by an increase in total ubiquitinated E5 (Figure S6D).

In addition, in neuronal cultures the EB1 induced decrease in

E5 protein expression is blocked if the proteasome inhibitor lac-

tacystin is added prior to EB1 addition (Figure 6E). Notably, in the

presence of lactacystin, E5 is ubiquitinated, further supporting

the idea that E5 is degraded by the proteasome.

To test whether E5 is ubiquitinated in the brain, we incubated

wild-type or E5�/�brain lysates with a-C-E5 and after immuno-

precipitation and SDS-PAGE, probed with a-ubiquitin anti-

bodies. This analysis detected the presence of ubiquitinated

species in a-C-E5 immunoprecipitates prepared from wild-type

but not E5�/� brain lysates (Figure 6F). These findings indicate

that E5 is ubiquitinated in the brain.

EphB2-Mediated Degradation of Ephexin5Requires Ube3ADuring proteasome-dependent degradation of proteins, speci-

ficity is conferred by E3 ligases or E2 conjugating enzymes

that recognize the substrate to be degraded. The E3 ligase binds

to the substrate and catalyzes the addition of polyubiquitin side

chains to the substrate thereby promoting degradation via the

proteasome (Hershko and Ciechanover, 1998). We considered

several E3 ligases that have recently been implicated in synapse

development as candidates that catalyze E5 degradation. One of

these E3 ligases, Cbl-b, has previously been implicated in the

degradation of EphAs and EphBs (Fasen et al., 2008; Sharfe

et al., 2003). A second E3 ligase, Ube3A, has been shown to

(F) E5 can suppress an EphB2-mediated increase in excitatory synapse number. At DIV10, control plasmid (�) or EphB2-expressing plasmid (+) were coex-

pressed in dissociated mouse hippocampal neurons with GFP and either control (Ctrl) plasmid or E5-Myc plasmid. At DIV14 excitatory synapses were measured

as described in methods. Error bars indicate ± SEM; **p < 0.01, ***p < 0.005, ANOVA.

See also Figure S4 and Figure S1.

Cell 143, 442–455, October 29, 2010 ª2010 Elsevier Inc. 449

Page 138: Cell 101029

A

E

G

F

B

C

D

Figure 5. EphB2 Mediates Phosphorylation of Ephexin5 at Tyrosine-361

(A) Dissociated mouse hippocampal neurons were stimulated with either a -Fc IgG (Ctrl) or preclustered Fc-EB1 for 15 min. Neuronal lysates were immunopre-

cipitated with a-N-E5, followed by immunoblotting for panphosphotyrosine (a-pTyr) or E5 with a-N-E5. EB1 stimulation was determined by immunoblotting

neuronal lysates for phospho-Eph (pEph). Input protein levels and a-Actin loading control are shown (bottom).

(B) E5-Y361 is a conserved residue with E1-Y87 (Sahin et al., 2005).

(C) Immunoprecipitation with a-Myc from 293 cell lysates previously transfected with various combinations of overexpressing plasmids containing E5-Myc, E5

(Y361F)-Myc and/or EphB2-Flag, followed by immunoblotting with a-pTyr, a-Myc, a-pY361 or a-Flag. Input EphB2 levels are shown (bottom).

(D) Neurons were treated and lysates prepared as in panel (A) followed by immunoblotting with a-pY361 or a-N-E5. Representative immunoblot with input phos-

pho-Eph (pEph) levels is shown (top). Quantification of three independent experiments is shown as a percent increase in pY361 over Ctrl stimulation (bottom).

Error bars indicate ± SEM; *p < 0.05.

(E) Dissociated rat hippocampal neurons were transfected with GFP (gray) and stimulated as in panel (A), followed by fixing and staining for endogenous phos-

phorylated E5 using a-pY361 (Red). Representative image shown (left). White rectangle outlines magnified dendritic region showing examples of phospho-E5

staining (left bottom). Four independent experiments were imaged and analyzed for pY361 (bar graph). Error bars indicate ± SEM; *p < 0.05.

(F) Lysates from 293 cells transfected with empty vector (-) or increasing concentrations of E5-Myc with or without Flag-EphB2 were assessed for endogenous

RhoA activity by RBD assay (previously described). GTPgS lane is a positive control for inducing RhoA. Input protein levels and a-Actin loading control are shown

(bottom).

(G) WT and EphB2�/� (B2�/�) brain lysates were immunoblotted with a-EphB2, a-N-E5, a-Actin, or a-pY361 according to methods (left). Quantification of a-N-E5

or a-pY361 signal from three independent experiments is normalized to a-Actin and represented as fold change compared to wild-type. Error bars indicate ±

SEM; *p < 0.05.

See also Figure S5.

450 Cell 143, 442–455, October 29, 2010 ª2010 Elsevier Inc.

Page 139: Cell 101029

regulate synapse number. To determine if Ube3A and/or Cbl-b

catalyze E5 degradation we first asked if either of these E3

ligases interacts with and degrades E5 in 293 cells. When these

E3 ligases were epitope-tagged and expressed in 293 cells

together with E5 we found that E5 coimmunoprecipitates with

Ube3A but not with Cbl-b (Figure 7A). The coimmunoprecipita-

tion of Ube3A with E5 was specific in that Ube3A was not

coimmunoprecipitated with two other neuronal proteins, E1 or

the transcription factor MEF2. In a previous study we have

shown that Ube3A binds to substrates via a Ube3A binding

domain (hereafter referred to as UBD [Greer et al., 2010]). Using

protein sequence alignment programs, ClustalW and ModBase,

we identified a UBD in E5, providing further support for the idea

that E5 might be a substrate of Ube3A (Figure S7A). Consistent

with this hypothesis, we found that the level of E5 expression

is reduced in 293 cells cotransfected with titrating amounts of

Ube3A compared to cells cotransfected with titrating amounts

of Cbl-b (Figure S7B).

We asked if EB1/EphB-mediated E5 degradation in neurons is

catalyzed by Ube3A. To inhibit Ube3A activity we introduced into

neurons a dominant interfering form of Ube3A (dnUbe3A) that

contains a mutation in the ubiquitin ligase domain rendering

Ube3A inactive. We have previously shown that even though

dnUbe3A is catalytically inactive it still binds to E2 ligases and

to its substrates and functions in a dominant negative manner

to block the ability of wild-type Ube3A to ubiquitinate its

substrates (Greer et al., 2010). We found that when introduced

into 293 cells dnUbe3A binds to E5 (Figure 7A). We also found

by immunofluorescence microscopy that when overexpressed

in neurons, dnUbe3A blocks EB1/EphB stimulation of E5 degra-

dation (Figure 7B). EB1/EphB stimulation of E5 degradation was

also attenuated when Ube3A expression was knocked down by

a shRNA that specifically targets the Ube3A mRNA (Figure 7B;

Greer et al., 2010). Notably, the presence of the dnUbe3A did

not affect E5 expression in neurons in the absence of EphrinB

stimulation, suggesting that EphrinB stimulation of E5 Y361

phosphorylation may be required for Ube3A-mediated degrada-

tion of E5 (Figure S7C).

To determine if Ube3A-dependent degradation of E5 might be

relevant to the etiology of Angelman syndrome we asked if the

absence of Ube3A in a mouse model of Angelman syndrome

affects the level of E5 expression in the brain. We compared

A

B D

F

C

E

Figure 6. EphB2-Mediated Degradation of Ephexin5 Is Kinase- and

Proteasome Dependent

(A) Dissociated mouse hippocampal neurons were incubated with preclustered

Fc, Fc-EB1 or Fc-EA1 for 60 min, lysed, and immunoprecipitated with a-C-E5

followed by immunoblotting with a-N-E5. Immunoblot of input with a-pEph or

a-Actin (loading control) are shown. Western is one representative image, and

quantification is of three separate experiments with samples normalized to

a-Actin (left). Error bars indicate ± SEM; *p < 0.05. Right, dissociated mouse

hippocampal neurons were transfected with GFP (gray) and stimulated with

either preclustered Fc (Ctrl) or Fc-EB1 (EB1) for 30 min, followed by fixing

and staining for endogenous E5 using a-N-E5 (red). White rectangle outlines

magnified dendritic region showing examples of E5 staining (right).

(B) Lysates from 293 cells previously transfected with various combinations of

overexpressing plasmids containing E5-Myc, E1-Myc and/or Flag-EphB2 were

immunoblotted with a-Myc, a-Flag, or a-Actin (loading control).

(C) Lysates from 293 cells previously transfected with various combinations of

overexpressing plasmids containing Flag-EphB2, Flag-EphB2KD and/or

E5-Myc were immunoblotted with a-Myc, a-Flag, or a-Actin (loading control).

(D) Lysates from 293 cells previously transfected with various combinations of

overexpressing plasmids containing E5-Myc, E5-Y361F-Myc and/or Flag-

EphB2 were immunoblotted with a-Myc, a-Flag, or a-Actin (loading control).

Representative immunoblot is shown (top). From three independent experi-

ments E5 levels were quantified and normalized to E5 expression in absence

of EphB2-Flag (bottom). Error bars indicate ± SEM; **p < 0.01.

(E) Dissociated mouse hippocampal neurons transfected with GFP (gray) were

stimulated similar to (B) in the absence or presence of lactacystin and immuno-

stained with a-N-E5. White rectangle outlines magnified dendritic region

showing examples of E5 staining (right).

(F) WT and E5�/� brains were lysed and immunoprecipitated with a-C-E5

followed by immunoblotting with a-ub or a-N-E5.

See also Figure S6.

Cell 143, 442–455, October 29, 2010 ª2010 Elsevier Inc. 451

Page 140: Cell 101029

the level of E5 protein expression in the brains of wild-type mice

to that expressed in the brains of mice in which the maternally in-

herited Ube3A was disrupted (Ube3Am-/p+). Because the pater-

nally inherited copy of Ube3A is silenced in the brain due to

imprinting, the level of Ube3A expression in Ube3Am-/p+ neurons

is very low. We found that the level of E5 expression in the brains

of Ube3Am-/p+ mice was significantly higher than that detected in

the brains of wild-type mice (Figure 7C). Moreover, the level of

ubiquitinated E5 in brains of Ube3Am-/p+ mice was significantly

reduced compared to the brains of litter mate controls

(Figure 7D). In addition we found that when neurons from wild-

A

HA-Ube3A

E5-MycE1-MycHA-DNUbe3AHA-MEF2AHA-Cbl-b

-

-++--

-

+--+-

-

+---+

-

+-+--

+

+----

-

+----

E5-Myc

E1-Myc

HA-Ube3AHA-Cbl-b

HA- MEF2A

Input

IB: α-Myc

IB: α-HA

α-Actin

α-Myc

IP: α-HA

DInput

WT

α-Ube3A

α-N-E5

WT

IB: α-Ub

Ub

-E5

IB:α-N-E5

250 kD

150 kD

100 kDE5

250 kD

150 kD

100 kD

IP:α-C-E5

0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

No

rmal

ized

E5

stai

nin

g in

ten

sity

EB1Ube3A shRNA

DNUbe3A

---

+--

+-+

++-

B

n=

48

n=

45

n=

40

n=

45

**

n.s.

E

0

100

200

300

400

500

600

700

Eph

exin

5 st

ain

ing

Inte

nsi

ty

Fc EB1EB1 Fc

Ube3Am-/p+

WT

n=

35

n=

35

n=

35

n=

35

**

**Ube3A

m-/p+Ube3Am-/p+

WT

α-N-E5

α-MEF2

α-EphB2

α-Actin

α-Ube3A0

0.5

1

1.5

2

2.5

3

EphB2 MEF2 E5P3 P8

Nor

mal

ized

E5

to α

-Act

in

P3

C

*n.s.

*

Ube3Am-/p+

WT

Ube3Am-/p+

Figure 7. EphB2-Mediated Degradation of

Ephexin5 Requires Ube3A

(A) Immunoprecipitation with a-HA from 293 cell

lysates previously transfected with various combi-

nations of plasmids containing E1-Myc, E5-Myc,

HA-DNUbe3A, HA-MEF2A, HA-Cbl-b, and/or HA-

Ube3A, followed by immunoblotting with a-HA or

a-Myc. Input protein levels and a-Actin loading

control are shown (bottom).

(B) Hippocampal mouse neurons were cotrans-

fected with GFP and control, HA-DNUbe3A or

Ube3A-shRNA at DIV10. At DIV14, neurons were

incubated with clustered Fc (�) or Fc-EB1 (+) for

30 min. Neurons were fixed and stained for E5

with a-N-E5 and quantified as described in the

methods. Quantification is of E5 staining intensity

normalized to Fc control. Error bars ± SEM; **p <

0.01, ANOVA.

(C) Ube3A wild-type and maternal-deficient

(Ube3Am-/p+) mouse brains were lysed and immu-

noblotted with a-N-E5, a-EphB2, a-MEF2, a-Actin

(loading control), or a-Ube3A (left). Samples were

normalized to a-Actin and quantified as described

in methods (right). Error bars indicate ± SEM; *p <

0.05, Mann-Whitney.

(D) Brain lysates from WT and Ube3Am-/p+ were

collected and treated similar to (C), immunoprecip-

itated with a-C-E5 and immunoblotted with a-N-E5

and a-ub. Input protein levels are shown (right).

(E) Neurons from WT and Ube3Am-/p+ mice were

dissociated, cultured and transfected with GFP at

DIV10. At DIV14, neurons were incubated with pre-

clustered Fc or Fc-EB1 for 30 min. Neurons were

fixed and stained for E5 with a-N-E5 and quantified

according to methods. Error bars indicate ± SEM;

**p < 0.01.

See also Figure S7.

type and Ube3Am-/p+ brains were cultured

and then treated with EB1 the level of E5

protein was reduced upon EB1 treatment

in wild-type but not in Ube3Am-/p+

neurons (Figure 7E). Taken together,

these findings suggest that in response

to EB1 treatment E5 is tyrosine phosphor-

ylated by an EphB-dependent mecha-

nism, and that this leads to E5 degrada-

tion by a Ube3A-dependent mechanism.

If E5 degradation is disrupted due to

a loss of Ube3A as occurs in Angelman

syndrome the result is an increase in E5 expression and a disrup-

tion of the proper control of excitatory synapse number during

brain development.

DISCUSSION

Previous studies have revealed a role for EphrinB/EphB signaling

in the development of excitatory synapses (Klein, 2009).

However, the regulatory constraints that temper EphB-depen-

dent synapse development so that excitatory synapses form at

the right time and place, and in the correct number were not

452 Cell 143, 442–455, October 29, 2010 ª2010 Elsevier Inc.

Page 141: Cell 101029

known. In this study we identify a RhoA GEF, E5, which functions

to restrict EphB-dependent excitatory synapse development. E5

interacts with EphB prior to EphrinB binding, and by activating

RhoA serves to inhibit synapse development. The binding of

EphrinB to EphB as synapses form triggers the phosphorylation

and degradation of E5 by a Ube3A-dependent mechanism. The

reduction in E5 expression may allow EphB to promote excit-

atory synapse development by activating Rac and other proteins

at the synapse.

The findings that E5 functions to restrict excitatory synapse

number suggests that, even though EphBs promote excitatory

synapse development, there are constraints on the activity of

EphB so that synapse number is effectively controlled. There

are several steps in the process of synapse development where

E5 may function to restrict synapse number. One possibility is

that E5 functions early in development as a barrier to excitatory

synapse formation by activating RhoA and restricting the motility

or growth of dendritic filopodia that are the sites of contact by the

presynaptic neuron. For example, by inhibiting dendritic filopo-

dia formation or motility, E5 may decrease the number of

contacts the filopodia make with the presynaptic neuron, thus

resulting in the formation of fewer synapses. An alternative

possibility is that E5 functions to restrict synapse number later

in development perhaps to counterbalance the positive effects

of EphB on Rac that promote dendritic spine development. An

additional possibility is that E5 functions after excitatory synapse

development as a regulator of synapse elimination.

Our analyses of E5 function are most consistent with the

possibility that E5 functions early in the process of synapse

development. First, we find that E5 is expressed, active, and

bound to EphB prior to synapse formation. Second, the interac-

tion of EphrinB with EphB, a process that is thought to be an early

step in excitatory synapse development, triggers the degrada-

tion of E5. Third, our preliminary time-lapse imaging studies

suggest that E5 is localized to newly formed filopodia prior to

synapse development where it appears to restrict filopodia

motility and growth (Margolis et al. unpublished). Thus, E5 might

function as an initial barrier to synapse formation until it is

degraded upon EphrinB binding to EphB.

It is possible that through its interaction with EphB, E5 marks

the sites where synapses will form, and that the degradation of

E5 is a critical early step in excitatory synapse development.

While the mechanisms by which E5 is degraded are not fully

understood, our studies suggest that the phosphorylation of

the N-terminus of E5 at Y361 triggers the Ube3A-mediated pro-

teasomal degradation of E5. One possibility is that prior to pY361

the N- and C-terminal portions of E5 interact, thereby protecting

E5 from degradation. The phosphorylation of E5 at Y361 may

relieve this inhibitory constraint allowing for E5 ubiquitination

and degradation. A similar mechanism has been shown to

regulate the activation of the Rac GEF Vav, (Aghazadeh et al.,

2000)). During EphrinA/EphA signaling it has been proposed

that Vav-mediated endocytosis of the EphrinA/EphA complex

may allow the conversion of the initial adhesive interaction

between EphrinA and EphA-expressing cells into a repulsive

interaction that results in growth cone collapse and axon repul-

sion. It is possible that E5 has a related function during EphB

signaling at synapses. Typically the EB/EphB interaction is

thought to be repulsive. This has been documented in studies

of EphB’s role in the process of axon guidance (Egea and Klein,

2007; Flanagan and Vanderhaeghen, 1998). However, during

synapse development the EphrinB/EphB interaction is thought

to result in synapse formation, a process that requires an interac-

tion between the developing pre- and postsynaptic specializa-

tion. One possibility is that when EphrinB and EphB mediate

the interaction between the incoming axon and the developing

dendrite, the interaction is facilitated by the degradation of E5

by Ube3A. Since E5 is a RhoA GEF, its presence might initially

lead to repulsion between the incoming axon and the dendrite.

However, the EphB-dependent degradation of E5 might convert

this initial repulsive interaction into an attractive one.

The finding that Ube3A is the ubiquitin ligase that controls

EphB-mediated E5 degradation is of interest given the role of

Ube3A in human cognitive disorders such as Angelman syn-

drome and autism. The absence of Ube3A function in Angelman

syndrome would be predicted to result in an increase in E5

protein expression, and thus a decrease in EphB-dependent

synapse formation. Consistent with this possibility, we find in a

mouse model for Angelman syndrome that the level of E5 protein

expression is elevated and that in response to EphrinB treatment

E5 is not degraded. Likewise, several studies have indicated that

synapse development and function is disrupted in these mice

(Jiang et al., 1998; Yashiro et al., 2009).

The recent finding that the Ube3A gene lies within a region of

chromosome 15 that is sometimes duplicated in autism raises

the possibility that altered levels of Ephexin5 and the resulting

defects in excitatory synapse restriction might also be a mecha-

nism relevant to the etiology of autism. If this is the case, a

possible therapy for treating autism might be to reduce the level

of Ube3A activity, and thus increase the level of Ephexin5 ex-

pression. It is important to consider that in addition to Ephexin5,

Ube3A regulates the abundance of other synaptic proteins.

Nevertheless, the ultimate effect of the aberrant expression of

Ephexin5 and other Ube3A substrates on synapse development

and function will require further study. It seems likely that such

studies will provide further understanding of the development

of human cognitive function and new insights into how this

process goes awry in disorders such as Angelman syndrome

and autism.

EXPERIMENTAL PROCEDURES

DNA Constructs

Details of DNA constructs can be found in Supplemental Information.

Generation of E5�/� Mice

An E5 targeting vector was electroporated into 129 J1 ES cells, and positive

clones were identified by Southern hybridization with two separate probes

(see Supplemental Information).

Antibodies

Details of antibodies can be found in Supplemental Information.

Mice, Cell Culture, Transfections, and Ephrin Stimulations

Ube3Am�/p+ mice were previously described (Greer et al., 2010). EphB2

knockout mice were previously described (Kayser et al., 2008). 293T cells

were cultured in DMEM + 10% FBS and transfected using the calcium phos-

phate method. Organotypic slice cultures were prepared from P6 mouse brains

Cell 143, 442–455, October 29, 2010 ª2010 Elsevier Inc. 453

Page 142: Cell 101029

and biolistically transfected. Acute slices were prepared from P12-14 mice.

Dissociated neurons were cultured in Neurobasal Medium supplemented

with B27 and transfected using the Lipofectamine method. For details on cell

culture, transfections, and Ephrin stimulations, see Supplemental Information.

Cell Lysis, Immunoprecipitations, GEF Pull-Down Assays,

and Western Blots

Whole rat or mouse brains or cultured cells were collected and homogenized in

RIPA buffer. For immunoprecipitations, lysed cells were centrifuged and

supernatants were incubated with appropriate antibody for 2 hr at 4�C, fol-

lowed by addition of Protein-A or Protein-G beads (Santa Cruz Biotechnology)

for 1 hr, and washed three times with ice-cold RIPA buffer. For the a-PY361

detection experiment in 293T cells, samples were boiled in SDS buffer to

disrupt the E5/EphB2 interaction and diluted 1:5 in 1.253 RIPA buffer prior

to immunoprecipitation of E5-Myc. RBD and PBD pull-down assays were

conducted according to the manufacture’s suggestions (Upstate Cell

Signaling Solutions). For details see Supplemental Information.

In Situ Hybridization

To generate probes for in situ hybridization, mouse E5 and EphB2 cDNA were

subcloned into pBluescript II SK (+). Bluescript plasmids containing E5 or

EphB2 cDNA were linearized using the restriction enzyme BssHII. Sense and

antisense probes were generated using DIG RNA labeling mix (Roche) accord-

ing to manufacturer’s instructions. Full-length DIG-labeled probes were

subjected to alkaline hydrolysis as described in Supplemental Information.

Immunocytochemistry

Neurons were paraformaldehyde fixed in PBS. For measuring synapse

density, fixed neurons were incubated with a-PSD-95 and a-Synapsin

antibodies followed by a-Cy3 and a-Cy5 antibodies to visualize the primary

antibodies. For protein colocalization experiments fixed neurons were similarly

treated using a-EphB2 antibodies and a-N-E5 antibodies or a-pY361-E5.

For overexpression studies fixed neurons were incubated using a-Myc or

a-Flag antibodies to visualize overexpressed E5-Myc or EphB2-Flag protein

in the context of the GFP-labeled neurons. For details see Supplemental

Information.

Synapse Assay, Image Analysis, and Quantification

Images were acquired on a Zeiss LSM5 Pascal confocal microscope and spine

and synapse analysis was performed as previously described (see Supple-

mental Information).

Ube3Am�/p+ Cultures

Dissociated hippocampal neurons from Ube3Am�/p+ and wild-type mice were

prepared as previously described (Greer et al., 2010).

Array Tomography

Array tomography was performed as previously described (Micheva and

Smith, 2007) with modifications as described in the Supplemental Information.

Electrophysiology

Electrophysiology was performed using standard methods (see Supplemental

Information).

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures and

seven figures and can be found with this article online at doi:10.1016/j.cell.

2010.09.038.

ACKNOWLEDGMENTS

We thank M. Thompson, Y. Zhou, and H. Ye for assistance in generating mice;

E. Griffith, J. Zieg, S. Cohen, I. Spiegel, M. Andzelm, and the Greenberg lab for

critical discussions. This work was supported by National Institute of Neuro-

logical Disorders and Stroke grant RO1 5R01NS045500 (M.E.G); NRSA

Training grant 5T32AG00222-15 (S.S.M.); and Edward R. and Anne G. Lefler

postdoctoral fellowship (S.S.M.).

