18.04.2015 Views

Pharmacogenetics of Breast Cancer - JOHN J. HADDAD, Ph.D.

Pharmacogenetics of Breast Cancer - JOHN J. HADDAD, Ph.D.

Pharmacogenetics of Breast Cancer - JOHN J. HADDAD, Ph.D.

SHOW MORE
SHOW LESS

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

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

Translational Medicine Series 7<br />

<strong><strong>Ph</strong>armacogenetics</strong><br />

<strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong><br />

Towards the<br />

Individualization <strong>of</strong> Therapy<br />

Edited by<br />

Brian Leyland-Jones


TRANSLATIONAL MEDICINE SERIES<br />

1. Prostate <strong>Cancer</strong>: Translational and Emerging Therapies, edited by<br />

Nancy A. Dawson and W. Kevin Kelly<br />

2. <strong>Breast</strong> <strong>Cancer</strong>: Translational Therapeutic Strategies, edited by Gary<br />

Lyman and Harold Burstein<br />

3. Lung <strong>Cancer</strong>: Translational and Emerging Therapies, edited by<br />

Kishan J. Pandya, Julie R. Brahmer, and Manuel Hidalgo<br />

4. Multiple Myeloma: Translational and Emerging Therapies, edited by<br />

Kenneth C. Anderson and Irene Ghobrial<br />

5. <strong>Cancer</strong> Supportive Care: Advances in Therapeutic Strategies,<br />

edited by Gary H. Lyman and Jeffrey Crawford<br />

6. <strong>Cancer</strong> Vaccines: Challenges and Opportunities in Translation,<br />

edited by Adrian Bot and Mihail Obrocea<br />

7. <strong><strong>Ph</strong>armacogenetics</strong> <strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong>: Towards the Individualization<br />

<strong>of</strong> Therapy, edited by Brian Leyland-Jones


Informa Healthcare USA, Inc.<br />

52 Vanderbilt Avenue<br />

New York, NY 10017<br />

# 2008 by Informa Healthcare USA, Inc.<br />

Informa Healthcare is an Informa business<br />

No claim to original U.S. Government works<br />

Printed in the United States <strong>of</strong> America on acid-free paper<br />

10987654321<br />

International Standard Book Number-10: 1-4200-5293-4 (Hardcover)<br />

International Standard Book Number-13: 978-1-4200-5293-0 (Hardcover)<br />

This book contains information obtained from authentic and highly regarded sources. Reprinted material is<br />

quoted with permission, and sources are indicated. A wide variety <strong>of</strong> references are listed. Reasonable<br />

efforts have been made to publish reliable data and information, but the author and the publisher cannot<br />

assume responsibility for the validity <strong>of</strong> all materials or for the consequence <strong>of</strong> their use.<br />

No part <strong>of</strong> this book may be reprinted, reproduced, transmitted, or utilized in any form by any electronic,<br />

mechanical, or other means, now known or hereafter invented, including photocopying, micr<strong>of</strong>ilming, and<br />

recording, or in any information storage or retrieval system, without written permission from the publishers.<br />

For permission to photocopy or use material electronically from this work, please access www.copyright<br />

.com (http://www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC) 222 Rosewood<br />

Drive, Danvers, MA 01923, 978-750-8400. CCC is a not-for-pr<strong>of</strong>it organization that provides licenses and<br />

registration for a variety <strong>of</strong> users. For organizations that have been granted a photocopy license by the CCC,<br />

a separate system <strong>of</strong> payment has been arranged.<br />

Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used<br />

only for identification and explanation without intent to infringe.<br />

Library <strong>of</strong> Congress Cataloging-in-Publication Data<br />

<strong><strong>Ph</strong>armacogenetics</strong> <strong>of</strong> breast cancer : towards the individualization <strong>of</strong> therapy / edited by<br />

Brian Leyland-Jones.<br />

p. ; cm. — (Translational medicine series ; 7)<br />

Includes bibliographical references and index.<br />

ISBN-13: 978-1-4200-5293-0 (hardcover : alk. paper)<br />

ISBN-10: 1-4200-5293-4 (hardcover : alk. paper)<br />

1. <strong>Breast</strong>—<strong>Cancer</strong>—Genetic aspects. 2. <strong>Breast</strong>—<strong>Cancer</strong>—Chemotherapy.<br />

3. <strong><strong>Ph</strong>armacogenetics</strong>. I. Leyland-Jones, Brian. II. Series.<br />

[DNLM: 1. <strong>Breast</strong> Neoplasms—genetics. 2. <strong>Breast</strong> Neoplasms—drug therapy.<br />

WP 870 P536 2008]<br />

RC280.B8P43 2008<br />

616.99 0 449—dc22<br />

2008001013<br />

For Corporate Sales and Reprint Permissions call 212-520-2700 or write to: Sales Department,<br />

52 Vanderbilt Avenue, 16th floor, New York, NY 10017.<br />

Visit the Informa Web site at<br />

www.informa.com<br />

and the Informa Healthcare Web site at<br />

www.informahealthcare.com


Preface<br />

I write this preface on Christmas Day, 2007, in Costa Rica. Christmas is always a<br />

special time for me, since my parents were married on Christmas Day sixty<br />

nine years ago. It is almost unimaginable to think how far we have progressed<br />

since 1938. My parents were married in the year before the Second World<br />

War. The nitrogen mustards, from which arose the first chemotherapies, were<br />

developed during the war. I was fortunate during my training at Sloan-Kettering<br />

to work in the laboratories <strong>of</strong> Fred <strong>Ph</strong>illips, Joe Burchenal, and Jack Fox, which<br />

conducted some <strong>of</strong> the pivotal work on these compounds. Indeed, the first task<br />

force for childhood leukemia was set up in Washington in the year <strong>of</strong> my birth.<br />

However, we all have to remember how slowly treatments are disseminated to<br />

the community; I can remember, as a child, my next-door neighbor having a leg<br />

amputated for metastatic lung cancer since no other treatments were accessible<br />

locally!<br />

The rest is history, but again it is difficult to believe that the Nobel Prize<br />

for the discovery <strong>of</strong> the cellular origin <strong>of</strong> retroviral oncogenes was awarded to<br />

Bishop and Varmus only in 1989.<br />

Now we face the challenge <strong>of</strong> too few patients eligible for clinical trials to<br />

adequately test the myriad <strong>of</strong> targeted therapies available as well as the availability<br />

<strong>of</strong> high-quality starting material with associated clinical and phenotypic<br />

data. Various platforms are available for the identification <strong>of</strong> biomarkers: early<br />

technologies have focused on reverse transcriptase polymerase chain reaction<br />

and expression arrays. Now single nucleotide polymorphisms, array comparative<br />

genomic hybridization, methylation signatures, microRNAs, and proteomics are<br />

increasingly included in biomarker pr<strong>of</strong>iling; nanotechnology, molecular imaging,<br />

and the use <strong>of</strong> circulating tumor cells are advancing at a tremendous pace.<br />

By the time this book is released, the specimen collection guidelines for breast<br />

cancer, representing a joint effort between the North American cooperative<br />

iii


iv<br />

Preface<br />

groups and those comprising the <strong>Breast</strong> International Group (BIG), will have<br />

been published and be available for open Web access. We have attempted to<br />

cover the breadth <strong>of</strong> these diverse areas in this book.<br />

I am especially proud <strong>of</strong> three aspects <strong>of</strong> the book. The first is the timeliness,<br />

as so many new approaches to individualization <strong>of</strong> therapy have emerged, and<br />

when two very large randomized trials are aggressively accruing on the basis <strong>of</strong><br />

Oncotype and mammoprint. Second, the diversity <strong>of</strong> the topics covered, ranging<br />

from pathology, tissue handling, DNA-, expression-, and epigenetic-based<br />

approaches, nanotechnology, stem cells, circulating tumor cells, endocrine, and<br />

chemotherapeutic through to pharmacologic aspects and the key pivotal trials.<br />

Third, I am proud that so many <strong>of</strong> my most distinguished friends and colleagues<br />

from around the world have been so generous with their time and devotion to<br />

make this book outstanding.<br />

Finally, in writing this preface from one <strong>of</strong> the most beautiful places on<br />

earth, I am deeply reminded <strong>of</strong> the disparities in access to healthcare. We are all<br />

immensely grateful that my friends and colleagues John Seffrin and Otis<br />

Brawley at the American <strong>Cancer</strong> Society have been leaders in bringing this<br />

critical issue to the forefront <strong>of</strong> public consciousness, but we must never fail to<br />

remember that severe disparities in access to care remain, not only in the United<br />

States but globally. Individualized therapy is not only meant to direct therapy to<br />

those who will benefit most but also intended to reduce waste and improve cure<br />

rates dramatically, so that this therapy is available to all who suffer. Let us, the<br />

privileged, never move our focus from the most disadvantaged in our world and<br />

aim at using individualized therapy as a means to make our most sophisticated<br />

treatment available and affordable to all.<br />

Brian Leyland-Jones, M.D., <strong>Ph</strong>.D.


Acknowledgments<br />

Many thanks are owed on a personal level, especially to all my friends and<br />

coauthors who have been immensely generous in dedicating their nonexistent<br />

spare moments to contributing the various chapters. I am deeply grateful to the<br />

publisher, Informa Healthcare, who not only had the insight to partner with<br />

me on this book but also had the patience to wait for the “nonexistent spare<br />

moments” <strong>of</strong> the contributing authors to materialize. I must also pr<strong>of</strong>oundly<br />

apologize to several colleagues who were approached early on to contribute<br />

chapters but who were unable to because <strong>of</strong> the original time lines and now must<br />

be frustrated because the time lines were forced to be extended considerably.<br />

Most especially, I must thank my dear, dear colleague, Dr. Brian R. Smith, who<br />

was relentless in organization, follow-up, and advice, a man whose wisdom I<br />

rely upon at so many critical moments and to whose structured approach I aspire.<br />

I am deeply, deeply grateful to Charles Bronfman, whose chair in the memory <strong>of</strong><br />

his dear sister, Minda, sustained me over the past 17 years and who now has<br />

inspirationally established his own Institute <strong>of</strong> Personalized <strong>Cancer</strong> Therapy in<br />

New York. I am grateful also to the Flanders family who have shown me a depth<br />

<strong>of</strong> kindness, support, and friendship that is without equal. Finally, I am humbled<br />

by my research and clinical colleagues who instruct and guide me constantly,<br />

and especially by each and every patient who inspires us and teaches, in each and<br />

every encounter, just how personalized breast cancer therapy really is.<br />

v


Contents<br />

Preface . . . . . . . . . . iii<br />

Acknowledgments . . . . v<br />

Contributors . . . . . . . . xi<br />

1. <strong><strong>Ph</strong>armacogenetics</strong> <strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong>: Toward the Individualization<br />

<strong>of</strong> Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1<br />

Jenny C. Chang, Susan G. Hilsenbeck, and Suzanne A. W. Fuqua<br />

2. <strong>Ph</strong>armacology, <strong><strong>Ph</strong>armacogenetics</strong>, and <strong>Ph</strong>armacoepidemiology:<br />

Three P’s <strong>of</strong> Individualized Therapy . . . . . . . . . . . . . . . . . . . . 11<br />

Shaheenah Dawood<br />

3. Role <strong>of</strong> Genetic Variability in <strong>Breast</strong> <strong>Cancer</strong> Treatment<br />

Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

Kandace L. Amend, Ji-Yeob Choi, and Christine B. Ambrosone<br />

4. The Role <strong>of</strong> Epigenetics in <strong>Breast</strong> <strong>Cancer</strong>: Implications for<br />

Diagnosis, Prognosis, and Treatment . . . . . . . . . . . . . . . . . . . . 45<br />

Amy M. Dworkin, Tim H.-M. Huang, and Amanda E. Toland<br />

5. Comparative Genomic Hybridization and Copy Number<br />

Abnormalities in <strong>Breast</strong> <strong>Cancer</strong> . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

Nicholas Wang and Joe Gray<br />

vii


viii<br />

Contents<br />

6. Gene Expression Pr<strong>of</strong>iling as an Emerging Diagnostic Tool<br />

to Personalize Chemotherapy Selection for Early Stage<br />

<strong>Breast</strong> <strong>Cancer</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77<br />

Cornelia Liedtke and Lajos Pusztai<br />

7. Gene Expression–Based Predictors <strong>of</strong> Prognosis and Response<br />

to Chemotherapy in <strong>Breast</strong> <strong>Cancer</strong> . . . . . . . . . . . . . . . . . . . . . 97<br />

Soonmyung Paik<br />

8. Utilization <strong>of</strong> Genomic Signatures for Personalized<br />

Treatment <strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . 107<br />

P. Kelly Marcom, Carey K. Anders, Ge<strong>of</strong>frey S. Ginsburg,<br />

Joseph R. Nevins, and Anil Potti<br />

9. Personalized Medicine by the Use <strong>of</strong> Microarray Gene<br />

Expression Pr<strong>of</strong>iling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123<br />

Stella Mook, Laura J. van ’t Veer, and Fatima Cardoso<br />

10. Predictive Markers for Targeted <strong>Breast</strong> <strong>Cancer</strong> Treatment . . . 135<br />

Hans Christian B. Pedersen and John M. S. Bartlett<br />

11. Circulating Tumor Cells in Individualizing <strong>Breast</strong> <strong>Cancer</strong><br />

Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151<br />

James M. Reuben and Massimo Crist<strong>of</strong>anilli<br />

12. Individualization <strong>of</strong> Endocrine Therapy in <strong>Breast</strong> <strong>Cancer</strong> . . . . 167<br />

Amelia B. Zelnak, Ruth M. O’Regan, and Clodia Osipo<br />

13. TAILORx: Rationale for the Study Design . . . . . . . . . . . . . . . 185<br />

Joseph A. Sparano<br />

14. Taking Prognostic Signatures to the Clinic . . . . . . . . . . . . . . . 197<br />

Michail Ignatiadis and Christos Sotiriou<br />

15. Tissue is the Issue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213<br />

Adekunle Raji<br />

16. Acquisition and Preservation <strong>of</strong> Tissue for Microarray<br />

Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231<br />

Brian R. Untch and John A. Olson


Contents<br />

ix<br />

17. Tapping into the Formalin-Fixed Paraffin-Embedded (FFPE)<br />

Tissue Gold Mine for Individualization <strong>of</strong> <strong>Breast</strong><br />

<strong>Cancer</strong> Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243<br />

Mark Abramovitz and Brian Leyland-Jones<br />

18. <strong>Breast</strong> <strong>Cancer</strong> Stem Cells and Their Niche: Lethal Seeds in<br />

Lethal Soil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259<br />

Danuta Balicki, Brian Leyland-Jones, and Max S. Wicha<br />

19. Molecular Imaging in Individualized <strong>Cancer</strong> Management . . . 291<br />

David M. Schuster and Diego R. Martin<br />

20. <strong>Cancer</strong> Nanotechnology for Molecular Pr<strong>of</strong>iling and<br />

Individualized Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309<br />

May Dongmei Wang, Jonathan W. Simons, and Shuming Nie<br />

Index . . . . . . . 323


Contributors<br />

Mark Abramovitz<br />

VM Institute <strong>of</strong> Research, Montreal, Québec, Canada<br />

Christine B. Ambrosone Department <strong>of</strong> <strong>Cancer</strong> Prevention and Control,<br />

Roswell Park <strong>Cancer</strong> Institute, Buffalo, New York, U.S.A.<br />

Kandace L. Amend Department <strong>of</strong> Oncological Sciences, Mount Sinai School<br />

<strong>of</strong> Medicine, New York, New York, U.S.A.<br />

Carey K. Anders Duke Institute for Genome Sciences & Policy, Department<br />

<strong>of</strong> Medicine, Duke University Medical Center, Durham, North Carolina, U.S.A.<br />

Danuta Balicki Department <strong>of</strong> Medicine and Research Centre, Centre<br />

Hospitalier de l’Université de Montréal; Department <strong>of</strong> Medicine, Université de<br />

Montréal; and <strong>Breast</strong> <strong>Cancer</strong> Prevention and Genetics and <strong>Breast</strong> <strong>Cancer</strong> Risk<br />

Assessment Clinic, Ville Marie Multidisciplinary <strong>Breast</strong> <strong>Cancer</strong> Centre,<br />

Montréal, Québec, Canada<br />

John M. S. Bartlett Endocrine <strong>Cancer</strong> Group, <strong>Cancer</strong> Research Centre,<br />

Western General Hospital, Edinburgh, United Kingdom<br />

Fatima Cardoso<br />

Brussels, Belgium<br />

Department <strong>of</strong> Medical Oncology, Institute Jules Bordet,<br />

Jenny C. Chang <strong>Breast</strong> Center, Dan L. Duncan <strong>Cancer</strong> Center, Baylor College<br />

<strong>of</strong> Medicine, Houston, Texas, U.S.A.<br />

Ji-Yeob Choi Department <strong>of</strong> <strong>Cancer</strong> Prevention and Control, Roswell Park<br />

<strong>Cancer</strong> Institute, Buffalo, New York, U.S.A.<br />

xi


xii<br />

Contributors<br />

Massimo Crist<strong>of</strong>anilli Department <strong>of</strong> <strong>Breast</strong> Medical Oncology, The<br />

University <strong>of</strong> Texas M. D. Anderson <strong>Cancer</strong> Center, Houston, Texas, U.S.A.<br />

Shaheenah Dawood Department <strong>of</strong> Medical Oncology, Dubai Hospital,<br />

Dubai, United Arab Emirates<br />

Amy M. Dworkin Human <strong>Cancer</strong> Genetics Program, Department <strong>of</strong><br />

Molecular Virology, Immunology, and Medical Genetics, The Ohio State<br />

University, Columbus, Ohio, U.S.A.<br />

Suzanne A. W. Fuqua <strong>Breast</strong> Center, Dan L. Duncan <strong>Cancer</strong> Center, Baylor<br />

College <strong>of</strong> Medicine, Houston, Texas, U.S.A.<br />

Ge<strong>of</strong>frey S. Ginsburg Duke Institute for Genome Sciences & Policy,<br />

Department <strong>of</strong> Medicine, Duke University Medical Center, Durham,<br />

North Carolina, U.S.A.<br />

Joe Gray Life Sciences Division, Lawrence Berkeley National Laboratory,<br />

Berkeley, California, U.S.A.<br />

Susan G. Hilsenbeck <strong>Breast</strong> Center, Dan L. Duncan <strong>Cancer</strong> Center, Baylor<br />

College <strong>of</strong> Medicine, Houston, Texas, U.S.A.<br />

Tim H.-M. Huang Human <strong>Cancer</strong> Genetics Program, Department <strong>of</strong><br />

Molecular Virology, Immunology, and Medical Genetics, The Ohio State<br />

University, Columbus, Ohio, U.S.A.<br />

Michail Ignatiadis Department <strong>of</strong> Medical Oncology, Translational Research<br />

Unit, Institut Jules Bordet, Université Libre de Bruxelles, Brussels, Belgium<br />

Brian Leyland-Jones Winship <strong>Cancer</strong> Institute, Emory University School <strong>of</strong><br />

Medicine, Atlanta, Georgia, U.S.A.<br />

Cornelia Liedtke Department <strong>of</strong> <strong>Breast</strong> Medical Oncology, The University <strong>of</strong><br />

Texas, Anderson <strong>Cancer</strong> Center, Houston, Texas, U.S.A.<br />

P. Kelly Marcom Duke Institute for Genome Sciences & Policy, Department<br />

<strong>of</strong> Medicine, Duke University Medical Center, Durham, North Carolina, U.S.A.<br />

Diego R. Martin<br />

Georgia, U.S.A.<br />

Department <strong>of</strong> Radiology, Emory University, Atlanta,


Contributors<br />

xiii<br />

Stella Mook Department <strong>of</strong> Pathology, The Netherlands <strong>Cancer</strong> Institute,<br />

Amsterdam, The Netherlands<br />

Joseph R. Nevins Duke Institute for Genome Sciences & Policy, Department<br />

<strong>of</strong> Molecular Genetics and Microbiology, Duke University Medical Center,<br />

Durham, North Carolina, U.S.A.<br />

Shuming Nie Departments <strong>of</strong> Biomedical Engineering and Chemistry and<br />

the Winship <strong>Cancer</strong> Institute, Emory University and Georgia Institute <strong>of</strong><br />

Technology, Atlanta, Georgia, U.S.A.<br />

Ruth M. O’Regan<br />

Emory Winship <strong>Cancer</strong> Institute, Atlanta, Georgia, U.S.A.<br />

John A. Olson Institute for Genome Sciences and Policy and the Department<br />

<strong>of</strong> Surgery, Duke University Medical Center, Durham, North Carolina, U.S.A.<br />

Clodia Osipo<br />

Loyola University, Maywood, Illinois, U.S.A.<br />

Soonmyung Paik Division <strong>of</strong> Pathology, NSABP Foundation, Pittsburgh,<br />

Pennsylvania, U.S.A.<br />

Hans Christian B. Pedersen Endocrine <strong>Cancer</strong> Group, <strong>Cancer</strong> Research<br />

Centre, Western General Hospital, Edinburgh, United Kingdom<br />

Anil Potti Duke Institute for Genome Sciences & Policy, Department <strong>of</strong><br />

Medicine, Duke University Medical Center, Durham, North Carolina, U.S.A.<br />

Lajos Pusztai Department <strong>of</strong> <strong>Breast</strong> Medical Oncology, The University <strong>of</strong><br />

Texas, Anderson <strong>Cancer</strong> Center, Houston, Texas, U.S.A.<br />

Adekunle Raji Eastern Cooperative Oncology Group Pathology Coordinating<br />

Office, Northwestern University, Evanston, Illinois, U.S.A.<br />

James M. Reuben Department <strong>of</strong> Hematopathology, The University <strong>of</strong> Texas<br />

M. D. Anderson <strong>Cancer</strong> Center, Houston, Texas, U.S.A.<br />

David M. Schuster Division <strong>of</strong> Nuclear Medicine and Molecular Imaging,<br />

Department <strong>of</strong> Radiology, Emory University, Atlanta, Georgia, U.S.A.<br />

Jonathan W. Simons<br />

California, U.S.A.<br />

The Prostate <strong>Cancer</strong> Foundation, Santa Monica,


xiv<br />

Contributors<br />

Christos Sotiriou Department <strong>of</strong> Medical Oncology, Translational Research<br />

Unit, Institut Jules Bordet, Université Libre de Bruxelles, Brussels, Belgium<br />

Joseph A. Sparano Albert Einstein College <strong>of</strong> Medicine, Montefiore Medical<br />

Center, Bronx, New York, U.S.A.<br />

Amanda E. Toland Human <strong>Cancer</strong> Genetics Program, Department <strong>of</strong><br />

Molecular Virology, Immunology, and Medical Genetics, The Ohio State<br />

University, Columbus, Ohio, U.S.A.<br />

Brian R. Untch Institute for Genome Sciences and Policy and the Department<br />

<strong>of</strong> Surgery, Duke University Medical Center, Durham, North Carolina, U.S.A.<br />

Laura J. van ’t Veer Department <strong>of</strong> Pathology, The Netherlands <strong>Cancer</strong><br />

Institute, Amsterdam, The Netherlands<br />

May Dongmei Wang Departments <strong>of</strong> Biomedical Engineering and Electrical<br />

and Computer Engineering, Georgia Institute <strong>of</strong> Technology and Emory<br />

University, Atlanta, Georgia, U.S.A.<br />

Nicholas Wang Life Sciences Division, Lawrence Berkeley National<br />

Laboratory, Berkeley, California, U.S.A.<br />

Max S. Wicha Comprehensive <strong>Cancer</strong> Centre, Department <strong>of</strong> Internal<br />

Medicine, Division <strong>of</strong> Hematology-Oncology, University <strong>of</strong> Michigan, Ann<br />

Arbor, Michigan, U.S.A.<br />

Amelia B. Zelnak<br />

Emory Winship <strong>Cancer</strong> Institute, Atlanta, Georgia, U.S.A.


1<br />

<strong><strong>Ph</strong>armacogenetics</strong> <strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong>:<br />

Toward the Individualization <strong>of</strong> Therapy<br />

Jenny C. Chang, Susan G. Hilsenbeck, and Suzanne A. W. Fuqua<br />

<strong>Breast</strong> Center, Dan L. Duncan <strong>Cancer</strong> Center,<br />

Baylor College <strong>of</strong> Medicine, Houston, Texas, U.S.A.<br />

INTRODUCTION<br />

Mortality from breast cancer results from the ability <strong>of</strong> some tumors to metastasize<br />

to distant sites. Selecting patients with micrometastases at diagnosis is<br />

crucial for clinicians in deciding who should, and who should not, receive toxic<br />

and expensive adjuvant chemotherapy to eradicate these metastatic cells.<br />

Although many individual biomarkers were originally attractive, over the years<br />

most have failed to become clinically useful. In addition, the management <strong>of</strong><br />

breast cancer has changed, with the majority <strong>of</strong> node-negative patients now<br />

undergoing systemic adjuvant therapy because we cannot precisely determine an<br />

individual’s risk <strong>of</strong> recurrence. A majority <strong>of</strong> node-negative patients are being<br />

unnecessarily overtreated because if left systemically untreated, only about 25%<br />

<strong>of</strong> node-negative patients would ever develop recurrence. There is therefore a<br />

critical need to identify patients with sufficiently low risk <strong>of</strong> breast cancer<br />

recurrence so as to avoid further treatment. In addition, in patients at risk <strong>of</strong><br />

recurrence and in need <strong>of</strong> therapy, optimal therapeutic selection is an increasingly<br />

important objective. Recent developments in applying microarray technologies<br />

to breast tumor samples suggest that these new techniques may provide<br />

for the transition <strong>of</strong> molecular biological discoveries to clinical application and<br />

1


2 Chang et al.<br />

will generate clinically useful genomic pr<strong>of</strong>iles that more accurately predict the<br />

long-term outcome <strong>of</strong> individual breast cancer patients.<br />

BACKGROUND<br />

Until recently, evaluations <strong>of</strong> prognostic and predictive factors have considered<br />

one factor at a time or have used small panels <strong>of</strong> markers. However, with the<br />

advent <strong>of</strong> new genomic technologies such as microarrays, capable <strong>of</strong> simultaneously<br />

measuring thousands <strong>of</strong> genes or gene products, we are beginning to<br />

construct molecular fingerprints <strong>of</strong> individual tumors so that accurate prognostic<br />

and predictive assessments <strong>of</strong> each cancer may be made. Clinicians may one day<br />

base clinical management on each woman’s personal prognosis and predict the<br />

best individual therapies according to the genetic fingerprint <strong>of</strong> each individual<br />

cancer.<br />

<strong>Breast</strong> cancer is characterized by a very heterogeneous clinical course. A<br />

major goal <strong>of</strong> recent studies is to determine whether RNA microarray expression<br />

pr<strong>of</strong>iling or DNA array gene amplification or gene loss patterns can accurately<br />

predict an individual’s long-term potential for recurrence <strong>of</strong> breast cancer, so that<br />

appropriate treatment decisions can be made. Microarrays can be used to measure<br />

the mRNA expression <strong>of</strong> thousands <strong>of</strong> genes at one time or survey genomic<br />

alterations that may distinguish molecular phenotypes associated with long-term,<br />

recurrence-free survival or clinical response to treatment. These new technologies<br />

have been successfully applied to primary breast cancers and may<br />

eventually outperform currently used clinical parameters in predicting disease<br />

outcome.<br />

Since RNA expression microarray technology provided a method for<br />

monitoring the RNA expression <strong>of</strong> many thousands <strong>of</strong> human genes at a time,<br />

there was considerable anticipation that it would quickly and easily revolutionize<br />

our approaches to cancer diagnosis, prognosis, and treatment. The reality<br />

remains extremely promising but is also complex. A potential complication in<br />

the application <strong>of</strong> microarray technology to primary human breast tumor samples<br />

is the presence <strong>of</strong> variable numbers <strong>of</strong> normal cells, such as stroma, blood<br />

vessels, and lymphocytes, in the tumor. Indeed, it has been demonstrated using<br />

gross analysis <strong>of</strong> human breast cancer specimens compared with breast cancer<br />

cell lines that the tumors expressed sets <strong>of</strong> genes in common not only with these<br />

cell lines but also with cells <strong>of</strong> hematopoietic lineage and stromal origin (1,2).<br />

Laser capture microdissection has also been successfully used to isolate pure-cell<br />

populations from primary breast cancers for array pr<strong>of</strong>iling (3). Sgroi et al. (3)<br />

utilized laser capture microdissection to isolate morphologically “normal” breast<br />

epithelial cells, invasive breast cancer cells, and metastatic lymph node cancer<br />

cells from one patient and were able to demonstrate the feasibility <strong>of</strong> using<br />

microdissected samples for array pr<strong>of</strong>iling as well as following potential progression<br />

<strong>of</strong> cancer in this patient. However, with the emerging data supporting<br />

important roles for the surrounding stroma in breast cancer progression and the


<strong><strong>Ph</strong>armacogenetics</strong> <strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong> 3<br />

labor-intensive and technically challenging nature <strong>of</strong> laser capture technology<br />

with subsequent amplification <strong>of</strong> RNA for quantitation, most published investigations<br />

to date have evaluated total gene expression to identify prognostic<br />

pr<strong>of</strong>iles, as will be described in the next section.<br />

MOLECULAR CLASSIFICATION OF BREAST CANCER<br />

A study <strong>of</strong> sporadic breast tumor samples by Perou et al. (2) was the first to show<br />

that breast tumors could be classified into subtypes distinguished by differences<br />

in their expression pr<strong>of</strong>iles. Using 40 breast tumors and 20 matched pairs<br />

<strong>of</strong> samples before and after doxorubicin treatment, an “intrinsic gene set” <strong>of</strong><br />

476 genes was selected that was more variably expressed between the 40 sporadic<br />

tumors than between the paired samples. This intrinsic gene set was then used to<br />

cluster and segregate the tumors into four major subgroups: a “luminal cell-like”<br />

group expressing the estrogen receptor (ER); a “basal cell-like” group expressing<br />

keratins 5 and 17, integrinb4, and laminin, but lacking ER expression; an<br />

“Erb-B2-positive” group; and a “normal” epithelial group (Fig. 1).<br />

In a subsequent study with 38 additional cancers, the investigators found<br />

the same subgroups as before (4), except that the luminal, ER-positive group was<br />

further subdivided into subsets with distinctive gene expression pr<strong>of</strong>iles. In<br />

univariate survival analysis, performed on the 49 patients diagnosed with locally<br />

advanced disease but without evidence <strong>of</strong> distant metastasis, ER positivity was<br />

not a significant prognostic factor on its own, but the luminal-type group enjoyed<br />

a more favorable survival compared with the other groups. Conversely, the<br />

basal-like group had a significantly poorer prognosis. Although small and<br />

exploratory, this study suggests that important differences in outcome can be<br />

ascertained from microarray expression pr<strong>of</strong>iling.<br />

An interesting study was reported by Gruvberger et al. (5), who pr<strong>of</strong>iled 58<br />

grossly dissected primary invasive breast tumors and used artificial neural network<br />

analysis to predict the ER status <strong>of</strong> the tumors on the basis <strong>of</strong> their gene expression<br />

patterns. They then determined which specific genes were the most important for ER<br />

classification. By comparing to Serial Analysis <strong>of</strong> Gene Expression (SAGE) data<br />

from estradiol-stimulated breast cancer cells, they determined that only a few genes<br />

<strong>of</strong> the many genes that were associated with ER expression in tumors were indeed<br />

estrogen responsive in cell culture. This observation lent further support to the<br />

hypothesis developed by Perou et al. (1) that basic cell lineages, such as the luminal<br />

ER-positive cell type, can be partly explained by observed genomic gene expression<br />

patterns rather than downstream effectors <strong>of</strong> only one pathway, such as the ER.<br />

PROGNOSTIC IMPLICATIONS<br />

Microarrays have also been used to predict lymph node status and very shortterm<br />

relapse-free survival in two groups (n ¼ 37 and 52, respectively) <strong>of</strong> heterogeneously<br />

treated patients (6). Although prediction <strong>of</strong> nodal status is <strong>of</strong>


4 Chang et al.<br />

Figure 1 Supervised classification <strong>of</strong> prognosis signatures (van ’t Veer). The 78 tumors<br />

are listed vertically, and the 70 “prognostic” genes horizontally. (A) Study design. (B)Heatmap<br />

<strong>of</strong> prognostic signature.<br />

limited interest clinically, the study uses innovative statistical methods, rigorously<br />

generates estimates <strong>of</strong> future classifier performance, and further demonstrates<br />

the feasibility <strong>of</strong> accurate prediction <strong>of</strong> tumor biology using expression<br />

arrays. In a more focused and somewhat more clinically relevant study, van ’t Veer<br />

et al. (7) used RNA expression microarray analyses to identify a 70-gene<br />

prognostic gene signature (“classifier”) in young, untreated, axillary lymph<br />

node–negative patients using a training set <strong>of</strong> 44 good (disease free for more than<br />

5 years) and 34 poor (distant relapse in less than 5 years) outcome tumors and<br />

then tested the classifier in a validation set <strong>of</strong> 19 tumors. The same group (8) has<br />

now extended the study to a total <strong>of</strong> 295 young (


<strong><strong>Ph</strong>armacogenetics</strong> <strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong> 5<br />

70-gene classifier (7). The microarray-based predictions are consistent with, and<br />

perhaps better than, estimates that can be obtained with current prognostic<br />

indices.<br />

GENETIC SUSCEPTIBILITY<br />

A few studies have utilized new genomic approaches for the study <strong>of</strong> inherited<br />

breast cancer (8). There is accumulating evidence, both epidemiological and<br />

histological, that tumors arising as a result <strong>of</strong> mutations in the two breast cancer<br />

susceptibility gene families (BRCA1 and BRCA2) are biologically distinct. For<br />

instance, BRCA1 breast cancers are most <strong>of</strong>ten ER and PR negative, but BRCA2<br />

cancers more <strong>of</strong>ten tend to be positive for these receptors (9). In a seminal paper<br />

published by Hedenfalk et al. (10), seven tumors each from BRCA1 and BRCA2<br />

gene mutation carriers or sporadic breast cancers were compared by expression<br />

microarray analysis. They found that the gene expression pr<strong>of</strong>iles <strong>of</strong> the three<br />

tumor groups differed significantly from each other, underscoring the fundamental<br />

differences between BRCA1 and BRCA2 mutation-associated tumors. Of<br />

course a potential confounding issue was the differential distribution <strong>of</strong> ER<br />

between the BRCA1 and BRCA2 tumors. However, even after removal <strong>of</strong> ERor<br />

PR-associated genes from the analysis, the two inherited tumor groups were<br />

still discernable. Thus, ER status alone does not fully explain the observed<br />

differences in gene expression pr<strong>of</strong>iles. Although this study is obviously very<br />

small, and other confounding issues such as tumor stage, grade, and treatment<br />

cannot be considered, it does set a foundation for larger validation studies to<br />

confirm differential genes, which could then provide important clues to the<br />

etiology <strong>of</strong> inheritable breast cancer.<br />

Microarrays are also being studied as a way to predict response to systemic<br />

therapy. The neoadjuvant setting is especially attractive for these studies for<br />

several reasons, including early assessment <strong>of</strong> response to therapy, biopsiable<br />

access to the primary tumor, and considerable reduced sample sizes compared to<br />

those required in the adjuvant setting.<br />

PREDICTIVE IMPLICATIONS<br />

Methods for assessing response in neoadjuvant trials remain problematic. Clinical<br />

response to neoadjuvant chemotherapy is a validated surrogate marker for<br />

improved survival (11,12). Women who achieve pathologic complete response<br />

are most likely to have the best clinical outcome, although survival is still<br />

improved in those who clinically respond but do not achieve pathologic complete<br />

response.<br />

In an early study, Buchholz et al. (13) obtained sufficient RNA from core<br />

biopsies <strong>of</strong> five patients to perform serial microarray expression pr<strong>of</strong>iles and<br />

showed that despite differences in therapy, patients with good pathological


6 Chang et al.<br />

responses to neoadjuvant treatment appeared to have gene pr<strong>of</strong>iles that clustered<br />

distinctly than those <strong>of</strong> patients who were poor responders to treatment. More<br />

recently, Chang et al. (14) have shown that gene pr<strong>of</strong>iling can be used to<br />

accurately predict response to neoadjuvant docetaxel. The study enrolled<br />

24 subjects, extracted sufficient RNA from all core needle biopsies, and constructed<br />

a 92-gene predictor <strong>of</strong> response. In a complete cross-validation analysis, which<br />

gives an unbiased estimate <strong>of</strong> performance on future samples, the classifier<br />

correctly identified 10 <strong>of</strong> 11 responders and 11 <strong>of</strong> 13 nonresponders with an<br />

overall accuracy <strong>of</strong> 88%. In a small validation set, this 92-gene classifier successfully<br />

predicted response in six patients (Fig. 2). This compares very favorably<br />

with the best existing predictive factors for response to specific therapy and<br />

strongly suggests that after appropriately extensive validation, microarray<br />

pr<strong>of</strong>iling will be useful for treatment selection. A second neoadjuvant study was<br />

recently published that used cDNA arrays to develop predictors for paclitaxel,<br />

fluorouracil, doxorubicin, and cyclophosphamide, involving 24 samples. A<br />

classifier with 74 markers was developed, with 78% accuracy, suggesting that<br />

transcriptional pr<strong>of</strong>iling has the potential to identify a gene expression pattern in<br />

breast cancer that may lead to clinically useful predictors <strong>of</strong> chemotherapy<br />

response (15).<br />

In vitro drug sensitivity data in NCI60 cell lines sensitive and resistant to<br />

specific chemotherapeutic agents (adriamycin, cyclophosphamide, docetaxel,<br />

etoposide, 5-fluoruracil, paclitaxel, topotecan) were recently published (16).<br />

Gene set enrichment analysis was performed, and gene expression signatures<br />

predictive <strong>of</strong> sensitivity to individual chemotherapeutic drugs were reported.<br />

Each signature was validated with response data from an independent set <strong>of</strong> cell<br />

line studies. These signatures were then used to predict clinical response in<br />

individuals treated with these drugs. Notably, signatures developed to predict<br />

response to individual agents, when combined, were found to also predict<br />

response to multidrug regimens (16). These pr<strong>of</strong>iles are being validated in<br />

prospective clinical trials.<br />

However, we acknowledge that studies to construct and validate arraybased<br />

prognostic and predictive “markers” are complex. These studies must<br />

address all <strong>of</strong> the concerns associated with ordinary, single-gene markers as well<br />

as a number <strong>of</strong> considerations unique to array studies. Recommendations for<br />

development <strong>of</strong> array-based prognostic classifiers have recently been enunciated<br />

by Simon et al. (17). Among the most important points, they recommend that<br />

studies should include (1) adequately large sample sizes in both training and<br />

validation sets, (2) complete iteration <strong>of</strong> the entire classifier construction process<br />

in estimating cross-validated prediction rates, (3) head-to-head comparison <strong>of</strong><br />

alternative classifiers on the same dataset, and (4) the full diversity <strong>of</strong> cases in<br />

any validation set. In addition, gene expression patterns may be confounded by<br />

several other factors, including ovarian ablation in premenopausal, ER-positive<br />

patients and a different mechanism <strong>of</strong> action <strong>of</strong> combination chemotherapies.


<strong><strong>Ph</strong>armacogenetics</strong> <strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong> 7<br />

Figure 2 Hierarchical clustering <strong>of</strong> genes correlated with docetaxel response. Sensitive<br />

tumors (S) are defined as 25% residual disease, and resistant tumors (R) are defined as<br />

greater than 25% residual disease.<br />

Groups <strong>of</strong> patients with different characteristics, such as menopausal and ER<br />

status or human epidermal growth factor receptor 2 (HER2) overexpression, may<br />

be necessary to definitively determine classifying patterns in these subsets <strong>of</strong><br />

patients.


8 Chang et al.<br />

CONCLUSIONS<br />

The goal <strong>of</strong> comprehensive, genomewide approaches is to identify clinically<br />

useful genetic pr<strong>of</strong>iles that will accurately predict the outcome to therapy and<br />

prognosis <strong>of</strong> patients with breast cancer. Despite improvements in technology,<br />

complex mechanisms driving breast cancer evolution continue to present challenges<br />

for the use <strong>of</strong> genomic approaches to better understand breast and other<br />

cancers. This, combined with our use <strong>of</strong> different markers, methods, tumors (e.g.,<br />

differing ER and HER2), and measurements <strong>of</strong> clinical outcomes, impedes the<br />

development <strong>of</strong> a consensus regarding predictive and prognostic markers for<br />

breast cancer.<br />

As this field matures, genomic studies examining identical breast tumor<br />

sets with multiple complementary technologies (e.g., loss <strong>of</strong> heterozygosity,<br />

comparative genomic hybridization, gene expression array analysis) will prove<br />

essential to unraveling the genetic heterogeneity characteristic <strong>of</strong> this disease. A<br />

combined genomic approach is necessary to define the underlying heterogeneous<br />

complexity characteristic <strong>of</strong> breast cancer. These data should lead to the identification<br />

and characterization <strong>of</strong> breast cancer subtypes, define the malignant<br />

potential <strong>of</strong> a given lesion, and predict its sensitivity to specific therapies. These<br />

multidisciplinary approaches should contribute to a better biological understanding,<br />

and therefore improved clinical management, <strong>of</strong> breast cancer.<br />

ACKNOWLEDGMENTS<br />

This work was supported by funding provided by National Institutes <strong>of</strong> Health<br />

(NIH) grants P50 CA58183 and P01 CA30195, the <strong>Breast</strong> <strong>Cancer</strong> Research<br />

Foundation, the Emma Jacobs Clinical <strong>Breast</strong> <strong>Cancer</strong> Fund, and the U.S. Army<br />

Medical Research and Materiel Command DAMD17-01-0132 and W81XWH-<br />

04-1-0468.<br />

REFERENCES<br />

1. Perou CM, Jeffrey SS, van de Run M, et al. Distinctive gene expression patterns in<br />

human mammary epithelial cells and breast cancers. Proc Natl Acad Sci USA 1999;<br />

96:9212–9217.<br />

2. Perou CM, Sorlie T, Eisen MB, et al. Molecular portraits <strong>of</strong> human breast tumours.<br />

Nature 2000; 406:747–752.<br />

3. Sgroi DC, Teng S, Robinson G, et al. In vivo gene expression pr<strong>of</strong>ile analysis <strong>of</strong><br />

human breast cancer progression. <strong>Cancer</strong> Res 1999; 59:5656–5661.<br />

4. Sorlie T, Perou CM, Tibshirani R, et al. Gene expression patterns <strong>of</strong> breast carcinomas<br />

distinguish tumor subclasses with clinical implications. Proc Natl Acad Sci USA 2001;<br />

98:10869–10874.<br />

5. Gruvberger S, Ringner M, Chen Y, et al. Estrogen receptor status in breast cancer is<br />

associated with remarkably distinct gene expression patterns. <strong>Cancer</strong> Res 2001; 61:<br />

5979–5984.


<strong><strong>Ph</strong>armacogenetics</strong> <strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong> 9<br />

6. Huang E, Cheng SH, Dressman H, et al. Gene expression predictors <strong>of</strong> breast cancer<br />

outcomes. Lancet 2003; 361:1590–1596.<br />

7. van ’t Veer LJ, Dai H, van De Vijver MJ, et al. Gene expression pr<strong>of</strong>iling predicts<br />

clinical outcome <strong>of</strong> breast cancer. Nature 2002; 415:530–536.<br />

8. van de Vijver MJ, He YD, van ’t Veer LJ, et al. A gene-expression signature as a<br />

predictor <strong>of</strong> survival in breast cancer. N Engl J Med 2002; 347:1999–2009.<br />

9. Verhoog LG, Brekelmans CTM, Seynaeve C, et al. Survival in hereditary breast cancer<br />

associated with germline mutations <strong>of</strong> BRCA2. J Clin Oncol 1999; 17:3396–3402.<br />

10. Hedenfalk I, Duggan D, Chen Y, et al. Gene-expression pr<strong>of</strong>iles in hereditary breast<br />

cancer. N Engl J Med 2001; 344:539–548.<br />

11. Fisher B, Bryant J, Wolmark N, et al. Effect <strong>of</strong> preoperative chemotherapy on the<br />

outcome <strong>of</strong> women with operable breast cancer. J Clin Oncol 1998; 16:2672–2685.<br />

12. Chang J, Powles TJ, Allred DC, et al. Biologic markers as predictors <strong>of</strong> clinical<br />

outcome from systemic therapy for primary operable breast cancer. J Clin Oncol<br />

1999; 17:3058–3063.<br />

13. Buchholz TA, Stivers DN, Stec J, et al. Global gene expression changes during<br />

neoadjuvant chemotherapy for human breast cancer. <strong>Cancer</strong> J 2002; 8:461–468.<br />

14. Chang JC, Wooten EC, Tsimelzon A, et al. Gene expression pr<strong>of</strong>iling predicts<br />

therapeutic response to docetaxel (Taxotere TM ) in breast cancer patients. Lancet<br />

2003; 362:280–287.<br />

15. Ayers M, Symmans WF, Stec J, et al. Gene expression pr<strong>of</strong>iles predict complete<br />

pathologic response to neoadjuvant paclitaxel and fluorouracil, doxorubicin, and<br />

cyclophosphamide chemotherapy in breast cancer. J Clin Oncol 2004; 22:2284–2293.<br />

16. Potti A, Dressman HK, Bild A, et al. Genomic signatures to guide the use <strong>of</strong> chemotherapeutics.<br />

Nat Med 2006; 12:1294–1300.<br />

17. Simon R, Radmacher MD, Dobbin K, et al. J Natl <strong>Cancer</strong> Inst 2003; 95:14–18.


2<br />

<strong>Ph</strong>armacology, <strong><strong>Ph</strong>armacogenetics</strong>,<br />

and <strong>Ph</strong>armacoepidemiology: Three<br />

P’s <strong>of</strong> Individualized Therapy<br />

Shaheenah Dawood<br />

Department <strong>of</strong> Medical Oncology, Dubai Hospital,<br />

Dubai, United Arab Emirates<br />

If it were not for the great variability among individuals medicine might as<br />

well be a science and not an art.<br />

Sir William Osler, 1892<br />

INTRODUCTION<br />

Over the last decade significant advances have been made in the management <strong>of</strong><br />

breast cancer. This is exemplified by the development <strong>of</strong> several new agents and<br />

significantly improved survival rates. However, despite our best efforts breast<br />

cancer remains the second most common cause <strong>of</strong> cancer-related death in the<br />

United States with approximately 40,910 women expected to die <strong>of</strong> this disease<br />

in 2007 (1). Unfortunately, such trends are not restricted to only breast cancer but<br />

have been observed across a range <strong>of</strong> solid tumors. A large portion <strong>of</strong> the<br />

underlying problem is attributed to the fact that the choice <strong>of</strong> a chemotherapeutic<br />

agent is <strong>of</strong>ten empirical based on guidelines that assumes all patients with a<br />

particular malignancy to be homogeneous. The unfortunate outcome <strong>of</strong> this<br />

11


12 Dawood<br />

erroneous belief is that approximately 40% <strong>of</strong> patients may be receiving the<br />

wrong drug (2).<br />

Numerous studies have consistently observed that the efficacy and toxicity<br />

<strong>of</strong> chemotherapeutic agents are heterogeneous rather than homogeneous across<br />

populations. Despite a multitiered approach to determining the efficacy and safety<br />

pr<strong>of</strong>iles <strong>of</strong> chemotherapeutic agents, spanning from in vitro assays, animal studies<br />

through to large human population studies, the current practice <strong>of</strong> administering<br />

the same dose <strong>of</strong> a given anticancer drug to a population <strong>of</strong> patients results in<br />

considerable variation in clinical outcome. These effects cover a spectrum <strong>of</strong><br />

possibilities ranging from success to failure when considering treatment outcome<br />

as an endpoint, and from no effect to a lethal event when treatment-associated<br />

toxicity is the variable under study (3). Complicating the problem further is the<br />

fact that currently available anticancer drugs show limited efficacy in up to 70% <strong>of</strong><br />

patients with adverse drug reactions accounting for as many as 100,000 patient<br />

deaths and two million hospitalizations in the United States every year (4).<br />

The key is to be able to, at the individual level, prospectively identify<br />

patients at risk for severe toxicity and those likely to attain maximum benefit<br />

from a given chemotherapeutic agent. Understanding that an individual does not<br />

live in isolation but is in fact a product <strong>of</strong> the gene-environment interaction is<br />

also vital to our understanding <strong>of</strong> the concept <strong>of</strong> individualized therapy (Fig. 1).<br />

Figure 1<br />

Sources <strong>of</strong> pharmacologic and pharmacogenetic variability.


<strong>Ph</strong>armacology, <strong><strong>Ph</strong>armacogenetics</strong>, and <strong>Ph</strong>armacoepidemiology 13<br />

Designing treatment strategies based on individual efficacy and toxicity predictions<br />

will shift clinical practice from simply an “art” to a “science” by<br />

incorporating a combination <strong>of</strong> physiological variables, genetic characteristics,<br />

and environmental factors (all known to alter the dose-concentration time pr<strong>of</strong>ile)<br />

as well as population epidemiological elements resulting in a practice<br />

paradigm <strong>of</strong> “personalized medicine.”<br />

THE COMPONENTS OF INDIVIDUALIZED THERAPY: THE THREE P’s<br />

When we administer a drug to a patient, our goal is to deliver a dose that will<br />

achieve maximum benefit with little, if any, associated toxicity. Seldom is this<br />

goal achieved. Individual differences in tumor response and normal tissue toxicity<br />

have been consistently observed with most anticancer agents. These differences<br />

are due to a variety <strong>of</strong> clinical variables (e.g., gender, age, diet, and<br />

drug-drug interactions) as well as variations in drug metabolism and transport.<br />

It is believed that 30% <strong>of</strong> patients treated with a particular drug will exhibit<br />

beneficial effects from the drug, 30% will not attain any form <strong>of</strong> beneficial<br />

effect, 10% will experience only side effects from the drug, and 30% are noncompliant<br />

with the drug <strong>of</strong>ten as a result <strong>of</strong> drug-related side effects or perceived<br />

insufficient effect (5). This results in a waste <strong>of</strong> allocated resources with large<br />

proportions <strong>of</strong> patients being unnecessarily exposed to adverse side effects. Here<br />

we will describe three components that aim to explain the observed differences<br />

in responses to individual drugs: pharmacology, pharmacoepidemiology, and<br />

pharmacogenetics.<br />

Current dosing <strong>of</strong> anticancer drugs is based on either a fixed quantity <strong>of</strong><br />

the drug or on a dose that is normalized to the individual body surface area.<br />

This method makes the assumption that within a group <strong>of</strong> individuals there<br />

will be a uniform degree <strong>of</strong> systemic exposure to the drug. With studies<br />

reporting 2- to 10-fold variations in drug clearance, this assumption is clearly<br />

not valid (6). With drugs used in the field <strong>of</strong> oncology having narrow therapeutic<br />

indices, it thus becomes imperative that we understand the mechanisms<br />

behind the observed variability in drug response and toxicity when treating a<br />

patient with cancer. At the clinical level, variability in drug response can be<br />

explained by its pharmacology that describes the pharmacokinetic and pharmacodynamic<br />

pr<strong>of</strong>iles <strong>of</strong> the drug. <strong>Ph</strong>armacokinetics explain “what the body<br />

does to the drug” by describing the relationship between time and plasma<br />

concentrations <strong>of</strong> the drug metabolites. Thisrelationshipisaffectedbyvariables<br />

such as absorption, distribution, metabolism, and excretion <strong>of</strong> the drug.<br />

The underlying objective <strong>of</strong> pharmacodynamics is to describe “what the drug<br />

does to the body” by correlating drug concentration to drug effect (both beneficial<br />

and adverse effects). Both the pharmacokinetic and pharmacodynamic<br />

components <strong>of</strong> a drug are not independent, but represent a spectrum <strong>of</strong> continuous<br />

events starting with the ingestion <strong>of</strong> the drug and culminating in the<br />

observed clinical effect.


14 Dawood<br />

Variability in any <strong>of</strong> the pharmacokinetic components will affect the<br />

observed drug effect. This is typically observed in patients with organ dysfunction<br />

where the inability to either metabolize or excrete the drug will lead to<br />

exaggerated drug effects. For example, the use <strong>of</strong> carboplatin in patients with<br />

renal dysfunction exaggerates the thrombocytopenic effect <strong>of</strong> this drug. Dosing<br />

formulas using creatinine clearance are used to avoid this problem while maximizing<br />

dose intensity (7). Another example is the use <strong>of</strong> taxanes, which are<br />

primarily metabolized and excreted by the liver. When administered to patients<br />

with hepatic dysfunction, prolonged bouts <strong>of</strong> severe neutropenia are observed.<br />

These examples illustrate how clinical variables can affect the pharmacokinetic<br />

pr<strong>of</strong>ile <strong>of</strong> a drug leading to disturbances in effect. Good clinical assessment <strong>of</strong><br />

the patients is therefore the first and most important step in designing an individual<br />

treatment regimen that will maximize drug benefit.<br />

<strong>Ph</strong>armacoepidemiology is a sub discipline <strong>of</strong> epidemiology that focuses on<br />

understanding why individuals respond differently to drug therapy. It is defined<br />

“as the study <strong>of</strong> the distribution and determinants <strong>of</strong> drug-related events in<br />

populations and the application <strong>of</strong> this study to efficacious drug treatments” (8).<br />

These drug-related events pertain to both beneficial and adverse events. <strong>Ph</strong>armacoepidemology,<br />

through the use <strong>of</strong> observational methods, aims to quantify<br />

drug exposure and to describe, as well as predict, the effectiveness and safety <strong>of</strong><br />

the drug within defined populations at specific places and time. Factors<br />

important in interpreting the variability in drug outcome include a patient’s<br />

health pr<strong>of</strong>ile, sex, age, diet, associated comorbidities, disease severity and<br />

prognosis, drug compliance, drug prescribing and dispensing quality, and genetic<br />

pr<strong>of</strong>ile <strong>of</strong> the patient and tumor (9,10). Depending on drug effect there are four<br />

main groups <strong>of</strong> patients: (i) responders, (ii) non-responders, (iii) toxic responders,<br />

and (iv) toxic non-responders (Fig. 2).<br />

<strong><strong>Ph</strong>armacogenetics</strong>, a third and important component <strong>of</strong> individualized<br />

therapy, focuses on describing the extent to which an individual’s genetic<br />

makeup is responsible for the observed differences in drug efficacy and toxicity<br />

pr<strong>of</strong>iles (11,12). This information is then used to make predictions about the<br />

safety, toxicity, and efficacy <strong>of</strong> drugs in individual patients. Inherited variability<br />

<strong>of</strong> drug targets, drug metabolizing enzymes, and drug transporters may all have a<br />

major impact on overall drug response, disposition, and associated adverse side<br />

effects by altering the pharmacokinetic and pharmacodynamic properties <strong>of</strong><br />

drugs (Table 1).<br />

Genetic variations include nucleotide repeats, insertions, deletions, and<br />

single nucleotide polymorphisms (SNPs). SNPs are the simplest and most<br />

commonly studied DNA polymorphism that occur once every 1900 base pairs in<br />

the 3 billion bases in the human genome (13) and account for more than 90% <strong>of</strong><br />

the genetic variation observed. More than 1.4 million SNPs have been identified<br />

in the human genome. Of these more than 60,000 SNPs occur in the coding<br />

region <strong>of</strong> genes with some SNPs being associated with variations in drug<br />

metabolism and effects (13). When polymorphisms occur in the promoter region


<strong>Ph</strong>armacology, <strong><strong>Ph</strong>armacogenetics</strong>, and <strong>Ph</strong>armacoepidemiology 15<br />

Figure 2 A group <strong>of</strong> individuals with the same diagnosis do not respond homogeneously<br />

to a single drug. Through the elements <strong>of</strong> phamacology, pharmacoepidemiology, and<br />

pharmacogenetics, the aim is to identify patients at risk for toxicity or reduced response to<br />

therapy prior to medication initiation.<br />

Table 1<br />

Examples <strong>of</strong> Genetic Variations and Their Associated <strong>Ph</strong>enotypic Changes<br />

Enzymes Substrate <strong>Ph</strong>enotype<br />

CYP3A4/3A5 D, P, E, T Interindividual variability in<br />

pharmacokinetics<br />

DPD 5-FU Increased toxicity<br />

P-glycoprotein Various anticancer agents Drug resistance<br />

TS 5-FU and antifolates Drug resistance<br />

TPMT 6-MP Increased hematological toxicity<br />

UGT1A1 Irinotecan Increased diarrhea and neutropenia<br />

Abbreviations: D, docetaxel, P, paclitaxel; E, etoposide, T, tamoxifen; 6-MP, mercaptopurine, TPMT,<br />

thiopurine methyltransferases, UGT1A1, UDP-glucuronosyltrasferase 1A1, DPD, dihydropyrimidine<br />

dehydrogenase.<br />

it can effect the transcription <strong>of</strong> the gene. When the polymorphism occurs in the<br />

exon (and intron) region or 3 0 UTR <strong>of</strong> the gene, it can affect translation and RNA<br />

stability, respectively, resulting in either reduced or enhanced activity <strong>of</strong> the<br />

encoded protein (3).<br />

Two approaches have been used to identify clinically relevant polymorphisms.<br />

In the reverse genetics approach, a drug is first given to a group<br />

individuals and the variability in response to and disposition <strong>of</strong> the drug is<br />

observed. <strong>Ph</strong>enotypes <strong>of</strong> clinical importance are prioritized and then experiments


16 Dawood<br />

are conducted to determine if a genetic component is present to explain the<br />

observed variability. Finally, the biochemical and molecular nature <strong>of</strong> the variability<br />

is determined through extensive resequencing techniques that will link the<br />

polymorphisms with the observed phenotypes. The second and more cost effective<br />

method is the forward genetics approach. This method involves first determining<br />

the presence <strong>of</strong> genetic variability at specific loci in a group <strong>of</strong> individuals. A<br />

hypothesis is then formulated regarding these polymorphisms causing phenotypic<br />

variability. Finally, an experiment is performed to investigate the clinical consequences<br />

<strong>of</strong> the genetic variability. Because <strong>of</strong> the multitude <strong>of</strong> polymorphisms,<br />

the challenge today is to be able to identify polymorphisms that occur in gene<br />

regulatory or coding regions that have clinical relevance. The field <strong>of</strong> oncology<br />

adds another level <strong>of</strong> complexity as issues such as clinically relevant tumor<br />

genome variations not found in the germline and somatic mutations acquired<br />

before or after chemotherapy can affect drug efficacy and toxicity. Moreover,<br />

different polymorphisms <strong>of</strong> certain genes may exhibit different phenotypes among<br />

different malignancies. Polymorphisms <strong>of</strong> glutathione S-transferases (GSTs) are a<br />

common example. GSTs facilitate the conjugation <strong>of</strong> glutathione to xenobiotics,<br />

including anticancer agents (e.g., cyclophosphamide and anthracyclines), environmental<br />

carcinogens, and reactive oxygen products (14). A variety <strong>of</strong> functionally<br />

important polymorphisms <strong>of</strong> GSTs have been described. The presence <strong>of</strong><br />

the GSTT1 polymorphism has been shown to predispose children treated intensively<br />

for acute myelogenous leukemia to greater toxicity and death during<br />

remission (15), while women with breast cancer treated with cyclophosphamide<br />

carrying the GSTP1 polymorphism are more likely to survive (16). This example<br />

illustrates how the phenotypic effect <strong>of</strong> polymorphisms depends on the underlying<br />

tumor as well as the intensity <strong>of</strong> regimen used to treat the malignancy. However,<br />

as we begin to unravel and accurately identify the mechanism <strong>of</strong> action <strong>of</strong> many<br />

anticancer drugs, polymorphisms in candidate genes likely to influence drug<br />

efficacy and associated adverse events can be identified. Indeed the field <strong>of</strong><br />

pharmacogenetics has transitioned from one that determined the genetic basis <strong>of</strong><br />

observed inter patient phenotypic variability to one that determines the phenotypic<br />

consequence <strong>of</strong> an observed genotypic variability.<br />

Linking the information obtained from the pharmacology and pharmacoepidemiology<br />

<strong>of</strong> a drug and the pharmacogenetics that describes the genetic<br />

makeup <strong>of</strong> the individual being treated with the drug is thus vitally important in<br />

developing an optimal approach to individualized therapy. The use <strong>of</strong> isoniazid<br />

in the treatment <strong>of</strong> tuberculosis is a good example. Epidemiological evidence<br />

shows that polymorphisms differ in frequency among ethnic and racial groups.<br />

This phenomenon was shown early on when interindividual and interethnic<br />

variations <strong>of</strong> efficacy and toxicity associated with isoniazid treatment was<br />

observed in a cohort <strong>of</strong> individuals with tuberculosis. The enzyme responsible<br />

for the metabolism <strong>of</strong> isoniazid is N-acetyltransferase, a phase II conjugating<br />

liver enzyme. Early studies <strong>of</strong> isoniazid revealed a bimodal distribution in its<br />

plasma concentration among individuals. High plasma concentrations were


<strong>Ph</strong>armacology, <strong><strong>Ph</strong>armacogenetics</strong>, and <strong>Ph</strong>armacoepidemiology 17<br />

observed among individuals with the slow acetylator phenotype <strong>of</strong>ten resulting<br />

in peripheral nerve damage, while low plasma concentrations <strong>of</strong> isoniazid were<br />

observed among individuals with fast acetylator phenotype and were not affected<br />

(17). Studies have shown large variations in the incidence <strong>of</strong> the slow acetylator<br />

phenotype (inherited as an autosomal recessive trait) among different ethnic<br />

groups (40–70% <strong>of</strong> Caucasians and African Americans, 10–20% <strong>of</strong> Japanese, and<br />

more than 80% <strong>of</strong> Egyptians) (18,19). Amonafide (20), a topoisomerase II<br />

inhibitor, is another substrate <strong>of</strong> N-acetyltransferase whose clinical development<br />

has been hampered by highly variable and unpredictable toxicity caused, at least,<br />

in part, by interindividual differences in N-acetylation with fast acetylators<br />

experiencing greater myelosuppression than slow acetylators. This example<br />

illustrates both the importance <strong>of</strong> pharmacoepidemiological evidence in identifying<br />

and characterizing phenotypic variance to individual drugs and how the<br />

pharmacological effects <strong>of</strong> these drugs depend on pharmacogenetic determinants<br />

that define sensitivity to the drug, and ultimately describes the drug’s behavior.<br />

THE LEUKEMIA STORY: WHAT DID WE LEARN?<br />

<strong>Ph</strong>armacological and pharmacogenetic studies are increasingly being used to<br />

optimize existing drug therapy with the goal <strong>of</strong> reducing toxicity while maximizing<br />

efficacy, leading to improved survival. By utilizing this method<br />

remarkable progress has been observed in the treatment <strong>of</strong> acute lymphoblastic<br />

leukemia (ALL) with cure rates increasing from a mere 10% (four decades ago)<br />

to nearly 80% in children and 40% in adults (21). This has been achieved<br />

through optimization <strong>of</strong> use <strong>of</strong> a variety <strong>of</strong> commonly used antileukemic drugs<br />

such as mercaptopurine and methotrexate.<br />

Thiopurine methyltransferase (TPMT) is an enzyme that converts thiopurine<br />

prodrugs, such as mercaptopurine, into inactive methylated metabolites.<br />

Prior to inactivation, mercaptopurine is converted into a variety <strong>of</strong> active<br />

nucleotide metabolites that are incorporated into DNA and RNA, thereby<br />

inducing an antileukemic effect. Patients with an inherited TPMT deficiency<br />

suffer severe, potentially fatal hematopoietic toxicity when exposed to standard<br />

doses <strong>of</strong> mercaptopurine (and other thiopurine prodrugs such as azathioprine and<br />

thioguanine). More than 95% <strong>of</strong> the clinically relevant TPMT mutations are<br />

accounted for by three nonsynonymous SNPs (TPMT*2, TPMT*3A, and<br />

TPMT*3C) (22). TPMT activity is inherited as an autosomal codominant trait<br />

with approximately 95% <strong>of</strong> the population homozygous for the wild-type allele<br />

TPMT*1 (that have full enzyme activity), 10% heterozygous for the polymorphism<br />

(that have intermediate levels <strong>of</strong> enzyme activity), and 1 in 300<br />

individuals will carry two mutant TPMT alleles (that do not express functional<br />

TPMT) (23). Using a pharmacogenetic test, developed at St. Jude Hospital,<br />

patients are <strong>of</strong>ten classified according to their levels <strong>of</strong> TPMT activity into<br />

normal, intermediate, and deficient. This test has a concordance rate <strong>of</strong> 100%<br />

between genotype and phenotype with studies revealing no long- term outcome


18 Dawood<br />

Figure 3 Individualization <strong>of</strong> dosing based on TPMT status does not affect cumulative<br />

incidence <strong>of</strong> relapse. Source: From Ref. 24.<br />

changes related to dosage individualization based on the TPMT status <strong>of</strong> a<br />

patient (24) (Fig. 3). This illustrates how the clinical application <strong>of</strong> pharmacogenetics<br />

can be used to minimize toxicity without compromising efficacy.<br />

Methyltetrahydr<strong>of</strong>olate reductase (MTHFR), a crucial enzyme in the<br />

metabolism <strong>of</strong> folic acid, catalyses the reduction <strong>of</strong> 5,10-methylenetetrahydr<strong>of</strong>olate<br />

to 5-methyltetrahydr<strong>of</strong>olate (the predominant circulatory form <strong>of</strong> folate<br />

and the carbon donor for the demethylation <strong>of</strong> homocysteine to methionine). As<br />

methotrexate increases serum levels <strong>of</strong> homocysteine and the active polyglutamate<br />

metabolites <strong>of</strong> methotrexate inhibit MTHFR function, patients with a<br />

low level <strong>of</strong> MTHFR function are at an increased risk <strong>of</strong> methotrexate-induced<br />

toxicities such as oral mucosities. Two polymorphisms associated with reduced<br />

MTHFR activity have been described (25): (i) 677C > T polymorphism that is<br />

homozygous in 10% <strong>of</strong> the population where it encodes the enzyme with 30% <strong>of</strong><br />

the wild-type enzyme activity and heterozygous in 40% <strong>of</strong> the population where<br />

it encodes an enzyme that has 60% <strong>of</strong> the wild type-enzyme activity and (ii)<br />

1298A > C (less common than 677C > T). Recent findings have also suggested<br />

that MTHFR polymorphisms may protect against the development <strong>of</strong> both adult<br />

and pediatric ALL (26).<br />

5-FLUOROURACIL<br />

5-Fuorouracil (5-FU) is a uracil analog widely used to treat solid tumors such as<br />

breast and colorectal cancer (27). 5-FU is inactivated by dihydropyrimidine<br />

dehydrogenase (DPD), an enzyme that exhibits up to 20-fold variation in activity<br />

among individuals (28). Mutations that result in low DPD activity predispose to<br />

potentially life-threatening gastrointestinal, hematopoietic, and neurological<br />

toxicities (29). Up to 20 mutations (30) have been identified that cause inactivation<br />

<strong>of</strong> DPD. In the general population it has been estimated that 3% to 5% <strong>of</strong><br />

individuals are heterozygous and 0.1% are homozygous carriers for the mutation<br />

that inactivates DPD, with DPYD*2A being the most inactivating allele (31).<br />

Although polymorphisms associated with DPD inactivation have been identified,<br />

a number <strong>of</strong> patients still exhibit severe 5-FU toxicity with no detectable


<strong>Ph</strong>armacology, <strong><strong>Ph</strong>armacogenetics</strong>, and <strong>Ph</strong>armacoepidemiology 19<br />

mutations in the coding region <strong>of</strong> the DPD gene explaining the difficulty in<br />

implementation <strong>of</strong> DPD pharmacogenetics in the prospective identification <strong>of</strong><br />

high-risk patients for severe 5-FU toxicity.<br />

One <strong>of</strong> the primary mechanisms <strong>of</strong> action <strong>of</strong> 5-FU is inhibition, via fluorodeoxyuridine<br />

monophosphate (an active metabolite <strong>of</strong> 5-FU), <strong>of</strong> thymidylate<br />

synthase (TS), a critical enzyme required for de novo thymidylate synthesis,<br />

which is required for DNA synthesis and repair (32). Polymorphisms in the gene<br />

encoding TS are one <strong>of</strong> the few examples <strong>of</strong> a polymorphism that is associated<br />

with the genetic makeup <strong>of</strong> the tumor with clinical studies showing an inverse<br />

relationship between TS levels and antitumor response (33). The TSER polymorphism<br />

<strong>of</strong> TS is characterized by tandem repeats in the TS promoter enhancer<br />

region <strong>of</strong> the gene. Higher levels <strong>of</strong> TS expression and enzyme activity have been<br />

observed with increasing copies <strong>of</strong> tandem repeats, which result in lower tumor<br />

response rates to 5-FU (34). In a study by Marsh et al (35), it was observed that in<br />

a cohort <strong>of</strong> patients with metastatic colorectal tumors receiving 5-FU-based chemotherapy<br />

2 tandem repeats was nearly twice as common in responders than in<br />

non-responders with longer median survival observed in patients with 2 tandem<br />

repeats compared to those with 3 tandem repeats (16 months vs. 12 months) (35).<br />

Polymorphisms <strong>of</strong> TS that result in 5-FU variability <strong>of</strong> its pharmacodynamic<br />

and pharmacokinetic properties illustrate the important fact that cancer<br />

being a polygenic disease will need a polygenic solution to guide therapy. To<br />

accurately select patients that are most likely to tolerate and respond to 5-FU<br />

therapy, one needs to use combined genotyping <strong>of</strong> DPYD and TSER functional<br />

variants as well as taking into account nongenetic factors.<br />

TAXANES<br />

Paclitaxel and Docetaxel are the two most commonly used taxanes in the<br />

treatment <strong>of</strong> both early and advanced stage breast cancer (36). They exert their<br />

cytotoxic effects by binding to B-tubulin that subsequently results in the stabilization<br />

<strong>of</strong> microtubules and disruption <strong>of</strong> mitotic spindle formation during cell<br />

division ultimately resulting in cell death. Genes involved in drug transport,<br />

metabolism, and drug target all play an important role in overall taxane efficacy,<br />

with their polymorphisms partly explaining the variability observed in toxicity<br />

and response to these drugs.<br />

The multidrug-resistant transporter, P-glycoprotein, encoded by ABCB1<br />

(MDR1) and expressed in multiple cell types, has been shown to mediate both<br />

paclitaxel and docetaxel influx (37,38). Three polymorphisms in ABCB1 have<br />

been commonly studied (3435C>T, 1236C>Y, and 2677G>T/A) (39), with<br />

most studies showing inconsistent results when trying to establish an association<br />

between these polymorphisms and drug efficacy/toxicity. Studies focusing on<br />

looking at ABCB1 polymorphism in combination may yield more useful clinical<br />

information. Some <strong>of</strong> the recent studies have reported an association <strong>of</strong> the<br />

various ABCB1 polymorphisms with tumor response to paclitaxel (ABCB1


20 Dawood<br />

2677G>T/A) (40), risk <strong>of</strong> neutropenia with docetaxel (ABCB1 3435C>T) (41),<br />

and risk <strong>of</strong> neuropathy with paclitaxel (ABCB1 3435C>T) (42). Oxidative<br />

taxane metabolism occurs via the CYP450 pathway, with CYP2C8 and CYP3A4<br />

primarily responsible for paclitaxel metabolism while docetaxel is mainly<br />

metabolized by CYP3A4 (43,44). The functional significance <strong>of</strong> the polymorphisms<br />

<strong>of</strong> these enzymes has not been established with associations seen in<br />

in vivo studies (45) not observed in studies <strong>of</strong> cancer patients (46).<br />

TAMOXIFEN<br />

Tamoxifen is a selective estrogen receptor modulator (SERM) that is widely<br />

used in the treatment and prevention <strong>of</strong> breast cancer. When given in the<br />

adjuvant setting, it reduces the annual risk <strong>of</strong> recurrence <strong>of</strong> breast cancer by half<br />

and the breast cancer mortality rate by one-third (47). The pharmacology and<br />

metabolism <strong>of</strong> tamoxifen is complex and undergoes extensive phase I and II<br />

metabolism with the production <strong>of</strong> endoxifen and to a lesser extent 4-hydroxytamoxifen<br />

(48). These metabolites play an important role in the anticancer<br />

effect <strong>of</strong> tamoxifen through their high-affinity binding to estrogen receptors and<br />

suppression <strong>of</strong> estradiol-stimulated cell proliferation (49).<br />

The clinical effects <strong>of</strong> tamoxifen are highly variable (50) due to the variability<br />

observed in the plasma levels <strong>of</strong> the more potent metabolites <strong>of</strong> tamoxifen.<br />

Tamoxifen is converted to endoxifen (4-hydroxy-N-desmethyltamoxifen) primarily<br />

by CYP2D6-mediated oxidation <strong>of</strong> N-desmethyl tamoxifen, while the<br />

conversion <strong>of</strong> tamoxifen to 4-hydroxy tamoxifen is catalyzed by multiple enzymes<br />

(51). Plasma concentrations <strong>of</strong> endoxifen is thus sensitive to variants <strong>of</strong> the<br />

CYP2D6 genotype with the most common variant making the enzyme inactive<br />

(common allele associated with CYP2D6 poor metabolizer phenotype is<br />

CYP2D6*4). One prospective trial revealed that women treated with tamoxifen<br />

who either carried the genetic variants associated with low or absent CYP2D6<br />

activity or were treated concomitantly with drugs that inhibited the activity <strong>of</strong><br />

CYP2D6 had lower levels <strong>of</strong> endoxifen (52). Subsequently, using tissue sample<br />

from a cohort <strong>of</strong> early stage breast cancer patients treated with tamoxifen, it was<br />

shown that women with a homozygous CYP2D6 phenotype tended to have a<br />

higher risk <strong>of</strong> disease relapse and a lower incidence <strong>of</strong> hot flashes (53).<br />

Polymorphisms associated with the activity <strong>of</strong> the metabolites <strong>of</strong> tamoxifen<br />

teach us two important lessons. First, the interindividual variability in the<br />

response to tamoxifen is explained both by genetic variation in CYP2D6 and<br />

the coadministration <strong>of</strong> drugs that inhibit the activity <strong>of</strong> CYP2D6. Second, the<br />

toxicity associated with drug can help identify responders.<br />

CONCLUSION<br />

At present time, dosages <strong>of</strong> anticancer drugs are administered based on the<br />

assumption <strong>of</strong> homogeneity among individuals and are rarely based on individual<br />

characteristics. As a result we encounter a situation where by a large


<strong>Ph</strong>armacology, <strong><strong>Ph</strong>armacogenetics</strong>, and <strong>Ph</strong>armacoepidemiology 21<br />

number <strong>of</strong> individuals experience unwanted adverse side effects after treatment<br />

without attaining maximum drug efficacy. Such adverse drug reactions impose a<br />

huge burden not only on the patient but also on the health care system. A better<br />

understanding <strong>of</strong> the pharmacoepidemiological, pharmacological, and pharmacogenetic<br />

determinants <strong>of</strong> a chemotherapeutic agent is needed to optimize drug<br />

dosing regimens to obtain maximal efficacy with minimal toxicity. Advancements<br />

in the field <strong>of</strong> pharmacogenetics hold promise in helping to identify which<br />

polymorphisms are likely to be important in clinical oncology thereby enabling<br />

the development <strong>of</strong> effective agents that will make their way to the clinic in a<br />

timely manner. Realizing the importance <strong>of</strong> each <strong>of</strong> the components <strong>of</strong> individualized<br />

therapy and incorporating each one into the infrastructure <strong>of</strong> clinical<br />

trials will allow us to further explore and define the relationships between different<br />

polymorphisms and observed phenotypes. Considering the fact that anticancer<br />

drugs have narrow a therapeutic range oncology is a field that will<br />

certainly benefit from the application <strong>of</strong> pharmacogenetic and pharmacological<br />

principles to dosage individualization. The idea <strong>of</strong> the “one model fits all” should<br />

be accepted as grossly flawed and, as we progress in the field <strong>of</strong> cancer therapeutics,<br />

will be replaced in the future with the application <strong>of</strong> “personalized<br />

cancer chemotherapy” into routine clinical practice.<br />

REFERENCES<br />

1. Jemal A, Siegal R, Ward E, et al. <strong>Cancer</strong> statistics 2007. CA <strong>Cancer</strong> J Clin 2007;<br />

57:43–66.<br />

2. Bordet R, Gautier S, Le Louet H, et al. Analysis <strong>of</strong> the direct cost <strong>of</strong> adverse drug<br />

reactions in hospitalized patients. Eur J Clin <strong>Ph</strong>armacol 2001; 56:935–941.<br />

3. Evans WE, Reilling MV. <strong>Ph</strong>armacogenomics: Translating functional genomics into<br />

rational therapeutics. Science 1999; 286:487–491.<br />

4. Jeffrey S, Ross MD, David P, et al. <strong>Ph</strong>armacogenomics. Adv Anat Pathol 2004;<br />

11:211–220.<br />

5. Maitland-van der Zee AH, de Boer A, Leufkens HG. The interface between pharmacoepidemiology<br />

and pharmacogenetics. Eur J <strong>Ph</strong>armacol 2000; 410:121–130.<br />

6. Gaplin AJ, Evans WE. Therapeutic drug monitoring in cancer management. Clin<br />

Chem 1993; 39:2419–2430.<br />

7. Calvert A, Newell D, Gumbrell L, et al. Carboplatin dosage: prospective evaluation<br />

<strong>of</strong> a simple formula based on renal function. J Clin Oncol 1989; 7:1748.<br />

8. Last JM. A Dictionary <strong>of</strong> Epidemiology. Oxford: Oxford University Press, 2000.<br />

9. Sander C. Genomic medicine and future <strong>of</strong> health care. Science 2000; 287:1977–<br />

1978.<br />

10. Vesell E.S. Therapeutic lessons from pharmacogenetics. Ann Intern Med 1997;<br />

126:653–655.<br />

11. Evans WE, Relling MV. <strong>Ph</strong>armacogenomics: Translating functional genomics into<br />

rational therapeutics. Science 1999; 286:487–491.<br />

12. Relling MV, Dervieux T. <strong><strong>Ph</strong>armacogenetics</strong> and cancer therapy. Nat Rev <strong>Cancer</strong><br />

2001; 1:99–108.


