+ All Categories
Home > Documents > Cell 100903

Cell 100903

Date post: 20-Nov-2014
Category:
Upload: prasath
View: 126 times
Download: 9 times
Share this document with a friend
Description:
This is the September 03, 2010 issue from Cell. A highly reputed journal for Biology.
Popular Tags:
212
Reorienting Planar Polarity Minireview Histone Crosstalk Volume 142 www.cell.com Number 5 September 3, 2010
Transcript
Page 1: Cell 100903

ReorientingPlanar Polarity

MinireviewHistone Crosstalk

Volum

e 142 Num

ber 5 Pages 649–822 S

eptember 3, 2010

Volume 142

www.cell.com

Number 5

September 3, 2010

INSERT ADVERT

cell142_5.c1.indd 1cell142_5.c1.indd 1 8/27/2010 8:23:38 PM8/27/2010 8:23:38 PM

Page 2: Cell 100903

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 3: Cell 100903

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 100903

015.

A1.0

101.

F.U

S ©

200

9 E

pp

end

orf A

G.

Bac

kgro

und

: ele

ctro

por

ated

L-m

ouse

cel

ls.

Electroporation of mammalian cells

Electroporation of bacteria

Electroporation of yeasts

Cell fusion

Outstanding transfection

rates!

www.eppendorf.com • Email: [email protected] • Application hotline: 516-515-2258

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

Our Multiporator® is a compact, flexible system designed for the electroporation of eukaryotic cells, bacteria and other microorganisms, as well as the electrofusion of cells.

Unique features include Soft Pulse microsecond pulses, patented* electronically regulated pulses, a gentle hypoosmolar buffer system and a choice of electroporation and electrofusion options.

A safe and easy method for performing cell fusion: Microfusion chamber for viewing and optimization

of fusion parameters Helix chamber for larger cell volumes Applications include: monoclonal antibodies, plant

protoplast fusion, tumor research, cancer vaccines, cloning mammals

For more information visit www.eppendorf.com

One electroporator–many applications.

*U.S. Pat. 6,008,038

Page 5: Cell 100903

DAWN HELEOS®. The most advanced multi-angle light scattering instrument for macromolecular characterization.

Optilab rEX®. The refractometer with the greatest sensitivity and range.

ViscoStar™. The viscometer with unparalleled signal-to-noise, stable baselines and a 21st-century interface.

Eclipse. The ultimate system for the separation of macromolecules and nanoparticles in solution.

DynaPro Plate Reader. Automated dynamic light scattering for 96 or384 or 1536 well plates.

If your standards and your samples are the same size,they must be the same mass, right?

©2008 Wyatt Technology. DAWN, Optilab, DynaPro and the Wyatt Technology logo are registered trademarks, and ViscoStar and Eclipse are trademarks of Wyatt Technology Corporation.

� at’s the problem with calibration-based measuring techniques. You don’t know—you can only assume. Which is why

every major pharmaceutical and biotechnology company, as well as most federal regulatory agencies are switching from relative

methods to Wyatt Technology’s absolute measurements. Our DAWN® multi-angle light scattering (MALS) instruments

allow you to determine absolute molecular weights and sizes without relying on so-called standards, or measurements made in

someone else’s lab. Wyatt instruments measure all of the quantities required for determining absolute molar masses directly.

If you’re ready to abandon the habit and hope of reference standards, call 805-681-9009 or visit wyatt.com to request our

free 28-page Ultimate Guide to Light Scattering. Our booklet will show you how

to make your reliance on standards a thing of the past. CORPORATIONCORPORATIONCORPORA

Page 6: Cell 100903

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 100903

Editor

Emilie Marcus

Senior Deputy Editor

Elena Porro

Scientific Editors

Karen Carniol

Connie M. Lee

Fabiola Rivas

Lara Szewczak

Leading Edge Editor

Orla M. Smith

Associate Editor

Robert Kruger

Assistant Editor

Michaeleen Doucleff

Senior Managing Editor

Meredith Adinolfi

Deputy Managing Editor

Jackie M. Smith

Illustration and

Design Supervisor

Nathan E. Bress

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 8: Cell 100903

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 T M

T h e E l e m e n t s o f D i s c o v e r y

S c i s s o r s • N e e d l e H o l d e r s • F o r c e p s • R e t r a c t o r s • C l a m p s • R o n g e u r s • P r o b e s

S c a l p e l s • S u r g i c a l & L a b o r a t o r y A c c e s s o r i e s • S t u d e n t Q u a l i t y I n s t r u m e n t s

R e q u e s t a c a t a l o g a t f i n e s c i e n c e . c o m o r c a l l 1 - 8 0 0 - 5 2 1 - 2 1 0 9

enrichyour

experience

Scopus is the largest abstract and citationdatabase of peer-reviewed literature and quality web sources with smart tools totrack, analyze and visualize research.

www.scopus.com

Page 9: Cell 100903

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 Supervisor

Meredith Adinolfi

Operations Supervisor

Daniel Fauxsmith

Press Officer

Cathleen Genova

Display Advertising

Midwest/Southeast/Eastern Canada:

Inez Herrero, ph: 585 678 4395; fax: 585 678 4722;

e-mail: [email protected]

Mid-Atlantic/New England: Vikki Macomber, ph: 508 928

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

West Coast/Western Canada/Asia: Lynne Stickrod, ph: 415

931 9782; fax: 415 520 6940; e-mail: [email protected]

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

6585 3136; e-mail: [email protected]

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 damage

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

wise, 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 advertising material is

expected to conform to ethical (medical) standards, inclusion 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 is-

sue number, payment (check or credit card, MasterCard, Visa, or Amer-

ican 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 www.cellpress.com.

Guide for Authors:

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

authors.

Page 10: Cell 100903

REMARKABLY FAST

www.neb.com

CLONING & MAPPING DNA AMPLIFICATION& PCR RNA ANALYSIS PROTEIN EXPRESSION

& ANALYSISGENE EXPRESSION

& CELLULAR ANALYSIS

Not all PCR polymerases are created equal

Amplifi cation of a 3.8 kb fragment from the human beta globin gene clearly illustrates the extreme speed and robustness offered by using Phusion DNA Polymerase. Reactions were performed according to the suppliers’ recommendations using varying extension times (shown above gel).

Phusion is a registered trademark of Finnzymes Oy.

Amplifi cation of a 3.8 kb fragment from the human beta globin gene clearly illustrates the

Phusion® High-Fidelity DNA Polymerase With Phusion® High-Fidelity DNA Polymerase, there is no need to compromise any aspect of your PCR performance.

A superior choice for cloning, this recombinant polymerase has an error rate 50-fold lower than Taq DNA Polymerase,

combining extreme precision with unparalleled speed and robustness.

Advantages:• Extreme fi delity

• High speed

• Robustness

Distributed byProduced by

Go to www.neb.com/phusion to fi nd out

how Phusion High-Fidelity DNA Polymerase can

improve your PCR performance.

• High yield

• Specifi city

Page 11: Cell 100903

Leading EdgeCell Volume 142 Number 5, September 3, 2010

IN THIS ISSUEMOLECULAR BIOLOGY SELECT

ESSAY

661 Phosphotyrosine Signaling: Evolving a NewCellular Communication System

W.A. Lim and T. Pawson

CORRESPONDENCE

668 What Controls T Cell ReceptorPhosphorylation?

R.A. Fernandes, C. Yu, A.M. Carmo, E.J. Evans,P.A. van der Merwe, and S.J. Davis

669 Response: Multilayered Control of T CellReceptor Phosphorylation

E. Gagnon, C. Xu, W. Yang, H.H. Chu, M.E. Call,J.J. Chou, and K.W. Wucherpfennig

PREVIEWS

672 Fishing Out a Sensor for Anti-inflammatory Oils A.R. Saltiel

674 A New Spin on Planar Cell Polarity P. Olguin and M. Mlodzik

676 Viable Rat-Mouse Chimeras:Where Do We Go from Here?

D. Solter

679 ‘Fore Brain:A Hint of the Ancestral Cortex

L.B. Sweeney and L. Luo

MINIREVIEW

682 The Language of Histone Crosstalk J.-S. Lee, E. Smith, and A. Shilatifard

SNAPSHOT

822 Nuclear Receptors I N.J. McKenna and B.W. O’Malley

Page 12: Cell 100903

© 2010 Cell Signaling Technology, Inc.

Cell Signaling Technology® and PhosphoSitePlus

® are trademarks of Cell Signaling Technology, Inc.

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.comUnparalleled product quality, validation and technical support.

PhosphoSitePlus® v3.0 (PSP) is an open web resource that integrates encyclopedic information on experimentally determined protein modifi cation sites, upstream and downstream regulation of these modifi cations, and powerful analytical tools for investigating the structural and biological signifi cance of protein modifi cations.

www.phosphosite.org

• Expansive and continuously curated content

• Molecular rendering to visualize the location of modifi cation sites

• On-the-fl y generation of kinase substrate sequence logos

• Browsing of high-throughput content by disease, cell line, and tissue

• New search interfaces that retrieve modifi cation sites and proteins by subcellular locations, sequence and motifs, domains, responsiveness to treatments, disease, tissue, and cell type

A Comprehensive Online Protein Modification Resource

provided by Cell Signaling Technology with grant support from the NIH

Page 13: Cell 100903

ArticlesCell Volume 142 Number 5, September 3, 2010

687 GPR120 Is an Omega-3 Fatty Acid ReceptorMediating Potent Anti-inflammatoryand Insulin-Sensitizing Effects

D.Y. Oh, S. Talukdar, E.J. Bae, T. Imamura,H. Morinaga, W. Fan, P. Li, W.J. Lu,S.M. Watkins, and J.M. Olefsky

699 Anti-CD47 Antibody Synergizes withRituximab to Promote Phagocytosis andEradicate Non-Hodgkin Lymphoma

M.P. Chao, A.A. Alizadeh, C. Tang, J.H. Myklebust,B. Varghese, S. Gill, M. Jan, A.C. Cha, C.K. Chan,B.T. Tan, C.Y. Park, F. Zhao, H.E. Kohrt, R. Malumbres,J. Briones, R.D. Gascoyne, I.S. Lossos, R. Levy,I.L. Weissman, and R. Majeti

714 A C-Type Lectin Collaborates with aCD45 Phosphatase Homolog to FacilitateWest Nile Virus Infection of Mosquitoes

G. Cheng, J. Cox, P. Wang, M.N. Krishnan,J. Dai, F. Qian, J.F. Anderson, and E. Fikrig

726 The ATAC Acetyltransferase ComplexCoordinates MAP Kinasesto Regulate JNK Target Genes

T. Suganuma, A. Mushegian, S.K. Swanson,S.M. Abmayr, L. Florens, M.P. Washburn,and J.L. Workman

737 A Bacterial mRNA Leader that EmploysDifferent Mechanisms to SenseDisparate Intracellular Signals

S.-Y. Park, M.J. Cromie, E.-J. Lee,and E.A. Groisman

749 Structural Basis of Semaphorin-PlexinRecognition and Viral Mimicry fromSema7A and A39R Complexes with PlexinC1

H. Liu, Z.S. Juo, A.H.-R. Shim, P.J. Focia, X. Chen,K.C. Garcia, and X. He

762 Photoadaptation in Neurospora byCompetitive Interaction of Activatingand Inhibitory LOV Domains

E. Malzahn, S. Ciprianidis, K. K�aldi, T. Schafmeier,and M. Brunner

773 Cell Flow Reorients the Axisof Planar Polarity in theWing Epithelium of Drosophila

B. Aigouy, R. Farhadifar, D.B. Staple, A. Sagner,J.-C. R€oper, F. J€ulicher, and S. Eaton

787 Generation of Rat Pancreas in Mouseby Interspecific Blastocyst Injectionof Pluripotent Stem Cells

T. Kobayashi, T. Yamaguchi, S. Hamanaka,M. Kato-Itoh, Y. Yamazaki, M. Ibata, H. Sato,Y.-S. Lee, J.-i. Usui, A.S. Knisely, M. Hirabayashi,and H. Nakauchi

(continued)

Page 14: Cell 100903

Let the Countess™ do the counting.

THE MOST

Wow! Consistent, fast, and easy.

The Countess™ Automated Cell Counter takes the subjectivity and tedium out of one of the fundamental steps of cell culture—counting live and dead cells. The Countess™ automated technology reduces user error and speeds up cell counting with less sample, less hassle, and a lower cost than any automated cell counter on the market. Start counting today at www.invitrogen.com/countcells.

For research use only. Not intended for any animal or human therapeutic or diagnostic use, unless otherwise stated. © 2009 Life Technologies Corporation. All rights reserved. The trademarks mentioned herein are the property of Life Technologies Corporation or their respective owners. These products may be covered by one or more Limited Use Label Licenses (see Invitrogen catalog or www.invitrogen.com). By use of these products you accept the terms and conditions of all applicable Limited Use Label Licenses.

Page 15: Cell 100903

800 Profiling by Image RegistrationReveals Common Origin of AnnelidMushroom Bodies and Vertebrate Pallium

R. Tomer, A.S. Denes, K. Tessmar-Raible,and D. Arendt

RESOURCE

810 The Protein Composition of MitoticChromosomes Determined UsingMulticlassifier Combinatorial Proteomics

S. Ohta, J.-C. Bukowski-Wills, L. Sanchez-Pulido,F. de Lima Alves, L. Wood, Z.A. Chen, M. Platani,L. Fischer, D.F. Hudson, C.P. Ponting, T. Fukagawa,W.C. Earnshaw, and J. Rappsilber

POSITIONS AVAILABLE

On the cover: Many epithelial tissues orient external structures such as cilia and hairs, and to

do so they use the planar cell polarity (PCP) proteins, which form polarized cortical domains

at epithelial cell junctions. The mechanisms that globally orient planar polarity have been

mysterious. In this issue, Aigouy et al. (pp. 773–786) show that the pattern of PCP domain

orientation in the Drosophila wing is a consequence of the oriented cell rearrangements and

divisions that shape this tissue.

Page 16: Cell 100903

Proteins

Antibodies

ELISAs

Assay Services

MultiAnalyte Pro� ling

Activity Assays

Stem Cells

ELISpot Kits

Flow Cytometry

Cell Selection

α1-Acid Glycoprotein Angiopoietin-like 3 Cathepsin V Clusterin Dkk-1 EGF R/ErbB1 EG-VEGF/PK1 Fetuin A FGF-21 Galectin-3 Gas 6 GDF-15 IL-17A/F Heterodimer IL-19 Lipocalin-2/NGAL MBL Proprotein Convertase 9/PCSK9 Periostin/OSF-2 Progranulin ST2/IL-1 R4 Thrombomodulin/CD141 Tie-1 TIM-1/KIM-1

NEW Quantikine ELISA Kits

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.

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

A direct measure of product quality is the frequency of citations in the scienti� c literature. R&D Systems has more than 20 years of experience designing, testing, and optimizing the most cited ELISA kits in the world. Find out why scientists trust R&D Systems ELISAs more than any other brand.

The Most Referenced Immunoassays

1.4%1.6%2.1%2.5%

3.7%

7.9%

17.3%

41.8%R&D Systems

R&D Systems is the Most Referenced ELISA Manufacturer

R&D Systems Quantikine® ELISAs

Approximately 42% of Referenced Immunoassays are Developed and Manufactured by R&D Systems. A survey of 860 manuscripts from 44 journals was conducted to compare the number of citations specifying the use of R&D Systems ELISAs to the number citing ELISAs from other commercial sources. A total of 433 ELISA citations referencing immunoassays from 66 di� erent vendors were identi� ed in the survey.

R&D Systems Tools for Cell Biology Research™

55870_Cell Press.indd 1 6/25/10 10:49 AM

Page 17: Cell 100903

Leading Edge

In This Issue

Fish Oil Turns the Tide on Insulin ResistancePAGE 687

Inflammation mediated by macrophages promotes insulin resistance in obesity. Oh et al. now identify the G protein-coupledreceptor 120 (GPR120) on macrophages and fat cells as a receptor for omega-3 fatty acids (u-3 FAs). The authors show thatGPR120 activation by u-3 FAs inhibits inflammation pathways in macrophages and can reverse insulin resistance in mice.These results provide a molecular basis for the anti-inflammatory effects of u-3 FAs and suggest that anti-inflammatorytreatments may ameliorate insulin resistance in obesity.

West Nile Virus Stopped at Its SourcePAGE 714

West Nile virus, a potentially deadly virus in humans, propagates in mosquitoes.Cheng et al. now find that infection of West Nile virus triggers mosquito cells toproduce the lectin protein mosGCTL-1. This C-type lectin enhances entry of thevirus into additional mosquito cells through its interaction with a protein tyrosinephosphatase receptor, which is homologous to human CD45. BlockingmosGTCL-1 with an antibody disrupts the infective cycle of West Nile virus inmosquitoes, suggesting a new strategy for controlling viral dissemination.

Antibodies Are Double-Trouble for CancerPAGE 699

Monoclonal antibodies are standard therapeutics for several cancers, including theanti-CD20 antibody rituximab for B cell non-Hodgkin lymphoma (NHL). However,antibodies are not curative and must be combined with cytotoxic chemotherapyfor clinical benefits. Now, Chao et al. identify CD47 as an antibody target in NHLand demonstrate that combining anti-CD47 antibody with the rituximab antibody eradicates human NHL in mice. The synergisticmechanism used by these two antibodies may be applicable to combined antibody treatments for many types of cancers.

ATAC Wears Two HATs in MAPK SignalingPAGE 726

Extracellular cues often trigger MAP kinase (MAPK) signaling pathways, whichthen activate downstream transcription factors like c-Jun. Here, Suganumaet al. demonstrate that the ATAC histone acetyltransferase (HAT) governs theresponse to MAPK signaling by serving as both a coactivator of transcriptionand a suppressor of upstream signaling in the MAPK pathway. The authorsshow that ATAC acetylates histone H4 at JNK target genes, which then servesas a positive cofactor for basal transcription. In addition, ATAC directs upstreamMAPKs to the site of c-Jun binding and restricts the levels of JNK activation.

A Multitasking Leader mRNAPAGE 737

Bacterial mRNAs often contain leader sequences that regulate transcriptionof the adjacent coding region by binding metabolites and ions. For example,the leader of the mRNA for the Salmonella Mg2+ transporter gene mgtA responds

to Mg2+. Now, Park et al. demonstrate that this leader also contains a short open reading frame with many proline codons, trans-lation of which places mgtA expression under the control of cytoplasmic proline concentrations as well as Mg2+. Thus, leadermRNAs can use distinct mechanisms to sense multiple intracellular signals.

Vivid Memories of Days Gone ByPAGE 762

Light responses and photoadaptation in the fungus Neurospora depend on the circadian transcription factor White CollarComplex (WCC) and its negative regulator Vivid (VVD). Mazhan et al. now demonstrate how WCC and VVD cooperate todiscriminate light intensities over more than five orders of magnitude during the day. At night, previously synthesized VVDserves as a molecular memory of the sun’s brightness during the preceding day and suppresses responses to light cuesof lower intensity, such as moonlight.

Cell 142, September 3, 2010 ª2010 Elsevier Inc. 649

Page 18: Cell 100903

Roche Diagnostics GmbHRoche Applied Science82372 Penzberg, Germany

Cancer Research Detect and Characterize Genomic Variations

Use powerful Roche instruments and consumables for high-throughput analysis of copy number variation using NimbleGen CGH microarrays. Rely on 454 Sequencing Systems, with their long read lengths (400 to 500 bp). Or combine both technologies using a Sequence Capture workfl ow.

Detect chromosomal imbalances down to approximately 5 kb, commonly found in tumor progression, using high-density NimbleGen CGH microarrays.

Identify all important genomic and transcriptomic variants by ultra-deep sequencing of PCR products using the 454 Sequencing System.

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

454, 454 SEQUENCING, LIGHTCYCLER, and NIMBLEGEN are trademarks of Roche.

Other brands and product names are trademarks of their respective holders.

© 2010 Roche Diagnostics GmbH. All rights reserved.

Verifi cation of molecular mutations. The JAK2 V617F mutation is characteristic in hematopoietic malignancies. Ultra-deep 454 Sequencing detected the mutation in just 1.16% of reads (top panel). Sensitive Carousel-Based LightCycler® melting curve analysis confirmed the mutation in approximately 1% of cells (lower panel).

Data provided by A. Kohlmann, MLL Munich Leukemia Laboratory GmbH, Germany. View the entire Application Note No. 7 at www.cancer-research.roche.com.

454-Ultra Deep Sequencing

qPCR Analysis using the LightCycler® System

Wild type V617F and LOH

Mutation

Visit www.cancer-research.roche.com for application notesand more detailed information about using Roche products incancer research.

Page 19: Cell 100903

Cell Polarity Goes with the FlowPAGE 773

Planar polarization of epithelial cells allows the uniform alignment of hairs, cilia, and other cellular structures with tissue shape.Now, Aigouy et al. combine experimental and theoretical approaches to show that polarity patterns in the Drosophila wing ariseduring growth. Specifically, cell polarity is reoriented from a radial to a proximal-distal axis when mechanical stresses duringgrowth cause the cells to rotate or ‘‘flow’’ with respect to each other. Linking planar polarity to morphogenesis provides a simplemechanism for coordinating the global polarity pattern with tissue shape.

Com-plexin’ with SemaphorinsPAGE 749

Semaphorins and their receptors, Plexins, are widely expressed protein familiesthat mediate repulsive signaling during cell guidance. Here, Liu et al. present twoX-ray crystal structures of PlexinC1, one in complex with the SemaphorinSema7A and another in complex with the Semaphorin mimetic A39R from thesmallpox virus. In both structures, the Semaphorin interacts with a PlexinC1dimer in a novel edge-on, orthogonal geometry.These findings suggest that Plex-ins are activated by ligand-induced dimerization during cell guidance.

Interspecies OrganogenesisPAGE 787

A goal of regenerative medicine is to derive organs from a patient’s pluripotentstem cells (PSCs), but in vitro organogenesis is complex. Here, Kobayashi et al.generate a functioning rat pancreas in mice without a pancreas by injecting ratPSCs into mouse blastocysts. These interspecific chimeras provide proof ofprinciple for in vivo generation of organs derived from donor PSCs and for inter-specific blastocyst complementation.

A Wormhole to the Origin of the CortexPAGE 800

The cerebral cortex or pallium controls the highest-order processing in mammals, but its evolution remains enigmatic. Now,Tomer et al. develop an expression profiling technique to generate a gene expression map for the developing brain of thesegmented worm Platynereis dumerilii. Comparison of this map with that observed for the developing cerebral cortexsuggests a common evolutionary origin for the mammalian cortex and the worm’s mushroom body, a brain region in inver-tebrates that processes sensory input.

Will the Real Chromosomal Proteins PleaseStand UpPAGE 810

Proteomic analysis of large cellular structures is frequently hindered by thepresence of contaminants. In their analysis of mitotic chromosomes, Ohtaet al. overcome this problem by integrating proteomics with additional quanti-tative and bioinformatic data—effectively adding a final purification step in silico.This approach successfully pinpoints hitchhikers from amidst the �4,000 iden-tified proteins and provides insight into the functional relationships among thegenuine constituents, including evidence that many more kinetochore-associ-ated proteins exist than recognized previously.

Cell 142, September 3, 2010 ª2010 Elsevier Inc. 651

Page 20: Cell 100903

Next-gen GWAS. NOW.

It’s a content revolution.Up to 50% more coverage of common and rare

variants than all other arrays. Maximum power for

any population.

The Omni family of microarrays can propel

your studies into true next-gen GWAS. With

a clearly defined product path to the future.

Immediate utility. Future flexibility.

Get on the path to next-gen GWAS.

Now is the time. Get started at

www.illumina.com/GWAS

Page 21: Cell 100903

Leading Edge

Molecular Biology Select

The central importance of the tumor suppressor Retinoblastoma protein (Rb) in cell-cycle progression makesits regulation a focal point for diverse biological processes, as evidenced by recent work described in thisissue’s Molecular Biology Select. These findings reveal new insight into Rb’s involvement in tissue regener-ation and differentiation, as well as previously unrecognized mechanisms of Rb regulation.

Short-Term Inactivation, Lasting BenefitsTissue regeneration in mammals is greatly restricted, whereasmany other types of vertebrates display astonishing feats of tissuereplacement, including regrowth of major structures such aslimbs. According to new findings by Pajcini et al. (2010), mamma-lian innovations in the realm of tumor suppression are one factorthat is likely to contribute to this pronounced deficiency. Theauthors show that combined inactivation of the tumor suppressorsretinoblastoma (Rb) and ARF reverses the differentiation of mousemuscle cells, turning postmitotic cells into proliferating myoblasts,and when these induced myoblasts are transplanted into donormice they successfully fuse into existing myofibers. Prior work innewts has shown that Rb phosphorylation, which inactivates theprotein, promotes the reentry of myotubes into the cell cycle, a crit-ical initiating event in limb regeneration. The inspiration to addition-ally inactivate ARF was motivated by available evidence suggestingits potential exclusivity to mammals and birds, and thus the authors

reasoned that ARF might be particularly important in mediating differences in regeneration potential betweenvertebrate species. Could a similar strategy of Rb and ARF inactivation be used to promote cell-cycle entry ofendogenous cells for regeneration therapy? Although this remains to be seen, the observation that only tran-sient inactivation of these tumor suppressors is needed for successful creation of regenerative cells may gosome way toward allaying concerns that such an approach would invariably promote cancer. Future effortsare also likely to address whether this intervention triggers cell-cycle reentry for a range of mammalian celltypes.K.V. Pajcini et al. (2010). Cell Stem Cell 7, 198–213.

Fat Chance for Bone FormationPrevious in vitro studies suggested that Rb plays a critical role in thedecision of meschenchymal cells to become adipocyte or bonecells, but in vivo evidence for this hypothesis has been lacking.Calo et al. (2010) now show that Rb gives meschenchymal cellsthe extra nudge they need to commit to becoming bone-formingosteoblasts. Without Rb, these cells are more likely to differentiateinto brown fat cells, leading to reduced levels of calcified bonesand increased levels of brown adipose tissue in mice. To sort outhow Rb regulates the fate of the meschenchymal cells during devel-opment, the authors engineered mice with the RB1 gene and/orthe p53 gene deleted only in uncommitted meschenchymal cells.As expected, animals missing the tumor suppressor p53 develop multiple types of tumors, including osteosar-comas. Combining the p53 mutation with a deletion of one RB1 allele increases the frequency of osteosarcomas,whereas deletion of both RB1 alleles shifts the tumor distribution away from osteosarcomas towards brown fattumors. Thus, Rb regulates the fate of meschenchymal cells in a dose-dependent manner. These results aresurprising given that the majority of human osteosarcomas contain RB1 mutations. The authors speculatethat osteosarcoma tumors most likely arise from cells already committed to becoming osteoblasts, and muta-tions in RB1 promote dedifferentiation of these cells and thus tumorigenesis.E. Calo et al. (2010). Nature. Published online August 4, 2010. 10.1038/nature09264.

Dedifferentiated myocytes redifferentiate and fuse

into existing muscles in vivo (visualized with green

fluorescent protein).

Deletion of Rb in the embryo proper using Meox2-

Cre reduces the level of calcified bone as detected

by Alizarin Red staining. Image courtesy of J. Lees.

Cell 142, September 3, 2010 ª2010 Elsevier Inc. 653

Page 22: Cell 100903

careers.cell.com

Reach Your Ideal Candidate!

Page 23: Cell 100903

Rb Gets Caught in a Custody BattleRb is caught in a tug-of-war between cyclin-dependent kinases(Cdks) that inactivate it by phosphorylation to permit cell-cycleprogression and phosphatases that remove the modificationto promote cell-cycle arrest. New findings of Hirschi et al. (2010)demonstrate that this conflict is waged over the same bindinginterface of Rb. They provide structural evidence that the phospha-tase PP1 interacts with Rb in a region previously shown critical forthe interaction of Rb with Cdks. They further show that PP1 cansuppress the activity of Cdk2-cyclin A toward Rb to block cell-cycle progression in a human osteosarcoma cell line, and thatthe complex of PP1 and Rb appears to be stable, or at leastmore prevalent, at mitotic exit. Among the interesting questionsthis work raises is, what factors determine the outcome of PP1and Cdk competition? The authors propose that concentrationand subcellular localization of the competing proteins likely playa role, but it remains unclear how this molecular dispute is settledunder specific biological circumstances, for example after DNAdamage.A. Hirschi et al. (2010). Nat. Struct. Mol. Bio. Published online August8, 2010. 10.1038/nsmb.1868.

Methylation Moves to the Front of the LineAlthough first reported more than 30 years ago, methylationof protein N termini (a-N-methylation) remains a poorly understoodprotein modification. Thus, the recent identification by Schaner-Too-ley et al. (2010) of an enzyme that catalyzes the reaction promises toaccelerate understanding of the modification’s functions. Theauthors use methylation of a known target of a-N-methylation calledRCC1 (a Ran guanine nucleotide exchange factor) as an indicator ofthe presence or absence of the enzymatic activity from fractionatedHeLa cell nuclear extracts. Fractions with methyltransferase activitywere subjected to mass spectrometry, leading to the identification ofthe methyltransferase responsible, which the authors name NMRT(N-terminal RCC1 methyltransferase). The NMRT crystal structurefacilitated the modeling of substrate recognition, and further analysisdefining the consensus sequence for target recognition suggestedRb as a potential substrate. Subsequent assays provide evidencethat Rb is modified by a-N-methylation, at least in some cell types.a-N-methylation of RCC1 promotes stable association with chro-matin, and loss of NMRT or the absence of the RCC1 methylation results in defects in mitosis. In contrast, thepurpose of Rb a-N-methylation remains a tantalizing mystery. Regardless of whether Rb a-N-methylation is rele-vant to its roles in cell-cycle control, the discovery of NMRT opens a door through which others will likely follow.Are there other a-N-methyltransferases? And if so, is their substrate specificity similar to NMRT? The answer maybe an indicator of whether this modification is an exotic posttranslational event or might instead be considerablymore common than previously appreciated. Another compelling question is whether there are a-N-demethylasesthat reverse the modification.C.E. Schaner-Tooley et al. (2010). Nature. Published online July 29, 2010. 10.1038/nature09343.

Robert P. Kruger

X-ray crystal structure of Rb (magenta) in complex

with the PP1 catalytic domain (gray). Image cour-

tesy of S. Rubin.

NMRT structure with RCC peptide modeled in the

active site. Image courtesy of I. Macara.

Cell 142, September 3, 2010 ª2010 Elsevier Inc. 655

Page 24: Cell 100903
Page 25: Cell 100903

Find out more at www.millipore.com/amicon

Dr. Michael WestCEO, BioTime, Inc.

Renowned thought leader in stem cell therapeutics

TRUSTED PARTNERSLIKE MILLIPORE ARE RARE BUT CENTRAL TO OUR SUCCESS.”

Go to the source at www.millipore.com/trustStericup1

Millipore, Advancing Life Science Together, Stericup and Steritop are registered trademarks of Millipore Corporation.The M mark is a trademark of Millipore Corporation. ©2010 Millipore Corporation. All rights reserved. LS-SBU-10-03205

Millipore is the source for sterile filtration you trust.Stericup® and Steritop® sterile filters combine the superior flow rates and low non-specific binding you need to protect everything you’ve poured into your cell culture work. With over 50 years of filtration expertise, Millipore continues to set the industry standard for high-performance membrane technology and its application in sterile filtration.

10MLP016_TrustedSupplier_CELL.indd 1 6/10/10 10:53:41 AM

Page 26: Cell 100903

Benefits of Membership:

• Subscription to AJHG, now published by Cell Press

• Reduced registration fees for the Annual Meeting

• Discounted subscriptions to Cell Press journals

• Access to educational materials

• Influence on public policy process

Stay ConnectedJoin ASHG Today!

www.ashg.org

Page 27: Cell 100903

ADVANCING LIFE SCIENCE TOGETHER®

Research. Development. Production.

Millipore and Advancing Life Science Together are registered trademarks of Millipore Corporation. The M Markis a trademark of Millipore Corporation. MAGPIX and xMAP are registered trademarks of Luminex Corporation.©2010 Millipore Corporation. All rights reserved.

Catch the wave atwww.millipore.com/MAGPIX2

The magnetic-bead, low-cost platform that attracts multiplexing to every lab. There’s never been a better time to jump in.Accelerate your research at a price you can afford, using the xMAP® technology you trust. The new MAGPIX instrument easily detects up to 50 analytes simultaneously in a small sample. To complete this flexible solution, choose from Millipore’s wide selection of magnetic bead kits encompassing metabolic disease, inflammation, toxicity, neuroscience and cell signaling.

MAGPIX®—THE NEW WAVE IN MULTIPLEXING

Page 28: Cell 100903
Page 29: Cell 100903

Leading Edge

Essay

Phosphotyrosine Signaling: Evolving a NewCellular Communication SystemWendell A. Lim1,2,* and Tony Pawson3,4,*1Howard Hughes Medical Institute2Department of Cellular and Molecular PharmacologyUniversity of California, San Francisco, San Francisco, CA 94158, USA3Samuel Lunenfeld Research Institute, Mount Sinai Hospital, Toronto, Ontario M5G 1X5, Canada4Department of Molecular Genetics, University of Toronto, Toronto, Ontario M5S 1A8, Canada

*Correspondence: [email protected] (W.A.L.), [email protected] (T.P.)DOI 10.1016/j.cell.2010.08.023

Tyrosine phosphorylation controls many cellular functions. Yet the three-part toolkit that regulatesphosphotyrosine signaling—tyrosine kinases, phosphotyrosine phosphatases, and Src Homology2 (SH2) domains—is a relatively new innovation. Genomic analyses reveal how this revolutionarysignaling system may have originated and why it rapidly became critical to metazoans.

Throughout human history, new technolo-

gies and technological platforms have

constantly been invented. Only a small

fraction of these technologies go on to

be widely adopted, but these can

ultimately have transformational conse-

quences. In the evolutionary history of

living organisms, we know that innovative

molecular systems have appeared at key

points in time, and these are thought to

have played a transformative role in major

evolutionary transitions in the tree of life.

But how do such innovative molecular

systems emerge, and how and why do

some proliferate and become stably

adopted by subsequent lineages?

An example of such an innovative

molecular system is phosphotyrosine

(pTyr)-based signal transduction. This

molecular system for transmitting cellular

regulatory information is estimated to

have appeared relatively recently in the

history of life—�600 million years ago,

just prior to the emergence of multicellular

animals (King et al., 2003; Pincus et al.,

2008; Manning et al., 2008). The pTyr sig-

naling system has become an essential

part of metazoan biology. For example,

pTyr signaling plays a central role in many

cell-to-cell communication pathways,

including those that regulate proliferation,

differentiation, adhesion, hormone res-

ponses, and immune defense (Hunter,

2009).

In modern metazoans, pTyr signaling is

mediated by a toolkit of three distinct

functional modules: tyrosine kinases

(TyrK) phosphorylate specific target

tyrosine residues, phosphotyrosine phos-

phatases (PTP) remove the phosphates,

and Src Homology 2 (SH2) domains

recognize the modifications (Pawson

1995). Together, these three modules

form the ‘‘writer,’’ ‘‘eraser,’’ and ‘‘reader’’

toolkit that is common to many diverse

cellular information processing platforms

(Figure 1A). A rich array of diverse and

complex regulatory schemes can be

achieved through the dynamic interplay

of these three modular functions (Pawson

et al., 1993; Pawson, 1995; Bhattachar-

yya et al., 2006; Kholodenko, 2006).

A combination of these modules can

lead to higher-order functions (Figure 1B).

For example, there are several proteins

containing a combination of SH2 and

kinase domains that can generate

positive feedback (phosphorylation of

tyrosine sites leads to SH2-mediated

recruitment of the kinase, and subse-

quently, more extensive phosphorylation)

(Pawson, 2004). Similarly SH2-phospha-

tase domain combinations can generate

negative feedback (Tonks and Neel,

2001).

The three-part pTyr signaling toolkit

thus raises a classic question in evolu-

tionary biology: how do complex, interde-

pendent systems arise? It is clear why

a new system encompassing a writer,

eraser, and reader might be extremely

useful. But, given their interdependence,

how could these individual components

arise in a stepwise fashion consistent

with an evolutionary process? Proteins

that bind or remove a posttranslational

modification would seem useless without

an enzyme to generate the modification

and, in principle, would not provide

a fitness advantage leading to its retention

and spread. The pTyr signaling platform

provides a case study to look for plausible

stepwise pathways of the evolution of

a multipart system.

Here, we reconstruct a possible history

for the evolution of pTyr signaling. This

reconstruction is based on the recent

sequencing of the genomes of a number

of organisms that originated both before

and after the emergence of metazoans

from single-celled eukaryotic ancestors

(King et al., 2008). The genome sequence

of the choanoflagellate, Monosiga brevi-

collis, has been particularly illuminating

as choanoflagellates are thought to be

one of the closest single-celled relatives

of metazoans. We present a model for

how this three-part signaling system

could have plausibly evolved in a stepwise

manner. We propose that once the

complete three-part system was in place,

it may have rapidly taken hold in subse-

quent lineages because it could generate

new regulatory behaviors without signifi-

cant cross-interference with existing

regulatory circuits. We also discuss the

possible role of this new communication

system in facilitating the transformative

evolutionary shift to multicellularity.

Given the incomplete record, however,

such an evolutionary reconstruction is

Cell 142, September 3, 2010 ª2010 Elsevier Inc. 661

Page 30: Cell 100903

highly speculative. For example, we

cannot rule out more complex paths

involving cycles of evolutionary gain and

loss of components, nor the possibility

that similar components in distinct line-

ages have independent origins. Nonethe-

less, this model may provide a useful

framework for focusing studies of pTyr

signaling origins and the origins of analo-

gous multicomponent signaling plat-

forms.

We describe three possible stages in

the emergence of the modern pTyr

signaling toolkit, each represented by an

extant model organism (Figure 2). These

stages are representative; we do not

claim to define the exact path of evolution,

but rather focus on identifying the domi-

nant classes of stable intermediates that

can exist in the broader evolutionary land-

scape.

PTPs in a Pre-TyrosineKinase WorldWhat came first, TyrK, PTP, or SH2

domains? Sequence analysis suggests

that it was PTP domains. The genome of

a simple single-celled eukaryote like the

budding yeast Saccharomyces cerevisiae

shows no TyrK proteins and one proto-

SH2 domain, but a handful of PTP

proteins (Pincus et al., 2008) (Figure 2,

Stage 1). Most fungi have no more than

five PTP proteins, and several of these

have tyrosine phosphatase activity. We

refer to the single putative ‘‘SH2’’ domain

in yeast, found within the gene Spt6, as

a proto-SH2 domain because it does not

show pTyr binding (the domain has been

reported to show phospho-Ser/Thr

binding; Dengl et al., 2009). Thus, func-

tionally, it cannot be considered a

‘‘reader’’ domain that is part of a pTyr

regulatory system. These observations

suggest a simple model: the first step in

the evolution of the three-part pTyr

signaling machinery was likely to have

been the emergence of a functional tyro-

sine phosphatase. But why would PTPs

arise in the pre-tyrosine kinase world?

What functional use and fitness advan-

tage would this eraser domain provide in

organisms lacking a writer domain?

The answer may lie in the fact that some

Ser/Thr kinase domains, which are more

ancient than tyrosine kinases (dating

back close to the origins of eukaryotes),

can carry out sporadic but functionally

important phosphorylation of tyrosines.

Phosphoamino acid analysis of yeast

reveals a small but significant population

of pTyr (Schieven et al., 1986). Moreover,

certain events, such as the activation

of mitogen-activated protein kinases

(MAPKs) and inhibition of the cell-cycle

kinase Cdk1, are known to involve phos-

phorylation of tyrosine residues (for acti-

vation, a MAPK must be phosphorylated

by an upstream Ser/Thr kinase on both

a Thr and Tyr residue within its activation

loop; Cdk1 is phosphorylated on Tyr14

by the inhibitory kinase Wee1). These

tyrosine modifications are clearly not

recognized by SH2 domains, but they

exert direct allosteric effects within the

proteins in which they occur. Thus, PTP

domains may have provided a fitness

benefit by negatively modulating these

rare but functionally important phosphor-

ylation events. Consistent with this model,

the proteins PTP2 and PTP3 in yeast

clearly have a functionally important role

in downregulating MAPK-mediated sig-

naling in response to pheromones or

osmolarity changes, explaining their

fitness benefit (Pincus et al., 2008). In

addition, PTPs may have played a general

role buffering against the occasional

harmful stray phosphorylation of function-

ally important tyrosines.

Where did these PTPs come from?

PTPs are likely to have arisen from

a common ancestor of the related dual-

specificity phosphatases, which are also

found in most single-celled eukaryotes

(Kennelly, 2001; Alonso et al., 2004).

Dual-specificity phosphatases are cata-

lytic domains that can dephosphorylate

both pSer/Thr and pTyr substrates. The

PTP and dual-specificity phosphatase

catalytic domains are distinct but are

evolutionarily related. They share a

common fold and the core catalytic motif

HC(X)5R, in which a phosphocysteine

enzyme intermediate is generated during

catalysis. (Sometimes, both dual-speci-

ficity phosphatases and classical PTPs

are referred to as PTPs; here, we use

this nomenclature only for the classical

PTP domains that act only on pTyr). The

domains of dual-specificity phosphatases

have a shallower active site than classical

PTPs, which may explain why they can

dephosphorylate either Tyr or Thr/Ser

residues. In some lineages, dual-speci-

ficity phosphatases have functionally

diverged further, giving rise to members

that can act on lipid substrates, such as

the phosphoinositide phosphatases

PTEN and the myotubularins (Alonso

et al., 2004). Thus, the PTPs appear to

have arisen from a somewhat promis-

cuous class of multifunctional phospha-

tases.

Despite the presence of PTP proteins in

fungi, there are striking differences

Figure 1. The Writer, Reader, Eraser pTyr Toolkit(A) In pTyr signaling, the tyrosine kinase (TyrK), Src Homology 2 (SH2), and phosphotyrosine phosphatase(PTP) domains form a highly interdependent signaling platform. This platform serves as the writer, reader,and eraser modules, respectively, for processing pTyr marks.(B) Components of pTyr signaling can be used to build complex circuits. For example, recruitment of anSH2-TyrK protein to an initiating pTyr site can lead to amplification of tyrosine phosphorylation througha positive feedback loop.

662 Cell 142, September 3, 2010 ª2010 Elsevier Inc.

Page 31: Cell 100903

between these proteins and those

found in metazoans (Figure 2). For

example, there are far fewer PTPs in

fungi (�5/genome versus �40/genome

in metazoans) and they are considerably

less complex in domain architecture

(Pincus et al., 2008). Metazoan PTP

proteins tend to be large multidomain

proteins in which the PTP module has

been functionally recombined with multi-

ple other signaling modules. In contrast,

in fungi, the PTP domains are all either

in simple single-domain proteins or in

combination with a single rhodanase-like

domain (a putative regulatory domain

that is homologous to a class of sulfur

transfer enzymes; Bordo and Bork,

2002). Thus, fungal PTP proteins are

very simple (one to two domains) and

lack the combinatorial complexity of

metazoan PTP proteins. The simplicity

and low number of PTPs in yeast sug-

gests that in early single-celled eukary-

otes, PTP domains had fairly limited

Figure 2. Evolution of pTyr SignalingShown is a possible path for the emergence of phosphotyrosine (pTyr) signaling. We postulate three successive stages, each represented by what is observed ina modern organism. The thickness of the tree reflects the approximate degree of usage of pTyr signaling (thicker lines mean more usage). Stage 1 (exemplified bythe budding yeast Saccharomyces cerevisiae) reflects the situation in early eukaryotes, in which PTPs emerged but were limited in number and complexity. Theywere most likely used to reverse or process sporadic cross-phosphorylation of tyrosine residues by Ser/Thr kinases. S. cerevisiae has fewer than five PTPproteins and no functional SH2 or TyrK domains. Stage 2 reflects systems in which functional SH2 domains emerged that were able to bind to pTyr motifs.Together with Ser/Thr kinases with increased cross-reactivity for Tyr (such as tyrosine kinase-like or dual specificity Ser/Thr kinases), these systems may reflectthe most primitive of pTyr writer/reader/eraser systems. However, the lack of a dedicated Tyr kinase may have limited the utility and expansion of this toolkit. Thisstage is potentially represented by the slime mold, Dictyostelium discoideum. Stage 3 reflects systems that evolved after the emergence of the modern TyrKdomain. We postulate that the full writer/reader/eraser system was of so much greater utility that its use expanded dramatically. This likely resulted in manymore proteins in these families, as well as much more complex, multidomain architectures than those seen in the earlier stages. This stage is represented byboth the multicellular metazoan and unicellular choanoflagellate lineages.

Cell 142, September 3, 2010 ª2010 Elsevier Inc. 663

Page 32: Cell 100903

functional utility, especially when com-

pared to their broad and complex usage

in metazoans.

Unlike PTPs, there are no known pTyr-

binding SH2 domains in fungi, although

there is one clearly homologous domain

found in the yeast protein SPT6. This

protein, which has a domain with an

SH2-like sequence and fold, is involved

in the regulation of transcription elonga-

tion, and the SH2 domain binds to the

Ser/Thr phosphorylated C-terminal tail of

RNA polymerase II. The domain does

not bind to pTyr (Dengl et al., 2009). Inter-

estingly, a single SPT6 ortholog, with the

same overall domain architecture, is

found in all eukaryotes, including all fungi

and metazoans (but not prokaryotes). This

finding suggests that in early eukaryotes,

a proto-SH2 domain emerged to perform

a highly specialized function—one that

was unrelated to the flexible modular

pTyr recognition function of the modern

SH2 domain. This proto-SH2 domain

most likely did not ‘‘read’’ pTyr modifica-

tions, but instead recognized a special-

ized related modification. Thus, although

SPT6 is likely to represent an early

ancestor or relative that eventually gave

rise to modern SH2 domains, it cannot

be considered a functional part of a pTyr

regulatory toolkit. We therefore postulate

that early eukaryotes had only a pTyr

eraser function (mediated by PTPs) with

no specialized complementary reader or

writer functions.

In summary, the PTP domain and

a structural ancestor of the SH2 domain

appear to have arisen in early single-

celled eukaryotes, but are likely to have

functional origins that are not directly

related to their later function in modern

pTyr regulatory systems. These compo-

nents may have provided a limited but

incremental fitness advantage, even in

the absence of a specialized tyrosine

kinase domain.

Toward a Write/Read/Erase SystemIn the early days of a more sophisticated

pTyr-signaling system, we suggest that

a proto-SH2 domain (mostly likely a

homolog of the yeast Spt6 protein) in

a single-celled organism acquired the

new and functionally beneficial ability to

bind to pTyr-containing peptide motifs.

The slime mold Dictyostelium discoideum

has the simplest repertoire of bona fide

pTyr-binding SH2 domains and may

therefore provide a living representative

of this second evolutionary stage (Fig-

ure 2, Stage 2). Dictyostelium is a soil-

living amoeba that has a unicellular

lifestyle in the presence of bacterial

food. However, when food is depleted,

individual cells aggregate in response to

the chemoattractant cAMP to form

a multicellular structure, which then

develops into a fruiting body through the

differentiation of stalk and spore cells.

The rudimentary pTyr-SH2 system in

Dictyostelium is important for aspects of

this differentiation process, including

intracellular responses to both cAMP

and the morphogen differentiation in-

ducing factor or DIF (which induces the

differentiation of prestalk cells), as well

as for transcriptional regulation in res-

ponse to hyperosmotic stress. These

observations are consistent with early

pTyr-SH2 signaling playing a role in

cellular responses to changing environ-

mental conditions.

The Dictyostelium genome specifies 13

proteins with SH2 domains (as well as

a single Spt6 homolog). These 13 proteins

cluster into five basic domain architec-

tures, two of which are homologous to

metazoan SH2 proteins. Notably, Dic-

tyostelium has four STAT (signal trans-

ducers and activators of transcription)

proteins that are very similar to metazoan

STAT transcription factors (Kay, 1997;

Kawata et al., 1997). For example, they

all have an SH2 domain juxtaposed to

a DNA-binding region; they are inducibly

phosphorylated on tyrosine residues in

response to stress or the extracellular

signaling molecule DIF; they undergo

pTyr/SH2-mediated dimerization and

then translocate to the nucleus to regulate

the expression of specific genes. Dictyos-

telium also has an ortholog of the

mammalian E3 ubiquitin ligase Cbl, which

uses SH2 and Ring domains to couple

pTyr signals to the ubiquitination ma-

chinery (Langenick et al., 2008). The re-

maining three domain architectures of

Dictyostelium SH2 proteins are distinct

from those found in other sequenced

organisms. The LrrB protein has an SH2

domain linked to a leucine-rich repeat

domain (Sugden et al., 2010), whereas

the FbxB protein has an F-box followed

by an SH2 domain and ankyrin repeats.

In addition, the Shk proteins have a protein

kinase domain followed by an SH2

domain, a domain combination that is

somewhat similar to metazoan cyto-

plasmic tyrosine kinases like Src (Monia-

kis et al., 2001). The Shk catalytic domain,

however, lacks motifs characteristic of

bona fide tyrosine kinases and biochemi-

cally displays dual specificity toward

serine/threonine and tyrosine residues.

Indeed, Dictyostelium differs from

metazoans and choanoflagellates in that

its genome does not encode any modern

tyrosine-specific protein kinases. For

example, metazoan STAT proteins are

usually phosphorylated by Janus tyrosine

kinases (JAKs), but there are no JAKs in

Dictyostelium (Kay, 1997). This suggests

the possibility that signaling proteins con-

taining SH2 domains such as STATs

evolved before the modern tyrosine

kinases with which they are associated

in metazoans. The identity of the kinase

responsible for STAT tyrosine phosphory-

lation, and the consequent formation of

SH2-binding sites, remains mysterious.

How, then, is tyrosine phosphorylation

of Dictyostelium proteins such as the

STATs controlled? Thus far, genetic anal-

ysis has not identified a specific relevant

kinase, and it has been proposed that, in

contrast to mammalian STATs, there

may be basal constitutive phosphoryla-

tion of Dictyostelium STAT tyrosine sites,

which is regulated by changes in PTP

activity in response to extracellular

signals (Langenick et al., 2008). One of

the PTPs in Dictyostelium, PTP3, binds

and dephosphorylates STATc, thereby

blocking SH2-mediated dimerization and

STATc accumulation in the nucleus.

Signaling induced by the DIF morphogen

appears to transmit signals by inhibiting

PTP3 activity and consequently boosting

STATc tyrosine phosphorylation and

STATc-dependent gene expression.

Although Dictyostelium lacks true tyro-

sine kinases, it is noteworthy that its

genome has a significant expansion in

the number of putative dual-specificity

protein kinases (there are�70, also known

as tyrosine kinase-like or TKL kinases)

(Manning et al., 2008). This set includes

the Shk catalytic domain, described

above. It is unlikely that any of these

kinases are precursors of modern tyrosine

kinases. However, it is plausible that these

represent the first evolutionary form of

the ‘‘writer’’ function in a prototype pTyr

664 Cell 142, September 3, 2010 ª2010 Elsevier Inc.

Page 33: Cell 100903

three-part regulatory system. The union of

an SH2 domain and a dual specificity

kinase domain, as found in the Shk

proteins, may be an early example of link-

ing ‘‘reader’’ and ‘‘writer’’ modules to

achieve more complex functions such as

positive feedback. Nonetheless, the

limited functionality of the dual specificity

kinases in carrying out tyrosine phosphor-

ylation may have limited the capabilities of

this early system. This may explain the

very modest expansion of pTyr signaling

in organisms such as Dictyostelium.

These observations paint the following

picture of Dictyostelium pTyr signaling

and, by extension, of an early phase in

the evolution of pTyr communication.

SH2 domains have acquired pTyr-binding

activity and are found in several distinct

combinations with other types of signaling

domains. Among these, the STAT and Cbl

proteins are shared with metazoans,

whereas the LrrB, FbxB, and Shk proteins

are unique to Dictyostelium. But no dedi-

cated modern tyrosine kinases have

been found, and the dynamic control of

tyrosine phosphorylation may be primarily

regulated by PTPs. Although functionally

important for aggregation and differentia-

tion, the pTyr signaling system has not

acquired the pervasive influence evident

in M. brevicollis and metazoans, perhaps

because of the lack of an efficient tyrosine

kinase. Put another way, Dicytostelium

has effective pTyr readers and erasers,

but the writer is poorly developed.

Invention of TyrK and Expansionof the pTyr ToolkitCurrent analysis suggests that the

modern tyrosine kinases arose just prior

to the evolution of the metazoans. Aside

from metazoans, canonical tyrosine

kinases have thus far only been observed

in the choanoflagellates, which appear to

be the closest known single-celled rela-

tives of metazoans (King et al., 2008).

The absence of significant numbers of

such tyrosine kinases in any other branch

of life suggests that this new catalytic

domain evolved in a recent common

ancestor of choanoflagellates and meta-

zoans, most likely as a branch of the older

Ser/Thr kinases. Some bacteria do have

specialized tyrosine kinases (BY kinases),

but these resemble P loop NTPases

(nucleotide triphosphatases) and are

structurally unrelated to eukaryotic tyro-

sine kinases (Lee and Jia, 2009). It is

therefore probable that BY kinases

evolved separately from metazoan tyro-

sine kinases and operate in a different

fashion.

The new eukaryotic tyrosine kinase

domain appears to have been a game

changing innovation (Figure 2, Stage 3).

The total number of tyrosine kinase

proteins in both choanoflagellate and

metazoan species is in the range of 30–

150 per genome (Pincus et al., 2008;

Manning et al., 2008). Among sequenced

genomes, there is a striking absence of

species with only a small number of TyrK

proteins. This all-or-none sudden jump in

the number of TyrK proteins suggests

their importance as they appear to have

undergone rapid expansion and subse-

quent retention.

What is perhaps more striking is the

observation that the emergence of the

TyrK domain and its rapid expansion

correlates with an equally rapid expansion

of PTP and SH2 domains within the same

genomes (Pincus et al., 2008). Although

fungi and Dictyostelium have �5 PTP

proteins, metazoans, and choanoflagel-

lates have 30–40 per genome. Similarly,

Dictyostelium has approximately ten

SH2 domain-containing proteins (fungi

have none), whereas metazoans and

choanoflagellates have �100 each.

Thus, both PTP and SH2 proteins

undergo a roughly 10-fold increase in

number per genome after the emergence

of the TyrK domain. Moreover, the

proteins containing SH2 and PTP

domains become far more complex and

varied (Jin et al., 2009). For example, in

yeast and Dictyostelium, SH2 and PTP

proteins normally are very simple one or

two domain proteins. However, in line-

ages that have modern TyrK proteins,

SH2 and PTP proteins almost always

comprise three to ten domains.

These observations are consistent with

the following model. When a far more

efficient TyrK domain—or ‘‘writer’’ func-

tion—emerged, this dramatically increased

the functional utility of the pre-existing PTP

(eraser) and SH2 (reader) domains. As

a three-part toolkit—a catalytic domain to

generate pTyr, an interaction domain to

bind to these pTyr sites, and an enzyme

to dephosphorylate them—this domain

set could be used to encode and execute

a far wider and diverse range of regulatory

functions, thus leading to the subsequent

expansion of the complete set. Although

the PTP and SH2 domains had utility in

simpler organisms, their much larger

functional potential was not unleashed

until the emergence of the TyrK domain.

The rapid expansion of the pTyr

signaling machinery in the ancestors of

choanoflagellates and animals is reminis-

cent of how technology expands in

quantum jumps, especially in situations

involving codependent technologies. For

example, the value of the laser expanded

dramatically after the later invention of

the complementary technology of fiber

optics. This codependent technology al-

lowed lasers to be repurposed to rapidly

displace electrical transmission via

copper wires as the backbone of global

communication (Alwayn, 2004). Thus,

although lasers had standalone utility,

their major application had to await the

introduction of complementary tech-

nology. The expansion of molecular com-

ponents in biology is likely to be similar.

A toolkit of writer, reader, and eraser

functions may be of full use only when all

components are present. Thus, it may

be common for any system of this type

to show a quantum ‘‘all-or-none’’ expan-

sion only when the final piece of the toolkit

emerges.

Applying the New pTyr Toolkitto Different FunctionsAlthough both choanoflagellate and

metazoan lineages show a large expan-

sion of the three-part pTyr regulatory

machinery, the way in which these

components are used appears to be quite

different. When one examines the domain

types that co-occur with TyrK, SH2, or

PTP domains, one finds many distinct

combinations that are unique to each

lineage (Pincus et al., 2008; Manning

et al., 2008). These differences in domain

combinations imply distinct functions for

proteins containing these domains in the

choanoflagellate and metazoan lineages

(Li et al., 2009). Assuming that the evolu-

tion of new TyrK, SH2, and PTP proteins

occurred by recombination with new

accessory domains (Jin et al., 2009; Pei-

sajovich et al., 2010), this observation

also implies that the complete signaling

toolkit emerged only shortly before the

divergence of metazoans and choanofla-

gellates (i.e., shortly before the evolution

Cell 142, September 3, 2010 ª2010 Elsevier Inc. 665

Page 34: Cell 100903

of metazoan multicellularity) and that

much of the divergent expansion of these

domain families occurred after the lineage

split.

Thus, earlier assumptions that pTyr

signaling is only used in metazoan cell-

cell communication are clearly incorrect.

Choanoflagellates do not form the com-

plex and permanent cell-cell organization

that metazoans do, yet surprisingly they

have a comparable (if not greater) number

of pTyr signaling proteins (Manning et al.,

2008). Sequencing of other organisms

that arose near the origins of metazoans

is ongoing. Preliminary data also suggest

a large number of pTyr signaling proteins

in other single-celled relatives of meta-

zoans. Thus, it may be more reasonable

to view the pTyr signaling system as an

innovative but generic information pro-

cessing system that could potentially be

used for transmitting many different types

of information.

Orthogonal Signaling: A Platformfor Biological InnovationWhen the three-part pTyr system first

emerged, it presented a new platform

with which to transmit information that

was orthogonal to pre-existing signaling

systems. Because it was based on a

distinct covalent modification, new regu-

latory circuits could be assembled with

these components without significant

cross-interference with pre-existing net-

works. Thus, this brand new signaling

apparatus probably had a high encoding

potential for evolving dramatically new

functions, such as those involved in multi-

cellularity. One possible problem that

could be caused by the expansion of the

new pTyr signaling enzymes might be

excessive general phosphorylation of

tyrosine residues throughout the pro-

teome. Interestingly, however, organisms

using pTyr signaling may have developed

a simple solution to deal with this

problem—a decrease in the tyrosine

content of proteins across the proteome

is observed to correlate with tyrosine

kinase expansion (Tan et al., 2009).

A new orthogonal signaling system like

the pTyr signaling platform can be viewed

as analogous to a newly opened region in

the telecommunications spectrum. New

frequencies provide the opportunity for

transmitting large amounts of information

as there is little interference from existing

communication. Because of this valuable

high encoding potential, there is extreme

pressure to quickly fill this region of the

spectrum. Moreover, the exact type of

information carried by each region of the

spectrum is flexible—for example, the

same region of the spectrum can be as-

signed to different functions in different

countries. We hypothesize that the new

pTyr signaling system that emerged prior

to metazoans presented similar new

opportunities to transmit more informa-

tion. This virgin system was rapidly

exploited, though the way it has been

used appears to be different in the two

branches (metazoans and choanoflagel-

lates) that emerged after the complete

toolkit was established.

It is tempting to speculate that the

emergence of a new signaling system

with high encoding potential may have

played a key role in the emergence of

a new, complex biological function such

as metazoan multicellularity. Such large-

scale phenotypic evolutionary innova-

tions may require and coincide with

innovations in basic molecular compo-

nents (King, 2004; Rokas, 2008).

Indeed, we speculate that pTyr

signaling may provide a more general

model for the generation of multicompo-

nent biological systems, involving first a

limited stepwise development of elements

that together have a rudimentary biolog-

ical utility, followed by an explosive

spread, once all of the components of

the mature system are in place. Explora-

tion of this concept, and further analysis

of the evolution of pTyr signaling, will be

assisted by the increasing sequence infor-

mation being gathered for both unicellular

and multicellular eukaryotes (Srivastava

et al., 2010), which will no doubt yield

surprises akin to the discovery of exten-

sive pTyr signaling in M. brevicollis. More-

over, as genomic information bracketing

other major evolutionary transitions

becomes available, it will be interesting

to see whether these innovations are

also associated with the explosive expan-

sion of new molecular toolkits.

A key point here is that the specific

emergence of the pTyr toolkit may not

have been essential for the evolution of

multicellularity, but rather, any number of

new orthogonal signaling toolkits with

the same high encoding potential could

have served a similar role. Other analo-

gous new molecular information curren-

cies could have, in principle, been able

to serve as the substrate for dramatic

phenotypic innovation. In this context,

plants make extensive use of protein

phosphorylation and have numerous

transmembrane receptor Ser/Thr kinases,

but they lack conventional tyrosine

kinases, indicating that pTyr-based

signaling is not the only mechanism of

information transfer through which organ-

isms can achieve multicellularity.

Is pTyr Signaling Saturated?How close is the pTyr signaling system to

being saturated? Is there still available en-

coding potential that could be tapped for

the evolution of new pathways and

behaviors? It is difficult to answer these

questions. However, the fact that new

pTyr signaling proteins appear to be asso-

ciated with advanced processes like

adaptive immunity suggests that there

was still some remaining encoding poten-

tial in the system as late as the evolution of

mammals. The evolutionary history re-

constructed here begs many questions.

Are there new regulatory toolkits evolving

now or in the future? Will these new tool-

kits be the substrate required for the

next big evolutionary innovation?

The importance of new molecular tool-

kits is conversely also very relevant to

the emerging field of synthetic biology, in

which the goal is to engineer cellular

systems with new functions. A major

potential limitation is how to build such

new functions in a reliable fashion that

does not cross-interfere in unanticipated

ways with existing systems (Lim, 2010).

Can we develop new synthetic molecular

signaling currencies that are orthogonal

to existing natural ones, and would these

systems dramatically facilitate our ability

to reliably and predictably endow cells

with innovative new functions?

ConclusionsCurrent data suggest that PTP and SH2

domains evolved before modern TyrK

domains, most likely to process pTyr

modifications sporadically catalyzed by

Ser/Thr kinases. However, the PTP and

SH2 domain protein families did not

expand dramatically until the emergence

of an efficient TyrK. We postulate that

only with the complete toolkit of writer

(TyrK), reader (SH2), and eraser (PTP)

666 Cell 142, September 3, 2010 ª2010 Elsevier Inc.

Page 35: Cell 100903

domains, was the full encoding potential

of this system unleashed, leading to rapid

expansion and elaboration of these

domain families. This type of explosive

increase in component usage may prove

to be common to all multipart molecular

systems. The emergence of the modern

TyrK maps just prior to the split between

metazoans and choanoflagellates. These

two lineages appear to have used this

new molecular communication system in

distinct ways—multicellular metazoans

used it for cell-cell coordination, whereas

unicellular choanoflagellates used it for

distinct but as yet uncharacterized

functions.

Thus, we are able to reconstruct a plau-

sible model by which the pTyr signaling

machinery could have evolved in a rela-

tively simple stepwise manner into what

today is a complex and highly interdepen-

dent system. In this model, evolution is

opportunistic and forward looking,

borrowing, and repurposing machinery

that pre-exists. The first simple PTP

proteins likely arose from the more

ancient Ser/Thr phosphatase family and

may have been maintained initially as

a way to reverse the unavoidable occa-

sional tyrosine phosphorylation event

catalyzed by a Ser/Thr kinase. In some

cases, like the MAPKs, which are present

in all eukaryotes, these tyrosine phos-

phorylation events appear to have

become exploited and fixed as actual

parts of signal transmission, alongside

Ser/Thr phosphorylation events. SH2

domains also likely arose from a pre-ex-

isting fold in the SPT6 protein, which is

found in all eukaryotes but has no pTyr

binding activity. But this fold, once co-

opted for this function, began to expand,

most likely because of its ability to

contribute to a wider range of modular

signaling events. But the full utility of these

components was only unleashed upon

the emergence of the modern TyrK

domain, which led to the highly expanded

three-part system. One cannot help but

wonder what other simple pieces of

molecular machinery may be lying around

in today’s biological systems, of limited

utility now, but awaiting the emergence

of some as yet unknown complementary

component that will generate a complete

toolkit that will help to drive future evolu-

tionary innovation.

ACKNOWLEDGMENTS

We thank D. Pincus, B. Mayer, P.Beltrao,O. Hoeller,

R.Linding, G. Superti-Furga, N.King, N.Helman, L.

Holt, A. Horwitz, T. Miller, G.Manning, T. Hunter,

D.Morgan, J. Williams, H. Bourne, and I. Ernberg

for helpful comments. This work was supported by

the Howard Hughes Medical Institute (W.L.), the

National Institutes of health (GM55040, GM62583,

GM081879, and EY016546—W.L.), the Packard

Foundation (W.L.), the National Science Foundation

Synthetic Biology Engineering Research Center

(W.L.), the Canadian Institutes for Health Research

(T.P.—MOP-6849), Genome Canada (T.P.), and

the Canadian Cancer Society Research Institute

(T.P.).

REFERENCES

Alonso, A., Sasin, J., Bottini, N., Friedberg, I.,

Friedberg, I., Osterman, A., Godzik, A., Hunter,

T., Dixon, J., and Mustelin, T. (2004). Cell 117,

699–711.

Alwayn, V. (2004). Optical Network Design and Im-

plementation (Indianapolis, IN: Cisco Press).

Bhattacharyya, R.P., Remenyi, A., Yeh, B.J., and

Lim, W.A. (2006). Annu. Rev. Biochem. 75,

655–680.

Bordo, D., and Bork, P. (2002). EMBO Rep. 3, 741–

746.

Dengl, S., Mayer, A., Sun, M., and Cramer, P.

(2009). J. Mol. Biol. 389, 211–225.

Hunter, T. (2009). Curr. Opin. Cell Biol. 21,

140–146.

Jin, J., Xie, X., Chen, C., Park, J.G., Stark, C.,

James, D.A., Olhovsky, M., Linding, R., Mao, Y.,

and Pawson, T. (2009). Sci. Signal. 2, ra76.

Kawata, T., Shevchenko, A., Fukuzawa, M.,

Jermyn, K.A., Totty, N.F., Zhukovskaya, N.V., Ster-

ling, A.E., Mann, M., and Williams, J.G. (1997). Cell

89, 909–916.

Kay, R.R. (1997). Curr. Biol. 7, R723–R725.

Kennelly, P.J. (2001). Chem. Rev. 101, 2291–2312.

Kholodenko, B.N. (2006). Nat. Rev. Mol. Cell Biol.

7, 165–176.

King, N. (2004). Dev. Cell 7, 313–325.

King, N., Hittinger, C.T., and Carroll, S.B. (2003).

Science 301, 361–363.

King, N., Westbrook, M.J., Young, S.L., Kuo, A.,

Abedin, M., Chapman, J., Fairclough, S., Hellsten,

U., Isogai, Y., Letunic, I., et al. (2008). Nature 451,

783–788.

Langenick, J., Araki, T., Yamada, Y., and Williams,

J.G. (2008). J. Cell Sci. 121, 3524–3530.

Lee, D.C., and Jia, Z. (2009). Trends Biochem. Sci.

34, 351–357.

Li, W., Scarlata, S., and Miller, W.T. (2009).

Biochemistry 48, 5180–5186.

Lim, W.A. (2010). Nat. Rev. Mol. Cell Biol. 11,

393–403.

Manning, G., Young, S.L., Miller, W.T., and Zhai, Y.

(2008). Proc. Natl. Acad. Sci. USA 105, 9674–9679.

Moniakis, J., Funamoto, S., Fukuzawa, M.,

Meisenhelder, J., Araki, T., Abe, T., Meili, R.,

Hunter, T., Williams, J., and Firtel, R.A. (2001).

Genes Dev. 15, 687–698.

Pawson, T. (1995). Nature 373, 573–580.

Pawson, T. (2004). Cell 116, 191–203.

Pawson, T., Olivier, P., Rozakis-Adcock, M.,

McGlade, J., and Henkemeyer, M. (1993). Philos.

Trans. R. Soc. Lond. B Biol. Sci. 340, 279–285.

Peisajovich, S.G., Garbarino, J.E., Wei, P., and

Lim, W.A. (2010). Science 328, 368–372.

Pincus, D., Letunic, I., Bork, P., and Lim, W.A.

(2008). Proc. Natl. Acad. Sci. USA 105, 9680–9684.

Rokas, A. (2008). Annu. Rev. Genet. 42, 235–251.

Schieven, G., Thorner, J., and Martin, G.S. (1986).

Science 231, 390–393.

Srivastava, M., Simakov, O., Chapman, J., Fahey,

B., Gauthier, M.E., Mitros, T., Richards, G.S., Con-

aco, C., Dacre, M., Hellsten, U., et al. (2010).

Nature 466, 720–726.

Sugden, C., Ross, S., Bloomfield, G., Ivens, A.,

Skelton, J., Mueller-Taubenberger, A., and Wil-

liams, J.G. (2010). J. Biol. Chem. 285, 22927–

22935.

Tan, C.S., Pasculescu, A., Lim, W.A., Pawson, T.,

Bader, G.D., and Linding, R. (2009). Science 325,

1686–1688.

Tonks, N.K., and Neel, B.G. (2001). Curr. Opin. Cell

Biol. 13, 182–195.

Cell 142, September 3, 2010 ª2010 Elsevier Inc. 667

Page 36: Cell 100903

Leading Edge

Correspondence

What Controls T Cell ReceptorPhosphorylation?In a study published in Cell, Xu et al. (2008)

raised the intriguing possibility that tyro-

sine residues present in the immunore-

ceptor tyrosine-based activation motifs

(ITAMs) of the cytoplasmic domain of the

CD33 subunit of the T cell receptor (TCR)

are sequestered in the membrane. Here,

they are inaccessible to tyrosine kinases

that would otherwise drive signaling.

As a consequence, the authors proposed

that ‘‘triggering’’ of the TCR—that is, the

events leading to phosphorylation of the

receptor—must be dependent upon

some, as yet undefined, process that

liberates the ITAMs from the membrane.

In an accompanying Preview, Kuhns and

Davis (2008) invoked the idea that mem-

brane association constitutes a ‘‘safety

catch’’ that prevents untoward triggering

of the TCR. Such effects add an additional

layer of complexity to the TCR triggering

problem.

Using equilibrium dialysis, Xu et al.

showed that the cytoplasmic domain of

CD33 (CD33CD) partitioned into acidic

rather than zwitterionic lipids. They then

went on to derive a structural model for

ITAMs bound by acidic lipids using

nuclear magnetic resonance (NMR)-

based methods. A new in vivo FRET-

based assay suggested that the associa-

tion of the ITAM with the membrane was

dependent on a cluster of positively

charged residues in the membrane-proxi-

mal amino-terminal region of CD33CD;

mutation of these residues reduced the

association of CD33CD with the mem-

brane as detected by the FRET assay.

The obvious question was, does mem-

brane association prevent kinases from

accessing ITAMs in vivo? Xu et al. showed

that the wild-type CD33 ITAM is not phos-

phorylated in resting T cells, implying that

it may indeed be completely inaccessible

to kinases. Somewhat surprisingly, how-

ever, they did not go on to clinch the argu-

ment by showing that there was increased

phosphorylation of the membrane nonas-

sociating CD33CD mutant. The results

of this experiment turn out to be inconsis-

tent with the predictions of Xu and

colleagues.

To show that CD33CD is not phosphor-

ylated in resting Jurkat T cells, Xu et al.

used a chimeric protein consisting of

murine CD33CD fused to the extracellular

and transmembrane domains of the

human natural killer cell inhibitory recep-

tor KIR2DL3. We generated FLAG-tagged

forms of this protein and a second

chimera consisting of the extracellular

and transmembrane domains of KIR2DL3

fused with CD33CD mutated at the set of

clustered positively charged residues

proposed to mediate membrane asso-

ciation, that is, the mutant Emut1+2

described by Xu et al. (see Experimental

Procedures). We expressed the two

chimeras at similar levels in Jurkat

T cells using a lentivirus expression

system (Figure S1A) and probed for

phosphorylation of CD33CD by western

blotting (Figure S1B). Neither KIR2DL3/

CD33CD nor KIR2DL3/CD33CDEmut1+2

immunoprecipitated with anti-FLAG anti-

body showed detectable phosphorylation

in resting T cells (Figure S1B, lanes 1

and 2). Even after very long exposures of

the blots, we were unable to detect phos-

phorylation over background levels of

cross-reactivity of the antibody and we

saw no differences in signal for the wild-

type versus the mutant proteins (Fig-

ure S1B). This was also the case for

wild-type and Emut1+2-mutated forms

of full-length CD33 incorporated into the

TCR complex (data not shown).

These findings suggest either that

the Emut1+2 mutations do not release

the ITAMs from the membrane or that,

in vivo, the association with the mem-

brane is weaker than expected and

some other process prevents CD33CD

phosphorylation in resting cells. We sus-

pected that tyrosine phosphatases might

be responsible for the lack of phosphory-

lation of the wild-type and mutant forms of

CD33CD. To test this, we incubated cells

expressing the chimeric proteins with

the tyrosine phosphatase inhibitor, perva-

nadate (Swarup et al., 1982). Both the

wild-type and mutated forms of CD33CD

were heavily phosphorylated following

incubation of the cells for 20 min with

the inhibitor (Figure S1B, lanes 3 and 4).

Time course experiments indicated that

rather than enhancing it, the Emut1+2

mutations somewhat slowed the kinetics

of CD33CD phosphorylation (data not

shown). In vivo, therefore, the wild-type

form of CD33CD seemed to be at least

as accessible to tyrosine kinases as

the mutated, membrane nonassociating

form of the protein when kinase activity

could be observed by inhibiting phospha-

tase activity. Finally, CD33 was among the

more heavily tyrosine phosphorylated

proteins present in whole-cell lysates

from pervanadate-treated wild-type

Jurkat T cells (Figure S1C), emphasizing

its accessibility to kinases relative to other

kinase targets.

How can our observations be recon-

ciled with those of Xu et al.? Our experi-

ments suggest that membrane associa-

tion, if it occurs, has less impact on

CD33CD phosphorylation than might

have been expected on the basis of the

Xu et al. work. The strongest evidence for

lipid association is, of course, the NMR

structure of CD33CD stably bound to

acidic lipids. It should be borne in mind,

however, that we know little about the

actual distribution of inner leaflet

lipids around receptors embedded in the

membrane. Recent work by Zech et al.

(2009) shows that the acidic phospholipid

phosphatidylserine is enriched in mem-

brane sheets isolated from activated

T cells using beads coated with anti-CD3

antibodies, but the extent to which this

reflects the distribution of lipids in the

immediate vicinity of the TCR in resting

or activated cells is unclear. The density

or properties of the acidic lipids adjacent

to CD33 in resting cells might be insuffi-

cient to sustain an interaction of the type

demonstrated in vitro by Xu et al. Whatever

interaction ITAMs have with membranes

in vivo might be too weak and dynamic to

prevent access of the kinases.

The reasons for the low levelsofCD33CD

phosphorylation in resting T cells remain to

be fully worked out. In any discussion of the

possible constraints on tyrosine phosphor-

ylation of the TCR, some consideration

must be given to the protein tyrosine phos-

phatases (PTPs) present at the cell surface,

such as CD45 (Hermiston et al., 2003).

Leukocytes express as many as 105 CD45

molecules, i.e., 2–3 for every TCR (Wil-

liams and Barclay, 1986), each with very

668 Cell 142, September 3, 2010 ª2010 Elsevier Inc.

Page 37: Cell 100903

broad substrate specificity (Barr et al.,

2009); the catalytic activities of PTPs have

been estimated to be 10- to 1000-fold

higher than that of tyrosine kinases

(discussed by Fischer et al., 1991). Another

contributing factor is the activity of inhibi-

tory kinases such as Csk, which also

constrain T cell activation (Schoenborn

et al., 2009). Progress in understanding

TCR triggering will require the teasing

apart of these and other competing

factors and the identification of those

factors with the largest impact on net

phosphorylation of the TCR. The effects

of pervanadate on phosphorylation levels

in resting T cells observed by ourselves

(Figures S1B and S1C) and by others

many years ago (e.g., O’Shea et al., 1992)

suggest that PTPs could be key. For us,

the sheer weight of the numbers also

warrants their serious consideration.

SUPPLEMENTAL INFORMATION

Supplemental Information includes Experimental

Procedures and one figure and can be found with

this article online at doi:10.1016/j.cell.2010.08.018.

ACKNOWLEDGMENTS

This work was funded by the Wellcome Trust, the

United Kingdom Medical Research Council, and

the Portuguese Fundacao para a Ciencia e a Tec-

nologia.

Ricardo A. Fernandes,1,4 Chao Yu,1,4

Alexandre M. Carmo,2

Edward J. Evans,1

P. Anton van der Merwe,3

and Simon J. Davis1,*1Nuffield Department of Clinical Medicine

and Medical Research Council HumanImmunology Unit, The Weatherall Institute of

Molecular Medicine, University of Oxford,

Oxford OX3 9DU, UK2Group of Cell Activation and Gene

Expression, Institute for Molecular and

Cellular Biology, 4150-180 Porto, Portugal

and Abel Salazar Institute for BiomedicalSciences, University of Porto,

Porto 4099-003, Portugal3Sir William Dunn School of Pathology,

The University of Oxford, Oxford OX1 3RE,UK4These authors contributed equally to this

work*Correspondence: [email protected].

ac.uk

DOI 10.1016/j.cell.2010.08.018

REFERENCES

Barr, A.J., Ugochukwu, E., Lee, W.H., King, O.N.,

Filippakopoulos, P., Alfano, I., Savitsky, P.,

Burgess-Brown, N.A., Muller, S., and Knapp, S.

(2009). Cell 136, 352–363.

Fischer, E.H., Charbonneau, H., and Tonks, N.K.

(1991). Science 253, 401–406.

Hermiston, M.L., Xu, Z., and Weiss, A. (2003).

Annu. Rev. Immunol. 21, 107–137.

Kuhns, M.S., and Davis, M.M. (2008). Cell 135,

594–596.

O’Shea, J.J., McVicar, D.W., Bailey, T.L., Burns,

C., and Smyth, M.J. (1992). Proc. Natl. Acad. Sci.

USA 89, 10306–10310.

Schoenborn, J.R., Zhang, C., Shokat, K.M., and

Weiss, A. (2009). J. Immunol. 182, Abstract 35.4.

Swarup, G., Cohen, S., and Garbers, D.L. (1982). Bi-

ochem. Biophys. Res. Commun. 107, 1104–1109.

Williams, A.F., and Barclay, A.N. (1986). Glycopro-

tein antigens of the lymphocyte surface and their

purification by antibody affinity chromatography.

In Handbook of Experimental Immunology (Oxford:

Blackwell Scientific Publications), pp. 22.1–22.24.

Xu, C., Gagnon, E., Call, M.E., Schnell, J.R., Schwi-

eters, C.D., Carman, C.V., Chou, J.J., and

Wucherpfennig, K.W. (2008). Cell 135, 702–713.

Zech, T., Ejsing, C.S., Gaus, K., de Wet, B., Shev-

chenko, A., Simons, K., and Harder, T. (2009).

EMBO J. 28, 466–476.

Response

Multilayered Control of T CellReceptor Phosphorylation

We would like to respond to the Corre-

spondence by Fernandes et al., but first

we will summarize the data in our Cell

paper (Xu et al., 2008). Our study used

a biophysical approach to examine the

binding of the CD33 cytoplasmic domain

(CD33CD) to the plasma membrane. We

reported a new FRET assay to examine

CD33CD binding to the inner leaflet of

the plasma membrane in live Jurkat cells,

a transformed T cell line. Furthermore, we

developed an approach to determine

the NMR structure of CD33CD bound in

a lipid bilayer environment. The structure

showed that the two tyrosines of the im-

munoreceptor tyrosine-based activation

motif (ITAM) partitioned into the hydro-

phobic core of the bilayer in a dynamic

manner, with substantial movement of

the two tyrosines and other elements of

the cytoplasmic domain (Figure 6, Xu

et al., 2008). Microscopy data yielded

similar FRET values for CD33CD tagged

with C-terminal teal fluorescent protein

(TFP) and for a positive control in which

TFP was in close proximity to the plasma

membrane (there was a three amino acid

linker between the transmembrane and

TFP domains). These data indicated that

most of the cytoplasmic domain was

membrane bound in these transfected

nonstimulated Jurkat cells. However, we

did not claim that all CD33 cytoplasmic

domains in a given T cell are completely

bound to the plasma membrane. This

would be impossible in a biological sys-

tem because binding events always

follow equilibrium conditions, with the

size of the bound versus free fractions

being determined by the ratio of on and

off rates. It follows that changes in equilib-

rium can result in dissociation of CD33CD

from the membrane. Our data showed

a structure with substantial mobility in

which changes in equilibrium, such as

recruitment of the tyrosine kinase Lck,

could enable phosphorylation.

We wish to point out that we did not

claim that binding of CD33CD to the

membrane was the only mechanism that

prevents spontaneous T cell signaling.

It is well established that there is

multilayered control of T cell receptor

(TCR) signaling because complete

Cell 142, September 3, 2010 ª2010 Elsevier Inc. 669

Page 38: Cell 100903

phosphorylation of even one or a few TCRs

is sufficient to result in T cell activation

(Bergman et al., 1992; Davis and van der

Merwe, 2006). Without highly effective

mechanisms to prevent spontaneous

signaling, rampant chronic inflammation

and autoimmunity would result. Sponta-

neous signaling is inhibited by the Csk

kinase, which phosphorylates Lck kinase

at inhibitory tyrosine 505, as well as

multiple phosphatases (Bergman et al.,

1992; Davis and van der Merwe, 2006). In

agreement with Fernandes et al., we did

not detect phosphorylation of the

CD33CD Emut1+2 mutant protein in Jurkat

cells (Figure S1B, CD33-Mut time 0), which

was a negative control for the FRET exper-

iments. However, if binding of CD33

to the membrane together with other inhib-

itory mechanisms guards against sponta-

neous phosphorylation, this result is not

surprising. Phosphorylation may only be

favored with appropriate TCR localization

into microclusters that exclude the CD45

phosphatase (Varma et al., 2006), as well

as simultaneous colocalization of active

Lck bound to CD4/CD8 coreceptors close

to the clustered TCR-CD3 complexes.

Fernandes et al. assume that the

CD33CD Emut1+2 mutant protein would

normally interact with Lck. However, the

loss of six basic residues substantially

changes its charge properties. We

therefore directly compared the phos-

phorylation of the mutant and wild-type

CD33CD proteins in an in vitro phosphor-

ylation assay with the purified cyto-

plasmic domains of the wild-type and

Emut1+2 proteins and purified Lck in the

absence of liposomes (Figure S1A). We

observed greatly reduced phosphoryla-

tion of the mutant protein, even though

equal amounts of wild-type and mutant

proteins were used. The mutant protein

was not detectably phosphorylated with

100 ng of Lck kinase and was only phos-

phorylated at low levels with 400 ng Lck;

the wild-type protein was robustly phos-

phorylated under both conditions. Fer-

nandes et al. also mention delayed phos-

phorylation kinetics for this mutant

protein in cells treated with pervanadate,

which inhibits a broad range of phospha-

tases (data not shown). We confirmed

this finding and observed substantially

delayed phosphorylation of the mutant

protein after pervanadate treatment: the

wild-type protein was phosphorylated

after 2 min, whereas phosphorylation of

the mutant was delayed, with a 3.5- to

5-fold reduction at 10 min (Figure S1B).

However, both proteins were phosphory-

lated to a similar degree at a later time

point (30 min), as Fernandes et al. also

showed. The greatly reduced interaction

of the mutant protein with Lck as well as

the multiple control mechanisms regu-

lating T cell activation could account for

the absence of observable phosphoryla-

tion of this particular CD33 mutant in non-

stimulated Jurkat cells. It is not known

why mutation of six basic residues in the

N-terminal part of CD33CD reduces the

interaction with Lck, but conformational

changes in the cytoplasmic domain or

loss of long-range charge interactions

with Lck could be involved.

Pervanadate oxidizes the active site

cysteines of many different phosphatases

and is generated by mixing hydro-

gen peroxide with sodium orthovanadate

(Huyer et al., 1997). Given the potential

for unanticipated effects, we used our

FRET assay to examine the conse-

quences of pervanadate treatment on

CD33CD membrane binding (Xu et al.,

2008). The assay measures the interac-

tion of a TFP domain attached to the

C terminus of CD33CD with a lipophilic

dye R18 (octadecyl rhodamine B) incor-

porated in the plasma membrane; label-

ing the cells with R18 caused a decrease

in the TFP signal (quenching) due to

energy transfer from TFP to R18 (Fig-

ure S1C). In the absence of pervanadate,

the FRET efficiency was high (average

�50%; Figure S1D), which is similar to

the FRET signal obtained with the positive

control in which TFP is positioned close

to the plasma membrane (a three amino

acid linker between the transmembrane

domain and TFP) (Xu et al., 2008). After

20 min of pervanadate treatment, the

FRET signal was reduced on average

to �15% (Figure S1D), similar to the

negative control with a 50 amino acid flex-

ible linker between the transmembrane

domain and TFP (Xu et al., 2008). Perva-

nadate treatment thus results in dissocia-

tion of the CD33 cytoplasmic domain from

the membrane.

We next examined the potential mech-

anisms that could contribute to this effect.

We had previously shown that basic resi-

dues of CD33CD are critical for its binding

to the plasma membrane. Therefore, we

assessed whether the negative charge of

the inner leaflet of the plasma membrane

could be affected. Phosphatidylserine is

the most abundant negatively charged

lipid in the inner leaflet. We used a highly

specific calcium-independent phosphati-

dylserine probe, the Lactadherin C2

domain (Lact-C2), to study the dynamics

of phosphatidylserine distribution (Yeung

et al., 2008). Jurkat cells were transduced

with a lentivirus containing the Lact-C2

probe, and the probe’s localization in live

cells was analyzed by confocal micros-

copy (Figure S1E). Under resting condi-

tions, the Lact-C2 probe predominantly

associated with the plasma membrane in

75% of cells. In contrast, pervanadate

treatment reduced the Lact-C2 signal

at the plasma membrane and increased

the amount of this phosphatidylserine

probe in internal membranes in 92%

of cells, suggesting that the plasma

membrane inner leaflet had a reduced

negative charge (Figures S1E and S1F).

Phosphatidylserine at the cell surface

was also slightly increased, as measured

by annexin-V labeling (Figure S1G). We

conclude that pervanadate treatment

reduces the negative charge of the inner

leaflet and induces release of CD33CD

from the plasma membrane. These find-

ings complicate the interpretation of

experiments using pervanadate to assess

whether CD33CD is bound to the plasma

membrane.

A recent study confirmed that CD33CD

binds to the plasma membrane (Deford-

Watts et al., 2009), and prior work has

shown that CD3z is membrane bound

(Aivazian and Stern, 2000). Furthermore,

other cytoplasmic peptides have been

shown to interact with the inner leaflet

by similar mechanisms. McLaughlin and

colleagues showed that the MARCKS

peptide binds to the inner leaflet using

clusters of basic amino acids and five

phenylalanine residues (Zhang et al.,

2003). NMR measurements demon-

strated directly that these phenylalanine

residues are located in the hydrophobic

core of the lipid bilayer, similar to the tyro-

sines of CD33CD (Zhang et al., 2003).

Also, the cytoplasmic small GTPase Rit

localizes to the plasma membrane using

three clusters of basic amino acids

and interspersed hydrophobic residues;

mutation of a tryptophan in this segment

results in loss of membrane binding

670 Cell 142, September 3, 2010 ª2010 Elsevier Inc.

Page 39: Cell 100903

(Heo et al., 2006). Therefore, usage of

both basic and hydrophobic residues

for plasma membrane binding is a more

general theme that extends beyond the

CD33 and z cytoplasmic domains of the

TCR-CD3 complex.

In their Correspondence, Fernandes

et al. raise questions about the functional

relevance of the binding of CD33CD to the

plasma membrane during TCR activation.

It will be important to further study the

functional significance of the binding of

CD33CD to the plasma membrane in

a physiological setting using primary

T cells because the early events in TCR

triggering are very complex, and subtle

changes may result in receptor activation

after ligand binding. Analysis of multiple

CD33CD mutant proteins will be useful,

including mutants with reduced rather

than complete loss of membrane binding.

SUPPLEMENTAL INFORMATION

Supplemental Information includes Experimental

Procedures and one figure and can be found with

this article online at doi:10.1016/j.cell.2010.08.019.

Etienne Gagnon,1 Chenqi Xu,2

Wei Yang,2 H. Hamlet Chu,1

Matthew E. Call,3 James J. Chou,3

and Kai W. Wucherpfennig1,4,*1Department of Cancer Immunology & AIDS,

Dana-Farber Cancer Institute, Boston,

MA 02115, USA2Institute of Biochemistry and Cell Biology,

Shanghai Institutes for Biological Sciences,

Chinese Academy of Sciences,

Shanghai 200031, China3Department of Biological Chemistry and

Molecular Pharmacology4Department of Neurology and Program in

ImmunologyHarvard Medical School, Boston, MA 02115,

USA

*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.08.019

REFERENCES

Aivazian, D., and Stern, L.J. (2000). Nat. Struct.

Biol. 7, 1023–1026.

Bergman, M., Mustelin, T., Oetken, C., Partanen,

J., Flint, N.A., Amrein, K.E., Autero, M., Burn, P.,

and Alitalo, K. (1992). EMBO J. 11, 2919–2924.

Davis, S.J., and van der Merwe, P.A. (2006). Nat.

Immunol. 7, 803–809.

Deford-Watts, L.M., Tassin, T.C., Becker, A.M.,

Medeiros, J.J., Albanesi, J.P., Love, P.E., Wulfing,

C., and van Oers, N.S. (2009). J. Immunol. 183,

1055–1064.

Heo, W.D., Inoue, T., Park, W.S., Kim, M.L., Park,

B.O., Wandless, T.J., and Meyer, T. (2006).

Science 314, 1458–1461.

Huyer, G., Liu, S., Kelly, J., Moffat, J., Payette, P.,

Kennedy, B., Tsaprailis, G., Gresser, M.J., and

Ramachandran, C. (1997). J. Biol. Chem. 272,

843–851.

Varma, R., Campi, G., Yokosuka, T., Saito, T., and

Dustin, M.L. (2006). Immunity 25, 117–127.

Xu, C., Gagnon, E., Call, M.E., Schnell, J.R., Schwi-

eters, C.D., Carman, C.V., Chou, J.J., and

Wucherpfennig, K.W. (2008). Cell 135, 702–713.

Yeung, T., Gilbert, G.E., Shi, J., Silvius, J., Kapus,

A., and Grinstein, S. (2008). Science 319, 210–213.

Zhang, W., Crocker, E., McLaughlin, S., and Smith,

S.O. (2003). J. Biol. Chem. 278, 21459–21466.

Cell 142, September 3, 2010 ª2010 Elsevier Inc. 671

Page 40: Cell 100903

Leading Edge

Previews

Fishing Out a Sensor for Anti-inflammatory OilsAlan R. Saltiel1,*1Life Sciences Institute, Departments of Internal Medicine and Molecular and Integrative Physiology, University of Michigan Medical School,

Ann Arbor, MI 48109, USA*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.08.022

The u-3 fatty acids have anti-inflammatory and antidiabetic effects in humans. Now, Oh et al. (2010)demonstrate that the G protein-coupled receptor GPR120 is a receptor for u-3 fatty acids onmacrophages and fat cells. Activation of GPR120 by u-3 fatty acids inhibits multiple inflammationcascades in macrophages and reverses insulin resistance in obese mice.

Evidence providing an inflammatory link

between obesity and type 2 diabetes is

accumulating. In numerous animal and

clinical studies, obesity is associated

with a state of low-grade, chronic inflam-

mation in liver and adipose tissue,

which includes activation of the innate

immune system and the appearance of

proinflammatory immune cells (Hotamisli-

gil, 2006; Shoelson and Goldfine, 2009).

Most notably, macrophages conspire

with increased levels of inflammatory

cytokines to attenuate insulin action and

increase lipid accumulation.

Recent studies indicate that the NF-kB

and JNK (JUN N-terminal kinase) path-

ways play important roles in the com-

munication among macrophages, adipo-

cytes, and liver cells (Arkan et al., 2005;

Chiang et al., 2009; Solinas et al., 2007).

However, key questions still remain about

the initial establishment of this inflamma-

tory state. Do fat and liver cells first sense

an overload of energy and respond by

secreting chemokines, which then recruit

macrophages to the liver and fat? Or do

fatty acids in the diet directly initiate the

inflammatory cascade, and, if so, which

ones? In this issue of Cell, Oh et al.

(2010) address the latter question by iden-

tifying a sensor for u-3 fatty acids that is

upregulated in obese mice. Furthermore,

activation of this receptor exerts potent

anti-inflammatory effects that improve

insulin resistance and other symptoms of

metabolic syndrome in mice.

Previous studies suggest that saturated

fatty acids promote inflammation by acti-

vating Toll-like receptor 4 (TLR4) on fat

cells and macrophages (Shi et al., 2006).

In contrast, most unsaturated fats are

metabolically neutral. However, recent

evidence indicates that u-3 fatty acids

derived from fish oils, such as docosahex-

anoic acid (DHA) and eicosapentanoic

acid (EPA), have an anti-inflammatory

effect (Serhan et al., 2008).

To sort out the metabolic impact of

various types of fatty acids, Oh et al.

characterized the tissue expression pat-

terns of five G protein-coupled receptors

(GPCRs) known to bind and respond

to fatty acids. Among these receptors,

GPR120 was the only one with an expres-

sion profile that correlated well with a

potential role in regulating metabolism.

They found that GPR120 is highly ex-

pressed in adipose tissue macrophages,

fat cells, and specialized macrophages

in the liver called Kupffer cells. Moreover,

a high-fat diet increases the expression of

this receptor on macrophages, suggest-

ing that GPR120 might be controlled by

inflammatory signals.

GPR120 is an orphan receptor for

which no endogenous ligands are known.

Using a heterologous reporter system, Oh

et al. now find that the u-3 fatty acids

DHA, EPA, and palmitoleate are agonists

of GPR120. Furthermore, activation of

GPR120 by DHA antagonizes the proin-

flammatory effects of TNFa and lipopoly-

saccharide in a macrophage cell line.

DHA not only blocks the NFkB and JNK

pathways but also prevents expression

of cytokines (Figure 1, bottom right).

GPR120 is known to couple with the

Gq/11 familyof G proteins. After ligandbinds

and Gq/11 is released, G protein receptor

kinases phosphorylate the receptor. This

generates binding sites for b-arrestins,

which mediate internalization and down-

regulation of the receptors. However,

b-arrestins can also interact with down-

stream signaling molecules (Rajagopal

et al., 2010). In a series of gene silencing

experiments, Oh and colleagues demon-

strate that b-arrestin2 is essential for

the anti-inflammatory effects of u-3 fatty

acids in macrophage cells, but Gq is

surprisingly dispensable for this process.

Moreover, b-arrestin2 inhibits both the

JNK and NF-kB pathways by seques-

tering the TAK1 binding protein TAB1.

The inhibition of TAB1 prevents phos-

phorylation and thus activation of IkB

kinase upstream of NFkB and MKK4

(mitogen-activatedproteinkinasekinase4)

upstream of JNK.

These new insights into the sensing and

signaling of GPR120 offered Oh and

colleagues a unique opportunity to study

the molecular mechanisms underlying

the metabolic benefits of u-3 fatty acids

in vivo. First, Oh and colleagues establish

that the high-fat diet given to mice in the

laboratory is low in u-3 fatty acids. They

then show that supplementing this diet

with DHA and EPA reverses the delete-

rious effects that the high-fat diet has on

glucose homeostasis and lipid storage

(i.e., steatosis). Although the authors do

not address whether u-3 fatty acids can

prevent insulin resistance or glucose intol-

erance in mice, they do demonstrate that

DHA and EPA reverse insulin resistance

caused by the high-fat diet. Moreover,

disruption of the GPR120 gene abolishes

the benefits of u-3 fatty acids on glucose

homeostasis and insulin sensitivity in

mice. These results demonstrate the

crucial role GPR120 plays in the meta-

bolic benefits of DHA and EPA. Interest-

ingly, mice receiving bone marrow

transplants from the GPR120-deficient

mice are also resistant to the beneficial

672 Cell 142, September 3, 2010 ª2010 Elsevier Inc.

Page 41: Cell 100903

properties of DHA and EPA. Thus, the u-3

fatty acids appear to act primarily through

macrophages.

Taken together with previous data,

these findings from Oh and colleagues

support a model in which dietary fatty

acids control the inflammatory properties

of macrophages in adipose tissue by

regulating the activity and expression

of opposing receptors (Figure 1). With

a normal diet containing a balanced

ratio of saturated and u-3 unsaturated

fatty acids, anti-inflammatory M2 macro-

phages protect adipose cells by damp-

ening excess inflammation and maintain-

ing insulin sensitivity of fat cells (Lumeng

et al., 2007). When mice are given

a high-fat diet with excess calories

and little u-3 fatty acids, TLR4 is left

unchecked in fat cells. The activated

receptor induces expression and release

of chemokines, such as MCP-1 (mono-

cyte chemotactic protein-1), which then

recruit proinflammatory M1 macrophages

into adipose tissue (Lumeng et al., 2007).

These cells produce cytokines, such as

TNFa, which further activate the macro-

phages and attenuate insulin action in

adipocytes. Eventually, this leads to

local and then systemic insulin resistance.

However, these activated M1 macro-

phages also express elevated levels of

GPR120. Thus, addition of u-3 fatty acids

to the diet stimulates GPR120 and

initiates a signaling pathway through

b-arrestin2, which blocks the effects of

TLR4 and inflammatory cytokine recep-

tors. This reduces the inflammatory state

of these cells and simultaneously pro-

motes the return of anti-inflammatory

M2 macrophages to adipose tissue,

which leads to the restoration of insulin

sensitivity.

This model raises many new questions

about the crosstalk between adipose

tissue and immune cells in obesity. First,

could GRP120 activation serve as

a homeostatic mechanism in macro-

phages to resolve inflammation in obese

animals? If so, then what controls

GPR120 expression, and what is its phys-

iological role in adipose and liver tissue?

Also, does GPR120 respond to endoge-

nous ligands to control macrophage

activity? Although the new study by Oh

and colleagues explains how activation

of GPR120 inhibits inflammatory path-

ways, it is still unknown how this receptor

increases the presence of anti-inflamma-

tory M2 macrophages in adipose tissue

and from where these cells arise. Do M2

macrophages (or their precursors)

express GPR120 in order to develop or

maintain the anti-inflammatory pheno-

type, or does activation of GPR120

induce the transdifferentiation of inflam-

matory M1 macrophages to anti-inflam-

matory M2 macrophages in situ?

Finally, previous studies demonstrate

that fish oils have diverse benefits on

multiple tissues. For example, u-3 fatty

acids can inhibit the production of proin-

flammatory eicosanoids and serve as

precursors for resolvins, (i.e., protective

lipids that help reduce inflammation)

(Serhan et al., 2008). In addition, fish oils

help prevent cardiovascular disease and

have positive effects on many inflamma-

tory disorders, such as arthritis, asthma,

and ulcerative colitis. Is GPR120 the only

receptor responsible for these various

benefits? Future studies are also needed

to determine whether dietary supple-

ments and ingestion of fatty fish can

provide high enough concentrations of

circulating u-3 fatty acids to promote

GPR120 activation. Nevertheless, the

new insights presented by Oh and

colleagues into the anti-inflammatory

mechanisms of u-3 fatty acids provide

a platform for investigating these impor-

tant questions. Plus, the identification of

GPR120 pinpoints a new therapeutic

target for treating the inflammatory state

associated with obesity and type 2 dia-

betes. This alone is cause for a bit of

excitement.

REFERENCES

Arkan, M.C., Hevener, A.L., Greten, F.R., Maeda,

S., Li, Z.W., Long, J.M., Wynshaw-Boris, A., Poli,

G., Olefsky, J., and Karin, M. (2005). Nat. Med.

11, 191–198.

Chiang, S.H., Bazuine, M., Lumeng, C.N., Geletka,

L.M., Mowers, J., White, N.M., Ma, J.T., Zhou, J.,

Qi, N., Westcott, D., et al. (2009). Cell 138,

961–975.

Hotamisligil, G.S. (2006). Nature 444, 860–867.

Figure 1. A Model for How u-3 Fatty Acids Preserve Insulin Sensitivity through GPR120A high-fat diet with a disproportionate ratio of saturated fatty acids to u-3 fatty acids triggers activation ofToll-like receptor 4 (TLR4) in adipocytes and circulating immune cells. This launches an inflammatorycascade that results in the recruitment of proinflammatory M1 macrophages, increased secretion ofTNFa, and insulin resistance in adipocytes. The addition of u-3 fatty acids to the diet activates theG protein-coupled receptor GPR120 on proinflammatory M1 macrophages (Oh et al., 2010), which inturn attenuates the inflammatory response and recruits anti-inflammatory M2 macrophages to adiposetissue. Eventually, these M2 macrophages restore secretion of interleukin-10 and improve insulinsensitivity.

Cell 142, September 3, 2010 ª2010 Elsevier Inc. 673

Page 42: Cell 100903

Lumeng, C.N., Bodzin, J.L., and Saltiel, A.R.

(2007). J. Clin. Invest. 117, 175–184.

Oh, D.Y., Talukdar, S., Bae, E.J., Imamura, T.,

Morinaga, H., Fan, W., Li, P., Lu, W.J., Watkins,

S.M., and Olefsky, J.W. (2010). Cell 142, this issue,

687–698.

Rajagopal, S., Kim, J., Ahn, S., Craig, S., Lam,

C.M., Gerard, N.P., Gerard, C., and Lefkowitz,

R.J. (2010). Proc. Natl. Acad. Sci. USA 107,

628–632.

Serhan, C.N., Chiang, N., and Van Dyke, T.E.

(2008). Nat. Rev. Immunol. 8, 349–361.

Shi, H., Kokoeva, M.V., Inouye, K., Tzameli, I., Yin,

H., and Flier, J.S. (2006). J. Clin. Invest. 116, 3015–

3025.

Shoelson, S.E., and Goldfine, A.B. (2009). Nat.

Med. 15, 373–374.

Solinas, G., Vilcu, C., Neels, J.G., Bandyopad-

hyay, G.K., Luo, J.L., Naugler, W., Grivennikov,

S., Wynshaw-Boris, A., Scadeng, M., Olefsky,

J.M., and Karin, M. (2007). Cell Metab. 6,

386–397.

A New Spin on Planar Cell PolarityPatricio Olguin1 and Marek Mlodzik1,*1Department of Developmental and Regenerative Biology, Mount Sinai School of Medicine, New York, NY 10029, USA

*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.08.025

The generation of planar cell polarity (PCP) and tissue shape during morphogenesis is tightly linked,but it is not clear how. Aigouy et al. (2010) now show in the developing Drosophila wing that PCPinitially has a radial orientation that becomes realigned to the proximal-distal axis of organ shapeby mechanical forces and cell rearrangements mediated by Dachsous.

Most tissues and organs composed of and

organized as epithelial cell layers display,

in addition to the common apical-basal

epithelial polarity, a polarity within the

epithelial plane. This is commonly referred

to as planar cell polarity (PCP). Genetic

studies in the fruit fly Drosophila mela-

nogaster have established that there are

two molecular systems coordinating the

cellular asymmetries in the plane of

tissues. These include the Frizzled/PCP

signaling pathway containing the Van-

Gogh (Vang, also known as Strabismus/

Stbm) protein and other factors (Strutt,

2003; Seifert and Mlodzik, 2007) and

a pathway mediated by the protocadher-

ins Fat and Dachsous (Lawrence et al.,

2007). Although the molecular relation-

ships between these two systems are

unclear, there is strong evidence that

they act in parallel, probably affect

different effectors, and may compensate

for each other in some tissues (Lawrence

et al., 2007; Wu and Mlodzik, 2009).

In the developing Drosophila wing,

cellular asymmetries in the plane of the

epithelium are first detected at later pupal

stages along the proximal-distal axis

(at around 24–30 hr after puparium forma-

tion). The core Frizzled/PCP factors form

two distinct complexes that become

localized asymmetrically to either the

proximal (Vang/Stbm and associated

proteins) or the distal side of pupal wing

cells (Frizzled and associated proteins).

These complexes are stabilized by feed-

back loop interactions among themselves

(Seifert and Mlodzik, 2007; Strutt, 2003). In

addition, Frizzled and Vang/Stbm protein

complexes may be required earlier to

coordinate global tissue polarity/PCP

within the wing epithelium (Classen et al.,

2005; Wu and Mlodzik, 2009).

Supporting this idea, the new study by

Aigouy et al. (2010) in this issue of Cell

helps to establish that subcellular asym-

metries among the Frizzled/PCP core

group proteins are already present at

early pupal stages during wing develop-

ment (14–15 hr after puparium formation

or earlier). Strikingly, the Frizzled/PCP

complexes display radial polarity that is

perpendicular to the wing margin

(Figure 1A), confirming that coordination

of global Frizzled/PCP signaling is estab-

lished early in pupal fly wings. The authors

further demonstrate that these early

asymmetries indeed depend on Frizzled-

Vang/Stbm signaling as the nonautono-

mous behavior of frizzled mutant cell

patches (clones) affecting the polarity of

wild-type cells flanking the frizzled mutant

cells (Vinson and Adler, 1987) is already

observed at this stage. Strikingly, in

contrast to the nonautonomous effects

observed at the distal side of frizzled

mutant cell patches in late pupal and adult

wings (Vinson and Adler, 1987), early friz-

zled clones influence the polarity of wild-

type cells residing between the wing

margin and the clone within the radial

polarity axis. This confirms a ‘‘signaling

axis’’ toward the wing margin at early

stages. Taken together, the observations

of Aigouy et al. (2010) indicate that (1)

PCP, mediated by Frizzled-Vang/Stbm

signaling, is established during late-

larval/early-pupal stages in a radial axis

perpendicular to the margin and (2) the

polarity/PCP seen in the adult wing is

a result of cellular rearrangements during

wing morphogenesis within the prox-

imal-distal axis that are dependent on

Dachsous.

How is polarity realigned along the

proximal-distal axis as morphogenesis

proceeds? As PCP is already established

674 Cell 142, September 3, 2010 ª2010 Elsevier Inc.

Page 43: Cell 100903

at early stages, its final align-

ment from the radial orienta-

tion to the proximal-distal

axis must be redirected

through active relocalization

of Frizzled and Vang/Stbm

complexes and/or through

the shifting or moving of the

cells as a whole. For

example, polarity could be

achieved by rotation of the

cells toward the distal axis

as happens with the rotation

of photoreceptor cell clusters

(ommatidia) in the Drosophila

eye toward the anterior-

posterior axis (Seifert and

Mlodzik, 2007). But how

would such a rotation be

regulated? At early stages of

pupal development, the

proximal half of the wing

epithelium (hinge) and the

wing blade are similar in size

(Figure 1A). Subsequently,

preceding and coinciding

with PCP realignment, the

hinge contracts and gener-

ates an anisotropic mechan-

ical stress on the blade,

which leads to its elongation

in the proximal-distal axis

(Figure 1B). Strikingly, epithelial cell elon-

gation could drive the realignment of

cortical microtubules with the proximal-

distal axis, which appears essential for

the delivery of Frizzled to the distal side

of cells (Shimada et al., 2006). New work

appearing in Developmental Cell by

Uemura and colleagues (Harumoto et al.,

2010) shows that, prior to hinge contrac-

tion, cortical microtubules align perpen-

dicular to the margin in the proximal

region of the wing blade. This supports

a general role for the orientation of cortical

microtubules in PCP and in the realign-

ment of PCP later in development.

Quantitative analyses of time-lapse

imaging of the pupal wing by Aigouy et al.

(2010) show that, in response to the aniso-

tropic stress, cells move with respect to

each other in a proximal direction and

inwardswithdifferentvelocities (Figure1B).

This behavior causes shear and the local

rotation of cells, mainly clockwise in the

anterior and anticlockwise in the posterior

half of the blade (Figure 1B). As a conse-

quence, PCP is reoriented from a radial to

a proximal-distal axis within the wing (Fig-

ure 1C). Interestingly, during this remodel-

ing process, the global coordination/long-

range coherence of PCP is diminished,

which may be the reason why previous

studies of PCP during development have

missed the early asymmetry/polarity of

PCP core proteins. As pupal wing cells re-

pack as they acquire a hexagonal shape,

the global coordination/long-range coher-

ence of PCP increases again. This

phenomenon is a consequence of the

persistence of the Vang/Stbm and Frizzled

complexes at boundaries formed in the

early stages of development and of the

proximal-distal alignment of new bound-

aries. On the other hand, the Frizzled/

PCP core factors are required for hexag-

onal cell packing, probably by polarizing

membrane trafficking along the proximal-

distal axis (Classen et al., 2005). Thus,

both early polarity and cellular packing

would feed in to one another to shape

and repolarize the epithelia.

It has been proposed that the Fat/

Dachsous system would provide ‘‘global

cues’’ that orient the initial

polarity of Frizzled/PCP

complexes, but this hypoth-

esis has been challenged by

strong genetic evidence

showing that the Fat/Dachs-

ous and Frizzled/PCP

systems act in parallel (Law-

rence et al., 2007). During fly

larval stages, the Fat/Dachs-

ous system is required to

regulate the growth and

shape of the wing, the latter

(at least in part) by orienting

the axis of cell division

perpendicular to the margin

(Baena-Lopez et al., 2005).