Received: November 25, 2009

Revised: July 19, 2010

Accepted: September 23, 2010

Published: October 28, 2010

REFERENCES

Aghazadeh, B., Lowry, W.E., Huang, X.Y., and Rosen, M.K. (2000). Structural

basis for relief of autoinhibition of the Dbl homology domain of proto-oncogene

Vav by tyrosine phosphorylation. Cell 102, 625–633.

Dalva, M.B., Takasu, M.A., Lin, M.Z., Shamah, S.M., Hu, L., Gale, N.W., and

Greenberg, M.E. (2000). EphB receptors interact with NMDA receptors and

regulate excitatory synapse formation. Cell 103, 945–956.

Dalva, M.B., McClelland, A.C., and Kayser, M.S. (2007). Cell adhesion mole-

cules: signalling functions at the synapse. Nat. Rev. Neurosci. 8, 206–220.

Egea, J., and Klein, R. (2007). Bidirectional Eph-ephrin signaling during axon

guidance. Trends Cell Biol. 17, 230–238.

Ethell, I.M., Irie, F., Kalo, M.S., Couchman, J.R., Pasquale, E.B., and

Yamaguchi, Y. (2001). EphB/syndecan-2 signaling in dendritic spine morpho-

genesis. Neuron 31, 1001–1013.

Fasen, K., Cerretti, D.P., and Huynh-Do, U. (2008). Ligand binding induces

Cbl-dependent EphB1 receptor degradation through the lysosomal pathway.

Traffic 9, 251–266.

Flanagan, J.G., and Vanderhaeghen, P. (1998). The ephrins and Eph receptors

in neural development. Annu. Rev. Neurosci. 21, 309–345.

Fu, W.Y., Chen, Y., Sahin, M., Zhao, X.S., Shi, L., Bikoff, J.B., Lai, K.O., Yung,

W.H., Fu, A.K., Greenberg, M.E., et al. (2007). Cdk5 regulates EphA4-mediated

dendritic spine retraction through an E1-dependent mechanism. Nat. Neuro-

sci. 10, 67–76.

Greer, P.L., Hanayama, R., Bloodgood, B.L., Mardinly, A.R., Lipton, D.M.,

Flavell, S.W., Kim, T.K., Griffith, E.C., Waldon, Z., Maehr, R., et al. (2010).

The Angelman Syndrome protein Ube3A regulates synapse development by

ubiquitinating arc. Cell 140, 704–716.

Grunwald, I.C., Korte, M., Wolfer, D., Wilkinson, G.A., Unsicker, K., Lipp, H.P.,

Bonhoeffer, T., and Klein, R. (2001). Kinase-independent requirement of

EphB2 receptors in hippocampal synaptic plasticity. Neuron 32, 1027–1040.

Grunwald, I.C., Korte, M., Adelmann, G., Plueck, A., Kullander, K., Adams,

R.H., Frotscher, M., Bonhoeffer, T., and Klein, R. (2004). Hippocampal

plasticity requires postsynaptic ephrinBs. Nat. Neurosci. 7, 33–40.

Henkemeyer, M., Itkis, O.S., Ngo, M., Hickmott, P.W., and Ethell, I.M. (2003).

Multiple EphB receptor tyrosine kinases shape dendritic spines in the

hippocampus. J. Cell Biol. 163, 1313–1326.

Hershko, A., and Ciechanover, A. (1998). The ubiquitin system. Annu. Rev.

Biochem. 67, 425–479.

Jiang, Y.H., Armstrong, D., Albrecht, U., Atkins, C.M., Noebels, J.L., Eichele,

G., Sweatt, J.D., and Beaudet, A.L. (1998). Mutation of the Angelman ubiquitin

ligase in mice causes increased cytoplasmic p53 and deficits of contextual

learning and long-term potentiation. Neuron 21, 799–811.

Jontes, J.D., Buchanan, J., and Smith, S.J. (2000). Growth cone and dendrite

dynamics in zebrafish embryos: early events in synaptogenesis imaged in vivo.

Nat. Neurosci. 3, 231–237.

Kayser, M.S., McClelland, A.C., Hughes, E.G., and Dalva, M.B. (2006).

Intracellular and trans-synaptic regulation of glutamatergic synaptogenesis

by EphB receptors. J. Neurosci. 26, 12152–12164.

Kayser, M.S., Nolt, M.J., and Dalva, M.B. (2008). EphB receptors couple

dendritic filopodia motility to synapse formation. Neuron 59, 56–69.

Kishino, T., Lalande, M., and Wagstaff, J. (1997). UBE3A/E6-AP mutations

cause Angelman syndrome. Nat. Genet. 15, 70–73.

Klein, R. (2009). Bidirectional modulation of synaptic functions by Eph/ephrin

signaling. Nat. Neurosci. 12, 15–20.

454 Cell 143, 442–455, October 29, 2010 ª2010 Elsevier Inc.

Page 143: Cell 101029

Lai, K.O., and Ip, N.Y. (2009). Synapse development and plasticity: roles of

ephrin/Eph receptor signaling. Curr. Opin. Neurobiol. 19, 275–283.

Lim, B.K., Matsuda, N., and Poo, M.M. (2008). Ephrin-B reverse signaling

promotes structural and functional synaptic maturation in vivo. Nat. Neurosci.

11, 160–169.

Micheva, K.D., and Smith, S.J. (2007). Array tomography: a new tool for

imaging the molecular architecture and ultrastructure of neural circuits.

Neuron 55, 25–36.

Murai, K.K., Nguyen, L.N., Irie, F., Yamaguchi, Y., and Pasquale, E.B. (2003).

Control of hippocampal dendritic spine morphology through ephrin-A3/

EphA4 signaling. Nat. Neurosci. 6, 153–160.

Ogita, H., Kunimoto, S., Kamioka, Y., Sawa, H., Masuda, M., and Mochizuki, N.

(2003). EphA4-mediated Rho activation via Vsm-RhoGEF expressed specifi-

cally in vascular smooth muscle cells. Circ. Res. 93, 23–31.

Penzes, P., Beeser, A., Chernoff, J., Schiller, M.R., Eipper, B.A., Mains, R.E.,

and Huganir, R.L. (2003). Rapid induction of dendritic spine morphogenesis

by trans-synaptic ephrinB-EphB receptor activation of the Rho-GEF kalirin.

Neuron 37, 263–274.

Rossman, K.L., Der, C.J., and Sondek, J. (2005). GEF means go: turning on

RHO GTPases with guanine nucleotide-exchange factors. Nat. Rev. Mol.

Cell Biol. 6, 167–180.

Sahin, M., Greer, P.L., Lin, M.Z., Poucher, H., Eberhart, J., Schmidt, S., Wright,

T.M., Shamah, S.M., O’Connell, S., Cowan, C.W., et al. (2005). Eph-dependent

tyrosine phosphorylation of ephexin1 modulates growth cone collapse.

Neuron 46, 191–204.

Shamah, S.M., Lin, M.Z., Goldberg, J.L., Estrach, S., Sahin, M., Hu, L.,

Bazalakova, M., Neve, R.L., Corfas, G., Debant, A., et al. (2001). EphA

receptors regulate growth cone dynamics through the novel guanine

nucleotide exchange factor ephexin. Cell 105, 233–244.

Sharfe, N., Freywald, A., Toro, A., and Roifman, C.M. (2003). Ephrin-A1

induces c-Cbl phosphorylation and EphA receptor down-regulation in

T cells. J. Immunol. 170, 6024–6032.

Snyder, J.T., Worthylake, D.K., Rossman, K.L., Betts, L., Pruitt, W.M.,

Siderovski, D.P., Der, C.J., and Sondek, J. (2002). Structural basis for the

selective activation of Rho GTPases by Dbl exchange factors. Nat. Struct.

Biol. 9, 468–475.

Tashiro, A., Minden, A., and Yuste, R. (2000). Regulation of dendritic spine

morphology by the rho family of small GTPases: antagonistic roles of Rac

and Rho. Cereb. Cortex 10, 927–938.

Yashiro, K., Riday, T.T., Condon, K.H., Roberts, A.C., Bernardo, D.R., Prakash,

R., Weinberg, R.J., Ehlers, M.D., and Philpot, B.D. (2009). Ube3a is required

for experience-dependent maturation of the neocortex. Nat. Neurosci. 12,

777–783.

Ziv, N.E., and Smith, S.J. (1996). Evidence for a role of dendritic filopodia

in synaptogenesis and spine formation. Neuron 17, 91–102.

Cell 143, 442–455, October 29, 2010 ª2010 Elsevier Inc. 455

Page 144: Cell 101029

Imaging Activity-Dependent Regulationof Neurexin-Neuroligin Interactions Usingtrans-Synaptic Enzymatic BiotinylationAmar Thyagarajan1 and Alice Y. Ting1,*1Department of Chemistry, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02139, USA

*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.09.025

SUMMARY

The functions of trans-synaptic adhesion molecules,such as neurexin and neuroligin, have been difficultto study due to the lack of methods to directly detecttheir binding in living neurons. Here, we use biotinlabeling of intercellular contacts (BLINC), a methodfor imaging protein interactions based on interac-tion-dependent biotinylation of a peptide by E. colibiotin ligase, to visualize neurexin-neuroligin trans-interactions at synapses and study their role insynapse development. We found that both develop-mental maturation and acute synaptic activity stimu-late the growth of neurexin-neuroligin adhesioncomplexes via a combination of neurexin and neuro-ligin surface insertion and internalization arrest. Bothmechanisms require NMDA receptor activity. Wealso discovered that disruption of activity-inducedneurexin-neuroligin complex growth prevents re-cruitment of the AMPA receptor, a hallmark ofmaturesynapses. Our results provide support for neurexin-neuroligin function in synapse maturation and intro-duce a general method to study intercellularprotein-protein interactions.

INTRODUCTION

During brain development, axons grow toward dendrites and

form initial contacts, and the contacts then stabilize, mature,

and differentiate into excitatory or inhibitory synapses (Sudhof,

2008). Both the initial contact and maturation phases of synapse

development are mediated by an assortment of adhesion

proteins, including neurexin, neuroligin, cadherins, and ephrins

(Sudhof and Malenka, 2008). Due to the lack of nonperturbative

methods to detect and study trans-synaptic protein-protein

interactions, however, the timing of these adhesion events, the

size and stability of adhesion complexes, and the relationship

between adhesion events and synaptic properties are largely

unknown.

In this work, we describe a method to image trans-synaptic

protein-protein interactions and use it to study the molecular

mechanisms of synapse development, specifically through the

lens of the neurexin-neuroligin adhesion complex. Both neurexin

and neuroligin are single-pass trans-membrane proteins, and

presynaptic neurexin binds to postsynaptic neuroligin in

a Ca2+-dependent manner with �21 nM affinity (Arac et al.,

2007). Knockout (Varoqueaux et al., 2006) and overexpression

(Chubykin et al., 2007) studies indirectly suggest that the neu-

rexin-neuroligin interaction functions in synapse maturation but

is not crucial for initial synapse formation. Direct evidence for

involvement of the neurexin-neuroligin interaction in synapse

maturation is lacking, however, as are mechanistic details.

Neurexin-neuroligin interactions are most commonly detected

in neurons via gain-of-function or overexpression assays (Chih

et al., 2005; Chubykin et al., 2007), but these methods are non-

physiological and lack specificity due to the numerous alternative

binding partners for both neurexin (Ko et al., 2009; Uemura et al.,

2010) and neuroligin (Xu et al., 2010). Colocalization imagingmay

also be used but has a high false-positive rate because imaging

resolution exceeds protein-protein interaction distances.

The most direct strategy to visualize trans-synaptic protein

binding is GRASP, or ‘‘GFP reconstitution across synaptic part-

ners’’ (Feinberg et al., 2008). Used to detect neuroligin-neuroligin

contacts at synapses of C. elegans, this technique involves

fusion of GFP fragments to the proteins of interest. trans-binding

triggers GFP reconstitution and hence fluorescence onset.

However, because fluorophore maturation takes hours, GFP

recombination is irreversible, and GFP fragments have high

intrinsic affinity that might lead to false positives, GRASP is

better suited to synapse detection and circuit mapping than

minimally invasive study of neurexin-neuroligin interactions and

their dynamics.

RESULTS

BLINC Visualization of trans-SynapticNeurexin-Neuroligin InteractionsTo address the need for new methodology to noninvasively

detect trans-synaptic protein-protein interactions, we turned to

E. coli biotin ligase (BirA) and its 15 amino acid acceptor peptide

(AP) substrate. 35 kD BirA catalyzes the ATP-dependent cova-

lent biotinylation of the central lysine in AP with a kcat of

12 min�1 and Km of 25 mM (Fernandez-Suarez et al., 2008).

Due to the orthogonal specificity of this enzyme-peptide pair in

456 Cell 143, 456–469, October 29, 2010 ª2010 Elsevier Inc.

Page 145: Cell 101029

mammalian cells, we previously used protein fusions to BirA and

a modified form of AP for detection of intracellular protein-

protein interactions via interaction-dependent biotinylation

(Fernandez-Suarez et al., 2008). To image intercellular protein-

protein interactions, we envisioned fusing BirA to neurexin1b

(NRX) and AP to neuroligin1 (NLG), as shown in Figure 1A. Inter-

action-dependent biotinylation would be initiated with the addi-

tion of biotin and ATP or synthetic biotin-AMP ester, which can

be used at lower concentrations to reduce the risk of purine

receptor activation (Howarth et al., 2006). Biotinylated AP would

be detected on the surface of live neurons with fluorophore-

conjugated monovalent streptavidin (mSA), which, unlike wild-

type streptavidin, cannot induce crosslinking (Howarth et al.,

2006). We named this methodology BLINC for ‘‘biotin labeling

of intercellular contacts.’’

The N-terminal ends of both NRX and NLG are extracellular

and face outward from the heterotetrameric complex (Arac

et al., 2007), such that fusion to AP and BirA would not be ex-

pected to disrupt oligomerization. Based on our estimates

from the BirA structure (Weaver et al., 2001) and the length of

AP, the fusion sites must be within �50 A in order to allow BirA

to contact AP. This distance should be easily spanned by the

N-terminal ends of NRX and NLG within the heterotetramer.

We prepared all four fusion constructs: BirA-NRX, AP-NRX,

BirA-NLG, and AP-NLG (Figure S1A available online). We started

with tests in HEK cell cultures and then COS-neuron mixed

cultures and found that BLINC was possible, and site specific,

in these systems (Figures S1B and S1C).

A

B

Figure 1. Imaging Neurexin-Neuroligin

Contacts between Hippocampal Neurons

Using BLINC

(A) Detection scheme. Biotin ligase (BirA) biotiny-

lates proximal acceptor peptide (AP) with biotin-

AMP ester. Ligated biotin is then detected using

AlexaFluor-conjugated monovalent streptavidin

(mSA).

(B) Labeling of contacts between neurons ex-

pressing AP-NLG1 andVenus transfectionmarker,

and neurons expressing BirA-NRX1b and Ceru-

lean transfection marker. Negative controls are

shown with noninteracting mutants of NRX

(second row) and NLG (third row). The red channel

was overlayed on the blue and green channels.

BLINC using reporters with swapped BirA and AP

tags (BirA-NLG1 + AP-NRX1b) is shown in

Figure S2.

See also Figure S1, Figure S3, and Figure S4 for

additional reporter characterization.

To perform BLINC in neuron cultures,

we separately transfected two pools of

suspended neurons immediately after

dissection and dissociation. One pool

expressed BirA-NRX and a fluorescent

protein marker, Cerulean. The other pool

expressed AP-NLG and Venus fluores-

cent protein marker. The two pools of

neurons were then plated together and allowed to form synaptic

contacts over 16 days. At DIV16 (16 days in vitro), cultures were

labeled with biotin-AMP for 15 min and then Alexa568-conju-

gated mSA for 3 min. Images in Figure 1B show Alexa568 signal

(‘‘BLINC signal’’) at sites of BirA-NRX/AP-NLG contact, as indi-

cated by the overlap between Cerulean and Venus markers.

As a measure of the specificity of BLINC labeling, > 97% of all

BLINC puncta were found to overlap with both Venus and Ceru-

lean markers.

To test whether BLINC was specific for NRX-NLG interactions

over neighboring (but noninteracting) NRX and NLG molecules,

we repeated the experiment using noninteracting mutants of

NRX and NLG. We separately confirmed that these mutants still

traffick to synapses (Figure S1A). Figure 1B shows that, with

these mutants, BLINC signal disappears. This is a somewhat

surprising result because for intracellular protein-protein interac-

tions, we previously found that full-length 15 amino acid AP did

not give interaction-dependent biotinylation; rather, we had to

use a shortened AP sequence called AP(�3), with greatly

reduced affinity for BirA (Km > 300 mM) to eliminate interaction-

independent signal (Fernandez-Suarez et al., 2008). By contrast,

the interaction-dependent labeling seen here with full-length AP

(Km 25 mM) may reflect the lower effective concentration of AP at

the synapse with respect to NRX-bound BirA, compared to AP in

the cytosol. We note that, with other fusion constructs, we have

observed weak interaction-independent biotinylation between

contacting cells when the biotinylation time is extended to > 1

hr (data not shown). With the constructs and labeling protocol

Cell 143, 456–469, October 29, 2010 ª2010 Elsevier Inc. 457

Page 146: Cell 101029

described here, however, biotinylation is strictly interaction

dependent.

Figure 1 shows BLINC with the BirA-NRX + AP-NLG reporter

pair. We also tested the other reporter pair with the BirA and

AP tags swapped: AP-NRX + BirA-NLG. Figure S2 shows that

this reporter pair also gives interaction-dependent BLINC signal

and 96% overlap with Venus and Cerulean markers. However,

under identical labeling conditions, mean BLINC intensities at

single puncta were 10-fold lower than with the BirA-NRX + AP-

NLG reporter pair (Figure S4A). We therefore used the latter

pair for nearly all of the experiments in this study.

We performed a panel of control experiments to test the

expression levels of our reporter constructs in neurons and to

examine whether our BirA, AP, or mSA tags affected trafficking

or function. First, we determined that our reporter constructs

are probably expressed at a fraction of the level of their endog-

enous counterparts (Figure S3). Second, by colocalization

analysis, we determined that our tagged NRX and NLG

constructs traffick like previously characterized HA-NLG (Chih

et al., 2005) and HA-NRX (Taniguchi et al., 2007) (Figure S1A).

Third, we used the gain-of-function/overexpression assay to

determine that mSA-labeled AP-NLG has the same ability to

recruit presynaptic VGLUT as HA-NLG (Figures S3D and S3E),

and BirA-NRX has the same ability to recruit postsynaptic

PSD-95 marker as HA-NRX (Figures S3F and S3G).

Characterization of BLINC MethodologyWe wished to determine whether BLINC could differentiate

between larger and smaller NRX-NLG adhesion complexes

and therefore be used to study changes in complex size at

different stages of synapse development. Note that the term

‘‘complex size’’ refers to the number of NRX-NLG interactions

at a synapse, and not the physical dimensions of the adhesion

complex, which we are unable to measure. Previous work has

suggested that overexpression of NRX or NLG may mediate

synaptic effects by artificially enhancing NRX-NLG adhesion

complexes (Graf et al., 2004; Chih et al., 2005). Figure S4B

shows that overexpression of BLINC reporters does increase

BLINC signal at single puncta by 2.5-fold on average, suggesting

that BLINC is semiquantitative.

We compared BLINC to colocalization imaging for detection of

NRX-NLG interactions by transfecting neurons with CFP-NRX

and YFP-NLG alongwith the BLINC reporters. Figure S4C shows

that 95% ± 6% of BLINC puncta overlap with CFP-YFP colocal-

ization sites, whereas only 68% ± 9% of CFP-YFP colocalization

sites overlap with BLINC puncta. Further analysis (Figure S4C)

showed that the mismatch likely results from a high false-posi-

tive rate for the colocalization assay, rather than a high false-

negative rate for BLINC.

We also analyzed the overlap of BLINC signal with synaptic

markers (Figure 2A). In mature DIV16 cultures, we found that

96% ± 5% and 91% ± 4% of BLINC puncta overlapped with

the postsynaptic marker protein Homer and presynaptic marker

protein Bassoon, respectively. 83% ± 5%of BLINC puncta over-

lappedwith FM1-43, a dye that labels recycling presynaptic vesi-

cles. Thus, the majority of BLINC-labeled NRX-NLG interactions

in mature cultures are synaptic.

Neurexin-Neuroligin BLINC Signal Is Correlated withMarkers of Developmental MaturationSynapse maturation is an activity-dependent process during

which synaptic features such as neurotransmitter vesicles and

ion channels assemble, leading to a stronger and more stable

synapse (Garner et al., 2006). If the NRX-NLG interaction is

involved in this process, we would expect a correlation between

NRX-NLG adhesion complex properties (such as size) and

synapse developmental stage. To test this, we simultaneously

imaged NRX-NLG BLINC signal and various markers of synapse

maturation at two different culture ages. Previous studies have

shown that our culturing conditions for hippocampal neurons

allow spontaneous activity that mimics the in vivo developmental

process (Mazzoni et al., 2007). Between DIV5 (‘‘immature’’

cultures) and DIV16 (‘‘mature’’ cultures), for example, dendritic

spines develop, synaptic markers such as Bassoon and Homer

accumulate, glutamate receptors arrive at the postsynaptic

membrane, and synaptic transmission increases significantly

(Kaech and Banker, 2006).

Figure 2B shows that BLINC signal correlates well in mature

DIV16 cultures with presynaptic marker Bassoon, postsynaptic

marker Homer, and FM1-43. The correlation is much poorer in

immature DIV5 cultures for Homer and FM1-43, although

improvement is seen for Homer after cultures are acutely stimu-

lated with high K+ for 1 min to induce synaptic activity. Bassoon

correlation with BLINC signal is high at both DIV5 and DIV16,

perhaps because Bassoon is an early-arriving protein in synapse

development (Friedman et al., 2000).

These observations suggest that NRX-NLG interactions are

linked with synapse maturation and lead to a working mecha-

nistic model for our study. Like Sudhof et al. (Chubykin et al.,

2007), we hypothesize that synaptic activity expands the size of

NRX-NLG adhesion complexes, perhaps via activity-dependent

regulation of NRX and NLG trafficking. The larger NRX-NLG

complexes may then, in turn, promote the recruitment or stabili-

zation of synaptic proteins, perhaps via multivalency or confor-

mational changes, leading to stronger andmore stable synapses.

Synaptic Activity Increases Neurexin-Neuroligin BLINCSignalTo experimentally test the first part of our model—that synaptic

activity expands the size of the NRX-NLG adhesion complex—

we analyzed NRX-NLG BLINC signal at two culture ages.

Figure 2C shows that BLINC intensities at single puncta are

7.4-fold larger on average at DIV16 compared to DIV5. Chronic

incubation of cultures with APV, an NMDA receptor blocker,

fromDIV5-DIV16 abolishes the signal increase at DIV16.We per-

formed controls to show that the different BLINC intensities did

not result from a changing ratio of recombinant-to-endogenous

NLG1 between DIV5 and DIV16 (Figure S4D).

In addition, we examined the effect of acute chemical stimulus

on BLINC signal. We found that 1 min depolarization with high K+

to trigger global neurotransmitter release (Wittenmayer et al.,

2009) caused BLINC to increase 7.2-fold on average in DIV5

cultures (Figure 2C). This effect was suppressed when KCl was

added together with APV to block NMDA receptor activity.

One possible artifact in the interpretation of the data in

Figure 2C arises from the two-step nature of BLINC labeling.

458 Cell 143, 456–469, October 29, 2010 ª2010 Elsevier Inc.

Page 147: Cell 101029

A B

C

Figure 2. Neurexin-Neuroligin BLINC Signal Increases with Culture Age and Activity and Also Correlates with Pre- and Postsynaptic Markers

of Synapse Maturation

(A) Correlation of BLINC signal with Homer, Bassoon, and FM1-43 markers. In the top row, neurons coexpressing AP-NLG and Homer1b-CFP are plated with

neurons coexpressing BirA-NRX and Bassoon-GFP. In the bottom row, BLINC labeling was performed before FM1-43 loading in 50 mM KCl.