22 Dawood<br />

13. Sachidanadam R, Weissman D, Schmidt SC, et al. A map <strong>of</strong> human genome<br />

sequence variation containing 1.42 million single nucleotide polymorphisms. Nature<br />

2001; 409:928–933.<br />

14. Ketterer B. Protective role <strong>of</strong> glutathione and glutathione transferases in mutagenesis<br />

and carcinogenesis. Mutat Res 1988; 202:343–361.<br />

15. Davies SM, Robison LL, Buckley JD, et al. Glutathione S-transferase polymorphisms<br />

and outcome <strong>of</strong> chemotherapy in childhood acute myeloid leukemia.<br />

J Clin Oncol 2001; 19:1279–1287.<br />

16. Sweeney C, McClure GY, Fares MY, et al. Association between survival after<br />

treatment for breast cancer and glutathione S-transferase P1 Ile105Val polymorphism.<br />

<strong>Cancer</strong> Res 2000; 60:5621–5624.<br />

17. Evans DA, Manley KA, Mckusch VA, et al. Genetic control <strong>of</strong> isoniazid metabolism<br />

in man. Br Med J 1960; 2:485–491.<br />

18. Knight RA, Selin MJ, Harris HW, et al. Genetic factors influencing isoniazid blood<br />

levels in humans. Trans Conf Chemother Tuberc 1959; 8:52–56.<br />

19. Evans Da. Survey <strong>of</strong> human acetylator polymorphism in spontaneous disorders.<br />

J Med Genet 1984; 21:243–253.<br />

20. Ratain MJ, Mick R, Berezin F, et al. Paradoxical relationship between acetylator<br />

phenotype and amonafide toxicity. Clin <strong>Ph</strong>armacol Ther 1991; 50:573–579.<br />

21. Pui C-H, Evans WE. Acute lymphoblastic leukemia. N Engl J Med 1998; 339: 605–615.<br />

22. McLeod HL, Coulthard S, Thomas, AE et al. Analysis <strong>of</strong> thiopurine methyltransferase<br />

variant alleles in childhood acute lymphoblastic leukemia. Br J Hematol<br />

1999; 105:696–700.<br />

23. Krynetski EY, Tai HL, Yates CR, et al. Genetic polymorphisms <strong>of</strong> thiopurine<br />

s-methyltransferase: clinical importance and molecular mechanisms. <strong><strong>Ph</strong>armacogenetics</strong><br />

1996; 6:279–290.<br />

24. Relling MV, Pui C-H, Cheng C, et al. Thiopurine methyltransferase in acute lymphoblastic<br />

leukemia. Blood 2006; 107:843–844.<br />

25. Molloy AM, Daly S, Mills JL, et al. Thermolabile variant <strong>of</strong> 5,10-methyenetetrahydr<strong>of</strong>olate<br />

reductase associated with low red-cell folates: implications for folate<br />

intake recommendations. Lancet 1997; 349:1591–1593.<br />

26. Franco RF, Simoes BP, Tone LG, et al. The methylenetetrahydr<strong>of</strong>olate reductase<br />

C677T gene polymorphism decreases the risk <strong>of</strong> childhood acute lymphocytic leukemia.<br />

Br J Haematol 2001; 115:616–618.<br />

27. Heidelberger C, Chaudhuri NK, Danneberg P, et al. Fluorinated pyrimidines, a new<br />

class <strong>of</strong> tumor – inhibitory compounds. Nature 1957; 179:663–666.<br />

28. Etienne MC, Lagrange JL, Dassonville O, et al. Population study <strong>of</strong> dihydropyrimidine<br />

dehydrogenase in cancer patients. J Clin Oncol 1994; 12:2248–2253.<br />

29. Raida M, Schwabe W, Hausler P, et al. Prevalence <strong>of</strong> a common point mutation in<br />

the dihydropyrimidine dehydrogenase (DPD) gene within the 5’-splice donor site <strong>of</strong><br />

intron 14 in patients with severe 5-fluorouracil (5-FU)-related toxicity compared<br />

with controls. Clin <strong>Cancer</strong> Res 2001; 7:2832–2839.<br />

30. McLeod HL, Collie-Duguid ES, Vreken P, et al. Nomenclature for human DPYD<br />

alleles. <strong><strong>Ph</strong>armacogenetics</strong> 1998; 8:455–459.<br />

31. Lu A, Zhang R, Diasio RB. Dihydropyrimadine dehydrogenase activity in human<br />

peripheral blood mononuclear cells and liver: population characteristics, newly<br />

identified deficient patients, and clinical implication in 5-fluorouracil chemotherapy.<br />

<strong>Cancer</strong> Res 1993; 53:5433–5438.


<strong>Ph</strong>armacology, <strong><strong>Ph</strong>armacogenetics</strong>, and <strong>Ph</strong>armacoepidemiology 23<br />

32. Grem JL. 5-Fluorouracil: forty-plus and still ticking. A review <strong>of</strong> its preclinical and<br />

clinical development. Invest New Drugs 2000; 18:299–313.<br />

33. Leichman CG, Lenz HJ, Leichman L, et al. Quantitation <strong>of</strong> intratumoral thymidylate<br />

synthase expression predicts for disseminated colorectal cancer response and resistance<br />

to protracted-infusion fluorouracil and weekly leucovorin. J Clin Oncol 1997;<br />

15:3223–3229.<br />

34. Horie N, Aiba H, Oguro K, et al. Functional analysis and DNA polymorphism <strong>of</strong> the<br />

tandemly repeated sequences in the 5’-terminal regulatory region <strong>of</strong> the human gene<br />

for thymidylate synthase. Cell Struct Funct 1995; 20:191–197.<br />

35. Marsh S, Mckay JA, Cassidy J, et al. Polymorphism in the thymidylate synthase<br />

promoter enhancer region in colorectal cancer. Int J Oncol 2001; 19:383–386.<br />

36. Nabholtz JM, Gligorov J. The role <strong>of</strong> taxanes in the treatment <strong>of</strong> breast cancer.<br />

Expert Opin <strong>Ph</strong>armacother 2005; 6:1073–1094.<br />

37. Shirakawa K, Takara K, Tanigawara Y, et al. Interaction <strong>of</strong> docetaxel (‘Taxotere’)<br />

with human P-glycoprotein. Jpn J <strong>Cancer</strong> Res 1999; 90:1380–1386.<br />

38. Sparreboom A, Van Asperen J, Mayer U, et al. Limited oral bioavailability and<br />

active epithelial excretion <strong>of</strong> paclitaxel (Taxol) caused by P-glycoprotein in the<br />

intestine. Proc Natl Acad Sci U S A 1997; 94:2031–2035.<br />

39. Lepper ER, Nooter K, Verweij J, et al. Mechanisms <strong>of</strong> resistance to anticancer drugs:<br />

the role <strong>of</strong> the polymorphic ABC transporters ABCB1 and ABCG2. <strong>Ph</strong>armacogenomics<br />

2005; 6:115–138.<br />

40. Green H, Soderkvist P, Rosenberg P, et al. mdr-1 Single nucleotide polymorphisms<br />

in ovarian cancer tissue: G2677T/A correlates with response to paclitaxel chemotherapy.<br />

Clin <strong>Cancer</strong> Res 2006; 12:854–859.<br />

41. Tran A, Jullien V, Alexandre J, et al. <strong>Ph</strong>armacokinetics and toxicity <strong>of</strong> docetaxel:<br />

role <strong>of</strong> CYP3A, MDR1, and GST polymorphisms. Clin <strong>Ph</strong>armacol Ther 2006;<br />

79:570–580.<br />

42. Sissung TM, Mross K, Steinberg SM, et al. Association <strong>of</strong> ABCB1 genotypes with<br />

paclitaxel-mediated peripheral neuropathy and neutropenia. Eur J <strong>Cancer</strong> 2006;<br />

42:2893–2896.<br />

43. Taniguchi R, Kumai T, Matsumoto N, et al. Utilization <strong>of</strong> human liver microsomes<br />

to explain individual differences in paclitaxel metabolism by CYP2C8 and CYP3A4.<br />

J <strong>Ph</strong>armacol Sci 2005; 97:83–90.<br />

44. Leyseng-Williamson KA, Fenton C. Docetaxel a review <strong>of</strong> its use in metastatic<br />

breast cancer. Drugs 2005; 65:2513–2531.<br />

45. Bahadur N, Leathart JB, Mutch E, et al. CYP2C8 polymorphisms in Caucasions and<br />

their relationship with paclitaxel 6 á- hydroxylase activity in human liver micorsomes.<br />

Biochem <strong>Ph</strong>armacol 2002; 64:1579–1589.<br />

46. Henningsson A, Marsh S, Loos WJ, et al. Association <strong>of</strong> CYP2C8, CYP3A4,<br />

CYP3A5, and ABCB1 polymorphisms with pharmacokinetics <strong>of</strong> paclitaxel. Clin<br />

<strong>Cancer</strong> Res 2005; 11:8097–8104.<br />

47. Effects <strong>of</strong> chemotherapy and hormonal therapy for early breast cancer on recurrence and<br />

15-year survival: An overview <strong>of</strong> the randomized trials. Lancet. 2005; 365:1687–1717.<br />

48. Lonning PE, Lien EA, Lundgren S, et al. Clinical pharmacokinetics <strong>of</strong> endocrine<br />

agents used in advanced breast cancer. Clin <strong>Ph</strong>armacokinet 1992; 22:327–358.<br />

49. Stearns V, Johnson MD, Rae JM, et al. Active tamoxifen metabolite plasma concentrations<br />

after co administration <strong>of</strong> tamoxifen and the selective serotonin reuptake<br />

inhibitor paroxetine. J Natl <strong>Cancer</strong> Inst 2003; 95:1758–1764.


24 Dawood<br />

50. Osborne CK. Tamoxifen in the treatment <strong>of</strong> breast cancer. N Engl J Med 1998;<br />

339:1609–18.<br />

51. Desta Z, Ward BA, Soukhova NV, et al. Comprehensive evaluation <strong>of</strong> tamoxifen<br />

sequential biotransformation by the human cytochrome P450 system in vitro: Prominent<br />

roles for CYP3A and CYP2D6. J <strong>Ph</strong>armacol Exp Ther 2004; 310:1062–1075.<br />

52. Jin Y, Desta Z, Stearns V, et al. CYP2D6 genotype, antidepressant use, and tamoxifen<br />

metabolism during adjuvant breast cancer treatment. J Natl <strong>Cancer</strong> Inst 2005; 97:30–39.<br />

53. Goetz MP, Rae JM, Suman VJ, et al. <strong><strong>Ph</strong>armacogenetics</strong> <strong>of</strong> tamoxifen biotransformation<br />

is associated with clinical outcomes <strong>of</strong> efficacy and hot flashes. J Clin<br />

Oncol 2005; 23:9312–9318.


3<br />

Role <strong>of</strong> Genetic Variability in <strong>Breast</strong><br />

<strong>Cancer</strong> Treatment Outcomes<br />

Kandace L. Amend<br />

Department <strong>of</strong> Oncological Sciences, Mount Sinai School <strong>of</strong> Medicine,<br />

New York, New York, U.S.A.<br />

Ji-Yeob Choi and Christine B. Ambrosone<br />

Department <strong>of</strong> <strong>Cancer</strong> Prevention and Control, Roswell Park<br />

<strong>Cancer</strong> Institute, Buffalo, New York, U.S.A.<br />

INTRODUCTION<br />

Variability in toxicity and response to the same dose <strong>of</strong> medication may occur<br />

among cancer patients and contribute to the unpredictability <strong>of</strong> clinical outcomes.<br />

In addition to the potential importance <strong>of</strong> clinical factors in determining<br />

drug effects (1), inherited differences in drug metabolism are likely to have a<br />

pr<strong>of</strong>ound effect on treatment outcomes (2). <strong><strong>Ph</strong>armacogenetics</strong>, the study <strong>of</strong> the<br />

contribution <strong>of</strong> genetic variation to these interindividual differences, may allow<br />

for the individualization <strong>of</strong> therapy and improved prediction <strong>of</strong> pharmacodynamic<br />

events, including host toxicity and treatment efficacy. To date, the role <strong>of</strong><br />

pharmacogenetics in breast cancer therapy has not been clearly defined. Herein<br />

we discuss the drugs most commonly used to treat breast cancer and describe<br />

their metabolic pathways, as well as provide epidemiologic support for the role<br />

<strong>of</strong> pharmacogenetics in breast cancer outcomes.<br />

25


26 Amend et al.<br />

BREAST CANCER THERAPEUTICS<br />

The most frequently used agents to treat breast cancer include cyclophosphamide<br />

(CP) in combination with doxorubicin [Adriamycin 1 (A)] or with methotrexate<br />

(MTX) and 5-flourouracil (5FU) (3). The taxanes paclitaxel (Taxol 1 ) and<br />

docetaxel (Taxotere 1 ) have an important role in the treatment <strong>of</strong> early-stage<br />

breast cancer as well as metastatic disease. In patients with estrogen receptor<br />

(ER)-positive tumors, tamoxifen (TAM) reduces the risk <strong>of</strong> recurrence by 31%<br />

and breast cancer morality by one-third (4), and randomized trials have indicated<br />

that treatment with third-generation aromatase inhibitors (AIs) is now an<br />

important option for the adjuvant treatment <strong>of</strong> hormone receptor–positive breast<br />

cancer (5,6).<br />

Multiple enzymatic pathways are involved in the metabolism and detoxification<br />

<strong>of</strong> chemotherapeutic agents, and common genetic polymorphisms in<br />

drug metabolizing enzymes are likely to contribute to interindividual variability<br />

in toxicity and cancer recurrence (7).<br />

Cyclophosphamide<br />

Cyclophosphamide requires multistep activation before it can function as an<br />

antitumor alkylating agent. The drug is first oxidized in the liver to the active<br />

4-hydroxy metabolite mainly by CYP2B6, although other enzymes, including<br />

CYP2A6, CYP3A4, CYP2C8, CYPC29, and CYP2C19, are involved. Drug<br />

intermediates are transported in blood and decompose spontaneously to phosphoramide<br />

mustard and acrolein, both <strong>of</strong> which can react with nucleophilic<br />

groups, forming DNA cross-links and inhibition <strong>of</strong> protein synthesis (8). Inactivating<br />

glutathionyl conjugation reactions <strong>of</strong> C intermediates are catalyzed by<br />

glutathione S-transferases (GSTs) GSTA1 and GSTP1 (9). Variability in the<br />

activity <strong>of</strong> these activating and detoxifying enzymes is likely to affect ultimate<br />

tumor tissue dose and efficacy <strong>of</strong> treatment.<br />

The human cytochrome P450 CYP2B6 gene is involved in the metabolic<br />

activation <strong>of</strong> a number <strong>of</strong> clinically important chemotherapeutic drugs for breast<br />

cancer, including CP (10,11) as well as the antioestrogen TAM (12). Several<br />

single nucleotide polymorphisms (SNPs) in CYP2B6 have been described, some<br />

having functional significance (13) and affecting the pharmacokinetic parameters<br />

<strong>of</strong> CP. In cell culture, enzymatic activities in microsomes from COS-1 cells<br />

expressing a K262R amino acid substitution (CYP2B6*4, CYP2B6*6, and<br />

CYP2B6*7) showed increased values for V max and V max /K m compared with that<br />

<strong>of</strong> the wild type (CYP2B6*1) (14), while a second study reported that the amino<br />

acid substitution Q172H resulting from the CYP2B6*6 allele enhanced<br />

7-ethoxycoumarin O-deethylase activity (15). A study by Lang et al. (2001)<br />

identified a total <strong>of</strong> nine SNPs by sequencing DNA from 35 subjects and<br />

reported extensive variability in the expression and activity <strong>of</strong> CYP2B6, where<br />

CYP2B6 expression was significantly reduced in carriers <strong>of</strong> the R487T, Q172H,


Role <strong>of</strong> Genetic Variability in <strong>Breast</strong> <strong>Cancer</strong> Treatment Outcomes 27<br />

and K262R polymorphisms (16). Ethnic variation in allelic variants and CYP2B6<br />

expression has been also been observed (17,18).<br />

CYP3A4 also plays a large role in the metabolism <strong>of</strong> CP, and the CYP3A4<br />

gene has a number <strong>of</strong> known variant alleles, including some having functional<br />

significance (19). A single base substitution in the 5 0 promoter region <strong>of</strong> the gene<br />

(20), present in 9% <strong>of</strong> European-Americans and 53% <strong>of</strong> African-Americans (21),<br />

with the CYP3A4*1B allele associated with expression levels is 1.4-fold higher<br />

than that in common alleles (22). Using enzyme kinetic analyses to evaluate<br />

intersample variation in activation <strong>of</strong> CP by CYP3A and 2B in human liver, both<br />

Chang et al. (1993) and Roy et al. (1999) found considerable interindividual<br />

differences in CYP3A and CYP2B levels and variation over a wide range in CP<br />

hydroxylation (10,23). Variation in CYP3A enzyme levels and expression may<br />

be related, in part, to genetic differences or to other chemotherapeutic and<br />

natural agents that have been shown to increase CYP3A4 expression (24).<br />

CYP3A5 accounts for at least 50% <strong>of</strong> the total CYP3A content in most<br />

Caucasians and African-Americans and is thus considered an important contributor<br />

to interindividual differences in CYP3A-dependent drug clearance (25).<br />

Although CYP3A5 is polymorphically expressed at high levels in a minority <strong>of</strong><br />

Caucasians, the expression in human liver is more frequent in African-<br />

Americans (60%) than in Caucasians (33%) (25). Reduced expression in some<br />

individuals may be due to a common polymorphism in intron 3 <strong>of</strong> CYP3A5<br />

(A6986G), which generates a cryptic splice site (CYP3A*6) resulting in a<br />

truncated amino acid (25,26).<br />

Genetic polymorphisms have been described in several <strong>of</strong> the CYP2C<br />

family members, including CYP2C8, CYP2C29, CYP2C18, and CYP2C19,<br />

although the catalytic contribution <strong>of</strong> CYP2C8 and CYP2C18 to activation <strong>of</strong> CP<br />

is likely minor compared with the contributions <strong>of</strong> CYP2C9 and CYP2C19. In<br />

vitro, CYP2C19 was found to bioactivate CP in human liver microsomes (27),<br />

and while both CYP2C9 and CYP2C19 were found to possess CP 4-hydroxylase<br />

activity in vitro, CYP2C19 exhibited higher CP 4-hydroxylase activity following<br />

only CYP2B6 (23). Although known variants in CYP2C9 and CYP2C19 exist,<br />

little is known regarding their contribution to pharmacokinetic variability in<br />

relation to breast cancer prognosis. In a study by Xie et al., CYP2C9*2,<br />

CYP2C9*3, CYP2C19*2, and CYP2C19*3 were not significantly correlated with<br />

the metabolism <strong>of</strong> CP in 19 human liver specimens (11).<br />

<strong>Ph</strong>ase II drug metabolism is catalyzed by GSTs, which consist <strong>of</strong> several<br />

isoenzymes, including mu (GSTM1), pi (GSTP1), alpha (GSTA1), theta<br />

(GSTT1), and zeta. The GSTs detoxify many chemotherapeutic drugs, including<br />

CP, are induced by oxidative stress, and detoxify the reactive oxygen species<br />

(ROS) that are produced by chemotherapeutic drugs and radiation therapy.<br />

GSTA1, the enzyme most active in glutathione conjugation reactions with CP<br />

intermediates, has a polymorphism in the 5 0 promoter region <strong>of</strong> the gene (28).<br />

Coles et al. (2001) identified substitutions designated as GSTA1*A ( 567, 69,<br />

and 52) and GSTA1*B ( 631, 567, and 69) (29), with the GSTA1*B variant


28 Amend et al.<br />

showing decreased hepatic expression and increased GSTA2 expression compared<br />

with GSTA1*A homozygotes. Given that GSTA1 increases the conjugation<br />

<strong>of</strong> CP intermediates (9), these differences in hepatic expression may affect the<br />

detoxification <strong>of</strong> CP metabolites and, ultimately, the outcomes <strong>of</strong> treatment and<br />

survival.<br />

For the GSTP1 gene, two variant alleles, GSTP1*B and GSTP1*C, have<br />

been detected (30). A single amino acid change from isoleucine to valine occurs<br />

at codon 105 (Ile105Val) <strong>of</strong> the GSTP1 gene, with genotype frequencies in<br />

Caucasians 51% homozygous for Ile/Ile, 43% heterozygous for Ile/Val, and 6%<br />

homozygous for the variant allele Val/Val (31). The allelic variant 105Val<br />

exhibited lower glutathione conjugation <strong>of</strong> the alkylating agent thiopeta compared<br />

with the isoleucine allele (32). GSTP1 variants may also differ in activity<br />

toward CP, structurally similar to thiotepa, and patients with GSTP1 allelic<br />

variants may have reduced activity in removal <strong>of</strong> active CP.<br />

Anthracyclines<br />

The commonly used anthracyclines include doxorubicin, daunorubicin, and<br />

epirubicin, although doxorubicin and epirubicin are used more frequently for the<br />

treatment <strong>of</strong> breast cancer. Several different mechanisms have been proposed for<br />

anthracycline tumor activity, including free radical generation, DNA binding and<br />

alkylation, DNA cross-linking, interference with DNA unwinding or strand<br />

separation, direct membrane effects, inhibition <strong>of</strong> topoisomerase II, and induction<br />

<strong>of</strong> apoptosis (33). Doxorubicin, like other anthracyclines, results in the<br />

formation <strong>of</strong> quinone-mediated free radicals, which have the capacity to cause<br />

oxidative damage and cytotoxicity, resulting in significant cardiotoxicity as a<br />

result <strong>of</strong> the production <strong>of</strong> ROS (34). Cellular oxidoreductases reduce Adriamycin<br />

to a semiquinone radical that is subsequently reoxidized by oxygen to a<br />

superoxide anion and the parent quinine (35–37). Superoxide anions can dismutate<br />

to form hydrogen peroxide and/or react with nitric oxide to form peroxynitrite.<br />

Lipid peroxides resulting from doxorubicin can further break down to<br />

yield hydroxyalkenals, which are substrates for glutathione-conjugating isozymes<br />

(38), indicating the potential importance <strong>of</strong> GSTs in mediating therapeutic<br />

outcomes.<br />

Carbonyl reductase (CBR) is a monomeric, NADPH-dependent oxidoreductase<br />

that catalyzes the reduction <strong>of</strong> a variety <strong>of</strong> carbonyl compounds,<br />

including doxorubicin and daunorubicin. In humans, two CBRs have been<br />

identified, CBR1 and CBR3. An SNP has been identified in CBR3 (CBR3<br />

V244M), where the M244 is<strong>of</strong>orm results in higher V max and is 100% more<br />

efficient in catalyzing the reduction <strong>of</strong> substrate (39). To our knowledge, CBRs<br />

have not been investigated in relation to breast cancer outcomes following<br />

treatment with anthracyclines.<br />

In addition to its mechanistic role in cytotoxicity <strong>of</strong> anthracyclines, it is well<br />

documented that oxidative stress plays a role in the cell-killing abilities <strong>of</strong> many


Role <strong>of</strong> Genetic Variability in <strong>Breast</strong> <strong>Cancer</strong> Treatment Outcomes 29<br />

other chemotherapy agents, as well as in radiation therapy. Specifically for breast<br />

cancer, CP, doxorubicin (Adriamycin), and paclitaxel and radiation therapy generate<br />

ROS and induce apoptosis, which may contribute to treatment efficacy and<br />

increased survival. ROS can cause mitochondrial permeability, and once the<br />

mitochondrial membrane barrier function is lost, a number <strong>of</strong> other factors contribute<br />

to cell death. While ROS, among other factors, induce or facilitate mitochondrial<br />

permeability, glutathione and antioxidant enzymes such as manganese<br />

superoxide dismutase (MnSOD), catalase (CAT), and glutathione peroxidase<br />

(GPX1) inhibit it (40). In the mitochondrion, MnSOD catalyzes the dismutation <strong>of</strong><br />

two superoxide radicals, producing H 2 O 2 and oxygen. If H 2 O 2 is not neutralized, it<br />

may contribute to further generation <strong>of</strong> ROS by a reaction catalyzed by myeloperoxidase<br />

(MPO). Thus, levels <strong>of</strong> MnSOD, CAT, GPX1, and MPO are important<br />

in determining the amount <strong>of</strong> ROS or detoxification <strong>of</strong> H 2 O 2 . It is plausible that<br />

interindividual variability resulting from polymorphisms in these genes may<br />

influence the efficacy <strong>of</strong> cancer treatment and survival.<br />

Two genetic polymorphisms appear to affect mitochondrial levels <strong>of</strong><br />

MnSOD. A single-base substitution variant in the MnSOD gene results in an<br />

amino acid substitution from alanine to valine at position 9 (ala-9) in the<br />

mitochondrial targeting sequence (41). The substitution alters the secondary<br />

structure <strong>of</strong> the protein, affecting the cellular distribution <strong>of</strong> the enzyme and<br />

transport <strong>of</strong> MnSOD in the mitochondrion, where it would be biologically<br />

available (41). In the mitochondria, the MnSOD Ala variant has been shown to<br />

generate more active MnSOD protein compared with the Val allele (42). A<br />

second polymorphic variant has been identified in MnSOD, and Ile or Thr at<br />

amino acid position 58 (43) and the Thr 58 MnSOD variant was shown to exhibit<br />

only half the enzymatic activity <strong>of</strong> Ile 58 MnSOD (43,44). It can be hypothesized<br />

that the variant (ala-9 and Thr 58 MnSOD) would confer reduced protection <strong>of</strong><br />

breast tumor cells from ROS produced during therapy. Thus, lesser MnSOD<br />

activity due to polymorphisms could increase treatment efficacy but also result<br />

in greater toxicity due to ROS.<br />

CAT is most abundant in the liver, kidneys, and erythrocytes, where it<br />

decomposes H 2 O 2 to H 2 O and O 2 . A common functional polymorphism<br />

(C-262T) alters transcription factor binding and the expression <strong>of</strong> CAT in<br />

erythrocytes. We previously found that individuals with C alleles had significantly<br />

higher CAT levels compared with those with T/T genotypes (45).<br />

H 2 O 2 , if not neutralized, may contribute to further generation <strong>of</strong> ROS by a<br />

reaction catalyzed by MPO. MPO generates ROS endogenously by functioning<br />

as an antimicrobial enzyme, catalyzing a reaction between H 2 O 2 and chloride to<br />

generate hypochlorous acid (HOCl), a potent oxidizing agent. HOCl further<br />

reacts with other biological molecules to generate secondary radicals (46). The<br />

promoter region <strong>of</strong> the MPO gene has a G-to-A base substitution at position 463,<br />

which creates high- (G) and low-expression (A) alleles (47). Variability in genes<br />

in this oxidative stress-related pathway may influence ultimate levels <strong>of</strong> ROS<br />

and, thereby, tumor cell kill.


30 Amend et al.<br />

5-Fluorouracil and Methotrexate<br />

The agents 5FU and MTX are involved in folate metabolism, in the conversion<br />

<strong>of</strong> homocysteine to methionine, and in purine and pyrimidine synthesis. The<br />

activated metabolites <strong>of</strong> 5FU, similar to pyrimidine nucleotides, competitively<br />

and reversibly bind fluoropyrimidine thymidylase synthase (TS) inhibitors,<br />

forming covalently bound complexes with the TS enzyme, thus inhibiting the<br />

enzyme’s activity (48). Folate antagonists, such as MTX, are structurally similar<br />

to folate and inhibit the action <strong>of</strong> various enzymes such as dihydr<strong>of</strong>olate<br />

reductase (DHFR) (49). Their action is accomplished through incorporation as<br />

false precursors in DNA or RNA or through inhibition <strong>of</strong> proteins involved in<br />

nucleotide metabolism.<br />

Common polymorphisms have been reported in many enzymes and transport<br />

proteins involved in intracellular folate metabolism. Using a candidate gene<br />

approach, pharmacogenetic research, to date, has focused on TS because <strong>of</strong> its role<br />

as a drug target, dihydropyrimidine dehydrogenase (DPD) because <strong>of</strong> its<br />

involvement in pyrimidine degradation, and methylenetetrahydr<strong>of</strong>olate reductase<br />

(MTHFR) because <strong>of</strong> its key involvement in the control <strong>of</strong> the flux <strong>of</strong> intracellular<br />

folate metabolites (50). A genetic polymorphism in the 5 0 regulatory region <strong>of</strong> the<br />

TS gene promoter, consisting <strong>of</strong> either double (2R) or triple (3R) repeats <strong>of</strong> a 28-bp<br />

sequence (51), has been shown to influence TS expression, with higher expression<br />

in 3R/3R tumors (52). A G>C polymorphism has been identified within the<br />

second repeat <strong>of</strong> 3R alleles that alters the transcriptional activation <strong>of</strong> the TS gene<br />

constructs bearing this genotype (53). A 6-bp deletion (1494del6) within the<br />

3 0 UTR (54) has been also associated with decreased mRNA stability (55). Several<br />

clinical studies have shown the potential contribution <strong>of</strong> the above TS germinal<br />

polymorphism in the prediction <strong>of</strong> tumor responsiveness and/or toxicity following<br />

treatment by fluoropyrimidines (56). However, the role <strong>of</strong> TS polymorphism as a<br />

predictor <strong>of</strong> treatment outcome is still unclear.<br />

MTHFR catalyzes the conversion <strong>of</strong> 5,10-methyltetrahydr<strong>of</strong>olate to 5-methyltetrahydr<strong>of</strong>olate,<br />

which is an essential c<strong>of</strong>actor in the biosynthesis <strong>of</strong> deoxythymidine<br />

monophosphate. A common C677T transition in exon 4 and an A1298C<br />

transition in exon 7 <strong>of</strong> the MTHFR gene result in lower specific activity (57,58).<br />

Most studies, to date, have examined the association between 5FU and MTHFR<br />

polymorphisms in the context <strong>of</strong> colorectal cancer, but T<strong>of</strong>foli et al. (59) reported<br />

that <strong>of</strong> six patients who developed severe acute toxicity in the first cycle <strong>of</strong> adjuvant<br />

CMF, five had the MTHFR 677 T/T genotype and one had the C/C genotype.<br />

Because 5FU has a relatively narrow therapeutic index, toxicity increases<br />

as the dose is increased, and more than 80% <strong>of</strong> administered 5FU is catabolized<br />

by DPD (1). Patients with complete or partial deficiency <strong>of</strong> DPD <strong>of</strong>ten experience<br />

severe or even life-threatening toxicity after the administration <strong>of</strong> 5FU (60).<br />

At least 11 <strong>of</strong> 33 mutations identified in the dihydropyrimidine dehydrogenase<br />

(DPYD) gene have been detected in patients suffering from severe 5FU-associated<br />

toxicity (61–63).


Role <strong>of</strong> Genetic Variability in <strong>Breast</strong> <strong>Cancer</strong> Treatment Outcomes 31<br />

There have been few studies <strong>of</strong> MTHFR and TS polymorphisms in relation<br />

to breast cancer treatment outcomes. In a study with 19 human cancer cell lines,<br />

including breast cancer, Etienne et al. (64) found that there was a marked trend<br />

for greater 5FU efficacy in mutated MTHFR 1298C cell lines compared with<br />

those with A/A genotypes (P ¼ 0.055), but MTHFR C677T and TS 2R/3R<br />

polymorphisms were not associated with 5FU sensitivity. Largillier et al. (65)<br />

noted that genetic polymorphisms <strong>of</strong> MTHFR (677C>T and 1298A>C), DPD<br />

(IVS14þ1G>A), and TS were associated with efficacy and toxicity associated<br />

with capecitabine, an oral 5FU prodrug, in 105 breast cancer patients receiving<br />

capecitabine as monotherapy.<br />

Although MTX has been widely used to treat many cancers, including<br />

leukemia, lymphomas, and breast cancer, as well as some autoimmune diseases,<br />

there has been limited research to evaluate the effects <strong>of</strong> genetic variability on<br />

treatment outcomes. Associations have been noted for polymorphisms in folaterelated<br />

genes in leukemia (66–69), and Sohn et al. (70) found that in vitro<br />

transfection <strong>of</strong> mutant 677T MTHFR cDNA in breast and colon cancer cells<br />

increased the chemosensitivity <strong>of</strong> these cells to 5FU but decreased the chemosensitivity<br />

<strong>of</strong> breast cancer cells to MTX, suggesting at least a modulating role <strong>of</strong><br />

this genotype on cellular fluouropyrimidine sensitivity.<br />

Taxanes<br />

The taxanes paclitaxel and docetaxel interfere with microtubular disassembly,<br />

ultimately resulting in DNA fragmentation and features <strong>of</strong> apoptosis (71). The<br />

contributions <strong>of</strong> genetic variability to paclitaxel metabolism and breast cancer<br />

outcomes is still unclear. CYP2C8 is thought to be the main CYP-metabolizing<br />

enzyme <strong>of</strong> paclitaxel, while CYP3A4 and CYP3A5 play more minor roles in drug<br />

metabolism (19). At least three known variants in CYP2C8 exist, and the CYP2C8*3<br />

variant allele has been associated with decreased paclitaxel 6-ahydroxylase activity<br />

in human cell lines and human liver microsomes (72). In a recent study <strong>of</strong><br />

paclitaxel metabolism in breast cancer, however, CYP2C8 genotypes (CYP2C8*3<br />

and CYP2C8*4) were not significantly associated with paclitaxel clearance (73).<br />

Ethnic variation exists in the distribution <strong>of</strong> CYC2C8 alleles, with the CYP2C8*2<br />

allele more common in African-Americans than Caucasians (18% vs. 0%,<br />

respectively) and the CYP2C*3 allele more common in whites (13%) than<br />

African-Americans (2%). The CYPC2*5 allele is found in 9% <strong>of</strong> Asians (74).<br />

Antiestrogens<br />

The pharmacology and metabolism <strong>of</strong> TAM is complex, and differences in<br />

patient outcomes could be due to individual variation in the metabolism <strong>of</strong> the<br />

drug. TAM is metabolized primarily by CYP3A4, CYP2D6, CYP2C9,<br />

SULT1A1, and UGT2B15 (75–78). The main metabolites <strong>of</strong> TAM are<br />

N-desmethyltamoxifen (formed by CYP3A4) and endoxifen and 4-OH-TAM


32 Amend et al.<br />

(formed by CYP2D6) (8,76). CYP2D6 is polymorphic with 46 known allelic<br />

variants, which are categorized as resulting in abolished, decreased, normal,<br />

increased, or qualitatively altered catalytic activity (79). The major polymorphic<br />

variant CYP2D6 alleles include CYP2D6*2, CYP2D6*4, CYP2D6*5,<br />

CYP2D6*10, CYP2D6*17, and CYP2D6*41 (79).<br />

Human sulfotransferase 1A1 (SULT1A1) catalyzes the sulfation <strong>of</strong> a<br />

variety <strong>of</strong> phenolic and estrogenic compounds, including endogenous and<br />

environmental estrogens and the 4-hydroxy metabolite <strong>of</strong> TAM, 4-OH TAM<br />

(80). An exon 7 polymorphism in SULT1A1 (SULT1A1*2) results in a translated<br />

protein with approximately tw<strong>of</strong>old lower catalytic activity and decreased<br />

thermostability than the common allele (SULT1A1*1) (81). We found, as<br />

hypothesized, that women treated with TAM who were homozygous for the<br />

variant SULT1A1*2/*2 allele had approximately three times the risk <strong>of</strong> death<br />

compared with those homozygous for the common allele or heterozygotes (82).<br />

A wide range <strong>of</strong> drugs are substrates for glucuronosyltransferases (UGTs)<br />

and this pathway has been estimated to account for approximately one-third <strong>of</strong><br />

all drugs metabolized by phase II drug metabolizing enzymes (2). SULT1A1 and<br />

UGT2B15 have overlapping substrate specificity for 4-OH TAM. A functional<br />

polymorphism in UGT2B15, a G?T substitution at codon 85 within the putative<br />

substrate recognition site <strong>of</strong> the gene, results in higher V max compared with the<br />

common allele (83). In breast cancer patients treated with TAM, the combined<br />

effects <strong>of</strong> SULT1A1 and UGT2B15 were evaluated, and an increasing number <strong>of</strong><br />

variant SULT1A1 and UGT2B15 alleles was associated with poorer survival (84).<br />

Aromatase Inhibitors<br />

AIs, including exemestane, anastrozole, and letrozole, have come into clinical<br />

use more recently, developed primarily for hormone-dependent breast cancer in<br />

postmenopausal women. AIs act by inhibiting the cytochrome P450 enzyme,<br />

aromatase (CYP19), which catalyzes the conversion <strong>of</strong> androgens to estrogens.<br />

AIs block this enzyme, decreasing circulating levels <strong>of</strong> estrogens. It has been<br />

demonstrated that exemestane is metabolized by CYP3A4 but is also metabolized<br />

by aldoketoreductases (85). Letrazole is metabolized by CYP2A6 and<br />

CYP3A4 to a pharmacologically inactive carbinol metabolite (86), and anastrazole<br />

is metabolized by N-deaklylation and hydroxylation reactions followed<br />

by glucuronidation. However, specific is<strong>of</strong>orms involved and the effects on drug<br />

levels have not yet been identified.<br />

MOLECULAR EPIDEMIOLOGIC STUDIES OF PHARMACOGENETICS<br />

AND BREAST CANCER OUTCOMES<br />

To date, most analytical studies <strong>of</strong> pharmacogenetics and breast cancer outcomes<br />

have been conducted among patients receiving heterogeneous chemotherapy<br />

treatments, although most chemotherapy regimens included CP. Several studies


Role <strong>of</strong> Genetic Variability in <strong>Breast</strong> <strong>Cancer</strong> Treatment Outcomes 33<br />

have evaluated polymorphisms in phase II or GST enzymes in relation to<br />

prognosis after treatment for breast cancer. In a study <strong>of</strong> 92 women with<br />

advanced breast cancer, the GSTM1-null genotype was not significantly associated<br />

with tumor characteristics, disease-free survival (DFS), or overall survival<br />

(OS) (87), but in a retrospective study with a larger sample size, we found that<br />

women with low-activity homozygous genotypes for the GSTA1*B allele had<br />

better survival, with the hazard <strong>of</strong> death reduced by 30% (88,89), and women<br />

with low-activity GSTP1 Val/Val genotype had a similar risk reduction (89).<br />

Deletion polymorphisms in GSTM1 and GSTT1, genes responsive to ROS and<br />

resulting from radiation or chemotherapy and subsequent lipid peroxidation,<br />

were associated with a reduced hazard <strong>of</strong> death compared with women with both<br />

alleles present [HR ¼ 0.59; 95% confidence interval (CI), 0.36–0.97; and HR ¼<br />

0.51; 95% confidence interval (CI), 0.29–0.90] (90). In a study in China (91),<br />

women with GSTP1 Val/Val genotypes experienced a 60% reduction in<br />

overallmortality (HR ¼ 0.4; 95% confidence interval (CI), 0.2–0.8), although<br />

no associations were found with GSTM1 or GSTT1 genotypes. A smaller study<br />

(n ¼ 85) evaluated polymorphisms in both phase I and phase II metabolism <strong>of</strong><br />

CP among women with metastatic or inflammatory breast cancer treated with<br />

high-dose CP, cisplatin, and carmustine. Patients with the GSTM1-null genotype<br />

had prolonged OS, while those with variants in CYP3A4*1B and CYP3A5*1 had<br />

shortened survival. To our knowledge, only one genotype association study has<br />

focused on paclitaxel metabolism. In that study <strong>of</strong> 93 patients (73), there were no<br />

significant associations between ABCB1, ABCG2, CYP1B1, CYP3A4, CYP3A5,<br />

and CYP2C8 and paclitaxel clearance. However, patients homozygous for<br />

CYP1B1*3 Leu/Leu genotype had significantly longer progression-free survival<br />

(PFS) compared with patients with at least one Val allele (Table 1).<br />

Few studies have examined the effects <strong>of</strong> MTHFR genotypes on breast<br />

cancer survival. In the Shanghai <strong>Breast</strong> <strong>Cancer</strong> Study (92), no significant<br />

associations were found between MTHFR genotypes and the risk <strong>of</strong> death. A<br />

second study evaluated these same genotypes in African-American and Caucasian<br />

women from the Baltimore area (93) and found that carriers <strong>of</strong> the lowactivity<br />

allele at codon 1298 (A/C or C/C) had significantly poorer survival,<br />

while those with the variant allele at codon 677 (C/T or T/T) had improved<br />

survival. Associations were stronger in patients with ER-negative tumors, and<br />

African-American women had improved survival if they were carriers <strong>of</strong> either<br />

variant allele (A1298C, P interaction ¼ 0.088, C677T, P interaction ¼ 0.026).<br />

Molecular epidemiologic studies have indicated that genetic variation in<br />

phases I and II conjugating enzymes can influence the efficacy <strong>of</strong> TAM therapy<br />

for breast cancer. In a recent study in Sweden (94), patients homozygous for<br />

CYP2D6*4 alleles experienced significantly better DFS compared with patients<br />

with at least *1 allele (P ¼ 0.05). In a group <strong>of</strong> women randomized to 5, rather<br />

than two years <strong>of</strong> TAM, patients homozygous for CYP3A5*3 experienced<br />

improved relapse-free survival (RFS) (HR ¼ 0.02; CI, 0.07–0.55). However, this<br />

finding should be interpreted with caution since the number <strong>of</strong> patients within


34 Amend et al.<br />

Table 1 Epidemiologic Studies <strong>of</strong> Genetic Variation in Chemotherapeutic Metabolizing Genes and <strong>Breast</strong> <strong>Cancer</strong> Outcomes<br />

Genes SNP Participants Treatment Outcome Risk (95% CI) Reference<br />

Cyclophosphamide<br />

GSTM1 GSTM1-null<br />

genotype<br />

92 patients CAF OS or DFS GSTM1 not significant 87<br />

GSTA1 GSTA1*B/*B 245 patients CP based 5-yr survival GSTA1*B/*B HR ¼ 0.3 (0.1–0.8) 88<br />

GSTP1 GSTP Val/Val 240 patients CP based OS GSTP Val/Val HR ¼ 0.3 (0.1–1.0) 89<br />

GSTP1 GSTP Val/Val 1034 patients CP based OS GSTP Val/Val HR ¼ 0.4 (0.2–0.8) 91<br />

6 genes, CYPs,<br />

ABCB1,<br />

ABCG2<br />

17 genes, CYPs,<br />

METs, GSTs<br />

8 SNPs 93 patients High-dose<br />

paclitaxel and<br />

CP and A<br />

29 SNPs 85 patients CP and cisplatin<br />

and carmustine<br />

PFS CYP1B1*3 Leu/Leu Longer PFS (P ¼ 0.037) 73<br />

PK <strong>of</strong> CP<br />

and OS<br />

GSTM1-null genotype; better OS; at least one<br />

variant allele in CYP3A4*1B or CYP3A5*1<br />

(98)<br />

Methotrexate and 5-fluorouracil<br />

MTHFR MTHFR C677T;<br />

MTHFR<br />

A1298C<br />

MTHFR MTHFR C677T;<br />

Tamoxifen<br />

CYP3A5;<br />

CYP2D6;<br />

SULT1A1;<br />

UGT2B15<br />

CYP3A5;<br />

CYP2D6<br />

MTHFR<br />

A1298C<br />

CYP3A5*3;<br />

CYP2D6*4;<br />

SULT1A1*2;<br />

UGT2B15*2<br />

CYP3A5*3;<br />

CYP2D6*4;<br />

1067 patients Combination<br />

(most received<br />

5FU)<br />

OS MTHFR genotypes not significant with OS 92<br />

248 patients Combination OS MTHFR 1298 A/C or C/C HR ¼ 2.05<br />

(1.05–4.0);<br />

MTHFR 677 C/T or T/T HR ¼ 0.65 (0.31–1.35)<br />

677 patients<br />

(238<br />

randomized<br />

to TAM)<br />

TAM RFS CYP3A5 *3/*3 in 5-year<br />

TAM group HR ¼ 0.20 (0.07–0.55);<br />

CYP2D6 *4/*4 better<br />

RFS (n ¼ 677) (P ¼ 0.05)<br />

223 patients TAM RFS or DFS<br />

or OS<br />

CYP2D6 *4/*4<br />

HR ¼ 2.71 (1.15–6.41) for RFS;<br />

HR ¼ 2.44 (1.22–4.90) for DFS (both ratios<br />

unadjusted)<br />

(93)<br />

94<br />

(95)


Role <strong>of</strong> Genetic Variability in <strong>Breast</strong> <strong>Cancer</strong> Treatment Outcomes 35<br />

Table 1 Epidemiologic Studies <strong>of</strong> Genetic Variation in Chemotherapeutic Metabolizing Genes and <strong>Breast</strong> <strong>Cancer</strong> Outcomes (Continued )<br />

Genes SNP Participants Treatment Outcome Risk (95% CI) Reference<br />

CYP2D6;<br />

UGT2B15;<br />

SUL1A1<br />

CYP2D6*4;<br />

UGT2B15*2;<br />

SULT1A1*2<br />

337 patients<br />

(162 patients<br />

treated with<br />

TAM)<br />

SULT1A1 SULT1A1*2 377 patients<br />

(160 treated<br />

with TAM)<br />

TAM OS or PFS CYP2D6 NS;<br />

UGT2B15*2 and SULT1A1*2<br />

HR ¼ 4.4 (1.17–16.55) in TAM group<br />

TAM OS SULT1A1 *2/*2<br />

HR ¼ 2.9 (1.1–7.6) in TAM group<br />

84<br />

(82)<br />

Oxidative stress<br />

GSTM1/GSTT1 GSTM1 and<br />

GSTT1-null<br />

genotypes<br />

MPO, MnSOD,<br />

CAT<br />

MnSOD Ala-9Val,<br />

CAT-262C?T,<br />

MPO-463G?A<br />

MnSOD MnSOD-102 C?T 291 patients<br />

(230 treated<br />

with chemotherapy)<br />

251 patients Combination OS and DFS GSTM1-null genotype<br />

HR ¼ 0.59 (0.36–0.97;<br />

GSTT1-null genotype<br />

HR ¼ 0.51 (0.29–0.90)<br />

279 patients Combination OS MPO GG genotype<br />

HR ¼ 0.60 (0.38–0.95),<br />

MnSOD CC genotype HR ¼ 0.70 (0.40–1.24),<br />

genotypes combined HR ¼ 0.33 (0.13–0.80)<br />

Combination RFS MnSOD CT þ TT HR ¼ 0.65 (0.40–1.05)<br />

for chemotherapy group<br />

90<br />

96<br />

97<br />

Abbreviations: CAF, Cyclophosphamide/doxorubicin/fluorouracil; A, doxorubicin; CP, cyclophosphamide; T, DFS, disease-free survival; NS, not significant; OS,<br />

overall survival; PFS, progression-free survival; RFS, relapse-free survival; TAM, tamoxifen; 5FU, 5-fluorouracil; GST, glutathione S-transferase; SNP, single<br />

nucleotide polymorphism; METs, Metallothioneins; MTHFR, methylenetetrahydr<strong>of</strong>olate reductase; MPO, myeloperoxidase; MnSOD, manganese superoxide dismutase;<br />

CAT, catalase; PT, <strong>Ph</strong>armacokinetics.


36 Amend et al.<br />

this group was small. No statistically significant differences were found for<br />

CYP2D6, SULT1A1, or UGT2B15 and the length <strong>of</strong> TAM treatment. In contrast,<br />

Goetz et al. (95) found no significant associations between CYP3A5*3 and breast<br />

cancer outcomes; however, women with the CYP2D6 *4/*4 genotype (n ¼ 13)<br />

had statistically significant worse RFS (P ¼ 0.023) and DFS (P ¼ 0.12) but not<br />

OS (P ¼ 0.169). However, after adjustment for nodal status and tumor size,<br />

associations were no longer statistically significant. These results are in agreement<br />

with our previous report, wherein we noted no significant associations<br />

between the CYP2D6*4 genotype and OS or PFS among breast cancer patients<br />

treated with TAM (84).<br />

In addition to CYP is<strong>of</strong>orms, phase II enzymes, including sulfotransferase<br />

1A1 (SULT1A1) and UDP-glucuronosyltransferase is<strong>of</strong>orm 2B15 (UGT2B15),<br />

have been implicated in the metabolism <strong>of</strong> TAM, and genetic variability in these<br />

genes together may influence breast cancer survival. In a cohort <strong>of</strong> women with<br />

breast cancer who were treated with TAM, those who were homozygous for the<br />

low-activity SULT1A1*2 allele had a threefold higher risk <strong>of</strong> death compared<br />

with those with at least *1 common allele (HR ¼ 2.9; 1.1–7.6), while no significant<br />

associations were found among women who did not receive TAM (82).<br />

When considering the independent and combined effects <strong>of</strong> SULT1A1 and<br />

UGT2B15 genotypes on OS and recurrence among women treated with TAM,<br />

there was a nonsignificant trend toward increased risk <strong>of</strong> disease recurrence and<br />

poorer survival with increasing number <strong>of</strong> UGT2B15*2 alleles (P ¼ 0.11) (84).<br />

However, there was a significant trend toward decreased OS among women<br />

treated with TAM, with increasing number <strong>of</strong> variant alleles when UGT2B15 and<br />

SULT1A1 genotypes were combined (P ¼ 0.03), although these risk estimates<br />

may be unstable because only 16 cases had two variant alleles.<br />

We previously found that genotypes for MnSOD and MPO resulting in<br />

higher levels <strong>of</strong> ROS were associated with better survival in a study <strong>of</strong> women<br />

from Arkansas, and the combined effects <strong>of</strong> MnSOD and MPO were greater than<br />

those for either polymorphism alone (HR ¼ 0.33; CI, 0.13–0.80) (96). An<br />

additional MnSOD variant, located in the promoter region <strong>of</strong> the gene, was also<br />

associated with a significant decrease in RFS (CTþTT genotype HR ¼ 0.65, CI,<br />

0.40–1.05), although no significant associations were found with OS (97).<br />

In the context <strong>of</strong> a clinical trial for breast cancer with CAF versus CMF<br />

(SWOG 8897) (unpublished data, presented at the San Antonio <strong>Breast</strong> <strong>Cancer</strong><br />

Symposium, 2006), as in the Arkansas study, MPO genotypes encoding for<br />

higher activity were associated with better survival.<br />

FUTURE DIRECTIONS<br />

Studies, to date, indicate that interindividual variability in drug metabolism may<br />

affect treatment-related toxicities and DFS. However, much <strong>of</strong> the data are<br />

inconsistent, and it is likely that the evaluation <strong>of</strong> single genes in complex<br />

pathways presents only part <strong>of</strong> a complicated picture. Thus, future studies <strong>of</strong>


Role <strong>of</strong> Genetic Variability in <strong>Breast</strong> <strong>Cancer</strong> Treatment Outcomes 37<br />

pharmacogenetics and treatment outcomes should focus on entire drug metabolism<br />

pathways in large sample sets receiving homogeneous treatment regimens.<br />

Furthermore, assessment <strong>of</strong> variability across an entire gene, through construction<br />

<strong>of</strong> haplotype blocks and tag SNPs, may give more complete information<br />

than a study <strong>of</strong> single SNPs. It is likely that, in the future, other, more systemic<br />

markers <strong>of</strong> inherited drug metabolism capabilities may better predict treatment<br />

outcomes, and scans <strong>of</strong> variability across the entire genome may reveal factors<br />

that are likely to determine treatment outcomes. The consideration <strong>of</strong> individual<br />

genotypes, as well as tumor characteristics, in therapeutic decision making holds<br />

promise for individualized treatment for breast cancer in the future.<br />

REFERENCES<br />

1. Heggie GD, Sommadossi JP, Cross DS, et al. Clinical pharmacokinetics <strong>of</strong> 5-<br />

fluorouracil and its metabolites in plasma, urine, and bile. <strong>Cancer</strong> Res 1987; 47(8):<br />

2203–2206.<br />

2. Evans WE, Relling MV. <strong>Ph</strong>armacogenomics: translating functional genomics into<br />

rational therapeutics. Science 1999; 286(5439):487–491.<br />

3. DeVita VT Jr., Hellman S, Rosenberg SA (Eds.). <strong>Cancer</strong>: Principles and Practice <strong>of</strong><br />

Oncology. 5th ed. New York, NY: Lippincott-Raven, 1997.<br />

4. Early <strong>Breast</strong> <strong>Cancer</strong> Trialists’ Collaborative Group (EBCTG). Effects <strong>of</strong> chemotherapy<br />

and hormonal therapy for early breast cancer on recurrence and 15-year<br />

survival: an overview <strong>of</strong> the randomised trials. Lancet 2005; 365:1687–1717.<br />

5. Cuzick J. Aromatase inhibitors for breast cancer prevention. J Clin Oncol 2005;<br />

23(8):1636–1643.<br />

6. Conte P, Frassoldati A. Aromatase inhibitors in the adjuvant treatment <strong>of</strong> postmenopausal<br />

women with early breast cancer: putting safety issues into perspective.<br />

<strong>Breast</strong> J 2007; 13(1):28–35.<br />

7. Choi JY, Nowell SA, Blanco JG, et al. The role <strong>of</strong> genetic variability in drug<br />

metabolism pathways in breast cancer prognosis. <strong>Ph</strong>armacogenomics 2006; 7(4):<br />

613–624.<br />

8. Scripture CD, Sparreboom A, Figg WD. Modulation <strong>of</strong> cytochrome P450 activity:<br />

implications for cancer therapy. Lancet Oncol 2005; 6(10):780–789.<br />

9. Dirven HA, van Ommen B, van Bladeren PJ. Involvement <strong>of</strong> human glutathione<br />

S-transferase isoenzymes in the conjugation <strong>of</strong> cyclophosphamide metabolites with<br />

glutathione. <strong>Cancer</strong> Res 1994; 54(23):6215–6220.<br />

10. Chang TK, Weber GF, Crespi CL, et al. Differential activation <strong>of</strong> cyclophosphamide<br />

and ifosphamide by cytochromes P-450 2B and 3A in human liver microsomes.<br />

<strong>Cancer</strong> Res 1993; 53(23):5629–5637.<br />

11. Xie HJ, Yasar U, Lundgren S, et al. Role <strong>of</strong> polymorphic human CYP2B6 in<br />

cyclophosphamide bioactivation. <strong>Ph</strong>armacogenomics J 2003; 3(1):53–61.<br />

12. White IN, De MF, Gibbs AH, et al. Species differences in the covalent binding <strong>of</strong><br />

[14C]tamoxifen to liver microsomes and the forms <strong>of</strong> cytochrome P450 involved.<br />

Biochem <strong>Ph</strong>armacol 1995; 49(8):1035–1042.<br />

13. http://www.imm.ki.se/CYPalleles/cyp2b6.htm (accessed January 2007). 2007. 7. Ref<br />

Type: Electronic Citation


38 Amend et al.<br />

14. Jinno H, Tanaka-Kagawa T, Ohno A, et al. Functional characterization <strong>of</strong> cytochrome<br />

P450 2B6 allelic variants. Drug Metab Dispos 2003; 31(4):398–403.<br />

15. Ariyoshi N, Miyazaki M, Toide K, et al. A single nucleotide polymorphism <strong>of</strong><br />

CYP2b6 found in Japanese enhances catalytic activity by autoactivation. Biochem<br />

Biophys Res Commun 2001; 281(5):1256–1260.<br />

16. Lang T, Klein K, Fischer J, et al. Extensive genetic polymorphism in the human<br />

CYP2B6 gene with impact on expression and function in human liver. <strong><strong>Ph</strong>armacogenetics</strong><br />

2001; 11(5):399–415.<br />

17. Hiratsuka M, Takekuma Y, Endo N, et al. Allele and genotype frequencies <strong>of</strong><br />

CYP2B6 and CYP3A5 in the Japanese population. Eur J Clin <strong>Ph</strong>armacol 2002; 58(6):<br />

417–421.<br />

18. Lamba V, Lamba J, Yasuda K, et al. Hepatic CYP2B6 expression: gender and ethnic<br />

differences and relationship to CYP2B6 genotype and CAR (constitutive androstane<br />

receptor) expression. J <strong>Ph</strong>armacol Exp Ther 2003; 307(3):906–922.<br />

19. Daly AK. <strong><strong>Ph</strong>armacogenetics</strong> <strong>of</strong> the major polymorphic metabolizing enzymes.<br />

Fundam Clin <strong>Ph</strong>armacol 2003; 17(1):27–41.<br />

20. Rebbeck TR, Jaffe JM, Walker AH, et al. Modification <strong>of</strong> clinical presentation <strong>of</strong><br />

prostate tumors by a novel genetic variant in CYP3A4. J Natl <strong>Cancer</strong> Inst 1998;<br />

90(16):1225–1229.<br />

21. Walker AH, Jaffe JM, Gunasegaram S, et al. Characterization <strong>of</strong> an allelic variant<br />

in the nifedipine-specific element <strong>of</strong> CYP3A4: ethnic distribution and implications<br />

for prostate cancer risk. Mutations in brief no. 191. Hum Mutat 1998; 12(4):289<br />

(online).<br />

22. Amirimani B, Walker AH, Weber BL, et al. RESPONSE: re: modification <strong>of</strong> clinical<br />

presentation <strong>of</strong> prostate tumors by a novel genetic variant in CYP3A4. J Natl <strong>Cancer</strong><br />

Inst 1999; 91(18):1588–1590.<br />

23. Roy P, Yu LJ, Crespi CL, et al. Development <strong>of</strong> a substrate-activity based approach<br />

to identify the major human liver P-450 catalysts <strong>of</strong> cyclophosphamide and ifosfamide<br />

activation based on cDNA-expressed activities and liver microsomal P-450<br />

pr<strong>of</strong>iles. Drug Metab Dispos 1999; 27(6):655–666.<br />

24. Raucy JL. Regulation <strong>of</strong> CYP3A4 expression in human hepatocytes by pharmaceuticals<br />

and natural products. Drug Metab Dispos 2003; 31(5):533–539.<br />

25. Kuehl P, Zhang J, Lin Y, et al. Sequence diversity in CYP3A promoters and characterization<br />

<strong>of</strong> the genetic basis <strong>of</strong> polymorphic CYP3A5 expression. Nat Genet<br />

2001; 27(4):383–391.<br />

26. Hustert E, Haberl M, Burk O, et al. The genetic determinants <strong>of</strong> the CYP3A5<br />

polymorphism. <strong><strong>Ph</strong>armacogenetics</strong> 2001; 11(9):773–779.<br />

27. Griskevicius L, Yasar U, Sandberg M, et al. Bioactivation <strong>of</strong> cyclophosphamide: the<br />

role <strong>of</strong> polymorphic CYP2C enzymes. Eur J Clin <strong>Ph</strong>armacol 2003; 59(2):103–109.<br />

28. Morel F, Rauch C, Coles B, et al. The human glutathione transferase alpha locus:<br />

genomic organization <strong>of</strong> the gene cluster and functional characterization <strong>of</strong> the<br />

genetic polymorphism in the hGSTA1 promoter. <strong><strong>Ph</strong>armacogenetics</strong> 2002; 12(4):<br />

277–286.<br />

29. Coles BF, Morel F, Rauch C, et al. Effect <strong>of</strong> polymorphism in the human glutathione<br />

S-transferase A1 promoter on hepatic GSTA1 and GSTA2 expression. <strong><strong>Ph</strong>armacogenetics</strong><br />

2001; 11(8):663–669.<br />

30. Ali-Osman F, Akande O, Antoun G, et al. Molecular cloning, characterization, and<br />

expression in Escherichia coli <strong>of</strong> full-length cDNAs <strong>of</strong> three human glutathione


Role <strong>of</strong> Genetic Variability in <strong>Breast</strong> <strong>Cancer</strong> Treatment Outcomes 39<br />

S-transferase Pi gene variants. Evidence for differential catalytic activity <strong>of</strong> the<br />

encoded proteins. J Biol Chem 1997; 272(15):10004–10012.<br />

31. Harries LW, Stubbins MJ, Forman D, et al. Identification <strong>of</strong> genetic polymorphisms<br />

at the glutathione S-transferase Pi locus and association with susceptibility to<br />

bladder, testicular and prostate cancer. Carcinogenesis 1997; 18(4):641–644.<br />

32. Srivastava SK, Singhal SS, Hu X, et al. Differential catalytic efficiency <strong>of</strong> allelic<br />

variants <strong>of</strong> human glutathione S-transferase Pi in catalyzing the glutathione conjugation<br />

<strong>of</strong> thiotepa. Arch Biochem Biophys 1999; 366(1):89–94.<br />

33. Minotti G, Menna P, Salvatorelli E, et al. Anthracyclines: molecular advances and<br />

pharmacologic developments in antitumor activity and cardiotoxicity. <strong>Ph</strong>armacol<br />

Rev 2004; 56(2):185–229.<br />

34. Ravid A, Rocker D, Machlenkin A, et al. 1,25-Dihydroxyvitamin D3 enhances the<br />

susceptibility <strong>of</strong> breast cancer cells to doxorubicin-induced oxidative damage.<br />

<strong>Cancer</strong> Res 1999; 59(4):862–867.<br />

35. Kong Q, Lillehei KO. Antioxidant inhibitors for cancer therapy. Med Hypotheses<br />

1998; 51(5):405–409.<br />

36. Kalyanaraman B, Perez-Reyes E, Mason RP. Spin-trapping and direct electron spin<br />

resonance investigations <strong>of</strong> the redox metabolism <strong>of</strong> quinone anticancer drugs.<br />

Biochim Biophys Acta 1980; 630(1):119–130.<br />

37. Conklin KA. Dietary antioxidants during cancer chemotherapy: impact on chemotherapeutic<br />

effectiveness and development <strong>of</strong> side effects. Nutr <strong>Cancer</strong> 2000; 37(1):<br />

1–18.<br />

38. Sellin S, Holmquist B, Mannervik B, et al. Oxidation and reduction <strong>of</strong> 4-hydroxyalkenals<br />

catalyzed by isozymes <strong>of</strong> human alcohol dehydrogenase. Biochemistry<br />

1991; 30(9):2514–2518.<br />

39. Lakhman SS, Ghosh D, Blanco JG. Functional significance <strong>of</strong> a natural allelic<br />

variant <strong>of</strong> human carbonyl reductase 3 (CBR3). Drug Metab Dispos 2005; 33(2):<br />

254–257.<br />

40. Costantini P, Jacotot E, Decaudin D, et al. Mitochondrion as a novel target <strong>of</strong><br />

anticancer chemotherapy. J Natl <strong>Cancer</strong> Inst 2000; 92(13):1042–1053.<br />

41. Shimoda-Matsubayashi S, Matsumine H, Kobayashi T, et al. Structural dimorphism<br />

in the mitochondrial targeting sequence in the human manganese superoxide dismutase<br />

gene. A predictive evidence for conformational change to influence mitochondrial<br />

transport and a study <strong>of</strong> allelic association in Parkinson’s disease. Biochem<br />

Biophys Res Commun 1996; 226(2):561–565.<br />

42. Sutton A, Khoury H, Prip-Buus C, et al. The Ala16Val genetic dimorphism modulates<br />

the import <strong>of</strong> human manganese superoxide dismutase into rat liver mitochondria.<br />

<strong><strong>Ph</strong>armacogenetics</strong> 2003; 13(3):145–157.<br />

43. Ho YS, Crapo JD. Isolation and characterization <strong>of</strong> complementary DNAs encoding<br />

human manganese-containing superoxide dismutase. FEBS Lett 1988; 229(2):<br />

256–260.<br />

44. Borgstahl GE, Parge HE, Hickey MJ, et al. Human mitochondrial manganese<br />

superoxide dismutase polymorphic variant Ile58Thr reduces activity by destabilizing<br />

the tetrameric interface. Biochemistry 1996; 35(14):4287–4297.<br />

45. Forsberg L, Lyrenas L, de FU, et al. A common functional C-T substitution polymorphism<br />

in the promoter region <strong>of</strong> the human catalase gene influences transcription<br />

factor binding, reporter gene transcription and is correlated to blood catalase levels.<br />

Free Radic Biol Med 2001; 30(5):500–505.


40 Amend et al.<br />

46. Kleban<strong>of</strong>f SJ. Myeloperoxidase: occurrence and biological function. In: Everse J,<br />

Everse KE, Grisham MB, eds. Peroxides in Chemistry. Boca Raton: CRC<br />

Press, 1991.<br />

47. Piedrafita FJ, Molander RB, Vansant G, et al. An Alu element in the myeloperoxidase<br />

promoter contains a composite SP1-thyroid hormone-retinoic acid response<br />

element. J Biol Chem 1996; 271(24):14412–14420.<br />

48. Chabner BA, Ryan DP, Paz-Ares LP, et al. Antineoplastic agents. In: Hardman JG,<br />

Limbird LE, Gilman AG, eds. Goodman and Gilman’s the <strong>Ph</strong>armacological Basis <strong>of</strong><br />

Therapeutics. New York, NY: McGraw-Hill, 2001:1389–1460.<br />

49. Calvert H. An overview <strong>of</strong> folate metabolism: features relevant to the action and<br />

toxicities <strong>of</strong> antifolate anticancer agents. Semin Oncol 1999; 26(2 suppl 6):3–10.<br />

50. Robien K, Boynton A, Ulrich CM. <strong><strong>Ph</strong>armacogenetics</strong> <strong>of</strong> folate-related drug targets in<br />

cancer treatment. <strong>Ph</strong>armacogenomics 2005; 6(7):673–689.<br />

51. Horie N, Aiba H, Oguro K, et al. Functional analysis and DNA polymorphism <strong>of</strong> the<br />

tandemly repeated sequences in the 5 0 -terminal regulatory region <strong>of</strong> the human gene<br />

for thymidylate synthase. Cell Struct Funct 1995; 20(3):191–197.<br />

52. Pullarkat ST, Stoehlmacher J, Ghaderi V, et al. Thymidylate synthase gene polymorphism<br />

determines response and toxicity <strong>of</strong> 5-FU chemotherapy. <strong>Ph</strong>armacogenomics<br />

J 2001; 1(1):65–70.<br />

53. Mandola MV, Stoehlmacher J, Muller-Weeks S, et al. A novel single nucleotide<br />

polymorphism within the 5 0 tandem repeat polymorphism <strong>of</strong> the thymidylate synthase<br />

gene abolishes USF-1 binding and alters transcriptional activity. <strong>Cancer</strong> Res<br />

2003; 63(11):2898–2904.<br />

54. Ulrich CM, Bigler J, Velicer CM, et al. Searching expressed sequence tag databases:<br />

discovery and confirmation <strong>of</strong> a common polymorphism in the thymidylate synthase<br />

gene. <strong>Cancer</strong> Epidemiol Biomarkers Prev 2000; 9(12):1381–1385.<br />

55. Mandola MV, Stoehlmacher J, Zhang W, et al. A 6 bp polymorphism in the thymidylate<br />

synthase gene causes message instability and is associated with decreased<br />

intratumoral TS mRNA levels. <strong><strong>Ph</strong>armacogenetics</strong> 2004; 14(5):319–327.<br />

56. Danenberg PV. <strong>Ph</strong>armacogenomics <strong>of</strong> thymidylate synthase in cancer treatment.<br />

Front Biosci 2004; 9:2484–2494.<br />

57. Stern LL, Bagley PJ, Rosenberg IH, et al. Conversion <strong>of</strong> 5-formyltetrahydr<strong>of</strong>olic acid<br />

to 5-methyltetrahydr<strong>of</strong>olic acid is unimpaired in folate-adequate persons homozygous<br />

for the C677T mutation in the methylenetetrahydr<strong>of</strong>olate reductase gene.<br />

J Nutr 2000; 130(9):2238–2242.<br />

58. Weisberg I, Tran P, Christensen B, et al. A second genetic polymorphism in<br />

methylenetetrahydr<strong>of</strong>olate reductase (MTHFR) associated with decreased enzyme<br />

activity. Mol Genet Metab 1998; 64(3):169–172.<br />

59. T<strong>of</strong>foli G, Veronesi A, Boiocchi M, et al. MTHFR gene polymorphism and severe<br />

toxicity during adjuvant treatment <strong>of</strong> early breast cancer with cyclophosphamide,<br />

methotrexate, and fluorouracil (CMF). Ann Oncol 2000; 11(3):373–374.<br />

60. Etienne MC, Lagrange JL, Dassonville O, et al. Population study <strong>of</strong><br />

dihydropyrimidine dehydrogenase in cancer patients. J Clin Oncol 1994; 12(11):<br />

2248–2253.<br />

61. van Kuilenburg AB, Haasjes J, Richel DJ, et al. Clinical implications <strong>of</strong> dihydropyrimidine<br />

dehydrogenase (DPD) deficiency in patients with severe 5-<br />

fluorouracil-associated toxicity: identification <strong>of</strong> new mutations in the DPD gene.<br />

Clin <strong>Cancer</strong> Res 2000; 6(12):4705–4712.