Aigouy et al. (2010) show

that high expression of

Dachsous in the hinge region

is not required for its contrac-

tion, but correct Dachsous

levels in the wing blade are

required for the wing blade

to respond to the anisotropic

mechanical stress that

orients cell elongation along

the proximal-distal axis.

Moreover, cell polarity

defects correlate with the

inversion of local tissue rota-

tion in wild-type wings, with

altered levels of Dachsous in the posterior

compartment, suggesting that Dachsous

imbues cells with the ability to respond

coordinately to mechanical stress. Inter-

estingly, Harumoto et al. (2010) show

that Dachsous and Fat are required to

align cortical microtubules along the prox-

imal-distal axis and that Dachsous biases

the direction of microtubule growth from

high to low Dachsous levels, similar to its

role in cell orientation. Whether the Fat/

Dachsous system regulates cell remodel-

ing by controlling the polarity of cortical

microtubules or vice versa remains to be

resolved.

In their elegant new study, Aigouy et al.

(2010) analyzed the timeline of events for

the establishment of PCP in the devel-

oping Drosophila wing. They have

provided evidence that the early Friz-

zled/PCP core polarization toward the

wing margin (in a radial orientation) is real-

igned along the proximal-distal axis by

anisotropic mechanical stress and

Dachsous-mediated tissue remodeling.

These conclusions are consistent with,

Figure 1. Planar Cell Polarity in the Fly Wing(A) During fly pupal development, the initial axis of planar cell polarity (PCP) isradial, that is, oriented toward the wing margin (black arrows).(B) Asdevelopment proceeds, the hinge region (blue) contracts creatingan aniso-tropic mechanical stress on the wing blade, resulting in movements of wing cellsand realignment of PCP to the proximal-distal axis. Green arrows indicate thedirectionofcellmovement, and redarrows showthe directionofcellular rotations.These processes take place between 14 and 24 hr after puparium formation.(C) The final orientation of PCP is in the proximal-distal axis in late pupal/adult flywings (black arrows).

Cell 142, September 3, 2010 ª2010 Elsevier Inc. 675

Page 44: Cell 100903

and supported by, the phenotypic PCP

features of Frizzled/PCP core group

genes on one side and that of the Fat/

Dachsous system on the other. Flies

carrying mutations in Frizzled/PCP core

proteins exhibit defects in PCP

throughout the wing. In contrast, the

Fat/Dachsous system mainly affects

polarity in the proximal half of the wing,

as this area strongly depends on cellular

realignment and rotation during the

switch to the proximal-distal PCP axis.

Together, these observations provide an

exciting new framework for under-

standing the generation of PCP and its

relation to new mechanisms that sculpt

the shape of organs in general.

REFERENCES

Aigouy, B., Farhadifar, R., Staple, D.B., Sagner, A.,

Roper, J., Julicher, F., and Eaton, S. (2010). Cell

142, this issue, 773–786.

Baena-Lopez, L.A., Baonza, A., and Garcıa-Bel-

lido, A. (2005). Curr. Biol. 15, 1640–1644.

Classen, A.K., Anderson, K.I., Marois, E., and Ea-

ton, S. (2005). Dev. Cell 9, 805–817.

Harumoto, T., Ito, M., Shimada, Y., Kobayashi,

T.J., Ueda, H.R., Lu, B., and Uemura, T. (2010).

Dev. Cell 19. Published online September 2,

2010. 10.1016/j.devcel.2010.08.004.

Lawrence, P.A., Struhl, G., and Casal, J. (2007).

Nat. Rev. Genet. 8, 555–563.

Seifert, J.R., and Mlodzik, M. (2007). Nat. Rev.

Genet. 8, 126–138.

Shimada, Y., Yonemura, S., Ohkura, H., Strutt, D.,

and Uemura, T. (2006). Dev. Cell 10, 209–222.

Strutt, D. (2003). Development 130, 4501–4513.

Vinson, C.R., and Adler, P.N. (1987). Nature 329,

549–551.

Wu, J., and Mlodzik, M. (2009). Trends Cell Biol. 19,

295–305.

Viable Rat-Mouse Chimeras:Where Do We Go from Here?Davor Solter1,*1Institute of Medical Biology, A*STAR, 138648 Singapore, Republic of Singapore

*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.08.021

In a tour-de-force study, Kobayashi et al. (2010) describe the first viable rat-mouse chimeras anddemonstrate that rat induced pluripotent stem (iPS) cells can rescue organ deficiency in mice.Rat iPS cells formed a fully functional pancreas when injected into mouse blastocysts lacking thePdx1 gene required for pancreas formation.

Experimentally produced chimeras

between different mouse strains (Tarkow-

ski, 1961) have been an exceedingly

useful tool for developmental biologists,

contributing to our understanding of the

establishment of cell lineages, cell deter-

mination, and the development of the

immune system and other organs. In this

issue of Cell, Kobayashi et al. (2010)

dramatically extend the potential of mam-

malian chimeras with their report of viable

rat-mouse chimeras that can develop to

term and become fully functional adults.

In their study, Kobayashi and col-

leagues relied on previous knowledge

but also added a few new wrinkles. They

first derived mouse and rat embryonic

stem (ES) cells and induced pluripotent

stem (iPS) cells using standard methods

for the mouse and capitalizing on the

recent isolation of rat ES and iPS cells

(Buehr et al., 2008; Li et al., 2008). Both

mouse and rat cells were tagged with

different fluorescent dyes, allowing the

authors to follow their distribution in the

developing chimeras. The authors wanted

to prove, at least in principle, that xenoge-

neic organ complementation could be

achieved, that is, that donor cells of one

species could rescue a defect in organ

development in a recipient of a different

species. So, as a first step, they set out

to produce viable chimeras between rats

and mice, even though many previous

efforts to make such chimeras had failed.

The only viable intergeneric chimera—

that is, a hybrid between animals from

different genera—reported so far is

the geep between a sheep (Ovis aries)

and a goat (Capra hircus) (Fehilly et al.,

1984).

To test the possibility that viable rat-

mouse chimeras could be formed,

Kobayashi et al. (2010) injected fluores-

cently labeled mouse or rat iPS cells into

rat or mouse blastocysts, respectively,

and returned them to blastocyst-compat-

ible pseudopregnant females (that is,

foster mothers of the same species as

the blastocysts). The authors then exam-

ined the resulting fetuses, newborns, and

adults and found evidence of a substantial

contribution of donor stem cells to tis-

sues and organs of the host (Figure 1A).

Despite a big contribution of donor cells,

the size of newborn and adult chimeras

(with one exception) was determined by

676 Cell 142, September 3, 2010 ª2010 Elsevier Inc.

Page 45: Cell 100903

the species of the host blastocyst. It is not

clear whether it is the embryo itself or the

uterine environment that determines the

extent of chimera growth. To distinguish

between these possibilities, one would

have to transfer chimeric embryos into

the uterus of pseudopregnant females of

the same species as the donor stem cells

(not the blastocysts). Previous studies

suggest that such experiments would fail

because of the need for compatibility

between the fetal part of the placenta and

the uterus (Rossant et al., 1982).

Besides controlling the size and growth

of the chimera, the host blastocyst seems

to impose additional morphogenetic reg-

ulation. The postimplantation develop-

ment of normal rat and mouse embryos

is very similar, but there are differences

in organ morphogenesis. One of the

most noticeable differences is the pres-

ence of a gall bladder in mice and its

absence in rats. In all adult chimeras

produced by combining rat stem cells

and mouse blastocysts, the resulting

‘‘mouse-like’’ chimeras had a gall bladder

despite the significant contribution of rat

cells to abdominal organs. Reciprocal

chimeras were ‘‘rat-like’’ and, again,

despite a significant contribution from

mouse cells to abdominal organs, the

gall bladder was absent. These results

suggest that cells of the blastocyst inner

cell mass possess a ‘‘morphogenetic’’

capacity that controls the behavior of

injected stem cells at all developmental

stages.

This may explain why Kobayashi et al.

were able to successfully inject rat stem

cells into mouse blastocysts, whereas

insertion of the rat inner cell mass into

the mouse blastocyst cavity did not result

in viable rat-mouse chimeras (Gardner

and Johnson, 1973). This notion could

be tested further using tetraploid comple-

mentation (that is, donor ES or iPS cells

are injected into tetraploid blastocysts)

to produce rat-mouse chimeras (Nagy

et al., 1993). Tetraploid blastocyst cells

cannot participate in formation of the

embryo proper; thus, the resulting fetus

(and adult) is derived entirely from the

injected cells, whereas the placenta and

extraembryonic membranes are derived

from the tetraploid blastocyst. It remains

to be seen whether this approach could

produce a fetus derived entirely from

mouse ES cells after their injection into

a rat tetraploid blastocyst that then

develops in the uterus of a pseudopreg-

nant rat female.

A major goal of the Kobayashi et al.

study was to determine whether stem cells

from a xenogeneic donor mammal could

correct a genetic defect in a recipient

mammal of a different species. So, in their

next set of experiments, the authors in-

jected rat iPS cells into recipient mouse

blastocysts that lacked the Pdx1 gene,

Figure 1. Generating Rat-Mouse Chimeras(A) Induced pluripotent stem (iPS) cells were derived from adult mouse and rat cells and were labeled with different fluorescent proteins. Rat (blue) and mouse(red) iPS cells were injected into reciprocal blastocysts (mouse into rat and vice versa) to produce intergeneric chimeras. From these blastocysts, severalchimeras were born and some survived to adulthood. The contribution of injected donor stem cells was observed throughout the body of the host. The sizeand morphology of the newborn and adult chimeras was determined by the host blastocyst.(B) Fluorescently labeled rat iPS cells (blue) were injected into normal mouse blastocysts (left) or blastocysts lacking the Pdx1 gene (right), which encodes thetranscription factor pancreatic and duodenal homeobox 1 that is required for pancreas development. Chimeras derived from normal or Pdx1-deficient mouseblastocysts showed an extensive contribution of rat cells to all tissues. However, in the Pdx1-deficient chimeras, the entire pancreas was derived from donorrat cells (inset, blue) and was fully functional, including production of insulin by b islet cells.

Cell 142, September 3, 2010 ª2010 Elsevier Inc. 677

Page 46: Cell 100903

which encodes a transcription factor

(pancreatic and duodenal homeobox 1)

that is essential for development of the

pancreas and formation of insulin-

producing b islet cells. Although they

observed a substantial contribution of rat

cells to different organs and tissues, most

importantly, the pancreas of the rat-mouse

chimeras was composed exclusively of rat

cells (Figure 1B). Thus, cells derived from

rat iPS cells were able to completely

rescue the genetic deficiency of the host

mouse blastocyst. These rat-mouse

chimeras developed into adult animals

witha normal functional pancreas, demon-

strating that xenogeneic organ comple-

mentation is achievable. This is a remark-

able accomplishment.

So where do we go from here? Although

human ES and iPS cells offer hope for

tissue and cell replacement therapies in

the not too distant future, the replacement

of complex organs—lung, kidney, liver,

gut, and, of course, pancreas—is likely to

be much more difficult. Several strategies

for organ replacement are being tested.

These include the growth of organs

in vitro with mixtures of different stem cells

and biocompatible scaffolds or the gener-

ation of ‘‘humanized’’ pigs engineered to

lack certain antigens so that their organs

can be used for transplantation in human

patients with a reduced chance of immune

rejection. Could production of human

organs in, for example, human-pig

chimeras be an alternative approach?

Although production of viable rat-mouse

chimeras could be viewed as a first step

in this direction, as Kobayashi et al.

propose, there are huge biological and

technical challenges. For example, the

mouse and rat are developmentally very

similar (apart from size), but it is not clear

that chimeras between animals belonging

to different phylogenetic families or orders

would be viable. Indeed, the only attempts

to make such chimeras (between a mouse

and a bank vole) have failed (Mystkowska,

1975). In this experiment, the mouse-bank

vole chimeras were made by aggregation

of embryos; it is possible that injection of

bank vole stem cells into mouse blasto-

cysts, followed by their development

in the uterus of mouse foster mothers

might yield positive results. Successful

chimerism between members of different

orders (the pig and human, for example)

seems very unlikely, and attempts to

produce early postimplantation human-

mouse chimeras have not been encour-

aging (James et al., 2006). Even if we

succeed in developing organ-deficient

pigs by genetic manipulation and pro-

ducing pig-human chimeras in which the

parenchymal cells of the specific organ

are entirely derived from human cells,

immune rejection will still be a problem

because the human organ carried by the

pig will contain pig-derived stromal cells

and blood vessels.

Finally, there are huge legal and ethi-

cal barriers to creating human-animal

chimeras and, indeed, their production is

forbidden in most countries. However, it

is possible that injecting human ES or

iPS cells into a mouse blastocyst and al-

lowing limited (early postimplantation)

development of human-mouse chimeras

would be approved for the specific

purpose of testing the differentiation

potential of human stem cells. Yet such

experiments will be complicated, time

consuming, difficult to interpret, and, I

suspect, will never become part of the

standard protocols regulating the medical

use of human stem cells. Although xeno-

geneic organ complementation is unlikely

to be a viable strategy for regenerative

medicine, the elegant work of Kobayashi

et al. is a boon for researchers seeking

to better understand the biology of stem

cells and mammalian development.

REFERENCES

Buehr, M., Meek, S., Blair, K., Yang, J., Ure, J.,

Silva, J., McLay, R., Hall, J., Ying, Q.-L., and Smith,

A. (2008). Cell 135, 1287–1298.

Fehilly, C.B., Willadsen, S.M., and Tucker, E.M.

(1984). Nature 307, 634–636.

Gardner, R.L., and Johnson, M.H. (1973). Nat. New

Biol. 246, 86–89.

James, D., Noggle, S.A., Swigut, T., and Brivanlou,

A.H. (2006). Dev. Biol. 295, 90–102.

Kobayashi, T., Yamaguchi, T., Hamanaka, S.,

Kato-Itoh, M., Yamazaki, Y., Ibata, M., Sato, H.,

Lee, Y.-S.,, Usui, J.-i., Knisely, A.S., et al. (2010).

Cell 142, this issue, 787–799.

Li, P., Tong, C., Mehrian-Shai, R., Jia, L., Wu, N.,

Yan, Y., Maxson, R.E., Schulze, E.N., Song, H.,

Hsieh, C.-L., et al. (2008). Cell 135, 1299–1310.

Mystkowska, E.T. (1975). J. Embryol. Exp. Mor-

phol. 33, 731–744.

Nagy, A., Rossant, J., Nagy, R., Abramow-

Newerly, W., and Roder, J.C. (1993). Proc. Natl.

Acad. Sci. USA 90, 8424–8428.

Rossant, J., Mauro, V.M., and Croy, B.A. (1982).

J. Embryol. Exp. Morphol. 69, 141–149.

Tarkowski, A.K. (1961). Nature 190, 857–860.

678 Cell 142, September 3, 2010 ª2010 Elsevier Inc.

Page 47: Cell 100903

Leading Edge

Previews

‘Fore Brain:A Hint of the Ancestral CortexLora B. Sweeney1,2,3 and Liqun Luo1,2,3,*1Neurosciences Program2Department of Biology3Howard Hughes Medical Institute

Stanford University, Stanford, CA 94305, USA

*Correspondence: [email protected] 10.1016/j.cell.2010.08.024

By combining gene expression profiling with image registration, Tomer et al. (2010) find that themushroom body of the segmented worm Platynereis dumerilii shares many features with themammalian cerebral cortex. The authors propose that the mushroom body and cortex evolvedfrom the same structure in the common ancestor of vertebrates and invertebrates.

The mammalian cerebral cortex underlies

many higher-order processes, such as

perception, memory, language, and ad-

vanced motor skills. With its intricate

furrows and ridges (i.e., the sulci and

gyri), the complexity of the cerebral cortex

is evident even at its surface. Beneath the

surface, the cerebral cortex is separated

into layers of densely packed neurons

with axons reaching deep into the white

matter. These layers are divided further

into functional regions that correspond to

the body plan. How does such a complex

structure develop such precise organiza-

tion? Many researchers have approached

this question from a developmental

perspective by identifying and perturbing

molecular components that contribute

to the structure of the cerebral cortex

(Hebert and Fishell, 2008). Alternatively,

one can take an evolutionary approach

and search for an ancestral precursor to

the cortex. In this issue of Cell, Tomer

et al. (2010) follow the latter approach

and identify a brain region of the seg-

mented worm Platynereis dumerilii (an

annelid) called the mushroom body that

shares the same ‘‘molecular fingerprint’’

as the developing mammalian cortex.

The cerebral cortex derives from the

pallium. From the Latin for ‘‘cloak,’’

pallium refers to the outer layer of the

brain. The mammalian palium consists of

the cerebral cortex, olfactory cortex, and

the hippocampus. The evolutionary origin

of the vertebrate pallium has fascinated

biologists for centuries because our high-

est mental functions originate from it. To

identify structures in the developing brain

of Platynereis that are possibly related

to the vertebrate pallium, Tomer et al.

characterized the expression patterns of

many genes at different stages of neural

development in the worm. The standard

methods for characterizing gene expres-

sion are in situ hybridizations and immuno-

staining with antibodies. However, these

techniques are generally restricted to

one or two genes at a time and thus are

unable to provide direct comparisons of

expression patterns for a large number of

genes.

To overcome this technical hurdle,

Tomer et al. use advanced image registra-

tion methods (including linear transforma-

tion and nonlinear warping), in which

multiple microscopy images are aligned

to one coordinate system, standardized

to one size, and smoothed to correct arti-

facts due to stretches during sample

preparation (Rueckert et al., 1999; Rohlf-

ing et al., 2001; Kurylas et al., 2008).

Specifically, Tomer et al. use the axon

scaffold (i.e., bundles of axons in the

developing brain) as a landmark to align

individual gene expression patterns from

multiple worms onto a standard brain

template. This allows the simultaneous

mapping of expression profiles for an

unlimited number of genes during neural

development. In this manner, the authors

define the spatial relationship for the

expression of more than fifteen genes in

Platynereis that were previously shown

to regulate the patterning of the cortex in

mammals. These include many transcrip-

tion factors such as Bf1 (brain factor 1,

also known as Foxg1) and Pax 6 (paired

box gene 6) (Hebert and Fishell, 2008).

With this technique, Tomer et al. iden-

tify a structure in Platynereis with gene

patterns that mirror those observed in

the vertebrate pallium. The authors then

follow this structure during the develop-

ment of Platynereis and find that its

neurons develop into the mushroom

body, a brain region in insects and worms

involved in sensory processing and

memory (Figure 1). Moreover, this embry-

onic structure in the worm gives rise

to similar types of neurons as found

in the pallium (e.g., glutamatergic and

GABAergic neurons).

Similarities between the mammalian

cortex and the invertebrate mushroom

body have been noted previously, but

such clear parallels in their gene pattern-

ing have not been observed (Strausfeld

et al., 1998; Farris, 2008). Why not? Mush-

room bodies are well characterized in in-

sects, especially in the fruit fly Drosophila,

and choosing to study an annelid was

critical to making this new connection.

Based on the arrangement of the axon

fibers and the function of its neurons, the

mushroom body of insects has been

loosely compared to the mammalian

cerebral cortex, the hippocampus, or the

cerebellum (Strausfeld et al., 1998).

However, insects are fast-evolving inver-

tebrates and thus may deviate quickly

from their ancestors. By contrast, anne-

lids evolve slowly and therefore make

excellent organisms for comparing the

Cell 142, September 3, 2010 ª2010 Elsevier Inc. 679

Page 48: Cell 100903

evolution of invertebrates with that of

vertebrates. For example, an analysis

of intron abundance across the entire

genomes of vertebrates, annelids, and

insects found that two-thirds of the

human introns were also present in

annelids but were largely absent from

insects, confirming that insects have

evolved faster than annelids (Raible

et al., 2005). Moreover, annelids are ex-

tremely amenable to experimental anal-

ysis because they are easy to breed and

observe in vivo. Further, they develop

with highly stereotyped structures that

vary little between individuals.

Given the similarity in the gene expres-

sion patterns of the developing annelid

mushroom body and the mammalian

cerebral cortex, do these structures

share similar functions? The cerebral

cortex transforms sensory information

into action. The mushroom body of anne-

lids also processes sensory information,

receiving input from chemosensory cells

on their mouthparts. In insects, the analo-

gous structure is best known for process-

ing olfactory inputs and for memory

formation (Heisenberg, 2003). Indeed,

Tomer and colleagues found that a subset

of genes that share similar expression

patterns in the annelid mushroom body

and the mammalian cortex are also

expressed in the Drosophila mushroom

body.

Two distinct mechanisms can explain

the molecular similarities of the inverte-

brate mushroom body and the vertebrate

cerebral cortex. They could arise inde-

pendently through convergent evolution,

or they could evolve from the same

structure in a common ancestor. Tomer

and colleagues present a statistical anal-

ysis indicating that the similarities in

gene expression that they observed are

unlikely to occur by convergent evolution.

Although these arguments suffer from

the caveat that expression patterns of

different genes may not have evolved

independently, the number of genes

studied by Tomer and colleagues do

indeed favor the possibility of a common

origin. Together, their data suggest that

the vertebrate pallium and invertebrate

mushroom body evolved from the same

structure in the last common ancestor

of these organisms (Figure 1). Moreover,

this structural precursor may also have

processed sensory information, as the

cortex and mushroom body do.

That said, however, the vertebrate

pallium and invertebrate mushroom body

are considerably different. First, they have

different shapes (Figure 1). This may result

in part from considerable differences

in neural developmental programs in

vertebrates and invertebrates. Second,

it is unclear whether the mushroom body

exhibits the same type of layering and

functional specialization as the cortex.

Third, gene expression patterns appear

to remain constant over time during

development of the annelid mushroom

body, but the patterns are quite dynamic

during the development of the mamma-

lian cortex. This temporal variation com-

plicates the interpretation of the homolo-

gous gene expression patterns detected

for the two structures. Finally, the expres-

sion profiles are not identical for the

two structures. For example, expression

of Wnt3a was present in the vertebrate

pallium but not in the developing

mushroom body. Such differences are

probably not surprising, given that verte-

brates and annelids diverged from a

common precursor �600 million years

ago (Figure 1).

In summary, the data presented by

Tomer and colleagues provide a tanta-

lizing hint that one of the highest-order

processing centers of the human brain

shares an evolutionary origin with the

mushroom body of worms and insects.

In principle, the techniques presented

by the authors for profiling expression

patterns of numerous genes simulta-

neously can be used in other organisms

to garner further support for this hypoth-

esis. Furthermore, comparative struc-

ture-function analyses may also provide

additional insights into the function of

this ancestral structure, including how its

form has adapted over the last 600 million

years.

REFERENCES

Farris, S.M. (2008). Brain Behav. Evol. 72, 106–122.

Hebert, J.M., and Fishell, G. (2008). Nat. Rev.

Neurosci. 9, 678–685.

Figure 1. A Common Origin for the Mushroom Body and Cerebral Cortex?In mammals, the pallium develops into the cerebral cortex, olfactory cortex, and hippocampus. Geneexpression profiling combined with image registration in the developing brain of the segmented wormPlatynereis dumerilii identified a brain region with gene expression patterns similar to that of the mousecortex (Tomer et al., 2010). This region in the worm develops into the mushroom body, which is involvedin sensory processing and memory formation in insects and likely worms. These results suggest that themushroom body and the cerebral cortex evolved from the same structure in a common ancestor of verte-brates and annelids �600 million years ago (mya). The estimates of divergence follow Peterson et al.(2004).

680 Cell 142, September 3, 2010 ª2010 Elsevier Inc.

Page 49: Cell 100903

Heisenberg, M. (2003). Nat. Rev. Neurosci. 4,

266–275.

Kurylas, A.E., Rohlfing, T., Krofczik, S., Jenett, A.,

and Homberg, U. (2008). Cell Tissue Res. 333,

125–145.

Peterson, K.J., Lyons, J.B., Nowak, K.S., Takacs,

C.M., Wargo, M.J., and McPeek, M.A. (2004).

Proc. Natl. Acad. Sci. USA 101, 6536–6541.

Raible, F., Tessmar-Raible, K., Osoegawa, K.,

Wincker, P., Jubin, C., Balavoine, G., Ferrier, D.,

Benes, V., de Jong, P., Weissenbach, J., et al.

(2005). Science 310, 1325–1326.

Rohlfing, T., Brandt, R., Maurer, C.R., Jr., and

Menzel, R. (2001). IEEE Workshop on Mathemat-

ical Methods in Biomedical Image Analysis,

L. Staib, ed. Kauai, HI, IEEE Computer Society,

Los Alamitos, CA, pp. 187–194.

Rueckert, D., Sonoda, L.I., Hayes, C., Hill, D.L.,

Leach, M.O., and Hawkes, D.J. (1999). IEEE Trans.

Med. Imaging 18, 712–721.

Strausfeld, N.J., Hansen, L., Li, Y., Gomez, R.S.,

and Ito, K. (1998). Learn. Mem. 5, 11–37.

Tomer, R., Denes, A.S., Tessmar-Raible, K., and

Arendt, D. (2010). Cell 142, this issue, 800–809.

Cell 142, September 3, 2010 ª2010 Elsevier Inc. 681

Page 50: Cell 100903

Leading Edge

Minireview

The Language of Histone CrosstalkJung-Shin Lee,1 Edwin Smith,1 and Ali Shilatifard1,*1Stowers Institute for Medical Research, 1000 East 50th Street, Kansas City, MO 64110, USA

*Correspondence: [email protected] 10.1016/j.cell.2010.08.011

It has been suggested that a specific pattern of histone posttranslational modifications and theircrosstalk may constitute a code that determines transcriptional outcomes. However, recent studiesindicate that histone modifications have context-dependent effects, making their interplay morelike a language within the chromatin signaling pathway than a code.

For almost two decades, a primary focus in the field of transcrip-

tional regulation was to identify DNA elements that control the

expression of genes. These efforts were in part motivated by the

expectation that it would one day be possible to look at a gene

and its regulatory sequences and predict when and where

a gene was going to be transcribed. We then learned that along

with the sequence-specific binding of activators and repressors,

there is an additional world of factors that modify, interact, and

remodel chromatin to regulate gene expression. The identification

of a multitude of histone modifications—some correlated with

activation, some with repression—led to the proposal that the

modifications constitute a code that could be recognized by tran-

scription factors to determine the transcriptional state of a gene.

However, additional research has since added layers of

complexity, revealing a nuanced and intriguing language, not

a strict code, as the basis for transcriptional regulation through

the chromatin signaling pathway. Here, we review the complex

crosstalk among histone modifications, including recent studies

that illustrate how the context and timing of these modifications

are critical for a particular transcriptional readout.

Histone Crosstalk and Gene ActivityIn eukaryotic cells, gene expression can be regulated at the level

of chromatin structure. Numerous residues within the histone

tails and several residues within the histone globular domains

can be modified in a variety of ways, including acetylation,

phosphorylation, ubiquitination, and methylation. A well-charac-

terized posttranslational modification regulating chromatin

structure is the acetylation of histone N-terminal tails, which is

thought to facilitate transcriptional activation either by charge

neutralization of the tails’ interaction with DNA or by forming

a binding site for bromodomain-containing transcription factors,

some of which can remodel nucleosomes. Another well-studied

histone modification is the methylation of lysine 4 of histone H3

(H3K4), a modification generally associated with transcriptionally

active genes and a binding site for a variety of factors that include

histone-modifying and -remodeling activities (Shilatifard, 2006).

More complex scenarios arise when histone modifications act

combinatorially in a context-dependent manner to facilitate or

repress chromatin-mediated transcription. In some cases the

modification of one residue can alter the ability of a second

residue to be implemented by its modifying enzyme(s) (Figure 1).

The first example of histone crosstalk falls into this category: the

phosphorylation of serine 10 on histone H3 stimulates the ability

of Gcn5 to acetylate histone H3 at lysine 14 (H3K14) (Cheung

et al., 2000; Lo et al., 2000) (Figure 1A). Another well-character-

ized example is the requirement of histone H2B monoubiquitina-

tion for proper H3K4 methylation by the H3K4 methylase

complex COMPASS (Figure 1B) (Shilatifard, 2006). This process,

initially discovered in yeast (Shilatifard, 2006), is now known to

be highly conserved among eukaryotes (Kim et al., 2009). Addi-

tionally, a recent comprehensive mutation analysis of all histone

residues reveals that a single point mutation in histone H3K14,

a site of acetylation, results in the specific loss of H3K4 trimethy-

lation, but not mono- or dimethylation (Nakanishi et al., 2008).

This screen also demonstrated that H3K4 trimethylation is regu-

lated by both monoubiquitination-dependent and monoubiquiti-

nation-independent processes (Nakanishi et al., 2008).

Given that histone-modifying enzymes are often found in mul-

tisubunit complexes, modification of nearby residues can create

binding sites for the components of the complex helping to

anchor an enzyme to a nucleosome. For example, the PHD finger

of Yng1, a subunit of the NuA3 histone acetyltransferase

complex, recognizes methylated H3K4 and helps recruit this

histone acetyltransferase complex for acetylation of H3K14

(Martin et al., 2006; Taverna et al., 2006). The Yng1-related

ING2 can also bind methylated H3K4; however, it is present in

a histone deacetylase complex (Shi et al., 2006). Therefore,

H3K4 methylation can serve as a landing platform for a variety

of histone-modifying enzymes with opposing activities.

Modifications of nearby residues can also prevent the recog-

nition of a substrate by an enzyme, as recently reported to occur

when methylation of histone H3 arginine 2 (H3R2) interferes with

H3K4 methylation by Set1/COMPASS in yeast and COMPASS-

like complexes in mammalian cells (Guccione et al., 2007;

Kirmizis et al., 2007) (Figure 1B). Histone modifications can

also prevent the recruitment of factors other than enzymes.

For example, heterochromatin protein 1 (HP1), which binds

methylated H3K9, cannot do so when the adjacent serine 10

(H3S10) is phosphorylated during mitosis or during gene activa-

tion (Fischle et al., 2005; Mateescu et al., 2008).

Multiple types of histone crosstalk, involving numerous histone-

modifying complexes, can occur at any one gene. A major chal-

lenge is to understand the events that regulate changes in gene

expression through these modifications/crosstalk. One strategy

has been to profile histone modifications genomewide, with the

682 Cell 142, September 3, 2010 ª2010 Elsevier Inc.

Page 51: Cell 100903

expectation that a given pattern will indicate a transcriptional

outcome due to the recruitment of specific proteinsby these modi-

fications. However, some recent examples of trans-histone cross-

talk illustrate that transcriptional readout depends on context and

timing by which these modifications are introduced. Simply put,

just looking at the pattern of chromatin modifications at a locus

is not sufficient to determine its gene expression status. These

studies provide new insight into the language of histone crosstalk.

From Histone Phosphorylation to TranscriptionElongationA novel form of crosstalk was recently discovered by Zippo and

colleagues, who studied the transcriptional control of FOSL1,

a gene activated in response to serum (Zippo et al., 2009) (Fig-

ure 2A). They present evidence for a transcription activation

pathway in which the phosphorylation of H3 tails leads to the

acetylation of H4 tails. In turn, acetylation of H4 tails is required

for the recruitment of the RNA Pol II positive transcription elonga-

tion factor, P-TEFb (Figure 2A). Previously, the authors found

that activation of the FOSL1 gene requires PIM1, a proto-onco-

gene whose kinase activity is activated through MAP kinase

signaling. Numerous cellular substrates of PIM1 have been iden-

tified, including H3S10. Other H3S10 kinases, such as MSK1 and

MSK2 (MSK1/2), are also implicated in the phosphorylation of

histones at serum-responsive genes, including FOSL1. Zippo

and colleagues find that the spatiotemporal pattern of H3S10

phosphorylation differs for PIM1 and MSK1/2. MSK1/2 mediates

the phosphorylation of H3S10 at the promoter of FOSL1 at early

time points of gene expression, whereas PIM1 is required for

H3S10 phosphorylation at a FOSL1 enhancer after the MSK1/

2-mediated phosphorylation of H3S10 (Figure 2A).

Screening for other histone modifications specifically associ-

ated with the FOSL1 enhancer shows that the acetylation of

H4K16 coincides with H3S10 phosphorylation. RNA interfer-

ence-mediated knockdown of PIM1 results in loss of H4K16

acetylation, suggesting a trans-tail crosstalk from H3S10 phos-

phorylation to H4K16 acetylation. Zippo and colleagues asked

whether 14-3-3 g, 3, and z proteins, previously shown to bind

phosphorylated H3S10, are recruited to the promoter and the

enhancer of FOSL1 in response to serum. They find that 14-3-

33 and 14-3-3z are recruited to both the promoter and enhancer

of FOSL1 after serum stimulation. However, 14-3-3 is required

for recruiting the H4K16 histone acetyltransferase MOF only to

the enhancer and not to the promoter of FOSL1. Recruitment

of MOF to the enhancer results in H4K16 acetylation, which

can be bound by the bromodomain-containing protein, Brd4.

Brd4 associates with P-TEFb, a kinase that phosphorylates Pol

Figure 1. Examples of Histone Crosstalk(A) The first characterized example of histone crosstalk is the stimulation ofacetyltransferase activity of GCN5 toward the histone H3 tail by prior phos-phorylation (P) of serine 10. Acetylation, Ac.(B) Crosstalk among histone modifications can span more than one histone.Monoubiquitination of histone H2B on lysine 120 of the C-terminal helix canlead to the trimethylation of lysine 4 in the histone 3 tail (H3K4) by Set1/COMPASS. However, H3K4 methylation by COMPASS and COMPASS-likecomplexes can be blocked if the nearby arginine of H3 is already methylated.

Figure 2. Context-Dependent Outcomes of

Histone Crosstalk(A) Zippo and colleagues (Zippo et al., 2009)uncover a new form of histone crosstalk bystudying the transcriptional control of FOSL1,a gene activated in response to serum. Activationrequires the binding of PIM1 to the FOSL1enhancer. PIM1 is a kinase responsible for phos-phorylation (P) of serine 10 on the histone H3 tail(H3S10). Phosphorylated H3S10 creates a bindingsite for 14-3-3, a phosphoserine binding protein.Acetylation (Ac) of lysine 16 on the H4 (H4K16)tail occurs through interaction of 14-3-3 with thehistone acetyltransferase MOF. AcetylatedH4K16 can in turn form a binding site for the bro-modomain-containing protein Brd4, a componentof P-TEFb, a kinase that phosphorylates theC-terminal domain of RNA Pol II to facilitate tran-scription elongation. However, at an earlier stageof serum stimulation, an MSK1/2 kinase is re-cruited to the promoter where it phosphorylatesH3S10. 14-3-3 is then recruited to the promoter,but unlike the situation at the enhancer, MOF isnot recruited to the promoter. Thus, the timing,

location, and perhaps identity of the H3 kinase, and not the H3S10 modification alone, determines downstream events.(B) Another example of how the order of implementation of histone modifications can affect transcription comes from work from Wang et al. (2009). They reportthat despite correlations between histone acetylation and H3K4 methylation, artificially increasing acetylation through treatment of cells with deacetylase inhib-itors (HDACs) does not lead to productive transcription, despite the presence of H3K4 methylation and Pol II recruitment. Therefore, patterns of histone modi-fications cannot simply be ‘‘read’’ but instead have distinct effects depending on the cellular context and upstream signaling events.

Cell 142, September 3, 2010 ª2010 Elsevier Inc. 683

Page 52: Cell 100903

II to facilitate transcription elongation (Figure 2A). Thus, Zippo

and colleagues propose that crosstalk between modifications

on two different histone tails regulate productive transcription

elongation through the sequential recruitment of proteins that

bind these modifications.

One question raised by this study is why H3S10 phosphoryla-

tion produces different results at the enhancer than at the

promoter of FOSL1 even though 14-3-3 is recruited to both sites.

At the enhancer, 14-3-3 recruits the histone acetyltransferase

MOF. At the promoter, it does not. What is the difference

between 14-3-3 at the promoter and at the enhancer? Interest-

ingly, 14-3-33 and 14-3-3z are thought to be regulated via lysine

acetylation (Choudhary et al., 2009) and an acetyltransferase,

Tip60, is preferentially recruited to the promoter of FOSL1. One

possibility is that Tip60 acetylates 14-3-3 and prevents its inter-

action with MOF.

Another intriguing aspect of the study by Zippo and colleagues

is the link between H3S10 phosphorylation and H4K16 acetyla-

tion. These two modifications were previously linked in studies of

dosage compensation in the fruit fly Drosophila. In Drosophila

dosage compensation, MOF is recruited to the coding region

of X-linked genes in males where it mediates H4K16 acetylation

in a process thought to facilitate transcription elongation. Coloc-

alizing with MOF on the male X chromosome is the JIL-1 kinase,

an MSK1/2-related kinase that mediates the phosphorylation of

H3S10 on this chromosome. In the case of Drosophila dosage

compensation, recruitment of the JIL-1 kinase to the male X

chromosome occurs later than H4K16 acetylation (Wang et al.,

2001), a reversal of the order of the addition of these marks at

FOSL1 in response to serum. Concordantly, the MOF complex

that mediates acetylation in coding regions is likely to be distinct

from the MOF complex that mediates promoter/enhancer acet-

ylation of genes (Li et al., 2009). Thus, by all appearances, these

two examples of the coexistence of both H3S10 phosphorylation

and H4K16 acetylation are unrelated in their order of implemen-

tation and in their biological meaning. This suggests that

descriptions of histone modification patterns, without under-

standing the mechanisms leading to the implementation of these

marks, should be interpreted with caution. Importantly, the study

by Zippo and colleagues begins to determine the role of histone

modifications in the activation of FOSL1, with a spatial and

temporal dissection of how a cascade of histone modifications

can lead to a particular transcriptional outcome.

Trimethylation Converses with AcetylationAnother example of trans-tail crosstalk was proposed by Wang

et al. (2009). In this case the communication takes place between

the H3 and H4 tails and, like the example provided by Zippo et al.

(2009), involves H4K16 acetylation and FOSL1 transcription. By

analyzing genomewide profiles of several histone acetyltrans-

ferases, deacetylases, and modifications, these investigators

find a link between H3K4 methylation and H4K16 acetylation at

some inducible genes, including FOSL1. The authors show

that a subset of transcriptionally quiescent genes, marked by

the presence of H3K4 methylation, display a marked increase

in histone acetylation at H3K9 and H4K16 after addition of de-

acetylase inhibitors. In contrast, quiescent genes not marked

with H3K4 methylation rarely show this increase in acetylation

in response to deacetylase inhibitors.

In order to determine whether H3K4 methylation is functionally

linked to H4K16 acetylation, Wang and colleagues use RNA

interference-mediated knockdown of WDR5, a common compo-

nent of the Set1 and MLL (mixed-lineage leukemia) COMPASS-

like H3K4 methyltransferase complexes. Upon knockdown of

WDR5, they observe reduced levels of histone acetylation at

the subset of transcriptionally quiescent genes marked by

H3K4 methylation. Based on this information, Wang and

colleagues suggest that H3K4 methylation primes certain genes

for an increase in H3K9 and H4K16 acetylation. Interestingly,

what was not considered by Wang and colleagues is the fact

that WDR5 is also a component of complexes that can acetylate

H4K16 or H3K9. WDR5 associates with the H4K16 acetyltrans-

ferase MOF in the NSL/MSL1v complex (Cai et al., 2010; Li

et al., 2009) as well as with the H3K9/K14 acetyltransferase

Gcn5 in the ATAC complex (Suganuma et al., 2008; Wang

et al., 2008). As such, the effect of WDR5 knockdown could be

a consequence of WDR5’s role as a subunit of the H3K4 methyl-

ases, or WDR5’s role as a subunit of the H3 and H4 acetyltrans-

ferase complexes, or a combination of the two. Given that WDR5

is part of the H3 and H4 acetyltransferase complexes, the exis-

tence of crosstalk between H3K4 and H4K16 needs to be further

characterized.

One surprising finding of the study by Wang and colleagues is

that transcription is rarely induced at the genes tested, although

Pol II is recruited following treatment with a deacetylase inhibitor

(Figure 2B). Thus, Pol II recruitment does not lead to the antici-

pated increase in transcription, despite the fact that H3K9 and

H4K16 acetylation are increased. These modifications coincide

with transcriptional activation at FOSL1 upon serum treatment

(Zippo et al., 2009). Together with the studies of H3S10 phos-

phorylation and H4K16 acetylation at the FOSL1 enhancer, it is

clear that knowing the mechanism and timing of these modifica-

tions is necessary for determining the transcriptional outcome.

The Emerging Grammar of Histone CrosstalkThe existence of a histone modification code was proposed

10 years ago as a way to approach the study of the quickly

growing number of histone modifications involved in the regula-

tion of gene expression and other DNA-templated processes,

such as replication, repair, and recombination. New ‘‘words’’ of

histone modifications are being discovered, and they continue

to appear in interesting combinations. However, discovering

the exact roles these modifications play in gene expression

has been complicated by finding a counterexample for almost

every example of crosstalk, such as the case of H3K4 and

H3K36 methylation recruiting both histone acetyltransferases

and deacetylases.

A common theme of recent research on histone crosstalk is

that the order and mechanism of the addition and removal of

histone modifications are important for the transcriptional

readout of a gene. The recent examples of histone crosstalk

that we have addressed here illustrate this point. In one study,

the implementation of H3S10 phosphorylation at two different

locations, by two different enzymes, and at two different times

after serum stimulation had disparate effects on subsequent

684 Cell 142, September 3, 2010 ª2010 Elsevier Inc.

Page 53: Cell 100903

histone acetylation at the respective locations (Zippo et al., 2009)

(Figure 2A). Zippo and colleagues were able to propose a mech-

anism of gene activation by identifying the histone-modifying

enzymes, the histone modifications, and a set of proteins that

recognized these modifications on the FOSL1 gene after serum

stimulation. In another study, Wang and colleagues found

that artificially recreating histone modifications that correlate

with gene expression could result in the recruitment of RNA

Pol II, but this was not sufficient for transcription (Figure 2B).

Thus, simply mapping histone modification patterns without

understanding the recruitment, regulation, and interactions of

the complexes implementing these marks is not sufficient

to understand the mechanisms regulating gene expression.

Genomewide profiling techniques have now become widely

adopted, providing the ability to map histone modifications,

the enzymes implementing these marks, and the factors that

recruit them under different experimental conditions.

The study of the regulation of gene expression has grown from

identifying transcription factors and their binding sites to include

a wide variety of other binding events associated with the modi-

fications of histones that package the DNA. Future progress will

require us to learn much more about how the words comprising

the dictionary of histone crosstalk are used in a particular order

to provide the grammar of this complex language.

ACKNOWLEDGMENTS

We thank Dr. E. Park for the critical reading of the manuscript. We also thank

L. Shilatifard for editorial assistance. Studies in the Shilatifard laboratory are

supported in part by grants from the NIH: R01CA89455, R01GM069905, and

R01CA150265.

REFERENCES

Cai, Y., Jin, J., Swanson, S.K., Cole, M.D., Choi, S.H., Florens, L., Washburn,

M.P., Conaway, J.W., and Conaway, R.C. (2010). J. Biol. Chem. 285, 4268–

4272.

Cheung, P., Tanner, K.G., Cheung, W.L., Sassone-Corsi, P., Denu, J.M., and

Allis, C.D. (2000). Mol. Cell 5, 905–915.

Choudhary, C., Kumar, C., Gnad, F., Nielsen, M.L., Rehman, M., Walther, T.C.,

Olsen, J.V., and Mann, M. (2009). Science 325, 834–840.

Fischle, W., Tseng, B.S., Dormann, H.L., Ueberheide, B.M., Garcia, B.A., Sha-

banowitz, J., Hunt, D.F., Funabiki, H., and Allis, C.D. (2005). Nature 438, 1116–

1122.

Guccione, E., Bassi, C., Casadio, F., Martinato, F., Cesaroni, M., Schuchlautz,

H., Luscher, B., and Amati, B. (2007). Nature 449, 933–937.

Kim, J., Guermah, M., McGinty, R.K., Lee, J.S., Tang, Z., Milne, T.A., Shilati-

fard, A., Muir, T.W., and Roeder, R.G. (2009). Cell 137, 459–471.

Kirmizis, A., Santos-Rosa, H., Penkett, C.J., Singer, M.A., Vermeulen, M.,

Mann, M., Bahler, J., Green, R.D., and Kouzarides, T. (2007). Nature 449,

928–932.

Li, X., Wu, L., Corsa, C.A., Kunkel, S., and Dou, Y. (2009). Mol. Cell 36,

290–301.

Lo, W.S., Trievel, R.C., Rojas, J.R., Duggan, L., Hsu, J.Y., Allis, C.D., Marmor-

stein, R., and Berger, S.L. (2000). Mol. Cell 5, 917–926.

Martin, D.G., Baetz, K., Shi, X., Walter, K.L., MacDonald, V.E., Wlodarski, M.J.,

Gozani, O., Hieter, P., and Howe, L. (2006). Mol. Cell. Biol. 26, 7871–7879.

Mateescu, B., Bourachot, B., Rachez, C., Ogryzko, V., and Muchardt, C.

(2008). EMBO Rep. 9, 267–272.

Nakanishi, S., Sanderson, B.W., Delventhal, K.M., Bradford, W.D., Staehling-

Hampton, K., and Shilatifard, A. (2008). Nat. Struct. Mol. Biol. 15, 881–888.

Shi, X., Hong, T., Walter, K.L., Ewalt, M., Michishita, E., Hung, T., Carney, D.,

Pena, P., Lan, F., Kaadige, M.R., et al. (2006). Nature 442, 96–99.

Shilatifard, A. (2006). Annu. Rev. Biochem. 75, 243–269.

Suganuma, T., Gutierrez, J.L., Li, B., Florens, L., Swanson, S.K., Washburn,

M.P., Abmayr, S.M., and Workman, J.L. (2008). Nat. Struct. Mol. Biol. 15,

364–372.

Taverna, S.D., Ilin, S., Rogers, R.S., Tanny, J.C., Lavender, H., Li, H., Baker, L.,

Boyle, J., Blair, L.P., Chait, B.T., et al. (2006). Mol. Cell 24, 785–796.

Wang, Y., Zhang, W., Jin, Y., Johansen, J., and Johansen, K.M. (2001). Cell

105, 433–443.

Wang, Y.L., Faiola, F., Xu, M., Pan, S., and Martinez, E. (2008). J. Biol. Chem.

283, 33808–33815.

Wang, Z., Zang, C., Cui, K., Schones, D.E., Barski, A., Peng, W., and Zhao, K.

(2009). Cell 138, 1019–1031.

Zippo, A., Serafini, R., Rocchigiani, M., Pennacchini, S., Krepelova, A., and

Oliviero, S. (2009). Cell 138, 1122–1136.

Cell 142, September 3, 2010 ª2010 Elsevier Inc. 685

Page 54: Cell 100903
Page 55: Cell 100903

GPR120 Is an Omega-3 Fatty Acid ReceptorMediating Potent Anti-inflammatoryand Insulin-Sensitizing EffectsDa Young Oh,1,4 Saswata Talukdar,1,4 Eun Ju Bae,1 Takeshi Imamura,2 Hidetaka Morinaga,1 WuQiang Fan,1 Pingping Li,1

Wendell J. Lu,1 Steven M. Watkins,3 and Jerrold M. Olefsky1,*1Department of Medicine, Division of Endocrinology and Metabolism, University of California, San Diego, La Jolla, CA 92093, USA2Division of Pharmacology, Shiga University of Medical Science, Tsukinowa, Seta, Otsu-city, Shiga, 520-2192 Japan3Tethys Bioscience, 3410 Industrial Boulevard, Suite 103, West Sacramento, CA 95691, USA4These authors contributed equally to this work

*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.07.041

SUMMARY

Omega-3 fatty acids (u-3 FAs), DHA and EPA, exertanti-inflammatory effects, but the mechanisms arepoorly understood. Here, we show that the Gprotein-coupled receptor 120 (GPR120) functions asan u-3 FA receptor/sensor. Stimulation of GPR120with u-3 FAs or a chemical agonist causes broadanti-inflammatory effects in monocytic RAW 264.7cells and in primary intraperitoneal macrophages.All of these effects are abrogated by GPR120 knock-down. Since chronic macrophage-mediated tissueinflammation is a key mechanism for insulin resis-tance in obesity, we fed obese WT and GPR120knockout mice a high-fat diet with or without u-3 FAsupplementation. The u-3 FA treatment inhibitedinflammation and enhanced systemic insulin sensi-tivity in WT mice, but was without effect in GPR120knockout mice. In conclusion, GPR120 is a functionalu-3 FA receptor/sensor and mediates potent insulinsensitizing and antidiabetic effects in vivo by repres-sing macrophage-induced tissue inflammation.

INTRODUCTION

Chronic activation of inflammatory pathways plays an important

role in the pathogenesis of insulin resistance and the macro-

phage/adipocyte nexus provides a key mechanism underlying

the common disease states of decreased insulin sensitivity

(Schenk et al., 2008). This involves migration of monocytes/

macrophages to adipose tissue (including intramuscular fat

depots) and liver with subsequent activation of macrophage

proinflammatory pathways and cytokine secretion. Through

paracrine effects, these events promote inflammation and

decreased insulin sensitivity in nearby insulin target cells

(Shoelson et al., 2007; Schenk et al., 2008). In these studies,

we explore the interlocking biology between proinflammatory

and anti-inflammatory molecules within the specialized popula-

tion of proinflammatory tissue macrophages.

G protein-coupled receptors (GPCRs) are important signaling

molecules for many aspects of cellular function. They are

members of a large family that share common structural motifs

such as seven transmembrane helices and the ability to acti-

vate heterotrimeric G proteins, such as Gas, Gai, and Gaq.

Ligands bind specifically to GPCRs to stimulate and induce

a variety of cellular responses via several second messenger

pathways; e.g., modulation of cAMP production, the phospho-

lipase C pathway, ion channels, and mitogen-activated protein

kinases (Ulloa-Aguirre et al., 1999; Gether, 2000; Schulte and

Fredholm, 2003. Recently, several groups reported that five

orphan receptors, GPR40, GPR41, GPR43, GPR84, and

GPR120, can be activated by free fatty acids (FFAs). Short-

chain fatty acids (FAs) are specific agonists for GPR41 and

GPR43 (Tazoe et al., 2008) and medium-chain FAs for GPR84

(Wang et al., 2006a). Long-chain FAs can activate GPR40

(Itoh et al., 2003) and GPR120 (Hirasawa et al., 2005). Hirasawa

et al. showed that the stimulation of GPR120 by FFAs resulted

in elevation of [Ca2+]i and activation of the ERK cascade which

suggests interactions with the Gaq family of G proteins.

However, other physiological functions of GPR120 remain to

be explored. In the current study, we found that GPR120 was

highly expressed in adipose tissue, and proinflammatory

macrophages. The high expression level of GPR120 in mature

adipocytes and macrophages indicates that GPR120 might

play an important role in these cell types. We stimulated

GPR120 with a synthetic agonist (GW9508) and omega-3 fatty

acids (u-3 FAs) and examined whether activation of GPR120

affected LPS- and TNF-a-induced inflammatory signaling

responses. While SFAs are proinflammatory and unsaturated

FAs are generally neutral, we found that u-3 FAs (docosahex-

aenoic acid (C22:6n3, DHA) and eicosapentaenoic acid

(C20:5n3, EPA)), the major ingredients in fish oil, exert potent

anti-inflammatory effects through GPR120.

b-arrestins can serve as scaffold or adaptor proteins for a wide

range of GPCRs, as well as a selected group of other receptor

subtypes (Miller and Lefkowitz, 2001). After ligand binding,

b-arrestins can associate with the cytoplasmic domains of

Cell 142, 687–698, September 3, 2010 ª2010 Elsevier Inc. 687

Page 56: Cell 100903

GPCRs and couple the receptor to specific downstream

signaling pathways, as well as mediate receptor endocytosis

(Luttrell and Lefkowitz, 2002). Here we find that b-arrestin2 asso-

ciates with ligand-stimulated GPR120 and participates in down-

stream signaling mechanisms.

Since chronic inflammation is a mechanistic feature of

obesity-related insulin resistance, we postulated that the anti-

inflammatory effect of GPR120 stimulation could promote insulin

sensitization. In the present study, we elucidate the role of

GPR120 activation in integrating anti-inflammatory and insulin

sensitizing effects in vitro and in vivo.

RESULTS

GPR120 ExpressionFatty acids (FAs) can function as endogenous ligands modu-

lating inflammatory responses, but not all FAs work in the

same way. In general, saturated FAs (SFAs) are proinflammatory,

unsaturated FAs are weakly proinflammatory or neutral, and u3-

FAs can be anti-inflammatory (Lee et al., 2003; Calder, 2005;

Solinas et al., 2007). Because of the importance of inflammation

in a number of chronic human diseases including insulin resis-

tance, obesity, and type 2 diabetes mellitus, we surveyed the

family of FA sensing GPCRs (GPR40, 41, 43, 84, and 120). Based

on its tissue expression pattern, GPR120 emerged as a receptor

of particular interest. As seen in Figure 1, GPR120 is the only lipid

sensing GPCR which is highly expressed in adipose tissue,

proinflammatory CD11c+ macrophages (BMDCs), mature adipo-

cytes, and monocytic RAW 264.7 cells (Figure 1A and 1B).

GPR120 is induced in the stromal vascular fraction (SVF) of

adipose tissue (which contains the macrophages), as well as in

hepatic Kupffer cells, during high-fat diet (HFD) feeding in mice

(Figure 1C). GPR120 is also expressed in enteroendocrine L cells

with negligible expression in muscle (Figure S4C available

online), hepatocytes or other cell types (Hirasawa et al., 2005;

Gotoh et al., 2007).

Ligand-Stimulated GPR120 Exerts Anti-inflammatoryEffectsIt has been previously reported that GPR120 signals via

a Gaq/11-coupled pathway and can respond to long chain FAs

(Hirasawa et al., 2005). To pursue the biology of GPR120,

a tool compound was needed, and, some years ago, Glaxo pub-

lished GW9508 as a GPR40 selective agonist. However, this

compound was not specific and also stimulated GPR120 (Bris-

coe et al., 2006). Since macrophages and adipocytes do not

express GPR40 (this was confirmed by repeated q-PCR and

RT-PCR measures, Figure S1A), GW9508 is a functional

GPR120 specific compound in these cell types. Using this

approach, we found that GW9508 treatment broadly and mark-

edly repressed the ability of the TLR4 ligand LPS to stimulate

inflammatory responses in RAW 264.7 cells (Figure 1D and E).

Figure 1. Expression Level of GPR120 and

GPR120-Mediated Anti-inflammatory Res-

ponse in RAW 264.7 Cells

(A and B) (A) The mRNA expression pattern of

various lipid sensing GPCRs is shown in adipose

tissue, (B) CD11c+ bone marrow-derived dendritic

cells (BMDCs), bone marrow-derived macro-

phages (BMDMs), IPMacs, 3T3-L1 preadipocytes,

differentiated 3T3-L1 adipocytes, RAW 264.7

cells, and L6 myocytes. Ribosomal protein S3

(RPS3) was used as internal control.

(C) Expression of GPR120 in SVF, adipocytes and

hepatic Kupffer cells from chow (NC)- or HFD-fed

mice was examined by q-PCR. Data are ex-

pressed as the mean ± SEM of at least three

independent experiments in triplicate. *p < 0.05

versus NC.

(D) RAW 264.7 cells, transfected with scrambled

(Scr) or GPR120 #2 siRNA (GPR120 KD), were

treated with 100 mM of GW9508 for 1 hr prior to

LPS (100 ng/ml) treatment for 10 min and then

subjected to western blotting. Left panel is a repre-

sentative image from three independent experi-

ments, and the scanned bar graph (right panel)

shows fold induction over basal conditions.

Knockdown efficiency of GPR120 siRNA is shown

in Figure S1.

(E) Cytokine secretion level was measured in RAW

264.7 cells by ELISA. Data are expressed as the

mean ± SEM of three independent experiments.

*p < 0.05 versus LPS treatment in scrambled

siRNA transfected cells. See also Figures S1

and S2.

688 Cell 142, 687–698, September 3, 2010 ª2010 Elsevier Inc.

Page 57: Cell 100903

Thus, GW9508 inhibited LPS stimulated phosphorylation of IKKb

and JNK, prevented IkB degradation, and inhibited TNF-a and

IL-6 secretion. All of these effects of GW9508 were completely

abrogated by siRNA mediated knockdown of GPR120 (Figures

1D and 1E and Figure S1F).

Based on these remarkable anti-inflammatory effects of

GPR120 stimulation, we established a cell based reporter

system by transfecting HEK293 cells with constructs for

GPR120 along with a serum response element-luciferase

promoter/reporter (SRE-luc). Since GPR120 is a Gaq/11-

coupled receptor, it stimulates both PKC and MAP kinase, and

both of these biologic effects are detected by the SRE-driven

reporter system (Oh et al., 2005). The reporter cells were treated

with various FAs and the synthetic GW9508 ligand. We found

that GW9508, the u-3 FAs (DHA and EPA) and palmitoleate

(C16:1n7), all activated the SRE-luc reporter with an EC50 of

1-10 mM (Figure 2A), while SFAs were without effect. GW9508

and DHA were used at 100 mM in all subsequent studies to

achieve maximal action. The u-3 FAs (DHA and a-linolenic

acid), and SFA (palmitic acid (C16:0)) activated ERK phosphory-

Figure 2. Omega-3 FA Stimulates GPR120

and Mediates Anti-inflammatory Effects

(A–D) GPR120-mediated SRE-luc activity after

treatment with various FAs. Results are fold

activities over basal. Each data point represents

mean ± SEM of three independent experiments

performed in triplicate. Black lines indicate SRE-

luc activities without GPR120 transfection or with

non-stimulating FAs. DHA inhibits LPS-induced

inflammatory signaling (B), cytokine secretion (C),

and inflammatory gene mRNA expression level

(D) in RAW 264.7 cells, but not in GPR120 knock-

down cells.

(E and F) GPR120 stimulation inhibits LPS-induced

inflammatory response in WT primary macro-

phage. Data are expressed as the mean ± SEM

of three independent experiments. *p < 0.05

versus LPS treatment in scrambled siRNA trans-

fected cells or WT IPMacs. See also Figure S2.

lation in RAW 264.7 cells, but only DHA-

and a-linolenic acid-mediated ERK phos-

phorylation were abolished by GPR120

knockdown (Figure S2A). These results

indicate that u-3 FAs, but not SFAs,

specifically activated ERK via GPR120.

The activation of GPR120 by u-3 FAs,

as well as its expression in adipocytes

and macrophages, led us to study

whether DHA, a representative u-3 FA,

can modulate inflammation through

GPR120 in these cells. To examine this,

we pretreated RAW 264.7 cells and

3T3-L1 adipocytes with GW9508 or DHA

for 1 hr, followed by LPS (TLR4), TNF-a,

TLR2, or TLR3 stimulation, respectively.

We found that GW9508 and, more impor-

tantly, DHA, strongly inhibited LPS-

induced phosphorylation of JNK and IKKb, IkB degradation,

cytokine secretion and inflammatory gene expression level in

RAW 264.7 cells (Figures 2B–2D) as well as TNF-a, TLR2 and

TLR3-induced JNK and IKKb phosphorylation in 3T3-L1 adipo-

cytes (Figure S2B) or RAW 264.7 cells (Figure S2C). All of the

effects of GW9508 and DHA were completely prevented by

GPR120 knockdown, demonstrating that these anti-inflamma-

tory effects were specifically exerted through GPR120 (Figure 1,

Figure 2, Figure S1, and Figure S2). Similar results were seen in

primary wild-type (WT) intraperitoneal macrophages (IPMacs)

and GPR120 knockout (KO) IPMacs (Figures 2E and 2F). These

data argue that GPR120 is an u-3 FA receptor or sensor, and

provide a molecular mechanism for the anti-inflammatory effects

of this class of FAs.

Role of b-arrestin2 in GPR120 SignalingGiven these potent cell selective anti-inflammatory effects,

it was of interest to understand the specific mechanisms

whereby signals from GPR120 inhibit inflammatory pathways.

To further assess this, we used RNA interference to examine

Cell 142, 687–698, September 3, 2010 ª2010 Elsevier Inc. 689

Page 58: Cell 100903

molecules involved in generation of GPR120 signals. As seen in

Figures 3A and 3B, LPS signaling was not affected by b-ar-

restin1, -2, or Gaq/11 knockdown. However, with b-arrestin2

knockdown, DHA-mediated anti-inflammatory signaling was ex-

tinguished, while b-arrestin1 and Gaq/11 knockdown were

without effect (Figure 3A).

Figure 3A and Figure S2 show that GPR120 stimulation inhibits

both TLR4- and TNF-a mediated inflammatory responses. Since

the TNF-a and TLR signaling cascades converge downstream of

GPR120 activation, these results indicate that the site of

GPR120-induced inhibition is either at, or upstream, of JNK/

IKKb. LPS activates inflammation through the TLR4 pathway

by engaging the serine kinase IRAK, leading to phosphorylation

of transforming growth factor-b activated kinase 1 (TAK1) which

is upstream of MKK4/JNK and IKKb (Kawai and Akira, 2006,

Figure 3I). TNF-a and TLR2/3 also leads to stimulation of

TAK1, resulting in activation of IKKb and JNK (Takaesu et al.,

2003). Consequently, we determined whether DHA stimulation

of GPR120 inhibited TAK1 and MKK4. As seen in Figure 3C,

DHA treatment abrogated LPS-induced TAK1 and MKK4 phos-

Figure 3. GPR120 Internalization with

b-arrestin2 Mediates Anti-inflammatory

Effects

(A) RAW 264.7 cells were transfected with siRNA as

indicated and stimulated with or without 100 mM of

DHA 1 hr prior to LPS (100 ng/ml) treatment for

10 min and then subjected to western blotting.

(B) TNF-a secretion was measured in RAW 264.7

cell cultured media with or without RNA interfer-

ence as indicated.

(C) Phosphorylation of TAK1 and MKK4 in RAW

264.7 cells with or without siRNA transfection as

indicated.

(D) HEK293 cells were cotransfected with HA-

GPR120 and b-arrestin2$GFP to analyze

GPR120 internalization after DHA stimulation for

the indicated times. GPR120 (red) and b-arrestin2

(green) were localized by confocal microscopy.

(E–H) (E) Coimmunoprecipitation between GPR120

and b-arrestin2 with DHA stimulation for 30 min

in RAW 264.7 cells and, (F) HEK293 cells (HA-

GPR120 andb-arrestin2$GFP), respectively. Lysate

indicates 1/10 input in each experiment. Interaction

between TAB1 and b-arrestin2 (G) and interaction

between TAB1 and TAK1 (H) were detected by

coimmunoprecipitation and the scanned bar graph

quantitates the association in RAW 264.7 cells.

(I) Schematic diagram of the b-arrestin2 and

GPR120-mediated anti-inflammatory mechanism.

Red colored letters and arrows indicate the DHA-

mediated anti-inflammatory effect, and black

colored letters and arrows indicate the LPS- and

TNF-a-induced inflammatory pathway. See also

Figure S2.

phorylation in a GPR120 and b-arrestin2-

dependent manner. Since TLR2/3/4 and

TNF-a signaling were inhibited by

GPR120 activation, these results indicate

that DHA signaling intersects at TAK1 and

inhibits all upstream input activating signals via a GPR120/b-

arrestin2 interaction (Figure 3I).

After ligand stimulation, b-arrestin2 can translocate to

a number of GPCRs where it mediates receptor internalization

and signaling (Barak et al., 1997). We transfected HEK293 cells

with b-arrestin2$GFP to visualize intracellular trafficking of b-ar-

restin2 following activation of GPR120 (Figure 3D). In the basal

state, GPR120 was localized to the plasma membrane as

assessed by immunostaining (red fluorescence, Figure 3D),

while b-arrestin2 exhibited a diffuse, largely cytoplasmic stain-

ing pattern (green, Figure 3D). Following DHA treatment for

5 min, b-arrestin2$GFP translocated from the cytosol to the

plasma membrane and can be seen colocalized with GPR120

(merged, right fields). After 30 min of DHA treatment, much of

the GPR120 is internalized, as visualized by punctate intracel-

lular staining (lower left panel), and b-arrestin2$GFP is now

colocalized with the intracellular GPR120 (lower right panel,

Figure 3D). DHA-stimulated binding of b-arrestin2 to activated

GPR120 was also detected by coimmunoprecipitation

(Figure 3E and F).

690 Cell 142, 687–698, September 3, 2010 ª2010 Elsevier Inc.

Page 59: Cell 100903

LPS or TNF-a signaling activate TAK1 by causing the associ-

ation of TAK1 binding protein 1 (TAB1) with TAK1. Figure 3H

shows that LPS stimulation of RAW 264.7 cells causes TAB1/

TAK1 association. DHA treatment leads to the association of

b-arrestin2 with TAB1 (Figure 3G) and largely blocks TAK1/

TAB1 association (Figure 3H). To further examine the interaction

site of b-arrestin2 and GPR120 or TAB1, we pursued coimmuno-

precipitation with a series of b-arrestin2 truncation/deletion

mutants (Figure S2D). Full-length b-arrestin2 was able to bind

to GPR120 and TAB1 only in the presence of DHA, clearly

showing the DHA dependency of this interaction. Interestingly,

only the full-length b-arrestin2 coprecipitated with GPR120 and

TAB1, while a series of deletion/truncation b-arrestin2 mutants

did not, indicating that the interactions are dependent on the

complete tertiary structure of b-arrestin2 (Figure S2D; Luttrell

et al., 1999). Taken together, these results suggest that

GPR120 activation leads to association of b-arrestin2 with the

receptor and that this complex subsequently internalizes, where-

upon b-arrestin2 can bind to TAB1. The data further suggest that

association of b-arrestin2 with TAB1 blocks TAB1/TAK1 binding,

resulting in inhibition of TAK1 phosphorylation and activation

(Figure 3I).

Figure 4. GPR120 Activation Enhances

GLUT4 Translocation and Glucose Uptake

(A) 3T3-L1 adipocytes were transfected with

a dually tagged HA-GLUT4-GFP construct. Total

GLUT4 expression was determined by GFP fluo-

rescence, and GLUT4 translocation to the cell

surface after 100 ng/ml insulin or 100 mM DHA

stimulation for 30 min was determined by indirect

immunofluorescence of the HA-conjugated with

Alexa 594 in fixed cells. Translocation following

insulin stimulation was expressed as a percentage

of the maximum response. The bar graph

represents the mean ± SEM data from four inde-

pendent experiments. *p < 0.05 versus vehicle

treatment. (B) Glucose uptake was measured in

WT and GPR120 KO mouse primary adipose

tissue and in (C–H) 3T3-L1 adipocytes ± siRNA

with the indicated treatment. Data are expressed

as mean ± SEM of three independent experiments

in triplicate. *p < 0.05 versus basal activity. The

indicated siRNA knockdown efficiency was vali-

dated by western blotting. See also Figure S3.

GPR120 Activation EnhancesGlucose Uptake in 3T3-L1AdipocytesSince our data show that GPR120 is

expressed in mature adipocytes and sig-

nals through Gaq/11 in these cells, we

assessed the effects of GPR120 stimula-

tion on insulin sensitivity in primary adi-

pose tissue cultures and in 3T3-L1 adipo-

cytes. Primary adipose tissue explants

and 3T3-L1 adipocytes were pretreated

for 30 min with GW9508 or DHA, followed

by measurement of basal and insulin

stimulated GLUT4 translocation (Fig-

ure 4A) and 2-deoxyglucose (2-DOG) transport (Figures 4B–

4H). Ligand-stimulation of GPR120 led to an increase in glucose

transport and translocation of GLUT4 to the plasma membrane

in adipocytes, but was without effect in muscle cells (Figure S4D)

which don’t express GPR120 (Figure 1B and Figure S4C). This

stimulatory effect of DHA and GW9508 was blocked when

GPR120 or Gaq/11 was depleted by siRNA knockdown

(Figure 4D and Figure 4F). GLUT4 knockdown also blocked the

effects of DHA and GW9508, while the effects of insulin were

decreased by �90% (Figure 4E). This, along with the GLUT4

translocation data provided in Figure 4A, indicates that the

stimulatory effects of GPR120 are indeed working through

GLUT4. Further assessment of this pathway showed that DHA

had a modest effect to stimulate phosphorylation of Akt, but

that this was abrogated with GPR120 knockdown (Figure S3A).

The effects of DHA to stimulate Akt were blocked by inhibiting

PI3 kinase with LY294002 (Figure S3B). Finally, DHA did not

stimulate IRS-1 phosphorylation (Figure S3C), indicating that

its glucose transport stimulatory effects were downstream of

IRS-1. Knockdown of Gaq/11 also completely blocked the

effects of DHA to stimulate glucose transport (Figure 4F), while

b-arrestin1 or -2 knockdown was without effect (Figures 4G

Cell 142, 687–698, September 3, 2010 ª2010 Elsevier Inc. 691

Page 60: Cell 100903

and 4H). Interestingly, Gaq/11 knockdown not only inhibited

DHA and GW9508 stimulated glucose transport, but it also atten-

uated insulin stimulatory effects, and the latter is fully consistent

with previous publication (Imamura et al., 1999) showing the role

of Gaq/11 in insulin signaling to glucose transport in adipocytes.

This scheme is shown in Figure S3D.

In Vivo Metabolic Studies in GPR120 KO MiceSince chronic tissue inflammation can cause insulin resistance,

we hypothesized that deletion of GPR120 would enhance the

proinflammatory tone, promoting glucose intolerance and

decreased insulin sensitivity. To test this idea, GPR120

KO mice and WT littermates were evaluated on normal chow

diet (NC). Body weights were similar in both groups, and as

summarized in Figure 5, glucose tolerance tests (GTT) showed

a mild degree of impairment in GPR120 KO animals compared

to WTs (Figure 5A). More impressively, insulin secretion was

more than 2-fold greater in the KO animals, and the combination

of hyperinsulinemia and mild glucose intolerance indicates the

presence of insulin resistance (Figures 5B and 5C). This was

confirmed by performing hyperinsulinemic/euglycemic clamp

studies in the chow fed WT and KO mice (Figure 5D). These

studies revealed a 31% decrease in the glucose infusion rate

Figure 5. In Vivo Metabolic Studies in

GPR120 KO Mice

(A) GTT in NC-fed WT and GPR120 KO mice. n = 7

per group.

(B and C) Insulin concentration was measured at

the indicated time points and (C) area-under-curve

analysis of the insulindata shows a significantdiffer-

ence between WT and GPR120 KO mice on NC.

(D) Hyperinsulinemic/euglycemic clamp studies in

chow-fed WT and GPR120 KO mice.

(E) Clamp studies in HFD, u-3 FA supplemented

(+u3), and Rosiglitazone treated HFD mice

(+Rosi). n = 8 per group, *p < 0.05 compared to

HFD-fed WT group.

(F) Mean ± SEM plasma concentration (mole (%))

of DHA and EPA for each diet in WT and GPR120

KO mice. n = 7 per each group. *p < 0.05,

compared to NC, and **p < 0.05 compared to

HFD. Data are represented as mean ± SEM. See

also Figure S4, Figure S5, Figure S6, and Table S2.

(GIR) required to maintain euglycemia in

the KO mice. Since 70%–80% of total

body insulin stimulated glucose disposal

is accounted for by skeletal muscle

glucose uptake (Baron et al., 1988), the

decreased insulin stimulated (IS)-glucose

disposal rate (GDR) provides direct

evidence for skeletal muscle insulin resis-

tance in the KO mice. Likewise, the

GPR120 KO mice exhibited a marked

decrease in the ability of insulin to sup-

press hepatic glucose production (HGP),

demonstrating the presence of hepatic

insulin resistance. Thus, the decreased

GIR was �50% related to muscle and �50% due to liver insulin

resistance, respectively. Since the chow diet contains exoge-

nous u-3 FAs, we conclude that blunted u-3 FA signaling in the

KO mice, accounts for the decreased insulin sensitivity.

Since u-3 FA administration can improve insulin sensitivity in

rats (Buettner et al., 2006), we reasoned that u-3 FA supplemen-

tation could alleviate HFD/obesity-induced insulin resistance in

WT mice, but would be ineffective in GPR120 KOs. Accordingly,

WT and GPR120 KO mice were placed on 60% HFD for 15 weeks.

At this point, separate groups of 15 mice each, were treated for

five additional weeks with 60% HFD or an isocaloric HFD diet con-

taining 27% fish oil supplementation enriched inu-3 FAs. This diet

provided 50 and 100 mg of DHA and EPA, respectively, per

mouse, per day. Figure 5E shows that administration of the u-3

FA diet led to improved insulin sensitivity with increased glucose

infusion rates, enhanced muscle insulin sensitivity (increased IS-

GDR), greater hepatic insulin sensitivity (increased HGP suppres-

sion), and decreased hepatic steatosis (Figures S6A and S6B).

Importantly, the u-3 FA diet was completely without effect in the

GPR120 KO mice. A separate group of WT mice were treated

with the insulin sensitizing thiazolidinedione Rosiglitazone, and

the effects ofu-3 FAs were equal to or greater (HGP suppression)

than the effects of this clinically used insulin sensitizing drug.

692 Cell 142, 687–698, September 3, 2010 ª2010 Elsevier Inc.

Page 61: Cell 100903

In addition to improving hepatic insulin sensitivity, u-3 FA

treatment had a beneficial effect on hepatic lipid metabolism,

causing decreased liver triglycerides, DAGs, along with reduced

SFA and u-6 FA content in the various lipid classes (Figures

S6A–S6C and Table S2). The u-3 FA supplementation was

entirely without effect, or much less effective, at reducing hepatic

lipid levels in the GPR120 KOs.

Interestingly, in the absence of u-3 FA supplementation,

GPR120 KO mice were just as susceptible to HFD-induced

insulin resistance as were the WT mice. We hypothesize that

this was because the 60% HFD is relatively deficient of exoge-

nous u-3 FAs, so that ligands for GPR120 were relatively absent

in these animals. To assess this, we performed a lipomics anal-

ysis of the various fatty acid classes in the chow and HFD-fed WT

and KO mice. As predicted, circulating concentrations of u-3

FAs were much lower on HFD compared to chow diets, and

the administration of the u-3 FA supplement to the HFD led to

a large increase in plasma u-3 FA content in both genotypes

(Figure 5F). This would account for the relative lack of effect of

GPR120 KO on HFD alone, since u-3 FA ligand stimulation is

negligible, while the KO animals displayed an insulin resistant

phenotype on chow diets when a moderate level of u-3 FAs

was provided. Importantly, the GPR120 KO mice are completely

refractory to the insulin sensitizing effects of u-3 FA administra-

tion on HFD.

To address the contribution of macrophages to the overall

in vivo phenotype, we performed bone marrow transplantation

(BMT) from GPR120 KOs into irradiated WT mice (adoptive

transfer) to generate hematopoietic cell deletion of GPR120.

The studies in the BMT WT and BMT GPR120 KO mice on

chow diet revealed a highly significant 20%–30% decrease in

GIR in the KOs, with a more dramatic impairment in the ability

of insulin to suppress hepatic glucose production (Figure S4A).

Thus, the studies in the BMT animals on the chow diet are

comparable to the results (Figure 5D) observed in WT versus

whole body GPR120 KOs on chow diet. When studied on the

HFD ±u-3 FA supplementation (Figure S4B), the u-3 FA supple-

mented BMT GPR120 KO animals exhibited a 30% decrease in

GIR compared to the u-3 FA supplemented BMT WTs. This was

explained by skeletal muscle insulin resistance (decreased IS-

GDR) and hepatic insulin resistance (decreased HGP suppres-

sion) in the GPR120 KOs compared to the WT BMT mice on

the u-3 FA supplemented HFD. These data are fully consistent

with the results in the global KOs (Figure 5E) and reinforce the

concept that the in vivo phenotype we observed can be largely

traced to hematopoietic cells/macrophages.

Omega-3 FAs Reduce Inflammatory Macrophagesin Adipose TissueWe conducted histologic examination of adipose tissue

macrophages (ATMs) from WT and GPR120 KO mice on HFD

or the u-3 FA enriched HFD by immunostaining for the M1

macrophage marker F4/80 and the M2 macrophage marker

MGL1 (Lumeng et al., 2008) (Figure 6A). Consistent with previous

studies (Weisberg et al., 2003; Xu et al., 2003; Nguyen et al.,

2007), HFD induced a large increase in F4/80 positive ATMs,

which form crown-like structures (CLS) around adipocytes in

both WT and GPR120 KO mice. In contrast, MGL1 staining

was minimal in both groups on HFD (Figure 6A). On the u-3 FA

diet, we observed a decrease in F4/80 staining, along with

a marked increase in MGL1 positive cells in WT mice. Impor-

tantly, no change in F4/80 or MGL1 staining was noted in the

GPR120 KO mice on the u-3 FA diet. SVFs were prepared

from adipose tissue and analyzed by flow cytometry to quanti-

tate the total number of ATMs, as well as the content of

CD11b+ and CD11c+ and negative macrophage subpopulations

(Figure 6B). HFD led to a large but comparable increase in

CD11b+ and CD11c+ ATM content in WT and GPR120 KO

mice (Figure 6B, middle panel). Treatment with the u-3 FA-en-

riched HFD caused a striking decrease in CD11b+ and CD11c+

ATMs in WT mice, but was without effect in the GPR120 KO

group (Figure 6B, right panel). Thus, the FACS analysis was fully

consistent with the histological results. Interestingly, CD11c+

ATM content was also greater in the GPR120 KOs on the chow

diet relative to WT consistent with the insulin resistance in the

KO animals.

It seemed possible that the reduction in ATM content in WT

animals on the u-3 FA diet reflected decreased chemotaxis of

macrophages. To test this hypothesis, we measured the migra-

tory capacity of IPMacs from WT and GPR120 KO mice using

an in vitro transwell chemotaxis assay. As seen in Figure 6C,

macrophages from both groups readily migrated toward condi-

tioned media (CM) harvested from 3T3-L1 adipocytes. Pretreat-

ment of macrophages with DHA for 3 hr before exposure to CM

led to an 80% inhibition of chemotactic capacity in WT macro-

phages, but had no significant effect on IPMacs obtained from

the GPR120 KO mice. Similar experiments were performed

using the specific chemokine, monocyte chemotactic protein-1

(MCP-1) as a chemoattractant, rather than CM, and these exper-

iments yielded identical results (Figure 6D). These data indicate

that u-3 FAs cause decreased macrophage chemotaxis by

acting through the GPR120 receptor, contributing to the differ-

ences in ATM content seen in Figures 6A and 6B.

Omega-3 FAs Decrease M1 Proinflammatory Geneand Increase M2 Anti-inflammatory Gene Expressionin Adipose TissueAs shown in Figure 7A, expression of M1 inflammatory genes

such as IL-6, TNF-a, MCP-1, IL-1b, iNOS, and CD11c was

increased by HFD compared to chow diet in both genotypes,

and was reduced in the u-3 FA treated WT mice, but not in

the GPR120 KO mice. Even on chow diet, expression of several

inflammatory genes was higher in GPR120 KOs compared to

WT, consistent with the insulin resistance observed in the

chow-fed KO mice. Expression of the M2 anti-inflammatory

genes, arginase 1, IL-10, MGL1, Ym-1, Clec7a, and MMR

was increased by u-3 FAs in WT, but not in the GPR120 KO

adipose tissue (Figure 7B). These results are consistent with

Figure 6 and demonstrate that the dietary change from HFD

to u-3 FA supplemented HFD resulted in an overall decreased

proinflammatory profile in adipose tissue from WT, but not in

GPR120 KO mice. These changes in gene expression were

predominantly manifested in the SVF, except for MCP-1 and

IL-6, which are known to be readily expressed in adipocytes

(Figure S7). Qualitatively similar results were seen in the liver

(Figures S6D and S6E).

Cell 142, 687–698, September 3, 2010 ª2010 Elsevier Inc. 693

Page 62: Cell 100903

DISCUSSION

In this report we show that GPR120 functions as an u-3 FA

receptor/sensor in proinflammatory macrophages and mature

adipocytes. By signaling through GPR120, DHA and EPA (the

major natural u-3 FA constituents of fish oil), mediate potent

anti-inflammatory effects to inhibit both TLR and TNF-a inflam-

matory signaling pathways. The mechanism of GPR120-medi-

ated anti-inflammation involves inhibition of TAK1 through a

b-arrestin2/TAB1 dependent effect. Since chronic tissue inflam-

mation is an important mechanism causing insulin resistance

(Xu et al., 2003; Shoelson et al., 2007; Schenk et al., 2008), the

anti-inflammatory actions of u-3 FAs exert potent insulin sensi-

tizing effects. The in vivo anti-inflammatory and insulin sensi-

tizing effects of u-3 FAs are dependent on expression of

GPR120, as demonstrated in studies of obese GPR120 KO

animals and WT littermates. Thus, GPR120 is highly expressed

in proinflammatory macrophages and functions as an u-3 FA

receptor, mediating the anti-inflammatory effects of this class

Figure 6. Omega-3 FA Enriched Diet Decreases

Inflammatory Macrophage Infiltration in Adipose

Tissue

(A) Confocal merged images from epididymal fat pads

from HFD and u-3 FA enriched HFD (HFD+u3)-fed WT

and GPR120 KO mice, costained with anti-F4/80 (green)

and anti-Caveolin1 (blue) antibodies, left 4 panels, or

anti-MGL1 (green) and anti-Caveolin1 (red) antibodies,

right 4 panels. The image is representative of similar

results from three to four independent experiments. Scale

bar represents 100 mm.

(B) Dot plot representation of CD11b versus CD11c

expression for FACS data obtained from adipose tissue

SVF of NC, HFD or HFD+u3-fed WT and GPR120 KO.

Scattergram is representative from three independent

mice from each group.

(C and D) Migratory capacity of IPMacs from WT and

GPR120 KO mice as measured using an in vitro transwell

chemotaxis assay as described under supplemental

experimental procedures. Data are expressed as mean ±

SEM of three independent experiments in triplicate. *,

p<0.05 versus CM treatment.

of FAs to inhibit both the TLR2/3/4 and the

TNF-a response pathways and cause systemic

insulin sensitization.

GPR120 is a Gaq/11-coupled receptor, and

since it is expressed in enteroendocrine L cells,

past interest in this receptor has focused on its

potential ability to stimulate L cell GLP-1 secre-

tion. In the current study, we show that, in addi-

tion to L cells, GPR120 is highly expressed in

proinflammatory, M1-like macrophages and

mature adipocytes, with negligible expression

in muscle, pancreatic b-cells, and hepatocytes

(Gotoh et al., 2007). In the HFD/obese mouse

model, GPR120 expression is highly induced in

ATMs as well as resident liver macrophages

(Kupffer cells). To explore the biology around

GPR120, we established an artificial reporter

cell assay and found that theu-3 FAs, DHA, and EPA, are ligands

for GPR120, and comparable to the effects of a non-selective

GPR120 tool compound (GW9508), the u-3 FAs exert potent

anti-inflammatory effects in macrophages. Our results also re-

vealed the molecular mechanisms underlying these anti-inflam-

matory effects. Thus, DHA stimulation of GPR120 inhibits both

the TLR2/3/4 and TNF-a proinflammatory cascade. Since acti-

vation of IKKb and JNK are common to TLR and TNF-a signaling,

this indicates that the locus of GPR120 inhibition is at or proximal

to these kinases. TAK1 activation stimulates both the IKKb/NFkB

and JNK/AP1 pathways, and the TLR and TNF-a signaling path-

ways converge at this step. Our data show that stimulation of

GPR120 specifically inhibits TAK1 phosphorylation and activa-

tion providing a common mechanism for the inhibition of both

TLR and TNF-a signaling.

Beta-arrestins can serve as important adaptor and scaffold

molecules mediating the functions of a number of different

GPCRs, as well as other receptor subtypes (Miller and Lefkowitz,

2001). The C-terminal region of GPR120 contains several

694 Cell 142, 687–698, September 3, 2010 ª2010 Elsevier Inc.

Page 63: Cell 100903

putative b-arrestin2 binding motifs [(S/T)X4-5(S/T); Cen et al.,

2001], but whether b-arrestins play any role in GPR120 function

was unknown. Here we find that activation of GPR120 by DHA

stimulation leads to association of the receptor with b-arrestin2,

but not b-arrestin1, and that the anti-inflammatory effects of

GPR120 are completely b-arrestin2 dependent. Functional

immunocytochemical studies showed that DHA stimulation

leads to recruitment of b-arrestin2 to the plasma membrane

where it colocalizes with GPR120. This is followed by receptor

and b-arrestin2 internalization, where the two are now colocal-

ized in the cytoplasmic compartment. TAB1 is the activating

protein for TAK1 and our results show that following DHA-stim-

ulated internalization of the GPR120/b-arrestin2 complex,

b-arrestin2 can now associate with TAB1, as measured in coim-

munoprecipitation experiments; only full-length b-arrestin2 was

capable of interacting with GPR120 and TAB1. This apparently

blocks the association of TAB1 with TAK1, inhibiting TAK1

activation and downstream signaling to the IKKb/NFkB and

JNK/AP1 system. These results provide a mechanism for the

b-arrestin2-mediated inhibition of TLR4, TNF-a, and TLR2/3

action. Other studies in the literature are consistent with these

findings, since it has been shown that b-arrestin2 can inhibit

NFkB signaling in other systems (Gao et al., 2004; Wang et al.,

2006b). Furthermore, Lefkowitz’ group has recently published

an extensive proteomics analysis of b-arrestin2 interacting

partners, and among the 266 proteins they identified, TAB1

was represented on the list (Xiao et al., 2007).

Interestingly, the anti-inflammatory effects mediated by

GPR120 were entirely dependent on b-arrestin2, but indepen-

dent of Gaq/11, despite the fact that GPR120 can be a Gaq/

Figure 7. M1 and M2 Inflammatory Gene Expres-

sion Levels in Adipose Tissue from WT versus

GPR120 KO Mice

Relative mRNA levels for M1 proinflammatory genes (A)

and M2 anti-inflammatory genes (B) in NC, HFD, or

HFD+u3 (+u3)-fed WT and GPR120 KO mice, as measured

by q-PCR. Data are expressed as mean ± SEM of three

independent experiments in triplicate. n=7 per group, *,

p<0.05 compared to the HFD-fed WT group. **, p<0.05

compared to the WT versus GPR120 KO on NC. See also

Figure S7. Primer sequences are shown in Table S1.

11-coupled receptor in other contexts. This

provides further evidence demonstrating the

concept that a single GPCR can independently

signal through multiple pathways. In previous

studies, we have demonstrated that Gaq/11

activation can lead to stimulation of GLUT4

translocation in adipocytes. Since GPR120

was expressed in mature adipocytes, but not

preadipocytes, we explored the potential role

of GPR120 in glucose transport control. Inter-

estingly, we found that DHA stimulation of

GPR120 in 3T3-L1 adipocytes increased

GLUT4 translocation to the cell surface with

a subsequent increase in glucose transport

into the cells. RNA interference studies showed

that the DHA effect on glucose uptake was GPR120, GLUT4, and

Gaq/11 dependent, but independent of b-arrestin2. This effect

was about 30%–50% as great as the effect of insulin and the

actions of DHA on glucose uptake were additive to those of

a submaximally stimulating concentration of insulin. From this,

it is possible to propose that these insulinometic effects con-

tribute to the overall insulin sensitizing actions of u-3 FAs.

However, muscle glucose uptake accounts for the great majority

(70%–80%) of insulin stimulated glucose disposal. Furthermore,

GPR120 is not expressed in muscle, and DHA did not stimulated

glucose uptake in L6 myocytes (Figures S4C and S4D). In addi-

tion, acute administration of DHA had no stimulatory effects on

IS-GDR (Figure S4E). This reports the conclusion that the

in vivo stimulatory effects of DHA on GDR are related to

anti-inflammation, and that the glucose transport stimulatory

effects in adipocytes contribute little to the overall phenotype.

Since chronic, low grade tissue inflammation is an important

cause of obesity-related insulin resistance, we reasoned that

the anti-inflammatory effects of GPR120 stimulation should be

coupled to insulin sensitizing actions in vivo. This idea was

confirmed in studies of WT and GPR120 KO mice. On a chow

diet, the lean GPR120 KO mice were glucose intolerant, hyperin-

sulinemic and displayed decreased skeletal muscle and hepatic

insulin sensitivity, as measured during glucose clamp studies.

They also displayed increased ATM content, relative to WT

mice, and a 2- to 5-fold higher expression level of the M1 proin-

flammatory markers, MCP-1, iNOS, and IL-6 (Figure 7A). On

HFD, GPR120 KO and WT mice became equally obese and

insulin resistant. Importantly, u-3 FA supplementation markedly

increased insulin sensitivity in WT mice but was without effect in

Cell 142, 687–698, September 3, 2010 ª2010 Elsevier Inc. 695

Page 64: Cell 100903

the GPR120 KOs. Consistent with these results, u-3 FA treat-

ment led to a decrease in ATM accumulation with reduced

adipose tissue markers of inflammation in WT, but not in KO

mice. In addition to direct anti-inflammatory effects in macro-

phages, DHA treatment inhibited the ability of primary WT

macrophages to migrate toward adipocyte CM. This could be

due to DHA-induced decreased chemokine secretion or down-

regulation of chemokine receptors, or both. In addition, it is

possible that DHA, by signaling through GPR120, can mediate

heterologous desensitization of other GPCR chemokine recep-

tors. We also observed a concomitant increase in M2 markers,

such as IL-10, arginase 1, MGL1, Ym-1, Clec7a, and MMR.

This latter finding raises the possibility that u-3 FAs can redirect

ATMs from an M1 to an M2 polarization state. Taken together,

these mechanisms account for the decreased inflammatory

state. The in vivo anti-inflammatory actions of u-3 FAs are

consistent with the insulin sensitizing effects of these agents

and are fully dependent on the presence of GPR120, indicating

a causal relationship. Finally, the adoptive transfer studies

showed that hematopoietic cell GPR120 deletion results in

a comparable insulin resistant, u-3 FA non-responsive pheno-

type as seen in the global GPR120 KOs, indicating that this

phenotype can be traced back to inflammatory events in

macrophages.

We also performed a detailed in vivo lipidomic analysis of FAs

in the different lipid classes in the liver (Table S2). The results

showed that HFD leads to an increase in total TAGs, DAGs, total

SFAs, monounsaturated FAs andu-6 FAs in WT mice, while all of

these lipid changes are ameliorated withu-3 FA treatment. In the

GPR120 KO mice, all of these lipids are elevated on HFD to the

same extent as in WT mice, but, u-3 FA supplementation was

either ineffective or much less effective. These results are

consistent with the view that the reversal of steatosis/non-

alcoholic fatty liver disease (NAFLD) by u-3 FA treatment is

mediated, in large part, by GPR120 and that the GPR120 KO

mice are predisposed toward NAFLD even in the context of

u-3 FA supplementation.

Dietary DHA is rapidly esterified into chylomicrons during the

process of gastrointestinal absorption, and is also packaged

into VLDL triglycerides by the liver. DHA can also be esterified

into phospholipids and cholesterol esters associated with

circulating lipopoproteins and only a small proportion (�5%) of

total plasma DHA is found in the FFA pool. Through the action

of lipoprotein lipase bound to the luminal surface of endothelial

cells, u-3 FAs are cleaved from circulating triglycerides where

they can act as ligands or be taken up by peripheral tissues

(Polozova and Salem Jr., 2007). Recent studies have also indi-

cated that metabolic products derived from u-3 FAs, such as

17S-hydroxy-DHA, resolvins, and protectins may play a role in

the long term resolution of inflammation and this might attenuate

insulin resistance in the context of obesity (Gonzalez-Periz et al.,

2009). If this proves to be correct, then this could provide an

additional mechanism for long term u-3 FA-induced anti-inflam-

matory, insulin sensitizing effects. However, in the current

studies, we found that these u-3 FA derivatives were unable to

stimulate GPR120 activation in our reporter cell assay (data not

shown), indicating that GPR120 functions as a receptor for u-3

FAs and not their biochemical products. Resolution of inflamma-

tion versus anti-inflammatory actions are distinct processes, and

it is certainly possible that the products derived from u-3 FA

metabolism work on the former but not the latter.

In summary, we have found that GPR120 functions as an u-3

FA receptor/sensor and mediates robust and broad anti-inflam-

matory effects, particularly in macrophages. After ligand stimula-

tion, GPR120 couples to b-arrestin2 which is followed by

receptor endocytosis and inhibition of TAB1-mediated activation

of TAK1, providing a mechanism for inhibition of both the TLR

and TNF-a proinflammatory signaling pathways. Since chronic

tissue inflammation is linked to insulin resistance in obesity, we

used GPR120 KO mice to demonstrate that u-3 FAs cause

GPR120-mediated anti-inflammatory and insulin sensitizing

effects in vivo. Overall, these results strongly argue that anti-

inflammatory effects can ameliorate insulin resistance in obesity.

Taken together, GPR120 emerges as an important control point

in the integration of anti-inflammatory and insulin sensitizing

responses, which may prove useful in the future development

of new therapeutic approaches for the treatment of insulin resis-

tant diseases.

EXPERIMENTAL PROCEDURES

Chemicals and Reagents

GW9508 was purchased from Tocris bioscience (Ellisville, MO) and DHA was

from Cayman chemical (Ann Arbor, MI). All other chemicals were purchased

from Sigma unless mentioned otherwise.

Animal Care and Use

Male C57Bl/6 or GPR120 KO littermates were fed a normal chow (13.5% fat;

LabDiet) or high-fat diet (60% fat; Research Diet) ad libitum for 15–20 weeks

from 8 weeks of age. GPR120 KO mice and WT littermates were provided

by Taconic Inc. (Hudson, NY). After 15 weeks on HFD, WT and GPR120 KO

mice were switched to an isocaloric HFD-containing 27% menhaden fish oil

replacement (wt/wt; menhaden fish oil: 16% EPA (C20:5n3), 9%, DHA

(C22:6n3), Research Diet) (Jucker et al., 1999; Neschen et al., 2007) and fed

for 5 weeks. Mice received fresh diet every 3rd day, and food consumption

and body weight were monitored. Animals were housed in a specific path-

ogen-free facility and given free access to food and water. All procedures

were approved by the University of California, San Diego animal care and

use committee. In vivo metabolic studies were performed as described under

supplemental experimental procedures.

Data Analysis

Densitometric quantification and normalization were performed using the

ImageJ 1.42q software. The values presented are expressed as the means ±

SEM. The statistical significance of the differences between various treat-

ments was determined by one-way ANOVA with the Bonferroni correction

using GraphPad Prism 4.0 (San Diego, CA). The p < 0.05 was considered

significant.

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures, seven

figures, and two tables and can be found with this article online at doi:10.1016/

j.cell.2010.07.041.

ACKNOWLEDGMENTS

We thank Jachelle M. Ofrecio and Sarah Nalbandian for their help with animal

maintenance and Elizabeth J. Hansen for editorial assistance. We are grateful

to Dr. Robert Lefkowitz (Howard Hughes Medical Institute, Duke University) for

the gift of FLAG-tagged serial mutant b-arrestin2 constructs and to Dr. Maziyar

696 Cell 142, 687–698, September 3, 2010 ª2010 Elsevier Inc.

Page 65: Cell 100903

Saberi at NGM Bio Inc. (San Francisco, CA) for GLP-1 measurements. We

thank the Flow Cytometry Resource and Neal Sekiya for assistance with

FACS analysis at the VA San Diego hospital, the UCSD Histology Core lab

for technical help with processing liver specimens, and UCSD Microscope

Resource for microscopy analysis. This study was funded in part by the

National Institutes of Health grants NIDDK DK033651 (J.M.O.), DK063491

(J.M.O.), DK 074868 (J.M.O.), and the Eunice Kennedy Shriver NICHD/NIH

through a cooperative agreement U54 HD 012303-25 as part of the specialized

Cooperative Centers Program in Reproduction and Infertility Research.

Received: January 20, 2010

Revised: May 24, 2010

Accepted: July 19, 2010

Published: September 2, 2010

REFERENCES

Barak, L.S., Ferguson, S.S., Zhang, J., and Caron, M.G. (1997). A beta-

arrestin/green fluorescent protein biosensor for detecting G protein-coupled

receptor activation. J. Biol. Chem. 272, 27497–27500.

Baron, A.D., Brechtel, G., Wallace, P., and Edelman, S.V. (1988). Rates and

tissue sites of non-insulin- and insulin-mediated glucose uptake in humans.

Am. J. Physiol. 255, E769–E774.

Briscoe, C.P., Peat, A.J., McKeown, S.C., Corbett, D.F., Goetz, A.S., Littleton,

T.R., McCoy, D.C., Kenakin, T.P., Andrews, J.L., Ammala, C., et al. (2006).

Pharmacological regulation of insulin secretion in MIN6 cells through the fatty

acid receptor GPR40: identification of agonist and antagonist small molecules.

Br. J. Pharmacol. 148, 619–628.

Buettner, R., Parhofer, K.G., Woenckhaus, M., Wrede, C.E., Kunz-Schughart,

L.A., Scholmerich, J., and Bollheimer, L.C. (2006). Defining high-fat-diet rat

models: metabolic and molecular effects of different fat types. J. Mol. Endocri-

nol. 36, 485–501.

Calder, P.C. (2005). Polyunsaturated fatty acids and inflammation. Biochem.

Soc. Trans. 33, 423–427.

Cen, B., Xiong, Y., Ma, L., and Pei, G. (2001). Direct and differential interaction

of beta-arrestins with the intracellular domains of different opioid receptors.

Mol. Pharmacol. 59, 758–764.

Gao, H., Sun, Y., Wu, Y., Luan, B., Wang, Y., Qu, B., and Pei, G. (2004).

Identification of beta-arrestin2 as a G protein-coupled receptor-stimulated

regulator of NF-kappaB pathways. Mol. Cell 14, 303–317.

Gether, U. (2000). Uncovering molecular mechanisms involved in activation of

G protein-coupled receptors. Endocr. Rev. 21, 90–113.

Gonzalez-Periz, A., Horrillo, R., Ferre, N., Gronert, K., Dong, B., Moran-

Salvador, E., Titos, E., Martinez-Clemente, M., Lopez-Parra, M., Arroyo, V.,

and Claria, J. (2009). Obesity-induced insulin resistance and hepatic steatosis

are alleviated by omega-3 fatty acids: a role for resolvins and protectins.

FASEB J. 23, 1946–1957.

Gotoh, C., Hong, Y.H., Iga, T., Hishikawa, D., Suzuki, Y., Song, S.H., Choi,

K.C., Adachi, T., Hirasawa, A., Tsujimoto, G., et al. (2007). The regulation of

adipogenesis through GPR120. Biochem. Biophys. Res. Commun. 354,

591–597.

Hirasawa, A., Tsumaya, K., Awaji, T., Katsuma, S., Adachi, T., Yamada, M.,

Sugimoto, Y., Miyazaki, S., and Tsujimoto, G. (2005). Free fatty acids regulate

gut incretin glucagon-like peptide-1 secretion through GPR120. Nat. Med. 11,

90–94.

Imamura, T., Vollenweider, P., Egawa, K., Clodi, M., Ishibashi, K., Nakashima,

N., Ugi, S., Adams, J.W., Brown, J.H., and Olefsky, J.M. (1999). G alpha-q/11

protein plays a key role in insulin-induced glucose transport in 3T3-L1 adipo-

cytes. Mol. Cell. Biol. 19, 6765–6774.

Itoh, Y., Kawamata, Y., Harada, M., Kobayashi, M., Fujii, R., Fukusumi, S., Ogi,

K., Hosoya, M., Tanaka, Y., Uejima, H., et al. (2003). Free fatty acids regulate

insulin secretion from pancreatic beta cells through GPR40. Nature 422,

173–176.

Jucker, B.M., Cline, G.W., Barucci, N., and Shulman, G.I. (1999). Differential

effects of safflower oil versus fish oil feeding on insulin-stimulated glycogen

synthesis, glycolysis, and pyruvate dehydrogenase flux in skeletal muscle:

a 13C nuclear magnetic resonance study. Diabetes 48, 134–140.

Kawai, T., and Akira, S. (2006). TLR signaling. Cell Death Differ. 13, 816–825.

Lee, J.Y., Plakidas, A., Lee, W.H., Heikkinen, A., Chanmugam, P., Bray, G., and

Hwang, D.H. (2003). Differential modulation of Toll-like receptors by fatty

acids: preferential inhibition by n-3 polyunsaturated fatty acids. J. Lipid Res.

44, 479–486.

Lumeng, C.N., DelProposto, J.B., Westcott, D.J., and Saltiel, A.R. (2008).

Phenotypic switching of adipose tissue macrophages with obesity is gener-

ated by spatiotemporal differences in macrophage subtypes. Diabetes 57,

3239–3246.

Luttrell, L.M., Ferguson, S.S., Daaka, Y., Miller, W.E., Maudsley, S., Della

Rocca, G.J., Lin, F., Kawakatsu, H., Owada, K., Luttrell, D.K., et al. (1999).

Beta-arrestin-dependent formation of beta2 adrenergic receptor-Src protein

kinase complexes. Science 283, 655–661.

Luttrell, L.M., and Lefkowitz, R.J. (2002). The role of beta-arrestins in the

termination and transduction of G-protein-coupled receptor signals. J. Cell

Sci. 115, 455–465.

Miller, W.E., and Lefkowitz, R.J. (2001). Expanding roles for beta-arrestins as

scaffolds and adapters in GPCR signaling and trafficking. Curr. Opin. Cell Biol.

13, 139–145.

Neschen, S., Morino, K., Dong, J., Wang-Fischer, Y., Cline, G.W., Romanelli,

A.J., Rossbacher, J.C., Moore, I.K., Regittnig, W., Munoz, D.S., et al. (2007).

n-3 Fatty acids preserve insulin sensitivity in vivo in a peroxisome

proliferator-activated receptor-alpha-dependent manner. Diabetes 56,

1034–1041.

Nguyen, M.T., Favelyukis, S., Nguyen, A.K., Reichart, D., Scott, P.A., Jenn, A.,

Liu-Bryan, R., Glass, C.K., Neels, J.G., and Olefsky, J.M. (2007). A subpopula-

tion of macrophages infiltrates hypertrophic adipose tissue and is activated by

free fatty acids via Toll-like receptors 2 and 4 and JNK-dependent pathways. J.

Biol. Chem. 282, 35279–35292.

Oh, D.Y., Song, J.A., Moon, J.S., Moon, M.J., Kim, J.I., Kim, K., Kwon, H.B.,

and Seong, J.Y. (2005). Membrane-proximal region of the carboxyl terminus

of the gonadotropin-releasing hormone receptor (GnRHR) confers differential

signal transduction between mammalian and nonmammalian GnRHRs. Mol.

Endocrinol. 19, 722–731.

Polozova, A., and Salem, N., Jr. (2007). Role of liver and plasma lipoproteins

in selective transport of n-3 fatty acids to tissues: a comparative study of

14C-DHA and 3H-oleic acid tracers. J. Mol. Neurosci. 33, 56–66.

Schenk, S., Saberi, M., and Olefsky, J.M. (2008). Insulin sensitivity: modulation

by nutrients and inflammation. J. Clin. Invest. 118, 2992–3002.

Schulte, G., and Fredholm, B.B. (2003). Signaling from adenosine receptors to

mitogen-activated protein kinases. Cell Signal 15, 813–827.

Shoelson, S.E., Herrero, L., and Naaz, A. (2007). Obesity, inflammation, and

insulin resistance. Gastroenterology 132, 2169–2180.

Solinas, G., Vilcu, C., Neels, J.G., Bandyopadhyay, G.K., Luo, J.L., Naugler,

W., Grivennikov, S., Wynshaw-Boris, A., Scadeng, M., Olefsky, J.M., and

Karin, M. (2007). JNK in hematopoietically derived cells contributes to diet-

induced inflammation and insulin resistance without affecting obesity. Cell

Metab. 6, 386–397.

Takaesu, G., Surabhi, R.M., Park, K.J., Ninomiya-Tsuji, J., Matsumoto, K., and

Gaynor, R.B. (2003). TAK1 is critical for IkappaB kinase-mediated activation of

the NF-kappaB pathway. J. Mol. Biol. 326, 105–115.

Tazoe, H., Otomo, Y., Kaji, I., Tanaka, R., Karaki, S.I., and Kuwahara, A. (2008).

Roles of short-chain fatty acids receptors, GPR41 and GPR43 on colonic func-

tions. J. Physiol. Pharmacol. 59 (Suppl 2), 251–262.

Ulloa-Aguirre, A., Stanislaus, D., Janovick, J.A., and Conn, P.M. (1999).

Structure-activity relationships of G protein-coupled receptors. Arch. Med.

Res. 30, 420–435.

Cell 142, 687–698, September 3, 2010 ª2010 Elsevier Inc. 697

Page 66: Cell 100903

Wang, J., Wu, X., Simonavicius, N., Tian, H., and Ling, L. (2006a). Medium-

chain fatty acids as ligands for orphan G protein-coupled receptor GPR84.

J. Biol. Chem. 281, 34457–34464.

Wang, Y., Tang, Y., Teng, L., Wu, Y., Zhao, X., and Pei, G. (2006b). Association

of beta-arrestin and TRAF6 negatively regulates Toll-like receptor-interleukin 1

receptor signaling. Nat. Immunol. 7, 139–147.

Weisberg, S.P., McCann, D., Desai, M., Rosenbaum, M., Leibel, R.L., and

Ferrante, A.W., Jr. (2003). Obesity is associated with macrophage accumula-

tion in adipose tissue. J. Clin. Invest. 112, 1796–1808.

Xiao, K., McClatchy, D.B., Shukla, A.K., Zhao, Y., Chen, M., Shenoy, S.K.,

Yates, J.R., 3rd, and Lefkowitz, R.J. (2007). Functional specialization of

beta-arrestin interactions revealed by proteomic analysis. Proc. Natl. Acad.

Sci. USA 104, 12011–12016.

Xu, H., Barnes, G.T., Yang, Q., Tan, G., Yang, D., Chou, C.J., Sole, J., Nichols,

A., Ross, J.S., Tartaglia, L.A., and Chen, H. (2003). Chronic inflammation in fat

plays a crucial role in the development of obesity-related insulin resistance. J.

Clin. Invest. 112, 1821–1830.

698 Cell 142, 687–698, September 3, 2010 ª2010 Elsevier Inc.

Page 67: Cell 100903

Anti-CD47 Antibody Synergizes withRituximab to Promote Phagocytosis andEradicate Non-Hodgkin LymphomaMark P. Chao,1,10,* Ash A. Alizadeh,1,2,3,10 Chad Tang,1 June H. Myklebust,3,9 Bindu Varghese,3 Saar Gill,5 Max Jan,1

Adriel C. Cha,1 Charles K. Chan,1 Brent T. Tan,4 Christopher Y. Park,1,4 Feifei Zhao,1 Holbrook E. Kohrt,2,3

Raquel Malumbres,6 Javier Briones,7 Randy D. Gascoyne,8 Izidore S. Lossos,6 Ronald Levy,3 Irving L. Weissman,1,4,10

and Ravindra Majeti1,2,10

1Institute for Stem Cell Biology and Regenerative Medicine, Stanford Cancer Center, and Ludwig Center at Stanford2Department of Internal Medicine, Division of Hematology3Division of Oncology4Department of Pathology5Division of Blood and Bone Marrow Transplantation

Stanford University, Palo Alto, CA 94304, USA6Department of Medicine, Division of Hematology-Oncology, University of Miami Miller School of Medicine, Miami, FL 33136, USA7Department of Hematology, Hospital Santa Creu i Sant Pau, Autonomous University of Barcelona, Barcelona 08193, Spain8Department of Pathology, British Columbia Cancer Agency, Vancouver V5Z 1L3, Canada9Department of Immunology, Institute for Cancer Research, Oslo University Hospital/Centre for Cancer Biomedicine, University of Oslo,

Oslo 0310, Norway10These authors contributed equally to this work*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.07.044

SUMMARY

Monoclonal antibodies are standard therapeutics forseveral cancers including the anti-CD20 antibodyrituximab for B cell non-Hodgkin lymphoma (NHL).Rituximab and other antibodies are not curativeand must be combined with cytotoxic chemotherapyfor clinical benefit. Here we report the eradicationof human NHL solely with a monoclonal antibodytherapy combining rituximab with a blocking anti-CD47 antibody. We identified increased expressionof CD47 on human NHL cells and determined thathigher CD47 expression independently predictedadverse clinical outcomes in multiple NHL subtypes.Blocking anti-CD47 antibodies preferentially enabledphagocytosis of NHL cells and synergized with ritux-imab. Treatment of human NHL-engrafted mice withanti-CD47 antibody reduced lymphoma burden andimproved survival, while combination treatmentwith rituximab led to elimination of lymphoma andcure. These antibodies synergized through a mecha-nism combining Fc receptor (FcR)-dependent andFcR-independent stimulation of phagocytosis thatmight be applicable to many other cancers.

INTRODUCTION

Emerging evidence has demonstrated that monoclonal anti-

bodies (mAbs) either alone or in combination are an effective

modality for cancer treatment (Adams and Weiner, 2005).

Although therapies combining a mAb with chemotherapeutic

agents are effective in several human cancers, antibodies alone

are not curative. Antibodies effective against cancer are believed

to function by several mechanisms including antibody-depen-

dent cellular cytotoxicity (ADCC), stimulation of complement-

dependent cytotoxicity (CDC), inhibition of signal transduction,

or direct induction of apoptosis (Cheson and Leonard, 2008).