(B) Graphs of data in (A) show correlation between BLINC intensity and Bassoon, Homer, or FM1-43 intensity at single puncta. Sites without BLINC signal

were excluded from this analysis. For Homer and Bassoon, data are also shown after 1 min stimulation with 50 mM KCl. For each synaptic marker and each

condition, > 400 puncta from five different experiments were pooled and analyzed.

(C) Histograms comparing BLINC intensity at single puncta before and after 1 min stimulation (with or without 50 mM APV) at different culture ages. Pink lines

indicate the 25%–75% interquartile ranges. Insets show representative images andmean BLINC signal intensities (± SEM). APV is an NMDA receptor antagonist.

Chronic APV indicates incubation with 50 mM APV from DIV5 to DIV16.

See also Figure S4D.

Cell 143, 456–469, October 29, 2010 ª2010 Elsevier Inc. 459

Page 148: Cell 101029

A decrease in the mobility of AP-NLG, rather than an increase in

NRX-NLG complex size, could potentially lead to a larger BLINC

signal due to increased retention of biotinylated AP at the cell

surface, which would lead to stronger mSA staining. To check

for this artifact, we repeated all of the comparisons in

Figure 2C with a shorter biotinylation time of 5 min rather than

15 min. If changes in mobility played a role, we would expect

the fold change in BLINC to be less pronounced with 5 min

labeling, but this was not observed (data not shown).

In addition, we probed activity-dependent changes in NRX

and NLG surface levels by a separate assay. We prepared

fusions of NRX andNLG to super ecliptic pHluorin (SEP) (Sankar-

anarayanan et al., 2000), which is dark in acidic vesicles but

bright at the cell surface pH of 7.4. Figure S5A shows that these

constructs exhibit proper trafficking. We found that KCl stimula-

tion increases the abundance of both NRX and NLG at the cell

surface (vide infra; Figure 3C). Figure S5B shows that surface

NRX and NLG levels are also higher at DIV16 than at DIV5, but

not when cells are cultured in APV. Combined with the BLINC

measurements, these observations suggest that synaptic

activity—both acute and developmental—increases NRX-NLG

interactions, and such increase depends on the activity of the

NMDA receptor.

Activity Induces Surface Insertion of Neurexinand Neuroligin and New Interaction FormationWhat is the mechanism of activity-induced increase in NRX-NLG

complex size? One possibility is that activity induces the addition

of new NRX-NLG interactions to each synapse. Another

possibility is that turnover/removal of NRX-NLG interactions

from each synapse is slowed or arrested. Our single time point

BLINC assay above does not distinguish between these mecha-

nisms, so we developed new BLINC assays to probe these

mechanisms separately. In this section, we describe a pulse-

chase labeling assay to detect new NRX-NLG interaction addi-

tion to single synapses (Figure 3). In the next section, we

describe a time-lapse/surface quenching assay to detect NRX-

NLG interaction removal from single synapses (Figure 4).

A

C

D

BFigure 3. Pulse-Chase BLINC and pHluorin

Imaging Detect Activity-Dependent Addi-

tion of New Neurexin-Neuroligin Interac-

tions to the Synapse

(A) Pulse-chase labeling scheme and representa-

tive epifluorescence images. BirA-NRX/AP-NLG

contacts were first labeled to saturation using

Alexa568 and then stimulated with KCl in the

presence of FM1-43. A second round of BLINC

with Alexa647 labeled newly formed NRX-NLG

interactions.

(B) Correlation of Alexa647 and Alexa568 intensi-

ties at single puncta, under basal conditions, and

with KCl stimulation at DIV5 (top graph) or DIV16

(bottom graph).

(C) Time-lapse imaging of pHluorin (SEP) fusions to

neurexin, neuroligin, and the GluR1 subunit of the

AMPA receptor to visualize activity-induced

surface insertion. The graph on the right shows

the mean fold change in SEP intensity at single

puncta relative to prestimulus levels. Each value

is averaged from > 900 puncta from three indepen-

dent experiments. Error bars represent SEM. See

Figure S5 for additional characterization of SEP

fusion constructs.

(D) Activity-induced neurexin-neuroligin interac-

tion formation requires recycling endosomes. A

dominant-negative Rab11a mutant, Rab11aS25N-

GFP (Park et al., 2004), was introduced at DIV12

to cultures expressing BLINC reporters. At

DIV14, pulse-chase labeling was performed as in

(A) except mSA-Alexa568 was used for both steps,

and the same field of view was imaged repeatedly.

The graph on the right shows the mean fold

change (± SEM) in BLINC puncta intensity upon

KCl stimulation, with and without Rab11aS25N-

GFP coexpression.

460 Cell 143, 456–469, October 29, 2010 ª2010 Elsevier Inc.

Page 149: Cell 101029

Figure 3A shows the pulse-chase labeling scheme. First,

BLINC is performed with mSA-Alexa568 and incubation time

is extended to ensure saturation labeling of all cell surface

NRX-NLG interactions. Then, cultures are stimulated with KCl

in the presence of FM1-43 to confirm mobilization of synaptic

vesicles. Newly formed NRX-NLG interactions are detected

with a second round of BLINC labeling using mSA-Alexa647.

Figures 3A and 3B show that KCl induces robust addition

of NRX-NLG interactions in DIV5 cultures, in contrast to

untreated cultures that exhibit much less NRX-NLG interaction

addition during the same time period. Coapplication of

NMDA receptor blocker APV with KCl completely stopped

interaction addition, even to a level below that of untreated

cultures. These data suggest that NMDA receptor activity is

crucial for NRX-NLG interaction addition, both in the stimulated

and basal states. Similar trends, though less pronounced, were

observed in older DIV16 cultures (Figure 3B). Bicuculline with

4-aminopyridine to elicit acute action potentials (Hardingham

A C

B

Figure 4. Time-Lapse Imaging and Surface Quenching Reveal Activity-Dependent Arrest of Neurexin and Neuroligin Internalization at

Synapses

(A) Assay scheme and representative images. After BLINC labeling and incubation at 37�C for 15min, internalized biotinylated AP-NLG is selectively visualized by

quenching surface fluorescence with trypan blue (Howarth et al., 2008). To visualize neurexin internalization, the same assay was performed with AP on NRX

instead of NLG. Percent internalization values for single puncta were calculated by taking the ratio of pre- and postquench BLINC intensities.

(B) Histograms showing the percent internalization of biotinylatedNLG (left) or biotinylated NRX (right) at single puncta. Values in the upper right of each graph give

the percent of puncta showing > 5% internalization.

(C)Model describing the turnover of NRX-NLG interactions under basal conditions and how it changes in response to stimulation to give larger adhesion complexes.

Cell 143, 456–469, October 29, 2010 ª2010 Elsevier Inc. 461

Page 150: Cell 101029

et al., 2002) also had the same effect as KCl at DIV5 (data not

shown).

What is the source of new NRX-NLG interactions? Do they

arise from new surface NRX and NLG molecules, delivered

from internal pools? Do NRX and NLGmolecules diffuse laterally

on the cell surface into the synaptic cleft? Or are excess mole-

cules of NRX and NLG already present at the synaptic cleft,

and stimulus causes a rearrangement that leads to new binding

interactions? To investigate this question, we performed two

assays. First, we imaged SEP-NRX and SEP-NLG during KCl

stimulation (Figure 3C). We found that both undergo activity-

induced surface insertion but with different kinetics. SEP-NRX

displayed gradual surface insertion over 15 min to �3.8-fold

above prestimulus levels, whereas SEP-NLG first inserted

strongly (7-fold increase) and then decreased to�3.5-fold above

prestimulus levels after 15min. As a control, we also imaged SEP

fused to the GluR1 subunit of the AMPA receptor, which has

previously been shown to undergo activity-dependent surface

insertion (Kopec et al., 2006). SEP-GluR1 displayed similar

kinetics to SEP-NLG1, inserting strongly and then decreasing

to a level �2.4-fold above prestimulus levels. This observation

suggested that NLG1 and AMPA receptor trafficking may be

mechanistically linked.

Controls with acidification of external media confirmed that

SEP fluorescence was indeed from the cell surface

(Figure S5D). At the end of each experiment, we also increased

intracellular pH with NH4Cl and saw that internal SEP-NRX and

SEP-NLG pools colocalized with their surface counterparts

(Figure S5D).

GluR1-containing AMPA receptors are delivered to the post-

synaptic membrane from recycling endosomes (Wang et al.,

2008). To determine whether NLG1 is similarly delivered from

recycling endosomes, we performed pulse-chase labeling

in the presence of a dominant-negative Rab11a mutant,

Rab11aS25N (Park et al., 2004), to disrupt activity-induced

mobilization of recycling endosomes. Figure 3D shows that

BLINC signal growth requires NLG delivery from recycling endo-

somes. We conclude that surface insertion of NRX and NLG is

a likely mechanism for new NRX-NLG interaction formation.

Activity Also Arrests the Internalization of SynapticNeurexin and NeuroliginNRX-NLG complex growth could also be caused by activity-

dependent slowing or arrest of NRX-NLG interaction removal

from synapses. To test this hypothesis, we performed BLINC

labeling of NRX-NLG interactions, stimulated the cultures, incu-

bated for 15min, and then addedmembrane-impermeant trypan

blue to quench cell surface fluorescence (Howarth et al., 2008)

(Figure 4A). By comparing the BLINC signal before and after

quenching, we could quantify the fraction of internalized biotiny-

lated AP-NLG at each synapse.

Figures 4A and 4B show that, without stimulation, 75%of DIV5

synapses show > 5% internalization of biotinylated AP-NLG.

After 1min KCl stimulation, however, internalization is essentially

arrested; only 1% of DIV5 synapses show > 5% internalization.

Note that this arrested behavior was observed 15 min after KCl

stimulation; separate experiments showed that the ‘‘memory’’

of stimulation persisted for up to 45 min after stimulation (data

not shown). As with the pulse-chase assay, addition of APV

with KCl blocked the effect; AP-NLG internalization arrest was

no longer observed (Figure 4B).

We also used the other BLINC reporter pair, AP-NRX andBirA-

NLG, to examine the activity-dependent internalization of bioti-

nylated NRX. The same trends were observed (Figure 4B).

Without stimulation, AP-NRX displayed a wide range of internal-

ization extents. With 1 min KCl stimulation, internalization of

biotinylated AP-NRX was completely arrested. The effect was

mostly removed when APV was added with KCl, which is

particularly interesting given that NMDA receptors are on the

postsynaptic membrane, whereas AP-NRX is on the presynaptic

membrane. A retrograde signal must connect NMDA receptor

activity to NRX trafficking—possibly the NRX-NLG interaction

itself.

These internalization assays were also repeated in mature

DIV16 cultures (Figure 4B). The same trends were observed,

with one notable difference. Biotinylated AP-NLG internalized

to a lesser extent under basal conditions at DIV16 compared

to DIV5. In contrast, AP-NRX internalization was mostly

unchanged. This suggests that both acute stimulus and develop-

mental activity can alter the kinetics of NLG, but not NRX,

turnover.

The model in Figure 4C consolidates our observations from

single time point BLINC, pulse-chase BLINC, time-lapse/surface

quenching BLINC, and SEP fusion imaging. Under basal

conditions, we envision slow turnover of NRX-NLG interactions

at the synapse, with new interaction formation balanced by

NRX-NLG internalization/removal. With acute stimulus or devel-

opmental activity, however, more NRX and NLG molecules are

delivered to the cell surface to form trans-interactions, and

removal of NRX-NLG pairs is also arrested. Both processes

seem to require the activity of the NMDA receptor. These

changes lead to a net increase in the number of NRX-NLG inter-

actions at each synapse, i.e., larger NRX-NLG adhesion

complexes.

Activity-Dependent Growth of the Neurexin-NeuroliginComplex Is Correlated with AMPA Receptor InsertionHaving observed activity-dependent growth of the NRX-NLG

adhesion complex, we wondered whether this could, in turn,

promote synapse maturation via recruitment or stabilization of

specific molecules at the synaptic membrane. To investigate

this, we used one of the most established markers of mature

or potentiated synapses, the AMPA receptor (Groc et al.,

2006). pHluorin (SEP) fused to the GluR1 subunit of the AMPA

receptor (SEP-GluR1) has been shown to insert robustly into

postsynaptic membranes upon synaptic stimulation (Kopec

et al., 2006) (Figure 3C). Figure 5A shows our protocol for simul-

taneous time-lapse imaging of NRX-NLG complex growth and

SEP-GluR1 insertion, in which two rounds of BLINC staining

are performed, before and after stimulation, with the same

mSA-Alexa568 reagent. Figure 5C shows that BLINC signal

increases by 3.7-fold on average upon KCl stimulation and that

prestimulus BLINC intensity is correlated with poststimulus

BLINC intensity at each synapse. It can be seen in the first two

rows of Figure 5A that synapses that exhibit BLINC signal growth

also recruit the AMPA receptor. Figures 5D and 5E show that the

462 Cell 143, 456–469, October 29, 2010 ª2010 Elsevier Inc.

Page 151: Cell 101029

A

C

E

D

B

Figure 5. Disrupting Activity-Induced Neurexin-Neuroligin Complex Growth Disrupts AMPA Receptor Recruitment

(A) Simultaneous imaging of NRX-NLG complex growth and AMPA receptor recruitment. Neurons coexpressing AP-NLG and SEP-GluR1 were plated with

neurons expressing BirA-NRX. BLINC labeling was performed twice on each sample—both before and after KCl stimulus—with mSA-Alexa568. The top two

rows show the same field of view before and after stimulus. Bottom rows show the same experiment with coexpressed perturbing mutants BirA-NRX(D137A)

or AP-NLG(AChE swap). Arrowheads point to BLINC-positive sites at which SEP-GluR1 recruitment is disrupted. Arrows point to either extrasynaptic sites or

contacts between transfected dendrites and untransfected axons, at which activity-induced SEP-GluR1 insertion is still observed. Scale bars, 5 mm.

(B) Coexpression of BirA-NRX(D137A) or AP-NLG(AChE swap) with BLINC reporters decreases BLINC signal. Neurons were transfected with the indicated ratios

of expression plasmids. Each mean BLINC intensity (± SEM) was calculated from > 500 single puncta. See also Figure S6 for additional characterization of NRX

and NLG mutants.

(C) Correlation of prestimulus BLINC intensity with poststimulus BLINC intensity at single DIV5 puncta.

(D) Correlation of change in BLINC intensity with change in SEP-GluR1 intensity upon stimulus of DIV5 cultures. Inset shows zoom.

(E) Correlation of BLINC and SEP-GluR1 intensities before and after KCl stimulation for single puncta at DIV5.

See also Figure S7 for analogous experiments performed with glycine stimulation instead of KCl.

Cell 143, 456–469, October 29, 2010 ª2010 Elsevier Inc. 463

Page 152: Cell 101029

magnitude of AMPA receptor recruitment (DSEP-GluR1) is

correlated with the magnitude of NRX-NLG complex expansion

(DBLINC) at single synapses.

Disruption of Neurexin-Neuroligin Complex GrowthInhibits AMPA Receptor RecruitmentHaving established a correlation between NRX-NLG complex

growth and AMPA receptor recruitment, we next asked whether

NRX-NLG complex growth was required for AMPA receptor

recruitment. To investigate this, we perturbed NRX-NLG

complex growth by coexpressing BirA-NRX(D137A), a noninter-

acting NRX mutant (Graf et al., 2006), along with our standard

BLINC reporters. Figure 5B shows that this mutant has a domi-

nant-negative effect on the BLINC signal. Introduction of all three

plasmids at a 1:1:1 ratio leads to a 4.7-fold reduction in BLINC

signal compared to just the two reporter plasmids alone.

Notably, the effect on BLINC signal is even more pronounced

after stimulation (Figure 5C and Figure S6); the NRX mutant

appears to promote the removal of wild-type NRX from the

synapse by an unknown mechanism (Figure S6D).

When the perturbing mutant BirA-NRX(D137A) is introduced,

Figures 5A–5E show that BLINC signal no longer increases at

single synapses upon KCl stimulation. At these same synapses,

SEP-GluR1 surface recruitment is now blocked. One conse-

quence of our experimental setup is that, in addition to BLINC-

positive contacts between transfected axons and transfected

dendrites, each culture also contains BLINC-negative contacts

between untransfected axons and transfected dendrites. It can

be seen in Figure 5A (see arrows) and also Figure S7A (which

uses a CFP marker to highlight transfected axons) that these

BLINC-negative synapses lacking the perturbing mutant BirA-

NRX(D137A) do show robust activity-dependent SEP-GluR1

recruitment, which serves as an internal positive control.

The same set of experiments with and without BirA-NRX

(D137A) coexpressed were also performed using glycine stim-

ulus in the absence of magnesium to activate NMDA receptors

in a cell culture model of LTP (Park et al., 2004) (Figures S7B

and S7C). Similar results were obtained. We also performed

the flipped experiment, with the noninteracting mutant AP-NLG

(AChE swap) coexpressed with the BLINC reporters to perturb

NRX-NLG complex growth from the postsynaptic rather than

presynaptic side. Similar results were again obtained (Figure 5).

To examine the relationship between NRX-NLG complex

growth and AMPA receptor recruitment during development,

we analyzed synapses at DIV16 with and without the perturbing

NRX and NLG mutants coexpressed. Figure 6A shows that,

whereas SEP-GluR1 and BLINC signals are correlated at

DIV16, mutant NRX or NLG coexpression drastically reduces

BLINC signal, prevents SEP-GluR1 recruitment, and removes

the correlation between BLINC and SEP-GluR1 signals. Chronic

APV treatment to block NMDA receptor activity from DIV5-16

has a similar effect (Figure 6A).

We also examined the effect of increasing network activity

with bicuculline from DIV3-DIV5 (Ehlers, 2003) in an attempt to

artificially accelerate synapse development. Analysis of SEP-

NRX and SEP-NLG shows that these conditions promote NRX

and NLG surface insertion (Figure S5C), similar to the nonaccel-

erated developmental process from DIV5 to DIV16 (Figure S5B).

Analysis of bicuculline-treated cultures at DIV5 shows correlated

increase in BLINC and SEP-GluR1. Perturbation of NRX-NLG

complex growth with BirA-NRX(D137A) both reduces BLINC

signal and prevents SEP-GluR1 recruitment to synapses

(Figure 6B). APV addition from DIV3-DIV5 has a similar effect.

One caveat is that, because we are expressing the perturbing

mutants along with the BLINC reporters from DIV0, it is possible

that other effects, such as downregulation of NMDA receptors,

may contribute to the disruption of AMPA receptor recruitment.

Future experiments with more temporally restricted perturba-

tions will address this concern. In aggregate, our results suggest

that activity-dependent NRX-NLG complex expansion and

NMDA receptor activity are together required for AMPA receptor

recruitment during development and in response to acute simu-

lation.

DISCUSSION

Technology for Imaging trans-Synaptic Protein-ProteinInteractionsOur BLINCmethod for imaging intercellular protein-protein inter-

actions should be generally extensible to a wide variety of

protein-protein pairs and to many cell types, such as the HEK

and COS cells shown in Figures S1B and S1C. Our previous

work showed that this strategy is extensible to intracellular

protein-protein interactions (Fernandez-Suarez et al., 2008),

but after live-cell biotinylation, cells must be fixed in order to

be stained by membrane-impermeant streptavidin.

We applied BLINC to image the trans-synaptic neurexin-neu-

roligin interaction. Compared to the alternative detection

strategy of GFP complementation (GRASP) (Feinberg et al.,

2008), BLINC is nontrapping, much faster (providing signal in

as little as 8 min), and less prone to false positives. Via pulse-

chase labeling or time-lapse imaging with surface quenching,

the dynamics of interaction formation and destruction can be

studied.

BLINC in its current form does have limitations, however, and

design improvements are needed to fully exploit the power of

enzymatic probe ligation for protein interaction detection. First,

the two-step nature of the labeling adds complexity and poten-

tially introduces artifacts when, for example, biotinylated AP

internalizes into cells before streptavidin is able to stain it.

A one-step labeling protocol, such as with our coumarin fluoro-

phore ligase (Uttamapinant et al., 2010), would be preferable if

the kinetics could be improved. The other advantage of elimi-

nating the streptavidin-staining step would be better compati-

bility with labeling in live tissue, where delivery and washout of

large reagents is difficult (mSA is 56 kD). Second, biotinylation

and streptavidin staining are irreversible, so the BLINC signal

remains even after the protein pair has separated. It would be

better to have a reversible label, although, in themeantime, tricks

such as surface quenching can provide some information about

the dynamics of protein separation.

Here, we used BLINC as a tool to study the biology of the neu-

rexin-neuroligin interaction, but we also envision the use of

BLINC and related methodologies for general synapse labeling

and circuit mapping, similar to GRASP (Feinberg et al., 2008).

Depending on the proteins to which BirA and AP tags are fused,

464 Cell 143, 456–469, October 29, 2010 ª2010 Elsevier Inc.

Page 153: Cell 101029

very early synapse formation events could be detected even

before conventional synapse markers such as FM1-43 and

Bassoon are visible. In addition, BLINC with activity-dependent

proteins would allow one to distinguish between active versus

inactive synapses or newer versus older synapses.

Activity-Dependent Trafficking and Interactionsof Neurexin and NeuroliginNRX and NLG have been shown to travel in packets with other

synaptic proteins, and sites of stationary packets seem to

mark sites for development of apposing synaptic termini (Gerrow

et al., 2006; Fairless et al., 2008). Gutierrez et al. showed that

acute stimulus can slow NLG1 motion and that presynaptic

proteins accumulate at these stop sites (Gutierrez et al., 2009).

Whether these locations represent sites of surface insertion

and NRX-NLG interactions is unknown.

In general, due to lack of suitable technology, NRX-NLG inter-

actions have only been probed indirectly by gain-of-function and

loss-of-function assays. For example, overexpression of NRX in

nonneuronal cells (Graf et al., 2004), or NLG in neurons (Chih

et al., 2005), leads to enhanced recruitment of synaptic mole-

cules to apposing neuronal termini. The inference is that NRX-

NLG interactions mediated the effect. Conversely, NLG

knockout disrupts presynaptic recruitment of synaptophysin

and VGLUT (Varoqueaux et al., 2006) or Bassoon (Wittenmayer

et al., 2009), also presumably via disruption of the NRX-NLG

A

B

Figure 6. Disrupting Neurexin-Neuroligin Complex Growth during Developmental Maturation Disrupts AMPA Receptor Recruitment

(A) Neurons prepared as in Figure 5 were labeled and imaged at DIV5 or DIV16. Arrowheads point to BLINC sites that do not contain surface AMPA receptors

because of BirA-NRX(D137A) or AP-NLG(AChE swap) coexpression. Arrows point to surface AMPA receptors that may be localized to dendrites apposing

untransfected axons. Graphs on right show correlation of BLINC and SEP-GluR1 intensities at single puncta for each condition. Note that, with chronic 50

mMAPV treatment fromDIV5 to DIV16, we observed small SEP-GluR1 puncta (mean intensity 8.5 ± 2.5, compared to 62.2 ± 6.4 for the non-APV condition), which

may result from rapid AMPA receptor recruitment upon switching of cells to non-APV buffer (Liao et al., 2001). Scale bars, 5 mm.

(B) Neurons prepared as in (A) and Figure 5 were untreated or incubated with 40 mMbicuculline in the presence or absence of 50 mMAPV for 48 hr fromDIV3-DIV5

to increase network activity. Graphs on the right show correlation of BLINC and SEP-GluR1 intensities at single puncta. Scale bars, 5 mm.