Role <strong>of</strong> Genetic Variability in <strong>Breast</strong> <strong>Cancer</strong> Treatment Outcomes 41<br />

62. van Kuilenburg AB, De Abreu RA, van Gennip AH. <strong>Ph</strong>armacogenetic and clinical<br />

aspects <strong>of</strong> dihydropyrimidine dehydrogenase deficiency. Ann Clin Biochem 2003;<br />

40(pt 1):41–45.<br />

63. Collie-Duguid ES, Etienne MC, Milano G, et al. Known variant DPYD alleles do not<br />

explain DPD deficiency in cancer patients. <strong><strong>Ph</strong>armacogenetics</strong> 2000; 10(3):217–223.<br />

64. Etienne MC, Ilc K, Formento JL, et al. Thymidylate synthase and methylenetetrahydr<strong>of</strong>olate<br />

reductase gene polymorphisms: relationships with 5-fluorouracil<br />

sensitivity. Br J <strong>Cancer</strong> 2004; 90(2):526–534.<br />

65. Largillier R, Etienne-Grimaldi MC, Formento JL, et al. <strong><strong>Ph</strong>armacogenetics</strong> <strong>of</strong><br />

capecitabine in advanced breast cancer patients. Clin <strong>Cancer</strong> Res 2006; 12(18):<br />

5496–5502.<br />

66. Laverdiere C, Chiasson S, Costea I, et al. Polymorphism G80A in the reduced folate<br />

carrier gene and its relationship to methotrexate plasma levels and outcome <strong>of</strong><br />

childhood acute lymphoblastic leukemia. Blood 2002; 100(10):3832–3834.<br />

67. Costea I, Moghrabi A, Laverdiere C, et al. Folate cycle gene variants and chemotherapy<br />

toxicity in pediatric patients with acute lymphoblastic leukemia. Haematologica<br />

2006; 91(8):1113–1116.<br />

68. Seidemann K, Book M, Zimmermann M, et al. MTHFR 677 (C–>T) polymorphism<br />

is not relevant for prognosis or therapy-associated toxicity in pediatric NHL:<br />

results from 484 patients <strong>of</strong> multicenter trial NHL-BFM 95. Ann Hematol 2006;<br />

85(5):291–300.<br />

69. Jazbec J, Kitanovski L, Aplenc R, et al. No evidence <strong>of</strong> association <strong>of</strong> methylenetetrahydr<strong>of</strong>olate<br />

reductase polymorphism with occurrence <strong>of</strong> second neoplasms<br />

after treatment <strong>of</strong> childhood leukemia. Leuk Lymphoma 2005; 46(6):893–897.<br />

70. Sohn KJ, Croxford R, Yates Z, et al. Effect <strong>of</strong> the methylenetetrahydr<strong>of</strong>olate<br />

reductase C677T polymorphism on chemosensitivity <strong>of</strong> colon and breast cancer cells<br />

to 5-fluorouracil and methotrexate. J Natl <strong>Cancer</strong> Inst 2004; 96(2):134–144.<br />

71. Schiff PB, Fant J, Horwitz SB. Promotion <strong>of</strong> microtubule assembly in vitro by taxol.<br />

Nature 1979; 277(5698):665–667.<br />

72. Soyama A, Saito Y, Hanioka N, et al. Non-synonymous single nucleotide alterations<br />

found in the CYP2C8 gene result in reduced in vitro paclitaxel metabolism. Biol<br />

<strong>Ph</strong>arm Bull 2001; 24(12):1427–1430.<br />

73. Marsh S, Somlo G, Li X, et al. <strong>Ph</strong>armacogenetic analysis <strong>of</strong> paclitaxel transport and<br />

metabolism genes in breast cancer. <strong>Ph</strong>armacogenomics J 2007.<br />

74. van Schaik RH. Implications <strong>of</strong> cytochrome P450 genetic polymorphisms on the<br />

toxicity <strong>of</strong> antitumor agents. Ther Drug Monit 2004; 26(2):236–240.<br />

75. Crewe HK, Ellis SW, Lennard MS, et al. Variable contribution <strong>of</strong> cytochromes P450<br />

2D6, 2C9 and 3A4 to the 4-hydroxylation <strong>of</strong> tamoxifen by human liver microsomes.<br />

Biochem <strong>Ph</strong>armacol 1997; 53(2):171–178.<br />

76. Dehal SS, Kupfer D. CYP2D6 catalyzes tamoxifen 4-hydroxylation in human liver.<br />

<strong>Cancer</strong> Res 1997; 57(16):3402–3406.<br />

77. Stearns V, Johnson MD, Rae JM, et al. Active tamoxifen metabolite plasma concentrations<br />

after coadministration <strong>of</strong> tamoxifen and the selective serotonin reuptake<br />

inhibitor paroxetine. J Natl <strong>Cancer</strong> Inst 2003; 95(23):1758–1764.<br />

78. Nishiyama T, Ogura K, Nakano H, et al. Reverse geometrical selectivity in glucuronidation<br />

and sulfation <strong>of</strong> cis- and trans-4-hydroxytamoxifens by human liver<br />

UDP-glucuronosyltransferases and sulfotransferases. Biochem <strong>Ph</strong>armacol 2002;<br />

63(10):1817–1830.


42 Amend et al.<br />

79. Ingelman-Sundberg M. Genetic polymorphisms <strong>of</strong> cytochrome P450 2D6<br />

(CYP2D6): clinical consequences, evolutionary aspects and functional diversity.<br />

<strong>Ph</strong>armacogenomics J 2005; 5(1):6–13.<br />

80. Falany CN, Wheeler J, Oh TS, et al. Steroid sulfation by expressed human cytosolic<br />

sulfotransferases. J Steroid Biochem Mol Biol 1994; 48(4):369–375.<br />

81. Raftogianis RB, Wood TC, Otterness DM, et al. <strong>Ph</strong>enol sulfotransferase pharmacogenetics<br />

in humans: association <strong>of</strong> common SULT1A1 alleles with TS PST phenotype.<br />

Biochem Biophys Res Commun 1997; 239(1):298–304.<br />

82. Nowell S, Sweeney C, Winters M, et al. Association between sulfotransferase 1A1<br />

genotype and survival <strong>of</strong> breast cancer patients receiving tamoxifen therapy. J Natl<br />

<strong>Cancer</strong> Inst 2002; 94(21):1635–1640.<br />

83. Levesque E, Beaulieu M, Green MD, et al. Isolation and characterization <strong>of</strong><br />

UGT2B15(Y85): a UDP-glucuronosyltransferase encoded by a polymorphic gene.<br />

<strong><strong>Ph</strong>armacogenetics</strong> 1997; 7(4):317–325.<br />

84. Nowell SA, Ahn J, Rae JM, et al. Association <strong>of</strong> genetic variation in tamoxifenmetabolizing<br />

enzymes with overall survival and recurrence <strong>of</strong> disease in breast<br />

cancer patients. <strong>Breast</strong> <strong>Cancer</strong> Res Treat 2005; 91(3):249–258.<br />

85. Buzdar AU, Robertson JF, Eiermann W, et al. An overview <strong>of</strong> the pharmacology and<br />

pharmacokinetics <strong>of</strong> the newer generation aromatase inhibitors anastrozole, letrozole,<br />

and exemestane. <strong>Cancer</strong> 2002; 95(9):2006–2016.<br />

86. Sioufi A, Gauducheau N, Pineau V, et al. Absolute bioavailability <strong>of</strong> letrozole in<br />

healthy postmenopausal women. Biopharm Drug Dispos 1997; 18(9):779–789.<br />

87. Lizard-Nacol S, Coudert B, Colosetti P, et al. Glutathione S-transferase M1 null<br />

genotype: lack <strong>of</strong> association with tumour characteristics and survival in advanced<br />

breast cancer. <strong>Breast</strong> <strong>Cancer</strong> Res 1999; 1(1):81–87.<br />

88. Sweeney C, Ambrosone CB, Joseph L, et al. Association between a glutathione<br />

S-transferase A1 promoter polymorphism and survival after breast cancer treatment.<br />

Int J <strong>Cancer</strong> 2003; 103(6):810–814.<br />

89. Sweeney C, McClure GY, Fares MY, et al. Association between survival after<br />

treatment for breast cancer and glutathione S-transferase P1 Ile105Val polymorphism.<br />

<strong>Cancer</strong> Res 2000; 60(20):5621–5624.<br />

90. Ambrosone CB, Sweeney C, Coles BF, et al. Polymorphisms in glutathione<br />

S-transferases (GSTM1 and GSTT1) and survival after treatment for breast cancer.<br />

<strong>Cancer</strong> Res 2001; 61(19):7130–7135.<br />

91. Yang G, Shu XO, Ruan ZX, et al. Genetic polymorphisms in glutathione-S-transferase<br />

genes (GSTM1, GSTT1, GSTP1) and survival after chemotherapy for invasive<br />

breast carcinoma. <strong>Cancer</strong> 2005; 103(1):52–58.<br />

92. Shrubsole MJ, Shu XO, Ruan ZX, et al. MTHFR genotypes and breast cancer survival<br />

after surgery and chemotherapy: a report from the Shanghai <strong>Breast</strong> <strong>Cancer</strong><br />

Study. <strong>Breast</strong> <strong>Cancer</strong> Res Treat 2005; 91(1):73–79.<br />

93. Martin DN, Boersma BJ, Howe TM, et al. Association <strong>of</strong> MTHFR gene polymorphisms<br />

with breast cancer survival. BMC <strong>Cancer</strong> 2006; 6:257.<br />

94. Wegman P, Elingarami S, Carstensen J, et al. Genetic variants <strong>of</strong> CYP3A5,<br />

CYP2D6, SULT1A1, UGT2B15 and tamoxifen response in postmenopausal patients<br />

with breast cancer. <strong>Breast</strong> <strong>Cancer</strong> Res 2007; 9(1):R7.<br />

95. Goetz MP, Rae JM, Suman VJ, et al. <strong><strong>Ph</strong>armacogenetics</strong> <strong>of</strong> tamoxifen biotransformation<br />

is associated with clinical outcomes <strong>of</strong> efficacy and hot flashes. J Clin<br />

Oncol 2005; 23(36):9312–9318.


Role <strong>of</strong> Genetic Variability in <strong>Breast</strong> <strong>Cancer</strong> Treatment Outcomes 43<br />

96. Ambrosone CB, Ahn J, Singh KK, et al. Polymorphisms in genes related to oxidative<br />

stress (MPO, MnSOD, CAT) and survival after treatment for breast cancer. <strong>Cancer</strong><br />

Res 2005; 65(3):1105–1111.<br />

97. Martin RC, Ahn J, Nowell SA, et al. Association between manganese superoxide<br />

dismutase promoter gene polymorphism and breast cancer survival. <strong>Breast</strong> <strong>Cancer</strong><br />

Res 2006; 8(4):R45.<br />

98. Petros WP, Hopkins PJ, Spruill S, et al. Associations between drug metabolism<br />

genotype, chemotherapy pharmacokinetics, and overall survival in patients with<br />

breast cancer. J Clin Oncol 2005; 23(25):6117–6125.


4<br />

The Role <strong>of</strong> Epigenetics in <strong>Breast</strong> <strong>Cancer</strong>:<br />

Implications for Diagnosis, Prognosis,<br />

and Treatment<br />

Amy M. Dworkin, Tim H.-M. Huang, and Amanda E. Toland<br />

Human <strong>Cancer</strong> Genetics Program, Department <strong>of</strong> Molecular Virology,<br />

Immunology, and Medical Genetics, The Ohio State University,<br />

Columbus, Ohio, U.S.A.<br />

INTRODUCTION<br />

In the past 10 years, the field <strong>of</strong> epigenetics has expanded into an essential<br />

component <strong>of</strong> clinical cancer research and therapeutics. Epigenetic modifications<br />

in DNA or chromatin are stably transmitted over rounds <strong>of</strong> cell division but do<br />

not result in any permanent alterations to the underlying DNA sequence. These<br />

changes, such as histone modifications and DNA methylation <strong>of</strong> tumor suppressor<br />

genes, are considered a hallmark <strong>of</strong> certain cancers, including breast<br />

cancer. Unlike germline mutations, epigenetic modifications can potentially be<br />

reversed, making them very appealing for preventative care and therapeutics<br />

against cancers (1). Because <strong>of</strong> this, demethylating agents are currently being<br />

evaluated for efficacy in the treatment <strong>of</strong> breast cancer.<br />

There are distinct mechanisms that initiate and sustain epigenetic modifications<br />

(2–8). Of these, DNA methylation and posttranslational modifications<br />

<strong>of</strong> histone proteins are the best understood. DNA methylation is a covalent<br />

addition <strong>of</strong> a methyl group to DNA, usually to a cytosine located 5 0 to guanasine<br />

(CpG dinucleotide). CpG dinucleotides are underrepresented in the genome,<br />

45


46 Dworkin et al.<br />

Figure 1 Methylation <strong>of</strong> CpG islands. Unmethylated CpG islands (in light gray) allow<br />

for transcription, while methylated CpG islands (in dark gray) repress transcription.<br />

except for small clusters, referred to as CpG islands, located in or near the<br />

promoter <strong>of</strong> approximately half <strong>of</strong> all genes (1,9–11). In addition, other epigenetic<br />

modifications have recently been explored, including the Polycomb<br />

group (PcG) proteins, which repress gene function and can only be overcome by<br />

germline differentiation processes, and small noncoding RNA molecules, which<br />

regulate gene expression by targeting RNA degradation (3). Small noncoding<br />

RNA molecules have recently been found to also target gene promoters and<br />

result in transcriptional gene silencing (11,12).<br />

DNA methylation plays a variety <strong>of</strong> roles in normal cells: silencing <strong>of</strong><br />

transposable elements and viral sequences, maintenance <strong>of</strong> chromosomal<br />

integrity, X chromosome inactivation, and transcriptional regulation <strong>of</strong> genes<br />

(1). Methylation is established and maintained by at least four DNA methyltransferases<br />

(DNMTs), DNMT1, DNMT3A, DNMT3B, and DNMT3L, which<br />

catalyze the covalent addition <strong>of</strong> methyl groups to cytosine in the CpG dinucleotide.<br />

Methylation <strong>of</strong> CpG islands causes transcriptional inactivation (Fig. 1).<br />

These methyltransferases and their is<strong>of</strong>orms have been demonstrated to be<br />

essential for development; a knockout <strong>of</strong> any <strong>of</strong> these enzymes in mouse<br />

embryos results in embryonic or perinatal lethality (1). Functionally, there are<br />

two types <strong>of</strong> methyltransferases, maintenance and de novo. Maintenance methyltransferases<br />

are essential for copying DNA methylation patterns during cell<br />

division, and de novo methyltransferases introduce new patterns <strong>of</strong> methylation<br />

during early development. There are duplications <strong>of</strong> functions between these<br />

methylases so that perturbation <strong>of</strong> both types may be necessary to change overall<br />

methylation patterns (6). DNMT1 functions as a maintenance methyltransferase<br />

by copying DNA methylation patterns to daughter strands, following replication.<br />

It is the most abundant and catalytically active form. DNMT3A and DNMT3B<br />

act on both unmethylated and hemimethylated DNA. However, they are thought<br />

to have different methylation targets depending on the cell type and stage <strong>of</strong><br />

development (13). In vitro knockouts <strong>of</strong> DNMT3a or DNMT3b each exhibit<br />

distinct phenotypes. DNMT3a-null cells lack methylation at imprinting loci,<br />

while DNMT3b-null cells result in loss <strong>of</strong> DNA methylation on endogenous


The Role <strong>of</strong> Epigenetics in <strong>Breast</strong> <strong>Cancer</strong> 47<br />

Figure 2 Histone modification and chromatin remodeling. Histone modifications, like<br />

acetylation, change the affinity <strong>of</strong> the histones to DNA. The acetylated histones on the left<br />

are in an open euchromatin structure and can be transcribed. The unacetylated histones on<br />

the right are in a heterochromatin structure and repress transcription. Source: Cartoon<br />

compliments <strong>of</strong> Kristi Bennett.<br />

C-type retroviral DNA. Deletion <strong>of</strong> both DNMT3a and DNMT3b abolishes de<br />

novo methylation, while DNMT1 deletion produces bulk DNA demethylation<br />

(7). Of the methyltransferases, DNMT3L is unique. It stimulates de novo<br />

methylation by DNMT3a, does not enhance its binding to DNA, and alone does<br />

not bind to DNA (14).<br />

Along with methylation, posttranslational modification <strong>of</strong> histones plays<br />

an important role in epigenetics. Histone proteins associate with DNA to form<br />

nucleosomes, permitting large amounts <strong>of</strong> DNA to be neatly packaged into the<br />

nucleus. There are two configurations <strong>of</strong> chromatin, heterochromatin and<br />

euchromatin. Heterochromatin is condensed and transcriptionally inactive, while<br />

euchromatin has an “open” configuration and is favorable for gene transcription<br />

(Fig. 2). The N-terminal tails <strong>of</strong> histones “stick out” <strong>of</strong> the nucleosome and are<br />

subject to posttranslational modification such as acetylation, phosphorylation,<br />

methylation, ubiquitination, and sumoylation (2,4,7,15). The pattern <strong>of</strong> histone<br />

modification creates a “code,” which is read by proteins involved in chromatin<br />

remodeling and the dynamics <strong>of</strong> gene transcription (6,11). Acetylation and<br />

deacteylation are the most common types <strong>of</strong> histone modification. Histone<br />

deacetylases (HDACs) are a class <strong>of</strong> enzymes that remove acetyl groups from a<br />

e-N-acetyl lysine amino acid on a histone. Its action is opposite to that <strong>of</strong> histone<br />

acetyltransferases (HATs). Deacetylation removes acetyl groups from histone<br />

tails, causing the histones to wrap more tightly around the DNA and interfering<br />

with transcription by blocking access to transcription factors. The overall result<br />

<strong>of</strong> histone deacetylation is a global (nonspecific) reduction in gene expression<br />

(Table 1).<br />

METHYLATION IN CANCER<br />

<strong>Cancer</strong> development is complex due to progressive changes <strong>of</strong> the genome that<br />

result in altering normal cellular processes such as cell growth, apoptosis, and<br />

angiogenesis (12). Both genetic and epigenetic modifications play a strong role<br />

in tumorigenesis (7). In classic tumor biology, cells lose both copies <strong>of</strong> tumor


48 Dworkin et al.<br />

Table 1<br />

Histone Modifiers<br />

Histone modifier Mechanism Catalysis Role in transcription<br />

Acetylation<br />

Deacetylation<br />

<strong>Ph</strong>osphorylation<br />

Dephosphorylation<br />

Methylation<br />

Demethylation<br />

Ubiquitination<br />

Deubiquitination<br />

Sumoylation<br />

Introduction <strong>of</strong> the<br />

acetyl group to<br />

the lysine residue<br />

at the N terminus<br />

<strong>of</strong> the histone<br />

protein<br />

Removal <strong>of</strong> the<br />

acetyl group<br />

from the lysine<br />

residue at the N<br />

terminus <strong>of</strong> the<br />

histone protein<br />

Addition <strong>of</strong><br />

phosphate group<br />

Removal <strong>of</strong><br />

phosphate group<br />

Addition <strong>of</strong> 1, 2, or<br />

3 methyl groups<br />

on either lysine<br />

or arginine<br />

residues<br />

Removal <strong>of</strong> methyl<br />

groups<br />

Covalent<br />

attachment<br />

<strong>of</strong> ubiquitin<br />

Removal <strong>of</strong><br />

ubiquitin<br />

Covalent<br />

attachment<br />

<strong>of</strong> SUMO to<br />

lysine residues<br />

HATs<br />

HDACs<br />

Histone kinases<br />

Histone<br />

phosphatases<br />

HMTs<br />

Cytochrome P450<br />

family <strong>of</strong> liver<br />

enzymes<br />

Enzymatic cascade,<br />

including<br />

ubiquitin ligase,<br />

E1 ubiquitin<br />

activating<br />

enzyme, and<br />

ATP<br />

Isopeptidases<br />

Enzymatic cascade,<br />

including SUMOactivating<br />

enzyme<br />

and ATP<br />

Enhances transcription<br />

(removes positive<br />

charge and reduces<br />

affinity between<br />

histones and DNA)<br />

Represses transcription<br />

Enhances transcription<br />

(neutralizes histones’<br />

basic charge and<br />

reduces affinity to<br />

DNA)<br />

Represses transcription<br />

Silences genes<br />

Activates genes<br />

Labels protein for<br />

degradation<br />

Reverses<br />

ubiquitination<br />

Can mediate gene<br />

silencing through<br />

recruitment <strong>of</strong> HDAC<br />

and heterochromatin<br />

protein 1<br />

Abbreviations: HATs, histone acetyltransferases; HDACs, histone deacetylases; HMT, histone<br />

methyltranferases; SUMO, small ubiquitin-like modifier; ATP, adenosine triphosphate.


The Role <strong>of</strong> Epigenetics in <strong>Breast</strong> <strong>Cancer</strong> 49<br />

suppressor genes through mutation or deletion. Epigenetic modifications, such<br />

as DNA methylation, serve a similar role to somatic mutations, leading to<br />

gene silencing. Overall, tumors tend to exhibit hypomethylation <strong>of</strong> repeat<br />

elements and pericentromeric regions and hypermethylation <strong>of</strong> normally<br />

unmethylated CpG islands. Gene-specific hypermethylation <strong>of</strong> CpG islands is<br />

associated with chromosomal instability, resulting in increased mutation rates<br />

and abnormal gene expression. For example, hypermethylation <strong>of</strong> critical<br />

genes, such as tumor suppressors, adhesion molecules, DNA repair genes,<br />

inhibitors <strong>of</strong> angiogenesis, and inhibitors <strong>of</strong> metastasis, is observed in<br />

tumorigenesis (7). Hypermethylation <strong>of</strong> DNA repair genes (e.g., hMLH1) is<br />

associated with genomic instability through disruption <strong>of</strong> chromosome replication<br />

control (7). Unlike gene mutations, DNA methylation can be reversed<br />

by DNMT inhibitors.<br />

An important component <strong>of</strong> the cancer epigenome is an altered DNA<br />

methylation pattern <strong>of</strong> global demethylation and localized promoter hypermethylation.<br />

Aberrant methylation patterns are seen in almost every cancer type<br />

(7,16). In addition to aberrant methylation, tumors also show altered expression<br />

<strong>of</strong> HATs, making them another potential target for therapy (11). It is believed<br />

that histone modifications precede promoter hypermethylation to initiate transcriptional<br />

silencing <strong>of</strong> tumor suppressor genes. Repressive histone modifications<br />

are first established. Over time the density <strong>of</strong> DNA methylation<br />

increases in the gene promoter regions, which creates a “permanent” mode <strong>of</strong><br />

silencing (17).<br />

THE ROLE OF EPIGENETICS IN BREAST CANCER<br />

In the United States, breast cancer is one <strong>of</strong> the most common cancers, with<br />

approximately 200,000 women diagnosed with breast cancer each year (http://<br />

www.cancer.org). In addition to well-understood DNA mutations in breast<br />

tumors, epigenetic alterations <strong>of</strong> gene expression are key contributors (16,18–24).<br />

A large number <strong>of</strong> genes are found to be hypermethylated in breast tumors but<br />

not in normal breast tissue. Many <strong>of</strong> these genes have well-characterized functions<br />

in tumor development. Genes commonly methylated in breast cancer and<br />

their potential roles in tumor development are listed in Table 2. Four key genes,<br />

RASSF1A, BRCA1, ESR1, and ESR2, are further discussed in detail.<br />

RASSF1A<br />

The CpG island <strong>of</strong> the Ras-associated domain family 1A gene (RASSF1) is<br />

hypermethylated in 60% to 77% <strong>of</strong> breast cancers, and its transcripts are frequently<br />

inactivated in cancer cell lines and primary tissues (13,22). When cells<br />

lacking RASSF1A expression are treated with a DNMT inhibitor, such as 5-aza-<br />

2 0 -deoxycytidine, expression can be reactivated (25). RASSF1A is involved in


50 Dworkin et al.<br />

Table 2<br />

Genes Methylated in <strong>Breast</strong> <strong>Cancer</strong><br />

Gene <strong>Cancer</strong> hallmark contribution Reference<br />

RASSF1A Apoptosis 15<br />

CYP26A1 Microenvironment, encodes for cytochrome p450 enzymes 17<br />

KCNAB1 Microenvironment 17<br />

SNCA Microenvironment 17<br />

RARb Growth regulation 15<br />

TWIST Cell lineage and differentiation 15<br />

Cyclin D2 Cell cycle regulation 15<br />

APC Cell cycle regulation and apoptosis 8<br />

BRCA1 Cell cycle regulation 8,15<br />

BRCA2 Tumor suppressor gene 15<br />

p16 Cell cycle regulation 8,15<br />

DAPK Apoptosis 8<br />

GSTP1 Cellular metabolism and carcinogen detoxification 8<br />

TMS1 Apoptosis 8<br />

PR Growth regulation 8<br />

CDH1 Angiogenesis 8<br />

WWOX Apoptosis 38<br />

growth regulation and apoptotic pathways. Specifically, RASSF1A is a Ras<br />

effector and induces apoptosis through its interactions with pro-apoptotic kinase<br />

MST1. Mouse knockout studies show that RASSF1a –/– mice are prone to<br />

spontaneous development <strong>of</strong> lung adenomas, lymphomas, and breast adenocarcinomas.<br />

Knockout RASSF1A mice are prone to early spontaneous tumorigenesis<br />

and show a severe tumor susceptibility phenotype compared to wild-type<br />

mice (25). In addition to breast tumors, hypermethylation <strong>of</strong> RASSF1A can be<br />

detected in nonmalignant breast cells and patient sera. In one study, RASSF1A<br />

hypermethylation in sera <strong>of</strong> breast cancer patients was detected in six out <strong>of</strong> 26<br />

cases. These data suggest that RASSF1A hypermethylation may be a promising<br />

biomarker to screen putative cancer patients (25).<br />

BRCA1<br />

Germline mutations <strong>of</strong> BRCA1 and BRCA2 are responsible for familial breast<br />

cancers. Somatic mutations in sporadic cases are rare, but chromosomal losses<br />

occur in 30% to 50% <strong>of</strong> sporadic tumors, respectively (26). BRCA1 acts as a<br />

tumor suppressor gene for both breast and ovarian cancer (16). It encodes a<br />

multifunctional protein involved in DNA repair, cell cycle checkpoint control,<br />

protein ubiquitinylation, and chromatin remodeling (19). In vitro studies showed<br />

that decreased BRCA1 expression in cells led to increased levels <strong>of</strong> tumor<br />

growth, while increased expression <strong>of</strong> BRCA1 led to growth arrest and apoptosis.<br />

Inactivation <strong>of</strong> BRCA1 by promoter methylation is seen in 7% to 31% <strong>of</strong>


The Role <strong>of</strong> Epigenetics in <strong>Breast</strong> <strong>Cancer</strong> 51<br />

sporadic breast cancer cases (27). It is believed that aberrant methylation pairs<br />

with loss <strong>of</strong> heterozygosity to reduce BRCA1 expression in invasive sporadic<br />

breast tumors (16,19,22,26,28). The magnitude <strong>of</strong> the decrease <strong>of</strong> functional<br />

BRCA1 protein correlates with disease prognosis (19,22). <strong>Ph</strong>enotypically,<br />

BRCA1-methylated tumors are similar to tumors with germline mutations.<br />

Higher-grade tumors tend to show complete loss <strong>of</strong> BRCA1 protein. In contrast to<br />

BRCA1, BRCA2 does not show as high a degree <strong>of</strong> promoter methylation (16).<br />

ESR1 and ESR2<br />

Along with other genes, the estrogen receptors (ERs) ERa and ERb have been<br />

implicated in breast cancer development. ERa is encoded by ESR1, and when<br />

estrogen is activated, it stimulates cell proliferation. ERb is encoded by ESR2<br />

and is known to inhibit the proliferation and invasion <strong>of</strong> breast cancer cells. ERa<br />

and ERb both have promoter-associated CpG islands that can be abnormally<br />

methylated in breast cancer, but ESR2 methylation has been less well studied<br />

(16,18). Almost all breast cancers show some degree <strong>of</strong> DNA methylation <strong>of</strong> the<br />

ESR1 gene promoter, but this methylation is only associated with gene repression<br />

in ~30% <strong>of</strong> tumors (16,18). Approximately 66% <strong>of</strong> breast cancers express<br />

ERa. A fraction <strong>of</strong> breast cancers that are initially ERa positive lose ER<br />

expression during tumor progression, but it is unclear if this is due to methylation<br />

or other causes (18).<br />

The presence or absence <strong>of</strong> ERs plays an important role in both therapy<br />

and survival (16,18). ER-positive breast cancer can be treated with antiestrogenic<br />

drugs, such as tamoxifen, which binds to the ER, preventing estrogen from<br />

binding. This results in a decrease in growth <strong>of</strong> ER-positive breast cancer cells.<br />

ER-negative cells are no longer responsive to estrogen; therefore, antiestrogenic<br />

drugs have no effect (18). Reversal <strong>of</strong> ER-negative status is an attractive target<br />

for breast cancer therapy.<br />

BIOMARKERS AND DETECTION<br />

Gene-specific epigenetic changes for breast cancer are likely to occur early in<br />

tumorigenesis and have potential to be used as early detection and prevention<br />

methods (2). As epigenetic modifications become clear biomarkers for breast<br />

cancer treatment and therapeutic intervention, it is important to understand the<br />

many different techniques available for studying and detecting the presence <strong>of</strong><br />

methylation, histone modifications, and microRNAs (11).<br />

New, promising high-throughput methylation detection methods are<br />

available, which enable researchers and clinicians to identify an “epigenetic<br />

signature” specific to breast cancer. Biomarkers based on unique DNA methylation<br />

and histone modification patterns <strong>of</strong> breast tumors may be used in the<br />

future for diagnosis and early detection. Currently, breast cancer detection relies<br />

on various screening methods. Cytology is the main standard for identification


52 Dworkin et al.<br />

<strong>of</strong> abnormalities typical <strong>of</strong> cellular transformation. Quantitative multiplex<br />

methylation-specific polymerase chain reaction (PCR) (QM-MSP) is a highly<br />

sensitive method to quantitate cumulative gene promoter hypermethylation in<br />

samples where DNA is limiting. First, gene-specific primer pairs are used to<br />

coamplify genes, independent <strong>of</strong> their methylation status. Next, gene-specific<br />

primers (methylated and unmethylated) conjugated with fluorescent labels are<br />

combined with the first PCR product. The signal generated is proportional to the<br />

extent <strong>of</strong> DNA amplification (15). Since it is known that multigene methylation<br />

<strong>of</strong> CpG islands is common in early breast cancer, a panel <strong>of</strong> nine genes were<br />

tested for their sensitivity and specificity <strong>of</strong> detecting premalignant changes<br />

using QM-MSP. These genes included RASSF1A, RARb, TWIST, HIN1, Cyclin<br />

D2, APC, BRCA1, BRCA2, and p16. Tests were done using ductal lavage, nipple<br />

aspiration fluids, and fine needle aspirates.<br />

Using this detection panel, the rate <strong>of</strong> detection <strong>of</strong> breast cancer cells rises<br />

from 43% sensitivity with cytologic examination alone to 71.4% sensitivity (15).<br />

Hypermethylation <strong>of</strong> genes that are commonly methylated in breast cancer in<br />

sera in breast cancer patients has also been used to detect early malignant<br />

changes. Sera methylation can be detected using methylation-specific PCR (MS-<br />

PCR), a noninvasive PCR-based technique. Early studies indicate that MS-PCR<br />

is a promising approach to screen putative cancer patients (25). This technique<br />

utilizes primers designed for methylated or unmethylated bisulfite-modified<br />

DNA (29). RASSF1A gene hypermethylation has been detected using MS-PCR<br />

in sera <strong>of</strong> ovarian cancer patients, with 100% specificity (25).<br />

Recurrence rates for breast cancer in the same breast have been recorded as<br />

high as 40% despite negative pathological margins. Researchers believe that a<br />

primary tumor may serve as a locus from which methylation density progressively<br />

diffuses outward to surrounding tissues. CYP26A1, KCNAB1, RASSF1A,<br />

and SNCA have found to be frequently methylated in adjacent normal tissue. A<br />

methylation pr<strong>of</strong>ile containing genes such as these, or those identified in other<br />

studies (Table 2), could be used by clinicians to determine whether patients<br />

should undergo radiological therapy or other treatments to prevent future<br />

recurrence (17).<br />

MassARRAY and pyrosequencing, two additional methods <strong>of</strong> detecting<br />

methylation, are currently being developed for use in clinical settings.<br />

MassARRAY is a highly sensitive technique that uses base-specific cleavage and<br />

matrix-assisted laser desorption/ionization time-to-flight mass spectrometry<br />

(MALDI-TOF MS). It is capable <strong>of</strong> detecting DNA methylation levels as low as<br />

5%. It is suitable for testing methylation patterns on various sources, including<br />

archival tissues and laser capture microdissected specimens (1). Pyrosequencing<br />

is a locus-specific quantitative method that utilizes the detection <strong>of</strong> pyrophosphate,<br />

which is liberated from incorporated nucleotides by DNA polymerase<br />

during strand elongation. Free pyrophosphates are converted to adenosine triphosphate<br />

(ATP), which provides energy for the oxidation <strong>of</strong> luciferin to then<br />

generate light. Nucleotides are added sequentially to enable base calling.


The Role <strong>of</strong> Epigenetics in <strong>Breast</strong> <strong>Cancer</strong> 53<br />

Pyrosequencing has two major advantages over MS-PCR. First, the data are<br />

actual nucleotide sequences rather than fluorescence data. Second, pyrosequencing<br />

can detect partially methylated sequences that are outside <strong>of</strong> the priming<br />

sites (30).<br />

Commonly used methods for detecting histone modifications usually rely<br />

on chromatin immunoprecipitation (ChIP). In ChiP, DNA is cross-linked to<br />

chromatin, followed by co-immunoprecipitation using antibodies against specific<br />

histone residues or chromatin proteins. “ChIP” DNA can be used in microarrays,<br />

cloning, and sequencing (11). Techniques for detecting histone modifications<br />

have not yet been adapted for clinical application.<br />

CLINICAL APPLICATION<br />

<strong>Breast</strong> cancer prevention, treatments, and diagnostics are being developed to<br />

target epigenetic changes leading to breast cancer. Treatments for breast cancer<br />

currently being studied extensively focus on reversing aberrant DNA methylation<br />

and histone acetylation <strong>of</strong> tumor suppressor genes and genes involved in<br />

therapeutic response. Combinations <strong>of</strong> epigenetic-targeted therapies with conventional<br />

chemotherapeutic agents may provide a way to resensitize drugresistant<br />

tumors (Table 3).<br />

Methylation Therapy<br />

DNA methylation in proliferating cells is dependent on continued expression <strong>of</strong><br />

DNMTs. Treatments that focus on inhibiting the expression <strong>of</strong> these proteins<br />

would therefore result in progressive reduction in DNA methylation in newly<br />

divided cells. There are three main classes <strong>of</strong> DNA methylation inhibitors:<br />

nucleoside inhibitors, nonnucleoside weak inhibitors, and rationally designed<br />

inhibitors. Nucleoside inhibitors trap DNMTs for degradation. However, they<br />

require DNA incorporation and active DNA synthesis, which limits the drug in<br />

hypoproliferating cells. 5-Azacytidine (5-aza-dC) and its deoxy analog are two<br />

<strong>of</strong> the first hypomethylating agents approved by the U.S. Food and Drug<br />

Administration for cancer treatment (31). They work by two mechanisms: (i) as<br />

demethylators by incorporating into CpG sites opposite a methylated CpG site on<br />

the template strand and (ii) by binding to the cytosine in the catalytic domain <strong>of</strong><br />

the DNMT1 enzyme. Within a few hours <strong>of</strong> treatment with 5-aza-dC, there is<br />

rapid passive loss <strong>of</strong> methylation (32). Nonnucleoside inhibitors are not as well<br />

understood and are limited by their low level <strong>of</strong> hypomethylation induction.<br />

Nonnucleoside inhibitors were first discovered for their use in reducing autoreactivity<br />

for autoimmune diseases. During this study researchers discovered their<br />

potential as demethylating agents. Researchers are currently using clinically<br />

approved nonnucleoside inhibitors as a starting point to study their use as<br />

demethylating agents. Rationally designed inhibitors are also in development.<br />

These drugs are designed on the basis <strong>of</strong> knowledge <strong>of</strong> specific chemical


54 Dworkin et al.<br />

Table 3<br />

Drugs in Use for <strong>Breast</strong> <strong>Cancer</strong> Treatment<br />

Nucleoside inhibitors<br />

Drug Use Research level Reference<br />

5-Azacytidine Trap DNMTs for degradation Clinical use 31,40<br />

5-aza-<br />

Trap DNMTs for degradation Clinical use 31<br />

2 0 deoxycytidine<br />

Zebularine Unknown mechanism Shown promise in vitro 31<br />

5 0 -fluoro-<br />

2 0 deoxycytidine<br />

Unknown mechanism In clinical trials 31<br />

Nonnucleoside inhibitors<br />

Drug Use Research level Reference<br />

Procainamide<br />

Hydralazine<br />

Mechanism not well<br />

understood but proven to<br />

decrease expression <strong>of</strong><br />

DNMT1 and DNMT3a<br />

Mechanism not well<br />

understood but proven to<br />

decrease expression <strong>of</strong><br />

DNMT1 and DNMT3a<br />

Rationally designed inhibitors<br />

Clinically used as<br />

an antihypertensive<br />

drug<br />

Clinically used as a<br />

vasodilator<br />

31,41,42<br />

31,41,42<br />

Drug Use Research level Reference<br />

NSC303530 Blocks the active site <strong>of</strong> DNMTs Basic research 31,43<br />

NSC401077 Blocks the active site <strong>of</strong> DNMTs Basic research 31,43<br />

HDAC inhibitors<br />

Drug Use Advantages Disadvantages<br />

Research<br />

level<br />

Reference<br />

LAQ824<br />

Magnesium<br />

Valporate<br />

and<br />

Hydralazine<br />

HDAC<br />

inhibitor<br />

N-hydroxy-2- HDAC<br />

propenamide inhibitor<br />

Unknown<br />

HDAC and Increases<br />

methyltransferase<br />

conventional<br />

efficacy <strong>of</strong><br />

inhibitor cytotoxic<br />

agents<br />

Suppresses<br />

mammary<br />

tumor growth<br />

in cells<br />

Important to<br />

pay attention<br />

to schedule <strong>of</strong><br />

administration<br />

Unknown<br />

Unknown<br />

Preclinical<br />

level; studies<br />

on hematological<br />

malignancies<br />

<strong>Ph</strong>ase III<br />

clinical<br />

trials<br />

Research<br />

level<br />

40<br />

44<br />

l45<br />

Abbreviations: DNMTs, DNA methyltransferases; HDAC, histone deacetylase.


The Role <strong>of</strong> Epigenetics in <strong>Breast</strong> <strong>Cancer</strong> 55<br />

Figure 3 Drug impact on histone modifications. HDAC modifies histones and leads to<br />

chromatin remodeling so that genes are transcriptionally inactive. HAT can reverse the<br />

effects <strong>of</strong> HDAC to acetylate histones and open the confirmation into an active state.<br />

Drugs such as Scriptaid, TSA, LAQ824, and N-hydroxy-2-propenamide can act on this<br />

pathway to block HDACs and keep DNA transcriptionally active. Abbreviations: HDAC,<br />

histone deactylase; HAT, histone acetyltransferase; TSA, trichostatin A. Source: Cartoon<br />

compliments <strong>of</strong> Kristi Bennett.<br />

responses in the body, tailoring combinations <strong>of</strong> these to fit a treatment pr<strong>of</strong>ile.<br />

An example <strong>of</strong> a rational design involves the use <strong>of</strong> three-dimensional information<br />

to target the drug. Limitations <strong>of</strong> DNA methylation inhibitors include the<br />

cooperation between different DNMTs such that the drugs must inhibit several <strong>of</strong><br />

them simultaneously. All demethylating agents suffer from nonspecificity and<br />

toxicity. The next phase <strong>of</strong> drug development is how to target gene-specific<br />

hypomethylation (e.g., unmethylated oligonucleotides) (31).<br />

Histone Modification Therapy<br />

Histone deacetylases (HDACs) are also targets for therapy (Fig. 3). HDACs are<br />

important because <strong>of</strong> their role in cell cycle arrest and inhibition <strong>of</strong> apoptosis.<br />

These drugs have been studied in isolation and synergistically with DNMTs.<br />

Therapies that target both DNMTs and HDACs act globally, which can cause<br />

reactivation <strong>of</strong> other genes that are typically silenced, such as oncogenes (18,33–37).<br />

ER Therapy<br />

Drugs are in development for different types <strong>of</strong> breast cancer. A therapy that could<br />

restore gene expression <strong>of</strong> ERS1 (for ER-negative cancer patients) could reestablish<br />

cancer cell growth regulation through estrogen. After reexpression <strong>of</strong><br />

ERS1, antiestrogenic drugs could subsequently be used. A number <strong>of</strong> drugs,<br />

including DNMT inhibitors and HDAC inhibitors, are being used to reactivate ER<br />

expression (Table 4). DNMTs and HDACs work synergistically to silence gene<br />

expression <strong>of</strong> ER. One problem found in treatments that target DNMTs and HDACs<br />

is that alone they may not be enough to reverse ER-promoter hypermethylation.


56 Dworkin et al.<br />

Table 4<br />

Drugs Involved in Reactivation <strong>of</strong> ER Expression<br />

Treatment Result Reference<br />

5-aza-Dc<br />

(methyltransferase<br />

inhibitor)<br />

Antisense oligonucleotides<br />

(for DNMT1)<br />

Partial demethylation <strong>of</strong> ER CpG islands,<br />

reexpression <strong>of</strong> ER mRNA, and synthesis<br />

<strong>of</strong> functional ER protein<br />

ER gene reexpression and estrogen<br />

responsiveness reestablished<br />

18,36,37,46<br />

18,35<br />

TSA (HDAC inhibitor) Reexpression <strong>of</strong> ERa protein 18,33<br />

DNMT and HDAC Reactivation ERa protein and expression 18,37<br />

inhibitor<br />

Scriptaid (HDAC inhibitor) Reexpression <strong>of</strong> functional ER in vivo 18,34<br />

Scriptaid and 5-aza-Dc More effective in inducing ER expression<br />

than just Scriptaid or 5-aza-Dc alone<br />

18,34<br />

Abbreviations: ER, estrogen receptor; DNMT, DNA methyltransferase; HDAC, histone deacetylase;<br />

TSA, trichostatin A.<br />

Tamoxifen is a commonly used drug for ERa-positive breast cancers.<br />

Unfortunately, acquired resistance <strong>of</strong> tamoxifen occurs in nearly 40% <strong>of</strong> patients.<br />

Therefore, it is important to identify patients most likely to respond to<br />

Tamoxifen and to identify individuals likely to acquire resistance. One promising<br />

biomarker for Tamoxifen response is WWOX. Preliminary studies show<br />

that WWOX expression levels predict tamoxifen resistance better than the two<br />

previously known biomarkers, PR and HER2 (38). WWOX appears to mediate<br />

tamoxifen sensitivity, and its expression is reduced in a large fraction (63.2%) <strong>of</strong><br />

breast cancers (39). The primary mechanism <strong>of</strong> downregulation <strong>of</strong> WWOX is<br />

through DNA hypermethylation <strong>of</strong> its regulatory region.<br />

FUTURE DIRECTIONS<br />

The next 10 years are likely to bring significant advances for breast cancer<br />

treatment. Therapeutics that target methylation and histone modifications may<br />

play a leading role in treating breast cancer. Since epigenetic modifications can<br />

also be used as biomarkers, targeted therapies may some day be used as preventative<br />

measures.<br />

REFERENCES<br />

1. Brena RM, Huang TH, Plass C. Quantitative assessment <strong>of</strong> DNA methylation:<br />

potential applications for disease diagnosis, classification, and prognosis in clinical<br />

settings. J Mol Med 2006; 84(5):365–377.<br />

2. Balch C, Montgomery JS, Paik HI, et al. New anti-cancer strategies: epigenetic<br />

therapies and biomarkers. Front Biosci 2005; 10:1897–1931.<br />

3. Luczak MW, Jagodzinski PP. The role <strong>of</strong> DNA methylation in cancer development.<br />

Folia Histochem Cytobiol 2006; 44(3):143–154.


The Role <strong>of</strong> Epigenetics in <strong>Breast</strong> <strong>Cancer</strong> 57<br />

4. Miyamoto K, Ushijima T. Diagnostic and therapeutic applications <strong>of</strong> epigenetics. Jpn<br />

J Clin Oncol 2005; 35(6):293–301.<br />

5. Reamon-Buettner SM, Borlak J. A new paradigm in toxicology and teratology:<br />

altering gene activity in the absence <strong>of</strong> DNA sequence variation. Reprod Toxicol<br />

2007; 24(1):20–30.<br />

6. Szyf M. Therapeutic implications <strong>of</strong> DNA methylation. Future Oncol 2005; 1(1):<br />

125–135.<br />

7. Ting AH, McGarvey KM, Baylin SB. The cancer epigenome—components and<br />

functional correlates. Genes Dev 2006; 20(23):3215–3231.<br />

8. Widschwendter M, Jones PA. DNA methylation and breast carcinogenesis. Oncogene<br />

2002; 21(35):5462–5482.<br />

9. Hellebrekers DM, Griffioen AW, van Engeland M. Dual targeting <strong>of</strong> epigenetic<br />

therapy in cancer. Biochim Biophys Acta 2007; 1775(1):76–91.<br />

10. Hellebrekers DM, Jair KW, Viré E, et al. Angiostatic activity <strong>of</strong> DNA methyltransferase<br />

inhibitors. Mol <strong>Cancer</strong> Ther 2006; 5(2):467–475.<br />

11. Curt Balch TH-MH, Nephew KP. High-throughput assessments <strong>of</strong> epigenetics in<br />

human disease. In: Kim S, Tang H, Mardis ER, eds. Advances in Genome Sequencing<br />

Technology and Algorithms, Norwood, MA: Artech House Publishers, 2007.<br />

12. Han J, Kim D, Morris KV. Promoter-associated RNA is required for RNA-directed<br />

transcriptional gene silencing in human cells. Proc Natl Acad Sci U.S.A. 2007; 104(30):<br />

12422–12427.<br />

13. Oka M, Rodic N, Graddy J, et al. CpG sites preferentially methylated by Dnmt3a in<br />

vivo. J Biol Chem 2006; 281(15):9901–9908.<br />

14. Ooi SK, Qiu C, Bernstein E, et al. DNMT3L connects unmethylated lysine 4 <strong>of</strong><br />

histone H3 to de novo methylation <strong>of</strong> DNA. Nature 2007; 448(7154):714–717.<br />

15. Fackler MJ, Malone K, Zhang Z, et al. Quantitative multiplex methylation-specific<br />

PCR analysis doubles detection <strong>of</strong> tumor cells in breast ductal fluid. Clin <strong>Cancer</strong> Res<br />

2006; 12(11 Pt 1):3306–3310.<br />

16. Campan M, Weisenberger DJ, Laird PW. DNA methylation pr<strong>of</strong>iles <strong>of</strong> female steroid<br />

hormone-driven human malignancies. Curr Top Microbiol Immunol 2006;<br />

310:141–178.<br />

17. Yan PS, Venkataramu C, Ibrahim A, et al. Mapping geographic zones <strong>of</strong> cancer risk<br />

with epigenetic biomarkers in normal breast tissue. Clin <strong>Cancer</strong> Res 2006; 12(22):<br />

6626–6636.<br />

18. Giacinti L, Claudio PP, Lopez M, et al. Epigenetic information and estrogen receptor<br />

alpha expression in breast cancer. Oncologist 2006; 11(1):1–8.<br />

19. Mirza S, Sharma G, Prasad CP, et al. Promoter hypermethylation <strong>of</strong> TMS1, BRCA1,<br />

ERalpha and PRB in serum and tumor DNA <strong>of</strong> invasive ductal breast carcinoma<br />

patients. Life Sci 2007; 81(4):280–287.<br />

20. Sharma D, Blum J, Yang X, et al. Release <strong>of</strong> methyl CpG binding proteins and histone<br />

deacetylase 1 from the Estrogen receptor alpha (ER) promoter upon reactivation in<br />

ER-negative human breast cancer cells. Mol Endocrinol 2005; 19(7):1740–1751.<br />

21. Sui M, Huang Y, Park BH, et al. Estrogen receptor alpha mediates breast cancer cell<br />

resistance to paclitaxel through inhibition <strong>of</strong> apoptotic cell death. <strong>Cancer</strong> Res 2007;<br />

67(11):5337–5344.<br />

22. Vincent-Salomon A, Ganem-Elbaz C, Manié E, et al. X inactive-specific transcript<br />

RNA coating and genetic instability <strong>of</strong> the X chromosome in BRCA1 breast tumors.<br />

<strong>Cancer</strong> Res 2007; 67(11):5134–5140.


58 Dworkin et al.<br />

23. Visvanathan K, Sukumar S, Davidson NE. Epigenetic biomarkers and breast cancer:<br />

cause for optimism. Clin <strong>Cancer</strong> Res 2006; 12(22):6591–6593.<br />

24. Zhou Q, Davidson NE. Silencing estrogen receptor alpha in breast cancer cells.<br />

<strong>Cancer</strong> Biol Ther 2006; 5(7):848–849.<br />

25. Pfeifer GP, Dammann R. Methylation <strong>of</strong> the tumor suppressor gene RASSF1A in<br />

human tumors. Biochemistry (Mosc) 2005; 70(5):576–583.<br />

26. Esteller M, Silva JM, Dominguez G, et al. Promoter hypermethylation and BRCA1<br />

inactivation in sporadic breast and ovarian tumors. J Natl <strong>Cancer</strong> Inst 2000; 92(7):<br />

564–569.<br />

27. Wei M, Grushko TA, Dignam J, et al. BRCA1 promoter methylation in sporadic<br />

breast cancer is associated with reduced BRCA1 copy number and chromosome 17<br />

aneusomy. <strong>Cancer</strong> Res 2005; 65(23):10692–10699.<br />

28. Catteau A, Morris JR. BRCA1 methylation: a significant role in tumour development?<br />

Semin <strong>Cancer</strong> Biol 2002; 12(5):359–371.<br />

29. Herman JG, Graff JR, Myöhänen S, et al. Methylation-specific PCR: a novel PCR<br />

assay for methylation status <strong>of</strong> CpG islands. Proc Natl Acad Sci U.S.A. 1996; 93(18):<br />

9821–9826.<br />

30. Shames DS, Minna JD, Gazdar AF. Methods for detecting DNA methylation in<br />

tumors: from bench to bedside. <strong>Cancer</strong> Lett 2007; 251(2):187–198.<br />

31. Issa JP. DNA methylation as a therapeutic target in cancer. Clin <strong>Cancer</strong> Res 2007; 13(6):<br />

1634–1637.<br />

32. Christman JK. 5-Azacytidine and 5-aza-2 0 -deoxycytidine as inhibitors <strong>of</strong> DNA<br />

methylation: mechanistic studies and their implications for cancer therapy. Oncogene<br />

2002; 21(35):5483–5495.<br />

33. Kawai H, Li H, Avraham S, et al. Overexpression <strong>of</strong> histone deacetylase HDAC1<br />

modulates breast cancer progression by negative regulation <strong>of</strong> estrogen receptor<br />

alpha. Int J <strong>Cancer</strong> 2003; 107(3):353–358.<br />

34. Keen JC, Yan L, Mack KM, et al. A novel histone deacetylase inhibitor, scriptaid,<br />

enhances expression <strong>of</strong> functional estrogen receptor alpha (ER) in ER negative<br />

human breast cancer cells in combination with 5-aza 2 0 -deoxycytidine. <strong>Breast</strong> <strong>Cancer</strong><br />

Res Treat 2003; 81(3):177–186.<br />

35. Yan L, Nass SJ, Smith D, et al. Specific inhibition <strong>of</strong> DNMT1 by antisense oligonucleotides<br />

induces re-expression <strong>of</strong> estrogen receptor-alpha (ER) in ER-negative<br />

human breast cancer cell lines. <strong>Cancer</strong> Biol Ther 2003; 2(5):552–556.<br />

36. Yang X, Ferguson AT, Nass SJ, et al. Transcriptional activation <strong>of</strong> estrogen receptor<br />

alpha in human breast cancer cells by histone deacetylase inhibition. <strong>Cancer</strong> Res<br />

2000; 60(24):6890–6894.<br />

37. Yang X, <strong>Ph</strong>illips DL, Ferguson AT, et al. Synergistic activation <strong>of</strong> functional<br />

estrogen receptor (ER)-alpha by DNA methyltransferase and histone deacetylase<br />

inhibition in human ER-alpha-negative breast cancer cells. <strong>Cancer</strong> Res 2001; 61(19):<br />

7025–7029.<br />

38. Guler G, Iliopoulos D, Guler N, et al. Wwox and Ap2g expression levels predict<br />

tamoxifen response. Clin <strong>Cancer</strong> Res 2007; 13(20).<br />

39. Guler G, Uner A, Guler N, et al. The fragile genes FHIT and WWOX are inactivated<br />

coordinately in invasive breast carcinoma. <strong>Cancer</strong> 2004; 100(8):1605–1614.<br />

40. Hurtubise A, Momparler RL. Effect <strong>of</strong> histone deacetylase inhibitor LAQ824 on<br />

antineoplastic action <strong>of</strong> 5-Aza-2 0 -deoxycytidine (decitabine) on human breast carcinoma<br />

cells. <strong>Cancer</strong> Chemother <strong>Ph</strong>armacol 2006; 58(5):618–625.


The Role <strong>of</strong> Epigenetics in <strong>Breast</strong> <strong>Cancer</strong> 59<br />

41. Chuang JC, Yoo CB, Kwan JM, et al. Comparison <strong>of</strong> biological effects <strong>of</strong> nonnucleoside<br />

DNA methylation inhibitors versus 5-aza-2 0 -deoxycytidine. Mol <strong>Cancer</strong><br />

Ther 2005; 4(10):1515–1520.<br />

42. Cornacchia E, Golbus J, Maybaum J, et al. Hydralazine and procainamide inhibit<br />

T cell DNA methylation and induce autoreactivity. J Immunol 1988; 140(7):<br />

2197–2200.<br />

43. Siedlecki P, Boy RG, Musch T, et al. Discovery <strong>of</strong> two novel, small-molecule<br />

inhibitors <strong>of</strong> DNA methylation. J Med Chem 2006; 49(2):678–683.<br />

44. Arce C, Pérez-Plasencia C, Gonzälez-Fierro A, et al. A pro<strong>of</strong>-<strong>of</strong>-principle study <strong>of</strong><br />

epigenetic therapy added to neoadjuvant doxorubicin cyclophosphamide for locally<br />

advanced breast cancer. PLoS ONE 2006; 1:e98.<br />

45. Joung KE, Min KN, An JY, et al. Potent in vivo anti-breast cancer activity <strong>of</strong> IN-<br />

2001, a novel inhibitor <strong>of</strong> histone deacetylase, in MMTV/c-Neu mice. <strong>Cancer</strong> Res<br />

2006; 66(10):5394–5402.<br />

46. Ferguson AT, Vertino PM, Spitzner JR, et al. Role <strong>of</strong> estrogen receptor gene<br />

demethylation and DNA methyltransferase. DNA adduct formation in 5-aza-<br />

2 0 deoxycytidine-induced cytotoxicity in human breast cancer cells. J Biol Chem<br />

1997; 272(51):32260–32206.


5<br />

Comparative Genomic Hybridization<br />

and Copy Number Abnormalities in<br />

<strong>Breast</strong> <strong>Cancer</strong><br />

Nicholas Wang and Joe Gray<br />

Life Sciences Division, Lawrence Berkeley National Laboratory,<br />

Berkeley, California, U.S.A.<br />

INTRODUCTION<br />

The progression <strong>of</strong> normal breast epithelial cells from a normal state toward one<br />

characterized by uncontrolled growth and metastatic behavior is caused by the<br />

deregulation <strong>of</strong> key cellular processes and signaling pathways. These alterations<br />

in normal cellular behavior are rooted in the accumulation <strong>of</strong> genomic and<br />

epigenomic lesions that impact hallmarks <strong>of</strong> cancer, such as the ability <strong>of</strong> the cell<br />

to control proliferation, undergo apoptosis, increase motility leading to invasion,<br />

and alter angiogenesis. A suite <strong>of</strong> technologies has now been developed to assess<br />

genomic and epigenomic aberrations that contribute to cancer progression. The<br />

application <strong>of</strong> these has shown that genome copy number abnormalities (CNAs)<br />

are among the most frequent genomic aberrations. Remarkably, these studies<br />

have revealed that 10% to 15% <strong>of</strong> the genes in a typical carcinoma tumor may be<br />

deregulated by recurrent genome CNAs and regions. Some <strong>of</strong> these genes<br />

influence disease progress and so may be assessed to facilitate prognosis. Others<br />

influence response to therapy and so may be assessed as predictive markers.<br />

Some <strong>of</strong> these enable oncogenic processes, on which tumors depend for survival,<br />

and so are candidate therapeutic targets. In most cases, array comparative<br />

61


62 Wang and Gray<br />

genomic hybridization (CGH) is the method <strong>of</strong> choice for their discovery and<br />

may be used in some setting for clinical assessments <strong>of</strong> these abnormalities.<br />

Accordingly, we review here several <strong>of</strong> the current array CGH technologies<br />

available and the considerations needed when determining the technology most<br />

applicable to a given study.<br />

CGH analyses <strong>of</strong> thousands <strong>of</strong> primary breast tumors and cell lines have<br />

now defined recurrent CNAs throughout the genome and identified genes<br />

encoded in these aberrant regions that are deregulated by the abnormalities (1,2–4).<br />

In breast cancer, for example, 10% to 15% <strong>of</strong> known genes are deregulated by<br />

these CNAs (2–4). The aberrations tend to fall into two general classes that may<br />

contribute in different ways to cancer pathophysiology: low-level CNAs that<br />

occur through the gain or loss <strong>of</strong> extended regions <strong>of</strong> single chromosomes or<br />

extended regions there<strong>of</strong> and high-level amplification or homozygous deletions<br />

that tend to be narrower in genomic extent.<br />

Most genes are deregulated by low-level aberrations. The contributions <strong>of</strong><br />

these genes to cancer pathophysiology are not well understood; however, evidence<br />

is now emerging that these may provide, en ensemble, a general metabolic<br />

advantage (4–6). These aberrations typically occur relatively early in the carcinogenesis<br />

process. It has been proposed that these aberrations are selected during<br />

successful transition through a telomere crisis (7).<br />

High-level amplification and narrow homozygous deletions provide strong<br />

evidence <strong>of</strong> active selection during cancer progression and seem to occur later<br />

during cancer progression. In breast cancer, regions <strong>of</strong> high-level amplification<br />

encode well-studied oncogenes, such as ERBB2 (17q12) (8), FGFR1 (8p11) (9),<br />

PVT1 and MYC (8q24) (10,11), CCND1 and EMS1 (11q13) (12), and ABI1 and<br />

ZNF217 (20q13) (13,14), while deletions contribute to the inactivation <strong>of</strong> tumor<br />

suppressor genes such as TP53 (17p13) (15), CDKN2A (9p21) (16), and BRCA1<br />

(13q14) (17). While it is <strong>of</strong>ten assumed that these oncogenes and tumor suppressor<br />

genes are the sole targets for these aberrations, a growing body <strong>of</strong><br />

functional data suggests that several genes in each region <strong>of</strong> recurrent abnormality<br />

may be coselected by the event. In breast cancer, for example, regions <strong>of</strong><br />

amplification in which more than one gene appears to be functionally important<br />

include 8p11, 8q24, 11q13, 17q21, and 20q13 (4). Table 1 summarizes published<br />

studies for genes in these regions <strong>of</strong> amplification. The concept <strong>of</strong> multigene<br />

selection is also supported by studies in mouse models, where the extents <strong>of</strong><br />

amplicons in mice and humans are similar. Similarities in the extent <strong>of</strong> amplification<br />

around ERBB2 in humans and in polyomavirus middle T-induced<br />

mammary tumors in mice are particularly striking (18). Similar studies are<br />

emerging at other organ sites (19). As a result, it is critically important that the<br />

extent <strong>of</strong> each region be defined as precisely as possible. High-resolution CGH<br />

analyses are now contributing to this process.<br />

Several recurrent high-level CNAs are associated with clinical endpoints,<br />

including reduced survival duration and response to aberration-targeted therapies.<br />

Assays for recurrent aberrations or genes deregulated by CNA abnormalities have


Comparative Genomic Hybridization in <strong>Breast</strong> <strong>Cancer</strong> 63<br />

Table 1<br />

Array CGH Analysis Platform Summary<br />

Platform<br />

composition Reference Density<br />

Allele<br />

specific<br />

Source<br />

cDNA 5 40K No Academic<br />

BAC 68 25K No Academic/Spectral<br />

Genomics<br />

Oligonucleotide http://affymetrix.com 950K Yes Affymetrix<br />

Oligonucleotide http://www.<br />

400K No Nimblegen<br />

nimblegen.com<br />

Oligonucleotide http://www.chem. 240K No Agilent<br />

agilent.com<br />

Oligonucleotide http://www.illumina. 1M Yes Illumina<br />

com<br />

MIP 41 20K Yes Academic<br />

Abbreviations: CGH, comparative genomic hybridization; BAC, bacterial artificial chromosome;<br />

MIP, molecular inversion probe.<br />

already been incorporated into molecular indicators <strong>of</strong> outcome. In addition,<br />

several therapeutic agents are now being developed to attack genes deregulated<br />

fully or in part by recurrent genome abnormalities so that assays for these targets<br />

may be useful as predictors <strong>of</strong> response. In breast cancer, the most prominent<br />

examples are trastuzumab (20,21) and lapatinib (22) (NEJM PMID 17192538),<br />

developed to attack tumors driven by amplification <strong>of</strong> ERBB2. In addition,<br />

therapies have been developed to treat tumors arising from deletion or inactivation<br />

<strong>of</strong> TP53 (23) and activation <strong>of</strong> elements <strong>of</strong> the PI3-kinase pathway, including<br />

PIK3CA, PTEN, AKT, and ZNF217 (24–27).<br />

The ability to detect and define the extents <strong>of</strong> CNAs has steadily improved<br />

with the evolution <strong>of</strong> CGH analysis technologies (Table 2). The first CGH<br />

studies mapped CNAs onto metaphase chromosomes, but the resolution <strong>of</strong> such<br />

analyses was limited to about 10 Mbp by the complex packaging <strong>of</strong> the DNA<br />

into chromosomes (28). The introduction <strong>of</strong> array CGH in which CNAs were<br />

mapped onto arrays <strong>of</strong> cloned probes substantially increased resolution. Early<br />

arrays comprised hundreds to thousands <strong>of</strong> probes and improved resolution to<br />

about 1 Mbp (29,30). Current cloned probe arrays interrogate the genome at<br />

approximately single-gene resolution (5). Today, several commercial platforms<br />

map abnormalities onto arrays <strong>of</strong> oligonucleotide probes with subgene resolution<br />

(either on flat substrates or attached to microbeads) (31), and some allow<br />

analyses in an allele-specific manner (32). Applications <strong>of</strong> these platforms are<br />

further refining the extents <strong>of</strong> recurrent genomic aberrations but also reveal the<br />

somewhat surprising existence <strong>of</strong> germline copy number polymorphisms that<br />

must be taken into account when interpreting CGH measurements. Several<br />

aspects <strong>of</strong> these important array CGH analysis technologies are reviewed in the<br />

following sections.


64 Wang and Gray<br />

Table 2<br />

Data Analysis Algorithms<br />

Platform Program Reference Type<br />

Required<br />

s<strong>of</strong>tware<br />

Two-color arrays aCGH 54 HMM R<br />

GLAD 56 Adaptive weight R<br />

smoothing<br />

CLAC 57 Clustering along R<br />

chromosomes<br />

DNAcopy 55 Segmentation R<br />

Single-color arrays dChipSNP 37 HMM none<br />

(Affymetrix)<br />

CARAT 69 none<br />

CNAG 70 Regression none<br />

GEMCA 71 Regression none<br />

Abbreviations: aCGH, array comparative genomic hybridization; HMM, hidden Markov model;<br />

GLAD, Gaussian model–based approach; CART, copy number analysis with regression and tree; CLAC,<br />

clustering along each chromosome; SNP, single nucleotide polymorphism; CNAG, Copy Number<br />

Analyzer for GeneChip; GEMCA, Genotyping microarray based copy number variation analysis.<br />

COMPARATIVE GENOMIC HYBRIDIZATION<br />

CGH is the primary technology used for assessment <strong>of</strong> genome copy number.<br />

Typically, analysis proceeds by hybridizing differentially labeled tumor and normal<br />

DNA to a representation <strong>of</strong> the normal genome, although some platforms<br />

perform tumor and normal reference hybridization in separate representations. The<br />

ratio <strong>of</strong> tumor label to normal label, hybridized to each locus on the representation<br />

is then measured as an estimate <strong>of</strong> relative copy number at that locus.<br />

The first CGH experiments employed metaphase chromosomes as the<br />

representation onto which information was mapped (28,33). While useful, the<br />

genomic resolution was limited by the nonlinear organization <strong>of</strong> DNA along<br />

metaphase chromosomes. Thus, chromosome CGH has now been largely supplanted<br />

by array CGH, in which information is mapped onto arrays <strong>of</strong> DNA<br />

probes (Fig. 1). The initial array CGH platforms comprised large genomic cloned<br />

probes, such as yeast artificial chromosomes (YACs) (34), bacterial artificial<br />

chromosomes (BACs), and P1-derived artificial chromosomes (PACs) (29), or<br />

shorter cDNA sequences spotted on glass slides (35). More recently, synthetic<br />

oligonucleotide probes have been used successfully (36–39). Several platforms<br />

are summarized in Table 1 and in the following sections.<br />

Cloned probe arrays are typically available from large academic cores with<br />

the experience and resources to maintain and print the arrays. These arrays are<br />

developed using the large collections <strong>of</strong> cDNAs (35) or BACs (29,30,40)<br />

developed during the course <strong>of</strong> the human genome program. The advantages <strong>of</strong>


Comparative Genomic Hybridization in <strong>Breast</strong> <strong>Cancer</strong> 65<br />

Figure 1 The use <strong>of</strong> a dual-color hybridization protocol for the detection <strong>of</strong> genome copy<br />

number aberrations. Both a normal sample (representing two copies <strong>of</strong> the genome) and a<br />

tumor sample are randomly labeled with differential fluorescent dyes. They are then<br />

cohybridized onto an array platform (BACs or oligonucleotides) with the subsequent log 2<br />

dye ratios <strong>of</strong> the tumor sample compared with the normal sample used to detect copy<br />

number changes. Abbreviation: BACs, bacterial artificial chromosomes.<br />

this approach are that the clones are readily available in the public domain and<br />

the signals produced during hybridization are large so that measurement precision<br />

is relatively high. The disadvantages are that the clones are not perfectly<br />

curated and mapped, so some may be mismapped. In addition, managing the<br />

large number is daunting, and the ultimate resolution <strong>of</strong> the analysis platforms is<br />

limited by the relatively genomic extents <strong>of</strong> the cloned probes.<br />

Commercial vendors, such as Affymetrix, Agilent, Illumina, and NimbleGen,<br />

have begun <strong>of</strong>fering oligonucleotide-based arrays with which to study<br />

genomewide CNAs (41–43). The Illumina arrays are unique in that the oligonucleotide<br />

probes are attached to microbeads (39), while the NimbleGen arrays<br />

are synthesized on demand and so can be easily customized (44). The smaller<br />

target sequences, 25 to 85 bp, also <strong>of</strong>fer the possibility <strong>of</strong> higher resolution than<br />

is possible using cloned probe arrays. Some protocols generate relative copy<br />

number information by comparing tumor and normal hybridization signals using<br />

two-color cohybridization approaches, while others rely on hybridizations <strong>of</strong><br />

tumor and normal reference DNA to separate arrays. The latter approach is<br />

technically simpler but requires additional normalization steps to account for<br />

array-to-array differences, introducing additional experimental variables when<br />

analyzing data sets.