Non-Hodgkin lymphoma (NHL) is the fifth most common

cancer in the United States consisting of indolent and aggressive

subtypes with a 5 year overall survival ranging from 25%–75%

(The International Non-Hodgkin’s Lymphoma Prognostic

Factors Project, 1993). The anti-CD20 antibody, rituximab (Rit-

uxan), is a standard therapy for many CD20-positive B cell

lymphomas and significantly improves long-term survival in

combination with conventional chemotherapy (Cheson and Leo-

nard, 2008). As a single agent or in combination with chemo-

therapy, rituximab is not curative in the majority of B cell NHL

patients and rituximab resistance has been observed (reviewed

in Cheson and Leonard, 2008). Multiple lines of evidence indicate

that rituximab acts at least in part by engaging Fc receptors

(FcRs) on immune effector cells, such as NK cells and macro-

phages, and stimulating effector functions such as ADCC (Glen-

nie et al., 2007; Nimmerjahn and Ravetch, 2007). Although resis-

tance has been reported to occur through several mechanisms

(Cartron et al., 2004), there has been limited development of

agents that can overcome this resistance.

Immune effector cells, including NK cells and phagocytes, are

critical to the efficacy of many anticancer antibodies. Phagocytic

cells, including macrophages and dendritic cells, express signal

regulatory protein alpha (SIRPa), which binds CD47, a widely

Cell 142, 699–713, September 3, 2010 ª2010 Elsevier Inc. 699

Page 68: Cell 100903

expressed transmembrane protein (Brown and Frazier, 2001).

CD47-mediated activation of SIRPa initiates a signal transduc-

tion cascade resulting in inhibition of phagocytosis (reviewed in

Jaiswal et al., 2010). In identifying a role for CD47 in cancer path-

ogenesis, we previously demonstrated that forced expression of

mouse CD47 on a human leukemia cell line facilitated tumor

engraftment in immunodeficient mice through the evasion of

phagocytosis (Jaiswal et al., 2009). We further demonstrated

that this mechanism could be targeted therapeutically in human

acute myeloid leukemia (AML) with a blocking anti-CD47 anti-

body that enabled phagocytosis and eliminated AML stem cells

(Majeti et al., 2009). Based on this antibody mechanism, we

hypothesized that the combination of a blocking anti-CD47

antibody with a second FcR-activating antibody would both

prevent an inhibitory signal and deliver a positive stimulus,

resulting in the synergistic phagocytosis and elimination of tar-

get cells. Here, we tested this antibody synergy hypothesis by

investigating the combination of a blocking anti-CD47 mAb

with rituximab against human NHL.

RESULTS

CD47 Expression Is Increased on NHL Cells Comparedto Normal B CellsWe examined CD47 protein expression on primary human NHL

samples and normal B cells by flow cytometry. Compared to

both normal peripheral blood and germinal center B cells,

CD47 was more highly expressed on a large subset of primary

patient samples from multiple B cell NHL subtypes (Figure 1A

and Figure S1A available online), including diffuse large B cell

lymphoma (DLBCL), B cell chronic lymphocytic leukemia

(B-CLL), mantle cell lymphoma (MCL), follicular lymphoma (FL),

marginal zone lymphoma (MZL), and pre-B acute lymphoblastic

leukemia (pre-B ALL). Across NHL subtypes, we found differing

levels of CD47 expression that also varied within each NHL

subtype (Figure 1B).

Increased CD47 Expression Correlates with a WorseClinical Prognosis and Adverse Molecular Featuresin Multiple NHL SubtypesHaving previously shown a correlation between CD47 mRNA

and protein expression (Majeti et al., 2009), we assessed CD47

mRNA expression across NHL subtypes for associations with

morphologic and molecular subgroups using gene expression

data from previously described patient cohorts (Table S1) and

investigated the prognostic implications of increased CD47

expression in disease outcome. Higher CD47 expression was

associated with adverse prognosis in DLBCL, B-CLL, and MCL

(Figures 1C–1E). In patients with DLBCL, whether treated with

or without rituximab-based combination chemotherapy (Fig-

ure 1C and Figure S1B), higher CD47 expression was signifi-

cantly associated with risk of death. This increased risk was

largely due to disease progression (Figure S1C), a finding vali-

dated in an independent cohort of patients using quantitative

RT-PCR on fixed archival specimens (Figure S1D).

We next investigated whether increased CD47 expression was

associated with known adverse molecular features in NHL.

In DLBCL, prior studies have identified two distinct subgroups

based on the presumed cell of origin of tumors: normal germinal

center B cells (GCB-like), which are associated with a favorable

clinical outcome, or activated blood memory B cells (ABC-like),

which are associated with a poor clinical outcome (Alizadeh

et al., 2000; Rosenwald et al., 2002). CD47 expression was sig-

nificantly higher in ABC-like DLBCL (Figure 1F and Figure S1E).

CD47 expression was not found to have independent prog-

nostic value within GCB and ABC subtypes, suggesting a strong

association with the cell-of-origin classification of DLBCL.

Higher CD47 expression was also associated with unmutated

immunoglobulin heavy chain variable regions (IgVH) in CLL

(Figure 1G and Figure S1E) and significantly correlated with

the proliferative index in MCL (Figure 1H), both known adverse

prognostic factors (Katzenberger et al., 2006; Oscier et al.,

2002; Rosenwald et al., 2003). In multivariate analyses, CD47

expression remained prognostic of disease progression inde-

pendent from the international prognostic index in DLBCL (The

International Non-Hodgkin’s Lymphoma Prognostic Factors

Project, 1993) and two major prognostic factors in CLL: IgVH

mutation status and ZAP-70 status (Figure S1G). Within the

small MCL cohort, a multivariate model did not find independent

prognostic value for CD47 when considering the proliferation

index (data not shown).

Blocking Anti-CD47 Antibodies Enable Phagocytosisof NHL Cells by Macrophages and Synergizewith Rituximab In VitroWe first tested the ability of anti-human CD47 antibodies to

enable phagocytosis of human NHL cell lines, primary NHL cells,

and normal peripheral blood (NPB) cells by human macrophages

in vitro. Incubation of NHL cells in the presence of IgG1 isotype

control or anti-CD45 IgG1 antibody did not result in significant

phagocytosis; however, two different blocking anti-CD47 anti-

bodies (B6H12.2 and BRIC126) enabled phagocytosis of NHL

cells but not NPB cells (Figures 2A and 2B).

Next, we repeated the in vitro phagocytosis assays with

mouse macrophages. Incubation of NHL cells in the presence

of IgG1 isotype control or anti-CD45 IgG1 antibody did not result

in significant phagocytosis; however, phagocytosis of NHL cells

was observed with blocking antibodies to CD47 (B6H12.2 and

BRIC126), whereas no phagocytosis was observed with a non-

blocking antibody (2D3) (Figure 2B). Disruption of the CD47-

SIRPa interaction with an anti-mouse SIRPa antibody also

resulted in significant phagocytosis (Figure 2B).

Given variable expression of CD47 on primary NHL, we inves-

tigated by two independent methods whether CD47 expression

levels correlated with the degree of anti-CD47 antibody-medi-

ated phagocytosis. First, lentiviral shRNA vectors were used to

knock down expression of CD47 in Raji cells. Several clones

were generated with a range of CD47 knockdown (Figures S2A

and S2B). Those clones with a greater than 50% reduction in

CD47 expression (shCD47-1 and shCD47-2) demonstrated

a significant reduction in anti-CD47 antibody-mediated phago-

cytosis (Figure S3C). In the second approach, a statistical anal-

ysis demonstrated a positive correlation between CD47 expres-

sion and degree of anti-CD47 antibody-mediated phagocytosis

with both mouse and human macrophage effector cells

(Figure S2D).

700 Cell 142, 699–713, September 3, 2010 ª2010 Elsevier Inc.

Page 69: Cell 100903

Figure 1. CD47 Expression Is Increased on NHL Cells Compared to Normal B Cells, Confers a Worse Clinical Prognosis, and Correlates with

Adverse Molecular Features in Multiple NHL Subtypes

(A) CD47 expression on normal peripheral blood (PB) B cells (CD19+), normal germinal center (GC) B cells (CD3�CD5�CD20+CD10+CD38+), and NHL B cells

(CD19+) was determined by flow cytometry, and mean fluorescence intensity was normalized for cell size. Each point represents a different patient sample:

DLBCL = 2, CLL = 15, MCL = 4, FL = 6, MZL = 2, and pre-B ALL = 1 (****p < 0.0001). Normalized mean expression (and range) for each population were: normal

PB B cells 420.9 (267.3–654.0), normal GC B cells 482.5 (441.1–519.9), and NHL 888.5 (270.1–1553).

(B) CD47 expression across NHL subtypes including DLBCL (DL, n = 15), MCL (n = 34), FL (n = 28), and B-CLL (n = 14) was determined as in (A). Normalized mean

expression (and range) for each population were: DL 725.7 (261.2–1344), MCL 1055 (444.2–2196), FL 825.1 (283.6–1546), CLL 713.6 (432.8–1086) (*p < 0.05).

(C–E) Prognostic influence of CD47 mRNA expression is shown on overall (C and E) and event-free (D) survival of patients with DLBCL, B-CLL, and MCL.

For DLBCL and CLL, stratification into low and high CD47 expression groups was based on an optimal threshold determined by microarray analysis; this cutpoint

was internally validated for both DLBCL and CLL and also externally validated in an independent DLBCL cohort. For MCL, stratification relative to the median was

employed as an optimal cutpoint could not be defined. Significance measures are based on log-likelihood estimates of the p value, when treating CD47 expres-

sion as a continuous variable, with corresponding dichotomous indices also provided in Table S1.

(F–H) CD47 mRNA expression is shown in relation to cell-of-origin classification for DLBCL (F), immunoglobulin heavy chain mutation status (IgVH) for CLL (G),

and proliferation index for MCL (H). Error bars represent upper and lower quartiles (F and G). Analyses for (C)–(H) employed publicly available datasets for NHL

patients (Table S1). NGC = normal germinal center, ABC = activated B cell-like, GCB = germinal center B cell-like.

See also Figure S1 and Table S1.

Cell 142, 699–713, September 3, 2010 ª2010 Elsevier Inc. 701

Page 70: Cell 100903

Figure 2. Blocking Antibodies against CD47 Enable Phagocytosis of NHL Cells by Macrophages and Synergize with Rituximab In Vitro

(A) CFSE-labeled NHL cells were incubated with human macrophages and the indicated antibodies and examined by immunofluorescence microscopy to detect

phagocytosis (arrows). Photomicrographs from a representative NHL sample are shown.

(B) Phagocytic indices of primary human NHL cells (blue), normal peripheral blood (NPB) cells (red), and NHL cell lines (purple, orange, and green) were deter-

mined using human (left) and mouse (middle) macrophages. Antibody-induced apoptosis (right panel) was tested by incubating NHL cells with the indicated anti-

bodies or staurosporine without macrophages and assessing the percentage of apoptotic and dead cells (% annexin V and/or PI positive).

702 Cell 142, 699–713, September 3, 2010 ª2010 Elsevier Inc.

Page 71: Cell 100903

It has been reported that immobilized or crosslinked anti-

bodies against CD47 induce apoptosis of primary human

B-CLL cells, as well as several malignant lymphoid cell lines

(Kikuchi et al., 2004, 2005; Mateo et al., 1999; Uno et al.,

2007). Therefore, anti-CD47 antibodies might be predicted to

directly induce apoptosis of NHL cells that are then recognized

by macrophages and phagocytosed. Contrary to this prediction,

when NHL cells were incubated with anti-CD47 antibody in

the absence of macrophages, no induction of apoptosis was

observed when cells were incubated in suspension for either

2 hr (Figure 2B, right) or 8 hr (Figures S2E and S2F). Incubation

of NHL cells with immobilized anti-CD47 antibody resulted in

increased apoptosis compared to controls (Figures S2G and

S2H), consistent with prior reports (Mateo et al., 1999). Given

that phagocytosis of NHL cells occurs in the presence of soluble

anti-CD47 mAbs, it is unlikely that these mAbs induce apoptosis

of NHL cells that are then secondarily phagocytosed.

Next, we tested the ability of a blocking anti-CD47 mAb to syn-

ergize with rituximab in the phagocytosis of NHL cells. We exam-

ined CD20 expression on NHL cells and found no difference

between normal B cells and NHL cells (Figures S2I and S2J).

Incubation of NHL cells with rituximab in the presence of mouse

or human macrophages resulted in significant phagocytosis

(Figures 2D and S2E). We then tested the synergy hypothesis

by isobologram analysis (Chou, 2008; Tallarida, 2006). Using

Raji, SUDHL4, and NHL17 cells, which express varying levels

of both CD47 and CD20 (Figure S2K), anti-CD47 antibody

synergized with rituximab or anti-human CD20 (mouse IgG2a)

antibody, as indicated by combination indices less than 1 (Fig-

ure 2C). In a second approach, in vitro phagocytosis assays

were conducted with primary NHL cells incubated with either

anti-CD47 antibody or rituximab alone, or both in combination

at half of the single agent dose. NHL cells exhibited a significant

increase in phagocytosis when incubated with the combination

compared to either antibody alone when using mouse (Fig-

ure 2D) or human (Figure 2E) macrophage effectors. No phago-

cytosis of NPB cells was observed with either antibody alone or

in combination with human macrophages (Figure 2E).

Ex Vivo Coating of NHL Cells with an Anti-CD47 AntibodyInhibits Tumor EngraftmentNext, the ability of blocking anti-CD47 antibody to eliminate NHL

in vivo either alone or in combination with rituximab was explored

by two separate treatment strategies. First, the effect of ex vivo

anti-CD47 antibody coating on the engraftment of human NHL

cells was tested. Luciferase-expressing Raji and SUDHL4 cell

lines were precoated ex vivo with anti-CD47, IgG1 isotype

control, or anti-CD45 antibody and transplanted intravenously

into SCID mice. Coating with anti-CD47 antibody prevented

engraftment of both cell lines (Figures 3A–3F). Coating of Raji

cells with rituximab also inhibited engraftment when trans-

planted into SCID mice (Figure S3). In addition to these cell lines,

we identified a primary NHL patient specimen that engrafted in

NSG mice in the bone marrow upon intravenous transplantation

(Figures S5A and S5B). As with the cell lines, ex vivo coating of

these primary cells with anti-CD47 antibody, but not controls

(Figure 3G), resulted in complete inhibition of bone marrow

engraftment (Figure 3H).

Combination Therapy with Anti-CD47 Antibody andRituximab Eliminates Lymphoma in Both Disseminatedand Localized Human NHL Xenotransplant ModelsIn the second treatment strategy, mice were first engrafted with

NHL and then administered single or combination antibody

therapy. To best model NHL, we established disseminated and

localized xenotransplantation models in NSG mice that are defi-

cient in T, B, and NK cells (Shultz et al., 2005) but retain phago-

cyte effector cells. In the disseminated model, luciferase-

expressing Raji cells were transplanted intravenously into adult

NSG mice. Two weeks later, these mice were administered daily

injections of either control mouse IgG, anti-CD47 antibody, ritux-

imab, or anti-CD47 antibody and rituximab. Anti-CD47 antibody

treatment decreased the lymphoma burden in these mice (Fig-

ures 4A and 4B) and significantly prolonged survival compared

to control IgG, although all mice eventually died (Figure 4C and

Table S2). Similar results were seen with rituximab and were

not statistically different compared to anti-CD47 antibody (Fig-

ures 4A–4C and Table S2). In contrast, combination therapy

with anti-CD47 antibody and rituximab eliminated lymphoma in

60% of mice as indicated by long-term survival (Figure 4C) and

the absence of luciferase-positive lymphoma (data not shown)

more than 4 months after the end of treatment. In humans, ritux-

imab efficacy is thought to be primarily mediated by both macro-

phages and NK cells (Nimmerjahn and Ravetch, 2007; Taylor

and Lindorfer, 2008). Given that NSG mice lack NK cells, we con-

ducted a similar experiment in NK cell-containing SCID mice and

observed similar therapeutic responses as in NSG mice (Figures

S4A and S4B).

In the localized NHL model, luciferase-expressing Raji cells

were transplanted subcutaneously in the right flank of NSG

mice. Once tumors were palpable (approximately 2 weeks),

mice were treated with antibody therapy. Mice treated with

anti-CD47 antibody or rituximab demonstrated a decreased

rate of lymphoma growth compared to control IgG-treated

mice as measured by both luciferase signal and tumor volume

(Figures 4D–4F) but, like controls, eventually had to be sacrificed

due to enlarged tumor growth. In contrast, mice treated with the

combination of anti-CD47 antibody and rituximab demonstrated

complete elimination of lymphoma, with 86% of treated mice

having no measurable mass or luciferase-positive lymphoma

4 weeks after the end of therapy (Figures 4D–4F and Figures

S4C–S4E). Moreover, all showed no evidence of tumor growth,

(C) Synergistic phagocytosis by anti-CD47 antibody (B6H12.2) and anti-CD20 mAbs was examined by isobologram analysis and determination of combination

indices (CI). CIobs indicates observed results, and the dashed line indicates the expected results if antibodies were additive.

(D and E) Synergy between anti-CD47 antibody and rituximab in the phagocytosis of NHL and NPB cells was assessed by determining the phagocytic index when

incubated with a combination of both antibodies compared to either antibody alone at twice the dose, with either mouse (D) or human (E) macrophages. NHL17*:

cell line derived from primary sample NHL17.

***p < 0.001, ****p < 0.0001, *****p < 0.00001. Figure 2B p values represent comparison against IgG1 isotype control antibody. See also Figure S2.

Cell 142, 699–713, September 3, 2010 ª2010 Elsevier Inc. 703

Page 72: Cell 100903

Figure 3. Ex Vivo Coating of NHL Cells with an Anti-CD47 Antibody Inhibits Tumor Engraftment

(A–F) Luciferase-expressing Raji (A) and SUDHL4 (C) cells were assessed for ex vivo antibody coating by flow cytometry. SCID mice transplanted with Raji (B) and

SUDHL4 (D) were subject to bioluminescent imaging (1-IgG1 isotype control, 2-anti-CD45, 3 and 4-anti-CD47, 5-luciferase negative control). Bioluminescence

for Raji (E) and SUDHL4 (F) engrafted mice was quantified (n = 3 per antibody condition). No tumor engraftment was observed in mice transplanted with anti-

CD47-coated cells compared to IgG-coated cells (p < 0.05) for both Raji and SUDHL4, as assessed by bioluminescent imaging. Data are represented as

mean ± standard deviation (SD).

704 Cell 142, 699–713, September 3, 2010 ª2010 Elsevier Inc.

Page 73: Cell 100903

remained relapse free, and were alive at over 197 days after

tumor engraftment. Out of six mice achieving a complete remis-

sion, five remained relapse free whereas one mouse died of non-

tumor-related causes (Figure 4E). For both disseminated and

localized xenograft models, expression of CD47 and CD20 in

transplanted Raji cells did not differ from Raji cells in culture

(Figure S4F).

Combination Therapy with Anti-CD47 Antibodyand Rituximab Eliminates Lymphoma in Primary HumanNHL Xenotransplant Mouse ModelsNHL cell lines have been valuable for the evaluation of therapeu-

tics, but they may not accurately recapitulate the heterogeneity

of the primary disease. We report here two new mouse xenograft

models for NHL in which intravenous transplantation of cells

from a DLBCL patient (NHL7/SUNHL7) and a FL patient

(NHL31/SUNHL31) give rise to robust lymphoma engraftment

in the bone marrow and/or peripheral blood (Figure S1A and

Figure S5A). Primary DLBCL cells were transplanted into mice,

which 2 weeks later were treated with daily injections of anti-

bodies for 14 days. Treatment with anti-CD47 antibody either

alone or in combination with rituximab eliminated human

lymphoma in the bone marrow, whereas treatment with rituxi-

mab resulted in a reduction of disease in 60% of mice (Figures

5A and 5B). Mouse hematopoietic cells were unaffected by anti-

body therapy (Figure S5E). Treatment was then discontinued,

and all mice were monitored for survival. Mice treated with either

anti-CD47 antibody or rituximab alone had a significantly longer

survival compared to mice treated with control IgG, but all even-

tually died secondary to disease due to widespread organ

dissemination on autopsy (Figure 5C, data not shown). Most

significantly, 8 out of 9 mice (89%) administered combination

antibody treatment were cured of lymphoma, as indicated by

long-term disease-free survival more than 4 months after the

end of treatment (Figure 5C and Table S3) with no detectable

lymphoma in the bone marrow (data not shown). In a second

primary model, FL cells were transplanted intravenously in a

similar manner. CD20+CD10+ lymphoma engraftment in the

peripheral blood and bone marrow was detected after 8 weeks.

At this time, mice were treated with a single intraperitoneal injec-

tion of either control IgG, anti-CD47 antibody, rituximab, or the

combination and then followed for disease progression. A single

dose of anti-CD47 antibody alone or in combination with rituxi-

mab eliminated lymphoma both in the peripheral blood (Fig-

ure 5D) and in bone marrow (Figure 5E). In contrast, a single

dose of rituximab enabled a partial reduction in tumor burden

that rebounded back to baseline levels in the peripheral blood

with no tumor reduction observed in the bone marrow. The differ-

ence in anti-CD47 antibody clearance of lymphoma as a single

dose in FL-engrafted mice (Figures 5D and 5E) compared to

multiple dose therapy in mice engrafted with DLBCL (Fig-

ure 5B) or Raji (Figure 4) may be due to cell-intrinsic differences

in antibody sensitivity between different NHL subtypes or due to

different anti-CD47 antibody potencies in distinct tissue com-

partments (peripheral blood versus bone marrow (BM) versus

soft tissue compartments).

To assess the ability of these two primary NHL xenotransplant

models to model the disease, we compared histological sections

of the primary NHL specimen and the transplanted tumor. Similar

DLBCL and FL morphology was observed for NHL7 and NHL31,

respectively (Figure S5B). We next determined whether the per-

centages of macrophages infiltrating the tumor differed between

the primary patient and the xenografted tumor. For NHL31, the

percentage of infiltrating human macrophages (CD68+) in the

primary lymph node was similar to the percentage of infiltrating

mouse macrophages (F4/80+) in bone marrow of transplanted

mice (Figure S5C). Analyzing infiltrating macrophage frequency

by flow cytometry, no difference was observed between the

primary specimen and xenograft for either NHL7 or NHL31

(Figure S5D).

Synergy between Anti-CD47 Antibody and RituximabDoes Not Occur through NK Cells or ComplementRituximab can eliminate malignant cells via apoptosis, NK cell-

mediated ADCC, and CDC (reviewed in Smith, 2003). However,

it is not known whether anti-CD47 antibody also enables ADCC

or CDC in addition to phagocytosis. Therefore, we investigated

whether anti-CD47 antibody alone could induce antitumor

effects by macrophage-independent mechanisms, and whether

synergy with rituximab could occur through these modalities.

First, to investigate possible synergy in direct apoptosis, NHL

cells were incubated with either anti-CD47 antibody/rituximab

alone or in combination without macrophages, and cell death

was measured. No synergistic apoptosis was observed when

NHL cells were incubated with soluble (Figure 6A and Fig-

ure S6A) or immobilized (Figures S6B and S6C) antibodies.

Furthermore, crosslinking of soluble anti-CD47 antibody alone

or in combination with rituximab by macrophages did not induce

increased apoptosis of nonphagocytosed NHL cells compared

to IgG1 isotype control, whereas rituximab induced a slight

increase in apoptosis (Figure S6D). No synergistic apoptosis

was observed in this context (Figure S6D). The small increase

in apoptosis upon antibody treatment was not FcR dependent

given that results were similar with macrophages lacking the

Fcg subunit (Takai et al., 1994) (Figure S6D). Second, we inves-

tigated whether NK cells could mediate tumor elimination by

anti-CD47 antibody alone or in synergy with rituximab. A prior

report observed increased NK cell cytotoxicity of cancer cell

lines with an anti-CD47 antibody in vitro, though the mechanism

of targeting was not elucidated (Kim et al., 2008). We first deter-

mined whether human or mouse NK cells expressed SIRPa and

found that both human NK cells, CD3�CD56+CD7+ (Milush et al.,

2009), as well as mouse NK cells, CD3�DX5+, expressed minimal

to no SIRPa (Figure 6B). Next, the ability of anti-CD47 antibody

to induce NK cell-mediated ADCC through FcRs or by CD47-

SIRPa blockade was investigated. Utilizing human NK cells as

(G) Bulk lymphoma cells from a human NHL patient were assessed for ex vivo antibody coating by flow cytometry.

(H) Compared to IgG1 isotype control, anti-CD47 antibody pretreatment inhibited engraftment of NHL cells (p < 0.0001) whereas anti-CD45-coated cells

engrafted similarly to controls (p = 0.54). All p values were determined using Fisher’s exact test.

Error bars represent SD (E and F). See also Figure S3.

Cell 142, 699–713, September 3, 2010 ª2010 Elsevier Inc. 705

Page 74: Cell 100903

Figure 4. Combination Therapy with Anti-CD47 Antibody and Rituximab Eliminates Lymphoma in Both Disseminated and Localized Human

NHL Xenotransplant Mouse Models

(A) NSG mice transplanted intravenously with luciferase-expressing Raji cells were treated with the indicated antibodies (n = 8 per treatment group). Luciferase

imaging of representative mice from pre- and 10 days post-treatment are shown (A) and averaged for all mice in each treatment group (B).

(C) Kaplan-Meier survival analysis was performed (Table S2). p values compare IgG control to combination antibody treatment or anti-CD47 antibody/rituximab

single antibody to combination. Arrows indicate start (day 14) and stop (day 35) of treatment.

(D) Luciferase-expressing Raji cells were transplanted subcutaneously in the flank of NSG mice. When palpable tumors (�0.1 cm3) formed, treatment

began with the indicated antibodies. Luciferase imaging of representative mice from each treatment group is shown before (day 0) and during (day 14)

treatment.

(E) Quantified bioluminescence was determined and averaged for all mice in each treatment group (n = 7).

(F) Tumor volume was measured with average volume shown. p values were derived from a two-way ANOVA and compared to IgG treatment.

*p < 0.05, ***p < 0.001, ****p < 0.0001. Complete remission was defined as the number of mice with no evidence of tumor at the indicated date. Relapse was

defined as the number of mice achieving a complete remission that later developed recurrence of tumor growth. For (E), one mouse that achieved a complete

remission died of non-tumor-related causes. Data presented in (B), (E), and (F) are mean values ± SD. See also Figure S4 and Table S2.

706 Cell 142, 699–713, September 3, 2010 ª2010 Elsevier Inc.

Page 75: Cell 100903

Figure 5. Combination Therapy with Anti-CD47 Antibody and Rituximab Eliminates Lymphoma in Primary Human NHL Xenotransplant

Mouse Models(A and B) Sublethally irradiated NSG mice were transplanted intravenously with cells from a primary DLBCL patient (NHL7) and treated with the indicated anti-

bodies. Flow cytometry of human lymphoma engraftment in the bone marrow of two representative mice is shown pre- and 14 days post-antibody treatment in

(A). Data from all mice are included in (B). **p < 0.01, comparing pre- and post-treatment values for each respective antibody treatment.

(C) Kaplan-Meier survival analysis (Table S3) of DLBCL-engrafted mice from each antibody treatment cohort is shown (n R 10 per antibody group), with p values

calculated comparing control IgG to combination antibody treatment or anti-CD47 antibody/rituximab single antibody to combination treatment. Arrows indicate

start (day14) and stop (day 28) of treatment.

(D and E) Mice engrafted intravenously with a primary FL patient sample (NHL31) were treated with a single dose of the indicated antibodies. Compared to IgG

control and rituximab, anti-CD47 antibody alone or in combination with rituximab eliminated tumor burden in the peripheral blood (p = 0.04, two-way ANOVA) and

bone marrow (p < 0.001, t test).

(E) Lyphoma engraftment in the bone marrow was determined 14 days post-treatment. Each antibody treatment group consisted of three mice.

For mice reported in (D) and (E), human lymphoma chimerism was between 5% and 25% in the peripheral blood and bone marrow. Error bars represent SD

(D and E). See also Figure S5 and Table S3.

Cell 142, 699–713, September 3, 2010 ª2010 Elsevier Inc. 707

Page 76: Cell 100903

Figure 6. Synergy between Anti-CD47 Antibody and Rituximab Does Not Occur through NK Cells or Complement

(A) NHL cells were incubated with the indicated soluble antibodies for 2 hr and the percentage of dead cells was calculated (% annexin V+ and/or 7-AAD+). No

statistically significant difference in % dead cells was observed with the combination of anti-CD47 antibody and rituximab compared to either anti-CD47 antibody

alone (p = 0.24) or rituximab alone (p = 0.95).

(B) SIRPa expression is shown for both human and mouse NK cells as determined by flow cytometry.

(C and D) Chromium release assays measuring ADCC were performed in triplicate with human (C) and mouse (D) at an effector:target ratio of 17.5:1, and percent-

specific lysis is reported. Antibodies were incubated at 10 mg/ml except anti-CD47 full-length or F(ab0)2 antibody+rituximab (5 mg/ml).

(E) CDC assay with human complement was performed in duplicate. Compared to IgG1 isotype control, anti-CD47 antibody did not enable CDC (p > 0.2),

whereas rituximab did (p < 0.001) by two-way ANOVA for both SUDHL4 and NHL17*. Combination treatment with anti-CD47 antibody and rituximab did not

enable greater levels of CDC compared to rituximab (p = 0.78).

(F) CDC assay with mouse complement was performed in duplicate. Compared to IgG1 isotype control, anti-CD47 antibody did not enable CDC (p > 0.25)

whereas rituximab did (p = 0.03, p = 0.08, respectively) for both SUDHL4 and NHL17*. No difference in CDC between CD47 antibody+rituximab and rituximab

alone was observed (p > 0.13) for both SUDHL4 and NHL17*.

*p < 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001. Error bars represent SD (C–F). NHL17* = primary NHL17 cells expanded in culture. See also Figure S6.

708 Cell 142, 699–713, September 3, 2010 ª2010 Elsevier Inc.

Page 77: Cell 100903

effectors, anti-CD47 antibody did not induce increased ADCC of

Raji or SUDHL4 cells compared to IgG1 isotype control (Fig-

ure 6C). Although rituximab did enable ADCC of these two

NHL cell lines, no synergistic effect between anti-CD47 antibody

and rituximab was observed (p = 0.77, Figure 6C). As anti-CD47

antibody (B6H12.2) is a mouse IgG1 isotype, we repeated these

assays with mouse NK cells. Anti-CD47 antibody caused

increased ADCC of these two NHL cell lines compared to isotype

control, whereas rituximab induced ADCC to a lesser degree

(Figure 6D). To test whether anti-CD47 antibody-mediated

ADCC was Fc dependent, we generated a F(ab0)2 fragment of

the anti-CD47 antibody (Figures S7B–S7I). The F(ab0)2 fragment

did not enable ADCC, indicating that the increased ADCC was

likely mediated through FcRs (Figure 6D). The combination of

anti-CD47 antibody or F(ab0)2 fragment with rituximab did not

induce a statistically significant increase in ADCC compared to

single agent therapy, indicating no synergistic effect. Third, we

investigated the role of complement in anti-CD47 antibody-

mediated cytotoxicity. Using either human (Figure 6E) or mouse

(Figure 6F) complement, anti-CD47 antibody did not induce CDC

of either an NHL cell line or a primary NHL sample, whereas

rituximab did induce significant CDC against both of these

samples. Moreover, the combination of anti-CD47 antibody

and rituximab did not induce increased CDC compared to ritux-

imab alone. Fourth, we investigated the relative contribution of

major components of macrophages, NK cells, and complement

in mediating the therapeutic effects of anti-CD47 antibody and

rituximab in vivo. Luciferase-labeled Raji cells were engrafted

intravenously into SCID mice, which have functional macro-

phages, NK cells, and complement. Mice were then separated

into cohorts receiving selective depletion of either macrophages

by clodronate, NK cells by anti-asialoGM1 antibody, comple-

ment by cobra venom factor, or a vehicle control. These cohorts

were then treated with combination anti-CD47 antibody and rit-

uximab therapy for 3 days, and tumor burden was measured

by bioluminescent imaging pre- and post-treatment. Compared

to vehicle control, NK cell and complement depletion had no

effect on tumor elimination by combination antibody therapy

(Figure S6E). In contrast, macrophage depletion significantly

abrogated the therapeutic effect, indicating that macrophages,

and not NK cells or complement, are required for combination

anti-CD47 antibody and rituximab-mediated elimination of NHL

in vivo.

Anti-CD47 Antibody Synergizes with Rituximab throughFcR-Independent and FcR-Dependent MechanismsWe hypothesize that the observed synergy between an anti-

CD47 antibody and rituximab occurs through the combination

of two separate mechanisms for stimulating phagocytosis: (1)

FcR-independent through blockade of an inhibitory phagocytic

signal by anti-CD47 antibody, and (2) FcR-dependent through

delivery of a positive phagocytic signal by rituximab. We utilized

four independent methods to investigate this hypothesis. First,

synergistic phagocytosis was observed with the combination

of anti-SIRPa antibody and rituximab by isobologram analysis

(Figure 7A), and with a large panel of primary NHL samples (Fig-

ure 7B). Second, mouse macrophages lacking the Fcg receptor,

thereby incapable of enabling FcR-dependent phagocytosis

(Takai et al., 1994), but still expressing SIRPa (Figure S7A),

were utilized as effector cells for phagocytosis of NHL cells incu-

bated with either anti-CD47 antibody, anti-SIRPa antibody, ritux-

imab, or anti-CD47/anti-SIRPa antibody in combination with

rituximab. As predicted, anti-CD47 antibody and anti-SIRPa

antibody, but not rituximab, enabled phagocytosis of NHL cells,

without evidence of synergistic phagocytosis (Figure 7C).

Third, F(ab0)2 fragments of both anti-CD47 antibody and

rituximab were generated (Figures S7B–S7I) and utilized in

phagocytosis assays with NHL cells and wild-type macro-

phages. Anti-CD47 F(ab0)2 antibody synergized with rituximab

as demonstrated by isobologram analysis (Figure 7D). Addition-

ally, anti-CD47 F(ab0)2, but not rituximab F(ab0)2, enabled phago-

cytosis of NHL cells (Figure 7E). Consistent with the proposed

mechanism, synergistic phagocytosis was observed with the

combination of either full-length anti-CD47 or anti-CD47 F(ab0)2with full-length rituximab, but not with any combination involving

rituximab F(ab0)2 (Figure 7E).

Fourth, synergistic phagocytosis was investigated in vivo

using GFP+ Raji cells engrafted in NSG mice. As single agents,

anti-CD47 antibody and rituximab enabled phagocytosis of

Raji cells engrafted in the liver as evidenced by an increased

percentage of mouse macrophages containing phagocytosed

GFP+ Raji cells (Figure 7F). Most significantly, combination

anti-CD47 antibody and rituximab treatment enabled signifi-

cantly increased phagocytosis compared to either single agent

demonstrating that synergistic phagocytosis occurred in vivo

(Figure 7F).

DISCUSSION

In this report, we identify a distinct mechanism of synergy

between mAbs in cancer therapy leading to cures in the absence

of chemotherapy. Specifically, we utilized a blocking anti-CD47

antibody in combination with the anti-CD20 antibody rituximab

to eradicate human NHL through a mechanism of synergy

involving FcR-independent enabling of phagocytosis by anti-

CD47 antibody and FcR-dependent stimulation of phagocytosis

by rituximab. In addition, the identification of CD47 expression

as a prognostic factor could be incorporated into standard clin-

ical prognostic considerations across multiple subtypes of NHL

and may be useful in risk-adapted therapy decision-making.

Although it is thought that many therapeutic mAbs for human

malignancies, including rituximab, function primarily through

NK cell-mediated ADCC, several lines of evidence indicate that

the therapeutic effect of anti-CD47 antibody alone or in combi-

nation with rituximab is mediated primarily through macrophage

phagocytosis. First, synergistic macrophage phagocytosis was

observed with combination anti-CD47 antibody and rituximab

in vitro, whereas no synergy was observed for direct apoptosis,

ADCC, or CDC (Figures 2C–2E, Figure S6, Figure 6A, and Figures

6C–6F). Second, phagocytosis of NHL cells in vivo was observed

with either anti-CD47 antibody or rituximab alone and, most

importantly, significantly increased with combination therapy

(Figure 7F). Third, the therapeutic effect of combination antibody

treatment was similar in an NHL xenotransplant model using

complement and NK cell-deficient NSG mice (Figure 4C) as in

complement and NK cell-competent SCID mice (Figures S4A

Cell 142, 699–713, September 3, 2010 ª2010 Elsevier Inc. 709

Page 78: Cell 100903

Figure 7. Anti-CD47 Antibody Synergizes with Rituximab through FcR-Independent and FcR-Dependent Mechanisms

(A) Isobologram analysis of phagocytosis induced by anti-SIRPa antibody and rituximab is shown for Raji cells and mouse macrophages.

(B and C) NHL cells were incubated in vitro with the indicated antibodies in the presence of wild-type (B) or Fcgr�/� (C) mouse macrophages, and the phagocytic

index was determined.

(D) Isobologram analysis of phagocytosis induced by anti-CD47 F(ab0 )2 antibody and rituximab is shown for Raji cells and mouse macrophages.

(E) NHL cells were incubated with wild-type mouse macrophages in the presence of the indicated full-length or F(ab0)2 antibodies (single antibodies at 10 mg/ml,

combination antibodies at 5 mg/ml each) and the phagocytic index was determined.

(F) The level of in vivo phagocytosis measured as the percentage of mouse macrophages containing phagocytosed GFP+ Raji cells (hCD45�GFP+F4/80+) was

determined by flow cytometry of livers from mice engrafted with GFP+ Raji cells and then treated with the indicated antibodies (see Experimental Procedures),

with each treatment group performed in duplicate. Compared to IgG control, anti-CD47 antibody and rituximab enabled increased levels of phagocytosis.

Compared to anti-CD47 antibody alone, combination anti-CD47 antibody and rituximab enabled higher levels of phagocytosis.

*p < 0.05, **p < 0.01, ****p < 0.0001. Error bars represent SD (E and F). See also Figure S7.

710 Cell 142, 699–713, September 3, 2010 ª2010 Elsevier Inc.

Page 79: Cell 100903

and S4B), suggesting that macrophages alone are sufficient to

mediate the therapeutic effect. Fourth, depletion of macro-

phages, but not complement or NK cells, abrogated the syner-

gistic effect of anti-CD47 antibody in combination with rituximab

(Figure S6E). These studies highlight the importance of macro-

phages as effectors of anti-CD47 antibody therapy in human

NHL.

This study describes a mechanism of antibody synergy in the

elimination of NHL in the absence of chemotherapy. Combina-

tion antibody therapies for NHL have previously been investi-

gated, mostly in combination with rituximab, with some pro-

gressing to clinical trials. These include a humanized antibody

targeting the B cell antigen CD22 (epratuzumab) and galiximab,

a chimeric antibody targeting the costimulatory ligand CD80.

Phase I/II studies with either epratuzumab or galiximab in combi-

nation with rituximab demonstrate relative safety and clinical

responses equal to or greater than single agent therapy alone

(Leonard et al., 2005, 2007, 2008). Based on these results, phase

III trials are underway. Antibody combinations involving anti-

CD20 antibodies and antibodies to proapoptotic receptors are

also being explored preclinically (Daniel et al., 2007; Maddipatla

et al., 2007). These studies highlight the clinical potential of

combination antibody approaches in NHL patients.

Combination therapy with two or more mAbs possesses

several advantages compared to monotherapies in NHL or other

malignancies. First, therapy solely with monoclonal antibodies

targeting cancer-specific antigens could result in decreased

off-target toxicity compared to current therapeutic regimens

that utilize chemotherapy. Second, synergy between two distinct

antibody effector mechanisms, FcR-independent and FcR-

dependent as shown here, would result in increased therapeutic

efficacy. Third, antibody targeting of two distinct cell-surface

antigens would be more likely to eliminate cancer cells with

pre-existing epitope variants or epitope loss, such as those

reported in rituximab-refractory/resistant NHL patients (Foran

et al., 2001; Hiraga et al., 2009; Kennedy et al., 2004). Fourth,

a bispecific FcR-engaging antibody with one arm binding and

blocking CD47 and the other binding to a validated cancer anti-

body target (CD20) could reduce potential antibody toxicity,

while retaining the synergy effect, especially as CD47 is

expressed in multiple normal tissue types. Although we demon-

strated that an anti-mouse CD47 antibody is relatively nontoxic

to wild-type mice (Majeti et al., 2009), a clinical anti-human

CD47 antibody may have a different human toxicity profile that

could be overcome by a bispecific antibody.

In addition to its application in NHL, the reported mechanism

of antibody synergy provides proof-of-principle that a blocking

mAb directed against CD47 can synergize with an FcR-acti-

vating antibody to provide superior therapeutic efficacy. This

finding raises the possibility of potential synergy between an

anti-CD47 antibody and other clinically approved therapeutic

antibodies that may activate FcRs on immune effector cells for

the treatment of diverse human malignancies, including trastu-

zumab (Herceptin) for HER2-positive breast carcinomas, cetux-

imab (Erbitux) for colorectal carcinomas and head and neck

squamous cell carcinomas, alemtuzumab (Campath) for CLL

and T cell lymphoma, and others in development (Finn, 2008).

To date, we have demonstrated effective anti-CD47 antibody

targeting of several human cancers including AML (Majeti

et al., 2009), bladder cancer (Chan et al., 2009), and now NHL,

leading us to speculate that CD47 targeting will be effective

against a wide range of human cancers.

EXPERIMENTAL PROCEDURES

Cell Lines

A Burkitt’s lymphoma cell line (Raji) and a DLBCL cell line (SUDHL4) were

obtained from the American Type Culture Collection or generated in the lab.

The NHL17* cell line was generated from a patient with DLBCL by culturing

bulk cells in vitro with IMDM supplemented with 10% human AB serum for

1.5 months.

Human Samples

Normal human peripheral blood and human NHL samples were obtained from

the Stanford University Medical Center (Stanford, CA, USA) with informed

consent, according to an IRB-approved protocol (Stanford IRB# 13500) or

with informed consent from the Norwegian Radium Hospital (Oslo, Norway)

according to a Regional Ethic Committee (REK)-approved protocol (REK#

2.2007.2949). Normal tonsils for germinal center B cell analysis were obtained

from discarded tonsillectomy specimens from consented pediatric patients at

Stanford University Medical Center according to an IRB-approved protocol

(Stanford IRB# 13500).

Flow Cytometry Analysis

For analysis of normal peripheral blood cells, germinal center B cells, and

primary NHL cells, the following antibodies were used: CD19, CD20, CD3,

CD10, CD45, CD5, CD38 (Invitrogen, Carlsbad, CA, USA and BD Biosciences,

San Jose, CA, USA). Analysis of CD47 expression was performed with an anti-

human CD47 FITC antibody (clone B6H12.2, BD Biosciences). Cell staining

and flow cytometry analysis was performed as previously described (Majeti

et al., 2009).

Evaluation of Prognostic Value of CD47 in NHL

Gene expression and clinical data were analyzed for eight previously

described cohorts of adult NHL patients, including four studies of patients

with DLBCL, three with B-CLL, and one with MCL (detailed in Table S1). The

clinical end points analyzed included overall (OS), progression free (PFS),

and event-free survival (EFS), with events defined as the interval between

study enrollment and need for therapy or death from any cause, with data

censored for patients who did not have an event at the last follow-up visit.

See Extended Experimental Procedures for a detailed description of the

analyses.

Therapeutic Antibodies

Rituximab (anti-CD20, human IgG1) was obtained from the Stanford University

Medical Center, mouse anti-human CD20, IgG2a from Beckman Coulter

(Miami, FL, USA), and anti-CD47 antibody BRIC126, IgG2b from AbD Serotec

(Raleigh, NC, USA). Other anti-CD47 antibodies were used as in Majeti et al.

(2009). All in vivo antibody experiments were performed using the anti-CD47

B6H12.2 antibody.

In Vitro Isobologram Studies

In vitro phagocytosis assays were conducted with NHL cells incubated with

anti-CD47 antibody (B6H12.2), anti-CD20 IgG2a, or rituximab either alone or

in combination at concentrations from 1 mg/ml to 10 mg/ml. The concentration

of each antibody required to produce a defined single-agent effect (phagocytic

index) was determined for each cell type. Concentrations of the two antibodies

combined to achieve this same phagocytic index were then plotted on an

isobologram and the combination index (CI) determined. The CI was calcu-

lated from the formula CI = (d1/D1) + (d2/D2), whereby d1 = dose of drug

1 in combination to achieve the phagocytic index, d2 = dose of drug 2 in

combination to achieve the phagocytic index, D1 = dose of drug 1 alone to

achieve the phagocytic index, D2 = dose of drug 2 alone to achieve the

Cell 142, 699–713, September 3, 2010 ª2010 Elsevier Inc. 711

Page 80: Cell 100903

phagocytic index. A CI of less than, equal to, and greater than 1 indicates

synergy, additivity, and antagonism, respectively.

Annexin V Apoptosis Assays

Assays were performed as previously described (Majeti et al., 2009).

Preparation of Human and Mouse Immune Effector Cells, Immune

Effector Cytotoxicity Assays, and In Vivo Depletion of Immune

Effector Cells

See Extended Experimental Procedures.

Preparation of F(ab0)2 Fragments

See Extended Experimental Procedures.

Generation of Luciferase-Positive Cell Lines and Luciferase Imaging

Analysis

See Extended Experimental Procedures.

In Vivo Precoating Engraftment Assay

Assay were performed as previously described (Majeti et al., 2009). Precoated

cells were transplanted intravenously into SCID mice or sublethally irradiated

(200 rads) NOD.Cg-PrkdcscidIl2rgtm1Wjl/SzJ (NSG). All experiments involving

mice were performed according to Stanford University institutional animal

guidelines.

In Vivo Antibody Treatment in a Disseminated Lymphoma

Xenograft Model

1.5 3 106 luciferase-labeled Raji cells were injected intravenously into the

retro-orbital sinus of 6- to 10-week-old SCID or NSG mice. Those mice with

luciferase-positive lymphoma were given daily intraperitoneal injections of

200 mg mouse IgG control, anti-CD47 antibody, rituximab, or 200 mg anti-

CD47 antibody + 200 mg rituximab for 3 weeks. Antibody treatment was

then stopped and mice were followed for survival analysis. A complete remis-

sion (CR) was defined as no evidence of lymphoma by bioluminescence at the

end of treatment. A relapse was defined as evidence of lymphoma by biolumi-

nescence after the end of treatment in a mouse with a prior CR.

In Vivo Antibody Treatment in a Localized Lymphoma

Xenograft Model

3 3 106 luciferase-labeled Raji cells were injected subcutaneously into the

right flank of 6- to 10-week-old NSG mice. Those mice with luciferase-positive

lymphoma were given daily intraperitoneal injections of 400 mg mouse IgG

control, 400 mg anti-CD47, 200 mg rituximab, or 400 mg anti-CD47 + 200 mg rit-

uximab for 4 weeks. Tumor volume was measured every 3–4 days using the

formula (length*width)/2. Antibody treatment was then stopped and mice

were followed for survival analysis.

In Vivo Antibody Treatment of Primary NHL-Engrafted Mice

2 3 106 NHL cells were transplanted intravenously via retro-orbital plexus into

sublethally irradiated NSG mice. Two to ten weeks later, bone marrow was

aspirated from these mice and those mice with evidence of human lymphoma

engraftment (hCD45+CD19/CD10+ bone marrow cells) were then treated with

the same antibody regimen as in the disseminated lymphoma model. After

14 days, bone marrow cells from these mice were aspirated and antibody

treatment was stopped and mice followed for survival analysis. A CR was

defined as no evidence of lymphoma in the BM at end of treatment. A relapse

was defined as evidence of lymphoma in the BM after end of treatment in

a mouse with a prior CR.

In Vivo Phagocytosis

In vivo phagocytosis was performed as previously described (Majeti et al.,

2009) analyzing mice transplanted with GFP+ Raji cells into adult NSG mice.

Mice were given a single dose of antibody and analyzed 4 hr later for in vivo

phagocytosis.

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures, seven

figures, and three tables and can be found with this article online at doi:10.

1016/j.cell.2010.07.044.

ACKNOWLEDGMENTS

The authors acknowledge Dr. Christopher Contag for providing luciferase

constructs, Dr. Robert Negrin for assistance, Dr. Yaso Natkunam for immuno-

histochemistry, Libuse Jerabek and Theresa Storm for lab management, and

Adriane Mosely for animal husbandry. We also acknowledge the patients and

surgeons including Drs. Wapnir, Chang, Cheng, Janisiewicz, Koltai, Liebowitz,

and Messner for providing research specimens. M.P.C. is supported by the

HHMI and the Stanford Cancer Biology Program, A.A.A. by an NIH T32 Ruth

L. Kirschstein National Research Service Award (HL007970), R. Malumbres

by a fellowship from Fundacion Caja Madrid, I.S.L. by NIH grant CA122105,

and R. Majeti by a grant from the AACR. R. Majeti holds a Career Award for

Medical Scientists from the Burroughs Wellcome Fund. I.L.W., M.P.C.,

A.A.A., and R. Majeti have filed U.S. Patent Application Serial No. 12/321,215

entitled ‘‘Methods For Manipulating Phagocytosis Mediated by CD47.’’ This

research is supported by NIH grant P01CA139490 to I.L.W. and funding from

the Smith Family Fund.

M.P.C., A.A.A., I.L.W., and R. Majeti designed the experiments, and M.P.C.,

A.A.A., I.L.W., and R. Majeti wrote the manuscript. M.P.C., A.A.A., C.T.,

J.H.M., S.G., M.J., A.C.C., C.K.C., B.T.T., C.Y.P, F.Z., R. Malumbres, and

I.S.L. performed the experiments and analyzed data. J.B., R.D.G., R.L., and

I.S.L provided patient samples and clinical data. H.E.K. provided reagents.

B.V. generated F(ab0)2 fragments of anti-CD47 antibody and rituximab. All

authors endorse the full content of this work.

Received: January 7, 2010

Revised: April 23, 2010

Accepted: July 6, 2010

Published: September 2, 2010

REFERENCES

Adams, G.P., and Weiner, L.M. (2005). Monoclonal antibody therapy of cancer.

Nat. Biotechnol. 23, 1147–1157.

Alizadeh, A.A., Eisen, M.B., Davis, R.E., Ma, C., Lossos, I.S., Rosenwald, A.,

Boldrick, J.C., Sabet, H., Tran, T., Yu, X., et al. (2000). Distinct types of diffuse

large B-cell lymphoma identified by gene expression profiling. Nature 403,

503–511.

Brown, E.J., and Frazier, W.A. (2001). Integrin-associated protein (CD47) and

its ligands. Trends Cell Biol. 11, 130–135.

Cartron, G., Watier, H., Golay, J., and Solal-Celigny, P. (2004). From the bench

to the bedside: ways to improve rituximab efficacy. Blood 104, 2635–2642.

Chan, K.S., Espinosa, I., Chao, M., Wong, D., Ailles, L., Diehn, M., Gill, H.,

Presti, J., Jr., Chang, H.Y., van de Rijn, M., et al. (2009). Identification, molec-

ular characterization, clinical prognosis, and therapeutic targeting of human

bladder tumor-initiating cells. Proc. Natl. Acad. Sci. USA 106, 14016–14021.

Cheson, B.D., and Leonard, J.P. (2008). Monoclonal antibody therapy for

B-cell non-Hodgkin’s lymphoma. N. Engl. J. Med. 359, 613–626.

Chou, T.C. (2008). Preclinical versus clinical drug combination studies. Leuk.

Lymphoma 49, 2059–2080.

Daniel, D., Yang, B., Lawrence, D.A., Totpal, K., Balter, I., Lee, W.P., Gogineni,

A., Cole, M.J., Yee, S.F., Ross, S., and Ashkenazi, A. (2007). Cooperation of the

proapoptotic receptor agonist rhApo2L/TRAIL with the CD20 antibody rituxi-

mab against non-Hodgkin lymphoma xenografts. Blood 110, 4037–4046.

Finn, O.J. (2008). Cancer immunology. N. Engl. J. Med. 358, 2704–2715.

Foran, J.M., Norton, A.J., Micallef, I.N., Taussig, D.C., Amess, J.A., Rohatiner,

A.Z., and Lister, T.A. (2001). Loss of CD20 expression following treatment with

rituximab (chimaeric monoclonal anti-CD20): a retrospective cohort analysis.

Br. J. Haematol. 114, 881–883.

712 Cell 142, 699–713, September 3, 2010 ª2010 Elsevier Inc.

Page 81: Cell 100903

Glennie, M.J., French, R.R., Cragg, M.S., and Taylor, R.P. (2007). Mechanisms

of killing by anti-CD20 monoclonal antibodies. Mol. Immunol. 44, 3823–3837.

Hiraga, J., Tomita, A., Sugimoto, T., Shimada, K., Ito, M., Nakamura, S., Kiyoi,

H., Kinoshita, T., and Naoe, T. (2009). Down-regulation of CD20 expression in

B-cell lymphoma cells after treatment with rituximab-containing combination

chemotherapies: its prevalence and clinical significance. Blood 113, 4885–

4893.

Jaiswal, S., Jamieson, C.H., Pang, W.W., Park, C.Y., Chao, M.P., Majeti, R.,

Traver, D., van Rooijen, N., and Weissman, I.L. (2009). CD47 is upregulated

on circulating hematopoietic stem cells and leukemia cells to avoid phagocy-

tosis. Cell 138, 271–285.

Jaiswal, S., Chao, M.P., Majeti, R., and Weissman, I.L. (2010). Macrophages

as mediators of tumor immunosurveillance. Trends Immunol. 31, 212–219.

Katzenberger, T., Petzoldt, C., Holler, S., Mader, U., Kalla, J., Adam, P., Ott,

M.M., Muller-Hermelink, H.K., Rosenwald, A., and Ott, G. (2006). The Ki67

proliferation index is a quantitative indicator of clinical risk in mantle cell

lymphoma. Blood 107, 3407.

Kennedy, A.D., Beum, P.V., Solga, M.D., DiLillo, D.J., Lindorfer, M.A., Hess,

C.E., Densmore, J.J., Williams, M.E., and Taylor, R.P. (2004). Rituximab infu-

sion promotes rapid complement depletion and acute CD20 loss in chronic

lymphocytic leukemia. J. Immunol. 172, 3280–3288.

Kikuchi, Y., Uno, S., Yoshimura, Y., Otabe, K., Iida, S., Oheda, M., Fukushima,

N., and Tsuchiya, M. (2004). A bivalent single-chain Fv fragment against CD47

induces apoptosis for leukemic cells. Biochem. Biophys. Res. Commun. 315,

912–918.

Kikuchi, Y., Uno, S., Kinoshita, Y., Yoshimura, Y., Iida, S., Wakahara, Y.,

Tsuchiya, M., Yamada-Okabe, H., and Fukushima, N. (2005). Apoptosis

inducing bivalent single-chain antibody fragments against CD47 showed anti-

tumor potency for multiple myeloma. Leuk. Res. 29, 445–450.

Kim, M.J., Lee, J.C., Lee, J.J., Kim, S., Lee, S.G., Park, S.W., Sung, M.W., and

Heo, D.S. (2008). Association of CD47 with natural killer cell-mediated cytotox-

icity of head-and-neck squamous cell carcinoma lines. Tumour Biol. 29,

28–34.

Leonard, J.P., Coleman, M., Ketas, J., Ashe, M., Fiore, J.M., Furman, R.R.,

Niesvizky, R., Shore, T., Chadburn, A., Horne, H., et al. (2005). Combination

antibody therapy with epratuzumab and rituximab in relapsed or refractory

non-Hodgkin’s lymphoma. J. Clin. Oncol. 23, 5044–5051.

Leonard, J.P., Friedberg, J.W., Younes, A., Fisher, D., Gordon, L.I., Moore, J.,

Czuczman, M., Miller, T., Stiff, P., Cheson, B.D., et al. (2007). A phase I/II study

of galiximab (an anti-CD80 monoclonal antibody) in combination with rituxi-

mab for relapsed or refractory, follicular lymphoma. Ann. Oncol. 18, 1216–

1223.

Leonard, J.P., Schuster, S.J., Emmanouilides, C., Couture, F., Teoh, N.,

Wegener, W.A., Coleman, M., and Goldenberg, D.M. (2008). Durable complete

responses from therapy with combined epratuzumab and rituximab: final

results from an international multicenter, phase 2 study in recurrent, indolent,

non-Hodgkin lymphoma. Cancer 113, 2714–2723.

Maddipatla, S., Hernandez-Ilizaliturri, F.J., Knight, J., and Czuczman, M.S.

(2007). Augmented antitumor activity against B-cell lymphoma by a combina-

tion of monoclonal antibodies targeting TRAIL-R1 and CD20. Clin. Cancer Res.

13, 4556–4564.

Majeti, R., Chao, M.P., Alizadeh, A.A., Pang, W.W., Jaiswal, S., Gibbs, K.D.,

Jr., van Rooijen, N., and Weissman, I.L. (2009). CD47 is an adverse prognostic

factor and therapeutic antibody target on human acute myeloid leukemia stem

cells. Cell 138, 286–299.

Mateo, V., Lagneaux, L., Bron, D., Biron, G., Armant, M., Delespesse, G., and

Sarfati, M. (1999). CD47 ligation induces caspase-independent cell death in

chronic lymphocytic leukemia. Nat. Med. 5, 1277–1284.

Milush, J.M., Long, B.R., Snyder-Cappione, J.E., Cappione, A.J., 3rd, York,

V.A., Ndhlovu, L.C., Lanier, L.L., Michaelsson, J., and Nixon, D.F. (2009). Func-

tionally distinct subsets of human NK cells and monocyte/DC-like cells identi-

fied by coexpression of CD56, CD7, and CD4. Blood 114, 4823–4831.

Nimmerjahn, F., and Ravetch, J.V. (2007). Antibodies, Fc receptors and

cancer. Curr. Opin. Immunol. 19, 239–245.

Oscier, D.G., Gardiner, A.C., Mould, S.J., Glide, S., Davis, Z.A., Ibbotson, R.E.,

Corcoran, M.M., Chapman, R.M., Thomas, P.W., Copplestone, J.A., et al.

(2002). Multivariate analysis of prognostic factors in CLL: clinical stage,

IGVH gene mutational status, and loss or mutation of the p53 gene are inde-

pendent prognostic factors. Blood 100, 1177–1184.

Rosenwald, A., Wright, G., Chan, W.C., Connors, J.M., Campo, E., Fisher, R.I.,

Gascoyne, R.D., Muller-Hermelink, H.K., Smeland, E.B., Giltnane, J.M., et al.

(2002). The use of molecular profiling to predict survival after chemotherapy

for diffuse large-B-cell lymphoma. N. Engl. J. Med. 346, 1937–1947.

Rosenwald, A., Wright, G., Wiestner, A., Chan, W.C., Connors, J.M., Campo,

E., Gascoyne, R.D., Grogan, T.M., Muller-Hermelink, H.K., Smeland, E.B.,

et al. (2003). The proliferation gene expression signature is a quantitative inte-

grator of oncogenic events that predicts survival in mantle cell lymphoma.

Cancer Cell 3, 185–197.

Shultz, L.D., Lyons, B.L., Burzenski, L.M., Gott, B., Chen, X., Chaleff, S., Kotb,

M., Gillies, S.D., King, M., Mangada, J., et al. (2005). Human lymphoid and

myeloid cell development in NOD/LtSz-scid IL2R gamma null mice engrafted

with mobilized human hemopoietic stem cells. J. Immunol. 174, 6477–6489.

Smith, M.R. (2003). Rituximab (monoclonal anti-CD20 antibody): mechanisms

of action and resistance. Oncogene 22, 7359–7368.

Takai, T., Li, M., Sylvestre, D., Clynes, R., and Ravetch, J.V. (1994). FcR

gamma chain deletion results in pleiotrophic effector cell defects. Cell 76,

519–529.

Tallarida, R.J. (2006). An overview of drug combination analysis with isobolo-

grams. J. Pharmacol. Exp. Ther. 319, 1–7.

Taylor, R.P., and Lindorfer, M.A. (2008). Immunotherapeutic mechanisms of

anti-CD20 monoclonal antibodies. Curr. Opin. Immunol. 20, 444–449.

The International Non-Hodgkin’s Lymphoma Prognostic Factors Project.

(1993). A predictive model for aggressive non-Hodgkin’s lymphoma. N. Engl.

J. Med. 329, 987–994.

Uno, S., Kinoshita, Y., Azuma, Y., Tsunenari, T., Yoshimura, Y., Iida, S., Kiku-

chi, Y., Yamada-Okabe, H., and Fukushima, N. (2007). Antitumor activity of

a monoclonal antibody against CD47 in xenograft models of human leukemia.

Oncol. Rep. 17, 1189–1194.

Cell 142, 699–713, September 3, 2010 ª2010 Elsevier Inc. 713

Page 82: Cell 100903

A C-Type Lectin Collaborates with aCD45 Phosphatase Homolog to FacilitateWest Nile Virus Infection of MosquitoesGong Cheng,1 Jonathan Cox,1 Penghua Wang,1 Manoj N. Krishnan,1 Jianfeng Dai,1 Feng Qian,2 John F. Anderson,3

and Erol Fikrig1,4,*1Section of Infectious Diseases2Section of Rheumatology

Department of Internal Medicine, Yale University School of Medicine, New Haven, CT 06520, USA3Department of Entomology, Connecticut Agricultural Experiment Station, New Haven, CT 06504, USA4Howard Hughes Medical Institute, Chevy Chase, MD 20815, USA

*Correspondence: [email protected] 10.1016/j.cell.2010.07.038

SUMMARY

West Nile virus (WNV) is the most commonarthropod-borne flavivirus in the United States;however, the vector ligand(s) that participate in infec-tion are not known. We now show that an Aedesaegypti C-type lectin, mosGCTL-1, is induced byWNV, interacts with WNV in a calcium-dependentmanner, and facilitates infection in vivo and in vitro.A mosquito homolog of human CD45 in A. aegypti,designated mosPTP-1, recruits mosGCTL-1 toenable viral attachment to cells and to enhance viralentry. In vivo experiments show that mosGCTL-1 andmosPTP-1 function as part of the same pathwayand are critical for WNV infection of mosquitoes.A similar phenomenon was also observed in Culexquinquefasciatus, a natural vector of WNV, furtherdemonstrating that these genes participate in WNVinfection. During the mosquito blood-feeding pro-cess, WNV infection was blocked in vivo withmosGCTL-1 antibodies. A molecular understandingof flaviviral-arthropod interactions may lead to strat-egies to control viral dissemination in nature.

INTRODUCTION

West Nile virus (WNV) is maintained in a bird-mosquito transmis-

sion cycle and has become the most common arthropod-borne

flavivirus in the United States. Humans, horses, and other nona-

vian vertebrates are incidental hosts (Monath and Heinz, 1996).

Infection in man can result in fever, meningitis, or encephalitis,

among other symptoms (Hubalek and Halouzka, 1999; Leis

et al., 2002). Approved human vaccines or therapeutics are not

available and preventive measures largely focus upon mosquito

control (Reisen and Brault, 2007).

The ability of different mosquito species to transmit WNV

varies widely. Culex spp. are the major vectors for WNV world-

wide (Nir et al., 1968; Turell et al., 2000). In the United States,

the most important vectors are Culex pipiens in the east, Culex

tarsalis in the midwest and west, and Culex quinquefasciatus in

the southeast (Hayes et al., 2005). WNV has also been isolated

from Aedes, Ochlerotatus, and Culisetam mosquitoes (http://

www.cdc.gov/ncidod/dvbid/westnile/mosquitoSpecies.htm).

Aedes aegypti, a member of the Culicinae subfamily, is a major

vector for numerous flaviviruses (Gould and Solomon, 2008).

A. aegypti is ideal for viral pathogenesis studies because these

mosquitoes are easy to cultivate and the genome has been char-

acterized (Gubler, 1998; Nene et al., 2007; Halstead, 2008).

A. aegypti is readily susceptibility to infection with WNV in

the laboratory, and the virus rapidly disseminates throughout

most of the mosquito after the blood meal. As with Culex spp.,

A. aegypti is a threat for the transmission of WNV to humans

(Vanlandingham et al., 2007).

C-type lectins are a group of carbohydrate-binding proteins

(Zelensky and Gready, 2005). Several members of this family

are highly expressed by immune cells, including monocytes,

macrophages, and dendritic cells (DCs), and play a central

role in activating host defenses (Robinson et al., 2006). Human

mannose-binding lectin (MBL) is a pattern recognition molecule

of the innate immune system that binds to sugars on the surface

of invading pathogens, leading to opsonization, phagocytosis,

and activation of the complement pathway (Neth et al., 2002).

In contrast, some C-type lectins are recruited to facilitate flavivi-

ral infection. In mammals, two membrane C-type lectins,

DC-SIGN (CD209) and L-SIGN (CD209L), interact with flavivi-

ruses via high mannose glycans on viral glycoproteins (Klimstra

et al., 2003) and are essential host cell factors exploited by

dengue virus (DENV) and WNV to invade immature DCs and

macrophages (Geijtenbeek et al., 2000; Soilleux et al., 2002;

Tassaneetrithep et al., 2003; Davis et al., 2006). Another

C-type lectin, the mannose receptor (MR), also interacts with

the DENV envelope protein and may enhance viral attachment

to phagocytes (Miller et al., 2008). A recent study identified

C-type lectin domain family 5, member A (CLEC5A), as

a DENV receptor. The association between CLEC5A and

DENV does not result in viral entry, but rather stimulates the

714 Cell 142, 714–725, September 3, 2010 ª2010 Elsevier Inc.

Page 83: Cell 100903

release of proinflammatory cytokines, potentially contributing to

the pathogenesis of dengue hemorrhagic fever (Chen et al.,

2008).

Protein tyrosine phosphatases (PTPs) remove phosphate

groups from phosphorylated tyrosine residues and play critical

roles in cell communication, shape, motility, proliferation, and

differentiation (Alonso et al., 2004; Mustelin et al., 2005). One

well-known PTP, protein tyrosine phosphatase receptor type C

(PTPRC, CD45), is important for thymocyte development and

T cell activation (Byth et al., 1996; Trowbridge and Thomas,

1994) and is expressed on all nucleated cells of hemopoietic

origin (Thomas, 1989). The association between MBL and

CD45 in immature T cells influences thymocyte development

(Baldwin and Ostergaard, 2001).

Figure 1. Effect of WNV Infection on mosGCTL-1 Expression

(A) Viral distribution in A. aegypti. WNV (1 3 103 M.I.D50) was inoculated into the mosquito thorax.

(B–E) mosGCTL-1 messenger RNA (mRNA) was induced by viral infection in whole A. aegypti (B), salivary glands (C), hemolymph (D), and midgut (E). Total RNA

was isolated from various tissues or whole mosquitoes at 5 time points after viral infection.

(A–E) The viral load and mosGCTL-1 mRNA levels were determined by Taqman RT-QPCR and normalized with A. aegypti actin (AAEL011197). WNV (1 3 103

M.I.D50) was microinjected into each mosquito. Data are shown as the mean ± standard error (SEM). The experiments were repeated three times.

(F) Immunoblot to detect mosGCTL-1 in WNV-infected mosquitoes. Three WNV infected or control mosquitoes were pooled and homogenized. The supernatant

was then isolated, separated by SDS-PAGE, and probed with rabbit mosGCTL-1 antisera. UI, uninfected mosquitoes; I, WNV-infected mosquitoes; dpi, days

postinfection. Fifty micrograms of protein from mosquito lysates was loaded into each lane.

See also Figure S1 and Table S1.

Cell 142, 714–725, September 3, 2010 ª2010 Elsevier Inc. 715

Page 84: Cell 100903

Figure 2. The Function of mosGCTL-1 in WNV Infection

(A and B) mosGCTL-1 silencing. The mock group was treated with the same amount of GFP dsRNA. mosGCTL-1- or GFP-dsRNA-treated mosquitoes were used

to isolate total RNA at several time points post dsRNA injection. mRNA levels were determined by SYBR Green RT-QPCR (A). mosGCTL-1-dsRNA-treated or

mock-treated mosquitoes were collected at 6 days after gene silencing. The supernatant was separated by SDS-PAGE and probed with rabbit mosGCTL-1

antisera (B).

(C and D) Silencing mosGCTL-1 impairs WNV (C), but not dengue virus (D), infection. The viral burden was examined at day 6 after infection. WNV or dengue virus

(10 M.I.D50) was used to challenge mosquitoes. The viral load was determined by Taqman RT-QPCR and normalized with A. aegypti actin. The result shown was

representative of four independent experiments.

(E) The role of the mosGCTL paralogues in WNV infection. The sample number was no less than 12 mosquitoes in each group. The viral burden was determined by

Taqman RT-QPCR and normalized with A. aegypti actin. *p < 0.05. The results were pooled from two independent experiments.

(F) WNV induces mosGCTL-1 homolog expression in C. quinquefasciatus. WNV-infected and mock mosquitoes were collected at 6 days after infection. Culex

mosGCTL-1 mRNA was determined by SYBR Green RT-QPCR and normalized with C. quinquefasciatus actin (CPIJ012570). The experiment was repeated three

times with similar results.

(G) Efficiency of Culex mosGCTL-1 RNA interference. The mock group was treated with the same amount of GFP dsRNA. mRNA levels were determined by SYBR

Green RT-QPCR.

(H) Silencing of Culex mosGCTL-1 decreases the WNV burden in C. quinquefasciatus. The viral burden was determined by Taqman RT-QPCR. The results were

combined from three independent experiments.

716 Cell 142, 714–725, September 3, 2010 ª2010 Elsevier Inc.

Page 85: Cell 100903

Genome-wide RNA interference (RNAi) screening studies

have revealed several hundred host factors that influence WNV

or DENV infection in human or Drosophila cell lines and have

identified cellular pathways that have a role in viral internaliza-

tion, replication, assembly, or secretion (Krishnan et al., 2008;

Sessions et al., 2009). However, the relationship between flavivi-

ruses and mosquitoes is not well understood. We now examine

WNV-mosquito interactions by using A. aegypti and Culex quin-

quefasciatus, characterize the expression profile of mosquito

homologs of human susceptibility proteins (Krishnan et al.,

2008) in response to viral infection, and identify a lectin-based

pathway that is critical for viral infection of mosquitoes.

RESULTS

A recent RNAi screening study characterized 283 human

proteins that facilitate WNV infection (Krishnan et al., 2008).

We have now identified 215 homologs of these genes in

A. aegypti. The expression of 32 genes was altered by the pres-

ence of WNV in mosquitoes, potentially suggesting a role in viral

infection of the arthropod (Table S1 available online). To examine

their function, we therefore silenced each of these 32 genes in

mosquitoes with RNAi and assessed the effect on viral load.

Knockdown of 13 genes significantly altered the viral burden in

the vector (Table S1 and Figure S1). One of these genes,

AAEL000563, exhibited the most dramatic reduction in viral

load after silencing and was selected as the target for further

investigation. AAEL000563 belongs to the A. aegypti galac-

tose-specific binding C-type lectin family, shares 26% amino

acid identity with human mannose-binding lectin, and was

designated as mosGCTL-1 (mosquito galactose-specific binding

C-type lectin).

mosGCTL-1 Expression Increases during WNV Infectionof A. aegyptiTo further investigate the relationship between WNV infection

and mosGCTL-1 expression, we determined the viral load and

mosGCTL-1 level in selected A. aegypti tissues after the inocula-

tion of WNV into female mosquitoes. WNV was readily detect-

able 4 days after infection, and the viral level subsequently

increased. The salivary glands and hemolymph had the highest

levels of WNV, while the viral load in the midgut was substantially

lower (Figure 1A). mosGCTL-1 expression was induced in

diverse tissues over time, including the salivary glands, hemo-

lymph, and midgut (Figures 1B–1E). Immunoblots also demon-

strated an increased amount of mosGCTL-1 in WNV-infected

mosquitoes (Figure 1F).

mosGCTL-1 Facilitates WNV Infection in A. aegypti

Double-stranded RNA (dsRNA)-mediated gene silencing studies

then assessed the role of mosGCTL-1 in WNV infection of

mosquitoes. Total RNA from mock-injected and mosGCTL-1

dsRNA-injected mosquitoes was analyzed by RT-QPCR. Com-

pared to the mock group, the target gene was silenced 6- to

10-fold from day 3 through 9 (Figures 2A and 2B). WNV was

therefore inoculated into mosquitoes on day 3 after dsRNA treat-

ment, and the viral load was quantified on day 9. There was

a 3-fold reduction in the WNV burden in mosGCTL-1 silenced

A. aegypti (Figure 2C, p < 0.0001) compared to controls.

In contrast, the burden of a related flavivirus, dengue virus,

was unaffected by mosGCTL-1 silencing (Figure 2D), suggesting

that mosGCTL-1 has a specific role in WNV infection of

mosquitoes.

As mosGCTL-1 is part of a multigene family with 25 members

and shares 22%–48% identity with many of these genes, we

examined the role of the mosGCTL paralogues in WNV and

DENV infection. Twenty-one of these 25 genes were expressed

in adult female A. aegypti (Tables S1 and S2). Since AAEL014382

and AAEL014390 share greater than 95% identity with

AAEL011607 and AAEL011619, respectively, AAEL011607 and

AAEL011619 were selected for functional assays. We silenced

each of the expressed mosGCTL genes and then infected the

mosquitoes with virus. In the WNV challenge, two additional

mosGCTL subtypes (AAEL000556 and AAEL011610) showed

the similar phenotype to mosGCTL-1 (Figure 2E), and in DENV

infection, silencing of 5 mosGCTL-related genes reduced the

DENV burden (Figure S2).

Culex mosquitoes are a common vector for WNV in nature.

We therefore extended the studies to Culex quinquefasciatus,

a example of this species that is abundant in the southeast

U.S. We identified a mosGCTL-1 homolog (CPIJ010995, Culex

mosGCTL-1) in C. quinquefasciatus that has a higher degree

of similarity (64%) with mosGCTL-1 than other homologs.

We then determined whether Culex mosGCTL-1 had a similar

role in WNV infection. Culex mosGCTL-1 was upregulated by

WNV infection (Figure 2F). Silencing of Culex mosGCTL-1 in

C. quinquefasciatus reduced the WNV burden compared to the

control group (Figures 2G and 2H, p < 0.0001), suggesting that

Culex mosGCTL-1 plays a role in facilitating WNV infection in

Culex mosquitoes.

mosGCTL-1 Interacts with WNVOur results show that silencing mosGCTL-1 in A. aegypti and

C. quinquefasciatus reduces WNV infection. The mechanism

by which mosGCTL-1 facilitates viral infection in mosquitoes

was, therefore, investigated. We first generated A. aegypti

mosGCTL-1 protein in a Drosophila expression system (Fig-

ure 3A) and investigated whether mosGCTL-1 associates with

WNV envelope (E) protein. Immunoprecipitation experiments

revealed that these two proteins strongly interacted in a

calcium-dependent manner (Figure 3B). We then determined,

by ELISA, that mosGCTL-1 bound to WNV virions. mosGCTL-1

(A, C, D and H) Statistical analysis was done with the Mann-Whitney test in all experiments. Each dot represents the mRNA levels in an individual mosquito.

The horizontal line depicts the medians.

(E–G) The Mann-Whitney test was used for statistical analysis. Data are shown as the mean ± standard error (SEM).

See also Figure S2 and Table S2.

Cell 142, 714–725, September 3, 2010 ª2010 Elsevier Inc. 717

Page 86: Cell 100903

Figure 3. mosGCTL-1 Interacts with WNV

(A) Recombinant mosGCTL-1, produced in Drosophila cells, and purified with a Ni-His column (left). Glycosylated mosGCTL-1 was detected with a V5-HRP mAb

(right). The control was the supernatant of mock-infected S2 cells.

(B) mosGCTL-1 interacted with WNV E protein in a coimmunoprecipitation assay. The protein complex was pulled down with a flavivirus E mAb and probed with

anti-V5-HRP mAb. The experiments were repeated three times.

(C) mosGCTL-1 captured West Nile virions by ELISA. The interaction was determined with a flavivirus E mAb. Data are expressed as the mean ± standard error

from three independent experiments.

(D) Comicroinjection of the purified mosGCTL-1 with WNV enhanced virus infection in A. aegypti. The various amount of purified mosGCTL-1 were premixed with

WNV (10 M.I.D50 per mosquito) for 1 hr at 4�C. The mixture was microinjected into mosquito and compared to the control group inoculated with the same amount

WNV. The mosquitoes were collected at 3 days (i), 6 days (ii), and 9 days (iii) after WNV inoculation. Total RNA was isolated to determine the viral burden by

Taqman RT-QPCR and normalized with A. aegypti actin. Each dot represents the mRNA level in one mosquito. The horizontal line represents the medians.

n.s., nonsignificance (p > 0.05). The Mann-Whitney test was used for analysis. Three independent experiments yielded similar data.

(E) The association between mosGCTL-1 and WNV in A .aegypti hemolymph. Hemolymph was collected from WNV-infected or mock-infected mosquitoes

for immunofluorescence staining. WNV E protein was stained with anti-mouse IgG Alexa-488 (green), and mosGCTL-1 was identified with anti-rabbit

IgG Alexa-546 (red). Nuclei were stained blue with To-Pro-3 iodide. The white arrow represents the induced expression of mosGCTL-1 in WNV-infected

718 Cell 142, 714–725, September 3, 2010 ª2010 Elsevier Inc.

Page 87: Cell 100903

efficiently captured virus in the presence of calcium, and

this interaction was inhibited by EDTA (Figure 3C). Since this

association suggests a role during infection, we premixed

mosGCTL-1 and WNV and then coinjected the combination

into A. aegypti and determined the viral burden. mosGCTL-1

protein significantly enhanced the viral load at days 3 (Fig-

ure 3D, i, p = 0.0001) and 6 (Figure 3D, ii, p < 0.0001) after viral

challenge.

To further characterize the association between mosGCTL-1

and WNV infection in vivo, we used immunofluorescence to

examine the hemolymph and salivary glands of WNV-infected

A. aegypti at different time points. Colocalization of mosGCTL-1

and WNV was clearly observed in both tissues at various inter-

vals after WNV infection (Figure 3E and Figure S3). In the hemo-

lymph, some hemocytes were highly infected by WNV, and

mosGCTL-1 was induced in these infected hemocytes

(Figure 3E). We counted the number of mosGCTL-1-positive

and WNV-infected hemocytes by microscopy. All these infected

cells stained positive for mosGCTL-1. The virus spread rapidly to

the salivary glands after inoculation (Figure S3), consistent with

the Q-PCR data (Figure 1A). mosGCTL-1 was identified on the

basement membrane of salivary glands in uninfected mosqui-

toes, and throughout this tissue after infection, suggesting that

mosGCTL-1 was induced by WNV infection of the salivary

glands (Figure S3).

mosPTP-1 Captures mosGCTL-1 onto the Cell SurfaceHuman mannose-binding lectin (MBL) (Zelensky and Gready,

2005) and mosGCTL-1 are both secreted. Human MBL is

thought to interact with several surface receptors to exert pleo-

trophic effects (Baldwin and Ostergaard, 2001; Arnold et al.,

2006). We therefore postulated that secreted mosGCTL-1

captures WNV and presents it to a ligand on the cell surface,

thereby facilitating viral entry. To assess this, we identified 11

A. aegypti homologs of human proteins that putatively interact

with human MBL (Table S3) from the Human Protein Reference

Database (http://www.hprd.org/) and examined their roles in

WNV infection of A. aegypti (Figure S4A). One mosquito CD45

homolog (AAEL013105, mosPTP-1) exhibited a phenotype

similar to that of mosGCTL-1. The viral burden was significantly

decreased in the mosPTP-1 dsRNA-treated mosquitoes (p <

0.002) (Figures 4A and 4B). To further determine whether

mosPTP-1 interacts with mosGCTL-1, we cloned and expressed

the extracellular region of mosPTP-1 (mosPTP-1-Ex) in a

Drosophila cell line (Figure 4C). mosGCTL-1 strongly interacted

with mosPTP-1-Ex in a coimmunoprecipitation assay (Figure 4D).

We then expressed the mosPTP-1 gene, including the trans-

membrane region (91 bp–2202 bp), in S2 cells (Figure 4C) and

examined the ability of mosPTP-1 to bind mosGCTL-1 on the

cell surface. A mock DNA vector transfected stable S2 cell line

was used as control. More than 40% of the mosPTP-1-express-

ing cells bound mosGCTL-1 and showed double-positive stain-

ing in FACS (Figure 4E, i) compared with controls (Figure 4E, ii,

iii). Confocal microscopy demonstrated that cells expressing

high levels of mosPTP-1 in the cytoplasm also had substantial

mosGCTL-1 on the cell surface (Figure 4F). These results

suggest that mosGCTL-1 binds mosPTP-1.

To test whether mosPTP-1 has a conserved role in Culex spp.,

we identified the mosPTP-1 homolog from the Culex quinquefas-

ciatus genome (CPIJ014098, Culex mosPTP-1). Silencing of this

gene in Culex also influenced the viral burden (Figures S4B and

S4C), similar to mosPTP-1 in A. aegypti (Figure 4B), suggesting

that mosPTP-1 in both these mosquito species have a similar

role in WNV infection.

The mosGCTL-1/mosPTP-1 Pathway Has a DominantRole in WNV InfectionmosGCTL-1 and mosPTP-1 each facilitate WNV infection of

A. aegypti in vivo and in vitro. We therefore assessed whether

mosGCTL-1 and mosPTP-1 cooperate to enable viral infection.

WNV E protein, mosGCTL-1, and mosPTP-1-Ex form a complex

in which mosGCTL-1 is the key factor linking the other two

proteins (Figure 5A). WNV E protein did not interact directly

with mosPTP-1-Ex (Figure 5A). We then examined whether

mosPTP-1-expressing cells could recruit more WNV in the pres-

ence of mosGCTL-1. mosGCTL-1 and WNV were added to cells

and incubated at 4�C for membrane attachment. After washing,

the cells were transferred to room temperature and collected at

different time points to determine the viral burden (Krishnan

et al., 2007). mosPTP-1-expressing cells incorporated up to

5- to 10-fold more WNV in the presence of mosGCTL-1, com-

pared to control groups (Figure 5B). We then examined the role

of the mosGCTL-1/mosPTP-1 pathway for WNV infection in

A. aegypti. We knocked down the mosPTP-1 gene with dsRNA

and then inoculated the mosGCTL-1/WNV mixture at 3 days

after RNAi silencing. Silencing of mosPTP-1 interfered with the

ability of mosGCTL-1 to facilitate WNV infection (shown in Fig-

ure 3D) at 3 and 6 days after challenge (Figure 5C). The viral

burden of the mosPTP-1 RNAi/mosGCTL-1/WNV group was

2- to 3-fold lower than that of the mock/mosGCTL-1/WNV group

(p < 0.001) and decreased to a level similar to that of the group

that did not receive mosGCTL-1, suggesting that mosGCTL-1

and mosPTP-1 cooperate in the same pathway to enhance

WNV infection. To further determine the relationship between

mosGCTL-1 and mosPTP-1 in infection, we silenced both of

these genes with dsRNA. These genes were successfully

knocked down in mosquitoes in the cosilenced group (Figures

S5A and S5B). The decrease in the viral burden was similar in

the cosilenced and individually silenced groups (Figure S5C),

suggesting that mosPTP-1 is the dominant downstream

receptor for mosGCTL-1 in the process of WNV infection of

A. aegypti.

To better understand the association between mosGCTL-1,

mosPTP-1, and WNV in vivo, we examined the distribution of

mosPTP-1 in A. aegypti. mosPTP-1 was highly expressed in

various mosquito tissues but not induced by WNV infection

(Figure S4D). The salivary glands and hemolymph were sites of

abundant mosPTP-1 expression, while expression in midgut

hemocytes. The yellow arrows show the infected hemocytes. Images were examined with a Zeiss LSM 510 Meta Confocal Microscope 633 objective

lens.

See also Figure S3.

Cell 142, 714–725, September 3, 2010 ª2010 Elsevier Inc. 719

Page 88: Cell 100903

Figure 4. mosPTP-1 Captures mosGCTL-1 onto the Cell Surface

(A) RNAi efficiency for the mosPTP-1 gene at 6 days after dsRNA treatment. The amount of mosPTP-1 mRNA was determined by SYBR Green RT-QPCR and

normalized with A. aegypti actin. Data are represented as the mean ± standard error.

(B) Silencing mosPTP-1 decreased WNV infection. The viral burden was measured at day 6. WNV (10 M.I.D50) was injected into each mosquito. The viral load was

determined by Taqman RT-QPCR and normalized with A. aegypti actin. Statistical analysis was done with the Mann-Whitney test. The horizontal line depicts the

medians. The result shown is a combination of three independent experiments.

(C) Expression of recombinant mosPTP-1 and mosPTP-1-Ex in S2 cells. mosPTP-1 and mosPTP-1-Ex genes were isolated from complementary DNA (cDNA)

library of A. aegypti and expressed as recombinant proteins with an HA tag at the N terminus. The left panel is mosPTP-1, and the right panel is mosPTP-1-EX,

probed by anti-HA tag mAb in western blot. The control was the products from S2 cells transfected with mock DNA vector.

(D) mosGCTL-1 interacts with mosPTP-1-Ex peptide by coimmunoprecipitation. The protein complex was pulled down with a rabbit HA antibody and probed with

a V5-HRP mAb. The experiments were repeated three times.

720 Cell 142, 714–725, September 3, 2010 ª2010 Elsevier Inc.

Page 89: Cell 100903

was comparatively lower (Figures S4E–S4G). We next generated

mosPTP-1 antibody in mice, which recognized native mosPTP-1

protein (Figure S5D) and the mosPTP-1-Ex expressed by S2

cells (Figure S5E). We then determined the relationship between

mosGCTL-1, mosPTP-1, and WNV in salivary glands by immu-

nofluorescence. mosPTP-1 was copiously expressed on the

cell surface of salivary glands (Figure 5D), thereby providing

additional data to complement and extend the initial QPCR

expression data (Figure S4E). Several regions in the salivary

glands with substantial mosPTP-1 also demonstrated staining

for mosGCTL-1 and WNV (Figure 5D).

mosGCTL-1 Antisera Interferes with WNV Infectionof MosquitoesDisruption of the transfer of WNV from the vertebrate to

arthropod host could theoretically diminish viral dissemination

in nature. We therefore investigated whether mosGCTL-1 anti-

sera reduces WNV infection in mosquitoes during the blood

meal. We generated mosGCTL-1 antisera in rabbits and showed

that mosGCTL-1 antisera interfered with mosGCTL-1 binding to

WNV E protein in vitro (Figure 6A). Then, we examined whether

mosGCTL-1 antisera influenced the ability of WNV to infect

mosquitoes during a blood meal. We mixed mosGCTL-1 antisera

and WNV with fresh whole blood and performed membrane

blood feeding with a Hemotek. Seven days later, mosquitoes

were sacrificed to determine the infectivity rate. mosGCTL-1

antisera efficiently blocked WNV infection of A. aegypti. The

number of infected mosquitoes was reduced in mosGCTL-1

antisera treated groups, compared to mock group, by QPCR

(Figure 6B) or a 50% tissue culture infective doses (TCID50) assay

(Figure 6C). Hence, a humoral response against mosGCTL-1 in

a vertebrate host may alter WNV infection of mosquitoes during

the feeding process. This hypothetically affords a strategy to

develop a transmission blocking vaccine to control WNV

dissemination in nature.

DISCUSSION

As mosquitoes are prominent vectors for flaviviruses, specific

interactions between the virus and arthropod likely enhance

pathogen survival. In the mammalian host, C-type lectins such

as DC-SIGN and the mannose receptor augment viral entry

into specific DCs and macrophages (Tassaneetrithep et al.,

2003; Davis et al., 2006; Miller et al., 2008). Our results show

that a secreted mosquito C-type lectin, mosGCTL-1, binds to

WNV in a calcium-dependent manner and enhances viral infec-

tion. A mosquito homolog of human CD45 (mosPTP-1) recruits

mosGCTL-1 to facilitate viral attachment to cells. Based on our

findings, we envision a model whereby WNV that is inoculated

into mosquitoes binds to secreted mosGCTL-1 in the hemo-

lymph, thereby forming a complex in the extracellular milieu

that has the ability to interact with the membrane protein,

mosPTP-1, to facilitate cellular entry. The virus rapidly replicates

in the mosquito thorax. This induces additional mosGCTL-1

expression, which accelerates formation of the mosGCTL-1/

WNV complex—enabling WNV to invade different mosquito

tissues and enhancing viral spread throughout the mosquito

body. This mechanism, which involves WNV associating with

mosGCTL-1 and then being captured by mosPTP-1 onto the

cell surface in mosquitoes, suggests that an extracellular soluble

protein is an important receptor for flavivirus in arthropods.

mosGCTL-1 shares homology with human MBL. In mammals,

MBL is a pattern recognition molecule that recognizes carbohy-

drate moieties on invading microbes (Neth et al., 2002). As exam-

ples, MBL interacts with HIV envelope protein (gp120) (Saifuddin

et al., 2000) and HBV surface antigen (HBsAg) (Chong et al., 2005)

and has a role in the opsonization of HIV (Ezekowitz et al., 1989).

In these processes, MBL associates with serine proteases,

MASPs, and activates the complement system (Neth et al.,

2002). Homologs of the proteins that associate with mammalian

MBL have not been found in A. aegypti (Table S3), suggesting that

the A. aegypti mosGCTL-1 may have different physiological func-

tions than mammalian MBL. Invertebrates lack antibody- and

interferon-based immune responses (Cheng et al., 2009). Since

lectin expression is significantly upregulated by microbial infec-

tion, these molecules are presumed to participate in nonself

recognition and pathogen resistance (Wilson et al., 1999; Tanji

et al., 2006). Indeed, recent studies have shown that a comple-

ment-like system exists in the hemolymph of Anopheles gambiae

and mediates parasite killing (Blandin et al., 2004; Povelones

et al., 2009). It is possible that mosGCTL-1 and other subtypes

in this family, similar to their mammalian homologs, may normally

recognize most pathogens and be involved in the arthropod

complement-like system. Nevertheless, our studies showed

that the expression of mosGCTL-1 is induced by WNV infection.

The induced mosGCTL-1 that then binds to virus amplifies WNV

infection. Overall, these suggest a critical role for mosGCTL-1 in

WNV infection of mosquitoes.

In the mammalian host, the association between MBL and the

CD45 external domain primarily occurs in immature T cells and

affect the development of thymocytes (Baldwin and Ostergaard,

2001). Mammalian CD45 is expressed on the hemopoietic-orig-

inated nucleated cells (Thomas, 1989); however, the mosquito

CD45 homolog, mosPTP-1, does not appear to be restricted to

particular cells. As a transmembrane protein, mosPTP-1 was

abundantly detected in the salivary glands and hemolymph of

mosquitoes. The pattern of mosPTP-1 expression correlated

with the distribution of WNV in A. aegypti. In our model, after

(E) mosPTP-1 captured mosGCTL-1 to the cell surface by flow cytometry. A stable cell line was generated to express mosPTP-1 in S2 cells. The purified

mosGCTL-1 was inoculated with mosPTP-1-expressing cells at 4�C. An empty DNA vector transfected stable S2 cell line was used as the control. The interaction

between mosPTP-1 and mosGCTL-1 was investigated by FACS. mosPTP-1 was stained by Alexa-488; mosGCTL-1 was stained by Phycoerythrin (PE). Three

independent experiments yielded similar results, and one representative study is shown in this figure.

(F) Confocal microscopy to examine for mosPTP-1 and mosGCTL-1. mosGCTL-1 was stained with Alexa-488 (green) and mosPTP-1 was identified with Alexa-

546 (red). Nuclei were stained by To-Pro-3 iodide (blue). The images were collected using a Zeiss LSM 510 Meta Confocal Microscope 633 objective lens.

The arrows represent the overlap between mosPTP-1 and mosGCTL-1.

See also Figure S4 and Table S3.

Cell 142, 714–725, September 3, 2010 ª2010 Elsevier Inc. 721

Page 90: Cell 100903

Figure 5. The mosGCTL-1/mosPTP-1 Pathway in WNV Infection

(A) mosGCTL-1 associated with WNV E protein bound to mosPTP-1-Ex. The protein complex was pulled down with an HA antibody and probed with V5-HRP and

flavivirus E protein antibody. The experiment was repeated three times.

(B) mosPTP-1-expressing cells recruit WN virions in the presence of mosGCTL-1. mosGCTL-1 and WNV were added to cells and incubated at 4�C for 1 hr, for

membrane attachment. Cells were gently washed three times with cold PBS and were then moved to room temperature. At different time points—0, 15, and

60 min—cells were collected for total RNA isolation. Control cells were the empty DNA vector transfected stable S2 cell line. Viral attachment was determined

by SYBR Green RT-QPCR and normalized with Drosophila actin 5C (CG4027). Statistical analysis was done with ANOVA. Data are represented as the mean ±

standard error. The results are representative of three independent experiments.

(C) Silencing of mosPTP-1 impairs the function of mosGCTL-1. The mosPTP-1 gene was knocked down with dsRNA treatment. Then the mosGCTL-1/WNV

mixture was inoculated at 3 days after RNAi silencing. The virus burden was measured at 3 (i) and 6 (ii) days after the introduction of virus by Taqman RT-QPCR.

WNV (10 M.I.D50) and mosGCTL-1 (100 pg) were inoculated into each mosquito. Each dot represents the mRNA level in an individual mosquito. Statistical analysis

was done with the Mann-Whitney test. The horizontal line depicts the medians. The result shown is the combination of five independent experiments.

(D) Immunostaining of mosGCTL-1, mosPTP-1, and WNV in A. aegypti salivary glands. mosPTP-1 was stained with anti-mouse IgG Alexa-488 (green),

mosGCTL-1 was identified by anti-rabbit IgG Alexa-546 (red), and WNV E protein was probed with horse anti-E protein IgG and detected by anti-horse IgG

Alexa-633 (blue). The arrows show the sites of overlap between mosGCTL-1, mosPTP-1, and WNV at 6 days after infection. LL, lateral lobe; ML, median lobe

in female A.aegypti salivary glands. Images were examined with a Zeiss LSM 510 Meta Confocal Microscope 253 objective lens.

See also Figure S5.

722 Cell 142, 714–725, September 3, 2010 ª2010 Elsevier Inc.

Page 91: Cell 100903

binding to mosGCTL-1, WNV binds to membrane bound

mosPTP-1. This implies that mosGCTL-1 and mosPTP-1 are

recruited as receptors to facilitate cellular invasion by WNV.

Mosquito control is a common strategy to influence

WNV numbers in nature (van der Meulen et al., 2005; Dauphin

and Zientara, 2007). The increase in viral spread and fatalities

over the last decade (Reisen and Brault, 2007; Lindsey et al.,

2009) (http://www.cdc.gov/ncidod/dvbid/westnile/index.htm)

suggests that additional strategies could assist in combating

WNV. For arthropod-borne microbes, vector ligands that interact

with pathogens are potential targets for interfering with

the successful acquisition of the microbe from the vertebrate

host. As an example, blocking the tick gut receptor for the

Lyme disease agent limits the colonization of ticks by Borrelia

burgdorferi (Pal et al., 2004). Our studies show that blocking

mosGCTL-1 within A. aegypti reduced the vector competence

for WNV and interrupted the infective cycle of WNV. These

results indicate that it is theoretically possible to develop a trans-

mission-blocking vaccine to interfere with the migration of WNV

from vertebrates to mosquito, thereby restricting viral dissemi-

nation in the environment.

In summary, we identified a lectin-based pathway to facilitate

flaviviral entry, in which mosGCTL-1 and mosPTP-1 are cascade

receptors in WNV infection of A. aegypti and C. quinquefascia-

tus. Characterization of mosquito ligands for WNV enhances

Figure 6. mosGCTL-1 Antiserum Interferes

with WNV Infection of Mosquitoes

(A) mosGCTL-1 antisera blocked the interaction

between mosGCTL-1 and WNV E protein.

mosGCTL-1 antisera or mock sera was diluted

1000-fold. The protein complex was pulled down

with a V5 mAb, and WNV E protein was detected

with an E protein antibody.

(B and C) mosGCTL-1 antisera interrupted WNV

infection during the blood meal. The antisera or

control sera were diluted 100- or 2000-fold with

fresh whole blood containing 5 3 106 pfu/ml

WNV. Membrane blood-feeding was then per-

formed with a Hemotek. Seven days later, mosqui-

toes were sacrificed to determine the infectivity

rate by Taqman RT-QPCR (B) and TCID50 (C).

Each group included 50 mosquitoes in the QPCR

assay and 32 mosquitoes in TCID50 assay. One

dot corresponds to a mosquito. n, the number of

mosquitoes in each group. The result is represen-

tative of three independent experiments.

our understanding of flavivirus-arthropod

interactions and may aid in the develop-

ment of strategies to target selected

points in the flaviviral life cycle and inter-

fere with these pathogens in nature.

EXPERIMENTAL PROCEDURES

Mosquitoes and Viruses

A. aegypti and C. quinquefasciatus mosquitoes

were maintained in a sugar solution at 27�C and

80% humidity according to standard rearing

procedures (Keene et al., 2004; Xi et al., 2008).

WNV strain 2471 and DENV-2 (DENV New Guinea C strain) were passaged

in mosquito C6/36 cells (Hanna et al., 2005; Lin et al., 2007; Krishnan et al.,

2008). The titer of WNV for cell culture was determined by a plaque formation

assay as described previously (Bai et al., 2007). The viruses for in vivo exper-

iments were titrated in mosquitoes through thoracic microinjection. The dose

of viruses was determined by 10-fold serial dilutions (i.e., 10�4, 10�5, 10�6, and

10�7) in PBS. The mosquitoes (12 in each group) were inoculated in the thorax

by microinjection with 300 nl of diluted virus. On day 6, the mosquitoes were

sacrificed, total RNA was isolated, and the viral load determined by RT-

QPCR (Figure S1). The 50% mosquito infective dose (M.I.D50) was estimated

by the Reed-Muench method (Pizzi, 1950).

Gene Silencing and Viral Challenge in Mosquitos

dsRNA synthesis was performed as described previously (Brackney et al.,

2008). The primers are shown in Table S4. For silencing the target genes, adult

female mosquitoes were kept on ice for 15 min and then transferred to a cold

tray to receive a systemic injection of dsRNA into the hemocoele. Two micro-

grams of dsRNA/300 nl in PBS was microinjected into the thorax of each

mosquito. After a 3 day recovery period, the mosquitoes were microinjected

with either WNV 10 M.I.D50/ 300 nl for functional studies or 1000 M.I.D50/

300 nl for expression profile assays.

Purification of mosGCTL-1 with the Drosophila Expression System

mosGCTL-1 was amplified by RT-PCR from adult female A. aegypti. The

primers are shown in Table S4. The PCR product was subcloned into the

pMT/BiP/V5-His A vector (Invitrogen, Carlsbad, CA) and transfected into

Drosophila S2 cells in combination with the hygromycin selection vector

pCo-Hygro for stable transfection. The cells were selected through the

use of 300 mg/ml Hygromycin-B (Invitrogen) for 4 weeks. The resistant cells

Cell 142, 714–725, September 3, 2010 ª2010 Elsevier Inc. 723

Page 92: Cell 100903

were grown in spinner flasks, switched to Express Five serum-free medium

(GIBCO, Invitrogen) for 3 days, and induced with copper sulfate at a final

concentration of 500 mM for 4 days. The culture medium was cleared by

centrifugation at 1000 3 g for 5 min and collected for protein purification

with the Talon metal affinity resin (Clontech, Mountain View, CA). The

protein was eluted with 150 mM imidazole, extensively dialyzed against

PBS (pH 7.8), and concentrated by centrifugal filtration through a 5 kDa

filter (Millipore, Bedford, MA). The protein purity was checked by SDS-

PAGE, and immunoblots were performed with an anti-V5-HRP mouse

mAb (Invitrogen).

Isolation and Image of Mosquito Tissues

Salivary glands were dissected as the previously described (Coleman et al.,

2007). Tissues were isolated, placed on sialylated slides (PGC Scientific,

Gaithersburg, MD), washed in PBS, and fixed in 4% PFA at 37�C for 1 hr.

For hemolymph isolation, the mosquitoes were anesthetized on ice tray, and

then the proboscis was removed with forceps. Hemolymph was expelled in

a droplet using the tip of the proboscis upon pressure to the thorax. Only clear

droplets were collected, to avoid contamination by the fat body (Han et al.,

1999). The liquid was allowed to be dried on sialylated slides and fixed in

4% PFA. Samples were blocked in PBS with 1% BSA and 0.1% Triton

X-100 at room temperature for 2 hr before antibody incubations. After staining

by primary and secondary antibodies, slides were imaged with the Multi-Track

mode of a Zeiss LSM 510 Meta Confocal Microscope.

Membrane Blood Feeding

Fresh whole blood was obtained from C57BL/6J mice, placed in tubes with

anticoagulant, and centrifuged at 3000 rpm for 10 min to separate the serum

and cells. The serum was collected and heat inactivated at 57�C for 1 hr.

The blood cells were washed in PBS three times to remove the anticoagulant.

The cells were resuspended with serum. Five million plaque forming units per

milliliter WNV and antisera (or preimmune sera) were added to the treated

blood. About 1 ml of the virus/sera/blood mixture was used on each Hemotek

feeder. The feeder was placed on the mosquito container and the mosquitoes

were allowed to feed for 1 hr. The container was transferred to 4�C for 20 min to

anesthetize the mosquitoes. Carefully selected fed mosquitoes were placed

into new containers and reared for 7 days. The mosquitoes were sacrificed;

the whole mosquitoes were homogenized in RLT buffer (RNeasy Mini kit,

QIAGEN) to measure the viral load by RT-QPCR, or grinded in PBS for virologic

assays.

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures, five

figures, and four tables and can be found with this article online at doi:10.

1016/j.cell.2010.07.038.

ACKNOWLEDGMENTS

This work was supported by National Institutes of Health grants AI 50031 and

AI 070343. We thank Yue Zhang, Lei Liu, Lili Zhang, and Sukanya Narasimhan

for help with the biochemical studies, Debbie Beck for technical assistance,

and Michel Ledizet for providing the purified WNV E protein and the flaviviral

E protein antibodies. G.C. is a James Hudson Brown-Alexander Brown

Coxe Postdoctoral Fellow at the Yale School of Medicine. P.W. is supported

by a Career Development Award from Northeast Biodefense Center (U54-

AI057158-Lipkin). E.F. is an investigator with the Howard Hughes Medical

Institute. G.C. and E.F. designed the experiments and wrote the manuscript;

G.C. performed the majority of the experiments and analyzed data; J.C. assis-

ted the mosquito techniques; J.F.A. provided the A. aegypti and C. quinque-

fasciatus mosquitoes and helped with the mosquito feeding; J.D. and F.Q.

assisted in FACS and confocal microscopy; and P.W., M.N.K. and J.D.

provided suggestions for the project. All authors reviewed, critiqued, and

provided comments on the text.

Received: September 12, 2009

Revised: April 5, 2010

Accepted: July 27, 2010

Published online: August 26, 2010

REFERENCES

Alonso, A., Sasin, J., Bottini, N., Friedberg, I., Friedberg, I., Osterman, A., God-

zik, A., Hunter, T., Dixon, J., and Mustelin, T. (2004). Protein tyrosine phospha-

tases in the human genome. Cell 117, 699–711.

Arnold, J.N., Dwek, R.A., Rudd, P.M., and Sim, R.B. (2006). Mannan binding

lectin and its interaction with immunoglobulins in health and in disease. Immu-

nol. Lett. 106, 103–110.

Bai, F.W., Town, T., Pradhan, D., Cox, J., Ashish, Ledizet, M., Anderson, J.F.,

Flavell, R.A., Krueger, J.K., Koski, R.A., and Fikrig, E. (2007). Antiviral peptides

targeting the west nile virus envelope protein. J. Virol. 81, 2047–2055.

Baldwin, T.A., and Ostergaard, H.L. (2001). Developmentally regulated

changes in glucosidase II association with, and carbohydrate content of, the

protein tyrosine phosphatase CD45. J. Immunol. 167, 3829–3835.

Blandin, S., Shiao, S.H., Moita, L.F., Janse, C.J., Waters, A.P., Kafatos, F.C.,

and Levashina, E.A. (2004). Complement-like protein TEP1 is a determinant

of vectorial capacity in the malaria vector Anopheles gambiae. Cell 116,

661–670.

Brackney, D.E., Foy, B.D., and Olson, K.E. (2008). The effects of midgut serine

proteases on dengue virus type 2 infectivity of Aedes aegypti. Am. J. Trop.

Med. Hyg. 79, 267–274.

Byth, K.F., Conroy, L.A., Howlett, S., Smith, A.J., May, J., Alexander, D.R., and

Holmes, N. (1996). CD45-null transgenic mice reveal a positive regulatory role

for CD45 in early thymocyte development, in the selection of CD4+CD8+

thymocytes, and B cell maturation. J. Exp. Med. 183, 1707–1718.

Chen, S.T., Lin, Y.L., Huang, M.T., Wu, M.F., Cheng, S.C., Lei, H.Y., Lee, C.K.,

Chiou, T.W., Wong, C.H., and Hsieh, S.L. (2008). CLEC5A is critical for

dengue-virus-induced lethal disease. Nature 453, 672–676.

Cheng, G., Zhao, X., Li, Z., Liu, X., Yan, W., Zhang, X., Zhong, Y., and Zheng, Z.

(2009). Identification of a putative invertebrate helical cytokine similar to the

ciliary neurotrophic factor/leukemia inhibitory factor family by PSI-BLAST-

based approach. J. Interferon Cytokine Res. 29, 461–468.

Chong, W.P., To, Y.F., Ip, W.K., Yuen, M.F., Poon, T.P., Wong, W.H., Lai, C.L.,

and Lau, Y.L. (2005). Mannose-binding lectin in chronic hepatitis B virus infec-

tion. Hepatology 42, 1037–1045.

Coleman, J., Juhn, J., and James, A.A.A. (2007). Dissection of midgut and

salivary glands from Ae. aegypti mosquitoes. J. Vis. Exp. 5, 228.

Dauphin, G., and Zientara, S. (2007). West Nile virus: recent trends in diagnosis

and vaccine development. Vaccine 25, 5563–5576.

Davis, C.W., Nguyen, H.Y., Hanna, S.L., Sanchez, M.D., Doms, R.W., and Pier-

son, T.C. (2006). West Nile virus discriminates between DC-SIGN and

DC-SIGNR for cellular attachment and infection. J. Virol. 80, 1290–1301.

Ezekowitz, R.A., Kuhlman, M., Groopman, J.E., and Byrn, R.A. (1989).

A human serum mannose-binding protein inhibits in vitro infection by the

human immunodeficiency virus. J. Exp. Med. 169, 185–196.

Geijtenbeek, T.B., Torensma, R., van Vliet, S.J., van Duijnhoven, G.C., Adema,

G.J., van Kooyk, Y., and Figdor, C.G. (2000). Identification of DC-SIGN, a novel

dendritic cell-specific ICAM-3 receptor that supports primary immune

responses. Cell 100, 575–585.

Gould, E.A., and Solomon, T. (2008). Pathogenic flaviviruses. Lancet 371,

500–509.

Gubler, D.J. (1998). Dengue and dengue hemorrhagic fever. Clin. Microbiol.

Rev. 11, 480–496.

Halstead, S.B. (2008). Dengue virus-mosquito interactions. Annu. Rev. Ento-

mol. 53, 273–291.

Han, Y.S., Chun, J., Schwartz, A., Nelson, S., and Paskewitz, S.M. (1999).

Induction of mosquito hemolymph proteins in response to immune challenge

and wounding. Dev. Comp. Immunol. 23, 553–562.

724 Cell 142, 714–725, September 3, 2010 ª2010 Elsevier Inc.

Page 93: Cell 100903

Hanna, S.L., Pierson, T.C., Sanchez, M.D., Ahmed, A.A., Murtadha, M.M., and

Doms, R.W. (2005). N-linked glycosylation of west nile virus envelope proteins

influences particle assembly and infectivity. J. Virol. 79, 13262–13274.

Hayes, E.B., Komar, N., Nasci, R.S., Montgomery, S.P., O’Leary, D.R., and

Campbell, G.L. (2005). Epidemiology and transmission dynamics of West

Nile virus disease. Emerg. Infect. Dis. 11, 1167–1173.

Hubalek, Z., and Halouzka, J. (1999). West Nile fever—a reemerging mosquito-

borne viral disease in Europe. Emerg. Infect. Dis. 5, 643–650.

Keene, K.M., Foy, B.D., Sanchez-Vargas, I., Beaty, B.J., Blair, C.D., and Olson,

K.E. (2004). RNA interference acts as a natural antiviral response to O’nyong-

nyong virus (Alphavirus; Togaviridae) infection of Anopheles gambiae. Proc.

Natl. Acad. Sci. USA 101, 17240–17245.

Klimstra, W.B., Nangle, E.M., Smith, M.S., Yurochko, A.D., and Ryman, K.D.

(2003). DC-SIGN and L-SIGN can act as attachment receptors for alphaviruses

and distinguish between mosquito cell- and mammalian cell-derived viruses.

J. Virol. 77, 12022–12032.

Krishnan, M.N., Sukumaran, B., Pal, U., Agaisse, H., Murray, J.L., Hodge,

T.W., and Fikrig, E. (2007). Rab 5 is required for the cellular entry of dengue

and West Nile viruses. J. Virol. 81, 4881–4885.