See also Figure S5C for SEP-NRX and SEP-NLG imaging under the same conditions and Figure S6 for characterization of NRX and NLG perturbing mutants.

Cell 143, 456–469, October 29, 2010 ª2010 Elsevier Inc. 465

Page 154: Cell 101029

interaction. How activity affects the trafficking and interactions

of NRX and NLG, and ultimately the function of this trans-

synaptic complex, is therefore currently unknown.

Here, we directly and noninvasively imaged the trafficking and

interactions of NRX and NLG by BLINC and also by pHluorin

tagging. We found that NRX-NLG interactions are dynamic and

turn over steadily under basal conditions. Synaptic activity

induces the expansion of NRX-NLG complexes via a combina-

tion of new NRX and NLG surface insertion and arrest of NRX

and NLG internalization. Both activity-dependent surface inser-

tion and internalization arrest require the activity of the NMDA

receptor. Of interest, in our BLINC and pHluorin experiments,

we did not observe any long-range (>500 nm) lateral trafficking

of NRX and NLG into or out of synapses along the surface

membrane. One question raised but not answered by our study

is where in the synaptic cleft NRX-NLG interactions are found.

BLINC in combination with super-resolution imaging techniques

should help to determine whether NRX-NLG interactions are

located in the center or periphery of synapses.

Colocalization analysis of BLINC with synaptic markers

showed that nearly all NRX-NLG interactions are found at

Homer- and Bassoon-containing synapses at DIV16 (Figure 2).

At DIV5, however, we observed many BLINC puncta that did

not overlap with either Homer or FM1-43, although overlap

with Bassoon was high (Figure 2). This suggests that, even if

NRX-NLG interactions are not required for synapse initiation,

they still could represent one of the earliest events in the matu-

ration of nascent contacts, arriving even before functional

synaptic vesicles. Also supporting this idea is that, during our

two-color pulse-chase labeling experiments, we observed

numerous Alexa647 puncta (new NRX-NLG interactions) that

did not overlap with Alexa568 (old NRX-NLG interactions) or

FM1-43 (Figure 3). These may represent new or unsilenced

synapses that have NRX-NLG interactions, but not synaptic

vesicle activity. An interesting but unanswered question is

whether all BLINC-positive sites eventually become functional

synapses with vesicle release activity or whether formation of

NRX-NLG interactions does not represent a committed step.

BLINC also revealed several interesting differences between

immature DIV5 cultures andmature DIV16 cultures. For instance,

DIV5 neurons gave larger responses to chemical stimulation than

older DIV16 neurons in pulse-chase BLINC labeling (Figure 3 and

Figure 5). In our surface quenching assay (Figure 4), we found

a higher degree of AP-NLG internalization at DIV5 than at

DIV16. As neurons mature, decreased dendritic endocytic

capacity (Blanpied et al., 2003) may stabilize NLG at synapses,

contributing to the maturation process. Such plasticity in

younger neurons suggests a role for NRX and NLG in the early

phases of synapse maturation and possibly circuit refinement.

However, the observation that DIV16 neurons also show

activity-dependent changes in NRX-NLG complexes (Figure 3)

suggests that these proteins may also modulate plasticity in

mature neurons (Gutierrez et al., 2009).

We found that inhibition of postsynaptic NMDA receptor

activity affected both surface levels (Figure S5B) and internaliza-

tion kinetics of presynaptic neurexin (Figure 4), suggesting

retrograde signaling. Hayashi et al. previously observed that

overexpression of NLG1 and its intracellular binding partner

PSD-95 results in accumulation of presynaptic proteins and

concluded that the PSD-95-NLG1 complex may regulate

presynaptic release probability via retrograde signaling, possibly

via the NRX-NLG complex itself (Futai et al., 2007). The retro-

grade signaling that we observe may also be mediated by NRX

binding to NLG. For example, NMDA receptor activity may

lead to NLG surface insertion and, hence, more trans-binding

to NRX, which then undergoes a conformational change that

reduces its association with clathrin adaptor proteins.

Relationship between Neurexin-Neuroligin Interactionand AMPA Receptor RecruitmentPrevious studies have linked NRX-NLG signaling with the

AMPA receptor. For example, Nam et al. observed that NRX in

nonneuronal PC12 cells induces clustering of PSD-95, NMDA

receptors, and AMPA receptors (after glutamate application) in

contacting dendrites of cocultured hippocampal neurons (Nam

and Chen, 2005). Heine et al. plated NRX-coated beads on top

of neurons and observed recruitment of PSD-95 and GluR2

containing, but not GluR1 containing, AMPA receptors to the

contact sites (Heine et al., 2008).

NRX-NLG interactions and AMPA receptors have also been

linked via their shared connection to NMDA receptors.

Numerous studies have demonstrated the importance of

NMDA receptor activity for the synaptic functions of NRX and

NLG (Chubykin et al., 2007; Wittenmayer et al., 2009) and for

stable recruitment of AMPA receptors (Groc et al., 2006).

Here, we used both pHluorin imaging and BLINC to probe the

relationship between NRX-NLG interactions and AMPA recep-

tors in pure neuron cultures, without overexpression. First,

pHluorin imaging showed that NLG1 andGluR1 AMPA receptors

undergo activity-induced surface insertion with similar kinetics

(Figure 3). Second, we found that surface NLG1 is delivered

from Rab11a-containing recycling endosomes (Figure 3), from

which GluR1 AMPA receptors also originate (Park et al., 2004).

Simultaneous imaging of NRX-NLG complex growth and

GluR1 recruitment at single synapses revealed that both

processes are correlated (Figure 5). Furthermore, perturbation

of NRX-NLG complex growth, using NRX or NLG noninteracting

mutants, prevented GluR1 recruitment at those specific

synapses (Figure 5 and Figure 6).

An intriguing aspect of our study was the effect of interaction-

deficient mutants of NRX and NLG on NRX-NLG complex

dynamics. For example, coexpression of NRX(D137A) seems

to destabilize surface wild-type NRX, abolish activity-dependent

growth by removal of wild-type NRX from the synapse surface,

and consequently abolish AMPA receptor recruitment

(Figure S6, Figure 5, and Figure 6). This could be partly explained

by NRX oligomerization, although there is no current data

supporting direct or indirect (via scaffolding proteins) oligomeri-

zation of NRX. These results also raise the possibility that muta-

tions in NRX and NLG genes associated with autism spectrum

disorders (ASD) may not only affect trafficking, but also influence

surface dynamics of the NRX-NLG complex and, hence, trans-

synaptic signaling.

Figure 7 shows a proposed model for the trafficking and inter-

actions of NRX, NLG, and AMPA receptors during synapse

maturation. The link between NRX-NLG interactions and AMPA

466 Cell 143, 456–469, October 29, 2010 ª2010 Elsevier Inc.

Page 155: Cell 101029

receptors provides a molecular mechanism to rapidly and

efficiently couple structural changes at the synapse to modula-

tion of synaptic function. Aside from stabilizing AMPA receptors,

the NRX-NLG interaction may contribute to synapse maturity in

other ways as well, such as by increasing synapse adhesive

force or promoting the recruitment or stabilization of other

molecules.

Whether such NRX-NLG complex growth is a general mech-

anism for maturation of all synapses and how this phenomenon

functions cooperatively with other adhesion systems during

development is unknown. For example, because NLG1 and

NLG2 seem to specify the excitatory and inhibitory properties

of synapses, respectively, our studies raise the question of

whether inhibitory synapses mature via similar NRX-NLG2

signaling that recruits the GABA receptor. Recently, several

new binding partners for NRX and NLG have been discovered

(Siddiqui et al., 2010; Xu et al., 2010). How these interactions

influence NRX-NLG dynamics and function is unknown. It

will be intriguing to use BLINC to probe these and related

questions.

EXPERIMENTAL PROCEDURES

Neuron Culture

Dissociated hippocampal neurons were prepared from E18 rat pups. Separate

populations of suspended neurons were electroporated using a Nucleofector

apparatus (Amaxa) with AP-NLG or BirA-NRX. 1 mg of each reporter plasmid

was used for 4 million neurons. The two neuron pools were then plated

together and allowed to form synaptic contacts for 5–16 days in vitro.

BLINC Labeling

Neurons were washed twice with Tyrode’s buffer (see Extended Experimental

Procedures for recipe) and incubated with 10 mM biotin-AMP ester (Howarth

et al., 2006) in Tyrode’s buffer for 5–20 min at room temperature. Cells were

then washed once with Tyrode’s buffer and twice with TC buffer (Tyrode’s

buffer plus 0.5% biotin-free casein) before incubation with 5–7 mg/ml

mSA-Alexa568 (Howarth et al., 2006) in TC buffer for 3 min at room

temperature in the dark. Cells were washed with Tyrode’s buffer and either

imaged live in Tyrode’s buffer or fixed, depending on the downstream

experiments.

Acute Chemical Synaptic Stimulus

KCl stimulation was performed for 1 min using 50 mM KCl, 78.5 mM NaCl,

2 mM CaCl2, 2 mMMgCl2, 30 mM glucose, and 25 mM HEPES (pH 7.4). Bicu-

culline stimulation was performed for 5 min using 50 mM bicuculline (Tocris)

and 250 mM 4-amino-pyridine (4-AP, Tocris) in Tyrode’s buffer.

Two-Color Pulse-Chase BLINC Labeling

The first round of BLINC was performed as described above, with 20 min

biotin-AMP and 3 min mSA-Alexa568. Neurons were then stimulated with

KCl as described above in the presence of 10 mM FM1-43. Cells were washed

with Tyrode’s buffer once and immediately incubated with biotin-AMP for

5 min followed by mSA-Alexa647 for 3 min. Cells were then washed with Ty-

rode’s buffer and imaged live immediately within 5–10 min at room tempera-

ture. For experiments with APV, 50 mMAPV (Sigma) was added to the Tyrode’s

wash buffer and to the stimulation buffer after the first BLINC labeling.

Single-Color Pulse-Chase BLINC Labeling

Neurons were labeled and imaged in a RC21B chamber using a PM-2 heated

platform (Warner Instruments, Hamden CT). Cells were constantly perfused

with Tyrode’s buffer running through an in-line heater set at 37�C. All labelingreagents and stimulants were delivered by perfusion. The first round of BLINC

was performed as described above, with 20 min biotin-AMP and 3 min mSA-

Alexa568, and prestimulus images were acquired. Neurons were then stimu-

lated with KCl as described above, washed, and labeled a second time with

biotin-AMP for 5 min and mSA-Alexa568 for 3 min. Seven minutes after the

second BLINC labeling, poststimulus images were acquired. This delay was

to match the 15 min time window between stimulus and surface quenching

steps in Figure 4.

For glycine stimulus experiments, neurons were initially perfused with

Tyrode’s buffer containing 10 mM CNQX, 50 mM APV, and 1 mM strychnine.

The first round of BLINC labeling was performed in this same buffer. Neurons

were then stimulated for 3 min with 200 mM glycine, 1 mM strychnine, and

20 mM bicuculline in Tyrode’s buffer (without magnesium) (Park et al., 2004).

Neurons were then switched to Tyrode’s buffer containing 2 mM MgCl2,

0.5 mM tetrodotoxin, 10 mM CNQX, 50 mM APV, and 1 mM strychnine for the

second round of BLINC labeling for 8 min. Imaging was performed in this

same buffer after 7 min.

BLINC Internalization Assay via Surface Fluorescence Quenching

After BLINC labeling as described above, neurons were stimulated with KCl as

described above and then incubated in Tyrode’s buffer for 15 min at 37�C. Forsurface fluorescence quenching, Tyrode’s buffer was replacedwith pre-chilled

(4�C) quench buffer (20 mM trypan blue [VWR International] in Tyrode’s buffer)

Figure 7. Model for Activity-Dependent

Trafficking and Interactions of Neurexin,

Neuroligin, and AMPA Receptor during

Synapse Maturation

Nascent synapses have few NRX-NLG interac-

tions, few NMDA receptors, and probably few

AMPA receptors. If present, these AMPA recep-

tors are considered labile (Groc et al., 2006).

Synaptic activity causes robust insertion of both

NLG1 and AMPA receptors into the postsynaptic

membrane via NMDA receptor activity and mobi-

lization of recycling endosomes (Park et al., 2004).

Synaptic activity also arrests internalization of

both neurexin and neuroligin. Gradually, some

NLG1 molecules are stabilized by binding to

NRX on the presynaptic membrane. NRX-bound

NLGs stabilize AMPA receptors, whereas the

ones not bound by NRX may endocytose back

along with AMPA receptors. Note that, while

surface NRX levels increase gradually in this model, surface NLG1 increases strongly and then decreases again to a level higher than basal. In this model,

activity-dependent recruitment of the AMPA receptor requires both NMDA receptor activation and activity-dependent NRX-NLG complex growth. These

processes ultimately lead to unsilencing or maturation of the synapse.

Cell 143, 456–469, October 29, 2010 ª2010 Elsevier Inc. 467

Page 156: Cell 101029

for 1 min. Images were acquired immediately before and after surface quench-

ing. Where indicated, 100 mM APV was added during stimulation and in all

subsequent steps to block NMDA receptor activity.

Detailed protocols for neuron culture preparation, pHluorin imaging, cell

fixation, immunostaining, fluorescence microscopy, and image analysis can

be found in the Extended Experimental Procedures.

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures and

seven figures and can be found with this article online at doi:10.1016/j.cell.

2010.09.025.

ACKNOWLEDGMENTS

We thank the late Alaa El-Husseini (University of British Columbia), Michael

Ehlers (Duke), Craig Garner (Stanford), and Joshua Sanes (Harvard) for their

advice, plasmids, and critical comments on the manuscript. Masahito

Yamagata (Harvard), Brian Chen (McGill), Yasunori Hayashi (RIKEN), Miguel

Bosch (MIT), and Ann-Marie Craig (University of British Columbia) provided

plasmids and antibodies. Mark Howarth made key intellectual contributions,

performed preliminary experiments, and provided biotin-AMP. Daniel Dai

and Yi Zheng assisted with neuron cultures and provided mSA protein.

Funding was provided by the NIH (DP1 OD003961-01), McKnight Foundation,

Sloan Foundation, and MIT. A. Thyagarajan was supported by an Autism

Speaks postdoctoral fellowship.

Received: November 30, 2009

Revised: May 26, 2010

Accepted: August 17, 2010

Published online: October 7, 2010

REFERENCES

Arac, D., Boucard, A.A., Ozkan, E., Strop, P., Newell, E., Sudhof, T.C., and

Brunger, A.T. (2007). Structures of neuroligin-1 and the neuroligin-1/neu-

rexin-1 beta complex reveal specific protein-protein and protein-Ca2+ interac-

tions. Neuron 56, 992–1003.

Blanpied, T.A., Scott, D.B., and Ehlers, M.D. (2003). Age-related regulation of

dendritic endocytosis associated with altered clathrin dynamics. Neurobiol.

Aging 24, 1095–1104.

Chih, B., Engelman, H., and Scheiffele, P. (2005). Control of excitatory and

inhibitory synapse formation by neuroligins. Science 307, 1324–1328.

Chubykin, A.A., Atasoy, D., Etherton, M.R., Brose, N., Kavalali, E.T., Gibson,

J.R., and Sudhof, T.C. (2007). Activity-dependent validation of excitatory

versus inhibitory synapses by neuroligin-1 versus neuroligin-2. Neuron 54,

919–931.

Ehlers, M.D. (2003). Activity level controls postsynaptic composition and

signaling via the ubiquitin-proteasome system. Nat. Neurosci. 6, 231–242.

Fairless, R., Masius, H., Rohlmann, A., Heupel, K., Ahmad, M., Reissner, C.,

Dresbach, T., and Missler, M. (2008). Polarized targeting of neurexins to

synapses is regulated by their C-terminal sequences. J. Neurosci. 28,

12969–12981.

Feinberg, E.H., Vanhoven, M.K., Bendesky, A., Wang, G., Fetter, R.D., Shen,

K., and Bargmann, C.I. (2008). GFP Reconstitution Across Synaptic Partners

(GRASP) defines cell contacts and synapses in living nervous systems. Neuron

57, 353–363.

Fernandez-Suarez, M., Chen, T.S., and Ting, A.Y. (2008). Protein-protein inter-

action detection in vitro and in cells by proximity biotinylation. J. Am. Chem.

Soc. 130, 9251–9253.

Friedman, H.V., Bresler, T., Garner, C.C., and Ziv, N.E. (2000). Assembly of

new individual excitatory synapses: time course and temporal order of

synaptic molecule recruitment. Neuron 27, 57–69.

Futai, K., Kim, M.J., Hashikawa, T., Scheiffele, P., Sheng, M., and Hayashi, Y.

(2007). Retrograde modulation of presynaptic release probability through

signaling mediated by PSD-95-neuroligin. Nat. Neurosci. 10, 186–195.

Garner, C.C., Waites, C.L., and Ziv, N.E. (2006). Synapse development: still

looking for the forest, still lost in the trees. Cell Tissue Res. 326, 249–262.

Gerrow, K., Romorini, S., Nabi, S.M., Colicos, M.A., Sala, C., and El-Husseini,

A. (2006). A preformed complex of postsynaptic proteins is involved in excit-

atory synapse development. Neuron 49, 547–562.

Graf, E.R., Kang, Y., Hauner, A.M., and Craig, A.M. (2006). Structure function

and splice site analysis of the synaptogenic activity of the neurexin-1 beta LNS

domain. J. Neurosci. 26, 4256–4265.

Graf, E.R., Zhang, X., Jin, S.X., Linhoff, M.W., and Craig, A.M. (2004). Neurex-

ins induce differentiation of GABA and glutamate postsynaptic specializations

via neuroligins. Cell 119, 1013–1026.

Groc, L., Gustafsson, B., and Hanse, E. (2006). AMPA signalling in nascent

glutamatergic synapses: there and not there! Trends Neurosci. 29, 132–139.

Gutierrez, R.C., Flynn, R., Hung, J., Kertesz, A.C., Sullivan, A., Zamponi, G.W.,

El-Husseini, A., and Colicos, M.A. (2009). Activity-driven mobilization of post-

synaptic proteins. Eur. J. Neurosci. 30, 2042–2052.

Hardingham, G.E., Fukunaga, Y., and Bading, H. (2002). Extrasynaptic

NMDARs oppose synaptic NMDARs by triggering CREB shut-off and cell

death pathways. Nat. Neurosci. 5, 405–414.

Heine, M., Thoumine, O., Mondin, M., Tessier, B., Giannone, G., and Choquet,

D. (2008). Activity-independent and subunit-specific recruitment of functional

AMPA receptors at neurexin/neuroligin contacts. Proc. Natl. Acad. Sci. USA

105, 20947–20952.

Howarth, M., Chinnapen, D.J., Gerrow, K., Dorrestein, P.C., Grandy, M.R., Kel-

leher, N.L., El-Husseini, A., and Ting, A.Y. (2006). A monovalent streptavidin

with a single femtomolar biotin binding site. Nat. Methods 3, 267–273.

Howarth, M., Liu, W., Puthenveetil, S., Zheng, Y., Marshall, L.F., Schmidt,

M.M., Wittrup, K.D., Bawendi, M.G., and Ting, A.Y. (2008). Monovalent,

reduced-size quantum dots for imaging receptors on living cells. Nat. Methods

5, 397–399.

Kaech, S., and Banker, G. (2006). Culturing hippocampal neurons. Nat. Protoc.

1, 2406–2415.

Ko, J., Fuccillo, M.V., Malenka, R.C., and Sudhof, T.C. (2009). LRRTM2 func-

tions as a neurexin ligand in promoting excitatory synapse formation. Neuron

64, 791–798.

Kopec, C.D., Li, B., Wei, W., Boehm, J., and Malinow, R. (2006). Glutamate

receptor exocytosis and spine enlargement during chemically induced long-

term potentiation. J. Neurosci. 26, 2000–2009.

Liao, D., Scannevin, R.H., and Huganir, R. (2001). Activation of silent synapses

by rapid activity-dependent synaptic recruitment of AMPA receptors. J. Neu-

rosci. 21, 6008–6017.

Mazzoni, A., Broccard, F.D., Garcia-Perez, E., Bonifazi, P., Ruaro, M.E., and

Torre, V. (2007). On the dynamics of the spontaneous activity in neuronal

networks. PLoS ONE 2, e439.

Nam, C.I., and Chen, L. (2005). Postsynaptic assembly induced by neurexin-

neuroligin interaction and neurotransmitter. Proc. Natl. Acad. Sci. USA 102,

6137–6142.

Park, M., Penick, E.C., Edwards, J.G., Kauer, J.A., and Ehlers, M.D. (2004).

Recycling endosomes supply AMPA receptors for LTP. Science 305,

1972–1975.

Sankaranarayanan, S., De Angelis, D., Rothman, J.E., and Ryan, T.A. (2000).

The use of pHluorins for optical measurements of presynaptic activity. Bio-

phys. J. 79, 2199–2208.

Siddiqui, T.J., Pancaroglu, R., Kang, Y., Rooyakkers, A., and Craig, A.M.

(2010). LRRTMs and neuroligins bind neurexins with a differential code to

cooperate in glutamate synapse development. J. Neurosci. 30, 7495–7506.

Sudhof, T.C. (2008). Neuroligins and neurexins link synaptic function to cogni-

tive disease. Nature 455, 903–911.

468 Cell 143, 456–469, October 29, 2010 ª2010 Elsevier Inc.

Page 157: Cell 101029

Sudhof, T.C., and Malenka, R.C. (2008). Understanding synapses: past,

present, and future. Neuron 60, 469–476.

Taniguchi, H., Gollan, L., Scholl, F.G., Mahadomrongkul, V., Dobler, E.,

Limthong, N., Peck, M., Aoki, C., and Scheiffele, P. (2007). Silencing of neuro-

ligin function by postsynaptic neurexins. J. Neurosci. 27, 2815–2824.

Uemura, T., Lee, S.J., Yasumura, M., Takeuchi, T., Yoshida, T., Ra, M.,

Taguchi, R., Sakimura, K., and Mishina, M. (2010). Trans-synaptic interaction

of GluRdelta2 and Neurexin through Cbln1 mediates synapse formation in the

cerebellum. Cell 141, 1068–1079.

Uttamapinant, C., White, K.A., Baruah, H., Thompson, S., Fernandez-Suarez,

M., Puthenveetil, S., and Ting, A.Y. (2010). A fluorophore ligase for site-specific

protein labeling inside living cells. Proc. Natl. Acad. Sci. USA 107, 10914–

10919.

Varoqueaux, F., Aramuni, G., Rawson, R.L., Mohrmann, R., Missler, M.,

Gottmann, K., Zhang, W., Sudhof, T.C., and Brose, N. (2006). Neuroligins

determine synapse maturation and function. Neuron 51, 741–754.

Wang, Z., Edwards, J.G., Riley, N., Provance, D.W., Jr., Karcher, R., Li, X.D.,

Davison, I.G., Ikebe, M., Mercer, J.A., Kauer, J.A., and Ehlers, M.D. (2008).

Myosin Vb mobilizes recycling endosomes and AMPA receptors for postsyn-

aptic plasticity. Cell 135, 535–548.

Weaver, L.H., Kwon, K., Beckett, D., andMatthews, B.W. (2001). Corepressor-

induced organization and assembly of the biotin repressor: a model for allo-

steric activation of a transcriptional regulator. Proc. Natl. Acad. Sci. USA 98,

6045–6050.

Wittenmayer, N., Korber, C., Liu, H., Kremer, T., Varoqueaux, F., Chapman,

E.R., Brose, N., Kuner, T., and Dresbach, T. (2009). Postsynaptic Neuroligin1

regulates presynaptic maturation. Proc. Natl. Acad. Sci. USA 106, 13564–

13569.

Xu, J., Xiao, N., and Xia, J. (2010). Thrombospondin 1 accelerates synaptogen-

esis in hippocampal neurons through neuroligin 1. Nat. Neurosci. 13, 22–24.