66 Wang and Gray<br />

Since each oligonucleotide probe is synthesized to be complementary to a<br />

known sequence, oligonucleotide array platforms do not suffer as much from<br />

incorrect mapping, and probes can be selected to be uniquely represented in the<br />

genome, thereby minimizing cross-hybridization artifacts. The smaller sequences<br />

also are useful in detecting single nucleotide polymorphisms (SNPs), since<br />

hybridization rates to 25 mer are significantly reduced by a single nucleotide<br />

mismatch (37,45). The additional information provides researchers the ability to<br />

detect loss <strong>of</strong> heterozygosity (LOH) in regions <strong>of</strong> neutral copy change and also<br />

provide data for use in association studies (42). Regions <strong>of</strong> LOH that do not carry<br />

imbalances in copy number may provide evidence <strong>of</strong> selective pressures duplicating<br />

one allele coupled with the loss <strong>of</strong> the other. This is an important<br />

mechanism enabling removal <strong>of</strong> wild-type alleles while selecting mutations.<br />

Allele-specific selection can be either positive or negative. For example, loss <strong>of</strong><br />

the wild-type allele and duplication <strong>of</strong> an inactivating mutation might lead to<br />

inactivation <strong>of</strong> a tumor suppressor gene. The high frequency <strong>of</strong> LOH on 17p<br />

where p53 is located is an example <strong>of</strong> this type <strong>of</strong> selection (Fig. 2). On the other<br />

hand, amplification <strong>of</strong> a polymorphism that increases gene activity may contribute<br />

an advantage to the tumor. Amplification <strong>of</strong> a specific polymorphism in<br />

the aurora kinase STK15 is one example (46). The utility <strong>of</strong> allele-specific copy<br />

number analysis may also contribute to the identification <strong>of</strong> polymorphisms that<br />

confer increased susceptibility to cancer, since evidence is now emerging that<br />

tumors may increase the number <strong>of</strong> copies <strong>of</strong> polymorphisms that confer<br />

increased risk and may decrease the number <strong>of</strong> copies <strong>of</strong> polymorphisms that<br />

confer decreased risk <strong>of</strong> developing cancer (47). The frequency <strong>of</strong> allele-specific<br />

copy increases and decreases in individual tumors is substantial, as illustrated in<br />

Figure 3.<br />

While commercial oligonucleotide arrays <strong>of</strong>fer advantages, the smaller<br />

target sequences also present some disadvantages. Foremost among these is the<br />

relatively low signal-to-noise ratio generated during hybridization. Because the<br />

hybridization sequences are smaller, there is less nucleic acid sequence available<br />

to label, leading to less intense labeling. This typically leads to higher variability<br />

between probe sets in oligonucleotide arrays, and the need to combine multiple<br />

data points to estimate copy number, which in turn reduces the actual resolution<br />

<strong>of</strong> these arrays (38,48). The increased variability compared to BAC arrays also<br />

makes the detection <strong>of</strong> low-level changes, such as single-copy gains and losses,<br />

more difficult to detect, although performance with oligonucleotide arrays is<br />

improving steadily. Typically, microgram quantities <strong>of</strong> starting DNA are needed<br />

for an analysis.<br />

The most recent array CGH analysis platform does not operate through<br />

hybridization <strong>of</strong> tumor and normal DNA to a normal representation. Instead, it<br />

utilizes a molecular inversion probe (MIP) interrogation strategy to perform<br />

allele-specific copy number analysis (45). In this approach, each locus is interrogated<br />

by a linear MIP probe that has sequence homology to approximately<br />

20-bp regions flanking a single base to be tested. Each MIP also contains a unique


Comparative Genomic Hybridization in <strong>Breast</strong> <strong>Cancer</strong> 67<br />

Figure 2 LOH detection using Affymetrix GeneChip 1 Human Mapping 50K Array Xba<br />

240. A panel <strong>of</strong> breast cancer cell lines is individually hybridized to oligonucleotide<br />

arrays, and copy number changes and LOH are determined using dChipSNP. (A) The<br />

expected result <strong>of</strong> a genomic deletion is the presence <strong>of</strong> LOH, as seen on chromosome 8.<br />

(B) However, chromosome 17 shows much higher levels <strong>of</strong> LOH compared with deletions,<br />

suggesting that these regions have undergone genomic changes such as a deletion or<br />

duplication event or genomic conversion <strong>of</strong> one homolog so that you maintain normal<br />

copy number but not allelic information. This may result in the loss <strong>of</strong> a functional allele<br />

<strong>of</strong> a tumor suppressor gene or an active gain-<strong>of</strong>-function allele in the case <strong>of</strong> an oncogene.<br />

These genomic changes are only evident using a platform capable <strong>of</strong> detecting allelic<br />

changes. Abbreviation: LOH, loss <strong>of</strong> heterozygositiy.<br />

tag sequence and universal polymerase chain reaction (PCR) primer-binding<br />

sites facing away from each other so that PCR with primers to the universal<br />

sequences will not amplify. The first step in MIP CGH is to hybridize the linear<br />

MIPs to the DNA sample to be interrogated. After hybridization, the reaction<br />

mix is separated into multiple tubes, each containing polymerase, ligase, and one<br />

<strong>of</strong> the four nucleotides. The polymerase and ligase then fill the gaps in each<br />

reaction mix that contains a nucleotide complementary to the base at the test<br />

locus. The linear DNA and probes are then eliminated, reactions are mixed, and<br />

the ligated probes are amplified and labeled using PCR with the universal primers.<br />

The resulting PCR products are hybridized to an array comprising oligonucleotides<br />

homologous to the tag sequences in the MIPs. To date, MIP analyses<br />

interrogating over 20,000 loci have been reported (45). This approach reports


68 Wang and Gray<br />

Figure 3 Allelic copy number changes in breast cancer cell lines from a 50K MIP<br />

array platform. Allele copy numbers are represented as the higher copy allele and<br />

lower copy allele. (A) Allele copy number data from HCC38 from chromosome 5.<br />

The lack <strong>of</strong> copies <strong>of</strong> the lower allele is indicative <strong>of</strong> LOH. The array detects a region<br />

<strong>of</strong> LOH without copy number change. (B) Allele copy number data from ZR75.30<br />

from chromosome 8. 8q displays a region <strong>of</strong> monoallelic amplification in the region<br />

that contains MYC. Abbreviations: MIP, molecular inversion probe; LOH, loss <strong>of</strong><br />

heterozygositiy.


Comparative Genomic Hybridization in <strong>Breast</strong> <strong>Cancer</strong> 69<br />

copy number in an allele-specific manner and works well with FFPE material<br />

since only ~40 bp <strong>of</strong> intact DNA is needed at each locus for proper analysis.<br />

Typically, only about 75 ng <strong>of</strong> DNA is needed for an analysis since the reaction<br />

itself contains a PCR amplification step. The approach has the advantage that the<br />

tag sequences and arrays can be designed to have uniform hybridization conditions<br />

to copy number and the hybridization does not depend on the quality <strong>of</strong><br />

the hybridized DNA.<br />

DNA REQUIREMENTS<br />

One <strong>of</strong> the most important factors when considering a clinical CGH tool is the<br />

quality and amount <strong>of</strong> DNA available for analysis. Depending on which array<br />

platform is chosen, the amounts needed for successful hybridization can range<br />

from 50 ng to 3 mg. In addition, some protocols have been developed that allow<br />

analysis <strong>of</strong> material from formalin-fixed, paraffin-embedded (FFPE) samples<br />

(39,49). The largest amounts <strong>of</strong> starting material generally are used in protocols<br />

in which entire genomes are labeled and hybridized to the array, as is usual for<br />

array CGH using BAC or Agilent oligonucleotide arrays. In these experiments,<br />

the starting DNA is typically labeled with fluorescent nucleotides using random<br />

priming. For BAC arrays, unlabeled repeat-rich DNA is included to suppress<br />

hybridization <strong>of</strong> repeated sequences to repetitive sequences that are present in<br />

the BAC probes. Larger amounts <strong>of</strong> DNA are needed to “drive” these wholegenome<br />

hybridization reactions. The amount <strong>of</strong> starting material needed for<br />

analysis is reduced if the DNA is amplified or reduced in complexity—for<br />

example, through a PCR amplification step preceded by restriction digestion that<br />

reduces the complexity <strong>of</strong> the hybridized DNA (50). This complexity reduction<br />

results in an increase <strong>of</strong> template sequences, which can utilized in downstream<br />

hybridization steps. Both academically and commercially developed platforms<br />

result in data sets that are able to detect gains and losses with the caveat that<br />

variations and noise may differ between platforms (Fig. 4).<br />

While there is a range <strong>of</strong> starting DNA amounts between the various<br />

platforms, recent protocols have been developed that can amplify a large<br />

quantity <strong>of</strong> DNA from small starting amounts. These applications can be critical<br />

when dealing with samples that are limiting. These whole-genome amplifications<br />

are useful for both intact, high-quality DNA and degraded samples and<br />

involve the use <strong>of</strong> j29 enzyme or PCR (51,52). Alternately, the MIP platform has<br />

a targeted PCR amplification inherent to the protocol, so it is particularly tolerant<br />

<strong>of</strong> small amounts <strong>of</strong> starting DNA (45). To date, successful data sets have been<br />

generated using as little material as 1 ng or a single cell (53). However, as with<br />

any random amplification procedure, the protocols may artificially create misrepresentations<br />

<strong>of</strong> the genome, resulting in false positives for amplifications and<br />

deletions. Determining the amount and quality <strong>of</strong> starting material along with the


70 Wang and Gray<br />

Figure 4 Chromosomal copy number changes in the breast cancer cell line MDAMB453. DNA from MDAMB453 is applied to three separate<br />

platforms: (A) a 3K BAC array, (B) a 20K MIP array, and (C) a 50K XbaI Affymetrix SNP array. Data from each platform are scaled using adjacent data<br />

points so that the resulting resolutions are comparable. The data are then smoothed using the GLAD s<strong>of</strong>tware package. Abbreviations: BAC, bacterial<br />

artificial chromosome; MIP, molecular inversion probe; SNP, single nucleotide polymorphism; GLAD, Gaussian model–based approach.


Comparative Genomic Hybridization in <strong>Breast</strong> <strong>Cancer</strong> 71<br />

type <strong>of</strong> information required will help determine the appropriate array CGH<br />

platform and DNA preparation technique to utilize.<br />

ABNORMALITY DETECTION AND INTERPRETATION<br />

Computational strategies to identify copy number gains and losses are a critical<br />

part <strong>of</strong> array CGH. The first step in the analysis <strong>of</strong> data is the identification <strong>of</strong><br />

any features with anomalous intensity readings. These could result from a high<br />

background, errors in printing, contamination, or air bubbles prohibiting<br />

hybridization <strong>of</strong> the sample to the array. After removal <strong>of</strong> these spurious data<br />

points, a normalization step is applied to calculate the tumor hybridization<br />

intensity along the genome, relative to the normal hybridization intensity. A<br />

statistical analysis is then applied to identify genome segments with different<br />

copy number levels. Several strategies have been developed for this purpose and<br />

are summarized in Table 2. These include the hidden Markov model (HMM)<br />

(54), circular binary segmentation (55), a Gaussian model–based approach<br />

(GLAD) (56), hierarchical clustering along each chromosome (CLAC) (57), an<br />

expectation maximization–based method (58), a Bayesian model (59), and a<br />

wavelet approach (60). Several <strong>of</strong> these algorithms have been made available as<br />

R packages through Bioconductor (www.bioconductor.org). Comparsions <strong>of</strong><br />

several <strong>of</strong> these methods have now been published (61,62).<br />

Detection <strong>of</strong> the differences in copy number between tumor and normal<br />

samples is increasingly straightforward and precise as CGH procedures improve.<br />

However, the interpretation <strong>of</strong> CGH analyses requires additional interpretation.<br />

First, high-resolution array CGH analyses are important now that focal, germline<br />

genome copy number polymorphisms are relatively common (36). These may be<br />

scored as CNAs if the samples for tumor versus normal CGH analyses are not<br />

from the same individual. In addition, it is tempting to think that the regions <strong>of</strong><br />

the genome not detected as aberrant during CGH analyses are present in cancer<br />

genomes in two copies. This is <strong>of</strong>ten not the case since tumor genomes may be<br />

highly aneuploid. Thus, a CGH comparison <strong>of</strong> a perfectly tetraploid tumor with a<br />

diploid normal sample will not show any evidence <strong>of</strong> abnormality. Thus, if<br />

absolute copy number assessments are important, it will be necessary to normalize<br />

the CGH analyses using information about absolute copy number at one or a few<br />

loci measured using fluorescent in situ hybridization (FISH) or quantitative PCR.<br />

CONCLUSIONS AND FUTURE DIRECTIONS<br />

Array CGH technologies are becoming increasingly powerful tools for the<br />

identification <strong>of</strong> recurrent genome CNAs that are associated with outcome or<br />

response to therapy in breast and other cancers (63). The ability to detect genetic<br />

copy number changes <strong>of</strong> both large and small scales will be critical in evaluating<br />

the most effective therapies on an individual patient basis. In the short term,<br />

array CGH will likely be accomplished using commercial oligonucleotide<br />

analysis platforms that allow assessment <strong>of</strong> allele-specific copy number using


72 Wang and Gray<br />

material from FFPE samples. In the long term, however, array CGH may be<br />

supplanted by high-throughput sequencing technologies that allow assessment <strong>of</strong><br />

both copy number and DNA sequence through deep sequencing. It is already clear<br />

that shotgun sequencing, if done at sufficient redundancy, <strong>of</strong>fers the same or better<br />

resolution than CGH platforms (64,65). Next-generation sequencing platforms<br />

based on massively parallel single-molecule sequencing promise sufficiently lowcost<br />

sequencing, so that these platforms may become a viable alternative to array<br />

CGH (66,67). Today, this is not cost effective, but costs <strong>of</strong> high-throughput<br />

sequencing are falling rapidly, so these approaches should be followed closely.<br />

REFERENCES<br />

1. Knuutila S, Autio K, Aalto Y. Online access to CGH data <strong>of</strong> DNA sequence copy<br />

number changes. Am J Pathol 2000; 157:689.<br />

2. Neve RM, Chin K, Fridlyand J, et al. A collection <strong>of</strong> breast cancer cell lines for the<br />

study <strong>of</strong> functionally distinct cancer subtypes. <strong>Cancer</strong> Cell 2006; 10:515–527.<br />

3. Pollack JR, Sorlie T, Perou CM, et al. Microarray analysis reveals a major direct role<br />

<strong>of</strong> DNA copy number alteration in the transcriptional program <strong>of</strong> human breast<br />

tumors. Proc Natl Acad Sci U.S.A. 2002; 99:12963–12968.<br />

4. Chin K, DeVries S, Fridlyand J, et al. Genomic and transcriptional aberrations linked<br />

to breast cancer pathophysiologies. <strong>Cancer</strong> Cell 2006; 10:529–541.<br />

5. Bergamaschi A, Kim YH, Wang P, et al. Distinct patterns <strong>of</strong> DNA copy number<br />

alteration are associated with different clinicopathological features and gene-expression<br />

subtypes <strong>of</strong> breast cancer. Genes Chromosomes <strong>Cancer</strong> 2006; 45:1033–1040.<br />

6. Callagy G, <strong>Ph</strong>aroah P, Chin SF, et al. Identification and validation <strong>of</strong> prognostic<br />

markers in breast cancer with the complementary use <strong>of</strong> array-CGH and tissue<br />

microarrays. J Pathol 2005; 205:388–396.<br />

7. Chin K, de Solorzano CO, Knowles D, et al. In situ analyses <strong>of</strong> genome instability in<br />

breast cancer. Nat Genet 2004; 36:984–988.<br />

8. Slamon DJ, Godolphin W, Jones LA, et al. Studies <strong>of</strong> the HER-2/neu proto-oncogene<br />

in human breast and ovarian cancer. Science 1989; 244:707–712.<br />

9. Theillet C, Adelaide J, Louason G, et al. FGFRI and PLAT genes and DNA<br />

amplification at 8p12 in breast and ovarian cancers. Genes Chromosomes <strong>Cancer</strong><br />

1993; 7:219–226.<br />

10. Escot C, Theillet C, Lidereau R, et al. Genetic alteration <strong>of</strong> the c-myc protooncogene<br />

(MYC) in human primary breast carcinomas. Proc Natl Acad Sci U.S.A. 1986;<br />

83:4834–4838.<br />

11. Guan Y, Kuo WL, Stilwell JL, et al. Amplification <strong>of</strong> PVT1 contributes to the<br />

pathophysiology <strong>of</strong> ovarian and breast cancer. Clin <strong>Cancer</strong> Res 2007; 13:5745–5755.<br />

12. Schuuring E, Verhoeven E, Mooi WJ, et al. Identification and cloning <strong>of</strong> two<br />

overexpressed genes, U21B31/PRAD1 and EMS1, within the amplified chromosome<br />

11q13 region in human carcinomas. Oncogene 1992; 7:355–361.<br />

13. Anzick SL, Kononen J, Walker RL, et al. AIB1, a steroid receptor coactivator<br />

amplified in breast and ovarian cancer. Science 1997; 277:965–968.<br />

14. Tanner MM, Tirkkonen M, Kallioniemi A, et al. Amplification <strong>of</strong> chromosomal<br />

region 20q13 in invasive breast cancer: prognostic implications. Clin <strong>Cancer</strong> Res<br />

1995; 1:1455–1461.


Comparative Genomic Hybridization in <strong>Breast</strong> <strong>Cancer</strong> 73<br />

15. Varley JM, Brammar WJ, Lane DP, et al. Loss <strong>of</strong> chromosome 17p13 sequences and<br />

mutation <strong>of</strong> p53 in human breast carcinomas. Oncogene 1991; 6:413–421.<br />

16. Kamb A, Gruis NA, Weaver-Feldhaus J, et al. A cell cycle regulator potentially<br />

involved in genesis <strong>of</strong> many tumor types. Science 1994; 264:436–440.<br />

17. Friedman LS, Ostermeyer EA, Lynch ED, et al. The search for BRCA1. <strong>Cancer</strong> Res<br />

1994; 54:6374–6382.<br />

18. Hodgson JG, Malek T, Bornstein S, et al. Copy number aberrations in mouse breast<br />

tumors reveal loci and genes important in tumorigenic receptor tyrosine kinase<br />

signaling. <strong>Cancer</strong> Res 2005; 65:9695–9704.<br />

19. Maser RS, Choudhury B, Campbell PJ, et al. Chromosomally unstable mouse tumors<br />

have genomic alterations similar to diverse human cancers. Nature 2007; 447:966–971.<br />

20. Yeon CH, Pegram MD. Anti-erbB-2 antibody trastuzumab in the treatment <strong>of</strong> HER2-<br />

amplified breast cancer. Invest New Drugs 2005; 23:391–409.<br />

21. Vogel CL, Cobleigh MA, Tripathy D, et al. First-line Herceptin monotherapy in<br />

metastatic breast cancer. Oncology 2001; 61(suppl 2):37–42.<br />

22. Geyer CE, Forster J, Lindquist D, et al. Lapatinib plus capecitabine for HER2-<br />

positive advanced breast cancer. N Engl J Med 2006; 355(26):2733–2743.<br />

23. Kumar S, Walia V, Ray M, et al. p53 in breast cancer: mutation and countermeasures.<br />

Front Biosci 2007; 12:4168–4178.<br />

24. Jonsson G, Staaf J, Olsson E, et al. High-resolution genomic pr<strong>of</strong>iles <strong>of</strong> breast cancer<br />

cell lines assessed by tiling BAC array comparative genomic hybridization. Genes<br />

Chromosomes <strong>Cancer</strong> 2007; 46:543–558.<br />

25. Wu G, Xing M, Mambo E, et al. Somatic mutation and gain <strong>of</strong> copy number <strong>of</strong><br />

PIK3CA in human breast cancer. <strong>Breast</strong> <strong>Cancer</strong> Res 2005; 7:R609–R616.<br />

26. Huang G, Krig S, Kowbel D, et al. ZNF217 suppresses cell death associated with<br />

chemotherapy and telomere dysfunction. Hum Mol Genet 2005; 14:3219–3125.<br />

27. Bellacosa A, de Feo D, Godwin AK, et al. Molecular alterations <strong>of</strong> the AKT2<br />

oncogene in ovarian and breast carcinomas. Int J <strong>Cancer</strong> 1995; 64:280–285.<br />

28. Kallioniemi A, Kallioniemi OP, Sudar D, et al. Comparative genomic hybridization<br />

for molecular cytogenetic analysis <strong>of</strong> solid tumors. Science 1992; 258:818–821.<br />

29. Pinkel D, Segraves R, Sudar D, et al. High-resolution analysis <strong>of</strong> DNA copy number<br />

variation using comparative genomic hybridization to microarrays. Nat Genet 1998;<br />

20:207–211.<br />

30. Snijders AM, Nowak N, Segraves R, et al. Assembly <strong>of</strong> microarrays for genomewide<br />

measurement <strong>of</strong> DNA copy number. Nat Genet 2001; 29:263–264.<br />

31. Toruner GA, Streck DL, Schwalb MN, et al. An oligonucleotide based array-CGH<br />

system for detection <strong>of</strong> genome wide copy number changes including subtelomeric<br />

regions for genetic evaluation <strong>of</strong> mental retardation. Am J Med Genet A 2007;<br />

143:824–829.<br />

32. George RE, Attiyeh EF, Li S, et al. Genome-wide analysis <strong>of</strong> neuroblastomas using<br />

high-density single nucleotide polymorphism arrays. PLoS ONE 2007; 2:e255.<br />

33. Kallioniemi A, Kallioniemi OP, Piper J, et al. Detection and mapping <strong>of</strong> amplified<br />

DNA sequences in breast cancer by comparative genomic hybridization. Proc Natl<br />

Acad Sci U.S.A. 1994; 91:2156–2160.<br />

34. Solinas-Toldo S, Lampel S, Stilgenbauer S, et al. Matrix-based comparative genomic<br />

hybridization: biochips to screen for genomic imbalances. Genes Chromosomes<br />

<strong>Cancer</strong> 1997; 20:399–407.


74 Wang and Gray<br />

35. Pollack JR, Perou CM, Alizadeh AA, et al. Genome-wide analysis <strong>of</strong> DNA copynumber<br />

changes using cDNA microarrays. Nat Genet 1999; 23:41–46.<br />

36. Lucito R, Healy J, Alexander J, et al. Representational oligonucleotide microarray<br />

analysis: a high-resolution method to detect genome copy number variation. Genome<br />

Res 2003; 13:2291–2305.<br />

37. Zhao X, Li C, Paez JG, et al. An integrated view <strong>of</strong> copy number and allelic<br />

alterations in the cancer genome using single nucleotide polymorphism arrays.<br />

<strong>Cancer</strong> Res 2004; 64:3060–3071.<br />

38. Barrett MT, Scheffer A, Ben-Dor A, et al. Comparative genomic hybridization using<br />

oligonucleotide microarrays and total genomic DNA. Proc Natl Acad Sci U.S.A.<br />

2004; 101:17765–17770.<br />

39. Oosting J, Lips EH, van Eijk R, et al. High-resolution copy number analysis <strong>of</strong><br />

paraffin-embedded archival tissue using SNP BeadArrays. Genome Res 2007; 17:<br />

368–376.<br />

40. Hodgson G, Hager JH, Volik S, et al. Genome scanning with array CGH delineates<br />

regional alterations in mouse islet carcinomas. Nat Genet 2001; 29:459–464.<br />

41. Iljin K, Wolf M, Edgren H, et al. TMPRSS2 fusions with oncogenic ETS factors in<br />

prostate cancer involve unbalanced genomic rearrangements and are associated with<br />

HDAC1 and epigenetic reprogramming. <strong>Cancer</strong> Res 2006; 66:10242–10246.<br />

42. Zhou X, Mok SC, Chen Z, et al. Concurrent analysis <strong>of</strong> loss <strong>of</strong> heterozygosity (LOH)<br />

and copy number abnormality (CNA) for oral premalignancy progression using the<br />

Affymetrix 10K SNP mapping array. Hum Genet 2004; 115:327–330.<br />

43. Sharp AJ, Hansen S, Selzer RR, et al. Discovery <strong>of</strong> previously unidentified genomic<br />

disorders from the duplication architecture <strong>of</strong> the human genome. Nat Genet 2006;<br />

38:1038–1042.<br />

44. Selzer RR, Richmond TA, P<strong>of</strong>ahl NJ, et al. Analysis <strong>of</strong> chromosome breakpoints in<br />

neuroblastoma at sub-kilobase resolution using fine-tiling oligonucleotide array<br />

CGH. Genes Chromosomes <strong>Cancer</strong> 2005; 44:305–319.<br />

45. Wang Y, Moorhead M, Karlin-Neumann G, et al. Allele quantification using<br />

molecular inversion probes (MIP). Nucleic Acids Res 2005; 33:e183.<br />

46. Ewart-Toland A, Briassouli P, de Koning JP, et al. Identification <strong>of</strong> Stk6/STK15 as a<br />

candidate low-penetrance tumor-susceptibility gene in mouse and human. Nat Genet<br />

2003; 34:403–412.<br />

47. Balmain A, Gray J, Ponder B. The genetics and genomics <strong>of</strong> cancer. Nat Genet 2003;<br />

33:238–244.<br />

48. Carvalho B, Ouwerkerk E, Meijer GA, et al. High-resolution microarray comparative<br />

genomic hybridisation analysis using spotted oligonucleotides. J Clin Pathol 2004;<br />

57:644–646.<br />

49. Nassiri M, Gugic D, Olczyk J, et al. Preservation <strong>of</strong> skin DNA for oligonucleotide<br />

array CGH studies: a feasibility study. Arch Dermatol Res 2007; 299:353–357.<br />

50. Hicks J, Muthuswamy L, Krasnitz A, et al. High-resolution ROMA CGH and FISH<br />

analysis <strong>of</strong> aneuploid and diploid breast tumors. Cold Spring Harb Symp Quant Biol<br />

2005; 70:51–63.<br />

51. Guillaud-Bataille M, Valent A, Soularue P, et al. Detecting single DNA copy number<br />

variations in complex genomes using one nanogram <strong>of</strong> starting DNA and BAC-array<br />

CGH. Nucleic Acids Res 2004; 32:e112.


Comparative Genomic Hybridization in <strong>Breast</strong> <strong>Cancer</strong> 75<br />

52. Lage JM, Leamon JH, Pejovic T, et al. Whole genome analysis <strong>of</strong> genetic alterations<br />

in small DNA samples using hyperbranched strand displacement amplification and<br />

array-CGH. Genome Res 2003; 13:294–307.<br />

53. Le Caignec C, Spits C, Sermon K, et al. Single-cell chromosomal imbalances<br />

detection by array CGH. Nucleic Acids Res 2006; 34:e68.<br />

54. Fridlyand J, Snijders A, Pinkel D, et al. Hidden Markov models approach to the<br />

analysis <strong>of</strong> array CGH data. J Multivar Anal 2004; 90:132–153.<br />

55. Olshen AB, Venkatraman ES, Lucito R, et al. Circular binary segmentation for the<br />

analysis <strong>of</strong> array-based DNA copy number data. Biostatistics 2004; 5:557–572.<br />

56. Hupe P, Stransky N, Thiery JP, et al. Analysis <strong>of</strong> array CGH data: from signal ratio<br />

to gain and loss <strong>of</strong> DNA regions. Bioinformatics 2004; 20:3413–3122.<br />

57. Wang P, Kim Y, Pollack J, et al. A method for calling gains and losses in array CGH<br />

data. Biostatistics 2005; 6:45–58.<br />

58. Myers CL, Dunham MJ, Kung SY, et al. Accurate detection <strong>of</strong> aneuploidies in array<br />

CGH and gene expression microarray data. Bioinformatics 2004; 20:3533–3543.<br />

59. Daruwala RS, Rudra A, Ostrer H, et al. A versatile statistical analysis algorithm to detect<br />

genome copy number variation. Proc Natl Acad Sci U.S.A. 2004; 101:16292–16297.<br />

60. Hsu L, Self SG, Grove D, et al. Denoising array-based comparative genomic<br />

hybridization data using wavelets. Biostatistics 2005; 6:211–226.<br />

61. Lai WR, Johnson MD, Kucherlapati R, et al. Comparative analysis <strong>of</strong> algorithms for<br />

identifying amplifications and deletions in array CGH data. Bioinformatics 2005;<br />

21:3763–3770.<br />

62. Willenbrock H, Fridlyand J. A comparison study: applying segmentation to array<br />

CGH data for downstream analyses. Bioinformatics 2005; 21:4084–4091.<br />

63. Gunn SR, Robetorye RS, Mohammed MS. Comparative genomic hybridization arrays<br />

in clinical pathology: progress and challenges. Mol Diagn Ther 2007; 11:73–77.<br />

64. Volik S, Raphael BJ, Huang G, et al. Decoding the fine-scale structure <strong>of</strong> a breast<br />

cancer genome and transcriptome. Genome Res 2006; 16:394–404.<br />

65. Volik S, Zhao S, Chin K, et al. End-sequence pr<strong>of</strong>iling: sequence-based analysis <strong>of</strong><br />

aberrant genomes. Proc Natl Acad Sci U.S.A. 2003; 100:7696–7701.<br />

66. Albert TJ, Molla MN, Muzny DM, et al. Direct selection <strong>of</strong> human genomic loci by<br />

microarray hybridization. Nat Methods 2007; 4:903–905.<br />

67. Bennett ST, Barnes C, Cox A, et al. Toward the 1,000 dollars human genome.<br />

<strong>Ph</strong>armacogenomics 2005; 6:373–382.<br />

68. Ishkanian AS, Mall<strong>of</strong>f CA, Watson SK, et al. A tiling resolution DNA microarray<br />

with complete coverage <strong>of</strong> the human genome. Nat Genet 2004; 36:299–303.<br />

69. Huang J, Wei W, Chen J, et al. CARAT: a novel method for allelic detection <strong>of</strong> DNA<br />

copy number changes using high density oligonucleotide arrays. BMC Bioinformatics<br />

2006; 7(83):83.<br />

70. Nannya Y, Sanada M, Nakazaki K, et al. A robust algorithm for copy number<br />

detection using high-density oligonucleotide single nucleotide polymorphism genotyping<br />

arrays. <strong>Cancer</strong> Res 2005; 65(14):6071–6079.<br />

71. Komura D, Shen F, Ishikawa S, et al. Genome-wide detection <strong>of</strong> human copy number<br />

variations using high-density DNA oligonucleotide arrays. Genome Res 2006; 16(12):<br />

1575–1584.


6<br />

Gene Expression Pr<strong>of</strong>iling as an Emerging<br />

Diagnostic Tool to Personalize<br />

Chemotherapy Selection for<br />

Early Stage <strong>Breast</strong> <strong>Cancer</strong><br />

Cornelia Liedtke and Lajos Pusztai<br />

Department <strong>of</strong> <strong>Breast</strong> Medical Oncology, The University <strong>of</strong> Texas,<br />

Anderson <strong>Cancer</strong> Center, Houston, Texas, U.S.A.<br />

INTRODUCTION<br />

The selection <strong>of</strong> adjuvant chemotherapy for breast cancer is currently more <strong>of</strong> an<br />

art than a science. Several combination chemotherapy regimens can improve<br />

disease-free and overall survival <strong>of</strong> stage I to stage III breast cancer when used as<br />

adjuvant (postoperative) or neoadjuvant (preoperative) therapy (1,2). However,<br />

no chemotherapy regimen is universally effective for all patients, and it is currently<br />

impossible to predict which <strong>of</strong> these regimens will actually work for a<br />

particular individual. Choosing the best possible adjuvant or neoadjuvant chemotherapy<br />

regimen for an individual is important because the right choice may<br />

make the difference between cure and relapse. Results from traditionally<br />

designed randomized clinical trials <strong>of</strong>fer little guidance for individualized<br />

treatment selection. These studies compare the activity <strong>of</strong> different treatments to<br />

establish which is the most effective for the entire study population. This<br />

approach ignores the fact that many patients are cured with earlier generation<br />

adjuvant regimens and therefore are not well served by the administration <strong>of</strong><br />

77


78 Liedtke and Pusztai<br />

longer, potentially more toxic, and more expensive second- or third-generation<br />

combination therapies. These studies also cannot reveal whether some cancers<br />

are particularly sensitive to a specific drug (or a combination <strong>of</strong> drugs) and<br />

therefore can only be cured by that treatment. The most effective treatment for<br />

an individual may or may not be the same regimen that is the most effective for<br />

the whole population. These theoretical concerns and the increasing number <strong>of</strong><br />

adjuvant chemotherapy options, as well as the advent <strong>of</strong> new molecular analytical<br />

techniques, fuel a renewed interest in developing chemotherapy response<br />

predictors that may lead to more personalized treatment recommendations.<br />

THE ROLE OF NEOADJUVANT CHEMOTHERAPY<br />

IN THE TREATMENT OF BREAST CANCER<br />

Administration <strong>of</strong> chemotherapy before surgical resection <strong>of</strong> the cancer (preoperative<br />

or neoadjuvant chemotherapy) provides a unique opportunity to<br />

identify molecular markers <strong>of</strong> response. Preoperative chemotherapy is the current<br />

standard <strong>of</strong> care for locally advanced breast cancer because this treatment<br />

<strong>of</strong>ten renders previously inoperable cancers amenable to surgery due to frequent<br />

and substantial tumor responses (3). Preoperative chemotherapy is also<br />

increasingly used in the management <strong>of</strong> operable breast cancer because it can<br />

reduce tumor size and allow for more limited surgery and possibly improved<br />

cosmetic outcome (4). Importantly, there is no statistically significant difference<br />

in survival between those who receive chemotherapy before surgery (i.e., neoadjuvant)<br />

and those who are treated with the same regimen postoperatively (i.e.,<br />

adjuvant) (5–7). Therefore, preoperative chemotherapy is an appropriate option<br />

for most patients who present with a breast cancer that requires systemic therapy.<br />

Furthermore, eradication <strong>of</strong> all invasive cancer from the breast and regional<br />

lymph nodes by the preoperative therapy [i.e., pathologic complete response,<br />

(pCR)], is associated with excellent long-term, disease-free, and overall survival<br />

(7,8). Therefore, one could propose that the regimen most likely to induce pCR is<br />

the most effective treatment for a particular individual, and that such treatment<br />

could be ideally suited to that patient either as adjuvant or neoadjuvant therapy.<br />

This is the rationale why pCR is a commonly used clinical endpoint for breast<br />

cancer response predictors. However, it is important to realize that most patients<br />

with less than pCR had at least some clinical response to chemotherapy, and<br />

therefore cannot be considered truly resistant to treatment, but rather, cases with<br />

pCR represent extreme chemotherapy sensitivity.<br />

CURRENT CHEMOTHERAPY RESPONSE PREDICTORS<br />

There are several clinical and pathologic variables <strong>of</strong> breast cancer that are consistently<br />

associated with a greater likelihood <strong>of</strong> response to neoadjuvant chemotherapy.<br />

These include estrogen receptor (ER)-negative status, smaller tumor<br />

size, increased proliferative activity, and high nuclear or histological grade (9–12).


Gene Expression Pr<strong>of</strong>iling to Personalize Chemotherapy Selection 79<br />

Tumor size, grade, and ER-status are routinely available parameters that can be<br />

combined into a multivariate response prediction model (13). One such model<br />

is freely available at our institutional Web site, http://www.mdanderson.org/<br />

Care_Centers/<strong>Breast</strong>Center/dIndex.cfm?pn¼448442B2-3EA5-4BAC-<br />

98310076A9553E63. However, clinical variable–based predictions are limited to<br />

predict response to chemotherapy in general and lack regimen specificity.<br />

A large number <strong>of</strong> molecular markers were proposed as potential predictors<br />

<strong>of</strong> response to various chemotherapy agents (11,12,14–16). However, currently<br />

there are no molecular markers <strong>of</strong> response that could be used in the clinic to select<br />

patients for a particular therapy. Biomarker studies were traditionally conducted as<br />

correlative science projects attached to therapeutic trials as optional components.<br />

This has imposed serious limitations on these studies. Sample sizes for marker<br />

discovery were limited by the availability <strong>of</strong> specimens that has lead to underpowered<br />

observations and introduced potential bias due to subset analysis. Most<br />

commonly, only associations between marker and response were reported, but no<br />

formal prediction models were developed. The methodologies to detect a marker<br />

were rarely standardized, and definitions for test-positive and test-negative cases<br />

<strong>of</strong>ten differed from study to study, even for the same marker.<br />

In addition to these technical limitations, there are important theoretical<br />

considerations, which also suggest that single gene molecular markers expressed<br />

by the cancer may have limited predictive value. Factors that reside outside the<br />

neoplastic cells could influence drug response. The importance <strong>of</strong> interactions<br />

between stroma and neoplastic cells, including neovascularization, changes in<br />

interstitial fluid pressure, and immune/inflammatory response are increasingly<br />

recognized as potential determinants <strong>of</strong> prognosis and possibly response to<br />

therapy (17,18). Rates <strong>of</strong> drug metabolism are also variable and can have an<br />

effect on drug activity. Cellular response to toxic insults is a complex molecular<br />

process that involves simultaneous induction <strong>of</strong> several proapoptotic and survival<br />

pathways, including activation <strong>of</strong> repair and drug elimination mechanisms<br />

(19,20). Tumor response may depend on the balance <strong>of</strong> these pathways, and<br />

therefore any single gene may contain only limited information about the<br />

complexity <strong>of</strong> the mechanisms that determine response to chemotherapy.<br />

GENE EXPRESSION PROFILING AS A NOVEL TOOL<br />

FOR RESPONSE PREDICTION<br />

Gene expression pr<strong>of</strong>iling with DNA microarrays represents a relatively recent<br />

tissue analysis tool that was developed in the early 1990s (21). This technology<br />

enables investigators to measure, in a semiquantitative manner in a single<br />

experiment, the expression <strong>of</strong> almost all mRNAs expressed in a small piece <strong>of</strong><br />

cancer tissue. The rationale behind pharmacogenomic a research is that by using<br />

a <strong>Ph</strong>armacogenomics is defined in this review as the use <strong>of</strong> mRNA expression pr<strong>of</strong>iles <strong>of</strong> the tumor to<br />

predict drug efficacy. <strong><strong>Ph</strong>armacogenetics</strong> is defined as the study <strong>of</strong> DNA variations in the germline to<br />

predict toxicity or treatment response.


80 Liedtke and Pusztai<br />

this technology, one can identify individually predictive genes and combine<br />

these into a multivariate prediction model that will be more accurate than any<br />

single gene alone. Multivariate outcome prediction models are not new to<br />

medicine. Independent and individual weakly predictive clinical variables have<br />

been combined into useful and validated prognostic or predictive models. There<br />

are several established statistical methods for combining variables into prediction<br />

models, including the commonly used logistic regression analysis and<br />

the Cox proportional hazards regression model (22). Predictors could be<br />

binary—that is, predicting response versus no response—or could predict the<br />

probability <strong>of</strong> response on a continuous scale. The Adjuvant Online (http://www<br />

.adjuvantonline.com) s<strong>of</strong>tware represents a freely available validated multivariate<br />

prognostic prediction tool for breast cancer.<br />

The general strategy for multigene marker development in neoadjuvant<br />

studies is to compare the transcriptional pr<strong>of</strong>iles <strong>of</strong> cancers that responded to<br />

therapy with those that did not and to identify differentially expressed genes that<br />

can be combined into a prediction model. There are several caveats that make<br />

pharmacogenomic predictor discovery challenging.<br />

1. The multiple comparison problem inherent to microarray analysis is well<br />

known. It refers to the large number <strong>of</strong> variables (i.e., genes) that are<br />

compared between two small data sets. As a consequence, observation <strong>of</strong><br />

many small p values may be due to chance. This could result in spuriously<br />

identifying genes as associated with response when in fact they are not.<br />

Several methods used to adjust for this phenomenon, at least to some extent,<br />

by calculating false discovery rates for particular p values were described (23).<br />

2. Another confounder relates to the coordinated expression <strong>of</strong> thousands <strong>of</strong> genes.<br />

Individual transcripts do not represent independent variables, but rather their<br />

expression is highly correlated with one another. Many phenotypic characteristics<br />

<strong>of</strong> breast cancer are also associated with, and likely caused by, the coordinated<br />

expression <strong>of</strong> thousands <strong>of</strong> genes. For example, ER-positive cancers<br />

differ from ER-negative tumors in the expression <strong>of</strong> thousands <strong>of</strong> genes (24). The<br />

gene expression pattern <strong>of</strong> high-grade tumors is also different from low-grade<br />

tumors (25). These large-scale gene expression patterns associated with clinical<br />

phenotypes have a pr<strong>of</strong>ound influence on the pharmacogenomic discovery<br />

process, because ER-negative, high-grade breast cancers are more sensitive<br />

to many different types <strong>of</strong> chemotherapies compared with ER-positive and<br />

low-grade tumors. Unadjusted pharmacogenomic comparison <strong>of</strong> responders<br />

with nonresponders will yield gene lists that are dominated by ER- and gradeassociated<br />

genes, and the resulting pharmacogenomic predictor may represent<br />

a predictor <strong>of</strong> clinical phenotype. When the discovery sample size is large, it<br />

is possible to adjust for these phenotypic confounders; however, in small<br />

studies, these adjustments may be difficult to do.<br />

3. The probability that a supervised pharmacogenomic discovery approach, when<br />

responders are compared with nonresponders, will lead to regimen-specific


Gene Expression Pr<strong>of</strong>iling to Personalize Chemotherapy Selection 81<br />

predictors depends on to what extent the response groups are balanced for<br />

strong phenotypic markers and on the extent <strong>of</strong> molecular differences that<br />

determine drug-specific response. If drug sensitivity is influenced by the<br />

two- to threefold higher or lower expression <strong>of</strong> a few dozen genes, these<br />

differences may not be readily discovered through supervised pharmacogenomic<br />

analysis <strong>of</strong> small data sets. The modest gene expression differences<br />

between responders and nonresponders can be masked by the larger scale<br />

molecular differences due to any phenotypic imbalance between the<br />

response groups and the technical noise <strong>of</strong> microarray experiments. For<br />

example, when 24,000 measurements are performed (e.g., Affymetrix<br />

U133A gene chip) and the overall concordance is 97.98% in a technical<br />

replicate (i.e., same RNA pr<strong>of</strong>iled twice), 1.31% <strong>of</strong> all measurements can<br />

have tw<strong>of</strong>old or greater variation. This means that 314 genes decreased or<br />

increased tw<strong>of</strong>old from one experiment to another due to technical noise<br />

alone (26). If the sample size is large, this technical noise can be separated<br />

from the true signal, but when sample sizes are small it is more difficult to<br />

separate noise from the true signal.<br />

PHARMACOGENOMIC MARKER DEVELOPMENT PROCESS<br />

Ideally, the development process <strong>of</strong> molecular markers should resemble the<br />

development stages <strong>of</strong> therapeutic agents (Fig. 1) (27). In this context, a phase I<br />

marker discovery study would demonstrate the technical feasibility <strong>of</strong> developing a<br />

predictor and assess the reproducibility <strong>of</strong> marker results. The second phase<br />

Figure 1<br />

Schema <strong>of</strong> the developmental stages <strong>of</strong> molecular markers.


82 Liedtke and Pusztai<br />

includes phase II validation studies <strong>of</strong> statistically adequate power to estimate<br />

more precisely the predictive performance <strong>of</strong> the test in independent validation<br />

cases. The next level <strong>of</strong> marker evaluation would include the demonstration <strong>of</strong><br />

clinical value by showing that a particular clinical endpoint improves when the<br />

new test is used compared with the current standard, which could be no testing<br />

or another type <strong>of</strong> testing. This may require a prospective phase III marker<br />

validation study. Currently, no pharmacogenomic response predictors have<br />

completed all three levels <strong>of</strong> clinical evaluation. Most published studies represent<br />

phase I marker discovery and phase II marker validation studies.<br />

FEASIBILITY OF MICROARRAY-BASED DIAGNOSTIC TESTS<br />

Technical Reproducibility<br />

The reproducibility <strong>of</strong> microarray measurements has been examined systematically<br />

by the Microarray Quality Control Project (MAQC), which included the US<br />

Food and Drug Administration and 51 academic and industry collaborators. The<br />

goal <strong>of</strong> this project was to assess the within- and between-laboratory technical<br />

reproducibility <strong>of</strong> microarray measurements and compare the results across<br />

different microarray platforms. Four different RNA samples were pr<strong>of</strong>iled in five<br />

replicates on seven different array platforms, three different laboratories representing<br />

each platform. The same RNA samples were also pr<strong>of</strong>iled by three<br />

different reverse-transcription polymerase chain reaction (RT-PCR) methods.<br />

The most important finding <strong>of</strong> this collaborative effort was that the microarray<br />

measurements were highly reproducible within and across the different commercially<br />

available microarray platforms (28,29). The median coefficient <strong>of</strong><br />

variation (CV) for within-laboratory replicates ranged from 5% to 15% for the<br />

various platforms and it was 10% to 20% for between-laboratory replicates. Most<br />

importantly, the concordance between the microarray based mRNA measurements<br />

and RT-PCR results were also high. The correlation coefficients for gene<br />

expression ratios measured by arrays and by RT-PCR in two different samples<br />

ranged from 0.79 to 0.92 for several hundreds <strong>of</strong> genes.<br />

Between-platform variation in gene expression measurements is due to<br />

sequence variations in the probe sets that target the same gene at different<br />

locations. Factors that determine signal intensities generated by distinct probes<br />

for the same gene include GC content <strong>of</strong> the sequence, sequence length, crossmatch<br />

opportunities, and the location <strong>of</strong> the probe sequence in relation to the<br />

3 0 end <strong>of</strong> the target gene. The concordance between signal intensity for probes<br />

that target the same gene on different platforms is directly related to sequence<br />

homology (30,31). Probes with complete sequence match yield highly concordant<br />

results across platforms.<br />

Few studies examined the within- and cross-platform reproducibility <strong>of</strong><br />

multigene prediction scores, as opposed to the reproducibility <strong>of</strong> single gene<br />

measurements. Two reports indicated high reproducibility <strong>of</strong> pharmacogenomic


Gene Expression Pr<strong>of</strong>iling to Personalize Chemotherapy Selection 83<br />

prediction results in replicate experiments from the same RNA, when the same<br />

experimental procedures were used (26,32). Furthermore, when random noise<br />

was added to corrupt the gene expression data, a remarkable degree <strong>of</strong> robustness<br />

was observed in the face <strong>of</strong> noise.<br />

However, caution must be exercised when trying to reproduce prediction<br />

results across different platforms. There are multiple sources <strong>of</strong> variation that<br />

will decrease reproducibility including differences in probe set and different<br />

normalization methods. Essentially, all published results that attempted crossplatform<br />

testing <strong>of</strong> gene signatures reported diminished, but not completely<br />

lost, classification accuracy on data generated by other than the original<br />

platform (33).<br />

Preanalytical Sources <strong>of</strong> Variation<br />

Another potentially important source <strong>of</strong> variation in microarray data is preanalytical<br />

variations, including differences in tissue sampling and RNA degradation.<br />

There are several different techniques to obtain tissue from a patient for<br />

pharmacogenomic studies, including excisional biopsy (with or without microdissection),<br />

fine needle aspiration (FNA), and core needle biopsy (CNB). Both<br />

needle biopsy techniques yield sufficient RNA for gene expression pr<strong>of</strong>iling with<br />

DNA microarrays (1–3 mg) (34–36). However, different sampling methods may<br />

result in different cellular composition and distinct gene expression pr<strong>of</strong>iles (37).<br />

FNA specimens contain mostly neoplastic cells with few infiltrating leukocytes,<br />

whereas the tumor cell content <strong>of</strong> CNB can vary from 10% to 90%. It is<br />

important to consider the different amount <strong>of</strong> stromal cell that might be present<br />

in a specimen, because the presence or absence <strong>of</strong> stromal genes might influence<br />

prognostic or predictive gene signatures. Also, in studies comparing pre- and<br />

postchemotherapy samples, an important source <strong>of</strong> variation may be the different<br />

cellular composition <strong>of</strong> the specimens.<br />

RNA is an easily degradable molecule and therefore was considered<br />

suboptimal target for marker measurements compared with DNA or proteins.<br />

The development <strong>of</strong> single step universal RNA stabilizing and preserving<br />

reagents (RNA later, Ambion) has completely changed the field <strong>of</strong> RNA-based<br />

diagnostics (38). Today, RNA can be stabilized within seconds after a fresh<br />

tissue biopsy has been obtained and stored at room temperature for several days<br />

(if needed) due to readily available commercial RNA stabilizing solutions.<br />

Under these circumstances, RNA degradation from properly collected fresh<br />

tissues is now rare. Interestingly, it is now more cumbersome to preserve proteins<br />

for future analysis than it is to preserve RNA. A single step, universal<br />

protein preserving solution is yet to be discovered. However, it is important to<br />

realize that surgical resection methods (i.e., length <strong>of</strong> warm ischemia) and the<br />

length <strong>of</strong> time between removing a piece <strong>of</strong> tissue from the body and placing it<br />

into RNA preservative have a major effect on RNA quality (39).


84 Liedtke and Pusztai<br />

STATISTICAL DESIGN OF PHARMACOGENOMIC<br />

DISCOVERY AND VALIDATION STUDIES<br />

Clinical biomarker discovery and validation is governed by similar statistical<br />

rules as therapeutic clinical trials (40). Estimation <strong>of</strong> discovery sample size is<br />

essential to develop a reliable pharmacogenomic predictor, and appropriately<br />

powered independent validation studies are needed to convincingly demonstrate<br />

its predictive values. There are several statistical methods that could be used to<br />

estimate the number <strong>of</strong> patients needed to develop a clinical outcome predictor<br />

(41). One simple approach is to power the study to discover individual genes that<br />

show a predefined level <strong>of</strong> differential expression (<strong>of</strong>ten expressed in statistical<br />

terms as standard effect size) between outcome groups. However, the number<br />

and extent <strong>of</strong> differentially expressed genes are <strong>of</strong>ten unknown in advance, and<br />

the multiple comparisons (i.e., thousands <strong>of</strong> different measurements on each<br />

case) represent a major bias in this approach. Another approach takes into<br />

account accumulating preliminary results to estimate the sample size needed for<br />

discovery (42). With multivariate prediction algorithms, it is expected that the<br />

predictive accuracy increases as the predictor is developed from and trained on<br />

larger and larger data sets until it reaches a plateau (i.e., best possible predictor<br />

within the constrains <strong>of</strong> the technology). The general principle is to fit a learning<br />

curve into existing data by incrementally increasing the training set size and<br />

extrapolate results to larger sample sizes. This is an intuitively appealing<br />

approach, but it requires a substantial amount <strong>of</strong> preliminary results from many<br />

dozens <strong>of</strong> cases. It also implies that discovery sample size will vary depending<br />

on the difficulty <strong>of</strong> separating the two response groups on the basis <strong>of</strong> gene<br />

expression results. When we applied this method to human breast cancer gene<br />

expression data to estimate the number <strong>of</strong> cases needed for discovery <strong>of</strong> a<br />

chemotherapy response predictor, the results indicated that 80 to 100 cases<br />

would yield a predictor as good as extending the sample size to 200 (32).<br />

To assess the accuracy <strong>of</strong> the predictor, it is necessary to test it on independent<br />

patient cohorts. Often, there are not enough cases available for validation,<br />

and investigators report cross-validation results. In cross-validation, some<br />

samples are left out from the marker development process and later on used to<br />

test the accuracy <strong>of</strong> the predictor, and this process is repeated many times. This<br />

leave-one-out cross-validation method gives a reliable estimate <strong>of</strong> the overall<br />

accuracy <strong>of</strong> the marker development process, but it is not the validation <strong>of</strong> a<br />

particular predictor. The genes that contribute to the predictors differ from<br />

iteration to iteration. The final single best predictor that is developed from the<br />

entire training data must be tested on completely separate cases to assess its<br />

predictive value. The goal <strong>of</strong> an independent validation study is to (i) define the<br />

true sensitivity, specificity, and the positive and negative predictive values (PPV<br />

and NPV) with narrow confidence intervals (CI) and (ii) to prove clinical utility<br />

<strong>of</strong> the test. Since the estimated performance characteristics <strong>of</strong> the test are known<br />

from the cross-validation results, the design <strong>of</strong> validation studies is more


Gene Expression Pr<strong>of</strong>iling to Personalize Chemotherapy Selection 85<br />

straightforward. Sample size for a single arm validation studies can be calculated<br />

on the basis <strong>of</strong> the known historical response rate in unselected patients and the<br />

estimated PPV <strong>of</strong> the genomic predictor. For example, if we assume that the<br />

response rate in unselected patients is 30% and the true PPV <strong>of</strong> the test is 60%<br />

(two-fold greater chance <strong>of</strong> response in “test-positive cases”), the sample<br />

size could be calculated as follows: The lower bound <strong>of</strong> a two-sided 95% CI<br />

should not be less than 0.3 (since the expected maximum pCR rate is 30% in<br />

unselected cases). This implies that the standard error for the PPV is less than 0.1<br />

and therefore the study must include at least 24 “test-positive” patients. The<br />

proportion <strong>of</strong> patients with response is expected to be around 30%, and an<br />

accurate prediction algorithm should generate a proportion that test positive also<br />

about 0.20 to 0.40, this equates to a total sample size <strong>of</strong> 80 to 120 patients for<br />

validation. If the predictor performs better and the true PPV is higher than 60%,<br />

fewer patients may be sufficient. A randomized trial design could simultaneously<br />

address response rates in selected and unselected patients and could also assess<br />

treatment specificity <strong>of</strong> the predictor.<br />

BREAST CANCER PHARMACOGENOMIC RESPONSE PREDICTORS<br />

Supervised <strong>Ph</strong>armacogenomic Predictors Developed from Human Data<br />

Results from several phase I genomic response marker discovery studies were<br />

reported that compared gene expression pr<strong>of</strong>iles <strong>of</strong> cases with response and noresponse,<br />

and used the genes that were differentially expressed to construct<br />

multigene predictors. Some <strong>of</strong> these markers were also assessed in small independent<br />

validation cohorts. Table 1 summarizes the currently available literature<br />

in this field. Chang et al. presented the first study that suggested the feasibility <strong>of</strong><br />

developing multigene predictors <strong>of</strong> response to neoadjuvant chemotherapy. They<br />

developed a 92-gene predictor for clinical response to docetaxel using gene<br />

expression data from 20 patients with locally advanced breast cancer (36).<br />

Leave-one-out partial cross-validation indicated a predictive accuracy <strong>of</strong> 88%<br />

(95% CI, 68–97%), sensitivity <strong>of</strong> 85% (95% CI, 55–98%), and specificity <strong>of</strong> 91%<br />

(95% CI, 59–100%). Larger scale independent validation <strong>of</strong> this marker set has<br />

not yet been reported. Another small study examined 24 breast cancer patients<br />

treated with neoadjuvant paclitaxel, 5-fluoruracil, doxorubicin, and cyclophosphamide<br />

(T/FAC) chemotherapy and developed a 74-gene predictor for<br />

pCR (43). This multigene predictor model, tested on an independent cohort <strong>of</strong><br />

12 patients, showed a predictive accuracy <strong>of</strong> 78% (95% CI, 52–94%), PPV <strong>of</strong> 100%<br />

(95% CI, 29–100%), NPV <strong>of</strong> 73% (95% CI, 45–92%), sensitivity <strong>of</strong> 43% (95%<br />

CI, 10–82%), and specificity <strong>of</strong> 100% (95% CI, 72–100%). A larger follow-up<br />

study included 82 cases for marker discovery and examined 780 distinct pharmacogenomic<br />

predictors in cross-validation (32). Many <strong>of</strong> the distinct predictors<br />

demonstrated statistically equal performance in complete cross-validation. This<br />

(text continues on page 91)


86 Liedtke and Pusztai<br />

Table 1 Overview <strong>of</strong> Published Studies on <strong>Ph</strong>armacogenomic Predictors <strong>of</strong> Response to Neoadjuvant Chemotherapy in <strong>Breast</strong> <strong>Cancer</strong><br />

Patient characteristics and<br />

treatment Method/hypothesis<br />

Gene expression<br />

platform Results Comment<br />

Chang et al.,<br />

Lancet 2003<br />

(36)<br />

24 (¼dev.) and 6 (¼val.) pts.<br />

with locally advanced breast<br />

cancer<br />

? 4 cycles 3-weekly<br />

docetaxel<br />

Development,<br />

LOOCV and<br />

independent<br />

validation <strong>of</strong> a<br />

gene expression<br />

pr<strong>of</strong>ile<br />

predictive for<br />

clinical response<br />

(25% RD)<br />

HgU95-Av2<br />

(12,625 genes)<br />

92 differentially<br />

expressed genes (p <<br />

0.001) from different<br />

functional classes;<br />

Combination into<br />

prediction model<br />

LOOCV: 88% accuracy,<br />

92% PPV, 83%<br />

NPV, 85% sensitivity,<br />

91% specificity<br />

Independent validation:<br />

correct classification<br />

<strong>of</strong> six pts. with sensitive<br />

tumors<br />

First study developing<br />

a multigene<br />

predictor for<br />

response to<br />

chemotherapy in<br />

breast cancer<br />

patients<br />

Ayers et al., JCO<br />

2004 (43)<br />

24 (¼dev.) and 12 (¼val.) pts.<br />

with primary breast cancer<br />

? 3 cycles 3-weekly or 12<br />

cycles weekly paclitaxel<br />

followed by 4 cycles <strong>of</strong> FAC<br />

(T/FAC)<br />

Development and<br />

independent<br />

validation <strong>of</strong><br />

gene expression<br />

predictor for<br />

pCR<br />

cDNA arrays<br />

(30,721 genes)<br />

Development <strong>of</strong><br />

74-gene-multigene<br />

predictor<br />

Independent validation:<br />

78% accuracy, 100%<br />

PPV, 73% NPV,<br />

43% sensitivity,<br />

100% specificity<br />

No single gene sufficiently<br />

associated<br />

with pCR to serve as<br />

single valid marker<br />

Pro<strong>of</strong><strong>of</strong>principle<br />

for building multigene<br />

predictors for<br />

chemotherapy<br />

response


Gene Expression Pr<strong>of</strong>iling to Personalize Chemotherapy Selection 87<br />

Table 1 Overview <strong>of</strong> Published Studies on <strong>Ph</strong>armacogenomic Predictors <strong>of</strong> Response to Neoadjuvant Chemotherapy in <strong>Breast</strong> <strong>Cancer</strong> (Continued )<br />

Patient characteristics and<br />

treatment Method/hypothesis<br />

Gene expression<br />

platform Results Comment<br />

Iwao-Koizumi,<br />

JCO 2005 (44)<br />

44 (¼dev.) and 26 (¼val.) pts.<br />

with stage II/IV breast cancer<br />

? 4 cycles <strong>of</strong> docetaxel<br />

Development and<br />

independent<br />

validation <strong>of</strong><br />

gene expression<br />

predictor for<br />

clinical response<br />

(WHO criteria)<br />

ATAC-PCR (2,453<br />

selected genes)<br />

Development <strong>of</strong><br />

85-gene signature<br />

Independent validation:<br />

81% accuracy, 73%<br />

PPV, 91% NPV,<br />

92% sensitivity, 71%<br />

specificity<br />

Genes controlling the<br />

cellular redox<br />

environment associated<br />

with docetaxel<br />

resistance<br />

? supported by<br />

in vitro results<br />

Hannemann et al.,<br />

JCO 2005 (45)<br />

48 pts. with LABC<br />

? cycles <strong>of</strong> doxorubicin/<br />

docetaxel (AD, n ¼ 24)<br />

or doxorubicin/cyclophosphamide<br />

(AC, n ¼ 24)<br />

Identification <strong>of</strong><br />

gene expression<br />

predictor for<br />

pCR or nPCR<br />

for response to<br />

AD and/or AC<br />

Proprietary array<br />

(18,432 genes)<br />

Failure to identify differentially<br />

expressed<br />

genes or define multigene<br />

predictor<br />

Hierarchical clustering:<br />

pCR/npCR samples<br />

did not cluster distinctly<br />

compared<br />

to RD<br />

Comparison <strong>of</strong> gene<br />

expression between<br />

pre- and<br />

postchemotherapy<br />

samples: more<br />

transcriptional<br />

changes among<br />

sensitive compared<br />

to resistant tumors<br />

Gianni et al., JCO<br />

2005 (46)<br />

89 (¼dev.) and 82 (¼val.) pts.<br />

with LABC<br />

? 3 cycles <strong>of</strong> 3-weekly<br />

doxorubicin/paclitaxel<br />

followed by 12 cycles<br />

weekly paclitaxel<br />

? 3 cycles 3-weekly or<br />

12 cycles weekly paclitaxel<br />

followed by 4 cycles <strong>of</strong><br />

FAC (T/FAC)<br />

Identification <strong>of</strong><br />

gene expression<br />

markers that<br />

predict response<br />

Correlation<br />

between RS and<br />

response<br />

RT-PCR (384<br />

genes)<br />

Affymetrix HG-<br />

U133 Plus 2.0<br />

86 genes correlating<br />

with pCR (p < 0.05)<br />

forming three clusters<br />

(ER, proliferation<br />

and immunerelated<br />

gene cluster)<br />

RS correlated with pCR<br />

86-gene model developed<br />

and successfully<br />

validated<br />

(Continued)


88 Liedtke and Pusztai<br />

Table 1 Overview <strong>of</strong> Published Studies on <strong>Ph</strong>armacogenomic Predictors <strong>of</strong> Response to Neoadjuvant Chemotherapy in <strong>Breast</strong> <strong>Cancer</strong> (Continued )<br />

Patient characteristics and<br />

treatment Method/hypothesis<br />

Gene expression<br />

platform Results Comment<br />

Folgueira et al.,<br />

Clin <strong>Cancer</strong> Res<br />

2005 (47)<br />

38 (¼dev.) and 13 (¼val.)<br />

31 (¼dev.) and 13 (¼val.)<br />

? 4 cycles 3-weekly<br />

doxorubicin/cyclophosphamide<br />

(AC)<br />

Development and<br />

independent<br />

validation <strong>of</strong><br />

gene expression<br />

predictor for<br />

pCR<br />

692 cDNA<br />

microarray<br />

4,608 gene<br />

platform<br />

25 differentially<br />

expressed transcripts<br />

between responders<br />

and nonresponders,<br />

3 gene classifier could<br />

not be validated;<br />

successful development<br />

and validation<br />

(LOOCV and independent<br />

dataset) <strong>of</strong><br />

distinct three-gene<br />

classifier<br />

Dressman et al.,<br />

Clin <strong>Cancer</strong> Res<br />

2006 (48)<br />

37 pts. with stage IIB/III<br />

breast cancer<br />

? 4 cycles <strong>of</strong> liposomal<br />

doxorubicin/paclitaxel þ<br />

local whole breast<br />

hyperthermia<br />

Definition <strong>of</strong> gene<br />

signature for<br />

(a) inflammatory<br />

breast cancer<br />

(IBC), (b) tumor<br />

hypoxia, and (c)<br />

clinical response<br />

and validation in<br />

two independent<br />

datasets<br />

Affymetrix<br />

HG-U133<br />

Plus 2.0<br />

Identification and validation<br />

(LOOCV) <strong>of</strong><br />

(a) 22-gene signature<br />

for IBC<br />

(b) 18-gene signature<br />

for tumor hypoxi<br />

Failure to identify gene<br />

signature predicting<br />

clinical response<br />

Development and<br />

successful validation<br />

(2 independent<br />

datasets)<strong>of</strong>a<br />

signature predictive<br />

for lymph node<br />

persistence


Gene Expression Pr<strong>of</strong>iling to Personalize Chemotherapy Selection 89<br />

Table 1 Overview <strong>of</strong> Published Studies on <strong>Ph</strong>armacogenomic Predictors <strong>of</strong> Response to Neoadjuvant Chemotherapy in <strong>Breast</strong> <strong>Cancer</strong> (Continued )<br />

Patient characteristics and<br />

treatment Method/hypothesis<br />

Gene expression<br />

platform Results Comment<br />

Thuerigen et al.,<br />

JCO 2006 (49)<br />

52 (¼dev/GEsDoc) and 48<br />

(¼val./GEDoc) pts. with<br />

breast cancer<br />

? 5 cycles biweekly gemcitabine/epirubicin<br />

followed by<br />

4 cycles biweekly docetaxel<br />

(GEsDoc)<br />

? 6 cycles <strong>of</strong> 3-weekly<br />

gemcitabine/epirubicin/<br />

docetaxel (GEDoc)<br />

Development and<br />

independent<br />

validation <strong>of</strong><br />

gene expression<br />

predictor for<br />

pCR<br />

Oligonucleotide<br />

array (21329<br />

genes)<br />

512-gene-signature<br />

predictive <strong>of</strong> pCR<br />

with: independent<br />

validation: 88%<br />

accuracy, 64% PPV,<br />

95% NPV, 78%<br />

sensitivity, 90%<br />

specificity<br />

HER2/neu expression<br />

independently<br />

predictive <strong>of</strong><br />

response<br />

Park et al., <strong>Breast</strong><br />

<strong>Cancer</strong> Res<br />

Treat 2006 (50)<br />

21 pts. with stage II/III breast<br />

cancer<br />

? 4 cycles <strong>of</strong> 3-weekly<br />

FEC followed by 12 cycles<br />

<strong>of</strong> weekly paclitaxel<br />

Development and<br />

LOOCV <strong>of</strong><br />

ABC transporter<br />

gene expression<br />

pr<strong>of</strong>iles<br />

predictive for<br />

pCR<br />

Affymetrix HG-<br />

U133 Plus 2.0<br />

(49 ABC transporters)<br />

11 differentially expressed<br />

ABC transporters multigene<br />

predictor model <strong>of</strong><br />

differentially expressed<br />

ABC transporters predictive<br />

<strong>of</strong> pCR<br />

LOOCV: 93% accuracy,<br />

93% PPV, 94%<br />

NPV, 88% sensitivity,<br />

96% specificity<br />

Cleator et al.,<br />

<strong>Breast</strong> <strong>Cancer</strong><br />

Res Treat 2006<br />

(51)<br />

40 pts. with primary breast<br />

cancer<br />

? 6 cycles doxorubicin/<br />

cyclophosphamide (AC)<br />

Development and<br />

LOOCV <strong>of</strong><br />

predictor for<br />

clinical response<br />

Affymetrix<br />

HG-U133A<br />

253 differentially<br />

expressed genes;<br />

75 and 178 genes overexpressed<br />

in resistant<br />

and sensitive tumors,<br />

respectively<br />

Development and<br />

LOOCV <strong>of</strong> gene predictor<br />

for response<br />

Previously identified<br />

92-gene docetaxel<br />

classifier (36) predicted<br />

only 39–61%<br />

tumors correctly<br />

? gene expression<br />

pattern for docetaxel<br />

is not predictive for<br />

AC response<br />

(Continued)


90 Liedtke and Pusztai<br />

Table 1 Overview <strong>of</strong> Published Studies on <strong>Ph</strong>armacogenomic Predictors <strong>of</strong> Response to Neoadjuvant Chemotherapy in <strong>Breast</strong> <strong>Cancer</strong> (Continued )<br />

Patient characteristics and<br />

treatment Method/hypothesis<br />

Gene expression<br />

platform Results Comment<br />

Hess et al., JCO<br />

2006 (32)<br />

82 (¼dev.) and 51 (¼val.) pts.<br />

with primary breast cancer<br />

? 3 cycles 3-weekly or<br />

12 cycles weekly paclitaxel<br />

followed by 4 cycles <strong>of</strong> FAC<br />

(T/FAC)<br />

Development and<br />

independent<br />

validation <strong>of</strong><br />

gene expression<br />

predictor for<br />

pCR<br />

Affymetrix<br />

HG-U133A<br />

Identification <strong>of</strong> 30-<br />

gene predictor<br />

Validation fivefold<br />

cross-validation and<br />

permutation testing<br />

Independent validation:<br />

76% accuracy, 52%<br />

PPV, 96% NPV,<br />

92% sensitivity, 71%<br />

specificity<br />

Follow-up study<br />

<strong>of</strong> (43)<br />

Predictor preserved its<br />

predictive accuracy<br />

when applied to<br />

replicate samples<br />

Largest study to date<br />

Mina et al., <strong>Breast</strong><br />

<strong>Cancer</strong> Res<br />

Treat 2006 (52)<br />

45 pts. with stage II/III breast<br />

cancer (including 6 pts. with<br />

inflammatory BC)<br />

? 3 cycles biweekly doxorubicin<br />

and 6 cycles weekly<br />

docetaxel<br />

Identification <strong>of</strong><br />

genes correlating<br />

with pCR<br />

Correlation <strong>of</strong> pCR<br />

with RS or ERrelated<br />

genes<br />

Identification <strong>of</strong><br />

genes correlating<br />

with inflammatory<br />

phenotype<br />

RT-PCR assay<br />

(274 genes)<br />

22 <strong>of</strong> 274 candidate genes<br />

correlated with pCR (p <<br />

0.05) forming three large<br />

clusters (angiogenesis,<br />

proliferation, and invasion-related<br />

genes)<br />

Lack <strong>of</strong> correlation<br />

between pCR and<br />

RS/ER-related genes<br />

24/274 genes correlated<br />

with inflammatory BC<br />

Abbreviations: ABC transporter, ATP-binding cassette transporter; dev., development/training set; IBC, inflammatory breast cancer; LABC, locally advanced breast<br />

cancer; LOOCV, leave-one-out cross-validation; npCR, near pCR; NPV, negative predictive value; pCR, pathological complete response; PPV, positive predictive<br />

value; pts., patients; RD, residual disease; RS, recurrence score (Oncotype DX # ) (53); RT-PCR, reverse transcription polymerase chain reaction; val., validation set.


Gene Expression Pr<strong>of</strong>iling to Personalize Chemotherapy Selection 91<br />

observation is confirmed by similar reports in the genomic literature, which<br />

indicates that several different but equally good predictors can be developed<br />

from the same data. This is due to the coordinated expression <strong>of</strong> many thousands<br />

<strong>of</strong> genes. Any combination from a particular group <strong>of</strong> coexpressed genes can<br />

carry similar information. However, a nominally best 30-probe set genomic<br />

predictor <strong>of</strong> pCR (DLDA-30) was selected for independent validation on 51<br />

cases, and its performance was compared with a multivariate clinical variable–<br />

based predictor including age, estrogen receptor status, and grade. The genomic<br />

predictor showed a sensitivity <strong>of</strong> 92% that was significantly higher than that <strong>of</strong><br />

the clinical predictor (61%). The NPV (96% vs. 86%) and area under the<br />

receiving operator curve (0.877 vs. 0.811) were also slightly better. Several other<br />

smaller studies demonstrated similar pro<strong>of</strong> <strong>of</strong> concept that this supervised<br />

approach to marker development is feasible (Table 1).<br />

Genomic Predictors from Cell Line Models<br />

An interesting alternative approach is to use cancer models to develop chemotherapy<br />

response predictors and test the performance <strong>of</strong> these in human data. The<br />

gene expression pr<strong>of</strong>iles <strong>of</strong> breast cancer cell lines are different from those <strong>of</strong><br />

human breast cancers, but nevertheless, they recapitulate many biological and<br />

genomic properties <strong>of</strong> primary breast cancers (54). Several groups reported drugspecific<br />

transcriptional changes in response to chemotherapy exposure in vitro<br />

that might be exploited as predictors <strong>of</strong> therapy response in patients (55).<br />

However, the predictive value <strong>of</strong> these in vitro generated signatures in patients<br />

has not yet been studied extensively. Nevertheless, this is an important research<br />

direction that if working can have important implications for individualizing<br />

treatment regimens for patients. In fact one could envisage custom-assembled<br />

chemotherapy regimens based on sensitivity pr<strong>of</strong>iles for individual drugs. It is<br />

rapidly becoming feasible to test such concepts in silico. Several groups have<br />

made their complete gene expression data from neoadjuvant biopsies publicly<br />

available. Cell line-based predictive signatures can be tested on these already<br />

existing and clinically annotated human data sets (56).<br />

Gene Expression Changes in Serial Biopsies<br />

Similar to cell lines, transcriptional response to chemotherapy was also observed<br />

in serial biopsies before and during chemotherapy. The rationale is that these<br />

changes may be more predictive <strong>of</strong> response than baseline gene expression<br />

pr<strong>of</strong>iles. However, the few small studies that examined this approach reported<br />

very heterogeneous transcriptional response to chemotherapy (57). In essence,<br />

different genes change in different individuals. Sotiriou et al. examined gene<br />

expression changes after one complete cycle <strong>of</strong> neoadjuvant chemotherapy and<br />

demonstrated that tissue samples, obtained from patients with good response,<br />

exhibited much greater changes in gene expression than those from poor


92 Liedtke and Pusztai<br />

responders (34). These observations were corroborated by results obtained by<br />

Hannemann et al., who compared gene expression pr<strong>of</strong>iling on pretreatment<br />

biopsies (n ¼ 46) and posttreatment tumors (n ¼ 17) obtained at time <strong>of</strong> surgery.<br />

They also demonstrated that neoadjuvant chemotherapy causes major changes in<br />

gene expression in those cases who responded to treatment compared with those<br />

who did not (58,59). Several technical and theoretical concerns limit our<br />

enthusiasm for this approach to predictor development. It is difficult to obtain<br />

serial biopsies at regular intervals from patients who undergo chemotherapy.<br />

There is substantial variation in the actual timing <strong>of</strong> the biopsy compared with<br />

the planned time <strong>of</strong> obtaining the follow-up biopsy. For example, biopsies<br />

planned at 24 hours after chemotherapy are in fact usually collected during an<br />

18- to 36-hour window. Since the kinetics <strong>of</strong> important gene expression changes<br />

are unknown, this introduces unpredictable variation into already highly variable<br />

data. Also, there are altogether missing time points, which further reduce the<br />

already small sample size for discovery. These factors increase the risk for false<br />

discovery from serially collected small pharmacogenomic data sets. More<br />

importantly, the practical value <strong>of</strong> a response predictor that is applied to a biopsy<br />

after chemotherapy was given is not particularly appealing.<br />

FUTURE POTENTIAL OF PHARMACOGENOMICS<br />

<strong>Ph</strong>armacogenomic predictive marker discovery is a new science. It is relatively<br />

less mature than the field <strong>of</strong> prognostic gene signatures because <strong>of</strong> the lack <strong>of</strong><br />

preexisting tissue banks from patients who received preoperative chemotherapy.<br />

All human biopsies had to be collected in prospective clinical trials in the past<br />

few years, unlike the readily available frozen and archived tumor banks that were<br />

suitable for prognostic signature discovery and validation. However, several<br />

such data sets now exist and the field has likely entered an accelerated phase <strong>of</strong><br />

development.<br />

We have learned several important lessons. (i) It is clear that the gene<br />

expression pr<strong>of</strong>iles <strong>of</strong> breast cancers with extreme chemotherapy sensitivity are<br />

different from those with lesser sensitivity. Unfortunately, these gene expression<br />

differences correspond to major clinical pathological phenotypes such as ERnegative<br />

status and high grade that somewhat limits the clinical value <strong>of</strong> the<br />

current predictors. However, as larger data sets become available, it will be<br />

possible to attempt to develop predictors after stratification by known clinical<br />

pathological variables. (ii) It is also clear from many different reports in the<br />

genomic literature that there is no unique, single best signature for any particular<br />

classification problem. There are several distinct predictors that can perform<br />

equally well (60). (iii) The availability <strong>of</strong> large, public, clinically annotated,<br />

comprehensive gene expression databases started to revolutionize the biomarker<br />

discovery process. It is now possible to test in silico any mRNA-based marker<br />

for prognostic or chemotherapy or endocrine therapy predictive values, respectively,<br />

on data sets that were prospectively assembled for this purpose (61). As


Gene Expression Pr<strong>of</strong>iling to Personalize Chemotherapy Selection 93<br />

more neoadjuvant data sets become available from patient cohorts treated with<br />

different treatment regimens, it will be possible to examine the regimen specificity<br />

<strong>of</strong> the various predictors. (iv) The most important long-term benefit from<br />

pharmacogenomic research may not be the current predictive signatures or the<br />

ones that will be developed in the near future, which still represent relatively<br />

crude attempts to match patients with existing drugs, but rather the large<br />

molecular data sets that this research generates. These databases could allow the<br />

identification <strong>of</strong> novel drug targets in human data and the development <strong>of</strong> drugs<br />

that target subsets <strong>of</strong> patients with particular genomic abnormalities.<br />

REFERENCES<br />

1. National Comprehensive <strong>Cancer</strong> Network. Available at: http://www.nccn.org. Accessed<br />

June 2007.<br />

2. Goldhirsch A, Wood WC, Gelber RD, et al. Meeting highlights: updated international<br />

expert consensus on the primary therapy <strong>of</strong> early breast cancer. J Clin Oncol<br />

2003; 21(17):3357–3365.<br />

3. Scholl SM, Fourquet A, Asselain B, et al. Neoadjuvant versus adjuvant chemotherapy<br />

in premenopausal patients with tumours considered too large for breast<br />

conserving surgery: preliminary results <strong>of</strong> a randomised trial: S6. Eur J <strong>Cancer</strong> 1994;<br />

30A(5):645–652.<br />

4. Boughey JC, Peintinger F, Meric-Bernstam F, et al. Impact <strong>of</strong> preoperative versus<br />

postoperative chemotherapy on the extent and number <strong>of</strong> surgical procedures in<br />

patients treated in randomized clinical trials for breast cancer. Ann Surg 2006; 244(3):<br />

464–470.<br />

5. Makris A, Powles TJ, Ashley SE, et al. A reduction in the requirements for mastectomy<br />

in a randomized trial <strong>of</strong> neoadjuvant chemoendocrine therapy in primary<br />

breast cancer. Ann Oncol 1998; 9(11):1179–1184.<br />

6. Mauriac L, Durand M, Avril A, et al. Effects <strong>of</strong> primary chemotherapy in conservative<br />

treatment <strong>of</strong> breast cancer patients with operable tumors larger than 3 cm.<br />

Results <strong>of</strong> a randomized trial in a single centre. Ann Oncol 1991; 2(5):347–354.<br />

7. Fisher B, Bryant J, Wolmark N, et al. Effect <strong>of</strong> preoperative chemotherapy on the<br />

outcome <strong>of</strong> women with operable breast cancer. J Clin Oncol 1998; 16(8):2672–2685.<br />

8. Kuerer HM, Newman LA, Smith TL, et al. Clinical course <strong>of</strong> breast cancer patients<br />

with complete pathologic primary tumor and axillary lymph node response to<br />

doxorubicin-based neoadjuvant chemotherapy. J Clin Oncol 1999; 17(2):460–469.<br />

9. Abu-Farsakh H, Sneige N, Atkinson EN, et al. Pathologic predictors <strong>of</strong> tumor<br />

response to preoperative chemotherapy in locally advanced breast carcinoma. <strong>Breast</strong><br />

J 1995; 1(2):96–101.<br />

10. Colleoni M, Minchella I, Mazzarol G, et al. Response to primary chemotherapy in<br />

breast cancer patients with tumors not expressing estrogen and progesterone receptors.<br />

Ann Oncol 2000; 11(8):1057–1059.<br />

11. Chang J, Powles TJ, Allred DC, et al. Biologic markers as predictors <strong>of</strong> clinical<br />

outcome from systemic therapy for primary operable breast cancer. J Clin Oncol<br />

1999; 17(10):3058–3063.<br />

12. Ellis P, Smith I, Ashley S, et al. Clinical prognostic and predictive factors for<br />

primary chemotherapy in operable breast cancer. J Clin Oncol 1998; 16(1):107–114.