Krishnan, M.N., Ng, A., Sukumaran, B., Gilfoy, F.D., Uchil, P.D., Sultana, H.,

Brass, A.L., Adametz, R., Tsui, M., Qian, F., et al. (2008). RNA interference

screen for human genes associated with West Nile virus infection. Nature

455, 242–245.

Leis, A.A., Stokic, D.S., Polk, J.L., Dostrow, V., and Winkelmann, M. (2002).

A poliomyelitis-like syndrome from West Nile virus infection. N. Engl. J. Med.

347, 1279–1280.

Lin, C.C., Yang, C.F., Tu, C.H., Huang, C.G., Shih, Y.T., Chuang, C.K., and

Chen, W.J. (2007). A novel tetraspanin C189 upregulated in C6/36 mosquito

cells following dengue 2 virus infection. Virus Res. 124, 176–183.

Lindsey, N.P., Hayes, E.B., Staples, J.E., and Fischer, M. (2009). West Nile

virus disease in children, United States, 1999-2007. Pediatrics 123, e1084–

e1089.

Miller, J.L., de Wet, B.J., deWet, B.J., Martinez-Pomares, L., Radcliffe, C.M.,

Dwek, R.A., Rudd, P.M., and Gordon, S. (2008). The mannose receptor medi-

ates dengue virus infection of macrophages. PLoS Pathog. 4, e17.

Monath, T.P., and Heinz, F.X. (1996). Flaviviruses (Philadelphia: Lippincott-

Raven Publishers).

Mustelin, T., Vang, T., and Bottini, N. (2005). Protein tyrosine phosphatases

and the immune response. Nat. Rev. Immunol. 5, 43–57.

Nene, V., Wortman, J.R., Lawson, D., Haas, B., Kodira, C., Tu, Z.J., Loftus, B.,

Xi, Z., Megy, K., Grabherr, M., et al. (2007). Genome sequence of Aedes

aegypti, a major arbovirus vector. Science 316, 1718–1723.

Neth, O., Jack, D.L., Johnson, M., Klein, N.J., and Turner, M.W. (2002).

Enhancement of complement activation and opsonophagocytosis by

complexes of mannose-binding lectin with mannose-binding lectin-associ-

ated serine protease after binding to Staphylococcus aureus. J. Immunol.

169, 4430–4436.

Nir, Y., Goldwasser, R., Lasowski, Y., and Margalit, J. (1968). Isolation of West

Nile virus strains from mosquitoes in Israel. Am. J. Epidemiol. 87, 496–501.

Pal, U., Li, X., Wang, T., Montgomery, R.R., Ramamoorthi, N., Desilva, A.M.,

Bao, F., Yang, X., Pypaert, M., Pradhan, D., et al. (2004). TROSPA, an Ixodes

scapularis receptor for Borrelia burgdorferi. Cell 119, 457–468.

Pizzi, M. (1950). Sampling variation of the fifty percent end-point, determined

by the Reed-Muench (Behrens) method. Hum. Biol. 22, 151–190.

Povelones, M., Waterhouse, R.M., Kafatos, F.C., and Christophides, G.K.

(2009). Leucine-rich repeat protein complex activates mosquito complement

in defense against Plasmodium parasites. Science 324, 258–261.

Reisen, W., and Brault, A.C. (2007). West Nile virus in North America: perspec-

tives on epidemiology and intervention. Pest Manag. Sci. 63, 641–646.

Robinson, M.J., Sancho, D., Slack, E.C., LeibundGut-Landmann, S., and Reis

e Sousa, C. (2006). Myeloid C-type lectins in innate immunity. Nat. Immunol. 7,

1258–1265.

Saifuddin, M., Hart, M.L., Gewurz, H., Zhang, Y., and Spear, G.T. (2000). Inter-

action of mannose-binding lectin with primary isolates of human immunodefi-

ciency virus type 1. J. Gen. Virol. 81, 949–955.

Sessions, O.M., Barrows, N.J., Souza-Neto, J.A., Robinson, T.J., Hershey,

C.L., Rodgers, M.A., Ramirez, J.L., Dimopoulos, G., Yang, P.L., Pearson,

J.L., and Garcia-Blanco, M.A. (2009). Discovery of insect and human dengue

virus host factors. Nature 458, 1047–1050.

Soilleux, E.J., Morris, L.S., Leslie, G., Chehimi, J., Luo, Q., Levroney, E., Trow-

sdale, J., Montaner, L.J., Doms, R.W., Weissman, D., et al. (2002). Constitutive

and induced expression of DC-SIGN on dendritic cell and macrophage

subpopulations in situ and in vitro. J. Leukoc. Biol. 71, 445–457.

Tanji, T., Ohashi-Kobayashi, A., and Natori, S. (2006). Participation of a galac-

tose-specific C-type lectin in Drosophila immunity. Biochem. J. 396, 127–138.

Tassaneetrithep, B., Burgess, T.H., Granelli-Piperno, A., Trumpfheller, C.,

Finke, J., Sun, W., Eller, M.A., Pattanapanyasat, K., Sarasombath, S., Birx,

D.L., et al. (2003). DC-SIGN (CD209) mediates dengue virus infection of human

dendritic cells. J. Exp. Med. 197, 823–829.

Thomas, M.L. (1989). The leukocyte common antigen family. Annu. Rev.

Immunol. 7, 339–369.

Trowbridge, I.S., and Thomas, M.L. (1994). CD45: an emerging role as a protein

tyrosine phosphatase required for lymphocyte activation and development.

Annu. Rev. Immunol. 12, 85–116.

Turell, M.J., O’Guinn, M., and Oliver, J. (2000). Potential for New York mosqui-

toes to transmit West Nile virus. Am. J. Trop. Med. Hyg. 62, 413–414.

van der Meulen, K.M., Pensaert, M.B., and Nauwynck, H.J. (2005). West Nile

virus in the vertebrate world. Arch. Virol. 150, 637–657.

Vanlandingham, D.L., McGee, C.E., Klinger, K.A., Vessey, N., Fredregillo, C.,

and Higgs, S. (2007). Relative susceptibilties of South Texas mosquitoes to

infection with West Nile virus. Am. J. Trop. Med. Hyg. 77, 925–928.

Wilson, R., Chen, C.W., and Ratcliffe, N.A. (1999). Innate immunity in insects:

the role of multiple, endogenous serum lectins in the recognition of foreign

invaders in the cockroach, Blaberus discoidalis. J. Immunol. 162, 1590–1596.

Xi, Z., Ramirez, J.L., and Dimopoulos, G. (2008). The Aedes aegypti toll

pathway controls dengue virus infection. PLoS Pathog. 4, e1000098.

Zelensky, A.N., and Gready, J.E. (2005). The C-type lectin-like domain super-

family. FEBS J. 272, 6179–6217.

Cell 142, 714–725, September 3, 2010 ª2010 Elsevier Inc. 725

Page 94: Cell 100903

The ATAC Acetyltransferase ComplexCoordinates MAP Kinasesto Regulate JNK Target GenesTamaki Suganuma,1 Arcady Mushegian,1,2 Selene K. Swanson,1,3 Susan M. Abmayr,1,3 Laurence Florens,1

Michael P. Washburn,1,4 and Jerry L. Workman1,*1Stowers Institute for Medical Research, 1000 East 50th Street, Kansas City, MO 64110, USA2Department of Microbiology, Molecular Genetics, and Immunology3Department of Anatomy and Cell Biology4Department of Pathology and Laboratory Medicine

University of Kansas Medical Center, Kansas City, KS 66160, USA

*Correspondence: [email protected] 10.1016/j.cell.2010.07.045

SUMMARY

In response to extracellular cues, signal transductionactivates downstream transcription factors likec-Jun to induce expression of target genes. Wedemonstrate that the ATAC (Ada two A containing)histone acetyltransferase (HAT) complex serves asa transcriptional cofactor for c-Jun at the JunN-terminal kinase (JNK) target genes Jra and chick-adee. ATAC subunits are required for c-Jun occu-pancy of these genes and for H4K16 acetylation atthe Jra enhancer, promoter, and transcribed se-quences. Under conditions of osmotic stress, ATACcolocalizes with c-Jun, recruits the upstream kinasesMisshapen, MKK4, and JNK, and suppresses furtheractivation of JNK. Relocalization of these MAPKsand suppression of JNK activation by ATAC aredependent on the CG10238 subunit of ATAC. Thus,ATAC governs the transcriptional response to MAPkinase signaling by serving as both a coactivatorof transcription and as a suppressor of upstreamsignaling.

INTRODUCTION

Histone acetyltransferase complexes have been isolated from

multiple organisms and shown to be involved in nuclear events

that relate to chromatin biology (Kimura et al., 2005; Lee and

Workman, 2007). The Drosophila ATAC (Ada two A containing)

complex consists of 13 proteins and includes two distinct

histone acetyltransferases, Gcn5/KAT2 and Atac2/KAT14 (Fig-

ure 1A). Whereas Gcn5/KAT2 preferentially acetylates histones

H3K9 and H3K14, Atac2/KAT14 acetylates H4K16 (Ciurciu

et al., 2006; Guelman et al., 2006; Suganuma et al., 2008). The

Gcn5/KAT2, Ada3, and CG30390 (Sgf29 in yeast) subunits of

ATAC are shared with SAGA (Spt-Ada-Gcn5 acetyltransferase)

(Suganuma et al., 2008), an important transcriptional coactivator

complex (Lee and Workman, 2007). However, it was unknown

whether ATAC functions as a transcription cofactor like SAGA.

Mitogen-activated protein kinases (MAPKs) and their target

kinases in the pathway lead to phosphorylation of target

transcription factors (Edmunds and Mahadevan, 2004; Thomson

et al., 1999). Exposure of cells to increases in extracellular osmo-

lality results in rapid activation of stress-activated MAPKs

(SAPKs), including c-Jun-NH2-terminal kinase (JNK/basket in

Drosophila) and p38 MAPKs (de Nadal et al., 2002; Edmunds

and Mahadevan, 2004; Kyriakis and Avruch, 2001). Osmotic

stress causes activation of JNK by phosphorylation which, in

turn, phosphorylates c-Jun and enhances its transcriptional

activity (Kayali et al., 2000; Thomson et al., 1999). MAPKs are

more intimately involved in the regulation of downstream target

genes than just phosphorylation of target transcription factors

(Dioum et al., 2009). For example, the yeast SAPK Hog1 is

recruited to target genes in chromatin and interacts with tran-

scription factors, cofactors, and RNA polymerase II (Alepuz

et al., 2001, 2003; de Nadal et al., 2004; Mas et al., 2009; Pokho-

lok et al., 2006; Proft et al., 2006; Zapater et al., 2007). Recently,

the ERK2 MAPK was found to bind with sequence specificity

to DNA and act as a transcriptional repressor of interferon-g-

induced genes (Hu et al., 2009). Moreover, multiple MAPKs,

including ERK1/2, p38, and JNK, bind to and regulate transcrip-

tion of the insulin gene (Lawrence et al., 2009). Thus, MAPKs play

important chromatin-associated functions in the regulation of

gene expression.

MBIP (MAPK upstream kinase [MUK] binding inhibitory

protein), a component of the human ATAC complex, was

identified as a MUK binding partner in a yeast two-hybrid screen

(Fukuyama et al., 2000). The Drosophila ATAC component

CG10238 encodes the Drosophila homolog of MoaE, a subunit

of molybdopterin (MPT) synthase, an essential enzyme involved

in the synthesis of the molybdenum cofactor (Moco). Moco binds

molybdenum in the active site of molybdenum enzymes, which

catalyze redox reactions as part of ancient and conserved

biosynthetic pathways (Iyer et al., 2006; Leimkuhler et al.,

2003; Rudolph et al., 2001; Schwarz and Mendel, 2006).

However, CG10238 also contains C-terminal sequences that

726 Cell 142, 726–736, September 3, 2010 ª2010 Elsevier Inc.

Page 95: Cell 100903

are homologous to human MBIP. This observation, coupled with

the presence of MAPK signaling proteins in ATAC purifications,

led us to ask whether CG10238 and ATAC play a role in MAPK

signaling.

RESULTS

ATAC Interacts with Proteins Related to the MAPKPathwayAffinity purifications of the ATAC complex revealed proteins that

are part of the MAPK signaling pathway (Figure 1A; see Fig-

ure S1A available online). Peptides from these proteins were

found in purifications via the CG10238, CHRAC14, and D12

ATAC subunits. Peptides were identified from the transcription

factor Jra (Jun-related antigen), the Drosophila homolog of

c-Jun, and Misshapen (MSN), the Drosophila homolog of

Ste-20 kinase (Figure 1A) (Morrison et al., 2000; Su et al., 1998;

Treisman et al., 1997). We also found peptides from other STE

kinases, such as MKK4, slik, and MEK3/MKK3, as well as

peptides from Chickadee, which has been shown to interact

genetically with the JNK pathway (Jasper et al., 2001; Morrison

et al., 2000).

The ATAC Subunit CG10238 Functions as an Inhibitorof JNK Activation in Response to Osmotic StressWe tested whether CG10238 plays a role in MAPK signaling like

MBIP (Fukuyama et al., 2000). MAPK cascades can be activated

by osmotic stress, resulting in activation of JNK by phosphoryla-

tion (Kayali et al., 2000; Yang et al., 2003). We therefore exam-

ined whether expression of CG10238 affected the activation of

JNK under conditions of osmotic stress (Kayali et al., 2000).

We first titrated the cellular response to osmotic stress

stimulated by sorbitol in S2 cells by western blot (Figures S2A

and S2B). Maximum activation of JNK was observed between

7 and 30 min after treatment with a minimum concentration of

500 mM sorbitol (Figure S2A, lane 5; Figure S2B, lanes 2–4).

We then examined the level of JNK activation in cells expressing

CG10238 in the presence or absence of osmotic stress induced

by 500 mM sorbitol for 12 min. Expression of CG10238 was

inducible by CuSO4, and parental S2 cells were treated with

A

HA-CG10238CuSO4 (x10µM)

SorbitolParental cells

- - - -

- - - - - + + + + + + + - - - + + - - -

- 15 - 5 15 - 15 0 5 15

anti-ActiveJNK53.9

1 2 3 4 5 6 7 8 9 10

anti-HA 53.9

anti-JNK53.9

BDrosophila melanogaster

Homo sapiens

MoaE MBIP367 a.a.1 137 171

Mocs2B

MBIP

1 188 a.a.

MoaE1 344 a.a.

N

C

-HF1 137a.a.

-HF137a.a. 367a.a.

Design of deletion mutants of CG10238

FL1 367a.a.

C

Purifications

CG10238 Atac2 CHRAC14 Ada2a WDS D12 ControlProteins PEP (SC%) PEP (SC%) PEP (SC%) PEP (SC%) PEP (SC%) PEP (SC%) PEP (SC%)

ATAC CG10238 29 (72.21) 5 (21.25) 2 (9.81) 7 (25.61) 6 (30.25) 16 (53.13) 0 (0)Gcn5/KAT 35 (46.69) 4 (6.15) 4 (6.4) 5 (8.24) 4 (7.38) 33 (33.95) 0 (0)

Atac1 14 (53.37) 2 (5.34) 2 (7.58) 6 (16.3) 0 (0) 9 (20.71) 0 (0)Ada3 21 (58.89) 1 (1.98) 1 (1.98) 2 (7.19) 1 (3.24) 10 (16.37) 0 (0)Ada2a 20 (48.2) 2 (4.93) 2 (4.93) 9 (18.6) 0 (0) 18 (36.43) 0 (0)HCF 66 (68.87) 11 (11.93) 16 (17.6) 17 (19.87) 19 (18.07) 54 (49.4) 1 (2.6)D12 37 (45.86) 7 (9.08) 11 (19.81) 10 (13.62) 5 (9.7) 31 (38.29) 0 (0)

CG30390 14 (56.60) 1 (5.88) 0 (0) 6 (26.64) 0 (0) 13 (55.36) 0 (0)Atac2/KAT14 26 (47.16) 10 (22.09) 7 (16.28) 5 (9.95) 5 (13.18) 19 (39.41) 0 (0)

CHARC14 5 (62.5) 1 (10.16) 1 (10.16) 3 (27.34) 1 (14.84) 7 (71.88) 0 (0)Atac3 14 (40.88) 1 (3.89) 0 (0) 3 (7.61) 0 (0) 9 (20.71) 0 (0)

NC2 beta 6 (60.66) 2 (14.21) 1 (7.1) 0 (0) 2 (14.21) 5 (27.32) 0 (0)WDS 18 (66.48) 7 (25.76) 5 (18.28) 8 (26.59) 20 (63.43) 13 (53.19) 1 (3.6)

Proteins Jra 2 (10.03) 0 (0) 3 (19.03) 0 (0) 0 (0) 2 (13.84) 0 (0)related to Misshapen 16 (21.34) 0 (0) 5 (6.12) 0 (0) 0 (0) 5 (5.85) 0 (0)

JNK pathway MKK4 4 (15.8) 0 (0) 0 (0) 0 (0) 0 (0) 0 (0) 0 (0)slik 5 (4.82) 0 (0) 0 (0) 0 (0) 0 (0) 0 (0) 0 (0)

MEK3/MKK3 2 (8.08) 0 (0) 0 (0) 0 (0) 0 (0) 0 (0) 0 (0)Chickadee 2 (20.63) 0 (0) 0 (0) 0 (0) 0 (0) 0 (0) 0 (0)

Figure 1. Peptides from Proteins Related to

the JNK Pathway Were Detected in Purifica-

tions of ATAC Subunits, and CG10238

Inhibits JNK Activation by Osmotic Stress

in Drosophila Schneider’s S2 Cells

(A) Peptides of MAPK pathway proteins in addition

to ATAC subunits detected by MudPIT analysis for

affinity purifications of ATAC subunits with FLAG-

HA-tagged CG10238, Atac2/KAT14, Ada2a,

CHRAC14, WDS, and D12. Parental S2 cells

were used in mock purification as a control. PEP,

peptide count; SC%, sequence coverage (%).

(B) Homology domains of CG10238 and experi-

mental design of truncation mutants. The diagram

shows the domains of CG10238 that are homolo-

gous to human MOCS2B (hMoaE) and to human

MBIP, which are separate proteins. The full-length

(FL), N-terminal that contains only MoaE domain

(N), and C-terminal that contains only MBIP protein

sequences (C) were stably expressed in S2 cells

(Figure 2; see Figure S1B).

(C) CG10238 inhibited the activation of JNK stimu-

lated by osmotic stress in S2 cells. Stably ex-

pressed FLAG-HA-tagged CG10238 was induced

by 0, 50, or 150 mM CuSO4 in S2 cells. Cells were

stimulated with osmotic stress by addition of

500 mM sorbitol for 12 min before harvesting cells

(C, lanes 6–10; Figure 2C, lanes 5–14; Figure 2D,

lanes 1–12; see Figures S2A–S2C). The nuclear

extracts were examined by western blot probing

with anti-HA, anti-Active JNK, and anti-JNK

(C; Figure 2C). Parental S2 cells of stable cell lines

cultured with 0 or 150 mM CuSO4 were used as

controls (C, lanes 1, 2, 6, and 7; see Figure S3).

See also Figures S1, S2, and S3.

Cell 142, 726–736, September 3, 2010 ª2010 Elsevier Inc. 727

Page 96: Cell 100903

the same amount of CuSO4 and sorbitol as controls. We first

confirmed by western blot that the expression level of HA-

tagged CG10238 in the stable cell line without induction was

similar to that of endogenous CG10238 in the parental cells

(Figure S2C). Active JNK was not detected in the absence of

sorbitol treatment, or upon induction of CG10238 by addition

of CuSO4 (Figure 1C, lanes 1 and 2). In the presence of sorbitol,

JNK was activated in the parental cells (Figure 1C, lanes 6 and 7).

However, induction of CG10238 expression by CuSO4 inhibited

the activation of JNK in a dose-dependent manner (Figure 1C,

lanes 8–10). Thus, CG10238 inhibits JNK activation by osmotic

stress in vivo. Because JNK is also activated by ultraviolet light

(UV), we examined whether the inhibitory activity of CG10238

extended to this JNK activation mechanism (Angel et al., 1988;

Devary et al., 1991; Rozek and Pfeifer, 1995). Expression of

CG10238 also inhibited activation of JNK by UV (Figure S3).

The MBIP Domain of CG10238 Is Requiredfor Incorporation into the ATAC ComplexThe finding that CG10238 prevented JNK activation when

expressed in vivo led us to ask whether CG10238 served this

function in isolation or as part of the ATAC complex. We first

examined the two parts of CG10238 to determine which might

incorporate it into the ATAC complex. We generated S2 cell lines

that stably expressed tagged truncated forms of CG10238 that

included only the N-terminal MoaE domain or the C-terminal

MBIP domain (Figure 1B). These tagged proteins were affinity

purified from the stable cell lines after their expression levels

were normalized (Figure S1B). Proteins that copurified with

each domain of CG10238 were identified by multidimensional

protein identification technology (MudPIT) analysis and

confirmed by western blots (Figures 2A and 2B; Figure S1A).

All ATAC subunits except CHRAC14 associated with the MBIP

domain. A few peptides from two of the ATAC subunits were de-

tected by purification with the MoaE domain and three subunits

were detected weakly by westerns (Figure 2B). Thus, the MBIP

domain of CG10238 is responsible for incorporating the protein

into the ATAC complex.

The MoaE Domain of CG10238 Inhibits JNK ActivationAnalysis of the domain structure of MoaE and MBIP orthologs in

a variety of organisms indicates that many insects encode these

two domains as a translational fusion like that of Drosophila

CG10238. By contrast, orthologs in nematodes, plants, fungi,

- F N C F N C - Input nuclear extracts Elution

anti-HA

anti-CG30390

anti-Gcn5

anti-Atac2

anti-D12

anti-Ada2a

anti-NC2 beta

anti-Hcf

anti-WDS

anti-Atac3

anti-Atac1

29.4

37.7

(KDa)

54.8

37.6

97.3

118.2

54.8

118.2

97.3

206.3

29.4

54.8

97.3

118.2

17.5

A B

CG10238-Domain N CATAC subunits PEP( SC%) PEP( SC%)

CG10238-PA 7 (25.6) 17 (50.95)Gcn5/KAT 0 (0) 8 (12.67)

Atac1 0 (0) 3 (9.55)Ada3 0 (0) 3 (8.99)

Ada2a 0 (0) 5 (13.09)HCF 5 (1.47) 24 (26.8)D12 0 (0) 15 (21.26)

CG30390 0 (0) 3 (13.15)Atac2/KAT14 0 (0) 5 (10.98)

CHRAC14 0 (0) 0 (0)Atac3 0 (0) 2 (6.55)

NC2beta 0 (0) 3 (14.21)WDS 1 (3.88) 6 (27.42)

D

Active JNK

0

20

40

60

80

100

120

140

160

1 2 3 4 5 6 7 8 9 10 11 12

Rat

io (%

)

P PHA-CHA-FL HA-N

P

C

anti-HA

anti-JNK

- + - - - - - - - - - HA-FL

Sorbitol - - - - + + + + + + + + + + + - - - + - - - - - - - - -

HA-NHA-C

Parental cells

- - - + - - - - - - -

1 2 3 4 5 6 7 8 9 10 11 12 13 14

anti-ActiveJNK

- - + - - - - - - - -

Figure 2. MBIP-Related Sequences Incor-

porate CG10238 into ATAC; However, the

MoaE Domain Inhibits JNK Activation

(A) FLAG-HA-tagged N terminus that contains only

the MoaE domain (N) and C terminus that contains

only MBIP protein sequences (C) were stably ex-

pressed in S2 cells purified by FLAG affinity beads.

The copurified proteins were examined by MudPIT

analysis (see Figures S1A and S1B). The observed

PEP and SC (%) of each sample are indicated as in

Figure 1.

(B) The copurified proteins of CG10238-full-length

(F), -N, or -C (Figure 1B; Figure S1) were analyzed

by western blotting with antibodies against HA tag

and ATAC subunits. As a control, untransfected

parental S2 cells were mock purified and analyzed

by western blotting (-).

(C and D) The MoaE domain inhibits JNK activa-

tion. The ability to inhibit JNK activation was

compared between FLAG-HA-tagged CG10238

(FL), CG10238-N (MoaE domain), and CG10238-

C (MBIP domain) after the expression levels were

normalized (see also Figure S1B). Parental S2 cells

cultured with 150 mM CuSO4 were used as a posi-

tive control (C, lanes 1 and 5; D, lanes 1, 5, and 9).

Osmotic stress was induced as in Figure 1C

(see also Figures S2A and S2B). For sorbitol-stim-

ulated samples, the intensity of each band for

Active JNK in western blotting was quantified,

and each sample from CG10238 (FL), CG10238-N,

and CG10238-C was individually compared with

the intensity of parental cells (P), and the ratios

(%) are shown in (D). The average of four indepen-

dent experiments is graphed. Error bars represent

standard deviation.

See also Figures S1 and S2.

728 Cell 142, 726–736, September 3, 2010 ª2010 Elsevier Inc.

Page 97: Cell 100903

and all prokaryotes except parasitic bacteria have a MoaE

homology domain but are missing the MBIP domain. Sequence

database searches using the PSI-BLAST program (Altschul

et al., 1997) with human MBIP (GenBank accession number

119586267) as query revealed sequence matches between the

N-terminal portion of MBIP orthologs from Metazoa and MoaE

proteins. Thus, the N-terminal sequences of mammalian MBIPs

are related to the MoaE sequences. Hence, our bioinformatic

analysis suggests the N-terminal sequences of MBIP evolved

from MoaE and raises the possibility that the MoaE domain of

CG10238 contributes to the JNK inhibition activity.

We generated S2 cell lines that stably expressed tagged

truncated forms containing the MoaE domain (N-terminal) or

MBIP domain (C-terminal) (Figure S1B), and then compared

A

(KDa)- + - +dsRNA-Atac2

dsRNA-Atac2

dsRNA-Cont

dsRNA-Cont

118.1

54.8

54.8

54.8

SorbitolTransfected with

anti-Atac2

anti-ActiveJNK

anti-tubulin

anti-JNK

0

100

200

300

400

500

600

700

800

1 2 3 40

20

40

60

80

100

120

JNKActJNKAtac2

dsRNA-Atac2dsRNA-Cont.

Sorbitol - + - +

dsRNA-Atac2

Rat

io-J

NK

or R

atio

-Act

JNK

C

B

dsRNA-NC2beta

dsRNA-NC2beta

0

500

1000

1500

2000

2500

3000

1 2 3 40

20

40

60

80

100

120

JNKActJNK

NC2beta

dsRNA-Cont.Sorbitol - + - +

Rat

io-J

NK

or R

atio

-Act

JNK

D

S2 Cells S2 Cells

293T Cells 293T Cells

dsRNA-CSRP2BP

0

100

200

300

400

500

600

1 2 3 40

20

40

60

80

100

120

140

JNK

ActJNK

CSRP2BP

Sorbitol - + - +dsRNA-Cont dsRNA-CSRP2BP

Rat

io-C

SR

P2B

P

Rat

io-J

NK

or R

atio

-Act

JNK

dsRNA-MBIP

0

50

100

150

200

250

300

350

1 2 3 40

20

40

60

80

100

120

140

JNK

ActJNK

MBIP

Rat

io-M

BIP

Sorbitol - + - +

dsRNA-Cont dsRNA-MBIP

Rat

io-A

tac2

Rat

io-N

C2

beta

Rat

io-J

NK

or R

atio

-Act

JNK

Transfected with

anti-NC2beta

dsRNA-NC2beta

anti-ActiveJNK

Sorbitol - + - +

anti-tubulin

anti-JNK

dsRNA-NC2beta

dsRNA-Cont

dsRNA-Cont

54.8

54.8

54.8

29.4

(KDa)

Sorbitol - + - +Transfected with dsRNA-Cont

dsRNA-Cont

dsRNA-CSRP2BP

dsRNA-CSRP2BP

anti-Atac2/CSRP2BP

anti-Active JNK

anti-tubulin

anti-JNK

80

46

46

46

(KDa)

anti-MBIP

anti-tubulin

anti-JNK

anti-Active JNK

Sorbitol - + - +dsRNA-Cont

dsRNA-Cont

Transfected with dsRNA-MBIP

dsRNA-MBIP

46

46

46

46

(KDa)

Figure 3. ATAC Inhibits JNK Signaling

S2 cells were transfected with dsRNA-Atac2, dsRNA-NC2

b, or dsRNA-Control (dsRNA-Cont) (A and B) (see

Figure S4). 293T cells were transfected with dsRNA-

hAtac2/CSRP2BP, dsRNA-MBIP, or dsRNA-Control

(dsRNA-Cont) (C and D). The knockdown cells were then

stimulated with or without 500 mM (for S2 cells) or

400 mM (for 293T cells) sorbitol as in Figure 1C (see

Figure S2D). The nuclear extracts were examined by

western blot probing with anti-Active JNK, anti-JNK,

anti-Atac2 (A and C), anti-NC2 b (B), anti-MBIP (D), or

anti-tubulin antibodies as a loading control (A–D, top).

The intensities of each band of JNK, Active JNK, Atac2,

NC2 b, hAtac2, or MBIP in western blotting were quantified

from four independent experiments and plotted as ratios to

the untreated control cells (A–D, bottom). Dark bars show

the ratio of JNK intensities, light bars show the ratio of

Active JNK intensities, and the dotted line shows the ratio

of Atac2 (A), NC2 b (B), Atac2/CSRP2BP (C), or MBIP (D)

intensities. Error bars represent standard deviation.

See also Figures S2 and S4.

the inhibitory activities of the full-length protein

to that of the MoaE and MBIP domains (Fig-

ures 2C and 2D). Full-length CG10238 and the

MoaE domain alone inhibited activation of JNK

(Figure 2C, lanes 6–11; Figure 2D, bars 1–8).

However, the MBIP domain of CG10238 failed

to prevent the activation of JNK (Figure 2C,

lanes 12–14; Figure 2D, bars 9–12). Thus, the

active portion of CG10238 in inhibition of JNK

activation is the MoaE domain, whereas the

C-terminal sequences connect this activity to

the ATAC complex.

ATAC Inhibits JNK ActivationThe MoaE domain of CG10238 was sufficient to

prevent JNK activation when expressed in vivo,

whereas the MBIP domain was required to

incorporate CG10238 into ATAC (Figures 2A–

2D). These data raised the question of whether

ATAC itself plays a role in inhibition of the JNK

pathway. To address this question, we exam-

ined JNK activation in S2 cells where endoge-

nous subunits of ATAC were knocked down by

dsRNA interference. The expression level of Atac2 was reduced

60% in cells expressing dsRNA-Atac2 (Figure 3A). Interestingly,

JNK was partially activated upon reduction of Atac2 even in the

absence of osmotic stress (Figure 3A). Moreover, activation of

JNK by osmotic stress was enhanced in the Atac2 knockdown

cells (Figure 3A). We observed similar results upon knockdown

of NC2 b or CG10238 subunits of ATAC by dsRNA (Figure 3B;

Figure S4A). Although JNK activation was not observed in D12

knockdown cells without osmotic stress, its activation was

also increased in these cells under conditions of osmotic stress

(Figure S4B). Because human MBIP was recently shown to be a

component of the human ATAC complex (Fukuyama et al., 2000;

Wang et al., 2008), it was of interest to determine whether human

ATAC inhibited the activation of JNK by osmotic stress in human

Cell 142, 726–736, September 3, 2010 ª2010 Elsevier Inc. 729

Page 98: Cell 100903

293T cells (Figure S2D). Indeed, JNK was activated to higher

levels upon knockdown of MBIP or CSRP2BP (human homolog

of Atac2) in 293T cells (Figures 3C and 3D). These data indicate

that the ATAC complex is required for complete inhibition of JNK

activation in both Drosophila and human cells.

ATAC Functions as Transcriptional Cofactorfor JNK Target GenesThe presence of peptides from the Jra transcription factor in

MudPIT analyses of the affinity-purified ATAC complex (Fig-

ure 1A) suggests that ATAC interacts with Jra and may serve

as a transcriptional cofactor. Indeed, endogenous Jra was able

to coimmunoprecipitate with Atac2, CG10238, and NC2 b, sub-

units of ATAC (Figure 4A). We also sought to confirm the interac-

tion of MSN (Figure 1A), the Drosophila Ste-20 kinase, with ATAC

A B

C

IP with

anti-CG10238

- IgG Jra

IP Input

anti-Jra

1 2 3

anti-Atac2

anti-NC2 beta

Jra Sorbitol (+)

0

1

2

3

4

1 2 3 4 5 6

Jra/

Con

trol

Jra Sorbitol (-)

00.20.40.60.8

11.21.41.61.8

2

1 2 3 4 5 6

Jra/

Con

trol

Chickadee Sorbitol (-)

00.20.40.60.8

11.21.41.61.8

2

1 2 3 4 5 6

Chi

ckad

ee/C

ontr

ol

Chickadee Sorbitol (+)

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1 2 3 4 5 6

Chi

ckad

ee/C

ontr

ol

LacZAtac2

NC2b Jra

CG10238MSNRNAi

LacZAtac2

NC2b Jra

CG10238MSNRNAi

LacZAtac2NC2b Jra

CG10238MSNRNAi

LacZAtac2

NC2b Jra

CG10238MSNRNAi

D

IP with MSN- IgG

IP Input

anti-Atac2

anti-NC2beta

anti-MSN

anti-D12

anti-CG10238

anti-JNK

1 2 3

ChIP-Jra on Jra Sorbitol (-)

0.00%

10.00%

20.00%

30.00%

40.00%

50.00%60.00%

70.00%

80.00%

90.00%

100.00%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2 dsRNA-CG10238 dsRNA-MSN

Probe

ChIP-Jra on Jra Sorbitol (+)

0.00%

10.00%

20.00%

30.00%

40.00%

50.00%

60.00%

70.00%

80.00%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2 dsRNA-CG10238 dsRNA-MSN

Probe

Figure 4. ATAC Is Required for the Transcription

Regulation of JNK Target Genes

(A and B) Extracts from S2 cells endogenously expressing

ATAC subunits and JNK (B) were immunoprecipitated with

antibodies to endogenous Jra (A) or Misshapen (MSN) (B).

The presence of ATAC subunits, MSN, and JNK in the

input and immunoprecipitates (IP) was detected by

western blots.

(C) Quantitative real-time RT-PCR (qRT-PCR) of Jra

and Chickadee mRNA in S2 cells transfected with

dsRNA-LacZ (Control), dsRNA-Atac2, dsRNA-CG10238,

dsRNA-NC2 b (NC2b), dsRNA-Jra, or dsRNA-MSN (see

Figure S5A). Gene-specific mRNA levels from the cells

without sorbitol stimulation (C, top) and 30 min after

500 mM sorbitol stimulation (C, bottom) were measured

and normalized to RpL-32 expression (Figure S5B), and

are represented as ratios by qRT-PCR measurements.

The average of three independent experiments is graphed.

Error bars represent standard deviation.

(D) ChIP assays were performed with an antibody against

Jra from S2 cells without sorbitol stimulation (D, left) or with

sorbitol stimulation (D, right) expressing dsRNA-LacZ

(Cont), dsRNA-Atac2, dsRNA-CG10238, or dsRNA-MSN

(D) (Figures 5B–5M). The association of Jra on the probed

region of the Jra gene as indicated in Figure 5A, and input

chromatin was measured by qRT-PCR and normalized to

input, and the ratios of quantities are represented (see

Figure S6). The average of three independent experiments

is graphed. Error bars represent standard deviation.

See also Figures S2, S5, and S6.

in vivo. Endogenous Atac2, CG10238, NC2 b,

and D12 coimmunoprecipitated with endoge-

nous MSN (Figure 4B).

If ATAC serves as a cofactor for Jra, then

Jra-dependent transcription would require

ATAC and ATAC should be localized to Jra target

genes. Because the gene encoding c-Jun, the

mammalian homolog of Jra, is positively

regulated by c-Jun protein (Angel et al., 1988),

we examined whether ATAC functions in Jra

transcription. We measured the levels of Jra

transcripts by quantitative real-time RT-PCR

(qRT-PCR) of control and ATAC knockdown S2

cells. As a control, we confirmed that Jra tran-

script levels were reduced in cells expressing dsRNA-Jra

compared with cells expressing control-dsRNA-LacZ (Figure 4C,

upper left). Importantly, Jra transcript levels were significantly

reduced in Atac2 and CG10238 knockdown cells and slightly

reduced in NC2 b knockdown cells (Figure 4C, upper left). We

did not detect effects on Jra transcripts in MSN knockdown cells.

We confirmed that the transcripts of atac2, CG10238, nc2 b, and

msn were reduced by expression of each dsRNA-Atac2, dsRNA-

CG10238, dsRNA-NC2 b, and dsRNA-MSN by qRT-PCR (Fig-

ure S5A). We also observed that the expression of Atac2 was

increased in Jra knockdown (see below).

We next examined chickadee, another Jra target gene

(Jasper et al., 2001). Transcript levels of chickadee decreased

significantly upon knockdown of Atac2 and somewhat more

modestly upon knockdown of NC2 b or Jra (Figure 4C).

730 Cell 142, 726–736, September 3, 2010 ª2010 Elsevier Inc.

Page 99: Cell 100903

Transcript levels of chickadee were not reduced significantly in

CG10238 knockdown cells. Of note, we did not observe changes

in the transcript levels of target genes of a different H4K16

acetyltransferase complex (MSL) upon knockdown of ATAC,

indicating that requirement for ATAC is specific for Jra and

chickadee (Figure S5B).

The JNK cascade is activated by osmotic stress and increases

c-Jun expression in mammals. Thus, we examined Jra and

chickadee transcript levels in ATAC knockdown cells upon

treatment with sorbitol. Interestingly, Jra transcripts were not

reduced in Atac2 knockdown cells upon osmotic stress but

were increased in CG10238 knockdown cells. The expression

of Chickadee was also increased in ATAC and MSN knockdown

cells under osmotic stress. Thus, ATAC positively impacts the

expression of JNK target genes in the absence of osmotic stress

but plays a predominantly negative role at JNK target genes

under conditions of osmotic stress. Moreover, the positive

effects of ATAC on expression of JNK target genes are

mediated primarily by Atac2, whereas its negative effects are

most strongly dependent on CG10238.

ATAC Functions as a Cofactor for the JNK Target GeneJra, and Coordinates Upstream MAPKs onto Jra

Previous studies have shown the direct association of MAPKs

with gene loci that they regulate (de Nadal and Posas, 2010;

Dioum et al., 2009). Thus, we measured the occupancy of

ATAC and upstream kinases directly on JNK target genes. We

examined the occupancy of ATAC on the Jra gene in ATAC

and MSN knockdown cells by chromatin immunoprecipitation

(ChIP) assay. Previous studies have shown that Jun can form a

homodimer or heterodimerize with ATF or Fos to form the AP-1

transcription factor. The c-Jun promoter contains AP-1-like

sequences (TTACCTCA and TGACATCA) in addition to an

NF-jun binding sequence (GGAGTCTCC) (Rozek and Pfeifer,

1993). c-Jun was found to occupy the c-Jun promoter without

external stimulus (Angel and Karin, 1991). We confirmed that

Jra actually bound to the Jra gene by ChIP assay (Figure 4D,

left). Jra was found to significantly occupy a site in the middle

of the coding region that contains an NF-jun recognition

sequence (Figure 4D, left). We analyzed regions of the Jra

gene that are comparable to the regulatory regions of the human

c-Jun gene (Figure 5A). We probed ATAC occupancy, with and

without osmotic stress, at an upstream enhancer (primer A),

the promoter containing the AP-1 binding sequence (primer B),

the 50 region of the coding sequence (primer C), the middle

of the coding sequence that contained an NF-jun binding

sequence (GGAGGCACC) (primer D), and a region 30 of the

coding sequence (primer E) (FlyBase and UCSC Genome

Browser) (Figure 5A) (Rozek and Pfeifer, 1993).

In the absence of osmotic stress, Atac2 was found to

occupy the enhancer and promoter regions of Jra (Figure 5B,

dsRNA-Cont). As expected, Atac2 occupancy of these regions

was reduced in the Atac2 knockdown cells (Figure 5B). Atac2

occupancy of the promoter was reduced in CG10238 but not

MSN knockdown cells. As controls for the specificity of ATAC

occupancy of Jra, we performed an ATAC-ChIP assay on two

target genes of the MSL complex (Figures S5 and S6). Back-

ground signals on these genes were not affected by the

knockdown of Atac2 or CG10238. We next investigated

whether ATAC occupancy affected acetylation of the nucleo-

some on the Jra gene by a ChIP assay for H4K16 acetylation.

In control cells without osmotic stress, both the enhancer and

promoter regions were enriched in H4K16 acetylation, which

was significantly reduced in the Atac2 and CG10238 knock-

down cells (Figure 5D). Thus, in the absence of osmotic

stress, ATAC occupied the enhancer and promoter of Jra,

and H4K16 acetylation at these sites was dependent on

ATAC.

In the presence of osmotic stress, ATAC occupancy and

H4K16 acetylation of the promoter were similar but were

reduced at the upstream enhancer (Figures 5C and 5E). Interest-

ingly, ATAC occupancy and H4K16 acetylation of the coding

region at primer D were stimulated by osmotic stress (Figures

5C and 5E). Jra was clearly localized on this region with or

without osmotic stress (Figure 4D). Jra occupancy at primer D

was strongly reduced in Atac2, CG10238, and MSN knockdown

cells in the absence of sorbitol. However, under osmotic stress,

Jra occupancy required primarily CG10238 (Figure 4D). Thus,

under conditions of osmotic stress, ATAC occupancy at the

upstream enhancer was reduced and instead increased in the

coding region where Jra was bound.

We next tested whether stress-activated upstream kinases

were recruited to the Jra gene. We probed Active JNK (phos-

pho-JNK) and phospho-MKK4/SEK1 (Ser257/Thr261) occu-

pancy on the Jra gene. We also tested for the occupancy of

the MSN kinase. In the absence of osmotic stress in control cells,

each of these kinases could be found on the Jra gene although

their locations varied. Active JNK was found at the enhancer

(primer A), the promoter (primer B), and downstream of the

coding region (primer E) (Figure 5F). Phospho-MKK4 was local-

ized to the enhancer, promoter, and downstream of the coding

region (primer E) and MSN was localized to the promoter (primer

B) (Figures 5H and 5J). Strikingly, like ATAC upon osmotic stress,

occupancy of each of these kinases was increased in the coding

region (primer D), the site of Jra binding (Figures 5G, 5I, and 5K).

This was most significant for Active JNK, which bound strongly

to the primer D region upon osmotic stress (note the difference

in scale between Figures 5F and 5G), and for MSN, which relo-

calized from the promoter to the coding region (primer D)

(Figures 5J and 5K).

The ChIP results from the knockdown cell lines indicate a

dynamic interplay between ATAC and the upstream kinases.

For example, in the absence of osmotic stress, H4K16 acetyla-

tion at the promoter (primer B) was independent of MSN (Fig-

ure 5D); however, upon osmotic stress, MSN suppressed

H4K16 acetylation at the promoter but was required for

maximum acetylation in the coding region (primer D) (Figure 5E).

In the absence of osmotic stress, the binding of Active JNK to the

enhancer, promoter, and downstream of the coding region

required ATAC but was suppressed by MSN (Figure 5F). By

contrast, under osmotic stress, the strong binding of Active

JNK to the coding region (primer D) required both ATAC and

MSN (Figure 5G). In the absence of osmotic stress, binding of

phospho-MKK4 to the enhancer was suppressed by ATAC and

MSN (Figure 5H) but upon osmotic stress, binding to the

promoter required MSN and CG10238 specifically (Figure 5I).

Cell 142, 726–736, September 3, 2010 ª2010 Elsevier Inc. 731

Page 100: Cell 100903

Finally, binding of MSN to the promoter is suppressed by ATAC

in the absence of osmotic stress (Figure 5J); however, when it

moves to the coding region where Jra is bound (primer D),

upon osmotic stress its binding is independent of ATAC

(Figure 5K).

A previous study linked the acetylation of H4K16 with

H3S10 phosphorylation in the activation of the FOSL1 gene

(Zippo et al., 2009). At this gene, H3S10 phosphorylation was

reported to recruit MOF-dependent H4K16 acetylation, which

then recruited the elongation factor P-TEFb to release paused

polymerase (Zippo et al., 2009). As ATAC is also an H4K16 ace-

tyltransferase, we tested for a relationship between H3S10

phosphorylation and H4K16 acetylation, ATAC, or MSN at Jra.

H3S10P was enriched in the enhancer and promoter regions

(primer B) in the absence of osmotic stress (Figure 5L). H3S10

phosphorylation levels were suppressed by ATAC and MSN at

ChIP-ActiveJNK on Jra Sorbitol (+)

0.00%

10.00%

20.00%

30.00%

40.00%

50.00%

60.00%

70.00%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2 dsRNA-CG10238 dsRNA-MSN

ChIP-ActiveJNK on Jra Sorbitol (-)

0.00%

0.20%

0.40%

0.60%

0.80%

1.00%

1.20%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2 dsRNA-CG10238 dsRNA-MSN

ChIP-Missahpen on Jar Sorbitol (+)

0.00%

0.05%

0.10%

0.15%

0.20%

0.25%

0.30%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2 dsRNA-CG10238 dsRNA-MSN

ChIP-Misshapen on Jra Sorbitol (-)

0.00%

0.05%

0.10%

0.15%

0.20%

0.25%

0.30%

0.35%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2 dsRNA-CG10238 dsRNA-MSN

ChIP-Atac2 on Jra Sorbitol (-)

0.00%

0.20%

0.40%

0.60%

0.80%

1.00%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2 dsRNA-CG10238 dsRNA-MSN

A

B

ChIP-H3S10P on Jra Sorbitol (-)

0.00%

0.50%

1.00%

1.50%

2.00%

2.50%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2 dsRNA-CG10238 dsRNA-MSN

ChIP-H4K16ac on Jra Sorobitol (-)

0.00%

0.10%

0.20%

0.30%

0.40%

0.50%

0.60%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2 dsRNA-CG10238 dsRNA-MSN

C

L

K

G

H I

ED

F

J

M

Probes

Jra gene span

CodingPromoterA C DB

Enhancer ECoding

Jra CDS

Enhancer100bp

AP-1NF-jun

0.00%

0.05%

0.10%

0.15%

0.20%

0.25%

0.30%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2 dsRNA-CG10238 dsRNA-MSN

ChIP-pMKK4 on Jra Sorbitol (-)

0.00%

0.05%

0.10%

0.15%

0.20%

0.25%

0.30%

0.35%

0.40%

0.45%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2 dsRNA-CG10238 dsRNA-MSN

ChIP-pMKK4 on Jra Sorbitol (+)

ChIP-H3S10P on Jra Sorbitol (+)

0.00%

2.00%

4.00%

6.00%

8.00%

10.00%

12.00%

14.00%

16.00%

18.00%

20.00%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2 dsRNA-CG10238 dsRNA-MSN

ChIP-Atac2 on Jra Sorbitol (+)

0.00%

0.20%

0.40%

0.60%

0.80%

1.00%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2 dsRNA-CG10238 dsRNA-MSN

ChIP-H4K16ac on Jra Sorbitol (+)

0.00%

1.00%

2.00%

3.00%

4.00%

5.00%

6.00%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2 dsRNA-CG10238 dsRNA-MSN

ChIP-Activevv JNK on Jra Sorbitol (+)

0.00%

10.00%

20.00%

30.00%

40.00%

50.00%

60.00%

70.00%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-AtAA ac2 dsRNA-CG10238 dsRNA-MSN

ChIP-Activevv JNK on Jra Sorbitol (-)

B C D E

dsRNA-AtAA ac2 dsRNA-CG10238 dsRNA-MSN

ChIP-Missahpen on Jar Sorbitol (+)

0.00%

0.05%

0.10%

0.15%

0.20%

0.25%

0.30%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-AtAA ac2 dsRNA-CG10238 dsRNA-MSN

ChIP-Misshapen on Jra Sorbitol (-)

A B C D E

t dsRNA-AtAA ac2 dsRNA-CG10238 dsRNA-MSN

B C D E

t dsRNA-AtAA ac2 dsRNA-CG10238 dsRNA-MSN

ChIP-H3S10P on Jra Sorbitol (-)

t dsRNA-AtAA ac2 dsRNA-CG10238 dsRNA-MSN

K

G

I

E

M

A B C D E

t dsRNA-AtAA ac2 dsRNA-CG10238 dsRNA-MSN

ChIP-pMKK4 on Jra Sorbitol (-)

0.00%

0.05%

0.10%

0.15%

0.20%

0.25%

0.30%

0.35%

0.40%

0.45%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-AtAA ac2 dsRNA-CG10238 dsRNA-MSN

ChIP-pMKK4 on Jra Sorbitol (+)

ChIP-H3S10P on Jra Sorbitol (+)

0 00%

2.00%

4.00%

6.00%

8.00%

10.00%

12.00%

14.00%

16.00%

18.00%

20.00%

ChI

P/In

put

dsRNA-Cont dsRNA-AtAA ac2 dsRNA-CG10238 dsRNA-MSN

0.00%

0.20%

0.40%

0.60%

0.80%

1.00%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-AtAA ac2 dsRNA-CG10238 dsRNA-MSN

ChIP-H4K16ac on Jra Soror bitol (-)

B C D E

dsRNA-AtAA ac2 dsRNA-CG10238 dsRNA-MSN

ChIP-H4K16ac on Jra Sorbitol (+)

0.00%

1.00%

2.00%

3.00%

4.00%

5.00%

6.00%

A B C D E

ChI

P/In

put

dsRNA-Cont dsRNA-AtAA ac2 dsRNA-CG10238 dsRNA-MSN

Figure 5. Occupancy of ATAC, H4K16

Acetylation, MAPKs, and H3S10P on the

Jra Gene with and without Osmotic Stress

(A) Schematic representation of the Jra gene

indicating the position of different probes tested

(A–E on x axis for B–M). The gray box indicates

the coding sequence (CDS), the gray circles indi-

cate the enhancer, the white circle indicates pre-

dicted AP-1 sites, and the black square indicates

the NF-jun-like sequence.

(B–M) ChIP assays were performed with anti-

bodies against Atac2, H4K16ac, Active JNK,

phospho-MKK4 (S257T261) (pMKK4), MSN, and

phospho-H3S10 (H3S10P) from ATAC and MSN

knockdown S2 cells (as in Figure 4D) without

sorbitol stimulation (B, D, F, H, J, and L) (Figures

6B, 6D, 6F, 6H, and 6J) or with sorbitol stimulation

(C, E, G, I, K, M) (Figures 6C, 6E, 6G, 6I, and 6K)

on the probed region of the Jra gene as indicated

in (A), and input chromatin was measured by

qRT-PCR and normalized to input. The ratios of

measured quantities are represented (see Figures

S5A and S6). The average of three independent

experiments is graphed. Error bars represent

standard deviation.

See also Figures S2, S5, and S6.

the promoter and strongly suppressed

downstream of the coding region

(Figure 5L). Osmotic stress substantially

increased H3S10 phosphorylation levels

across the Jra gene except for the

upstream enhancer (note the different

scales in Figures 5L and 5M). Induction

of H3S10 phosphorylation at the

promoter, the coding region (primer D),

and downstream of the coding region de-

pended on ATAC (especially CG10238);

however, MSN suppressed H3S10 phos-

phorylation across the gene (Figure 5M).

These results suggest that the relation-

ship between H3S10 phosphorylation

and H4K16 acetylation by ATAC at Jra

differs from what was observed with H3S10 phosphorylation

and H4K16 acetylation by MOF at the FOSL1 gene (Zippo

et al., 2009). Acetylation of H4K16 at the enhancer of FOSL1

by MOF required H3S10 phosphorylation; however, ATAC is

required for H3S10 phosphorylation in the presence of osmotic

stress. Moreover, ATAC-dependent H4K16 acetylation at the

Jra enhancer in the absence of osmotic stress (Figure 5D)

occurred independently of significant H3S10 phosphorylation

(Figure 5L).

ATAC Acetylates Nucleosomes and Coordinatesthe JNK Upstream Kinases with Jra on chickadee

The most striking aspect of our ChIP analysis of the Jra gene is the

movement upon osmotic stress of ATAC and MAPKs to the

coding region (primer D), which contains an NF-jun binding

sequence and is occupied by the Jra transcription factor. This

732 Cell 142, 726–736, September 3, 2010 ª2010 Elsevier Inc.

Page 101: Cell 100903

raised the question as to whether ATAC and MAPKs move to

this location because Jra is bound there or because they need

to associate with the coding region. Thus, we performed a ChIP

assay on another JNK target gene, chickadee. Chickadee does

not have an apparent NF-jun binding site in the coding region. It

does, however, have several AP-1-like sequences upstream of

the promoter and at a far-upstream enhancer (Figure 6A). We

analyzed 7 kb upstream of the chickadee gene and found five

AP-1-like sequences. We probed a region containing an AP-1-

like site, which was also near two CCAAT motifs that are 6.4 kb

upstream from the coding sequence (ChicA primer). We also

probed adjacent to the AP-1-like site 2 kb upstream from the

coding sequence close to the predicted promoter (ChicB primer)

and the middle of the coding region (ChicC primer) (Figure 6A).

ChIP analysis revealed that Jra was localized at the ChicA and

ChicB regions, suggesting Jra binds AP-1 sites on chickadee

(Figure S7A). Jra occupancy was reduced in CG10238 and

A

C

K

G

I

E

B

H

D

F

J

ChIP-Atac2 on Chickadee Sorbitol (+)

0.00%

0.10%

0.20%

0.30%

0.40%

0.50%0.60%

0.70%

0.80%

0.90%

1.00%

ChicA ChicB ChicC

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2dsRNA-CG10238 dsRNA-MSNdsRNA-Jra

ChIP-Atac2 on Chickadee Sorbitol (-)

0.00%

0.02%

0.04%

0.06%

0.08%

0.10%

0.12%

0.14%

ChicA ChicB ChicC

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2dsRNA-CG10238 dsRNA-MSNdsRNA-Jra

ChIP-H4K16ac on Chickadee Sorbitol (-)

0.00%

0.10%

0.20%

0.30%

0.40%

0.50%

0.60%

0.70%

0.80%

0.90%

ChicA ChicB ChicC

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2dsRNA-CG10238 dsRNA-MSNdsRNA-Jra

ChIP-Misshapen on Chickadee Sorbitol (+)

0.00%

0.05%

0.10%

0.15%

0.20%

0.25%

ChicA ChicB ChicC

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2dsRNA-CG10238 dsRNA-MSNdsRNA-Jra

ChIP-Misshapen on Chickadee Sorbitol (-)

0.00%

0.05%

0.10%

0.15%

0.20%

0.25%

0.30%

ChicA ChicB ChicC

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2dsRNA-CG10238 dsRNA-MSNdsRNA-Jra

ChIP-H4K16ac on Chickadee Sorbitol (+)

0.00%

0.10%

0.20%

0.30%

0.40%

0.50%

0.60%

0.70%

0.80%

0.90%

ChicA ChicB ChicC

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2dsRNA-CG10238 dsRNA-MSNdsRNA-Jra

ChIP-H3S10P on Chickadee Sorbitol (-)

0.00%0.20%0.40%0.60%0.80%1.00%1.20%1.40%1.60%1.80%2.00%

ChicA ChicB ChicC

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2dsRNA-CG10238 dsRNA-MSNdsRNA-Jra

ChIP-H3S10P on Chickadee Sorbitol (+)

0.00%

2.00%

4.00%

6.00%

8.00%

10.00%

12.00%

14.00%

16.00%

18.00%

ChicA ChicB ChicC

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2dsRNA-CG10238 dsRNA-MSNdsRNA-Jra

0.00%

0.20%

0.40%

0.60%

0.80%

1.00%

1.20%

ChicA ChicB ChicC

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2dsRNA-CG10238 dsRNA-MSNdsRNA-Jra

ChIP-pMKK4 on Chickadee Sorbitol (+)

0.00%

0.10%

0.20%

0.30%

0.40%

0.50%

0.60%

ChicA ChicB ChicC

ChI

P/In

put

dsRNA-Cont dsRNA-Atac2dsRNA-CG10238 dsRNA-MSN

K

G

I

E

0.00%

0.10%

0.20%

0.30%

0.40%

0.50%0.60%

0.70%

0.80%

0.90%

1.00%

ChicA ChicB ChicC

ChI

P/In

put

dsRNA-AA CG10238 dsRNA-AA MSNdsRNA-AA Jra

0%

2%

4%

6%

8%

0%

2%

4%

ChicA ChicB ChicC

dsRNA-AA CG10238 dsRNA-AA MSNdsRNA-AA Jra

ChIP-H4K16ac on Chickadee Sorbitol (-)

.00%

.10%

.20%

.30%

.40%

.50%

.60%

.70%

.80%

.90%

ChicA ChicB ChicC

dsRNA-AA Cont dsRNA-AA AtAA ac2dsRNA-AA CG10238 dsRNA-AA MSNdsRNA-AA Jra

ChIP-Misshapen on Chickadee Sorbitol (+)

0.00%

0.05%

0.10%

0.15%

0.20%

0.25%

ChicA ChicB ChicC

ChI

P/In

put

dsRNA-AA Cont dsRNA-AA AtAA ac2dsRNA-AA CG10238 dsRNA-AA MSNdsRNA-AA Jra

ChIP-Misshapen on Chickadee Sorbitol (-)

0.00%

0.05%

0.10%

0.15%

0.20%

0.25%

0.30%

ChicA ChicB ChicC

dsRNA-AA Cont dsRNA-AA AtAA ac2dsRNA-AA CG10238 dsRNA-AA MSNdsRNA-AA Jra

ChIP-H4K16ac on Chickadee Sorbitol (+)

0.00%

0.10%

0.20%

0.30%

0.40%

0.50%

0.60%

0.70%

0.80%

0.90%

ChicA ChicB ChicC

ChI

P/In

put

dsRNA-AA Cont dsRNA-AA AtAA ac2dsRNA-AA CG10238 dsRNA-AA MSNdsRNA-AA Jra

ChIP-H3S10P on Chickadee Sorbitol (-)

0 00%0.20%0.40%0.60%0.80%1.00%1.20%1.40%1.60%1.80%2.00%

dsRNA-AA Cont dsRNA-AA AtAA ac2dsRNA-AA CG10238 dsRNA-AA MSNdsRNA-AA Jra

ChIP-H3S10P on Chickadee Sorbitol (+)

0 00%

2.00%

4.00%

6.00%

8.00%

10.00%

12.00%

14.00%

16.00%

18.00%

ChI

P/In

put

dsRNA-AA Cont dsRNA-AA AtAA ac2dsRNA-AA CG10238 dsRNA-AA MSNdsRNA-AA Jra

0.00%

0.20%

0.40%

0.60%

0.80%

1.00%

1.20%

ChicA ChicB ChicC

ChI

P/In

put

dsRNA-AA Cont dsRNA-AA AtAA ac2dsRNA-AA CG10238 dsRNA-AA MSNdsRNA-AA Jra

ChIP-pMKK4 on Chickadee Sorbitol (+)

.00%

.10%

.20%

.30%

.40%

.50%

.60%

ChicA ChicB ChicC

dsRNA-AA Cont dsRNA-AA AtAA ac2dsRNA-AA CG10238 dsRNA-AA MSNdsRNA-Jra

ChIP-pMKK4 on Chicadee Sorbitol (-)

1K bp

AP-1Enhancer

Probes ChicCChicA ChicB

Chickadee gene spanchic CDS

Figure 6. Occupancy of ATAC, H4K16

Acetylation, MAPKs, and H3S10P on the

chickadee Gene with and without Osmotic

Stress

(A) Schematic representation of the chickadee

gene including 7 kb upstream of the transcription

start site indicating the position of different probes

on the gene (ChicA-C on x axis for B–K). The gray

box indicates the coding sequence (CDS), the gray

circles indicate the enhancer, and the white circles

indicate predicted AP-1 sites.

(B–K) ChIP assays were performed with antibodies

against Atac2, H4K16ac, pMKK4 (S257T261),

MSN, and phospho-H3S10 (H3S10P) from ATAC

and MSN knockdown S2 cells with or without

sorbitol stimulation (as in Figure 4D) on the probed

region of the chickadee gene and upstream region

as indicated in (A), and input chromatin was

measured by qRT-PCR and normalized to input

(see Figure S7A). The ratios of the measured

quantities are represented. The average of three

independent experiments is graphed. Error bars

represent standard deviation.

See also Figures S2, S5, S6, and S7.

MSN knockdown cells but not in Atac2

knockdown cells (Figure S7A). These

two regions were co-occupied by ATAC

and H4K16 acetylation (Figures 6B and

6D). The presence of Atac2 and H4K16

acetylation at these regions was not

affected by the knockdown of Jra, sug-

gesting ATAC does not require Jra to

bind the chickadee gene in the absence

of osmotic stress. Interestingly, in the

presence of osmotic stress, ATAC occu-

pancy and H4K16 acetylation were

increased at the ChicA region, and this

increased occupancy was reduced in

Jra knockdown cells (Figures 6C and 6E).

Phospho-MKK4 and MSN also occupied

the ChicA and ChicB regions, and these

kinases were also found in the open reading frame (ChicC) in

the absence of osmotic stress (Figures 6F and 6H). In contrast

to what was seen at the Jra gene, on chickadee the kinases

move to the upstream ChicA region upon osmotic stress and their

occupancy is largely (MKK4) or slightly (MSN) dependent on Jra

(Figures 6G and 6I). Occupancy of ChicA by the kinases under

these conditions is slightly suppressed by Atac2 but completely

dependent on CG10238. Indeed, CG10238 was crucial for the

occupancy of ATAC, MKK4, and MSN on the far-upstream

ChicA region under conditions of osmotic stress (Figures 6C,

6G, and 6I). At the Jra gene, occupancy of the primer D region

containing the NF-Jun binding site by ATAC, Active JNK, and

phospho-MKK4 upon osmotic stress was dependent on

CG10238. As seen on Jra, H3S10P on chickadee was sup-

pressed by ATAC and MSN in the absence of osmotic stress

and, although H3S10P increased upon osmotic stress, these

levels were suppressed by MSN specifically (Figures 6J and 6K).

Cell 142, 726–736, September 3, 2010 ª2010 Elsevier Inc. 733

Page 102: Cell 100903

The models presented in Figure 7 summarize the molecular

events occurring on Jra and chickadee as revealed by the ChIP

data. Although differing in details having to do with the architec-

ture of each gene, comparison of the expression and ChIP data

from Jra and chickadee reveals strikingly common features

about the role of ATAC at these genes. First, in the absence of

osmotic stress, ATAC is required for the basal expression of

these genes, which is most highly dependent on the Atac2 ace-

tyltransferase subunit. Second, upon osmotic stress, ATAC

suppresses induced transcription levels of these genes, which

is most highly dependent on the CG10238 subunit. Third, upon

osmotic stress, ATAC and the MAPKs move to the region of

the gene where Jra is bound. And fourth, the relocalization of

ATAC and MAPKs is dependent on CG10238, the MAP kinase

upstream kinase binding inhibitory protein. Thus, ATAC coordi-

nates the occupancy of Jra and upstream kinases while it

controls the levels of JNK activation and target gene induction.

DISCUSSION

ATAC shares four subunits with the SAGA transcription coactiva-

tor complex and was anticipated to function as a coactivator.

However, preliminary studies indicated that ATAC did not

interact with the activation domains of VP16 or p53 under

conditions in which SAGA bound (Kusch et al., 2003). Prompted

by the identification of peptides from the Jra (Drosophila c-Jun)

transcription factor in ATAC affinity purifications, we demon-

strated that Jra coimmunoprecipitates with ATAC subunits. We

further demonstrate that ATAC is localized to Jra target genes

(Jra and chickadee) and is required for H4K16 acetylation at

the Jra and chickadee enhancer and promoter. ATAC was

required for basal levels of expression of these genes, which

depended strongly on the Atac2 acetyltransferase subunit.

Thus, to our knowledge, this is the first report to demonstrate

that ATAC functions as a transcription cofactor.

Our pursuit of the functions of the CG10238 subunit of ATAC

and the potential role of its MoaE domain led us to discover

that ATAC is intimately integrated into MAP kinase signaling

and target gene expression. CG10238 was found to inhibit

JNK activation by osmotic stress, an activity mediated through

the MoaE domain. Thus, structural features of this protein that

are utilized for molybdopterin synthesis have been conscripted

in ATAC to function in the regulation of MAP kinase signaling.

The MSN kinase is the most likely direct target of CG10238, as

it coimmunoprecipitates with ATAC subunits. However, we

cannot rule out potential interactions with other kinases; for

example, peptides from MKK4 have been found in ATAC purifi-

cations (Figure 1). Under conditions of osmotic stress, ATAC

suppresses the level of target gene expression in a manner

strongly dependent on CG10238. We believe this is due to inhi-

bition of JNK activation rather than a direct negative effect of

ATAC on transcription. In fact, ATAC may still function as a posi-

tive cofactor for induced Jra expression, as the levels of H3S10P

on the gene, considered a positive mark for transcription

A

B

D

C

Jra

Jra

Osmotic stress

MSN

CG10238

ATACAtac2

MKK4-P

JNK-P

K16acS10P K16ac

Jra-P

MSN

E AP-1 NF-jun TATA-like

E AP-1 NF-jun TATA-likeJra

JNKMKK4Jra-P

S10PK16ac

JNK

MKK4

CG10238

Atac2

ATAC

Jra

MSN

MKK4

MSN

Jra

CG10238

JNK

K16acS10P

Chickadee

E

AP-1

AP-1E

Osmotic stress

MSN

MKK4-P

JNK-PK16acS10PS10PS10PJra-P

ATAC

E

AP-1

E

Chickadee

AP-1

K16acS10P

MKK4

MSN

JNKJra

CG10238

K16ac

Jra

ATAC

Figure 7. Models Summarizing ATAC Functions and the Molecular

Events Occurring on Jra and chickadee as Revealed by the ChIP

Data

Summaries of ChIP data on the Jra gene without and with osmotic stress are

shown in (A) and (B), respectively. Summaries of ChIP data on the chickadee

gene without and with osmotic stress are shown in (C) and (D), respectively.

(A) In the absence of osmotic stress, ATAC occupies and acetylates H4K16

(K16ac) at the enhancer (E) and promoter regions, containing AP-1 sites

(AP-1), of the Jra gene and is required for basal levels of Jra transcription (small

arrow). ATAC blocks the recruitment of MKK4 to the enhancer and MSN to the

promoter while inhibiting the activation of JNK. The phosphorylation of H3S10

(S10P) on the promoter region is suppressed by MSN.

(B) When cells are exposed to osmotic stress, the stress-activated kinases in

the JNK cascade are recruited to the Jra binding motif (NF-jun) on the Jra gene

by ATAC, especially dependent on CG10238. In addition to interacting with the

promoter region, ATAC also interacts with the NF-jun site by interaction with

Jra and further acetylates H4K16 at this site. ATAC continues to limit the extent

of JNK activation. Phosphorylation of H3S10 is increased across the Jra gene

by osmotic stress but is still suppressed by MSN. Jra transcription is rapidly

and transiently induced (large arrow).

(C) The chickadee gene does not have an apparent NF-jun site. In the absence

of osmotic stress, Jra, ATAC, and acetyl-H4K16 are found at two AP-1 sites,

which are 2 and 7 kb upstream from the coding sequence. Upstream kinases

occupy these sites and the downstream coding region.

(D) In the presence of osmotic stress, ATAC preferentially relocalizes to the

far-upstream AP-1 site and recruits Jra to this location. Phospho-MKK4

(MKK4-P) and MSN are also relocalized to this far-upstream AP-1 site in

a CG10238- and Jra-dependent manner. The CG10238 subunit of ATAC is

required for ATAC, Jra, MKK4-P, and MSN relocalization to the far-upstream

enhancer. Transcription of chickadee is activated (large arrow) whereas

the levels of phosphorylated H3S10 are suppressed by MSN across the

chickadee gene.

See also Figure S7.

734 Cell 142, 726–736, September 3, 2010 ª2010 Elsevier Inc.

Page 103: Cell 100903

(Zippo et al., 2009), are dependent on ATAC (Figure 5M). In

a manner strongly dependent on CG10238, ATAC plays a striking

role in the binding of MAPKs to the Jra and chickadee genes.

Upon osmotic stress, ATAC was required for the recruitment of

the JNK, MKK4, and MSN kinases to the site of Jra binding.

The only exception was the binding of MSN to the Jra gene

(Figure 5K). However, MSN required CG10238 for binding chick-

adee under inducing conditions (Figure 6I). The crucial role of

CG10238 in both the recruitment of the kinases and the inhibition

of signaling is most consistent with a model by which ATAC

inhibits JNK activation at the downstream target genes. The

fact that ATAC and the MAPKs all move to the site of Jra binding

suggests that a localized signaling network is organized at the

site occupied by the downstream transcription factor.

Our data indicate that a positive cofactor of JNK target gene

transcription, ATAC, also serves to negatively feed back for inhi-

bition of the JNK signaling pathway. Why would a downstream

activator also repress upstream signaling? The JNK pathway

and c-Jun have been shown to suppress p53-dependent

apoptosis and cellular senescence (Alexaki et al., 2008; Butter-

field et al., 1997; Das et al., 2007; Fan et al., 2010; Ham et al.,

2000; Takekawa and Saito, 1998; Yang et al., 1997). Moreover,

c-Jun has been shown to repress transcription of the p53 gene

and hence its downstream targets (Schreiber et al., 1999).

Thus, a delicate balance of JNK signaling and p53 expression

controls cell growth, proliferation, and apoptosis. If ATAC

suppression of JNK is involved in this process, we might expect

loss of ATAC to result in decreased expression of p53 due to

increased JNK activity (Schreiber et al., 1999). Indeed, whereas

knockdown of Jra increased p53 expression in S2 cells, knock-

down of ATAC subunits resulted in a decrease in p53 expression

consistent with ATAC suppression of JNK activation (Figure S7B).

Our results reveal a mechanism for governing the transcrip-

tional response to signaling pathways. By serving as both a

downstream transcriptional cofactor and as an inhibitor of

upstream signaling, ATAC serves as a master regulator which

administers the appropriate level of response to the inducing

signals.

EXPERIMENTAL PROCEDURES

Cell Lines, Extract Preparation, and Complex Purification

The nuclear extracts from S2 cell lines were generated as described in

Extended Experimental Procedures and were used for FLAG affinity purifica-

tion as previously described (Suganuma et al., 2008).

Chromatin Immunoprecipitation Assay

The crosslinked S2 cell pellets were washed, resuspended, and sonicated,

and the DNA was immunoprecipitated with antibodies using Dynabeads

(Invitrogen). The bound DNA and input DNA were incubated with RNaseA

and the crosslinking was reversed. The ethanol-precipitated DNA pellets

were resuspended and then were analyzed by quantitative real-time RT-PCR.

Antibodies, buffer content, and primers for ChIP-qRT-PCR are described in

the Extended Experimental Procedures.

Osmotic Stress Response in dsRNAi Knockdown S2 Cells

and siRNA Knockdown 293T Cells

The dsRNAi was transfected in S2 cells. The siRNA was transfected in 293T

cells as described in the Extended Experimental Procedures. The cells were

incubated with 500 mM sorbitol for 12 min before harvest.

Multidimensional Protein Identification Technology Analysis

MudPIT analysis was performed as described 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.

2010.07.045.

ACKNOWLEDGMENTS

We thank the Workman Lab members, Bjoern Gaertner in the Zeitlinger Lab,

and Hidehisa Takahashi in the Conaway Lab for advice and support during

this project. This research was supported by the Stowers Institute for Medical

Research.

Received: December 30, 2009

Revised: May 14, 2010

Accepted: July 1, 2010

Published: September 2, 2010

REFERENCES

Alepuz, P.M., Jovanovic, A., Reiser, V., and Ammerer, G. (2001). Stress-

induced MAP kinase Hog1 is part of transcription activation complexes. Mol.

Cell 7, 767–777.

Alepuz, P.M., de Nadal, E., Zapater, M., Ammerer, G., and Posas, F. (2003).

Osmostress-induced transcription by Hot1 depends on a Hog1-mediated

recruitment of the RNA Pol II. EMBO J. 22, 2433–2442.

Alexaki, V.I., Javelaud, D., and Mauviel, A. (2008). JNK supports survival in

melanoma cells by controlling cell cycle arrest and apoptosis. Pigment Cell

Melanoma Res 21, 429–438.

Altschul, S.F., Madden, T.L., Schaffer, A.A., Zhang, J., Zhang, Z., Miller, W.,

and Lipman, D.J. (1997). Gapped BLAST and PSI-BLAST: a new generation

of protein database search programs. Nucleic Acids Res. 25, 3389–3402.

Angel, P., and Karin, M. (1991). The role of Jun, Fos and the AP-1 complex in

cell-proliferation and transformation. Biochim. Biophys. Acta 1072, 129–157.

Angel, P., Hattori, K., Smeal, T., and Karin, M. (1988). The jun proto-oncogene

is positively autoregulated by its product, Jun/AP-1. Cell 55, 875–885.

Butterfield, L., Storey, B., Maas, L., and Heasley, L.E. (1997). c-Jun

NH2-terminal kinase regulation of the apoptotic response of small cell lung

cancer cells to ultraviolet radiation. J. Biol. Chem. 272, 10110–10116.

Ciurciu, A., Komonyi, O., Pankotai, T., and Boros, I.M. (2006). The Drosophila

histone acetyltransferase Gcn5 and transcriptional adaptor Ada2a are

involved in nucleosomal histone H4 acetylation. Mol. Cell. Biol. 26, 9413–9423.

Das, M., Jiang, F., Sluss, H.K., Zhang, C., Shokat, K.M., Flavell, R.A., and

Davis, R.J. (2007). Suppression of p53-dependent senescence by the JNK

signal transduction pathway. Proc. Natl. Acad. Sci. USA 104, 15759–15764.

de Nadal, E., and Posas, F. (2010). Multilayered control of gene expression by

stress-activated protein kinases. EMBO J. 29, 4–13.

de Nadal, E., Alepuz, P.M., and Posas, F. (2002). Dealing with osmostress

through MAP kinase activation. EMBO Rep. 3, 735–740.

de Nadal, E., Zapater, M., Alepuz, P.M., Sumoy, L., Mas, G., and Posas, F.

(2004). The MAPK Hog1 recruits Rpd3 histone deacetylase to activate

osmoresponsive genes. Nature 427, 370–374.

Devary, Y., Gottlieb, R.A., Lau, L.F., and Karin, M. (1991). Rapid and

preferential activation of the c-jun gene during the mammalian UV response.

Mol. Cell. Biol. 11, 2804–2811.

Dioum, E.M., Wauson, E.M., and Cobb, M.H. (2009). MAP-ping unconven-

tional protein-DNA interactions. Cell 139, 462–463.

Edmunds, J.W., and Mahadevan, L.C. (2004). MAP kinases as structural

adaptors and enzymatic activators in transcription complexes. J. Cell Sci.

117, 3715–3723.

Cell 142, 726–736, September 3, 2010 ª2010 Elsevier Inc. 735

Page 104: Cell 100903

Fan, Y., Lee, T.V., Xu, D., Chen, Z., Lamblin, A.F., Steller, H., and Bergmann, A.

(2010). Dual roles of Drosophila p53 in cell death and cell differentiation. Cell

Death Differ. 17, 912–921.

Fukuyama, K., Yoshida, M., Yamashita, A., Deyama, T., Baba, M., Suzuki, A.,

Mohri, H., Ikezawa, Z., Nakajima, H., Hirai, S., et al. (2000). MAPK upstream

kinase (MUK)-binding inhibitory protein, a negative regulator of MUK/dual

leucine zipper-bearing kinase/leucine zipper protein kinase. J. Biol. Chem.

275, 21247–21254.

Guelman, S., Suganuma, T., Florens, L., Swanson, S.K., Kiesecker, C.L.,

Kusch, T., Anderson, S., Yates, J.R., III, Washburn, M.P., Abmayr, S.M.,

et al. (2006). Host cell factor and an uncharacterized SANT domain protein

are stable components of ATAC, a novel dAda2A/dGcn5-containing histone

acetyltransferase complex in Drosophila. Mol. Cell. Biol. 26, 871–882.

Ham, J., Eilers, A., Whitfield, J., Neame, S.J., and Shah, B. (2000). c-Jun and

the transcriptional control of neuronal apoptosis. Biochem. Pharmacol. 60,

1015–1021.

Hu, S., Xie, Z., Onishi, A., Yu, X., Jiang, L., Lin, J., Rho, H.S., Woodard, C.,

Wang, H., Jeong, J.S., et al. (2009). Profiling the human protein-DNA interac-

tome reveals ERK2 as a transcriptional repressor of interferon signaling. Cell

139, 610–622.

Iyer, L.M., Burroughs, A.M., and Aravind, L. (2006). The prokaryotic anteced-

ents of the ubiquitin-signaling system and the early evolution of ubiquitin-like

b-grasp domains. Genome Biol. 7, R60.

Jasper, H., Benes, V., Schwager, C., Sauer, S., Clauder-Munster, S., Ansorge,

W., and Bohmann, D. (2001). The genomic response of the Drosophila embryo

to JNK signaling. Dev. Cell 1, 579–586.

Kayali, A.G., Austin, D.A., and Webster, N.J. (2000). Stimulation of MAPK

cascades by insulin and osmotic shock: lack of an involvement of p38

mitogen-activated protein kinase in glucose transport in 3T3-L1 adipocytes.

Diabetes 49, 1783–1793.

Kimura, A., Matsubara, K., and Horikoshi, M. (2005). A decade of histone acet-

ylation: marking eukaryotic chromosomes with specific codes. J. Biochem.

138, 647–662.

Kusch, T., Guelman, S., Abmayr, S.M., and Workman, J.L. (2003). Two

Drosophila Ada2 homologues function in different multiprotein complexes.

Mol. Cell. Biol. 23, 3305–3319.

Kyriakis, J.M., and Avruch, J. (2001). Mammalian mitogen-activated protein

kinase signal transduction pathways activated by stress and inflammation.

Physiol. Rev. 81, 807–869.

Lawrence, M.C., Shao, C., McGlynn, K., Naziruddin, B., Levy, M.F., and Cobb,

M.H. (2009). Multiple chromatin-bound protein kinases assemble factors that

regulate insulin gene transcription. Proc Natl Acad Sci USA 106, 22181–22186.

Lee, K.K., and Workman, J.L. (2007). Histone acetyltransferase complexes:

one size doesn’t fit all. Nat. Rev. Mol. Cell Biol. 8, 284–295.

Leimkuhler, S., Freuer, A., Araujo, J.A., Rajagopalan, K.V., and Mendel, R.R.

(2003). Mechanistic studies of human molybdopterin synthase reaction and

characterization of mutants identified in group B patients of molybdenum

cofactor deficiency. J. Biol. Chem. 278, 26127–26134.

Mas, G., de Nadal, E., Dechant, R., Rodriguez de la Concepcion, M.L., Logie,

C., Jimeno-Gonzalez, S., Chavez, S., Ammerer, G., and Posas, F. (2009).

Recruitment of a chromatin remodelling complex by the Hog1 MAP kinase

to stress genes. EMBO J. 28, 326–336.

Morrison, D.K., Murakami, M.S., and Cleghon, V. (2000). Protein kinases and

phosphatases in the Drosophila genome. J. Cell Biol. 150, F57–F62.

Pokholok, D.K., Zeitlinger, J., Hannett, N.M., Reynolds, D.B., and Young, R.A.

(2006). Activated signal transduction kinases frequently occupy target genes.

Science 313, 533–536.

Proft, M., Mas, G., de Nadal, E., Vendrell, A., Noriega, N., Struhl, K., and Posas,

F. (2006). The stress-activated Hog1 kinase is a selective transcriptional elon-

gation factor for genes responding to osmotic stress. Mol. Cell 23, 241–250.

Rozek, D., and Pfeifer, G.P. (1993). In vivo protein-DNA interactions at the c-jun

promoter: preformed complexes mediate the UV response. Mol. Cell. Biol. 13,

5490–5499.

Rozek, D., and Pfeifer, G.P. (1995). In vivo protein-DNA interactions at the c-jun

promoter in quiescent and serum-stimulated fibroblasts. J. Cell. Biochem. 57,

479–487.

Rudolph, M.J., Wuebbens, M.M., Rajagopalan, K.V., and Schindelin, H. (2001).

Crystal structure of molybdopterin synthase and its evolutionary relationship to

ubiquitin activation. Nat. Struct. Biol. 8, 42–46.

Schreiber, M., Kolbus, A., Piu, F., Szabowski, A., Mohle-Steinlein, U., Tian, J.,

Karin, M., Angel, P., and Wagner, E.F. (1999). Control of cell cycle progression

by c-Jun is p53 dependent. Genes Dev. 13, 607–619.

Schwarz, G., and Mendel, R.R. (2006). Molybdenum cofactor biosynthesis and

molybdenum enzymes. Annu. Rev. Plant Biol. 57, 623–647.

Su, Y.C., Treisman, J.E., and Skolnik, E.Y. (1998). The Drosophila Ste20-

related kinase misshapen is required for embryonic dorsal closure and acts

through a JNK MAPK module on an evolutionarily conserved signaling

pathway. Genes Dev. 12, 2371–2380.

Suganuma, T., Gutierrez, J.L., Li, B., Florens, L., Swanson, S.K., Washburn,

M.P., Abmayr, S.M., and Workman, J.L. (2008). ATAC is a double histone

acetyltransferase complex that stimulates nucleosome sliding. Nat. Struct.

Mol. Biol. 15, 364–372.

Takekawa, M., and Saito, H. (1998). A family of stress-inducible GADD45-like

proteins mediate activation of the stress-responsive MTK1/MEKK4 MAPKKK.

Cell 95, 521–530.

Thomson, S., Mahadevan, L.C., and Clayton, A.L. (1999). MAP kinase-

mediated signalling to nucleosomes and immediate-early gene induction.

Semin. Cell Dev. Biol. 10, 205–214.

Treisman, J.E., Ito, N., and Rubin, G.M. (1997). misshapen encodes a protein

kinase involved in cell shape control in Drosophila. Gene 186, 119–125.

Wang, Y.L., Faiola, F., Xu, M., Pan, S., and Martinez, E. (2008). Human ATAC is

a GCN5/PCAF-containing acetylase complex with a novel NC2-like histone

fold module that interacts with the TATA-binding protein. J. Biol. Chem. 283,

33808–33815.

Yang, D.D., Kuan, C.Y., Whitmarsh, A.J., Rincon, M., Zheng, T.S., Davis, R.J.,

Rakic, P., and Flavell, R.A. (1997). Absence of excitotoxicity-induced

apoptosis in the hippocampus of mice lacking the Jnk3 gene. Nature 389,

865–870.

Yang, S.H., Sharrocks, A.D., and Whitmarsh, A.J. (2003). Transcriptional

regulation by the MAP kinase signaling cascades. Gene 320, 3–21.

Zapater, M., Sohrmann, M., Peter, M., Posas, F., and de Nadal, E. (2007).

Selective requirement for SAGA in Hog1-mediated gene expression depend-

ing on the severity of the external osmostress conditions. Mol. Cell. Biol. 27,

3900–3910.

Zippo, A., Serafini, R., Rocchigiani, M., Pennacchini, S., Krepelova, A., and

Oliviero, S. (2009). Histone crosstalk between H3S10ph and H4K16ac

generates a histone code that mediates transcription elongation. Cell 138,

1122–1136.

736 Cell 142, 726–736, September 3, 2010 ª2010 Elsevier Inc.

Page 105: Cell 100903

A Bacterial mRNA Leader that EmploysDifferent Mechanisms to SenseDisparate Intracellular SignalsSun-Yang Park,1,3,4 Michael J. Cromie,1,2,3,5 Eun-Jin Lee,1,2,5 and Eduardo A. Groisman1,2,5,*1Department of Molecular Microbiology2Howard Hughes Medical Institute

Washington University School of Medicine, 660 South Euclid Avenue, Campus Box 8230, St. Louis, MO 63110, USA3These authors contributed equally to this work4Present address: Section of Microbial Pathogenesis, Yale University School of Medicine, Boyer Center for Molecular Medicine,

295 Congress Avenue, New Haven, CT 06536-0812, USA5Present address: Section of Microbial Pathogenesis, Howard Hughes Medical Institute, Yale University School of Medicine,Boyer Center for Molecular Medicine, 295 Congress Avenue, New Haven, CT 06536-0812, USA

*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.07.046

SUMMARY

Bacterial mRNAs often contain leader sequencesthat respond to specific metabolites or ions byaltering expression of the associated downstreamprotein-coding sequences. Here we report that theleader RNA of the Mg2+ transporter gene mgtA ofSalmonella enterica, which was previously knownto function as a Mg2+-sensing riboswitch, harborsan 18 codon proline-rich open reading frame—termed mgtL—that permits intracellular proline toregulate mgtA expression. Interfering with mgtLtranslation by genetic, pharmacological, or environ-mental means was observed to increase the mRNAlevels from the mgtA coding region. Substitution ofthe mgtL proline codons by other codons abolishedthe response to proline and to hyperosmotic stressbut not to Mg2+. Our findings show that mRNA leadersequences can consist of complex regulatoryelements that utilize different mechanisms to senseseparate signals and mediate an appropriate cellularresponse.

INTRODUCTION

The leader region (LR) of many bacterial mRNAs has the ability

to form mutually exclusive secondary structures that determine

whether transcription will continue into the adjacent coding

region (CR). Which secondary structure forms is governed

by the binding of specific metabolites, ions, tRNAs, or proteins

to the LR, by pausing of RNA polymerase at particular sequences

within the LR and by translation of short open reading frames

(ORFs) located in the leader RNA (Grundy and Henkin, 2006;

Henkin, 2008; Henkin and Yanofsky, 2002; Landick et al., 1996;

Merino and Yanofsky, 2005; Turnbough and Switzer, 2008).

These mRNA leaders typically control the expression of the

proteins that synthesize and/or transport the metabolites or

ions to which they respond. Whereas leaders known as ribos-

witches sense metabolites or ions directly (Henkin, 2008; Winkler

and Breaker, 2005), the presence of stretches of certain nucleo-

tides in a leader sequence or particular codons in a leader ORF

enables cells to probe the cellular levels of the corresponding

nucleotides and amino acids indirectly. Here we identify a leader

mRNA that controls gene expression by responding to Mg2+

directly as a riboswitch and to proline levels indirectly, via trans-

lation of a proline-codon-rich open reading frame.

The mRNA for the Mg2+ transporter gene mgtA from Salmo-

nella enterica serovar Typhimurium includes a 264 nucleotide

long leader sequence that functions as a Mg2+-responding ribos-

witch determining whether transcription proceeds into the mgtA

CR (Cromie et al., 2006). In low Mg2+, the mgtA leader RNA

adopts a conformation (i.e., stem loop C) that favors transcrip-

tion of the full-length mgtA mRNA (Cromie et al., 2006) (Figure 1)

and production of the MgtA protein (Cromie and Groisman,

2010). MgtA-mediated Mg2+ uptake restores cytoplasmic Mg2+

to prestress levels, which promotes a different conformation in

the mgtA leader RNA (i.e., stem loop B) that hinders transcription

elongation into the mgtA CR (Cromie et al., 2006) (Figure 1).

In addition to being regulated by a Mg2+-sensing riboswitch,

mgtA transcription is controlled at the initiation step by two

DNA binding proteins: PhoP, which activates transcription

when its cognate sensor PhoQ detects low extracytoplasmic

Mg2+ (Garcia Vescovi et al., 1996), acid pH (Prost et al., 2007),

or antimicrobial peptides (Bader et al., 2005); and Rob, the over-

expression of which stimulates mgtA transcription from a site

located downstream of the PhoP-dependent start site (Barchiesi

et al., 2008) (Figure 1). Thus, there is PhoP/PhoQ-dependent

production of the mgtA leader mRNA following bacterial growth

in media with <1 mM Mg2+ (Cromie et al., 2006) or pH 5.7 (Choi

et al., 2009); however, due to the riboswitch action, the mRNA

corresponding to the mgtA CR is observed primarily upon growth

in media containing very low (i.e., 10 mM) Mg2+ (Choi et al., 2009;

Cell 142, 737–748, September 3, 2010 ª2010 Elsevier Inc. 737

Page 106: Cell 100903

Cromie and Groisman, 2010; Cromie et al., 2006). This raises the

possibility of additional signals and/or genetic elements acting on

the mgtA leader mRNA synthesized when Salmonella experi-

ences inducing conditions for PhoP/PhoQ and/or Rob to

promote transcription elongation into the mgtA CR.

Here we demonstrate that the mgtA LR harbors an ORF rich in

proline codons that enables Salmonella to control transcription

elongation into the mgtA CR in response to changes in cytosolic

proline levels. We establish that high osmolarity promotes tran-

scription of the mgtA CR in a fashion dependent on the proline

codons present in the identified ORF. Also, we show that the

low Mg2+ signal and the low proline signal act synergistically to

increase the mRNA levels for the mgtA CR. Our findings highlight

the complexity of RNA regulatory sequences and provide

a singular example of a leader RNA that utilizes two distinct

mechanisms to detect two different signals.

RESULTS

Identification of a Translated ORF in the mgtA LRIn a search for new factors that might regulate mgtA expression

by acting upon the mgtA leader mRNA, we introduced a plasmid

library made in the multicopy number vector pUC19 into strain

YS774 (see Table S1 available online), which harbors the

PhoP-independent plac1-6 promoter driving transcription of

a lac fusion in the mgtA CR from the normal (i.e., PhoP-depen-

dent) start site (Cromie et al., 2006). Ampicillin-resistant trans-

Figure 1. Expression of the Salmonella

Mg2+ Transporter Gene mgtA Is Regulated

by the PhoP/PhoQ System, the Rob Protein,

and the mgtA LR

When the sensor PhoQ detects low extracytoplas-

mic Mg2+, acid pH, or antimicrobial peptides,

it promotes phosphorylation of the PhoP protein,

which binds to the mgtA promoter resulting in

transcription initiation. Rob promotes mgtA tran-

scription in response to a yet unidentified signal.

Transcription elongation into the mgtA CR is regu-

lated by the mgtA leader via a Mg2+-sensing

riboswitch and by translation of an 18 codon

proline-rich ORF designated mgtL. The mgtL

ribosome binding site is denoted by RBS adjacent

to a black line. Positions and sequences of stop

codon mutations in the strains used in the experi-

ments presented in Figure 4 are denoted below the

linear mgtL RNA sequence. High-Mg2+ and high-

proline conditions promote formation of stem

loop B, hindering transcription elongation into the

mgtA CR. Low Mg2+ and/or low proline favor

formation of stem loop C, resulting in transcription

of the mgtA CR. See also Figure S1.

formants recovered on X-Gal-containing

Luria-Bertani (LB) ampicillin agar plates

that were of a darker or lighter shade of

blue than those that received a plasmid

control were purified and their plasmid

DNA extracted and used to transform

strain YS809, a derivative of strain

YS774 with the same mgtA-lac transcriptional fusion. One of

the plasmids—designated pSL55—promoted higher levels of

mgtA-lac expression than the plasmid control (Figure 2A), like

it did in strain YS774, indicative that the observed phenotype

was plasmid linked. Plasmid pSL55 appears to promote mgtA-

lac expression specifically because it had no effect on the Lac

phenotype of strains harboring lac transcriptional fusions in the

CRs of the PhoP-dependent mgtC gene or the PhoP-indepen-

dent corA gene (data not shown). The mgtC and corA transcripts

also include long leaders (Lejona et al., 2003) (T. Latifi and E.A.G.,

unpublished results) and specify proteins participating in Mg2+

homeostasis (Blanc-Potard and Groisman, 1997; Maguire,

2006; Soncini et al., 1996).

Sequencing of the insert in pSL55 revealed the presence

of 1579 nucleotides corresponding to position 78347–79925 in

the chromosome of strain 14028s, which is a portion of the

3228 nucleotide long carB gene (see Table S2 for primers).

The insert operates in an orientation-dependent manner

because it did not alter mgtA-lac expression when cloned into

pUC18 where the opposite strand is transcribed (data not

shown). Although this suggested that pSL55 might promote

mgtA-lac expression by functioning as an antisense RNA for

the carB transcript, we ruled out this possibility because

pSL55 upregulated mgtA transcription even in a strain deleted

for the carB gene (Figure 2A). Given that a transcript correspond-

ing to the strand opposite the carB gene was not detected when

RNA was harvested from wild-type Salmonella grown in LB or in

738 Cell 142, 737–748, September 3, 2010 ª2010 Elsevier Inc.

Page 107: Cell 100903

N-minimal media (data not shown), and that mgtA-lac was not

derepressed in the DcarB strain (Figure 2A), we concluded that

pSL55 upregulates mgtA-lac expression by a mechanism that

does not involve the carB gene.

We then examined the nucleotide sequence of the insert in

pSL55 and identified a 15 nucleotide segment that is comple-

mentary to nucleotides 61–75 in the mgtA leader RNA

(Figure 2B). Interestingly, this region corresponds to the potential

start codon and ribosome binding site for a previously unidenti-

fied 18 codon ORF—hereafter designated mgtL—that begins at

position 71 and ends at position 124 (Figure 1). The presence of

an ORF and ribosome binding site is conserved in the mgtA LRs

from Escherichia coli, Shigella flexneri, Klebsiella pneumoniae,

Citrobacter koseri, Enterobacter sp. 638, Dickeya zeae, Dickeya

dadantii, and Serratia proteamaculans (Figure 3A). The mgtA

leader sequences from all these species have the potential ability

to form the alternative stem loop structures B and C identified in

the Salmonella mgtA leader (Figure 1) (Cromie et al., 2006) (data

not shown).

We determined that mgtL is translated in vivo because wild-

type Salmonella carrying the medium-copy-number plasmid

pmgtL-0lacZ, with the full mgtL coding sequence and its putative

ribosome binding site fused in frame to the E. coli lacZ gene start-

ing at the ninth codon, produced high levels of b-galactosidase,

whereas an isogenic strain harboring a modified plasmid with

a stop codon after the 17th mgtL codon did not (Figure S1). (There

is a potential start codon at position 26–28 in the Salmonella mgtA

leader sequence that is in frame with the mgtL ORF and is

preceded by a possible ribosome binding site [Figure 1]. However,

the AUG at this position does not appear to be a true translation

start site because first, the sequence between nucleotides 26 and

70 is not conserved in the other examined species, and second,

wild-type Salmonella harboring plasmid pmgtLD2nt-0 lacZ, a

derivative of plasmid pmgtL-0lacZ with two nucleotides deleted

at position 29–30, which creates a stop codon at position

79–81 for the ORF starting at position 26, still expressed high

levels of b-galactosidase [Figure S1].) The strains harboring plas-

mids pmgtL-0lacZ and pmgtLD2nt-0lacZ displayed similar levels

of b-galactosidase following growth in low and high Mg2+

(Figure S1). These data indicate that mgtL translation is not regu-

lated by changes in the Mg2+ concentration in the media. More-

over, they reflect that in the pmgtL-0lacZ and pmgtLD2nt-0lacZ

plasmids, mgtL-lac transcription is initiated from a Mg2+-blind

promoter and the Mg2+-responding riboswitch is disrupted.

Inhibiting mgtL Translation Promotes Expressionof the mgtA CRPlasmid pSL55 appears to upregulate mgtA-lac expression by

producing a trans-acting RNA that interferes with mgtL transla-

tion because first, the region of complementarity with the mgtA

leader corresponds to the ribosome binding site and start codon

of mgtL (Figure 2B). Second, inactivation of the hfq gene, which

codes for the RNA chaperone usually required for pairing

between trans-acting regulatory RNAs and their targets

(Brennan and Link, 2007), eliminated pSL55’s regulatory effect

(Figure 2A). And third, pSL55-Mut, a pSL55 derivative with nucle-

otide substitutions in the region of complementarity with the

mgtA LR (Figure 2B), could not upregulate mgtA expression

(Figure 2A).

If plasmid pSL55 derepresses mgtA-lac transcription by

inhibiting mgtL translation, we reasoned that disrupting mgtL

translation by other means might have the same effect. To test

this hypothesis, we introduced mutations in plasmid pYS1010,

which harbors the plac1-6 derivative of the lac promoter driving

transcription of the full-length mgtA LR fused to a promoterless

lacZ gene and confers Mg2+-regulated synthesis of b-galactosi-

dase (Cromie et al., 2006). Wild-type Salmonella harboring

pYS1010 derivatives with stop codons at positions 80–82,

89–91, or 98–100 (Figure 1) produced high levels of b-galactosi-

dase when grown in high Mg2+ (Figure 4A). These levels were 46-

to 89-fold higher than those displayed by the strain carrying the

original plasmid pYS1010.

The high expression levels exhibited by strains harboring

pYS1010 derivatives with stop codon mutations in mgtL are

probably due to lack of translation of the full-length mgtL

(as opposed to resulting from major alterations in the mgtA

leader RNA structure that lock the riboswitch in an ON state)

because first, a plasmid expressing the amber suppressor

supF restored normal levels of mgtA-lac transcription to the

strain harboring a pYS1010 derivative with an amber stop codon

at position 98–100 whereas the plasmid vector had no effect

A

B

Figure 2. Plasmid pSL55 Promotes mgtA-lac Transcription by Inter-

fering with mgtL Translation

(A) b-galactosidase activity (Miller units) from a chromosomal mgtA-lac tran-

scriptional fusion driven by the plac1-6 promoter was determined in wild-type

(YS809), carB (SP31), and hfq (SP54) strains harboring plasmids pSL55,

pSL55-Mut, or a control plasmid. Bacteria were grown in LB medium for

4.5 hr. Shown are the mean and standard deviation (SD) from two independent

experiments.

(B) Nucleotide sequences corresponding to position 61–75 from the mgtA LR

including the mgtL ribosome binding sequence (RBS; horizontal black line) and

start codon (bold AUG) (top); the sequence of the pSL55-derived transcript

that is complementary to the mgtA leader mRNA (middle); and the sequence

of the pSL55-Mut-derived transcript that is no longer complementary to

the mgtA leader mRNA (bottom).

Cell 142, 737–748, September 3, 2010 ª2010 Elsevier Inc. 739

Page 108: Cell 100903

(Figure 4C). Furthermore, the supF-promoted suppression was

specific because the supF-carrying plasmid did not reduce

expression in a strain carrying a pYS1010 derivative with an

opal stop codon at the same position (Figure 4C), nor did

it modify it in a strain with the original pYS1010 plasmid carrying

the wild-type mgtA leader sequence (Figure 4C). Second,

the structures of the wild-type and stop codon mutant mgtA

leaders were comparable when examined by in-line probing

(Figure S2A). And third, in vitro transcription assays demon-

strated that Mg2+ regulated transcription elongation beyond

the mgtA LR similarly in DNA templates corresponding to the

wild-type and stop codon mutant mgtA leaders (Figure S2B).

We next tested whether inhibiting mgtL translation by means

other than mutation of the mgtA LR affected the mRNA levels

for the mgtA CR. The protein synthesis inhibitor tetracycline

promoted an �28-fold increase in the mgtA CR mRNA 15 min

after its addition to wild-type Salmonella (Figure 4D). Thus, tetra-

cycline could overcome the transcriptional silencing of the mgtA

CR that normally takes place when wild-type Salmonella experi-

ences 500 mM Mg2+ (Cromie and Groisman, 2010; Cromie et al.,

2006). Tetracycline appears to affect the mRNA levels produced

from the mgtA CR specifically because there was little change in

the mRNA levels of the mgtA LR or the phoP CR, which were

examined as controls (Figure 4D). Addition of chloramphenicol

also increased the mRNA levels of the mgtA CR in wild-type

Salmonella (data not shown), implying that the derepression

promoted by protein synthesis inhibitors is not limited to a partic-

ular mechanism of action. Tetracycline exerts its effect on the

mgtA LR (as opposed to the mgtA promoter) because

it increased lacZ mRNA levels in wild-type Salmonella harboring

plasmid pYS1010 (data not shown). Furthermore, tetracycline

acts by inhibiting mgtL translation because it did not promote

a significant increase in lacZ mRNA levels in wild-type Salmo-

nella carrying the pYS1010 derivative with a stop codon at posi-

tion 98–100 (data not shown). Taken together, our results

indicate that interfering with mgtL translation by genetic or phar-

macological means increases the mRNA levels for the mgtA CR.

The mgtL-encoded peptide does not appear to act in trans

because a plasmid carrying mgtL DNA failed to restore normal

mgtA-lac expression in strains harboring pYS1010 derivatives

with stop codons at different positions in mgtL (Figure S2C).

This suggested that mgtL exerts its regulatory effect in cis and

raised the question as to the physiological signal controlling

mgtL translation.

Proline Limitation Enhances Transcriptionof the mgtA CRThe Salmonella mgtL sequence includes four proline codons

(Figure 1), which is a disproportionately high frequency for an

18 codon ORF. The number and location of the proline codons

(at the third, fifth, seventh, and ninth positions) are largely

conserved in the nine examined species (Figures 3A and 3B).

A

B

Figure 3. The Presence of an Open Reading Frame and Ribosome-Binding Site Is Conserved in the mgtA LR from Salmonella enterica,

Escherichia coli, Shigella flexneri, Citrobacter koseri, Klebsiella pneumoniae, Enterobacter sp. 638, Dickeya zeae, Dickeya dadantii, and

Serratia proteamaculans

(A) Alignment of the RNA sequences corresponding to the ribosome binding site and mgtL from the species listed above. Sequences in blue correspond to mgtL.

Asterisks correspond to nucleotides conserved in all species. The number of nucleotides shown in each line is indicated to the right of the nucleotide sequences.

(B) Alignment of the deduced amino acid sequences corresponding to mgtL from the species listed above. Sequences in yellow correspond to proline residues.

Asterisks correspond to positions conserved in all species. The number of sense codons in the mgtL ORFs is indicated to the right of the deduced amino acid

sequences.

740 Cell 142, 737–748, September 3, 2010 ª2010 Elsevier Inc.

Page 109: Cell 100903

Because the proline codons are located in the mgtL region where

introduction of stop codons heightens expression downstream of

the mgtA leader (Figure 1), we hypothesized that a drop in the

cytosolic proline levels might decrease the availability of proline-

charged tRNAs, causing ribosome stalling at mgtL proline

codons, leading to formation of stem loop C and an increase in

the levels of the mgtA CR mRNA. By contrast, when cytosolic

proline is abundant, coupling of mgtA leader transcription and

mgtL translation would be restored, thereby favoring formation

of stem loop B and reducing the production of mgtA CR mRNA.

To explore this hypothesis, we grew a proline auxotroph in the

presence of 1 mM proline for 1 hr, washed the cells, split

the culture into two different media—one containing and one

lacking proline—and then harvested the mRNA 15 min later.

The mRNA level for the mgtA CR was 8-fold higher in organisms

grown in the absence of proline than in those grown in its pres-

ence (Figure 4E). This effect was unique to the mgtA CR, as no

differences were detected in the mRNA levels corresponding

to the mgtA LR or the phoP CR (Figure 4E). Furthermore,

it was specific for proline, as limitation for histidine, which does

not have codons in the mgtL sequence (Figure 1), failed to

increase the mRNA level for the mgtA CR (data not shown).

Indeed, despite the presence of three arginine codons in mgtL

(Figure 1), arginine limitation did not promote an increase in the

mRNA levels of the mgtA CR in an arginine auxotroph (data not

shown). Perhaps this reflects the location of the arginine codons,

one of which is in stem loop C, and that the remaining arginine

codons may not be sufficient to mediate mgtA derepression.

A B

Fo

ld c

han

ge

(T15

/ T

0)

40m

gtA le

ader

mgt

A cod

ing

phoP

cod

ing

0

10

20

30

D

0

2000

4000

6000

8000

10000

pYS1

010

psto

p 80

-82

psto

p 89

-91

psto

p 98

-100

psto

p 10

7-10

9ps

top

110-

112

psto

p 10

7-10

9 11

0-11

2pY

S101

0-G

120C

psto

p 80

-82-

G12

0CpY

S101

0-C

145G

psto

p 98

-100

-C14

5G

-gal

acto

sid

ase

acti

vity

(Mill

er u

nit

s)

0

2000

4000

6000

8000

10000

pYS1

010

psto

p 80

-82

psto

p 89

-91

psto

p 98

-100

psto

p 10

7-10

9ps

top

110-

112

psto

p 10

7-10

9 11

0-11

2pY

S101

0-G

120C

psto

p 80

-82-

G12

0CpY

S101

0-C

145G

psto

p 98

-100

-C14

5G

-gal

acto

sid

ase

acti

vity

(Mill

er u

nit

s)

C

vect

orpU

Hsu

pF

-gal

acto

sid

ase

acti

vity

(Mill

er u

nit

s)

pstop 98-100(amber)

pstop 98-100(opal)

pYS1010

0

1000

3000

4000

5000

2000

vect

orpU

Hsu

pF

vect

orpU

Hsu

pF

F

0

0.5

1.0

1.5

2.0

Fo

ld c

han

ge

(no

pro

line

/ pro

line)

mgt

A lead

erm

gtA c

odin

gph

oP c

odin

g

E

Fo

ld c

han

ge

(no

pro

line

/ pro

line)

2

4

6

8

10

12

0

mgt

A lead

erm

gtA c

odin

gph

oP c

odin

g

Figure 4. Translation of mgtL Governs

Transcription Elongation beyond the

mgtA LR

(A) b-galactosidase activity (Miller units) produced

by wild-type Salmonella (14028s) harboring

plasmid pYS1010 with the wild-type mgtA leader

(Cromie et al., 2006), or derivatives with stop

codons at different positions in mgtL and/or

mutations that hinder stem loop C formation:

pstop 80–82; pstop 89–91; pstop 98–100; pstop

107–109; pstop 110–112; pstop 107–109 D110–

112; pYS1010-G120C; pstop 80–82-G120C;

pYS1010-C145G; and pstop 98–100-C145G.

Bacteria were grown in N-minimal medium with

10 mM Mg2+ for 4 hr. Shown are the mean and

SD from at least three independent experiments.

(B) b-galactosidase activity produced by the

strains listed in (A). Bacteria were grown in

N-minimal medium with 10 mM Mg2+ for 4 hr.

Shown are the mean and SD from at least three

independent experiments.

(C) b-galactosidase activity produced by wild-type

Salmonella (14028s) harboring plasmid pUHsupF

or the plasmid vector pUH21-2lacIq, and

pYS1010 or its derivatives with UAG (pstop

98–100; amber) or UGA (pstop 98–100; opal)

stop codons at position 98–100 in the mgtA leader.

Bacteria were grown in N-minimal medium with

500 mM Mg2+ for 3 hr and with 1 mM IPTG for

1 hr. Shown are the mean and SD from three inde-

pendent experiments.

(D) Fold change in the mRNA levels of the mgtA LR

and the mgtA and phoP CRs produced by wild-

type Salmonella (14028s) treated with the protein

synthesis inhibitor tetracycline (25 mg/ml). Expres-

sion levels of target genes were normalized to that

of the 16S ribosomal RNA rrs gene. Fold change

was calculated by dividing the mRNA levels from

samples taken 15 min after treatment with tetracy-

cline (T150) by that of samples taken before addi-

tion of the antibiotic (T00). Shown are the mean

and SD from three independent experiments.

(E) Fold change in the mRNA levels of the mgtA LR and the mgtA and phoP CRs produced by a proline auxotrophic strain (EG19886) grown in modified N-minimal

media with 500 mM Mg2+ in the presence of 1 mM proline for 1 hr, and then grown for 15 min in media containing or lacking proline. Expression levels of target

genes were normalized as described in (D). Fold change was calculated by dividing the mRNA levels of cells grown in the absence of proline by that of cells grown

in the presence of proline. Shown are the mean and SD from three independent experiments.

(F) Fold change in the mRNA levels of the mgtA LR and the mgtA and phoP CRs produced by wild-type Salmonella (14028s) grown and analyzed as described in

(D). Shown are the mean and SD from two independent experiments.

See also Figure S2.

Cell 142, 737–748, September 3, 2010 ª2010 Elsevier Inc. 741

Page 110: Cell 100903

Proline limitation promoted a 2-fold increase in the mRNA

levels of the mgtA CR in wild-type (i.e., prototrophic) Salmonella

(Figure 4F). Although this is lower than the increase observed in

the proline auxotroph (Figure 4E), it is similar to the gene expres-

sion changes promoted by branched-chain amino acids in the

attenuation-regulated ilvGMEDA operon (Chen et al., 1991).

Moreover, it is higher than what has been reported for the atten-

uation-regulated trp operon, where starvation for tryptophan

does not alter trp expression in a tryptophan prototroph

(Yanofsky and Horn, 1994).

The mgtL Proline Codons Are Necessaryfor the Response to Proline LimitationWe determined that proline limitation failed to promote an

increase in the mRNA levels for the mgtA CR in a derivative of

the proline auxotroph with all four mgtL proline codons replaced

by codons specifying other amino acids (Figure 5A). As found

with the isogenic mgtL+ strain, proline limitation had little effect

on the mRNA levels for the mgtA leader and phoP CRs

(Figure 5A). Importantly, the mutant still responded to changes

in the Mg2+ concentration (Figure 5B) (albeit not as well as the

isogenic strain with a wild-type mgtA leader), indicating that

the proline codon substitutions did not lock the mutant into an

inactive conformation. These data indicate that the mgtL proline

codons are essential for the response to proline limitation, and

they suggest that this ability is distinct from the mgtA leader’s

role as a Mg2+-sensing device.

We hypothesized that the proline codon at the third position of

mgtL might not be required for regulation of mgtA expression by

proline because D. dadantii does not have a proline codon at this

position (Figures 3A and 3B). Indeed, a Salmonella strain in which

the mgtL proline codon at the third position was replaced by

a leucine codon derepressed the mgtA CR mRNA when limited

for proline, like the strain with the wild-type mgtL (Figure 5A).