Cell 143, 456–469, October 29, 2010 ª2010 Elsevier Inc. 469

Page 158: Cell 101029

Nucleosome-Interacting ProteinsRegulated by DNAand Histone MethylationTill Bartke,1 Michiel Vermeulen,2,3 Blerta Xhemalce,1 Samuel C. Robson,1 Matthias Mann,2 and Tony Kouzarides1,*1The Gurdon Institute and Department of Pathology, Tennis Court Road, Cambridge CB2 1QN, UK2Department of Proteomics and Signal Transduction, Max-Planck-Institute for Biochemistry, D-82152 Martinsried, Germany3Present address: Department of Physiological Chemistry and Cancer Genomics Centre, University Medical Center Utrecht,

3584 CX Utrecht, The Netherlands*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.10.012

SUMMARY

Modifications on histones or on DNA recruit proteinsthat regulate chromatin function. Here, we use nucle-osomes methylated on DNA and on histone H3 in anaffinity assay, in conjunction with a SILAC-basedproteomic analysis, to identify ‘‘crosstalk’’ betweenthese two distinct classes of modification. Ouranalysis reveals proteins whose binding to nucleo-somes is regulated by methylation of CpGs, H3K4,H3K9, and H3K27 or a combination thereof. We iden-tify the origin recognition complex (ORC), includingLRWD1 as a subunit, to be a methylation-sensitivenucleosome interactor that is recruited cooperativelyby DNA and histone methylation. Other interactors,such as the lysine demethylase Fbxl11/KDM2A,recognize nucleosomes methylated on histones,but their recruitment is disrupted by DNA methyla-tion. These data establish SILAC nucleosome affinitypurifications (SNAP) as a tool for studying thedynamics between different chromatin modificationsand provide a modification binding ‘‘profile’’ forproteins regulated by DNA and histone methylation.

INTRODUCTION

Most of the genetic information of eukaryotic cells is stored in

the nucleus in the form of a nucleoprotein complex termed

chromatin. The basic unit of chromatin is the nucleosome,

which consists of 147 bp of DNA wrapped around an octamer

made up of two copies each of the core histones H2A, H2B,

H3, and H4 (Luger et al., 1997). Nucleosomes are arranged

into higher-order structures by additional proteins, including

the linker histone H1, to form chromatin. Because chromatin

serves as the primary substrate for all DNA-related processes

in the nucleus, its structure and activity must be tightly

controlled.

Two key mechanisms known to regulate the functional state

of chromatin in higher eukaryotes are the C5 methylation of

DNA at cytosines within CpG dinucleotides and the posttransla-

tional modification of amino acids of histone proteins. Whereas

DNA methylation is usually linked to silent chromatin and is

present in most regions of the genome (Bernstein et al., 2007),

the repertoire and the location of histone modifications are

much more diverse, with different modifications associated

with different biological functions (Kouzarides, 2007). Most

modifications can also be removed from chromatin, thus

conferring flexibility in the regulation of its activity. Due to the

large number of possible modifications and the enormous diver-

sity that can be generated through combinatorial modifications,

epigenetic information can be stored in chromatin modification

patterns. Several chromatin-regulating factors have recently

been identified that recognize methylated DNA or modified

histone proteins. Such effector molecules use a range of

different recognition domains such as methyl-CpG-binding

domains (MBD), zinc fingers (ZnF), chromo-domains, or plant

homeodomains (PHD) in order to establish and orchestrate

biological events (Sasai and Defossez, 2009; Taverna et al.,

2007). However, most of these studies were conducted using

isolated DNA or histone peptides and cannot recapitulate the

situation found in chromatin. Considering the three-dimensional

organization of chromatin in the nucleus, DNA methylation and

histone modifications most likely act in a concerted manner by

creating a ‘‘modification landscape’’ that must be interpreted

by proteins that are able to recognize large molecular assem-

blies (Ruthenburg et al., 2007).

In an effort to increase our understanding of how combinatorial

modifications on chromatin might modulate its activity, we set

out to identify factors that recognize methylated DNA and

histones in the context of nucleosomes. We reasoned that using

whole nucleosomes would enable us to find factors that

integrate the folded nucleosomal structure with modifications

on the DNA and on histones. Here, we describe a SILAC nucle-

osome affinity purification (SNAP) approach for the identification

of proteins that are influenced by CpG methylation and histone

H3 K4, K9, or K27 methylation (or a combination thereof) in the

context of a nucleosome. Our results reveal many proteins and

complexes that can read the chromatin modification status.

These results establish SNAP as a valuable approach in defining

the chromatin ‘‘interactome.’’

470 Cell 143, 470–484, October 29, 2010 ª2010 Elsevier Inc.

Page 159: Cell 101029

RESULTS

The SILAC Nucleosome Affinity PurificationProteins recognize modifications of chromatin in the context of

a nucleosome. However, to date, modification-interacting

proteins have been identified using modified DNA or modified

histone peptides as affinity columns. We set out to identify

proteins that can sense the presence of DNA and histone meth-

ylation within the physiological background of a nucleosome.

To this end, we reconstituted recombinant nucleosomes con-

taining combinations of CpG-methylated DNA and histone H3

trimethylated at lysine residues 4, 9, and 27 (H3K4me3,

H3K9me3, or H3K27me3). These modified nucleosomes were

immobilized on beads and used to affinity purify interacting

proteins from SILAC-labeled HeLa nuclear extracts (Figure 1A).

Bound proteins regulated by the different modification patterns

were identified by mass spectrometry (MS).

The methylation of lysines in H3 was accomplished by native

chemical ligation (Muir, 2003). An existing protocol (Shogren-

Knaak et al., 2003) was adapted to develop an improved method

that allows the purification of large quantities of recombinant tail-

less human H3.1 (Figure 1B). This method employs the

coexpression of tobacco etch virus (TEV) protease and a modi-

fied TEV cleavage site (Tolbert and Wong, 2002) to expose

a cysteine in front of the histone core sequence in E. coli (Figur-

e S1A available online). The tail-less H3.1 starting with a cysteine

at position 32 was ligated to thioester peptides spanning the

N terminus of histone H3.1 (residues 1–31) and containing the

above-mentioned methylated lysines (Figure S1B). The resulting

full-length modified H3.1 proteins (Figure S1C) were subse-

quently refolded into histone octamers together with recombi-

nant human histones H2A, H2B, and H4 (Figure 1C).

As nucleosomal DNAs, we used two biotinylated 185 bp DNA

fragments containing either the 601 or the 603 nucleosome posi-

tioning sequences (Lowary and Widom, 1998). Both DNAs have

similar nucleosome-forming properties, albeit with different

sequences (Figure S1D), which allows us to test for sequence

specificities of methyl-CpG interactors. The nucleosomal DNAs

were treated with recombinant prokaryotic M.SssI DNA methyl-

transferase, which mimics the methylation pattern found at CpG

dinucleotides in eukaryotic genomic DNA (Figures S1E and S1F).

Finally, nucleosomal core particles were reconstituted from the

nucleosomal DNAs and octamers and were immobilized on

streptavidin beads via the biotinylated DNAs. All assembly reac-

tions were quality controlled on native PAGE gels (Figure S1G).

The immobilized modified nucleosomes were incubated in

HeLaS3 nuclear extracts and probed for the binding of known

modification-interacting factors to make sure that the nucleo-

somal templates were functional. Figure 1D shows that, as

expected, PHF8, HP1a, and the polycomb repressive complex

2 (PRC2) subunit SUZ12 (Bannister et al., 2001; Hansen et al.,

2008; Kleine-Kohlbrecher et al., 2010) specifically bind to

H3K4me3-, H3K9me3-, and H3K27me3-modified nucleosomes,

respectively. In addition, we did not detect any modification of

the immobilized nucleosomal histones by modifying activities

present in the nuclear extract (Figure S1H).

In order to identify proteins that bind to chromatin in a modifi-

cation-dependent manner, we utilized a SILAC pull-down

approach that we have developed to identify interactors of

histone modifications (Vermeulen et al., 2010). We simply

replaced immobilized peptides with complete reconstituted

modified nucleosomes (Figure 2A). All pull-downs were repeated

in two experiments. In a ‘‘forward’’ experiment, the unmodified

nucleosomes were incubated with light (R0K0) extracts, and the

modified nucleosomes were incubated with heavy-labeled

(R10K8) extracts, as depicted in Figure 2A. In an independent

‘‘reverse’’ experiment, the extracts were exchanged. Bound

proteins were identified and quantified by high-resolution MS

for both pull-down experiments. A logarithmic (Log2) plot of the

SILAC ratios heavy/light (ratio H/L) of the forward (x axis) and

reverse (y axis) experiments for each identified protein allows

the unbiased identification of proteins that specifically bind to

the modified or the unmodified nucleosomes. Proteins that

preferentially bind to the modified nucleosomes show a high

ratio H/L in the forward and a low ratio H/L in the reverse exper-

iment and can, therefore, be identified as outliers in the bottom-

right quadrant. Proteins that are excluded by the modification

have a low ratio H/L in the forward experiment and a high ratio

H/L in the reverse experiment and appear in the top-left quad-

rant. Background binders have a ratio H/L of around 1:1 and

cluster around the intersection of the x and y axes. Outliers in

the bottom-left quadrant are contaminating proteins. Outliers in

the top-right quadrant are false positives. An enrichment/exclu-

sion ratio of 1.5 in both directions generally identifies outliers

outside of the background cluster. We consider a protein to be

significantly regulated when it is enriched/excluded at least

2-fold. Higher ratios H/L in the forward and lower ratios H/L in

the reverse experiments indicate stronger binding, whereas

stronger exclusion is indicated by lower ratios H/L in the forward

and higher ratios H/L in the reverse experiments.

Proteins Identified by SNAPThe SNAP approach was used to identify proteins that are

recruited or excluded by DNA methylation, histone H3 methyla-

tion, or a combination of both (Figures 2B and 2C and Figure S2).

In Table 1, Table 2, and Table S2, we summarize the proteins that

display a regulation of at least 1.5 in both the forward and reverse

experiments, thus defining the proteins that are enriched or

excluded by the modified nucleosomes. The complete MS

analysis defining all interacting proteins in all pull-down reactions

is summarized in Table S1.

The data set includes a number of proteins (about 20%) that

are already known to bind methyl-DNA and methyl-H3, as well

as many proteins whose regulation by modifications had not

been previously defined. The presence of many known methyl-

binding proteins validates our approach. The database provides

a complex ‘‘profile’’ for the modulation of proteins by DNA and

histone methylation that have the potential to recognize specific

‘‘chromatin landscapes.’’ Below, we highlight several interac-

tions with modified nucleosomes, which exemplify the different

modes of regulation that we observe (summarized in Figures

2D and 2E).

Regulation by CpG MethylationTable 1 shows DNA- and nucleosome-binding proteins regu-

lated by CpG methylation. The two different methylated DNAs

Cell 143, 470–484, October 29, 2010 ª2010 Elsevier Inc. 471

Page 160: Cell 101029

were subjected to SNAP analysis either on their own (601me DNA

and 603me DNA) or assembled into nucleosomes (601me Nuc and

603me Nuc). We identify several well-characterized methyl-

binding proteins such as MBD2 (Sasai and Defossez, 2009) to

be enriched on the 601me and 603me DNAs. MBD2 is enriched

on both DNAs and exemplifies a form of methyl-CpG binding

that is not sequence selective. In contrast, other proteins (e.g.,

ZNF295) display sequence specificity toward only one of the

methylated DNAs, suggesting that they may recognize CpG

methylation in a sequence-specific manner.

Figure 1. Preparation of Reconstituted Modified Nucleosomes

(A) Experimental strategy for the preparation of immobilized and modified nucleosomes for pull-down studies.

(B) The native chemical ligation strategy for generating posttranslationally modified histone H3.1. We bacterially express an IPTG-inducible truncated histone

precursor containing a modified TEV-cleavage site (ENLYFQYC) followed by the core sequence of histone H3.1 starting from glycine 33. The plasmid also

contains TEV-protease under the control of the AraC/PBAD promoter. TEV-protease accepts a cysteine instead of glycine or serine as the P10 residue of its recog-

nition site, and upon arabinose induction, it processes the precursor histone into the truncated form (H3.1D1-31 T32C), which is purified and ligated to modified

thioester peptides spanning the N-terminal residues 1 to 31 of histone H3.1. All ligated histones contain the desired modification and a T32C mutation.

(C) Summary of the modified histone octamers. The top panel shows 1 mg of each octamer separated by SDS-PAGE and stained with Coomassie. For the bottom

panel, octamers were dot blotted on PVDF membranes and probed with modification-specific antibodies as indicated. The anti-H3K27me3 antibody shows slight

cross-reactivity with H3K4me3 and H3K9me3.

(D) Functional test of the nucleosome affinity matrix. R10K8-labeled nuclear extract was incubated with immobilized modified nucleosomes as indicated. Binding

of PHF8, HP1a, and SUZ12 was detected by immunoblot. Equal loading was confirmed by silver and Coomassie staining. Modification of histone H3 was verified

by immunoblot against H3 trimethyl lysine marks. All three antibodies show slight cross-reactivity with the other histone marks.

See also Figure S1.

472 Cell 143, 470–484, October 29, 2010 ª2010 Elsevier Inc.

Page 161: Cell 101029

We also identify many proteins that preferentially recognize

nonmethylated DNA and are excluded by CpG methylation.

The most prominent example is the general RNA polymerase III

transcription factor TFIIIC. All subunits of the TFIIIC complex

show specific exclusion from the 603me DNA (e.g., GTF3C5

shown in Figure 2D), most likely because this DNA (unlike the

601me DNA) contains two putative B box elements (Figure S1D),

sequences that are known TFIIIC-binding sites. This defines

a form of methyl-CpG-dependent exclusion that is sequence

specific.

CpG methylation can have a distinct influence on protein

binding when it is present within a nucleosomal background.

Factors such as MeCP2 are specifically enriched on CpG-meth-

ylated DNA only in the context of a nucleosome, but not on free

DNA (Figure 2D). Other factors, such as L3MBTL3, show

nucleosome-dependent exclusion by CpG methylation. These

two factors are influenced by DNA methylation regardless of

DNA sequence. Several proteins, such as the DNA-binding

factor USF2, are specifically excluded only from 601me nucleo-

somes. This is most likely due to an E box motif in the 601

DNA (Figure S1D), which is recognized by USF2.

One final example of the effect of nucleosomes on DNA-

binding proteins is demonstrated by the observation that many

proteins such as TFIIIC bind free DNA but cannot recognize

the DNA when it is assembled into nucleosomes. This is probably

due to binding motifs (such as the B box motif) being occluded

by the histone octamer (Figure 2D and Table S2). This type of

interaction may identify proteins that need nucleosome-remod-

eling activities to bind their DNA element. Together, these exam-

ples highlight the additional constraints forced on protein-DNA

interactions by the histone octamer.

Regulation by H3 Lysine MethylationTable 2 shows a summary of the proteins enriched or excluded

by nucleosomes trimethylated at H3K4, H3K9, or H3K27 in the

presence or absence of DNA methylation. Trimethylation of

H3K4 is primarily associated with active promoters, whereas

trimethyl H3K9 and H3K27, as well as methyl-CpG, are hallmarks

of silenced regions of the genome (Kouzarides, 2007).

We identify several known histone methyl-binding proteins in

our screen, such as the H3K4me3-interactor CHD1, the

H3K9me3-binder UHRF1, and the H3K27me3-interacting poly-

comb group protein CBX8 (Hansen et al., 2008; Karagianni

et al., 2008; Pray-Grant et al., 2005). In addition, a number of

uncharacterized factors were identified. For example, Spindlin1

binds strongly to H3K4me3. Spindlin1 is a highly conserved

protein consisting of three Spin/Ssty domains that have recently

been shown to fold into Tudor-like domains (Zhao et al., 2007),

motifs known to bind methyl lysines on histone proteins. Most

notably, we identify the origin recognition complex (Orc2,

Orc3, Orc4, Orc5, and to a lesser extent Orc1) to be enriched

on both H3K9me3- and H3K27me3-modified nucleosomes.

Because no binding was detected on H3K4me3 nucleosomes,

the origin recognition complex (ORC) seems to specifically

recognize heterochromatic modifications (Figure 2E). One

protein, PHF14, and, to a lesser extent, HMG20A and

HMG20B are excluded by the H3K4me3 modification. Of

interest, these factors represent the only significant examples

of proteins excluded from nucleosomes by methylation of

histones, including methylation at H3K9 and H3K27.

Crosstalk between DNA and Histone MethylationThe SNAP approach allows us to investigate cooperative

effects between DNA methylation and histone modifications

on the recruitment of proteins to chromatin. Analysis of our

data reveals several examples of such a regulation (Figures

2E and 2F). We observe a cooperative stronger binding of

UHRF1 to H3K9me3-modified nucleosomes in the presence

of CpG methylation. Similarly, the ORC (as shown for the

Orc2 subunit) can recognize nucleosomes more effectively if

CpG methylation coincides with the repressive histone marks

H3K9me3 or H3K27me3. This might explain its preferential

localization to heterochromatic regions in the nucleus (Pak

et al., 1997; Prasanth et al., 2004). In contrast, the H3K36 deme-

thylase Fbxl11/KDM2A is enriched by H3K9 methylation but

excluded by DNA methylation. Finally, the PRC2 complex is

enriched on H3K27me3 nucleosomes (and to a lesser extent

on H3K9me3 nucleosomes), but incorporation of methyl-CpG

DNA counteracts this recruitment, as shown for the EED

(Figure 2E) and the SUZ12 (Figure 2F) subunits. These findings

demonstrate the ability of these factors to simultaneously

monitor the methylation status of both histones and DNA on

a single nucleosome.

Identification of Complexes Regulated by ChromatinModificationsThe proteins regulated by nucleosome modifications in the

SNAP experiments were subjected to a cluster analysis in order

to define common features of regulation. In this analysis, the

SILAC enrichment values are represented as a heat map in

which proteins with similar interaction profiles group into clus-

ters that may be indicative of protein complexes. Figure 3

shows that members of several known complexes cluster

together in this analysis, including the BCOR and the NuRD

corepressor complexes (Gearhart et al., 2006; Le Guezennec

et al., 2006).

Identification of LRWD1 as an ORC-Interacting ProteinThe cluster analysis also identifies the ORC based on the similar

interaction profiles of the ORC subunits. Of interest, an unchar-

acterized protein termed LRWD1 closely associates with the

ORC cluster (see also Figures 2B and 2C and Figures S2G and

S2H), suggesting that this protein may be a component of

ORC. To test this hypothesis, we raised an antibody against

LRWD1 (Figure S3A) and used it to probe for colocalization

with the ORC by immunofluorescence (IF) staining of MCF7 cells.

Figure 4A indicates that LRWD1 colocalizes with the ORC at

a subset of nuclear foci marked by strong staining with an

antibody against the Orc2 subunit. As previously shown for

Orc2 (Prasanth et al., 2004), these foci often colocalize with

HP1a, a marker for H3K9me3-containing heterochromatin (Fig-

ure S3B). In addition, endogenous LRWD1 and Orc2 can be

coimmunoprecipitated from extracts prepared from MCF7 and

HelaS3 cells (Figure 4B and Figure S3C). We further expressed

various truncated variants of FLAG-tagged LRWD1 in 293T cells

and immunoprecipitated them using an anti-FLAG antibody. The

Cell 143, 470–484, October 29, 2010 ª2010 Elsevier Inc. 473

Page 162: Cell 101029

Figure 2. Identification of Nucleosome-Interacting Proteins Regulated by DNA and Histone Methylation Using SNAP

(A) Experimental design of the SILAC nucleosome affinity purifications. Nuclear extracts are prepared from HeLaS3 cells grown in conventional ‘‘light’’ medium or

medium containing stable isotope-labeled ‘‘heavy’’ amino acids. The resulting ‘‘light’’ and ‘‘heavy’’ labeled proteins can be distinguished and quantified by MS.

474 Cell 143, 470–484, October 29, 2010 ª2010 Elsevier Inc.

Page 163: Cell 101029

coimmunoprecipitation of Orc1 and Orc2 indicates that LRWD1

interacts with ORC via its WD40 domain (Figures 4C and 4D and

Figure S3D). Similar to Orc3 (Prasanth et al., 2004), expression of

LRWD1 depends on Orc2 because reducing Orc2 expression in

MCF7 cells by siRNA treatment also reduces LRWD1 protein

levels (Figure 4E) without perturbing its transcription (data not

shown). These experiments establish LRWD1 as an ORC

component and demonstrate the potential of the modification

interaction profiling for the identification of protein complex

subunits.

Recognition of Nucleosome Modification Statusby Fbxl11/KDM2ATo provide independent validation of the SNAP approach, we

investigated in greater detail the modulation of binding of

Fbxl11/KDM2A by DNA and histone methylation. This enzyme

is a JmjC domain protein that demethylates lysine 36 on histone

H3 (Tsukada et al., 2006). Our data show that KDM2A is enriched

on H3K9me3-modified nucleosomes, but its recruitment is

disrupted by CpG-methylation on either free or nucleosomal

DNA (Figure 2E).

KDM2A has several described isoforms, and in our initial

SNAP experiments, some identified KDM2A peptides showed

a markedly lower enrichment than others. The H3K9me3-nucle-

osome SILAC pull-down was repeated to assign the identified

peptides to gel bands covering different molecular weights.

Most peptides were detected in a band corresponding to

a molecular weight of 60–75 kDa and mapped to the C-terminal

half of KDM2A (Figures S4A and S4B). Probing for the binding of

KDM2A to modified nucleosomes by immunoblot also showed

enrichment of a lower molecular weight isoform (Figure 2F and

Figure S4C). Immunoprecipitating KDM2A from nuclear extracts

confirmed the presence of this isoform (Figure S4D). This variant

corresponds to the recently described 70 kDa isoform KDM2ASF

that is transcribed from an alternative promoter and spans the

C-terminal half of KDM2A from position 543 (Tanaka et al.,

2010).

We next sought to verify the recruitment of KDM2A to the

H3K9me3 modification seen by SNAP in a different biochemical

assay. To this end, various methylated and unmethylated nucle-

osomes or histone H3 peptides were used to isolate FLAG-

tagged full-length KDM2A from transfected 293T cell extracts.

The SILAC experiments indicated a moderate enrichment of

KDM2A on H3K9me3-nucleosomes (Figure 2E). However, we

could not detect substantial binding to either H3K9me3-modi-

fied nucleosomes (Figure 5A, lane 5) or peptides (Figure 5A,

lane 8) with the overexpressed protein. This result suggested

the possibility that KDM2A may need a second factor in order

to recognize H3K9me3. A recent study reporting the interaction

of KDM2A with all HP1 isoforms (Frescas et al., 2008) prompted

us to test whether the binding was mediated by HP1. Indeed,

addition of purified HP1a to the pull-down reactions strongly

stimulated the association of KDM2A to H3K9me3 nucleo-

somes (Figure 5A, lane 13). Using HP1a, -b, and -g showed

that the interaction could be mediated by all HP1 isoforms

(Figure 5B).

We next verified the disruptive effect of DNA methylation seen

in the SNAP experiments. KDM2A harbors a DNA-binding

module consisting of a CXXC-type zinc finger domain that was

recently demonstrated to bind unmethylated CpG residues and

to be sensitive to DNA methylation (Blackledge et al., 2010).

When FLAG-tagged KDM2A was isolated from extracts with

immobilized 601 DNA (Figure S4E), binding was abolished by

CpG methylation as expected. We also sought to establish

whether the recruitment of KDM2A to H3K9me3 nucleosomes

in the presence of HP1 could be disrupted by DNA methylation.