94 Liedtke and Pusztai<br />

13. Rouzier R, Pusztai L, Delaloge S, et al. Nomograms to predict pathologic complete<br />

response and metastasis-free survival after preoperative chemotherapy for breast<br />

cancer. J Clin Oncol 2005; 23(33):8331–8339.<br />

14. Rozan S, Vincent-Salomon A, Zafrani B, et al. No significant predictive value <strong>of</strong><br />

c-erbB-2 or p53 expression regarding sensitivity to primary chemotherapy or<br />

radiotherapy in breast cancer. Int J <strong>Cancer</strong> 1998; 79(1):27–33.<br />

15. Petit T, Borel C, Ghnassia JP, et al. Chemotherapy response <strong>of</strong> breast cancer depends<br />

on HER-2 status and anthracycline dose intensity in the neoadjuvant setting. Clin<br />

<strong>Cancer</strong> Res 2001; 7(6):1577–1581.<br />

16. Zhang F, Yang Y, Smith T, et al. Correlation between HER2 expression and response<br />

to neoadjuvant FAC chemotherapy in breast cancer. <strong>Cancer</strong> 2003; 97(7):1758–1765.<br />

17. Silberstein GB. Tumour-stromal interactions. Role <strong>of</strong> the stroma in mammary<br />

development. <strong>Breast</strong> <strong>Cancer</strong> Res 2001; 3(4):218–223.<br />

18. Coussens LM, Werb Z. Inflammation and cancer. Nature 2002; 420(6917):860–867.<br />

19. Burns TF, El-Deiry WS. Microarray analysis <strong>of</strong> p53 target gene expression patterns<br />

in the spleen and thymus in response to ionizing radiation. <strong>Cancer</strong> Biol Ther 2003; 2(4):<br />

431–443.<br />

20. Reed JC. Dysregulation <strong>of</strong> apoptosis in cancer. J Clin Oncol 1999; 17(9):2941–2953.<br />

21. Fodor SP, Rava RP, Huang XC, et al. Multiplexed biochemical assays with biological<br />

chips. Nature 1993; 364(6437):555–556.<br />

22. Kattan MW, Scardino PT. Prediction <strong>of</strong> progression: nomograms <strong>of</strong> clinical utility.<br />

Clin Prostate <strong>Cancer</strong> 2002; 1(2):90–96.<br />

23. Pounds S, Morris SW. Estimating the occurrence <strong>of</strong> false positive and false negatives<br />

in microarray studies by approximating and partitioning the empirical distribution <strong>of</strong><br />

p-values. Bioinformatics 2003; 19(10):1236–1242.<br />

24. Pusztai L, Ayers M, Stec J, et al. Gene expression pr<strong>of</strong>iles obtained from fine-needle<br />

aspirations <strong>of</strong> breast cancer reliably identify routine prognostic markers and reveal<br />

large-scale molecular differences between estrogen-negative and estrogen-positive<br />

tumors. Clin <strong>Cancer</strong> Res 2003; 9(7):2406–2415.<br />

25. Sotiriou C, Wirapati P, Loi S, et al. Gene expression pr<strong>of</strong>iling in breast cancer:<br />

understanding the molecular basis <strong>of</strong> histologic grade to improve prognosis. J Natl<br />

<strong>Cancer</strong> Inst 2006; 98(4):262–272.<br />

26. Anderson K, Hess KR, Kapoor M, et al. Reproducibility <strong>of</strong> gene expression signature<br />

based predictions in replicate experiments. Clin <strong>Cancer</strong> Res 2006; 12(6):1721–1727.<br />

27. Pusztai L. Perspectives and challenges <strong>of</strong> clinical pharmacogenomics in cancer.<br />

<strong>Ph</strong>armacogenomics 2004; 5(5):451–454.<br />

28. MAQC Consortium, . The MicroArray Quality Control (MAQC) project shows interand<br />

intraplatform reproducibility <strong>of</strong> gene expression measurements. Nat Biotechnol<br />

2006; 24(9):1151–1161.<br />

29. Klebanov L, Yakovlev A. How high is the level <strong>of</strong> technical noise in microarray<br />

data? Biol Direct 2007; 2(1):9.<br />

30. Mecham BH, Klus GT, Strovel J, et al. Sequence-matched probes produce increased<br />

cross-platform consistency and more reproducible biological results in microarraybased<br />

gene expression measurements. Nucleic Acids Res 2004; 32(9):e74.<br />

31. Ji Y, Coombes K, Zhang J, et al. RefSeq refinements <strong>of</strong> UniGene-based gene<br />

matching improve the correlation <strong>of</strong> expression measurements between two microarray<br />

platforms. Appl Bioinformatics 2006; 5(2):89–98.


Gene Expression Pr<strong>of</strong>iling to Personalize Chemotherapy Selection 95<br />

32. Hess KR, Anderson K, Symmans WF, et al. <strong>Ph</strong>armacogenomic predictor <strong>of</strong> sensitivity<br />

to preoperative chemotherapy with paclitaxel and fluorouracil, doxorubicin,<br />

and cyclophosphamide in breast cancer. J Clin Oncol 2006; 24(26):4236–4244.<br />

33. Stec J, Wang J, Coombes K, et al. Comparison <strong>of</strong> the predictive accuracy <strong>of</strong> DNA<br />

array-based multigene classifiers across cDNA arrays and Affymetrix GeneChips.<br />

J Mol Diagn 2005; 7(3):357–367.<br />

34. Sotiriou C, Powles TJ, Dowsett M, et al. Gene expression pr<strong>of</strong>iles derived from fine<br />

needle aspiration correlate with response to systemic chemotherapy in breast cancer.<br />

<strong>Breast</strong> <strong>Cancer</strong> Res 2002; 4(3):R3.<br />

35. Assersohn L, Gangi L, Zhao Y, et al. The feasibility <strong>of</strong> using fine needle aspiration<br />

from primary breast cancers for cDNA microarray analyses. Clin <strong>Cancer</strong> Res 2002;<br />

8(3):794–801.<br />

36. Chang JC, Wooten EC, Tsimelzon A, et al. Gene expression pr<strong>of</strong>iling for the prediction<br />

<strong>of</strong> therapeutic response to docetaxel in patients with breast cancer. Lancet<br />

2003; 362(9381):362–369.<br />

37. Symmans WF, Ayers M, Clark EA, et al. Total RNA yield and microarray gene<br />

expression pr<strong>of</strong>iles from fine-needle aspiration biopsy and core-needle biopsy<br />

samples <strong>of</strong> breast carcinoma. <strong>Cancer</strong> 2003; 97(12):2960–2671.<br />

38. Mutter GL, Zahrieh D, Liu C, et al. Comparison <strong>of</strong> frozen and RNALater solid tissue<br />

storage methods for use in RNA expression microarrays. BMC Genomics 2004; 5(1):88.<br />

39. Dash A, Maine IP, Varambally S, et al. Changes in differential gene expression<br />

because <strong>of</strong> warm ischemia time <strong>of</strong> radical prostatectomy specimens. Am J Pathol<br />

2002; 161(5):1743–1748.<br />

40. Simon R. Roadmap for developing and validating therapeutically relevant genomic<br />

classifiers. J Clin Oncol 2005; 23(29):7332–7341.<br />

41. Pusztai L, Hess KR. Clinical trial design for microarray predictive marker discovery<br />

and assessment. Ann Oncol 2004; 15(12):1731–1737.<br />

42. Mukherjee S, Tamayo P, Rogers S, et al. Estimating dataset size requirements for<br />

classifying DNA microarray data. J Comput Biol 2003; 10(2):119–142.<br />

43. Ayers M, Symmans WF, Stec J, et al. Gene expression pr<strong>of</strong>iles predict complete<br />

pathologic response to neoadjuvant paclitaxel and fluorouracil, doxorubicin, and<br />

cyclophosphamide chemotherapy in breast cancer. J Clin Oncol 2004; 22(12):<br />

2284–2293.<br />

44. Iwao-Koizumi K, Matoba R, Ueno N, et al. Prediction <strong>of</strong> docetaxel response in human<br />

breast cancer by gene expression pr<strong>of</strong>iling. J Clin Oncol 2005; 23(3):422–431.<br />

45. Hannemann J, Oosterkamp HM, Bosch CA, et al. Changes in gene expression<br />

associated with response to neoadjuvant chemotherapy in breast cancer. J Clin Oncol<br />

2005; 23(15):3331–3342.<br />

46. Gianni L, Zambetti M, Clark K, et al. Gene expression pr<strong>of</strong>iles in paraffin-embedded<br />

core biopsy tissue predict response to chemotherapy in women with locally advanced<br />

breast cancer. J Clin Oncol 2005; 23(29):7265–7277.<br />

47. Folgueira MA, Carraro DM, Brentani H, et al. Gene expression pr<strong>of</strong>ile associated<br />

with response to doxorubicin-based therapy in breast cancer. Clin <strong>Cancer</strong> Res 2005;<br />

11(20):7434–7443.<br />

48. Dressman HK, Hans C, Bild A, et al. Gene expression pr<strong>of</strong>iles <strong>of</strong> multiple breast<br />

cancer phenotypes and response to neoadjuvant chemotherapy. Clin <strong>Cancer</strong> Res<br />

2006; 12(3 pt 1):819–826.


96 Liedtke and Pusztai<br />

49. Thuerigen O, Schneeweiss A, Toedt G, et al. Gene expression signature predicting<br />

pathologic complete response with gemcitabine, epirubicin, and docetaxel in primary<br />

breast cancer. J Clin Oncol 2006; 24(12):1839–1845.<br />

50. Park S, Shimizu C, Shimoyama T, et al. Gene expression pr<strong>of</strong>iling <strong>of</strong> ATP-binding<br />

cassette (ABC) transporters as a predictor <strong>of</strong> the pathologic response to neoadjuvant<br />

chemotherapy in breast cancer patients. <strong>Breast</strong> <strong>Cancer</strong> Res Treat 2006; 99(1):9–17.<br />

51. Cleator S, Tsimelzon A, Ashworth A, et al. Gene expression patterns for doxorubicin<br />

(Adriamycin) and cyclophosphamide (cytoxan) (AC) response and resistance. <strong>Breast</strong><br />

<strong>Cancer</strong> Res Treat 2006; 95(3):229–233.<br />

52. Mina L, Soule SE, Badve S, et al. Predicting response to primary chemotherapy:<br />

gene expression pr<strong>of</strong>iling <strong>of</strong> paraffin-embedded core biopsy tissue. <strong>Breast</strong> <strong>Cancer</strong><br />

Res Treat 2007; 103(2):197–208.<br />

53. Paik S, Shak S, Tang G, et al. A multigene assay to predict recurrence <strong>of</strong> tamoxifentreated,<br />

node-negative breast cancer. N Engl J Med 2004; 351(27):2817–2826.<br />

54. Neve RM, Chin K, Fridlyand J, et al. A collection <strong>of</strong> breast cancer cell lines for the<br />

study <strong>of</strong> functionally distinct cancer subtypes. <strong>Cancer</strong> Cell 2006; 10(6):515–527.<br />

55. Troester MA, Hoadley KA, Sorlie T, et al. Cell-type-specific responses to chemotherapeutics<br />

in breast cancer. <strong>Cancer</strong> Res 2004; 64(12):4218–4226.<br />

56. Potti A, Dressman HK, Bild A, et al. Genomic signatures to guide the use <strong>of</strong> chemotherapeutics.<br />

Nat Med 2006; 12(11):1294–1300.<br />

57. Buchholz TA, Stivers DN, Stec J, et al. Global gene expression changes during<br />

neoadjuvant chemotherapy for human breast cancer. <strong>Cancer</strong> J 2002; 8(6):461–468.<br />

58. Hannemann J, Oosterkamp HM, Bosch CAJ, et al. Changes in gene expression<br />

pr<strong>of</strong>iling due to primary chemotherapy in patients with locally advanced breast<br />

cancer. Proc Am Soc Clin Oncol 2004; 22(14S):502 (abstract).<br />

59. Modlich O, Prisack HB, Munnes M, et al. Immediate gene expression changes after<br />

the first course <strong>of</strong> neoadjuvant chemotherapy in patients with primary breast cancer<br />

disease. Clin <strong>Cancer</strong> Res 2004; 10(19):6418–6431.<br />

60. Fan C, Oh DS, Wessels L, et al. Concordance among gene-expression-based predictors<br />

for breast cancer. N Engl J Med 2006; 355(6):560–569.<br />

61. Andre F, Hatzis C, Anderson K, et al. Microtubule-associated protein-tau is a<br />

bifunctional predictor <strong>of</strong> endocrine sensitivity and chemotherapy resistance in<br />

estrogen receptor-positive breast cancer. Clin <strong>Cancer</strong> Res 2007; 13(7):2061–2067.


7<br />

Gene Expression–Based Predictors<br />

<strong>of</strong> Prognosis and Response to<br />

Chemotherapy in <strong>Breast</strong> <strong>Cancer</strong><br />

Soonmyung Paik<br />

Division <strong>of</strong> Pathology, NSABP Foundation, Pittsburgh, Pennsylvania, U.S.A.<br />

INTRODUCTION<br />

With the explosion <strong>of</strong> genomic technology, numerous studies have demonstrated<br />

the potential clinical utility <strong>of</strong> gene expression pr<strong>of</strong>iles as a predictor <strong>of</strong> prognosis<br />

and response to therapy. Of particular interest is the demonstration that<br />

there is a general link between poor prognostic features and response to chemotherapy.<br />

In addition, linearity <strong>of</strong> RNA-based assays shows a strong promise<br />

that they can be very useful as predictive assays for targeted therapies.<br />

Since many excellent reviews have been written on this subject, this<br />

chapter will concentrate mainly on conceptual background to aid the clinical<br />

readers understand what is going on in the field rather than going into details.<br />

GENE EXPRESSION PROFILING PROVIDES<br />

PROGNOSTIC INFORMATION<br />

Gene expression pr<strong>of</strong>iling using microarray platforms allows interrogation <strong>of</strong><br />

expression levels <strong>of</strong> essentially all known genes in the tumor cells (1). In most<br />

studies, gross tumor rich area is homogenized to extract RNA. Therefore, the<br />

pr<strong>of</strong>ile that is achieved is mean expression levels <strong>of</strong> each gene among different<br />

97


98 Paik<br />

cell population in the tumor, including host stromal cells or inflammatory infiltrates.<br />

It is also important to realize that current methods do not have sensitivity to<br />

allow absolute measurement <strong>of</strong> gene expression levels <strong>of</strong> all genes printed on the<br />

microarray. Instead only about 50% <strong>of</strong> genes will have enough signal above<br />

background level (so-called present call). Different microarray platforms use DNA<br />

probes <strong>of</strong> various lengths (ranging from 22-mer oligonucleotide in Affymetrix<br />

GeneChip, 60-mer oligonucleotides for Agilent arrays, and average 1 kilobase for<br />

cDNA arrrays) representing different positions in each gene between the arrays.<br />

Sensitivity and dynamic range for measuring a specific gene could be widely<br />

different among different array platforms. Thus different studies using different<br />

platforms may result in identifying different set <strong>of</strong> genes representing same clinical<br />

phenotype (2). Obviously most studies also differ in the study population.<br />

Until recently microarray gene expression pr<strong>of</strong>iling required fresh or snap frozen<br />

tissue, seriously limiting the scale <strong>of</strong> studies conducted so far.<br />

It is not entirely surprising to find gene expression pr<strong>of</strong>iling to provide<br />

prognostic information, since after all morphology-based prognostic factors such<br />

as tumor histological grade is essentially end result <strong>of</strong> gene expression changes in<br />

the tumor cells. Two seminal papers using microarray gene expression pr<strong>of</strong>iling,<br />

one by Sorlie et al. (3) and the other by van ’t Veer et al. (4), clearly demonstrated<br />

the association <strong>of</strong> gene expression pr<strong>of</strong>iles and prognosis <strong>of</strong> patients diagnosed<br />

with early breast cancer. However, these two studies used very different<br />

approaches. In the seminal paper published in 2000, Perou et al. used unsupervised<br />

analysis <strong>of</strong> gene expression data and identified distinct biological subtypes<br />

(normal-like, luminal A and B, basal-like, and HER2) (5). Subsequently Sorlie<br />

et al. demonstrated that these subtypes confer differing prognosis with basal-like<br />

and HER2 associated with poor clinical outcome (3). On the other hand, van ’t<br />

Veer et al. conducted a supervised analysis to develop prognostic index based on<br />

expression levels <strong>of</strong> 70 genes identified from microarray analysis (4). This was<br />

then validated in a larger cohort (6). Recently Dr. Perou’s group has demonstrated<br />

that there is a significant overlap <strong>of</strong> information between the two approaches<br />

using same dataset—meaning those classified as good prognosis using 70-gene<br />

van ’t Veer set was found to be largely populated by luminal A subtype (2).<br />

While studies described above had enormous impact on how we view<br />

breast cancer, despite recent clearance <strong>of</strong> 70-gene Mammaprint 1 assay by the<br />

U.S. Food and Drug Administration (FDA), actual clinical implementation has<br />

been impeded by the need for fresh or snap frozen tumor tissue because <strong>of</strong> the<br />

requirement for highest quality RNA as a starting material for the microarray<br />

gene expression pr<strong>of</strong>iling.<br />

DEVELOPMENT AND VALIDATION OF OncotypeDx TM<br />

RECURRENCE SCORE ASSAY<br />

RNA isolated from formalin-fixed, paraffin-embedded tissue (FFPET) samples<br />

is a poor material for gene expression pr<strong>of</strong>iling. Masuda et al. analyzed the RNA<br />

extracted from FFPET for chemical modifications (2,7). In freshly made FFPET


Predictors <strong>of</strong> Prognosis and Response to Chemotherapy 99<br />

that were fixed and processed in ideal conditions (fixed in 10% buffered formalin<br />

at 48C), RNAs are fairly well preserved. Although the extracted RNAs did not<br />

show any sign <strong>of</strong> degradation compared with fresh samples, they were poor<br />

substrates for cDNA synthesis and subsequent polymerase chain reaction (PCR)<br />

amplification, so that only PCR amplification <strong>of</strong> short targets was possible. The<br />

investigators found addition <strong>of</strong> monomethylol (CH 2 OH) groups to all four bases<br />

to varying degrees and some adenines dimerized through metheylene bridging.<br />

In addition to the chemical modification, however, RNAs in FFPET continue to<br />

be degraded or fragmented over time during storage for reasons unclear. Cronin<br />

et al. systematically examined the quality <strong>of</strong> RNA extracted from breast cancer<br />

FFPET specimens taken at different times. RNA from FFPET archived for<br />

approximately 1 year were less fragmented than RNA archived for nearly 6 or<br />

17 years (8).<br />

There have been reports on the use <strong>of</strong> reverse transcription polymerase<br />

chain reaction (RT-PCR) to measure expression levels <strong>of</strong> few genes using RNA<br />

isolated from FFPET, but large-scale RT-PCR was not reported until recently.<br />

Cronin et al. from Genomic Health Inc. has invested on developing high quality<br />

controlled robotic assays for RT-PCR and published pro<strong>of</strong> <strong>of</strong> concept study<br />

(8). In the latter study, while RNAs isolated from old formalin fixed tissue were<br />

heavily fragmented with average size under 200 base pairs and absolute signal<br />

was much less for older samples, normalization <strong>of</strong> signal using a set <strong>of</strong> reference<br />

genes could largely correct these problems.<br />

Encouraged by these results, the National Surgical Adjuvant <strong>Breast</strong> and<br />

Bowel Project (NSABP) has engaged in a collaborative study with Genomic<br />

Health Inc. to develop prognostic gene expression assay for node negative (N )<br />

estrogen receptor–positive (ERþ) breast cancer treated with tamoxifen (9).<br />

While there are many other outstanding clinical questions that need to be<br />

addressed, we were compelled to first address the question <strong>of</strong> who among N<br />

ERþ tamoxifen treated patients need chemotherapy. In addition, there were two<br />

independent tamoxifen treated cohorts (tamoxifen arms <strong>of</strong> B-14 and B-20 trials)<br />

in the NSABP tissue bank so that discovery and validation could be performed<br />

without difficulty. We wanted to make sure that by the time <strong>of</strong> reporting, the<br />

assay was validated in a completely independent cohort. At that time, technology<br />

and cost limited Genomic Health investigators to examine only 250 genes<br />

maximum in RNA extracted from 60 mm thick section from a paraffin block.<br />

Therefore, existing cancer literature and microarray data were surveyed and<br />

250 candidate genes were selected. Before tumor blocks from B-20 tamoxifen<br />

treated cohort were examined, two other cohorts, one from mixed patient<br />

population and the other from those with more than 10 positive nodes were<br />

examined with the same assay (10). We did not believe that there will be different<br />

prognostic genes based on nodal status and therefore selected the prognostic<br />

genes that were equally prognostic across all three cohorts as a basis for<br />

prognostic model building. From the beginning, because <strong>of</strong> the continuous nature<br />

<strong>of</strong> RNA expression data, we believed strongly that a linear prediction model<br />

needed to be developed.


100 Paik<br />

Correlating clinical outcome with expression levels yielded many genes<br />

from these three studies, and 16 top-performing genes were identified for final<br />

model building and validation. Relative expression levels <strong>of</strong> the 16 genes are<br />

measured in relationship to average expression levels <strong>of</strong> 5 reference genes.<br />

While a majority <strong>of</strong> genes comprising 16 genes are ER (ER, PGR, BCL2,<br />

SCUBE2) and proliferation (Ki67, STK15, Survivin, CCNB1, MYBL2) related,<br />

there are other genes (HER2, GRB7, MMP11, CTSL2, GSTM1, CD68, BACG1)<br />

(9). The unscaled recurrence score (RSu) was calculated with the use <strong>of</strong> coefficient<br />

that are defined on the basis <strong>of</strong> regression analysis <strong>of</strong> gene expression.<br />

Recurrence score (RS) was rescaled from the unscaled recurrence score (Rsu) as<br />

follows: RS ¼ 0 if Rsu < 0; RS ¼ 20 (Rsu–6.7) if 0 Rsu 100; and RS ¼<br />

100 if Rsu > 100 (9). This resulted in RS ranging from 0 to 100.<br />

Final validation <strong>of</strong> the RS was achieved by examination <strong>of</strong> its performance<br />

in a completely independent cohort from NSABP trial B-14, which was not used<br />

in the model building process (9). The validation study was conducted with<br />

rigorous predefined statistical analysis plan with prespecified outcome end<br />

points and cut<strong>of</strong>fs for RS. The predefined low-risk group (RS below 18) had<br />

significant better prognosis compared with higher risk groups. More importantly,<br />

as shown in the published report, there is a linear relationship between<br />

RS and risk <strong>of</strong> distant failure at 10 years among N ERþ tamoxifen-treated<br />

patients (9).<br />

To test the applicability <strong>of</strong> RS in community oncology setting, Habel et al.<br />

conducted a nested case-control study in breast cancer patients diagnosed from<br />

1985 to 1994 at 14 Kaiser hospitals (11). Eligibility included negative nodes, age<br />

less than 75 years, and no chemotherapy. Cases died <strong>of</strong> breast cancer prior to<br />

2002. Up to three controls were matched to each case on age, race, tamoxifen<br />

treatment, facility, diagnosis year, and follow-up time. Among 4964 potentially<br />

eligible patients, 220 cases and 570 matched controls were identified. Median<br />

age was 59 years (range, 28–74 years); 30.9% were tamoxifen treated (mostly<br />

after 1988). Median follow-up time was 4.9 years for cases (time to death) and<br />

12.9 years for controls. RS was significantly associated with breast death in ERþ<br />

patients treated with or without tamoxifen (p ¼ 0.002). This study is important<br />

since it provides validation <strong>of</strong> RS assay used in a community setting.<br />

PREDICTION OF RESPONSE TO CHEMOTHERAPY<br />

Among gene expression–based markers, only RS has been tested whether it<br />

predicts response to chemotherapy in adjuvant setting (12). Since many <strong>of</strong> the<br />

genes comprising 21 genes in the RS assay are ER or proliferation related, we<br />

hypothesized early on that RS would be predictive <strong>of</strong> response to chemotherapy.<br />

Existence <strong>of</strong> chemotherapy arm in trial B-20 allowed us to ask this<br />

question (in fact this exactly was the reason why we first decided to work with<br />

N ERþ tamoxifen-treated cohort since we knew that we could not only discover<br />

and validate the prognostic index, but also test whether it predicts response


Predictors <strong>of</strong> Prognosis and Response to Chemotherapy 101<br />

to chemotherapy using banked materials readily available to us) (12). There were<br />

651 evaluable patients (227 randomized to tamoxifen and 434 randomized to<br />

tamoxifen plus chemotherapy). Patients with tumors who had high RS greater<br />

than 30 had a large absolute benefit <strong>of</strong> chemotherapy [with an absolute increase<br />

in distant recurrence free survival (DRFS) at 10 years <strong>of</strong> 27.6 8.0%, mean <br />

SE]. Patients with tumors who had low RS less than 18 derived minimal, if any,<br />

benefit from chemotherapy (with an absolute increase in DRFS at 10 years <strong>of</strong><br />

–1.1 2.2%, mean SE). The test for interaction between chemotherapy<br />

treatment and RS was statistically significant (p ¼ 0.038). Of particular<br />

importance is the finding that there is a linear relationship between RS and<br />

degree <strong>of</strong> benefit from chemotherapy when RS was examined as a continuous<br />

variable in Cox Model.<br />

While the relationship between RS and chemotherapy response has not<br />

been validated in an independent cohort as was done for prognostic aspect <strong>of</strong> the<br />

assay, supportive evidences exist to allow its use in clinical decision making.<br />

The most important data are from a cohort <strong>of</strong> patients treated with neoadjuvant<br />

chemotherapy at Milan reported by Gianni et al. (13). In this cohort, higher RS<br />

correlated with higher incidence <strong>of</strong> complete pathological response (pCR). In<br />

addition, due to the fact that many genes in RS are proliferation and ER related, we<br />

can make reasonable assumption that B-20 data are true. Since the decision to get<br />

chemotherapy has to be based on evaluation <strong>of</strong> baseline risk and assessment <strong>of</strong><br />

the degree <strong>of</strong> potential benefit from adding chemotherapy, it is reasonable to<br />

conclude that those with lower RS may not need chemotherapy since their<br />

baseline risk is low on the basis <strong>of</strong> B-14 data, and expected benefit is low on the<br />

basis <strong>of</strong> B-20 data. One headache is the fact that because <strong>of</strong> the continuous nature<br />

<strong>of</strong> the RS, it is difficult to pinpoint from which RS score patients start to gain<br />

benefit from chemotherapy. Clinical oncologists are not used to this kind <strong>of</strong> data.<br />

Uncertainties regarding the degree <strong>of</strong> benefit from chemotherapy in patients with<br />

intermediate range <strong>of</strong> RS resulted in launching a large randomized clinical trial<br />

named Trial Assigning IndividuaLized Options for Treatment (TAILORx) by<br />

North American <strong>Breast</strong> <strong>Cancer</strong> Intergroup (14). Many questioned about the<br />

reasons for the definition <strong>of</strong> the intermediate risk group in TAILORx trial being<br />

different from the definition used in the original publication in B-14 cohort. This<br />

was due to the fact that primary clinical end point in TAILORx is disease-free<br />

survival instead <strong>of</strong> distant disease-free survival and consideration <strong>of</strong> 95% confidence<br />

intervals in the baseline risk assessment. Again one has to realize that<br />

these cut<strong>of</strong>fs are totally arbitrary.<br />

PREDICTION OF RESPONSE TO NEOADJUVANT CHEMOTHERAPY<br />

Preoperative systemic therapy trials provide unique opportunities to assess<br />

sensitivity <strong>of</strong> tumor cells to systemic therapy in vivo while treatment is ongoing.<br />

Numerous studies have demonstrated the correlation <strong>of</strong> pCR and long-term<br />

outcome in breast cancer patients, potentially eliminating the need for long-term


102 Paik<br />

followup. Together with ease <strong>of</strong> obtaining fresh pretreatment core biopsies<br />

allowing high throughput assessment <strong>of</strong> molecular markers, preoperative trials<br />

may allow speedy discovery and validation <strong>of</strong> predictive markers for response to<br />

systemic therapy. So it is not surprising that many investigators attempted to<br />

utilize neoadjuvant setting to develop predictive markers <strong>of</strong> response to chemotherapy<br />

using microarray-based gene expression pr<strong>of</strong>iling. Regrettably the<br />

cumulative experience has so far been somewhat disappointing. No pr<strong>of</strong>ile has<br />

been found to have enough accuracy to be directly clinically applicable, although<br />

Hess et al. demonstrated that gene expression–based predictor still provided<br />

better prediction than clinical predictors (15). But more importantly, there has<br />

been no study with large enough sample size to provide conclusive answers as to<br />

whether one could reliably predict pCR using gene expression levels.<br />

It probably is somewhat naïve to expect gene expression pr<strong>of</strong>ile alone to<br />

predict pCR. For example, one would expect a smaller tumor would have higher<br />

chance <strong>of</strong> getting into pCR than a bigger tumor with same gene expression<br />

phenotype. Host factors such as drug metabolism and distribution may influence<br />

the chance for pCR. Prediction <strong>of</strong> combination regimens also may be more<br />

difficult than predicting response to single regimen. However, in the latter situation,<br />

because <strong>of</strong> low pCR rate, it is very difficult to develop predictor unless<br />

sample size is very high. In this regard, perhaps what investigators have achieved<br />

so far is as far as they can get, unless these other factors are integrated into<br />

algorithm development. In addition, pCR itself is not a perfect measure <strong>of</strong><br />

response either. As many as 80% <strong>of</strong> tumors show clinical response on taxanebased<br />

neoadjuvant chemotherapy, and therefore among those not achieving pCR,<br />

there is a continuous spectrum <strong>of</strong> residual burden in the breast. So it may make<br />

more sense to correlate gene expression with residual tumor burden than pCR,<br />

which has a somewhat arbitrary cut<strong>of</strong>f.<br />

Despite these limitations, certain consensus has emerged that are in<br />

agreement with what we found in B-20 study—that the tumors with poor<br />

prognostic signatures have higher chance <strong>of</strong> responding to chemotherapy. In<br />

addition to above described study by Gianni et al. (13), Rouzier et al. (using gene<br />

expression array) (16) and Carey et al. (using IHC surrogates <strong>of</strong> intrinsic subtypes)<br />

(17) has examined whether Perou’s intrinsic subtypes correlate with<br />

chance <strong>of</strong> pCR. As expected from B-20 data, basal and HER2 subtypes have<br />

higher chance <strong>of</strong> pCR compared to luminal subtypes. So there is a general<br />

agreement in the field that genetic program that determines the prognosis is<br />

tightly linked with general sensitivity to chemotherapy.<br />

ASSESSMENT OF RESIDUAL RISK AFTER CHEMOTHERAPY<br />

Despite the fact that gene expression data alone have not been shown to be able<br />

to predict pCR with high accuracy, gene expression pr<strong>of</strong>iling can be still<br />

important in patients treated with neoadjuvant chemotherapy. In an unpublished<br />

study, using pretreatment core biopsy samples from NSABP B-27, we found that


Predictors <strong>of</strong> Prognosis and Response to Chemotherapy 103<br />

the combination <strong>of</strong> prognostic pr<strong>of</strong>ile with pCR may be used to define residual<br />

clinical risk after chemotherapy, since those with pCR had a very good prognosis<br />

despite having a poor prognostic gene signature. This means that patients with<br />

poor prognostic gene signature who did not achieve pCR are at high risk <strong>of</strong><br />

recurrence after chemotherapy. Such data will allow identification <strong>of</strong> patients<br />

who will require more than chemotherapy without the need for waiting until they<br />

develop recurrences and opens door for a concept for post-neoadjuvant targeted<br />

therapy trials.<br />

CLINICAL UTILITY OF RS ASSAY<br />

How Is RS Compared to NCCN or St. Gallen Clinical Practice Guidelines?<br />

In B-14 study, less than 10% <strong>of</strong> the tamoxifen treated patients are classified as<br />

low risk by either NCCN or St. Gallen clincial guidelines. These patients did<br />

have excellent prognosis with less than 10% distant failure within 10 years. In<br />

contrast, RS classified 51% <strong>of</strong> the same cohort into low-risk category with very<br />

similar 10-year failure rate. Therefore 40% more ERþ tamoxifen-treated<br />

patients would be spared from chemotherapy if RS is used in comparison to<br />

NCCN or St. Gallen clinical guidelines. In addition, unlike NCCN and St. Gallen<br />

guidelines, which assumed that treatment benefit is equal in all subgroups,<br />

NSABP B-20 data showed that the benefit from chemotherapy is nearly zero in<br />

low-risk group defined by RS.<br />

How Is RS Compared to Adjuvant on Line Algorithm?<br />

The high cost <strong>of</strong> the RS assay (estimated at $3400) can only be justified by the<br />

additional information that it provides compared with other assays. Bryant and<br />

colleagues, in collaboration with Peter Ravdin, compared RS with Adjuvant on<br />

Line (AOL) in the NSABP B-14 cohort in a presentation made at the 2005<br />

St. Gallen <strong>Breast</strong> <strong>Cancer</strong> Symposium (personal communications with the<br />

late Dr. John Bryant). The AOL program provides individualized estimates <strong>of</strong><br />

recurrence risk on the basis <strong>of</strong> information such as patient age, tumor size, nodal<br />

involvement, and histological grade. It is actually not that straightforward to<br />

compare RS and AOL, since AOL does not provide risk scores. Therefore, it is<br />

difficult to utilize the numerical output from AOL when comparisons are made<br />

to RS. To compare the two, Dr. Bryant rank ordered output from AOL so that<br />

similar proportion <strong>of</strong> patients belong to low- (50%), intermediate- (25%), and<br />

high-risk (25%) categories as defined by RS. Rank-ordered AOL (R-AOL)<br />

performed remarkably well as a prognostic factor (in B-14) as well as predictive<br />

factor for benefit from chemotherapy (in B-20). RS and R-AOL were both<br />

independently prognostic in B-14 in multivariate analysis, suggesting that both<br />

gene expression signature and clinical features influence patient prognosis.<br />

However, agreement between the two assays was only 48%. RS was able to tease


104 Paik<br />

out patients with poor prognosis from the R-AOL low-risk category, so that the<br />

10-year distant relapse rate was 5.6% for patients categorized as low risk by both<br />

R-AOL and RS. R-AOL low-risk but RS intermediate- and high-risk patients had<br />

a 12.9% distant relapse rate at 10 years in the tamoxifen-treated cohort (p ¼<br />

0.04). For R-AOL intermediate- or high-risk patients, the distant relapse rate was<br />

8.9% when the RS risk was low, whereas the distant relapse rate was 30.7%<br />

when the RS risk was intermediate or high (p < 0.0001). These data suggest that<br />

the RS adds to existing prognostic markers and it would be ideal to incorporate<br />

RS and AOL into a single algorithm. However, at this point we don’t have a<br />

clear way to actually implement this finding into AOL in clinical use.<br />

Applicability in Node-Positive Patients<br />

One question about RS is whether it will work for Nþ ERþ patients. There is no<br />

reason why it should not work when it actually works even in 10þ node patients.<br />

However, risk estimates for RS will be different between N and Nþ patients<br />

although overall direction <strong>of</strong> the data would be the same, so one cannot simply<br />

apply the current RS algorithm to Nþ patients without a study.<br />

Categorical Vs. Continuous Variable<br />

One <strong>of</strong> the biggest problems in the field currently in my view is the use <strong>of</strong> RS as<br />

a categorical variable (low-, intermediate-, and high-risk categories), which was<br />

arbitrarily defined for validation study to ease the understanding <strong>of</strong> the data,<br />

because clinicians are not used to seeing a continuous data (9). It is very<br />

important to recognize the continuous nature <strong>of</strong> the RS. For example, a patient<br />

with score 17 (categorized as low risk) has nearly identical risk as a patient with<br />

score 19 (categorized as intermediate risk). On the other hand, a patient with RS<br />

<strong>of</strong> 3 would have much different prognosis than the one with RS <strong>of</strong> 17, although<br />

they are both categorized as low risk in the published report.<br />

FUTURE PERSPECTIVE<br />

The problem <strong>of</strong> current gene expression pr<strong>of</strong>iling methods is the need for fresh or<br />

snap frozen tissue and/or high cost. There are alternative methods that may be<br />

able to overcome these barriers. For example, Quantiplex assay (Panomics)<br />

based on combination <strong>of</strong> branched DNA amplification and fluorescence-labeled<br />

beads may allow multiplexing as many as 30 genes at a time in a 96 well plate<br />

with reasonable amount <strong>of</strong> input RNA extracted from FFPET. cDNA-mediated<br />

annealing, selection, extension, and ligation (DASL) assays from Illumina<br />

is another promising example (18). These assays cost fraction <strong>of</strong> the cost for<br />

RT-PCR or microarray-based methods and can utilize FFPET as starting materials.<br />

Replacement <strong>of</strong> standard markers such as ER and HER2 by these assays or<br />

QRT-PCR is desirable since they will provide more linear prediction <strong>of</strong> response.


Predictors <strong>of</strong> Prognosis and Response to Chemotherapy 105<br />

Availability <strong>of</strong> gene expression pr<strong>of</strong>iling methods for FFPET opens the<br />

possibility that archived materials from adjuvant chemotherapy trials can be<br />

analyzed, and therefore current approach <strong>of</strong> using gene expression analysis in<br />

neoadjuvant setting as a mean to identify predictive markers will become less<br />

important. The latter approach always will require validation in adjuvant setting<br />

since the FDA will not approve marker based on pCR as a surrogate—therefore<br />

it makes more sense to perform discovery and validation <strong>of</strong> predictive markers<br />

purely in the adjuvant setting.<br />

REFERENCES<br />

1. Schena M, Shalon D, Davis RW, et al. Quantitative monitoring <strong>of</strong> gene expression<br />

patterns with a complementary DNA microarray. Science 1995; 270:467–470.<br />

2. Fan C, Oh DS, Wessels L, et al. Concordance among gene-expression-based predictors<br />

for breast cancer. N Engl J Med 2006; 355:560–569.<br />

3. Sorlie T, Perou CM, Tibshirani R, et al. Gene expression patterns <strong>of</strong> breast carcinomas<br />

distinguish tumor subclasses with clinical implications. Proc Natl Acad Sci U S A<br />

2001; 98:10869–10874.<br />

4. van ’t Veer LJ, Dai H, van de Vijver MJ, et al. Gene expression pr<strong>of</strong>iling predicts<br />

clinical outcome <strong>of</strong> breast cancer. Nature 2002; 415:530–536.<br />

5. Perou CM, Sorlie T, Eisen MB, et al. Molecular portraits <strong>of</strong> human breast tumours.<br />

Nature 2000; 406:747–752.<br />

6. van de Vijver MJ, He YD, van ’t Veer LJ, et al. A gene-expression signature as a<br />

predictor <strong>of</strong> survival in breast cancer. N Engl J Med 2002; 347:1999–2009.<br />

7. Masuda N, Ohnishi T, Kawamoto S, et al. Analysis <strong>of</strong> chemical modification <strong>of</strong> RNA<br />

from formalin-fixed samples and optimization <strong>of</strong> molecular biology applications for<br />

such samples. Nucleic Acids Res 1999; 27:4436–4443.<br />

8. Cronin M, <strong>Ph</strong>o M, Dutta D, et al. Measurement <strong>of</strong> gene expression in archival<br />

paraffin-embedded tissues: development and performance <strong>of</strong> a 92-gene reverse<br />

transcriptase-polymerase chain reaction assay. Am J Pathol 2004; 164:35–42.<br />

9. Paik S, Shak S, Tang G, et al. A multigene assay to predict recurrence <strong>of</strong> tamoxifentreated,<br />

node-negative breast cancer. N Engl J Med 2004; 351:2817–2826.<br />

10. Cobleigh MA, Tabesh B, Bitterman P, et al. Tumor gene expression and prognosis in<br />

breast cancer patients with 10 or more positive lymph nodes. Clin <strong>Cancer</strong> Res 2005;<br />

11:8623–8631.<br />

11. Habel LA, Shak S, Jacobs MK, et al. A population-based study <strong>of</strong> tumor gene<br />

expression and risk <strong>of</strong> breast cancer death among lymph node-negative patients.<br />

<strong>Breast</strong> <strong>Cancer</strong> Res 2006; 8:R25.<br />

12. Paik S, Tang G, Shak S, et al. Gene expression and benefit <strong>of</strong> chemotherapy in women<br />

with node-negative, estrogen receptor-positive breast cancer. J Clin Oncol 2006;<br />

24:3726–3734.<br />

13. Gianni L, Zambetti M, Clark K, et al. Gene expression pr<strong>of</strong>iles in paraffin-embedded<br />

core biopsy tissue predict response to chemotherapy in women with locally advanced<br />

breast cancer. J Clin Oncol 2005; 23:7265–7277.<br />

14. Sparano JA. TAILORx: trial assigning individualized options for treatment (Rx).<br />

Clin <strong>Breast</strong> <strong>Cancer</strong> 2006; 7:347–350.


106 Paik<br />

15. Hess KR, Anderson K, Symmans WF, et al. <strong>Ph</strong>armacogenomic predictor <strong>of</strong> sensitivity<br />

to preoperative chemotherapy with paclitaxel and fluorouracil, doxorubicin, and cyclophosphamide<br />

in breast cancer. J Clin Oncol 2006; 24:4236–4244.<br />

16. Rouzier R, Perou CM, Symmans WF, et al. <strong>Breast</strong> cancer molecular subtypes respond<br />

differently to preoperative chemotherapy. Clin <strong>Cancer</strong> Res 2005; 11:5678–5685.<br />

17. Carey LA, Dees EC, Sawyer L, et al. The triple negative paradox: primary tumor<br />

chemosensitivity <strong>of</strong> breast cancer subtypes. Clin <strong>Cancer</strong> Res 2007; 13:2329–2334.<br />

18. Bibikova M, Talantov D, Chudin E, et al. Quantitative gene expression pr<strong>of</strong>iling in<br />

formalin-fixed, paraffin-embedded tissues using universal bead arrays. Am J Pathol<br />

2004; 165:1799–1807.


8<br />

Utilization <strong>of</strong> Genomic Signatures for<br />

Personalized Treatment <strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong><br />

P. Kelly Marcom, Carey K. Anders,<br />

Ge<strong>of</strong>frey S. Ginsburg, and Anil Potti<br />

Duke Institute for Genome Sciences & Policy, Department <strong>of</strong> Medicine,<br />

Duke University Medical Center, Durham, North Carolina, U.S.A.<br />

Joseph R. Nevins<br />

Duke Institute for Genome Sciences & Policy, Department <strong>of</strong> Molecular<br />

Genetics and Microbiology, Duke University Medical Center, Durham,<br />

North Carolina, U.S.A.<br />

INTRODUCTION<br />

The challenge <strong>of</strong> personalized cancer treatment lies in matching the most effective<br />

therapy to the characteristics <strong>of</strong> an individual patient’s tumor. The selection<br />

<strong>of</strong> therapy for both early- and late-stage breast cancer, as is the case for most<br />

cancer therapy, is still largely empiric and guided by large, population-based,<br />

randomized clinical trials. Although widely adopted, this approach is inadequate for<br />

the selection <strong>of</strong> individualized chemotherapy regimens. Estimates <strong>of</strong> benefit are<br />

extrapolations from the effects observed in these large trials and are not necessarily<br />

applicable to individual patients.<br />

The use <strong>of</strong> genomic data, in particular gene expression pr<strong>of</strong>iles derived from<br />

DNA microarray analysis, has the promise <strong>of</strong> providing the basis for truly individualized<br />

therapy. This approach has the potential to transform the treatment <strong>of</strong><br />

107


108 Marcom et al.<br />

cancer from a population-based approach to an individualized one directed toward<br />

unique patient characteristics. This includes the use <strong>of</strong> gene expression data to<br />

identify subtypes <strong>of</strong> cancer not previously recognized by traditional methods <strong>of</strong><br />

analysis. Pr<strong>of</strong>iling genomic patterns to identify new subclasses <strong>of</strong> tumors has<br />

previously been described in distinguishing differences between acute myeloid<br />

leukemia and acute lymphoblastic leukemia (1,2). Similar studies have employed<br />

gene expression data to identify distinct subtypes <strong>of</strong> breast cancer with prognostic<br />

significance (3).<br />

A landmark study in breast cancer published in 2002 described the use <strong>of</strong> a<br />

DNA microarray-based, 70-gene expression pr<strong>of</strong>ile as a prognostic factor in<br />

breast cancer (4). This genomic marker has demonstrated predictive value<br />

stratifying stage I and II breast cancer patients into two broad subgroups, relatively<br />

high-risk versus low-risk groups as defined by a 10-year risk <strong>of</strong> breast<br />

cancer recurrence. Various other studies have described the development <strong>of</strong><br />

expression pr<strong>of</strong>iles that have prognostic value in breast cancer (5–9).<br />

The Netherlands 70-gene predictor has provided the basis for an expanded<br />

European study that aims to measure its performance in a more diverse set <strong>of</strong><br />

patients (10). The parallel, prospective clinical trial MINDACT (Microarray In<br />

Node-negative Disease may Avoid ChemoTherapy) aims to measure the effectiveness<br />

<strong>of</strong> this genomic predictor in guiding adjuvant chemotherapy when<br />

compared to predictions based solely on the traditional clinicopathologic features.<br />

Likewise, a large prospective trial in the United States, TAILORx (Trial<br />

Assigning IndividuaLized Options for Treatment), is currently evaluating the<br />

Oncotype DX test (Genomic Health, Inc., Redwood City, California, U.S.),<br />

a polymerase chain reaction (PCR)-based assay <strong>of</strong> 21 genes correlated with<br />

10-year breast cancer recurrence risk (11), to guide the administration <strong>of</strong> hormonal<br />

therapy with or without chemotherapy among patients with estrogen<br />

receptor (ER)-positive, early-stage breast cancer.<br />

Of equal, if not greater, importance in achieving the goal <strong>of</strong> personalized<br />

treatment <strong>of</strong> breast cancer is the ability to predict response to specific chemotherapeutics,<br />

particularly the standard-<strong>of</strong>-care cytotoxic regimens incorporated<br />

into routine clinical practice. Various studies have described approaches to<br />

developing genomic predictors <strong>of</strong> response to chemotherapeutic agents promising<br />

a more effective strategy <strong>of</strong> assigning chemotherapeutic regimens to<br />

patients, as guided by unique characteristics <strong>of</strong> their breast tumors (12–14).<br />

The overall aim <strong>of</strong> this chapter is to highlight both opportunities presented<br />

by recent advances in the development <strong>of</strong> genomic-based approaches within the<br />

field <strong>of</strong> breast cancer and corresponding strategies for implementation <strong>of</strong> these<br />

techniques into clinical practice. Rather than relying on historic paradigms <strong>of</strong><br />

chemotherapeutic selection as guided by broad, population-based studies, the<br />

goal <strong>of</strong> this unique methodology is to improve patient outcome, and ultimately<br />

patient survival, through the optimal selection <strong>of</strong> patient-specific chemotherapeutic<br />

agents as guided by genomic-based strategies.


Genomic Signatures for Personalized Treatment <strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong> 109<br />

CYTOTOXIC THERAPIES FOR EARLY-STAGE BREAST CANCER<br />

Currently, the standard regimens for the adjuvant treatment <strong>of</strong> early-stage breast<br />

cancer combine anthracyclines, taxanes, and cyclophosphamide in limited permutations<br />

<strong>of</strong> combination, dose, and administration schedule. The addition <strong>of</strong> a<br />

taxane, either sequentially or in combination, to a core anthracycline/cyclophosphamide<br />

regimen in the setting <strong>of</strong> node-positive breast cancer has been<br />

shown to improve disease-free survival (DFS) by 4% to 7% in four separate<br />

randomized trials (15–18). However, it is unclear whether all breast cancer<br />

subtypes are equally affected by this therapeutic approach. Recent work examining<br />

a series <strong>of</strong> breast cancer adjuvant trials conducted over the past two decades<br />

further confirms the importance <strong>of</strong> defining breast cancer subtypes and their<br />

relationship to treatment benefit. This work demonstrated a lack <strong>of</strong> benefit for<br />

this broad treatment approach across a subset <strong>of</strong> ER-positive patients (19).<br />

Optimizing chemotherapy outcomes through the manipulation <strong>of</strong> dosing<br />

schedules remains controversial. In CALGB 9741 (<strong>Cancer</strong> and Leukemia Group<br />

B Intergroup 9741), decreasing the dosing interval from every three weeks to<br />

every two weeks [i.e., dose-dense (DD) scheduling] was shown to improve DFS<br />

and overall survival (OS) at three-year median follow-up [risk ratio (RR) ¼ 0.74,<br />

p ¼ 0.010] (20). However, an update at five-year median follow-up showed a<br />

less impressive overall improvement in DFS (RR ¼ 0.80, p ¼ 0.012) and<br />

perhaps no improvement in ER-positive disease (21). Moreover, a randomized<br />

<strong>Ph</strong>ase III trial from the Gruppo Oncologico Nord Ovest–Mammella Inter<br />

Gruppo (GONO–MIG) examining DD scheduling <strong>of</strong> a non-taxane-containing<br />

regimen did not show a benefit in DFS or OS (22). In an accompanying editorial<br />

to this publication, the ambiguity regarding the benefit <strong>of</strong> DD therapy<br />

was acknowledged and emphasis was placed on the need to define biologically<br />

relevant breast cancer subtypes so that individualized treatment strategies could<br />

be developed (23).<br />

A ROLE FOR PREOPERATIVE SYSTEMIC CHEMOTHERAPY<br />

TO EVALUATE GENOMIC STRATEGIES<br />

Preoperative, systemic chemotherapy (PST), also known as neoadjuvant therapy,<br />

is an effective means for assessing the sensitivity <strong>of</strong> an individual breast tumor<br />

to chemotherapy. Multiple trials <strong>of</strong> preoperative systemic chemotherapy have<br />

been performed. Initially, this approach was primarily used for locally advanced,<br />

inoperable, or inflammatory breast cancer. However, subsequent trials have<br />

evaluated PST in the setting <strong>of</strong> smaller, operable breast tumors both to increase<br />

breast conservation rates and to assess sensitivity to chemotherapy.<br />

Over the past two decades, PST for early-stage breast cancer has been<br />

shown to be a safe and effective method for assessing the chemosensitivity <strong>of</strong><br />

primary breast cancers. Sequencing systemic therapy first, with the consequent


110 Marcom et al.<br />

three- to six-month delay in surgery, has not been shown to adversely affect<br />

outcomes (24). Assessment <strong>of</strong> pathologic response to PST, both in the breast and<br />

regional lymph nodes, provides the most powerful prognostic information<br />

available. Patients who obtain a complete pathologic response in the breast and<br />

lymph nodes have a greater than 95% survival at seven years (24,25). Although<br />

PST provides a powerful tool for in vivo assessment <strong>of</strong> chemosensitivity,<br />

exploiting the information gained in the aforementioned trials to improve<br />

management <strong>of</strong> individual patients has been more difficult.<br />

Studies employing a PST approach have yielded conflicting results illustrating<br />

the heterogeneity <strong>of</strong> primary breast tumors and responsiveness to different<br />

chemotherapeutic agents. This observation begs the question on whether<br />

or not breast tumors commonly possess intrinsic ‘pan-chemotherapy’ responsiveness<br />

or resistance. For example, patients responsive to preoperative treatment<br />

with CVAP (Cytoxan, vincristine, Adriamycin, and prednisone) were<br />

either continued on the same regimen or crossed over to docetaxel (26).<br />

Responding patients who received docetaxel had a significantly higher pathologic<br />

complete response (pCR) rate than those who continued on the CVAP<br />

regimen, supporting differential breast tumor sensitivity to anthracyclines and<br />

taxanes. Additionally, the pCR rate in the NSABP (National Surgical Adjuvant<br />

<strong>Breast</strong> and Bowel Project) B-27 clinical trial was doubled by the addition <strong>of</strong><br />

preoperative docetaxel to preoperative doxorubicin/cyclophosphamide (AC)<br />

chemotherapy (26.1% vs. 12.9%, p < 0.0001) supporting this conclusion.<br />

However, the addition <strong>of</strong> preoperative docetaxel did not yield a statistically<br />

significant difference in overall DFS when added to preoperative AC alone<br />

[hazard ratio (HR) ¼ 0.9, p ¼ 0.22]. The question remains open on whether<br />

patients only partially responsive to AC, who achieved a pCR following<br />

administration <strong>of</strong> docetaxel, might have achieved a pCR with docetaxel alone,<br />

thus avoiding the risk <strong>of</strong> cardiotoxicity inherent to anthracyclines. Unfortunately,<br />

this study is somewhat flawed by the concurrent administration <strong>of</strong> tamoxifen<br />

with chemotherapy in patients with known ER-positive breast tumors (25). An<br />

additional study by the GEPARDUO (German Preoperative Adriamycin Docetaxel)<br />

study compared DD doxorubicin/docetaxel (AT) with AC followed by<br />

docetaxel, illustrating improved response among patients receiving cyclophosphamide.<br />

This study, however, was more widely interpreted as a comparison<br />

<strong>of</strong> sequential versus DD therapy (27).<br />

The largest trials investigating response to PST with AC come from the<br />

NSABP B-18 and the aforementioned B-27 clinical trials, where the pCR rates,<br />

allowing for residual in situ disease, were consistently approximately 13%<br />

(25,28). Data for single-agent paclitaxel are more limited; however, a study<br />

conducted at the M.D. Anderson <strong>Cancer</strong> Center reported pCR rates <strong>of</strong> 6% when<br />

excluding in situ diseases (29). A comprehensive review <strong>of</strong> these and other PST<br />

trials supports that the pCR rate in the breast with most doublet chemotherapies<br />

is approximately 10% to 15% (24). The National Comprehensive <strong>Cancer</strong>


Genomic Signatures for Personalized Treatment <strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong> 111<br />

Network (NCCN) guidelines support that standard adjuvant regimens can also be<br />

used as preoperative therapy (http://www.nccn.org).<br />

Although response to PST is a powerful tool for assessing tumor chemosensitivity,<br />

it is a post hoc evaluation and requires the exposure <strong>of</strong> individual<br />

patients to toxic and potentially ineffective therapies. Consequently, it has been<br />

difficult to exploit the knowledge gained about both tumor sensitivity and<br />

resistance to the patient’s benefit. In addition, to our knowledge, no studies have<br />

examined the influence <strong>of</strong> PST on the postoperative management <strong>of</strong> patients who<br />

have minimal or no response to PST.<br />

The relative importance <strong>of</strong> including either an anthracycline or taxane or<br />

both as part <strong>of</strong> the adjuvant treatment regimen <strong>of</strong> early stage breast cancer is an<br />

area <strong>of</strong> active research. These two classes <strong>of</strong> agents, in combination with<br />

cyclophosphamide, provide the foundation for the treatment <strong>of</strong> early-stage breast<br />

cancer, and several clinical trials provide insight into the current use <strong>of</strong> these<br />

agents. A recently published randomized phase III trial compared AC with<br />

docetaxel/cyclophosphamide (TC) (30). While this trial demonstrated improvement<br />

in DFS favoring TC (86% vs. 80%, p ¼ 0.015), the confidence interval (CI)<br />

approached 1.0 (CI ¼ 0.50–0.94) and there was no difference in OS despite<br />

median follow-up <strong>of</strong> 5.5 years. Interestingly, a previous trial comparing AT with<br />

AC showed no difference in DFS at four years (31). Although both <strong>of</strong> these trials<br />

enrolled a lower-risk population <strong>of</strong> patients, reconciling differences in outcome<br />

across these two studies is difficult. An ongoing trial in patients with fewer than<br />

four positive axillary nodes is comparing AC with single-agent paclitaxel as<br />

adjuvant therapy (CALGB 40101), in part on the basis <strong>of</strong> responses observed in<br />

the aforementioned PST trials. In summary, although many active chemotherapeutic<br />

agents suitable for the treatment <strong>of</strong> early-stage breast cancer exist,<br />

data available in 2007 do not provide the necessary guidance to select the most<br />

effective chemotherapeutic agent for each unique patient. This observation<br />

speaks to the importance <strong>of</strong> developing novel methods employing genomic<br />

analyses <strong>of</strong> tumor biology to individualize and optimize patient care.<br />

GENE EXPRESSION PROFILES TO PREDICT THERAPEUTIC RESPONSE<br />

Given the current status <strong>of</strong> neoadjuvant and adjuvant chemotherapy for breast<br />

cancer, an opportunity exists to use genomic information to improve the selection<br />

<strong>of</strong> patients who would derive the greatest benefit from treatment with cytotoxic<br />

chemotherapy. That is, prognostic tools that can identify patients clearly in need <strong>of</strong><br />

further treatment can help to refine the decisions based now on clinical information.<br />

Additionally, the development <strong>of</strong> predictive signatures specific for<br />

response to individual cytotoxic agents would provide an invaluable clinical tool<br />

to enable the better use <strong>of</strong> available standard-<strong>of</strong>-care agents. Historically, predictive<br />

biomarkers in the setting <strong>of</strong> breast cancer have played an important role in<br />

selecting patients for targeted therapies (i.e., trastuzumab and endocrine therapies


112 Marcom et al.<br />

including tamoxifen and aromatase inhibitors). However, validated biomarkers for<br />

commonly used cytotoxic agents are not yet clinically available. The importance<br />

<strong>of</strong> cytotoxic agents, even with the advent <strong>of</strong> the targeted molecular therapy era,<br />

cannot be discounted. To date, many targeted therapies demonstrate greatest<br />

efficacy when combined with chemotherapeutic agents. To address this relevant<br />

clinical need, a variety <strong>of</strong> studies have been performed with the aim <strong>of</strong> developing<br />

predictors <strong>of</strong> sensitivity for a spectrum <strong>of</strong> cytotoxic chemotherapies.<br />

Gene expression pr<strong>of</strong>iling <strong>of</strong> breast cancers in the context <strong>of</strong> preoperative<br />

chemotherapy has shown promise in providing predictive signatures <strong>of</strong> response<br />

to specific agents (12–14). The availability <strong>of</strong> clinical datasets provides an<br />

opportunity to move ahead with the prospective assessment <strong>of</strong> gene expression<br />

pr<strong>of</strong>iling as a means for guiding PST. This methodology can also simultaneously<br />

provide insight into breast cancer biology, as the investigation <strong>of</strong> pr<strong>of</strong>iles will<br />

shed light on the deregulation <strong>of</strong> molecular pathways.<br />

Several studies have correlated breast cancer gene expression pr<strong>of</strong>iles with<br />

response to preoperative chemotherapy. Chang et al. performed gene expression<br />

pr<strong>of</strong>iling on locally advanced breast cancers treated with docetaxel to identify<br />

genes predictive <strong>of</strong> response to this therapeutic modality (32). Similar work has<br />

been done with preoperative doxorubicin therapy and sequential taxane/anthracycline<br />

therapy (33,34). While some studies have suggested that single markers<br />

such as NF-kB may predict resistance to anthracycline therapy, most have found<br />

that a multigene classifier is needed. In comparing a signature for doxorubicin/<br />

cyclophosphamide sensitivity with a signature for docetaxel sensitivity, Cleator<br />

et al. concluded that unique signatures for sensitivity to different chemotherapy<br />

regimens are likely to exist (33).<br />

More recently, studies <strong>of</strong> PST have examined the pCR rates for clinically<br />

and molecularly defined breast cancer subtypes. These retrospective analyses <strong>of</strong><br />

PST trials demonstrate marked differences in pCR rates depending on breast<br />

cancer subtype (35,36). A recent publication by the M.D. Anderson <strong>Breast</strong> Group<br />

has retrospectively defined gene expression pr<strong>of</strong>iles that predict response to a<br />

sequential regimen <strong>of</strong> paclitaxel (Taxol) followed by 5-fluorouracil, doxorubicin<br />

(Adriamycin), and cyclophosphamide (T-FAC) (13).<br />

As an alternate strategy, drug sensitivity data from the NCI-60 panel <strong>of</strong><br />

cancer cell lines, coupled with baseline Affymetrix gene expression data unique<br />

to these cells, has been utilized to generate a series <strong>of</strong> gene expression signatures<br />

with the potential to predict sensitivity to various chemotherapeutic agents (37)<br />

(Fig. 1). To test the capacity <strong>of</strong> the in vitro docetaxel sensitivity predictor to<br />

accurately identify the patients who responded to docetaxel, we evaluated data<br />

from a previously published neoadjuvant study that linked gene expression data<br />

with clinical response to docetaxel (32). The in vitro–generated pr<strong>of</strong>ile correctly<br />

predicted docetaxel response in 22 <strong>of</strong> 24 patient samples, achieving an overall<br />

accuracy <strong>of</strong> 91.6%. Findings were further validated in an ovarian cancer dataset,<br />

where accuracy for predicting the response to salvage docetaxel therapy<br />

exceeded 85% (37).


Genomic Signatures for Personalized Treatment <strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong> 113<br />

Figure 1 Signatures that predict response to cytotoxic chemotherapies. Heatmaps <strong>of</strong><br />

expression pr<strong>of</strong>iles selected to predict sensitivity to the indicated chemotherapies. Each<br />

column in the figure represents individual samples. Each row represents an individual<br />

gene, ordered from top to bottom according to regression coefficients. Shown below is<br />

prediction <strong>of</strong> single agent therapy response in patient samples using cell line based<br />

expression signatures <strong>of</strong> chemosensitivity. Source: Adapted from Potti et al., Nature<br />

Medicine 12:1294 –1300, 2006.<br />

Given that the in vitro developed gene expression pr<strong>of</strong>iles predicted<br />

clinical response to docetaxel, the approach has been extended to test the ability<br />

<strong>of</strong> additional signatures to predict response to commonly used salvage therapies<br />

in an independent dataset <strong>of</strong> ovarian cancer patients previously treated with<br />

Adriamycin. As shown in Figure 1B, each <strong>of</strong> these predictors was capable <strong>of</strong><br />

accurately predicting the response to the drugs in patient samples, achieving an<br />

accuracy in excess <strong>of</strong> 81% overall. Importantly, in these examples, the overall<br />

clinical response rate varied from 25% to 45%. Using the expression signatures<br />

to predict patients likely to respond would essentially increase the “effective”<br />

response rate (the positive predictive value for chemosensitivity for any given<br />

drug) to greater than 85%, by selecting those patients likely to respond, based on<br />

the drug sensitivity predictor. These data suggest that the genomic-based predictors<br />

<strong>of</strong> chemotherapeutic response provide a unique opportunity to guide the<br />

selection <strong>of</strong> a cytotoxic agent most effective for an individual patient—the<br />

essence <strong>of</strong> personalized medicine.<br />

Importantly, it is also evident from further analyses that while there are<br />

overlaps in the predicted sensitivities to the chemotherapeutic agents among<br />

ovarian patients, there are also distinct groups <strong>of</strong> patients that are predicted to be<br />

sensitive to various single-agent salvage therapies (Fig. 2). Many therapeutic<br />

regimens make use <strong>of</strong> combinations <strong>of</strong> chemotherapeutic cytotoxic agents,<br />

raising a question on the extent that signatures <strong>of</strong> individual therapeutic response<br />

will also predict response to a combination <strong>of</strong> agents. To address this question,


114 Marcom et al.<br />

Figure 2 Patterns <strong>of</strong> predicted sensitivity to commonly used chemotherapies. A<br />

collection <strong>of</strong> breast cancer samples represented by microarray data was used to predict<br />

the likely sensitivity to the indicated agents based on the signatures at the top <strong>of</strong> the<br />

figure. Predictions are plotted as a heatmap representing probability <strong>of</strong> sensitivity.<br />

Adapted from Potti et al., Nature Medicine 12:1294 –1300, 2006.<br />

data from a breast neoadjuvant trial were utilized to evaluate the ability <strong>of</strong> singleagent<br />

signatures for T-FAC to predict patient response measured by pCR (34).<br />

Utilization <strong>of</strong> individual chemotherapy sensitivity signatures to assess patient<br />

response illustrated a significant distinction between the responders (n ¼ 13) and<br />

nonresponders (n ¼ 38), with the exception <strong>of</strong> 5-flourouracil. Importantly, the<br />

combined probability <strong>of</strong> sensitivity to all four agents in the T-FAC neoadjuvant<br />

regimen was calculated using the probability theorem. It was clear from this<br />

analysis that the prediction <strong>of</strong> response based on a combined probability <strong>of</strong><br />

sensitivity, built from individual chemosensitivity predictions, yielded a statistically<br />

significant (p < 0.0001, Mann–Whitney U test) distinction between the<br />

responders and nonresponders.<br />

A STRATEGY FOR USING CHEMOTHERAPY<br />

RESPONSE PREDICTORS IN PST<br />

Although promising, studies to date have included post hoc gene expression<br />

pr<strong>of</strong>iling <strong>of</strong> tumor samples. The ability to effectively select the most appropriate<br />

preoperative systemic cytotoxic agents for the treatment <strong>of</strong> early-stage breast<br />

cancer must ultimately be tested and validated in prospective clinical trials. This<br />

rationale approach will not only provide an opportunity to confirm biologically<br />

relevant subtypes, but also establish their predictive capacities regarding<br />

response to specific, conventional cytotoxic agents.


Genomic Signatures for Personalized Treatment <strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong> 115<br />

The development <strong>of</strong> genomic signatures predictive <strong>of</strong> response to commonly<br />

employed chemotherapeutics <strong>of</strong>fers an opportunity to significantly<br />

expand the identification <strong>of</strong> combinations <strong>of</strong> cytotoxic agents best suited for an<br />

individual patient. With this in mind, several prospective clinical studies have<br />

been proposed to test the capability <strong>of</strong> genomic predictors to most effectively<br />

match cytotoxic agents with individual patients. This concept is best evaluated in<br />

the context <strong>of</strong> neoadjuvant trials where hypotheses regarding the effectiveness <strong>of</strong><br />

genomic predictions can be accurately assessed through pathologic measurements<br />

<strong>of</strong> tumor response.<br />

We note that the availability <strong>of</strong> chemotherapy response predictors provides<br />

an opportunity to employ genomic predictors in present-day practice, where<br />

patients are commonly treated with one or more therapeutic regimens <strong>of</strong> equal<br />

efficacy (Fig. 3). The ability to select the most effective therapy, among a panel<br />

<strong>of</strong> standard-<strong>of</strong>-care agents, is a rational strategy because standard treatment<br />

procedures would not be greatly altered. The goal <strong>of</strong> these trials would be to<br />

more effectively utilize currently available standard agents. For example, a<br />

breast neoadjuvant study has been designed and is currently being implemented<br />

by employing this approach. Patients with early-stage breast cancer will be<br />

treated with one <strong>of</strong> two current standard-<strong>of</strong>-care therapies for which genomic<br />

predictors are available. This study design would compare the response rates<br />

for the randomly or doctor-selected A or B regimens versus assignment to<br />

therapy A or B via genomic predictor <strong>of</strong> response. The success <strong>of</strong> this study<br />

could then set the stage for larger studies utilizing genomic predictors <strong>of</strong><br />

chemotherapy sensitivity to more accurately predict response to cytotoxic<br />

therapies. The next phase <strong>of</strong> clinical studies would then focus on novel therapeutics<br />

(i.e., targeted therapies) for the individual patients unlikely to respond<br />

to either regimen.<br />

Figure 3 Schematic diagram <strong>of</strong> prospective clinical trial that would evaluate the<br />

capacity <strong>of</strong> predictors <strong>of</strong> chemotherapy response to improve response rates.