However, this mutant still responded to changes in the Mg2+

concentration, but to a smaller degree than the mutant

substituted in all four mgtL proline codons (Figure 5B).

The results presented above suggested that replacement of

the mgtL proline codons at the fifth, seventh, and ninth positions

by codons specifying other amino acids might be sufficient to

eliminate the response to proline. As predicted, there was little

derepression of the mgtA CR mRNA upon proline limitation in

a derivative of the proline auxotroph with the last three mgtL

proline codons substituted by the same codons as in the mutant

with substitutions in all four mgtL proline codons (Figure 5A). As

with the other isogenic strains, proline limitation did not affect the

mRNA levels of the mgtA leader and phoP CRs (Figure 5A).

The response to Mg2+ of the mgtL derivative with substitutions

in the last three proline codons was intermediate to that

A

Fo

ld c

han

ge

(10

µµM

Mg

2+ /

500 µM

Mg

2+)

B

Fo

ld c

han

ge

(no

pro

line

/ pro

line)

C

0

10

20

30

40

50

60

70

80

Rel

ativ

e m

RN

A le

vels

mgtA leader mgtA coding phoP coding

[Mg2+]H H H H H HL L LL LLProline+ - + + - + -- + - -+

0

2

4

6

8

10

12

14 mgtA leader mgtA coding phoP coding

0

10

20

30

40

50

60

70

80 mgtA leader mgtA coding phoP coding

Pro3

Pro3

,5,7

,9

mgt

L+

Pro5

,7,9

Pro3

Pro3

,5,7

,9

mgt

L+

Pro5

,7,9

Pro3

Pro3

,5,7

,9

mgt

L+

Pro5

,7,9

Pro3

Pro3

,5,7

,9

mgt

L+

Pro5

,7,9

Pro3

Pro3

,5,7

,9

mgt

L+

Pro5

,7,9

Pro3

Pro3

,5,7

,9

mgt

L+

Pro5

,7,9

Figure 5. The mgtA Leader Responds to Proline and Mg2+ Indepen-

dently and Requires the mgtL Proline Codons for Regulation of the

mgtA CR by Changes in the Proline Concentration

(A) Fold change in the mRNA levels of the mgtA LR and the mgtA and phoP

CRs produced by a proline auxotroph (SP1; mgtL+), or derivatives in which

the mgtL proline codon at the third position was substituted (SP2; Pro3), the

last three mgtL proline codons were substituted (SP61; Pro5,7,9), or in which

all mgtL proline codons were substituted (SP8; Pro3,5,7,9). Bacteria were

grown and mRNA was analyzed as described in Figure 4E. Shown are the

mean and SD from three independent experiments.

(B) Fold change in the mRNA levels of the mgtA LR and the mgtA and phoP

CRs produced by the strains described in (A). Bacteria were grown in modified

N-minimal medium with 10 mM or 500 mM Mg2+ for 3.5 hr. Expression levels of

target genes were normalized as described in (A). Fold change was calculated

by dividing the mRNA levels of cells grown in N-minimal medium with 10 mM

Mg2+ by those present in cells grown in 500 mM Mg2+. Shown are the mean

and SD from three independent experiments.

(C) Relative mRNA levels of the mgtA LR and the mgtA and phoP CRs

produced by a proline auxotroph (EG19886) that experienced 500 mM Mg2+

and 1 mM proline, 500 mM Mg2+ and no proline, 5 mM Mg2+ and 1 mM proline,

or 5 mM Mg2+ and no proline for 15 min. Relative mRNA levels were calculated

by dividing the mRNA levels of cells grown in a given condition by those

present in cells grown in 500 mM Mg2+ and 1 mM proline. Shown are the means

and SD from three independent experiments.

742 Cell 142, 737–748, September 3, 2010 ª2010 Elsevier Inc.

Page 111: Cell 100903

exhibited by the one with the third proline codon substituted and

the one with all four proline codons substituted (Figure 5B).

Synergism between the Low Mg2+ and the ProlineLimitation SignalsTo examine whether direct Mg2+ sensing by the mgtA leader RNA

affects indirect proline sensing by the mgtL ORF and vice versa,

we grew the mgtL+ proline auxotrophic strain in the presence of

500mM Mg2+ and 1 mM proline for 1 hr, washed the cells, and split

the culture into four different media containing either 500 mM

Mg2+ and 1 mM proline, 500 mM Mg2+ and no proline, 5 mM

Mg2+ and 1 mM proline, or 5 mM Mg2+ and no proline. Fifteen

minutes later, we determined the mRNA levels corresponding

to the mgtA and phoP CRs as well as to the mgtA LR. We found

that the mRNA levels corresponding to the mgtA CR were higher

in cells experiencing both low Mg2+ and low proline than in cells

limited only for Mg2+ or for proline (Figure 5C). As expected, the

mRNA levels corresponding to the mgtA LR and the phoP CR

were higher in cells experiencing 5 mM Mg2+ than in those

exposed to 500 mM Mg2+ (Figure 5C), reflecting that the amount

of activated PhoP protein increases as the Mg2+ concentration

in the media decreases (Shin et al., 2006). Our results support

the notion that Mg2+ and proline are sensed independently

by the mgtA leader RNA. Furthermore, they suggest synergism

between the two inducing signals because the mRNA levels for

the mgtA CR were higher in cells experiencing both low Mg2+

and low proline than the sum of the mRNA levels produced by

cells experiencing only low Mg2+ or low proline (Figure 5C).

In agreement with this notion, wild-type Salmonella harboring

plasmid pYS1010 derivatives with early stop codons in the

mgtL sequence produced more b-galactosidase when grown in

low Mg2+ than the sum of theb-galactosidase activities produced

by the same strains grown in high Mg2+ plus the b-galactosidase

activity produced by wild-type Salmonella carrying pYS1010

grown in low Mg2+ (Figures 4A and 4B).

Hyperosmotic Shock Promotes an Increase in the mgtA

CR mRNA LevelsProline plays two major functions in bacterial cells: it is a compo-

nent of proteins and it can function as an osmoprotectant

(Csonka and Leisinger, 2007). Thus, we hypothesized that

when bacteria experience hyperosmotic shock, the increased

requirement for proline in osmoprotection might decrease its

availability to charge proline tRNAs. This could potentially lead

to ribosome stalling at the mgtL proline codons and result in

derepression of the mgtA CR. We tested this hypothesis by

comparing the mRNA levels produced by wild-type Salmonella

experiencing hyperosmotic shock in the presence or absence

of osmoprotectants.

We determined that the mRNA levels corresponding to the

mgtA CR were �6-fold higher when Salmonella experienced

500 mM Mg2+ + 0.3 M NaCl for 1 hr than in organisms grown in

500 mM Mg2+ (Figure 6A). By contrast, the mRNA levels for the

mgtA LR and the phoP CR were similar under the two growth

conditions (Figure 6A), indicative that the induction of the mgtA

CR mRNA levels promoted by high osmolarity is not mediated

by the PhoP-dependent mgtA promoter. Addition of the osmo-

protectant glycine betaine together with NaCl compromised

the induction of the mgtA CR mRNA promoted by NaCl but

had negligible effects when added in the absence of NaCl

(Figure 6A). Proline had a similar (albeit not as strong) effect as

glycine betaine (Figure 6A), presumably because it is not as

effective as glycine betaine in osmoprotection (Cayley et al.,

1992). The increase in the mgtA CR mRNA levels provoked by

hyperosmotic shock requires an intact mgtL ORF, because it

was not observed in an isogenic strain substituted in all four

mgtL proline codons (Figure 6B). Cumulatively, these data

demonstrate that hyperosmotic shock promotes transcription

of the mgtA CR in an mgtL-dependent manner.

Formation of Stem Loop C Is Necessaryfor Transcription of the mgtA CRBecause transcription and translation are coupled in bacteria,

when cytosolic proline is abundant, a ribosome translating the

complete mgtL sequence is likely to occlude the left arm of

A

B

mgtA leader mgtA coding phoP coding

0

1

2

3

4

5

6

7

Rel

ativ

e m

RN

A le

vels

NPNGNG-NPNGNG-NPNGNG-

0

1

2

3

4

5

6

7mgtA leader mgtA coding phoP coding

Rel

ativ

e m

RN

A le

vels

NPNGNG-NPNGNG-NPNGNG-

mgtL+

Pro3,5,7,9

Figure 6. Hyperosmotic Shock Promotes Transcription of the mgtA

CR in a Process that Requires the mgtL Proline Codons

Relative mRNA levels of the mgtA LR and the mgtA and phoP CRs produced

by wild-type Salmonella (YS957) (A), or a derivative in which all mgtL proline

codons were substituted (EG19870) (B). Bacteria were grown for 1 hr in modi-

fied N-minimal medium without casamino acids containing 500 mM Mg2+ and

either no additional supplements (-), 1 mM glycine betaine (G), 0.3 M NaCl (N),

0.3 M NaCl and 1 mM glycine betaine (NG), or 0.3 M NaCl and 1 mM proline

(NP). Relative mRNA levels were calculated by dividing the mRNA levels of

cells grown under the specified condition by the mRNA levels present in cells

grown in 500 mM Mg2+ (i.e., with no additional supplement). Shown are the

mean and SD from three independent experiments.

Cell 142, 737–748, September 3, 2010 ª2010 Elsevier Inc. 743

Page 112: Cell 100903

stem loop C (Figure 1 and Figure 7A). This would favor formation

of stem loop B, which has been shown to hamper transcription

elongation beyond the mgtA LR (Cromie et al., 2006).

By contrast, conditions that reduce the levels of free cytosolic

proline would promote ribosome stalling at the mgtL proline

codons, thereby advancing formation of stem loop C and result-

ing in transcription elongation into the mgtA CR (Figure 1 and

Figure 7A). Therefore, the position that a translating ribosome

reaches in the mgtL ORF should determine whether transcription

continues into the mgtA CR (Figure 1 and Figure 7A).

We tested our model by investigating the phenotype of wild-

type Salmonella harboring pYS1010 derivatives with stop

codons at different positions within mgtL. A derivative with

a stop codon at position 110–112, which is only six nucleotides

upstream of the left arm of stem loop C (Figure 1), produced

very low levels of b-galactosidase when grown in 10 mM Mg2+

(Figure 4A), like the isogenic strain with plasmid pYS1010

harboring the wild-type mgtA leader (Figure 4A). By contrast,

there were high levels of b-galactosidase in a pYS1010 derivative

with a stop codon at position 107–109 (Figure 4A), which is nine

nucleotides from the left arm of stem loop C (Figure 1), similar to

the behavior of strains with stop codons at positions 80–82,

89–91, or 98–100 (Figure 4A). Furthermore, the pYS1010 deriva-

tive with a stop codon at position 107–109 no longer conferred

high levels of b-galactosidase upon wild-type Salmonella when

nucleotides 110–112 were deleted (Figure 4A), which brought

the mgtL stop codon only six nucleotides away from the left

arm of stem loop C. Cumulatively, these results indicate that

the distance between the ribosome translating mgtL and the

nucleotides forming the left arm of stem loop C (as opposed to

the size of the translated mgtL product) is critical for mgtL-medi-

ated gene control.

If the lack of translation of the full-length mgtL stimulates tran-

scription elongation beyond the mgtA LR by favoring formation of

stem loop C (Figure 1 and Figure 7A), hindering formation of stem

loop C should abolish this stimulation. As predicted, the G120C

and C145G substitutions in the mgtA leader (Figure 1), which

were previously shown to impede formation of stem loop C

(Cromie et al., 2006), thwarted derepression in constructs

harboring stop codons at either of two mgtL positions (Figure 4A).

DISCUSSION

The LR of many mRNAs can respond to specific nutritional

and/or physical signals by modifying expression of the associ-

ated downstream coding sequences. Some of these leader

UUUUUUU

UUUUUUU

UUUUUUU

OFFstate

OFF state

OFF state

ON state

ON state

ON state

Ribosome is stalled atproline/tryptophan

codons

Ribosome loads butthere is no mgtL or

trpL translation

Leader ORF iscompletely translated

mgtA leader region trp leader region

AUG

AUG

AUG

AUG

AUG

AUG

STOP

STOP

STOP

STOP

STOP

STOP

RBS

RBS

RBS

RBS

RBS

AU

RBS

RibosomeRibosome

Stem-loop C

Stem-loop B

Stem-loop C

Terminator

Terminator

Antiterminator

Anti-antiterminator

STOP STOP

G

AU

RBS

A B

RBS

Figure 7. Regulation of Transcription Elon-

gation into the mgtA and trp CRs by ORFs

Located within Their Respective LRs

(A) Regulation of transcription elongation by the

mgtA leader RNA. Top: when proline levels are

limiting in the cytosol, a ribosome translating

mgtL stalls at proline codons, which favors forma-

tion of stem loop C and results in transcription of

the mgtA CR. Middle: when proline levels are not

limiting in the cytosol, a ribosome can translate

the complete mgtL ORF, which favors formation

of stem loop B and hinders transcription elonga-

tion into the mgtA CR by an unknown mechanism.

Bottom: a ribosome loads but cannot translate

a mutant mgtA leader with nucleotide substitutions

in the mgtL start codon, which favors formation of

stem loop C and results in transcription of the

mgtA CR.

(B) Regulation of transcription elongation by the trp

leaderRNA. Top: when tryptophan levelsare limiting

in the cytosol, a ribosome translating trpL stalls at

tryptophan codons, which favors formation of an

antiterminator structure and results in transcription

of the trp operon. Middle: when tryptophan levels

are not limiting in the cytosol, a ribosome can trans-

late the complete trpL ORF, which favors formation

ofan intrinsic transcription terminatorand there isno

transcription of the trp operon. Bottom: a ribosome

loads but cannot translate a mutant trp leader with

nucleotide substitutions in the trpL start codon,

which favors formation of both the anti-antitermina-

tor and terminator structures and there is no tran-

scription of the trp operon. Note the different posi-

tion of the mgtL and trpL ORFs relative to the

sequences formingstem loopC and the anti-antiter-

minator structures, respectively, which determines

the different phenotypes resulting from mutation of

the mgtL and trpL start codons.

744 Cell 142, 737–748, September 3, 2010 ª2010 Elsevier Inc.

Page 113: Cell 100903

RNAs rely on the same mechanism to detect more than one

signal whereas others utilize different means of controlling

gene expression in response to one signal. For example, two

riboswitches located in tandem, one responding to S-adenosyl-

methionine and the other to coenzyme B12, control transcription

of the metE gene in Bacillus clausii (Sudarsan et al., 2006). In

other cases, a cis-acting riboswitch can also function as

a trans-acting regulatory RNA to modulate the expression of

genes located somewhere else in the genome (Loh et al.,

2009). The mgtA leader described here constitutes a singular

example of an mRNA leader that utilizes different mechanisms

to sense different signals.

We determined that the mRNA leader corresponding to the

Mg2+ transporter gene mgtA harbors a translated ORF rich in

proline codons designated mgtL (Figure 1; Figure S1), which is

conserved in other bacterial species (Figures 3A and 3B) and

enables Salmonella to regulate transcription elongation into the

mgtA CR in response to the levels of cytoplasmic proline.

Whereas the mgtA leader RNA senses Mg2+ directly (Cromie

et al., 2006), it monitors proline levels in an indirect fashion, via

mgtL translation (Figure 4A), using an attenuation-like mecha-

nism (Landick et al., 1996; Henkin and Yanofsky, 2002).

The dual sensing ability of the mgtA leader allows Salmonella

to transcribe the mgtA CR not only when Mg2+ levels drop below

a certain threshold (Cromie and Groisman, 2010; Cromie et al.,

2006) but also under conditions resulting in a decrease in cyto-

solic proline levels (Figures 4E and 4F).

The discovery of a translatable ORF in the mgtA leader

(Figure 1; Figure S1) suggests a possible explanation for the

recent observation that the mRNA levels corresponding to the

50-most 127 nucleotides of the mgtA leader are much higher

than those for the 30-most 94 nucleotides of the mgtA leader

when Salmonella is grown in high-Mg2+ media (Spinelli et al.,

2008). This is because the RNA-processing enzyme RNase E,

which has been implicated in mgtA mRNA degradation during

growth in high Mg2+ (Spinelli et al., 2008), can target nascent

transcripts as well as complete messages (Hammarlof and

Hughes, 2008). Therefore, the ribosome translating mgtL, which

extends to position 124 in the mgtA leader (Figure 1), could

protect the 50 region of the mgtA leader from RNase E action, re-

sulting in higher levels of this portion of the RNA.

mgtL Translation Regulates Transcriptionof the mgtA CRWe determined that interfering with mgtL translation increased

the mRNA levels for the mgtA CR (Figures 4A and 4D–4F). This

appears to result from the formation of a particular secondary

structure (i.e., stem loop C) in the mgtA leader (Figure 1 and

Figure 7A) that promotes transcription elongation into the mgtA

CR, as nucleotide substitutions impeding formation of stem

loop C (Cromie et al., 2006) prevented derepression provoked

by stop codon mutations in mgtL (Figure 4A). Because transcrip-

tion and translation are coupled in bacteria and because the

ribosome occupies �30 nucleotides, covering 12–15 nucleo-

tides from the P site (Laursen et al., 2005), the position that

a translating ribosome reaches in mgtL relative to the sequences

that make up stem loop C would determine whether transcription

continues into the mgtA CR (Figure 1 and Figure 7A). If the ribo-

some is too close to the sequences that make up stem loop C,

then the alternative stem loop B would form (Figure 1 and

Figure 7A), and transcription would not continue into the mgtA

CR. For instance, stop codons at positions 80–82, 89–91,

98–100, or 107–109 in the mgtA leader resulted in derepression

whereas a stop codon at position 110–112 did not (Figure 4A).

Yet, the stop codon at position 107–109 no longer derepressed

expression if nucleotides 110–112 were also deleted (Figure 4A),

which moved the mgtL stop codon only six nucleotides away

from the left arm of stem loop C (Figure 1).

It was recently reported that a chromosomal C98T mutation in

the leader RNA causes constitutive high expression of the mgtA

CR (O’Connor et al., 2009). In the original description of this

mutation, it was not clear why this nucleotide substitution should

alter mgtA expression, but we now see that it created a stop

codon in mgtL that is identical to the mutation we had previously

constructed, which resulted in constitutive high expression of

mgtA (Figures 4A and 4B).

The mgtL proline codons are located at positions 77–79,

83–85, 89–91, and 95–97 (Figure 1), which correspond to the

region where placing a stop codon results in derepression (Fig-

ure 4A). Therefore, we propose that when cytosolic proline levels

are low, there would be less proline-charged tRNAs, resulting in

ribosome stalling at mgtL proline codons. This would favor

formation of stem loop C and transcription elongation into the

mgtA CR (Figure 7A). In contrast, when cytosolic proline levels

are high, the ribosome would translate the complete mgtL

sequence, thereby occluding the left arm of stem loop C and,

in this manner, favor formation of stem-loop B (Figure 7A) and

hinder transcription of the mgtA CR. Interestingly, despite

utilizing different mechanisms to sense proline and Mg2+, the

mgtA leader relies on formation of the same structure—stem

loop C—to promote transcription elongation into the mgtA CR,

because mutations that hinder formation of stem loop C pre-

vented the derepression caused by stop codons in mgtL

(Figure 4A) and by low Mg2+ (Figure 4B) (Cromie et al., 2006).

The regulation of mgtA by changes in the cytosolic proline

concentration is reminiscent of class I transcription attenuation,

where certain amino acid biosynthetic operons are regulated by

translation of a short ORF in the LR. These ORFs are rich in

codons specifying the amino acid(s) synthesized by the enzymes

encoded in the operon (Landick et al., 1996). Yet, there are signif-

icant differences between the two systems. First, mgtA codes for

a transporter rather than for an amino acid biosynthetic enzyme.

This provides experimental support for the proposal, based on

genomic analysis, that classical transcriptional attenuators

might direct the expression of operons mediating functions other

than nutrient biosynthesis (Merino and Yanofsky, 2005). Second,

the position of the leader ORF relative to the potential RNA

secondary structures that can be adopted differs between the

mgtA leader (Figure 7A) and the leaders of attenuation-regulated

amino acid biosynthetic operons such as the trp operon

(Figure 7B). This provides a plausible explanation for the oppo-

site effects that mutation of the leader ORF start codon has in

these two classes of leaders. It results in superattenuation for

the trp operon (Landick and Yanofsky, 1987) because there

would be no ribosomes to stall at Trp codons, which would result

in formation of a terminator structure (Figure 7B). However,

Cell 142, 737–748, September 3, 2010 ª2010 Elsevier Inc. 745

Page 114: Cell 100903

it gives rise to derepression of the mgtA CR (our unpublished

results), because formation of stem loop C would be favored

when mgtL is not translated (Figure 7A). Third, an intrinsic

terminator structure forms in the trp leader when the levels of

charged tRNATrp are high and the ribosome does not pause at

the consecutive Trp codons in the leader ORF (Landick and

Yanofsky, 1987) (Figure 7B). This is in contrast to the absence

of an intrinsic transcription terminator in the mgtA leader, where

sequences located downstream of the stem loop B structure

(Figure 1) are necessary to control transcription elongation into

the mgtA CR (Cromie et al., 2006) by an unknown mechanism.

And fourth, apart from utilizing a classical transcription attenua-

tion-like mechanism to mediate the response to proline, the

mgtA leader senses Mg2+ directly functioning as a riboswitch

(Cromie et al., 2006). This raises the intriguing possibility of clas-

sical transcriptional attenuators also sensing physical and/or

chemical signals directly.

High Osmolarity Enhances the mRNA Levels of the Mg2+

Transporter Gene mgtA

Why does proline limitation promote transcription of the Mg2+

transporter gene mgtA? We contemplated the possibility of the

MgtA protein being a conduit for proline whereby cells experi-

encing low proline would derepress expression of a proline

import system. However, we found that [14C]proline uptake

was similar between wild-type and mgtA Salmonella, and

between a triple mutant lacking all known proline uptake

systems (ProP, ProU, and PutP) (Csonka and Leisinger, 2007)

and the isogenic mgtA derivative (our unpublished results). We

also considered that MgtA-mediated Mg2+ uptake might be

necessary for proline biosynthesis, because the activity of the

proline biosynthetic enzyme glutamate 5-kinase is Mg2+ depen-

dent (Perez-Arellano et al., 2005). Yet, inactivation of the mgtA

gene did not render Salmonella auxotrophic for proline (our

unpublished results).

We determined that hyperosmotic shock promoted an

increase in the mRNA levels corresponding to the mgtA CR

(Figure 6A). This appears to result from a transient decrease in

the levels of charged-proline tRNAs taking place when proline

is used as an osmoprotectant, as there would be less cytosolic

proline available to charge the tRNAPros. Indeed, the increase

in mgtA mRNA levels was severely compromised if the osmotic

shock was alleviated by the osmoprotectant glycine betaine,

which did not influence mgtA expression when added in the

absence of NaCl (Figure 6A). The osmotic shock-promoted

increase in the mRNA levels for the mgtA CR requires the mgtL

proline codons, because a mutant with substitutions in all four

proline codons in mgtL lost the ability to derepress the mgtA

CR mRNA in response to hyperosmotic shock (Figure 6B).

Why does hyperosmotic shock induce transcription of the

Mg2+ transporter gene mgtA when Salmonella also harbors

the constitutively expressed Mg2+ transporter CorA? On the

one hand, high osmolarity promotes excretion of putrescine

(Schiller et al., 2000), which constitutes the major organic diva-

lent cation in bacterial cells and is normally bound to nucleic

acids (Wortham et al., 2007). This might create an increased

need for Mg2+ in order to neutralize the negatively charged

DNA and RNA, and to stabilize structures such as membranes

and ribosomes, which could be affected during hyperosmotic

stress. On the other hand, CorA-mediated Mg2+ uptake may

be compromised because it is driven by the membrane potential

(Froschauer et al., 2004), which decreases in cells experiencing

high osmolarity (Csonka, 1989). Yet, this stress would not affect

Mg2+ uptake by the P-type ATPase MgtA protein because it is

energized by ATP hydrolysis (Maguire, 1992).

That proline limitation promotes expression of the MgtA Mg2+

transporter suggests that there is a physiological connection

between proline and Mg2+. In agreement with this notion, tran-

scription of the proline transporter gene proP is promoted during

growth in low Mg2+ by the PhoP protein (Eguchi et al., 2004),

which directs transcription initiation of the mgtA gene (Garcia

Vescovi et al., 1996).

EXPERIMENTAL PROCEDURES

Bacterial Strains, Plasmids, Primers, and Growth Conditions

Bacterial strains and plasmids used in this study are listed in Table S1. Primers

used in this study are listed in Table S2. Unless otherwise stated, bacteria were

grown at 37�C in LB broth or in N-minimal medium (pH 7.4) (Snavely et al.,

1991) supplemented with 0.1% casamino acids, 38 mM glycerol, and the indi-

cated concentration of MgCl2.

Construction of Strains with Chromosomal Mutations

Deletion strains were constructed using the one-step inactivation method

(Datsenko and Wanner, 2000). See Extended Experimental Procedures for

detailed protocols.

Treatment with Bacteriostatic Protein Synthesis Inhibitors

An overnight bacterial culture grown in N-minimal medium with 10 mM MgCl2was used to inoculate 125 ml flasks containing 10 ml of the same medium

(1:50 dilution) and grown for 3 hr at 37�C with shaking. To promote PhoP-

dependent mgtA transcription initiation, bacteria were washed and resus-

pended in 10 ml of N-minimal medium with 500 mM MgCl2. Following growth

for 1 hr, a 500 ml aliquot was removed to determine the mRNA levels before

treatment. Subsequently, tetracycline (to 25 mg/ml final concentration) or

chloramphenicol (to 200 mg/ml final concentration) was added and bacteria

were grown for 15 min, when they were harvested and their RNA was isolated

for analysis.

Effect of Proline Limitation

Bacteria were grown overnight in modified N-minimal medium containing

0.2% glucose, 10 mM MgCl2, and 1 mM proline. The overnight culture was

used to inoculate 125 ml flasks containing 10 ml of the same medium (1:50 dilu-

tion) and grown for 3 hr at 37�C with shaking. The harvested bacteria were

washed in the modified N-minimal medium containing 500 mM MgCl2, and

grown in 10 ml of the same medium with 1 mM proline for 1 hr. After removing

a 250 ml aliquot to determine the mRNA levels before treatment, the harvested

bacteria were washed with the modified N-minimal medium containing 500 mM

MgCl2, and suspended in 100 ml of the same medium. To see the effect of

proline limitation, the suspended bacterial cells were split into two flasks con-

taining the modified N-minimal media with 500 mM MgCl2 and a mixture of

each of the 19 essential amino acids (50 mM final concentration for each of

them): one flask contained 1 mM proline and the other had no proline added.

Bacteria were grown for 15 min, when they were harvested and their RNA was

isolated for analysis.

Testing Effect of Growth in High versus Low Mg2+

Bacteria were grown overnight in modified N-minimal medium containing

0.2% glucose, 10 mM MgCl2, and 1 mM proline. The harvested bacteria

were washed in the modified N-minimal medium containing 500 mM MgCl2,

and this culture was used to inoculate 125 ml flasks containing 10 ml of the

modified N-minimal medium containing either 500 mM MgCl2 or 10 mM

746 Cell 142, 737–748, September 3, 2010 ª2010 Elsevier Inc.

Page 115: Cell 100903

MgCl2 (1:50 dilution) and grown for 3.5 hr at 37�C with shaking. The bacteria

were then harvested and their RNA was isolated for analysis.

Testing Effect of Mg2+ and/or Proline Limitation Treatment

Bacteria were grown overnight in modified N-minimal medium containing

0.2% glucose, 10 mM MgCl2, and 1 mM proline. The overnight culture

was used to inoculate 125 ml flasks containing 20 ml of the same medium

(1:50 dilution) and grown for 3 hr at 37�C with shaking. The harvested bacteria

were washed in the modified N-minimal medium containing 500 mM MgCl2,

and grown in 20 ml of the same medium with 1 mM proline for 1 hr. The

harvested bacteria were washed with the modified N-minimal medium

containing 500 mM MgCl2, and suspended in 200 ml of the same medium.

This cell suspension was used to inoculate 125 ml flasks containing 5 ml of

media containing either 500 mM MgCl2 and 1 mM proline, 500 mM MgCl2and no proline, no MgCl2 and 1 mM proline, or no MgCl2 and no proline

(1:100 dilution). Bacteria were grown for 15 min, when they were harvested

and their RNA was isolated for analysis.

Testing Effect of Hyperosmotic Shock

Bacteria were grown overnight in modified N-minimal medium containing

0.2% glucose, 0.1% casamino acids, and 10 mM MgCl2. The overnight culture

was used to inoculate 125 ml flasks containing 20 ml of the same medium

(1:50 dilution) and grown for 3.5 hr at 37�C with shaking. The harvested

bacteria were washed in modified N-minimal medium containing 500 mM

MgCl2 without casamino acids and suspended in 250 ml of the same medium.

This cell suspension was used to inoculate 125 ml flasks containing 5 ml of the

modified N-minimal medium without casamino acids containing 500 mM MgCl2and either no additional supplements, 1 mM glycine betaine, 0.3 M NaCl, 0.3 M

NaCl and 1 mM glycine betaine, or 0.3 M NaCl and 1 mM proline. Bacteria were

grown for 1 hr at 37�C with shaking, when they were harvested and their RNA

was isolated for analysis.

RNA Isolation and Determination of Transcript Levels

Total RNA was extracted using an RNeasy Mini Kit (QIAGEN). cDNA was

synthesized using TaqMan reverse-transcription reagents (Applied Biosys-

tems) following the manufacturer’s instructions. Quantification of transcripts

was performed by real-time PCR using Fast SYBR Green Master Mix (Applied

Biosystems) in a 7500 Sequence Detection System (Applied Biosystems). A list

of primers used for quantitative RT-PCR is presented in Extended Experi-

mental Procedures. See Extended Experimental Procedures for a detailed

protocol.

b-Galactosidase Assays

b-galactosidase activity was determined as described (Cromie et al., 2006).

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures, two

figures, and two tables and can be found with this article online at doi:10.

1016/j.cell.2010.07.046.

ACKNOWLEDGMENTS

We thank Laszlo Csonka, Kerry Hollands, Robert Landick, and Charles

Yanofsky for comments on the manuscript; Henry Huang for the plasmid

harboring supF; and John Roth for strains. This work was supported, in part,

by grant AI49561 from the NIH to E.A.G., who is an Investigator of the Howard

Hughes Medical Institute.

Received: March 19, 2010

Revised: May 24, 2010

Accepted: July 14, 2010

Published: September 2, 2010

REFERENCES

Bader, M.W., Sanowar, S., Daley, M.E., Schneider, A.R., Cho, U., Xu, W.,

Klevit, R.E., Le Moual, H., and Miller, S.I. (2005). Recognition of antimicrobial

peptides by a bacterial sensor kinase. Cell 122, 461–472.

Barchiesi, J., Castelli, M.E., Soncini, F.C., and Vescovi, E.G. (2008). mgtA

expression is induced by rob overexpression and mediates a Salmonella en-

terica resistance phenotype. J. Bacteriol. 190, 4951–4958.

Blanc-Potard, A.B., and Groisman, E.A. (1997). The Salmonella selC locus

contains a pathogenicity island mediating intramacrophage survival. EMBO

J. 16, 5376–5385.

Brennan, R.G., and Link, T.M. (2007). Hfq structure, function and ligand

binding. Curr. Opin. Microbiol. 10, 125–133.

Cayley, S., Lewis, B.A., and Record, M.T., Jr. (1992). Origins of the osmopro-

tective properties of betaine and proline in Escherichia coli K-12. J. Bacteriol.

174, 1586–1595.

Chen, J.W., Bennett, D.C., and Umbarger, H.E. (1991). Specificity of attenua-

tion control in the ilvGMEDA operon of Escherichia coli K-12. J. Bacteriol. 173,

2328–2340.

Choi, E., Groisman, E.A., and Shin, D. (2009). Activated by different signals, the

PhoP/PhoQ two-component system differentially regulates metal uptake.

J. Bacteriol. 191, 7174–7181.

Cromie, M.J., and Groisman, E.A. (2010). Promoter and riboswitch control of

the Mg2+ transporter MgtA from Salmonella enterica. J. Bacteriol. 192,

604–607.

Cromie, M.J., Shi, Y., Latifi, T., and Groisman, E.A. (2006). An RNA sensor for

intracellular Mg2+. Cell 125, 71–84.

Csonka, L.N. (1989). Physiological and genetic responses of bacteria to

osmotic stress. Microbiol. Rev. 53, 121–147.

Csonka, L.N., and Leisinger, T. (2007). Biosynthesis of proline. In EcoSal—Es-

cherichia coli and Salmonella: Cellular and Molecular Biology, G. Cohen, ed.

(Washington, DC: ASM Press), http://www.ecosal.org.

Datsenko, K.A., and Wanner, B.L. (2000). One-step inactivation of chromo-

somal genes in Escherichia coli K-12 using PCR products. Proc. Natl. Acad.

Sci. USA 97, 6640–6645.

Eguchi, Y., Okada, T., Minagawa, S., Oshima, T., Mori, H., Yamamoto, K., Ish-

ihama, A., and Utsumi, R. (2004). Signal transduction cascade between EvgA/

EvgS and PhoP/PhoQ two-component systems of Escherichia coli. J. Bacter-

iol. 186, 3006–3014.

Froschauer, E.M., Kolisek, M., Dieterich, F., Schweigel, M., and Schweyen,

R.J. (2004). Fluorescence measurements of free [Mg2+] by use of mag-fura 2

in Salmonella enterica. FEMS Microbiol. Lett. 237, 49–55.

Garcia Vescovi, E., Soncini, F.C., and Groisman, E.A. (1996). Mg2+ as an extra-

cellular signal: environmental regulation of Salmonella virulence. Cell 84,

165–174.

Grundy, F.J., and Henkin, T.M. (2006). From ribosome to riboswitch: control of

gene expression in bacteria by RNA structural rearrangements. Crit. Rev. Bio-

chem. Mol. Biol. 41, 329–338.

Hammarlof, D.L., and Hughes, D. (2008). Mutants of the RNA-processing

enzyme RNase E reverse the extreme slow-growth phenotype caused by

a mutant translation factor EF-Tu. Mol. Microbiol. 70, 1194–1209.

Henkin, T.M. (2008). Riboswitch RNAs: using RNA to sense cellular metabo-

lism. Genes Dev. 22, 3383–3390.

Henkin, T.M., and Yanofsky, C. (2002). Regulation by transcription attenuation

in bacteria: how RNA provides instructions for transcription termination/anti-

termination decisions. Bioessays 24, 700–707.

Landick, R., and Yanofsky, C. (1987). Transcription attenuation. In Escherichia

coli and Salmonella Cellular and Molecular Biology, J.L. Ingraham, ed. (Wash-

ington, DC: ASM Press).

Landick, R., Turnbough, C.L., and Yanofsky, C. (1996). Transcription attenua-

tion. In Escherichia coli and Salmonella Cellular and Molecular Biology, F.C.

Neidhardt, ed. (Washington, DC: ASM Press).

Cell 142, 737–748, September 3, 2010 ª2010 Elsevier Inc. 747

Page 116: Cell 100903

Laursen, B.S., Sorensen, H.P., Mortensen, K.K., and Sperling-Petersen, H.U.

(2005). Initiation of protein synthesis in bacteria. Microbiol. Mol. Biol. Rev. 69,

101–123.

Lejona, S., Aguirre, A., Cabeza, M.L., Garcia Vescovi, E., and Soncini, F.C.

(2003). Molecular characterization of the Mg2+-responsive PhoP-PhoQ regu-

lon in Salmonella enterica. J. Bacteriol. 185, 6287–6294.

Loh, E., Dussurget, O., Gripenland, J., Vaitkevicius, K., Tiensuu, T., Mandin, P.,

Repoila, F., Buchrieser, C., Cossart, P., and Johansson, J. (2009). A trans-

acting riboswitch controls expression of the virulence regulator PrfA in Listeria

monocytogenes. Cell 139, 770–779.

Maguire, M.E. (1992). MgtA and MgtB: prokaryotic P-type ATPases that

mediate Mg2+ influx. J. Bioenerg. Biomembr. 24, 319–328.

Maguire, M.E. (2006). Magnesium transporters: properties, regulation and

structure. Front. Biosci. 11, 3149–3163.

Merino, E., and Yanofsky, C. (2005). Transcription attenuation: a highly

conserved regulatory strategy used by bacteria. Trends Genet. 21, 260–264.

O’Connor, K., Fletcher, S.A., and Csonka, L.N. (2009). Increased expression of

Mg2+ transport proteins enhances the survival of Salmonella enterica at high

temperature. Proc. Natl. Acad. Sci. USA 106, 17522–17527.

Perez-Arellano, I., Rubio, V., and Cervera, J. (2005). Dissection of Escherichia

coli glutamate 5-kinase: functional impact of the deletion of the PUA domain.

FEBS Lett. 579, 6903–6908.

Prost, L.R., Daley, M.E., Le Sage, V., Bader, M.W., Le Moual, H., Klevit, R.E.,

and Miller, S.I. (2007). Activation of the bacterial sensor kinase PhoQ by acidic

pH. Mol. Cell 26, 165–174.

Schiller, D., Kruse, D., Kneifel, H., Kramer, R., and Burkovski, A. (2000). Poly-

amine transport and role of potE in response to osmotic stress in Escherichia

coli. J. Bacteriol. 182, 6247–6249.

Shin, D., Lee, E.J., Huang, H., and Groisman, E.A. (2006). A positive feedback

loop promotes transcription surge that jump-starts Salmonella virulence

circuit. Science 314, 1607–1609.

Snavely, M.D., Miller, C.G., and Maguire, M.E. (1991). The mgtB Mg2+ trans-

port locus of Salmonella typhimurium encodes a P-type ATPase. J. Biol.

Chem. 266, 815–823.

Soncini, F.C., Garcia Vescovi, E., Solomon, F., and Groisman, E.A. (1996).

Molecular basis of the magnesium deprivation response in Salmonella typhi-

murium: identification of PhoP-regulated genes. J. Bacteriol. 178, 5092–5099.

Spinelli, S.V., Pontel, L.B., Garcia Vescovi, E., and Soncini, F.C. (2008). Regu-

lation of magnesium homeostasis in Salmonella: Mg2+ targets the mgtA tran-

script for degradation by RNase E. FEMS Microbiol. Lett. 280, 226–234.

Sudarsan, N., Hammond, M.C., Block, K.F., Welz, R., Barrick, J.E., Roth, A.,

and Breaker, R.R. (2006). Tandem riboswitch architectures exhibit complex

gene control functions. Science 314, 300–304.

Turnbough, C.L., Jr., and Switzer, R.L. (2008). Regulation of pyrimidine biosyn-

thetic gene expression in bacteria: repression without repressors. Microbiol.

Mol. Biol. Rev. 72, 266–300.

Winkler, W.C., and Breaker, R.R. (2005). Regulation of bacterial gene expres-

sion by riboswitches. Annu. Rev. Microbiol. 59, 487–517.

Wortham, B.W., Patel, C.N., and Oliveira, M.A. (2007). Polyamines in bacteria:

pleiotropic effects yet specific mechanisms. Adv. Exp. Med. Biol. 603,

106–115.

Yanofsky, C., and Horn, V. (1994). Role of regulatory features of the trp operon

of Escherichia coli in mediating a response to a nutritional shift. J. Bacteriol.

176, 6245–6254.

748 Cell 142, 737–748, September 3, 2010 ª2010 Elsevier Inc.

Page 117: Cell 100903

Structural Basis of Semaphorin-PlexinRecognition and Viral Mimicry fromSema7A and A39R Complexes with PlexinC1Heli Liu,1 Z. Sean Juo,2,3,4 Ann Hye-Ryong Shim,1 Pamela J. Focia,1 Xiaoyan Chen,1 K. Christopher Garcia,2,3,4

and Xiaolin He1,*1Department of Molecular Pharmacology and Biological Chemistry, Northwestern University Feinberg School of Medicine, Searle 8-417,

303 East Chicago Avenue, Chicago, IL 60611, USA2Howard Hughes Medical Institute3Department of Molecular and Cellular Physiology4Department of Structural Biology

Stanford University School of Medicine, Beckman B171B, 279 Campus Drive, Stanford, CA 94305, USA*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.07.040

SUMMARY

Repulsive signaling by Semaphorins and Plexins iscrucial for the development and homeostasis of thenervous, immune, and cardiovascular systems.Sema7A acts as both an immune and a neural Sem-aphorin through PlexinC1, and A39R is a Sema7Amimic secreted by smallpox virus. We report thestructures of Sema7A and A39R complexed withthe Semaphorin-binding module of PlexinC1. Bothstructures show two PlexinC1 molecules symmetri-cally bridged by Semaphorin dimers, in which theSemaphorin and PlexinC1 b propellers interact inan edge-on, orthogonal orientation. Both bindinginterfaces are dominated by the insertion of the Sem-aphorin’s 4c-4d loop into a deep groove in blade 3 ofthe PlexinC1 propeller. A39R appears to achieveSema7A mimicry by preserving key Plexin-bindingdeterminants seen in the mammalian Sema7Acomplex that have evolved to achieve higher affinitybinding to the host-derived PlexinC1. The complexstructures support a conserved Semaphorin-Plexinrecognition mode and suggest that Plexins are acti-vated by dimerization.

INTRODUCTION

Semaphorins and their receptors, Plexins, are two families of

widely expressed proteins whose respective structures and

functions are conserved across the animal kingdom. Originally

identified as ligand-receptor pairs that control axon guidance

by repulsion during central nervous system (CNS) development

(Kolodkin et al., 1993; Luo et al., 1993; Tamagnone et al.,

1999; Tessier-Lavigne and Goodman, 1996), Semaphorins and

Plexins have been shown to serve as path-finding controls for

a diverse array of additional functions in physiology, including

vascularization and angiogenesis (normal and pathological),

organogenesis, and immune responses (Kruger et al., 2005;

Suzuki et al., 2008). Semaphorins are also associated with tumor

progression and other diseases (reviewed in Kruger et al., 2005;

Tamagnone and Comoglio, 2000). Within the neural system,

Semaphorin-Plexin signaling is implicated well beyond axon

guidance, ranging from axon pruning to synaptic formation,

specificity, and plasticity (reviewed in He et al., 2002; Waimey

and Cheng, 2006). While Plexins are the predominant receptors

for Semaphorins, the alternative Semaphorin receptors, Neuro-

pilin-1 and -2, differ from Plexins in structure and function

and appear to serve as obligate coreceptors for either Sema-

phorin signaling or VEGF signaling in a variety of specific cellular

(neuronal, vascular, etc.) contexts (Pellet-Many et al., 2008;

Uniewicz and Fernig, 2008).

Semaphorins have been grouped into eight classes on the

basis of primary sequence and source, where classes 3–7 are

vertebrate Semaphorins and class V are viral Semaphorins

(Semaphorin Nomenclature Committee, 1999). The Semaphor-

ins are characterized by an N-terminal �500 amino acid (aa)

Sema domain that is essential for signaling through Plexins

(Koppel et al., 1997). The crystal structures of the Sema domains

of Sema3A and Sema4D (also known as CD100) have been

determined, showing that the Sema domain is a seven-bladed

b propeller, and appear to exist as a homodimer by virtue of

conserved structural elements (Antipenko et al., 2003; Love

et al., 2003). There are four subfamilies of vertebrate Plexins

(A, B, C, and D), which are all type I transmembrane glycopro-

teins featuring an extracellular segment containing an N-terminal

Sema domain, followed by variable numbers of PSI (‘‘found in

Plexins, Semaphorins and Integrins’’) domains, immunoglob-

ulin-like (Ig) domains (Bork et al., 1999), and an intracellular

GTPase-activating (GAP) domain that regulates Rho family

GTPases (Oinuma et al., 2004). The structures of the intracellular

domains of two Plexins have been solved (He et al., 2009; Tong

et al., 2009), but three-dimensional structural information does

not exist for the Plexin extracellular segment or for the Sema-

phorin-Plexin interaction.

In the immune system, a subset of Semaphorins are active

and are designated as ‘‘immune Semaphorins.’’ These include

Cell 142, 749–761, September 3, 2010 ª2010 Elsevier Inc. 749

Page 118: Cell 100903

Sema4D, Sema4A, and Sema7A, which appear to modulate

a variety of immune responses ranging from thymic selection

to B cell homing (Suzuki et al., 2008). Sema7A is a GPI-anchored

cell-surface glycoprotein expressed on activated lymphocytes

and thymocytes (Xu et al., 1998; Yamada et al., 1999), and its

receptor is PlexinC1 (also named VESPR or CD232) (Tamagnone

et al., 1999). The Sema7A/PlexinC1 interaction was originally

shown to induce the activation of monocytes (Holmes et al.,

2002), but recently this receptor-ligand pair has been shown to

regulate melanocyte adhesion, and silencing of PlexinC1 is

seen during the development and progression of melanoma

(Scott et al., 2009). Sema7A also has been suggested to have

neuronal functions, by promoting axon growth through integrins

(Pasterkamp et al., 2003). Underscoring the importance of Sem-

aphorin/Plexin interactions in the immune system, viruses have

evolved proteins that engage the immune Semaphorin/Plexin

system, presumably to enhance virus survival in the host. The

most well-characterized examples derive from Vaccinia

(smallpox) virus and Alcelaphine herpesvirus, which encode

secreted Semaphorin homologs A39R and AHVsema (Comeau

et al., 1998; Kolodkin et al., 1993). These viral Semaphorins

show sequence similarity to Sema7A and can also bind to and

activate PlexinC1 (Comeau et al., 1998). The viral Semaphorins

probably play an immunomodulatory role by mimicking Sema7A;

an antibody against PlexinC1 inhibits A39R-induced induction of

inflammatory cytokines by monocytes (Comeau et al., 1998).

We present here the crystal structures of both the Sema7A/

PlexinC1 complex and the A39R/PlexinC1 complex, revealing

the basic architecture of a Semaphorin/Plexin recognition com-

plex. Our data point to a conserved mode of recognition of

Plexins by Semaphorins and provide insights into Plexin activa-

tion and viral mimicry.

RESULTS

Biochemical Analysis of PlexinC1 Bindingby the Mammalian Semaphorin Sema7Aand the Viral Semaphorin A39RThe extracellular domains of Semaphorins and Plexins are large

and highly glycosylated; thus we expressed recombinant forms

of PlexinC1, Sema7A and A39R using baculovirus-mediated

mammalian cell gene transduction (BacMam) (Dukkipati et al.,

2008) and purified the secreted proteins from HEK293H cells

or N-acetylglucosaminyltransferase I-deficient HEK293 (GnTI-

HEK293) cells (Reeves et al., 2002). Our construct designs

were informed by prior structural analysis of Sema4D (Love

et al., 2003), which showed that the first membrane-distal PSI

domain is integrally associated with the Sema domain. Deletion

of the PSI domains in either PlexinC1 or Sema7A destabilized the

proteins. Ultimately, after testing a variety of constructs, for

PlexinC1, we expressed the Sema domain plus the first PSI

domain (SemaPSI), and the entire extracellular segment (ECD).

For Sema7A, we expressed the entire extracellular segment

excluding the GPI anchor; for A39R, we expressed the full-length

protein, which is a ‘‘minimal’’ secreted Semaphorin that does not

contain a PSI domain (Figure 1A).

To analyze the solution binding of PlexinC1 to its A39R and

Sema7A ligands, we used isothermal titration calorimetry (ITC)

(Figures 1B–1E). In HEPES-buffered saline (HBS), Sema7A

bound PlexinC1-SemaPSI with a KD of �290 nM (Figure 1B).

The ITC measurements for Sema7A/PlexinC1 were complicated

by the small enthalpy change (��2 kcal/mol), which resulted in

shallow titration curves with poor signal-to-noise ratio. There-

fore, we used high concentrations of the samples and injected

large amounts of protein so as to maximize the heat per injection,

and also to fully saturate the titration curve. The Sema7A/Plex-

inC1 KD measured here by calorimetry using soluble proteins,

albeit in the nanomolar range, is weaker than the reported KD

of 2.1nM from Scatchard analysis using intact receptors in the

cell membrane (Tamagnone et al., 1999). Previously reported

Semaphorin/Plexin affinity measurements are primarily cell

based and generally in the low nanomolar KD range. We attribute

the lower-affinity values of the soluble recombinant proteins to

the lack of membrane confinement in the ITC format, not to

structural differences in the recombinant proteins. Similar differ-

ences between solution-based and cell-based affinities have

been seen for ligand binding to other receptors, such as SCF/

KIT (55 nM versus 2 nM) (Lemmon et al., 1997; Lev et al.,

1992), CNP/NPRB (2.7 nM versus 0.03 nM) (He et al., 2006; Kol-

ler and Goeddel, 1992), and MHC/TCR (micromolar versus nano-

molar) (Huppa et al., 2010; Krogsgaard et al., 2003). We also

determined that the affinity of Sema7A for the truncated Plex-

inC1-SemaPSI and full-length PlexinC1 ECD were nearly iden-

tical (Figures 1B and 1C), indicating the Ig and PSI domains C-

terminal to the PlexinC1-SemaPSI module are not significantly

involved in Semaphorin recognition. Thus, for structural studies

we proceeded with crystallization of the Sema7A/PlexinC1-

SemaPSI complex.

The viral-derived A39R bound PlexinC1-SemaPSI with a KD of

�9.4 nM (Figure 1D). The large enthalpy change (�14.6 kcal/mol)

facilitated the obtainment of high-quality binding curves. We also

characterized A39R for the effect of various PlexinC1 ECD trun-

cations on affinity. A39R affinity and thermodynamic parameters

for full-length PlexinC1-ECD (KD �8.9 nM) were nearly identical

to that of A39R binding to PlexinC1-SemaPSI (Figures 1D and

1E). Thus, similar to Sema7A/PlexinC1 binding, this indicated

that the N-terminal Sema-PSI domains of PlexinC1 were suffi-

cient for Semaphorin binding and represented an appropriate

form for cocrystallization of both the Sema7A/PlexinC1 and

A39R/PlexinC1 complexes.

Structures of the Sema7A/PlexinC1 Complex, FreeA39R, and the A39R/PlexinC1 ComplexTo obtain diffraction quality crystals of the heavily glycosylated

Sema7A/PlexinC1-SemaPSI complex (four N-linked glycosy-

lated sites on Sema7A and seven on PlexinC1-SemaPSI), we

used Endo-H to trim the N-linked glycans attached to the GnTI�

HEK293-expressed Sema7A and PlexinC1-SemaPSI proteins.

This treatment leaves a single N-acetylglucosamine (GlcNAc)

residue remaining at each occupied N-linked glycan site,

preserving the core glycosylation of the proteins. The glycan-

trimmed proteins exhibited the same solution behavior to wild-

type proteins, e.g., the Semaphorins existed exclusively as

dimers after glycan-trimming, as assessed by gel filtration chro-

matography (Figure S1 available online). The Sema7A/PlexinC1-

SemaPSI complex structure was determined to a resolution of

750 Cell 142, 749–761, September 3, 2010 ª2010 Elsevier Inc.

Page 119: Cell 100903

2.4 A via the method of single isomorphous replacement with

anomalous scattering (SIRAS) (Table S1, Figure S2). The asym-

metric unit of the crystal contains one dimeric complex in a 2:2

stoichiometry, comprised of a central Sema7A dimer and two

monomeric PlexinC1-SemaPSI molecules (Figure 2A).

Overall, the shape of the complex resembles a crab where

Sema7A comprises the body, and the two PlexinC1 molecules

resemble pincers emanating radially from the body at four and

eight o’clock. The approximate dimensions of the complex are

80 A 3 140 A 3 160 A (Figure 2A). The head-to-head docking

architecture of the complex is in an ideal orientation for interac-

tion of two cell surface-associated proteins across a cell-cell

junction. In the complex, both Sema7A and PlexinC1 contain

large, disk-shaped Sema domains composed of seven-bladed

b propellers, each of which is intimately associated with its

respective PSI domain and, in the case of Sema7A, a single Ig

domain that would lead to the membrane. In the complex, the

Sema domains mediate the vast majority of the receptor-ligand

contact. The center-to-center distances between the Sema

propellers are 50 A for Sema7A-Sema7A, 55 A for Sema7A-Plex-

inC1, and 80 A for PlexinC1-PlexinC1.

Interestingly, the Sema domains of Sema7A and PlexinC1

interact ‘‘edge-on’’ using their sides to contact one another,

rather than the top or bottom faces, as was generally expected.

The planes of the Sema7A and PlexinC1 b propellers are orthog-

onally related. The orientation shown in Figure 2 places the

C-termini of Sema7A close to the cell membrane, leading to

the GPI anchor. The C terminus of the PSI domain of PlexinC1,

which is �160 A from the Semaphorin C terminus, leads to the

opposing cell membrane, albeit through one additional PSI and

four Ig domains not present in the current structure. Because

the remaining PSI and Ig domains of PlexinC1 would appear to

emanate away from the central dyad axis of the complex, they

are unlikely to contact Sema7A.

We expressed A39R in baculovirus and determined the

unliganded A39R structure, to a resolution of 2.0 A, by SIRAS

(Table S1). The structure of A39R shows a dimer of minimal

Sema domains with no PSI or Ig domains (Figure 2D). The

Figure 1. Characterization of the Binding of Sema7A and A39R to PlexinC1

(A) Domain diagram of PlexinC1, Sema7A, and A39R and the constructs used in characterization. The cartoon key for the diagrams is shown to the right.

(B–E) Profiles of the binding of PlexinC1 to Sema7A and A39R as measured by isothermal titration calorimetry. The binding parameters are indicated below the

traces.

See also Figure S1 for gel filtration profiles showing that Sema7A and A39R exist as dimers.

Cell 142, 749–761, September 3, 2010 ª2010 Elsevier Inc. 751

Page 120: Cell 100903

Sema domains of A39R and Sema7A are clearly built on the

same scaffold, with a root-mean-square deviation (rmsd) of

1.5 A for matching Ca atoms (Figure S3). Major structural differ-

ences are located only at the N-terminal segment and several

loops remote from the Plexin-binding site of Sema7A. Sema7A,

like Sema3A and Sema4D, has a long N-terminal segment that

provides an additional, fifth b strand on the outer edge of blade

6, whereas A39R lacks this segment (Figure S3).

We followed the same approach as Sema7A/PlexinC1-

SemaPSI to obtain the crystals of the A39R/PlexinC1-SemaPSI

complex, which we solved by molecular replacement (Table

S1). The architecture of the A39R/PlexinC1 complex is similar

to that of Sema7A/PlexinC1, except that as a result of the lack

of PSI and Ig domains in A39R, it is shorter in its longest dimen-

sion (Figure 2C). As in the Sema7A/PlexinC1 complex, the Sema

domains mediate all the A39R-PlexinC1 contact. The orienta-

tions of the Sema propellers in the complex are nearly identical

to those in Sema7A/PlexinC1, as is the ‘‘edge-on’’ interaction

mode between the ligand and receptor b propellers.

Dimerization of Sema7A and A39R Are Mediated byConserved Structural Elements with Varied ChemistryDimerization appears to be a general and probably important

property of Semaphorins (Antipenko et al., 2003; Love et al.,

2003) that we can now examine within the context of a receptor

complex. The Sema7A molecules in the complex, consisting of

Sema, PSI, and Ig domains, form a dimer roughly similar to the

Sema4D dimer (Love et al., 2003) and the Sema3A dimers that

only contain the Sema domains (Antipenko et al., 2003)

(Figure S2C). This is consistent with our gel filtration analysis

Figure 2. Structures of the Sema7A/PlexinC1-SemaPSI and A39R/PlexinC1-SemaPSI Complexes

(A) Ribbon models of the Sema7A/PlexinC1-SemaPSI complex in front view (left) and side view (right), with the Sema7A protomers colored in cyan and blue, and

the PlexinC1-SemaPSI protomers in pink and magenta. The N-linked glycans are depicted as sticks with carbon atoms colored in green. A cartoon of a membrane

is drawn above and below the complex to indicate where the respective proteins would be attached to the cell surfaces.

(B) The structure of an individual PlexinC1-SemaPSI molecule from the complex in two orthogonal views, with each of the seven b -propeller blades, the extrusion,

the flap, and the PSI domain individually colored.

(C) Ribbon model of the A39R/PlexinC1-Sema-PSI complex in front view (left) and side view (right), with the A39R protomers colored in yellow and wheat and the

other components colored similarly to (A).

(D) Ribbon model of an A39R protomer from the free A39R dimer, with the structural modules colored in the same format as PlexinC1 shown in (B).

See also Table S1 and Figure S2 for crystallographic statistics and structural comparisons.

752 Cell 142, 749–761, September 3, 2010 ª2010 Elsevier Inc.

Page 121: Cell 100903

(Figure S1). The Sema7A dimer interface buries 2860 A2 of

solvent-accessible surface area. Two sets of three loops on

the top face of the Sema domain, 4b-4c, 4d-5a, and 5b-5c are

intertwined at the dimer interface. In comparison, both Sema4D

and Sema3A use an additional loop for dimerization

(Figure S2C). The Sema7A Sema-Sema interface is primarily

hydrophobic. The Ig domain of Sema7A is also involved in the

dimer interface but only contributes �15% to the total buried

surface area. The Ig domain of Sema7A, unlike the Sema4D Ig

domain, which is canonical, is an unusual variation of the Ig

fold, with a 5-on-2 topology in which strand D switches from

the BED b sheet to join the AFGC b sheet (Figure S2D).

The viral Semaphorin A39R also exists as a dimer both in solu-

tion and in the crystal (Figure S1). Dimerization of A39R is exclu-

sively mediated by its Sema domain, since this protein does not

have PSI or Ig domains, and the dimer interface is considerably

smaller (�1990 A2) than that seen on the canonical Semaphorins

(Figure S2C). A39R uses the same set of loops as the Sema7A

Sema for dimerization, but in contrast to the hydrophobic

Sema7A dimer interface, the A39R dimer interface is composed

almost exclusively of hydrophilic interactions involving six salt

bridges.

The Sema7A and A39R structures, together with the previous

Sema3A and Sema4D structures, indicate that although the

dimer interfaces of Semaphorins can vary significantly in

sequence and chemical nature, the general structural mode of

domain dimerization is conserved, and this is likely to be impor-

tant for the appropriate dimerization geometry of the bound

Plexin receptors for signaling. The preservation of the dimeriza-

tion geometry despite vastly different interface chemistries

suggests that there is evolutionary pressure to achieve this

dimerization mode for Semaphorin function, presumably to

orient bound Plexins for signaling.

The Structure of PlexinC1-SemaPSI Has UniqueFeaturesGiven that the extracellular segments of Plexins have not been

structurally characterized, the fold of the PlexinC1 Sema-PSI

domains merits some description. The Sema domain of PlexinC1

in the complex is generally similar to other seven-bladed

b propeller domains in topology but has unique features that

are distinct from Semaphorins or MET (Stamos et al., 2004)

(Figure 2B). After the common nomenclature used by other

propeller proteins, blade 7 of the PlexinC1 Sema domain is

C-terminally adjacent to blade 1; each of the seven blades is

formed by four antiparallel b strands with strands a-d from the

inside to the outside of the b propeller (Figure 2B). The surface

bears the loops linking strands b and c (e.g., loop 4b-4c) and link-

ing strands d and a (e.g., loop 5d-6a) on the top face. Loop 1d-2a

traverses the top of the propeller like a flap, sealing the central

channel of the propeller, which is hollow in Semaphorins, MET,

and integrins (Figure S2E) (Antipenko et al., 2003; Love et al.,

2003; Stamos et al., 2004; Xiao et al., 2004; Xiong et al., 2002).

A long insertion, termed the ‘‘extrusion’’ (Love et al., 2003), is

located between strands 5c and 5d. A major structural difference

is that the extrusion of PlexinC1 is shorter than that of Semaphor-

ins. The structure of the complex shows that this extrusion in

Semaphorins is critical for binding Plexins, but the extrusion in

Plexins is not a central part of the interface. It appears, then,

that Plexins have lost the prominence of this structural element

as the relative functions of the Sema domain in Semaphorins

versus Plexins specialized over the course of evolution

(Figure S2E) (Antipenko et al., 2003; Love et al., 2003; Stamos

et al., 2004).

The PSI domain of PlexinC1 is a small cysteine-rich domain

similar to that of Semaphorins and integrins, but its orientation

relative to the Sema domain is different from that in Semaphorins

by an �20� rotation (Figure S2E). The PSI domain is intimately

packed against the Sema domain by a broad array of primarily

hydrophobic interactions, which would facilitate the sensitive

structural transmission of Sema binding from the membrane-

distal to the membrane-proximal regions of PlexinC1.

The Sema7A-PlexinC1 Interaction Featuresa ‘‘Loop-in-Groove’’ Recognition ModeThe edge-on, orthogonal stacking of the respective b propellers

in the Sema7A-PlexinC1 interface (Figure 3A) buries a total of

�2100 A2 solvent-accessible surface area. The extensive inter-

face can be divided into three principal regions: (1) The 4c-4d

loop of Sema7A inserting into the groove on PlexinC1 that is

bounded by walls composed of the PlexinC1 loop 3b-3c and

the bulged strand 3d—this PlexinC1 groove has an open and

a closed (obstructed) end (Figure 3B); (2) the extrusion helix 2

of Sema7A contacting the loop 3b-3c of PlexinC1 at one side

of the groove; and (3) a small area of contact between the

Sema7A blade 3, and the PlexinC1 strand 3d, that forms the

opposing wall of the groove (Figures 3C–3E).

The central ‘‘loop-in-groove’’ interaction (Figure 3D) features

ten hydrogen bonds (all hydrogen bonds discussed are pre-

dicted from geometry) between the protruding 4c-4d loop of

Sema7A and the PlexinC1 groove (Table S2). The polar interac-

tions are supplemented by two hydrophobic residues (Leu276

and Val278) in the Sema7A 4c-4d loop, which contact several

hydrophobic residues in the PlexinC1 groove through van der

Waals interactions. Overall, the PlexinC1/Sema7A loop-in-

groove interaction includes a mixture of hydrophobic residues

lining the groove wall, interspersed with hydrophilic residues,

to presumably provide ligand specificity through polar interac-

tions. Importantly, at the edge of this groove in PlexinC1,

Sema7A Lys280 forms a salt bridge with PlexinC1 Asp200, and

this interaction appears to be a key component of A39R mimicry

(discussed below).

At the obstructed (closed) end of the groove (Figure 3B), the

Sema7A extrusion helix 2 interacts in a roughly parallel manner

with the PlexinC1 3b-3c loop, also showing several ancillary

long-range interactions with the tips of the PlexinC1 2b-2c and

2d-3a loops (Figure 3C). The interaction is intimate only at the

PlexinC1 Ala197-Ala198-Ser199 bulge. The obstructed end of

the PlexinC1 groove is largely due to the steric bulk sidechain

of Arg131, which forms a salt bridge with Sema7A Glu376

(Figure 3C).

In the third region of the interface, at the open end of the Plex-

inC1 groove (Figure 3B), the outer edge of Sema7A blade 3 forms

extended interactions with the bulged strand 3d of PlexinC1

(Figure 3E). There are two spatially separate clusters of salt

bridges in this region of the interface. The first cluster involves

Cell 142, 749–761, September 3, 2010 ª2010 Elsevier Inc. 753

Page 122: Cell 100903

Sema7A residues Arg204 and Arg202 both forming salt bridges

with PlexinC1 residue Glu219. Sema7A Tyr213 occupies the

space between this salt bridge and a salt bridge involving the

important Sema7A residue Lys280 and PlexinC1 Asp200 (Fig-

ure S2A). The second cluster of charged interactions involve

Sema7A Asp216, which forms bifurcated salt bridges with

Arg222 and Lys224 of PlexinC1. Mutagenesis data support that

the interactions at this region are important for Sema7A-PlexinC1

binding (Figure S4). In the complex structure, the location of the

ridges flanking the groove in the Semaphorin binding site is

consistent with a previous mutagenesis study implicating

the region between residues 166–235 of the Sema domain of

Sema3A in its Plexin-binding specificity (Koppel et al., 1997). In

summary, the Sema7A/PlexinC1 interface is extensive and varied

in chemical and structural character and is dominated by the

insertion of a long loop in Sema7A into a deep groove in PlexinC1.

Figure 3. The Interface between Sema7A and PlexinC1-SemaPSI

(A) The overall structure of an interacting pair of Sema7A and PlexinC1-SemaPSI protomers, highlighting the structural elements involved in their binding, colored

cyan for Sema7A and pink for PlexinC1-SemaPSI.

(B) Close-up view of the interface with Sema7A depicted as ribbons and PlexinC1-SemaPSI in a surface representation, showing how the protruding loop 4c-4d

of Sema7A inserts into a groove in PlexinC1 blade 3, and is flanked by the contacts between the Sema7A extrusion helix 2 and the PlexinC1 loop 3b-3c, and the

contacts between the blade 3 of Sema7A and the bulged strand 3d of PlexinC1.

(C) The interactions of residues near the obstructed end of the PlexinC1 groove, with Sema7A in chocolate and PlexinC1 in pink.

(D) The interaction between residues of the 4c-4d loop of Sema7A (cyan) and the PlexinC1 groove (pink).

(E) The interaction between the residues of blade 3 of Sema7A (green) and strand 3d of PlexinC1 (pink). Note that residues 218–220 of PlexinC1 comprise a bulge

from strand 3d.

See also Table S2 and Figure S4 for a list of the interactions and for mutagenesis/binding data on this interface, respectively.

754 Cell 142, 749–761, September 3, 2010 ª2010 Elsevier Inc.

Page 123: Cell 100903

The A39R-PlexinC1 Interaction Globally Resemblesthe Sema7A-PlexinC1 InteractionWith strikingly similar orientations of the respective b propellers

in the two complexes, the A39R-PlexinC1 interface involves

the same set of structural elements as the Sema7A-PlexinC1

interface (Figure 4). The A39R-PlexinC1 interface buries a

total of 1890 A2 solvent-accessible surface area, slightly smaller

than the Sema7A-PlexinC1 interface. The protruding loop 4c-4d

of A39R is in an almost identical conformation as that of Sema7A,

inserting into the groove of the blade 3 surface of PlexinC1

(Figures 4A, 4B, and 5). The neighboring segments of this loop,

including the extrusion helix 2 of A39R that contacts the loop

3b-3c of PlexinC1, and the A39R blade 3 that contacts the Plex-

inC1 strand 3d, have undergone some relatively minor structural

accommodations at the periphery of the interface (Figure 4A).

These small movements result in remodeled pairwise interac-

tions by these surrounding structural elements, relative to the

Sema7A/PlexinC1 interface, but still preserving several key

contacts that presumably are important for the cross-reactivity

(discussed below).

The ‘‘loop-in-groove’’ interaction in A39R/PlexinC1 (Fig-

ure 4D) has six hydrogen bonds between the A39R 4c-4d loop

and the PlexinC1 groove, four less than in Sema7A/PlexinC1

(Table S3). Only one of these hydrogen bonds is conserved

between Sema7A/PlexinC1 and A39R/PlexinC1. The diversity

of amino acid contacts with PlexinC1 formed by the Sema7A

versus A39R 4c-4d loop indicates that the PlexinC1 pocket

has the capacity for highly degenerate interactions, which

Figure 4. The Interface between A39R and PlexinC1-SemaPSI

(A) Structural superposition of the A39R/PlexinC1-SemaPSI and Sema7A/PlexinC1-SemaPSI complexes indicating the closely related docking modes of the viral

and mammalian Semaphorins to the PlexinC1 receptor. The complexes have been aligned on the PlexinC1 component in order to visualize the respective overlap

of the two different Semaphorins. The A39R is in yellow, Sema7A in cyan.

(B) Close-up view of the interface with A39R depicted as ribbons and PlexinC1-SemaPSI in a surface representation.

(C) The interactions of residues near the obstructed end of the PlexinC1 groove, with A39R in chocolate and PlexinC1-SemaPSI in pink.

(D) The interaction between residues of the 4c-4d loop of A39R (blue) and the PlexinC1 groove (pink).

(E) The interaction between the residues of blade 3 of A39R (green) and strand 3d of PlexinC1-SemaPSI (pink).

See also Table S3 for a list of the interactions.

Cell 142, 749–761, September 3, 2010 ª2010 Elsevier Inc. 755

Page 124: Cell 100903

facilitates cross-reactivity. At the edge of this groove in PlexinC1,

A39R presents an arginine (Arg207), as opposed to a lysine

(Lys280) in Sema7A, to form a salt bridge with PlexinC1

Asp200 (discussed below).

Peripheral to the 4c-4d loop, the structural chemistry of the

A39R/PlexinC1 interactions at the obstructed (closed) end of

the PlexinC1 groove (Figures 4B and 4C) are quite different

than what is seen for Sema7A/PlexinC1 as a result of a slight

rigid-body repositioning (Figure 4A). Compared to Sema7A, the

A39R extrusion helix 2 is shifted outwards relative to the center

of the interface and tilted away from PlexinC1 (Figures 4A

and 4C). Consequently, the A39R/PlexinC1 interaction is not as

intimate as the Sema7A/PlexinC1 interaction at this region.

However, the shifting of the helix also allows A39R Asp300,

corresponding to Sema7A Gln379 (Figures 3C and 4C), to

move into the position to form a salt bridge with PlexinC1

Arg131, which is not present in Sema7A/PlexinC1.

The A39R/PlexinC1 interactions at the open end of the

PlexinC1 groove (Figures 4B and 4E) involve the same set of

PlexinC1 residues as in Sema7A/PlexinC1, but the pattern

of interaction is altered. While there are four salt bridges in

two clusters in Sema7A/PlexinC1, there are only two (A39R

Arg132/Arg134 to PlexinC1 Glu219) in A39R/PlexinC1 at this

region (Figure 4E). In A39R/Sema7A, the loss of the other cluster

of salt bridges seen in Sema7A/PlexinC1 (PlexinC1 Arg222/

Lys224 to Sema7A Asp216) is due to the raised A39R blade 3.

While PlexinC1 Lys224 is not directly bonded to A39R in the

A39R/PlexinC1 complex, PlexinC1 Arg222 manages to form

hydrogen bonds with the hydroxyl of A39R Tyr145. The loss

of these two salt bridges from the mammalian complex may

be further compensated by the strengthening of the A39R

Arg132-PlexinC1 Glu219 salt bridge, which is more deeply

buried than in Sema7A/PlexinC1, because of the neighboring

hydrophobic interaction between A39R Ile125 and PlexinC1

Leu220 (Figure 4E). Collectively, The A39R/PlexinC1 interface

shows variations from the Sema7A/PlexinC1 interface due to

some intermolecular repositioning around the central 4c-4d

loop. Nevertheless, the general scheme of inserting the long

Figure 5. The Conformation of the 4c-4d Loop of Semaphorins Is Central for Plexin Recognition

(A–D) Stick models of the isolated 4c-4d loops of Sema7A (cyan), A39R (yellow), Sema4D (green), and Sema3A (pink), showing its conserved conformation at the

bases (top) and the midpoints of the loops, and variable conformation at the tips of the loops (bottom).

(E) Sequence comparison of the 4c-4d loop of Semaphorins, with the key conserved structural determinants highlighted.

See also Figure S5 for the RGD motif at the base of the Sema7A 4c-4d loop.

756 Cell 142, 749–761, September 3, 2010 ª2010 Elsevier Inc.

Page 125: Cell 100903

4c-4d loop into the blade 3 groove in PlexinC1 is identical for

Sema7A and A39R.

The Similarities and Variations of the CentralRecognition Loops of SemaphorinsThe structural features of the 4c-4d interaction loop of Sema7A

and A39R are similar in all Semaphorin structures determined

to date (Figure 5). The conformation of this loop is stabilized

by an extensive intraloop hydrogen-bonding network, which

involves residues highly conserved in most Semaphorins (Fig-

ures 5A–5E). At the base of the loop, Asp269 (Sema7A

numbering), a buried aspartate conserved in all Semaphorins

(highlighted in Figures 5A–5E), forms a network of four hydrogen

bonds with main chain amides, which play a central role in orga-

nizing the loop conformation. At the midpoint, a serine (274 in

Sema7A and 201 in A39R) hydrogen bonds with two main-chain

amides on the opposing strand to narrow the loop. The kinking

of the loop is also facilitated by the presence of two consecutive

glycines (Gly271-Gly272) conserved in all Semaphorins (Fig-

ure 5E). At the tip of the loop, the Sema7A Ser274 main-chain

carbonyl forms hydrogen bonds with the main-chain amides

of Ser277 and Val278, a pattern reminiscent of reverse-turns,

which may help to maintain a rigid and protruding conformation.

Similar conformation is observed for the tip of this A39R loop

(Figures 5A and 5B).

In Sema3A and Sema4D, however, the tip of this loop has

a variation from Sema7A/A39R (Figures 5C and 5D). Class 3–5

Semaphorins, including Sema3A and Sema4D, have a one-

residue deletion at the Ser274 position. Sema4D and Sema3A

also lack the main-chain hydrogen bonds at the tip seen in

Sema7A and A39R. Consequently, the tips of the 4c-4d loops

of these Semaphorins can adopt different main chain conforma-

tions, probably reflecting the structural requirements of their

specific Semaphorin-Plexin interactions, yet all appear to

contain a polar residue at the loop tips for hydrogen bonding.

Because the residues that reside at the apex of the loop are

the most deeply embedded in the base of the Plexin groove,

as exemplified by Sema7A/PlexinC1, it is likely that these are

major determinants for Semaphorin-Plexin binding specificity

between classes. Indeed, there is extensive sequence diver-

gence between different classes of Semaphorins at these

positions, but few within the same class of Semaphorins

(Figure 5E).

The Mimicry of the Mammalian Semaphorin Sema7Aby the Viral Semaphorin A39RThe mammalian and viral Semaphorins have arrived at nearly

identical binding modes despite substantial variations in

sequence (31% identity between Sema7A and A39R). Visualiza-

tion of the PlexinC1 binding surfaces of Sema7A and A39R

(Figures 6A and 6B) indicates that there are several key corre-

sponding regions of structural mimicry that likely serve as the

foundation for their cross-reactivity with a common receptor.

There is obvious structural conservation of the centrally located

4c-4d loop and the presence of identical residues at the tips of

the loop (Ser-Leu) that engage the deepest portion of the

PlexinC1 groove. The preservation of the di-Glycine pair in

both Sema7A and A39R 4c-4d loops (Figure 5) ensures similar

loop conformations presented to PlexinC1. There is mimicry of

an array of interactions peripheral to the tip of the 4c-4d loop,

as is clear from comparison of the respective binding surfaces.

For example, the Lys280 in Sema7A that salt bridges to

Asp200 in PlexinC1 is mimicked by Arg207 in A39R that also

salt bridges to Asp200 on PlexinC1. Also found in both Sema-

phorins is a cluster of salt bridges on blade 3 that interact with

residues on strand d of PlexinC1. Here, Arg202 of Sema7A salt

bridges with Glu219 in PlexinC1, and this interaction is mimicked

by Arg134 of A39R salt bridging to PlexinC1 Glu219. Arg202

and Asp216 in Sema7A are spatially mimicked by Arg132 and

Asp148 in A39R, but the hydrogen bonding network with the

same PlexinC1 residues is remodeled. Key to this cluster is

a Tyrosine residue (Tyr213 in Sema7A, Tyr145 in A39R) (Fig-

ure 6F) that is in a nearly identical position in both complexes,

in the center of the cluster of salt bridges.

In order to ask whether the residues A39R uses to ‘‘mimic’’

Sema7A were energetically important for the viral Semaphorin

binding, we mutated several of them and tested binding by ITC

(Figure 6). Mutations of A39R Arg132 or Tyr145 (corresponding

to Sema7A Arg202 and Tyr213) to glutamate or serine abolished

PlexinC1 binding (Figures 6C and 6D), whereas mutation of A39R

Arg207 (corresponding to Sema7A Lys280) to glutamate

reduced PlexinC1 binding by >60-fold (Figure 6E). The A39R

Arg207 has a stronger charge and a larger head group than

a Lys residue, and its aliphatic stem is kinked, contacting the

neighboring hydrophobic moieties more intimately. Arg207 is a

more complementary fit in the A39R/PlexinC1 interface com-

pared to the Lys280 in the Sema7A-PlexinC1 interface, where

there is more solvent-occupied space adjacent to the side chain

(Figures S2A and S2B). Our data (Figure 6E) suggest that the

difference at this position could be an important contributing

factor to the higher-affinity binding to PlexinC1 achieved by

A39R, although additional interactions no doubt also contribute.

We also mutated PlexinC1 Arg222, which is hydrogen bonded to

A39R Tyr145 and serves as shared contact with both Semaphor-

ins. We found a 9-fold affinity loss in A39R-PlexinC1 binding

(Figure 6G). While Sema7A and A39R have exactly the same resi-

dues at the local region contacting PlexinC1, PlexinC1 Arg222

adopts very different rotamers in the two complexes (Figure 6F).

The 9-fold affinity loss upon PlexinC1 Arg222Ser mutation

suggests that this is an important determinant in A39R/PlexinC1

binding (Figure 6G). Therefore, the residues we mutated, which

are strikingly coincident in the viral and mammalian Sema

complexes, while certainly not giving us a comprehensive ener-

getic map of the A39R/PlexinC1 interface landscape, are indeed

energetically important in ligand-receptor binding.

Collectively, our structures suggest that A39R, built on a

smaller scaffold, has used a limited set of key, highly coincident

amino acid contacts to evolve a more efficient binding interac-

tion with PlexinC1 through both the enhancement of particularly

important polar interactions, such as the substitution of A39R

Arg207 for the Sema7A Lys280, and slight adjustments of side

rotamer and main chain positions throughout the binding

surface. These adjustments sum to a substantial optimiza-

tion of binding energetics and a concomitant gain in binding

affinity.

Cell 142, 749–761, September 3, 2010 ª2010 Elsevier Inc. 757

Page 126: Cell 100903

Figure 6. Viral Mimicry of Sema7A by A39R

(A and B) The respective PlexinC1-binding surfaces of Sema7A (cyan, left) and A39R (yellow, right). The residues mapped at the interface are colored red for acidic

residues, blue for basic residues, gray for polar, noncharged residues, and green for apolar residues.

(C–E) Mutations of the A39R residues mapped to the Sema7A-PlexinC1 interface abolished or dramatically reduced A39R-PlexinC1 binding.

(F) The different conformation of PlexinC1 Arg222 in binding A39R and Sema7A.

(G) Calorimetric measurement of the binding between A39R and the Arg222Ser mutant of PlexinC1-SemaPSI.

See also Figure S3 for a structural and sequence comparison of Sema7A and A39R.

758 Cell 142, 749–761, September 3, 2010 ª2010 Elsevier Inc.

Page 127: Cell 100903

DISCUSSION

We present here the structures of two Semaphorin-Plexin

complexes, establishing the general recognition paradigm

between two protein families with important and wide-ranging

physiological functions involving deliverance of repulsive guid-

ance cues. Unexpectedly, the two b propellers of the Sema-

phorin and Plexin Sema domains contact each other edge-on.

In comparison, MET uses the bottom face of the propeller for

HGF-b chain interaction (Stamos et al., 2004), whereas integrins

bind their ligands with the top face of the propeller (Xiao et al.,

2004; Xiong et al., 2002). Thus, as for other recognition modules

in biology, such as leucine-rich repeats, immunoglobulin folds,

and even MHC-folds, the b propeller reveals itself to be capable

of a diverse array of ligand interaction modes. The moderate-to-

low sequence identity between different classes of Semaphorins

and Plexins, especially at the binding segments, suggests that

there is variability in the specific interatomic contacts in Sema-

phorin/Plexin interfaces between families that is buttressed by

a framework of conserved interactions. This variability may be

necessary for coding the complex pattern of specificity in the

growth- or motility-guiding activities. Nevertheless, the orienta-

tion of each component in the complex, the Sema-Sema dock-

ing geometry, and the global usage of structural elements should

be generally conserved for Semaphorins and Plexins, as has

been found in other ligand-receptor families with similar extents

of sequence identity.

The poxvirus-derived A39R has evolved a more efficient

PlexinC1 binding surface, yet it is smaller than that of Sema7A,

with fewer overall contacts. Our structural analysis showed

that the central binding elements of Sema7A are rigorously main-

tained in the viral complex; therefore, how does A39R achieve

higher affinity binding? This is very difficult to clearly answer,

as the sources of free energy differences between interacting

proteins are not obviously identified by inspection of structures

and are contributed by many subtle effects such as intramolec-

ular side chain cooperativity, solvent restructuring at the binding

surface, and enthalpy-entropy compensation. Thus, we suspect

that it is due to a collective reduction of energetic penalties

across the entire Semaphorin/Plexin interface, perhaps in

combination with an optimization of protein dynamics, such as

lowering overall temperature factors (Table S1), to gain favor in

entropic change. The high similarity of Sema7A and A39R

suggests the likelihood that the poxviruses have acquired the

A39R gene by hijacking and mutating the Sema7A gene, rather

than through convergent evolution where the virus would have

evolved a structurally distinct scaffold for PlexinC1 binding

(Gewurz et al., 2001). Thus, this does not seem to be a case of

molecular ‘‘mimicry’’ as much as simply modifying an existing

binding scaffold with subtle improvements. Given the lack of

involvement of the PSI and Ig domains in Plexin interactions,

these domains may have been lost in A39R over the course of

evolution.

The dimerized Sema7A/PlexinC1 and A39R/PlexinC1 com-

plexes support a model of PlexinC1 activation via dimerization

akin to many other cell-surface receptors. This, in itself, is an

important advance since the downstream mechanisms of Sem-

aphorin/Plexin repulsive signaling are not well understood, and

the knowledge that Plexins are active as dimers should help

refine our thinking about possible downstream effectors. The

quiescent PlexinC1 possibly exists as preformed oligomers,

given the evidence that some other Plexins, e.g., plexin B, exist

as stable homodimers in the absence of ligands (Tong et al.,

2007). We have also found that unliganded PlexinC1 ECD exists

as a mixture of oligomeric states, as assessed by gel filtration

chromatography and crosslinking, although we have not quanti-

tated the exact stochiometries of these states. We propose that

constitutively dimeric ligands such as Sema7A and A39R orient

the inactive PlexinC1 into a dimerized state, in which the

distance, geometry and conformation enables the activation of

the intracellular GAP activity. This PlexinC1 activation scheme

is simpler than a previously proposed model for Sema3A-

induced PlexinA1 activation. That model proposed that the

Plexin Sema domain associates intramolecularly with its addi-

tional extracellular domains in an autoinhibited conformation,

analogous to the LDL receptor (Rudenko et al., 2002), and that

Sema3A disrupts this intramolecular association (Antipenko

et al., 2003; Takahashi and Strittmatter, 2001). Although we

have no unequivocal data to support or refute such a model,

our ITC experiments of A39R binding to full-length versus trun-

cated PlexinC1 ECD show identical binding energetics and

monophasic association, indirectly implying that PlexinC1 is in

the same conformation in the full-length ECD and truncated

forms (Figures 1D and 1E). An important caveat to our experi-

ments is that thermodynamic measurements would not detect

a potential biphasic binding kinetics indicative of a two-step

binding event involving a Plexin conformational change. Never-

theless, whether Plexin undergoes a conformational change

prior to Semaphorin binding, clearly a Semaphorin-induced

arrangement of Plexin into a specifically orientated dimer is the

means of receptor activation. Future structural and mechanistic

studies can now address whether the preactivated Plexin is

involved in an intramolecular association that contributes to

the maintenance of an autoinhibited state.

The Sema7A/PlexinC1 and A39R/PlexinC1 complexes evoke

several important questions. First, where is the Neuropilin-

docking site in a class 3 Semaphorin-Plexin complex? Although

predictions could be made on the basis of our structures

and the previous mutagenesis mapping data (Antipenko et al.,

2003), accurate insights may require a complex structure

encompassing Neuropilin. Second, does Sema7A bind integrins

directly? It has been suggested Sema7A promotes axon growth

and initiates T cell-mediated inflammation responses through

integrins, but not PlexinC1, using an RGD-dependent mechanism

(Pasterkamp et al., 2003; Suzuki et al., 2007). The RGD motif of

Sema7A (Arg267-Gly268-Asp269) is buried (Figure S5) and is

unlikely to be recognized by integrins. Finally, the complex struc-

tures can now be used as a basis to interrogate Sema/Plexin

interactions in vivo with high precision. While mutation of the

key determinants of both Sema7A-PlexinC1 and A39R-PlexinC1

interfaces clearly compromise binding (Figure 6, Figure S4), fur-

ther work will be necessary to demonstrate the consequences

of such perturbation in a cellular or organismal context. Such

functional validation should determine whether modulation of

Semaphorin-Plexin recognition holds clinical promise for applica-

tions involving cell guidance, e.g., directional nerve regeneration.

Cell 142, 749–761, September 3, 2010 ª2010 Elsevier Inc. 759

Page 128: Cell 100903

EXPERIMENTAL PROCEDURES

Cloning, Cell Culture, and Baculovirus Generation

Human Sema7A, PlexinC1-SemaPSI, and PlexinC1-ECD, were expressed in

the BacMam system as described previously (Dukkipati et al., 2008), with

a 7-His tag directly attached to the C terminus of each expression construct.

Full-length Vaccinia virus A39R was expressed in both the BacMam system

and the baculovirus-insect cell system. Sf9 and hi5 cells were maintained in

the HyQ-SFX (HyClone) and InsectXpress (Lonza) media. HEK293H cells

(Invitrogen) and GnTI- HEK293 cells (Reeves et al., 2002) were maintained in

CDM4HEK293 media (HyClone). For generation of recombinant viruses, the

expression constructs and the BacVector-3000 baculovirus DNA (EMD

Chemicals) were used to cotransfect sf9 cells in 6-well plates. After 5 days,

the resulted low titer virus stock was harvested and was used to infect Sf9 cells

at 2 3 106 cell/ml for amplification.

Protein Preparation

The amplified baculoviruses were used to infect 1-6 liters of HEK293 cells

or hi5 cells at a density of 1.0–1.5 3 106 cells/ml. After 72 hr, the cells were

pelleted and the supernatants were concentrated. The recombinant proteins

in the supernatant were captured by Ni-NTA metal affinity resin (QIAGEN)

and eluted with 300 mM imidazole (pH 7.5). The eluted proteins were further

purified with gel filtration.

Glycan Trimming, Crystallization, and Data Collection

and Processing

For the Sema7A-PlexinC1 complex, GnTI� HEK293-expressed proteins were

trimmed with carboxypeptidase-A and Endo-H. The digested products were

further purified with size exclusion columns and concentrated to 10 mg/ml.

Crystallization was performed via the sitting-drop vapor-diffusion method.

The Sema7A and PlexinC1-SemaPSI proteins, at 1:1 molar ratio, were mixed

with an equal volume of reservoir solution and equilibrated against the reser-

voir solution containing 12% PEG1000, 0.1 M Tris (pH 7.5), and 0.2 M Li2SO4.

The A39R used for crystallization was expressed from hi5 cells using the

pAcGP67A vector. A39R was His-tagged at the N-terminus and expressed

in the presence of 10 mM kifunensine in the expression media (InsectXpress).

The His-tag was removed from A39R with TEV protease, the glycans were

trimmed with Endo-H, and the protein was further purified by gel filtration.

A39R crystallized in the following condition: 50 mM sodium citrate (pH 5.0),

0.2 M Li2SO4, 15% PEG3350, and 20% glycerol.

For the A39R-PlexinC1-SemaPSI complex, the A39R protein was expressed

from the BacMam system following the same procedure for PlexinC1-Sema-

PSI. The complex was crystallized in the following condition: 6% PEG8000,

0.1 M CaCl2, 0.1 M HEPES (pH 7.0), and 0.2 M NaCl.

The Sema7A/PlexinC1 and A39R/PlexinC1 data were measured at the

Life Sciences Collaborative Access Team (LS-CAT) beamlines 21-ID-D and

21-ID-G at the Advanced Photon Source (APS) and were processed with

HKL2000 (Otwinowski and Minor, 1997) (Table S1). The A39R data were

measured at both the Stanford Synchrotron Research Lab (SSRL) and the

Advanced Light Source (ALS) and were processed with Mosflm and scaled

with Scala of the CCP4 program suite (Collaborative Computational Project,

Number 4, 1994). The statistics on the data are presented in Table S1.

Structure Determination and Model Refinement

For structure determinations of Sema7A/PlexinC1-SemaPSI and free A39R,

experimental phases were calculated using heavy atom derivatives by the

SIRAS method, as implemented in the program SOLVE (Terwilliger and

Berendzen, 1999) and SHARP (Bricogne et al., 2003). For structure determina-

tion of the A39R/PlexinC1-SemaPSI complex, the components from the above

two structures were used as search models in molecular replacement. The

models were refined with CNS (Brunger et al., 1998). The final models include

seven N-linked glycans per PlexinC1-SemaPSI molecule (Asn86, Asn141,

Asn149, Asn241, Asn252, Asn386, and Asn407), four per Sema7A (Asn105,

Asn157, Asn258, and Asn330), and one per A39R (Asn51). Each glycan

was modeled with a single GlcNAc residue, as expected, due to the Endo-H

treatment of the proteins. A summary of the refinement statistics is given in

Table S1.

Isothermal Titration Calorimetry

The proteins used for calorimetry were all expressed from HEK293H cells. ITC

was carried out on a VP-ITC calorimeter (MicroCal, Northhampton, MA) at

30�C. All samples were thoroughly degassed before titration. The titration

data were processed with MicroCal Origin software, Version 5.0.

ACCESSION NUMBERS

The structure factors and coordinates for the Sema7A/PlexinC1-SemaPSI

complex, free A39R and the A39R/PlexinC1-SemaPSI complex have been

deposited in the Protein Data Bank (accession codes 3NVQ, 3NVX, and

3NVN).

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures, five

figures, and three tables and can be found with this article online at doi:

10.1016/j.cell.2010.07.040.

ACKNOWLEDGMENTS

We thank the staff at the LS-CAT beamlines at APS and the staffs of ALS and

SSRL for the support in X-ray data collection. We also thank L. Colf, and

N. Goriatcheva for help with the expression of A39R. X.H. is supported by

the NIH grant 1R01GM07805, and K.C.G. is an Investigator of the Howard

Hughes Medical Institute and is supported by NIH-R01-AI51321. The Struc-

tural Biology Facility is supported by the R.H. Lurie Comprehensive Cancer

Center of Northwestern University.

Received: March 19, 2010

Revised: June 8, 2010

Accepted: July 20, 2010

Published online: August 19, 2010

REFERENCES

Antipenko, A., Himanen, J.P., van Leyen, K., Nardi-Dei, V., Lesniak, J., Barton,

W.A., Rajashankar, K.R., Lu, M., Hoemme, C., Puschel, A.W., and Nikolov,

D.B. (2003). Structure of the semaphorin-3A receptor binding module. Neuron

39, 589–598.

Bork, P., Doerks, T., Springer, T.A., and Snel, B. (1999). Domains in plexins:

links to integrins and transcription factors. Trends Biochem. Sci. 24, 261–263.

Bricogne, G., Vonrhein, C., Flensburg, C., Schiltz, M., and Paciorek, W. (2003).

Generation, representation and flow of phase information in structure determi-

nation: recent developments in and around SHARP 2.0. Acta Crystallogr. D

Biol. Crystallogr. 59, 2023–2030.

Brunger, A.T., Adams, P.D., Clore, G.M., DeLano, W.L., Gros, P.,

Grosse-Kunstleve, R.W., Jiang, J.S., Kuszewski, J., Nilges, M., Pannu, N.S.,

et al. (1998). Crystallography & NMR system: a new software suite for macro-

molecular structure determination. Acta Crystallogr. D Biol. Crystallogr. 54,

905–921.

Collaborative Computational Project, Number 4. (1994). The CCP4 suite:

programs for protein crystallography. Acta Crystallogr. D Biol. Crystallogr.

50, 760–763.

Comeau, M.R., Johnson, R., DuBose, R.F., Petersen, M., Gearing, P.,

VandenBos, T., Park, L., Farrah, T., Buller, R.M., Cohen, J.I., et al. (1998). A

poxvirus-encoded semaphorin induces cytokine production from monocytes

and binds to a novel cellular semaphorin receptor, VESPR. Immunity 8,

473–482.

Dukkipati, A., Park, H.H., Waghray, D., Fischer, S., and Garcia, K.C. (2008).

BacMam system for high-level expression of recombinant soluble and

membrane glycoproteins for structural studies. Protein Expr. Purif. 62,

160–170.

760 Cell 142, 749–761, September 3, 2010 ª2010 Elsevier Inc.

Page 129: Cell 100903

Gewurz, B.E., Gaudet, R., Tortorella, D., Wang, E.W., and Ploegh, H.L. (2001).

Virus subversion of immunity: a structural perspective. Curr. Opin. Immunol.

13, 442–450.

He, Z., Wang, K.C., Koprivica, V., Ming, G., and Song, H.J. (2002). Knowing

how to navigate: mechanisms of semaphorin signaling in the nervous system.

Sci. STKE 2002, re1.

He, X.L., Dukkipati, A., and Garcia, K.C. (2006). Structural determinants of

natriuretic peptide receptor specificity and degeneracy. J. Mol. Biol. 361,

698–714.

He, H., Yang, T., Terman, J.R., and Zhang, X. (2009). Crystal structure of the

plexin A3 intracellular region reveals an autoinhibited conformation through

active site sequestration. Proc. Natl. Acad. Sci. USA 106, 15610–15615.

Holmes, S., Downs, A.M., Fosberry, A., Hayes, P.D., Michalovich, D.,

Murdoch, P., Moores, K., Fox, J., Deen, K., Pettman, G., et al. (2002). Sema7A

is a potent monocyte stimulator. Scand. J. Immunol. 56, 270–275.

Huppa, J.B., Axmann, M., Mortelmaier, M.A., Lillemeier, B.F., Newell, E.W.,

Brameshuber, M., Klein, L.O., Schutz, G.J., and Davis, M.M. (2010). TCR-

peptide-MHC interactions in situ show accelerated kinetics and increased

affinity. Nature 463, 963–967.

Koller, K.J., and Goeddel, D.V. (1992). Molecular biology of the natriuretic

peptides and their receptors. Circulation 86, 1081–1088.

Kolodkin, A.L., Matthes, D.J., and Goodman, C.S. (1993). The semaphorin

genes encode a family of transmembrane and secreted growth cone guidance

molecules. Cell 75, 1389–1399.

Koppel, A.M., Feiner, L., Kobayashi, H., and Raper, J.A. (1997). A 70 amino

acid region within the semaphorin domain activates specific cellular response

of semaphorin family members. Neuron 19, 531–537.

Krogsgaard, M., Prado, N., Adams, E.J., He, X.L., Chow, D.C., Wilson, D.B.,

Garcia, K.C., and Davis, M.M. (2003). Evidence that structural rearrangements

and/or flexibility during TCR binding can contribute to T cell activation. Mol.

Cell 12, 1367–1378.

Kruger, R.P., Aurandt, J., and Guan, K.L. (2005). Semaphorins command cells

to move. Nat. Rev. Mol. Cell Biol. 6, 789–800.

Lemmon, M.A., Pinchasi, D., Zhou, M., Lax, I., and Schlessinger, J. (1997). Kit

receptor dimerization is driven by bivalent binding of stem cell factor. J. Biol.

Chem. 272, 6311–6317.

Lev, S., Yarden, Y., and Givol, D. (1992). Dimerization and activation of the kit

receptor by monovalent and bivalent binding of the stem cell factor. J. Biol.

Chem. 267, 15970–15977.

Love, C.A., Harlos, K., Mavaddat, N., Davis, S.J., Stuart, D.I., Jones, E.Y., and

Esnouf, R.M. (2003). The ligand-binding face of the semaphorins revealed

by the high-resolution crystal structure of SEMA4D. Nat. Struct. Biol. 10,

843–848.

Luo, Y., Raible, D., and Raper, J.A. (1993). Collapsin: a protein in brain that

induces the collapse and paralysis of neuronal growth cones. Cell 75,

217–227.

Oinuma, I., Ishikawa, Y., Katoh, H., and Negishi, M. (2004). The Semaphorin 4D

receptor Plexin-B1 is a GTPase activating protein for R-Ras. Science 305,

862–865.

Otwinowski, Z., and Minor, W. (1997). Processing of X-ray diffraction data

collected in oscillation mode. Methods Enzymol. 276, 307–326.

Pasterkamp, R.J., Peschon, J.J., Spriggs, M.K., and Kolodkin, A.L. (2003).

Semaphorin 7A promotes axon outgrowth through integrins and MAPKs.

Nature 424, 398–405.

Pellet-Many, C., Frankel, P., Jia, H., and Zachary, I. (2008). Neuropilins: struc-

ture, function and role in disease. Biochem. J. 411, 211–226.

Reeves, P.J., Callewaert, N., Contreras, R., and Khorana, H.G. (2002).

Structure and function in rhodopsin: high-level expression of rhodopsin with

restricted and homogeneous N-glycosylation by a tetracycline-inducible

N-acetylglucosaminyltransferase I-negative HEK293S stable mammalian cell

line. Proc. Natl. Acad. Sci. USA 99, 13419–13424.

Rudenko, G., Henry, L., Henderson, K., Ichtchenko, K., Brown, M.S.,

Goldstein, J.L., and Deisenhofer, J. (2002). Structure of the LDL receptor extra-

cellular domain at endosomal pH. Science 298, 2353–2358.

Scott, G.A., McClelland, L.A., Fricke, A.F., and Fender, A. (2009). Plexin C1,

a receptor for semaphorin 7a, inactivates cofilin and is a potential tumor

suppressor for melanoma progression. J. Invest. Dermatol. 129, 954–963.

Semaphorin Nomenclature Committee. (1999). Unified nomenclature for the

semaphorins/collapsins. Cell 97, 551–552.

Stamos, J., Lazarus, R.A., Yao, X., Kirchhofer, D., and Wiesmann, C. (2004).

Crystal structure of the HGF beta-chain in complex with the Sema domain of

the Met receptor. EMBO J. 23, 2325–2335.

Suzuki, K., Okuno, T., Yamamoto, M., Pasterkamp, R.J., Takegahara, N.,

Takamatsu, H., Kitao, T., Takagi, J., Rennert, P.D., Kolodkin, A.L., et al.

(2007). Semaphorin 7A initiates T-cell-mediated inflammatory responses

through alpha1beta1 integrin. Nature 446, 680–684.

Suzuki, K., Kumanogoh, A., and Kikutani, H. (2008). Semaphorins and their

receptors in immune cell interactions. Nat. Immunol. 9, 17–23.

Takahashi, T., and Strittmatter, S.M. (2001). Plexina1 autoinhibition by the

plexin sema domain. Neuron 29, 429–439.

Tamagnone, L., and Comoglio, P.M. (2000). Signalling by semaphorin recep-

tors: cell guidance and beyond. Trends Cell Biol. 10, 377–383.

Tamagnone, L., Artigiani, S., Chen, H., He, Z., Ming, G.I., Song, H., Chedotal,

A., Winberg, M.L., Goodman, C.S., Poo, M., et al. (1999). Plexins are a large

family of receptors for transmembrane, secreted, and GPI-anchored sema-

phorins in vertebrates. Cell 99, 71–80.

Terwilliger, T.C., and Berendzen, J. (1999). Automated MAD and MIR structure

solution. Acta Crystallogr. D Biol. Crystallogr. 55, 849–861.

Tessier-Lavigne, M., and Goodman, C.S. (1996). The molecular biology of

axon guidance. Science 274, 1123–1133.

Tong, Y., Chugha, P., Hota, P.K., Alviani, R.S., Li, M., Tempel, W., Shen, L.,

Park, H.W., and Buck, M. (2007). Binding of Rac1, Rnd1, and RhoD to a novel

Rho GTPase interaction motif destabilizes dimerization of the plexin-B1

effector domain. J. Biol. Chem. 282, 37215–37224.

Tong, Y., Hota, P.K., Penachioni, J.Y., Hamaneh, M.B., Kim, S., Alviani, R.S.,

Shen, L., He, H., Tempel, W., Tamagnone, L., et al. (2009). Structure and

function of the intracellular region of the plexin-b1 transmembrane receptor.

J. Biol. Chem. 284, 35962–35972.

Uniewicz, K.A., and Fernig, D.G. (2008). Neuropilins: a versatile partner of

extracellular molecules that regulate development and disease. Front. Biosci.

13, 4339–4360.

Waimey, K.E., and Cheng, H.J. (2006). Axon pruning and synaptic develop-

ment: how are they per-plexin? Neuroscientist 12, 398–409.

Xiao, T., Takagi, J., Coller, B.S., Wang, J.H., and Springer, T.A. (2004).

Structural basis for allostery in integrins and binding to fibrinogen-mimetic

therapeutics. Nature 432, 59–67.

Xiong, J.P., Stehle, T., Zhang, R., Joachimiak, A., Frech, M., Goodman, S.L.,

and Arnaout, M.A. (2002). Crystal structure of the extracellular segment of

integrin alpha Vbeta3 in complex with an Arg-Gly-Asp ligand. Science 296,

151–155.

Xu, X., Ng, S., Wu, Z.L., Nguyen, D., Homburger, S., Seidel-Dugan, C., Ebens,

A., and Luo, Y. (1998). Human semaphorin K1 is glycosylphosphatidylinositol-

linked and defines a new subfamily of viral-related semaphorins. J. Biol. Chem.

273, 22428–22434.

Yamada, A., Kubo, K., Takeshita, T., Harashima, N., Kawano, K., Mine, T.,

Sagawa, K., Sugamura, K., and Itoh, K. (1999). Molecular cloning of a

glycosylphosphatidylinositol-anchored molecule CDw108. J. Immunol. 162,

4094–4100.

Cell 142, 749–761, September 3, 2010 ª2010 Elsevier Inc. 761

Page 130: Cell 100903

Photoadaptation in Neurospora byCompetitive Interaction of Activatingand Inhibitory LOV DomainsErik Malzahn,1,3 Stilianos Ciprianidis,1,3 Krisztina Kaldi,2,3 Tobias Schafmeier,1 and Michael Brunner1,*1University of Heidelberg Biochemistry Center, Im Neuenheimer Feld 328, 69120 Heidelberg, Germany2Department of Physiology, Semmelweis University, POB 259 H-1444 Budapest, Hungary3These authors contributed equally to this work

*Correspondence: [email protected] 10.1016/j.cell.2010.08.010

SUMMARY

Light responses and photoadaptation of Neurosporadepend on the photosensory light-oxygen-voltage(LOV) domains of the circadian transcription factorWhite Collar Complex (WCC) and its negative regu-lator VIVID (VVD). We found that light triggers LOV-mediated dimerization of the WCC. The activatedWCC induces expression of VVD, which then dis-rupts and inactivates the WCC homodimers by thecompetitive formation of WCC-VVD heterodimers,leading to photoadaptation. During the day, expres-sion levels of VVD correlate with light intensity, allow-ing photoadaptation over several orders of magni-tude. At night, previously synthesized VVD servesas a molecular memory of the brightness of thepreceding day and suppresses responses to lightcues of lower intensity. We show that VVD is essen-tial to discriminate between day and night, even innaturally ambiguous photoperiods with moonlight.

INTRODUCTION

Molecular clocks are circadian timing systems that control

rhythmic transcription of a large number of genes (Dibner

et al., 2010; McClung, 2008; McDonald and Rosbash, 2001;

Siepka et al., 2007; Vitalini et al., 2006; Woelfle and Johnson,

2006). To ensure robust undisturbed circadian oscillation,

clock controlled transcription is in principal independent of light

and compensated for temperature and other environmental

stimuli. However, exogenous stimuli, such as light, are needed

as resetting cues to synchronize the clock with the environ-

mental cycles (Doyle and Menaker, 2007; Foster et al., 2007;

Helfrich-Forster, 2002; Merrow and Roenneberg, 2007; Nayak

et al., 2007). Hence, circadian systems must be able to sense

and properly interpret light intensity signals under a variety of

natural conditions.

Molecular mechanisms that allow circadian systems to

respond reliably to natural photoperiods with noisy and ambig-

uous light intensity characteristics are not known. In nature,

photoperiodic entrainment is further complicated by light

reflected from the moon since the photoreceptors generally

detect even low light levels (Bachleitner et al., 2007; Merrow

et al., 1999).

The response of Neurospora to light is mediated by WHITE

COLLAR-1 (WC-1) and VIVID (VVD). Together with WC-2,

WC-1 forms the circadian transcription factor White Collar

Complex (WCC), and VVD is expressed under control of light-

activated WCC. WC-1 and VVD contain a flavin-binding light-

oxygen-voltage (LOV) domain (He et al., 2002; Schwerdtfeger

and Linden, 2003). Flavin-binding LOV domains are blue-light

sensors in light receptors of plants and fungi (Krauss et al.,

2009). Transduction of the light signal depends on formation of

a covalent photoadduct between a conserved cysteinyl residue

and the bound flavin cofactor (Salomon et al., 2000; Swartz et al.,

2001; Zoltowski et al., 2007, 2009). The photoadduct interferes

with the interaction of the LOV core and an adjacent helix (Corch-

noy et al., 2003; Harper et al., 2003), but it is not clear how the

signal is further transduced to downstream effectors and tar-

gets. In particular, the molecular basis of light activation of

WCC and adaptation by VVD is not known. VVD has a propensity

to form rapidly exchanging dimers in response to light (Zoltowski

and Crane, 2008), but the physiological relevance of the dynamic

dimerization is obscure.

Once activated by light, LOV domains cannot sense additional

light cues. Spontaneous decay of the flavin-cysteinyl photoad-

duct resets the LOV domain to the dark state and restores its

light-sensing capacity. The photoadduct of VVD is extremely

stable (t1/2 �4.5 hr), but generally the kinetics of photoadduct

decay varies from seconds to hours in different LOV domains

(Corchnoy et al., 2003; Swartz et al., 2001; Zoltowski et al.,

2007, 2009).

WCC directly controls expression of a large number of genes

(Smith et al., 2010). The products of two of these genes,

frequency (frq) and vvd, inhibit WCC in negative feedback loops

(Brunner and Schafmeier, 2006; Loros et al., 2007). The clock

protein FRQ inhibits WCC in a light-independent fashion and

generates a daily rhythm of WCC activity that is essential for

self-sustained circadian rhythms (Schafmeier et al., 2008; Schaf-

meier et al., 2005). In contrast, VVD inhibits WCC in a light-

dependent fashion, and the oscillating feedback loop dampens

rapidly in constant darkness (Elvin et al., 2005; Heintzen et al.,

762 Cell 142, 762–772, September 3, 2010 ª2010 Elsevier Inc.

Page 131: Cell 100903

2001). VVD is not essential for clock function. Deletion of vvd has

a mild circadian phenotype, resulting in a slight phase delay of

the circadian clock (Heintzen et al., 2001). This raises a question

about the physiologically relevant function of VVD.

When dark-grown Neurospora is abruptly exposed to light,

WCC-dependent transcript levels increase rapidly and adapt to

a low steady state within 1–2 hr (Schwerdtfeger and Linden,

2001). In the photoadapted state, a second transcription spike

can be evoked by increasing the light intensity (Schwerdtfeger

and Linden, 2001). Photoadaptation and light-sensing in the

adapted state depend on VVD.

In nature, mycelia growing deeply concealed behind the

bark of a tree may receive the same light intensity level during

the day as openly exposed mycelia receive by moonlight in the

course of a clear night. How the photoreceptor system adapts

to different light intensities and facilitates photoperiodic entrain-

ment over a broad range of conditions (Tan et al., 2004) is not

known.

In the present study, we elucidated the molecular mechanism

of photoactivation and adaptation of WCC. We show that WCC

and VVD form highly dynamic homo- and heterodimers via their

activated LOV domains. In response to light, WCC initially forms

homodimers that efficiently activate transcription of vvd. VVD

accumulates with a delay after light induction and acts as

a competitive inhibitor of WCC homodimerization, leading to

an attenuation of light-induced transcription. Interaction with

VVD protects light-activated WCC from rapid degradation and

thus allows a sizable fraction of WCC to equilibrate with the

WCC dark form. In the photoadapted state, VVD synthesis trig-

gered by light-activated WCC is balanced by VVD-dependent

inhibition of WCC. The VVD-mediated desensitization and stabi-

lization of the light activated WCC explain on a molecular level

how Neurospora can robustly entrain to artificial and natural

photoperiods.

RESULTS

Neurospora Responds to Light over Several Ordersof MagnitudeTo characterize the light-sensing system of Neurospora,

we exposed dark-grown mycelia to increasing light intensities

(5 mmol 3 m�2 3 s�1 and 130 mmol 3 m�2 3 s�1) for two consec-

utive 4 hr periods, and expression of vvd RNA was analyzed by

qPCR (Figure 1A). In dark-grown mycelia, vvd RNA could not

Figure 1. Neurospora Responds to Light Intensities over Several Orders of Magnitude

(A) Photoadaptation and reactivation of vvd transcription. Dark-grown Neurospora was exposed for 4 hr to a light intensity of 5 mmol (photons) 3 m�2 3 s�1 and

then transferred to 130 mmol 3 m�2 3 s�1 (red line). vvd RNA was measured by qRT-PCR and normalized to actin RNA (black line).

(B) In the photoadapted state, levels of vvd RNA correlate with light intensity. Neurospora was grown for 48 hr in constant light of the indicated intensity, and vvd

RNA was quantified. Average of six RNA measurements of two independent experiments is shown; error bars indicate the standard deviation (SD).