Lane 14 in Figure 5A clearly shows that KDM2A cannot recognize

H3K9me3 nucleosomes when the DNA is methylated. The simul-

taneous recognition of DNA and HP1 leads to a stronger associ-

ation with nucleosomes. This is indicated by a more effective

recruitment of KDM2A to H3K9me3 nucleosomes compared to

H3K9me3-modified peptides in the presence of HP1 (compare

lanes 13 and 16 in Figure 5A).

To confirm that the recruitment of KDM2A to nucleosomes

through HP1 also occurs in a physiological context, we investi-

gated whether the recently reported localization of KDM2A to

ribosomal RNA genes (rDNA) in MCF7 cells (Tanaka et al.,

2010) is dependent on HP1. Indeed, downregulation of HP1a

by siRNA results in a specific decrease of HP1a and KDM2A

binding, as assessed by chromatin immunoprecipitation (ChIP)

analysis (Figures 5C and 5D).

Together, these experiments confirm the observations made

using SNAP and show that KDM2A recognizes H3K9me3 via

HP1 and that an additional interaction component is conferred

by its recognition of DNA, which is sensitive to the state of

methylation.

Immobilized unmodified or modified nucleosomes are separately incubated with light or heavy extracts, respectively. Both pull-down reactions are pooled, and

eluted proteins are separated by SDS-PAGE. After in-gel trypsin digestion, peptides are analyzed by high-resolution MS.

(B) Results of SNAP performed with H3K9me3-modified nucleosomes containing unmethylated 601 DNA. Shown are the Log2 values of the SILAC ratios (ratio H/

L) of each identified protein for the forward (x axis) and the reverse (y axis) experiments. The identities of several interacting proteins are indicated. Subunits of the

MBD2/NuRD complex are labeled in orange.

(C) Results of SNAP performed with H3K9me3-modified nucleosomes containing CpG-methylated 601 DNA. For additional SNAP results, see Figure S2 and

Table S1.

(D) Differential recognition of nucleosomes. The graphs show the forward SILAC enrichment values (ratio H/L forward) of MeCP2, L3MBTL3, USF2, and the TFIIIC

subunit GTF3C5 on CpG-methylated DNAs and modified nucleosomes. Binding to the modified nucleosomes or DNAs is indicated in red; exclusion is indicated in

blue. If proteins were not detected (n.d.), no value is assigned.

(E) Crosstalk between DNA and histone methylation. The graphs show the SILAC enrichment values of the proteins KDM2A, UHRF1, the PRC2 subunit EED, and

the ORC subunit Orc2 as described in (D).

(F) Immobilized modified nucleosomes were incubated with an independently prepared R0K0 nuclear extract as indicated. Binding of KDM2A, UHRF1, Orc2, and

the PRC2 subunit SUZ12 was detected by immunoblot. Equal loading and modification of histone H3 were verified as in Figure 1D. The asterisk marks a cross-

reactive band recognized by the KDM2A antibody.

Cell 143, 470–484, October 29, 2010 ª2010 Elsevier Inc. 475

Page 164: Cell 101029

DISCUSSION

Proteins are localized on chromatin depending on a complex set

of cues derived from the recognition of histones and DNA in

a modified or unmodified form. Here, we present an approach

(SNAP) that allows the identification of proteins that recognize

distinct chromatin modification patterns. The SNAP method

employs modified recombinant nucleosomes to isolate proteins

from SILAC-labeled nuclear extracts and to identify them by

mass spectrometry. In this study, we have used nucleosomes

containing a combination of methylation events on DNA (CpG)

and histone H3 (K4, K9, and K27). It is apparent from our results

that proteins recognizing methylated nucleosomes can be

Table 1. Proteins Enriched or Excluded by CpG-Methylated DNA

and Nucleosomes as Identified by SNAP

Enrichment/Exclusion

(Ratio H/L Forward)

601me

DNA

603me

DNA

601me

Nuc

603me

Nuc

Enriched

Proteins

very strong

enrichment

(>10)

ZBTB33 ZBTB33 ZHX2

strong

enrichment

(5–10)

ZHX1 ZHX1

MBD2b

HOMEZ

UHRF1

moderate

enrichment

(2–5)

ZBTB9

ZHX2

ZHX3

MBD2b

MTA2b

CDK2AP1b

GATAD2Ab

FOXA1

CHD4b

ZNF295

MTA3b

HOMEZ

MTA1b

GATAD2Bb

MBD4

ZHX2

MTA2b

GATAD2Ab

MTA3b

ZHX3

CDK2AP1b

FOXA1

CHD4b

GATAD2Bb

RFXANKd

RFXAPd

MTA1b

PBX1

RFX5d

PKNOX1

FIZ1

TRIM28

ZBTB40

MeCP2

PAX6

MTERF

MBD2b

GATAD2Ab

MTA2b

MBD2b

MBD4

ZBTB12

CHD4b

MeCP2

GATAD2Bb

ZHX3

ZHX1

C14orf93

RBBP4b

RBBP7b

MTERF

PAX6

LCOR

weak

enrichment

(1.5–2)

PAX9

CHD3b

CUX1

ZNF740*

RBBP7b

POGZ

KIAA1958

UHRF1

ZNF787

MBD4

CHD3b

ZFHX3

ZBTB9*

NR2C1

MAD2B

MTA2b

MBD4

CHD4b

GATAD2Ab

PPIB

ACTR5

ZBED5

AURKA

HOXC10

JUNB

Excluded

Proteins

weak exclusion

(0.5–0.67)

ANKRD32 Atherin*

SKP1*,a

RBBP5

NUFIP1

CBFB

MSH3

RBBP5

moderate

exclusion

(0.2–0.5)

RB1

TFEB

SIX4

HES7

ZFP161

YAF2

TIGD5

ARID4B

CXXC5

SKP1a

JRK

USF2

USF1

FBXW11

RAD1

ZBTB2

MLX

SP3

HES7

TCOF1*

TFDP1

ATF1

MLL

SKP1a

RECQL

ONECUT2

ZFP161

TIGD1

RB1

E2F3

CUX1

EEDc

RUNX

RNF2a

RING1a

BANP

PRDM11

SUZ12c

NAIF1

MYC

SUB1

RMI1

TOP3A

RPA2e

NAIF1

RPA1e

RPA3e

KIAA1553

TCF7L2

RNF2a

BCORa

RING1a

BANP*

Table 1. Continued

Enrichment/Exclusion

(Ratio H/L Forward)

601me

DNA

603me

DNA

601me

Nuc

603me

Nuc

BCORL1

ZNF639

strong

exclusion

(0.1–0.2)

ZBTB25

PURB

RPA1e

RPA3*,e

RPA2e

MNT

UBF1

UBF2

EEDc

SUZ12c

VHL

E2F4

BCORa

FBXL10a

FBXL11

SUZ12c

RPA3e

SSBP1

RPA2e

RPA1e

CGGBP1

UBF2

FBXL11

PURA

UBF1

ZBTB2

ZNF639

RAD1

HUS1

PURB

BCORL1

OLA1

MAX

L3MBTL3

BCORa

FBXL10a

PCGF1a

FBXL11

SUB1

FBXL10a

very strong

exclusion

(<0.1)

E2F1

PCGF1a

ZNF395

TIMM8A

KIAA1553

bHLHB2

CGGBP1

GMEB2

GTF3C2f

BCORa

GTF3C4f

FBXL10a

PCGF1a

GTF3C1f

E2F1

DEAF1

GTF3C3f

GTF3C6f

GTF3C5f

HIF1A

CXXC5

BCORL1*

FBXL11

Syntenin1

ARNT

HES7

USF2

bHLHB2

USF1

PCGF1a

Atherin

L3MBTL3

FLYWCH1

Syntenin1

ZFP161

Table 1 shows the proteins that were enriched or excluded by CpG-meth-

ylated DNA or nucleosomes compared to the respective unmodified

species at least 1.5-fold in both the forward and reverse pull-down exper-

iments. Proteins are grouped according to their ratio H/L in the forward

experiments. Proteins marked by an asterisk are just below the threshold.

For the values of the SILAC ratios, see Table S1 and Table S2.a BCOR complex.b NuRD complex.c PRC2 complex.d Regulatory factor X.e Replication factor A complex.f TFIIIC complex.

476 Cell 143, 470–484, October 29, 2010 ª2010 Elsevier Inc.

Page 165: Cell 101029

influenced by (1) the DNA sequence (in a modified and unmodi-

fied form), (2) the configuration of the histone octamer, and (3) the

precise combination of histone and DNA modifications. Below,

we discuss these modes of engagement.

(1) Recognition of DNAThe use of two distinct DNA sequences (601 or 603) in our SNAP

experiments has identified proteins that recognize methyl-CpGs

in a sequence-specific way (e.g., ZNF295) as well as proteins

that are not sequence selective (e.g., MBD2). This suggests

that some proteins may have a promiscuous methyl-DNA recog-

nition domain (i.e., recognizing methylated CpG dinucleotides

regardless of the surrounding DNA sequence), whereas others

require a specific motif surrounding the methylated CpG site.

Analysis of factors recognizing CpG methylation for the

presence of known domains identifies a striking number of zinc

finger-containing proteins (Table S2). Our data indicate that

around 50% of proteins binding to methyl-CpG and 20% of

proteins excluded from methylated DNA and nucleosomes

harbor a zinc finger domain, a motif already known to have

methyl-CpG binding potential (Sasai and Defossez, 2009).

Of interest, the second most prevalent domain in methyl-CpG-

binding proteins (20%) is a homeobox (e.g., in HOMEZ,

PKNOX1, and ZHX proteins). Homeoboxes are known DNA-

binding domains but have not previously been demonstrated

to bind methyl-CpG. These data raise the possibility that

homeoboxes may possess a methyl-CpG recognition function.

(2) Influence of NucleosomesWhen methylated 601 or 603 DNA is incorporated into nucleo-

somes, the histone octamer appears to have an effect on the

binding of certain proteins. The TFIIIC complex cannot bind

a B box effectively in the presence of an octamer, suggesting

the need for remodeling activities for full access. The methyl-

CpG-binding protein MeCP2 is seen to bind DNA-methylated

nucleosomes but showed no binding to methyl-DNA in the

absence of a histone octamer. The USF2 transcription factor is

excluded from its binding site in the 601 DNA more strongly in

the presence of histone octamers. These examples indicate

that the histone octamer may have a steric effect on the DNA

binding of such factors or that these factors contain additional

contact points with histones, which results in an increased

affinity to nucleosomes compared to free DNA.

(3) Regulation by a Combination of DNA and HistoneMethylationProteins are able to associate with nucleosomes depending on

the precise status of DNA and histone methylation. UHRF1,

which binds cooperatively to methyl-DNA and H3K9me3, may

represent a class of proteins that have an intrinsic capacity to

recognize both modifications directly because it contains an

SRA domain that binds methylated DNA and a tandem Tudor

and a PHD domain that can bind methylated H3K9 (Hashimoto

et al., 2009). In the case of protein complexes, the recognition

of each modification may reside on separate subunits. We iden-

tified two protein complexes, ORC and PRC2, that are

influenced by both types of modification in opposite ways. The

ORC, including the LRWD1 protein, recognizes H3K9 and

H3K27 methylation in a cooperative manner with DNA methyla-

tion. This may allow for a stronger interaction of ORC with

heterochromatic regions (Pak et al., 1997; Prasanth et al.,

2004). The PRC2 complex, which recognizes H3K27 methyla-

tion, is negatively regulated by DNA methylation. This may

enable this transcriptional repressor to associate preferentially

with a specific chromatin state that is not silenced completely

and can respond to external stimuli, such as poised genes.

Finally, the KDM2A histone H3K36 demethylase can recognize

H3K9me3 indirectly via its association with HP1, and recruit-

ment is blocked when DNA is methylated. This disruptive effect

would allow the demethylase to distinguish between distinct

chromatin landscapes: it will recognize silenced genes that are

marked by H3K9 methylation and HP1, but it will not dock on

heterochromatic regions that carry both H3K9me3 and DNA

methylation. Together, these examples provide evidence that

proteins can monitor the methylation state of both histones

and DNA in order to discriminate between distinct states of

repressed chromatin.

SNAP as a Tool for Studying Chromatin ModificationCrosstalkSNAP has several advantages over the current approaches

using peptides and oligonucleotides to identify chromatin-

binding factors. One advantage is that nucleosomes provide

a more physiological substrate. Proteins may have a number of

contact points to chromatin (histone tails, histone core, DNA)

and may recognize more than one histone at a time. As a result

of this multiplicity of possible interactions, SNAP will allow the

identification of proteins whose affinity may be too weak to be

selected for by the current methods. Our results clearly identify

proteins, such as KDM2A, whose binding depends on such

a physiological nucleosomal context. A second powerful advan-

tage of SNAP is that it allows the identification of proteins that

recognize multiple independent modifications on chromatin. In

this study, we have analyzed histone modifications in combina-

tion with DNA methylation. But it is equally possible to monitor

the binding of proteins to combinations of histone modifications

either on the same histone or on different histones or to use

multiple nucleosomes. The SNAP approach is also suitable for

modified histones generated using methyl-lysine analogs (Simon

et al., 2007). But because binding affinities might be crucial for

the identification of interacting proteins, natural modified amino

acids might be more desirable. In this regard, recent successful

attempts to genetically install modified amino acids in recombi-

nant histones are very promising (Neumann et al., 2009; Nguyen

et al., 2009). In summary, our findings demonstrate that chro-

matin modification-binding proteins can recognize distinct

modification patterns in a chromatin landscape. The SNAP

approach is therefore a valuable tool for studying the mecha-

nisms by which epigenetic information encoded in chromatin

modifications can be interpreted by proteins.

EXPERIMENTAL PROCEDURES

Extract Preparation and Immunoprecipitation

HeLa S3 cells were grown in suspension in RPMI 1640 medium containing 5%

FBS and normal arginine and lysine or 5% dialyzed FBS and heavy

Cell 143, 470–484, October 29, 2010 ª2010 Elsevier Inc. 477

Page 166: Cell 101029

Table 2. Nucleosome-Binding Proteins Regulated by CpG and Lysine Methylation as Identified by SNAP

Enrichment/Exclusion

(Ratio H/L Forward)

H3K4me3/601

Nuc

H3K4me3/601me

Nuc

H3K9me3/601

Nuc H3K9me3/601me Nuc

H3K27me3/601

Nuc

H3K27me3/601me

Nuc

Enriched

Proteins

very strong

enrichment (>10)

Spindlin1 IWS1h

Spindlin1

CBX5/HP1a

UHRF1

UHRF1

strong

enrichment (5–10)

PHF8

CHD1

PHF8 CBX3/HP1g

CDYL2

CBX5/HP1a

Orc4c

Orc2c

Orc3c

Orc5c

LRWD1

MeCP2

moderate

enrichment (2–5)

DIDO1

UBF1

Sin3Af

PAX6

CHD1

MeCP2

MTERF

MBD2b

DIDO1

Orc2c

Orc4c

MBD4

LRWD1

CDYL

FBXL11

UBF1

Orc2c

Orc4c

Orc5c

Orc3c

PAX6

CBX3/HP1g

CDYL

MTERF

MBD2b

Orc1c

C17orf96

LRWD1

EEDd

Orc4c

Orc5c

SUZ12d

Orc2c

Orc3c

EZH2d

MTF2

CBX8

LRWD1

Orc2c

Orc3c

Orc4c

Orc5c

MeCP2

CBX8

UHRF1

PAX6

MTERF

Orc1c

weak

enrichment (1.5–2)

SAP30f

WDR82

EMG1

TAF9B

PPIB

VRK2

HNRNPA1*

HNRNPA2B1*

ING4

WDR61

HNRNPA0*

FLYWCH1

BUB3

FUBP3

Orc5c

LRWD1

PPIB

ING4

TOX4

MTA2b

CHD4b

ZSCAN21

Orc3c

NONO

CDCA7L*

WDR82*

CHD1

SUZ12d

EEDd

PPIB

NONO

MTF2

SUB1

MTA2b

MBD4

ZSCAN21

CHD4b

NSD3

PPIB CDCA7L

BMI1

PPIB

MTA2b

MBD4*

Excluded

Proteins

weak exclusion

(0.5–0.67)

SKP1a

RCOR1

SKP1a

CREB1

HCFC1

PHF14

SKP1a

moderate

exclusion (0.2–0.5)

HMG20A

HMG20B

MTF2*

RING1a

SUB1

HMG20B

NAIF

MYC

IMP4 RCOR1

BANP

RING1a

SUB1

EEDd

TIGD5

RNF2a

MYC

NAIF1

ARNT

TCF7L2

HES7

SPTH16g

SSRP1g

TCF7L2

BANP*

PRDM11

NAIF1

RPA1e

BANP*

SUB1

strong

exclusion (0.1–0.2)

PHF14 FBXL10a

PHF14

BCORa

PCGF1a

MAX

CXXC5

L3MBTL3

FBXL10a

BCORa

RPA2e

BCORa

MYC

FBXL10a

PCGF1a

MAX

very strong

exclusion

(<0.1)

L3MBTL3

ARNT

FBXL11

PCGF1a

HIF1A

Syntenin1

L3MBTL3

HES7

Syntenin1

478 Cell 143, 470–484, October 29, 2010 ª2010 Elsevier Inc.

Page 167: Cell 101029

arginine-13C6, 15N4 and lysine-13C6, 15N2 (Isotec). Cells were harvested at

a density of 0.5–0.8 3 106 cells/ml, and nuclear extracts were essentially

prepared as described (Dignam et al., 1983). For both SILAC extracts, three

independent nuclear extracts were prepared and pooled to yield an ‘‘average’’

extract that compensates for differences in each individual preparation. 293T

and MFC7 cells were grown in DMEM medium supplemented with 10% FBS.

293T cells were transfected using a calcium phosphate protocol. Whole-cell

extracts were prepared �36 hr after transfection by rotating the cells in extrac-

tion buffer (20 mM HEPES [pH 7.5], 300 mM NaCl, 1 mM EDTA, 20% Glycerol,

0.5% NP40, 1 mM DTT, and complete protease inhibitors [Roche]) for 1 hr at

4�C. HeLa S3 nuclear extracts and 293T or MCF7 whole-cell extracts were

snap frozen and stored in aliquots at �80�C. For coimmunoprecipitations,

extracts were prepared without DTT and diluted 1:1 with 20 mM HEPES

(pH 7.5), 1 mM EDTA, and 20% Glycerol containing complete protease inhib-

itors. Extracts were precleared and proteins immunoprecipitated with typically

5 mg of antibody and Protein-G Sepharose (GE Healthcare) or 20 ml anti-FLAG

M2 agarose (Sigma).

Chromatin Immunoprecipitation and Immunofluorescence

For ChIPs, MCF7 cells were reverse transfected with siRNAs against HP1a or

negative control siRNA using Lipofectamine RNAiMAX (Invitrogen) according

to the manufacturer’s protocol. At 48 hr after transfection, cells were washed

twice with PBS, fixed with 1% formaldehyde (Sigma) in PBS at room temper-

ature for 10 min, and quenched with 125 mM Glycine for 5 min. After three

washes with 10 ml of cold PBS, cells were harvested in cold PBS supple-

mented with complete protease inhibitor cocktail by scraping. Pellets from

two 10 cm dishes were suspended in 1.6 ml of RIPA buffer (50 mM Tris-HCl

(pH 8), 150 mM NaCl, 2 mM EDTA, 1% NP-40, 0.5% sodium deoxycholate,

and 0.1% SDS supplemented with EDTA-free complete protease inhibitors),

sonicated in 15 ml conical tubes three times for 10 min at high 30 s on/off

cycles in a cooled Bioruptor (Diagenode), and cleared by centrifugation for

15 min at 13,000 rpm. ChIPs were then performed as described (Xhemalce

and Kouzarides, 2010). The PCR analysis was performed on a StepOnePlus

Real-Time PCR System using Fast SYBR Green (Applied Biosystems). For

IFs, MCF7 cells were grown in slide flasks, washed with PBS, treated for

5 min on ice with CSK buffer (10 mM PIPES [pH 6.8], 100 mM NaCl, 300 mM

sucrose, 3 mM MgCl2, 1 mM EGTA, and 0.5% Triton), washed again with

PBS, and fixed with 5% Formalin solution (Sigma) in PBS/2% sucrose. The

fixed cells were incubated O/N at 4�C with 0.5 mg/ml of each primary antibody

and for 1 hr at RT with DAPI and the secondary antibodies. Images were

acquired with an Olympus FV1000 Upright confocal microscope and pro-

cessed using Adobe Photoshop CS software.

Protein Expression and Purification

Recombinant histone proteins were expressed in E. coli BL21(DE3)/RIL cells

from pET21b(+) (Novagen) vectors and purified by denaturing gel filtration

and ion exchange chromatography essentially as described (Dyer et al.,

2004). Truncated H3.1D1-31T32C protein was generated in vivo by expressing

a H3.1D1-31T32C precursor in the presence of TEV-protease. For this

purpose, E. coli cells harboring the pET28a(+)-AraC-PBAD-His6TEV/pro-

H3.1D1-31T32C plasmid were grown in LB medium containing 0.25% L-arab-

inose to keep TEV-protease induced. At an OD600 of 0.6 the expression of

pro-hH3.1D1-31T32C was induced for 3 hr at 37�C with 50 mM IPTG. TEV-

protease processes the precursor histone H3.1 into tail-less H3.1D1-

31T32C. The insoluble protein was extracted from inclusion bodies with solu-

bilization buffer (20 mM Tris [pH 7.5], 7 M Guanidine HCl, and 100 mM DTT) for

1 hr at RT and passed over a Sephacryl S200 gel filtration column (GE Health-

care) in SAU-200 (20 mM NaAcetate [pH 5.2], 7 M Urea, 200 mM NaCl, and

1 mM EDTA) without any reducing agents. Positive fractions were directly

loaded onto a reversed-phase ResourceRPC column (GE Healthcare) and

eluted with a gradient of 0%–65% B (A: 0.1% TFA in water, B: 90% Acetoni-

trile; 0.1% TFA) over 20 column volumes. Fractions containing pure

H3.1D1-31T32C were pooled and lyophilized. All histone proteins were stored

lyophilized at �80�C. Recombinant HP1 GST-fusion proteins were expressed

in E. coli BL21(DE3)/RIL cells and purified by glutathione Sepharose

(GE Healthcare) chromatography. HP1 proteins were cleaved off the beads

with biotinylated thrombin (Novagen). After removal of thrombin with strepta-

vidin Sepharose, HP1 proteins were dialyzed into TBS/10% glycerol, snap

frozen, and stored at �80�C.

Preparation of Modified Histones and Nucleosomal DNAs

For native chemical ligations, lyophilized modified H3.1 1-31 thioester peptide

(Almac) was incubated at a concentration of 0.56 mg/ml (�0.167 mM) with

truncated H3.1D1-31T32C protein at 4 mg/ml (�0.333 mM) and thiophenol

at 2% (v/v) in ligation buffer (6 M Guanidine HCl and 200 mM KPO4 [pH 7.9]).