116 Marcom et al.<br />

OTHER GENOMIC OPPORTUNITIES TO GUIDE<br />

BREAST CANCER THERAPEUTICS<br />

The successful development and validation <strong>of</strong> predictors <strong>of</strong> response to standard<br />

chemotherapeutics will foster the more efficient use <strong>of</strong> these agents through the<br />

identification <strong>of</strong> patients with the greatest likelihood <strong>of</strong> response. Concurrently,<br />

these strategies will also identify a subgroup <strong>of</strong> patients resistant to all available<br />

standard-<strong>of</strong>-care chemotherapeutics, necessitating the implementation <strong>of</strong> targeted<br />

strategies. This is a critical point as it is <strong>of</strong> little benefit to patients to<br />

identify resistance if a plan for alternative therapeutic strategies does not exist.<br />

Currently, most patients resistant to standard chemotherapeutic agents are traditionally<br />

treated with second- or third-line chemotherapeutic agents or are<br />

enrolled on a clinical trial. The more effective alternative is to identify patients<br />

likely to be resistant prior to initiation <strong>of</strong> therapy so that opportunities for<br />

alternative treatments might be employed in the first-line setting.<br />

A viable option is to exploit the wealth <strong>of</strong> pathway-specific therapeutics<br />

that have been developed over the past two decades. To this goal, gene<br />

expression patterns reflecting deregulation <strong>of</strong> oncogenic pathways have been<br />

developed, as depicted in Figure 4 (38). These oncogenic pathway signatures<br />

have been shown to accurately predict the status <strong>of</strong> pathway deregulation in a<br />

series <strong>of</strong> both tumors derived from murine models and human tumors.<br />

These results demonstrate an ability to predict pathway pr<strong>of</strong>iles unique to<br />

individual human tumors, thus characterizing the genetic events associated with<br />

tumor development. Importantly, the assay <strong>of</strong> gene expression pr<strong>of</strong>iles provides<br />

a measure <strong>of</strong> the consequence <strong>of</strong> the oncogenic process, irrespective <strong>of</strong> how the<br />

pathway might have been altered. Thus, even if the known oncogene is not<br />

mutated, but rather another component <strong>of</strong> the pathway is altered, the gene<br />

expression pr<strong>of</strong>ile will detect this alteration.<br />

The analysis <strong>of</strong> oncogenic pathways through gene expression pr<strong>of</strong>iling<br />

<strong>of</strong>fers an opportunity to identify new therapeutic options for patients by providing<br />

a potential basis for guiding the use <strong>of</strong> pathway-specific drugs. To<br />

demonstrate the value <strong>of</strong> pathway prediction to guide drug selection, pathway<br />

deregulation was mapped among a series <strong>of</strong> breast cancer cell lines to screen<br />

potential therapeutic drugs. In parallel with pathway status mapping, the cell<br />

lines were treated with a variety <strong>of</strong> drugs known to target specific activities<br />

within given oncogenic pathways. A variety <strong>of</strong> available reagents known to<br />

target oncogenic pathways for which signatures have been developed [i.e., a farnesyl<br />

transferase inhibitor (FTS) and an Src inhibitor (SU6656)] were employed.<br />

In each case, a clear relationship was demonstrated between prediction <strong>of</strong> pathway<br />

deregulation and sensitivity to the respective targeted therapeutic. These preliminary,<br />

yet promising, results have been further validated in a larger set <strong>of</strong> 50 cancer<br />

cell lines, including lung, ovarian, and melanoma tumor types.<br />

Taken together, these data demonstrate a potential approach to identify<br />

therapeutic options for chemotherapy-resistant patients, as well as identify novel


Genomic Signatures for Personalized Treatment <strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong> 117<br />

Figure 4 Patterns <strong>of</strong> predicted cell signaling pathway activation. A collection <strong>of</strong><br />

breast cancer samples represented by microarray data was used to predict the likely state<br />

<strong>of</strong> activation <strong>of</strong> the indicated cell signaling pathways based on the signatures at the top <strong>of</strong><br />

the figure developed to represent pathway activation. Predictions are plotted as a heatmap<br />

representing probability <strong>of</strong> pathway activation. A Kaplan-Meier survival analysis is<br />

shown at the bottom reflecting the overall survival for patients in each cluster defined by<br />

pathway patterns. Source: Adapted from Bild et al., Nature 439:353–357, 2006.<br />

combinations for chemotherapy-sensitive patients. This unique approach represents<br />

a potential strategy to optimize effective treatments for cancer patients and<br />

warrants future, prospective, validations trials in both the adjuvant and metastatic<br />

settings.<br />

Pathway signatures have the potential to identify patients most sensitive<br />

to standard-<strong>of</strong>-care cytotoxic drugs, as well as those resistant to standard<br />

regimens. This approach creates an opportunity to combine chemotherapy<br />

sensitivity pr<strong>of</strong>iles with concurrent oncogenic pathway deregulation within an<br />

individual patient’s tumor. As depicted in Figure 5, predictions <strong>of</strong> sensitivity to


118 Marcom et al.<br />

Figure 5 Combining information regarding status <strong>of</strong> individual patient’s tumors.<br />

The figure depicts an analysis <strong>of</strong> a set <strong>of</strong> breast tumors for predicted sensitivity to various<br />

chemotherapies (top figure) and then the pattern <strong>of</strong> pathway activation in relation to one<br />

<strong>of</strong> these (docetaxel sensitivity).<br />

various cytotoxic agents can be combined with predictions <strong>of</strong> oncogenic<br />

pathway deregulation with the goal <strong>of</strong> defining the most optimal regimen. For<br />

example, a patient whose tumor is resistant to docetaxel may concurrently<br />

illustrate a high probability <strong>of</strong> Src pathway deregulation. Src inhibitors are<br />

currently being tested in clinical trials in a variety <strong>of</strong> solid tumors, including<br />

breast tumors. The selection <strong>of</strong> a therapy targeted against the Src pathway may<br />

prove more efficacious for this individual patient compared with the empiric<br />

selection <strong>of</strong> a cytotoxic agent. In addition, this strategy could be employed to<br />

identify novel combinations <strong>of</strong> therapies—either multiple cytotoxic agents or<br />

cytotoxic agents combined with targeted agents—as guided by a more rational<br />

and informed approach.<br />

IMPACT AND FUTURE<br />

The practice <strong>of</strong> oncology continually faces the challenge <strong>of</strong> matching the optimal<br />

therapeutic regimen with the most appropriate patient, balancing relative benefit<br />

with risk, to achieve the most favorable outcome. This challenge is <strong>of</strong>ten<br />

daunting with marginal success rates in many advanced disease contexts, likely<br />

reflecting the enormous complexity <strong>of</strong> the disease process coupled with an<br />

inability to properly guide the use <strong>of</strong> available therapeutics.<br />

The importance <strong>of</strong> selecting patients who will respond to a given therapeutic<br />

agent is perhaps best illustrated by the example <strong>of</strong> trastuzumab. In the<br />

absence <strong>of</strong> selection as guided by Her2 protein overexpression or gene amplification,<br />

the overall response rate among breast cancer patients is on the order <strong>of</strong>


Genomic Signatures for Personalized Treatment <strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong> 119<br />

10%. In contrast, for patients selected on the basis <strong>of</strong> Her2 overexpression or<br />

amplification, the overall response rate when combined with chemotherapy<br />

approached 50% to 60% (39–41). The use <strong>of</strong> trastuzumab in the treatment <strong>of</strong><br />

both early-stage and advanced breast cancer is guided by Her2 protein overexpression<br />

<strong>of</strong> gene amplification. Gene expression signatures predicting<br />

response to various cytotoxic chemotherapeutic agents provide an equally<br />

important opportunity to optimize response to various cytotoxic agents.<br />

Importantly, clinically-available, validated predictors <strong>of</strong> chemotherapy<br />

response could provide an invaluable tool in present day practice. Commonly,<br />

patients are treated with one or more therapeutic regimens that, on a population<br />

basis, have individually demonstrated equal efficacy. Validated chemotherapy<br />

predictive <strong>of</strong> chemotherapy sensitivity provides an opportunity to select the most<br />

effective therapy among a panel <strong>of</strong> standard-<strong>of</strong>-care agents individualized to a<br />

particular patient’s tumor biology. It is essential that this approach be incorporated<br />

into modern day clinical trials. Ultimately, this novel concept has the<br />

potential to augment future treatment strategies incorporating likelihood <strong>of</strong><br />

sensitivity to common cytotoxic agents, as well as various targeted therapeutic<br />

agents. Incorporating this approach into the routine management <strong>of</strong> patients with<br />

various malignancies has the potential to more rationally assign combination<br />

therapies, both cytotoxic and targeted agents, as rationally guided by each<br />

individual’s unique tumor biology.<br />

SUMMARY<br />

Treatment within the field <strong>of</strong> oncology, in particular that <strong>of</strong> breast cancer, is<br />

shifting away from a broad, population-based approach toward individualized<br />

care as driven by unique tumor biology. Validated signatures <strong>of</strong> chemotherapy<br />

sensitivity, coupled with an understanding <strong>of</strong> oncogenic pathway signaling<br />

within a given breast tumor, provides an opportunity to improve the utilization <strong>of</strong><br />

standard-<strong>of</strong>-care therapies. This approach also provides an opportunity to take<br />

the next step toward personalized cancer care—incorporation <strong>of</strong> targeted<br />

agents—as guided by advances in genomic technology.<br />

REFERENCES<br />

1. Alizadeh AA, Eisen MB, Davis RE, et al. Distinct types <strong>of</strong> diffuse large B-cell<br />

lymphoma identified by gene expression pr<strong>of</strong>iling. Nature 2000; 403:503–511.<br />

2. Golub TR, Slonim DK, Tamayo P, et al. Molecular classification <strong>of</strong> cancer: class discovery<br />

and class prediction by gene expression monitoring. Science 1999; 286:531–537.<br />

3. Perou CM, Jeffrey SS, van de Rijn M, et al. Distinctive gene expression patterns in<br />

human mammary epithelial cells and breast cancers 1999; 96:9212–9217.<br />

4. van ’t Veer L, Dai H, van de Vijver MJ, et al. Gene expression pr<strong>of</strong>iling predicts<br />

clinical outcome <strong>of</strong> breast cancer. Nature 2002; 415:530–536.


120 Marcom et al.<br />

5. Huang E, Cheng SH, Dressman H, et al. Gene expression predictors <strong>of</strong> breast cancer<br />

outcomes. Lancet 2003; 361:1590–1596.<br />

6. Pittman J, Huang E, Dressman H, et al. Integrated modeling <strong>of</strong> clinical and gene<br />

expression information for personalized prediction <strong>of</strong> disease outcomes. Proc Natl<br />

Acad Sci U S A 2004; 101:8431–8436.<br />

7. Ma XJ, Wang Z, Ryan PD, et al. A two-gene expression ratio predicts clinical outcome<br />

in breast cancer patients treated with tamoxifen. <strong>Cancer</strong> Cell 2004; 5:607–616.<br />

8. Chang HY, Sneddon JB, Alizadeh AA, et al. Gene expression signature <strong>of</strong> fibroblast<br />

serum response predicts human cancer progression: similarities between tumors and<br />

wounds. PLOS Biol 2004; 2:206–214.<br />

9. Perou CM, Sorlie T, Eisen MB, et al. Molecular portraits <strong>of</strong> human breast tumors.<br />

Nature 2000; 406:747–752.<br />

10. van de Vijver MJ, He YD, van’T Veer LJ, et al. A gene-expression signature as a<br />

predictor <strong>of</strong> survival in breast cancer. N Engl J Med 2002; 347:1999–2009.<br />

11. Paik S, Shak S, Tang G, et al. A multigene assay to predict recurrence <strong>of</strong> tamoxifentreated,<br />

node-negative breast cancer. N Engl J Med 2004; 351:2817–2826.<br />

12. Chang JC, Wooten EC, Tsimelzon A, et al. Gene expression pr<strong>of</strong>iling for the prediction<br />

<strong>of</strong> therapeutic response to docetaxel in patients with breast cancer. Lancet<br />

2003; 362:362–369.<br />

13. Hess KR, Anderson K, Symmans WF, et al. <strong>Ph</strong>armacogenomic predictor <strong>of</strong> sensitivity<br />

to preoperative chemotherapy with paclitaxel and fluorouracil, doxorubicin,<br />

and cyclophosphamide in breast cancer. J Clin Oncol 2006; 24:4236–4244.<br />

14. Modlich O, Prisack HB, Munnes M, et al. Predictors <strong>of</strong> primary breast cancers<br />

responsiveness to preoperative epirubicin/cyclophosphamide-based chemotherapy:<br />

translation <strong>of</strong> microarray data into clinically useful predictive signatures. J Transl<br />

Med 2005; 3:1–18.<br />

15. Mamounas EP, Bryant J, Lembersky B, et al. Paclitaxel after doxorubicin plus<br />

cyclophosphamide as adjuvant chemotherapy for node-positive breast cancer: results<br />

from NSABP B-28. J Clin Oncol 2005; 23:3686–3696.<br />

16. Henderson IC, Berry DA, Demetri GD, et al. Improved outcomes from adding<br />

sequential Paclitaxel but not from escalating Doxorubicin dose in an adjuvant chemotherapy<br />

regimen for patients with node-positive primary breast cancer. J Clin<br />

Oncol 2003; 21:976–983.<br />

17. Martin M, Pienkowski T, Mackey J, et al. Adjuvant Docetaxel for node-positive<br />

breast cancer 2005; 52:2302–2313.<br />

18. Roche H, Fumoleau P, Spielmann M, et al. Five years analysis <strong>of</strong> the PACS 01 trial:<br />

6 cycles <strong>of</strong> FEC100 vs 3 cycles <strong>of</strong> FEC100 followed by 3 cycles <strong>of</strong> docetaxel (D) for<br />

the adjuvant treatment <strong>of</strong> node positive breast cancer. <strong>Breast</strong> <strong>Cancer</strong> Res Treatment<br />

2004; 88:27 (abstr).<br />

19. Berry DA, Cirrincione C, Henderson IC, et al. Estrogen-receptor status and outcomes<br />

<strong>of</strong> modern chemotherapy for patients with node-positive breast cancer. JAMA 2006;<br />

295:1658–1667.<br />

20. Citron ML, Berry DA, Cirrincione C, et al. Randomized trial <strong>of</strong> dose-dense versus<br />

conventionally scheduled and sequential versus concurrent combination chemotherapy<br />

as postoperative adjuvant treatment <strong>of</strong> node-positive primary breast cancer:<br />

first report <strong>of</strong> Intergroup Trial C9741/<strong>Cancer</strong> and Leukemia Group B Trial 9741.<br />

J Clin Oncol 2003; 21:1431–1439.


Genomic Signatures for Personalized Treatment <strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong> 121<br />

21. Hudis C, Citron M, Berry D, et al. Five year follow-up <strong>of</strong> INT C9741: dose-dense<br />

chemotherapy is safe and effective. Presented at: The 28th Annual San Antonio<br />

<strong>Breast</strong> <strong>Cancer</strong> Symposium, December 8–11, 2005.<br />

22. Venturini M, Del Mastro L, Aitini E, et al. Dose-dense adjuvant chemotherapy in<br />

early breast cancer patients: results from a randomized trial. J Natl <strong>Cancer</strong> Inst 2005;<br />

97:1724–1733.<br />

23. Lin NU, Gelman R, Winer EP. Dose density in breast cancer: a simple message?<br />

J Natl <strong>Cancer</strong> Inst 2005; 97:1712–1714.<br />

24. Kaufmann M, Hortobagyi GN, Goldhirsch A, et al. Recommendations from an<br />

international expert panel on the use <strong>of</strong> neoadjuvant (primary) systemic treatment <strong>of</strong><br />

operable breast cancer: an update. J Clin Oncol 2006; 24:1940–1949.<br />

25. Bear HD, Anderson S, Smith RE, et al. Sequential preoperative or postoperative<br />

docetaxel added to preoperative doxorubicin plus cyclophosphamide for operable<br />

breast cancer: national Surgical Adjuvant <strong>Breast</strong> and Bowel Project Protocol B-27.<br />

J Clin Oncol 2006; JCO.2005.04:1665.<br />

26. Smith IC, Heys SD, Hutcheon AW, et al. Neoadjuvant chemotherapy in breast cancer:<br />

significantly enhanced response with docetaxel. J Clin Oncol 2002; 20:1456–1466.<br />

27. von Minckwitz G, Raab G, Caputo A, et al. Doxorubicin with cyclophosphamide<br />

followed by docetaxel every 21 days compared with doxorubicin and docetaxel every<br />

14 days as preoperative treatment in operable breast cancer: the GEPARDUO study<br />

<strong>of</strong> the German <strong>Breast</strong> Group. J Clin Oncol 2005; 23:2676–2685.<br />

28. Fisher B, Bryant J, Wolmark N, et al. Effect <strong>of</strong> preoperative chemotherapy on the<br />

outcome <strong>of</strong> women with operable breast cancer. J Clin Oncol 1998; 16:2672–2685.<br />

29. Buzdar AU, Singletary SE, Theriault RL, et al. Prospective evaluation <strong>of</strong> paclitaxel<br />

versus combination chemotherapy with fluorouracil, doxorubicin, and cyclophosphamide<br />

as neoadjuvant therapy in patients with operable breast cancer. J Clin<br />

Oncol 1999; 17:3412–3417.<br />

30. Jones SE, Savin MA, Holmes FA, et al. <strong>Ph</strong>ase III trial comparing doxorubicin plus<br />

cyclophosphamide with docetaxel plus cyclophosphamide as adjuvant therapy for<br />

operable breast cancer. J Clin Oncol 2006; 24:5381–5387.<br />

31. Goldstein L, O’Neill A, Sparano J, et al. E2197: <strong>Ph</strong>ase III AT (doxorubicin/docetaxel)<br />

vs. AC (doxorubicin/cyclophosphamide) in the adjuvant treatment <strong>of</strong> node positive and<br />

high risk node negative breast cancer. American Society <strong>of</strong> Clinical Oncology Annual<br />

Meeting, May 13–17, 2005.<br />

32. Chang JC, Wooten EC, Tsimelzon A, et al. Patterns <strong>of</strong> resistance and incomplete<br />

response to docetaxel by gene expression pr<strong>of</strong>iling in breast cancer patients. J Clin<br />

Oncol 2005; 23:1169–1177.<br />

33. Cleator S, Tsimelzon A, Ashworth A, et al. Gene expression patterns for doxorubicin<br />

(Adriamycin) and cyclophosphamide (Cytoxan) (AC) response and resistance. <strong>Breast</strong><br />

<strong>Cancer</strong> Res Treat 2006 95:229–233.<br />

34. Ayers M, Symmans WF, Stec J, et al. Gene expression pr<strong>of</strong>iles predict complete<br />

pathologic response to neoadjuvant paclitaxel and fluorouracil, doxorubicin, and<br />

cyclophosphamide chemotherapy in breast cancer. J Clin Oncol 2004; 22:2284–2293.<br />

35. Guarneri V, Frassoldati A, Giovannelli S, et al. Primary systemic therapy for<br />

operable breast cancer: a review <strong>of</strong> clinical trials and perspectives. <strong>Cancer</strong> Lett 2007;<br />

248:175–185.<br />

36. Rouzier R, Perou CM, Symmans WF, et al. <strong>Breast</strong> cancer molecular subtypes respond<br />

differently to preoperative chemotherapy. Clin <strong>Cancer</strong> Res 2005; 11(16):5678–5685.


122 Marcom et al.<br />

37. Potti A, Lancaster JM, Dressman HK, et al. A genomic strategy to guide the use <strong>of</strong><br />

chemotherapeutic drugs in solid tumors. Nat Med 2006; 12:1294–1300.<br />

38. Bild A, Yao G, Chang JT, et al. Oncogenic pathway signatures in human cancers as a<br />

guide to targeted therapies. Nature 2006; 439:353–357.<br />

39. Chan A, Martin M, Untch M, et al. Vinorelbine plus trastuzumab combination as<br />

first-line therapy for HER 2-positive metastatic breast cancer patients: an international<br />

phase II trial. Br J <strong>Cancer</strong> 2006; 95:788–793.<br />

40. Marty M, Cognetti F, Maraninchi D, et al. Randomized phase II trial <strong>of</strong> the efficacy<br />

and safety <strong>of</strong> trastuzumab combined with docetaxel in patients with human epidermal<br />

growth factor receptor 2-positive metastatic breast cancer administered as firstline<br />

treatment: the M77001 study group. J Clin Oncol 2005; 23:4265–4274.<br />

41. Robert N, Leyland-Jones B, Asmar L, et al. Randomized phase III study <strong>of</strong> trastuzumab,<br />

paclitaxel, and carboplatin compared with trastuzumab and paclitaxel in women with<br />

HER-2-overexpressing metastatic breast cancer. J Clin Oncol 2006; 24:2786–2792.


9<br />

Personalized Medicine by the Use <strong>of</strong><br />

Microarray Gene Expression Pr<strong>of</strong>iling<br />

Stella Mook and Laura J. van ’t Veer<br />

Department <strong>of</strong> Pathology, The Netherlands <strong>Cancer</strong> Institute,<br />

Amsterdam, The Netherlands<br />

Fatima Cardoso<br />

Department <strong>of</strong> Medical Oncology, Institute Jules Bordet, Brussels, Belgium<br />

INTRODUCTION<br />

For accurate management <strong>of</strong> early breast cancer, several important decisions<br />

need to be made by patients and clinicians regarding the type <strong>of</strong> surgery, the<br />

need for adjuvant radiotherapy, and particularly the type <strong>of</strong> adjuvant systemic<br />

therapy (chemotherapy and/or endocrine therapy). The decision about adjuvant<br />

systemic therapy is <strong>of</strong> paramount importance since breast cancer is mainly a<br />

systemic disease, with a majority <strong>of</strong> deaths occurring as a consequence <strong>of</strong> distant<br />

metastases. This decision depends on the likelihood <strong>of</strong> recurrence; patientrelated<br />

factors such as biological age, menopausal status, performance status,<br />

and co-morbidities; and tumor-related factors such as size, grade, lymph node<br />

status, and hormonal and human epidermal growth factor receptor 2 (HER2)<br />

status. Most <strong>of</strong> these variables are combined into prognostic models or guidelines<br />

such as the St. Gallen consensus or the National Comprehensive <strong>Cancer</strong> Network<br />

(NCCN) guidelines, in order to assist patients and clinicians in the decisionmaking<br />

process (1–3). Although these guidelines provide valuable information<br />

about prognosis in certain breast cancer populations, the prediction <strong>of</strong> disease<br />

123


124 Mook et al.<br />

outcome for the individual patient is far from accurate. In reality, patients with<br />

similar clinicopathological characteristics can have strikingly different outcomes.<br />

Because <strong>of</strong> these limitations, and because <strong>of</strong> the recognition <strong>of</strong> the incurable<br />

nature <strong>of</strong> metastatic breast cancer, clinicians consider prescribing adjuvant<br />

chemotherapy in the majority <strong>of</strong> early-stage breast cancer patients to reduce the<br />

risk <strong>of</strong> relapse. However, a large proportion <strong>of</strong> these patients is probably cured<br />

with locoregional treatment alone or in combination with adjuvant endocrine<br />

therapy, thus unnecessarily exposing many patients to chemotherapy side effects.<br />

The identification <strong>of</strong> robust and reliable prognostic markers that accurately select<br />

patients not requiring adjuvant chemotherapy is essential to decrease the problem<br />

<strong>of</strong> overtreatment. Additionally, a more accurate patient selection will also contribute<br />

to the decrease in undertreatment, a less common but dangerous problem.<br />

The introduction <strong>of</strong> high-throughput techniques, such as microarray-based<br />

gene expression pr<strong>of</strong>iling and sequencing <strong>of</strong> the human genome, has led to a novel<br />

approach in breast cancer research. Measuring the expression level <strong>of</strong> thousands <strong>of</strong><br />

genes in one experiment using microarray technology enables us to better<br />

understand the biology and heterogeneity <strong>of</strong> breast cancer. The first study to<br />

examine gene expression patterns in breast cancer showed the existence <strong>of</strong> at least<br />

four molecular subtypes <strong>of</strong> breast cancer—luminal A, luminal B, basal, and HER2<br />

positive—distinguished by extensive differences in gene expression (4). Several<br />

subsequent studies confirmed these findings (5–8). In addition, microarray techniques<br />

have been used to identify pr<strong>of</strong>iles that correlate with disease outcome<br />

(prognostic pr<strong>of</strong>ile) or with response to treatment (predictive pr<strong>of</strong>ile) (9–16).<br />

One <strong>of</strong> the first prognostic pr<strong>of</strong>iles was the 70-gene pr<strong>of</strong>ile (Mamma-<br />

Print TM ) identified by van ’t Veer et al. (13). This dichotomous microarray<br />

classifier can accurately distinguish breast cancer patients who have a high<br />

likelihood <strong>of</strong> remaining free <strong>of</strong> distant metastases (good pr<strong>of</strong>ile) from those who<br />

are at high risk <strong>of</strong> developing distant metastases (poor pr<strong>of</strong>ile) (Fig. 1).<br />

DEVELOPMENT AND FIRST RETROSPECTIVE VALIDATION<br />

OF THE 70-GENE PROFILE (MAMMAPRINT TM )<br />

The 70-gene prognostic pr<strong>of</strong>ile was developed in a training set <strong>of</strong> 78 lymph<br />

node–negative, invasive breast tumors less than 5 cm in diameter. Patients were<br />

younger than 55 years at diagnosis, had no prior malignancies, and were treated<br />

in the Netherlands <strong>Cancer</strong> Institute (NKI) with breast conserving therapy or<br />

mastectomy. Five out <strong>of</strong> 78 patients received adjuvant systemic treatment<br />

comprising chemotherapy (n ¼ 3) or endocrine therapy (n ¼ 2). Thirty-four<br />

patients developed distant metastases within five years <strong>of</strong> diagnosis, while the<br />

remaining 44 patients remained free <strong>of</strong> distant metastases. The mean follow-up<br />

<strong>of</strong> patients who remained free <strong>of</strong> distant metastases was 8.7 years, and the mean<br />

time to metastases was 2.5 years.<br />

For all 78 samples, the percentage <strong>of</strong> tumor cells in a hematoxylin- and<br />

eosin-stained section, before and after cutting sections for RNA isolation, was at


Microarray Gene Expression Pr<strong>of</strong>iling 125<br />

Figure 1 Steps in 70-gene microarray analysis. Source: From Ref. 32. copyright # 2002<br />

Massachusetts Medical Society. All rights reserved. Adopted with permission.<br />

least 50%. RNA was isolated and labeled with a fluorescent dye. Reference RNA<br />

consisted <strong>of</strong> an equal amount <strong>of</strong> RNA from all tumors. Both sample RNA and<br />

reference RNA were hybridized on an oligonucleotide microarray platform<br />

containing approximately 25,000 genes (produced by Agilent Technologies, Palo<br />

Alto, California, U.S.) (17). After hybridization, microarray slides were washed<br />

and scanned using a confocal laser scanner (Agilent). Fluorescence intensities<br />

were quantified, corrected for background noise, and normalized (13).<br />

Using a statistical method called “supervised classification,” 231 genes<br />

that were differentially expressed in tumors that metastasized compared with<br />

tumors that did not were selected. These 231 genes were then ranked on the basis<br />

<strong>of</strong> their correlation coefficient with disease outcome. Using a leave-one-out<br />

cross-validation procedure, the top 70 <strong>of</strong> these genes appeared to be the optimal<br />

set <strong>of</strong> genes to accurately discriminate between tumors with a good disease<br />

outcome (free from distant metastases for at least five years) and tumors with a<br />

poor disease outcome (distant metastases within five years <strong>of</strong> diagnosis).<br />

All 78 tumors were ranked according to their correlation with the average<br />

expression <strong>of</strong> the 70 genes <strong>of</strong> the patients who did not develop distant metastases.<br />

The sensitivity was optimized by setting a threshold, which resulted in the


126 Mook et al.<br />

misclassification <strong>of</strong> 3 out <strong>of</strong> 34 patients with a poor disease outcome being<br />

classified as good prognosis and therefore would erroneously have had chemotherapy<br />

withheld (9% misclassification). Subsequently, the 70-gene pr<strong>of</strong>ile was<br />

validated in 19 breast cancer tumors—7 tumors from patients with a good<br />

clinical outcome (no distant metastases within five years <strong>of</strong> diagnosis) and<br />

12 tumors from patients who did develop distant metastases within five years <strong>of</strong><br />

diagnosis. Using the 70-gene pr<strong>of</strong>ile, only 2 out <strong>of</strong> 19 tumors were classified<br />

incorrectly, thereby confirming the initial results (13,18).<br />

The same group <strong>of</strong> investigators performed a first retrospective validation<br />

study using a consecutive series <strong>of</strong> 295 breast tumors from NKI (144 lymph<br />

node–positive and 151 lymph node–negative tumors). All patients were younger<br />

than 53 years at diagnosis and were treated with locoregional therapy alone<br />

(56%) or in combination with adjuvant systemic therapy comprising chemotherapy<br />

alone (31%), hormonal therapy alone (7%), or a combination (7%). In<br />

this consecutive series <strong>of</strong> 295 patients, 61 were also part <strong>of</strong> the previous series<br />

used to develop the 70-gene pr<strong>of</strong>ile (14).<br />

All 295 tumors were classified as good or poor pr<strong>of</strong>ile by calculating the<br />

correlation coefficient with the established average expression level <strong>of</strong> the<br />

70 genes in the previously studied good-outcome group. For the 234 samples that<br />

were not included in the previous study, tumors with a correlation coefficient<br />

above the threshold <strong>of</strong> 0.4 were classified as good pr<strong>of</strong>ile; for the 61 tumors<br />

that were also part <strong>of</strong> the previous study, the threshold was adapted to 0.55 to<br />

correct for overfitting (18).<br />

The 70-gene pr<strong>of</strong>ile accurately classified 115 tumors (39%) as a goodprognosis<br />

group (good pr<strong>of</strong>ile) with a 10-year overall survival <strong>of</strong> 95% (2.6%)<br />

and 180 tumors (61%) as a poor-prognosis group (poor pr<strong>of</strong>ile) with a 10-year<br />

overall survival <strong>of</strong> 55% (4.4%). Interestingly, the 70-gene pr<strong>of</strong>ile was not only<br />

associated with disease outcome in the lymph node–negative patients [hazard<br />

ratio (HR) for distant metastases 5.5; 95% confidence interval (CI) 2.5–12.2; p <<br />

0.001], but also strongly associated with disease outcome in the group <strong>of</strong> 144<br />

lymph node–positive patients (HR for distant metastases 4.5; 95% CI 2.0–10.2;<br />

p < 0.001). Furthermore, the only significant independent factors for prediction<br />

<strong>of</strong> distant metastases were a poor-prognosis pr<strong>of</strong>ile, a larger tumor, the presence<br />

<strong>of</strong> vascular invasion, and no chemotherapy treatment. The 70-gene pr<strong>of</strong>ile was<br />

by far the most powerful predictor <strong>of</strong> distant metastases, with an adjusted HR <strong>of</strong><br />

4.6 (95% CI 2.3–9.2; p < 0.001) (14,18).<br />

THE PROGNOSTIC VALUE OF THE 70-GENE<br />

PROFILE IN A CLINICAL CONTEXT<br />

To assess the potential added value <strong>of</strong> this prognostic tool in a clinical context,<br />

the 70-gene pr<strong>of</strong>ile was compared with the St. Gallen consensus guidelines or<br />

the NIH criteria used at that time (3,19). In comparison with these guidelines, the<br />

70-gene pr<strong>of</strong>ile assigned more patients to the good-prognosis group than the<br />

St. Gallen consensus guidelines or the NIH criteria (40% vs. 15% and 7%,


Microarray Gene Expression Pr<strong>of</strong>iling 127<br />

respectively). Moreover, patients identified by the 70-gene pr<strong>of</strong>ile as having<br />

good prognosis were more likely to remain free <strong>of</strong> distant metastases than<br />

patients identified as having good prognosis according to the St. Gallen consensus<br />

guidelines or the NIH criteria. On the other hand, patients who were<br />

assigned to the poor-prognosis group by the 70-gene pr<strong>of</strong>ile had a higher risk <strong>of</strong><br />

developing distant metastases as compared with the poor-prognosis group<br />

classified by the St. Gallen consensus guidelines or the NIH criteria.<br />

The 70-gene pr<strong>of</strong>ile was built to have a minimum <strong>of</strong> misclassified patients<br />

with a poor disease outcome; in other words, the low-risk group should be<br />

maximized without causing considerable undertreatment. Consequently, in this<br />

retrospective validation series, the 70-gene pr<strong>of</strong>ile had a high negative predictive<br />

value <strong>of</strong> 96% for the 234 new patients and a positive predictive value <strong>of</strong> only<br />

38%. That is to say, <strong>of</strong> all patients with a good prognosis pr<strong>of</strong>ile, 4% will still<br />

develop distant metastases and would erroneously have had chemotherapy<br />

withheld, whereas <strong>of</strong> all patients classified as having a poor prognosis, 38% will<br />

indeed develop distant metastases. Consequently, treatment decisions based on<br />

the 70-gene pr<strong>of</strong>ile would still lead to overtreatment; but since the total proportion<br />

<strong>of</strong> patients identified as having a poor pr<strong>of</strong>ile is much smaller than the<br />

proportion <strong>of</strong> high-risk patients according to the St. Gallen consensus guidelines<br />

or the NIH criteria, overtreatment would be reduced by 25% to 30%. Furthermore,<br />

the overall prediction <strong>of</strong> disease outcome and consequent treatment<br />

decision according to the 70-gene pr<strong>of</strong>ile seems to be more accurate.<br />

INDEPENDENT RETROSPECTIVE VALIDATION<br />

BY THE TRANSBIG CONSORTIUM<br />

To further substantiate the prognostic value <strong>of</strong> the 70-gene pr<strong>of</strong>ile, the European<br />

Union’s sixth Framework Project TRANSBIG (http://www.breastinternationalgroup.org/TransBIG.aspx)<br />

conducted an independent retrospective validation<br />

study (20). Frozen tumor material and clinical data were collected from<br />

302 patients from five different cancer centers in the United Kingdom, Sweden,<br />

and France. Patients were younger than 61 years at diagnosis, had a lymph node–<br />

negative T1 or T2 tumor, and had not received adjuvant systemic therapy. The<br />

patients were treated before 1999, and the median follow-up in this series was<br />

13.6 years. The frozen tumor samples were sent to Agendia, the microarray facility<br />

in Amsterdam, where RNA was isolated and hybridized on a custom-designed<br />

microarray, measuring the gene expression level <strong>of</strong> the 70 genes in triplicate.<br />

Tumors were classified as poor pr<strong>of</strong>ile if the correlation coefficient <strong>of</strong> the average<br />

expression was under 0.4. The microarray facility had no access to the clinical<br />

data, and the researchers collecting the clinical data were blinded for the 70-gene<br />

pr<strong>of</strong>ile results. Only an independent statistical <strong>of</strong>fice had access to both the clinical<br />

and the microarray data simultaneously. Eighty percent <strong>of</strong> the samples were<br />

centrally reviewed for estrogen receptor (ER) status and tumor grade, and the<br />

centers were independently audited for quality <strong>of</strong> data control.


128 Mook et al.<br />

This independent study confirmed that the 70-gene pr<strong>of</strong>ile can accurately<br />

select patients who are at high risk <strong>of</strong> developing distant metastases (poor prognosis)<br />

from patients who are at low risk <strong>of</strong> developing distant metastases and<br />

hence can be safely spared chemotherapy (good pr<strong>of</strong>ile), with HRs <strong>of</strong> 2.79 (95%<br />

CI 1.60–4.87) and 2.32 (95% CI 1.35–4.0) for distant metastases–free survival and<br />

overall survival, respectively. Moreover, patients with a good pr<strong>of</strong>ile had a 10-year<br />

overall survival <strong>of</strong> 89% compared with 69% in patients with a poor pr<strong>of</strong>ile. The<br />

70-gene pr<strong>of</strong>ile outperformed several prognostic methods such as the Nottingham<br />

Prognostic Index and the St. Gallen consensus guidelines. The performance <strong>of</strong> the<br />

70-gene pr<strong>of</strong>ile was also compared with Adjuvant! Online. The Adjuvant! S<strong>of</strong>tware—available<br />

at www.adjuvantonline.com—calculated a 10-year overall survival<br />

probability based on clinicopathological criteria (the patient’s age, tumor<br />

size, tumor grade, ER status, and comorbidity) (21,22). Patients were considered<br />

as having low clinical risk when the 10-year breast cancer–specific survival<br />

probability calculated by Adjuvant! was more than 88% if their tumors were ER<br />

positive and more than 92% if they were ER negative. The rationale for these two<br />

cut<strong>of</strong>fs was the assumption that patients with ER-positive tumors would now<br />

receive hormonal therapy, with an average absolute benefit <strong>of</strong> 4%. In the discordant<br />

cases, the 70-gene pr<strong>of</strong>ile provided much stronger prognostic information,<br />

with almost identical survival rates for the low and high clinical risk groups within<br />

each 70-gene pr<strong>of</strong>ile risk, suggesting that the 70-gene pr<strong>of</strong>ile predicts disease<br />

outcome independently <strong>of</strong> clinical and pathological characteristics (Fig. 2) (20).<br />

Figure 2 Kaplan–Meier curve by clinical and 70-gene expression signature risk groups.<br />

Overall survival. Source: From Ref. 20. By permission <strong>of</strong> Oxford University Press.


Microarray Gene Expression Pr<strong>of</strong>iling 129<br />

Another important finding <strong>of</strong> this validation study is the time dependency<br />

<strong>of</strong> the 70-gene pr<strong>of</strong>ile. The median follow-up in this series was twice as long as<br />

the follow-up <strong>of</strong> the initial series. To determine the influence <strong>of</strong> the time <strong>of</strong><br />

distant metastases on the prognostic value <strong>of</strong> the pr<strong>of</strong>ile, HRs were calculated<br />

with arbitrary censoring <strong>of</strong> all observations at different time points. The adjusted<br />

HR for distant metastases at 5 years was 4.7 in comparison with an adjusted HR<br />

<strong>of</strong> 3.5 at 10 years, suggesting a better prediction <strong>of</strong> early (within five years <strong>of</strong><br />

diagnosis) metastases by the 70-gene pr<strong>of</strong>ile (18,23,24).<br />

OTHER PROGNOSTIC PROFILES IN BREAST CANCER<br />

Several other prognostic gene expression signatures were developed, such as the<br />

76-gene pr<strong>of</strong>ile, the wound signature, and the genomic grade index (15,25,26).<br />

The 76-gene prognostic pr<strong>of</strong>ile was developed on a different microarray platform<br />

(Affymetrix, Santa Clara, California, U.S.) and using a slightly different methodology;<br />

the genes associated with relapse <strong>of</strong> the disease were selected separately<br />

for ER-positive and ER-negative patients. The genes selected in each<br />

group were then combined in a 76-gene pr<strong>of</strong>ile (15). A retrospective validation<br />

study confirmed the prognostic value <strong>of</strong> this pr<strong>of</strong>ile in early-stage breast cancer<br />

patients (27). Recently, the 76-gene pr<strong>of</strong>ile has also been validated independently<br />

by the TRANSBIG consortium in the same patient series, using the same<br />

methodology as the validation <strong>of</strong> the 70-gene pr<strong>of</strong>ile (24). Although only<br />

three genes were common to both pr<strong>of</strong>iles, the biological pathways represented<br />

are the same, and the prognostic value <strong>of</strong> both the 70-gene pr<strong>of</strong>ile and the<br />

76-gene pr<strong>of</strong>ile was validated in this patient series (20,24). Since the pr<strong>of</strong>iles<br />

perform equally, and the 70-gene pr<strong>of</strong>ile has proven to be reproducible and<br />

robust and is available as a Food and Drug Administration (FDA)-cleared<br />

diagnostic test (28), the TRANSBIG consortium decided to use the 70-gene<br />

pr<strong>of</strong>ile in a large prospective clinical trial, the Microarray In Node-negative<br />

Disease may Avoid ChemoTherapy (MINDACT) trial.<br />

IMPLEMENTATION OF THE 70-GENE PROFILE:<br />

THE RASTER TRIAL AND THE LOGISTICS PILOT STUDY<br />

In addition to scientific evidence provided by validation studies, implementation<br />

<strong>of</strong> these microarray pr<strong>of</strong>iles in both prospective clinical trials and in daily practice<br />

requires logistical feasibility. The requirement <strong>of</strong> fresh tissue as the source <strong>of</strong> highquality<br />

RNA has several consequences for tumor sample handling. Traditional<br />

fixation <strong>of</strong> fresh tissue in formaldehyde, for example, causes degradation <strong>of</strong> RNA,<br />

thereby making tumor samples unsuitable for microarray experiments. In addition,<br />

to avoid interlaboratory variability, fresh tumor samples have to be shipped to one<br />

microarray facility. Consequently, for the implementation <strong>of</strong> microarray pr<strong>of</strong>iles,<br />

local procedures have to be adapted, and close collaboration between pathologists,<br />

surgeons, and medical oncologists is <strong>of</strong> outmost importance.


130 Mook et al.<br />

Recently, two multicenter feasibility studies were conducted, the RASTER<br />

study and a logistics pilot study (18,29). The first study was coordinated by NKI<br />

with financial support from the Dutch Health Care Insurance Board. The aim <strong>of</strong><br />

the study was to assess the feasibility <strong>of</strong> collecting good-quality tissue from<br />

several community hospitals in the Netherlands to perform the 70-gene pr<strong>of</strong>ile<br />

test. For all patients included in this study, tumor samples from excised specimens<br />

were obtained and placed in a commercially available preservation fluid<br />

(RNAlater 1 , Qiagen GmBH, Hilden, Germany) at room temperature. Subsequently,<br />

the samples were sent by conventional mail to NKI, where they<br />

were snap-frozen and stored at –808C. The 70-gene pr<strong>of</strong>ile was performed at the<br />

microarray facility in Amsterdam. Preliminary results showed that it is feasible<br />

to adapt local procedures and to collect good-quality tissue for microarray testing<br />

in a multicentric setting (18,29,30).<br />

In the RASTER study, tissue was preserved temporarily in RNAlater and<br />

sent by conventional mail at room temperature. Preservation <strong>of</strong> tumor tissue in<br />

RNAlater may influence several processes in the tissue, such as protein levels.<br />

Since one <strong>of</strong> the additional aims <strong>of</strong> the MINDACT trial is the establishment <strong>of</strong> a<br />

biological materials bank for future research, including proteomics, storage <strong>of</strong><br />

tissue in RNAlater is not suitable for the MINDACT trial and fresh, frozen tissue<br />

will be mandatory (31). Therefore, a second feasibility study was conducted to<br />

test the collection, storage, and shipment <strong>of</strong> fresh, frozen tissue in six European<br />

hospitals. Women younger than 71 years, with early-stage breast cancer, were<br />

included. After surgery, the pathologist obtained representative tumor samples<br />

using a 6-mm biopsy puncher. Tumor samples were snap-frozen in liquid<br />

nitrogen and stored at –808C within one hour <strong>of</strong> surgery. The samples were<br />

shipped to Amsterdam on dry ice by a contracted courier specializing in the<br />

transportation <strong>of</strong> frozen tissue. At Agendia, RNA was isolated when the samples<br />

were representative <strong>of</strong> the tumor (i.e. tumor cells 50%). After measurement <strong>of</strong><br />

the quality and quantity <strong>of</strong> the RNA, the samples were hybridized on the 70-gene<br />

pr<strong>of</strong>ile platform. Results showed that in general it is feasible to collect goodquality<br />

fresh, frozen tissue for microarray tests in a multicentric and multinational<br />

setting. Additionally, in the course <strong>of</strong> this feasibility study, procedures<br />

were adapted and optimized for use in the MINDACT trial (18).<br />

PROSPECTIVE VALIDATION OF THE 70-GENE<br />

PROFILE IN THE MINDACT TRIAL<br />

Despite the completion <strong>of</strong> all necessary technical validation procedures and the<br />

FDA clearance as the in vitro diagnostic multivariate index assay (IVDMIA),<br />

including approval <strong>of</strong> clinical utility, the clinical validation <strong>of</strong> the 70-gene<br />

pr<strong>of</strong>ile can still be further optimized in currently diagnosed, adjuvantly treated<br />

populations to achieve level-1 evidence. The TRANSBIG consortium has<br />

decided to further validate the 70-gene prognostic alongside the classical clinicopathological<br />

factors (31). This trial, the MINDACT trial [EORTC (European


Microarray Gene Expression Pr<strong>of</strong>iling 131<br />

Organization for Research and Treatment <strong>of</strong> <strong>Cancer</strong>) 10041 BIG (<strong>Breast</strong> International<br />

Group) 3-04], will enroll 6000 women between 18 and 71 years <strong>of</strong> age,<br />

with primary invasive breast cancer at diagnosis. Tumors should be unilateral,<br />

stage T1, T2, or operable T3; carcinomas in situ are eligible provided invasive<br />

cancer is present. Patients should be treated with breast conserving treatment or<br />

mastectomy and should have a negative sentinel node or axillary clearance.<br />

Patients with primary malignancies or previous chemotherapy or radiotherapy<br />

will not be included. The main goal <strong>of</strong> the trial is to prove that the 70-gene<br />

pr<strong>of</strong>ile will safely assign chemotherapy to a smaller proportion <strong>of</strong> node-negative<br />

breast cancer patients. Therefore, all patients included in MINDACT will have a<br />

risk assessment based on the 70-gene pr<strong>of</strong>ile and according to the currently used<br />

clinicopathological criteria. This last risk assessment is provided by an updated<br />

version <strong>of</strong> Adjuvant! Online. If, according to both risk assessments, the patients’<br />

risk <strong>of</strong> developing distant metastases is low (an estimated 13% <strong>of</strong> all patients), no<br />

chemotherapy will be given. If both methods classify the patients’ risk <strong>of</strong> developing<br />

distant metastases as high (an estimated 55% <strong>of</strong> the patients), chemotherapy<br />

will be proposed. If the risk assessments are discordant (an estimated 32% <strong>of</strong><br />

all patients), patients will be randomized to the clinicopathological criteria or<br />

the 70-gene pr<strong>of</strong>ile to be used for adjuvant chemotherapy decision making<br />

(Fig. 3). The group <strong>of</strong> most interest to answer the primary research question<br />

comprises patients with a low-risk 70-gene pr<strong>of</strong>ile (good pr<strong>of</strong>ile) and high-risk<br />

Figure 3 MINDACT study design. Source: From Ref. 18. With permission.<br />

Abbreviations: R-T, treatment decision randomization: CT, chemotherapy.


132 Mook et al.<br />

clinicopathological criteria who will be randomized to use the 70-gene pr<strong>of</strong>ile<br />

for the treatment decision and thus receive no chemotherapy. In this group a<br />

null hypothesis <strong>of</strong> a five-year distant metastases–free survival <strong>of</strong> 92% will be<br />

tested (31).<br />

Additional study objectives are related to the type <strong>of</strong> chemotherapy and<br />

endocrine therapy. Patients who are to receive chemotherapy may be randomized<br />

between an anthracycline-based regimen and a regimen <strong>of</strong> docetaxel and capecitabine<br />

in order to compare the efficacy and safety <strong>of</strong> both regimens. The latter<br />

regimen may have increased efficacy with less long-term side effects such as<br />

cardiotoxicity and leukemia. Post-menopausal (spontaneous or induced) patients<br />

with hormone receptor–positive tumors will be randomized for endocrine therapy<br />

with two years <strong>of</strong> tamoxifen, followed by five years <strong>of</strong> letrozole or seven<br />

years <strong>of</strong> letrozole upfront.<br />

In addition to these three research questions, whole genome arrays will be<br />

performed for all 6000 patients, and frozen and paraffin-embedded tumor samples<br />

together with blood samples will be collected from all patients. These<br />

microarray data and the biological materials will be stored in an independent<br />

biobank and will provide a valuable source <strong>of</strong> information for future research<br />

that may result in the identification <strong>of</strong> new prognostic or predictive biomarkers<br />

and new drug targets.<br />

FUTURE PROSPECTS<br />

The MINDACT trial is the first randomized clinical trial that will prospectively<br />

evaluate the prognostic value <strong>of</strong> a microarray-based prognostic tool, the 70-gene<br />

pr<strong>of</strong>ile or Mammaprint, in an adjuvantly treated population. The implementation<br />

<strong>of</strong> molecular diagnostics based on the true biology <strong>of</strong> a patient’s tumor will<br />

greatly expedite the introduction <strong>of</strong> personalized medicine.<br />

The details <strong>of</strong> the rationale, design, and logistics <strong>of</strong> the MINDACT trial are<br />

described in references 18,23, and 31, and the authors strongly recommend these<br />

articles to the reader.<br />

ACKNOWLEDGMENTS<br />

The studies mentioned in this chapter were supported by the European Commission<br />

Framework Programme VI, the Center <strong>of</strong> Biomedical Genetics, the<br />

Dutch Health Care Insurance Board, the Dutch National Genomic Initiative:<br />

<strong>Cancer</strong> Genomics Program, the <strong>Breast</strong> <strong>Cancer</strong> Research Foundation, the EBCC-<br />

<strong>Breast</strong> <strong>Cancer</strong> Working Group–association sans but lucratif (asbl), the Fondation<br />

Belge Contre Le <strong>Cancer</strong>, the S. G. Komen for the Cure Foundation, Jacqueline<br />

Seroussi Memorial Foundation for <strong>Cancer</strong> Research, and the European Organisation<br />

for Research and Treatment <strong>of</strong> <strong>Cancer</strong> (EORTC)–<strong>Breast</strong> <strong>Cancer</strong> Group.<br />

S. Mook was partially supported by the traineeship program <strong>of</strong> the TRANSBIG<br />

network. The authors thank the numerous individuals who have contributed


Microarray Gene Expression Pr<strong>of</strong>iling 133<br />

to the studies mentioned in this review, especially those from the TRANSBIG<br />

consortium and the EORTC, and all the patients who have and still are participating<br />

in these studies.<br />

CONFLICTS OF INTEREST<br />

Dr. L. J. van ’t Veer is a named inventor on a patent application for Mammaprint<br />

and reports holding equity in Agendia BV.<br />

REFERENCES<br />

1. National Comprehensive <strong>Cancer</strong> Network. NCCN Clinical Practice Guidelines in<br />

Oncology. <strong>Breast</strong> <strong>Cancer</strong> V.I.2008. Available at: http://www.nccn.org/pr<strong>of</strong>essionals/<br />

physician_gls/PDF/breast.pdf.<br />

2. Carlson RW, McCormick B. Update: NCCN breast cancer clinical practice guidelines.<br />

J Natl Compr Canc Netw 2005; 3(suppl 1):S7–S11.<br />

3. Goldhirsch A, Glick JH, Gelber RD, et al. Meeting highlights: international expert<br />

consensus on the primary therapy <strong>of</strong> early breast cancer 2005. Ann Oncol 2005;<br />

16(10):1569–1583.<br />

4. Perou CM, Sorlie T, Eisen MB, et al. Molecular portraits <strong>of</strong> human breast tumours.<br />

Nature 2000; 406(6797):747–752.<br />

5. Hu Z, Fan C, Oh DS, et al. The molecular portraits <strong>of</strong> breast tumors are conserved<br />

across microarray platforms. BMC Genomics 2006; 796.<br />

6. Sorlie T, Perou CM, Tibshirani R, et al. Gene expression patterns <strong>of</strong> breast carcinomas<br />

distinguish tumor subclasses with clinical implications. Proc Natl Acad Sci<br />

U S A. 2001; 98(19):10869–10874.<br />

7. Sorlie T, Tibshirani R, Parker J, et al. Repeated observation <strong>of</strong> breast tumor subtypes<br />

in independent gene expression data sets. Proc Natl Acad Sci U S A. 2003; 100(14):<br />

8418–8423.<br />

8. Sotiriou C, Neo SY, McShane LM, et al. <strong>Breast</strong> cancer classification and prognosis<br />

based on gene expression pr<strong>of</strong>iles from a population-based study. Proc Natl Acad Sci<br />

U.S.A. 2003; 100(18):10393–10398.<br />

9. Chang JC, Wooten EC, Tsimelzon A, et al. Gene expression pr<strong>of</strong>iling for the prediction<br />

<strong>of</strong> therapeutic response to docetaxel in patients with breast cancer. Lancet<br />

2003; 362(9381):362–369.<br />

10. Huang E, Cheng SH, Dressman H, et al. Gene expression predictors <strong>of</strong> breast cancer<br />

outcomes. Lancet 2003; 361(9369):1590–1596.<br />

11. Jansen MP, Foekens JA, van SI, et al. Molecular classification <strong>of</strong> tamoxifen-resistant<br />

breast carcinomas by gene expression pr<strong>of</strong>iling. J Clin Oncol 2005; 23(4):732–740.<br />

12. Paik S, Shak S, Tang G, et al. A multigene assay to predict recurrence <strong>of</strong> tamoxifentreated,<br />

node-negative breast cancer. New England J Med 2004; 351(27):2817–2826.<br />

13. van ’t Veer LJ, Dai H, van de Vijver MJ, et al. Gene expression pr<strong>of</strong>iling predicts<br />

clinical outcome <strong>of</strong> breast cancer. Nature 2002; 415(6871):530–536.<br />

14. van de Vijver MJ, He YD, van ’t Veer LJ, et al. A Gene-expression signature as a<br />

predictor <strong>of</strong> survival in breast cancer. New England J Med 2002; 347(25):1999–2009.<br />

15. Wang Y, Klijn JG, Zhang Y, et al. Gene-expression pr<strong>of</strong>iles to predict distant metastasis<br />

<strong>of</strong> lymph-node-negative primary breast cancer. Lancet 2005; 365(9460):671–679.


134 Mook et al.<br />

16. West M, Blanchette C, Dressman H, et al. Predicting the clinical status <strong>of</strong> human<br />

breast cancer by using gene expression pr<strong>of</strong>iles. Proc Natl Acad Sci U.S.A. 2001;<br />

98(20):11462–11467.<br />

17. Hughes TR, Mao M, Jones AR, et al. Expression pr<strong>of</strong>iling using microarrays fabricated<br />

by an ink-jet oligonucleotide synthesizer. Nat Biotech 2001; 19(4):342–347.<br />

18. Mook S, ’van ’t Veer LJ, Rutgers EJ, et al. Individualization <strong>of</strong> therapy using<br />

Mammaprint: from development to the MINDACT trial. <strong>Cancer</strong> Genomics Proteomics<br />

2007; 4(3):147–155.<br />

19. Eifel P, Axelson JA, Costa J, et al. National Institutes <strong>of</strong> Health Consensus Development<br />

Conference Statement: adjuvant therapy for breast cancer, November 1–3,<br />

2000. JNCI <strong>Cancer</strong> Spectrum 2001; 93(13):979–989.<br />

20. Buyse M, Loi S, van ’t Veer L, et al. Validation and clinical utility <strong>of</strong> a 70-gene<br />

prognostic signature for women with node-negative breast cancer. JNCI <strong>Cancer</strong><br />

Spectrum 2006; 98(17):1183–1192.<br />

21. Olivotto IA, Bajdik CD, Ravdin PM, et al. Population-based validation <strong>of</strong> the<br />

prognostic model ADJUVANT! for early breast cancer. J Clin Oncol 2005; 23(12):<br />

2716–2725.<br />

22. Ravdin PM, Simin<strong>of</strong>f LA, Davis GJ, et al. Computer program to assist in making<br />

decisions about adjuvant therapy for women with early breast cancer. J Clin Oncol<br />

2001; 19(4):980–991.<br />

23. Cardoso F, van ’t Veer L, Rutgers E, et al. Clinical application <strong>of</strong> the 70-gene pr<strong>of</strong>ile<br />

(Mammaprint TM ): the MINDACT trial. J Clin Oncol 2008; 26(5):729–735.<br />

24. Desmedt C, Piette F, Loi S, et al. Strong time dependence <strong>of</strong> the 76-gene prognostic<br />

signature for node-negative breast cancer patients in the TRANSBIG multicenter<br />

independent validation series. Clin <strong>Cancer</strong> Res 2007; 13(11):3207–3214.<br />

25. Chang HY, Nuyten DSA, Sneddon JB, et al. Robustness, scalability, and integration<br />

<strong>of</strong> a wound-response gene expression signature in predicting breast cancer survival.<br />

Proc Natl Acad Sci U S A. 2005; 102(10):3738–3743.<br />

26. Sotiriou C, Wirapati P, Loi S, et al. Gene expression pr<strong>of</strong>iling in breast cancer:<br />

understanding the molecular basis <strong>of</strong> histologic grade to improve prognosis. JNCI<br />

<strong>Cancer</strong> Spectrum 2006; 98(4):262–272.<br />

27. Foekens JA, Atkins D, Zhang Y, et al. Multicenter validation <strong>of</strong> a gene expressionbased<br />

prognostic signature in lymph node-negative primary breast cancer. J Clin<br />

Oncol 2006; 24(11):1665–1671.<br />

28. Couzin J. DIAGNOSTICS: amid debate, gene-based cancer test approved. Science<br />

2007; 315(5814):924.<br />

29. Bueno-de-Mesquita JM, van Hakten WH, Retel VP, et al. use <strong>of</strong> 70-gene signature to<br />

predict prognosis <strong>of</strong> patients with node-negative breast cancer: a prospective community-based<br />

feasibility study (RASTER). Lancet Oncol 2008; 8: 1079–1087.<br />

30. van de Vijver M. Gene-expression pr<strong>of</strong>iling and the future <strong>of</strong> adjuvant therapy.<br />

Oncologist 2005; 10(suppl 2):30–34.<br />

31. Bogaerts J, Cardoso F, Buyse M, et al. Gene signature evaluation as a prognostic<br />

tool: challenges in the design <strong>of</strong> the MINDACT trial. Nat Clin Pract Oncol 2006; 3(10):<br />

540–551.<br />

32. Sauter G, Simon R. Predictive molecular pathology. N Engl J Med 347 2002; (25):<br />

1995–1996.


10<br />

Predictive Markers for Targeted <strong>Breast</strong><br />

<strong>Cancer</strong> Treatment<br />

Hans Christian B. Pedersen a and John M. S. Bartlett<br />

Endocrine <strong>Cancer</strong> Group, <strong>Cancer</strong> Research Centre,<br />

Western General Hospital, Edinburgh, United Kingdom<br />

INTRODUCTION<br />

Currently selection <strong>of</strong> breast cancer treatment is guided predominantly by patient<br />

prognosis using classical pathological assessment <strong>of</strong> tumors, with higher risk<br />

patient groups being <strong>of</strong>fered more aggressive therapy. The choice <strong>of</strong> treatment<br />

regimen is guided by a very small number <strong>of</strong> predictive biomarkers (Table 1) (1).<br />

However, it is now clearly recognized that not only may the risks <strong>of</strong> treatment<br />

outweigh the benefits in some patient groups but that not all patients are equal<br />

with respect to their response to and benefit gained from exposure to novel and<br />

established therapies. There is increasing interest in the identification <strong>of</strong> biological<br />

markers to improve the targeting <strong>of</strong> established therapies to those cancers<br />

likely to respond. In addition, development <strong>of</strong> new molecular targeted drugs has<br />

led to an urgent need for research into predictive markers that can distinguish<br />

patients who benefit from treatment. New proteomic technologies are leading to<br />

markers being discovered at an increasing rate, with protein expression, cellular<br />

localization, and posttranslational modification being recognized as those <strong>of</strong><br />

potential importance. Many new immunohistochemical (IHC) markers are<br />

a Hans Christian B. Pedersen is employed on a Marie Curie Transfer <strong>of</strong> Knowledge project between<br />

Dako A/S, Glostrup, Denmark, and University <strong>of</strong> Edinburgh.<br />

135


136 Pedersen and Bartlett<br />

Table 1<br />

Current FDA-Approved Predictive Markers in <strong>Breast</strong><br />

Predictive Marker Method Drug<br />

ERa IHC Tamoxifen, AIs<br />

PR IHC Tamoxifen, AIs<br />

HER2/neu IHC Trastuzumab (Herceptin)<br />

HER2/neu FISH Trastuzumab (Herceptin)<br />

Abbreviations: ERa, estrogen receptor a; PR, progesterone receptor; AIs,<br />

aromatase inhibitors; FISH, fluorescence in situ hybridization.<br />

currently being tested for predictive potential for various therapies in breast<br />

cancer. Through large-scale gene expression studies using microarrays, it has<br />

been realized that tumors have molecular signatures that may determine response<br />

to treatment and patient prognosis. Many candidate-predictive and prognostic<br />

signatures in breast cancer are being explored, and some are already in use to aid<br />

patient management (e.g., OncoType Dx 1 , Genomic Health, Redwood City,<br />

California, U.S.). The challenge for the future is to identify both the biological<br />

and diagnostic determinates <strong>of</strong> treatment response in order to allow matching <strong>of</strong><br />

appropriate treatments to specific patients.<br />

Future breast cancer treatment is likely to be guided by predictive markers<br />

along with genomic signatures, treating patients with mixtures <strong>of</strong> molecular<br />

targeted therapeutics. This chapter will review current IHC and fluorescent in<br />

situ hybridization (FISH)-based predictive markers in breast and the challenges<br />

associated with the development and implementation <strong>of</strong> predictive markers in<br />

future.<br />

CURRENT PREDICTIVE MARKERS IN BREAST CANCER: LESSONS LEARNT<br />

A predictive marker is defined as a factor that indicates sensitivity or resistance<br />

to a specific treatment (2). In contrast, a prognostic marker predicts a tumor’s<br />

future behavior independent <strong>of</strong> systemic adjuvant treatment. Some markers can<br />

be both predictive and prognostic, as exemplified by the estrogen receptor a<br />

(ERa) expression, which both predicts response to endocrine treatment and<br />

correlates with prognosis.<br />

Hormonal Therapy and Hormone Receptors<br />

ERa and progesterone receptor (PR) status were the first predictive markers used<br />

in the clinic to select patients for breast cancer treatment. Today, IHC ERa or PR<br />

staining is done routinely on patient tumor samples to predict response to<br />

endocrine therapy. More correctly, the assay is used to exclude ERa- or PRnegative<br />

patients from ERa- or PR-targeted therapy, as they will have a low<br />

probability <strong>of</strong> responding to treatment (3). Historically however, ERa was first<br />

discovered in the late 1960s on its ability to bind radiolabeled ligand (4,5).


Predictive Markers for Targeted <strong>Breast</strong> <strong>Cancer</strong> Treatment 137<br />

Within 10 years <strong>of</strong> its initial discovery, assays had been developed that allowed<br />

ER to be used a marker for hormonal responsiveness and clinical aggressiveness<br />

<strong>of</strong> the tumor (6). Many years passed before it was realized that only ERapositive<br />

patients could benefit from treatment and for these finding to be applied<br />

as clinical practice (3). It is now almost universally accepted that ERa testing has<br />

a central role in breast cancer management, but it is worth noting that no prospective<br />

trial has yet tested the predictive value <strong>of</strong> ERa or PR as predictive<br />

biomarkers. Later evidence has emerged supporting testing for PR as well as<br />

ERa (7). PR is a transcriptional target <strong>of</strong> ligand-bound ERa, and it is has been<br />

proved that including the PR score in the ERa score provides stronger evidence<br />

for true ERa positivity and therefore makes patients more likely to respond to<br />

therapy (8–10). However, controversy remains over the value <strong>of</strong> supplementary<br />

markers to ERa, and in several countries, PR testing in not routinely performed.<br />

The drugs administered to ERa- or PR-positive patients have been through<br />

many generations through the last three decades, from pure ERa-antagonist<br />

(fulvestrant) to ERa–partial antagonist tamoxifen and drugs inhibiting the enzyme<br />

aromatase that synthesise oestrogen (11). Current clinical practice in breast cancer<br />

management utilizes a combination adjuvant tamoxifen followed up by aromatase<br />

inhibitors (AIs) or AIs alone. However, ER or PR do not select between these<br />

therapies, and despite evidence that PR is predictive <strong>of</strong> poor response on tamoxifen,<br />

there is no evidence, to date, that AIs circumvent this resistance.<br />

Despite the long history <strong>of</strong> the ERa or PR status as a predictive and<br />

prognostic marker, no standardized protocols for fixation and IHC staining are<br />

yet widely applied. Conflicting evidence exists for when patients are considered<br />

ER negative and therefore will not respond to treatment, but above 10% positivestaining<br />

tumor cells are generally considered ER positive (12). Perhaps because<br />

ERa has been available for testing as a predictive marker for such a long<br />

time, before emphasis was put on standardizing IHC testing, this has lead to the<br />

development <strong>of</strong> a multitude <strong>of</strong> different protocols using a variety <strong>of</strong> antibodies.<br />

ER IHC results have been shown to be highly influenced by the efficiency <strong>of</strong> the<br />

antigen-retrieval step (13). External quality assurance initiatives such as the<br />

United Kingdom National External Quality Assessment Scheme for Immunocytochemistry<br />

(UK NEQAS-ICC) exist now to improve quality across clinical<br />

laboratories through external assessment <strong>of</strong> analytical and interpretative performance;<br />

such schemes are however not uniformly successful.<br />

In addition to standardizing ERa IHC for scoring tumors as positive or<br />

negative, it has been suggested that quantification <strong>of</strong> ERa levels may be necessary<br />

in some clinical settings because tumors expressing low or intermediate<br />

ERa levels have been shown to benefit from chemotherapy in addition to<br />

endocrine therapy (14). Current ERa IHC protocols are semiquantitative, so<br />

alternatives such as radiolabeled ligand-binding assays or quantitative polymerase<br />

chain reaction (qPCR) have been suggested. Gene expression <strong>of</strong> ER and<br />

PR correlate well with protein expression, and it has been suggested to be an<br />

alternative to IHC as it is sensitive and quantitative over a large dynamic


138 Pedersen and Bartlett<br />

range (15). It has been suggested that more quantitative methods <strong>of</strong> measuring<br />

ERa and PR may provide additional information to guide choice <strong>of</strong> therapy. We<br />

will discuss this in detail later (12).<br />

HER2/neu and Trastuzumab (Herceptin 1 )<br />

HER2/neu or human epidermal growth factor receptor 2, a product <strong>of</strong> the protooncogene<br />

HER2 first discovered in 1985, belongs to the HER family <strong>of</strong> tyrosine<br />

kinase receptors (16–18). Numerous studies have documented HER2/neupositive<br />

breast cancers display more aggressive disease and shortened diseasefree<br />

survival (19). Using recombinant technologies, trastuzumab (Herceptin 1 ,<br />

Genentech, South San Francisco, California, U.S.), a monoclonal immunoglobulin<br />

G1 class, humanized murine antibody, was developed to specifically target<br />

patients with breast cancer that overexpressed the HER2/neu protein (20,21).<br />

Clinical trials demonstrated the efficiency <strong>of</strong> trastuzumab for treating metastatic<br />

breast cancers and subsequent trials in combination therapy (21–25) for HER2<br />

positive disease. Again, HER2/neu, like ERa, is more predictive <strong>of</strong> which<br />

patients will not respond to treatment.<br />

As in the case for endocrine therapy, most predictive marker development<br />

followed drug development, leading to some confusion. In initial clinical trials,<br />

patient selection was based on an IHC assay using the mouse monoclonal<br />

antibody that had been humanized to create trastuzumab (26,27). Today most<br />

patients are selected for trastuzumab therapy on the basis <strong>of</strong> HER2/neu<br />

expression levels in tumor cells (Fig. 1). In 2001, HER2/neu testing achieved<br />

standard care <strong>of</strong> practice in the American Society <strong>of</strong> Clinical Oncology (ASCO)<br />

breast cancer clinical practice guidelines. In most laboratories, tumors are<br />

screened with FDA-approved IHC tests, HercepTest 1 (Dako, Glostrup, Denmark)<br />

or HER2/neu (CB11) PATHWAY 1 (Ventana Medical Systems, Inc., Tucson,<br />

Arizona, U.S.). Ambiguous cases are then analyzed with FDA-approved HER2/<br />

neu FISH from Vysis (Downers Grove, Illinois, U.S.), Dako, or Ventana. Alternatively,<br />

in some laboratories, FISH is the only method used for HER2/neu testing.<br />

HER2/neu testing for trastuzumab therapy has undergone considerable scrutiny<br />

in numerous studies, because evidence emerged indicated that the IHC tests<br />

generated considerable amount <strong>of</strong> false positives and negatives (28–31). This<br />

discrepancy is due in part to lack <strong>of</strong> adherence to protocols, variable fixation<br />

methods, and subjectivity related to tumor scoring.<br />

Only 20% to 35% <strong>of</strong> HER2/neu positive tumors respond to treatment,<br />

depending on the context the drug is given in, and studies have demonstrated that<br />

this may be attributed to a combination <strong>of</strong> factors, including lack <strong>of</strong> pathway<br />

dependence, receptor being inactive despite overexpressed, or the induction <strong>of</strong><br />

resistance through various mechanisms (32). Controversy persists with regard to<br />

the optimal method for HER2/neu testing and interpretation <strong>of</strong> results (33). It has<br />

been reported that it may be more cost effective to test all tumors with HER2/neu<br />

FISH, avoiding false positives and false negatives (28). This is a possibility, as


Predictive Markers for Targeted <strong>Breast</strong> <strong>Cancer</strong> Treatment 139<br />

Figure 1 Current HER2/neu testing in breast cancer. IHC staining performed on patient<br />

tumor samples and scored using the HercepTest. Patient testing 0 to 1þ go on to receive<br />

other treatments. Patients testing 2þ go on to HER2/neu FISH testing and if amplified for<br />

HER2/neu receive trastuzumab. Patients scoring 3þ receive trastuzumab. Alternatively all<br />

patients undergo HER2/neu FISH and patients scored as amplified receive trastuzumab.<br />

Abbreviations: IHC, immunohistochemical; FISH, fluorescence in situ hybridization.<br />

FISH and, more critically, chromogenic in situ hybridization (CISH) technology<br />

become more widely distributed, and an agreement can be made on the best<br />

choice <strong>of</strong> HER2/neu testing.<br />

DEVELOPMENT OF PREDICTIVE MARKERS<br />

The successful targeting <strong>of</strong> HER2/neu with trastuzumab inspired the development<br />

<strong>of</strong> a wave <strong>of</strong> new therapeutic antibodies and kinase inhibitors that were<br />

aimed at inhibiting growth factor–regulated signaling. However, repeating the<br />

success <strong>of</strong> HER2/neu testing for trastuzumab therapy has proven difficult.<br />

Inhibitors <strong>of</strong> mammalian Target Of Rapamycin (mTOR, e.g., temsirolimus,<br />

Toricel 1 , Wyeth <strong>Ph</strong>armaceuticals, Inc., Madison, New Jersey, U.S.), epidermal<br />

growth factor receptor (EGFR, e.g., cetuximab, Erbitux 1 , ImClone Systems, Inc.,<br />

Branchburg, New Jersey, U.S.) and vascular endothelial growth factor receptor<br />

(VEGFR, e.g., bevacizumab, Avastin 1 , Genentech) have shown modest or no<br />

results in clinical trials as single agent therapies in breast cancer (34–36). This may<br />

be partly attributed to lack <strong>of</strong> predictive markers <strong>of</strong> response in combination with<br />

drugs simply not being effective as single agents. However, this failure may also


140 Pedersen and Bartlett<br />

reflect poorly designed approaches toward the discovery <strong>of</strong> predictive biomarkers.<br />

Predictive markers for EGFR-targeted therapy have been in focus for the past five<br />

years, and we will therefore briefly go through the history <strong>of</strong> EGFR-targeted<br />

therapy and what has been learnt for future predictive marker development. In<br />

addition, a model for codevelopment <strong>of</strong> predictive markers in relation to traditional<br />

drug development is presented.<br />

EGFR—A Cautionary Tale<br />

EGFR or HER1 belongs to the same family <strong>of</strong> tyrosine kinase receptors as<br />

HER2/neu and EGFR is considered an attractive target for therapeutic intervention<br />

in many cancer forms due to the receptor being frequently mutated and<br />

overexpressed (34). EGFR overexpression is associated with reduced diseasefree<br />

survival and resistance to chemotherapy (37). EGFR inhibitors (e.g., erlotinib<br />

and gefitinib) and therapeutic antibodies against the extracellular domain <strong>of</strong><br />

the receptor (e.g., cetuximab, matuzumab, and panitumumab) are currently in<br />

clinical trials or late-stage development (38). Most clinical trials have investigated<br />

EGFR inhibitors in cancers <strong>of</strong> the lung and colon, and cetuximab has<br />

been approved for treatment <strong>of</strong> colorectal cancer (34). Clinical trials have also<br />

been conducted in breast cancer, with limited success (39–47). In several <strong>of</strong> these<br />

trials, patients have been selected for trial on the basis <strong>of</strong> EGFR overexpression.<br />

However, this was later shown to be a mistake because EGFR expression, as<br />

currently evaluated, did not correlate with clinical response (48,49) or with<br />

breast cancer cell line inhibition (50). Clinical trials in breast cancer have clearly<br />

demonstrated that EGFR inhibitors and therapeutic antibodies are targeting<br />

EGFR receptors in both tumor and skin from patients, so the lack <strong>of</strong> effect is not<br />

due to the drug’s failure to reach the target or the target’s lack <strong>of</strong> inhibition (41).<br />

So EGFR inhibitors hit their target, and certain cancer patients clearly benefit<br />

from these drugs. Therefore, it is surprising that although EGFR-targeted therapy<br />

has been in focus for the past decade and drugs are now being administered to<br />

patients, no FDA-approved test or pr<strong>of</strong>ile exists at present to guide clinicians on<br />

the use <strong>of</strong> EGFR inhibitors. Studies have shown EGFR gene copy number,<br />

mutations, and downstream signalling to be associated with clinical response in a<br />

series <strong>of</strong> lung cancer trials (51–55). However, these were retrospective, poorly<br />

powered studies, which are <strong>of</strong> insufficient impact to influence molecular diagnostics.<br />