(C) VVD protein levels depend on light intensity. Mycelia were grown for 48 hr at the indicated light conditions and VVD protein levels were measured. A repre-

sentative western blot is shown (n = 3).

(D) Entrainment of the Neurospora clock during cycles of low light intensities. Race tubes were inoculated with WT conidia in light and subsequently entrained to

13/13 hr light-dark (LD) cycles with the indicated light intensities during the light phase. Conidial bands were analyzed densitometrically. Light and dark periods

are indicated by white and black boxes, respectively. Conidial densities of consecutive cycles (26 hr) are plotted below each other. Plots show representative

individual race tubes (n R 9).

(E) Subcellular localization of VVD-GFP. Mycelia of Dvvd strains expressing GFP or VVD-GFP under control of the cpc-1 promoter were analyzed by fluorescence

microscopy. DAPI staining reveals the nuclei.

See also Figure S1.

Cell 142, 762–772, September 3, 2010 ª2010 Elsevier Inc. 763

Page 132: Cell 100903

be detected. After the mycelia were transferred to the light

(5 mmol 3 m�2 3 s�1), RNA levels increased rapidly, reached

a peak after �20 min, and subsequently decreased to a lower

steady-state level. When the light intensity was increased after

4 hr to 130 mmol 3 m�2 3 s�1, a second spike of vvd RNA induc-

tion and adaptation was observed. The data confirm that

Neurospora adapts to constant light but can still sense light in

the adapted state (Schwerdtfeger and Linden, 2001). However,

we noted that levels of vvd RNA in the adapted state were higher

at the higher light intensity.

To determine whether steady-state levels of vvd RNA correlate

with the light intensity, we grew mycelia for 48 hr at a number of

different light intensities, and vvd RNA was quantified. Surpris-

ingly, vvd RNA levels correlated with light intensity over more

than two orders of magnitude (Figure 1B). Above �130 mmol 3

m�2 3 s�1, the light-dependent increase in transcript levels

was less pronounced than at lower fluence levels, suggesting

that the system approached saturation. VVD protein levels

increased in a similar fashion (Figure 1C). Transcript levels of

frq, al-2, and wc-1, which are also controlled by the WCC initially,

also rise with increasing light intensity. However, light induction

of these genes saturates readily and therefore does not reflect

light intensities over a wide range (Figure S1A available online).

At light intensities below 0.8 mmol 3 m�2 3 s�1, vvd RNA and

VVD protein levels were hardly above background and hence

difficult to quantify. To address whether the Neurospora photo-

receptor system can sense light of lower intensity, we examined

entrainment to light cues of different intensities in the race-tube

assay (Figure 1D). Entrainment of the conidiation rhythm of

Neurospora to light:dark (LD) cycles depends strictly on func-

tional WCC (Liu, 2003). Neurospora entrained to LD cycles with

light intensities as low as 2.8 nmol 3 m�2 3 s�1, the lowest level

of light tested. Assuming that the molecular mechanism of light

sensing and adaptation is the same at high and low light, the

data suggest that the photoreceptors of Neurospora can

discriminate light intensities over five orders of magnitude.

(2.8 nmol 3 m�2 3 s�1 to 750 mmol 3 m�2 3 s�1).

Subcellular Localization of VVDThe molecular mechanism by which VVD attenuates the activity

of the WCC is not known, and interaction partners of VVD have

not been found. VVD might directly inhibit WCC in the nucleus

or indirectly affect its activity via signaling molecules, such as

kinases or phosphatases. To determine whether a nuclear

pool of VVD could directly interfere with transcription activation

by the WCC, we constructed a strain, Dvvd cpc-1-vvd-gfp,

which constitutively expressed GFP-tagged VVD under control

of the cpc-1 promoter. Expression of VVD-GFP suppressed

the intense coloring of Dvvd mycelia and restored the wild-

type (WT) phase of conidiation (not shown) suggesting that

VVD-GFP was functional. Analysis of living mycelia by fluores-

cence microscopy revealed that VVD-GFP was located in

the cytosol and nuclei but displayed slightly greater nuclear

enrichment than GFP alone, which was uniformly distributed

(Figure 1E).

The observed nuclear localization of VVD-GFP appears to

contradict previous reports that VVD was not detected in the

nuclear fraction when mycelia were fractionated (Schwerdtfeger

and Linden, 2003). However, VVD is a small protein (21 kDa) and

might freely equilibrate between nuclei and cytosol. Hence, it

might readily be lost from nuclei during fractionation. To address

this issue, we fractionated mycelia that were untreated or treated

with 1% formaldehyde to induce low levels of nonspecific cross-

linking (Figure S1B). In untreated mycelia, VVD was found exclu-

sively in the cytosolic fraction, in agreement with previous

reports (Schwerdtfeger and Linden, 2003). In treated mycelia,

a substantial portion of VVD was recovered in the nuclear frac-

tion, in agreement with the localization of VVD-GFP in living

mycelia. These data suggest that nuclear VVD is readily lost

during subcellular fractionation procedures.

VVD Competes with Light-Induced Dimerization of WCCVVD has been reported to form light-dependent dimers that are,

however, highly unstable and dynamic (Zoltowski and Crane,

2008). We reasoned that WCC might dimerize in corresponding

fashion via the LOV domain of WC-1 and that VVD could

compete with this process by interacting with the LOV domain

of WC-1. Attempts to coimmunoprecipitate light or dark states

of VVD and WCC from total cell extracts failed in all possible

combinations (data not shown), indicating that VVD and WCC

do not interact in a stable manner. We constructed a strain

expressing a strep-tagged version of VVD (cpc-1-vvdStrep).

When VVDStrep was purified via a streptavidin affinity resin,

WCC copurified in low but significant amounts (Figure S2A).

To detect interactions of low affinity by independent means,

we examined the interaction of the LOV domains of WC-1 and

VVD in the yeast two-hybrid system. The DNA binding domain

(BD) of the Gal4 transcription factor was fused to the LOV

domain of WC-1, and the transactivation domain of Gal4 (AD)

was fused to the LOV domain of WC-1 and VVD, respectively.

Pairs of LOV domain fusion proteins with BD (‘‘bait’’) and with

the AD (‘‘prey’’) were expressed in a yeast reporter strain that

allows selection for the interaction of ‘‘bait’’ and ‘‘prey’’ via the

fused LOV domains. Since LOV domains are light receptors,

the yeast transformants were grown in light as well as in dark-

ness. The LOV domains of WC-1 displayed homodimerization

as well as heterodimeric interaction with VVD when the corre-

sponding yeast strains were grown on selective medium in light.

However, the strains did not grow in darkness, indicating that the

LOV domains did not interact in the absence of light (Figure 2A).

A positive control strain grew in light and in darkness while

a negative control did not grow at all (Figure 2A). The data

demonstrate that the LOV-domain of WC-1 has the capacity to

interact with itself and with VVD in a light-dependent fashion.

We then asked whether the interaction of VVD with WC-1

affects WCC function. WCC is a GATA-type transcription factor

that binds to light-responsive elements, such as the proximal

light regulatory element (LRE) of the frq promoter (pLRE).

EMSA analysis can discriminate between light and dark states

of the WCC (Froehlich et al., 2002; He and Liu, 2005). Light-acti-

vated and dark forms of WCC display predominantly low or high

electrophoretic mobility shifts (Figure 2B, lanes 1 and 2). Since

we have shown above that the LOV domains of WC-1 dimerize

in light-dependent fashion, the light-activated WCC should

represent a dimer of the WCC monomer that is the prevalent

form in the dark.

764 Cell 142, 762–772, September 3, 2010 ª2010 Elsevier Inc.

Page 133: Cell 100903

To demonstrate this directly in an EMSA, we generated an

antibody against the LOV domain of WC-1 that should interfere

with dimerization of WCC protomers. Nuclear extracts of dark-

grown mycelia were preincubated with an excess of LOV anti-

body over WCC to saturate each WC-1 molecule with one anti-

body. Subsequently, one aliquot was exposed to a saturating

light pulse, while the other was kept in the dark. By EMSA anal-

ysis, the dark form of WCC was super-shifted to a single discrete

band of lower mobility, indicating that one LOV antibody had

bound to the WC-1 subunit of the WCC (Figure 2B, lane 3).

However, a subsequent light pulse did not induce a further

super-shift (lane 4). This finding indicates that binding of LOV

antibodies to the dark form of WCC interfered with light-depen-

dent dimerization. In a further assay, nuclear extract was first

preincubated in light with the pLRE oligonucleotide to allow

binding of the light-activated WCC dimer to its DNA binding

site. Subsequently, LOV antibody was added and the sample

was then analyzed by EMSA (lane 5). Under these conditions,

the light-activated WCC was super-shifted, suggesting that the

WCC dimer was crosslinked by one antibody bound to two

LOV domains, a reaction that should be kinetically favored,

when WCC is a dimer. Together, the data demonstrate that

WCC dimerizes in a light-dependent fashion.

If WCC dimers are highly dynamic, the monomer-dimer ratio

should be sensitive to the WCC concentration. To vary the

concentration of WCC, we performed EMSA in a fixed volume

with increasing amounts of light-exposed nuclear extract. As

expected, the intensity of the EMSA signals increased due to

the increasing amounts of WCC (Figure S2B). At high concentra-

tions of extract, only the mobility shift corresponding to the

Figure 2. VVD Interferes with Dimerization

of the Light-Activated WCC and Thereby

Mimics the Monomeric Dark State of the

WCC

(A) Light-dependent homo- and heterodimeriza-

tion of the LOV domains of WC-1 and VVD

analyzed by yeast two-hybrid assay. The LOV

domain of WC-1 was fused to the DNA binding

(BD) of the Gal4 transcription factor. The transac-

tivation domain of Gal4 (AD) was fused to the LOV

domains of WC-1 and VVD, respectively as indi-

cated. The yeast strain PJ69-4A expressing HIS3

under control of a Gal4-dependent promoter was

transformed with the indicated combination of

constructs. Positive control, BD-p53/AD-SV40;

negative control, BD and AD vectors without

insert. Corresponding yeast transformants were

plated on either selective (SDC -Leu -Trp -His;

Two Hybrid) or nonselective medium (SDC -Leu

-Trp; Plating) in dilution series as indicated

(1 corresponds to 2.5 OD). Growth of transform-

ants after 3 day incubation is shown.

(B) Light-activated WCC forms dimers. Nuclear

extract of dark-grown mycelia (DD) or light-pulsed

extract (LP) was preincubated without or with

excess of affinity-purified antibody against the

LOV-domain of WC-1. Subsequently samples

were either kept in the dark (DD) or exposed to

a saturating light pulse (LP) when indicated, and

electrophoretic mobility shift assays (EMSA) of

radiolabeled pLRE oligonucleotide were per-

formed. The LOV antibody interferes with dimer-

ization of the light-activated WCC. Monomeric

and dimeric WCC with and without bound LOV

antibody are schematically depicted above the

corresponding lanes. The dark form of the WCC

is shown in gray, and the light-activated WCC is

shown in white.

(C) Concentration dependence of the monomer-

dimer equilibrium of light-activated WCC. EMSA

was performed with 1 and 9 ml light-pulsed nuclear

extract (total volume 20 ml). A long exposure (l.e.)

of the EMSA with low concentration of extract

(1 ml/assay) is shown to compare the ratio of monomeric and dimeric WCC to a high concentration of extract (9 ml/assay).

(D) VVD interferes with dimerization of the light activated WCC functionally mimicking the dark form. Nuclear extract (DD) was light pulsed (LP) and preincubated

with and without recombinant His6-VVD36. When indicated, His antibody was added to neutralize the His6-VVD36, and EMSAs were performed. All EMSAs were

performed at least three times with identical results.

See also Figure S2.

Cell 142, 762–772, September 3, 2010 ª2010 Elsevier Inc. 765

Page 134: Cell 100903

dimeric WCC was detected. However, a long exposure of the

X-ray film revealed monomeric and dimeric WCC at low con-

centrations of light-exposed nuclear extract (Figure 2C and

Figure S2B). This finding indicates that the light-induced dimer-

ization of WCC is dynamic and depends on the concentration of

WCC.

For functional characterization of the interaction of VVD and

WCC, nuclear extract was first exposed to a saturating light-

pulse, then incubated with recombinant light-activated His6-

VVD36 (Zoltowski et al., 2007), and finally, examined by EMSA

(Figure 2D). In the presence of His6-VVD36, the light-activated

WCC displayed a mobility shift characteristic for the dark form.

In its absence, WCC displayed the dimer-specific mobility shift.

These observations indicate that His6-VVD36 disrupted the WCC

dimers. In further characterizations, the nuclear extract was

exposed to light and preincubated with recombinant light-acti-

vated His6-VVD36 as described above. The sample was trans-

ferred to darkness, and anti-His antibody was added to oligo-

merize His6-VVD36 and thereby functionally neutralize the

protein. Although the sample did not receive additional light,

subsequent analysis by EMSA revealed that the dimer-specific

Figure 3. Formation and Decay of the Light Form of

the WCC

(A) Kinetics of formation of the light-activated WCC.

Nuclear extracts of dark grown mycelia (DD, time point

‘‘0’’) were irradiated with the indicated intensities and for

the indicated times and examined by EMSA.

(B) Upper (light complex) and lower (dark complex) bands

were quantified by densitometry and the ratio of upper/

(upper + lower) was calculated for each time point and

the halftime for conversion was determined.

(C) Decay kinetics of the WCC light form. Light-treated

nuclear extracts were incubated in the dark for the indi-

cated times. Left: Samples were analyzed by EMSA.

Open and filled arrowheads indicate the light-activated

and the dark form of the WCC, respectively. EMSAs

were performed at least three times with identical results.

Right: Densitometric analysis of shown EMSA. So that the

functionality of the WCC could be controlled, an aliquot of

each sample was light pulsed to recover the light form

(data not shown).

See also Figure S3.

mobility shift of the light-activated WCC was

fully restored. The data demonstrate that VVD

interacts with WCC and interferes with dimeriza-

tion of the WCC protomers. It thereby function-

ally resets the light-activated WCC to the dark

form.

Photochemical Properties of WCCand VVDActivation of WCC protomers by light-induced

dimerization and subsequent competitive inhibi-

tion of dimerization by newly synthesized VVD

readily explains the phenomenon of induction

and adaptation of WCC-dependent transcrip-

tion in response to light on a molecular level.

However, the model does not explain how thesystem maintains its light-sensing capacity over several orders

of magnitude, since a sizable pool of the WCC must be in the

dark state to sense a further increase in light intensity in constant

light conditions.

To estimate the equilibrium between WCC light and dark

states, we measured the kinetics of formation and decay of the

light complex. Nuclear extracts of dark-grown mycelia were

exposed to light of different intensities, and the kinetics of

WCC dimerization was analyzed by EMSA (Figures 3A and 3B).

At high light intensity (420 mmol 3 m�2 3 s�1), an exposure of

15 s was sufficient for complete light activation of WCC. At inter-

mediate intensity (80 mmol 3 m�2 3 s�1), an exposure of about

5 min was required, and at low light (3.6 and 6 mmol 3 m�2 3

s�1), 20–30 min was needed. When light-exposed nuclear

extract was transferred to the dark, WCC dimers were stable

for several hours (t1/2 �4 hr) (Figure 3C).

The ratio of forward versus reverse rates indicates that the

fraction of WCC in the dark state depends on light intensity.

In bright light (420 mmol 3 m�2 3 s�1), the fraction of WCC in

the dark form is rather small (6 s/14,400 s = 43 10�4). In constant

intermediate light (6 mmol 3 m�2 3 s�1), about 12.5% of the

766 Cell 142, 762–772, September 3, 2010 ª2010 Elsevier Inc.

Page 135: Cell 100903

WCC should be in the dark state (30 min/240 min). Extrapolating

to low light, we estimate that the fraction of WCC in the dark form

is substantial in dim light (>30% below 1 mmol 3 m�2 3 s�1).

Thus, the pool of WCC in the dark form is detectable and there-

fore of physiological relevance for light-sensing in the photoa-

dapted state. Recombinant His6-VVD36 displayed similar light

sensitivity and photoadduct stability (Figure S3).

VVD Stabilizes the Light-Activated WCCDNA-binding-competent, activated WCC is unstable (Schafme-

ier et al., 2008). Since VVD competes with dimerization of the

WCC and thereby reduces its activity, we asked whether VVD

also affects the degradation kinetics of WCC. WC-1 was rapidly

degraded when dark-grown mycelia were treated with cyclohex-

imide (CHX) and then exposed to saturating light (Figure 4A)

(Schafmeier et al., 2008). To assess the turnover of light-

activated WCC in the presence of VVD under corresponding

conditions, we analyzed the stability of WC-1 in the Dvvd, cpc-

1-vvd-gfp strain that constitutively expresses VVD-GFP in a

light-independent manner. WC-1 was stable when dark-grown

mycelia of Dvvd, cpc-1-vvd-gfp were treated with CHX and

exposed to light (Figure 4A), indicating that the previously

synthesized VVD-GFP interfered with the light-induced turnover

of the WCC. The acute induction of transcription by light was

attenuated in the Dvvd, cpc-1-vvd-gfp strain, demonstrating

that VVD-GFP was immediately active as a repressor of the

light-activated WCC (Figure S4A). In a vvd deletion mutant,

light-dependent gene expression is induced to a higher level

and adapts slower (Schwerdtfeger and Linden, 2001) (Fig-

ure S2B). In summary, VVD reduces the transcriptional activity

of the WCC, which results in a stabilization of the transcription

factor, since only the active WCC is subjected to rapid degrada-

tion (Schafmeier et al., 2008).

WCC levels and activity do not depend solely on VVD but are

regulated in a complex manner. On the posttranscriptional level,

FRQ-dependent phosphorylation inactivates and stabilizes

WCC (Schafmeier et al., 2008; Schafmeier et al., 2005). On the

transcriptional level, WCC supports expression of WC-1, the

limiting component of WCC (Kaldi et al., 2006).

To assess the contribution of FRQ and VVD to stabilizing the

light-activated WCC, we analyzed WC-1 levels in frq- and vvd-

deficient strains grown in light (Figure 4B). WC-1 levels were

distinctly lower in frq9 (nonfunctional allele of frq) and Dvvd

than in WT cells, demonstrating that FRQ and VVD indepen-

dently stabilize the light-activated WCC to a similar extent.

WC-1 levels were even lower in the double deficient frq9, Dvvd

strain. This observation indicates that the positive feedback of

WCC on wc-1 transcription cannot compensate its light-induced

degradation.

VVD Synthesized during the Day Suppresses Moonlightduring the NightWe reasoned that overexpression of VVD above levels expected

by the feedback system on the basis of the prevalent light inten-

sity should result in blindness of Neurospora. To test this predic-

tion, we compared the light-entrainment characteristics of Dvvd,

cpc-1-vvd-gfp, and Dvvd (Figure 5A). After transfer from bright

light to darkness (DD), both strains displayed a free-running con-

idiation rhythm, indicating that the circadian clock was functional

and synchronized by the LD transfer. Dvvd entrained to LD

cycles at low and high light intensities. In contrast, Dvvd, cpc-

1-vvd-gfp entrained to LD cycles with a light intensity of

180 mmol 3 m�2 3 s�1 but displayed a free-running rhythm at

lower light intensities (2.8 and 180 nmol 3 m�2 3 s�1). Hence,

overexpression of VVD-GFP desensitizes the WCC and renders

the strain blind to the lower levels of light intensity.

What is the physiological role of desensitization of the WCC

photoreceptor by VVD? Photoperiods in nature are quite noisy.

In particular, nights are not always entirely dark, and thus, abso-

lute light intensity is an ambiguous cue that does not allow

discrimination between day and night. On a clear night with

a full moon, the light intensity (�20 nmol 3 m�2 3 s�1) is well

above the sensitivity of the WCC photoreceptor (<2.8 nmol 3

m�2 3 s�1), and hence, moonlight should reset the clock.

However, light intensities during the day are more than 103-

fold higher than moonlight. VVD that is expressed at high levels

during the day is turned over in the dark with a half-time of about

2.5 hr (Figure 5B). Accordingly, VVD levels should remain

Figure 4. VVD Stabilizes the Light-Activated WCC

(A) Light-induced degradation of WC-1 in the WT and Dvvd, cpc-1-vvd-gfp

strain. Mycelia were grown in constant darkness. Cycloheximide (CHX)

(10 mg/ml) was added, and cultures were transferred to the light. Samples

were harvested after the indicated times, and WC-1 was analyzed by western

blotting.

(B) Positive feedback of FRQ and VVD on levels of light-activated WCC.

Steady-state expression levels of WC-1 in frq- and vvd-deficient strains grown

in light are shown in comparison to the WT. A long exposure is shown to illus-

trate differences in WC-1 levels in the low-expressing mutant strains. Repre-

sentative western blots are shown (n = 3).

See also Figure S4.

Cell 142, 762–772, September 3, 2010 ª2010 Elsevier Inc. 767

Page 136: Cell 100903

sufficiently high throughout the night to suppress responses

evoked by the considerably dim moonlight and thus allow

entrainment of Neurospora to natural photoperiods.

To test this hypothesis, we exposed race tubes inoculated

with the WT and Dvvd to natural photocycles. Incubation was

started on a dark new moon night and terminated 6 days later

at almost half moon (Figure 5C). The WT entrained to the natural

photoperiod throughout the entire time period. In contrast, Dvvd

entrained for 3 or 4 days, but the rhythm was then obviously

disturbed by the increasing moonlight, and the strain became

arrhythmic. Similarly, the WT entrained to artificial moonlight

(240 nmol 3 m�2 3 s�1) in a 12 hr high-light:12 hr low-light cycle

while entrainment of Dvvd was disturbed (Figure S5). Both

strains became arrhythmic when the intensity of the artificial

moonlight was raised to 800 nmol 3 m�2 3 s�1. The data

demonstrate that VVD is required for robust entrainment of

Neurospora to ambiguous natural photoperiods.

DISCUSSION

In this study, we showed that Neurospora has evolved an elabo-

rate molecular mechanism of light activation and adaptation of

transcription that allows discrimination between night and day

on the basis of relative differences in light intensity levels. The

system requires the interaction of two LOV domain photorecep-

tors, the activating photoreceptor WCC, and the inhibitory

photoreceptor VVD that constitute a negative feedback loop

(Elvin et al., 2005; Heintzen et al., 2001). We present here the

molecular mechanism of the feedback and describe properties

that are crucial for the dynamic behavior of this system that

allows light sensing in the photoadapted state over several

orders of light intensity (see model in Figure 6). Light activation

of the LOV domain of the WC-1 facilitates dimerization of proto-

mers of the WCC. Dimers of WCC display a higher activity than

the monomeric WCC protomers, which are prevalent in dark-

ness. Hence, abrupt light exposure of dark grown Neurospora

triggers a burst of transcription of vvd, frq, and other light-

induced genes. Subsequent photoadaptation is a complex pro-

cess dependent on one hand on rapid degradation of activated

WCC and on the other hand on inactivation and stabilization of

WCC by VVD and FRQ. Rapid degradation of activated WCC

would, in principle, lead to an efficient attenuation of light-

induced transcription. However, it would be detrimental if light-

induced degradation of WCC were the sole mechanism of adap-

tation, since this would result in severe depletion of the pool of

WCC. Although transcription of wc-1 is light induced by the

WCC (Kaldi et al., 2006), this positive feedback loop cannot

compensate for the light-induced degradation. Hence, WCC

levels are extremely low in a frq9, Dvvd strain.

VVD, which is rapidly synthesized in light, acts on the light-

activated WCC by inhibiting, in a competitive manner, the forma-

tion of highly active WCC dimers. This causes a rapid and effi-

cient attenuation of light-induced transcription. Since inactive

WCC is not subjected to rapid degradation (Schafmeier et al.,

2008) VVD promotes accumulation of the light-activated but

transcriptionally inactive WCC. Hence, VVD resets the light-acti-

vated WCC to a monomeric, inactive, and stable form that is

functionally equivalent to the dark state (Figure 6).

Figure 5. VVD Is a Critical Element for Controlling Neurospora

Response to Light

(A) Desensitization of Neurospora to light depends on VVD expression levels.

Densitometric analysis of race-tube assays of Dvvd and Dvvd, cpc-1-vvd-gfp

strains as shown in Figure 1D. Strains were grown in high light and then trans-

ferred to the dark and exposed to a 13/13 hr LD regime with the indicated light

intensity during the light phase. Overexpression of VVD-GFP caused blindness

to entraining light cues of lower intensity resulting in a free-running conidiation

rhythm. Plots from representative race tubes are shown (n R 9).

(B) VVD synthesized during the light phase is progressively degraded in the

dark. Light-grown cultures were transferred to darkness, and VVD levels

were determined after the indicated time periods by western blotting with an

antibody against the N terminus of VVD. Asterisk indicates unspecific reaction

of the VVD antibody. Representative western blot is shown (n = 3).

(C) VVD is essential for the entrainment of Neurospora to natural photoperiods

with moonlight. Analysis of race-tube assays of the WT and Dvvd. Race tubes

were exposed to natural day-night cycles for 6 days beginning at a dark new-

moon night and ending at almost half moon. Incubation was performed in an

isolated room at 21�C. Conidial density traces of the race-tubes of the WT

and Dvvd are shown. The moon phases are indicated below the traces.

Dvvd became arrhythmic with upcoming moonlight.

See also Figure S5.

768 Cell 142, 762–772, September 3, 2010 ª2010 Elsevier Inc.

Page 137: Cell 100903

In addition to VVD, the circadian clock protein FRQ also

supports inactivation and stabilization of WCC (Schafmeier

et al., 2008; Schafmeier et al., 2005). FRQ inactivates WCC in

a light-independent fashion by facilitating its phosphorylation

by casein kinase 1a. However, since FRQ expression is sup-

ported by WCC, FRQ gradually accumulates to an elevated level

in constant light, resulting in an enhanced feedback of FRQ that

partially compensates for the increased activity of light-activated

WCC. Thus, vvd-deficient but frq-proficient strains display

a slow and partial attenuation of light-induced transcription

(Schwerdtfeger and Linden, 2001).

Notably, the circadian clock is slightly phase delayed but

otherwise fully functional in vvd-deficient (frq-proficient) strains.

This mild phenotype provokes the question of whether VVD

has a physiologically relevant specific function that cannot

be provided by FRQ. While FRQ is a general inhibitor of the

WCC, VVD inhibits specifically the light-activated WCC. Under

entrained conditions VVD accumulates to a high level during

the day and efficiently attenuates transcription induced by the

light-activated WCC. Although VVD is degraded during the night,

the levels remain sufficiently high to suppress activation of the

WCC by moonlight (>1000-fold lower light intensity than

daylight). In the absence of VVD (Dvvd), the highly sensitive

WCC is activated by moonlight, and the clock cannot entrain

because of the ambiguous light cues. Hence, in natural photope-

riods, VVD is an essential component of the circadian clock.

It modulates the sensitivity and responsiveness of the system

Figure 6. Model of the Molecular Interactions of WCC

and VVD upon Light Induction and Adaptation

W and V denote the dark forms of the WCC and VVD. Accord-

ingly, W* and V* denote the light-activated forms of the photo-

receptors. Upon light exposure, W is converted to W*, which

dimerizes (W*W*) and efficiently activates transcription of

vvd. The DNA-bound W*W* complex is unstable and rapidly

degraded (Deg.). As V* accumulates, it interacts with W* and

thereby competes with the homodimerization of W*. The

W*V* complex is transcriptionally inactive and stable, leading

to light adaptation of vvd transcription. The stable W*V* accu-

mulates gradually and equilibrates with the dark forms of W

and V. According to law of mass action, the equilibrium is

established by light-dependent conversion of W to W* and

light-independent decay of the W* photoadduct. Stabilization

of W*V* and equilibration with the dark forms lead to a substan-

tial population of W, which allows light sensing in the photoa-

dapted state.

to light cues and thus allows proper interpretation

of ambiguous light signals. Desensitization of

WCC correlates with the expression level of VVD,

which accumulates as a function of duration and

intensity of light exposure. The accumulated VVD

provides a molecular memory of the previous light

history, which is gradually fading in darkness. Since

the extent of desensitization is gradually lost during

the night, the system remains blind for levels of light

that are low in relation to the previous day and can

thus ignore moonlight as a resetting cue. However,

the adapted system remains responsive to stimuli

of higher light intensity. The capacity to respond to light cues

of higher intensity depends on accumulation of a significant

pool of the dark form of WCC. This is facilitated by VVD, which

inactivates the light-activated WCC and thereby blocks its

degradation. Since the light and dark forms of the WCC are in

equilibrium, stabilization of the light-activated WCC allows accu-

mulation of WCC in the dark form. According to law of mass

action, the equilibrium is established by formation of the photo-

adduct, which depends on light intensity, and its spontaneous,

light independent decay. When this equilibrium is disturbed by

a sudden increase in light intensity, the accumulated pool of

WCC in the dark state is activated. It transiently outcompetes

the inhibitory capacity of VVD until sufficient additional VVD is

synthesized to establish a new steady state that reflects adapta-

tion to the higher light intensity. Since the WCC/VVD system can

discriminate light intensities over more than five orders of magni-

tude, Neurospora readily adapts to habitats that differ in lighting

conditions.

In constant darkness, rhythmic expression of VVD dampens

rapidly and levels of VVD drop below the limit of detection (Elvin

et al., 2005). The absence of VVD after several days in darkness

renders the WCC highly sensitive to light cues. This can

be viewed as dark adaptation of the light-sensing system of

Neurospora.

Finally, the flavin-cysteinyl photoadducts of WCC and VVD are

stable for several hours while the photoadducts of other LOV

domains decay in the range of minutes. This extraordinary

Cell 142, 762–772, September 3, 2010 ª2010 Elsevier Inc. 769

Page 138: Cell 100903

photoadduct stability should smoothen the activity profile of the

WCC such that noisy fluctuations in light intensity (e.g., as a result

of moving clouds concealing the sun) are not transduced by the

photoreceptor system. Rather, the photoreceptors integrate the

light signals and keep a slowly decaying molecular memory of

the maximal light level of the previous time period.

We showed that light-sensing and adaptation (desensitization)

in Neurospora depend on homo- and heterodimerization of the

activating and inhibitory LOV domains of WC-1 and VVD,

respectively. Flavin-binding LOV domains are also found in plant

phototropins (phot1 and phot2). These are membrane-associ-

ated receptor kinases that dimerize in response to blue light

and activate a number of different processes, such as phototro-

pism and chloroplast movement (Christie, 2007). Phototropins

always contain a pair of LOV domains, LOV1 and LOV2. LOV2

is a dark-state inhibitor of receptor kinase autophosphorylation,

and its activation by light is essential for signaling (Cho et al.,

2007; Christie et al., 2002; Kaiserli et al., 2009). The role of

LOV1 is less clear. It is not essential for light signaling. Photoac-

tivation of LOV1 appears to have an inhibitory role at high light

intensities (Kaiserli et al., 2009), but the underlying molecular

mechanism is not understood. The photochemical properties

of phot1 suggest an interaction between LOV1 and LOV2

(Kaiserli et al., 2009). It is tempting to speculate that, like the

WCC-VVD system, LOV2 might act by facilitating light-depen-

dent dimerization of two phototropins, while heterodimerization

of LOV2 with LOV1 could be inhibitory by interfering with

receptor dimerization. In principle, heterodimerization of LOV1

and LOV2 should be kinetically favored since both domains

occur in the same molecule. However, the quantum yield for

photoadduct formation (Kasahara et al., 2002) and the rapid

decay of the photoadducts of LOV1 and LOV2 (Salomon et al.,

2000) suggest that light of considerable intensity is required to

simultaneously activate both LOV domains in one phototropin

molecule. Hence, an intramolecular inhibitory dimerization of

LOV1 and LOV2 could desensitize the system at high light inten-

sities and attenuate the strength of downstream signaling. Thus,

competitive homo- and heterodimerization of activating and

inhibitory LOV domains might be a general mechanism for

light-sensing and adaptation.

In summary, we report a molecular mechanism of photo-acti-

vation and adaptation in Neurospora. In response to light, WCC

forms homodimers that induce transcription of vvd. VVD, in turn,

competes for binding with WCC, disrupts the WCC dimers, and

attenuates light-induced transcription. In the photoadapted

state, VVD synthesis by light activated WCC is balanced by

VVD-dependent inhibition of WCC. The VVD-mediated desensi-

tization of the light-activated WCC provides a molecular mecha-

nism for how Neurospora robustly entrains to photoperiods.

EXPERIMENTAL PROCEDURES

Neurospora Strains and Culture Conditions

Neurospora strains used in this study (WT; Dvvd; frq9, Dvvd; Dvvd, cpc-1-gfp;

Dvvd, cpc-1-vvd-gfp; and Dvvd, cpc-1-vvdStrep) carried the ras-1bd mutation

(Belden et al., 2007). Standard growth medium contained 2% glucose, 0.5%

L-arginine, 1 3 Vogel’s and 10 ng/ml biotin. Cultures were incubated 48 hr

at 25�C if not indicated otherwise. In the race-tube medium, glucose was

omitted, and arginine was reduced to 0.17%. Race tubes contained 2.2%

agar. Analysis of race tubes was performed by densitometry with the ChronoII

software (T. Roenneberg, Ludwig-Maximilians-Universitat Munich).

Plasmid Construction

cpc-1-gfp was constructed by replacement of the ccg-1 promoter with the

cpc-1 promoter in pMF272 (generous gift of Michael Freitag) and insertion of

an AscI site between the promoter and the SpeI site. For construction of

cpc-1-vvd-gfp, a genomic PCR product, including the coding region of vvd

and 46 bp upstream of the start codon, was inserted as an AscI-SpeI fragment

into cpc-1-gfp. For construction of cpc-1-vvd-Strep, a PCR fragment contain-

ing a genomic region of vvd from position�46 to +691, a short linker and nucle-

otides coding for a C-terminal Strep-tag (Herzberg et al., 2007), followed by

a stop codon, was amplified and inserted between the AscI and SpeI sites

of pcpc-1-gfp.

pET28a-vvd36 for the expression of a soluble N-terminal truncated version of

VVD fragment (Zoltowski et al., 2007) was kindly provided by Brian R. Crane.

Plasmids for the yeast two-hybrid assay were based on pGBD-C1 and

pGAD-C1 (James et al., 1996). The VVD36 fragment and the LOV domain of

WC-1 were amplified by PCR from Neurospora total complementary DNA

(cDNA) and inserted in the EcoRI and BamHI sites of the respective vectors.

Neurospora Transformation

Neurospora conidia (5–7 days old) were transformed by electroporation as

described (Schafmeier et al., 2006). The his-3 locus was used for homologous

recombination. Insertion of the target gene was verified by PCR and

sequencing. Homokaryotic strains were generated by purification by several

plating steps and checked by quantitative PCR.

RNA Analysis

Total RNA was extracted with the peqGOLD TriFast reagent (Peqlab). cDNA

was synthesized from 1 mg total RNA with the QuantiTect Reverse Transcrip-

tion Kit (QIAGEN). Transcript levels were quantified by RT-PCR with TaqMan-

Probes in a StepOne system (Applied Biosystems). Triplicate reactions (20 ml)

containing cDNA equivalent to 0.05 mg RNA were analyzed.

Protein Analysis

Extraction of Neurospora protein, subcellular fractionation, and extraction of

nuclear proteins were performed as described (Schafmeier et al., 2005).

Protein concentration was determined by Bradford assay, and 350 mg (total

extract, cytosol) or 100 mg (nuclei or nuclear extract) were analyzed by SDS-

PAGE. Western blots were performed as described (Schafmeier et al.,

2005). Enhanced chemiluminescence signals were detected with X-ray films.

Series of exposures were generated and signals were quantified with Adobe

Photoshop as described previously (Schafmeier et al., 2008). The VVD anti-

body used for Western blotting was directed against the N terminus.

Fluorescence Microscopy

Neurospora cultures were grown in liquid medium for 1–2 days. Pieces of

mycelia were placed on an object slide and embedded in mounting medium

(13% Mowiol [Hoechst], 30% glycerol [pH 8.5]) containing 0.25 mg/ml DAPI.

Microscopy was performed with a Leica DMI 6000B device. GFP and DAPI

were excited at 470 and 360 nm, respectively.

Purification of Recombinant VVD

Affinity purification of VVDStrep was performed with the One-STrEP System

(IBA). Protein extracts of light grown Dvvd and Dvvd, cpc-1-vvdStrep corre-

sponding to 150 mg protein were applied to 400 ml Strep-Tactin Superflow

and incubated for 1 hr at 4�C under constant rotation. After several washing

steps, the resin was packed into a column and proteins were eluted with

desthiobiotin in 250 ml fractions.

His6-VVD36 was expressed in Eschericia coli BL21 (DE3) cells under control

of the T7 promoter. T7 RNA polymerase was induced by addition of 0.1 mM

IPTG for 18 hr at 16�C. His6-VVD36 was purified by affinity chromatography with

HisTrap HP columns (GE Healthcare) as recommended by the manufacturer.

770 Cell 142, 762–772, September 3, 2010 ª2010 Elsevier Inc.

Page 139: Cell 100903

Yeast Transformation and Two-Hybrid Assay

The yeast strain PJ69-4A was transformed with 500 ng of plasmid DNA (Gietz

and Woods, 2002; Ito et al., 1983). Transformed cells were plated on SDC solid

medium lacking leucin and tryptophane (SDC -Leu -Trp). Single colonies were

picked, plated again, and checked for positive transformation by analysis of

protein expression.

For the yeast two-hybrid assay, cells were transformed with WC-1LOV-

GAL4-BD (bait) and either GAL4-AD-VVD or GAL4-AD- WC-1LOV (prey). Plas-

mids containing GAL4-BD-P53 and GAL4-AD-SV40 were used as positive

control, and GAL4-BD and GAL4-AD were used as negative control. Single

clones were inoculated on a sector of plates containing SDC -Leu -Trp and

SDC -Leu -Trp -His medium. Dilution series of positive clones were spotted

on plates for direct comparison of growth rate.

Electrophoretic Mobility Shift Assay

EMSA with g32P-ATP-labeled double-stranded oligonucleotide corresponding

to the proximal LRE (50-CGCAGAGGACCCTGAACTTTTCGATCCGCTCGA

TCCCCTG GAA-30 ) was performed as described (Froehlich et al., 2002). Puri-

fied oligonucleotide (180 fmol) and 10–20 mg nuclear extract were incubated in

binding buffer (20 mM HEPES, 40 mM KCl, 1 mM EDTA, 10% glycerol, 0.2 mM

DTT, 10 mM MgCl2, 0.1% N-p40) on ice for 30 min. Poly dIdC (1 mg) was used

as unspecific competitor. Samples were loaded on a gel containing 4% acryl-

amide in TBE buffer. Gels were fixed in 10% acetic acid, dried, and analyzed

by autoradiography. Antibodies used for EMSA specifically were a-His6

(QIAGEN) and a-LOV, which was generated against a peptide obtained from

the WC-1 LOV-domain (RFLQAPDGNVEAGTKREF).

Analysis of Photoadduct Formation and Decay

Purified His6-VVD36 or Neurospora nuclear extract were prepared as

described above. For analysis of photoadduct formation, samples were

prepared in red light. His6-VVD36 and nuclear extracts were irradiated with

the indicated light intensities. Conditions with different light intensities were

established through the use of a variable number of neon tubes in combination

with light absorber foil. Light intensity was measured with a LI-250A Light

Meter (LI-COR Biosciences). For decay analysis, His6-VVD36 or nuclear

extracts were irradiated with saturating light (10 min, 130 mmol 3 m�2 3 s�1)

and transferred to darkness. WC-1 photoadduct formation and decay were

analyzed by EMSA, and the His6-VVD36 photoadduct was analyzed photomet-

rically by measurementof the absorbance spectrum of the FAD cofactor as

described (Zoltowski et al., 2007).

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.

2010.08.010.

ACKNOWLEDGMENTS

We thank Johanna Scholz and Krisztina Makara for excellent technical

assistance. We are grateful to Cornelia Ulbrich for her experienced help with

setting up the yeast two-hybrid system. This work was supported by the

DFG grant BR 1375-3-1 to M.B. and T.S., SFB 638 and the FCI to M.B.,

GP1188 to E.M. and by an EMBO-HHMI startup grant by the Hungarian

OTKA (K 68960) and ETT (477-05) to K.K., M.B. is a member of CellNetworks

and S.C. is a member of HBIGS.

Received: February 16, 2010

Revised: June 16, 2010

Accepted: August 7, 2010

Published: September 2, 2010

REFERENCES

Bachleitner, W., Kempinger, L., Wulbeck, C., Rieger, D., and Helfrich-Forster,

C. (2007). Moonlight shifts the endogenous clock of Drosophila melanogaster.

Proc. Natl. Acad. Sci. USA 104, 3538–3543.

Belden, W.J., Larrondo, L.F., Froehlich, A.C., Shi, M., Chen, C.H., Loros, J.J.,

and Dunlap, J.C. (2007). The band mutation in Neurospora crassa is a domi-

nant allele of ras-1 implicating RAS signaling in circadian output. Genes Dev.

21, 1494–1505.

Brunner, M., and Schafmeier, T. (2006). Transcriptional and post-transcrip-

tional regulation of the circadian clock of cyanobacteria and Neurospora.

Genes Dev. 20, 1061–1074.

Cho, H.Y., Tseng, T.S., Kaiserli, E., Sullivan, S., Christie, J.M., and Briggs, W.R.

(2007). Physiological roles of the light, oxygen, or voltage domains of phototro-

pin 1 and phototropin 2 in Arabidopsis. Plant Physiol. 143, 517–529.

Christie, J.M. (2007). Phototropin blue-light receptors. Annu. Rev. Plant Biol.

58, 21–45.

Christie, J.M., Swartz, T.E., Bogomolni, R.A., and Briggs, W.R. (2002). Photo-

tropin LOV domains exhibit distinct roles in regulating photoreceptor function.

Plant J. 32, 205–219.

Corchnoy, S.B., Swartz, T.E., Lewis, J.W., Szundi, I., Briggs, W.R., and Bogo-

molni, R.A. (2003). Intramolecular proton transfers and structural changes

during the photocycle of the LOV2 domain of phototropin 1. J. Biol. Chem.

278, 724–731.

Dibner, C., Schibler, U., and Albrecht, U. (2010). The mammalian circadian

timing system: organization and coordination of central and peripheral clocks.

Annu. Rev. Physiol. 72, 517–549.

Doyle, S., and Menaker, M. (2007). Circadian photoreception in vertebrates.

Cold Spring Harb. Symp. Quant. Biol. 72, 499–508.

Elvin, M., Loros, J.J., Dunlap, J.C., and Heintzen, C. (2005). The PAS/LOV

protein VIVID supports a rapidly dampened daytime oscillator that facilitates

entrainment of the Neurospora circadian clock. Genes Dev. 19, 2593–2605.

Foster, R.G., Hankins, M.W., and Peirson, S.N. (2007). Light, photoreceptors,

and circadian clocks. Methods Mol. Biol. 362, 3–28.

Froehlich, A.C., Liu, Y., Loros, J.J., and Dunlap, J.C. (2002). White Collar-1,

a circadian blue light photoreceptor, binding to the frequency promoter.

Science 297, 815–819.

Gietz, R.D., and Woods, R.A. (2002). Transformation of yeast by lithium

acetate/single-stranded carrier DNA/polyethylene glycol method. Methods

Enzymol. 350, 87–96.

Harper, S.M., Neil, L.C., and Gardner, K.H. (2003). Structural basis of a photo-

tropin light switch. Science 301, 1541–1544.

He, Q., and Liu, Y. (2005). Molecular mechanism of light responses in Neuros-

pora: from light-induced transcription to photoadaptation. Genes Dev. 19,

2888–2899.

He, Q., Cheng, P., Yang, Y., Wang, L., Gardner, K.H., and Liu, Y. (2002). White

collar-1, a DNA binding transcription factor and a light sensor. Science 297,

840–843.

Heintzen, C., Loros, J.J., and Dunlap, J.C. (2001). The PAS protein VIVID

defines a clock-associated feedback loop that represses light input, modu-

lates gating, and regulates clock resetting. Cell 104, 453–464.

Helfrich-Forster, C. (2002). The circadian system of Drosophila melanogaster

and its light input pathways. Zoology (Jena) 105, 297–312.

Herzberg, C., Weidinger, L.A., Dorrbecker, B., Hubner, S., Stulke, J., and

Commichau, F.M. (2007). SPINE: a method for the rapid detection and analysis

of protein-protein interactions in vivo. Proteomics 7, 4032–4035.

Ito, H., Fukuda, Y., Murata, K., and Kimura, A. (1983). Transformation of intact

yeast cells treated with alkali cations. J. Bacteriol. 153, 163–168.

James, P., Halladay, J., and Craig, E.A. (1996). Genomic libraries and a host

strain designed for highly efficient two-hybrid selection in yeast. Genetics

144, 1425–1436.

Kaiserli, E., Sullivan, S., Jones, M.A., Feeney, K.A., and Christie, J.M. (2009).

Domain swapping to assess the mechanistic basis of Arabidopsis phototropin

1 receptor kinase activation and endocytosis by blue light. Plant Cell 21, 3226–

3244.

Cell 142, 762–772, September 3, 2010 ª2010 Elsevier Inc. 771

Page 140: Cell 100903

Kaldi, K., Gonzalez, B.H., and Brunner, M. (2006). Transcriptional regulation of

the Neurospora circadian clock gene wc-1 affects the phase of circadian

output. EMBO Rep. 7, 199–204.

Kasahara, M., Swartz, T.E., Olney, M.A., Onodera, A., Mochizuki, N., Fuku-

zawa, H., Asamizu, E., Tabata, S., Kanegae, H., Takano, M., et al. (2002).

Photochemical properties of the flavin mononucleotide-binding domains of

the phototropins from Arabidopsis, rice, and Chlamydomonas reinhardtii.

Plant Physiol. 129, 762–773.

Krauss, U., Minh, B.Q., Losi, A., Gartner, W., Eggert, T., von Haeseler, A., and

Jaeger, K.E. (2009). Distribution and phylogeny of light-oxygen-voltage-blue-

light-signaling proteins in the three kingdoms of life. J. Bacteriol. 191, 7234–

7242.

Liu, Y. (2003). Molecular mechanisms of entrainment in the Neurospora circa-

dian clock. J. Biol. Rhythms 18, 195–205.

Loros, J.J., Dunlap, J.C., Larrondo, L.F., Shi, M., Belden, W.J., Gooch, V.D.,

Chen, C.H., Baker, C.L., Mehra, A., Colot, H.V., et al. (2007). Circadian output,

input, and intracellular oscillators: insights into the circadian systems of single

cells. Cold Spring Harb. Symp. Quant. Biol. 72, 201–214.

McClung, C.R. (2008). Comes a time. Curr. Opin. Plant Biol. 11, 514–520.

McDonald, M.J., and Rosbash, M. (2001). Microarray analysis and organiza-

tion of circadian gene expression in Drosophila. Cell 107, 567–578.

Merrow, M., and Roenneberg, T. (2007). Circadian entrainment of Neurospora

crassa. Cold Spring Harb. Symp. Quant. Biol. 72, 279–285.

Merrow, M., Brunner, M., and Roenneberg, T. (1999). Assignment of circadian

function for the Neurospora clock gene frequency. Nature 399, 584–586.

Nayak, S.K., Jegla, T., and Panda, S. (2007). Role of a novel photopigment,

melanopsin, in behavioral adaptation to light. Cell. Mol. Life Sci. 64, 144–154.

Salomon, M., Christie, J.M., Knieb, E., Lempert, U., and Briggs, W.R. (2000).

Photochemical and mutational analysis of the FMN-binding domains of the

plant blue light receptor, phototropin. Biochemistry 39, 9401–9410.

Schafmeier, T., Haase, A., Kaldi, K., Scholz, J., Fuchs, M., and Brunner, M.

(2005). Transcriptional feedback of Neurospora circadian clock gene by

phosphorylation-dependent inactivation of its transcription factor. Cell 122,

235–246.

Schafmeier, T., Kaldi, K., Diernfellner, A., Mohr, C., and Brunner, M. (2006).

Phosphorylation-dependent maturation of Neurospora circadian clock

protein from a nuclear repressor toward a cytoplasmic activator. Genes Dev.

20, 297–306.

Schafmeier, T., Diernfellner, A., Schafer, A., Dintsis, O., Neiss, A., and Brunner,

M. (2008). Circadian activity and abundance rhythms of the Neurospora clock

transcription factor WCC associated with rapid nucleo-cytoplasmic shuttling.

Genes Dev. 22, 3397–3402.

Schwerdtfeger, C., and Linden, H. (2001). Blue light adaptation and desensiti-

zation of light signal transduction in Neurospora crassa. Mol. Microbiol. 39,

1080–1087.

Schwerdtfeger, C., and Linden, H. (2003). VIVID is a flavoprotein and serves as

a fungal blue light photoreceptor for photoadaptation. EMBO J. 22, 4846–

4855.

Siepka, S.M., Yoo, S.H., Park, J., Lee, C., and Takahashi, J.S. (2007). Genetics

and neurobiology of circadian clocks in mammals. Cold Spring Harb. Symp.

Quant. Biol. 72, 251–259.

Smith, K.M., Sancar, G., Dekhang, R., Sullivan, C.M., Li, S., Tag, A.G., Sancar,

C., Bredeweg, E.L., Priest, H.D., McCormick, R.F., et al. (2010). Transcription

factors in light and circadian clock signaling networks revealed by genome-

wide mapping of direct targets for Neurospora WHITE COLLAR COMPLEX.

Eukaryot. Cell, in press. Published online July 30, 2010. 10.1128/EC.00154-10.

Swartz, T.E., Corchnoy, S.B., Christie, J.M., Lewis, J.W., Szundi, I., Briggs,

W.R., and Bogomolni, R.A. (2001). The photocycle of a flavin-binding domain

of the blue light photoreceptor phototropin. J. Biol. Chem. 276, 36493–36500.

Tan, Y., Merrow, M., and Roenneberg, T. (2004). Photoperiodism in Neuros-

pora crassa. J. Biol. Rhythms 19, 135–143.

Vitalini, M.W., de Paula, R.M., Park, W.D., and Bell-Pedersen, D. (2006). The

rhythms of life: circadian output pathways in Neurospora. J. Biol. Rhythms

21, 432–444.

Woelfle, M.A., and Johnson, C.H. (2006). No promoter left behind: global circa-

dian gene expression in cyanobacteria. J. Biol. Rhythms 21, 419–431.

Zoltowski, B.D., and Crane, B.R. (2008). Light activation of the LOV protein

vivid generates a rapidly exchanging dimer. Biochemistry 47, 7012–7019.

Zoltowski, B.D., Schwerdtfeger, C., Widom, J., Loros, J.J., Bilwes, A.M.,

Dunlap, J.C., and Crane, B.R. (2007). Conformational switching in the fungal

light sensor Vivid. Science 316, 1054–1057.

Zoltowski, B.D., Vaccaro, B., and Crane, B.R. (2009). Mechanism-based

tuning of a LOV domain photoreceptor. Nat. Chem. Biol. 5, 827–834.

772 Cell 142, 762–772, September 3, 2010 ª2010 Elsevier Inc.

Page 141: Cell 100903

Cell Flow Reorients the Axisof Planar Polarity in theWing Epithelium of DrosophilaBenoıt Aigouy,1 Reza Farhadifar,2 Douglas B. Staple,2 Andreas Sagner,1 Jens-Christian Roper,1 Frank Julicher,2,*and Suzanne Eaton1,*1Max Planck Institute of Molecular Cell Biology and Genetics, Pfotenhauerstrasse 108, Dresden 01307, Germany2Max Planck Institute for the Physics of Complex Systems, Noethnitzer Strasse 38, Dresden 01187, Germany

*Correspondence: [email protected] (F.J.), [email protected] (S.E.)DOI 10.1016/j.cell.2010.07.042

SUMMARY

Planar cell polarity (PCP) proteins form polarizedcortical domains that govern polarity of externalstructures such as hairs and cilia in both vertebrateand invertebrate epithelia. The mechanisms thatglobally orient planar polarity are not understood,and are investigated here in the Drosophila wingusing a combination of experiment and theory.Planar polarity arises during growth and PCPdomains are initially oriented toward the well-charac-terized organizer regions that control growth andpatterning. At pupal stages, the wing hinge con-tracts, subjecting wing-blade epithelial cells toanisotropic tension in the proximal-distal axis. Thisresults in precise patterns of oriented cell elongation,cell rearrangement and cell division that elongate theblade proximo-distally and realign planar polaritywith the proximal-distal axis. Mutation of the atypicalCadherin Dachsous perturbs the global polaritypattern by altering epithelial dynamics. This mecha-nism utilizes the cellular movements that sculpttissues to align planar polarity with tissue shape.

INTRODUCTION

The PCP pathway coordinates tissue planar polarity in vertebrate

and invertebrate epithelia (Simons and Mlodzik, 2008; Vladar

et al., 2009). PCP proteins localize to apical junctions and form

intracellularly polarized domains with different compositions.

In the Drosophila wing, these domains link proximal and distal

cell boundaries and orient wing hair outgrowth distally. Distal

complexes containing Flamingo (Fmi), a.k.a. Starry night (Stan),

Frizzled (Fz), Dishevelled (Dsh) and Diego (Dgo) interact across

cell boundaries with proximal complexes containing Fmi, Stra-

bismus (Stbm), a.k.a. Van Gogh (Vang) and Prickle (Pk). In the

absence of any single PCP protein, the others do not polarize

well and hair outgrowth is misoriented (Seifert and Mlodzik,

2007; Strutt and Strutt, 2005; Uemura and Shimada, 2003).

Feedback loops arising from interactions of PCP proteins within

and between cells may be sufficient for the local alignment of

PCP complexes between small groups of cells (Amonlirdviman

et al., 2005; Tree et al., 2002a; Tree et al., 2002b). Proposed

cellular mechanisms underlying local feedback loops include

preferential interactions between Fmi/Fz and Fmi/Stbm com-

plexes across cell boundaries, and the decreased susceptibility

of these complexes to endocytosis (Chen et al., 2008; Strutt

and Strutt, 2008; Wu and Mlodzik, 2008). Also, alignment of

microtubules may bias the delivery of Fz to the distal side of

the cell (Shimada et al., 2006).

While these mechanisms can polarize PCP domains intracellu-

larly and align them locally, less is understood about mecha-

nisms that specify global alignment of PCP domains with the

proximal-distal (PD) axis of the wing. The atypical Cadherins

Fat (Ft) and Dachsous (Ds), and a regulatory golgi kinase Four-

jointed (Fj) influence global orientation of PCP domains by

a mechanism that is unclear (Adler et al., 1998; Ma et al., 2003;

Matakatsu and Blair, 2004; Strutt and Strutt, 2002; Zeidler

et al., 2000). The Fat ligand Ds is expressed highly in proximal

regions that give rise to the hinge, but at lower levels in the

more distal wing blade, during both larval and pupal stages

(Cho and Irvine, 2004; Clark et al., 1995; Ma et al., 2003; Mata-

katsu and Blair, 2004; Strutt and Strutt, 2002). Fj, which regulates

their activity (Ishikawa et al., 2008) is expressed with an opposite

pattern (Villano and Katz, 1995; Zeidler et al., 2000). These

opposing patterns have been proposed to cause intracellular

asymmetries that directly bias accumulation of core PCP

proteins in wing cells (Ma et al., 2008, 2003; Strutt and Strutt,

2002; Tree et al., 2002a). However this is not the case in the

abdomen where Ft/Ds and PCP proteins act independently to

control trichome polarity (Casal et al., 2006). Even in the wing,

Ft and Ds do not act at the time that PCP domains actually align

with the PD axis, but much earlier (during larval or prepupal

stages), to influence the global PCP pattern (Matakatsu and

Blair, 2004; Strutt and Strutt, 2002). Ft and Ds regulate both

the amount and orientation of proliferation in the larval wing

epithelium (Baena-Lopez et al., 2005; Bryant et al., 1988; Clark

et al., 1995; Garoia et al., 2005, 2000). It is not yet clear whether

this activity is relevant to the PD alignment of PCP proteins

during pupal stages.

Cell 142, 773–786, September 3, 2010 ª2010 Elsevier Inc. 773

Page 142: Cell 100903

Our previous experiments suggested that PCP domain

polarity does not develop de novo during pupal stages; PCP

domains are polarized in both larval wing discs and prepupal

wings, but polarity is not aligned with the PD axis. At early pupal

stages, the PCP axis is oriented at an angle to the PD axis

(Classen et al., 2005). Later, by the time wing hairs form, PCP

domains are oriented along the PD axis of the wing. Alignment

with the PD axis occurs during a phase in which wing epithelial

cells are exchanging their cell contacts (Classen et al., 2005).

One consequence of remodelling is an increase in hexagonal

order of the wing epithelium. Theoretical analysis suggested

that different types of fluctuations could guide epithelial cells

initially disordered by proliferation toward a hexagonal state

(Farhadifar et al., 2007). Here, we investigate the relationship

between these cell rearrangements and the temporal evolution

of PCP orientation patterns. We quantitatively analyze time-

lapses of pupal wing epithelia and combine these data with theo-

retical analysis to extract key mechanisms that couple cell rear-

rangements to PCP reorientation.

RESULTS

Planar Polarity Points Initially toward the Wing Marginand Reorients DistallyTo investigate how PCP order evolves during pupal develop-

ment, we imaged wings expressing Stbm:YFP between 15 and

34 hr after puparium formation (hAPF). We developed a method

to quantify planar polarity based on the cell perimeter intensity of

Stbm:YFP. This method quantifies the axis and magnitude of

polarity, but not its vector orientation (Supplemental Theoretical

Procedures, 1.1 and Figures S1A–S1E available online). We

started by quantifying the axis of PCP at 15 hAPF (Figure 1A);

we also determined the polarity vectors in these wings by

creating Fz:YFP-expressing clones (Figures S1F–S1H). As

A B

C D E

20 µm

F

PCP order elongation16 hAPF 20h50 APF 32h20 APF

30 hAPF15 hAPF50 µm

Tim

e [h

AP

F]

Figure 1. Time Evolution of Planar Polarity

(A and B) The magnitude and axis of average nematic order (yellow bars) overlaid on Stbm:YFP-expressing wings, at 15 hAPF (A) and 30 hAPF (B). Red dots

indicate Fz:YFP domain polarity. Green ellipses indicate the anterior crossvein.

(C–E) Stbm:YFP at the indicated times between veins L1 and L3. Yellow bars show the PCP nematic for each cell.

(F) Magnitude (arbitrary units) and axis of average nematic order (left) and cell elongation (right) as a function of time (indicated on the y axis and by color) derived

from the wing shown in (C–E).

In all figures, anterior is top and proximal is left.

See also Figure S1.

774 Cell 142, 773–786, September 3, 2010 ª2010 Elsevier Inc.

Page 143: Cell 100903

suggested previously (Classen et al., 2005), PCP at early stages

is correlated over long distances (Figure S1M) and is oriented

such that Fz:YFP domains face the wing margin (Figures 1A

and 1C and Figures S1F–S1I). Starting at about 18 hAPF, the

magnitude of the average alignment of groups of cells (local

nematic order, see Supplemental Theoretical Procedures, 1.2)

begins to decrease (Figures 1D and 1F and Figure S1J), and

the average axis of alignment begins to rotate toward the PD

axis (Figure 1F and Figures S1J and S1K). Reduced alignment

reflects both reduced polarity of individual cells and reduced

polarity correlation over distance (Figures S1L and S1M). After

reaching a minimum about 20 hAPF, the magnitude of average

nematic order increases; by 26 hAPF nematic order is again

maximal and aligned parallel to the PD axis (Figures 1B, 1E,

and 1F, Figures S1J and S1K, and Movie S1). Where we have

determined the direction of polarity, we quantify average polar

order by a measure ranging from zero to one (one corresponding

to perfect alignment) (Supplemental Theoretical Procedures,

1.3). At 15 hAPF, polar order averaged over the whole wing blade

is only 0.60, because polarity is oriented differently in anterior

and posterior wing regions (Figure 1A). By 32 hAPF average polar

order reaches 0.98 (Figure 1B). Thus, PCP domains do not

develop polarity de novo during pupal stages, but rather reorient

pre-existing polarity.

Polarity reorientation is reflected in the influence of fz mutant

clones on the polarity of adjacent wild-type tissue. At 15 hAPF,

fz clones perturb polarity of tissue lying between the clone and

the wing margin (Figures S1N and S1O) consistent with the early

polarity pattern. After reorientation, polarity is disturbed in wild-

type tissue lying distal to the clone (Vinson and Adler, 1987).

We noted that reorientation of PCP correlated with a transient

cell shape deformation that elongates cells in the PD axis

(Figure 1F). Quantifying elongation (Supplemental Theoretical

Procedures, 2) reveals that average polarity shrinks in magnitude

and reorients toward the PD axis as cells elongate. The magni-

tude of average polarity along the PD axis then increases as

cell elongation decreases (Figure 1F). The precise temporal

correlation of these events suggests that they are controlled by

the same underlying mechanisms.

Dramatic Morphogenetic Movements Reshape the Wingduring Pupal DevelopmentTo explore global changes occurring in the wing as cells change

shape and PCP reorients, we imaged ECadherin:GFP-express-

ing wings at low magnification from 15 to 32 hAPF. Dramatic

morphogenetic movements reshape the wing during this time

(Figures 2A–2C and Movie S2). At 15–16 hAPF, the hinge and

blade regions are roughly equal in size and the shape of the hinge

is not yet recognizable (Figure 2A). Between 15 and 24 hAPF, the

hinge undergoes patterned contractions that halve its area

and shape the allula and costa (Figures 2A–2C, Movie S2, and

Movie S3A). While wing-blade area remains constant during

this time, wing shape changes – elongating in the PD axis and

narrowing in the AP axis (Figures 2A–2C and Movie S2).

To quantify tissue movements underlying this shape change,

we tracked different regions and calculated local velocity vectors

(Supplemental Theoretical Procedures, 3). As the hinge con-

tracts, wing-blade cells flow proximally toward the hinge at

different velocities (Figures 2D–2F, Movie S3B). Between 15

and 18 hAPF, velocities have both inward and proximal compo-

nents (Figure 2D). Later, flow is proximally oriented and fastest in

the middle of the wing blade (Figure 2E). These inhomogeneous

velocities define local compression, shear and rotation as cells

move with respect to each other (Supplemental Theoretical

Procedures, 4). The shear axis in a particular region can be

thought of as the local axis of tissue deformation. In the

wing, local shear axes are oriented in a fan-like pattern that is

roughly symmetric about the third wing vein (Figures 2G–2I,

Figures S2J and S2K, and Movie S3C). Although the wing as

a whole does not rotate, local rotation does occur and is mainly

clockwise anterior to the third wing vein and anti-clockwise

posterior to it (Figures 2J–2L, Figure S2L, and Movie S3D). These

patterns of shear and rotation lengthen the wing blade in the

PD axis and narrow it in the anterior-posterior (AP) axis. Strik-

ingly, the observed antisymmetric rotation pattern would shift

early margin-oriented polarity toward the distal direction.

To investigate whether the tissue flows that reshape the wing

blade were driven by hinge contraction, we completely severed

the hinge from the blade before contraction occurred (Figure 3A).

After severing, wing-blade tissue flows distally rather than prox-

imally (Figure 3C and Movie S4A). Shear is reduced and is mainly

perpendicular to the PD axis (Figure 3D and Movie S4B), rather

than forming a fan-like pattern. Rotation is also reduced, and

its orientation is reversed compared to unperturbed wings

(compare Figure 3E with Figure 2K, and Movie S4C with Movie

S3D). As a result, the wing blade undergoes opposite shape

changes to those that normally occur. Nevertheless, wing hairs

(not shown) and longitudinal veins (Figure 3B) form normally.

These data suggest that hinge contraction exerts mechanical

stresses that contribute to the observed cell flow patterns that

change the shape of the wing blade.

Cell Boundary Tension Is Elevated along the PD Axisduring Hinge ContractionTo ask how cell boundary tension changes during hinge contrac-

tion, we used a pulsed UV laser to sever cell boundaries lying at

different angles to the PD axis, and monitored the movements

of adjacent vertices. The initial velocity with which vertices move

apart is a measure of cell boundary tension. Before hinge contrac-

tion, tension is similar on all cell boundaries. When the hinge

contracts, tension increases specifically on those cell boundaries

lying at angles close to that of the PD axis (Figures S3A and S3B).

This suggests that hinge contraction exerts forces on the blade

that direct cellular flows via an anisotropic stress profile.

Tissue Flow in the Wing Blade Results from OrientedCell Divisions, Cell Elongation, and Neighbor ExchangesTo investigate the cellular events underlying tissue flow,

shear, and rotation, we tracked groups of cells between veins 3

and 4, starting 15 hAPF. This revealed that local tissue shape

change occurs in two distinct phases. The first is dominated by

oriented cell division and PD cell elongation, and the second by

oriented cell rearrangements (Figure 4 and Movie S5A).

During phase I (about 15–24 hAPF), cell elongation increases

in the PD axis (Figures 4A–4C and 4G). Cell boundary loss

exceeds new boundary formation (Figure 4G), as confirmed by

Cell 142, 773–786, September 3, 2010 ª2010 Elsevier Inc. 775

Page 144: Cell 100903

the decrease in average neighbor number (Figure 4F). These new

cell boundaries have no preferred direction during phase I

(Figure 4I). Cells undergo between one and two rounds of

oriented cell division that reduce cell size. Between veins 3

and 4, these divisions are on average oriented 20� to the PD

axis (Figure 4H). As phase I ends, cell elongation peaks (Fig-

ure 4G) and cell division stops (Figure 4F). At the end of this

phase, tracked tissue patches have deformed in a way that

can be accounted for by the combination of cell elongation

and oriented cell division. During phase II (about 24–32 hAPF),

cells assemble new contacts parallel to the PD axis (Figures

4G and 4I). These oriented neighbor exchanges relieve cell elon-

gation (Figure 4G) and stabilize the tissue shape change caused

by cell elongation during phase I (Figures 4C–4E). They also

increase average neighbor number and the fraction of hexago-

nally packed cells (Figure 4F).

Severing the Hinge from the Wing Blade Blocks PDElongation and Misorients Cell Divisions and NeighborExchangesTo investigate the cellular basis of altered shear and rotation in

severed wing blades, we quantified cell elongation and the

orientation of neighbor exchanges and cell divisions after

severing from the hinge (Figures S3C–S3G and Movie S5B).

24h00' 33h30'

24h00' 31h40'

A B C

D E F

G H I

*

80 µm

80 µm

50 µm/h

0.3 h-1

J K L 0.1 h 0.05 h 0.01 h

clockwiseanti-clockwise

-1 -1 -1

*

ubi-Ecad:GFP

ubi-Ecad:GFP

L3L3L3

15h30'

Figure 2. Tissue Shape Changes during Hinge Contraction(A–C) Wing images from a movie spanning roughly 15-33 hAPF. The hinge (blue) contracts, the wing blade (red) elongates. Allula (arrowhead) and costa (asterisk)

are indicated. The area of the wing hinge is 1023 103 mm2 in (A) and 503 103 mm2 in (C). The area of the wing blade is 1213 103 mm2 in (A) and 1273 103mm2 in (C).

(D–L) Locally averaged features of the cellular flow field at indicated times.

(D–F) Flow velocity vectors. Speed corresponds to arrow length as indicated.

(G–I) Rate and axis of pure shear. Bar length indicates shear rate.

(J–L) Local rotation rate (indicated by circle size) in radians per hour.

L3 indicates the third longitudinal vein.

Results shown are representative of five separate experiments.

See also Figure S2.

776 Cell 142, 773–786, September 3, 2010 ª2010 Elsevier Inc.

Page 145: Cell 100903

Cell divisions are strongly oriented in unperturbed wings. When

the hinge is severed from the blade, cell division continues

(Figure S3H), but the angle of cell division is more disperse

(Figure S3J). In addition, these cells do not elongate further along

the PD axis (Figure S3I), becoming instead more isotropic. While

neighbor exchanges still occur, their axis is shifted with respect

to unwounded wings, lying at ±45 degrees with respect to the PD

axis (compare Figure S3K to Figure 4I). Finally, wing severing

reduces the final number of hexagonally packed cells

(62.44% ± 3.5% in wounded wings [n = 3] versus 77% ± 4.8%

in unperturbed wings [n = 5]). Thus, severing the wing dramati-

cally changes the observed patterns of cell elongation, cell divi-

sion, and rearrangements. This suggests that anisotropic stress

caused by hinge contraction has a key role in guiding the tissue

flow that reshapes the wing blade. It also suggests that cell flow

increases hexagonal packing, similar to the annealing effect of

cell boundary fluctuations (Farhadifar et al., 2007).

Severing the Hinge Alters the Global Patternof Planar PolarityReorientation of planar polarity in the intact wing occurs at the

same time that the wing blade is reshaped by hinge contraction,

suggesting that these events depend on one another. The

observed pattern of clockwise and anti-clockwise tissue rotation

tends to reorient early margin-directed polarity toward the PD

axis. In addition, the shear pattern presages the final orientation

of planar polarity (Figure 2H). To ask whether tissue movements

induced by hinge contraction guided polarity reorientation, we

examined the distribution of Stbm:YFP in wing blades that had

been severed from the hinge before contraction. In severed

wings, the global PCP pattern differs from that of the intact

wing (compare Figure 1B with Figure 3F), suggesting that polarity

15h30' 32h00'

22h30'

A B

D

80 µm

80 µm

50 µm/h 0.3 h

K0.1 h 0.05 h 0.01 h

clockwiseanti-clockwise

-1 -1 -1

Stbm:YFP

E

C -1

L3

F

Stbm:YFP32h00'

22h30'

22h30'

Figure 3. Tissue Shape Changes in Severed Wings

(A and B) Images taken at the indicated times from a movie

spanning 15–32 hAPF. The hinge was severed from the blade

along the line indicated.

(C–E) Locally averaged features of the cellular flow field at

22h300 APF.

(C) Flow velocity vectors. Speed corresponds to arrow length

as indicated.

(D) Rate and axis of pure shear. Shear rate corresponds to bar

length as indicated.

(E) Local rotation rate (indicated by circle size) in radians

per hour.

(F) Magnitude and axis of average nematic order (yellow bars)

overlaid on a severed Stbm:YFP-expressing wing at 32 hAPF.

Green ellipse = anterior crossvein. L3, third longitudinal vein.

Results shown are representative of four separate experi-

ments.

is altered by the changed patterns of cell rearrange-

ments, cell division, and cell elongation. This

implies that cell flows guide reorientation of PCP.

Oriented Cell Boundary Assembly Increasesthe Visibility of Planar PolarityWhile the majority of polarity reorientation occurs

during phase I, average PCP order increases in

magnitude during phase II, as cells extend new contacts mostly

parallel to the PD axis and develop hexagonal packing geometry

(Figure 1F and Figures 4F and 4I). Could the increase in polar

order be related to these processes? To investigate this, we

imaged PCP protein localization as new cell boundaries formed

(Figures 5A–5F and Movie S5C). We noted that Stbm:YFP had

a clustered rather than a uniform distribution on cell boundaries

throughout pupal development. Even at early stages, these clus-

ters contain both Stbm:YFP and Fz:CFP, suggesting that they

constitute PCP protein complexes (Figures S4A–S4C). Consis-

tent with this, Stbm:YFP clustering is reduced in fz mutant tissue

(Figures S4D and S4E). Individual clusters are persistent and can

be tracked for several hours. However, proteins within them turn

over more rapidly (Figures S4G–S4I). Clusters do not form rapidly

on new cell boundaries derived either from cell division or

neighbor exchange (Figure 5 and Movie S5C). During phase II,

new cell boundaries, expanding parallel to the PD axis within

minutes, typically remain devoid of these clusters for hours

(Figures 5A–5H). Consequently, Stbm:YFP clusters become

restricted mainly to older cell interfaces on the proximal and

distal sides of the cell. Thus, oriented cell boundary expansion

and the concomitant increase in hexagonal packing order ensure

that PCP clusters become well separated within cells and better

aligned between cells; even though PCP has already reoriented

during phase I, oriented cell boundary expansion during phase II

tends to increase the average nematic order (compare

Figure S4F with Figures S1A—S1E).

A Theoretical Analysis of Polarity ReorientationWhile local tissue rotation clearly implies local rotation of the

polarity axis, it is less simple to understand how tissue shear

affects the axis of planar polarity. Tissue shear is caused by

Cell 142, 773–786, September 3, 2010 ª2010 Elsevier Inc. 777

Page 146: Cell 100903

oriented cell divisions, cell elongation, and oriented cell rear-

rangements. Because microtubules have been shown to align

with the axis of cell elongation (Cortes et al., 2006; Daga and

Nurse, 2008; Haase and Lew, 2007; Strauss et al., 2006), cell

shape may directly influence the distribution of PCP within the

cell. Indeed, microtubules in wing epithelial cells align with the

PD axis at this time (Shimada et al., 2006). We turned to theoret-

ical analysis to explore how each of these processes might affect

the axis of planar polarity. Previously, we used a vertex model to

study the effects of proliferation on cell packing geometry

(Farhadifar et al., 2007). Here, we add to this model a simplified

description of the dynamics of PCP order.

B

19h30 APF

C

24 hAPF

A

15 hAPF 20 µm

phase I phase II

28h30 APF

D

F G

0

90

180

270

0

90

180

270

E

31h30 APF

cell number (fraction of maximum)

average neighbor number

fraction hexagons

PD cell elongationnew cell contacts (fraction of maximum)cell contacts that will disappear (fraction of maximum)

H

orientation of division axisphase I phase II

orientation of new boundary assemblyphase I phase II

I

P DP DP DP DP D

0

90

180

270

0

90

180

270

0

0.2

0.4

0.6

0.8

1

14 16 18 20 22 24 26 28 30 32

0.08

0.12

0.16

0.2

0.24

0.28

0.32

fract

ion

PD

elo

ngat

ion

time (in hAPF)

0

0.2

0.4

0.6

0.8

1

14 16 18 20 22 24 26 28 30 325.4

5.5

5.6

5.7

5.8

5.9

6

fract

ion

aver

age

neig

hbor

num

ber

time (in hAPF)

phase IIphase I

phase IIphase I

persistent disappearing new boundary from neighbor exchange new boundary from cell division

Figure 4. Cell Elongation, Division, and Rearrangement

(A–E) A group of cells anterior to the posterior crossvein was tracked between 15h and 31h300 APF. Cell boundaries are color-coded depending on their fate:

red boundaries persist indefinitely, yellow boundaries disappear, green boundaries form during the movie as a result of neighbor exchange, and blue boundaries

form between daughter cells upon division.

(F and G) Quantification of cellular changes in the patch of tissue tracked in (A–E). Fraction of maximal cell number (dark blue), average neighbor number (light

blue, averaged over five frames), fraction of hexagonal cells (brown, averaged over five frames), PD cell elongation (magenta, average maximum value = 0.388 ±

0.053, n = 5), fraction of boundaries that will disappear (yellow), and the fraction of newly formed boundaries from neighbor exchange (green).

(H and I) Angular distribution of cell divisions (blue) (H) and new cell boundaries (green) (I) at the end of phase I and phase II. Yellow and magenta bars indicate

average angle of nematic order of cell division (yellow) and new cell boundary formation (magenta) (see Supplemental Theoretical Procedures). Bar length indi-

cates the degree of focus (magnitude of average nematic order, ranging from 0 to 1, see Supplemental Theoretical Procedures, 1.2). The outer circles in (H) and (I)

correspond to magnitude = 0.4. Average magnitude of phase I cell division order is 0.44+/�0.12, n = 4. Average magnitude of new boundary formation order in

phase II = 0.35+/�0.04, n = 4.

Data are presented as mean ± SD.

See also Figure S3.

778 Cell 142, 773–786, September 3, 2010 ª2010 Elsevier Inc.

Page 147: Cell 100903

A Physical Description of PCP in the Vertex ModelCell shape, as defined by the network of adherens junctions, can

be described as force-balanced configurations in a vertex

model, in which an arrangement of cells is represented by a

set of cell bonds. This model takes into account cell mechanics,

cell adhesion and cell division (Farhadifar et al., 2007). Our

description of PCP order is based on attractive interactions

between proximal and distal PCP domains across cell bound-

aries and repulsive interactions within cells. This differs from

previous approaches (Amonlirdviman et al., 2005; Le Garrec

et al., 2006), in that we do not attempt to describe all details of

interactions between PCP proteins, but rather consider effective

interactions between proximal and distal domains. Each cell

bond i, separating cells a and b, is assigned two variables, saiand s

bi , which describe the level of PCP proteins on either side

of this bond (Figure 6A). The variables sai can take continuously

varying values; sai = 1 corresponds to a high level of proximal

proteins (blue), while sai = � 1 corresponds to a high level of

distal proteins (red). We introduce a potential function for interac-

tions of PCP domains:

E = J1

X

i

sai s

bi � J2

X

fi;jgsa

i saj � J3

X

a

3a$Qa; (1)

where J1 and J2 are PCP interaction parameters and J3

describes coupling of PCP to cell elongation (for J3 = 0, PCP

does not couple to cell elongation). For J1>0 the potential favors

interactions between proximal and distal proteins across cell

bonds. For J2>0, accumulation of proximal and distal proteins

on neighboring bonds within a given cell is disfavored. For

J3>0, alignment of PCP with the long axis of the cell is favored:

this is motivated by the idea that aligned microtubules bias the

transport of PCP proteins (Shimada et al., 2006). The first sum

is over all bonds i, the second sum is over all pairs i; jgf of adja-

cent bonds in the network, and the third sum is over all cells a.

The cell shape and PCP nematic order are described by the

tensors 3a and Qa, respectively (Supplemental Theoretical

Procedures, 5). We introduce a time evolution of PCP level on

cell bonds: dsai =dt = � gvE=vsai , where t is time and g is a kinetic

coefficient that determines the PCP relaxation time t = 1=ðgJ1Þ.We solve these dynamic equations numerically, imposing two

constraints: (1) in each cell, amounts of proximal and distal

proteins are equal and (2) these amounts do not change with

time (Supplemental Theoretical Procedures, 5).

Global Polar Order Generated by Network GrowthTo examine how different types of shear reorient polarity, we

must start from a cellular configuration in the vertex model with

21 hAPF

25 hAPF

29 hAPF

Stbm:YFP

8 µ

0

90

180

270

B

D

A

C

E F

J K

L M

N O

I

P Q

R S

T U

G H90

270

0 180

1

1

1

2

2

2

3

3

3

4

4

4

5

5

5

8

7

6

6

6

7b7a

7b

8b

8a

8a8b

9

9

9

10

10

10

11

11

11

12

12

12

13

13

13

14

14

14

15

15

15

16

16

16

17

17

17

18

18

18

19

19

19

20

20

20

m

Figure 5. Cell Rearrangements Increase the Visibility of PCP

(A–F) Images from a 21–29 hAPF movie of a Stbm:YFP-expressing wing. In (A),

(C), and (E), cell identity is indicated by numbers and cell boundaries are

colored according to their fate as in Figure 4. (B), (D), and (F) show PCP

nematics for each cell (yellow).

(G and H) Angular distribution of newly formed (green, [G]) and permanent

boundaries (red, [H]).

(I) Each kymograph line depicts the linearized perimeter of a selected cell at

time t.

(J–O) Perimeter intensity kymographs for two cells of the movie (frame rate =

60, movie length = 8 hr) shown in (A–F). In (L) and (M), each boundary is

identified by a random color. Colors in (N) and (O) indicate boundary fate

(see Figure 4). Permanent boundaries (red in [N] and [O]) carry PCP clusters

that persist for 8 hr. New boundaries (green in [N] and [O]) remain devoid of

Stbm:YFP clusters.

(P-U) Kymographs of a nondividing cell (P), (R), and (T) or a dividing cell (Q), (S),

and (U) between 20-21 hAPF (300 frame rate). In (R) and (S), each boundary is

identified by a random color. Colors in (T) and (U) indicate boundary fate (see

Figure 4). After cell division (Q), (S), and (U) depicts the joint perimeter of the

two daughter cells and the boundary created by cell division (blue in [U])

separately.

See also Figure S4.

Cell 142, 773–786, September 3, 2010 ª2010 Elsevier Inc. 779

Page 148: Cell 100903

globally aligned PCP variables. To generate such a configuration

without an external bias is nontrivial (Burak and Shraiman, 2009).

Surprisingly, a simple and general way to generate large

networks with long-range polarity is to start from a small group

of cells with an initially random PCP configuration (random

values of sai , Supplemental Theoretical Procedures, 6). This

network is then expanded by simulating stochastic cell divisions

at a rate kd. Simultaneously, the dynamic equations for the PCP

variables are solved. Interactions of PCP variables generate

local order, which aligns over the whole network when it is

still small. Order is maintained during growth and leads to glob-

ally aligned polarity in the resulting large networks if the PCP

relaxation time t is sufficiently fast (kdt<1, Figures 6B, Fig-

α

β

σi

α

σi

β

A

E

C

G

B

F

D

H

Figure 6. Theoretical Analysis of Shear and

Rotation on the Reorientation of Planar

Polarity

(A) Schematic representation of PCP in the vertex

model. The variables, sai and sbi , represent the level

and type of PCP domains on bond i shared by cells

a and b.

(B) Cell packing with PCP order generated by

simulated proliferation in the vertex model with

PCP dynamics, starting from 36 randomly polar-

ized cells with kdt = 0.01, J2/J1 = 0.5 and J3 = 0.

Arrows in (B)–(D) show the direction of polarity.

(C and D) Reorientation of polar order by shear due

to external forces (C) or oriented cell division (D).

Starting from the network (B), and simulating shear

along the horizontal axis, leads to the networks

shown. Parameter values for (C) are kdt = 0.01

and J3/J1 = 0.05. For (D), kdt = 0.01, J3 = 0 and

�3 rounds of cell division were simulated. The

shear generated corresponds to an aspect ratio

of 4 in (C) and 1.5 in (D).

(E and F) Calculations of polarity reorientation

based on Equation 2 and measured patterns of

shear and rotation. Starting with the observed

early polarity (E), the time evolution described by

Equation 2 with n= � 3 generates a final PD

pattern (F). The wing is the same as that shown

in Figures 2D–2L.

(G and H) same procedure as in (E and F) applied to

the wounded wing shown in Figure 3 with n= 2. We

start with an initial pattern mimicking the early

polarity (G).

See also Figure S5.

ure S5A, and Movie S6A). This is consis-

tent with the observation of global PCP

patterns in the growing wing disc

(Classen et al., 2005), and suggests that

these patterns arise early during

development.

Shear Reorients the Polarity Axis inthe Vertex ModelStarting with a globally polarized

network, we used the vertex model to

examine how different types of shear

could influence the average polarity

orientation of groups of cells. We first studied the effects of

shear caused by cell elongation and oriented cell rearrange-

ments. In our simulations, we induced pure shear at a rate ks

by forcing the network to elongate along one axis at constant

total area (Supplemental Theoretical Procedures, 6). We

oriented the initial axis of average polarity at an angle of 45

degrees to this shear axis (Figure 6B). As the network is

deformed, cells elongate and undergo T1 transitions that are

oriented along the stretch axis. We find that shear reorients

polarity either parallel or perpendicular to the shear axis, de-

pending on the value of kst and on the strength J3 of the

coupling of cell shape to PCP distribution (Figures 6C, Figures

S5B–S5D, and Movies S6B—S6C). Reorientation occurs during

780 Cell 142, 773–786, September 3, 2010 ª2010 Elsevier Inc.

Page 149: Cell 100903

a characteristic time 1=j2nksj, where n is a dimensionless coef-

ficient. We define n to be negative if polarity aligns with the

shear axis, and positive if it aligns perpendicular to this axis.

Positive values of J3 promote the alignment of the polarity

vector with the shear axis, as do faster relaxation times t of

PCP (Figures S5B–S5D). Network configurations generated by

this process are irregular and cells are stretched. We relaxed

these networks by introducing fluctuations of cell bond tension

(Supplemental Theoretical Procedures, 6). This reduced cell

stretch and increased both the fraction of hexagons and the

magnitude of average polar order, similar to what is observed

during phase II in vivo (Figures 4F and 4G, and Figure S5E).

We next studied whether shear generated by oriented cell divi-

sions reorients polarity. We imposed cell divisions in the vertex

model at rate kd oriented at an angle of about 45 degrees to

the initial polarity axis (Supplemental Theoretical Procedures,

6). Oriented cell division generates shear parallel to the cell divi-

sion axis (defined as the line connecting the centers of the two

daughter cells) at a rate ks =mkd, where m depends on network

mechanics. We find again that polarity is reoriented either

parallel or perpendicular to the shear axis and that the dynamics

of reorientation can be characterized by the coefficient n defined

above (Figure 6D, Figure S5F, and Movie S6D). Again, faster

relaxation times t of PCP promote alignment of the polarity

vector with the shear axis (Figure S5F).

An Effective Description of Large-Scale PolarityReorientation by FlowThe reorientation of average polarity by shear flow that we

observe in the vertex model is well known in complex fluids

such as liquid crystals. In these systems, polarity reorientation

is determined by local rotation and shear rates of the flow field

and can be described by the following equation (de Gennes

and Prost, 1993; Joanny et al., 2007)

vq

vt= nkssin2q+u: (2)

Here, u is the local rotation rate, and ks is the rate of pure shear

(Supplemental Theoretical Procedures, 7). The orientation of

polarity is described by the angle q, relative to the axis of pure

shear. The coefficient n describes how pure shear influences

reorientation of polarity. In the absence of rotation, u= 0, pure

shear aligns polarity either with the shear axis ðq= 0Þ or perpen-

dicular to it ðq=p=2Þ depending on the sign of n. In liquid crystals,

the value of n depends on molecular geometries and intermolec-

ular interactions (de Gennes and Prost, 1993). In the absence of

pure shear, ks = 0, local tissue rotations reorient polarity as

vq=vt =u.

Equation 2 for u= 0 describes the main features of polarity

reorientation by pure shear in the vertex model described above.

The coefficient n that characterizes how shear affects reorienta-

tion in the vertex model is the same as the parameter n in Equa-

tion 2. We can thus use Equation 2 to determine quantitatively

how the observed patterns of rotation and shear reorient the

global pattern of planar polarity in the wing.

Starting with the observed polarity pattern at 15 hAPF, we

locally applied Equation 2, with n being the only free parameter,

to calculate the corresponding time-dependent reorientation of

polar order. Local rates of rotation and shear were obtained

from the experimentally determined flow field between 15 and

32 hAPF. To ask to what extent local rotation alone could repro-

duce the observed polarity reorientation, we first considered

n= 0, i.e., shear does not influence reorientation. In this case

polarity realigns toward the PD axis, however, the final average

polar order achieved is only 0.76 (77% of the experimental value

at 32 hAPF). Thus, local rotation can account for only part of the

observed reorientation of the polarity axis.

If n is nonzero, shear also influences reorientation. Repeating

this calculation for different values of n shows that the final

average polar order is largest for nz� 3, reaching a value of

0.90 (92% of the experimental value, Figures S5G–S5J). For n

within the range �10<n< � 1 the polarity field reorients robustly

to reach an average polar order of at least 0.85 (Figure S5K,

6E,F and Movie S6E), tolerating small deviations from the exper-

imentally determined pattern of initial conditions (Figure S5K).

Note that the effects of local shear on polarity orientation depend

on PCP domain interactions and dynamics, as captured in the

vertex model. In the calculations using Equation 2, these effects

are represented via the value of the parameter n. However, in

writing Equation 2 we choose to neglect interactions of average

polar order between neighboring patches of tissue to highlight

the role of shear and rotation. We expect such additional interac-

tions could account for the remaining deviation between the

solution to Equation 2 and the observed final polarity. Note

also that positive values of n produce a final polarity pattern

with lower polar order than rotation alone (n= 0) (Figure S5K).

This analysis suggests that shear caused by the combination

of oriented cell divisions, cell elongation and cell rearrangement

plays an important role in aligning polarity parallel to the shear

axis by a process that is characterized by a negative value of n.

The close agreement between the polarity pattern generated

by Equation 2 (Figure 6F) and experimental observations

(Figure 1B) provides strong evidence that tissue remodelling is

responsible for reorienting the axis of planar polarity.

To ask whether Equation 2 could describe the polarity defects

observed when the hinge and wing blade are severed, we used it

to calculate polarity reorientation given the shear and rotation

patterns measured for these wings. For wounded wings, we

find that Equation 2 describes the main features of polarity reor-

ientation provided that n is positive (Figures 6G and 6H, Figures

S5L and S5N, and Movie S6F). This implies that, in wounded

wings, shear tends to align polarity perpendicular to the shear

axis, rather than parallel to it, as it does in unperturbed wings.

Why should shear affect polarity differently in wounded wings?

In simulations of the vertex model, we observed that shear gener-

ated by oriented cell divisions and by external stresses reorients

polarity with different characteristics (Figures S5B–S5D and S5F).

The value of n thus depends on the types of cell shape changes

and rearrangements occurring during shear. In wounded wings,

where mechanical boundary conditions are perturbed, the re-

modeling processes that dominate shear differ from those in

unperturbed wings (compare Figures S3C–S3K with Figure 4).

In early pupal wings, fz clones nonautonomously reorient

polarity of adjacent wild-type tissue lying between the clone

and the wing margin (Figure S1N and S1O). But after remodeling,

Cell 142, 773–786, September 3, 2010 ª2010 Elsevier Inc. 781

Page 150: Cell 100903

the location of affected wild-type tissue is distal to the fz clone

(Vinson and Adler, 1987). We wondered whether the shear and

rotation observed during remodeling could account for reposi-

tioning of affected wild-type tissue. To investigate this, we first

simulated growth in the vertex model mimicking the situation

of a fz mutant clone. Starting with a network of 36 cells, we chose

a single cell and set sai = 1 on all its bonds (corresponding to loss

of Fz domains). We then simulated growth of the network while

simultaneously solving the dynamic equation of the PCP network

(Supplemental Theoretical Procedures, 5), keeping sai = 1 in the

simulated clone (indicated in blue). In these simulations, the

polarity pattern becomes distorted on one side of the clone as

domains with sai <0 (red) are positioned to face the clone

(compare Figure S5O with S5Q). These simulations generate

PCP patterns that mimic domineering nonautonomy observed

in vivo for fz and stbm mutant clones. Such patterns suggest

the interesting possibility that domineering nonautonomy may

arise during growth of the wing and may explain why the polarity

of wild-type tissue is already reoriented by fz clones at 15 hAPF.

We use these patterns as initial conditions to study reorientation

during shear.

To investigate the effects of shear and rotation on the per-

turbed PCP polarity patterns, we solve Equation 2 with n= � 3

for both networks with and without simulated clones, and an

initial average polarity axis lying 45� to the horizontal (mimicking

margin-oriented polarity). We impose shear and rotation at

constant rate, with the shear axis oriented horizontally (corre-

sponding to the PD axis of the wing) (Supplemental Theoretical

Procedures, 8). This reorients the global polarity of both net-

works and shifts the perturbed region toward the ‘‘distal’’ side

of the clone (compare Figure S5P and R) consistent with exper-

iments (Vinson and Adler, 1987).

Ds Guides Global Planar Cell Polarity by InfluencingEpithelial DynamicsPerturbing the Ft/Ds pathway misaligns PCP domains, but the

underlying mechanism is controversial (Axelrod, 2009; Casal

et al., 2006; Lawrence et al., 2007; Ma et al., 2003; Tree et al.,

2002a). These mutations also cause shorter, broader wings –

at least in part by perturbing oriented cell divisions in larval discs

(Baena-Lopez et al., 2005). To investigate whether Ds also influ-

enced pupal epithelial remodelling, we analyzed time-lapses of

ds pupal wings expressing ECadherin:GFP. At 15 hAPF, ds

wings are already misproportioned; the hinge is overgrown and

the wing blade is short and broad (Figure 7A), consistent with

altered larval growth. Nevertheless, the hinge contracts as it

does in wild-type, reducing its area about twofold (Figures 7A

and 7B). Thus, Ds is not required for hinge contraction. Calcu-

lating shear and rotation over the whole blade is difficult in ds

mutants because the epithelium folds; in regions that can be

quantified, rotation rates are smaller than in wild-type (Figures

S2G–S2I0 and S2L and Movie S7A). However, differences in

the shear patterns could not be resolved (Figures S2D–S2F0).A clearer picture emerged from analyzing ds mutants at the

cellular level (Figures 7D–7H and Movie S5D). PD cell elongation

is reduced in ds wings (compare Figure 4G and Figure 7J), and

the axis of cell division is less focused than in wild-type (compare

Figures 4H and 7K). Thus Ds is required to tightly focus cell divi-

sion orientation in both larval and pupal wings. Simulations

suggest that both cell elongation and oriented cell divisions

can reorient polarity (Figures 6C and 6D). Interestingly, both

are perturbed in ds mutants.

Ds influences both larval and pupal epithelial dynamics—so

which is most important for development of PD polarity? To

investigate this, we altered Ds levels in the posterior compart-

ment of the wing either throughout development or only during

larval growth or after pupariation. Overexpressing Ds throughout

wing development causes strong polarity defects in both poste-

rior and anterior compartments (Figure S6A). Overexpressing Ds

during larval or pupal development alone causes much milder

polarity defects (Figures S6A–S6C). We obtain complementary

results by lowering Ds levels. Inducing RNAi against Ds in the

posterior compartment throughout development perturbs

polarity strongly in both anterior and posterior compartments

(Figure S6D). Ds knock-down during larval or pupal development

alone causes milder polarity defects (Figures S6E and S6F).

These data suggest that Ds is needed both during growth and

subsequent epithelial remodelling to ensure proper alignment

of PCP with the PD axis. Ds loss or overexpression throughout

development causes cumulative defects stronger than those

caused by stage-specific perturbation.

Loss of Ds during pupal stages may perturb PD polarity by

reducing cell elongation or misorienting cell divisions during

remodelling. How could loss of Ds during larval growth affect

evolution of PD polarity? One possibility is that Ds-dependent

oriented cell divisions in the wing disc guide development of

early margin-oriented polarity; complex global PCP patterns

are as already seen in the third instar, and prepupal wings

show a pattern consistent with margin-oriented polarity (Classen

et al., 2005). To test this, we quantified the Stbm:YFP pattern in

ds wings at 15 hAPF. The pattern of average nematic order in ds

wings deviates from that of wild-type over large regions,

although PCP domains are well aligned locally, (compare

Figure 7C and Figure 1A). These observations together with

simulations suggest that the larval pattern of oriented cell divi-

sions contributes to development of margin directed polarity.

Loss of Ds may perturb the early polarity pattern indirectly by

misaligning larval cell divisions.

To ask whether abnormal early polarity in combination with the

altered flow pattern could in principle produce normal PD

polarity, we started with the altered polarity observed in early

ds wings and applied the mutant velocity field using Equation

2. For all values of n, polarity fails to orient as it does in wild-

type (average polar order < 0.67, compare Figure S5M with

S5K). Taken together, these studies indicate that development

of PD polarity depends both on correct initial polarity and on

tissue rotation and shear produced during hinge contraction.

Although Ds is expressed at high levels in the hinge and at low

levels in the wing blade, this discontinuity along the PD axis is not

required to direct PD orientation of PCP domains. Uniform Ds

expression rescues polarity in a ds mutant background (Mata-

katsu and Blair, 2004 and Figure S6G-I). Furthermore, lowering

Ds levels in the wing blade perturbs polarity more than knock-

down in the hinge (Figure S6K,L). In contrast, discontinuities in

Ds levels along the AP axis do perturb polarity at a distance;

overexpression or loss of Ds in the posterior compartment

782 Cell 142, 773–786, September 3, 2010 ª2010 Elsevier Inc.

Page 151: Cell 100903

E

19h30 APF

F

24 hAPF

D

15h45 APF 20 µm

phase I phase II

28h30 APF

G H

31h APF

P DP DP DP DP D

0

90

180

270

0

90

180

270

K

orientation of division axisphase I phase II

orientation of new boundary assemblyphase I phase II

L

0

90

180

270

0

90

180

270

I J

cell number (fraction of maximum)

average neighbor number

fraction hexagons

PD cell elongationnew cell contacts (fraction of maximum)cell contacts that will disappear (fraction of maximum)

0

0.2

0.4

0.6

0.8

1

14 16 18 20 22 24 26 28 30 32

0.08

0.12

0.16

0.2

0.24

0.28

0.32

fract

ion

PD

elo

ngat

ion

time (in hAPF)

0

0.2

0.4

0.6

0.8

1

14 16 18 20 22 24 26 28 30 325.4

5.5

5.6

5.7

5.8

5.9

6

fract

ion

aver

age

neig

hbor

num

ber

time (in hAPF)

phase IIphase I

phase IIphase I

15 hAPF

A B15h30' APF 30 hAPF

212*103µm2 123*103µm2123*103µm2 146*103µm2

C

persistent disappearing new boundary from neighbor exchange new boundary from cell division

Figure 7. Early Polarity, Cell Elongation, Division, and Rearrangement Are Perturbed in ds05142 Wings

(A and B) A ds05142, ubi-ECad:GFP wing at 15h300 and 30 hAPF. The hinge is colored blue and the blade red. Numbers indicate blade and hinge areas in mm2.

(C) Average nematic order in a ds05142,act-stbm:YFP/ ds05142 wing at 15 hAPF (compare to Figure 1A).

(D–H) A group of cells anterior to the posterior crossvein in a ds05142, ubi-ECad:GFP wing was tracked between 15h450 and 31 hAPF. Cell boundaries are color-

coded as in Figure 4A.

(I and J) Quantification of cellular changes in the patch of tissue tracked in (D–H). Fraction of maximal cell number (dark blue), average neighbor number (light blue,

averaged over 8 frames), fraction of hexagonal cells (brown, averaged over 8 frames), PD cell elongation (magenta, average maximum values = 0.243, n = 2

(0.2245, 0.261), fraction of boundaries that will disappear (yellow), and fraction of new boundaries resulting from neighbor exchange (green).

(K and L) Angular distribution of cell divisions (blue) (K) and new boundaries (green) (L) at the end of Phase I and Phase II. Yellow and magenta bars indicate

average angle of nematic order of cell division (yellow) and new cell boundaries (magenta) (see Supplemental Theoretical Procedures and Figure 4). Average

magnitude of phase I cell division order is 0.145 (0.180, 0.109 n = 2). Average magnitude of new boundary formation order in phase II = 0.338 (0.278, 0.398;

n = 2).

See also Figure S6.

Cell 142, 773–786, September 3, 2010 ª2010 Elsevier Inc. 783

Page 152: Cell 100903

nonautonomously perturbs polarity anteriorly (Figures S6A–

S6F). Altering Ds levels changes the response of wing epithelial

cells to hinge-dependent pulling forces (Figure 7). We therefore

considered the possibility that local changes in cell rearrange-

ment and elongation could alter large-scale cellular flows and

thus the patterns of shear and rotation in wild-type regions.

Indeed, Ds overexpression in the posterior compartment nonau-

tonomously reverses local tissue rotation in anterior cells near

the compartment boundary (compare Figures S6M–S6O, Fig-

ures 2J–2L, and Movie S7B). These observations suggest that,

to generate normal flow patterns in response to hinge contrac-

tion, cells in the anterior and posterior compartments must react

similarly to mechanical stress.

DISCUSSION

The mechanisms that couple tissue shape to the planar polarity

of constituent cells are not well understood. Here, we describe

a novel morphogenetic event that both shapes the wing of

Drosophila, and orients global planar polarity of wing cells. We

show that the hinge region contracts shortly after the prepupal

to pupal transition. This contraction not only shapes the wing

hinge, but also exerts anisotropic tension on the adjacent

wing-blade region, causing it to elongate and narrow. Reshaping

of the wing blade occurs as constituent cells flow proximally with

different velocities, generating reproducible patterns of flow,

shear and rotation. These patterns result from a combination of

oriented cell division, cell elongation and neighbor exchanges

that are guided by the anisotropic stresses.

This scenario, where externally generated forces have an

important role in epithelial remodeling differs from other con-

vergent-extension events where autonomous cell move-

ments actively drive tissue shape changes (Bertet et al., 2004;

Blankenship et al., 2006; Keller et al., 2008; Rauzi et al., 2008;

Zallen and Wieschaus, 2004). Externally induced epithelial

stretching has been proposed to provoke stress-relieving rear-

rangements during Drosophila germband extension (Butler

et al., 2009), and in the amphibian ectoderm (Keller et al.,

1992, 2008). But in these cases movements of the underlying

mesoderm, rather than contraction of adjacent epithelial cells,

provides the external force. The Drosophila pupal wing will

be a powerful system for uncovering molecular mechanisms

underlying stress-induced epithelial remodelling.

The mechanisms specifying global PCP domain alignment

have been elusive and controversial (Axelrod, 2009; Casal

et al., 2006; Lawrence et al., 2007; Ma et al., 2003; Tree et al.,

2002a). It had been thought that PCP proteins in the Drosophila

wing are initially randomly distributed, but become intracellularly

polarized along the PD axis starting approximately 10 hr before

hair formation (Strutt and Strutt, 2005; Vladar et al., 2009; Wu

and Mlodzik, 2009; Zallen, 2007). This led to the search for

a global polarizing signal that operated at this time. Here, we

show that PD orientation of PCP complexes does not arise

from a random distribution; rather it evolves from a different

global polarity pattern present at the prepupal-pupal transition.

At this stage, PCP domains are organized such that Fz-contain-

ing domains face the wing margin rather than distally. Our

previous work identified a similar global polarity pattern in

prepupal wings, and showed that even larval wing discs show

a global pattern of PCP (Classen et al., 2005). Thus PCP domains

are always polarized, but their global orientation evolves dynam-

ically during development.

What causes global reorientation of PCP during phase I, and

the subsequent increase in the magnitude of polar order during

phase II of wing-blade remodeling? During phase I, margin-

oriented polarity is reoriented by specific patterns of local rotation

and shear due to cell flows caused by hinge contraction. Local

tissue rotation and shear reorient PCP domains by simple phys-

ical rules (see Equation [2] and Figure S7) similar to those that re-

orient molecular order in liquid crystal hydrodynamics. A signifi-

cant fraction of PCP reorientation is produced by local tissue

rotation alone. However, tissue shear also helps rotate polarity

toward the shear axis. Simulations suggest that shear caused

by oriented cell divisions, cell rearrangements and PD cell elonga-

tion can all contribute to polarity reorientation. One mechanism

that may underlie the effect of cell elongation on PCP is the

tendency of microtubules to align with the long axis of the cell.

Indeedmicrotubules are aligned with the PD axis of wing epithelial

cells and are essential for delivery of PCP proteins to the cortex

(Shimada et al., 2006). Significantly, under conditions where

normal PD polarity does not develop (e.g., in ds mutant wings

or in severed wings), oriented cell division, cell rearrangements

and cell elongation are disturbed. During phase II, oriented cell

boundary rearrangements increase hexagonal packing geometry

and improve PCP order. Simulations suggest this may occur via

an annealing process that relaxes irregular cell packing to a

more ordered hexagonal lattice with improved PCP order.

The atypical Cadherins Ft and Ds are needed to evolve the

global PD polarity pattern. Exactly how they influence the global

pattern has been controversial (Axelrod, 2009; Lawrence et al.,

2007). It was proposed that this pathway provides a PD polarity

cue that depends on higher relative levels of Ds expression in the

hinge. This inhomogeneity was thought to generate intracellular

asymmetry of Ft/Ds heterodimers throughout the wing, directly

generating a small bias in Fz activity within each cell that could

then be amplified to produce strong and stable alignment of

PCP domains with the PD axis (Ma et al., 2003; Tree et al.,

2002a). But the fact that uniform Ds expression suffices to

rescue polarity in ds wings argues against this view (Matakatsu

and Blair, 2004) (Figures S6G–S6I). Furthermore, Ft and Ds do

not appear to act directly on core PCP proteins in the Drosophila

abdomen (Casal et al., 2006). We propose that the Ft/Ds

pathway influences the global PCP pattern, not by generating

intracellular asymmetries that directly regulate PCP signaling,

but indirectly through its effects on epithelial dynamics during

both larval and pupal stages. Time-controlled loss of Ds shows

that it is required throughout development for evolution of PD

polarity (Figures S6D–S6F) (see also (Matakatsu and Blair,

2004)). During larval stages, Ds may guide the development of

early margin-oriented polarity through its influence on oriented

growth; our simulations suggest that oriented cell division

strongly influences the global axis of planar polarity. Later, during

hinge contraction and wing-blade remodeling, Ds is required

for the oriented cell division and cell elongation that occur in

response to anisotropic stresses. Both processes influence the

polarity axis in vertex model simulations.

784 Cell 142, 773–786, September 3, 2010 ª2010 Elsevier Inc.

Page 153: Cell 100903

In future, it will be interesting to examine the roles of aniso-

tropic growth and local morphogen signaling in establishing

the global PCP pattern as wing discs grow. Our theoretical anal-

ysis shows that large scale planar polarity can be stably main-

tained during growth even in the absence of global polarizing

signals. This is possible because polarity easily aligns over

a few cell diameters without orientation defects, and subsequent

cell rearrangements due to cell division do not destroy this order

during growth. Thus, polarity that was initially established when

the tissue was small can be expanded during growth, and any

anterior-posterior or dorsal-ventral boundary-derived signal

would not need to act directly over long distances.

The coupling of planar polarity to epithelial dynamics is

a robust and flexible mechanism for coordinating tissue shape

with planar polarity of constituent cells and may generalize to

other systems. Indeed, mutations in the Wnt pathway perturb

both convergent extension and the planar orientation of sensory

hair cells in the mouse cochlea (Wang et al., 2005). Furthermore,

the planar polarity of sensory hair cells in the lateral line organ of

zebrafish is oriented by the direction of cell migration rather than

long-range secreted cues (Lopez-Schier et al., 2004). Our theo-

retical analysis suggests that shear caused by stretch-induced

rearrangements or by oriented cell divisions can reorient polarity

either perpendicular or parallel to the shear axis, depending on

the relative rates of shear and PCP turnover. Thus, this process

is also highly versatile, and could be a fundamental principle of

tissue organization.

EXPERIMENTAL PROCEDURES

Imaging

Pupae were prepared for imaging as described (Classen et al., 2008). Images

were acquired with a Leica TCS-SP2 or an Olympus FV-1000 microscope.

203 and 403 oil immersion objectives were used to follow cell flows and

global morphological changes occurring in the wing. A 633 oil immersion

objective was used to image pupal wings with cellular resolution. Imaging

was performed at 25� ± 2�C, or at 29�C using a Bachhoffer chamber.

Wing Wounding

Wings were wounded by gently scratching the pupal cuticule in the hinge

region of the wing using forceps just before imaging (at around 15 hAPF).

Image Analysis and Figures

Z projections were created with ImageJ. For image analysis, we developed

‘‘packing analyzer v2.0’’, which measures cell areas, perimeter, packing, elon-

gation, polarity, cell divisions and tracks the cells and their boundaries. Briefly,

images are segmented using the watershed algorithm (Vincent and Soille,

1991). Each watershed catchment’s basin is defined as a cell. Vertices are

defined as pixels shared between three or more cells. Cell boundaries are

identified as pixels shared by exactly two cells. Cell boundaries smaller than

three pixels are reclassified as vertices. The polygon class is determined by

counting cell vertices. Cell area is defined as the number of pixels in each

cell. Cell perimeter is defined as the sum of the distance between all the pixels

surrounding a cell. To track cells, we assign a unique identity to each in the first

movie frame. Cells are then re-identified in the next images based on their

positions and neighborhoods. Boundaries are defined by the two cells that

share them. Division orientation is defined by a line connecting the centers

of the two daughter cells, and the angle it makes with the PD axis (which is

plotted as horizontal).

Movies created using our software were compressed using FFmpeg or

Quicktime Pro. Graphs were created using Microsoft Excel, Gnuplot and

Grace. Images were vectorized using Batik. Figures were composed with

Adobe Illustrator.

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures,

Supplemental Theoretical Procedures, seven figures, and seven movies and

can be found with this article online at doi:10.1016/j.cell.2010.07.042.

ACKNOWLEDGMENTS

This work represents a truly collaborative effort. Each author has contributed

significantly to the findings and regular group discussions guided the develop-

ment of the ideas presented here. The manuscript was written jointly by all

authors. S.E., A.S., J.R., R.F., and F.J. were supported by the Max Planck

Gesellschaft. B.A. was funded by the Fondation pour la Recherche Medicale.

D.S. acknowledges the Natural Sciences and Engineering Research Council of

Canada and the German Academic Exchange Service. J.R. was supported by

a predoctoral fellowship from the Boehringer Ingelheim Fonds. We thank Tony

Hyman and Stephan Grill for use of the laser cutting microscope. We are grate-

ful to Stephan Grill, Ewa Paluch, Elisabeth Knust, and Carl-Philipp Heisenberg

for critical comments on the manuscript. We thank David Strutt, Seth S.Blair,

VDRC and the Bloomington Stock Center for providing fly stocks, Barry Dick-

son for DNA constructs, Julia Gabel for help with cloning, Sven Ssykor for

transgenic injections.

Received: August 28, 2009

Revised: May 11, 2010

Accepted: July 23, 2010

Published: September 2, 2010

REFERENCES

Adler, P.N., Charlton, J., and Liu, J. (1998). Mutations in the cadherin super-

family member gene dachsous cause a tissue polarity phenotype by altering

frizzled signaling. Development 125, 959–968.

Amonlirdviman, K., Khare, N.A., Tree, D.N., Chen, W., Axelrod, J., and Tomlin,

C.J. (2005). Mathematical modeling of planar cell polarity to understand domi-

neering non-autonomy. Science 307, 423–426.