The cloudy mixture was left shaking vigorously at RT for 24 hr. The reaction

was stopped by adding DTT to a final concentration of 100 mM, dialyzed three

times against SAU-200 buffer containing 5 mM 2-Mercaptoethanol, and then

loaded onto a Hi-Trap SP HP column (GE-Healthcare). The ligated Histone

H3 was eluted with a linear gradient from SAU-200 to SAU-600 buffer

(20 mM NaAcetate [pH 5.2], 7 M Urea, 600 mM NaCl, 1 mM EDTA, and

5 mM 2-Mercaptoethanol). Positive fractions were pooled, diluted 3-fold in

SAU-0 buffer (20 mM NaAcetate [pH 5.2], 7 M Urea, 1 mM EDTA, and 5 mM

2-Mercaptoethanol) to reduce the NaCl concentration, and reloaded onto

the column. Three rounds of purification were needed to yield sufficiently

Table 2. Continued

Enrichment/Exclusion

(Ratio H/L Forward)

H3K4me3/601

Nuc

H3K4me3/601me

Nuc

H3K9me3/601

Nuc H3K9me3/601me Nuc

H3K27me3/601

Nuc

H3K27me3/601me

Nuc

Syntenin1

Atherin

USF2

USF1

HIF1A*

bHLHB2

FBXL11

Atherin

USF1

USF2

bHLHB2

HIF1A

Atherin

ARNT

FBXL11

USF1

USF2

bHLHB2

Table 2 shows the proteins that were enriched or excluded by modified nucleosomes compared to unmodified nucleosomes at least 1.5-fold in both

the forward and reverse pull-down experiments. Proteins are grouped according to their ratio H/L in the forward experiments. Proteins marked by an

asterisk are just below the threshold. For the values of the SILAC ratios, see Table S1 and Table S2. Fbxl11/KDM2A is italicized.a BCOR complex.b NuRD complex.c ORC complex.d PRC2 complex.e Replication factor A complex.f Sin3A complex.g FACT.h IWS should be treated with caution because it was found as a false positive outlier in the 601me-Nuc pull-down.

Cell 143, 470–484, October 29, 2010 ª2010 Elsevier Inc. 479

Page 168: Cell 101029

Figure 3. Interaction Profiles of Chromatin Modification-Binding Proteins

Agglomerative hierarchical clustering was performed on the SILAC enrichment values of proteins regulated by DNA and histone methylation to identify proteins

with related binding profiles. This analysis includes proteins based on an enrichment/exclusion of at least 1.5-fold in both directions in one of the nucleosome pull-

down experiments and excludes factors that were found solely in the DNA pull-downs. Log2(ratiofor/ratiorev) is the log2 ratio between the SILAC values (ratio H/L)

of the forward and reverse experiments. Enrichment by modifications is indicated in red; exclusion is indicated in blue. Gray bars indicate whether proteins were

not detected (n.d.) in particular experiments. These incidences were not included in the cluster analysis. Clusters of several known protein complexes and their

respective subunits are indicated on the right. For values, see Table S2.

480 Cell 143, 470–484, October 29, 2010 ª2010 Elsevier Inc.

Page 169: Cell 101029

pure ligated histone. Following ion exchange purification, the ligated histone

was dialyzed against water containing 1 mM DTT, lyophilized, and stored

at �80�C. Nucleosomal 601 or 603 DNAs were excised from purified plasmid

DNAs (Plasmid Giga Kit, QIAGEN) by digestion with EcoRV and separated

from the vector by PEG precipitation as described (Dyer et al., 2004). For

end biotinylation, the DNA was further digested with EcoRI and the overhangs

filled in with biotin-11-dUTP (Yorkshire Bioscience) using Klenow (30/50 exo�) polymerase (NEB). Nucleosomal biotinylated DNAs were then sepa-

rated by PEG precipitation or further methylated with M.SssI CpG Methyltrans-

ferase (NEB) and then PEG precipitated to remove small cleavage products.

Reconstitution of Nucleosomes and Nucleosome Pull-Downs

Octamers were refolded from purified histones and assembled into nucleo-

somes with biotinylated nucleosomal DNAs by salt deposition as described

(Dyer et al., 2004). Optimal reconstitution conditions were determined by titra-

tion and then kept constant for all nucleosome assembly reactions.

Figure 4. LRWD1 Interacts with the Origin Recognition Complex

(A) LRWD1 colocalizes with Orc2. IF staining of MCF7 cells with LWRD1 (2527) and Orc2 antibodies following pre-extraction shows colocalization at distinct

nuclear foci.

(B) LRWD1 and ORC coimmunoprecipitate. LRWD1 and Orc2 were immunoprecipitated from MCF7 whole-cell extracts, and interacting proteins were detected

by immunoblot as indicated. LRWD1 was immunoprecipitated using anti-LRWD1 (A301-867A) and detected using anti-LRWD1 (2527) antibodies. Anti-FLAG and

anti-GFP antibodies were used as IgG negative controls. Asterisks mark bands derived from antibody heavy chains.

(C) FLAG-tagged full-length and truncated versions of LRWD1 were overexpressed in 293T cells and immunoprecipitated using an anti-FLAG antibody. 1% of the

input and 10% of the IP were separated by SDS-PAGE, and Orc1, Orc2, and the FLAG fusions were detected by immunoblot. The asterisks mark bands derived

from the anti-FLAG IP antibody.

(D) Identities of the LRWD1 truncation constructs. Only deletions containing the WD40 repeats interact with ORC.

(E) LRWD1 expression is Orc2 dependent. Expression levels of LRWD1 and ORC proteins in MCF7 cells were detected by immunoblot after transfection with

siRNAs against LRWD1 and Orc2 as indicated. Cells were reverse transfected twice, 56 hr and 28 hr before harvesting. GAPDH serves as a loading control.

The asterisk marks a cross-reactive band detected by the anti-LRWD1 (2527) antibody.

See also Figure S3.

Cell 143, 470–484, October 29, 2010 ª2010 Elsevier Inc. 481

Page 170: Cell 101029

Nucleosomes were checked on 5% native PAGE gels. For SILAC pull-downs,

nucleosomes corresponding to 12.5 mg of octamer were immobilized on 75 ml

Dynabeads Streptavidin MyOne T1 (Invitrogen) in the final reconstitution buffer

(10 mM Tris [pH 7.5], 250 mM KCl, 1 mM EDTA, and 1 mM DTT; supplemented

with 0.1% NP40) and then rotated with 0.5 mg HeLa S3 nuclear extract in 1 ml

of binding buffer (20 mM HEPES [pH 7.9], 150 mM NaCl, 0.2 mM EDTA, 20%

Glycerol, 0.1% NP40, 1 mM DTT, and complete protease inhibitors) for 4 hr at

4�C. After five washes with 1 ml of binding buffer, the beads from both SILAC

pull-downs were pooled, and bound proteins were eluted in sample buffer and

analyzed on 4%–12% gradient gels by colloidal blue staining (NuPAGE/NO-

VEX, Invitrogen). For DNA and peptide pull-downs, streptavidin-coated

magnetic beads were saturated with either biotinylated 601 DNA or H3

peptides (residues 1–21) and then used as described for the nucleosome

beads.

Mass Spectrometry of Proteins and Computational Analyses

Nucleosome-bound proteins resolved on SDS-PAGE gels were subjected to

in-gel trypsin digestion as described (Vermeulen et al., 2010). Peptide identifi-

cation experiments were performed using an EASY nLC system (Proxeon)

connected online to an LTQ-FT Ultra mass spectrometer (Thermo Fisher,

Germany). Tryptic peptide mixtures were loaded onto a 15 cm long 75 mm

ID column packed in house with 3 mm C18-AQUA-Pur Reprosil reversed-

phase beads (Dr. Maisch GmbH) and eluted using a 2-h linear gradient from

8% to 40% acetonitrile. The separated peptides were electrosprayed directly

into the mass spectrometer, which was operated in the data-dependent mode

to automatically switch between MS and MS2. Intact peptide spectra were

acquired with 100,000 resolution in the FT cell while acquiring up to five

tandem mass spectra in the LTQ part of the instrument. Proteins were identi-

fied and quantified by analyzing the raw data files using the MaxQuant

Figure 5. Fbxl11/KDM2A Integrates DNA Methylation and H3K9me3 Modification Signals on Nucleosomes

(A) In vitro binding of KDM2A to modified nucleosomes. Whole-cell extracts prepared from transiently transfected 293T cells overexpressing FLAG-tagged

KDM2A were incubated with immobilized modified nucleosomes or modified H3 peptides as indicated. Binding reactions were supplemented with recombinant

purified HP1a or GST as a control. Binding was detected by immunoblot against the FLAG tag or HP1a. Equal loading of the nucleosomes and peptides and

modification of histone H3 were verified as in Figure 1D.

(B) KDM2A binding to H3K9me3 nucleosomes is mediated by HP1a, -b, and -g. Unmodified or H3K9me3-modified nucleosomes were immobilized on strepta-

vidin beads and incubated with 293T whole-cell extracts overexpressing FLAG-tagged KDM2A. Pull-down reactions were supplemented with recombinant puri-

fied HP1a, -b, or -g or GST as indicated. Binding of KDM2A was detected by immunoblot against the FLAG tag.

(C) Recruitment of KDM2A to the rDNA locus is augmented by HP1a. MCF7 cells were transfected with HP1a-specific siRNAs and analyzed for the enrichment of

the H13 region of the rDNA locus by ChIP using antibodies against KDM2A, HP1a, and histone H3K9me3. Shown are the mean ± SD of the signals normalized to

input of three independent experiments. KDM2A shows only little enrichment at the GAPDH locus.

(D) Analysis of KDM2A and HP1a expression in siRNA-treated MCF7 cells by immunoblot. GAPDH serves as loading control.

See also Figure S4.

482 Cell 143, 470–484, October 29, 2010 ª2010 Elsevier Inc.

Page 171: Cell 101029

software, version 1.0.12.5, in combination with the Mascot search engine

(Matrix Science), essentially as described (Vicent et al., 2009). The raw data

from all forward and reverse pull-downs were processed together and filtered

such that a protein was only accepted when it was quantified with at least two

peptides, both in the forward and the reverse pull-down. Results from the pull-

downs were visualized using the open-source software package R. For the

cluster analysis, the log2 ratio between the forward and reverse SILAC values

(ratio H/L) of each protein was calculated. These data were clustered to iden-

tify related clades of proteins. Clustering was performed in R using the hopach

package (van der Laan and Pollard, 2003). The distance between pairwise log2

ratio values was calculated using the absolute uncentered correlation

distance, and agglomerative hierarchical clustering using complete linkage

was performed.

Deposition of MS-Related Data

The MS raw data files for nucleosome pull-downs can be accessed via

TRANCHE (https://proteomecommons.org/) under the name ‘‘SILAC Nucleo-

some Affinity Purification.’’

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures, four

figures, and two tables and can be found with this article online at doi:10.

1016/j.cell.2010.10.012.

ACKNOWLEDGMENTS

We would like to thank Kevin Ford, Timothy Richmond, Bruce Stillman,

Jonathan Widom, and Yi Zhang for providing materials; Helder Ferreira and

Tom Owen-Hughes for advice on native chemical ligations; and Peter Tessarz

and Emmanuelle Vire for experimental help. This work was supported by post-

doctoral fellowships to T.B. from EMBO and HFSP and by a fellowship to

M.V. from the Dutch Cancer Society. The M.M. laboratory is supported by

the Max-Planck Society for the Advancement of Science and HEROIC, a grant

from the European Union under the 6th Research Framework Programme. The

T.K. lab is funded by grants from Cancer Research UK and the European Union

(Epitron, HEROIC, and SMARTER). T.K. is a director of Abcam Ltd.

Received: February 10, 2010

Revised: September 28, 2010

Accepted: October 8, 2010

Published: October 28, 2010

REFERENCES

Bannister, A.J., Zegerman, P., Partridge, J.F., Miska, E.A., Thomas, J.O.,

Allshire, R.C., and Kouzarides, T. (2001). Selective recognition of methylated

lysine 9 on histone H3 by the HP1 chromo domain. Nature 410, 120–124.

Bernstein, B.E., Meissner, A., and Lander, E.S. (2007). The mammalian epige-

nome. Cell 128, 669–681.

Blackledge, N.P., Zhou, J.C., Tolstorukov, M.Y., Farcas, A.M., Park, P.J., and

Klose, R.J. (2010). CpG islands recruit a histone H3 lysine 36 demethylase.

Mol. Cell 38, 179–190.

Dignam, J.D., Lebovitz, R.M., and Roeder, R.G. (1983). Accurate transcription

initiation by RNA polymerase II in a soluble extract from isolated mammalian

nuclei. Nucleic Acids Res. 11, 1475–1489.

Dyer, P.N., Edayathumangalam, R.S., White, C.L., Bao, Y., Chakravarthy, S.,

Muthurajan, U.M., and Luger, K. (2004). Reconstitution of nucleosome core

particles from recombinant histones and DNA. Methods Enzymol. 375, 23–44.

Frescas, D., Guardavaccaro, D., Kuchay, S.M., Kato, H., Poleshko, A., Basrur,

V., Elenitoba-Johnson, K.S., Katz, R.A., and Pagano, M. (2008). KDM2A

represses transcription of centromeric satellite repeats and maintains the

heterochromatic state. Cell Cycle 7, 3539–3547.

Gearhart, M.D., Corcoran, C.M., Wamstad, J.A., and Bardwell, V.J. (2006).

Polycomb group and SCF ubiquitin ligases are found in a novel BCOR complex

that is recruited to BCL6 targets. Mol. Cell. Biol. 26, 6880–6889.

Hansen, K.H., Bracken, A.P., Pasini, D., Dietrich, N., Gehani, S.S., Monrad, A.,

Rappsilber, J., Lerdrup, M., and Helin, K. (2008). A model for transmission of

the H3K27me3 epigenetic mark. Nat. Cell Biol. 10, 1291–1300.

Hashimoto, H., Horton, J.R., Zhang, X., and Cheng, X. (2009). UHRF1,

a modular multi-domain protein, regulates replication-coupled crosstalk

between DNA methylation and histone modifications. Epigenetics 4, 8–14.

Karagianni, P., Amazit, L., Qin, J., and Wong, J. (2008). ICBP90, a novel methyl

K9 H3 binding protein linking protein ubiquitination with heterochromatin

formation. Mol. Cell. Biol. 28, 705–717.

Kleine-Kohlbrecher, D., Christensen, J., Vandamme, J., Abarrategui, I., Bak,

M., Tommerup, N., Shi, X., Gozani, O., Rappsilber, J., Salcini, A.E., and Helin,

K. (2010). A functional link between the histone demethylase PHF8 and the

transcription factor ZNF711 in X-linked mental retardation. Mol. Cell 38,

165–178.

Kouzarides, T. (2007). Chromatin modifications and their function. Cell 128,

693–705.

Le Guezennec, X., Vermeulen, M., Brinkman, A.B., Hoeijmakers, W.A., Cohen,

A., Lasonder, E., and Stunnenberg, H.G. (2006). MBD2/NuRD and MBD3/

NuRD, two distinct complexes with different biochemical and functional prop-

erties. Mol. Cell. Biol. 26, 843–851.

Lowary, P.T., and Widom, J. (1998). New DNA sequence rules for high affinity

binding to histone octamer and sequence-directed nucleosome positioning. J.

Mol. Biol. 276, 19–42.

Luger, K., Mader, A.W., Richmond, R.K., Sargent, D.F., and Richmond, T.J.

(1997). Crystal structure of the nucleosome core particle at 2.8 A resolution.

Nature 389, 251–260.

Muir, T.W. (2003). Semisynthesis of proteins by expressed protein ligation.

Annu. Rev. Biochem. 72, 249–289.

Neumann, H., Hancock, S.M., Buning, R., Routh, A., Chapman, L., Somers, J.,

Owen-Hughes, T., van Noort, J., Rhodes, D., and Chin, J.W. (2009). A method

for genetically installing site-specific acetylation in recombinant histones

defines the effects of H3 K56 acetylation. Mol. Cell 36, 153–163.

Nguyen, D.P., Garcia Alai, M.M., Kapadnis, P.B., Neumann, H., and Chin, J.W.

(2009). Genetically encoding N(epsilon)-methyl-L-lysine in recombinant

histones. J. Am. Chem. Soc. 131, 14194–14195.

Pak, D.T., Pflumm, M., Chesnokov, I., Huang, D.W., Kellum, R., Marr, J.,

Romanowski, P., and Botchan, M.R. (1997). Association of the origin recogni-

tion complex with heterochromatin and HP1 in higher eukaryotes. Cell 91,

311–323.

Prasanth, S.G., Prasanth, K.V., Siddiqui, K., Spector, D.L., and Stillman, B.

(2004). Human Orc2 localizes to centrosomes, centromeres and heterochro-

matin during chromosome inheritance. EMBO J. 23, 2651–2663.

Pray-Grant, M.G., Daniel, J.A., Schieltz, D., Yates, J.R., III, and Grant, P.A.

(2005). Chd1 chromodomain links histone H3 methylation with SAGA- and

SLIK-dependent acetylation. Nature 433, 434–438.

Ruthenburg, A.J., Li, H., Patel, D.J., and Allis, C.D. (2007). Multivalent engage-

ment of chromatin modifications by linked binding modules. Nat. Rev. Mol.

Cell Biol. 8, 983–994.

Sasai, N., and Defossez, P.A. (2009). Many paths to one goal? The proteins

that recognize methylated DNA in eukaryotes. Int. J. Dev. Biol. 53, 323–334.

Shogren-Knaak, M.A., Fry, C.J., and Peterson, C.L. (2003). A native peptide

ligation strategy for deciphering nucleosomal histone modifications. J. Biol.

Chem. 278, 15744–15748.

Simon, M.D., Chu, F., Racki, L.R., de la Cruz, C.C., Burlingame, A.L., Panning,

B., Narlikar, G.J., and Shokat, K.M. (2007). The site-specific installation of

methyl-lysine analogs into recombinant histones. Cell 128, 1003–1012.

Tanaka, Y., Okamoto, K., Teye, K., Umata, T., Yamagiwa, N., Suto, Y., Zhang,

Y., and Tsuneoka, M. (2010). JmjC enzyme KDM2A is a regulator of rRNA tran-

scription in response to starvation. EMBO J. 29, 1510–1522.

Cell 143, 470–484, October 29, 2010 ª2010 Elsevier Inc. 483

Page 172: Cell 101029

Taverna, S.D., Li, H., Ruthenburg, A.J., Allis, C.D., and Patel, D.J. (2007). How

chromatin-binding modules interpret histone modifications: lessons from

professional pocket pickers. Nat. Struct. Mol. Biol. 14, 1025–1040.

Tolbert, T.J., and Wong, C.H. (2002). New methods for proteomic research:

preparation of proteins with N-terminal cysteines for labeling and conjugation.

Angew. Chem. Int. Ed. Engl. 41, 2171–2174.

Tsukada, Y., Fang, J., Erdjument-Bromage, H., Warren, M.E., Borchers, C.H.,

Tempst, P., and Zhang, Y. (2006). Histone demethylation by a family of JmjC

domain-containing proteins. Nature 439, 811–816.

van der Laan, M.J., and Pollard, K.S. (2003). A new algorithm for hybrid hierar-

chical clustering with visualization and the bootstrap. J. Statist. Plann. Infer-

ence 117, 275–303.

Vermeulen, M., Eberl, H.C., Matarese, F., Marks, H., Denissov, S., Butter, F.,

Lee, K.K., Olsen, J.V., Hyman, A.A., Stunnenberg, H.G., and Mann, M.

(2010). Quantitative interaction proteomics and genome-wide profiling of

epigenetic histone marks and their readers. Cell 142, 967–980.

Vicent, G.P., Zaurin, R., Nacht, A.S., Li, A., Font-Mateu, J., Le Dily, F., Vermeu-

len, M., Mann, M., and Beato, M. (2009). Two chromatin remodeling activities

cooperate during activation of hormone responsive promoters. PLoS Genet. 5,

e1000567.

Xhemalce, B., and Kouzarides, T. (2010). A chromodomain switch mediated by

histone H3 Lys 4 acetylation regulates heterochromatin assembly. Genes Dev.

24, 647–652.

Zhao, Q., Qin, L., Jiang, F., Wu, B., Yue, W., Xu, F., Rong, Z., Yuan, H., Xie, X.,

Gao, Y., et al. (2007). Structure of human spindlin1. Tandem tudor-like

domains for cell cycle regulation. J. Biol. Chem. 282, 647–656.

484 Cell 143, 470–484, October 29, 2010 ª2010 Elsevier Inc.

Page 173: Cell 101029

Retraction

Retraction Notice to: Assembly ofEndogenous oskar mRNA Particles forMotor-Dependent Transport inthe Drosophila OocyteAlvar Trucco, Imre Gaspar, and Anne Ephrussi**Correspondence: [email protected]

DOI 10.1016/j.cell.2010.10.011

(Cell 139, 983–998; November 25, 2009)

In this paper, we used cryoimmuno-electron microscopy and live-cell imaging to investigate the sequential assembly of oskar mRNA

into an mRNP competent for transport from the Drosophila nurse cells to the oocyte posterior pole. We have recently identified

instances in all of the figures where the cryoimmuno-EM data were inappropriately manipulated by the first author. The manipulations

do not affect the live-cell imaging data. We are in the process of reanalyzing the raw experimental cryoimmuno-EM data but can

already state that the published conclusions are not fully consistent with the raw data. We are therefore retracting the paper. We

sincerely apologize for any inconvenience that this might have caused.

Cell 143, 485, October 29, 2010 ª2010 Elsevier Inc. 485

Page 174: Cell 101029
Page 175: Cell 101029

Scientific Editor, Cell PressCell Press seeks to appoint three Scientific Editors with dual roles covering scientific editing and the review material. These positions will be associated with the Cell Press titles Cancer Cell, Current Biology, Developmental Cell, and Neuron, and expertise in any of the relevant areas covered by these journals will be considered. Working closely with the research community, you will be acquiring, managing, and developing new editorial content for the Cell Press research titles. These positions will also work closely with other aspects of the business, including production, business development, marketing, and commercial sales, and, therefore, provide an excellent entry opportunity to science publishing. You will work as part of a highly dynamic and collaborative editorial group in the Cambridge, MA office. These positions are an exciting opportunity to stay at the forefront of the latest scientific advances while developing a new career in an exciting publishing environment.

Minimum qualifications are a PhD in a relevant life science discipline, and additional postdoctoral or other experience is a plus. Ideal candidates would have a strong scientific background and broad research interests, excellent writing and communica-tion skills, strong organizational and interpersonal skills, as well as creative energy and enthusiasm for science and science communication. Prior publishing or editorial experience is an advantage but is not a requirement.

To apply Please submit to the url below a CV and cover letter explaining your interest in an editorial position and describing your qualifications, research interests, and reasons for pursuing a career in scientific publishing. Applications will be accepted on an ongoing basis through December 1, 2010.

http://reedelsevier.taleo.net/careersection/51/jobdetail.ftl?lang=en&job=SCI00063.

No phone inquiries. Elsevier-Cell Press is an Equal Opportunity Employer.

C

M

Y

CM

MY

CY

CMY

K

EditorAd_CP.pdf 1 10/12/10 4:40 PM

Page 176: Cell 101029

Scientific Editor, Molecular CellMolecular Cell is seeking a full-time scientific editor to join its editorial team. We will consider qualified candidates with scientific expertise in any area that the journal covers. The minimum qualification for this position is a PhD in a relevant area of biomedical research, although additional experience is preferred. This is a superb opportunity for a talented individual to play a critical role in the research community away from the bench.

As a scientific editor, you would be responsible for assessing submitted research papers, overseeing the refereeing process, and choosing and commissioning review material. You would also travel frequently to scientific conferences to follow develop-ments in research and establish and maintain close ties with the scientific community. The key qualities we look for are breadth of scientific interest and the ability to think critically about a wide range of scientific issues. The successful candidate will also be highly motivated and creative and able to work independently as well as in a team.

This is a full-time in-house editorial position, based at the Cell Press office in Cambridge, Massachusetts. Cell Press offers an attractive salary and benefits package and a stimulating working environment. Applications will be held in the strictest of confidence and will be considered on an ongoing basis until the position is filled. To apply Please submit a CV and cover letter describing your qualifications, research interests, and reasons for pursuing a career in scientific publishing, as soon as possible, to our online jobs site:http://www.elsevier.com/wps/find/job_search.careers. Click on “search for US jobs” and select “Massachusetts.” Or:http://reedelsevier.taleo.net/careersection/51/jobdetail.ftl?lang=en&job=SCI0005X.

No phone inquiries, please. Cell Press is an equal opportunity/affirmative action employer, M/F/D/V.