Taken together, what has emerged from clinical trials in breast and other<br />

cancers along with findings from in vitro studies is that EGFR inhibition is not<br />

similar to HER2/neu-targeted therapy in breast cancer, where a single, wellcharacterized<br />

molecular event can be used to target treatment. Therefore, what<br />

constitutes a good predictive marker for EGFR-targeted therapy in breast cancer<br />

is currently not clear. For future clinical trials with EGFR inhibitors to be successful<br />

and current, approved EGFR inhibitors to be administered to the right<br />

patients, it will be necessary to understand better the biology <strong>of</strong> EGFR-directed<br />

signaling and the relative importance <strong>of</strong> mutations, gene copy number, and


Predictive Markers for Targeted <strong>Breast</strong> <strong>Cancer</strong> Treatment 141<br />

downstream signaling pathways have on therapeutic response. These findings<br />

can then be used to setup adequately designed molecular pathology and translational<br />

studies within clinical trials that will address predictive markers for<br />

EGFR-targeted therapy.<br />

Predictive Markers in Drug Development<br />

Future clinical trials should utilize findings from HER2/neu and EGFR-targeted<br />

therapy. We have learnt that overexpression or amplification alone does not<br />

necessarily mean that protein can be targeted with drugs or that it is indicative <strong>of</strong><br />

aberrant activation. Posttranslational modifications such as phosphorylation,<br />

ubiquitination and acetylation, and localization are critical for protein function.<br />

Furthermore, what has become clear is that preclinical models do not reflect the<br />

complexity <strong>of</strong> human tumors well, and, therefore, predictive marker selection<br />

must be validated in clinical settings. Despite all these challenges, drug development<br />

must try and entail predictive marker development so that stratification<br />

<strong>of</strong> patients can be made possible, thus avoiding unnecessary side effects and<br />

improving the chance <strong>of</strong> a trial being successful. Without an increased focus on<br />

predictive marker–directed patient selection, many <strong>of</strong> the drugs currently in<br />

clinical trials run the risk <strong>of</strong> failing to reach the clinic despite being efficacious<br />

for a subpopulation <strong>of</strong> patients. Therefore, it should be cost effective for pharmaceutical<br />

companies to build predictive marker strategies in drug development,<br />

as this may improve the success rate <strong>of</strong> drug development in clinicals. In<br />

addition, it would be useful to develop predictive markers <strong>of</strong> response to many <strong>of</strong><br />

the drugs already in the market, thus improving patient care and reducing<br />

treatment costs by only treating patients who benefit from the treatment.<br />

The major risk we face at present is that we will continue to use poorly<br />

designed retrospective analyses <strong>of</strong> existing trial populations to seek to identify<br />

markers on the basis <strong>of</strong> incomplete understanding <strong>of</strong> the biology <strong>of</strong> drug<br />

response. This is, in part, due to the fact that we have no reference criteria<br />

against which to judge biomarker development. For ERa, despite no level-1<br />

evidence being available, it is unthinkable that this biomarker be regarded as<br />

anything less than essential to the management <strong>of</strong> breast cancer in the twentyfirst<br />

century. For HER2, only after the central importance <strong>of</strong> this biomarker was<br />

established were steps put in place to produce robust, accurate, and clinically<br />

viable test methods. However, clearly for future biomarkers, a structured<br />

approach will improve the likelihood <strong>of</strong> selecting appropriate markers. For this<br />

to succeed, we need to progress from the current model to one where, as with<br />

drug development, critical attention is paid to each step <strong>of</strong> the process by which<br />

novel biomarkers are developed (56).<br />

To develop and validate predictive markers, a number <strong>of</strong> hurdles have to be<br />

passed before they can be considered for clinical practice. These processes should<br />

be addressed in a structured manner to maximise the possibility <strong>of</strong> success. For<br />

example, the development <strong>of</strong> robust and accurate diagnostic assays should


142 Pedersen and Bartlett<br />

Figure 2 Development <strong>of</strong> IHC- and FISH-based predictive markers. Abbreviations: IHC,<br />

immunohistochemical; FISH, fluorescence in situ hybridization.<br />

precede, not follow, the identification <strong>of</strong> a potential clinical role for biomarkers<br />

(56). To illustrate this, we have adapted a biomarker development scheme from<br />

Pepe et al. and Hayes et al. (57,58) into four phases (Figure 2).<br />

l<br />

l<br />

<strong>Ph</strong>ase 1: Discovery stage. In order to discover predictive markers <strong>of</strong><br />

response, high-throughput gene expression microarrays or proteomic<br />

technologies are <strong>of</strong>ten used. The primary output <strong>of</strong> this phase is to develop<br />

a set <strong>of</strong> potential predictive markers that allow selecting patients who<br />

respond to treatment. Ideally, samples from the neoadjuvant setting should<br />

be used so that information about the influence that treatment has on the<br />

markers selected can be obtained.<br />

<strong>Ph</strong>ase 2: Development stage. In this stage, the predictive marker assay now<br />

needs make the transition from the research setting to the clinical diagnostic<br />

laboratory. At this stage, it is required to develop the assay for<br />

correct determination <strong>of</strong> predictive marker <strong>of</strong> response to a level where it


Predictive Markers for Targeted <strong>Breast</strong> <strong>Cancer</strong> Treatment 143<br />

can be used by clinical institutions, according to guidelines put forward by<br />

ASCO and EORTC (The European Organization for Research and<br />

Treatment <strong>of</strong> <strong>Cancer</strong>) (58,59). This means that proper instructions need to<br />

be in place for all aspects <strong>of</strong> predictive marker testing, including sample<br />

handling and pretreatment, protocols and standards for ensuring uniform<br />

testing across institutions, and external quality assurance. It is desirable<br />

that a set <strong>of</strong> positive controls are developed for interlaboratory variation to<br />

be accounted for in the same fashion as the HercepTest comes with positive<br />

control cell lines.<br />

l <strong>Ph</strong>ase 3: Validation phase. In order to validate markers found in phase 1,<br />

retrospective validation studies need to be set up. When setting up clinical<br />

trial at this stage, it is important that one <strong>of</strong> the endpoints <strong>of</strong> the trial is to<br />

assess predictive markers. One <strong>of</strong> the primary outcomes <strong>of</strong> this phase is to<br />

evaluate how well the predictive markers perform in independent<br />

validation studies and what predictive markers perform compared better<br />

with traditional diagnostic markers such as nodal status, grade, etc.<br />

l <strong>Ph</strong>ase 4: Clinical utility phase. If success is obtained in phase 3, the marker<br />

can be move to phase 4, with clinical studies utilizing randomized<br />

prospective screening with predictive markers. The primary outcome <strong>of</strong><br />

phase 4 is to evaluate if patients selected with predictive markers were the<br />

ones that benefited from treatment and if it is ethically defendable to select<br />

patients on the basis <strong>of</strong> the results <strong>of</strong> the clinical studies. The predictive<br />

marker then needs to go through regulatory approval for the defined<br />

indication before being applied in the clinic.<br />

Predictive marker development requires the joint effort <strong>of</strong> multiple academic<br />

institutions and industry to cover all the aspects <strong>of</strong> development. It takes<br />

many years and large investments, and the risk for failure is high because <strong>of</strong> the<br />

nature <strong>of</strong> drug development. That may be the reason why so few new markers<br />

have been introduced in recent years despite major research and development<br />

efforts and new techniques.<br />

Future Perspectives<br />

<strong>Cancer</strong> is a disease that involves the complex interplay <strong>of</strong> multiple genes in a<br />

framework <strong>of</strong> mutations in the cancer cell. In addition, tumor cells have interactions<br />

with other cells, thereby further promoting growth and invasiveness.<br />

Patients respond differently to treatment based on their tumor characteristics.<br />

Despite recent challenges, we foresee the strong likelihood that current research<br />

and drug development will introduce more predictive markers into IHC laboratories<br />

in the same fashion as trastuzumab has resulted in HER2/neu being<br />

routinely tested for (see above). This foresight is based on the premise that more<br />

care and more robust approaches are applied to the development <strong>of</strong> predictive<br />

biomarkers. It is likely that within two to five years, a panel <strong>of</strong> markers will


144 Pedersen and Bartlett<br />

be tested on every tumor sample in order to decide optimal therapy regimen.<br />

Currently only three predictive IHC markers (ERa, PR, and HER2/neu) are tested<br />

for routinely in breast, but already some pathologists use additional markers to<br />

inform decisions, e.g., p53, Ki67, and TOP2A. As more and more predictive IHC<br />

markers are introduced in clinical laboratories, this will lead to an increased<br />

pressure on pathology laboratories to deliver fast and accurate decisions. How can<br />

this be achieved and what technological advances can improve IHC testing?<br />

Current IHC testing in pathology laboratories is subjective and semiquantitative,<br />

as slides are scored manually and using chromogenic dyes (56–61).<br />

The use <strong>of</strong> computer-assisted analysis <strong>of</strong> predictive markers has been proven to<br />

reduce slide-scoring variability among pathologists (31). While FISH/CISH<br />

analysis is more robust, there remains room for improvement in this area also.<br />

We foresee the introduction <strong>of</strong> image analysis as a standard tool for IHC scoring,<br />

which will begin the process <strong>of</strong> moving the analytical phase from the microscope<br />

to the computer screen.<br />

As analytical challenges increase, accurate quantitation will become<br />

increasingly important. Already evidence suggests that additional information<br />

could be derived if it were possible to robustly quantify ERa and Ki67. There is,<br />

therefore, a strong opportunity for systems such as the Automated QUantitative<br />

Analysis (AQUA 1 ) from HistoRX, Inc., New Haven, Connecticut, U.S., to pave<br />

the way for a switch from traditional chromogenic staining techniques to fluorescent<br />

staining techniques which are potentially quantitative (wide dynamic<br />

range) (62). A further advantage is that fluorescent labels can be multiplexed<br />

together, yielding spatial information and quantization <strong>of</strong> proteins in different<br />

compartments <strong>of</strong> the cell. New techniques being developed, such as the proximity<br />

ligation in situ assay, show promising results for quantitative analysis <strong>of</strong><br />

specific protein interactions in FFPE, i.e., formalin-fixed, paraffin-embedded,<br />

tissue (63). In some cases, this may be superior to the measurement <strong>of</strong> protein<br />

concentrations or posttranslational modifications, as most proteins take part in<br />

multiprotein complexes and their inclusion in the complex indicates relevant<br />

activity. In addition, several protein-interaction inhibitors are currently in clinical<br />

trials, and, therefore, measurement <strong>of</strong> the amount <strong>of</strong> target (interacting proteins) in<br />

the tissue may be a predictive marker for responsiveness to the drug.<br />

There remains a key question: Will we be testing protein and DNA<br />

markers on slides at all in 5 to 10 years? There are a number <strong>of</strong> developments in<br />

process now miniaturizing assays onto small chips that will measure hundreds <strong>of</strong><br />

protein or DNA markers quantitatively and in a reproducible fashion. Such<br />

technological advances will change the pathology workflow in the future so that<br />

every cancer sample coming into the clinic will be analyzed by the pathologist,<br />

tumor cells microdissected into solution, and DNA, protein, and RNA extracted<br />

and applied onto analytical chips. The readout from these analyses could then be<br />

used to determine optimal therapy regimen for the patient and hopefully cure the<br />

disease. However, until these techniques have been properly validated and<br />

developed for clinical use and trials have proven their value, we are likely to rely


Predictive Markers for Targeted <strong>Breast</strong> <strong>Cancer</strong> Treatment 145<br />

on traditional in situ techniques for protein analysis (IHC) and molecular<br />

determination <strong>of</strong> gene amplifications (FISH/CISH).<br />

In summary, introduction <strong>of</strong> quality control, automated platforms, and<br />

computer-assisted scoring will lead to higher throughput and more accurate<br />

decisions with regard to treatment and facilitate the development <strong>of</strong> future<br />

predictive markers.<br />

CONCLUSION<br />

As our understanding <strong>of</strong> cancer as a disease progresses and the tools to investigate<br />

cancer biology develops, the complexity <strong>of</strong> the disease is unraveling.<br />

<strong>Breast</strong> cancers are now being classified not only on the basis <strong>of</strong> traditional<br />

pathology traits but also according to molecular phenotype and, more recently,<br />

gene expression pr<strong>of</strong>iles. New molecular targeted therapeutics will lead to more<br />

predictive markers being introduced, and these markers must be prospectively<br />

validated, internally and externally quality assured, and rigorously quantitative to<br />

allow correct decisions for treatment. Therefore, an evolution <strong>of</strong> the pathology<br />

laboratory is likely to take place in the near future, leading to a switch to<br />

fluorescent-based techniques, digital imaging <strong>of</strong> slides, and computerized scoring<br />

<strong>of</strong> predictive markers. On a longer time horizon, new technologies being<br />

developed and applied are likely to change cancer diagnostics so that molecular<br />

phenotyping <strong>of</strong> patients will take center stage in patient management. In the not<br />

too distant future, there is the possibility that pathologists will type and stage<br />

tumors and simultaneously identify key areas <strong>of</strong> tissue for automated microdissection<br />

and processing. The subsequent process <strong>of</strong> molecular phenotyping,<br />

which may include kinomic, proteomic, and transcriptomic assays performed<br />

simultaneously could be carried out in specialised laboratories using robotics and<br />

micr<strong>of</strong>luidics on high-density protein and gene arrays to determine optimal<br />

treatment regimens for tumors with specific molecular pr<strong>of</strong>iles linked to drug<br />

response. Such a future vision may represent the more radical view <strong>of</strong> what could<br />

be accomplished by linking <strong>of</strong> current concepts and technologies to maximize<br />

patient benefit from therapeutic options. However, we must dream <strong>of</strong> such<br />

futures if we are to maximize progress in the area <strong>of</strong> molecular pathology for the<br />

benefit <strong>of</strong> the most important people, our patients.<br />

REFERENCES<br />

1. Edwards BK, Brown ML, Wingo PA, et al. Annual report to the nation on the status<br />

<strong>of</strong> cancer, 1975-2002, featuring population-based trends in cancer treatment. J Natl<br />

<strong>Cancer</strong> Inst 2005; 97(19):1407–1427.<br />

2. Duffy MJ. Predictive markers in breast and other cancers: A review. Clin Chem<br />

2005; 51(3):494–503.<br />

3. Osborne CK. Steroid hormone receptors in breast cancer management. <strong>Breast</strong> <strong>Cancer</strong><br />

Res Treat 1998; 51(3):227–238.


146 Pedersen and Bartlett<br />

4. Jensen EV, DeSombre ER. Mechanism <strong>of</strong> action <strong>of</strong> the female sex hormones. Annu<br />

Rev Biochem 1972; 41:203–230.<br />

5. Gorski J, T<strong>of</strong>t D, Shyamala G, et al. Hormone receptors: studies on the interaction <strong>of</strong><br />

estrogen with the uterus. Recent Prog Horm Res 1968; 24:45–80.<br />

6. Osborne CK, Yochmowitz MG, Knight WA III, et al. The value <strong>of</strong> estrogen and<br />

progesterone receptors in the treatment <strong>of</strong> breast cancer. <strong>Cancer</strong> 1980; 46(12 suppl):<br />

2884–2888.<br />

7. Horwitz KB, McGuire WL. Predicting response to endocrine therapy in human<br />

breast cancer: a hypothesis. Science 1975; 189(4204):726–727.<br />

8. Elledge RM, Green S, Pugh R, et al. Estrogen receptor (ER) and progesterone<br />

receptor (PgR), by ligand-binding assay compared with ER, PgR and pS2, by<br />

immuno-histochemistry in predicting response to tamoxifen in metastatic breast<br />

cancer: a Southwest Oncology Group Study. Int J <strong>Cancer</strong> 2000; 89(2):111–117.<br />

9. Ravdin PM, Green S, Dorr TM, et al. Prognostic significance <strong>of</strong> progesterone<br />

receptor levels in estrogen receptor-positive patients with metastatic breast cancer<br />

treated with tamoxifen: results <strong>of</strong> a prospective Southwest Oncology Group study.<br />

J Clin Oncol 1992; 10(8):1284–1291.<br />

10. Bardou VJ, Arpino G, Elledge RM, et al. Progesterone receptor status significantly<br />

improves outcome prediction over estrogen receptor status alone for adjuvant endocrine<br />

therapy in two large breast cancer databases. J Clin Oncol 2003; 21(10): 1973–1979.<br />

11. Jordan VC, Brodie AM. Development and evolution <strong>of</strong> therapies targeted to the<br />

estrogen receptor for the treatment and prevention <strong>of</strong> breast cancer. Steroids 2007;<br />

72(1):7–25.<br />

12. Allred DC, Harvey JM, Berardo M, et al. Prognostic and predictive factors in breast<br />

cancer by immunohistochemical analysis. Mod Pathol 1998; 11(2):155–168.<br />

13. Rhodes A, Jasani B, Barnes DM, et al. Reliability <strong>of</strong> immunohistochemical demonstration<br />

<strong>of</strong> oestrogen receptors in routine practice: interlaboratory variance in the<br />

sensitivity <strong>of</strong> detection and evaluation <strong>of</strong> scoring systems. J Clin Pathol 2000; 53(2):<br />

125–130.<br />

14. Schnitt SJ. Estrogen receptor testing <strong>of</strong> breast cancer in current clinical practice:<br />

what’s the question? J Clin Oncol 2006; 24(12):1797–1799.<br />

15. Paik S, Shak S, Tang G, et al. A multigene assay to predict recurrence <strong>of</strong> tamoxifentreated,<br />

node-negative breast cancer. N Engl J Med 2004; 351(27):2817–2826.<br />

16. Coussens L, Yang-Feng TL, Liao YC, et al. Tyrosine kinase receptor with extensive<br />

homology to EGF receptor shares chromosomal location with neu oncogene. Science<br />

1985; 230(4730):1132–1139.<br />

17. King CR, Kraus MH, Aaronson SA. Amplification <strong>of</strong> a novel v-erbB-related gene in<br />

a human mammary carcinoma. Science 1985; 229(4717):974–976.<br />

18. Yamamoto T, Ikawa S, Akiyama T, et al. Similarity <strong>of</strong> protein encoded by the human<br />

c-erb-B-2 gene to epidermal growth factor receptor. Nature 1986; 319(6050):230–234.<br />

19. Ross JS, Fletcher JA, Bloom KJ, et al. Targeted therapy in breast cancer: the HER-2/<br />

neu gene and protein. Mol Cell Proteomics 2004; 3(4):379–398.<br />

20. Carter P, Presta L, Gorman CM, et al. Humanization <strong>of</strong> an anti-p185HER2 antibody<br />

for human cancer therapy. Proc Natl Acad Sci U S A 1992; 89(10):4285–4289.<br />

21. Hortobagyi GN. Overview <strong>of</strong> treatment results with trastuzumab (Herceptin) in<br />

metastatic breast cancer. Semin Oncol 2001; 28(6 suppl 18):43–47.<br />

22. Piccart-Gebhart MJ, Procter M, Leyland-Jones B, et al. Trastuzumab after adjuvant<br />

chemotherapy in HER2-positive breast cancer. N Engl J Med 2005; 353(16):1659–1672.


Predictive Markers for Targeted <strong>Breast</strong> <strong>Cancer</strong> Treatment 147<br />

23. Vogel CL, Cobleigh MA, Tripathy D, et al. Efficacy and safety <strong>of</strong> trastuzumab as a<br />

single agent in first-line treatment <strong>of</strong> HER2-overexpressing metastatic breast cancer.<br />

J Clin Oncol 2002; 20(3):719–726.<br />

24. Cobleigh MA, Vogel CL, Tripathy D, et al. Multinational study <strong>of</strong> the efficacy and<br />

safety <strong>of</strong> humanized anti-HER2 monoclonal antibody in women who have HER2-<br />

overexpressing metastatic breast cancer that has progressed after chemotherapy for<br />

metastatic disease. J Clin Oncol 1999; 17(9):2639–2648.<br />

25. Slamon DJ, Leyland-Jones B, Shak S, et al. Use <strong>of</strong> chemotherapy plus a monoclonal<br />

antibody against HER2 for metastatic breast cancer that overexpresses HER2.<br />

N Engl J Med 2001; 344(11):783–792.<br />

26. Pegram MD, Lipton A, Hayes DF, et al. <strong>Ph</strong>ase II study <strong>of</strong> receptor-enhanced<br />

chemosensitivity using recombinant humanized anti-p185HER2/neu monoclonal<br />

antibody plus cisplatin in patients with HER2/neu-overexpressing metastatic breast<br />

cancer refractory to chemotherapy treatment. J Clin Oncol 1998; 16(8):2659–2671.<br />

27. Baselga J, Tripathy D, Mendelsohn J, et al. <strong>Ph</strong>ase II study <strong>of</strong> weekly intravenous<br />

recombinant humanized anti-p185HER2 monoclonal antibody in patients with HER2/<br />

neu-overexpressing metastatic breast cancer. J Clin Oncol 1996; 14(3):737–744.<br />

28. Fornier M, Risio M, Van Poznak C, et al. HER2 testing and correlation with efficacy<br />

<strong>of</strong> trastuzumab therapy. Oncology (Williston Park) 2002; 16(10):1340–1342.<br />

29. Press MF, Slamon DJ, Flom KJ, et al. Evaluation <strong>of</strong> HER-2/neu gene amplification and<br />

overexpression: comparison <strong>of</strong> frequently used assay methods in a molecularly characterized<br />

cohort <strong>of</strong> breast cancer specimens. J Clin Oncol 2002; 20(14): 3095–3105.<br />

30. Lebeau A, Deimling D, Kaltz C, et al. Her-2/neu analysis in archival tissue samples<br />

<strong>of</strong> human breast cancer: comparison <strong>of</strong> immunohistochemistry and fluorescence in<br />

situ hybridization. J Clin Oncol 2001; 19(2):354–363.<br />

31. Wang S, Saboorian MH, Frenkel E, et al. Laboratory assessment <strong>of</strong> the status <strong>of</strong><br />

Her-2/neu protein and oncogene in breast cancer specimens: comparison <strong>of</strong> immunohistochemistry<br />

assay with fluorescence in situ hybridisation assays. J Clin Pathol<br />

2000; 53(5):374–381.<br />

32. Moasser MM. Targeting the function <strong>of</strong> the HER2 oncogene in human cancer<br />

therapeutics. Oncogene 2007; 26(46):6577–6592.<br />

33. Bartlett JM. <strong>Ph</strong>armacodiagnostic testing in breast cancer: focus on HER2 and trastuzumab<br />

therapy. Am J <strong>Ph</strong>armacogenomics 2005; 5(5):303–315.<br />

34. Scaltriti M, Baselga J. The epidermal growth factor receptor pathway: a model for<br />

targeted therapy. Clin <strong>Cancer</strong> Res 2006; 12(18):5268–5272.<br />

35. Chan S, Scheulen ME, Johnston S, et al. <strong>Ph</strong>ase II study <strong>of</strong> temsirolimus (CCI-779), a<br />

novel inhibitor <strong>of</strong> mTOR, in heavily pretreated patients with locally advanced or<br />

metastatic breast cancer. J Clin Oncol 2005; 23(23):5314–5322.<br />

36. Schneider BP, Sledge GW, Jr. Drug insight: VEGF as a therapeutic target for breast<br />

cancer. Nat Clin Pract Oncol 2007; 4(3):181–189.<br />

37. Tsutsui S, Kataoka A, Ohno S, et al. Prognostic and predictive value <strong>of</strong> epidermal<br />

growth factor receptor in recurrent breast cancer. Clin <strong>Cancer</strong> Res 2002; 8(11):<br />

3454–3460.<br />

38. Johnston JB, Navaratnam S, Pitz MW, et al. Targeting the EGFR pathway for cancer<br />

therapy. Curr Med Chem 2006; 13(29):3483–3492.<br />

39. Moulder SL, Arteaga CL. A <strong>Ph</strong>ase I/II Trial <strong>of</strong> trastuzumab and gefitinib in patients<br />

with Metastatic <strong>Breast</strong> <strong>Cancer</strong> that overexpresses HER2/neu (ErbB-2). Clin <strong>Breast</strong><br />

<strong>Cancer</strong> 2003; 4(2):142–145.


148 Pedersen and Bartlett<br />

40. Tan AR, Yang X, Hewitt SM, et al. Evaluation <strong>of</strong> biologic end points and pharmacokinetics<br />

in patients with metastatic breast cancer after treatment with erlotinib, an<br />

epidermal growth factor receptor tyrosine kinase inhibitor. J Clin Oncol 2004; 22(15):<br />

3080–3090.<br />

41. Baselga J, Albanell J, Ruiz A, et al. <strong>Ph</strong>ase II and tumor pharmacodynamic study <strong>of</strong><br />

gefitinib in patients with advanced breast cancer. J Clin Oncol 2005; 23(23):5323–5333.<br />

42. Fountzilas G, Pectasides D, Kalogera-Fountzila A, et al. Paclitaxel and carboplatin as<br />

first-line chemotherapy combined with gefitinib (IRESSA) in patients with advanced<br />

breast cancer: a phase I/II study conducted by the Hellenic Cooperative Oncology<br />

Group. <strong>Breast</strong> <strong>Cancer</strong> Res Treat 2005; 92(1):1–9.<br />

43. Gasparini G, Sarmiento R, Amici S, et al. Gefitinib (ZD1839) combined with weekly<br />

epirubicin in patients with metastatic breast cancer: a phase I study with biological<br />

correlate. Ann Oncol 2005; 16(12):1867–1873.<br />

44. Polychronis A, Sinnett HD, Hadjiminas D, et al. Preoperative gefitinib versus<br />

gefitinib and anastrozole in postmenopausal patients with oestrogen-receptor positive<br />

and epidermal-growth-factor-receptor-positive primary breast cancer: a double-blind<br />

placebo-controlled phase II randomised trial. Lancet Oncol 2005; 6(6):383–391.<br />

45. von Minckwitz G, Jonat W, Fasching P, et al. A multicentre phase II study on<br />

gefitinib in taxane- and anthracycline-pretreated metastatic breast cancer. <strong>Breast</strong><br />

<strong>Cancer</strong> Res Treat 2005; 89(2):165–172.<br />

46. Ciardiello F, Troiani T, Caputo F, et al. <strong>Ph</strong>ase II study <strong>of</strong> gefitinib in combination<br />

with docetaxel as first-line therapy in metastatic breast cancer. Br J <strong>Cancer</strong> 2006;<br />

94(11):1604–1609.<br />

47. Modi S, D’Andrea G, Norton L, et al. A phase I study <strong>of</strong> cetuximab/paclitaxel in<br />

patients with advanced-stage breast cancer. Clin <strong>Breast</strong> <strong>Cancer</strong> 2006; 7(3):270–277.<br />

48. Chung KY, Shia J, Kemeny NE, et al. Cetuximab shows activity in colorectal cancer<br />

patients with tumors that do not express the epidermal growth factor receptor by<br />

immunohistochemistry. J Clin Oncol 2005; 23(9):1803–1810.<br />

49. Vincenzi B, Santini D, Rabitti C, et al. Cetuximab and irinotecan as third-line<br />

therapy in advanced colorectal cancer patients: a single centre phase II trial. Br<br />

J <strong>Cancer</strong> 2006; 94(6):792–797.<br />

50. Campiglio M, Locatelli A, Olgiati C, et al. Inhibition <strong>of</strong> proliferation and induction<br />

<strong>of</strong> apoptosis in breast cancer cells by the epidermal growth factor receptor (EGFR)<br />

tyrosine kinase inhibitor ZD1839 (‘Iressa’) is independent <strong>of</strong> EGFR expression level.<br />

J Cell <strong>Ph</strong>ysiol 2004; 198(2):259–268.<br />

51. Uramoto H, Mitsudomi T. Which biomarker predicts benefit from EGFR-TKI<br />

treatment for patients with lung cancer? Br J <strong>Cancer</strong> 2007; 96(6):857–863.<br />

52. Lynch TJ, Bell DW, Sordella R, et al. Activating mutations in the epidermal growth<br />

factor receptor underlying responsiveness <strong>of</strong> non-small-cell lung cancer to gefitinib.<br />

N Engl J Med 2004; 350(21):2129–2139.<br />

53. Yun CH, Boggon TJ, Li Y, et al. Structures <strong>of</strong> lung cancer-derived EGFR mutants<br />

and inhibitor complexes: mechanism <strong>of</strong> activation and insights into differential<br />

inhibitor sensitivity. <strong>Cancer</strong> Cell 2007; 11(3):217–227.<br />

54. Benvenuti S, Sartore-Bianchi A, Di Nicolantonio F, et al. Oncogenic activation <strong>of</strong><br />

the RAS/RAF signaling pathway impairs the response <strong>of</strong> metastatic colorectal cancers<br />

to anti-epidermal growth factor receptor antibody therapies. <strong>Cancer</strong> Res 2007;<br />

67(6):2643–2648.


Predictive Markers for Targeted <strong>Breast</strong> <strong>Cancer</strong> Treatment 149<br />

55. Pao W, Wang TY, Riely GJ, et al. KRAS mutations and primary resistance <strong>of</strong> lung<br />

adenocarcinomas to gefitinib or erlotinib. PLoS Med 2005; 2(1):E17.<br />

56. Bartlett JM, Forsyth A. Detection <strong>of</strong> HER2 gene amplification by fluorescence in<br />

situ hybridization in breast cancer. Methods Mol Med 2006; 120:309–322.<br />

57. Pepe MS, Etzioni R, Feng Z, et al. <strong>Ph</strong>ases <strong>of</strong> biomarker development for early<br />

detection <strong>of</strong> cancer. J Natl <strong>Cancer</strong> Inst 2001; 93(14):1054–1061.<br />

58. Hayes DF, Bast RC, Desch CE, et al. Tumor marker utility grading system:<br />

a framework to evaluate clinical utility <strong>of</strong> tumor markers. J Natl <strong>Cancer</strong> Inst 1996;<br />

88(20):1456–1466.<br />

59. Sweep FC, Fritsche HA, Gion M, et al. Considerations on development, validation,<br />

application, and quality control <strong>of</strong> immuno(metric) biomarker assays in clinical<br />

cancer research: an EORTC-NCI working group report. Int J Oncol 2003; 23(6):<br />

1715–1726.<br />

60. Bartlett J, Mallon E, Cooke T. The clinical evaluation <strong>of</strong> HER-2 status: which test to<br />

use? J Pathol 2003; 199(4):411–417.<br />

61. Bartlett JM, Going JJ, Mallon EA, et al. Evaluating HER2 amplification and overexpression<br />

in breast cancer. J Pathol 2001; 195(4):422–428.<br />

62. Rimm DL. What brown cannot do for you. Nat Biotechnol 2006; 24(8):914–916.<br />

63. Soderberg O, Gullberg M, Jarvius M, et al. Direct observation <strong>of</strong> individual endogenous<br />

protein complexes in situ by proximity ligation. Nat Methods 2006; 3(12):995–1000.


11<br />

Circulating Tumor Cells in Individualizing<br />

<strong>Breast</strong> <strong>Cancer</strong> Therapy<br />

James M. Reuben<br />

Department <strong>of</strong> Hematopathology, The University <strong>of</strong> Texas<br />

M. D. Anderson <strong>Cancer</strong> Center, Houston, Texas, U.S.A.<br />

Massimo Crist<strong>of</strong>anilli<br />

Department <strong>of</strong> <strong>Breast</strong> Medical Oncology, The University <strong>of</strong> Texas<br />

M. D. Anderson <strong>Cancer</strong> Center, Houston, Texas, U.S.A.<br />

INTRODUCTION<br />

<strong>Breast</strong> cancer is the most frequently diagnosed nonskin cancer among women<br />

in the United States and is second only to lung cancer in causing cancer-related<br />

deaths among them (1). The vast majority <strong>of</strong> deaths are due to recurrent<br />

metastatic disease. Occult dissemination <strong>of</strong> tumor cells is the main cause <strong>of</strong><br />

recurrent metastatic breast cancer (MBC) in patients who have undergone<br />

resection <strong>of</strong> their primary tumor (2). Approximately 5% <strong>of</strong> patients with breast<br />

cancer have clinically detectable metastases at the time <strong>of</strong> initial diagnosis, and<br />

a further 30% to 40% <strong>of</strong> patients who appear clinically free <strong>of</strong> metastases<br />

harbor occult metastases (3,4). The formation <strong>of</strong> metastatic colonies is a<br />

continuous process, commencing early during the growth <strong>of</strong> the primary tumor.<br />

Metastasis occurs through a cascade <strong>of</strong> linked sequential steps involving<br />

multiple host-tumor interactions. This complex process requires the cells to<br />

enter the circulation, arrest at the distant vascular bed, extravasate into the<br />

organ interstitium and parenchyma, and proliferate as a secondary colony.<br />

151


152 Reuben and Crist<strong>of</strong>anilli<br />

Several experimental studies suggest that during each stage <strong>of</strong> the process, only<br />

the fittest tumor cells survive (2).<br />

The first report on tumor cells in the peripheral circulation was attributed<br />

to Ashworth in 1869 (5). Since then, the existence, origin, and clinical significance<br />

<strong>of</strong> circulating tumor cells (CTCs) have been debated. The introduction <strong>of</strong><br />

sensitive and specific immunohistochemical techniques in the late 1970s led to<br />

renewed interest in the detection <strong>of</strong> CTCs and their possible association with<br />

minimal residual disease in solid malignancies (6,7). However, there is little<br />

understanding <strong>of</strong> the genetic and phenotypic characteristics <strong>of</strong> CTCs in peripheral<br />

blood. It has been speculated that mammary stem cells represent the cellular<br />

origins <strong>of</strong> cancer because they exist quiescently over long periods and can<br />

accumulate multiple mutations over the lifetime, ultimately giving rise to tumors<br />

when stimulated to proliferate (8–10). It was recently reported that highly<br />

tumorigenic cells possessing properties consistent with those <strong>of</strong> stem or progenitor<br />

cells could be isolated from human breast cancers (9). Furthermore,<br />

transplantation <strong>of</strong> as many as several hundred <strong>of</strong> these cells into the cleared<br />

mammary fat pads <strong>of</strong> etoposide-treated mice resulted in the growth <strong>of</strong> breast<br />

tumors (10). <strong>Cancer</strong> stem cells not only exist in human breast cancer but also<br />

have different biologic features than the more differentiated cells that constitute<br />

the majority <strong>of</strong> cells in human breast cancers. These cancer stem cells may be<br />

detectable in the peripheral blood and bone marrow <strong>of</strong> patients with MBC and<br />

may account for their worse prognosis. Hence the development <strong>of</strong> therapies<br />

specifically targeting cancer stem cells may have great benefits for patients with<br />

CTCs who possess the properties <strong>of</strong> cancer stem cells.<br />

When CTCs migrate to distal sites, they can potentially give rise to<br />

microscopic lesions that result in metastasis.Microscopic disease may be present in<br />

lymph nodes, bone marrow (primary breast cancer), and peripheral blood (metastatic<br />

disease) at the time <strong>of</strong> breast cancer diagnosis (3,4,11,12). Most studies have<br />

demonstrated that the detection <strong>of</strong> microscopic disease in patients with breast<br />

cancer contributes prognostic information and, in selected cases, can predict the<br />

efficacy <strong>of</strong> treatments (12). In patients with primary breast cancer, the detection <strong>of</strong><br />

microscopic disease in lymph nodes and bone marrow has led to a better understanding<br />

<strong>of</strong> the role <strong>of</strong> minimal residual disease in establishing metastasis.<br />

Over the past few years, immunomagnetic separation technology, with its<br />

higher level <strong>of</strong> sensitivity and specificity, has been used to improve the detection<br />

<strong>of</strong> CTCs compared with the detection <strong>of</strong> occult CTCs by reverse transcriptasepolymerase<br />

chain reaction (RT-PCR) (13–21). In this chapter, we review the<br />

methods <strong>of</strong> detecting CTCs, the prognostic implications <strong>of</strong> CTCs, and the<br />

potential means <strong>of</strong> exploiting CTCs in developing individualized therapy.<br />

DETECTION OF CIRCULATING TUMOR CELLS<br />

Advances in technology have facilitated the detection <strong>of</strong> even very small numbers<br />

<strong>of</strong> CTCs in the peripheral blood <strong>of</strong> cancer patients. The principle <strong>of</strong> these<br />

methods is based on the fact that breast cancer cells express epithelial cell


Circulating Tumor Cells in Individualizing <strong>Breast</strong> <strong>Cancer</strong> Therapy 153<br />

adhesion molecule-1 (EpCAM). EpCAM, a 40-kDa glycoprotein, is present on<br />

most epithelial carcinomas, and recent studies indicate that this molecule has a<br />

major morphoregulatory function, relevant not only to epithelial tissue development<br />

but also to carcinogenesis and tumor progression (22,23).<br />

Detection <strong>of</strong> CTCs by the CellSearch TM System<br />

The only assay approved by the U.S. Food and Drug Administration is the<br />

CellSearch TM system (Veridex, LLC, Warren, New Jersey, U.S.). With this assay,<br />

a small sample <strong>of</strong> peripheral blood is allowed to react with ferr<strong>of</strong>luids coated with<br />

antibodies to EpCAM. Next, the sample is processed through a magnetic field to<br />

retain the EpCAM-positive cells, while the EpCAM-negative cells, primarily <strong>of</strong><br />

hematopoietic origin, are eluted from the column and discarded. Thereafter, the<br />

enriched EpCAM-positive cells are allowed to react with the nucleic acid dye<br />

4 0 ,6-doamidino-2-phenylindole (DAPI), monoclonal antibodies specific for leukocytes<br />

(anti-CD45 antibodies conjugated with allophycocyanin), and epithelial<br />

cells (phycoerythrin-conjugated antibodies to cytokeratin 8, 18, and 19) and<br />

analyzed by the CellSpotter TM Analyzer (Veridex). The CellSpotter Analyzer is a<br />

semiautomated fluorescence-based microscopy system that enables computergenerated<br />

reconstruction <strong>of</strong> cellular images. The analyzer interrogates each cell<br />

image to determine if it meets the very stringent requirements <strong>of</strong> an algorithm to<br />

be classified as a CTC. The algorithm insures that a CTC expresses EpCAM and<br />

not leukocyte lineage-specific antigens (represented by CD45), exhibits cytoplasmic<br />

expression <strong>of</strong> cytokeratin, and contains a nucleus that binds DAPI.<br />

Absence <strong>of</strong> any <strong>of</strong> these characteristics disqualifies a cell image as a CTC. In<br />

identifying CTCs, the procedure identifies cell objects that possess some, but not<br />

all, <strong>of</strong> the required characteristics to be a CTC and are labeled as “unclassified”<br />

objects (24). The stringent criteria <strong>of</strong> CellSearch for identifying a CTC are<br />

laudable; however, the merits <strong>of</strong> the assay continue to be debated as it minimizes<br />

the scientific or clinical significance <strong>of</strong> unclassified objects.<br />

Although CellSearch is a very reproducible and reliable method for the<br />

identification and enumeration <strong>of</strong> CTCs in the peripheral blood <strong>of</strong> patients with<br />

breast cancer, its clinical utility has been limited by its high cost. The semiautomated<br />

assay is also somewhat labor intensive. In addition, the need to permeabilize<br />

the cell membrane <strong>of</strong> a viable cell to introduce DAPI and anticytokeratin<br />

antibodies to label intracellular structures further deters the possibility <strong>of</strong> interrogating<br />

viable cells for their growth potential, colony formation, and genomics,<br />

among others. To address these issues, the CellPr<strong>of</strong>ile TM kit was introduced that<br />

permits the retrieval <strong>of</strong> EpCAM-positive cells prior to cell permeabilization and<br />

interrogation for genomic or proteomic pr<strong>of</strong>iles <strong>of</strong> the CTCs.<br />

Isolation <strong>of</strong> CTCs by the AutoMACS TM System<br />

EpCAM, also known as human epithelial cell antigen (HEA) or CD326, is<br />

expressed on the majority <strong>of</strong> tumor cells <strong>of</strong> epithelial cell origin but not on


154 Reuben and Crist<strong>of</strong>anilli<br />

circulating B cells, T cells, or monocytes. In another assay that exploits this<br />

characteristic <strong>of</strong> CTCs, peripheral blood mononuclear cells are incubated with<br />

magnetic beads coated with anti-CD326 (Miltenyi Biotec, Inc., Auburn, California,<br />

U.S.) prior to passage through a magnetic column to enrich for CTCs. Typically,<br />

mononuclear cells are isolated from peripheral blood, bone marrow, pleural<br />

fluid, paracentesis fluid, or other tissues, mixed with anti-CD326 microbeads, and<br />

incubated at 48C to88C for 15 minutes. After washing, the cell pellet is resuspended<br />

and loaded onto the magnetic column <strong>of</strong> an automatic magnetic cell<br />

sorting (AutoMACS TM ) system (Miltenyi Biotec) (25). CD326-positive CTCs are<br />

isolated using the positive selection protocol and can then be centrifuged onto<br />

glass slides to determine the morphology, viability, and purity <strong>of</strong> the preparation.<br />

In one study, Papanicolaou (Pap)-stained cytospin smears were analyzed in<br />

detail to identify tumor cells on the basis <strong>of</strong> conventional cytologic criteria. In<br />

cytospin slides <strong>of</strong> enriched CTCs, large cells with cytomorphological features<br />

suspicious for tumor cells were detected, and subsequently the cytospin slides<br />

were immunocytochemically stained with the avidin-biotin complex peroxidase<br />

technique using pancytokeratin antibody for the identification <strong>of</strong> tumor cells<br />

(Fig. 1A). (The immunostains <strong>of</strong> CTCs were performed and interpreted by<br />

Dr. Savitri Krishnamurthy, Department <strong>of</strong> Pathology, The University <strong>of</strong> Texas<br />

M. D. Anderson <strong>Cancer</strong> Center).<br />

A second cytospin slide <strong>of</strong> CD326-positive bone marrow tumor cells from<br />

the same patient was stained with a cocktail <strong>of</strong> anticytokeratin antibodies using<br />

the avidin-biotin complex peroxidase technique for the identification <strong>of</strong> tumor<br />

cells (Fig. 1B). In the majority <strong>of</strong> cases, the suspicious cells were positive for<br />

cytokeratin, thereby confirming the presence <strong>of</strong> micrometastatic carcinoma.<br />

In addition to bone marrow, it is possible to isolate CTCs from the<br />

paracentesis fluid <strong>of</strong> a patient with MBC. Figure 2 shows a cluster <strong>of</strong> CTCs<br />

Figure 1A A cytospin slide <strong>of</strong> CTCs from the bone marrow <strong>of</strong> a patient with primary<br />

breast cancer using anti-CD326 antibody–coated microbeads and reacted with Pap stain.<br />

The figure shows a cluster <strong>of</strong> CTCs with cellular morphology consistent with that <strong>of</strong><br />

epithelial cells. Abbreviations: CTCs, circulating tumor cells; Pap, Papanicolaou.


Circulating Tumor Cells in Individualizing <strong>Breast</strong> <strong>Cancer</strong> Therapy 155<br />

Figure 1B CD326-positive CTCs are cytokeratin-positive epithelial cells. Abbreviation:<br />

CTCs, circulating tumor cells.<br />

Figure 2 CD326-positive CTCs from the paracentesis fluid <strong>of</strong> a patient with MBC.<br />

Abbreviations: CTCs, circulating tumor cells; MBC, metastatic breast cancer.<br />

isolated from the paracentesis fluid <strong>of</strong> a patient with MBC that were reacted with<br />

Pap stain.<br />

Isolation <strong>of</strong> CTCs by the EasySep TM System<br />

Another immunomagnetic cell selection system is EasySep TM (StemCell Technologies,<br />

Inc., Vancouver, Canada), which combines the specificity <strong>of</strong> monoclonal<br />

antibodies with the simplicity <strong>of</strong> a column-free magnetic system. In this<br />

procedure, highly enriched cells are obtained by positively selecting the cells <strong>of</strong><br />

interest with antibodies to a specific cell-surface antigen. The cells targeted for


156 Reuben and Crist<strong>of</strong>anilli<br />

selection are cross-linked to EasySep nanoparticles via the formation <strong>of</strong> tetrameric<br />

antibody complexes (StemCell Technologies) in a standard centrifuge<br />

tube, which is then placed in the EasySep magnetic chamber. The handheld<br />

magnetic chamber is gently inverted to discard the cells that are not bound to<br />

the specific nanoparticle complexes that adhere to the walls <strong>of</strong> the tube in the<br />

magnetic chamber. To obtain a highly pure population, the residual cells in the<br />

tube are rinsed twice and then harvested by removing the tube from the chamber.<br />

Enriched CTCs can then be used for further interrogation for the differential<br />

expression <strong>of</strong> EpCAM or other markers <strong>of</strong> interest.<br />

Positive selection is most effective when a specimen has a generous<br />

complement <strong>of</strong> the desired cell population. When a specimen, such as peripheral<br />

blood, contains few or inadequate numbers <strong>of</strong> the cells <strong>of</strong> interest, a negativeselection<br />

approach can be taken to enrich for these cells. One negative-selection<br />

technique is the RosetteSep TM (StemCell Technologies) method that uses a<br />

highly purified combination <strong>of</strong> mouse and rat monoclonal antibodies. When<br />

reacted with peripheral whole blood, these antibodies bind in bispecific tetrameric<br />

antibody complexes directed against cell surface antigens on human<br />

hematopoietic cells (CD2, CD16, CD19, CD36, CD38, CD45, and CD66b) and<br />

glycophorin A on red blood cells, or erythrocytes. The unwanted cells <strong>of</strong> hematopoietic<br />

origin form rosettes by cross-linking to multiple erythrocytes that can<br />

be easily removed by performing a typical Ficoll-Paque TM (GE Amersham-<br />

<strong>Ph</strong>armacia Diagnostics, Uppsala, Sweden) density gradient procedure for the<br />

separation <strong>of</strong> mononuclear cells from peripheral blood. The rosettes, free<br />

erythrocytes, and granulocytes form a pellet, while the unlabeled, desired CTCs<br />

are collected from the interface between the plasma and the buoyant density<br />

medium. Figures 3A and 3B represent cytospin slides <strong>of</strong> CTCs isolated from the<br />

Figure 3A Pap stain <strong>of</strong> a cluster <strong>of</strong> CTCs isolated from the peripheral blood <strong>of</strong> a patient<br />

with MBC using RosetteSep TM . Abbreviations: Pap, Papanicolaou; CTCs, circulating<br />

tumor cells; MBC, metastatic breast cancer.


Circulating Tumor Cells in Individualizing <strong>Breast</strong> <strong>Cancer</strong> Therapy 157<br />

Figure 3B Cytokeratin-positive CTCs from the peripheral blood <strong>of</strong> a patient with MBC<br />

using RosetteSep TM . Abbreviations: CTCs, circulating tumor cells; MBC, metastatic<br />

breast cancer.<br />

peripheral blood <strong>of</strong> a patient with MBC that were reacted with Pap stain<br />

(Fig. 3A) and anticytokeratin antibodies (Fig. 3A).<br />

Detection <strong>of</strong> CTCs by AdnaGen’s Technology<br />

In addition to magnetic separation methodologies, PCR greatly facilitates the<br />

detection <strong>of</strong> occult tumor cells through the use <strong>of</strong> nucleic acid analysis. The<br />

PCR-based assay, Adna Test <strong>Breast</strong> <strong>Cancer</strong> Select, developed by AdnaGen AG,<br />

Langenhagen, Germany, makes use <strong>of</strong> RT-PCR to identify putative transcripts <strong>of</strong><br />

genes in EpCAM-positive cells that are isolated by a magnetic separation<br />

method. This approach takes advantage <strong>of</strong> the prevailing state <strong>of</strong> the art <strong>of</strong> CTC<br />

isolation and enhances the chance <strong>of</strong> detecting very low numbers <strong>of</strong> CTCs<br />

or occult CTCs on the basis <strong>of</strong> their expression <strong>of</strong> tumor-associated genes.<br />

AdnaGen’s two-step “combination-<strong>of</strong>-combinations principle” involves initially<br />

the enrichment <strong>of</strong> CTCs using an antibody mix linked to magnetic particles (26).<br />

Thereafter, mRNA is extracted from the CTCs and transcribed into cDNA before<br />

being subjected to multiplex RT-PCR for the detection <strong>of</strong> tumor-associated<br />

genes. Owing to the combination <strong>of</strong> different selection and tumor markers, both<br />

the heterogeneity <strong>of</strong> the tumor cells and the possible therapy-induced deviations<br />

in the expression patterns are taken into account. The AdnaGen system has high<br />

analytical sensitivity (two cells in 5 mL <strong>of</strong> blood) and high specificity (>90%)<br />

achieved through the combination <strong>of</strong> multimarker tumor cell enrichment and<br />

multiplex gene expression pr<strong>of</strong>iling.


158 Reuben and Crist<strong>of</strong>anilli<br />

CLINICAL SIGNIFICANCE AND PROGNOSTIC VALUE OF CTCs in MBC<br />

Tumor cells can be detected in the locoregional lymph nodes <strong>of</strong> women with<br />

early breast cancer, and the presence <strong>of</strong> these cells has been shown to have a<br />

negative effect on long-term prognosis (3,12). Despite evidence <strong>of</strong> the prognostic<br />

value <strong>of</strong> CTCs in some studies, the detection <strong>of</strong> micrometastases was never<br />

incorporated into cancer-staging protocols or considered a valuable clinical tool.<br />

This may be the result <strong>of</strong> a combination <strong>of</strong> factors, such as variable antigen<br />

expression in poorly differentiated tumors and reports <strong>of</strong> cytokeratin and epithelial<br />

membrane antigen positivity in cells that are not <strong>of</strong> epithelial origin,<br />

which demonstrated a need for more sensitive and specific methods <strong>of</strong> detection<br />

than were available at the time. This need was filled to some degree by sensitive<br />

PCR techniques (14,16,17). In the past decade, a few studies have shown that the<br />

detection <strong>of</strong> occult disease by PCR has prognostic significance in some solid<br />

tumors (13). However, PCR-based assays for the detection <strong>of</strong> occult tumor cells<br />

have limitations, particularly contamination <strong>of</strong> samples, specificity <strong>of</strong> the assays,<br />

and inability to quantify tumor cells. Moreover, PCR-based methods cannot be<br />

used to perform functional assays. These factors have precluded the widespread<br />

use <strong>of</strong> PCR in in vitro diagnostic applications.<br />

Using an immunomagnetic detection approach, Austrup et al. reported the<br />

prognostic significance <strong>of</strong> genomic alterations (e.g., c-erbB-2 overexpression)<br />

present in CTCs purified from the blood <strong>of</strong> patients with breast cancer (27). The<br />

authors investigated genomic imbalances, such as mutation, amplification, and<br />

loss <strong>of</strong> heterozygosity, <strong>of</strong> 13 tumor suppressor genes and 2 proto-oncogenes<br />

using DNA isolated from minimal residual cancer cells. The presence and the<br />

number <strong>of</strong> genomic imbalances measured in disseminated tumor cells were<br />

significantly associated with a worse prognosis (27).<br />

Subsequently, Meng et al. demonstrated that CTCs recapitulate the human<br />

epidermal growth factor receptor 2 (HER2) status <strong>of</strong> the primary tumor (28).<br />

Furthermore, the authors demonstrated that a fraction <strong>of</strong> patients with HER2-<br />

negative primary tumors had detectable HER2 gene amplification in their CTCs,<br />

suggesting acquisition or selection <strong>of</strong> this phenotype during cancer progression<br />

(37.5%; 95% confidence interval, 18.8–59.4%). Intriguingly, the initiation <strong>of</strong><br />

trastuzumab-based therapy in a few cases with altered HER2 status in CTCs was<br />

associated with a clinical response (29). These important observations support<br />

the argument that detection <strong>of</strong> CTCs and determination <strong>of</strong> their gene amplification<br />

or expression can be used for better tailoring <strong>of</strong> therapies.<br />

The predictive and prognostic roles <strong>of</strong> CTCs were investigated in a prospective<br />

multicenter clinical trial led by researchers at the University <strong>of</strong> Texas<br />

M. D. Anderson <strong>Cancer</strong> Center (11). In this study, the CellSearch system was<br />

used to prospectively determine the prognostic and predictive value <strong>of</strong> CTCs in<br />

patients with MBC who were about to start a new systemic treatment. The 177<br />

enrolled patients underwent peripheral blood collection at monthly intervals for<br />

up to six months after enrollment. A cut<strong>of</strong>f <strong>of</strong> 5 CTCs/7.5 mL was used to


Circulating Tumor Cells in Individualizing <strong>Breast</strong> <strong>Cancer</strong> Therapy 159<br />

stratify patients into positive and negative groups (positive, 5 CTCs/7.5 mL;<br />

negative, 18 months; P <<br />

0.0001). On multivariate Cox hazards regression analysis, CTC levels, both at<br />

baseline and at first follow-up, were the most significant predictors <strong>of</strong> progression-free<br />

and overall survival (11).<br />

An analysis restricted to patients who were about to start first-line systemic<br />

therapy after the diagnosis <strong>of</strong> MBC had similar findings (30). A subsequent<br />

investigation including patients with either measurable or evaluable MBC confirmed<br />

that CTCs are an independent predictor <strong>of</strong> survival that is more powerful<br />

than standard prognostic measures, such as hormone receptor status, and measures<br />

<strong>of</strong> tumor burden (e.g., the Swenerton score or level <strong>of</strong> CA27-29) (31).<br />

The detection <strong>of</strong> changes in CTC status may also have predictive utility.<br />

The CTC detection rate at first follow-up was lower than at baseline, particularly<br />

in patients undergoing first-line treatment (25% versus 52%) and those<br />

with visceral disease (28% versus 50%) (11). In patients with newly diagnosed<br />

disease treated with chemotherapy alone (n ¼ 37), the percentage <strong>of</strong> patients<br />

who were CTC positive decreased significantly from baseline to first follow-up<br />

(57% to 37%, P ¼ 0.001) and first imaging-visit (at eight to nine weeks) blood<br />

draw (57% to 31%, P ¼ 0.004) (31). More importantly, in patients receiving<br />

trastuzumab-containing regimens, CTC-positive patients decreased from<br />

baseline to first follow-up (59% to 7%, P ¼ 0.600) and first imaging-visit<br />

blood draw (59% to 0%, P ¼ not applicable). This data indicates that patients<br />

with newly diagnosed disease who are about to start first-line therapy can have<br />

detectable CTCs and that the changes in CTC status at three to four weeks may<br />

indicate a benefit from systemic treatment (particularly trastuzumab-based<br />

regimens). Moreover, the determination <strong>of</strong> CTCs at baseline and follow-up<br />

appears to be a superior prognostic marker in patients with measurable MBC<br />

compared with standard imaging assessments (32). CTCs could be used for<br />

assessment <strong>of</strong> the clinical benefit <strong>of</strong> systemic treatments, for prognostic<br />

stratification, and for evaluation <strong>of</strong> patients with nonmeasurable disease in<br />

future prospective validation trials.<br />

STAGING MBC USING CTCs<br />

The American Joint Committee on <strong>Cancer</strong> (AJCC) classification system, based<br />

on a schema developed by the International Union Against <strong>Cancer</strong> (UICC)<br />

(33–35), includes much <strong>of</strong> the traditional prognostic information used by clinicians<br />

when developing a comprehensive treatment plan. In brief, the AJCC


160 Reuben and Crist<strong>of</strong>anilli<br />

system attempts to define the disease by incorporating all aspects <strong>of</strong> cancer<br />

distribution in terms <strong>of</strong> the primary tumor (T), lymph nodes (N), and distant<br />

metastasis (M). A “stage” group (0, I, II, III, or IV) is then assigned on the basis<br />

<strong>of</strong> the possible TNM permutations, with 0 reflecting minimal involvement and<br />

IV either the maximum tumor involvement or distant metastasis. This basic<br />

TNM staging is then further subdivided according to the time the evaluation is<br />

performed: c ¼ clinical (before surgical treatment) and p ¼ pathologic (after<br />

surgical specimen analysis) (10,35).<br />

Over the past 45 years, the AJCC has regularly updated its staging standards<br />

to incorporate advances in prognostic technology. The committee’s current<br />

work concerns the development <strong>of</strong> prognostic indices based on molecular<br />

markers. However, until these changes are incorporated, TNM staging continues<br />

to quantify only the physical extent <strong>of</strong> the disease. Although it covers the<br />

approximately 2% <strong>of</strong> women who present with metastatic disease (stage IV), the<br />

prognostic information is applicable only to in situ, local, and regional primary<br />

breast cancers. Given the heterogeneity <strong>of</strong> the disease in primary and metastatic<br />

settings, the potential for continued mutation, and the variety <strong>of</strong> treatment<br />

options available, the information acquired at the time <strong>of</strong> the initial diagnosis <strong>of</strong><br />

breast cancer may not be as relevant to planning the treatment for recurrence in<br />

the approximately 30% <strong>of</strong> women who have recurrent MBC years after their<br />

initial diagnosis.<br />

The recent demonstration that the presence <strong>of</strong> CTCs predicts the prognosis<br />

<strong>of</strong> two subgroups <strong>of</strong> patients with MBC raises the possibility that this method<br />

will allow for a true “biologic staging” <strong>of</strong> breast cancer (e.g., stages IVA and<br />

IVB) (11). The main limitations <strong>of</strong> the previous study were the sample size (only<br />

177 patients); the inclusion only <strong>of</strong> patients with measurable disease, which does<br />

not reflect the heterogeneity <strong>of</strong> MBC; and the inclusion <strong>of</strong> patients at different<br />

points in treatment—those with newly diagnosed disease and those undergoing<br />

second- or third-line treatment. To overcome these limitations, a larger international<br />

validation trial, the International Stage IV Stratification Study, was<br />

begun. The objective <strong>of</strong> this study was to confirm the prognostic value <strong>of</strong> CTC<br />

detection in a larger and more homogeneous cohort <strong>of</strong> patients with newly<br />

diagnosed MBC. The study aimed at enrolling 660 patients and providing a more<br />

detailed analysis <strong>of</strong> the association between CTC detection and other factors<br />

such as ethnicity (black compared with white and Hispanic groups) and specific<br />

disease sites (visceral versus nonvisceral disease).<br />

CLINICAL APPLICATION OF CTCs<br />

Despite years <strong>of</strong> clinical research, the odds <strong>of</strong> patients with MBC achieving a<br />

complete response remain extremely low. Thus, the main goal in the management<br />

<strong>of</strong> MBC is palliation (36). Only a few patients who experience a complete<br />

response after chemotherapy remain in this state for prolonged periods, although


Circulating Tumor Cells in Individualizing <strong>Breast</strong> <strong>Cancer</strong> Therapy 161<br />

some have remained in remission for more than 20 years (37). These long-term<br />

survivors are usually young, have an excellent performance status, and, more<br />

importantly, have limited metastatic disease (37,38). Most patients with MBC<br />

respond only transiently to conventional therapies and develop evidence <strong>of</strong><br />

progressive disease within 12 to 24 months <strong>of</strong> starting treatment. For these<br />

patients, systemic treatment does not result in a significant improvement in<br />

survival time but may improve their quality <strong>of</strong> life.<br />

At present, clinicians use three different systemic treatment modalities for<br />

advanced breast cancer: endocrine therapy, chemotherapy, and biologic targeted<br />

therapy (39–42). Appropriate selection <strong>of</strong> patients for these modalities is based<br />

mostly on tissue assessment <strong>of</strong> hormone receptor status (estrogen and progesterone<br />

receptors) and c-erbB-2 status. In patients who lack expression <strong>of</strong> hormone<br />

receptors or who demonstrate no amplification <strong>of</strong> c-erbB-2, cytotoxic<br />

chemotherapy is used. However, no standard therapeutic regimen has been<br />

defined. For example, the optimal schedule <strong>of</strong> chemotherapy administration in<br />

MBC (i.e., concurrent vs. sequential) remains controversial, and the decision<br />

must be individualized. Sledge et al. addressed this issue in a prospective study<br />

that included 739 chemotherapy-naïve patients who were randomly assigned to<br />

receive, at progression, doxorubicin, paclitaxel, or both (39). Although the<br />

response rates and times to treatment failure were improved with the combination<br />

regimen, the overall survival rate was comparable in all groups. Other trials<br />

have demonstrated a survival advantage with combination regimens over singleagent<br />

chemotherapy, but all <strong>of</strong> these studies have shown differences in the actual<br />

median overall survival time between 10 and 13 months across studies, suggesting<br />

that even with comparable inclusion criteria, the heterogeneity that is<br />

typical <strong>of</strong> MBC cannot be eliminated (41,42). This heterogeneity is indeed one<br />

<strong>of</strong> the major limitations to the development <strong>of</strong> more personalized treatments for<br />

patients. Therefore, although the data demonstrate that patients can benefit from<br />

combination therapy, they do not clearly identify the subsets <strong>of</strong> patients who will<br />

most benefit or who will experience only additional toxicity.<br />

In this context, the use <strong>of</strong> CTCs to stratify patients at the time <strong>of</strong> disease<br />

recurrence may be an appropriate way to design personalized therapeutic<br />

approaches. CTC detection may enable a more rational selection <strong>of</strong> treatments<br />

for patients with newly recurrent disease, and this approach could maximize<br />

the chance <strong>of</strong> a particular combination or single new drug showing clinical<br />

benefit and, eventually, prolonging survival. It is possible that CTC detection<br />

could be used in the design <strong>of</strong> efficacy trials <strong>of</strong> different therapeutic<br />

approaches. The efficacy <strong>of</strong> these treatments could be more easily assessed if<br />

patients were stratified by their prognosis, leading to more tailored treatment<br />

strategies. We believe that the challenge for the next generation <strong>of</strong> clinical<br />

trials, and the responsibility for both clinical investigators and the pharmaceutical<br />

industry will be to incorporate these concepts into the process <strong>of</strong> drug<br />

development.


162 Reuben and Crist<strong>of</strong>anilli<br />

FUTURE USES OF CTCs<br />

The process <strong>of</strong> sorting cancer cells from other cellular components (e.g., blood<br />

and stromal cells) in clinical samples is fundamentally important for the future <strong>of</strong><br />

genomic and proteomic analysis (43–45). Collecting representative tissue from<br />

solid tumor metastases usually requires more invasive procedures that increase<br />

the risk <strong>of</strong> complications and discomfort. Furthermore, these procedures may not<br />

provide an adequate specimen for detailed analysis and typically cannot be<br />

repeated for dynamic evaluation <strong>of</strong> the biologic changes during treatments.<br />

Theoretically, CTC detection would allow specific genes [e.g., c-erbB-2, epidermal<br />

growth factor receptor (EGFR), and mammaglobin B (MGB)] or more global<br />

gene expression to be analyzed while using specific targeted treatments for MBC<br />

based on the expression <strong>of</strong> the CTCs (29,46–48). This information could then be<br />

used to design specific treatments that more appropriately reflect the dynamics and<br />

heterogeneity <strong>of</strong> MBC.<br />

CONCLUSIONS<br />

The detection <strong>of</strong> microscopic disease in the peripheral blood <strong>of</strong> patients with<br />

MBC provides prognostic information. This information will allow appropriate<br />

risk stratification and modification <strong>of</strong> the current staging system for advanced<br />

disease. Furthermore, the intriguing hypothesis that CTCs are, at least in part,<br />

representative <strong>of</strong> tumorigenic cancer stem cells may present an additional<br />

challenge for translational research. In fact, the phenotypic evaluation <strong>of</strong> CTCs is<br />

indicating the possibility <strong>of</strong> more critical design <strong>of</strong> targeted therapies in the<br />

setting <strong>of</strong> MBC. Moreover, with the introduction <strong>of</strong> this novel concept, it is<br />

expected that investigators and the pharmaceutical industry will derive benefits<br />

from these sorting technologies that will contribute to more tailored treatments<br />

and sophisticated trial designs.<br />

REFERENCES<br />

1. Surveillance Epidemiology and End Results (SEER). National <strong>Cancer</strong> Institute<br />

Website. Available at: http://seer.cancer.gov/statfacts/html/breast.htm.<br />

2. Folkman J. Tumor angiogenesis: therapeutic implications. N Engl J Med 1971; 285<br />

(21):1182–1186. Available at: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=<br />

Retrieve&db=PubMed&dopt=Citation&list_uids=4938153.<br />

3. Clare SE, Sener SF, Wilkens W, et al. Prognostic significance <strong>of</strong> occult lymph node<br />

metastases in node-negative breast cancer. Ann Surg Oncol 1997; 4(6):447–451.<br />

Available at http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=<br />

PubMed&dopt=Citation&list_uids=9309332.<br />

4. Braun S, Pantel K, Muller P, et al. Cytokeratin-positive cells in the bone marrow<br />

and survival <strong>of</strong> patients with stage I, II, or III breast cancer. N Engl J Med 2000; 342(8):<br />

525–533. Available at: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve<br />

&db=PubMed&dopt=Citation&list_uids=10684910.


Circulating Tumor Cells in Individualizing <strong>Breast</strong> <strong>Cancer</strong> Therapy 163<br />

5. Ashworth T. A case <strong>of</strong> cancer in which cells similar to those in the tumours were<br />

seen in the blood after death. Aust Med J 1869; 14:146.<br />

6. Roberts S, Watne A, McGrath R. Technique and results <strong>of</strong> isolation <strong>of</strong> cancer cells<br />

from the circulating blood. Am Med Assoc Arch Surg 1958; 76:334–346.<br />

7. Long L, Jonasson O, Roberts S, et al. <strong>Cancer</strong> cells in blood: results <strong>of</strong> a simplified<br />

isolation technique. Arch Surg 1960; 80:910–919. Available at http://www.ncbi.nlm.<br />

nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=<br />

14418112.<br />

8. Dontu G, Al-Hajj M, Abdallah WM, et al. Stem cells in normal breast development<br />

and breast cancer. Cell Prolif 2003; 36(suppl 1):59–72. Available at: http://www<br />

.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_<br />

uids=14521516.<br />

9. Al-Hajj M, Wicha MS, Benito-Hernandez A, et al. Prospective identification <strong>of</strong><br />

tumorigenic breast cancer cells. Proc Natl Acad Sci U S A 2003; 100(7):3983–3988.<br />

Available at: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=Pub<br />

Med&dopt=Citation&list_uids=12629218.<br />

10. Al-Hajj M, Becker MW, Wicha M, et al. Therapeutic implications <strong>of</strong> cancer stem<br />

cells. Curr Opin Genet Dev 2004; 14(1):43–47. Available at: http://www.ncbi.nlm.<br />

nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=<br />

15108804.<br />

11. Crist<strong>of</strong>anilli M, Budd GT, Ellis MJ, et al. Circulating tumor cells, disease progression,<br />

and survival in metastatic breast cancer. N Engl J Med 2004; 351(8):781–791.<br />

Available at: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=Pub<br />

Med&dopt=Citation&list_uids=15317891.<br />

12. Reed W, Bohler PJ, Sandstad B, et al. Occult metastases in axillary lymph nodes as a<br />

predictor <strong>of</strong> survival in node-negative breast carcinoma with long-term follow-up.<br />

<strong>Breast</strong> J 2004; 10(3):174–180. Available at: http://www.ncbi.nlm.nih.gov/entrez/<br />

query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=15125740.<br />

13. An XY, Egami H, Hayashi N, et al. Clinical significance <strong>of</strong> circulating cancer cells in<br />

peripheral blood detected by reverse transcriptase-polymerase chain reaction in<br />

patients with breast cancer. Tohoku J Exp Med 2001; 193(2):153–162. Available at:<br />

http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=<br />

Citation&list_uids=11318030.<br />

14. Bossolasco P, Ricci C, Farina G, et al. Detection <strong>of</strong> micrometastatic cells in breast cancer<br />

by RT-pCR for the mammaglobin gene. <strong>Cancer</strong> Detect Prev 2002; 26(1):60–63.<br />

Available at: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=Pub<br />

Med&dopt=Citation&list_uids=12088204.<br />

15. Gross HJ, Verwer B, Houck D, et al. Model study detecting breast cancer cells in<br />

peripheral blood mononuclear cells at frequencies as low as 10(-7). Proc Natl Acad<br />

Sci U S A 1995; 92(2):537–541. Available at: http://www.ncbi.nlm.nih.gov/entrez/<br />

query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=7831325.<br />

16. Hawes D, Neville AM, Cote RJ. Methods used for the detection <strong>of</strong> occult metastasis.<br />

Biomed <strong>Ph</strong>armacother 2001; 55(4):229–242. Available at: http://www.ncbi.nlm.nih.<br />

gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=<br />

11393810.<br />

17. Hu XC, Chow LW. Detection <strong>of</strong> circulating breast cancer cells with multiple-marker<br />

RT-PCR assay. Anticancer Res 2001; 21(1A):421–424. Available at: http://www


164 Reuben and Crist<strong>of</strong>anilli<br />

.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation<br />

&list_uids=11299772.<br />

18. Martin KJ, Graner E, Li Y, et al. High-sensitivity array analysis <strong>of</strong> gene expression<br />

for the early detection <strong>of</strong> disseminated breast tumor cells in peripheral blood. Proc<br />

Natl Acad Sci U S A 2001; 98(5):2646–2651. Available at http://www.ncbi.nlm.nih.<br />

gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=<br />

11226293.<br />

19. Pachmann K, Heiss P, Demel U, et al. Detection and quantification <strong>of</strong> small numbers<br />

<strong>of</strong> circulating tumour cells in peripheral blood using laser scanning cytometer (LSC).<br />

Clin Chem Lab Med 2001; 39(9):811–817. Available at http://www.ncbi.nlm.nih.<br />

gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=<br />

11601678.<br />

20. Racila E, Euhus D, Weiss AJ, et al. Detection and characterization <strong>of</strong> carcinoma<br />

cells in the blood. Proc Natl Acad Sci U S A 1998; 95(8):4589–4594. Available at<br />

http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=<br />

Citation&list_uids=9539782.<br />

21. Witzig TE, Bossy B, Kimlinger T, et al. Detection <strong>of</strong> circulating cytokeratin-positive<br />

cells in the blood <strong>of</strong> breast cancer patients using immunomagnetic enrichment and digital<br />

microscopy. Clin <strong>Cancer</strong> Res 2002; 8(5):1085–1091. Available at http://www.ncbi.nlm.<br />

nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=<br />

12006523.<br />

22. Kagan M, Howard D, Bendele T, et al. A sample preparation and analysis system for<br />

identification <strong>of</strong> circulating tumor cells J Clin Ligand Assay 2002; 25(1):104–110.<br />

23. Terstappen LW, Rao C, Gross S, et al. Peripheral blood tumor cell load reflects the<br />

clinical activity <strong>of</strong> the disease in patients with carcinoma <strong>of</strong> the breast. Int J Oncol<br />

2000; 17(3):573–578. Available at: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?<br />

cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=10938400.<br />

24. Riethdorf S, Fritsche H, Muller V, et al. Detection <strong>of</strong> circulating tumor cells in<br />

peripheral blood <strong>of</strong> patients with metastatic breast cancer: a validation study <strong>of</strong> the<br />

CellSearch system. Clin <strong>Cancer</strong> Res 2007; 13(3):920–928. Available at: http://<br />

clincancerres.aacrjournals.org/cgi/content/abstract/13/3/920.<br />

25. Taubert H, Blumke K, Bilkenroth U, et al. Detection <strong>of</strong> disseminated tumor cells in<br />

peripheral blood <strong>of</strong> patients with breast cancer: correlation to nodal status and<br />

occurrence <strong>of</strong> metastases. Gynecol Oncol 2004; 92:256–261.<br />

26. Lankiewicz S, Gutierrez B, Böcher O. Quantitative real-time RT-PCR <strong>of</strong> disseminated<br />

tumor cells in combination with immunomagnetic cell enrichment. Mol<br />

Biotechnol 2006; 34:15–27.<br />

27. Austrup F, Uciechowski P, Eder C. Prognostic value <strong>of</strong> genomic alterations in<br />

minimal residual cancer cells purified from the blood <strong>of</strong> breast cancer patients. Br J<br />

<strong>Cancer</strong> 2001; 83:1664–1673.<br />

28. Meng S, Tripathy D, Shete S, et al. HER-2 gene amplification can be acquired as<br />

breast cancer progresses. Proc Natl Acad Sci U S A 2004; 101(25):9393–9398.<br />

Available at: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=Pub<br />

Med&dopt=Citation&list_uids=15194824.<br />

29. Meng S, Tripathy D, Shete S, et al. uPAR and HER-2 gene status in individual<br />

breast cancer cells from blood and tissues. Proc Natl Acad Sci U S A 2006; 103<br />

(46):17361–17365. Available at: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=<br />

Retrieve&db= PubMed&dopt=Citation&list_uids=17079488.