Axelrod, J.D. (2009). Progress and challenges in understanding planar cell

polarity signaling. Semin. Cell Dev. Biol. 20, 964–971.

Baena-Lopez, L.A., Baonza, A., and Garcia-Bellido, A. (2005). The orientation

of cell divisions determines the shape of Drosophila organs. Curr. Biol. 15,

1640–1644.

Bertet, C., Sulak, L., and Lecuit, T. (2004). Myosin-dependent junction remod-

elling controls planar cell intercalation and axis elongation. Nature 429,

667–671.

Blankenship, J.T., Backovic, S.T., Sanny, J.S., Weitz, O., and Zallen, J.A.

(2006). Multicellular rosette formation links planar cell polarity to tissue

morphogenesis. Dev. Cell 11, 459–470.

Bryant, P.J., Huettner, B., Held, L.I., Jr., Ryerse, J., and Szidonya, J. (1988).

Mutations at the fat locus interfere with cell proliferation control and epithelial

morphogenesis in Drosophila. Dev. Biol. 129, 541–554.

Burak, Y., and Shraiman, B.I. (2009). Order and stochastic dynamics in

Drosophila planar cell polarity. PLoS Comput. Biol. 5, e1000628.

Butler, L.C., Blanchard, G.B., Kabla, A.J., Lawrence, N.J., Welchman, D.P.,

Mahadevan, L., Adams, R.J., and Sanson, B. (2009). Cell shape changes indi-

cate a role for extrinsic tensile forces in Drosophila germ-band extension. Nat.

Cell Biol. 11, 859–864.

Casal, J., Lawrence, P.A., and Struhl, G. (2006). Two separate molecular

systems, Dachsous/Fat and Starry night/Frizzled, act independently to confer

planar cell polarity. Development 133, 4561–4572.

Chen, W.S., Antic, D., Matis, M., Logan, C.Y., Povelones, M., Anderson, G.A.,

Nusse, R., and Axelrod, J.D. (2008). Asymmetric homotypic interactions of the

Cell 142, 773–786, September 3, 2010 ª2010 Elsevier Inc. 785

Page 154: Cell 100903

atypical cadherin flamingo mediate intercellular polarity signaling. Cell 133,

1093–1105.

Cho, E., and Irvine, K.D. (2004). Action of fat, four-jointed, dachsous and dachs

in distal-to-proximal wing signaling. Development 131, 4489–4500.

Clark, H.F., Brentrup, D., Schneitz, K., Bieber, A., Goodman, C., and Noll, M.

(1995). Dachsous encodes a member of the cadherin superfamily that controls

imaginal disc morphogenesis in Drosophila. Genes Dev. 9, 1530–1542.

Classen, A., Anderson, K., Marois, E., and Eaton, S. (2005). Hexagonal Packing

of Drosophila Wing Epithelial Cells by the Planar Cell Polarity Pathway. Dev.

Cell 9, 805–817.

Classen, A.K., Aigouy, B., Giangrande, A., and Eaton, S. (2008). Imaging

Drosophila pupal wing morphogenesis. Methods Mol. Biol. 420, 265–275.

Cortes, S., Glade, N., Chartier, I., and Tabony, J. (2006). Microtubule self-orga-

nisation by reaction-diffusion processes in miniature cell-sized containers and

phospholipid vesicles. Biophys. Chem. 120, 168–177.

Daga, R.R., and Nurse, P. (2008). Interphase microtubule bundles use global

cell shape to guide spindle alignment in fission yeast. J. Cell Sci. 121, 1973–

1980.

de Gennes, P.G., and Prost, J. (1993). The Physics of Liquid Crystals, Second

Edition (Gloucestershire, UK: Clarendon Press).

Farhadifar, R., Roper, J.C., Aigouy, B., Eaton, S., and Julicher, F. (2007). The

Influence of Cell Mechanics, Cell-Cell Interactions, and Proliferation on Epithe-

lial Packing. Curr. Biol. 17, 2095–2104.

Garoia, F., Grifoni, D., Trotta, V., Guerra, D., Pezzoli, M.C., and Cavicchi, S.

(2005). The tumor suppressor gene fat modulates the EGFR-mediated prolifer-

ation control in the imaginal tissues of Drosophila melanogaster. Mech. Dev.

122, 175–187.

Garoia, F., Guerra, D., Pezzoli, M.C., Lopez-Varea, A., Cavicchi, S., and Gar-

cia-Bellido, A. (2000). Cell behaviour of Drosophila fat cadherin mutations in

wing development. Mech. Dev. 94, 95–109.

Haase, S.B., and Lew, D.J. (2007). Microtubule organization: cell shape is

destiny. Curr. Biol. 17, R249–R251.

Ishikawa, H.O., Takeuchi, H., Haltiwanger, R.S., and Irvine, K.D. (2008). Four-

jointed is a Golgi kinase that phosphorylates a subset of cadherin domains.

Science 321, 401–404.

Joanny, J.F., Julicher, F., Kruse, K., and Prost, J. (2007). Hydrodynamic theory

for multi-component active polar gels. N. J. Phys. 9, 422.

Keller, R., Shih, J., and Sater, A. (1992). The cellular basis of the convergence

and extension of the Xenopus neural plate. Dev. Dyn. 193, 199–217.

Keller, R., Shook, D., and Skoglund, P. (2008). The forces that shape embryos:

physical aspects of convergent extension by cell intercalation. Phys. Biol. 5,

15007.

Lawrence, P.A., Struhl, G., and Casal, J. (2007). Planar cell polarity: one or two

pathways? Nat. Rev. Genet. 8, 555–563.

Le Garrec, J.F., Lopez, P., and Kerszberg, M. (2006). Establishment and main-

tenance of planar epithelial cell polarity by asymmetric cadherin bridges:

a computer model. Dev. Dyn. 235, 235–246.

Lopez-Schier, H., Starr, C.J., Kappler, J.A., Kollmar, R., and Hudspeth, A.J.

(2004). Directional cell migration establishes the axes of planar polarity in the

posterior lateral-line organ of the zebrafish. Dev. Cell 7, 401–412.

Ma, D., Amonlirdviman, K., Raffard, R.L., Abate, A., Tomlin, C.J., and Axelrod,

J.D. (2008). Cell packing influences planar cell polarity signaling. Proc. Natl.

Acad. Sci. USA 105, 18800–18805.

Ma, D., Yang, C.H., McNeill, H., Simon, M.A., and Axelrod, J.D. (2003). Fidelity

in planar cell polarity signalling. Nature 421, 543–547.

Matakatsu, H., and Blair, S.S. (2004). Interactions between Fat and Dachsous

and the regulation of planar cell polarity in the Drosophila wing. Development

131, 3785–3794.

Rauzi, M., Verant, P., Lecuit, T., and Lenne, P.F. (2008). Nature and anisotropy

of cortical forces orienting Drosophila tissue morphogenesis. Nat. Cell Biol. 10,

1401–1410.

Seifert, J.R., and Mlodzik, M. (2007). Frizzled/PCP signalling: a conserved

mechanism regulating cell polarity and directed motility. Nat. Rev. Genet. 8,

126–138.

Shimada, Y., Yonemura, S., Ohkura, H., Strutt, D., and Uemura, T. (2006).

Polarized transport of Frizzled along the planar microtubule arrays in

Drosophila wing epithelium. Dev. Cell 10, 209–222.

Simons, M., and Mlodzik, M. (2008). Planar cell polarity signaling: from fly

development to human disease. Annu. Rev. Genet. 42, 517–540.

Strauss, B., Adams, R.J., and Papalopulu, N. (2006). A default mechanism of

spindle orientation based on cell shape is sufficient to generate cell fate diver-

sity in polarised Xenopus blastomeres. Development 133, 3883–3893.

Strutt, H., and Strutt, D. (2002). Nonautonomous planar polarity patterning

in Drosophila: dishevelled-independent functions of frizzled. Dev. Cell 3,

851–863.

Strutt, H., and Strutt, D. (2005). Long-range coordination of planar polarity in

Drosophila. Bioessays 27, 1218–1227.

Strutt, H., and Strutt, D. (2008). Differential stability of flamingo protein

complexes underlies the establishment of planar polarity. Curr. Biol. 18,

1555–1564.

Tree, D.R., Ma, D., and Axelrod, J.D. (2002a). A three-tiered mechanism for

regulation of planar cell polarity. Semin. Cell Dev. Biol. 13, 217–224.

Tree, D.R., Shulman, J.M., Rousset, R., Scott, M.P., Gubb, D., and Axelrod,

J.D. (2002b). Prickle mediates feedback amplification to generate asymmetric

planar cell polarity signaling. Cell 109, 371–381.

Uemura, T., and Shimada, Y. (2003). Breaking cellular symmetry along planar

axes in Drosophila and vertebrates. J. Biochem. 134, 625–630.

Villano, J.L., and Katz, F.N. (1995). four-jointed is required for intermediate

growth in the proximal-distal axis in Drosophila. Development 121, 2767–

2777.

Vincent, L., and Soille, P. (1991). Watersheds in digital spaces: an efficient

algorithm based on immersion simulations. IEEE 13, 583–598.

Vinson, C., and Adler, P.N. (1987). Directional non-cell autonomy and the

transmission of polarity information by the frizzled gene of Drosophila. Nature

329, 549–551.

Vladar, E.K., Antic, D., and Axelrod, J.D. (2009). Planar cell polarity signaling:

The developing cell’s compass. Cold Spring Harb. Perspect. Biol. 1, a002964.

Wang, J., Mark, S., Zhang, X., Qian, D., Yoo, S.J., Radde-Gallwitz, K., Zhang,

Y., Lin, X., Collazo, A., Wynshaw-Boris, A., et al. (2005). Regulation of polarized

extension and planar cell polarity in the cochlea by the vertebrate PCP

pathway. Nat. Genet. 37, 980–985.

Wu, J., and Mlodzik, M. (2008). The frizzled extracellular domain is a ligand for

Van Gogh/Stbm during nonautonomous planar cell polarity signaling. Dev. Cell

15, 462–469.

Wu, J., and Mlodzik, M. (2009). A quest for the mechanism regulating global

planar cell polarity of tissues. Trends Cell Biol. 19, 295–305.

Zallen, J.A. (2007). Planar polarity and tissue morphogenesis. Cell 129, 1051–

1063.

Zallen, J.A., and Wieschaus, E. (2004). Patterned gene expression directs

bipolar planar polarity in Drosophila. Dev. Cell 6, 343–355.

Zeidler, M.P., Perrimon, N., and Strutt, D.I. (2000). Multiple roles for four-

jointed in planar polarity and limb patterning. Dev. Biol. 228, 181–196.

786 Cell 142, 773–786, September 3, 2010 ª2010 Elsevier Inc.

Page 155: Cell 100903

Generation of Rat Pancreas in Mouseby Interspecific Blastocyst Injectionof Pluripotent Stem CellsToshihiro Kobayashi,1,2 Tomoyuki Yamaguchi,1,2 Sanae Hamanaka,1,2 Megumi Kato-Itoh,2,3 Yuji Yamazaki,1,2

Makoto Ibata,2 Hideyuki Sato,1,2 Youn-Su Lee,1,2 Jo-ichi Usui,1,6 A.S. Knisely,5 Masumi Hirabayashi,3,4

and Hiromitsu Nakauchi1,2,*1Division of Stem Cell Therapy, Center for Stem Cell Biology and Regenerative Medicine, Institute of Medical Science, University of Tokyo,

4-6-1 Shirokanedai, Minato-ku, Tokyo 108-8639, Japan2Japan Science Technology Agency, ERATO, Nakauchi Stem Cell and Organ Regeneration Project, 4-6-1 Shirokanedai, Minato-ku,

Tokyo 108-8639, Japan3Center for Genetic Analysis of Behavior, National Institute for Physiological Sciences, Okazaki, Aichi 444-8585, Japan4School of Life Science, The Graduate University for Advanced Studies, Okazaki, Aichi 444-8585, Japan5Institute of Liver Studies, King’s College Hospital, London SE5 9RS, UK6Present address: Department of Nephrology, Graduate School of Comprehensive Human Sciences, University of Tsukuba, 1-1-1 Tennoudai,

Tsukuba, Ibaraki 305-8575, Japan*Correspondence: [email protected]

DOI 10.1016/j.cell.2010.07.039

SUMMARY

The complexity of organogenesis hinders in vitrogeneration of organs derived from a patient’s pluripo-tent stem cells (PSCs), an ultimate goal of regenera-tive medicine. Mouse wild-type PSCs injected intoPdx1�/� (pancreatogenesis-disabled) mouse blasto-cysts developmentally compensated vacancy ofthe pancreatic ‘‘developmental niche,’’ generatingalmost entirely PSC-derived pancreas. To examinethe potential for xenogenic approaches in blastocystcomplementation, we injected mouse or rat PSCs intorat or mouse blastocysts, respectively, generatinginterspecific chimeras and thus confirming thatPSCs can contribute to xenogenic developmentbetween mouse and rat. The development of thesemouse/rat chimeras was primarily influenced byhost blastocyst and/or foster mother, evident bybody size and species-specific organogenesis. Wefurther injected rat wild-type PSCs into Pdx1�/�

mouse blastocysts, generating normally functioningrat pancreas in Pdx1�/� mice. These data constituteproof of principle for interspecific blastocyst comple-mentation and for generation in vivo of organs derivedfrom donor PSCs using a xenogenic environment.

INTRODUCTION

Current stem cell therapy mainly targets diseases that can be

treated by cell replacement, such as Parkinson’s disease or dia-

betes mellitus. One of the ultimate goals of regenerative medi-

cine, however, is to grow organs using the patient’s own stem

cells and to transplant those organs into the patient. With the

development of induced pluripotent stem cell (iPSC) technology,

we are now able to obtain patient-derived PSCs (Takahashi

et al., 2007; Takahashi and Yamanaka, 2006), although actual

developmental potentials remain to be defined, as do risks asso-

ciated with somatic cell reprogramming. The real challenge is to

create a reproductive system for generation of PSC-derived

organs. The interactions among cells and tissues during devel-

opment and organogenesis are so complex that the recapitula-

tion of these interactions to generate organs in vitro is essentially

impractical. We have challenged this goal using the biology of

blastocyst complementation.

Blastocyst complementation was first reported by Chen et al.

They demonstrated that deficiency of T and B lymphocyte

lineages in Rag2-deficient (Rag2�/�) mice was complemented

by injecting normal mouse embryonic stem cells (mESCs) into

Rag2�/� mouse-derived blastocysts (Chen et al., 1993).

Because Rag2 is an indispensable enzyme for rearrangement

of immunoglobulin and T cell receptor genes, the T and B cells

generated in the complemented animals were mESC derived;

there were no host T or B lymphocytes. We assumed that this

complementation was possible because the Rag2�/� host, inca-

pable of generating mature T and B cells, provided a ‘‘develop-

mental niche’’ for ESC-derived T and B cells.

We hypothesized that with blastocysts derived from a mutant

mouse strain in which the gene necessary to form a particular

organ is deficient, the same principle might apply. To test

this hypothesis, we used, in this study, blastocyst complementa-

tion to generate functional pancreas from donor PSCs. The

pancreas, consisting of endocrine and exocrine glands, is formed

by early embryonic interactions of mesenchyme and epithelium

(Slack, 1995). Pdx1 (pancreatic and duodenal homeobox1) is

a Hox-type transcription factor that plays a critical role in pancre-

atic development and b cell maturation. Homozygous deficiency

of Pdx1 in the mouse results in death soon after birth due to

Cell 142, 787–799, September 3, 2010 ª2010 Elsevier Inc. 787

Page 156: Cell 100903

pancreatic insufficiency (Offield et al., 1996). Targeted disruption

of Pdx1 thus should empty a pancreatic ‘‘developmental niche’’

in embryos derived from Pdx1�/� blastocysts. Therefore, injec-

tion of mouse PSCs (mPSCs) into Pdx1�/� blastocysts should

result in generation of pancreata almost entirely derived from

injected mPSCs.

Although our results verified our initial hypothesis, this system

cannot be applied to generate human organs. A xenogenic, but

not allogenic, blastocyst complementation system must there-

fore be established. However, little is known about the nature

of the xenogenic barrier, how organogenesis by donor PSCs is

influenced, or how intrinsic developmental programs can be

regulated in xenogenic environments. To address these issues,

we attempted to generate interspecific chimeras between

mouse and rat using a blastocyst injection technique with mouse

and rat PSCs.

Generation of interspecific chimeras in livestock animals using

preimplantation embryos of each species is described; the

‘‘geep,’’ or chimera between goat and sheep, is famous as an

interspecific live chimera (Fehilly et al., 1984). In rodents,

however, methods like those producing the geep succeeded

only between mouse subspecies, such as Mus musculus and

Mus caroli, that are closely related but not capable of inter-

breeding (Rossant and Frels, 1980). Many groups have sought

to generate interspecific chimeras between mouse and rat,

successful with chimeric preimplantation embryos in vitro but

failing with live chimeric animals (Stern, 1973; Zeilmaker,

1973). Extraembryonic lineage cells like trophectoderm or prim-

itive endoderm, derived from xenogenic embryos, might suffer

inhibition of implantation on exposure to host uterus; inhibition

of further intrauterine development also is possible (Rossant

et al., 1982; Tarkowski, 1962). Only cells of pre-blastocyst origin

can contribute to extraembryonic lineage cells, and mESCs/

iPSCs do not thus contribute (Rossant, 2007). Therefore, we

rechallenged this old issue with new technology, using iPSCs

or ESCs that are not capable of contributing to extraembryonic

tissues. Using recently established culture conditions with a

combination of signaling inhibitors (Buehr et al., 2008; Li et al.,

2008; Ying et al., 2008), we generated rat-ESCs and -iPSCs

(rESCs, riPSCs). Our work has now confirmed the existence of

interspecific chimeras generated with PSCs, using injection not

only of mouse PSCs into rat embryos but also of rat PSCs into

mouse embryos.

Finally, by combining the principle of blastocyst complemen-

tation with the production of interspecific chimeras, we suc-

ceeded in generating rat pancreas in Pdx1�/� mice. These sets

of experiments provide proof-of-principle data for donor iPSC-

derived organ generation in a xenogenic environment.

RESULTS

Generation of Donor Mouse iPSC-Derived Pancreasin Pdx1-Deficient NeonatesOur first goal was to generate pancreas from mPSCs. We per-

formed a blastocyst complementation experiment using

Pdx1�/� blastocysts that would provide a niche for pancreatic

organogenesis. These mice exhibit pancreatic agenesis due to

the absence of Pdx1-specified interactions between prepancre-

atic epithelium and mesenchyme, a key step in pancreatic devel-

opment. Homozygote Pdx1-LacZ knockin mice (Pdx1�/�) are

born alive but die within 1 week after birth, presumably due to

pancreatic insufficiency (Offield et al., 1996).

For complementation donor cells, we used GT3.2 mouse

iPSCs (miPSCs) that had been generated from tail tip fibroblasts

(TTFs) of an adult EGFP-transgenic (TG) C57BL/6 mouse (Okabe

et al., 1997) using three factors (Oct3/4, Sox2, Klf4) in retroviral

vectors (Figure S1 available online). These GT3.2 miPSCs

were injected into blastocysts obtained by an intercross of

heterozygote Pdx1-LacZ knockin mice (Pdx1+/�), offspring of

C57BL/6(Pdx1+/�) 3 DBA(Pdx1+/+) or C57BL/6(Pdx1+/�) 3

BDF1(Pdx+/+) mice. G4.2 mESCs were also injected for compar-

ison. Neonates were assessed macroscopically and histologi-

cally for pancreatic development and were genotyped for

Pdx1. These neonates were highly chimeric: Due to silencing

of EGFP and to contamination by donor cells, accurate determi-

nation of the genotypes required analysis at the single-cell level.

To identify the genotypes of these neonates accurately, EGFP-

negative, c-Kit-positive, and Sca1-positive, lineage marker-

negative (EGFP�KSL) bone marrow cells were clone-sorted by

fluorescence-activated cell sorting (FACS) to obtain single-cell-

derived hematopoietic colonies (Figure S2A). After culture for

2 weeks, cells collected from each colony were subjected to

PCR analysis (Figure S2B). This ‘‘colony PCR’’ method unambig-

uously determined the host embryo genotype and proved that

pancreas had been formed in Pdx1�/� mice by blastocyst

complementation (Figure S2C).

In all postnatal chimeric mice derived from blastocysts

injected with either miPSCs or mESCs, pancreas was present

regardless of host genotype (including Pdx1�/�). As expected,

the pancreas in Pdx1+/� or Pdx1+/+ chimeric mice was a

composite of host-derived cells and EGFP-miPSC- or mESC-

derived cells, as in whole-body chimerism (Pdx1+/� + miPSCs

or mESCs in Figure 1A). In contrast, the pancreatic epithelium

in Pdx1�/� chimeric mice was almost entirely composed of

EGFP-marked miPSC- or mESC-derived cells (Pdx1�/� +

miPSCs or mESCs in Figure 1A). The pancreas of these mice

was grossly and histologically normal. We examined these

pancreatic tissues further for contributions of donor mPSCs to

different pancreatic lineages. Both miPSCs and mESCs supplied

all pancreatic cell lineages (exocrine and endocrine tissues,

including ductal epithelia), but pancreatic stromal elements—

vessels, nerves, and fibrocytes—were composites of host- and

mPSC-derived cells in all mice (Figure 1B).

miPSCs Rescued Pdx1�/� Mice by BlastocystComplementationWe next addressed whether mPSCs can rescue Pdx1�/�

lethality via blastocyst complementation. As we predicted based

on neonatal analysis (Figure 1), both mESCs and miPSCs

contributed to pancreatic organogenesis; Pdx1�/� chimeric

mice survived to adulthood (mESCs injection: n = 4, miPSCs

injection: n = 15 in Figure 2A). They even served as Pdx1�/�

founders, transmitting their genotype to the next generation.

Mating Pdx1�/� founder male mice with Pdx1+/� female mice

increased to 50% the proportion of pups of homozygously dis-

rupted genotype (Figure 2A). These data indicate that blastocyst

788 Cell 142, 787–799, September 3, 2010 ª2010 Elsevier Inc.

Page 157: Cell 100903

complementation can be used to generate a functional organ

and to yield homozygous founder mice even when uncomple-

mented homozygosity is fatal. The method described here thus

can be useful for efficient production of gene-targeted mice

that are embryonic lethal without introducing a conditional

gene targeting system.

As expected, in both Pdx1�/� and Pdx1+/ � mice injected with

EGFP-miPSC, miPSC-derived cells contributed to all the non-

pancreatic tissues of the body, including lung, heart, liver,

muscle, testis, and brain (data not shown). The extent of contri-

bution varied from tissue to tissue and also varied depending on

individual chimera. However, the pancreas in adult Pdx1�/� mice

was entirely derived from donor miPSCs (Pdx1�/� in Figure 2B).

Detailed analysis of EGFP distributions in adult miPSC-derived

pancreas demonstrated that pancreatic islets, exocrine tissues,

and duct epithelia were entirely derived from donor miPSCs, as

already shown in neonates (Pdx1�/� in Figures 2C and 2D). In

expected contrast, pancreas from Pdx1+/� mice was a com-

posite of host and donor derivatives (Pdx1+/� in Figures 2B

and 2C). Quantitative analysis of these sections by image J soft-

ware revealed percentages of EGFP-positive cells to be 27.4% ±

27.8% in the miPSC + Pdx1+/� setting, 95.6% ± 4.6% in the

miPSC + Pdx1�/� setting. Of note is that most individual islets

A

Pdx1-/- Pdx1+/-

Insulin Glucagon Somatostatin α-Amylase DBA-Lectin PECAM1Endocrine Exocrine Ductal cell Endothelial cell

Pancreatic tissues Blood vessel

Pha

se

EG

FP

B

+ miPSCs

Pdx1-/- Pdx1+/- Pdx1-/- + mESCs Figure 1. Mouse PSC-Derived Pancreas Gener-

ated by Way of Blastocyst Complementation

(A) mPSC-derived pancreas at neonatal stage. In

Pdx1�/� mice complemented with mPSCs, the pan-

creas was almost entirely positive for EGFP, indicating

mPSC origin (left panels in +miPSCs and +mESCs). On

the other hand, in Pdx1+/� mice complemented with

mPSCs, the pancreas was partially positive for EGFP,

indicating a composite of both mPSC- and host-derived

cells (right panels in +miPSCs and +mESCs). Right panel

shows absence of pancreas in a nonchimeric Pdx1�/�

mouse at the same developmental stage. See also

Figure S1 for detailed characteristics of miPSCs and

Figure S2 for genotyping of Pdx1 status.

(B) Histological analysis of the distribution of mESC-

derived cells in neonatal mouse pancreas. Sections of

pancreas generated in a Pdx1�/� mouse complemented

with mESCs were immunostained for EGFP; for pancre-

atic endocrine markers (insulin, glucagon, and somato-

statin); for the pancreatic exocrine marker a-amylase;

for the ductal marker rhodamine-conjugated DBA-

lectin; and for blood vessels with PECAM-1, with DAPI

nuclear counterstaining.

Sections were observed under confocal laser scanning

microscopy. Scale bars in (B), 20 mm.

were composed of both host-derived and

miPSC-derived cells (Pdx1+/� in Figure 2D),

indicating, as previously reported (Deltour

et al., 1991), nonclonal origin of pancreatic

islets.

To test whether miPSC-derived pancreata

are functional in Pdx1�/� chimeric mice, we

performed glucose tolerance testing (GTT).

Whereas streptozotocin (STZ)-induced dia-

betic mice failed to respond to GTT, Pdx1�/�

chimeric mice responded well, indicating that Pdx1�/� mice

with miPSC-derived pancreas secreted insulin in response to

glucose loading and maintained normal serum glucose levels

(Figure 2E). Histology and function of exogenously derived

pancreas were essentially the same for mice derived from

miPSC- and from mESC-complemented blastocysts (Figure S3).

These data demonstrate that the vacant ‘‘pancreatic niche’’

provided in Pdx1�/� mice can be occupied, with developmental

compensation, by miPSC- or mESC-derived cells after intraspe-

cific blastocyst complementation, generating functionally intact

pancreas.

Transplantation of miPSC-Derived Islets CorrectedHyperglycemia in Diabetic MiceTo assess the functionality of miPSC-derived pancreas, islets

from miPSC-derived pancreas were transplanted into mice of

the original strain (C57BL/6) in which STZ administration had

induced diabetes. As a control, islets from pancreas in Pdx1+/�

chimeric mice injected with miPSCs were transplanted. Blasto-

cysts were obtained by an intercross of Pdx1+/� mice (C57BL/

6 3 DBA2 or C57BL/6 3 BDF1 F1 strains) that were semi-

allogenic to miPSCs used in this study. EGFP-expressing

miPSC-derived islets (Figure 3A) were isolated conventionally

Cell 142, 787–799, September 3, 2010 ª2010 Elsevier Inc. 789

Page 158: Cell 100903

based on a method developed previously (Gotoh et al., 1985)

and were transplanted beneath the renal capsule of recipient

mice. Nonfasting blood glucose levels were then monitored. As

these islets were of donor origin (i.e., C57BL/6 strain), an immu-

nosuppressive regimen was not used.

To prevent nonspecific loss of islets due to inflammation,

anti-inflammatory cytokine monoclonal antibody (mAb) cocktails

were given at transplantation and 2 and 4 days thereafter

(arrows, Figure 3E), as described (Satoh et al., 2007). Two

months after transplant, EGFP-expressing miPSC-derived islets

were still detectable at the graft site (Figures 3B and 3C). Produc-

tion of insulin by the transplanted islets was confirmed immuno-

histologically (Figure 3D). The induced-diabetic recipients no

longer exhibited hyperglycemia; they maintained normal blood

glucose levels and responded normally to GTT (Figures 3E

and 3F). This is in contrast to the therapeutic effect conferred

by islets, composites of blastocyst- and miPSC-derived cells,

obtained from pancreas of Pdx1+/� chimeric mice (C57BL/6 3

DBA2 or C57BL/6 3 BDF1 F1 origin). This effect lasted only for

a short time, presumably due to immune rejection by the host

C57BL/6 diabetic mice (Figure 3E). These data strongly indicate

that the miPSC-derived pancreas, with islets, formed in an allo-

genic host is functional and that the ‘‘autologous’’ islets thus

formed can be used to treat diabetes, without rejection.

Blastocyst complementation thus permits demonstration of

proof-of-principle both for pancreas generation from PSCs and

for diabetic therapy using donor iPSC-derived syngenic islets.

Generation of Interspecific Chimeras between Mouseand RatOur second goal was to generate interspecific chimeras

between mouse and rat. To achieve this goal, we generated

not only miPSCs but also riPSCs and rESCs using an established

protocol (Hirabayashi et al., 2009; Ying et al., 2008). These

mouse and rat PSCs enabled us bidirectionally to generate inter-

specific chimeras.

We injected EGFP-marked GT3.2 miPSCs into rat blastocysts

(r-blastocysts) or EGFP-marked riPSCs (Figures S4A and S4B)

into mouse blastocysts (m-blastocysts). Because post-implan-

tation development reportedly is severely hampered after intra-

uterine transfer of xenogenic blastocysts (Rossant et al., 1982;

Tarkowski, 1962), injected r-blastocysts or m-blastocysts were

transferred, respectively, into the uteri of pseudopregnant rats

or mice. After intrauterine transfer of injected r-blastocysts or

m-blastocysts with development to the fetal stage, we evaluated

EGFP expression by fluorescence microscopy for each trans-

ferred embryo. EGFP-expressing cells were found in the body

of each injected conceptus, but never in placenta (Figure 4A).

B

D

C

DAPI InsulinDAPI / EGFP / Insulin

No. ofadult mice

No. of chimeras

Pdx1 genotype

+/-Injected

cell Parents+/+ -/-

A

Li

Pdx1-/- + miPSCs Pdx1

+/- + miPSCs

Ki

StPaSp

Pha

seE

GFP

Pdx1-/-

+ m

iPS

Cs

Pdx1+/

-+

miP

SC

s

Pdx1-/-+ miPSCs

Pdx1+/-+ miPSCs

(mg/dl)

0 30 60 90 120

Minutes after glucose administration

600

400

200

0

E

Pdx1-/- + miPSCs Pdx1

+/- + miPSCs

STZ-diabetic model

0 13 15GT3.2(miPSC)

(Pdx1-/- + mESCs)

x Pdx1+/- 29 28

( 0 : 1 : 1 )

27 18 9G4.2(mESC)

Pdx1+/-

x Pdx1+/-

5 4( 1 : 2 : 1 )

EGFP

DA

PI /

EG

FPH

E

Figure 2. Analysis of miPSC-Derived

Pancreas in Adult Mice: Derivation of

Pancreatic Functional Cells and Their

Response to Glucose Load

(A) Results of embryo manipulation and genotypes

of adult mice. Note that Pdx1�/� mice comple-

mented with mESCs served as founders and

generated knockout mice more efficiently.

(B) Macroscopic images at adult stage

(8–12 weeks) of miPSC-derived pancreas gener-

ated in a Pdx1�/� mouse (Pa in left panel) and of

composite pancreas generated in a Pdx1+/�

mouse (right). In left panel, Li = liver, Pa = pancreas,

St = stomach, Sp = spleen, and Ki = kidney.

(C) Immunohistological studies of sections ob-

tained from pancreas revealed clear differences

in the distributions of miPSC-derived cells.

Sections were stained with HE or for EGFP with

DAPI nuclear counterstaining.

(D) Immunohistological analysis of a generated

islet, showing the distribution of miPSC-derived

cells. Sections were stained for EGFP, with anti-

insulin antibodies, and with DAPI nuclear counter-

staining. Yellow lines show borders of endocrine

and exocrine tissues. See also Figure S3 for anal-

ysis of mESC-derived pancreas in adult mice.

(E) Results of GTT in Pdx1�/� (-) and Pdx1+/� (:)

mice (n = 6 each) complemented with miPSCs.

Mice with STZ-induced diabetes (d andA) served

as controls. Fasting after 20 hr, blood was sampled

via tail vein at intraperitoneal glucose administra-

tion (1 g/kg; 0 min) and 15, 30, 60, and 120 min

thereafter.

Sections in (C) were observed under fluorescence

microscopy and in (D) were observed under

confocal laser scanning microscopy. Scale bars

in (C), 100 mm; in (D), 50 mm. Error bars in (E) indi-

cate ± SD.

790 Cell 142, 787–799, September 3, 2010 ª2010 Elsevier Inc.

Page 159: Cell 100903

This finding indicates that injected mouse or rat iPSCs can

contribute to xenogenic development, with generation of inter-

specific chimeras.

Next, we tried to quantitate contribution of mouse or rat iPSCs

to these interspecific chimeras. Chimerism in interspecific

embryos appeared to vary individual-to-individual and organ-

to-organ. Since quantitation of PSC-derived cells was difficult

in organs, we analyzed embryonic fibroblasts and hematopoietic

cells. FACS analysis of embryonic fibroblasts revealed that

donor-derived EGFP+ cell percentages of about 28.0% and

26.5%, respectively, were detected in mouse and rat inter-

specific chimeras (representative FACS data shown in Fig-

ure 4B). We also examined chimerism in hematopoietic cells

by staining cells from livers of interspecific chimera fetuses

with antibodies specific for mouse and rat CD45 antigens.

Cells that expressed mouse or rat CD45 represented distinct

populations in interspecific chimeras, with only cells derived

from injected iPSCs expressing EGFP (Figure 4B). Whereas

a high proportion (28.3%) of mouse blood cells was detected

in r-blastocyst-derived chimeric fetal liver, rat blood cells were

B DA C

Minutes after glucose administration

F

1200 30 60 900

200

400

600(mg/dl)

(min)

Phase

EGFP

DAPI / EGFP / InsulinPhase

EGFP EGFPDAPI

HEHE

Pdx1+/- + miPSC derived islet transplantation Sham transplantation (NC)Pdx1-/- + miPSCs derived islet transplantation Syngenic islet transplantation (PC)

E

Days after islet transplantation

0

200

400

600

0 10 20 30 40 50 60

(mg/dl)

(days)

Blo

od g

luco

se

Blo

od g

luco

se

Figure 3. Transplantation of ‘‘Autologous’’ Islets from iPSC-Derived Pancreas: A Therapeutic Model

(A) EGFP-positive islets isolated from miPSC-derived pancreas. Via the common bile duct and pancreatic duct, miPSC-derived pancreas was perfused by colla-

genase. Density gradient centrifugation isolated a fraction enriched in EGFP-expressing islets.

(B) Kidney capsule 2 months after islet transplantation. At the site of transplantation (arrow), miPSC-derived islets still expressed EGFP.

(C) Renal subcapsular transplantation site (B, arrow); sections were stained with HE or immunostained for EGFP with DAPI nuclear counterstaining.

(D) Transplanted miPSC-derived islets in (C), immunostained for EGFP and insulin with DAPI nuclear counterstaining.

(E) Transplantation of miPSC-derived islets into mice with STZ-induced diabetes. Each mouse received 150 islets. Arrows indicate time points at which an mAb

cocktail (anti-INF-g, anti-TNF-a, anti-IL-1b) was administered. Nonfasting blood glucose levels were measured weekly for 2 months after transplantation.

Glucose levels are shown for STZ-induced diabetic mice transplanted with miPSC-derived islets (green; n = 3), with islets derived from Pdx1+/� chimeric

mice (orange; n = 2), with islets derived from syngenic strain of host strain (C57BL/6) as positive control (blue; n = 2), and with sham transplantation control

mice as negative control (purple; n = 3).

(F) Results of GTT 2 months after islet transplantation. Donors were Pdx1�/� chimeric mice and Pdx1+/� chimeric mice, with blood sampling at the same time

points as in Figure 2E.

Sections in (C) were observed under fluorescence microscopy and in (D) were observed under confocal laser scanning microscopy. Scale bar in (D), 50 mm. Error

bars in (E) and (F) indicate ± SD.

Cell 142, 787–799, September 3, 2010 ª2010 Elsevier Inc. 791

Page 160: Cell 100903

only rarely present (less than 3.3%) in m-blastocyst-derived

chimeric fetal liver (Figure 4B). This tendency was specific to

interspecific chimeras and not observed in intraspecific

chimeras (Figure 4E). It is not clear why the difference in contri-

bution of iPSCs to hematopoietic cells between mouse and rat

interspecific chimeras was so marked.

To further confirm interspecific chimerism, genomic DNA

extracted from FACS-sorted cells expressing CD45 was PCR-

amplified using primers common to the mouse and rat Oct3/4

loci (Figure 4C). PCR products of different lengths, indicating

origin in each species, were clearly present (Figure 4D). These

results strongly indicated that the animals harboring these cells

were mouse/rat interspecific chimeras.

To investigate the influence of iPSC contribution to xenogenic

development at the fetal stage, we assessed embryonic devel-

opment rate of interspecific chimeras, and the extent of chime-

rism, by FACS analysis using established embryonic fibroblasts.

Both embryonic development rate and degree of chimerism

were lower in interspecific chimeras than in intraspecific

chimeras (Figures 4F and 4G). In addition, high contributions

by xenogenic cells appeared to be associated with morpholog-

ical abnormalities and embryonic lethality (data not shown). To

exclude the possibility that these abnormalities were caused by

donor iPSCs, we also attempted to generate interspecific

chimeras using mouse or rat ESCs. DsRed-marked EB3DR

mESCs could also generate interspecific rat chimeras (top

panels in Figure S5A), with embryonic development rate and

degree of chimerism similar to those generated by miPSC

injection (Figure S5B). The Venus-marked WIv3i-1 and -5

rESC lines, with high contribution to rat embryo development

and germline competency (Hirabayashi et al., 2009), could

also generate interspecific chimeras after injection into m-blas-

tocyst (middle and bottom panels in Figure S5A), but embry-

onic development rate and degree of chimerism were lower

than reported for intraspecific chimeras (Figure S5C). These

results suggest that generation of interspecific chimeras

between mouse and rat is less efficient than generation of intra-

specific chimeras.

A

Pha

seE

GFP

rBL + miPSCs

4100 101 102 103 10

EGFP

28.0

4100 101 102 103 10

EGFP

26.5

mBL + riPSCs

100

101

102

103

104

100 101 102 103 104

19.3

30.0

1.7

49.0

98.8

100 101 102 103 104

mC

D45

-AP

C

rCD45-PE EGFP

100

101

102

103

104

100 101 102 103 104

6.68

93.1

0.02

0.23

87.0

100 101 102 103 104

mC

D45

-AP

C

rCD45-PE EGFP

Fetal liverEmbryonic fIbroblastB

rBL

+ m

iPS

Cs

mB

L +

riPS

Cs

100

75

0

25

50

(%)

Rat Mouse Mouse Rat

miPSC riPSC

Embryo

24 26 109 30No. of

transferredembryoa

Cell

ChimeraNon-chimeraAbortedR

elat

ive

frequ

ency

(tot

al tr

ansf

erre

d em

bryo

s =

100%

)

100

75

0

25

50

(%)

Rat Mouse Mouse Rat

miPSC riPSC

Embryo

Cell

n = 6 n = 12 n = 23 n = 19E F G

100

75

0

25

50

(%)

EF CD45+in FL EF CD45+

in FLrBL+riPSCsmBL+riPSCs

n = 16 n = 19

No te

mpla

te

Mar

ker

mCD45

+rC

D45+

mCD45

+rC

D45+

mCD45

+ in

mPB

rCD45

+ in

rPB

Exon 2 Exon 3 Exon 4

C DrBL

+ miPSCsmBL

+riPSCs

rOct3/4 locusmOct3/4 locus

Control

rOct3/4 locus = 966bpmOct3/4 locus = 861bp

Figure 4. Generation of Interspecific

Chimeras between Mouse and Rat

(A) Interspecific chimera fetuses generated by

injection of miPSCs into r-blastocyst (rBL: left

panels) and by injection of riPSCs into m-blasto-

cyst (mBL: right panels). With rat embryo manipu-

lation, fetuses were analyzed 12 days after embryo

transfer into uteri of 3.5 dpc pseudopregnant rats

(embryonic day (E) 15.5). With mouse embryo

manipulation, fetuses were analyzed 11 days after

embryo transfer into uteri of 2.5 dpc pseudopreg-

nant mice (E13.5). See also Figure S4 for charac-

teristics of riPSCs and interspecific embryos after

injection of riPSCs into mouse 8-cell/morula stage

embryos.

(B) Representative data of FACS analysis on cells

from fetal liver and on embryonic fibroblasts

derived from interspecific chimeras. Right panels

show cells from fetal liver immunostained with

anti-mCD45 and -rCD45 antibodies. Note that

anti-mouse or -rat monoclonal antibodies (mAb)

against CD45 can distinguish CD45-expressing

hematopoietic cells in a species-specific manner.

Almost all CD45-expressing cells derived from

injected cells express EGFP, indicating iPSC

origin.

(C) Schema of mouse and rat Oct3/4 loci (mOct3/4

and rOct3/4). Of note is that in the rat Oct3/4 locus

the 2 introns flanking exon 3 are longer than in the

mouse Oct3/4 locus. This difference in length can

distinguish origins of otherwise similar cells.

Arrowheads indicate common primers of each

species for PCR. PCR product sizes for both

species are shown below.

(D) Results of genotyping to identify origin of hematopoietic cells expressing mouse or rat CD45 in fetal liver of interspecific chimeras. Peripheral-blood CD45-

expressing cells from each species served as positive controls.

(E) Correlation of chimerism between embryonic fibroblasts and CD45+ hematopoietic cells in fetal liver. Cells were prepared from interspecific chimeras gener-

ated by injection of riPSCs into m-blastocysts or from intraspecific chimeras generated by injection of riPSCs into r-blastocysts.

(F) Relative frequencies of aborted, nonchimeras, and chimeras embryonic development (E13.5 and 15.5).

(G) Chimerism analysis of embryonic fibroblasts from chimeras generated by injection of miPSCs or riPSCs into mouse or rat blastocysts. Fibroblasts were

obtained from chimeras and analyzed for EGFP intensity by FACS. Plotted dots show degrees of chimerism for individual embryos. See also Figure S5 for inter-

specific chimeras using ESCs.

Scale bars in (A), 5 mm.

792 Cell 142, 787–799, September 3, 2010 ª2010 Elsevier Inc.

Page 161: Cell 100903

Interspecific Chimeras Were Live-born; Some Grewinto AdulthoodTo investigate the developmental potential of generated

chimeras and to assess the functionality of the cells, tissues, or

organs derived from injected cells, we analyzed interspecific

chimeras at neonatal and adult stages. Mouse- or rat-iPSC-

injected interspecific chimeras survived after birth and expressed

EGFP ubiquitously (as did intraspecific chimeras; Figure 5A,

r-blastocyst + miPSCs: n = 5, m-blastocyst + riPSCs: n = 10).

As these chimeras developed into adulthood, chimerism could

be judged by coat color because miPSCs (C57BL/6, black

coat) were injected into r-blastocyst (Wistar, white coat) or riPSCs

(Wistar) were injected into m-blastocyst (BDF11 3 C57BL/6,

black coat) (Figure 5C, r-blastocyst + miPSCs: n = 8, m-blasto-

cyst + riPSCs: n = 4). The full-term development rate of inter-

specific chimeras, either with miPSCs into r-blastocyst or with

riPSCs into m-blastocyst, was �20%. In both settings it was

lower than that for intraspecific chimeras (�50%).

Determination of Body Size in Interspecific ChimerasAdult rats typically are ten times bigger than adult mice, whereas

newborn rats are three times bigger than newborn mice.

Because mouse and rat gestations are of similar length (19 and

21 days, respectively), organogenesis requires more cell prolifer-

ation and differentiation during rat development than during

mouse development. What determines the size of interspecific

Heart

KidneyLiver

Pancreas

DAP

I / E

GFP

Heart

Liver Kidney

Pancreas

A BrBL + miPSCs mBL + riPSCs

C

E

D

7.5

0

2.5

5.0

n = 4 n = 4 n = 8 n = 7

rBL+ miPSCs

mBL+ riPSCs MouseRat

Body

wei

ght

(g)

DAPI / EGFP / MVH

Testis (P1)DAPI / EGFP / Insulin

Pancreas (P1)DAPI / EGFP / Albumin DAPI / EGFP / CK19

Liver (Adult)DAPI / EGFP / MVH

F

Gonad (♀ E13.5)rBL + miPSCs rBL + miPSCs mBL + rESCs

Phas

eEG

FP

rBL + miPSCs mBL + riPSCs

rBL + miPSCs mBL + riPSCs

Non-chimeraNon-chimera

Phase

Non-chimera

EGFP

Figure 5. Postnatal Analysis of Interspecific

Chimeras

(A) Photograph of newborn interspecific chimeras

derived from r-blastocyst injected with miPSCs

and m-blastocyst injected with riPSCs.

(B) Neonatal body weights of interspecific

chimeras were measured and plotted. One

chimera obtained after injection of miPSCs into

r-blastocyst showed a high contribution of mouse

cells as shown in the insert, with body weight and

size equivalent to those of newborn mouse (right

panel).

(C) Photograph of interspecific chimera at 8 weeks

of age.

(D) Analysis of chimerism in neonates by organ or

tissue. Interspecific chimeras generated by miPSC

injection into r-blastocyst (left panels) and by

riPSC injection into m-blastocyst (right panels)

are shown. Sections of representative organs

(heart, liver, pancreas, and kidney) after immunos-

taining for EGFP antibody with DAPI nuclear coun-

terstaining are shown.

(E) Neonatal pancreas and adult liver of inter-

specific chimera generated from r-blastocyst

injected with miPSC, immunostained for EGFP

and insulin or for EGFP and albumin or CK19

with DAPI nuclear counterstaining. In pancreas,

an islet-like cell cluster contains cells producing

insulin composed of both miPSC-derived cells

that express EGFP and r-blastocyst-derived cells

that do not. In liver, albumin-positive hepatocytes

and CK19-positive cholangiocytes (arrowhead)

also express EGFP, indicating miPSC origin.

(F) Distribution of xenogenic cells in neonatal testis

and in developing gonad. MVH-positive germ cells

did not express EGFP, indicating nonxenogenic

cell origin.

Sections in (D) were observed under fluorescence

microscopy and in (E) and (F) were observed under

confocal laser scanning microscopy. See also

Figure S6 for macroscopic images of neonates

and results of peripheral blood analysis of adults.

Scale bars in (A) and (B), 10 mm; in (D), 100 mm;

(E) and (F), 50 mm.

Cell 142, 787–799, September 3, 2010 ª2010 Elsevier Inc. 793

Page 162: Cell 100903

chimeras is an intriguing biological question. Interestingly, body

size and weight of interspecific chimeras born from rat foster

mothers were similar to those of normal newborn rats, whereas

for those born from mouse foster mothers they were similar to

those of newborn mice (Figures 5A and 5C). However, high

contributions by xenogenic cells may also affect interspecific

chimera size. One chimera obtained after injection of miPSCs

into r-blastocysts showed body weight and size equivalent to

those of a newborn mouse, although its birth mother was a rat

(Figure 5B). This particular chimera’s donor miPSC-derived cell

contribution was extremely high (Figure 5B, inlet). However,

chimeras with high contributions of xenogenic iPSC-derived

cells generally were not identified, suggesting an association

with embryonic lethality. In most chimeras, the size of newborns

seemed to conform with that in the species from which the blas-

tocyst originated. A correlation between body weight and contri-

bution of donor iPSC-derived cells as estimated from hair color

and peripheral blood cells is shown in Figure 6. The origins of

placenta and uterus may have a key role in size determination.

Injected xenogenic PSCs never developed into placenta, which

was always of host blastocyst origin (Figure 4A, Figure S5A).

Therefore it is difficult to determine whether it is placenta or

uterine environment that influences the size of embryos.

Distribution of Donor iPSC-Derived Cellsin the Xenogenic EnvironmentWe determined the distribution of mouse- or rat-iPSC-derived

EGFP-positive cells in neonatal interspecific chimeras. With

miPSC injection into r-blastocyst (Figure S6A) and with riPSC

injection into m-blastocyst, almost all organs contained EGFP-

mC

D45

-AP

C

rCD45-PE

0

200

400

600

(g)

4 9 14

Bod

y w

eigh

t

Weeks after birth(wk)

Littermate #1Littermate #2Chimera #2Chimera #4

0

200

400

600

(g)

4 9 14

Bod

y w

eigh

t

Weeks after birth(wk)

Littermate #3Littermate #4Chimera #1Chimera #3

B

A

0102

103

104

105

0 102 103 104 105

55.5

0.27

1.3

42.9

Chimera #4

0102

103

104

105

0 102 103 104 105

5.35

0.24

0.29

94.1

Chimera #1

0102

103

104

105

0 102 103 104 105

15.6

0.05

2.06

82.3

Chimera #2

0102

103

104

105

0 102 103 104 105

17.8

1.28

2.62

78.3

Chimera #3

Figure 6. Correlation between Body Weight

and Chimerism in Interspecific Chimeras

(A) Photographs of interspecific chimeras derived

from r-blastocyst injected with miPSCs, with FACS

analysis of their chimerisms in peripheral blood

mononuclear cells.

(B) Body weight change after weaning for inter-

specific chimeras (green) and their nonchimeric

littermates (orange).

positive cells (Figure S6B). In sections

immunostained with anti-EGFP antibody,

representative images demonstrated that

EGFP-positive cells become various

types of tissues (Figure 5D). Of particular

note was a pancreatic islet generated by

injection of miPSCs into r-blastocyst.

Cells in this islet marked immuno-

histochemically for insulin; the islet was

a composite of EGFP-positive and -nega-

tive cells, indicating that the islet con-

sisted of host rat cells and exogenous

donor mouse cells (Figure 5E). Islet poly-

clonality also was seen with intraspecific

chimeras (Figure 2E). The expression of

other functional molecules in xenogenic

cells (i.e., hepatocytes or cholangiocytes

in the liver or leukocytes in the peripheral blood) was also

detected (Figure 5E and Figure S6C).

It is known that rats do not have gallbladders, whereas mice

do. To date, when mPSCs were injected into r-blastocysts, the

interspecific chimeras produced (generally the size of rats)

have not had gallbladders (n = 8). In contrast, interspecific

chimeras generated from m-blastocysts complemented with

rPSCs (more like mice in size) have had gallbladders (n = 4).

To see the xenogenic contribution to germ cells, testis and

developing gonad of the interspecific chimeras were examined.

Cells were not found that coexpressed EGFP or Venus and the

germ cell marker mouse vasa homolog (MVH) (Figure 5F). As

the injected miPSCs or rESCs were both confirmed as germ-

line-competent pluripotent stem cells in intraspecific settings

(data not shown), germ cell development may be impaired or

more limited in the xenogenic environment.

Generation of Rat Pancreas in Mouse via InterspecificBlastocyst ComplementationOur last goal was to generate xenogenic rat pancreas in Pdx1�/�

mice by interspecific blastocyst complementation. For efficient

production of Pdx1�/� mice, embryos were generated by inter-

cross of Pdx1�/� founder male mice (in which pancreas arose

principally from exogenous miPSCs) with Pdx1+/� female mice.

With this founder system, half the mice born were Pdx1�/�

(in contrast to intercross of Pdx1+/� heterozygotes in which

only 25% of offspring were Pdx1�/�). For donor riPSCs, we

used the riPSC#3 line (Figure S4A). This line was selected among

11 established riPSC clones for embryonic developmental rate

and degree of chimerism after injection into mouse embryos

794 Cell 142, 787–799, September 3, 2010 ª2010 Elsevier Inc.

Page 163: Cell 100903

(data not shown). After injection, 139 embryos were transferred

into uteri of pseudopregnant mice and 34 mice were born.

They were analyzed at neonatal and adult stages.

The contribution of EGFP-marked riPSC-derived cells in the

pancreas of neonatal Pdx1+/� interspecific chimeras was small

relative to that of host-derived cells, as seen in whole-body

chimerism (Pdx1+/� + riPSCs in Figure 7A bottom panel; n = 5).

In contrast, the pancreatic epithelia in Pdx1�/� interspecific

chimeras were entirely composed of EGFP-marked riPSC-

derived cells (Pdx1�/� + riPSCs in Figure 7A top panel; n = 10).

Each genotype was confirmed by PCR using genomic DNA

extracted from FACS-sorted mouse CD45 (mCD45)-positive

splenocytes (Figure S7A). Neonates with entirely EGFP-positive

pancreata thus were clearly identified as of Pdx1�/� genotype

(Figure S7B). The existence of interspecific chimeras between

mouse and rat was also confirmed by FACS patterning, which

demonstrated distinct populations of mCD45- and rat CD45

(rCD45)-positive cells, with only rCD45-positive cells expressing

EGFP after riPSC injection (Figure S7A). FACS-sorted mCD45-

and rCD45-positive cells were also confirmed as, respectively,

mouse or rat in origin by genomic PCR testing using Oct3/4

locus primers (Figure S7C), which clearly identified cell origin.

On immunostaining, riPSC-derived pancreas expressed EGFP

almost universally (Figure 7B) and also expressed a-amylase

(an exocrine tissue marker) and insulin, glucagon, and somato-

statin (endocrine tissue markers; Figure 7B).

As in wild-type experiments, successful maturation into adult-

hood (8 weeks) was uncommon in Pdx1�/� mice complemented

with riPSCs; however, adult mice with riPSC-derived pancreas

(Figure 7C; n = 2) had intact pancreas expressing EGFP (Figures

7E and 7F). Quantitative analysis of the sections by image J soft-

ware revealed percentages of EGFP-positive cells to be 81.9% ±

3.4%. Additionally, on GTT in adulthood, insulin was secreted in

response to glucose loading and normal serum glucose levels

were maintained (Figure 7D). These results indicate that genera-

tion of a xenogenic iPSC-derived organ is possible via interspe-

cific blastocyst complementation.

DISCUSSION

We report three innovative observations, using proof-of-prin-

ciple approaches. (1) If an empty developmental niche for an

organ is provided (as with the Pdx1�/� mouse and the pancreatic

niche), PSC-derived cellular progeny can occupy that niche

and developmentally compensate for the missing contents of

the niche, forming an organ almost entirely composed of

cells derived from donor PSCs. (2) Generation of interspecific

chimeras between mouse and rat is possible with injection of

A

GlucagonDAPI

SomatostatinDAPI

EGFPDAPI

α-AmylaseDAPI

B

E

D

(min)

Pdx1-/-+ riPSCs

Pdx1+/-+ riPSCs

(mg/dl)

0 30 60 90 120

Minutes after glucose administration

600

400

200

0

Phase EGFP

Pdx1

+/- +

riPS

Cs

Pdx1

-/- +

riPS

Cs

C

Phase

Pdx1

+/- +

riPS

Cs

Pdx1

-/- +

riPS

Cs

PaKi

Li

SpDu

St

FEGFPDAPI

HE

Pdx1-/- + riPSCs

InsulinDAPI

EGFP

Figure 7. Generation of Rat Pancreas in

Pdx1-Deficient Mouse by Interspecific Blas-

tocyst Complementation

(A) riPSC-derived pancreas in neonatal Pdx1�/�

mouse. In Pdx1�/� mice complemented with

riPSCs, almost all pancreata expressed EGFP,

indicating riPSC origin (top panel). On the other

hand, in Pdx1+/� mice complemented with riPSCs,

pancreata only partially expressed EGFP, indi-

cating a composite of both riPSC- and mouse-

derived cells (bottom panel).

(B) Immunohistological studies of sections ob-

tained from riPSC-derived pancreas. Serial frozen

sections were immunostained for EGFP, a-

amylase, insulin, glucagon, and somatostatin

with DAPI nuclear counterstaining. See also

Figure S7 for genotyping to identify Pdx1 status

and origin of each species in interspecific chimera.

(C) Photograph of adult Pdx1�/� mouse comple-

mented with riPSCs.

(D) Results of GTT in Pdx1�/� (-) and Pdx1+/�

(:, A) mice complemented with riPSCs.

(E) riPSC-derived pancreas in adult Pdx1�/�

mouse. Compare similar results of neonatal anal-

ysis (Figure 5D).

(F) Histological studies of sections obtained from

riPSC-derived adult pancreas revealed clear

differences in the distributions of riPSC-derived

cells after staining by anti-EGFP antibody with

DAPI nuclear counterstaining (right panel).

Sections were also stained with HE.

Sections in (F) were observed under light or fluo-

rescence microscopy and in (B) were observed

under confocal laser scanning microscopy. Scale

bars in (B) and (F), 100 mm.

Cell 142, 787–799, September 3, 2010 ª2010 Elsevier Inc. 795

Page 164: Cell 100903

mouse or rat PSCs into embryos from the other species; injected

PSC-derived cells are distributed throughout the body and

appear to function normally. (3) The combination of (1) and (2)

successfully generates rat pancreas in mouse with injection of

riPSCs into Pdx1�/� mouse embryos, a technique that we term

‘‘interspecific blastocyst complementation.’’

We demonstrated that the Pdx1�/� mice derived from blasto-

cysts complemented with mPSCs were born with functional

pancreas almost entirely derived from donor PSCs and grew

into adulthood without showing any evidence of pancreatic

insufficiency. Several groups have used the same technique to

study the development of thymic epithelium (Muller et al.,

2005), to compensate for cardiac defects (Fraidenraich et al.,

2004), or to determine if yolk sac hematopoiesis and germ cell

development are of clonal or nonclonal origin (Ueno et al.,

2009; Ueno and Weissman, 2006). Although Stanger et al. tested

development of pancreas and liver in embryos to define organ

size determinants (Stanger et al., 2007), no study has exploited

this technique to produce donor-derived functional organs and

rescued a lethal phenotype to adulthood.

Direct in vitro differentiation of insulin-producing cells from

PSCs has been a major focus of stem cell therapy, as recently

demonstrated by a sophisticated protocol to generate pancre-

atic endoderm efficiently via stepwise endodermal differentia-

tion (Kroon et al., 2008). However, the in vitro generation

of insulin-producing cells still needs further improvement in

differentiation efficiency, in insulin production levels, or in speed

of insulin response to glucose changes. In addition, the risk

of tumor development due to contamination with undifferenti-

ated PSCs must be rigorously assessed before clinical use.

Compared with those generated in vitro, insulin-producing cells

obtained from pancreas that is formed in vivo by blastocyst

complementation must have gone through near-normal differ-

entiation processes with proper epigenetic changes. The

tissues obtained, such as insulin-producing cells, thus are

presumed to be fully functional and the risk of teratoma devel-

opment due to contamination of undifferentiated PSCs to be

negligible. Nonetheless, the oncogenicity of iPSC-derived cells

due to reactivation of introduced genes (Miura et al., 2009;

Nakagawa et al., 2008) or to genome abnormality due to long-

term culture remains to be assessed. To establish iPSCs without

genomic integration of a retroviral sequence should further

reduce the risk associated with the use of iPSCs (Okita et al.,

2008).

We assumed that aggregation of early embryos would lead

to the presence in trophectoderm of xenogenic cells that are

reported to be harmful to embryonic development after uterine

implantation. We therefore injected mESCs/iPSCs, but not blas-

tomere cells, into r-blastocysts. As predicted, EGFP-positive

PSC-derived cells were not detected in placentas (Figure 4A

and Figure S5A), and we succeeded in generating interspecific

chimeras. To confirm this further, we then attempted to generate

interspecific chimeras by injecting rESCs/iPSCs into m-blasto-

cysts. After injection into 8-cell/morula stage mouse embryos,

the riPSCs were eventually enclosed within the inner cell mass

of the m-blastocyst and were never detected in the m-blastocyst

trophectoderm (Figures S4C and S4D). Our study, consistent

with others, indicates that the presence of xenogenic cells

among extraembryonic lineage cells, with exposure of xeno-

genic cells to the uterine environment, is inhibitory to implanta-

tion and/or to further intrauterine development of interspecific

chimeras.

It is of prime interest that body size and weight of interspecific

chimeras conformed with those of the species of the foster

mother (Figures 5A–5C). What xenogenic components contribute

to the phenotypic determination of interspecific chimeras?

It seems that placenta and/or uterine environment are respon-

sible for size determination of adult interspecific chimeras

both as embryos and as adults. Given that placenta must be of

the same origin as the foster mother for successful generation

of interspecific chimeras, it is not clear which is primarily respon-

sible for this determination. Degree of chimerism may also

influence the phenotype. As observed in intraspecific chimeras,

contribution and distribution of xenogenic cells in interspecific

chimeras vary organ-to-organ. There is a negative correlation

between contribution of donor (mouse) PSC-derived cells and

body weight (Figure 6). Although we could obtain interspecific

chimeras consistently (Figure 4 and Figure 5), embryonic lethality

was high and postnatal development was poor. Some interac-

tions between mouse- and rat-derived cells indispensable

for organism survival may not work across species, resulting in

death during embryonic development or in retarded postnatal

development.

Another example is the formation of gallbladders in inter-

specific chimeras. Those derived from m-blastocysts have

gallbladders but those from r-blastocysts do not. It is conceiv-

able that the temporo-spatial development of donor PSC-

derived cells is regulated by the xenogenic host microenviron-

ment, which governs morphogenesis and organogenesis.

However, the data also indicate that the intrinsic develop-

mental program imprinted in PSCs may create competition

between host blastocyst-derived and donor PSC-derived cells

to form chimeric organs. The balance between host and donor

cells at certain critical points during embryonic development

thus may be important. Xenogenic developmental systems

may be useful to elucidate these and other developmental

questions and to help in generation of chimeras between

species evolutionarily more distant from one another than

are mouse and rat.

Most importantly, we succeeded in generating functional rat

pancreas in Pdx1�/� mice via interspecific blastocyst comple-

mentation. In all interspecific neonates derived from Pdx1�/�

blastocysts injected with riPSCs, pancreas was present.

Although full maturation into adulthood was not common, once

the mice matured into adulthood, the generated riPSC-derived

pancreas was morphologically and histologically normal and

was not associated with any sign of diabetes or other abnormal-

ities; GTT results strongly indicated normal function. Generation

of functional cells (sperm, hepatocytes) in xenogenic environ-

ments has been reported (Mercer et al., 2001; Shinohara et al.,

2006). In addition, hematopoietic xeno-chimeras have been

used commonly as a method to study functionality of hematopoi-

etic stem cells (Kamel-Reid and Dick, 1988; McCune et al.,

1988). No study, however, has demonstrated generation in

a xenogenic environment of a PSC-derived functional organ

that can rescue embryonic lethality to adulthood.

796 Cell 142, 787–799, September 3, 2010 ª2010 Elsevier Inc.

Page 165: Cell 100903

The organ generation system described may be applied to

treat organ failure in humans if pigs or other large animals are

used. There are, however, several issues that need to be

addressed to bring this principle into the clinic. For example,

though we were able to generate interspecific chimeras between

mouse and rat, their embryonic lethality is high and maturation

into adulthood is uncommon. The nature of this xenogenic barrier

is not clear, but it is evident that the evolutionary distance

accounts for this, as we do not see these problems in intraspecific

chimeras. Livestock animals such as pigs or sheep may be too

distant evolutionarily for successful complementation. In addi-

tion, as described in the allogenic system, vessels, nerves, and

some interstitial elements that are not under the influence of

Pdx1 expression were composites of host- and miPSC- or

mESC-derived cells. Although we showed in this study that islets

prepared from miPSC-derived pancreas generated in allogenic

hosts indeed were successfully transplanted into ‘‘autologous’’

mice with STZ-induced diabetes without rejection, whether the

same principle applies to transplantation of islets obtained from

pancreas generated in xenogenic animals remains to be seen.

Transplantation of islets from rat pancreas generated in Pdx1�/�

mice into diabetic rats should answer this question. However,

due to the size difference between mouse and rat, and to high

embryonic lethality and poor postnatal maturation of interspecific

chimeras, it is not possible to obtain sufficient numbers of islets to

treat diabetic rats. Generation of mouse pancreas in Pdx1�/� rats

will be necessary to do such experiments.

Production of organ-deficient livestock animals and genera-

tion of chimeras is another issue, but nuclear transfer technology

available for livestock animals may permit establishment of

organ-deficient pig lines, for example (Lai et al., 2002). The

successful generation of pig chimeras using blastocyst injection

has also been reported (Nagashima et al., 2004). The major diffi-

culty seems to be that primate and rodent PSCs differ (Nichols

and Smith, 2009); limits of primate PSCs in contributing to

embryo development have been suspected. Poor contribution

of human ESCs to embryo development after injection into

mouse blastocysts or chick embryo has been demonstrated

(Goldstein et al., 2002; James et al., 2006). Generation of inter-

specific chimera technology may prove a tool useful in assess-

ment of pluripotent stem cell potential, thereby addressing this

issue.

Another issue of concern is the fact that PSC-derived cells

are found not only in pancreas but in all organs and tissues,

including brain and gonads. Therefore, without proper control

of the differentiation potential of PSCs, generation of human

organs in livestock animals will face an ethical issue. There are

several approaches to address this. One is use of committed

stem or progenitor cells in place of PSCs. If they are introduced

into an appropriate microenvironment at an appropriate devel-

opmental time point, to restrict differentiation toward a particular

organ may be possible. An alternative is to use genetically modi-

fied PSCs whose differentiation potential is restricted to certain

tissues or organs.

In conclusion, the approach described here will be of use not

only for better understanding of the mechanism of organogen-

esis but also as an initial step toward the ultimate regenerative

medicine of the future.

EXPERIMENTAL PROCEDURES

Animals

C57BL/6NCrSlc, BDF1, DBA/2CrSlc, ICR mice, and Wistar rats were

purchased from SLC Japan (Shizuoka, Japan). Pdx1-LacZ heterozygous

mice (Offield et al., 1996), kindly provided by Dr. Y. Kawaguchi (Kyoto

University) and Dr. C.V. Wright (Vanderbilt University), were crossed with

C57BL/6, DBA2, or BDF1 strain mice. C57BL/6 mice were given STZ (Sigma,

St. Louis, MO, USA) to induce diabetes. Mice with nonfasting blood glucose

levels > 400 mg/dL 1 week after STZ administration (200 mg/kg) were regarded

as hyperglycemic and thus as diabetic mice. All experiments were performed

in accordance with the animal care and use committee guidelines of the Insti-

tute of Medical Science, University of Tokyo.

Culture of ESCs/iPSCs

Undifferentiated mESCs were maintained on gelatin-coated dishes without

feeder cells in Glasgow’s modified Eagle’s medium (Sigma) supplemented

with 10% fetal bovine serum (FBS; Nichirei Bioscience, Tokyo, Japan),

0.1 mM 2-mercaptoethanol (Invitrogen, San Diego, CA, USA), 0.1 mM nones-

sential amino acids (Invitrogen), 1 mM sodium pyruvate (Invitrogen), 1%

L-glutamine penicillin streptomycin (Sigma), and 1000 U/ml of mouse leukemia

inhibitory factor (LIF; Millipore, Bedford, MA, USA). The G4.2 mESCs and

EB3DR mESCs, kindly provided by Dr. H. Niwa (Center for Developmental

Biology, RIKEN), were derived from EB3 mESCs (Niwa et al., 2000) and carried

the CAG promoter-driven EGFP or DsRed gene. Undifferentiated miPSCs

were maintained on mitomycin-c treated mouse embryonic fibroblasts

(MEFs) in Dulbecco’s modified Eagle’s medium (Invitrogen) supplemented

with 15% knockout serum replacement (Invitrogen), 0.1 mM 2-mercaptoetha-

nol, 0.1 mM nonessential amino acids, 1 mM HEPES buffer solution (Invitro-

gen), 1% L-glutamine penicillin streptomycin, and 1000 U/ml of mouse LIF.

The GT3.2 miPSCs were generated from TTFs of a male EGFP Tg mouse

(kindly provided by Dr. M. Okabe, Osaka University) by introducing three

factors (Klf4, Sox2, Oct3/4) in retroviral vectors (Okabe et al., 2009). The

GT3.2 miPSCs ubiquitously express EGFP under the control of the CAG

promoter.

Undifferentiated rESCs/iPSCs were maintained on mitomycin-c treated

MEFs in N2B27 medium (Ying et al., 2003) containing 1 mM MEK inhibitor

PD0325901 (Axon Groeningen, The Netherlands), 3 mM GSK3 inhibitor

CHIR99021 (Axon), with or without FGF receptor inhibitor SU5402 (Calbio-

chem, La Jolla, CA, USA), and 1000 U/ml of rat LIF (Millipore). The riPSCs

were generated from Wistar rat embryonic fibroblast by introducing three

mouse factors (Oct3/4, Klf4, Sox2) in lentiviral vectors as a doxycycline-induc-

ible expression unit using N2B27 medium containing the three inhibitors

described above. Expression of representative ESC marker genes has been

confirmed and teratoma formation by these riPSCs after injection into immu-

nodeficient mice has also been confirmed (our unpublished data). The riPSCs

ubiquitously express EGFP under the control of the Ubiquitin-C promoter.

The WIv3i-1 or WIv3i-5 rESCs were generated from Venus Tg rat blastocyst

(Hirabayashi et al., 2009). These rESCs ubiquitously express Venus under

the control of the CAG promoter.

Embryo Culture and Manipulation

Preparation of wild-type or Pdx1 heterozygous intercrossing embryos was

carried out according to published protocols (Nagy et al., 2003). In brief,

mouse 8-cell/morula stage embryos were collected in Medium 2 (Millipore)

from oviduct and uterus of mice 2.5 days postcoitum (dpc). These embryos

were transferred into potassium simplex optimized medium with amino acids

(Millipore) and were cultured for 24 hr for blastocyst injection.

Rat blastocysts were collected in a bicarbonate-buffered medium

composed of Roswell Park Memorial Institute medium (RPMI) 1640, Eagle’s

solution, and Ham’s F12 containing 18% FBS medium (Ogawa et al., 1971)

from oviduct and uterus of rats 4.5 dpc. These embryos were transferred

into modified rat 1-cell embryo culture medium (Oh et al., 1998) containing

80 mM NaCl (Wako Pure Chemical Industries, Osaka, Japan) and 0.1% poly-

vinyl alcohol (Sigma) and were cultured for about 1 hr until injection.

For micromanipulation, ESCs or iPSCs were trypsinized and suspended in

ESC or iPSC culture medium. A piezo-driven micromanipulator (Prime Tech,

Cell 142, 787–799, September 3, 2010 ª2010 Elsevier Inc. 797

Page 166: Cell 100903

Tokyo, Japan) was used to drill zona pellucida and trophectoderm under the

microscope and 10–15 ESCs or iPSCs were introduced into blastocyst cavities

near the inner cell mass. After blastocyst injection, embryos underwent follow-

up culture for 1–2 hr. Mouse blastocysts then were transferred into the uteri

of pseudopregnant recipient ICR female mice (2.5 dpc) and rat blastocysts

were transferred into the uteri of pseudopregnant recipient Wistar female

rats (3.5 dpc).

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.

2010.07.039.

ACKNOWLEDGMENTS

We thank Ryo Sumazaki for critical advice on this work, Kazuya Ise and

Hirofumi Noguchi for technical advice on islet isolation and islet transplanta-

tion, and Naoki Iwamori for technical advice on embryo manipulation. This

work was supported by grants from JST, the Ministry of Education, Culture,

Sport, Science, and Technology, Japan. H.N. is a founder and shareholder

of ReproCELL Inc.

Received: January 25, 2010

Revised: May 5, 2010

Accepted: June 30, 2010

Published: September 2, 2010

REFERENCES

Buehr, M., Meek, S., Blair, K., Yang, J., Ure, J., Silva, J., McLay, R., Hall, J.,

Ying, Q.L., and Smith, A. (2008). Capture of authentic embryonic stem cells

from rat blastocysts. Cell 135, 1287–1298.

Chen, J., Lansford, R., Stewart, V., Young, F., and Alt, F.W. (1993). RAG-2-

deficient blastocyst complementation: an assay of gene function in lympho-

cyte development. Proc. Natl. Acad. Sci. USA 90, 4528–4532.

Deltour, L., Leduque, P., Paldi, A., Ripoche, M.A., Dubois, P., and Jami, J.

(1991). Polyclonal origin of pancreatic islets in aggregation mouse chimaeras.

Development 112, 1115–1121.

Fehilly, C.B., Willadsen, S.M., and Tucker, E.M. (1984). Interspecific chimae-

rism between sheep and goat. Nature 307, 634–636.

Fraidenraich, D., Stillwell, E., Romero, E., Wilkes, D., Manova, K., Basson,

C.T., and Benezra, R. (2004). Rescue of cardiac defects in id knockout

embryos by injection of embryonic stem cells. Science 306, 247–252.

Goldstein, R.S., Drukker, M., Reubinoff, B.E., and Benvenisty, N. (2002). Inte-

gration and differentiation of human embryonic stem cells transplanted to the

chick embryo. Dev. Dyn. 225, 80–86.

Gotoh, M., Maki, T., Kiyoizumi, T., Satomi, S., and Monaco, A.P. (1985). An

improved method for isolation of mouse pancreatic islets. Transplantation

40, 437–438.

Hirabayashi, M., Kato, M., Kobayashi, T., Sanbo, M., Yagi, T., Hochi, S., and

Nakauchi, H. (2009). Establishment of rat embryonic stem cell lines that can

participate in germline chimeras at high efficiency. Mol. Reprod. Dev. 77, 94.

James, D., Noggle, S.A., Swigut, T., and Brivanlou, A.H. (2006). Contribution of

human embryonic stem cells to mouse blastocysts. Dev. Biol. 295, 90–102.

Kamel-Reid, S., and Dick, J.E. (1988). Engraftment of immune-deficient mice

with human hematopoietic stem cells. Science 242, 1706–1709.

Kroon, E., Martinson, L.A., Kadoya, K., Bang, A.G., Kelly, O.G., Eliazer, S.,

Young, H., Richardson, M., Smart, N.G., Cunningham, J., et al. (2008). Pancre-

atic endoderm derived from human embryonic stem cells generates glucose-

responsive insulin-secreting cells in vivo. Nat. Biotechnol. 26, 443–452.

Lai, L., Kolber-Simonds, D., Park, K.W., Cheong, H.T., Greenstein, J.L., Im,

G.S., Samuel, M., Bonk, A., Rieke, A., Day, B.N., et al. (2002). Production of

alpha-1,3-galactosyltransferase knockout pigs by nuclear transfer cloning.

Science 295, 1089–1092.

Li, P., Tong, C., Mehrian-Shai, R., Jia, L., Wu, N., Yan, Y., Maxson, R.E.,

Schulze, E.N., Song, H., Hsieh, C.L., et al. (2008). Germline competent embry-

onic stem cells derived from rat blastocysts. Cell 135, 1299–1310.

McCune, J.M., Namikawa, R., Kaneshima, H., Shultz, L.D., Lieberman, M., and

Weissman, I.L. (1988). The SCID-hu mouse: murine model for the analysis of

human hematolymphoid differentiation and function. Science 241, 1632–1639.

Mercer, D.F., Schiller, D.E., Elliott, J.F., Douglas, D.N., Hao, C., Rinfret, A.,

Addison, W.R., Fischer, K.P., Churchill, T.A., Lakey, J.R., et al. (2001). Hepatitis

C virus replication in mice with chimeric human livers. Nat. Med. 7, 927–933.

Miura, K., Okada, Y., Aoi, T., Okada, A., Takahashi, K., Okita, K., Nakagawa,

M., Koyanagi, M., Tanabe, K., Ohnuki, M., et al. (2009). Variation in the safety

of induced pluripotent stem cell lines. Nat. Biotechnol. 27, 743–745.

Muller, S.M., Terszowski, G., Blum, C., Haller, C., Anquez, V., Kuschert, S.,

Carmeliet, P., Augustin, H.G., and Rodewald, H.R. (2005). Gene targeting of

VEGF-A in thymus epithelium disrupts thymus blood vessel architecture.

Proc. Natl. Acad. Sci. USA 102, 10587–10592.

Nagashima, H., Giannakis, C., Ashman, R.J., and Nottle, M.B. (2004). Sex

differentiation and germ cell production in chimeric pigs produced by inner

cell mass injection into blastocysts. Biol. Reprod. 70, 702–707.

Nagy, A., Gertsenstein, M., Vintersten, K., and Behringer, R. (2003). Manipu-

lating the Mouse Embryo: A Laboratory Manual, Third Edition (Cold Spring

Harbor, NY: Cold Spring Harbor Laboratory Press).

Nakagawa, M., Koyanagi, M., Tanabe, K., Takahashi, K., Ichisaka, T., Aoi, T.,

Okita, K., Mochiduki, Y., Takizawa, N., and Yamanaka, S. (2008). Generation of

induced pluripotent stem cells without Myc from mouse and human fibro-

blasts. Nat. Biotechnol. 26, 101–106.

Nichols, J., and Smith, A. (2009). Naive and primed pluripotent states. Cell

Stem Cell 4, 487–492.

Niwa, H., Miyazaki, J., and Smith, A.G. (2000). Quantitative expression of Oct-

3/4 defines differentiation, dedifferentiation or self-renewal of ES cells. Nat.

Genet. 24, 372–376.

Offield, M.F., Jetton, T.L., Labosky, P.A., Ray, M., Stein, R.W., Magnuson,

M.A., Hogan, B.L., and Wright, C.V. (1996). PDX-1 is required for pancreatic

outgrowth and differentiation of the rostral duodenum. Development 122,

983–995.

Ogawa, S., Sato, K., and Hashimoto, H. (1971). In vitro culture of rabbit ova

from the single cell to the blastocyst stage. Nature 233, 422–424.

Oh, S.H., Miyoshi, K., and Funahashi, H. (1998). Rat oocytes fertilized in modi-

fied rat 1-cell embryo culture medium containing a high sodium chloride

concentration and bovine serum albumin maintain developmental ability to

the blastocyst stage. Biol. Reprod. 59, 884–889.

Okabe, M., Ikawa, M., Kominami, K., Nakanishi, T., and Nishimune, Y. (1997).

‘Green mice’ as a source of ubiquitous green cells. FEBS Lett. 407, 313–319.

Okabe, M., Otsu, M., Ahn, D.H., Kobayashi, T., Morita, Y., Wakiyama, Y., Ono-

dera, M., Eto, K., Ema, H., and Nakauchi, H. (2009). Definitive proof for direct

reprogramming of hematopoietic cells to pluripotency. Blood 114, 1764–1767.

Okita, K., Nakagawa, M., Hyenjong, H., Ichisaka, T., and Yamanaka, S. (2008).

Generation of mouse induced pluripotent stem cells without viral vectors.

Science 322, 949–953.

Rossant, J. (2007). Stem cells and lineage development in the mammalian

blastocyst. Reprod. Fertil. Dev. 19, 111–118.

Rossant, J., and Frels, W.I. (1980). Interspecific chimeras in mammals:

successful production of live chimeras between Mus musculus and Mus caroli.

Science 208, 419–421.

Rossant, J., Mauro, V.M., and Croy, B.A. (1982). Importance of trophoblast

genotype for survival of interspecific murine chimaeras. J. Embryol. Exp.

Morphol. 69, 141–149.

Satoh, M., Yasunami, Y., Matsuoka, N., Nakano, M., Itoh, T., Nitta, T., Anzai,

K., Ono, J., Taniguchi, M., and Ikeda, S. (2007). Successful islet transplantation

798 Cell 142, 787–799, September 3, 2010 ª2010 Elsevier Inc.

Page 167: Cell 100903

to two recipients from a single donor by targeting proinflammatory cytokines in

mice. Transplantation 83, 1085–1092.

Shinohara, T., Kato, M., Takehashi, M., Lee, J., Chuma, S., Nakatsuji, N.,

Kanatsu-Shinohara, M., and Hirabayashi, M. (2006). Rats produced by inter-

species spermatogonial transplantation in mice and in vitro microinsemination.

Proc. Natl. Acad. Sci. USA 103, 13624–13628.

Slack, J.M. (1995). Developmental biology of the pancreas. Development 121,

1569–1580.

Stanger, B.Z., Tanaka, A.J., and Melton, D.A. (2007). Organ size is limited by

the number of embryonic progenitor cells in the pancreas but not the liver.

Nature 445, 886–891.

Stern, M.S. (1973). Letter: Chimaeras obtained by aggregation of mouse eggs

with rat eggs. Nature 243, 472–473.

Takahashi, K., Tanabe, K., Ohnuki, M., Narita, M., Ichisaka, T., Tomoda, K.,

and Yamanaka, S. (2007). Induction of pluripotent stem cells from adult human

fibroblasts by defined factors. Cell 131, 861–872.

Takahashi, K., and Yamanaka, S. (2006). Induction of pluripotent stem cells

from mouse embryonic and adult fibroblast cultures by defined factors. Cell

126, 663–676.

Tarkowski, A.K. (1962). Inter-specific transfers of eggs between rat and

mouse. J. Embryol. Exp. Morphol. 10, 476–495.

Ueno, H., and Weissman, I.L. (2006). Clonal analysis of mouse development

reveals a polyclonal origin for yolk sac blood islands. Dev. Cell 11, 519–533.

Ueno, H., Turnbull, B.B., and Weissman, I.L. (2009). Two-step oligoclonal

development of male germ cells. Proc. Natl. Acad. Sci. USA 106, 175–180.

Ying, Q.L., Stavridis, M., Griffiths, D., Li, M., and Smith, A. (2003). Conversion

of embryonic stem cells into neuroectodermal precursors in adherent mono-

culture. Nat. Biotechnol. 21, 183–186.

Ying, Q.L., Wray, J., Nichols, J., Batlle-Morera, L., Doble, B., Woodgett, J.,

Cohen, P., and Smith, A. (2008). The ground state of embryonic stem cell

self-renewal. Nature 453, 519–523.

Zeilmaker, G.H. (1973). Fusion of rat and mouse morulae and formation of

chimaeric blastocysts. Nature 242, 115–116.

Cell 142, 787–799, September 3, 2010 ª2010 Elsevier Inc. 799

Page 168: Cell 100903

Profiling by Image RegistrationReveals Common Origin of AnnelidMushroom Bodies and Vertebrate PalliumRaju Tomer,1,* Alexandru S. Denes,1,2 Kristin Tessmar-Raible,1,3 and Detlev Arendt1,*1Developmental Biology Unit, European Molecular Biology Laboratory, D-69117 Heidelberg, Germany2Present address: Biozentrum, University of Basel, Klingelbergstrasse 50/70, CH-4056 Basel, Switzerland3Present address: Max F. Perutz Laboratories, Campus Vienna Biocenter, Dr. Bohr-Gasse 9/4, A-1030 Vienna, Austria

*Correspondence: [email protected] (R.T.), [email protected] (D.A.)DOI 10.1016/j.cell.2010.07.043

SUMMARY

The evolution of the highest-order human braincenter, the ‘‘pallium’’ or ‘‘cortex,’’ remains enigmatic.To elucidate its origins, we set out to identify relatedbrain parts in phylogenetically distant animals, tothen unravel common aspects in cellular composi-tion and molecular architecture. Here, we comparevertebrate pallium development to that of the mush-room bodies, sensory-associative brain centers, inan annelid. Using a newly developed protocol forcellular profiling by image registration (PrImR), weobtain a high-resolution gene expression map for thedeveloping annelid brain. Comparison to the verte-brate pallium reveals that the annelid mushroombodies develop from similar molecular coordinateswithin a conserved overall molecular brain topologyand that their development involves conservedpatterning mechanisms and produces conservedneuron types that existed already in the proto-stome-deuterostome ancestors. These data indicatedeep homology of pallium and mushroom bodiesand date back the origin of higher brain centers toprebilaterian times.

INTRODUCTION

In human and other mammals, the pallium represents the most

highly developed part of the forebrain, the site of learning and

memory (Kandel et al., 2000). It harbors huge densities of inter-

neurons arranged in cortical layers around a central neuropil

(hence, ‘‘cortex’’; Figure 1). The pallium is less elaborate in other

vertebrates and generally thought to function as a sensory-asso-

ciative center integrating primarily olfactory information (Nieu-

wenhuys, 2002). Insects and spiders, but also annelids, form

similar brain centers with densely packed neurons that send

out thousands, in some species hundreds of thousands, of

approximately parallel processes that form a central, lobed

neuropil (Strausfeld et al., 1998) (Figure 1). These likewise repre-

sent sensory-associative brain centers implicated in olfactory

discrimination, as well as in olfactory learning and memory

(Heisenberg, 2003; Strausfeld et al., 2009). Anatomical, struc-

tural, and functional similarities of the evolutionarily more ancient

parts of the vertebrate pallium, the ‘‘paleopallium’’ and ‘‘archipal-

lium,’’ with, for example, mushroom bodies of neopteran insects

have long been noted and interpreted as convergent acquisi-

tions as a result of functional constraints in independent evolu-

tionary lineages (Farris, 2005, 2008; Strausfeld et al., 1998).

Also, the evolution of characteristic folds and fissures and the

subpartitioning into functional units have occurred indepen-

dently in the mammalian pallium and several times in insect

mushroom bodies (Farris, 2008). These multiple convergences,

however, do not rule out the possibility that vertebrate pallium

and invertebrate mushroom bodies ultimately trace back to

a common evolutionary precursor. Some less elaborate,

sensory-associative brain centers may have already existed in

the protostome ancestor (of, among others, insects and anne-

lids) or even in the last common ancestor of protostomes and

deuterostomes (to which vertebrates belong). These may have

given rise to vertebrate pallium and invertebrate mushroom

bodies alike.

We set out to decide between the two alternatives—indepen-

dent evolution versus common origin of pallium and mushroom

bodies from ancient sensory-associative brain centers—by the

detailed molecular comparison of their development and differ-

entiation. Previous studies revealed conserved expression of

various transcription factors in vertebrate and invertebrate fore-

brains, but a specific comparison of gene expression in associa-

tive brain centers has not been undertaken. For this, we investi-

gated mushroom body development in the marine annelid

Platynereis dumerilii (Fischer and Dorresteijn, 2004), to then

compare to available vertebrate data. Platynereis is an ‘‘errant’’

annelid with huge mushroom bodies (Muller, 1973) that actively

explores the environment and shows some sort of learning

(Evans, 1966). Fortunately, for the purpose of genetic compar-

ison, the Platynereis transcriptome has proven to be ‘‘slow

evolving,’’ exhibiting less evolutionary change than that of other

protostomes with mushroom bodies (Drosophila, Apis) (Raible

et al., 2005). Also, in line with a slower evolutionary pace of the

annelid lineage, we have shown that Platynereis neural

patterning exhibits ancient characteristics, many of which are

shared with the vertebrates but not with the faster-evolving fly

800 Cell 142, 800–809, September 3, 2010 ª2010 Elsevier Inc.

Page 169: Cell 100903

and nematode lineages (Denes et al., 2007; Tessmar-Raible

et al., 2007). Finally, Platynereis is easy to breed and thus acces-

sible to high-throughput approaches (Fischer and Dorresteijn,

2004), as a prerequisite for the analysis of multiple genes.

Previous comparisons of brain development between phylo-

genetically remote species have been restricted to gene-by-

gene comparisons that did not allow, or allowed only in a coarse

manner, relating expression data spatially for multiple genes. To

overcome this, we have established a new protocol for Profiling

by Image Registration (PrImR) that, for the first time, allows align-

ment of expression images with cellular resolution and thus

expression profiling for the whole brain at the single-cell level.

Our protocol builds on the highly stereotypical nature of Platyner-

eis neurodevelopment to align expression patterns for an unlim-

ited number of genes in an additive process.

Using PrImR, we first investigated whether the overall molec-

ular topography of the brain anlage is conserved between

annelid and vertebrate and whether the Platynereis mushroom

bodies develop from corresponding molecular regions when

compared to vertebrate pallium (as a test for homology) (Arendt,

2005; Woodger, 1945). We also tested whether the positioning of

both follows similar developmental patterning mechanisms.

Next, we used PrImR to determine the expression profile for

specific subregions, focusing on specifically expressed genes

making up the regulatory and differentiation signature of cells

developing from these regions and compared this ‘‘molecular

fingerprint’’ (Arendt, 2008) to that of vertebrate telencephalic

neuron types but also to that of the mushroom body neuroblasts

in Drosophila. We conclude that pallium and mushroom body

most likely evolved from the same (sensory associative) brain

center that was already present in the bilaterian ancestors.

RESULTS

In Silico Alignment of Gene Expression PatternsTo establish a high-resolution molecular topography of the

annelid brain and to investigate the molecular fingerprint of brain

subregions, we set out to obtain coexpression information for

a large number of specifically expressed genes. This is impos-

sible to achieve at large scale by two-color double whole-mount

in situ hybridization (Tessmar-Raible et al., 2005), because this

would require a nonlinearly increasing amount of technically

demanding experiments. To overcome this, we took advantage

of the observation that Platynereis development is highly stereo-

typic and synchronous within and across batches (Fischer and

Dorresteijn, 2004) and developed a computational protocol for

Figure 1. Comparison of Vertebrate Pallium and Annelid Mushroom

Body Development Based on Previous Work and on This Study

(A–E) Development of the mouse pallium from open neural plate early specifi-

cation (5 s; A and B) to E10.5 late specification stage (C and D) and adult (E).

(F–J) Development of the Platynereis mushroom bodies from 48 hpf specifica-

tion stage (F–I) to adult (J). Note that in mouse, elaborate morphogenesis

(neural tube and telencephalic vesicle formation) occurs between early and

late specification stages, which does not take place in the annelid.

(A) The mouse pallium (PAL) anlage with reference to the anterior-posterior

(otx, gbx, hox) and mediolateral (pax6, nkx2.1, nkx2.2) regional patterning

systems. (After Inoue et al. [2000] and Shimamura et al. [1995].)

(B) Expression of bf-1 (orange) outlining the telencephalon anlage including the

pallium. Wnt5/8 and Shh indicate spatially restricted activity of signaling

ligands implicated in telencephalon development.

(C) Regionalization of telencephalic vesicles into pallium (PAL) and lateral and

medial ganglionic eminences (LGE, MGE). Colored strip indicates level of

cross-section in (D). M, mesencephalon; HT, hypothalamus, Eye, position of

cutoff eye stalk.

(D) Gene expression along the telencephalic section in (C) indicated by black

bars and color code. For references, see the main text.

(E) Parasagittal section of the mouse cortex.

(F) The Platynereis mushroom body anlage with reference to the anterior-

posterior (otx, gbx, hox) and mediolateral (pax6, nkx2.1, nkx2.2) regional

patterning systems. MB, mushroom bodies.

(G) Expression of bf-1 (orange) in the annelid brain at 48 hpf. Wnt5/8 and Hh

indicate spatially restricted expression of conserved ligands.

(H) Brain molecular topography of the bf-1+ region at 48 hpf as revealed by this

study. Color code refers to (I).

(I) Gene expression in the Platynereis brain indicated by black bars and color

code referring to (H).

(J) The Platynereis adult mushroom bodies, parasagittal section.

See also Figure S2.

Cell 142, 800–809, September 3, 2010 ª2010 Elsevier Inc. 801

Page 170: Cell 100903

accurately aligning the expression patterns of different genes

acquired from different stained individuals of identical develop-

mental stage (Profiling by Image Registration, PrImR, j,primərj).In the first step, two-channel image stacks were acquired via

whole-mount reflection confocal laser-scanning microscopy

(Jekely and Arendt, 2007). One channel contained information

on the expression pattern of a given gene and the other on the

axonal scaffold of the Platynereis larval brain. More than three

biological replicas were acquired for every gene. Next, we

used the axonal scaffold channel to align these images to a

reference average axonal scaffold image (Figure 2A, Figure S1

available online), first via rigid alignment algorithms and then

by smooth nonrigid transformation (see the Experimental Proce-

dures) (Figures 2B–2D). For each gene, a normalized average

expression image from (in most cases) five individuals was

thus generated (Figure S1B) that was directly comparable to

those of an unlimited number of other genes. As a test for accu-

racy, we obtained almost perfect overlap of signal for single cells

stained for the same gene in different individuals (Figure 2E). We

also found that the coexpression images obtained by our PrImR

protocol fully reproduced those obtained by double-fluorescent

WMISH (Tessmar-Raible et al., 2005) (compare Figures 2F and

2G) but tended to be more ‘‘complete,’’ reflecting the higher

sensitivity of the NBT/BCIP staining and of the reflection micros-

copy technique (Jekely and Arendt, 2007). Finally, to challenge

the accuracy of the protocol, we systematically estimated the

extent of overlap between individually aligned scans by calcu-

lating the Pearson’s correlation coefficient for 171 individually

aligned scans with the gene-specific average image (see the

Experimental Procedures) and found that the probability is

highest to obtain a value above 0.9, which implies that the

average expression images used for this study should reliably

reproduce endogenous gene expression. Our protocol thus

allows the comparison of any newly added expression pattern

to all preexisting patterns at once, with high accuracy and in

cellular resolution.

The Molecular Topography of the Platynereis BrainWe used our PrImR protocol to determine the molecular topog-

raphy of the Platynereis brain after 2 days of development, to test

whether any of the Platynereis forebrain regions would show

‘‘telencephalon-like’’ coordinates. At this stage, major subre-

gions of the developing Platynereis nervous system are already

established and larger populations of neurons have started

differentiation (Tessmar-Raible et al., 2007). For our purposes,

it was sufficient to focus on one stage only, because in the

absence of drastic morphogenetic changes (such as neurulation

and vesicle formation in vertebrates), gene expression patterns

in the developing Platynereis brain remain spatially constant

over time (as shown for example for pax6, six1/2 [Arendt et al.,

2002] or dach and bf-1 [Figures 5 and 6]), indicating colocaliza-

tion and temporal co-occurrence of differentiated neurons and

their precursors.

We had previously shown that both the annelid and vertebrate

brains exhibit a basic subdivision into medial nk2.1+ and lateral

pax6+ subregions (Tessmar-Raible et al., 2007) (Figure 1A)

(transformed into a ventral-dorsal arrangement during vertebrate

neurulation; Figures 1A–1C). It is within this conserved frame-

work that the vertebrate telencephalon anlage is established,

with the pallium developing from the anterior part of the pax6+

region (Figure 1A). The outline of the whole telencephalon anlage

is demarcated early on by expression of bf-1 (Hebert and Fishell,

2008) (orange in Figure 1B), a crucial regulator of telencephalic

development (Danesin et al., 2009). In the course of telence-

phalic morphogenesis, gsx+ and emx+ (Kimura et al., 2005)

domains complement the nk2.1+ and pax6+ subregions (Figures

1C and 1D) such that the gsx+, pax6+ overlap demarcates the

future pallial-subpallial boundary (Danesin et al., 2009; Yun

et al., 2001) and pax6+, emx+ coexpression specifies pallial

telencephalic precursors (Kimura et al., 2005). Besides bf-1,

lhx2 (expressed broadly in the telencephalon and acting highly

up in the hierarchy of cortical induction (Hebert and Fishell,

2008), COUP-TF1/seven-up (contributing to telencephalic

Figure 2. In Silico Alignment of Expression

Patterns

(A) Maximum Z projection of average reference

axonal scaffold image of 48 hpf larval brain.

(B–D) Step-wise alignment of an axonal scaffold

image to the average reference scaffold (cyan)

using rigid, affine, and smoothened freeform

nonrigid transformations. Images are 20 mm thick

optical sections. See Movies S1 and S2 for details.

(E) A positive control experiment to test the accu-

racy of the alignment algorithm. All images are

maximum Z projections of aligned expression

patterns of r-opsin, acquired from three different

larvae, and an RGB merge of them. See Movie

S3 for details.

(F and G) Experimental validation of an in silico co-

expression (F) using double fluorescence WMISH

double staining (G). White, colocalization of green

and red pixels. Cyan and blue arrows point to veri-

fied coexpression.

(H) Violin plot illustrating the distribution of Pearson’s correlation coefficients of 171 gene expression images with their gene-specific average images. Magenta

shows the data density. Mean = 0.900, median = 0.904, standard deviation = 0.040.

Scale bars represent 50 mm. See also Figure S1 and Movies S1, S2, and S3.

802 Cell 142, 800–809, September 3, 2010 ª2010 Elsevier Inc.

Page 171: Cell 100903

patterning (Hebert and Fishell, 2008) and soxB family members

(sox1[Ekonomou et al., 2005] and sox2 [Bani-Yaghoub et al.,

2006]) play important roles in early telencephalic development

(Figure 3E), although the spatial distribution and coexpression

of these factors along the anterior-posterior and mediolateral

brain axes is not fully resolved for the vertebrates.

We accordingly examined bf-1, lhx2, svp, soxB, pax6, emx,

gsx, and nk2.1 expression in Platynereis and indeed identified

a brain region with telencephalon-like molecular topography

(Figures 3A–3I; summarized in Figure 1). First, we found bf-1

expressed in a horseshoe-shaped domain in the Platynereis

brain (Figure 3A), where it is specifically coexpressed with lhx2,

svp, and soxB (Figures 3C–3E). (The more lateral part of bf-1+

domain represents the eye anlage; Figure 3B.) Second, from

lateral to medial, the Platynereis bf-1 brain domain is subdivided

by spatially restricted coexpression of emx, pax6, gsx, and nk2.1

(Figures 3F–3J) in a telencephalon-like fashion. Nowhere else

in the developing Platynereis larva was a similar sequence of

emx+/pax6+, pax6+, pax6+/gsx+, gsx+, and gsx+/nk2.1+ subre-

gions (Figure 3J) detected. The emx+/pax6+ and pax6+ regions

within the bf1+ domain thus represented candidate counterpart

regions to the vertebrate pallium anlage, a notion that we

decided to explore further.

Comparison of Patterning MechanismsIn vertebrates, bf-1 coordinates the activity of two opposing

signaling centers that pattern the telencephalon anlage. bf-1

acts downstream of the ventral signal, Hh, to induce ventral

(subpallial) identities. At the same time, bf-1 inhibits dorsal

Wnt/b-catenin signaling through direct transcriptional repression

of Wnt8 (Danesin et al., 2009), which induces dorsal (pallial)

identities (Houart et al., 2002). Vertebrate telencephalic

patterning can be perturbed by ectopic b-catenin activation

(mimicking the dorsal signal), which accordingly triggers the

upregulation of dorsal pallial and downregulation of ventral sub-

pallial markers (Backman et al., 2005). Investigating the expres-

sion of orthologous signals and factors in Platynereis, we indeed

detected Hh in the medial (Figure 3K) and Wnt8 in the lateral brain

(Figure 3L), matching the vertebrate situation. We also observed

lateral expression of Wnt5 (Figure 3M), another dorsal Wnt signal

implicated in corticospinal axon guidance (Keeble et al., 2006), of

WntA (not found in vertebrates; Figure 3N) and Gli (Figure 3O),

a zinc finger transcription factor essential for dorsal telencephalic

fates (Tole et al., 2000). None of the other Platynereis Wnt genes

was expressed in the brain (data not shown; note that Wnt3 does

not exist in annelids). For Platynereis, exposure to the chemical

compound 1-Azakenpaullone has been established as an effi-

cient means to ectopically induceb-catenin activation (Schneider

and Bowerman, 2007). We accordingly tested the effect of 1-Aza-

kenpaullone on Platynereis brain regionalization and observed

upregulation of lateral emx expression (Figures 4B and 4G),

downregulation of intermediate gsx (Figures 4D and 4I), and

medial nk2.1 expression (Figures 4E and 4J), while pax6 ap-

peared unaffected (Figures 4C and 4H). These results mirror

vertebrate telencephalic patterning (Figure 1) and are thus

strongly indicative of evolutionary conservation.

The Molecular Fingerprint of Mushroom Body NeuronsIn vertebrates, distinct combinations of differentially expressed

transcription factors control the fate of the various telencephalic

subregions. We took advantage of PrImR to further determine

and compare molecular fingerprints. In mouse, dorsal telence-

phalic (pallial) pax6+ regions show differential activity of Dach

(Caubit et al., 1999) and ngn1/2, bHLH factors required for pallial

Figure 3. Coexpression of ‘‘Telencephalic’’

and Other Marker Genes, Red, with bf-1,

Green, in the Platynereis 48 hpf Larval Brain

(A and B) Complete expression of bf-1 (green) (A).

Blue, reference axonal scaffold. Dotted circle

marks expression in the eye field, as indicated

by r-opsin coexpression in (B).

(C–E) genes covering the almost entire bf-1

domain.

(F–I) Lateral-to-medial sequence of regionalisa-

tion genes.

(J) Five-micron-thick optical section showing

lateral-to-medial coexpression of emx (green),

pax6(red), and gsx (cyan).

(K–O) Genes involved in mediolateral signaling.

(P–T) Genes showing mediolaterally restricted

expression.

All images are Maximum Z projections of slice

numbers 13 to 40, i.e., 28 mm thick (except K:

8–28, i.e., 21 mm thick) in the reference scaffold

image stack. White marks colocalizing pixels.

Scale bars represent 50 mm. See also Figure S5.

Cell 142, 800–809, September 3, 2010 ª2010 Elsevier Inc. 803

Page 172: Cell 100903

neurogenesis (Nieto et al., 2001). In Platynereis, dach and ngn

are similarly restricted to emx+, pax6+ subregions (Figures 3P

and 3Q; compare Figures 1D and 1I). In mouse, ventral (subpal-

lial) gsx+ regions coexpress vax1, required for GABAergic inter-

neuron generation (Taglialatela et al., 2004), and coe/ebf1, an

HLH transcription factor affecting subpallial development (Garel

et al., 1999). We accordingly found both genes coexpressed in

Platynereis gsx+ regions (Figures 3R and 3S; compare Figures

1D and 1I). In mouse, telencephalic subregions express er81,

an ETS factor active in distinct subpallial domains (Yun et al.,

2001) and direct target of pax6 in the dorsal pallium (Tuoc

and Stoykova, 2008). Platynereis er81 is similarly expressed

in the pax6+/gsx+ and gsx+/nk2.1+ subregions but not in

between (Figure 3T; compare Figures 1D and 1I). Adding to

this, expression of Platynereis genes orthologous to ascl1/

mash1 (bHLH downstream effector of gsx1 [Wang et al., 2009];

expressed in ventral and dorsal telencephalon [Yun et al.,

2001]), to brn1/2/4 (POU domain transcription factors required

for cortical migration in dorsal and ventral telencephalon [Ekono-

mou et al., 2005; McEvilly et al., 2002]), and to many other genes

implicated in telencephalon development by function or by

expression is detailed for Platynereis in Figure S2 and has

been used to generate the molecular map in Figure 1H. The over-

all comparison suggests that the molecular fingerprints of

annelid and vertebrate emx+/pax6+, pax6+, pax6+/gsx+, gsx+,

and gsx+/nk2.1+ subregions are similar to large extent but may

also differ in the detail (for example, dlx is expressed more

broadly in annelid than in vertebrate; rx is expressed in the

emx+, pax6+, and gsx+ subregions in the annelid but excluded

in the vertebrate; Figure S2).

Positioning the Platynereis Mushroom Body AnlagenSince our molecular comparison had identified one unique

candidate region in the Platynereis brain, which, by position,

similar responsiveness to Wnt signaling, and similar molecular

fingerprint (bf1+, emx+, pax6+, dach+, svp+, tll+, ngn+, asc+)

qualified as possible evolutionary counterpart to the vertebrate

pallium anlage (encircled by stippled line in Figures 1C and 1D),

we set out to determine whether this paired region indeed

represented the Platynereis mushroom body anlagen by tracing

mushroom body development from adult (Figure 5, Figure S3,

Movie S4) and juvenile stages (10 days postfertilization [dpf],

when they are fully developed and can be easily identified

histologically) back to earlier larval stages. We combined

anatomical (the specific shape of mushroom bodies), histolog-

ical (the dense packing of neurons), and topological (the specific

spatial arrangement of mushroom bodies and palpal and

antennal nerves) evidence, as well as the expression of mush-

room body-specific marker genes (dach and pax6), to reidentify

and position the developing mushroom bodies at various stages

(6 weeks, 20, 10, 9, 8, 7, 6, 5, 4, 3, and 2 dpf; Figure 5 and data

not shown). This allowed unambiguous positioning of the mush-

room body anlagen to ventrolateral coordinates at 48 hr postfer-

tilization [hpf] that indeed matched the emx+, pax6+, dach+

candidate region.

In juvenile worms, the differentiating emx+, pax6+ mushroom

bodies and the more medial gsx+, nk2.1+ brain tissue (referred

to as ‘‘pars intercerebralis’’[Muller, 1973]) continued to express

bf-1 (Figure 6A) and started to specifically express arx (Fig-

ure 6B; expressed in mouse and fish telencephalon and ventral

thalamus [Miura et al., 1997]). We also detected specific expres-

sion of two conserved bilaterian microRNAs, miR-9 and mir-9*

(Christodoulou et al., 2010) (Figures 6C and 6D). In the verte-

brates, these microRNAs are expressed only in the telenceph-

alon, among all differentiated CNS tissues (Deo et al., 2006).

Regarding transmitter usage, large part of the Platynereis mush-

room bodies proved glutamatergic by expression of vglut,

encoding the vesicular glutamate transporter (Figure 6F). In

contrast, gad, a marker for GABAergic neurons, was restricted

to more medial brain tissue (Figure 6E), as was the dopaminergic

Figure 4. Differential Gene Regulation in the Developing Platynereis Brain

Expression of bf-1, emx, pax6, gsx, and nk2.1 in DMSO control (A–E) and 1-Azakenpaullone (1 mM final concentration) treated 48 hpf larval brain (F–J). Changes in

expression patterns are highlighted by dashed circles. Blue, nuclear stain DAPI; green, immunostaining against acetylated tubulin. n for bf-1: treated > 200,

control > 200; emx: treated > 380, control > 300; pax6: treated > 350, control > 310; gsx: treated > 150, control > 160; nk2.1: treated > 320, control > 200.

Scale bars represent 50 mm.

804 Cell 142, 800–809, September 3, 2010 ª2010 Elsevier Inc.

Page 173: Cell 100903

marker tyrosine hydroxylase (Figure 6H), again matching the

vertebrate situation (Marın and Rubenstein, 2001). However,

we also detected a difference in the spatial distribution of trans-

mitter type in that Platynereis mushroom bodies also harbored

cholinergic neurons (Figure 6G; not found in the vertebrate

pallium).

Molecular Fingerprint Comparison with InsectMushroom BodiesMushroom bodies of similar cellular composition also exist in

arthropods such as onychophorans (velvet worms), spiders,

and insects, yet ample molecular data exist only for the mush-

room bodies in insects. In Drosophila, mushroom body neuro-

blasts require pax6 and dach for their specification but do not

express sine oculis (six1/2) or eyes absent (eya) (Noveen et al.,

2000). We found that the only region of the Platynereis brain

coexpressing pax6 and dach but devoid of six1/2 and eya is

indeed the mushroom body anlage (Figures 7A and 7B and eya

Figure 5. Platynereis Mushroom Body

Development

Three-dimensional models were generated for

various stages according to confocal image

stacks (see Movie S4 for details). Depicted struc-

tures represent mushroom body pedunculi (red),

mushroom body perikarya (transparent red),

antennal nerve (dark blue), pars intercerebralis

(purple), eyes (yellow), palpae (green), and

a glomerulus-like structure (mediates the connec-

tion between palpae and mushroom bodies; light

green). Continuous expression of dach (top and

bottow row) in mushroom body neurons consis-

tent with anatomical backtracing. Blue arrows

point to the antennal nerve. A, P, D, V, L, and R

indicate body axes.

Scale bars represent 50 mm. See also Figure S3

and Movie S4.

Figure 6. Gene Expression in the Brain of

Platynereis Juvenile Worms

Transcription factors (A and B), microRNAs (C and

D) and marker genes (E–H) for GABAergic (gad, E),

glutamatergic (vglut, F), cholinergic (vacht, G),

and dopaminergic (tyrosine hydroxylase, H) neu-

rons. Dashed circles outline mushroom bodies.

A, P, L, and R indicate body axes. Scale bars

represent 50 mm.

data not shown), indicative of evolu-

tionary conservation. To challenge

homology of insect and annelid mush-

room bodies, we looked for a combination

of transcription factors that, by

coexpression, uniquely define Drosophila

mushroom body neuroblasts independent of pax6 and dach—to

then test whether a similar coexpression would locate to the

annelid mushroom body anlage in an equally specific manner.

Among the four Drosophila mushroom body neuroblasts, Pcd8

andPcd9 are uniquely identified by the specific coexpression

of asc/l’sc, tll, svp, slp/bf-1, and otx (Urbach and Technau,

2003). We determined by PrImR where in the Platynereis 48

hpf brain these five transcription factors would be coexpressed,

and indeed the only two bilateral spots of cells in the Platynereis

brain coexpressing them form part of the mushroom body

anlage (Figures 7C–7H). This is strong evidence in favor of an

evolutionary relatedness of insect and annelid mushroom

bodies. However, in contrast to annelid mushroom bodies

(Figure 7I) and to vertebrate pallium, the Drosophila dac+ mush-

room body precursors do not express emx, and the overall

molecular topography is conserved to lesser extent between

fly, annelid, and vertebrate (Urbach and Technau, 2003) than

between annelid and vertebrate. Our data suggest that this is

Cell 142, 800–809, September 3, 2010 ª2010 Elsevier Inc. 805

Page 174: Cell 100903

due to secondary modifications of mushroom body develop-

ment in the insect lineage.

DISCUSSION

In this study, we have established a new protocol, PrImR, that

allows expression profiling by image registration with cellular

resolution for the entire organism. Expression profiling at

single-cell resolution has been reported previously (Liu et al.,

2009; Sugino et al., 2006), but our protocol is the first to allow

simultaneous profiling for the whole central nervous system,

where cells are densely packed. Our protocol thus sets the stage

for a new chapter of evolutionary developmental research that

has so far been restricted to gene-by-gene comparisons of

expression patterns with restricted coexpression information,

necessarily remaining fragmentary and sometimes inconclusive.