C

M

Y

CM

MY

CY

CMY

K

EditorAd_MC.pdf 1 10/12/10 4:44 PM

Page 177: Cell 101029

Scientific Editor, Cell MetabolismCell Metabolism is seeking a full-time scientific editor to join its editorial team. Cell Metabolism publishes metabolic research with an emphasis on molecular mechanisms and translational medicine. The minimum qualification for this position is a PhD in a relevant area of biomedical research, although additional postdoctoral and/or editorial experience is preferred. This is a superb opportunity for a talented individual to play a critical role in promoting science by helping researchers shape and disseminate their findings to the wider community.

The scientific editor is responsible for assessing submitted research papers, overseeing the refereeing process, and choosing, commissioning, and editing review material. The scientific editor frequently travels to scientific conferences to follow developments in research and establish and maintain close ties with the scientific community. The key qualities we look for are breadth of scientific interest, the ability to think critically about a wide range of scientific issues, and strong communication skills. The successful candidate will also be highly motivated and creative and able to work independently as well as in a team and should have opportunities to pioneer and contribute to new trends in scientific publishing.

This is a full-time in-house editorial position, based at the Cell Press office in Cambridge, Massachusetts. Cell Press offers an attractive salary and benefits package and a stimulating working environment that encourages innovation.

Please submit a CV and cover letter describing your qualifications, general research interests, and motivation for pursuing a career in scientific publishing. Applications will be considered on an ongoing basis until the closing date of November 15th, 2010.

To apply, visit http://reedelsevier.taleo.net/careersection/51/jobdetail.ftl?lang=en&job=SCI0005Y.

No phone inquiries. Elsevier-Cell Press is an Equal Opportunity Employer.

C

M

Y

CM

MY

CY

CMY

K

CellMetEditorAd.pdf 1 10/8/10 2:05 PM

Page 178: Cell 101029

The American Society of Human Genetics is seeking an Editor for The American Journal of Human Genetics. The Editor leads one of the world’s oldest and most prestigious journals publishing pri-mary human genetics research.

Among the Editor’s responsibilities are determining the scope and direction of the scientific con-tent of The Journal, overseeing manuscripts submitted for review and their publication, selecting and supervising a staff consisting of an Editorial Assistant and doctoral-level Deputy Editor, direct-ing interactions with the publisher (currently Cell Press), reviewing quarterly reports provided by the publisher, evaluating the performance of the publisher, and if required, supervising the process of the selection a new publisher. The Editor serves as a member of the Board of Directors of the Ameri-can Society of Human Genetics (ASHG), as well as the ASHG Finance Committee, and presents semiannual reports to the Board. All Associate Editors of The Journal are appointed by the Editor, who also determines their duties. At the ASHG annual meeting, the Editor presides over a meeting of the Associate Editors and presents an annual report to the ASHG membership.

The term of the appointment is five years and includes a yearly stipend. The new Editor will be selected by the end of 2010 and will begin receiving manuscripts approximately in September 2011; there will be partial overlap with the Boston office. Applicants should be accomplished scientists in the field of human genetics and should have a broad knowledge and appreciation of the field. Nominations, as well as applications consisting of a letter of interest and curriculum vitae, should be sent to:

AJHG Editorial Search CommitteeAmerican Society of Human Genetics9650 Rockville PikeBethesda, MD 20814

The American Journal of Human Genetics Editor Position Available

editorad.indd 1 5/7/2010 12:25:11 PM

Page 179: Cell 101029

Cell Press is seeking a Business Project Editor to plan, develop, and implement projects that have commercial or sponsorship potential. By drawing on existing content or developing new material, the Editor will work with Cell Press’s commercial sales group to create collections of content in print or online that will be attractive to readers and sponsors. The Editor will also be responsible for leverag-ing new online opportunities for engaging the readers of Cell Press journals.

The successful candidate will have a PhD in the biological sciences, broad scientific interests, a

fascination with technology, good commercial instincts, and a true passion for both science and science communication. They should be highly organized and dedicated, with excellent written and oral communication skills, and should be willing to work to tight deadlines.

The position is full time and based in Cambridge, MA. Cell Press offers an attractive salary and

benefits package and a stimulating work environment. Applications will be considered on a rolling basis. For consideration, please apply online and include a cover letter and resume. To apply, visit the career page at http://www.elsevier.com and search on keywords “Business Project Editor.”

Cell Press Business Project Editor Position Available

businessprojecteditor.indd 1 8/4/2010 3:00:21 PM

Page 180: Cell 101029

23Brain Research take another look

www.elsevier.com/locate/brainres

One re-unified journal, nine specialist sections, 23 receiving Editors ←Authors receive first editorial decision within 30 days of submission ←

“Young Investigator Awards” for innovative work by a new generation of researchers ←

1

EDITOR-IN-CHIEFF.E. Bloom

La Jolla, CA, USA

SENIOR EDITORSJ.F. Baker

Chicago, IL, USAP.R. Hof

New York, NY, USAG.R. Mangun

Davis, CA, USAJ.I. Morgan

Memphis, TN, USAF.R. Sharp

Sacramento, CA, USAR.J.Smeyne

Memphis, TN, USAA.F. Sved

Pittsburgh, PA, USA

ASSOCIATE EDITORSG. Aston-Jones

Charleston, SC, USAJ.S. Baizer

Buffalo, NY, USAJ.D. Cohen

Princeton, NJ, USAB.M. Davis

Pittsburgh, PA, USAJ. De Felipe

Madrid, SpainM.A. Dyer

Memphis, TN, USAM.S. Gold

Pittsburgh, PA, USAG.F. Koob

La Jolla, CA, USA

T.A. Milner New York, NY, USA

S.D. Moore Durham, NC, USA

T.H. Moran Baltimore, MD, USA

T.F. Münte Magdeburg, Germany

K-C. Sonntag Belmont, MA, USA

R.J. Valentino Philadelphia, PA, USA

C.L. Williams Durham,NC, USA

Twenty-three tothe Power of One.

BresAd23_212X276:Ad 6/3/08 9:15 AM Page 1

Page 181: Cell 101029

cell1433cla.indd 1cell1433cla.indd 1 10/21/2010 10:59:24 PM10/21/2010 10:59:24 PM

Page 182: Cell 101029

cell1433cla.indd 2cell1433cla.indd 2 10/21/2010 10:59:44 PM10/21/2010 10:59:44 PM

Page 183: Cell 101029

Positions Available

TENURE TRACK FACULTY POSITION in NEUROSCIENCE

The Dept of Neurobiology & Anatomy at the University of Utah (http://www.neuro.utah.edu/) is seeking an outstanding scientist for a tenure track faculty position at the Assistant Professor level. After a successful faculty search this past year, we continue with our expansion of the department in the area of neuroscience.

We are interested in candidates who are using innovative combinations of molecular, genetic, and cellular approaches to pursue fundamental problems in neuroscience. Areas of interest include but are not limited to in vivo imaging, genetic and epigenetic mechanisms underlying neural circuitry plasticity, and behavior, as well as aging, regeneration and repair.

Individuals holding Ph.D. and/or M.D., or equivalent degrees, with two or more years of postdoctoral experience are encouraged to apply. Applicants should demonstrate excellence in research and strong potential for securing and sustaining independent and collaborative extramural funding.

The University of Utah offers excellent resources to support new faculty, including competitive salary and start-up support, a highly collegial research environment, core facilities and strong interdepartmental graduate training programs. A successful applicant will be expected to develop an innovative, independent research program, and to share our commitment to excellence in graduate and medical education.

Only electronic applications will be accepted. Please submit a single PDF document including: 1) cover letter, 2) curriculum vitae, 3) research statement 4) one recent publication. Email the application to: [email protected] Three letters of reference should be sent independently to: [email protected]

For full consideration, applications should be received by October 29, 2010.

The University of Utah is an Affirmative Action/Equal Opportunity employer and does not discriminate based upon race, national origin, color, religion, sex, age, sexual orientation, gender identity/expression, disability, or status as a Protected Veteran. Upon request, reasonable accommodations in the application process will be provided to individuals with disabilities. To inquire about the University’s nondiscrimination policy or to request disability accommodation, please contact: Director, Office of Equal Opportunity and Affirmative Action, 201 S. Presidents Circle, Rm 135, (801) 581-8365.

The University of Utah values candidates who have experience working in settings with students from diverse backgrounds, and possess a demonstrated commitment to improving access to higher education for historically underrepresented students.

The Department of Neurobiology invites applications for a tenure-track position with a rank of assistant professor. We seek an outstanding scientist addressing molecular or cellular mechanisms underlying behavior, sensation, and/or the function or development of neural circuits in vertebrates or invertebrates.

This position offers outstanding scholarly and scientific resources in a collegial and collaborative department with strong ties to related departments throughout Harvard University, the Harvard-affiliated teaching hospitals, and the Boston neuroscience community. The position provides the opportunity to join a growing coalition of researchers at Harvard Medical School interested in molecular and quantitative approaches to neuroscience and systems biology.

The position also offers the opportunity to teach exceptional graduate and medical students with strong interests in neuroscience and related fields. Candidates must have a Ph.D., M.D. or equivalent graduate degree.

Applicants should send a C.V., a 1-page summary of research contributions, and a 1-page description of plans for future work by Dec. 6, 2010. Applicants should arrange to have 3-5 letters of recommendation sent to the search committee.

Send all materials to:Faculty Search Committee, Department of Neurobiology, Harvard Medical School, 220 Longwood Ave., Boston, MA 02115

Or Email to:[email protected]

Harvard Medical School is an Equal Opportunity/Affirmative Action Employer. Women and minorities are

especially encouraged to apply.

Assistant Professor Department of Neurobiology

Harvard Medical School

cell1433cla.indd 3cell1433cla.indd 3 10/21/2010 10:59:52 PM10/21/2010 10:59:52 PM

Page 184: Cell 101029

Positions Available

Faculty PositionThe Department of Molecular and Cell Biology at the Boston University Henry M. Goldman School of Dental Medicine occupies a completely renovated floor adjacent to basic science departments of the Medical School. We have an opening at the Assistant, Associate or Full Professor level. We seek individuals with an outstanding publication record and an ongoing NIH RO1 or K99/ROO-funded research program as principal investigators. We seek qualified candidates with research interests in cell and developmental biology, molecular genetics, biochemistry, immunology or microbiology. Interest in craniofacial and or oral biology is encouraged but not necessary. Excellent laboratory facilities and start-up funds are available as well as joint appointments with appropriate departments at the Medical School and participation in the Bioinformatics Program at the School of Engineering. Email a c.v. including a 250 word summary of present and future research plans and names and email addresses of three to five references, no later than December 31, 2010 to:

Dr. P.W. Robbins, Search Committee Chair ([email protected]) or Dr. C.B. Hirschberg, Department Founding Chair([email protected]).

Please visit http://dentalschool.bu.edu/research/molecular/index.html.

Boston University is an Affirmative Action and Equal Opportunity Employer.

Assistant Professor – Tenure TrackTouchstone Diabetes Research Center

University of Texas Southwestern Medical Center of Dallas

The Touchstone Diabetes Center, Department of Internal Medicine at the UT Southwestern Medical Center is seeking an Investigator at the Assistant Professor level. The position requires a PhD, MD/PhD or MD degree and three to five years of postdoctoral experience in the field of metabolism research. The applicant will develop an independent and externally funded collaborative research program focused on metabolic dysfunction in peripheral organs. The individual will join a team of highly integrated research groups in the areas of obesity, diabetes, nutrition, cardiovascular disease and cancer and will have state of the art phenotyping and assay capabilities available.

Qualified applicants should submit a curriculum vitae, the names of three or more references, a brief summary of research accomplishments and an outline of future research directions to:

Philipp Scherer, PhD, UT Southwestern Medical Center at Dallas, Touchstone Diabetes Center, 5323 Harry Hines Blvd., Dallas, TX 75390-8549, [email protected]

UT Southwestern is an equal opportunity affirmative action employer

cell1433cla.indd 4cell1433cla.indd 4 10/21/2010 10:59:54 PM10/21/2010 10:59:54 PM

Page 185: Cell 101029

Positions Available

FACULTY POSITION — NEUROSCIENCEUNIVERSITY OF ROCHESTER

MEDICAL CENTERThe Center for Neural Development and Disease at the University of Rochester School of Medicine and Dentistry invites applications for a tenure-track faculty position at the Assistant, Associate, or Full Professor level. Candidates studying nervous system development and function using genetic approaches in vertebrate or invertebrate model systems are particularly encouraged to apply. Though not a prerequisite, a record of success in obtaining extramural research support is an asset.

The Center for Neural Development and Disease is a dynamic, interdepartmental group of investigators whose research spans a broad area of molecular and cellular neuroscience. Members of the Center hold faculty appointments in a variety of basic science and clinical departments, including Neurology and Biomedical Genetics, allowing multiple opportunities for interaction and collaboration.

Interested candidates should submit a cover letter, CV, and statement of research interests to [email protected]. Please also arrange for three letters of reference to be sent to this address. Review of applications will begin immediately.

The University of Rochester is an Equal Opportunity Employer, has a strong commitment to diversity and actively encourages applications from candidates from groups underrepresented in

higher education.

cell1433cla.indd 5cell1433cla.indd 5 10/21/2010 11:00:02 PM10/21/2010 11:00:02 PM

Page 186: Cell 101029

Positions Available

Tier 2 Canada Research Chair in Chemical Biology

Schulich School of Medicine & Dentistry and Faculty of Science

The University of Western Ontario

The Schulich School of Medicine & Dentistry and the Faculty of Science at The University of Western Ontario (UWO), one of Canada’s leading research intensive universities, seek applicants for a Tier 2 Canada Research Chair in Chemical Biology. In accordance with the regulations set for Tier 2 Canada Research Chairs (www.chairs-chaires.gc.ca/home-accueil-eng.aspx), the candidate will be an excellent emerging researcher who has demonstrated research creativity and innovation, and the potential to achieve international recognition in the field of Chemical Biology within the next five to ten years. The Candidate must propose an original and innovative research program of high quality which would attract excellent trainees, students and future researchers.

The Tier 2 CRC will be expected to establish an independent, externally funded research program in the area of Chemical Biology that will promote integration and synergy with existing areas of research strength in Proteomics & Protein Structure, Genomics & Bioinformatics, and/or Materials & Biomaterials within the Schulich School of Medicine & Dentistry and the Faculty of Science at UWO. Priority will be given to candidates with a strong record of productivity in chemical biology and interests in translational research. The candidate will have access to state-of-the-art facilities including the London Regional Proteomics Centre (www.lrpc.uwo.ca), the London Regional Genomics Centre (www.lrgc.ca) and the Western Nanofabrication Facility (http://www.uwo.ca/fab/). Furthermore, there will be excellent opportunities for collaboration with basic and clinical researchers at UWO and affiliated research institutes.

The successful applicant will hold a Ph.D. or an M.D., or equivalent, and will be a tenure track appointment at the position of Assistant Professor or at an Associate Professor level if qualifications and experience warrant. The appointment will be made to the Department of Biochemistry of Schulich School of Medicine & Dentistry and the Department of Chemistry of the Faculty of Science, with the opportunity for a cross-appointment to an appropriate Clinical Department, and an appointment as Scientist at the Robarts Research Institute and Lawson Health Research Institute.

With full time enrollment of about 32,000, The University of Western Ontario graduates students from a range of academic and professional programs. Further information about the Schulich School of Medicine & Dentistry can be found at www.schulich.uwo.ca, the Faculty of Science at www.uwo.ca/sci and/or at www.uwo.ca. Western’s Recruitment & Retention Office is available to assist in the transition of successful applications and their families.

Please send a detailed curriculum vitae, a brief description of current research program, accomplishments, and future plans, copies of representative publications, and the names of three references to:

Dr. Victor HanAssociate Dean, Research, Schulich School of Medicine & Dentistry

Room 3730-2, Clinical Skills BuildingThe University of Western Ontario

London, Ontario CANADA N6A [email protected]

Applications will be accepted until the position is filled. Review of applicants will begin after November 1, 2010.

Positions are subject to budget approval. Applicants should have fluent written and oral communication skills in English. All qualified candidates are encouraged to apply; however,

Canadians and permanent residents will be given priority. The University of Western Ontario is committed to employment equity and welcomes applications from all qualified women and men,

including visible minorities, aboriginal people and persons with disabilities.

cell1433cla.indd 6cell1433cla.indd 6 10/21/2010 11:00:06 PM10/21/2010 11:00:06 PM

Page 187: Cell 101029

Positions Available

WASHINGTON UNIVERSITY MEDICAL SCHOOL, ST LOUIS FACULTY SEARCH

The Renal Division at Washington University (WU) Medical School is recruiting scientists (MD/MD-PhD/PhD) with an emerging or established program in renal developmental biology/ stem cell biology/ organogenesis/ clinical genomics as reflected by high-quality publications and funding. We seek to fill a full time, tenure-track faculty position at the Assistant Professor level or tenured position as Associate Professor in the Department of Internal Medicine with a secondary appointment in an appropriate basic science department. The Renal Division (http://renal.wustl.edu) has 35 full-time faculty with diverse basic and clinical research interests. The WU George M. O’Brien Center for Kidney Disease Research supports Core facilities in Organogenesis; Transgenic Disease Models; and Renal Genomics. A NIH- Institutional Training Grant supports postdoctoral fellows. The WU Division of Biology and Biomedical Science (http://dbbs.wustl.edu) promotes interaction among diverse faculty groups. Clinical research within the Renal Division and integration with extensive programs of WU and Barnes-Jewish Hospital provide outstanding opportunities for translation of basic discoveries to the clinic.

State-of-the-art laboratory facilities house the Renal Division’s research program. Start-up and relocation packages will be provided. The expectation is to establish an internationally prominent, interactive and ultimately self-sustaining research program, or to advance an already successful program to the next level of excellence.

Faculty candidates should mail or email a statement of interest and CV, along with the names, telephone numbers and email addresses of 3 references to:

Marc R. Hammerman MDRenal Division, Box 8126

Washington University School of Medicine660 South Euclid Ave.St. Louis MO 63110

Attention: FACULTY SEARCHEmail: [email protected]

Washington University is an equal opportunity employer

Yale University – Institute for Chemical Biology

Yale University, to further the development of its West Campus research enterprise, is seeking faculty at both junior and senior ranks for a new multidisciplinary Institute for Chemical Biology. Faculty associated with this Institute will hold primary appointments in any of several life science and physical science departments within the Faculty of Arts and Sciences, the School of Engineering and Applied Science, or the Yale School of Medicine. We seek creative teacher-scholars with international reputations for outstanding research at the combined interface of chemistry, biology, engineering and medicine. Candidates must possess a Ph.D. in a relevant discipline. To apply, please submit to [email protected] in one pdf file with the subject heading “Chemical Biology Search” the following materials: a statement of research interests, complete CV, and up to five reprints of published work. In addition, arrange for three letters of recommendations to be sent to Chair, Chemical Biology Search Committee, c/o Kelly Locke, 1 Hillhouse Avenue, New Haven, CT 06520. The review of applications will begin on 1 November and proceed until suitable candidates are identified. Yale University is an affirmative action, equal opportunity employer. Yale values diversity among its faculty, students, and staff and strongly encourages applications from women and underrepresented minorities.

years of leadership in human genetics research,

education and service.

1948–2008www.ashg.org

60

cell1433cla.indd 7cell1433cla.indd 7 10/21/2010 11:00:07 PM10/21/2010 11:00:07 PM

Page 188: Cell 101029
Page 189: Cell 101029

careers.cell.com

Reach Your Ideal Candidate!

Page 190: Cell 101029
Page 191: Cell 101029

careers.cell.com

Find Your Ideal Job!

Page 192: Cell 101029
Page 193: Cell 101029

How does your institution measure up?Scopus is the optimal data source for research performance measurement. No other database has so much breadth of content covering so many authors.

“As financial resources become more scarce, it is more critical to identify research and researchers who are the most productive and on the right track.”Peter BrimblecombeProfessor, Atmospheric ChemistrySchool of Environmental SciencesUniversity of East Anglia, UK

With Scopus you can identify authors’ papers, tracktheir citations and analyze their influence using the Scopus h-index. And, to evaluate the performance of journals, research projects and groups of researchersyou can measure the performance of a specifiedcollection of articles.

Now it’s easy to:• Evaluate and prioritize resource allocation by

departments or fields• Make informed decisions about tenure and promotion • Promote your institution for funding and recruitment

www.scopus.com

Page 194: Cell 101029

See online version for legend and references.486 Cell 143, October 29, 2010 ©2010 Elsevier Inc. DOI 10.1016/j.cell.2010.10.025

SnapShot: Neural CrestTatjana Sauka-Spengler and Marianne BronnerCalifornia Institute of Technology, Pasadena, CA 91125, USA

Bronner.indd 1 10/21/10 2:42:37 PM

Page 195: Cell 101029

Announcing an innovative new textbook from Academic Cell Primer to The Immune Response, Academic Cell Update Edition

By Tak W. Mak and Mary Saunders

Facebook.com/academiccell Twitter.com/academiccell

Primer to The Immune ResponseAcademic Cell Update Edition

Tak W. MakThe Campbell Family Institute for Breast Cancer Research, Ontario, Canada

Mary SaundersThe Campbell Family Institute for Breast Cancer Research, Ontario, Canada

Paperback/456 pagesISBN: 9780123847430$79.95/£54.99/€64.95

Primer to The Immune Response, Academic Cell UpdateEdition, is an invaluable resource for students whoneed a concise but complete and understandableintroduction to immunology.

Academic Cell textbooks contain premium journal content from Cell Press and are part of a new cutting-edge textbook/journal collaboration designed to help today’s instructors teach students to “think like a scientist.”

academiccell.com

Academic Cell is a dynamic textbook publishing partnership between Academic Press and Cell Press, two market-leading publishers bringing scientific advances from the world of life science research into the classroom.

Order online now from: elsevierdirect.com/9780123847430Request and examination copy from textbooks.elsevier.com

Page 196: Cell 101029

ORDER 888.999.1371 • TECH 888.810.6168 • INQUIRIES [email protected]

Cell: trim 8.375x10.875, bleed .125, margin .25

Mouse TrueBlot®: Caspase-7 was immunoprecipitated from Jurkat cells using mouse anti-caspase-7. Immunoprecipitate was detected using either Mouse IgG TrueBlot® (left blot) or a conventional HRP (right blot) anti-mouse IgG second step reagent.

– Heavy Chain (55kD)

– Caspase 7Caspase 7 –

– Light Chain (25kD)

MouseTrueBlot®

ConventionalAnti-Mouse HRP

Mouse TrueBlot®

Mouse TrueBlot®: Caspase-7 was immunoprecipitated from Jurkat cells using mouse anti-caspase-7. Immunoprecipitate was detected using either Mouse IgG TrueBlot® (left blot) or a conventional HRP (right blot) anti-mouse IgG second step reagent.

– Heavy Chain (55kD)

– Caspase 7Caspase 7 –

– Light Chain (25kD)

MouseTrueBlot®

ConventionalAnti-Mouse HRP

Mouse TrueBlot®

Vist www.eBioscience.com and search TrueBlot for picture perfect data.

Picture PerfectTrueBlot® western blot detection system e� ectively eliminates the appearance of non-speci� c, reduced immunoglobulin (Ig) bands in SDS-PAGE following immunoprecipitation, giving you publication quality data, without additional heavy or light chain Ig bands.

TrueBlot® Advantages:

• Ease of useHorseradish Peroxidase (HRP)-conjugated TrueBlot® anti-Ig simply replaces your regular HRP-conjugated secondary antibody.

• Accurate target detectionPreferential detection of native Ig ensures youare only detecting your target protein.

• Publication-quality dataClean blots without irrelevant bands from heavy and light Ig contamination.

Publication Quality Data.

blot) anti-mouse IgG second step reagent.

Q310035-TrueBlot-Cell.indd 1 8/23/10 6:15 PM