Instead, PrImR allows much broader comparisons, involving

hundreds of transcription factors and differentiation genes, and

a much higher level of integration. For the first time, molecular

fingerprinting as introduced here allows testing homology

hypotheses such as the comparison of mushroom bodies

between annelids and insects and of mushroom bodies with

vertebrate pallium (Strausfeld et al., 1998) that could not be

resolved on merely anatomical and physiological grounds.

Taking advantage of this new protocol, our data reveal that

mushroom bodies in annelid and vertebrate pallium develop

from the same, molecularly defined, subregion (emx+/pax6+) at

similar anterior-posterior and mediolateral coordinates within an

overall conserved molecular topography of the brain (Figures 1A

and 1F), which is established by similar patterning mechanisms

(Figures 1B and 1G). They express similar and unique combina-

Figure 7. Coexpression of Insect Neuro-

blast Marker Genes, Red, with Dach, Green,

in the Platynereis 48 hpf Larval Brain

(A–J) Coexpression of dach (green) with pax6 (A),

six1/2 (B), ash2 (C), tll (D), svp (E), bf-1/slp (F), otx

(G), the intersection of ash2, tll, svp, bf-1/slp, and

otx (H), emx (I), and Wnt5 (J) (red). The intersection

of expression patterns was generated from

expression masks generated by smoothening

images with a Gaussian sigma of 2 and a signal

cutoff of 35. All images are Maximum Z projection

of slice numbers 13 to 40, i.e., 28 mm thick, in the

reference scaffold image stack. White marks

colocalizing pixels.

(K) Phylogenetic distribution of mushroom body

and pallium in Bilateria.

Scale bars represent 50 mm. See also Figure S4.

tions of additional transcription factors

(dach+, svp+, slp+, tll+, ngn+, asc+,

arx+) (Figures 1C, 1D, 1H, 1I, and 6).

They give rise to glutamatergic neurons,

while GABAergic and dopaminergic

neuron types emerge from more medial/

ventral brain regions. They are positive

for the same brain-specific microRNAs

(mir-9, -9*). We are confident that these

similarities are too complex to be plausibly explained by indepen-

dent (convergent) evolution. Instead, we propose that they were in

place already in the protostome-deuterostome ancestor, the last

common ancestor of annelids and vertebrates (Figure 7K).

To assess the statistical significance of our dataset, we con-

ducted permutation tests by randomly assigning gene expres-

sion to ‘‘pallium’’ or ‘‘mushroom body,’’ and we found the

observed similarity in molecular fingerprint extremely unlikely to

arise by chance (Figure S4; p value less than 2.0E-05). It should

be stressed, however, that this value might slightly overestimate

the significance of our comparison, because a subset of the

involved genes may be coregulated and should be treated as

a single observation.

Importantly, however, annelid mushroom body and vertebrate

pallium development and specification also differ to consider-

able extent, which is plausible given �600 million years of inde-

pendent evolution. First, vertebrate pallium development

involves neurulation and telencephalic vesicle formation (Figures

1B and 1C), which does not occur in the annelid. Second, impor-

tant differences are apparent in some expression domains. For

example, annelid nk2.1 expression exhibits lateral ‘‘wings’’

(Figure 1F), where it overlaps with the lateral pax6 domain (care-

ful 3D analysis, however, reveals that this overlap is outside of

the bf1+ mushroom body anlage). Also, the developing annelid

mushroom body is positive for the rx gene (Figure S2), which is

excluded from the vertebrate telencephalon anlage (Stigloher

et al., 2006). Finally, some genes are absent in the other species.

For example, Wnt3, expressed in the dorsomedial pallium

anlage, does not exist in annelids, and WntA, demarcating the

lateral mushroom body anlage (Figure 3N), is not found in the

vertebrate genomes.

806 Cell 142, 800–809, September 3, 2010 ª2010 Elsevier Inc.

Page 175: Cell 100903

What is the significance of these findings for our under-

standing of mushroom body and pallium/telencephalon evolu-

tion? We can infer that the last common ancestor of protostomes

and deuterostomes (and thus, of annelids and vertebrates)

already possessed distinct brain centers that later gave rise to

the pallium in the vertebrate and to mushroom bodies in the

annelid (and arthropod) lineage (Figure 7K). Based on our data,

it is likely that these centers harbored subsets of glutamatergic

neurons that presumably functioned in some sort of integration

of sensory signals. Input was most likely chemosensory, as

deduced from the primarily chemosensory input of pallium and

mushroom bodies alike, in insects, spiders, and onychophorans,

as well as in annelids (Strausfeld et al., 2006). (The Platynereis

mushroom bodies receive connections from the chemosensory

palpae [Movie S4].) Notably, however, the sensory input of

arthropod mushroom bodies is received by distinct olfactory

appendages (Strausfeld et al., 2006) and can even be visual

(in spiders, honeybee and cockroach [Strausfeld and Barth,

1993]), indicating some plasticity in mushroom body connec-

tivity. Apparently, the sensory input into sensory associative

centers can change according to the needs of changing ecolog-

ical niches.

We can speculate that the first function of these chemosen-

sory integrative brain centers was to distinguish between food

and nonfood, to decide about directed locomotion toward

identified food sources, and, ultimately, to integrate previous

experiences into some sort of ‘‘learning’’ (Evans, 1966). Extrap-

olating from the similar motor output of today’s pallium and

mushroom bodies, we can assume that these integrative centers

already produced coordinated locomotion of some sort (Farris,

2005).

There is a long-standing notion that the vertebrate pallium

comprises distinct parts of different evolutionary age, paleo-,

archi-, and neopallium. Roughly, the paleopallium corresponds

to olfactory cortex and the archipallium to the hippocampus,

and the neopallium comprises the remainder of the cerebral

hemispheres (Swanson, 2000). Paleo- and archipallium are

found in all vertebrates, while the neopallium is restricted to

amniotes. In line with this concept, the cellular organization of

the more ancient archi- and paleopallium strikingly resembles

that of annelid and arthropod mushroom bodies, in that densely

packed, basophilic somata send out parallel processes inter-

sected by afferent and efferent networks (Strausfeld et al.,

1998). Intriguingly, when compared to neopallium, the hippo-

campus and dentate gyrus each consist of only one layer of

principle neurons (Forster et al., 2006). This contrasts with the

six-layered architecture of the neopallium but rather resembles

the simple architecture of the mushroom bodies. We propose

that such simple architecture, ancient for vertebrates, might

have been in place in the urbilaterian mushroom body/pallium

sensory-associative precursor structure.

As mentioned in the introduction, it is unlikely that a mushroom

body-like shape was already in place in the protostome-deutero-

stome ancestor. Rather, it seems plausible that the common

evolutionary precursor structure of mushroom bodies and

pallium was morphologically less complex and that higher

degrees of histological (and thus, functional) complexity were

acquired independently in divergent evolutionary lineages, as

repeatedly proposed (Farris, 2005, 2008; Strausfeld et al.,

1998; Strausfeld et al., 2009). To settle this issue, it will be

rewarding to identify brain regions with similar molecular coordi-

nates and fingerprint in species lacking overt mushroom body

morphology such as crustaceans or mollusks on the protostome

side or amphioxus on the deuterostome side (Figure 7K) and

to investigate possible sensory input, integrative function and

locomotor output of the corresponding brain parts. If, on one

hand, such less complex, ‘‘mushroom body-related’’ brain parts

indeed exist, they will help defining a minimum set of anatomical

and physiological characteristics associated with the conserved

molecular fingerprints described here. This way, we would gain

more insight into anatomy and function of an ancient sensory-

integrative center present in the protostome-deuterostome

ancestor. If, on the other, such regions do not exist, we can infer

secondary loss of brain complexity. Possible evolutionary

scenarios leading to reduced brain complexity include the tran-

sition to sessile or hemisessile lifestyle with filter feeding, to inter-

stitial burrowing with minute body size, or to parasitism.

Once a better and more complete picture of mushroom body-

related brain parts in bilaterians is obtained, this will set the stage

for a more comprehensive view on the evolution of insect and

annelid mushroom bodies and vertebrate pallium alike. What

do these groups have in common that is not found in bilaterians

in which mushroom body-related brain parts are present but less

elaborate? This will ultimately help to link mushroom body/

pallium evolution to changes in life style, locomotion, or food

capture.

EXPERIMENTAL PROCEDURES

Imaging Gene Expression Patterns

We used whole-mount reflection confocal microscopy (Jekely and Arendt,

2007) to acquire gene expression pattern images. Typically, an oil immersion

403 objective on the Leica TCS SPE Confocal Microscope was used.

Alignment of Expression Patterns

For achieving highly accurate alignments of axonal scaffolds, we developed

a multistep multiresolution image registration algorithm using Mutual Informa-

tion (Viola and Wells, 1995) as the image similarity metric. In the first step,

images were transformed rigidly with six degrees of freedom (three for trans-

lation and three for rotation) to roughly align the images. In the second step,

affine transformation with 12 degrees of freedom (three for rotation, three for

translation, three for scaling, and three for shear) was used to compensate

for the shrinking and shearing artifacts introduced by in situ hybridization

and imaging procedures. In the last step, we used a uniform grid of control

points to model freeform nonrigid local deformations and used third-order

B-splines to smoothly interpolate the transformations. For the affine and

nonrigid registration steps, we used a multiresolution approach in which 3D

images were smoothened and downsampled to various degrees before the

alignment.

Average Reference Axonal Scaffold

We generated an average Platynereis larval axonal scaffold using images

acquired from 36 different individual 48 hpf larvae (Figure S1), which was

used as reference for all the alignments.

Averaging Expression Patterns

For direct comparison of the spatial expression patterns of different genes, we

generated normalized average expression patterns from three or more indi-

vidual larvae (see Movie S3 and Figure S1B for examples). We used

Cell 142, 800–809, September 3, 2010 ª2010 Elsevier Inc. 807

Page 176: Cell 100903

‘‘Colocalization Highlighter’’ plug-in in ImageJ 1.41e to highlight the colocaliz-

ing pixels as white pixels. Default settings were used for most of the cases.

Alignment Accuracy Estimates

We conducted a systematic analysis by testing images from 171 different

larvae. Pearson’s correlation coefficient was calculated for each aligned

expression image with the corresponding gene-specific average.

Modeling Platynereis Brain Anatomy

Amira 4.1.1 was used to segment various anatomical structures in the brain

(Movie S4) and to generate 3D surface renderings shown in Figure 5.

Ectopic Overexpression of Wnt Canonical Signaling

We used an established procedure (Schneider and Bowerman, 2007) of treat-

ment with 1-Azakenpaullone (cat. no. 191500, Calbiochem) to achieve ectopic

overexpression of canonical Wnt signaling in Platynereis dumerilii larvae.

A mixture of larvae from two to three different animal pairs was used for

treatment. The larvae were treated from 24 hpf to 48 hpf in 6 ml Natural Sea

Water at a final concentration of 1 mM 1-Azakenpaullone or DMSO and were

grown at 18�C. The larvae were fixed in 4% PFA/PTW, stored in 100% meth-

anol at �20�C, and proceeded for in situ hybridization. The staining reactions

were stopped simultaneously for control and treated larvae, much before the

saturation.

Cloning Genes

We used degenerate primers to clone fragments of brn1/2/4, dach, emx, svp,

tll, vglut, and gad and 50,30-RACE extensions to extend the fragments (see the

Extended Experimental Procedures for primer sequences). We used cDNA

templates generated with SMART technology (SMART RACE cDNA amplifica-

tion kit, Clonetech) from total RNA extracted from larvae at several develop-

mental stages. Other genes investigated in this study were obtained either

as EST sequences or from previous publications. The GenBank IDs of arx,

ash2, bf-1, brn1/2/4, coe, dach, emx, er81, ets3, lhx1/5, svp, tll, vglut, Wnt5,

Wnt8, gad, and tyr hyd are GU169412–GU169428, respectively.

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures, five

figures, and four movies and can be found with this article online at doi:

10.1016/j.cell.2010.07.043.

ACKNOWLEDGMENTS

We thank F. Christodoulou for providing larvae stained with mir-9, 9*,

H. Snyman for expert technical assistance, and the Arendt lab for support.

The electron micrograph in Figure 5 was taken with C. Nielsen during a prac-

tical course. R.T. and K.T.-R. were supported by fellowships of the Marie Curie

RTN ZOONET (MRTN-CT-2004-005624), R.T. by DFG Collaborative Research

Network SFB-488, and A.D. by a Louis Jeantet Foundation fellowship. R.T.

conceived and developed PrImR, designed and performed the experiments,

analyzed and interpreted the data, and wrote the paper. A.S.D. contributed

clones of vglut and gad. K.T.-R. contributed clones of brn124 and emx. D.A.

proposed to study annelid mushroom bodies, provided ideas, interpreted

the data, and wrote the paper.

Received: March 12, 2010

Revised: May 22, 2010

Accepted: July 14, 2010

Published: September 2, 2010

REFERENCES

Arendt, D. (2005). Genes and homology in nervous system evolution:

comparing gene functions, expression patterns, and cell type molecular

fingerprints. Theory Biosci. 24, 185–197.

Arendt, D. (2008). The evolution of cell types in animals: emerging principles

from molecular studies. Nat. Rev. Genet. 9, 868–882.

Arendt, D., Tessmar, K., de Campos-Baptista, M.I., Dorresteijn, A., and

Wittbrodt, J. (2002). Development of pigment-cup eyes in the polychaete

Platynereis dumerilii and evolutionary conservation of larval eyes in Bilateria.

Development 129, 1143–1154.

Backman, M., Machon, O., Mygland, L., van den Bout, C.J., Zhong, W.,

Taketo, M.M., and Krauss, S. (2005). Effects of canonical Wnt signaling

on dorso-ventral specification of the mouse telencephalon. Dev. Biol. 279,

155–168.

Bani-Yaghoub, M., Tremblay, R.G., Lei, J.X., Zhang, D., Zurakowski, B.,

Sandhu, J.K., Smith, B., Ribecco-Lutkiewicz, M., Kennedy, J., Walker, P.R.,

and Sikorska, M. (2006). Role of Sox2 in the development of the mouse

neocortex. Dev. Biol. 295, 52–66.

Caubit, X., Thangarajah, R., Theil, T., Wirth, J., Nothwang, H.G., Ruther, U.,

and Krauss, S. (1999). Mouse Dac, a novel nuclear factor with homology to

Drosophila dachshund shows a dynamic expression in the neural crest, the

eye, the neocortex, and the limb bud. Dev. Dyn. 214, 66–80.

Christodoulou, F., Raible, F., Tomer, R., Simakov, O., Trachana, K., Klaus, S.,

Snyman, H., Hannon, G.J., Bork, P., and Arendt, D. (2010). Ancient animal

microRNAs and the evolution of tissue identity. Nature 463, 1084–1088.

Danesin, C., Peres, J.N., Johansson, M., Snowden, V., Cording, A., Papalo-

pulu, N., and Houart, C. (2009). Integration of telencephalic Wnt and hedgehog

signaling center activities by Foxg1. Dev. Cell 16, 576–587.

Denes, A.S., Jekely, G., Steinmetz, P.R., Raible, F., Snyman, H., Prud’homme,

B., Ferrier, D.E., Balavoine, G., and Arendt, D. (2007). Molecular architecture of

annelid nerve cord supports common origin of nervous system centralization

in bilateria. Cell 129, 277–288.

Deo, M., Yu, J.Y., Chung, K.H., Tippens, M., and Turner, D.L. (2006). Detection

of mammalian microRNA expression by in situ hybridization with RNA oligonu-

cleotides. Dev. Dyn. 235, 2538–2548.

Ekonomou, A., Kazanis, I., Malas, S., Wood, H., Alifragis, P., Denaxa, M.,

Karagogeos, D., Constanti, A., Lovell-Badge, R., and Episkopou, V. (2005).

Neuronal migration and ventral subtype identity in the telencephalon depend

on SOX1. PLoS Biol. 3, e186.

Evans, S.M. (1966). Non-associative behavioural modifications in nereid poly-

chaetes. Nature 211, 945–948.

Farris, S.M. (2005). Evolution of insect mushroom bodies: old clues, new

insights. Arthropod Struct. Dev. 34, 211–234.

Farris, S.M. (2008). Evolutionary convergence of higher brain centers spanning

the protostome-deuterostome boundary. Brain Behav. Evol. 72, 106–122.

Fischer, A., and Dorresteijn, A.W.C. (2004). The polychaete Platynereis dumer-

ilii (Annelida): a laboratory animal with spiralian cleavage, lifelong segment

proliferation and a mixed benthic/pelagic life cycle. Bioessays 26, 314–325.

Forster, E., Zhao, S., and Frotscher, M. (2006). Laminating the hippocampus.

Nat. Rev. Neurosci. 7, 259–267.

Garel, S., Marın, F., Grosschedl, R., and Charnay, P. (1999). Ebf1 controls early

cell differentiation in the embryonic striatum. Development 126, 5285–5294.

Hebert, J.M., and Fishell, G. (2008). The genetics of early telencephalon

patterning: some assembly required. Nat. Rev. Neurosci. 9, 678–685.

Heisenberg, M. (2003). Mushroom body memoir: from maps to models. Nat.

Rev. Neurosci. 4, 266–275.

Houart, C., Caneparo, L., Heisenberg, C., Barth, K., Take-Uchi, M., and

Wilson, S. (2002). Establishment of the telencephalon during gastrulation by

local antagonism of Wnt signaling. Neuron 35, 255–265.

Inoue, T., Nakamura, S., and Osumi, N. (2000). Fate mapping of the mouse

prosencephalic neural plate. Dev. Biol. 219, 373–383.

Jekely, G., and Arendt, D. (2007). Cellular resolution expression profiling using

confocal detection of NBT/BCIP precipitate by reflection microscopy. Bio-

techniques 42, 751–755.

Kandel, E.R.S., James, H., and Jessell, T.M. (2000). Principles of Neural

Science, Fourth Edition (New York: McGraw-Hill Companies, Inc.).

808 Cell 142, 800–809, September 3, 2010 ª2010 Elsevier Inc.

Page 177: Cell 100903

Keeble, T.R., Halford, M.M., Seaman, C., Kee, N., Macheda, M., Anderson,

R.B., Stacker, S.A., and Cooper, H.M. (2006). The Wnt receptor Ryk is required

for Wnt5a-mediated axon guidance on the contralateral side of the corpus

callosum. J. Neurosci. 26, 5840–5848.

Kimura, J., Suda, Y., Kurokawa, D., Hossain, Z.M., Nakamura, M., Takahashi,

M., Hara, A., and Aizawa, S. (2005). Emx2 and Pax6 function in cooperation

with Otx2 and Otx1 to develop caudal forebrain primordium that includes

future archipallium. J. Neurosci. 25, 5097–5108.

Liu, X., Long, F., Peng, H., Aerni, S.J., Jiang, M., Sanchez-Blanco, A., Murray,

J.I., Preston, E., Mericle, B., Batzoglou, S., et al. (2009). Analysis of cell fate

from single-cell gene expression profiles in C. elegans. Cell 139, 623–633.

Marın, O., and Rubenstein, J.L. (2001). A long, remarkable journey: tangential

migration in the telencephalon. Nat. Rev. Neurosci. 2, 780–790.

McEvilly, R.J., de Diaz, M.O., Schonemann, M.D., Hooshmand, F., and

Rosenfeld, M.G. (2002). Transcriptional regulation of cortical neuron migration

by POU domain factors. Science 295, 1528–1532.

Miura, H., Yanazawa, M., Kato, K., and Kitamura, K. (1997). Expression of

a novel aristaless related homeobox gene ‘Arx’ in the vertebrate telenceph-

alon, diencephalon and floor plate. Mech. Dev. 65, 99–109.

Muller, W.A. (1973). [Autoradiographic studies on the synthetic activity of

neurosecretory cells in the brain of Platynereis dumerilii during sexual develop-

ment and regeneration]. Z. Zellforsch. Mikrosk. Anat. 139, 487–510.

Nieto, M., Schuurmans, C., Britz, O., and Guillemot, F. (2001). Neural bHLH

genes control the neuronal versus glial fate decision in cortical progenitors.

Neuron 29, 401–413.

Nieuwenhuys, R. (2002). Deuterostome brains: synopsis and commentary.

Brain Res. Bull. 57, 257–270.

Noveen, A., Daniel, A., and Hartenstein, V. (2000). Early development of the

Drosophila mushroom body: the roles of eyeless and dachshund. Develop-

ment 127, 3475–3488.

Raible, F., Tessmar-Raible, K., Osoegawa, K., Wincker, P., Jubin, C.,

Balavoine, G., Ferrier, D.E., Benes, V., de Jong, P., Weissenbach, J., et al.

(2005). Vertebrate-type intron-rich genes in the marine annelid Platynereis

dumerilii. Science 310, 1325–1326.

Schneider, S.Q., and Bowerman, B. (2007). beta-Catenin asymmetries after all

animal/vegetal- oriented cell divisions in Platynereis dumerilii embryos

mediate binary cell-fate specification. Dev. Cell 13, 73–86.

Shimamura, K., Hartigan, D.J., Martinez, S., Puelles, L., and Rubenstein, J.L.

(1995). Longitudinal organization of the anterior neural plate and neural tube.

Development 121, 3923–3933.

Stigloher, C., Ninkovic, J., Laplante, M., Geling, A., Tannhauser, B., Topp, S.,

Kikuta, H., Becker, T.S., Houart, C., and Bally-Cuif, L. (2006). Segregation of

telencephalic and eye-field identities inside the zebrafish forebrain territory is

controlled by Rx3. Development 133, 2925–2935.

Strausfeld, N.J., and Barth, F.G. (1993). Two visual systems in one brain:

neuropils serving the secondary eyes of the spider Cupiennius salei.

J. Comp. Neurol. 328, 43–62.

Strausfeld, N.J., Hansen, L., Li, Y., Gomez, R.S., and Ito, K. (1998). Evolution,

discovery, and interpretations of arthropod mushroom bodies. Learn. Mem. 5,

11–37.

Strausfeld, N.J., Strausfeld, C.M., Loesel, R., Rowell, D., and Stowe, S. (2006).

Arthropod phylogeny: onychophoran brain organization suggests an archaic

relationship with a chelicerate stem lineage. Proc Biol Sci 273, 1857–1866.

Strausfeld, N.J., Sinakevitch, I., Brown, S.M., and Farris, S.M. (2009). Ground

plan of the insect mushroom body: functional and evolutionary implications.

J. Comp. Neurol. 513, 265–291.

Sugino, K., Hempel, C.M., Miller, M.N., Hattox, A.M., Shapiro, P., Wu, C.,

Huang, Z.J., and Nelson, S.B. (2006). Molecular taxonomy of major neuronal

classes in the adult mouse forebrain. Nat. Neurosci. 9, 99–107.

Swanson, L.W. (2000). Cerebral hemisphere regulation of motivated behavior.

Brain Res. 886, 113–164.

Taglialatela, P., Soria, J.M., Caironi, V., Moiana, A., and Bertuzzi, S. (2004).

Compromised generation of GABAergic interneurons in the brains of Vax1-/-

mice. Development 131, 4239–4249.

Tessmar-Raible, K., Steinmetz, P.R.H., Snyman, H., Hassel, M., and Arendt, D.

(2005). Fluorescent two-color whole mount in situ hybridization in Platynereis

dumerilii (Polychaeta, Annelida), an emerging marine molecular model for

evolution and development. Biotechniques 39, 460, 462, 464.

Tessmar-Raible, K., Raible, F., Christodoulou, F., Guy, K., Rembold, M.,

Hausen, H., and Arendt, D. (2007). Conserved sensory-neurosecretory cell

types in annelid and fish forebrain: insights into hypothalamus evolution. Cell

129, 1389–1400.

Tole, S., Ragsdale, C.W., and Grove, E.A. (2000). Dorsoventral patterning of

the telencephalon is disrupted in the mouse mutant extra-toes(J). Dev. Biol.

217, 254–265.

Tuoc, T.C., and Stoykova, A. (2008). Er81 is a downstream target of Pax6 in

cortical progenitors. BMC Dev. Biol. 8, 23.

Urbach, R., and Technau, G.M. (2003). Molecular markers for identified neuro-

blasts in the developing brain of Drosophila. Development 130, 3621–3637.

Viola, P.A., and Wells, W.M., III. (1995). Alignment by maximization of mutual

information. International Journal of Computer Vision 24, 137–154.

Wang, B., Waclaw, R.R., Allen, Z.J., 2nd, Guillemot, F., and Campbell, K.

(2009). Ascl1 is a required downstream effector of Gsx gene function in the

embryonic mouse telencephalon. Neural Dev 4, 5.

Woodger, J.H. (1945). On biological transformations. In Essays on Growth and

Form, presented to D’Arcy Wentworth Thompson, W.E. le Gros Clark and P.B.

Medawar, eds. (Oxford: Oxford University Press), pp. 95–120.

Yun, K., Potter, S., and Rubenstein, J.L. (2001). Gsh2 and Pax6 play comple-

mentary roles in dorsoventral patterning of the mammalian telencephalon.

Development 128, 193–205.

Cell 142, 800–809, September 3, 2010 ª2010 Elsevier Inc. 809

Page 178: Cell 100903

Resource

The Protein Composition of MitoticChromosomes Determined UsingMulticlassifier Combinatorial ProteomicsShinya Ohta,1,6 Jimi-Carlo Bukowski-Wills,1,2,6 Luis Sanchez-Pulido,4 Flavia de Lima Alves,1 Laura Wood,1 Zhuo A. Chen,1

Melpi Platani,1 Lutz Fischer,1 Damien F. Hudson,1,5 Chris P. Ponting,4 Tatsuo Fukagawa,3 William C. Earnshaw,1,*and Juri Rappsilber1,*1Wellcome Trust Centre for Cell Biology2Centre for Systems Biology at Edinburgh, School of Biological SciencesUniversity of Edinburgh, Edinburgh EH9 3JR, UK3Department of Molecular Genetics, National Institute of Genetics and The Graduate University for Advanced Studies, Mishima,

Shizuoka 411-8540, Japan4MRC Functional Genomics Unit, Department of Physiology, Anatomy and Genetics, University of Oxford, Oxford OX1 3QX, UK5Present address: Murdoch Children’s Research Institute, Royal Children’s Hospital, Melbourne, Victoria 3052, Australia6These authors contributed equally to this work

*Correspondence: [email protected] (W.C.E.), [email protected] (J.R.)DOI 10.1016/j.cell.2010.07.047

SUMMARY

Despite many decades of study, mitotic chromo-some structure and composition remain poorly char-acterized. Here, we have integrated quantitative pro-teomics with bioinformatic analysis to generatea series of independent classifiers that describe the�4,000 proteins identified in isolated mitotic chro-mosomes. Integrating these classifiers by machinelearning uncovers functional relationships betweenprotein complexes in the context of intact chromo-somes and reveals which of the �560 uncharacter-ized proteins identified here merits further study.Indeed, of 34 GFP-tagged predicted chromosomalproteins, 30 were chromosomal, including 13 withcentromere-association. Of 16 GFP-tagged pre-dicted nonchromosomal proteins, 14 were confirmedto be nonchromosomal. An unbiased analysis of thewhole chromosome proteome from genetic knock-outs of kinetochore protein Ska3/Rama1 revealedthat the APC/C and RanBP2/RanGAP1 complexesdepend on the Ska complex for stable associationwith chromosomes. Our integrated analysis predictsthat up to 97 new centromere-associated proteinsremain to be discovered in our data set.

INTRODUCTION

As cells enter mitosis, chromosomes undergo a remarkable

series of physiological and structural transformations known as

chromosome condensation. This process involves individualiza-

tion of the chromosomal territories to create the characteristic

mitotic chromosome morphology and maturation of the kineto-

chores so that chromosomes can align and segregate on the

mitotic spindle. Our understanding of the mechanisms under-

lying chromosome condensation is still fragmentary. These

processes can be fully understood only when all components

of mitotic chromosomes have been identified and functional

relationships between them determined. We have developed

a new approach that we term multiclassifier combinatorial pro-

teomics (MCCP) to do this.

The current list of mitotic chromatin proteins reported in pro-

teomic studies is surprisingly short. Early analyses described

62 and 79 proteins, respectively, in mitotic chromosome scaf-

folds (Morrison et al., 2002; Gassmann et al., 2005). A later study

identified > 250 proteins that bound to sperm chromatin in

Xenopus egg extracts in vitro, revealing the kinetochore protein

Bod1 (Porter et al., 2007). Other studies identified �240 proteins,

subsequently corrected to roughly 50 bona fide putative struc-

tural proteins (Uchiyama et al., 2005; Takata et al., 2007). In a tar-

geted study, 98 proteins were identified as shared in isolated

telomeres from wild-type and ALT cells (Dejardin and Kingston,

2009). Despite these efforts, currently available proteomics

reports miss a significant fraction of known mitotic chromosomal

proteins, particularly kinetochore components.

Biochemical analysis of important chromosome substruc-

tures such as kinetochores is extremely challenging. The

kinetochore is one of the most complex cellular substructures

(Cheeseman and Desai, 2008), with over 120 constituents

described by a range of approaches (Earnshaw and Rothfield,

1985), including targeted proteomic studies (Obuse et al.,

2004; Foltz et al., 2006; Okada et al., 2006; Hori et al., 2008;

Amano et al., 2009). Biochemical dissection of kinetochores is

complicated by the fact that it is not known to what extent

the constituent protein complexes can be recovered in soluble

form from chromosomes with their relevant intermolecular

associations intact. As described here, those problems can

be circumvented by purifying and analyzing whole mitotic

chromosomes.

810 Cell 142, 810–821, September 3, 2010 ª2010 Elsevier Inc.

Page 179: Cell 100903

Purifying large cellular structures or organelles free of con-

taminants is virtually impossible. Genuine components have

been distinguished from contaminants in such preparations by

subtractive (Schirmer et al., 2003) or quantitative (Foster et al.,

2003) proteomics by determining the difference between two

near-identical fractions, one enriched and the other depleted of

the target structure. In protein correlation profiling, a set of known

components was used to define a common intensity profile

across neighboring biochemical fractions from sucrose gradients

during purification of organelles and this was used to select other

proteins that show a similar profile (Andersen et al., 2003). These

methods do not recognize cellular background proteins that

adhere to the structure of interest due to nonspecific hydro-

phobic or electrostatic interactions. This type of contamination

is particularly relevant for vertebrate mitotic chromosomes.

In the present study, we identified approximately4000 polypep-

tides in highly purified chromosomes. We developed a statistical

approach for analysis of proteomic data to confirm which known

and uncharacterized proteins from this long list are chromosomal.

An experimental test of our method led to the identification of 32

chromosomal proteins, including 13 kinetochore-associated

proteins. Key to our analysis is the innovative use of stable isotope

labeling with amino acids in cell culture (SILAC) (Ong et al., 2002),

plus development of a framework to integrate data from multiple

classifiers, including nonproteomic classifiers, to reveal proteins

of interest and determine functional relationships between protein

complexes in the context of whole chromosomes.

RESULTS

Identification of the Proteome of IsolatedMitotic ChromosomesWe isolated mitotic chromosomes from chicken DT40 cells by

a refinement of the polyamine method (Lewis and Laemmli,

1982) (Figures 1A and 1B). These preparations are negative for

porin in immunoblots (Figure 1C), indicating that mitochondrial

contamination (common in chromosome preparations) is low.

Proteomic analysis (Cox and Mann, 2008; de Godoy et al.,

2008) of 250 mg of total chromosomal protein identified 4029

proteins in 28 functional categories (Figure 1D and Table S1

available online), including essentially all previously described

chromosomal proteins.

When vertebrate cells enter mitosis, the nuclear envelope

breaks down and chromosomes are newly exposed to cyto-

plasmic proteins, organelles and cellular membranes. Since

highly positively charged histones contribute �38% of the chro-

mosome mass and an equivalent amount is highly negatively

charged DNA, many charged nonchromosomal proteins exhibit

strong adventitious binding to chromosomes. These ‘‘hitch-

hikers’’ differ from conventional contaminants (e.g., mitochon-

dria), as they are physically associated with the chromosomes

before cell lysis and apparently cannot be separated by conven-

tional purification protocols. Thus, the 1331 cytoskeletal, cyto-

plasmic, mitochondrial, membrane and receptor proteins found

in our preparations, may be physically associated with chromo-

somes following nuclear envelope disassembly, but many not be

functionally relevant.

A Classifier Approach to Identify Genuine MitoticChromosome ProteinsThe presence of hitchhiker proteins complicates the definition

of what constitutes a ‘‘true’’ chromosomal protein, as well as

the design of biochemical control experiments. For example,

comparing mitotic to interphase chromatin is of limited use,

since the latter is shielded from the cytoplasm by the nuclear

envelope. In such a comparison, cytoplasmic hitchhiker proteins

would be scored as mitosis-specific chromosomal proteins.

Here, we describe an approach to study the complex chromo-

somal proteome that both identifies proteins that merit further

study and reveals functional relationships between all chromo-

somal proteins. We quantify the chromosomal association of

each protein in a series of quantitative proteomics experiments,

mostly using SILAC technology (Ong et al., 2002). Each experi-

ment provides an independent measure of a protein’s associa-

tion with mitotic chromosomes, which we term a ‘‘classifier.’’

Integration of the data obtained with all classifiers enables us

to detect patterns in the behavior of groups of proteins that

reveal shared membership in protein complexes as well as func-

tional dependency relationships.

The experimental protocols that define the five proteomic

classifiers are shown in Figure 1E and all classifiers are summa-

rized below.

Classifier I: Abundance estimation

To estimate the amounts of individual proteins in mitotic chromo-

somes, we used an established protocol (Rappsilber et al., 2002;

Ishihama et al., 2005, 2008) to calculate a scaled protein abun-

dance index based on the number of peptides observed and

the number of times that each peptide is observed (spectral

count) for each protein. This calculation and its validation are dis-

cussed in Extended Experimental Procedures.

In the conventional pie chart of Figure 1D, all proteins are

weighted equally, independent of their actual abundance in iso-

lated chromosomes. A more informative representation of chro-

mosome composition is obtained by normalizing each class by

its mass, obtained by multiplying the estimated abundance by

the predicted molecular mass of each protein (Figure 1F). As

expected, histones comprise the bulk of mitotic chromosomal

protein (48%). Overall 68% of the protein mass is annotated as

chromosomal.

Classifier II: Enrichment in Chromosomes

We expected core chromosomal components like histones or

structural proteins would be more abundant in isolated chromo-

somes than in cytoplasmic extracts. The reverse would be true of

background proteins. We therefore mixed isolated chromo-

somes from mitotic DT40 cells grown in light medium with an

equal mass of protein from post-chromosomal extracts of

parallel cultures grown with heavy SILAC medium (Figure 1E).

Classifier II was calculated as the ratio of light/heavy peaks for

each protein. Among the 20% most enriched proteins, chromo-

somal proteins outnumbered nonchromosomal proteins by 3 to

1. Conversely, among the 20% least enriched proteins, back-

ground proteins outnumbered chromosomal proteins 6 to 1.

Classifier III: In Vitro Exchange on Chromosomes

We ranked proteins based on their ability to stably bind to chro-

mosomes during an incubation in cytosol. A crude light chromo-

some fraction obtained by gentle centrifugation was mixed with

Cell 142, 810–821, September 3, 2010 ª2010 Elsevier Inc. 811

Page 180: Cell 100903

an excess of heavy post-chromosomal extract and incubated

to allow proteins to exchange at 14�C for 30 min (Figure 1E).

The chromosomes were then subjected to rigorous purification.

All heavy proteins identified must have bound to the chromo-

somes during the incubation in vitro. Classifier III is the light/

heavy ratio for each protein identified in this experiment.

The most stable chromosomal-associated proteins were

histones (average classifier III = 63), topoisomerase IIa (classifier

Figure 1. Proteomic Analysis of Mitotic

Chromosome Proteins

(A) Outline of chromosome isolation procedure.

(B) Silver-stained gel of isolated chromosomes.

(C) Immunoblot with markers for mitochondria

(porin) and chromosomes.

(D) The 28 classes of proteins found in chromo-

somes.

(E) Experimental designs to prepare samples

for determination of classifiers I (abundance), II

(enrichment in chromosomes versus post-chro-

mosomal extract), III (exchange onto chromo-

somes from post-chromosomal cytosol), IV

(dependency on SMC2/condensin) and V (depen-

dency on Ska3/RAMA1).

(F–J) (F) Estimated percentages of total chromo-

somal protein mass in the major classes of

proteins. Sample spectra used to calculate classi-

fier III (exchange) values for (G) Ribosomal protein

SA (a hitchhiker protein), (H) MAD2, (I) BUBR1

and (J) MAD1. Spectra are plotted as the peak

intensity (a measure of abundance) against the

mass divided by charge. Peaks of light peptides

from chromosomes are indicated by blue bars

and heavy peptides bound from cell extracts

during the incubation are indicated by red bars.

See also Figure S1.

III = 61) and condensin I (average classi-

fier III = 59). Interestingly, condensin II

was more exchangeable (average classi-

fier III = 14). In contrast, the ratios for

�75% of ribosomal proteins were in the

range from 0.45 to 3.0 (Figure 1G), indi-

cating significant binding to chromo-

somes from cytosol during the incubation.

Although we sought to optimize purity

rather than preserve functionality of

chromosomes, the exchange experiment

revealed that at least one aspect of kinet-

ochore function was retained in purified

chromosomes. Similar to one recent

study (Kulukian et al., 2009), kinetochores

of the purified chromosomes can recruit

spindle checkpoint proteins Mad2, Bub3,

and BubR1 from cytosol (Figures 1H and

1I) but not Mad1 (Figure 1J).

Classifier IV: SMC2 Dependency

We used a conditional genetic knockout

of SMC2 in DT40 SMC2ON/OFF cells (Hud-

son et al., 2003) to compare the composi-

tion of mitotic chromosomes formed in the presence or absence

of condensin, which is required for structural integrity of mitotic

chromosomes. DT40 SMC2ON cells were cultured in SILACheavy

medium. To obtain chromosomes depleted of condensin, cells

grown in SILAClight medium were cultured with doxycycline for

30 hr to shut down SMC2 expression prior to the nocodazole

block (SMC2OFF). Equal numbers of mitotic cells from the two

different populations were mixed and mitotic chromosomes

812 Cell 142, 810–821, September 3, 2010 ª2010 Elsevier Inc.

Page 181: Cell 100903

isolated (Figure 1E). Classifier IV is the heavy/light ratio

(SMC2ON/SMC2OFF) for each of the proteins identified in this

experiment.

SMC2-depleted chromosomes contained 5.8% of the wild-

type level of SMC2. It is not possible to isolate chromosomes

from cultures completely lacking SMC2 as SMC2 is an essential

gene and dead cells do not accumulate in mitosis. These chro-

mosomes were similarly depleted of all condensin I and II

subunits.

Classifier V: Ska3/Rama1 Dependency

To demonstrate the targeted analysis possible with our

approach, we compared the association of kinetochore proteins

with chromosomes in cells with or without Ska3/Rama1/

C13orf3. Ska3/Rama1, which was identified in this analysis as

a chromosomal protein, was described in several recent publica-

tions (Daum et al., 2009; Gaitanos et al., 2009; Raaijmakers et al.,

2009; Theis et al., 2009; Welburn et al., 2009).

To obtain chromosomes depleted of Ska complex, we isolated

a genetic knockout of the Ska3/Rama1 gene (Figures S5A and

S5B). Ska3/Rama1�/� cells (homozygous knockouts are viable)

were grown in SILAClight medium (Figure 1E). Equal numbers of

Ska3/Rama1�/� and wild-type DT40 (cultured in SILACheavy

medium) mitotic cells from the two different populations were

mixed and mitotic chromosomes isolated. Classifier V is the

heavy/light ratio (wild-type/Ska3/Rama1�/�) for each protein

identified in this experiment.

Classifier VI: Domain Analysis

We added an additional nonproteomic classifier to our analysis

using the protein domains found in chromosomal and nonchro-

mosomal proteins (red/pink and green wedges in Figure 1D).

This made use of bioinformatic analysis in order to segregate

chromosomal from nonchromosomal proteins, but importantly

did not consider a protein’s relevance to mitosis. We counted

how often each domain was observed in chromosomal and

nonchromosomal proteins and assigned it this frequency as

a score (Table S2). Classifier VI was then determined for each

protein based on the sum of its domain scores.

Multiclassifier Combinatorial ProteomicsTraditional one-dimensional analysis (e.g., sorting the various

proteins according to their value for each classifier) was of

limited utility, as the data lacked a clear boundary between chro-

mosomal (red/pink in Figure 2A) and nonchromosomal proteins

(green in Figure 2A) for each classifier (Figure 2A and Figures

S2A–S2E).

By contrast, when classifiers were combined, our ability

to identify chromosomal proteins was vastly improved. For

example, an enrichment of centromeric or chromosomal pro-

teins relative to nonchromosomal proteins was obtained when

classifiers I (abundance) and II (enrichment) were plotted

(Figures 2C–2E). The clustering of protein complex subunits

in this plot (Figure 2D) reflects both their relative stoichiometry

Figure 2. Combining Classifiers Increases

Specificity

(A) The rank of all proteins observed for proteomic

classifiers I (abundance), II (enrichment), III (reten-

tion), IV (SMC2 dependency), V (Ska3/Rama1

dependency) and RF (combined random forest

ranking) is plotted as vertical color coded lines

(rug plots). Proteins are ranked from the highest

(left) to lowest values for each classifier. Centro-

mere proteins are indicated by longer bars

(indicated).

(B) Cumulative curves show that combining classi-

fiers by RF significantly increases the specificity

of identifying chromosomal proteins without com-

promising the completeness of the analysis.

(C) 2D scatter graph plotting classifiers I versus II

flanked by the relevant rug plots. Proteins most

enriched in chromosomes relative to cytoplasm

and proteins most abundant in chromosomes are

above and to the right, respectively. Each spot is

color-coded by category (Figure 1D).

(D) Zoom of the region indicated in (C) showing

members of the chromosomal passenger (CPC),

Mis12, Ndc80, Nup107-160 and Cohesin com-

plexes plus the constitutive centromere-associ-

ated network (CCAN) and anaphase-promoting

complex/cyclosome (APC/C).

(E) Enrichment for chromosomal proteins in the

region indicated in (C & D).

See also Figure S2.

Cell 142, 810–821, September 3, 2010 ª2010 Elsevier Inc. 813

Page 182: Cell 100903

(x axis), and the similar degree to which subunits in a complex are

all present either on or off chromosomes (y axis). Members of the

APC/C, Ndc80 and Mis12 complexes form closely knit clusters.

It is important to note that this was achieved in the context of

entire chromosomes and without requiring solubilization of the

complexes.

We used random forest (RF) analysis, a machine learning

approach, to progress beyond two-dimensional analyses and

integrate the information present in all proteomics classifiers.

This analysis offered two powerful benefits. First, it enabled us

to work with data sets that contain missing values. This is a signif-

icant advantage in proteomics studies where not every protein is

observed in every experiment, as seen in Figure 2A and Fig-

ure S2F. Second, RF analysis allowed us to use any descriptor

of our proteins as a classifier and integrate it into our overall

analysis. Here, we also included a bioinformatic analysis of the

distribution of protein domains in our analysis distinguishing

chromosomal from nonchromosomal proteins (classifier VI).

In brief, RF is a decision tree analysis that separates data sets

into ‘‘true’’ and ‘‘false’’ groups. The decision trees are trained on

defined data sets and randomly built to optimize the separation

between them. Analysis of the experimental data set then occurs

by running each protein through all trees and adding up its overall

RF score (i.e., the fraction of trees that scored it as ‘‘true’’). RFs

perform much better on training data than application data, so

their performance is evaluated by ten-fold cross-validation.

The training data are split into random sections of 90% for

training and 10% for evaluation, so that successively the entire

set is used for evaluation. Here, the two training data sets chosen

were ‘‘nonchromosomal’’ (green wedges in Figure 1D) and

‘‘chromosomal’’ (red + pink wedges in Figure 1D), and the RF

score for a given protein is the fraction of trees that scored it

as ‘‘chromosomal.’’

RF analysis readily discriminated chromosomal from nonchro-

mosomal proteins. In the RF rug plot of Figure 2A, which repre-

sents the ranked list of proteins generated by RF analysis, the

left side is predominantly red, while the right side is predomi-

nantly green. To reach the 500th chromosomal protein on the

RF-ranked list only 229 nonchromosomal proteins are included

(Figure 2B). In contrast, 410-671 nonchromosomal proteins

would be included when considering ranked lists from individual

classifiers. Note that the RF-based sorting was done on the

complete data set, including proteins that failed to be observed

with some classifiers. Therefore, adding information from addi-

tional experiments did not decrease the number of proteins

covered.

The advantage of combining classifiers can be statistically

expressed by ROC curves (Figure S3A), with increased area

under the curve (AUC) for our combined analysis when com-

pared to each of the individual classifiers (AUCRF(cI-V) = 0.81,

AUCscI-V = 0.41-0.76). The combined classifiers assigned 88.8%

of our gold standard, the 125 centromere proteins, correctly as

chromosomal, at the cut-off that minimizes misassignment of

chromosomal versus nonchromosomal proteins. This specificity

was further improved when bioinformatic domain analysis (cVI)

was integrated with our proteomic classifiers (AUCRF(cI-VI) =

0.97; identification of the first 500 chromosomal proteins yields

17 nonchromosomal proteins, Suppl. Figures S3B and S3C,

proteins lacking known domains are excluded from this boost,

Figure S3D). Now, 92.4% of the centromere proteins were

assigned as chromosomal. In summary, RF analysis provides

us with a tool for productively combining the outcomes of our

individual proteomics classifier experiments and further empow-

ering our analysis by including data from other sources.

If results of a random forest based on the five proteomic clas-

sifiers plus the bioinformatics-based classifier VI were plotted

against those from the initial random forest analysis (Fig-

ure 3A), a near-perfect separation of the training data was

achieved. Only a single chromosomal protein of the training set

and two nonchromosomal proteins were misassigned when

placing manually a separation line.

Using ten-fold cross-validation, we found that 118 centromere

proteins positioned right and only 7 left of the separation line

(Figure 3B). This compares to 14 and 9 centromere proteins

being missed when using the one-dimensional ranked lists by

classifiers I-V and I-VI, respectively. Accepting the line as a

threshold returns known centromere proteins with a yield of

94.4% and all other chromosomal proteins with 93.1% success.

In contrast, 83.1% of nonchromosomal proteins are rejected.

Thus, the classifier approach is sufficiently powerful to suggest

chromosomal proteins from among hitherto uncharacterized

proteins.

Identification of New Chromosomal ProteinsTo test the predictive power of our RF analysis, we cloned and

tested the location in mitosis of 50 previously uncharacterized

proteins including 15 without known domains: 34 predicted to

be and 16 predicted not to be on chromosomes in mitosis.

Reasoning that important proteins would be conserved, we

expressed GFP-tagged human homologs of these chicken

proteins in U2OS cells. Remarkably, 30 of 34 cloned proteins

from the chromosomal region were confirmed as chromosomal,

contrasting with only 2 of the 16 predicted ab initio to be

nonchromosomal. This confirms the power of our analysis and

indicates a success rate of 88%, with 44 of 50 tagged proteins

localizing as predicted. Of 50 newly cloned proteins, 13 were

associated with kinetochores in mitosis, 12 had a more general

distribution on mitotic chromosomes and 7 others were peri-

chromosomal, a class whose new members we propose to

term chromosome periphery proteins (cPERPs A-G) (Figures

S4B–S4G and Table S3). The chromosome periphery (perichro-

mosomal layer) is enriched in ribosomal and nucleolar hitchhiker

proteins, and is of unknown function (Van Hooser et al., 2005).

The new centromere proteins all appeared to localize to the

outer kinetochore, relative to CENP-C and HEC1 as standards

(Figures 4B–4J, Figure S4A). In keeping with established nomen-

clature, we propose to name these proteins CENP-Y, CENP-Z

and CENP-27 through CENP-33. Beyond ‘Z’, the 26th letter of

the basic modern Latin alphabet, we propose to designate the

new proteins with numbers starting with CENP-27.

Functional Analysis of New Kinetochore-AssociatedProteinsWe focused our initial functional analysis on kinetochore

proteins. Clustering analysis (Gentleman et al., 2004) allowed

us to combine data for proteins identified by all classifiers, and

814 Cell 142, 810–821, September 3, 2010 ª2010 Elsevier Inc.

Page 183: Cell 100903

look for informative groupings. This revealed a striking tendency

for functionally related proteins to form clusters, as exemplified

by members of the NDC80, CPC, Nup107-160 and APC/C

complexes (Figure 4). Interestingly, our clustering sorted

CENP-27 as a component of the APC/C. This was confirmed

and the protein named APC16 in three recent reports (Hutchins

et al., 2010; Kops et al., 2010; Hubner et al., 2010). To further

test the predictive value of cluster analysis for proteomic data

sets, we examined two kinetochore proteins in greater detail.

Ska3/Rama1 and Functional Analysis of Kinetochore

Subcomplexes

C13orf3 was located adjacent to Ska2 in the cluster analysis.

This protein now known as Ska3/Rama1 has been suggested

to be involved in microtubule attachment to kinetochores (Gaita-

nos et al., 2009; Raaijmakers et al., 2009; Welburn et al., 2009) or

coordination of the spindle checkpoint response (Daum et al.,

2009; Theis et al., 2009). We analyzed the kinetochore proteome

in the presence or absence of Ska3/Rama1 (defined as classifier

V) in order to determine the role of this protein in kinetochore

structure (Figure 5).

A map of the Ska3/Rama1 locus in DT40 is shown in Figures

S5A and S5B, together with a targeting strategy for inactivating

the gene. The Ska3/Rama1 gene is not essential for life in DT40

cells (Figure S5C). However, these cells struggle to achieve

a normal chromosome alignment, and show a �33 increase in

mitotic index (Figure S5D and S5E), a �33 increase in the

percentage of apoptotic cells and a �6x increase in the number

of bi-nucleated cells.

Proteomic analysis of isolated chromosomes revealed that

loss of Ska3/Rama1 was accompanied by the loss of Ska1 and

Ska2. Loss of the Ska complex caused no systematic changes

in the chromosomal association of proteins of the constitutive

centromere-associated network (CCAN), Knl-1/Mis12/Ndc80

(KMN), Mis18, Ndc80, CPC, and Nup107-160 kinetochore sub-

complexes. However, striking changes were seen in the levels

of the APC/C, RanBP2/RanGAP1, spindle checkpoint, Rod/

Zw10/Zwilch (RZZ), and dynein/dynactin complexes. We con-

firmed that the RanBP2/RanGAP1 complex is indeed depleted

from kinetochores when Ska3/Rama1 is deleted in HeLa cells

(Figure S5F and S5G). Attempts to confirm the specific kineto-

chore depletion of the APC/C were uninformative, as we were

unable to reproducibly obtain kinetochore staining for the

APC/C in HeLa cells using four independent antibodies.

We conclude that combining genetic and SILAC analysis

provides a powerful new method for analysis of multicomplex

protein superstructures.

A Protein Involved in Chromosome Alignment

and Spindle Organization

A second new kinetochore-associated protein, CENP-32/

C9orf114, was sorted on our kinetochore cluster diagram next

to CLASP1 and CLASP2, two paralogues known to be involved

in the regulation of microtubule dynamics. Like CLASPs,

CENP-32/C9orf114 mapped to the outer kinetochore (Figure 6A),

and its depletion caused a significant accumulation of cells in

later prometaphase (Figure S6A) with misaligned chromosomes

(Figure 6B). These cells frequently had bipolar spindles, however

60% of those spindles exhibited remarkable abnormalities

where centrosomes appeared to have detached from the poles

(Figures 6C and 6D and Figure S6B). In one remarkable case,

the centrosomes appeared at the midzone of a bipolar spindle

(Figure 6C9-12).

Figure 3. Random Forest Analysis Predicts New Proteins of Interest

(A) Separation of the two training sets (color-coded groups from Figure 1D) by

RF analysis. The line provides optimal separation of the two training sets.

x axis: RF analysis of classifiers I–V; y axis: RF analysis of classifiers I–VI,

including the bioinformatic domain analysis.

(B) Positions of previously known centromere proteins from ten-fold cross-

validation.

(C) Position of cloned proteins and remaining uncloned proteins in the same 2d

analysis.

(D) Classification of newly cloned proteins from regions to the left and right of

the dividing line, respectively.

See also Figure S3.

Cell 142, 810–821, September 3, 2010 ª2010 Elsevier Inc. 815

Page 184: Cell 100903

DISCUSSION

Multiclassifier Combinatorial ProteomicsThe approach described here for analysis of the proteome of

vertebrate mitotic chromosomes can be used to study any

complex proteome. The key approach of combining classifier

data can in principle be expanded indefinitely and can include

nonproteomics data sets such as the bioinformatic protein

domain analysis used here. Other classifiers that could

be used in the future include microarray, protein interaction

Figure 4. Cluster Analysis and Imaging of New Kinetochore Proteins

(A) Heat map and cluster dendrogram for 101 centromere-associated proteins identified by all 5 proteomic classifiers. Known complexes are color-marked (see

bottom) and show a tendency to cluster in this analysis. Kinetochore proteins identified in this study are also marked (stars).

(B–J) Localization of GFP-tagged kinetochore-associated proteins (panels 2) relative to DNA detected with DAPI (panels 1) and CENP-C (panels 3 in [B]–[D], [F]–

[H], and [J] or ACA (panels 3 in [E] and [I]). Insets show merged images with blow-ups of representative kinetochores (GFP-novelCENP, green and CENP-C or

ACA, red).

See also Figure S4.

816 Cell 142, 810–821, September 3, 2010 ª2010 Elsevier Inc.

Page 185: Cell 100903

(e.g., two hybrid screens or pull-down), protein phosphoryla-

tion, and localization data and, indeed, data from any experi-

mental approach in which the proteins of interest are sorted

systematically.

We first showed that plotting pairs of classifiers against one

another improved our ability to delineate potential chromosomal

proteins. As that approach could not be generalized when the

number of classifiers exceeded three, we adopted a random

forest (RF) analysis approach. This allowed us to integrate infor-

mation from all classifiers into decision trees on which known

and unknown proteins could be classified. Importantly, RF anal-

ysis handles missing values systematically. This is crucial when

not every protein is observed in every experiment. In contrast,

cluster analysis, which has been used both in this study and in

other recent work (Theis et al., 2009; Neumann et al., 2010),

can only integrate data for proteins that have a value for every

classifier.

Integrity of the Isolated ChromosomesOur methods focused on optimizing the purity of the chromo-

somes. Thus, our list of proteins is likely to represent the minimal,

stably associated components of mitotic chromosomes. None-

theless kinetochores of isolated chromosomes retain some

function, as judged by their ability to recruit components of

the mitotic checkpoint complex from cytoplasm. This may be

because chromosomes were isolated from nocodazole-treated

cells, with kinetochores actively engaged in spindle checkpoint

signaling.

New Insights into Kinetochore Functional OrganizationRemarkably, although no biochemical enrichment for centro-

meres was performed, our data set contained all known centro-

meric subcomplexes, with peptides from 125 reported cen-

tromere proteins (eight present as multiple isoforms) (Table S1).

We identified all members of the CCAN, KMN, Mis12 and Mis18

complexes and all members of the RZZ complex except Zwint

(which is not yet annotated in the chicken genome). Our success

in identifying centromere and telomere proteins may be

explained because 66 of the 78 chicken DT40 chromosomes

are microchromosomes whose purification provides a natural

enrichment for centromeres and telomeres since the chromo-

some arms are so short.

We combined genetics with whole proteome analysis in order

to identify complexes and structural dependencies in their

‘‘native environment’’ (e.g., kinetochore proteins in actual kinet-

ochores). Chromosomes lacking Ska3/Rama1 were depleted for

the entire Ska complex, confirming that these three proteins are

interdependent for chromosome binding. Similarly, depletion of

key condensin subunit SMC2 caused a loss of all seven subunits

of the condensin I and II complexes from chromosomes. Impor-

tantly, this analysis did not require tagging of any proteins or

attempts to solubilize functional complexes from large subcel-

lular structures.

In addition to these primary effects of depletion, loss of Ska3/

Rama1 also caused a significant secondary depletion of the

APC/C and RanBP2/RanGAP1 complexes from chromosomes

but had no consistent effect on most other kinetochore proteins.

Importantly, all members of the secondary-depleted complexes

Figure 5. Ska3/Rama1 Dependency for Chromosomal Association

and Analysis of CENP-27/APC16

Kinetochore proteins with increased abundance on chromosomes in the

absence of Ska3/Rama1 are shown with blue bars and proteins with

decreased abundance are shown with red bars. Note covariance of protein

complexes. See also Figure S5.

Cell 142, 810–821, September 3, 2010 ª2010 Elsevier Inc. 817

Page 186: Cell 100903

Figure 6. Initial Characterization of CENP-32

(A) Localization of GFP-CENP-32 (panel 2, green) in the kinetochore relative to DNA (panel 1, blue) and CENP-C (panel 3, red).

(B) CENP-32 RNAi causes problems with chromosome alignment. DNA, red; CENP-C, green; microtubules, white.

(C) CENP-32 RNAi causes detachment of centrosomes from bipolar spindles. DNA (panels 1,5,9,13), blue; pericentrin (panels 2,6) or g-tubulin (panels 10,14),

green; a-tubulin (panels 3,7,11,15), red.

(D) Quantitation of spindle phenotypes following CENP-32 RNAi.

(E) Schematic representation of domain architectures for new kinetochore proteins tagged in this study (drawn to approx. scale). The locations of domains are

according to Pfam and SMART family databases (Letunic et al., 2004; Finn et al., 2010), complemented by REP web server analysis (Andrade et al., 2000).

The localization of the PEHE domain has been assigned from (Marın, 2003). Phyletic distributions of proteins are indicated in blue, yellow, green and violet

for Metazoa, Fungi, Plantae and Archaea, respectively. CENP-35 has been truncated. Abbreviations: BCNT, Bucentaur or craniofacial development protein;

Fbox, cyclin-F motif; G-Patch, Glycine-rich nucleic binding domain; JmjC, jumonji C domain; LRR, Leucine-rich repeats; Nua4, Nucleosomal acetyltransferase

of H4; OB, Oligonucleotide/oligosaccharide Binding; PEHE, Pro-Glu-His-Glu conserved region; PHD, Plant HomeoDomain; RA, Ras associated (RalGDS/AF-6)

domain; SPOUT, spoU and trmD RNA methyltransferase; WD40, WD/beta-transducin repeats; Zf-AD, Zinc finger-associated domain; and Zf-C2H2, Classical

zinc finger Cys(2)His(2).

See also Figure S6.

818 Cell 142, 810–821, September 3, 2010 ª2010 Elsevier Inc.

Page 187: Cell 100903

also behaved coordinately. Our results (e.g., the behavior of

CENP-27/APC16 as a component of the APC/C) confirm the

utility of single protein depletion analysis for the identification

of protein complexes and determination of their mutual interde-

pendencies for association with chromosomes.

Our data suggest that the Ska complex may provide a docking

site for the APC/C in the outer kinetochore. Alternatively, sumoy-

lation by RanBP2 may have a role in APC/C binding to chromo-

somes. The RanBP2-RanGAP1 complex is known to be involved

in kinetochore-microtubule interactions and localization of

several spindle checkpoint proteins (Joseph et al., 2004). We

note that among the recent spate of publications on Ska3/

Rama1, our observations appear to support a role in integration

and regulation of the spindle checkpoint response (Daum et al.,

2009; Theis et al., 2009).

Our whole-proteome analysis revealed that Cdc20 behaved

like a member of the APC/C and was distinct from other compo-

nents of the spindle checkpoint pathway with respect to its Ska

complex dependency. Spindle checkpoint components asso-

ciate with one another in cytoplasm as a mitotic checkpoint

complex (MCC), containing BubR1, Bub3, Cdc20 and Mad2.

Our data suggest that once the MCC associates with chromo-

somes, Cdc20 stably associates with the APC/C.

What Classes of Kinetochore Proteins Remainto Be Discovered?We identified 13 kinetochore-associated proteins among previ-

ously uncharacterized proteins and, as discussed below, we

predict that many more remain to be described. We therefore

asked whether there is any functional relationship between these

new proteins. That is, what sorts of kinetochore proteins had

been missed in the many previous genetic and biochemical

screens? An interesting answer has emerged.

Since the new kinetochore proteins were identified solely

based on their occurrence in chromosomes, they could poten-

tially represent a wide range of functions. Nevertheless, it is

striking that five of the new centromere proteins (namely,

CENP-28, �29, �31, �35, and �36) are subunits of complexes

that modify and/or bind histones (Figure 6E). Yeast orthologs

of two of these proteins (namely, CENP-28/C1orf149 and

CENP-29/CFDP1) contribute to NuA4 histone acetyltransferase

(HAT) and SWR1 ATP-dependent chromatin remodelling com-

plexes, respectively. These complexes are known to share com-

ponents (Wu et al., 2005) and together stimulate the exchange of

histone H2A for H2A.Z, following acetylation of H2A or H4 (Altaf

et al., 2010). A third centromere-associated protein, CENP-36/

MSL1v1, is necessary for the activity of MOF, another HAT, on

nucleosomal H4 (Li et al., 2009). Finally, CENP-31/PHD6 and

CENP-35/PHF2 each contain PHD (plant homeodomain) zinc

fingers, which are usually associated with chromatin-mediated

transcriptional regulation. The PHD of CENP-35, which also

contains a JmjC (likely histone demethylase) domain, appears

to be required for demethylation of H3 at the promoters of ribo-

somal RNA genes (Wen et al., 2010).

Why were these proteins not discovered earlier as kineto-

chore-associated? One likely explanation is that they may have

essential functions in other chromatin regions as well. Thus,

mutations might have pleiotropic phenotypes not recognized

as specific for mitosis. Furthermore, their association might

depend on a fully assembled kinetochore and thus be lost

when attempting other than whole chromosome analysis.

Characterization of a Kinetochore ProteinCENP-32 is required both for chromosome alignment and for

association of the centrosomes with the poles of the bipolar

spindle during metaphase. This latter phenotype is very similar

to an unusual spindle morphology phenotype seen in Drosophila

cells following depletion of the CLASP homolog Mast/Orbit

(Maiato et al., 2002). Indeed, in our analysis, CENP-32 clusters

with CLASP1 and CLASP2. A yeast homolog of CENP-32 inter-

acts with CBF5, an enzyme involved in the posttranscriptional

modification of rRNA, that was shown to bind to budding yeast

centromeres and microtubules (Jiang et al., 1993). Bioinformatics

analysis suggests that CENP-32 is a member of the SPOUT family

of methyltransferases but is atypical in possessing a possible

RNA-binding OB fold inserted into its catalytic domain (Tkaczuk

et al., 2007). It is tempting to speculate that CENP-32 may func-

tion at kinetochores by interacting with an as-yet unknown RNA.

How Complex Is the Kinetochore?MCCP analysis allows us to predict how many more kinetochore

proteins remain to be identified in our data set. In the plot of

Figure 3, where the chromosomal proteome is displayed in two

dimensions, we found 35% of novel tagged proteins from region

R and 6% from region L to associate with the kinetochore during

mitosis. These regions have 224 and 287 as-yet uncharacterized

novel proteins, respectively. Assuming no bias in the proteins we

cloned, this suggests that approximately 97 more kinetochore

proteins remain to be discovered. Taking into account the 13

kinetochore-associated proteins confirmed in our work, this

roughly doubles the currently known protein complexity of the

kinetochore during mitosis, confirming it as one of the most

complex cellular substructures.

ConclusionsMulticlassifier combinatorial proteomics and the data sets

described here open the door to the identification of all functional

components of mitotic chromosomes despite the adventitious

binding of cellular background proteins during mitosis. Further-

more, MCCP can be extended by adding additional classifiers

to delineate protein complexes and define functional dependen-

cies between them in the context of intact mitotic chromosomes.

This will serve both as a starting point for systematic determina-

tion of the full range of functions involved in mitotic chromosome

segregation, and as a basis for the development of detailed

structural and functional interaction maps of key chromosomal

subdomains. MCCP should also prove useful for the analysis

of other cellular structures that lack defined boundaries, e.g.,

membrane associated complexes like the post-synaptic density.

EXPERIMENTAL PROCEDURES

Preparation of Mitotic Chromosomes

DT40 cells were incubated with Nocodazole for 13 hr, resulting in a mitotic

index of 70%–90%. Mitotic chromosomes were isolated in the polyamine-

EDTA buffer system optimized for chicken DT40 cells (Lewis and Laemmli,

Cell 142, 810–821, September 3, 2010 ª2010 Elsevier Inc. 819

Page 188: Cell 100903

1982). 19.3 OD260 units were obtained from pooling the material of four inde-

pendent preparations totaling 7.5 3 109 DT40 cells and solubilized in SDS-

polyacrylamide gel electrophoresis (SDS-PAGE) sample buffer.

Preparation of Chromosome-Free Mitotic Cell Extracts

Nocodazole blocked DT40 cells were dounce-homogenized under hypotonic

conditions. Mitotic chromosomes were removed by centrifuging the superna-

tant twice at 10,000 x g and discarding the pellets to prepare a cell extract free

of chromosomes.

To measure the ratios between chromosomal and nonchromosomal

proteins, SILAC based mass spectrometry was performed with 150 mg of

labeled cell extract from 7.03 106 cells and 150 mg of nonlabeled proteins con-

tained in isolated chromosomes from 2.0 3 109 cells.

To measure the exchange ratio, we isolated mitotic chromosomes from

roughly 1.0 3 109 cells. Mitotic chromosomes were pelleted after centrifuging

at 3000 3 g and mixed into 10 ml cell extract that were made from 3.0 3 108

cells. This mixture was incubated at 14�C for 30 min. Finally, we re-isolated the

chromosomes as described above.

Mass Spectrometric Analysis

Proteins were separated into a high and a low molecular weight fraction by

SDS-PAGE, in-gel digested using trypsin (Shevchenko et al., 2006), and frac-

tionated into 30 fractions each using SCX. The individual SCX fractions were

desalted using StageTips (Rappsilber et al., 2003) and analyzed using

LC-MS on a LTQ-Orbitrap (Thermo Fisher Scientific) coupled to HPLC via

a nanoelectrospray ion source. The six most intense ions of a full MS acquired

in the orbitrap analyzer were fragmented and analyzed in the linear ion trap.

The MS data were analyzed using MaxQuant (Cox and Mann, 2008) and

proteins identified by searching MS and MS/MS data using the MASCOT

search engine (Matrix Science, UK). For more details, see the Extended Exper-

imental Procedures.

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures, six

figures, and five tables and can be found with this article online at doi:10.

1016/j.cell.2010.07.047.

ACKNOWLEDGMENTS

We dedicate this paper to Uli Laemmli on the occasion of his 70th birthday. We

thank Mayumi Takahashi for assistance with preparation of the Ska3/Rama1

knockout, Frauke Melchior for anti-RanBP2 and David Tollervey, Margarete

Heck, Robin Allshire, Iain Cheeseman, Kumiko Samejima, Susana Ribeiro

and Jan Bergmann for critical reading of the manuscript. This work was sup-

ported by a European Community Marie Curie Excellence Grant (JR), the

MRC (CPP), EMBO (LSP) and the Wellcome Trust (WCE, JR). WCE and

JR are Principal and Senior Research Fellows of The Wellcome Trust,

respectively.

Received: December 7, 2009

Revised: May 20, 2010

Accepted: July 14, 2010

Published: September 2, 2010

REFERENCES

Altaf, M., Auger, A., Monnet-Saksouk, J., Brodeur, J., Piquet, S., Cramet, M.,

Bouchard, N., Lacoste, N., Utley, R.T., Gaudreau, L., et al. (2010). NuA4-

dependent acetylation of nucleosomal histone H4 and H2A directly stimulates

incorporation of H2A.Z by the SWR1 complex. J. Biol. Chem., in press.

Amano, M., Suzuki, A., Hori, T., Backer, C., Okawa, K., Cheeseman, I.M., and

Fukagawa, T. (2009). The CENP-S complex is essential for the stable assembly

of outer kinetochore structure. J. Cell Biol. 186, 173–182.

Andersen, J.S., Wilkinson, C.J., Mayor, T., Mortensen, P., Nigg, E.A., and

Mann, M. (2003). Proteomic characterization of the human centrosome by

protein correlation profiling. Nature 426, 570–574.

Andrade, M.A., Ponting, C.P., Gibson, T.J., and Bork, P. (2000). Homology-

based method for identification of protein repeats using statistical significance

estimates. J. Mol. Biol. 298, 521–537.

Cheeseman, I.M., and Desai, A. (2008). Molecular architecture of the kineto-

chore-microtubule interface. Nat. Rev. Mol. Cell Biol. 9, 33–46.

Cox, J., and Mann, M. (2008). MaxQuant enables high peptide identification

rates, individualized p.p.b.-range mass accuracies and proteome-wide

protein quantification. Nat. Biotechnol. 26, 1367–1372.

Daum, J.R., Wren, J.D., Daniel, J.J., Sivakumar, S., McAvoy, J.N., Potapova,

T.A., and Gorbsky, G.J. (2009). Ska3 is required for spindle checkpoint

silencing and the maintenance of chromosome cohesion in mitosis. Curr.

Biol. 19, 1467–1472.

de Godoy, L.M., Olsen, J.V., Cox, J., Nielsen, M.L., Hubner, N.C., Frohlich, F.,

Walther, T.C., and Mann, M. (2008). Comprehensive mass-spectrometry-

based proteome quantification of haploid versus diploid yeast. Nature 455,

1251–1254.

Dejardin, J., and Kingston, R.E. (2009). Purification of proteins associated with

specific genomic Loci. Cell 136, 175–186.

Earnshaw, W.C., and Rothfield, N. (1985). Identification of a family of human

centromere proteins using autoimmune sera from patients with scleroderma.

Chromosoma (Berl) 91, 313–321.

Finn, R.D., Mistry, J., Tate, J., Coggill, P., Heger, A., Pollington, J.E., Gavin,

O.L., Gunasekaran, P., Ceric, G., Forslund, K., et al. (2010). The Pfam protein

families database. Nucleic Acids Res. 38, D211–D222.

Foltz, D.R., Jansen, L.E., Black, B.E., Bailey, A.O., Yates, J.R., and Cleveland,

D.W. (2006). The human CENP-A centromeric nucleosome-associated

complex. Nat. Cell Biol. 8, 458–469.

Foster, L.J., De Hoog, C.L., and Mann, M. (2003). Unbiased quantitative pro-

teomics of lipid rafts reveals high specificity for signaling factors. Proc. Natl.

Acad. Sci. USA 100, 5813–5818.

Gaitanos, T.N., Santamaria, A., Jeyaprakash, A.A., Wang, B., Conti, E., and

Nigg, E.A. (2009). Stable kinetochore-microtubule interactions depend on

the Ska complex and its new component Ska3/C13Orf3. EMBO J. 28, 1442–

1452.

Gassmann, R., Henzing, A.J., and Earnshaw, W.C. (2005). Novel components

of human mitotic chromosomes identified by proteomic analysis of the chro-

mosome scaffold fraction. Chromosoma 113, 385–397.

Gentleman, R.C., Carey, V.J., Bates, D.M., Bolstad, B., Dettling, M., Dudoit, S.,

Ellis, B., Gautier, L., Ge, Y., Gentry, J., et al. (2004). Bioconductor: open soft-

ware development for computational biology and bioinformatics. Genome

Biol. 5, R80.

Hori, T., Amano, M., Suzuki, A., Backer, C.B., Welburn, J.P., Dong, Y.,

McEwen, B.F., Shang, W.H., Suzuki, E., Okawa, K., et al. (2008). CCAN makes

multiple contacts with centromeric DNA to provide distinct pathways to the

outer kinetochore. Cell 135, 1039–1052.

Hubner, N.C., Bird, A.W., Cox, J., Splettstoesser, B., Bandilla, P., Poser, I.,

Hyman, A., and Mann, M. (2010). Quantitative proteomics combined with

BAC TransgeneOmics reveals in vivo protein interactions. J. Cell Biol. 189,

739–754.

Hudson, D.F., Vagnarelli, P., Gassmann, R., and Earnshaw, W.C. (2003).

Condensin is required for nonhistone protein assembly and structural integrity

of vertebrate mitotic chromosomes. Dev. Cell 5, 323–336.

Hutchins, J.R., Toyoda, Y., Hegemann, B., Poser, I., Heriche, J.K., Sykora,

M.M., Augsburg, M., Hudecz, O., Buschhorn, B.A., Bulkescher, J., et al.

(2010). Systematic analysis of human protein complexes identifies chromo-

some segregation proteins. Science 328, 593–599.

Ishihama, Y., Oda, Y., Tabata, T., Sato, T., Nagasu, T., Rappsilber, J., and

Mann, M. (2005). Exponentially modified protein abundance index (emPAI)

for estimation of absolute protein amount in proteomics by the number of

sequenced peptides per protein. Mol. Cell. Proteomics 4, 1265–1272.

820 Cell 142, 810–821, September 3, 2010 ª2010 Elsevier Inc.

Page 189: Cell 100903

Ishihama, Y., Schmidt, T., Rappsilber, J., Mann, M., Hartl, F.U., Kerner, M.J.,

and Frishman, D. (2008). Protein abundance profiling of the Escherichia coli

cytosol. BMC Genomics 9, 102.

Jiang, W., Middleton, K., Yoon, H.J., Fouquet, C., and Carbon, J. (1993). An

essential yeast protein, CBF5p, binds in vitro to centromeres and microtu-

bules. Mol. Cell. Biol. 13, 4884–4893.

Joseph, J., Liu, S.T., Jablonski, S.A., Yen, T.J., and Dasso, M. (2004). The Ran-

GAP1-RanBP2 complex is essential for microtubule-kinetochore interactions

in vivo. Curr. Biol. 14, 611–617.

Kops, G.J., van der Voet, M., Manak, M.S., van Osch, M.H., Naini, S.M., Brear,

A., McLeod, I.X., Hentschel, D.M., Yates, J.R., 3rd, van den Heuvel, S., and

Shah, J.V. (2010). APC16 is a conserved subunit of the anaphase-promoting

complex/cyclosome. J. Cell Sci. 123, 1623–1633.

Kulukian, A., Han, J.S., and Cleveland, D.W. (2009). Unattached kinetochores

catalyze production of an anaphase inhibitor that requires a Mad2 template to

prime Cdc20 for BubR1 binding. Dev. Cell 16, 105–117.

Letunic, I., Copley, R.R., Schmidt, S., Ciccarelli, F.D., Doerks, T., Schultz, J.,

Ponting, C.P., and Bork, P. (2004). SMART 4.0: towards genomic data integra-

tion. Nucleic Acids Res. 32, D142–D144.

Lewis, C.D., and Laemmli, U.K. (1982). Higher order metaphase chromosome

structure: evidence for metalloprotein interactions. Cell 29, 171–181.

Li, X., Wu, L., Corsa, C.A., Kunkel, S., and Dou, Y. (2009). Two mammalian

MOF complexes regulate transcription activation by distinct mechanisms.

Mol. Cell 36, 290–301.

Maiato, H., Sampaio, P., Lemos, C.L., Findlay, J., Carmena, M., Earnshaw,

W.C., and Sunkel, C.E. (2002). MAST/Orbit has a role in microtubule-kineto-

chore attachment and is essential for chromosome alignment and mainte-

nance of spindle bipolarity. J. Cell Biol. 157, 749–760.

Marın, I. (2003). Evolution of chromatin-remodeling complexes: comparative

genomics reveals the ancient origin of ‘‘novel’’ compensasome genes.

J. Mol. Evol. 56, 527–539.

Morrison, C., Henzing, A.J., Jensen, O.N., Osheroff, N., Dodson, H., Kandels-

Lewis, S.E., Adams, R.R., and Earnshaw, W.C. (2002). Proteomic analysis of

human metaphase chromosomes reveals topoisomerase II alpha as an Aurora

B substrate. Nucleic Acids Res. 30, 5318–5327.

Neumann, B., Walter, T., Heriche, J.K., Bulkescher, J., Erfle, H., Conrad, C.,

Rogers, P., Poser, I., Held, M., Liebel, U., et al. (2010). Phenotypic profiling

of the human genome by time-lapse microscopy reveals cell division genes.

Nature 464, 721–727.

Obuse, C., Iwasaki, O., Kiyomitsu, T., Goshima, G., Toyoda, Y., and Yanagida,

M. (2004). A conserved Mis12 centromere complex is linked to heterochro-

matic HP1 and outer kinetochore protein Zwint-1. Nat. Cell Biol. 6, 1135–1141.

Okada, M., Cheeseman, I.M., Hori, T., Okawa, K., McLeod, I.X., Yates, J.R.,

Desai, A., and Fukagawa, T. (2006). The CENP-H-I complex is required for

the efficient incorporation of newly synthesized CENP-A into centromeres.

Nat. Cell Biol. 8, 446–457.

Ong, S.E., Blagoev, B., Kratchmarova, I., Kristensen, D.B., Steen, H., Pandey,

A., and Mann, M. (2002). Stable isotope labeling by amino acids in cell culture,

SILAC, as a simple and accurate approach to expression proteomics. Mol.

Cell. Proteomics 1, 376–386.

Porter, I.M., McClelland, S.E., Khoudoli, G.A., Hunter, C.J., Andersen, J.S.,

McAinsh, A.D., Blow, J.J., and Swedlow, J.R. (2007). Bod1, a novel kineto-

chore protein required for chromosome biorientation. J. Cell Biol. 179,

187–197.

Raaijmakers, J.A., Tanenbaum, M.E., Maia, A.F., and Medema, R.H. (2009).

RAMA1 is a novel kinetochore protein involved in kinetochore-microtubule

attachment. J. Cell Sci. 122, 2436–2445.

Rappsilber, J., Ishihama, Y., and Mann, M. (2003). Stop and go extraction tips

for matrix-assisted laser desorption/ionization, nanoelectrospray, and LC/MS

sample pretreatment in proteomics. Anal. Chem. 75, 663–670.

Rappsilber, J., Ryder, U., Lamond, A.I., and Mann, M. (2002). Large-scale pro-

teomic analysis of the human spliceosome. Genome Res. 12, 1231–1245.

Schirmer, E.C., Florens, L., Guan, T., Yates, J.R., 3rd, and Gerace, L. (2003).

Nuclear membrane proteins with potential disease links found by subtractive

proteomics. Science 301, 1380–1382.

Shevchenko, A., Tomas, H., Havlis, J., Olsen, J.V., and Mann, M. (2006). In-gel

digestion for mass spectrometric characterization of proteins and proteomes.

Nat. Protoc. 1, 2856–2860.

Takata, H., Uchiyama, S., Nakamura, N., Nakashima, S., Kobayashi, S., Sone,

T., Kimura, S., Lahmers, S., Granzier, H., Labeit, S., et al. (2007). A comparative

proteome analysis of human metaphase chromosomes isolated from two

different cell lines reveals a set of conserved chromosome-associated

proteins. Genes Cells 12, 269–284.

Theis, M., Slabicki, M., Junqueira, M., Paszkowski-Rogacz, M., Sontheimer,

J., Kittler, R., Heninger, A.K., Glatter, T., Kruusmaa, K., Poser, I., et al.

(2009). Comparative profiling identifies C13orf3 as a component of the Ska

complex required for mammalian cell division. EMBO J. 28, 1453–1465.

Tkaczuk, K.L., Dunin-Horkawicz, S., Purta, E., and Bujnicki, J.M. (2007). Struc-

tural and evolutionary bioinformatics of the SPOUT superfamily of methyltrans-

ferases. BMC Bioinformatics 8, 73.

Uchiyama, S., Kobayashi, S., Takata, H., Ishihara, T., Hori, N., Higashi, T.,

Hayashihara, K., Sone, T., Higo, D., Nirasawa, T., et al. (2005). Proteome anal-

ysis of human metaphase chromosomes. J. Biol. Chem. 280, 16994–17004.

Van Hooser, A.A., Yuh, P., and Heald, R. (2005). The perichromosomal layer.

Chromosoma 114, 377–388.

Welburn, J.P., Grishchuk, E.L., Backer, C.B., Wilson-Kubalek, E.M., Yates,

J.R., 3rd, and Cheeseman, I.M. (2009). The human kinetochore Ska1 complex

facilitates microtubule depolymerization-coupled motility. Dev. Cell 16,

374–385.

Wen, H., Li, J., Song, T., Lu, M., Kan, P.Y., Lee, M.G., Sha, B., and Shi, X.

(2010). Recognition of histone H3K4 trimethylation by the plant homeodomain

of PHF2 modulates histone demethylation. J. Biol. Chem. 285, 9322–9326.

Wu, W.H., Alami, S., Luk, E., Wu, C.H., Sen, S., Mizuguchi, G., Wei, D., and Wu,

C. (2005). Swc2 is a widely conserved H2AZ-binding module essential for

ATP-dependent histone exchange. Nat. Struct. Mol. Biol. 12, 1064–1071.

Cell 142, 810–821, September 3, 2010 ª2010 Elsevier Inc. 821

Page 190: Cell 100903
Page 191: Cell 100903

Cell is seeking a scientist to join its editorial team. The minimum qualification for this position is a PhD in a relevant area of biomedical research, although additional post-doctoral 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 disseminate their findings to the wider community.

As an editor, you would be responsible for assessing submitted research papers and overseeing the refereeing process, and you would commission and edit material for Cell’s Leading Edge. You would also travel frequently to scientific conferences to follow developments in research and to establish and maintain close ties with the scientific community. You would have opportunities to pioneer and contribute to new trends in scientific publishing. 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 be highly motivated and creative and able to work inde-pendently as well as in a team.

This is a full-time, in-house editorial position, based at the Cell Press office in Kendall Square near MIT in Cambridge, Massachusetts. Cell Press offers an attractive salary and benefits pack-age and a stimulating working environment that encourages innovation.

To apply, please send your CV and a letter describing your qualifications, general research inter-ests, and motivation for pursuing a career in scientific publishing as soon as possible to our online jobs site: http://reedelsevier.taleo.net/careersection/51/jobdetail.ftl?lang=en&job=CEL0000R.

No phone inquiries, please. Cell Press in an equal opportunity/affirmative action employer, m/f/d/v.

Scientific Editor Position Available

editorad.indd 1 3/24/2010 2:17:01 PM

Page 192: Cell 100903

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 193: Cell 100903

Cell Host & Microbe, the premier microbiology primary research journal, is seeking a full-time sci-entific editor to join its editorial team. 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 developments in research and establish and maintain close ties with the scientific community. This is a superb opportunity for a talented individual to play a critical role in the research community away from the bench.

This is a full-time in-house editorial position, based at the Cell Press office in Cambridge, Mas-

sachusetts. Cell Press offers an attractive salary and benefits package and a stimulating working environment that encourages innovation.

The minimum qualification for this position is a PhD in a relevant area of biomedical research, although additional postdoctoral or editorial experience is preferred. We will consider qualified can-didates with scientific expertise in any area that falls within the scope of the journal, including molec-ular and cell biology of microbes, host immune response to microbes, immune evasion by microbes, microbiology, therapeutics, and vaccine design. The key qualities we look for are breadth of scien-tific 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.

Applications will be accepted through August 31, 2010. To apply, please visit the Careers page at

http://www.elsevier.com and search on keywords “Cell Host & Microbe.” Please submit a resume and cover letter describing your qualifications, general research interests, and motivation for pursu-ing a career in scientific publishing. Applications will be held in the strictest confidence.

Cell Host & Microbe Scientific Editor Position Available

chm.indd 1 8/4/2010 2:58:37 PM

Page 194: Cell 100903

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 195: Cell 100903

Positions Available

cell1425cla.indd 1cell1425cla.indd 1 8/27/2010 9:07:34 PM8/27/2010 9:07:34 PM

Page 196: Cell 100903

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.

ASSISTANT PROFESSORDEPARTMENT OF GENETICS

HARVARD MEDICAL SCHOOL

The Department of Genetics at Harvard Medical School invites applicants for a tenure-track faculty position at the rank of Assistant Professor. The Department of Genetics consists of faculty working on diverse problems using a variety of approaches and model organisms, unified in their focus on the genome as an organizing principle for understanding biological phenomena. We are seeking outstanding applicants, with a demonstrated potential for imaginative research and a clear vision for the future, who are working on exciting problems in any area of genet-ics, broadly defined. The successful candidate is expected to direct innovative and independent research and participate in the teaching of graduate and medical students. Significant scholarly and scientific resources are readily available with this position. Our highly interactive Department provides the opportunity to interact and collaborate with other dedicated researchers within the diverse Harvard research community. For further informa-tion about our Department, please see our Web Page: http://genetics.med.harvard.edu.

Applicants should submit electronic copies of curriculum vitae, bibliography, a brief description of research accomplishments and future research interests (limit to 500 words) by October 31, 2010, and ask three refer-ences to provide letters of recommendation. These materials should be sent to the following email address: [email protected].

We strongly encourage applications from women and minority candidates.

Harvard University is an Equal Opportunity/Affirmative Action Employer.

cell1425cla.indd 2cell1425cla.indd 2 8/27/2010 9:07:37 PM8/27/2010 9:07:37 PM

Page 197: Cell 100903

Positions Available

THE UNIVERSITY OF TEXASSOUTHWESTERN MEDICAL CENTER

AT DALLAS

The Department of Molecular Biology at the University of Texas Southwestern Medical Center invites applications for Assistant Professor tenure-track faculty positions. We are seeking individuals with innovative research programs in the areas of stem cell biology, cell signaling, development and gene regulation. Successful applicants should be capa-ble of establishing a vigorous independent research program and teaching in one of several active graduate programs. Attractive start-up packages through the Endowed Scholars Program, state-of-the-art core facilities, and exceptional laboratory space are available.

Applicants should submit a curriculum vitae containing a summary of past research accomplishments, a statement of future objectives, and names of three references to:

Eric N. OlsonChairman

Department of Molecular Biology University of Texas

Southwestern Medical Center at Dallas5323 Harry Hines Boulevard Dallas, Texas 75390-9148

UTSW is an Equal Opportunity Employer

cell1425cla.indd 3cell1425cla.indd 3 8/27/2010 9:07:39 PM8/27/2010 9:07:39 PM

Page 198: Cell 100903

Positions Available

Department of Pharmacology(with potential joint appointment in the Institute for Translational Neuroscience)

University of Minnesota Medical SchoolTENURE/TENURE TRACK POSITION

(Assistant Professor, Associate Professor, Professor)

The Department of Pharmacology at the University of Minnesota invites applications for a tenure/tenure track faculty position (Assistant, Associate or Full Professor). Applicants using molecular, biochemical, cellular, and/or integrative translational approaches to study problems relevant to pharmacological sciences are encouraged to apply. Positions are also available in this joint recruitment with the Institute for Translational Neuroscience, whose interest is in how basic science discoveries may lead to new therapeutic principles of certain neurological disorders (see their website for more details). Qualifications include a Ph.D. in biomedical science, or an M.D. degree, and relevant postdoctoral research experience. Applicants must have a strong record of research accomplishments, as docu-mented by publications in leading peer-reviewed journals. The successful Assistant Professor candidate will be expected to develop an innovative, competitive research program supported by extramural funding and to participate in departmental teaching activities. Applicants for Associate Professor and Professor positions must demonstrate distinction in published research, evidence of consistent extramural funding, and a commitment to teaching. For additional information about the department and the Institute for Translational Neuroscience, visit: www.pharmacology.med.umn.edu and http://www.itn.umn.edu.

Interested applicants should apply online at http://employment.umn.edu/applicants/Central?quickFind=88959 for the Assistant Professor position or at http://employment.umn.edu/applicants/Central?quickFind=88961 for the Assoc. Prof. or Prof. position, and attach a letter of interest, curriculum vitae, brief statement of research interests, and contact information for three references.

Applications are invited for a tenure-track junior faculty appointment in the Department of Structural Biology, Stanford University School of Medicine. Candidates should have expertise and a commitment to future research in the broad area of structural biology and biophysics. The predominant criterion for appointment in the University Tenure Line is a major commitment to research and teaching.

To be considered for this position, interested individuals should:

1. Fully complete the Structural Biology Faculty Recruitment Form available at: https://med.stanford.edu/survey/sbio_fclty_recruit.2. Prepare the application materials as one .pdf fi le and e-mail the fi le to [email protected]. PDF fi le must contain

all of the following documents:

i) Cover letter ii) Curriculum vitaeiii) Description of research interests (4 page limit)iv) List of publications v) Maximum of 5 representative reprintsvi) Names and contact information of three referees

3. Arrange for three reference letters to be e-mailed by each referee directly to: [email protected]

or send by mail to:

Chair, Faculty Search Committee, Department of Structural Biology, Stanford University School of Medicine, 299 Campus Drive West, D100 Fairchild Bldg., Stanford, CA 94305-5126

Candidates must have a PhD and/ or MD degree and a minimum of two years of postdoctoral research experience. Application material and reference letters must be received by November 1, 2010.

Stanford University is an equal opportunity employer and is committed to increasing the diversity of its faculty. It welcomes nominations of and applications from women and members of minority groups, as well as others who would bring

additional dimensions to the university's research, teaching and clinical missions.

DEPARTMENT OF STRUCTURAL BIOLOGY

ASSISTANT PROFESSOR OF STRUCTURAL BIOLOGY

cell1425cla.indd 4cell1425cla.indd 4 8/27/2010 9:07:42 PM8/27/2010 9:07:42 PM

Page 199: Cell 100903

Positions Available

cell1425cla.indd 5cell1425cla.indd 5 8/27/2010 9:07:42 PM8/27/2010 9:07:42 PM

Page 200: Cell 100903

Positions Available

FACULTY POSITION IN MOLECULAR, CELLULAR, AND DEVELOPMENTAL BIOLOGY

University of Colorado at Boulder

The Department of MCD Biology invites applications for a tenure-track Assistant Professor in the area of molecular, cel-lular, or developmental biology with an emphasis on basic molecular biological problems. Applicants must have a Ph.D., M.D., or equivalent; and postdoctoral research experience. The candidate is expected to develop a vigorous and innova-tive research program, and have enthusiasm for teaching at the undergraduate and graduate levels.

Review of applications will begin on November 1, 2010 and continue until the position is filled. Application materials are accepted electronically at https://www.jobsatcu.com, posting number 810875. Applicants should submit a curriculum vitae and a concise statement of research and teaching interests, and arrange to have three reference letters sent.

For questions or concerns on submitting your materials elec-tronically, please contact [email protected].

The University of Colorado at Boulder is committed to diver-sity and equality in education and employment.

See www.colorado.edu/ArtsSciences/Jobs/ for full job description.

POSTDOCTORAL POSITIONS AVAILABLE AT SLOAN-KETTERING

INSTITUTE

Seeking research scientists with postdoctoral experi-ence to join a laboratory in order to study the molecular pathogenesis of multiple myeloma and utilize mouse models to develop new therapeutic approaches. Applicants must have a thorough knowledge of plasma cell biology and expertise using genetically modified mice to study cancer pathogenesis. Send a description of your research interests, curriculum vitae, and the names and phone numbers of three references to

Dr. Stephen D. NimerMemorial Sloan-Kettering Cancer Center

1275 York Avenue, Box 575 New York, NY 10065

The Memorial Sloan-Kettering Cancer Center is an Affirmative Action/Equal Opportunity Employer.

cell1425cla.indd 6cell1425cla.indd 6 8/27/2010 9:07:45 PM8/27/2010 9:07:45 PM

Page 201: Cell 100903

Positions Available

THE UNIVERSITY OF CALIFORNIA AT BERKELEY

Department of Molecular and Cell Biology, Helen Wills Neuroscience

Institute and Center for Computational Biology

The Department of Molecular and Cell Biology (MCB), the Helen Wills Neuroscience Institute (HWNI) and the Center for Computational Biology (CCB) are seeking applications for four faculty positions in the areas listed below. We seek candidates with Ph. D. and/or M.D. degrees who have a strong inter-est in undergraduate and graduate teaching and demonstrated excellence, originality and productivity in research.

Biomedical Sciences (MCB)We seek candidates interested in molecular and cellular mechanisms of disease processes. Areas of research could include cancer, human genetics, aging, molecular physiology, animal models of human disease, infectious and neurodegenerative diseases, or any other area of biomedical science. This position is open at the Assistant Professor level (tenure-track).

Molecular, Cellular or Developmental Neuroscience (MCB / HWNI)We seek candidates in any area of molecular, cellular, and developmental neurobiology, with a particular emphasis on molecular genetic approaches to understanding neural circuit function, development, or regeneration. Approaches could include (but are not limited to) human genetics, use of genetic model organisms, genomics, behavioral analysis, optical imaging, and/or computational modeling. This posi-tion is open at the Assistant Professor level (tenure-track).

Human Genome Variation (MCB / CCB)We seek candidates whose research focuses on an understanding of genome and epigenome variation of humans as it relates to phenotype, particularly disease predisposition, using both computational and laboratory approaches. This position is open at the Assistant Professor level (tenure-track).

Stem Cell Biology (MCB)We seek candidates whose research focuses on any aspect of stem cell biology, including (but not limited to) the use of stem cells to develop models of human biology or disease, molecular mecha-nisms of transcriptional regulation in pluripotent and differentiating stem cells, the regulation of stem cell renewal and differentiation during development, and the biology of cancer stem cells. This posi-tion is open at any level (tenured or tenure-track).

Applications and letters of reference should be submitted online through http://mcb.berkeley.edu. Applications should include a curriculum vitae; a list of publications; copies of three significant publications; a brief description of research accomplishments; and a statement of research objectives and teaching interests. In addition, junior applicants applying for a non-tenured position should arrange to have three letters of reference submitted online. Potential reviewers should be referred to the Statement of Confidentiality found at: http://apo.chance.berkeley.edu/evalltr.html. The deadline for applications is November 15, 2010.

We are interested in candidates who will contribute to diversity and equal opportunity in higher education through their teaching, research, and service. We are also committed to addressing the family needs of faculty. The University of California

is an Affirmative Action/Equal Opportunity Employer.

cell1425cla.indd 7cell1425cla.indd 7 8/27/2010 9:07:48 PM8/27/2010 9:07:48 PM

Page 202: Cell 100903

Positions Available

The Cell Biology Department seeks exceptional candidates at the rank of Assistant and Associate Professor who wish to engage in independent research programs in the field. Successful candidates will have PhD, and/or M.D. degrees and a demonstrated record of originality and productivity in research during gradu-ate training and/or post-doctoral research, as well as well-formulated plans for independent research. Although the focus is on tenure-track junior faculty appointments, applicants for senior faculty ranks may also be considered. The Department is undergoing a major expansion under the Chair, Dr. James E. Rothman. Exceptional candidates in all areas of cell biology will be considered.

Please email your CV with a list of publications, a summary of doctoral research [1 page], a summary of post-doctoral research [max. 2 pages], and a research plan [max. 3 pages], along with the names and addresses (including email) of three potential references by October 25, 2010 to: [email protected]

Applications from, or nominations of, women and minority scientists are encouraged. Yale is an

Affirmative Action/Equal Opportunity Employer.

years of leadership in human genetics research,

education and service.

1948–2008www.ashg.org

60

cell1425cla.indd 8cell1425cla.indd 8 8/27/2010 9:07:49 PM8/27/2010 9:07:49 PM

Page 203: Cell 100903

EuPA now has its own journal!

To receive more information register at:http://www.elsevier.com/locate/jprot

http://www.eupa.org/

Covered byPubMed

Editor in Chief:Juan J. Calvete, Valencia, Spain

Executive Editors:Proteomics in Cell BiologyJean-Jacques Diaz, Lyon, France

Proteomics in MicrobiologyConcha Gil, Madrid, Spain

Proteomics in Plant SystemsJesus V. Jorrín, Córdoba, Spain

Proteomics in Animal ModelsDario Neri, Zürich, Switzerland

Proteomics in Protein ScienceJasna Peter-Katalinic, Münster,Germany

Biomedical Applications ofProteomics and CongressProceedingsJean-Charles Sanchez, Geneva,Switzerland

Proteomics of Body Fluids andProteomic TechnologiesPier Giorgio Righetti, Milan, Italy

Bioinformatics in ProteomicsPeter Højrup, Odense, Denmark

For a complete listing of theeditorial board, visit thejournal’s homepage

Submitting AuthorsManuscripts can be submitted to the

Journal of Proteomics athttp://ees.elsevier.com/jprot

Page 204: Cell 100903

For the past 40 years FEBS Letters has

published award-winning research papers crafted

by distinguished experts in the field.

www.febsletters.com

Page 205: Cell 100903

John H. Byrne and James L. Roberts, who treated us to a superb textbook in their first edition of From Molecules to Networks, have now returned with another outstanding effort in the second edition.

— Eric R. Kandel, Howard Hughes Medical Institute; Columbia

University College of Physicians and Surgeons, New York, USA

Just Published!

From Molecules to Networks An Introduction to Cellular and Molecular

Neuroscience, Second Edition

Editors-in-Chief:

John H. Byrne, W.M. Keck Center for the

Neurobiology of Learning and Memory,

University of Texas Medical School at Houston;

James L. Roberts, Trinity University,

San Antonio, Texas

Now includes an introduction to

mathematical modeling and simulation

approaches and a chapter on the

Neurobiology of Disease

Full color, professional graphics throughout

Two expansive, frequently updated

support websites:

For students—study materials, updated -

further reading references, and quizzes

For instructors—all figures, solutions, -

PowerPoint lecture slides, and more

Feb 2009, hardback 656 pp.ISBN: 9780123741325$94.95 USD / €69.95 EUR / £47.99 GBP

Register at our instructor site: http://textbooks.elsevier.comThis site allows registered/approved instructors to browse our list of cutting-edge books at

student-friendly prices, request print review copies, access ancillaries, view selected books

electronically, and contact your textbook representative.

Secure online ordering at elsevierdirect.com/9780123741325

Page 206: Cell 100903

years of leadership in human genetics research,

education and service.

1948–2008www.ashg.org

60

Page 207: Cell 100903

FUNDAMENTAL NEUROSCIENCETHIRD EDITION

THE NEW EDIT ION OF FUNDAMENTAL NEUROSCIENCE I S NOW AVA ILABLE !

February 2008

Hardback

1,280 pp.

9780123740199

Print list price: $99.95

EDITED BY: Larry R. Squire, VA Medical Center and University of California, San Diego, La Jolla, California, USA

SENIOR EDITORS: Darwin Berg, Floyd E. Bloom, Sascha du Lac,Anirvan Ghosh, and Nicholas C. Spitzer

This comprehensive text reference presents the entire spectrum of modern neuroscience.

Addressing feedback on the previous edition, Fundamental Neuroscience, Third Edition is

now more concise and reader-friendly than ever before. Each chapter is heavily illustrated

and provides clinical boxes describing experiments, disorders, and methodological

approaches and concepts. A companion site contains an image bank of the figures for

use in poster presentations, slides, and handouts.

Capturing the promise and excitement of this fast-moving field, Fundamental Neuroscience,Third Edition is the book that you will reference throughout your neuroscience career!

NEW TO THIS EDITION:

• 30% new material including new chapters on Dendritic Development and Spine Morphogenesis; Chemical Senses; Cerebellum; Eye Movements; Circadian Timing; Sleep and Dreaming; and Consciousness

• Companion Web site with all figures and Web links to additional material

• Multiple model system coverage beyond rats, mice, and monkeys

• Extensively expanded index for easier referencing

WHY BUY THIS EDITION?

• It presents the entire spectrum of modern neuroscience as a cohesive reference rather than just a collection of review articles

• It is written at a level appropriate for graduate-level students and researchers with various science backgrounds who need a solid foundation in neuroscience principles

• Clinical conditions are discussed in neuroscience to help readers establish the connection between the basic biology and the neurological symptoms and syndromes

• It includes more than 650 illustrations, including color photographs, imaging, and micrographs, also available on the companion Web site

SECURE ONLINE ORDERING AT ELSEVIERDIRECT.COM

Page 208: Cell 100903

careers.cell.com

Find Your Ideal Job!

Page 209: Cell 100903
Page 210: Cell 100903

See online version for legend and references.822 Cell 142, September 3, 2010 ©2010 Elsevier Inc. DOI 10.1016/j.cell.2010.08.026

SnapShot: Nuclear Receptors INeil J. McKenna and Bert W. O’MalleyBaylor College of Medicine, Houston, TX 77030, USA

Receptor/Family* Symbols Ligands Major Functions

Disease Associations Target Genes

Estrogen receptors*

ERα/NR3A1; ERβ/NR3A2

Endogenous: 17β-estradiol Regulation of cell growth and proliferation in multiple tissues (e.g., female reproductive tissues, bone, and CNS)

Cancer, cardiovascular, immune and inflammatory, metabolic, neurological, reproductive

↑ MYC, NGF, BCL2, CXCL2, IGF1, TYMS; ↓ CD36, NDRG1, NCOR1, NCOA3Clinical: Mixed agonists (e.g. tamoxifen,

raloxifene, and toremifene in breast cancer)

Xenobiotics: Bisphenol A, PCBs

Androgen receptor

AR/NR3C4 Endogenous: Testosterone, dihydrotestosterone

Key role in male reproductive organs in addition to other systems (e.g., CNS)

Cancer, cardiovascular, immune, metabolic, neurological, reproductive

↑ MYC, VEGF, BCL2, IGF1, MUC1, P66(Shc), CCND1; ↓ TSHA, TSHB, PTEN, FAS, CASP2, CTNND2, ESR2, TMPRSS2

Clinical: Flutamide and bicalutamide for prostate cancer and alopecia

Glucocorticoid receptor

GR/NR3C1 Endogenous: Cortisol (hydrocortisone) Diverse developmental and physiological roles (e.g., antagonism of inflammatory signaling pathways, mediation of the stress response, and gluconeogenesis)

Metabolic, cardiovascular, immune and inflammatory, memory

↑ SCNN1A, GADD45B, GILZ, TAT; ↓ BGLAP, POMC, INS

Clinical: Fluticasone and prednisolonein inflammatory disorders

Vitamin D receptor

VDR/NR1I1 Endogenous: Calcitriol (1′,25′ dihdroxy vitamin D3)

Maintenance of serum calcium and phosphate levels for skeletal integrity; antiproliferative in many tissues

Bone, cancer, cardiovascular, metabolic, immune and inflammatory, renal, neurological

↑ FGF23, CYP24A1, CALB1, BGLAP, SPP1; ↓ IL2, PHEX

Clinical: Paracalcitol for 2o hyperparathyroidism in renal patients; Tacalcitol for psoriasis

Thyroid hormone receptors*

TRα/NR1A1; TRβ/NR1A2

Endogenous: Thyroxine (T4),Triiodothyronine (T3)

Regulation of oxygen consumption; protein, carbohydrate, lipid, and vitamin metabolism

Thyroid conditions, cancer

↑ ADRB1, PCK2, GH1, UCP1; ↓ DIO2, EHHADH, PRL, EGFR

Clinical: Levo-thyroxine, triiodothyroacetic acid (TRIAC) in resistance to thyroid hormone

Progesterone receptor

PR/NR3C3 Endogenous: Progesterone Diverse reproductive functions (e.g., establishing and maintaining pregnancy, developing breast tissue, and stopping proliferation in the uterus)

Cancer, metabolic, reproductive

↑ SERPINB14, FAS, MT2A, PGC, EGFR , IHH; ↓ ESR1, PGR, ANXA6Clinical: RU486 (Mifepristone) as an

abortifacient

Mineralocorticoid receptor

MR/NR3C2 Endogenous: Aldosterone Regulating electrolyte and fluid balance in the kidney; specific roles in the CNS

Metabolic ↑ SCNN1A, ATP1A1, ATP1B1, GILZ, SGK1, NDRG2

Clinical: Spironolactone in hypertensive cardiovascular disease

Peroxisome-proliferator-activated receptor-γ

PPARγ/NR1C3 Endogenous: FAs and FA intermediates Regulation of adipocytes, insulin sensitivity and lipogenesis, and broader integration of energy, lipid, and carbohydrate metabolism

Cardiovascular, metabolic, cancer, neurological

↑ FABP4, UCP1, AP2, PCK1, LPL, ADIPOQ, CD36, AQP7

Dietary: FAs and PUFAs

Clinical: Thiazolidinediones (e.g., rosiglitazone) in type II diabetes

Peroxisome-proliferator-activated receptor-α

PPARα/NR1C1

Endogenous: FAs and FA intermediates Regulating energy expenditure; modulating fatty acid oxidation systems (mitochondria), peroxisome β-oxidation, and microsomal ω-oxidation

Cardiovascular, metabolic, cancer, neurological

↑ ACBP, ACOX1, APOA1, CPT1A, CYP1A1, CYP4A1, CYP7A1, SLC27A1, LCAS, MLYCD, SCD, FADS2, RETN, MYC, CCND1, IGFBP1, UCP1, KRT23, IL6, TF, PEX11A

Clinical: Fibrates (e.g., fenofibrate) in hyperlipidemia

Dietary: FAs and PUFAs

Xenobiotics: DEHP, DEHA

Peroxisome-proliferator-activated receptor-δ (β)

PPARδ/NR1C2

Endogenous: FAs and FA intermediates Regulating cell proliferation, differentiation, and migration in wound healing and inflammatory processes

Cardiovascular, metabolic, cancer, neurological

↑ ACSL3, CPT1A, RGS3, RGS4, RGS5, ISG20, CXCL7, CCL21, RETN, CPT1ADietary: FAs and PUFAs

Retinoic acid receptors*

RARα/NR1B1; RARβ/NR1B2; RARγ/NR1B3

Endogenous: All-trans and 9-cis retinoic acid

Pleiotropic control of embryonic patterning and organogenesis, cell proliferation, differentiation, apoptosis and homeostatic control

Neurological and psychiatric, cancer

↑ Numerous HOX genes, STRA6, HNF3A, CRABP2, ACADM, MECOM; ↓ CYP1A1, HOXB9

Clinical: Tretinoin for treating acne and acute promyelocytic leukemia

Liver X receptors* LXRα/NR1H3; LXRβ/NR1H2

Endogenous: Oxysterols Cholesterol and steroid sensors with roles in lipid and carbohydrate metabolism

Metabolic ↑ SREBP1C, CYP7A1, ABC8, APOA1, APOE, LPL, PLTP

Retinoid X receptors*

RXRα/NR2B1; RXRβ/NR2B2; RXRγ/NR2B3

Endogenous: 9-cis retinoic acid Embryonic cell patterning and organogenesis, cell proliferation and differentiation, other functions as heterodimers with other nuclear receptors

Neurological and psychiatric, immune

↑↓ many genes as heterodimers with other receptors (e.g.,LXRs, PPARs, FXR, TRs, RARs; ↑ ABC1 (with LXR); ↓ CYP7A1 (with FXR)

Pregnane X receptor

PXR/NR1I2 Endogenous: Bile acids Metabolism and transport of pharmaceutical drugs, xenobiotics, and toxic bile acids in the liver and GI tract

Immune ↑ Multiple CYP2 and CYP3 gene family members, MDR1, MRP2, OATP2, UGT1A1, SULT, ↓ CYP7A1

Xenobiotic: St. John’s Wort (hyperforin), Taxol, rifampicin, phenobarbital

Dietary: Vitamin E, sulforaphane, Gugulipid

Constitutive androstane receptor

CAR/NR1I3 Endogenous: Androstanol, androstenol Metabolism of xenobiotics and endogenous lipids by regulating expression of cytochrome P450 genes

Involved in hepatotoxicity of acetaminophen

↑ CYP2B10, CYP311A, CYP3A4, CYP1A2, CYP2B6 THRSP, SLC21A6, MRP2, MDR1, OATP2Xenobiotics: Phenobarbital, DEHP,

Meclizine

Farnesoid X receptor

FXR/NR1H4 Endogenous: Bile acids (e.g., chenodeoxycholic acid)

A sensor for bile acid that helps regulate bile acid homeostasis

Metabolic ↑ SLC10A2, ABCB1, ABCB11, NR0B2, HSD3B2, FETUB, ABCB4, FGF19, NOS2; ↓ CYP7A1, HNF1A, HNF4A, SLCO1B1, SLC10A2

Dietary: Cafestol, guggulsterone

I[1].indd 1 8/27/10 10:24:39 AM

Page 211: Cell 100903
Page 212: Cell 100903

B: 218 mm

CLIENT: Santa Cruz Biotechnology, Inc. PROJECT: 2010 ChemCruz Biochemicals Ad VERSION: Size “B” DATE: 04.08.10

PUBLICATION: CELL (MASTER/DIGITAL DOWNLOAD w/MATCH PROOF)

BLEED: 218mm W x 282mm H TRIM: 212mm W x 276mm H LIVE: 200mm W x 264mm H

SAN TA CRUZ B I OT ECH N O LO GY, I N C . SANTA CRUZ

B I O T E C H N O L O G Y

The Power to Question

®

Leading (#1 worldwide) supplier of specialty biochemicals for proteomics research.

ChemCruz™

HEADQUARTERS

TOLL FREE: 800.457.3801 TEL: 831.457.3800FAX: 831.457.3801 E-MAIL: [email protected]

EUROPEAN SUPPORT

TOLL FREE: +00800.4573.8000 PHONE: +49.6221.4503.0 FAX: +49.6221.4503.45 E-MAIL: [email protected]

ASIAN SUPPORT

JAPAN TOLL FREE: (010) 800.40402026 S. KOREA TOLL FREE: 00798.1.1.002.0297 N. CHINA TOLL FREE: 10.800.711.0752 S. CHINA TOLL FREE: 10.800.110.0694 FAX: 831.457.3801 E-MAIL: [email protected]

Products include over 71,300 unique biochemicals sorted alphabetically, by CAS number, by formula and on the basis of functional activity.

See our web site for product details and ordering information.

www.scbt.com

150 linescreenmp


Recommended