Circulating Tumor Cells in Individualizing <strong>Breast</strong> <strong>Cancer</strong> Therapy 165<br />

30. Crist<strong>of</strong>anilli M, Hayes DF, Budd GT, et al. Circulating tumor cells: a novel prognostic<br />

factor for newly diagnosed metastatic breast cancer. J Clin Oncol 2005; 23(7):<br />

1420–1430. Available at: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=<br />

Retrieve&db=PubMed&dopt=Citation&list_uids=15735118.<br />

31. Crist<strong>of</strong>anilli M, Broglio KR, Guarneri V, et al. Circulating tumor cells in metastatic<br />

breast cancer: biologic staging beyond tumor burden. Clin <strong>Breast</strong> <strong>Cancer</strong> 2007; 7(6):<br />

471–479. Available at: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve<br />

&db=PubMed&dopt=Citation&list_uids=17386124.<br />

32. Budd GT, Crist<strong>of</strong>anilli M, Ellis MJ. Circulating tumor cells versus imaging:<br />

Predicting overall survival in metastatic breast cancer. Clin <strong>Cancer</strong> Res 2006; 12(21):<br />

6403–6409.<br />

33. Internation Union Against <strong>Cancer</strong> (UICC). TNM Classification <strong>of</strong> Malignant Tumours.<br />

1st ed. Geneva, Switzerland: UICC, 1968.<br />

34. Denoix P. Tumor, node, and metastasis (TNM). Bull Inst Nat Hyg (Paris) 1944; 1:1–69.<br />

35. Berndt H, Titze U. TNM clinical stage classification <strong>of</strong> breast cancer. Int J <strong>Cancer</strong><br />

1969; 4(6):837–844. Available at: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?<br />

cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=5378609.<br />

36. Greene FL, Page DI, Fleming ID, eds. AJCC <strong>Cancer</strong> Staging Manual. 6th ed. New<br />

York, NY: Springer-Verlag, 2002.<br />

37. Ellis MJ, Hayes DF, Lippman ME. Treatment <strong>of</strong> Metastatic Disease. <strong>Ph</strong>iladelphia,<br />

PA: Lippincott-Raven, 2000.<br />

38. Swenerton KD, Legha SS, Smith T, et al. Prognostic factors in metastatic breast<br />

cancer treated with combination chemotherapy. <strong>Cancer</strong> Res 1979; 39(5):1552–1562.<br />

Available at: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=Pub<br />

Med&dopt=Citation&list_uids=427797.<br />

39. Bernard-Marty C, Cardoso F, Piccart MJ. Facts and controversies in systemic<br />

treatment <strong>of</strong> metastatic breast cancer. Oncologist 2004; 9(6):617–632. Available at:<br />

http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=<br />

Citation&list_uids=15561806.<br />

40. Sledge GW, Neuberg D, Bernardo P, et al. <strong>Ph</strong>ase III trial <strong>of</strong> doxorubicin, paclitaxel,<br />

and the combination <strong>of</strong> doxorubicin and paclitaxel as front-line chemotherapy for<br />

metastatic breast cancer: an intergroup trial (E1193). J Clin Oncol 2003; 21(4):<br />

588–592. Available at: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve<br />

&db=PubMed&dopt=Citation&list_uids=12586793.<br />

41. O’Shaughnessy J, Miles D, Vukelja S, et al. Superior survival with capecitabine plus<br />

docetaxel combination therapy in anthracycline-pretreated patients with advanced<br />

breast cancer: phase III trial results. J Clin Oncol 2002; 20(12):2812–2823. Available<br />

at: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=<br />

Citation&list_uids=12065558.<br />

42. O’Shaughnessy J, Nag S, Calderillo-Ruiz G. Gemcitabine plus paclitaxel versus<br />

paclitaxel as first-line treatment for anthracycline pre-treated metastatic breast cancer:<br />

interim results <strong>of</strong> a global phase III study. Proc Am Soc Clin Oncol 2003; 22:7.<br />

43. Hancock W, Apffel A, Chakel J, et al. Integrated genomic/proteomic analysis. Anal<br />

Chem 1999; 71(21):742A–748A. Available at: http://www.ncbi.nlm.nih.gov/entrez/<br />

query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=10565269.<br />

44. Marshall A, Hodgson J. DNA chips: an array <strong>of</strong> possibilities. Nat Biotechnol 1998; 16(1):<br />

27–31. Available at: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&<br />

db=PubMed&dopt=Citation&list_uids=9447589.


166 Reuben and Crist<strong>of</strong>anilli<br />

45. Page MJ, Amess B, Townsend RR, et al. Proteomic definition <strong>of</strong> normal human<br />

luminal and myoepithelial breast cells purified from reduction mammoplasties. Proc<br />

Natl Acad Sci U S A 1999; 96(22):12589–12594. Available at: http://www.ncbi.nlm.<br />

nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=<br />

10535966.<br />

46. Smirnov DA, Zweitzig DR, Foulk BW, et al. Global gene expression pr<strong>of</strong>iling <strong>of</strong><br />

circulating tumor cells. <strong>Cancer</strong> Res 2005; 65(12):4993–4997. Available at: http://www<br />

.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&<br />

list_ uids=15958538.<br />

47. Speicher MR, Carter NP. The new cytogenetics: blurring the boundaries with<br />

molecular biology. Nat Rev Genet 2005; 6(10):782–792. Available at: http://www<br />

.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&<br />

list_uids=16145555.<br />

48. Reuben JM, Lee BN, Li C. Genomic <strong>of</strong> circulating tumor cells in metastatic breast<br />

cancer. Proc ASOC 25:32s (abstr 1002).


12<br />

Individualization <strong>of</strong> Endocrine Therapy<br />

in <strong>Breast</strong> <strong>Cancer</strong><br />

Amelia B. Zelnak and Ruth M. O’Regan<br />

Emory Winship <strong>Cancer</strong> Institute, Atlanta, Georgia, U.S.A.<br />

Clodia Osipo<br />

Loyola University, Maywood, Illinois, U.S.A.<br />

INTRODUCTION<br />

Discovery <strong>of</strong> estrogen receptors (ER) was critical for the development <strong>of</strong><br />

endocrine therapy in breast cancer. Expression <strong>of</strong> ERa, thepredominantis<strong>of</strong>orm,<br />

in breast tumors <strong>of</strong> both premenopausal and postmenopausal women is a<br />

highly predictive marker for response to antiestrogen treatment in women with<br />

ERa-positive breast cancer. Today, endocrine therapies include the use <strong>of</strong><br />

antiestrogens such as tamoxifen and aromatase inhibitors such as anastrozole,<br />

exemestane, and letrozole. Tamoxifen, the first antiestrogen to be used for the<br />

treatment <strong>of</strong> ERa-positive breast cancer, competitively blocks the actions <strong>of</strong><br />

17b-estradiol (E 2 ), the female hormone that binds and activates ERa in tumors.<br />

In postmenopausal women, peripheral aromatization <strong>of</strong> androgens to estrogens<br />

is the major source <strong>of</strong> plasma estrogen. Aromatase inhibitors inhibit this<br />

reaction and consequently suppress the production <strong>of</strong> circulating estrogen in<br />

postmenopausal women. Endocrine therapies are effective at reducing recurrence,<br />

increasing overall survival, and reducing contralateral breast cancer up<br />

to 50%. However, about 50% <strong>of</strong> patients with ERa-positive breast cancer have<br />

167


168 Zelnak et al.<br />

intrinsic resistance to antiestrogen therapy and therefore do not benefit. In<br />

contrast to patients with intrinsically resistant tumors, there are patients who do<br />

initially respond to antiestrogen therapy; however, most <strong>of</strong> these patients<br />

develop acquired resistance during the treatment regimen. Therefore, the<br />

current goal in breast cancer research is to elucidate the mechanisms <strong>of</strong> both<br />

intrinsic and acquired resistance to tamoxifen and the aromatase inhibitors in<br />

order to develop new therapeutic strategies to prevent and/or treat resistant<br />

breast cancer.<br />

ANTIESTROGENS<br />

Nonsteroidal antiestrogens were initially developed as contraceptives in the<br />

1960s. Walpole and colleagues synthesized tamoxifen (termed ICI 46, 474), a<br />

potent antiestrogen with antifertility properties in rats. However, in humans,<br />

tamoxifen induced ovulation in subfertile women. Therefore, the development<br />

<strong>of</strong> tamoxifen as a contraceptive was terminated. However, Walpole also<br />

patented the application <strong>of</strong> tamoxifen as a drug treatment for hormonedependent<br />

cancers. Thus, clinical trials were started to evaluate tamoxifen<br />

against the standard endocrine treatment at the time, diethylstilbestrol (DES),<br />

for the treatment <strong>of</strong> advanced breast cancer in postmenopausal women (1).<br />

Tamoxifen not only was as effective as DES for the treatment <strong>of</strong> advanced<br />

breast cancer but also had fewer side effects. Therefore, the advantage <strong>of</strong><br />

tamoxifen over DES was crucial for its subsequent evaluation as a treatment for<br />

all stages <strong>of</strong> breast cancer.<br />

In 1962, Jensen and colleagues discovered ERa. Jensen demonstrated<br />

that E 2 , the circulating female hormone that promoted breast cancer growth,<br />

binds to diverse tissue sites around a woman’s body but is retained in estrogentarget<br />

tissues, for example, the uterus and vagina (2). The identification <strong>of</strong><br />

ERa as the target <strong>of</strong> E 2 action in the breast and antiestrogens blocking the<br />

binding <strong>of</strong> E 2 to ERa provided a therapeutic target and an approach for the<br />

treatment <strong>of</strong> breast cancer. Although many antiestrogens were discovered and<br />

tested during the 1960s and 1970s, only tamoxifen was considered safe enough<br />

for extensive clinical evaluation (3). Clinical trials ultimately demonstrated<br />

that patients with ERa-positive breast cancer benefited the most from<br />

tamoxifen therapy, whereas women with ERa-negative breast cancer were<br />

found to be unaffected. During subsequent clinical trials, five years <strong>of</strong> adjuvant<br />

tamoxifen treatment was found to be more effective than less than five<br />

years <strong>of</strong> treatment in improving time to tumor recurrence and overall survival<br />

(4). In contrast to the beneficial effects <strong>of</strong> tamoxifen as a treatment for breast<br />

cancer, both laboratory and clinical results showed that tamoxifen increased<br />

the risk <strong>of</strong> endometrial cancer in the uterus fourfold in postmenopausal women<br />

compared with untreated women (5,6). These results strongly indicated that<br />

tamoxifen was not a pure antiestrogen but had selective functions depending<br />

on the target tissue.


Individualization <strong>of</strong> Endocrine Therapy in <strong>Breast</strong> <strong>Cancer</strong> 169<br />

Resistance to Tamoxifen<br />

Overview<br />

Five years <strong>of</strong> adjuvant tamoxifen therapy is a standard <strong>of</strong> care for early-stage ERapositive<br />

breast cancer. Although there is increasing use <strong>of</strong> aromatase inhibitors in<br />

postmenopausal women, tamoxifen is the only approved effective therapy in premenopausal<br />

women. Approximately 50% <strong>of</strong> patients with ERa-positive breast<br />

cancer benefit from tamoxifen treatment in the advanced- and early-stage settings<br />

(4,7). Unfortunately, the other 50% <strong>of</strong> patients do not respond, and the ones that do<br />

respond initially eventually progress. Thus, the problem with tamoxifen therapy is<br />

that breast tumors can either be resistant (intrinsic or de novo) prior to endocrine<br />

treatment or become resistant (acquired) during therapy. Numerous studies<br />

have looked at the mechanisms responsible for this resistant phenotype. They<br />

include the loss <strong>of</strong> the ERa gene and/or protein in tumors during short-term and<br />

long-term treatment, identification <strong>of</strong> mutations within the ERa gene, changes in<br />

the pharmacologic response <strong>of</strong> resistant breast cancer cells to tamoxifen, activation<br />

<strong>of</strong> ERa-mediated gene transcription in the absence <strong>of</strong> ligand, modifications in gene<br />

transcription by changes in interactions between ERa and coregulators, and the<br />

development <strong>of</strong> enhanced surface-to-intracellular signaling cross-talk between<br />

growth factor receptors and ERa (Fig. 1).<br />

Figure 1<br />

Mechanisms <strong>of</strong> hormone resistance.


170 Zelnak et al.<br />

ER-Negative <strong>Breast</strong> <strong>Cancer</strong> and Resistance<br />

<strong>Breast</strong> cancers initially lacking expression <strong>of</strong> ERa have intrinsic resistance to<br />

endocrine therapy. In addition, loss <strong>of</strong> ERa expression upon recurrence <strong>of</strong> breast<br />

cancer during tamoxifen therapy has been reported in as many as 25% <strong>of</strong> tumors<br />

(8–10). However, most <strong>of</strong> the acquired resistance during adjuvant tamoxifen<br />

treatment in ERa-positive tumors is not due to changes in ERa expression.<br />

Interestingly, most tamoxifen-resistant breast tumors retain ERa and respond to<br />

secondary endocrine treatments. Overall, a loss <strong>of</strong> ERa status does not seem to<br />

be the major mechanism for the development <strong>of</strong> acquired antiestrogen resistance.<br />

ER Mutations<br />

Several mutant forms <strong>of</strong> both the ERa and ERb, another ER is<strong>of</strong>orm, have been<br />

identified and previously reviewed (11–13). However, the functional significance<br />

<strong>of</strong> these mutants as a mechanism <strong>of</strong> antiestrogen resistance is yet to be<br />

elucidated. Most <strong>of</strong> the data demonstrate the existence <strong>of</strong> splice variants for ERa<br />

and ERb mRNAs without evidence <strong>of</strong> translation into functional proteins. Most<br />

<strong>of</strong> the tumors express both mutant and wild-type ER with the wild type being the<br />

predominant species. A single-point mutation in the ligand-binding domain<br />

(LBD) <strong>of</strong> ERa has been reported (D351Y) that converts tamoxifen and ralox<br />

ifene from antiestrogen to estrogens in in vivo and in vitro models <strong>of</strong> antiestrogenstimulated<br />

MCF-7 tumor cells (14,15). However, the D351Y ERa mutant has yet<br />

to be detected in either intrinsic or acquired resistant breast tumors from patients.<br />

In addition, mutations in the F-region <strong>of</strong> the ER have been shown to affect<br />

the activities <strong>of</strong> both E 2 and 4-hydroxytamoxifen (4OHT), the active metabolite<br />

<strong>of</strong> tamoxifen (16). Therefore, the clinical relevance <strong>of</strong> ER mutants is unclear to<br />

date as a mechanism <strong>of</strong> resistance to antiestrogens. However, our understanding<br />

<strong>of</strong> ER mutations in response to antiestrogens might lead to the development <strong>of</strong><br />

better antiestrogens or other drugs for breast cancer treatment.<br />

Role <strong>of</strong> <strong><strong>Ph</strong>armacogenetics</strong><br />

Tamoxifen undergoes extensive primary and secondary metabolism. The secondary<br />

metabolite 4-hydroxy-N-desmethyl tamoxifen (endoxifen) is formed by<br />

the CYP2D6-mediated oxidation <strong>of</strong> N-desmethyl tamoxifen and has been shown<br />

to play an important role in the anticancer effect <strong>of</strong> tamoxifen. CYP2D6 genotyping<br />

has revealed that women can be classified as either poor, intermediate, or<br />

extensive metabolizers <strong>of</strong> tamoxifen. Women who are poor metabolizers <strong>of</strong><br />

tamoxifen, and, therefore, have lower levels <strong>of</strong> endoxifen, have been shown to<br />

have worse disease-free survival. No difference in overall survival has been<br />

shown (17). Women who are extensive metabolizers <strong>of</strong> tamoxifen have also been<br />

shown to experience more hot flashes. A recent study showed that women who<br />

experience hot flashes while on tamoxifen had a hazard ratio <strong>of</strong> recurrence <strong>of</strong>


Individualization <strong>of</strong> Endocrine Therapy in <strong>Breast</strong> <strong>Cancer</strong> 171<br />

0.51 compared with women without hot flashes (18). This study suggests that<br />

vasomotor symptoms and efficacy <strong>of</strong> tamoxifen may be related to a women’s<br />

pharmacogenetics.<br />

Role <strong>of</strong> Coregulators in Resistance<br />

The recruitment <strong>of</strong> coactivators or corepressors to the ER determines the switch<br />

between ER activation and repression. In addition, the coordinated action <strong>of</strong><br />

ligand, ER, and coregulators determines which genes are transcribed or repressed<br />

depending on the cellular context and thus which cells will or will not proliferate.<br />

The overall data indicate that intricate modulation <strong>of</strong> the ER-to-coregulator ratio<br />

in breast cancer cells could determine resistance to antiestrogens. For example,<br />

the coactivator SRC-1 activates ER in a ligand-independent manner while<br />

increasing 4OHT’s agonist activity (19). On the other hand, the corepressor<br />

SMRT blocks the agonist activity <strong>of</strong> 4OHT-induced by SRC-1 (20). The other<br />

member <strong>of</strong> the p160 coactivator family is amplified in breast cancer (AIB1,<br />

SRC-3, and RAC3). AIB1 mRNA was found to be amplified in 60% <strong>of</strong> breast<br />

cancers, and the protein product was found to be overexpressed in about 10% <strong>of</strong><br />

tumors. Recently, an association was discovered between high AIB1 and human<br />

epidermal growth factor receptor 2 (HER2/neu), a member <strong>of</strong> the epidermal<br />

growth factor receptor (EGFR) family <strong>of</strong> receptor tyrosine kinases and tamoxifenresistant<br />

breast cancers (20). The potential mechanism for this resistance requires<br />

activation <strong>of</strong> the mitogen-activated protein kinase (MAPK) signaling cascade by<br />

HER2/neu, which in turn potentially phosphorylates and activates AIB1. Activated<br />

AIB1 in turn can convert 4OHT from an antiestrogen to an estrogen in<br />

regards to ERa activity. In contrast, low levels <strong>of</strong> N-CoR mRNA, a corepressor,<br />

have been implicated with resistance to tamoxifen (21). Therefore, the identification<br />

<strong>of</strong> abnormally expressed coregulators will probably assist in the design<br />

<strong>of</strong> future therapies that improve the efficacy <strong>of</strong> endocrine treatment without the<br />

development <strong>of</strong> resistance.<br />

EGFR and HER2/Neu and Antiestrogen Resistance<br />

Numerous laboratory and clinical studies indicate that overexpression and/or<br />

aberrant activity <strong>of</strong> the HER2/neu (erbB2) signaling pathway is associated with<br />

antiestrogen resistance in breast cancer (22). The HER2/neu receptor is a<br />

member <strong>of</strong> the EGFR family <strong>of</strong> receptor tyrosine kinases, which include HER3<br />

(erbB3) and HER4 (erbB4) (23–26). Upon dimerization, the tyrosine kinase<br />

domains located within the COOH-terminal regions <strong>of</strong> receptors are activated by<br />

an autophosphorylation cascade on specific tyrosine residues, which activate<br />

downstream effectors, such as MAPK and Akt, which promote cellular proliferation,<br />

survival, anti-apoptosis, and transformation. HER2/neu is overexpressed<br />

and/or amplified in 25% to 30% <strong>of</strong> breast tumors and is associated with a more<br />

aggressive phenotype and poor prognosis (27). Patients with breast tumors<br />

overexpressing HER2/neu exhibit much lower response rates to antiestrogen


172 Zelnak et al.<br />

therapy (27). Thus, it is suggested that one possible mechanism <strong>of</strong> resistance to<br />

tamoxifen is overexpression <strong>of</strong> HER2/neu in ERa-positive breast cancers. In<br />

addition, overexpression <strong>of</strong> EGFR and its ligands are observed in several human<br />

cancers including breast cancer (28–30). The increased expression <strong>of</strong> EGFR is<br />

frequently associated with tumor progression and resistance to antiestrogens.<br />

These data suggest that EGFR and/or HER2/neu are possible targets for preventing<br />

or treating antiestrogen resistance in breast cancer. Strategies to target<br />

EGFR and HER2/neu include the use <strong>of</strong> humanized monoclonal antibodies to<br />

the receptors (31), tyrosine kinase inhibitors that block reduction <strong>of</strong> adenosine<br />

triphosphate (ATP) to adenosine diphospohate (ADP) þ Pi, and receptor antisense<br />

molecules (32). Gefitinib, (ZD 1839, Iressa 1 ), an EGFR-specific tyrosine kinase<br />

inhibitor, has been shown to inhibit growth <strong>of</strong> breast cancer cell lines in vitro that<br />

are resistant to tamoxifen (33). More importantly, the combination <strong>of</strong> gefitinib<br />

and tamoxifen was shown to prevent resistance (34), demonstrating that (1)<br />

EGFR might be a key player in the development <strong>of</strong> antiestrogen resistance and<br />

(2) inhibiting the activity <strong>of</strong> this receptor might be therapeutically beneficial in<br />

preventing resistance to antiestrogens. In addition to gefitinib, trastuzumab<br />

(Herceptin 1 ) is a humanized monoclonal antibody directed against the ectodomain<br />

<strong>of</strong> the HER2/neu receptor. It has been shown to restore breast cancer cell<br />

sensitivity to tamoxifen in HER2/neu-overexpressing cells (35).<br />

Role <strong>of</strong> MAPK<br />

MAPK is a serine/threonine kinase that is activated by phosphorylation cascades<br />

originating from GTP-bound Ras, a downstream effector <strong>of</strong> EGFR and HER2/<br />

neu. MAPK has been shown to be hyperactivated in MCF-7 breast cancer cells<br />

overexpressing either EGFR or HER2/neu (36–38). This increased activity <strong>of</strong><br />

MAPK promoted an enhanced association <strong>of</strong> ERa with coactivators and<br />

decreased interaction with corepressors, thereby enhancing hormone-dependent<br />

gene transcription and possibly leading to tamoxifen-resistant breast cancer. The<br />

role <strong>of</strong> the MAPK signaling pathway in tamoxifen resistance has been demonstrated<br />

primarily in vitro using specific inhibitors such as U0126 (MAPKK,<br />

MEK1/2 inhibitor), AG1478 (EGFR and HER2/neu inhibitors), and PD98059<br />

(MAPK, ERK1/2 inhibitor) (36–38). The exact mechanism by which hyperactivated<br />

MAPK converts the antiestrogenic activity <strong>of</strong> the tamoxifen-ERa<br />

complex to more <strong>of</strong> an estrogenic one in resistant breast cancer cell is yet<br />

unclear. However, it has been demonstrated that activation <strong>of</strong> the Ras/MAPK<br />

pathway by EGFR/HER2/neu receptor activation could phosphorylate Ser-118 in<br />

the AF-1 domain <strong>of</strong> ERa, thus promoting ligand-independent transcription <strong>of</strong><br />

ERa and possibly loss <strong>of</strong> tamoxifen-induced inhibition <strong>of</strong> ERa-mediated gene<br />

transcription. On the basis <strong>of</strong> these preclinical models, several clinical trials are<br />

under way to determine whether treating patients with EGFR inhibitors (i.e.,<br />

Iressa) and/or MAPK-specific inhibitors (U0126) can effectively treat and/or<br />

prevent antiestrogen resistance in breast tumors.


Individualization <strong>of</strong> Endocrine Therapy in <strong>Breast</strong> <strong>Cancer</strong> 173<br />

Role <strong>of</strong> PI3-K/AKT Signaling<br />

<strong>Ph</strong>osphatidylinositol 3-kinase (PI3-K) is a heterodimer complex consisting <strong>of</strong><br />

a regulatory subunit, p85, and a catalytic subunit, p110. The regulatory subunit<br />

p85 is phosphorylated by either receptor tyrosine kinases such as EGFR/<br />

HER2/neu or intracellular adapter kinases such as Shc or Src and thereafter<br />

interacts and activates the catalytic subunit p110 (39). The activated PI3-K<br />

subsequently activates downstream effector kinases such as Akt (PKB), a<br />

serine/threonine kinase. Akt has been shown to activate either directly or<br />

indirectly proteins responsible for prosurvival and anti-apoptosis in cancer<br />

cells. Several reports indicate that the PI3-K/Akt pathway interacts with the<br />

ERa pathway, resulting in bidirectional cross talk. It has been shown that PI3-<br />

K/Akt can mediate E 2 -induced transcription <strong>of</strong> cyclin D1 and entry <strong>of</strong> MCF-7<br />

breast cancer cells into S-phase (40). In addition, the PI3-K/Akt pathway can<br />

induce ERa phosphorylation on Ser-167 to promote ERa-mediated transcription<br />

and thus protect cells against tamoxifen-induced apoptosis (41). The<br />

accumulating data suggest that ERa-positive breast tumors with alterations <strong>of</strong><br />

PI3-K and Akt signaling, whether dependent or independent <strong>of</strong> EGFR/HER2/<br />

neu, might be insensitive to antiestrogens and thus lead to resistance. Future<br />

studies are needed to confirm whether blocking PI3-K and/or Akt signaling in<br />

ERa-positive breast cancer will be beneficial in subverting resistance to<br />

antiestrogens.<br />

New SERMS<br />

New drug discovery for selective ER modulators (SERMs) is currently driven<br />

by the known side effects <strong>of</strong> tamoxifen, which include an increase in the<br />

incidence <strong>of</strong> endometrial cancer, development <strong>of</strong> resistance, and the recent<br />

report <strong>of</strong> the negative effects <strong>of</strong> hormone replacement therapy (HRT) on coronary<br />

heart disease (42). Several approaches are being pursued by altering the<br />

antiestrogenic side chain or improving the pharmacokinetics <strong>of</strong> existing molecules.<br />

Several novel antiestrogen compounds have been developed to replace<br />

tamoxifen for ERa-positive breast cancer without the agonist actions. These<br />

compounds have the potential to be more effective than tamoxifen, having<br />

the advantage <strong>of</strong> reducing the incidence <strong>of</strong> endometrial cancer and possibly<br />

acquired resistance during therapy. Two groups <strong>of</strong> antiestrogenic drugs exist<br />

today with the potential to replace tamoxifen: (1) The SERMs that include<br />

tamoxifen-like compounds (triphenylethylenes) such as toremifene, idoxifene,<br />

and GW 5638 and fixed-ring compounds (benzothiophenes) such as raloxifene,<br />

arzoxifene, EM 652, and CP 336,156, and (2) selective ER downregulators<br />

(SERDs and pure antiestrogen) such as ICI 182,780 (Fulvestrant 1 ). In phase II<br />

trials in patients with tamoxifen-resistant metastatic breast cancer, the new<br />

SERMs showed low response rates (0–15%) (43,44), suggesting crossresistance<br />

to tamoxifen. No difference in outcome was noted between<br />

tamoxifen and toremifene in patients with metastatic or early-stage breast


174 Zelnak et al.<br />

cancer (45,46). In contrast, few clinical trials exist today evaluating the<br />

effectiveness <strong>of</strong> the fixed-ring compounds to tamoxifen. The Study <strong>of</strong><br />

Tamoxifen and Raloxifene (STAR) trial compared the efficacy <strong>of</strong> raloxifene<br />

over tamoxifen as a chemopreventive for women at high risk for breast cancer.<br />

Both drugs reduced the risk <strong>of</strong> invasive breast cancer by approximately 50%.<br />

Tamoxifen also lowered the incidence <strong>of</strong> lobular carcinoma in situ (LCIS) and<br />

ductal carcinoma in situ (DCIS) by about 50 percent; however, raloxifene did<br />

not have an effect on the incidence <strong>of</strong> LCIS and DCIS (47). Of note, the<br />

incidence <strong>of</strong> uterine cancer was 36% lower, and the incidence <strong>of</strong> thromboembolic<br />

events was 29% lower in the raloxifene arm (47). Fulvestrant and<br />

other SERDs have been shown to have high affinity for ERs compared with<br />

tamoxifen with none <strong>of</strong> the agonist activities. In clinical trials, fulvestrant<br />

initially showed significant promise for the treatment <strong>of</strong> advanced breast<br />

cancer (48,49). However, in a phase III clinical trial versus tamoxifen for firstline<br />

therapy <strong>of</strong> advanced breast cancer, fulvestrant was not superior to<br />

tamoxifen in terms <strong>of</strong> time-to-treatment failure (50). Therefore, more work is<br />

needed to evaluate the efficacy <strong>of</strong> SERDs in comparison with SERMs for ERapositive<br />

breast cancer.<br />

AROMATASE INHIBITORS<br />

An alternate strategy to endocrine therapy, which specifically inhibits binding <strong>of</strong><br />

E 2 to the ER, is to inhibit the production <strong>of</strong> E 2 by blocking the cytochrome p450<br />

aromatase enzyme, the rate-limiting enzyme that converts androgens (i.e., testosterone<br />

and androstenedione) to estrogens (i.e., E 2 and estrone) in the adrenal<br />

gland, surrounding stroma, and adipose tissue <strong>of</strong> the breast tumor. The main<br />

drugs <strong>of</strong> this type are aromatase inhibitors, which include Type I (steroidal) or<br />

Type II (nonsteroidal) (Fig. 2). Steroidal inhibitors are competitive-substrate<br />

mimics <strong>of</strong> androstenedione. These include formestane and exemestane, which<br />

are irreversible inhibitors that bind with high affinity to the binding site <strong>of</strong><br />

aromatase and are converted to a covalently bound intermediate. Nonsteroidal<br />

inhibitors include the first-generation aromatase inhibitor aminoglutethimide and<br />

the second-generation compounds anastrozole and letrozole. All nonsteroidal<br />

aromatase inhibitors act by binding reversibly to the enzyme and competitively<br />

inhibiting binding <strong>of</strong> the substrate androstenedione. The benefits <strong>of</strong> using aromatase<br />

inhibitors over tamoxifen are believed to be the complete deprivation <strong>of</strong><br />

E 2 and thus better efficacy for ERa-positive breast cancer (51). Recent clinical<br />

data have clearly demonstrated that anastrozole (52), letrozole (53), and<br />

exemestane (54) are more effective than tamoxifen as first-line treatments in<br />

patients with metastatic breast cancer. On the basis <strong>of</strong> clinical results (52,53),<br />

currently both anastrozole and letrozole are approved by the Food and Drug<br />

Administration for first-line treatment <strong>of</strong> postmenopausal, ERa-positive<br />

advanced breast cancer. The data from advanced breast cancer trials provided the<br />

rationale to perform large-scale clinical trials to determine whether there is an


Individualization <strong>of</strong> Endocrine Therapy in <strong>Breast</strong> <strong>Cancer</strong> 175<br />

Figure 2<br />

Structures <strong>of</strong> selective aromatase inhibitors.<br />

advantage <strong>of</strong> using aromatase inhibitors over tamoxifen in the adjuvant setting.<br />

The anastrozole, tamoxifen, and the combination <strong>of</strong> anastrozole and tamoxifen<br />

(ATAC) trial has sufficient follow-up data to confirm that anastrozole is superior<br />

to tamoxifen as a first-line adjuvant therapy in ERa-positive breast cancer with<br />

regard to disease-free survival and incidence <strong>of</strong> contralateral breast cancer (55).<br />

The <strong>Breast</strong> International Group (BIG) 1-98 study demonstrated a similar<br />

improvement in event-free survival to confirm that letrozole is superior to<br />

tamoxifen as first-line adjuvant hormonal therapy (56). Other trials have demonstrated<br />

an improved outcome for postmenopausal patients with early-stage<br />

breast cancer treated with two to three years <strong>of</strong> tamoxifen, followed by an<br />

aromatase inhibitor, compared with five years <strong>of</strong> tamoxifen (57,58). The BIG 1-98<br />

trial will help to determine whether these sequencing/switching approaches are<br />

equivalent to upfront aromatase inhibitors.<br />

Mechanisms <strong>of</strong> Resistance<br />

Clinically, patients that relapse after a previous response to tamoxifen usually<br />

have a clinical response to aromatase inhibitors (59,60). These results strongly<br />

indicate that the ERa continues to be expressed and is functional in breast tumors<br />

that are resistant to antiestrogens. However, although estrogen deprivation<br />

treatment might be more effective than tamoxifen in delaying resistance, eventually<br />

resistance to aromatase inhibitors will also develop. To date, it is unclear


176 Zelnak et al.<br />

whether similar mechanisms <strong>of</strong> actions that have been identified for tamoxifen<br />

resistance are also involved in resistance to aromatase inhibitors. The exact<br />

mechanisms contributing to aromatase inhibitor resistance has yet to be fully<br />

elucidiated. However, in vitro studies have identified mutations within the aromatase<br />

gene that confers resistance to aromatase inhibitors (61). These mutations<br />

have not yet been identified in human breast carcinomas (62). Other studies have<br />

demonstrated that estrogen deprivation supersensitizes the breast cancer cell to<br />

low levels <strong>of</strong> estrogen, thus creating a hypersensitive environment to overcome<br />

estrogen deprivation resulting in resistance (63–65). In addition, results suggest<br />

that there is increased cross talk between growth factor receptor signaling<br />

pathways and ERa. ERa has been shown to become activated and supersensitized<br />

by several different intracellular kinases, including MAPKs, insulin-like growth<br />

factors, and the PI3-K/Akt pathway (41,66–69). Therefore, the data suggest that<br />

ERa continues to be an integral part <strong>of</strong> the breast cancer cell signaling pathway<br />

even after resistance to aromatase inhibitors has developed.<br />

Role <strong>of</strong> Progesterone Receptor and HER2/Neu<br />

There is emerging evidence to suggest that ER-positive cancers that do not<br />

express the progesterone receptor (PR) and/or HER2/neu are somewhat intrinsically<br />

resistant to tamoxifen and perhaps hormonal therapy in general. Arpino et<br />

al. have demonstrated an increased relapse rate in patients with ER-positive, PRnegative<br />

cancers compared with ER-positive, PR-positive cancers, treated with<br />

tamoxifen (70). Patients treated with tamoxifen with ER-positive, HER2-positive<br />

metastatic breast cancers have a shorter time to treatment failure compared with<br />

ER-positive, HER2-negative cancers (71). In fact, Arpino et al. have demonstrated<br />

an increase in both HER1 and HER2 in ER-positive, PR-negative cancers<br />

compared to ER-positive, PR-positive cancers, suggesting an interplay between<br />

the ER and epidermal growth factor pathways (70).<br />

Two very small trials demonstrated a significantly increased clinical<br />

response rate in patients with ER-positive, HER2-positive cancers treated with<br />

preoperative aromatase inhibitors compared to preoperative tamoxifen (72,73).<br />

This led to a widely accepted hypothesis that aromatase inhibitors were a better<br />

choice than tamoxifen in patients with ER-positive, HER2-positive cancers.<br />

However, an analysis <strong>of</strong> the BIG-1-98 trial demonstrates that letrozole improves<br />

outcome compared to tamoxifen in both ER-positive, HER2-positive cancers<br />

(HR 0.68) and in ER-positive, HER2-negative cancers (HR 0.72) (74). A recent<br />

subanalysis <strong>of</strong> the ATAC trial demonstrated a significantly improved outcome in<br />

ER-positive, HER2-negative cancers (HR 0.66) but not in ER-positive, HER2-<br />

positive cancers (HR 0.92), but this may have been due to the small number <strong>of</strong><br />

patients in the HER-positive group (75). Are HER2-positive cancers somewhat<br />

resistant to not just tamoxifen but also to aromatase inhibitors? As outlined<br />

above, a recent trial randomized patients with HR-positive, HER2-positive<br />

metastatic breast cancers to anastrozole alone or to anastrozole plus trastuzumab


Individualization <strong>of</strong> Endocrine Therapy in <strong>Breast</strong> <strong>Cancer</strong> 177<br />

(76). Although there was no significant difference in overall survival, possibly<br />

because patients randomized to anastrozole alone could receive trastuzumab at<br />

disease progression, the time to progression was doubled from 2.4 months in the<br />

anastrozole-alone arm to 4.8 months in the combined arm (p ¼ 0.0016) (76). The<br />

clinical benefit rate in the combination arm was 42%—significantly higher than<br />

in the anastrozole alone arm. A trial that evaluated single-agent trastuzumab as<br />

first-line therapy for patients with HER2-positive cancers demonstrated a clinical<br />

benefit rate <strong>of</strong> 48% (77). This suggests the intriguing possibility that HRpositive,<br />

HER2-positive cancers are driven by the HER2 pathway, which renders<br />

the cancers partly resistant to hormonal therapies.<br />

An initial evaluation <strong>of</strong> the ATAC trial using case report forms revealed<br />

that TTR was longer for anastrozole in both ER-positive/PR-positive and ERpositive/PR-negative<br />

subgroups, but the benefit was more pronounced in the ERpositive/PR-negative<br />

subgroup [HR 0.84, 95% confidence interval (CI) 0.69–<br />

1.02 vs. 0.43, 95% CI 0.31–0.61] (78). Importantly, the ER and PR analyses were<br />

not performed centrally. More recently, a central analysis <strong>of</strong> about 2000 patients<br />

on the ATAC trial demonstrated similar improvements with the use <strong>of</strong> anastrozole<br />

compared to tamoxifen, regardless <strong>of</strong> PR status (HR anastrozole vs.<br />

tamoxifen 0.72 for ER-positive, PR-positive subgroups and 0.66 for ER-positive,<br />

PR-negative subgroups) (75). In the BIG-1-98 trial, similar benefits for letrozole<br />

compared to tamoxifen were seen in the ER-positive/PR-positive and ERpositive/PR-negative<br />

subgroups (74).<br />

On the basis <strong>of</strong> this data, decisions regarding whether to start a patient on<br />

tamoxifen or an aromatase inhibtor should not be based on PR or HER2 status.<br />

Further molecular pr<strong>of</strong>iling may help in the future in making decisions regarding<br />

optimal hormonal therapies.<br />

Role <strong>of</strong> Gene Expression Analysis<br />

Recent studies have shown that that gene expression pr<strong>of</strong>iling <strong>of</strong> breast tumors<br />

can be used to predict outcome. The initial gene expression signatures were<br />

developed using snap-frozen tissue, limiting their clinical application (79). An<br />

assay (Oncotype DX, Genomic Health, Inc., Redwood City, California, U.S.) has<br />

been developed and validated in women with early-stage, node-negative breast<br />

cancer using fixed, paraffin-embedded tissue. The 21-gene panel includes genes<br />

involved with proliferation, invasion, and hormone response and classifies a<br />

tumor as being low, intermediate, or high risk <strong>of</strong> recurrence based on the level <strong>of</strong><br />

expression (80). Additional studies have shown that a woman’s recurrence score<br />

also predicts response to chemotherapy: women with low recurrence scores<br />

derive little benefit from adjuvant chemotherapy, whereas women with high<br />

recurrence scores benefit significantly from chemotherapy in addition to hormonal<br />

therapy. Most strikingly are the findings that breast cancers with a high<br />

recurrence score obtain minimal, if any, benefit from five years <strong>of</strong> tamoxifen<br />

(81). Given the fact that these cancers likely have higher expression <strong>of</strong> HER2


178 Zelnak et al.<br />

and lower quantitative hormone receptors, this is in keeping with the data outlined<br />

above, suggesting that HER2-positive cancers have intrinsic resistance to<br />

tamoxifen. It is not known if women with intermediate recurrence scores benefit<br />

from chemotherapy in addition to hormonal therapy (82). The TAILORx (Trial<br />

Assigning Individualized Options for Treatment) trial is under way, randomizing<br />

early-stage breast cancer patients with intermediate recurrence scores to chemotherapy<br />

plus hormonal therapy versus hormonal therapy alone.<br />

FUTURE RESEARCH AND THERAPEUTICS<br />

It is clear that the central regulator <strong>of</strong> growth for ERa-positive breast tumors is<br />

ERa. ER signaling is evidently not isolated from other cellular signaling pathways.<br />

Research studies demonstrate that there is bidirectional cross talk between<br />

EGFR/HER2/neu, MAPK, PI3-K/Akt, and ERa, which ultimately affects the<br />

cellular response <strong>of</strong> a cell to estrogens and mitogens. ERa integrates numerous<br />

signals from hormones, growth factors, and intracellular kinases along with the<br />

array <strong>of</strong> coregulators to modulate cellular physiology and tumor pathology. The<br />

first generation antiestrogen/SERM, tamoxifen, has been effective at increasing<br />

overall survival <strong>of</strong> patients with ERa-positive breast cancer. However, the<br />

eventual development <strong>of</strong> resistance to tamoxifen is common because <strong>of</strong> the<br />

multiple signaling networks affecting the function <strong>of</strong> ERa. As a result <strong>of</strong> this<br />

limitation, intense research over the past 25 years has revealed the need for<br />

alternate treatment strategies to tamoxifen such as newer and better SERMs with<br />

less agonist activity in the uterus while being full antagonists in the breast.<br />

In addition to SERMs, other therapeutic approaches have emerged such as<br />

the use <strong>of</strong> aromatase inhibitors to prevent the synthesis <strong>of</strong> E 2 and thus deprive<br />

breast cancer cells and other cells <strong>of</strong> E 2 . The strategy <strong>of</strong> estrogen deprivation<br />

may also have consequences such as resistance, bone loss, and/or dementia. In<br />

the ATAC and BIG 1-98 adjuvant trials, changes in bone mineral density (BMD)<br />

and increased incidence <strong>of</strong> bone fractures were reported. Longer follow-up in<br />

these studies will help determine if the decrease in BMD stabilizes over time or<br />

if it continues throughout the course <strong>of</strong> treatment.<br />

Therapeutic opportunities also exist from research studies investigating the<br />

role <strong>of</strong> growth factor receptors EGFR and HER2/neu and their intracellular<br />

signaling communicators MAPK and PI3-K/Akt in the development <strong>of</strong> tamoxifen<br />

resistance. Trastuzumab, the HER2/neu-specific antibody, has been shown<br />

to be highly effective in treating breast tumors overexpressing HER2/neu.<br />

Moreover, research data indicate a role for HER2/neu in antiestrogen resistance,<br />

demonstrating a rationale for using trastuzumab to treat or prevent antiestrogenresistant<br />

breast cancer. EGFR inhibitors and other tyrosine kinase inhibitors may<br />

also prove to be beneficial in preventing resistance when given in combination<br />

with tamoxifen or other SERMs or SERDs such as fulvestrant.<br />

In addition to EGFR and HER2/neu as important modulators <strong>of</strong> tamoxifen<br />

action, dissecting the specific interrelationship between ERa and its coregulators


Individualization <strong>of</strong> Endocrine Therapy in <strong>Breast</strong> <strong>Cancer</strong> 179<br />

may open the door for novel therapeutics. An understanding <strong>of</strong> how ERamediated<br />

gene transcription is dysregulated in breast cancer and how this<br />

changes in resistance during endocrine therapy will hopefully lead to improved<br />

strategies for the use <strong>of</strong> current and future antiestrogen therapies.<br />

REFERENCES<br />

1. Cole MP, Jones CT, Todd ID. A new anti-oestrogenic agent in late breast cancer. An<br />

early clinical appraisal <strong>of</strong> ICI46474. Br J <strong>Cancer</strong> 1971; 25:270–275.<br />

2. Jensen EV, Jacobson HI. Basic guides to the mechanism <strong>of</strong> estrogen action. Recent<br />

Prog Horm Res 1962; 18:387–414.<br />

3. Jordan VC, Dowse LJ. Tamoxifen as an anti-tumour agent: effect on oestrogen<br />

binding. J Endocrinol 1976; 68:297–303.<br />

4. Fisher B, Redmond C, Brown A, et al. Adjuvant chemotherapy with and without<br />

tamoxifen in the treatment <strong>of</strong> primary breast cancer: 5-year results from the National<br />

Surgical Adjuvant <strong>Breast</strong> and Bowel Project Trial. J Clin Oncol 1986; 4:459–471.<br />

5. Fisher B, Costantino JP, Redmond CK, et al. Endometrial cancer in tamoxifentreated<br />

breast cancer patients: findings from the National Surgical Adjuvant <strong>Breast</strong><br />

and Bowel Project (NSABP) B-14. J Natl <strong>Cancer</strong> Inst 1994; 86:527–537.<br />

6. Gottardis MM, Robinson SP, Satyaswaroop PG, et al. Contrasting actions <strong>of</strong><br />

tamoxifen on endometrial and breast tumor growth in the athymic mouse. <strong>Cancer</strong><br />

Res 1988; 48:812–815.<br />

7. Muss HB, Case LD, Atkins JN, et al. Tamoxifen versus high-dose oral medroxyprogesterone<br />

acetate as initial endocrine therapy for patients with metastatic breast<br />

cancer: a Piedmont Oncology Association study. J Clin Oncol 1994; 12:1630–1638.<br />

8. Kuukasjarvi T, Kononen J, Helin H, et al. Loss <strong>of</strong> estrogen receptor in recurrent<br />

breast cancer is associated with poor response to endocrine therapy. J Clin Oncol<br />

1996; 14:2584–2589.<br />

9. Bachleitner-H<strong>of</strong>mann T, Pichler-Gebhard B, Rudas M, et al. Pattern <strong>of</strong> hormone<br />

receptor status <strong>of</strong> secondary contralateral breast cancers in patients receiving adjuvant<br />

tamoxifen. Clin <strong>Cancer</strong> Res 2002; 8:3427–3432.<br />

10. Johnston SR, Saccani-Jotti G, Smith IE, et al. Changes in estrogen receptor, progesterone<br />

receptor, and pS2 expression in tamoxifen-resistant human breast cancer.<br />

<strong>Cancer</strong> Res 1995; 55:3331–3338.<br />

11. Hopp TA, Fuqua SA. Estrogen receptor variants. J Mammary Gland Biol Neoplasia<br />

1998; 3:73–83.<br />

12. Murphy LC, Dotzlaw H, Leygue E, et al. The pathophysiological role <strong>of</strong> estrogen<br />

receptor variants in human breast cancer. J Steroid Biochem Mol Biol 1998; 65:175–180.<br />

13. Fuqua SA. The role <strong>of</strong> estrogen receptors in breast cancer metastasis. J Mammary<br />

Gland Biol Neoplasia 2001; 6:407–417.<br />

14. Jiang SY, Langan-Fahey SM, Stella AL, et al. Point mutation <strong>of</strong> estrogen receptor<br />

(ER) in the ligand-binding domain changes the pharmacology <strong>of</strong> antiestrogens in ERnegative<br />

breast cancer cells stably expressing complementary DNAs for ER. Mol<br />

Endocrinol 1992; 6:2167–2174.<br />

15. Liu H, Park WC, Bentrem DJ, et al. Structure-function relationships <strong>of</strong> the raloxifeneestrogen<br />

receptor-alpha complex for regulating transforming growth factor-alpha<br />

expression in breast cancer cells. J Biol Chem 2002; 277:9189–9198.


180 Zelnak et al.<br />

16. Schwartz JA, Zhong L, Deighton-Collins S, et al. Mutations targeted to a predicted helix<br />

in the extreme carboxyl-terminal region <strong>of</strong> the human estrogen receptor-alpha alter its<br />

response to estradiol and 4-hydroxytamoxifen. J Biol Chem 2002; 277:13202–13209.<br />

17. Goetz MP, Rae JM, Suman VJ, et al. <strong><strong>Ph</strong>armacogenetics</strong> <strong>of</strong> tamoxifen biotransformation<br />

is associated with clinical outcomes <strong>of</strong> efficacy and hot flashes.<br />

J. Clin Onc 2005; 23:9312–8.<br />

18. Mortimer J, Flatt S, Parker B, et al. Tamoxifen, hot flashes, and breast cancer<br />

recurrence: support for pharmacogenetics. Proc ASCO 2007 (abstr 500).<br />

19. Smith CL, Nawaz Z, and O’Malley BW. Coactivator and corepressor regulation <strong>of</strong><br />

the agonist/antagonist activity <strong>of</strong> the mixed antiestrogen, 4-hydroxytamoxifen. Mol<br />

Endocrinol 1997; 11:657–666.<br />

20. Osborne CK, Bardou V, Hopp TA, et al. Role <strong>of</strong> the estrogen receptor coactivator<br />

AIB1 (SRC-3) and HER-2/neu in tamoxifen resistance in breast cancer. J Natl<br />

<strong>Cancer</strong> Inst 2003; 95:353–361.<br />

21. Girault I, Tozlu S, Lidereau R, et al. Expression analysis <strong>of</strong> DNA methyltransferases<br />

1, 3A, and 3B in sporadic breast carcinomas. Clin <strong>Cancer</strong> Res 2003; 9:4415–4422.<br />

22. Kurokawa H, Arteaga CL. ErbB (HER) receptors can abrogate antiestrogen action in<br />

human breast cancer by multiple signaling mechanisms. Clin <strong>Cancer</strong> Res 2003;<br />

9:511S–515S.<br />

23. Yarden Y, Sliwkowski MX. Untangling the ErbB signalling network. Nat Rev Mol<br />

Cell Biol 2001; 2:127–137.<br />

24. Yarden Y. The EGFR family and its ligands in human cancer. signalling mechanisms<br />

and therapeutic opportunities. Eur J <strong>Cancer</strong> 2001; 37(suppl 4):S3–S8.<br />

25. Yarden Y. Biology <strong>of</strong> HER2 and its importance in breast cancer. Oncology 2001;<br />

61(suppl 2):1–13.<br />

26. Rubin I, Yarden Y. The basic biology <strong>of</strong> HER2. Ann Oncol 2001; 12(suppl 1):S3–S8.<br />

27. Spigel DR, Burstein HJ. HER2 overexpressing metastatic breast cancer. Curr Treat<br />

Options Oncol 2002; 3:163–174.<br />

28. Klijn JG, Berns PM, Schmitz PI, et al. The clinical significance <strong>of</strong> epidermal growth<br />

factor receptor (EGF-R) in human breast cancer: a review on 5232 patients. Endocr<br />

Rev 1992; 13:3–17.<br />

29. Klijn JG, Look MP, Portengen H, et al. The prognostic value <strong>of</strong> epidermal growth<br />

factor receptor (EGF-R) in primary breast cancer: results <strong>of</strong> a 10 year follow-up<br />

study. <strong>Breast</strong> <strong>Cancer</strong> Res Treat 1994; 29:73–83.<br />

30. Tang CK, Gong XQ, Moscatello DK, et al. Epidermal growth factor receptor vIII<br />

enhances tumorigenicity in human breast cancer. <strong>Cancer</strong> Res 2000; 60:3081–3087.<br />

31. Slamon DJ, Leyland-Jones B, Shak S, et al. Use <strong>of</strong> chemotherapy plus a monoclonal<br />

antibody against HER2 for metastatic breast cancer that overexpresses HER2.<br />

N Engl J Med 2001; 344:783–792.<br />

32. Nahta R, Hortobagyi GN, Esteva FJ. Growth factor receptors in breast cancer:<br />

potential for therapeutic intervention. Oncologist 2003; 8:5–17.<br />

33. Morris C. The role <strong>of</strong> EGFR-directed therapy in the treatment <strong>of</strong> breast cancer.<br />

<strong>Breast</strong> <strong>Cancer</strong> Res Treat 2002; 75(suppl 1):S51–S55, S57–S59 (discussion).<br />

34. Johnston SR, Head J, Pancholi S, et al. Integration <strong>of</strong> signal transduction inhibitors<br />

with endocrine therapy: an approach to overcoming hormone resistance in breast<br />

cancer. Clin <strong>Cancer</strong> Res 2003; 9:524S–532S.<br />

35. Kurokawa H, Lenferink AE, Simpson JF, et al. Inhibition <strong>of</strong> HER2/neu (erbB-2)<br />

and mitogen-activated protein kinases enhances tamoxifen action against


Individualization <strong>of</strong> Endocrine Therapy in <strong>Breast</strong> <strong>Cancer</strong> 181<br />

HER2-overexpressing, tamoxifen-resistant breast cancer cells. <strong>Cancer</strong> Res 2000; 60:<br />

5887–5894.<br />

36. Oh AS, Lorant LA, Holloway JN, et al. Hyperactivation <strong>of</strong> MAPK induces loss <strong>of</strong><br />

ERalpha expression in breast cancer cells. Mol Endocrinol 2001; 15:1344–1359.<br />

37. Kato S, Endoh H, Masuhiro Y, et al. Activation <strong>of</strong> the estrogen receptor through<br />

phosphorylation by mitogen-activated protein kinase. Science 1995; 270:1491–1494.<br />

38. Bunone G, Briand PA, Miksicek RJ, et al. Activation <strong>of</strong> the unliganded estrogen<br />

receptor by EGF involves the MAP kinase pathway and direct phosphorylation.<br />

Embo J 1996; 15:2174–2183.<br />

39. Fry MJ. <strong>Ph</strong>osphoinositide 3-kinase signalling in breast cancer: how big a role might<br />

it play? <strong>Breast</strong> <strong>Cancer</strong> Res 2001; 3:304–312.<br />

40. Castoria G, Migliaccio A, Bilancio A, et al. PI3-kinase in concert with Src promotes the<br />

S-phase entry <strong>of</strong> oestradiol-stimulated MCF-7 cells. Embo J 2001; 20: 6050–6059.<br />

41. Campbell RA, Bhat-Nakshatri P, Patel NM, et al. <strong>Ph</strong>osphatidylinositol 3-kinase/<br />

AKT-mediated activation <strong>of</strong> estrogen receptor alpha: a new model for anti-estrogen<br />

resistance. J Biol Chem 2001; 276:9817–9824.<br />

42. Rossouw JE, Anderson GL, Prentice RL, et al. Risks and benefits <strong>of</strong> estrogen plus<br />

progestin in healthy postmenopausal women: principal results From the Women’s<br />

Health Initiative randomized controlled trial. JAMA 2002; 288:321–333.<br />

43. Vogel CL, Shemano I, Schoenfelder J, et al. Multicenter phase II efficacy trial <strong>of</strong><br />

toremifene in tamoxifen-refractory patients with advanced breast cancer. J Clin<br />

Oncol 1993; 11:345–350.<br />

44. Johnston SR, Gumbrell LA, Evans TR, et al. A cancer research (UK) randomized<br />

phase II study <strong>of</strong> idoxifene in patients with locally advanced/metastatic breast cancer<br />

resistant to tamoxifen. <strong>Cancer</strong> Chemother <strong>Ph</strong>armacol 2004; 53:341–348.<br />

45. Pyrhonen S, Ellmen J, Vuorinen J, et al. Meta-analysis <strong>of</strong> trials comparing toremifene<br />

with tamoxifen and factors predicting outcome <strong>of</strong> antiestrogen therapy in postmenopausal<br />

women with breast cancer. <strong>Breast</strong> <strong>Cancer</strong> Res Treat 1999; 56: 133–143.<br />

46. Pagani O, Gelber S, Price K, et al. Toremifene and tamoxifen are equally effective<br />

for early stage breast cancer: first results <strong>of</strong> International <strong>Breast</strong> <strong>Cancer</strong> Study Group.<br />

Ann Oncol 2004; 15:1749–1759.<br />

47. Wickerham DL, Costantino JP, Vogel V, et al. The study <strong>of</strong> tamoxifen and raloxifene<br />

(STAR): initial findings from the NSABP P-2 breast cancer prevention study. ASCO<br />

Annual Meeting Proceedings Part I. J Clin Oncol 2006; 24(suppl 18):LBA5.<br />

48. Howell SJ, Johnston SR, Howell A. The use <strong>of</strong> selective estrogen receptor modulators<br />

and selective estrogen receptor down-regulators in breast cancer. Best Pract<br />

Res Clin Endocrinol Metab 2004; 18:47–66.<br />

49. Osborne CK, Wakeling A, Nicholson RI. Fulvestrant: an oestrogen receptor antagonist<br />

with a novel mechanism <strong>of</strong> action. Br J <strong>Cancer</strong> 2004; 90(suppl 1):S2–S6.<br />

50. Howell A, Robertson JF, Abram P, et al. Comparison <strong>of</strong> fulvestrant versus tamoxifen<br />

for the treatment <strong>of</strong> advanced breast cancer in postmenopausal women previously<br />

untreated with endocrine therapy: a multinational, double-blind, randomized trial.<br />

J Clin Oncol 2004; 22:1605–1613.<br />

51. Johnston SR, Dowsett M. Aromatase inhibitors for breast cancer: lessons from the<br />

laboratory. Nat Rev <strong>Cancer</strong> 2003; 3:821–831.<br />

52. Bonneterre J, Buzdar A, Nabholtz JM, et al. Anastrozole is superior to tamoxifen as<br />

first-line therapy in hormone receptor positive advanced breast carcinoma. <strong>Cancer</strong><br />

2001; 92:2247–2258.


182 Zelnak et al.<br />

53. Mouridsen H, Gershanovich M, Sun Y, et al. Superior efficacy <strong>of</strong> letrozole versus<br />

tamoxifen as first-line therapy for postmenopausal women with advanced breast<br />

cancer: results <strong>of</strong> a phase III study <strong>of</strong> the International Letrozole <strong>Breast</strong> <strong>Cancer</strong><br />

Group. J Clin Oncol 2001; 19:2596–2606.<br />

54. Paridaens R, Dirix L, Lohrisch C, et al. Mature results <strong>of</strong> a randomized phase II<br />

multicenter study <strong>of</strong> exemestane versus tamoxifen as first-line hormone therapy for<br />

postmenopausal women with metastatic breast cancer. Ann Oncol 2003; 14:1391–1398.<br />

55. Baum M, Budzar AU, Cuzick J, et al. Anastrozole alone or in combination with<br />

tamoxifen versus tamoxifen alone for adjuvant treatment <strong>of</strong> postmenopausal women<br />

with early breast cancer: first results <strong>of</strong> the ATAC randomised trial. Lancet 2002;<br />

359:2131–2139.<br />

56. Thurlimann B, Keshaviah A, Coates AS, et al. A comparison <strong>of</strong> letrozole and<br />

tamoxifen in postmenopausal women with early breast cancer. N Engl J Med 2005;<br />

353:2747–2757.<br />

57. Coombes RC, Kilburn LS, Snowdon CF, et al. Survival and safety <strong>of</strong> exemestane<br />

versus tamoxifen after 2 to 3 years tamoxifen treatment (Intergroup Exenestane<br />

Study): a randomized controlled trial. Lancet 2007; 369:559–570.<br />

58. Kaufmann M, Jonat W, Hilfrich J, et al. Improved overall survival in postmenopausal<br />

women with early breast cancer after anastrozole initiated after treatment with<br />

tamoxifen compared with continued tamoxifen: the ARNO 95 Study. J Clin Oncol<br />

2007; 25:2664–2670.<br />

59. Buzdar A, Jonat W, Howell A, et al. Anastrozole, a potent and selective aromatase<br />

inhibitor, versus megestrol acetate in postmenopausal women with advanced breast<br />

cancer: results <strong>of</strong> overview analysis <strong>of</strong> two phase III trials. Arimidex Study Group.<br />

J Clin Oncol 1996; 14:2000–2011.<br />

60. Dombernowsky P, Smith I, Falkson G, et al. Letrozole, a new oral aromatase<br />

inhibitor for advanced breast cancer: double-blind randomized trial showing a dose<br />

effect and improved efficacy and tolerability compared with megestrol acetate. J Clin<br />

Oncol 1998; 16:453–461.<br />

61. Kao YC, Cam LL, Laughton CA, et al. Binding characteristics <strong>of</strong> seven inhibitors<br />

<strong>of</strong> human aromatase: a site-directed mutagenesis study. <strong>Cancer</strong> Res 1991; 56:<br />

3451–3460.<br />

62. Sourdaine P, Parker MG, Telford J, et al. Analysis <strong>of</strong> the aromatase cytochrome P450<br />

gene in human breast cancers. J Mol Endocrinol 1994; 13:331–337.<br />

63. Masamura S, Santner SJ, Heitjan DF, et al. Estrogen deprivation causes estradiol<br />

hypersensitivity in human breast cancer cells. J Clin Endocrinol Metab 1995;<br />

80:2918–2925.<br />

64. Santen R, Jeng MH, Wang JP, et al. Adaptive hypersensitivity to estradiol: potential<br />

mechanism for secondary hormonal responses in breast cancer patients. J Steroid<br />

Biochem Mol Biol 2001; 79:115–125.<br />

65. Chan CM, Martin LA, Johnston SR, et al. Molecular changes associated with the<br />

acquisition <strong>of</strong> oestrogen hypersensitivity in MCF-7 breast cancer cells on long-term<br />

oestrogen deprivation. J Steroid Biochem Mol Biol 2002; 81:333–341.<br />

66. Martin LA, Farmer I, Johnston SR, et al. Enhanced estrogen receptor (ER) alpha,<br />

ERBB2, and MAPK signal transduction pathways operate during the adaptation<br />

<strong>of</strong> MCF-7 cells to long term estrogen deprivation. J Biol Chem 2003; 278:<br />

30458–30468.


Individualization <strong>of</strong> Endocrine Therapy in <strong>Breast</strong> <strong>Cancer</strong> 183<br />

67. Jeng MH, Yue W, Eischeid A, et al. Role <strong>of</strong> MAP kinase in the enhanced cell<br />

proliferation <strong>of</strong> long term estrogen deprived human breast cancer cells. <strong>Breast</strong><br />

<strong>Cancer</strong> Res Treat 2000; 62:167–175.<br />

68. Shim WS, Conaway M, Masamura S, et al. Estradiol hypersensitivity and mitogenactivated<br />

protein kinase expression in long-term estrogen deprived human breast<br />

cancer cells in vivo. Endocrinology 2000; 141:396–405.<br />

69. Stephen RL, Shaw LE, Larsen C, et al. Insulin-like growth factor receptor levels are<br />

regulated by cell density and by long term estrogen deprivation in MCF7 human<br />

breast cancer cells. J Biol Chem 2001; 276:40080–40086.<br />

70. Arpino G, Weiss H, Lee AV, et al. Estrogen receptor-positive, progesterone receptornegative<br />

breast cancer: association with growth factor receptor expression and<br />

tamoxifen resistance. J Natl <strong>Cancer</strong> Inst 2005; 97:1254–1261.<br />

71. De Laurentiis M, Arpino G, Massarelli E, et al. A meta-analysis on the interaction<br />

between HER-2 expression and response to endocrine treatment in advanced breast<br />

cancer. Clin <strong>Cancer</strong> Res 2005; 11:4741–4748.<br />

72. Ellis MJ, Tao Y, Young O, et al. Estrogen-independent proliferation is present in<br />

estrogen-receptor HER2-positive primary breast cancer after neoadjuvant letrozole.<br />

J Clin Oncol 2006; 24:3019–3025.<br />

73. Smith IE, Dowsett M, Ebbs SR, et al. Neoadjuvant treatment <strong>of</strong> postmenopausal<br />

breast cancer with anastrozole, tamoxifen, or both in combination: the Immediate<br />

Preoperative Anastrozole, Tamoxifen, or Combined with Tamoxifen (IMPACT)<br />

multicenter double-blind randomized trial. J Clin Oncol 2005; 23:5108–5116.<br />

74. Viale G, Regan MM, Maiorano E, et al. Prognostic and predictive value <strong>of</strong> centrally<br />

reviewed expression <strong>of</strong> estrogen and progesterone receptors in a randomized trial<br />

comparing letrozole and tamoxifen adjuvant therapy for postmenopausal early stage<br />

breast cancer: BIG 1-98. J Clin Oncol 2007; 25:3846–3852.<br />

75. Dowsett M, Allred DC, for the TransATAC investigators. Relationship between<br />

quantitative ER and PgR expression and HER2 status with recurrence in the ATAC<br />

trial. SABCS 2006 (abstr 48).<br />

76. Mackey JR, Kaufman B, Clemens M, et al. Trastuzumab prolongs progression-free<br />

survival in hormone-dependent and HER2-positive metastatic breast cancer. SABCS<br />

2006 (abstr 3).<br />

77. Vogel CL, Cobleigh MA, Tripathy D, et al. Efficacy and safety <strong>of</strong> trastuzumab as a<br />

single agent in first-line treatment <strong>of</strong> HER2-overexpressing metastatic breast cancer.<br />

J Clin Oncol 2002; 20:719–726.<br />

78. Dowsett M, Cuzick J, Wale C, et al. Retrospective analysis <strong>of</strong> time to recurrence in<br />

the ATAC trial according to hormone receptor status: an hypothesis-generating<br />

study. J Clin Oncol 2005; 23:7512–7517.<br />

79. Van de Vijver MJ, He YD, van ’t Veer LJ, et al. A gene-expression signature as a<br />

predictor <strong>of</strong> survival in breast cancer. N Eng J Med 2002; 347:1999–2009.<br />

80. Paik S, Shak S, Tang G, et al. A multigene assay to predict recurrence <strong>of</strong> tamoxifentreated,<br />

node-negative breast cancer. N Engl J Med 2004; 351:2817–2826.<br />

81. Paik S, Shak S, Tang G, et al. Expression <strong>of</strong> the 21 genes in the recurrence score<br />

assay and prediction <strong>of</strong> clinical benefit from tamoxifen in NSABP study B-14 and<br />

chemotherapy in NSABP study B-20. <strong>Breast</strong> <strong>Cancer</strong> Res Treat 2004; 88 (abstr 24).<br />

82. Paik S, Tang G, Shak S, et al. Gene expression and benefit <strong>of</strong> chemotherapy in<br />

women with node-negative, estrogen receptor-positive breast cancer. J Clin Oncol<br />

2006; 24:3796–3834.


13<br />

TAILORx: Rationale for the Study Design<br />

Joseph A. Sparano<br />

Albert Einstein College <strong>of</strong> Medicine, Montefiore Medical Center,<br />

Bronx, New York, U.S.A.<br />

INTRODUCTION<br />

There have been substantial technical and analytical advances within the past<br />

decade that have facilitated high-throughput analysis <strong>of</strong> clinical specimens, a<br />

process that has been referred to as “molecular pr<strong>of</strong>iling.” This term, as applied<br />

here, may refer to evaluating various “markers,” including genomic, proteomic,<br />

and epigenomic expression patterns, or a combination <strong>of</strong> these patterns, in<br />

clinical specimens. Technological advances have led to the ability to measure<br />

thousands <strong>of</strong> markers with a variety <strong>of</strong> validated methods (1), and analytical<br />

models have been developed that facilitate analysis <strong>of</strong> the voluminous amount <strong>of</strong><br />

data that are generated (2). The technology has also led to an ability to perform<br />

“discovery-based research,” in which large volumes <strong>of</strong> data are generated and<br />

analyzed without a specific hypothesis, in contrast to the traditional scientific<br />

paradigm <strong>of</strong> “hypothesis-based research,” in which a limited number <strong>of</strong> genes or<br />

proteins are based upon a specific hypothesis and rationale (3). Discovery-based<br />

research and hypothesis-based research are not mutually exclusive; however,<br />

pr<strong>of</strong>iling may also be used to test specific hypotheses that are based on sound,<br />

scientific rationale. A series <strong>of</strong> studies have been reported over the past few<br />

years that have utilized gene expression pr<strong>of</strong>iling to discover “molecular<br />

markers,” which may identify (i) distinct molecular subtypes <strong>of</strong> breast cancer<br />

185


186 Sparano<br />

(i.e., genotypic-phenotypic correlation), (ii) molecular signatures–associated<br />

prognosis (i.e., prognostic factors), and (iii) molecular signatures that predict<br />

the benefit from specific therapies (i.e., predictive factors) (4). When a<br />

“molecular marker” is shown to perform more reliably than clinical features in<br />

predicting prognosis or the benefit from a specific intervention, then there is<br />

interest in further evaluating the utility <strong>of</strong> the marker in clinical practice (5).<br />

PROGRAM FOR THE ASSESSMENT OF CLINICAL CANCER TESTS<br />

In order to address the issue <strong>of</strong> integrating molecular diagnostic testing into<br />

clinical practice, the U.S. National <strong>Cancer</strong> Institute initiated the program for the<br />

assessment <strong>of</strong> clinical cancer tests (PACCT) (http://www.cancerdiagnosis.nci<br />

.nih.gov/assessment/index.html). The goals <strong>of</strong> PACCT are to ensure translation<br />

<strong>of</strong> new knowledge about cancer and new technologies to clinical practice and<br />

development <strong>of</strong> more informative laboratory tools to help maximize the impact<br />

<strong>of</strong> cancer treatments. This effort has led to the publication <strong>of</strong> the REMARK<br />

(REporting recommendations for tumour MARKer prognostic studies)<br />

guidelines (6) and to the TAILORx (Trial Assigning IndividuaLized Options<br />

for Treatment) trial developed by the North American <strong>Breast</strong> <strong>Cancer</strong> Intergroup<br />

(http://www.cancer.gov/clinicaltrials/digestpage/TAILORx). In reviewing<br />

the rationale for design <strong>of</strong> the TAILORx trial in breast cancer, the following<br />

questions were considered by the <strong>Breast</strong> <strong>Cancer</strong> Intergroup and are reviewed<br />

herein:<br />

l<br />

l<br />

l<br />

l<br />

What therapeutic intervention should the marker be used to select?<br />

What patient population should be evaluated?<br />

Which marker should be evaluated?<br />

What trial design would integrate current knowledge about the marker and<br />

address gaps in our knowledge about the current marker and future potential<br />

markers?<br />

WHICH TREATMENT INTERVENTION?<br />

Therapeutic options are used in an adjunctive manner for patients with earlystage<br />

breast cancer after surgery, irradiation, endocrine therapy, chemotherapy,<br />

and trastuzumab (Table 1). In contrast to irradiation, endocrine therapy, and<br />

trastuzumab, there are no clinical factors that may be used to predict the benefit<br />

from adjuvant chemotherapy. Chemotherapy is usually recommended on the<br />

basis <strong>of</strong> the risk <strong>of</strong> recurrence (i.e., prognostic factors) and age (7). In contrast<br />

to other therapies, the treatment effect <strong>of</strong> chemotherapy is modest, acute toxicities<br />

are frequent, and the cost is high (8). Therefore, evaluation <strong>of</strong> a marker<br />

that has the potential for predicting benefit from chemotherapy would be<br />

desirable.


TAILORx 187<br />

Table 1<br />

Standard Adjuvant Therapies for Operable <strong>Breast</strong> <strong>Cancer</strong><br />

Treatment<br />

Selection<br />

Percent<br />

eligible<br />

Treatment<br />

effect (%)<br />

Acute<br />

toxicity<br />

Relative<br />

cost<br />

Chemotherapy Recurrence risk Up to 90 25–35 High þþþ<br />

> 5–10%<br />

Endocrine ER- and/or PRpositive<br />

70 50 Low þþ<br />

therapy<br />

tumor<br />

Trastuzumab HER2/neu-positive 15 50 Low þþþþ<br />

tumor<br />

Radiation <strong>Breast</strong>-sparing surgery 40–60 90 Low þ<br />

Abbreviations: ER, estrogen receptor; PR, progesterone receptor; HER2, human epidermal growth<br />

factor receptor 2.<br />

WHICH PATIENT POPULATION?<br />

In 2006, approximately 124,000 women in the United States were diagnosed<br />

with estrogen receptor (ER)–positive breast cancer associated with negative<br />

axillary lymph nodes, accounting for about 50% <strong>of</strong> all breast cancer diagnosed<br />

each year and nearly 9% <strong>of</strong> all new cases <strong>of</strong> cancer (9). Treatment recommendations<br />

have generally been based on prognostic factors, with chemotherapy<br />

recommended if the residual risk <strong>of</strong> recurrence exceeds approximately 5% to<br />

10% despite adjuvant hormonal therapy (i.e., the tumor size exceeds 1 cm or<br />

smaller tumors are associated with unfavorable histologic features) (7). By these<br />

criteria, adjuvant chemotherapy could be recommended annually for up to<br />

75,000 women younger than 70 years in the United States. Currently, it is estimated<br />

that approximately 25% <strong>of</strong> all women, or about 30,000 annually, with earlystage<br />

ER-positive breast cancer receive adjuvant chemotherapy. Its use has<br />

increased substantially over the past 15 years (10) is highly dependent on age, and<br />

is estimated to be given to 75% <strong>of</strong> those younger than 50 years, 30% between the<br />

ages <strong>of</strong> 50 and 69 years, and about 5% <strong>of</strong> those 70 years or older.<br />

Approximately 85% <strong>of</strong> women with early-stage ER-positive breast cancer<br />

are therefore adequately treated with adjuvant hormonal therapy. Although<br />

adding chemotherapy reduces the risk <strong>of</strong> recurrence on average by about 30%,<br />

the absolute benefit for an individual patient is small, ranging from 1% to 5%<br />

(11). The vast majority <strong>of</strong> patients with ER-positive breast cancer are therefore<br />

overtreated with chemotherapy, since most would have been cured with hormonal<br />

therapy alone. Decision aids such as Adjuvant! Online (12), decision<br />

boards (13), and other tools, are <strong>of</strong>ten useful to assist patients and caregivers with<br />

information regarding absolute benefits that might be expected from chemotherapy<br />

(14). Although such decision aids may assist some patients in making a<br />

more informed decision regarding whether to accept adjuvant chemotherapy,<br />

when faced with a choice, many patients and their clinicians err on the side <strong>of</strong><br />

overtreatment because <strong>of</strong> the imprecise nature <strong>of</strong> predicting the treatment benefit


188 Sparano<br />

(15). Therefore, developing a clinical trial that employs a molecular marker to<br />

define the chemotherapy benefit was judged by the <strong>Breast</strong> <strong>Cancer</strong> Intergroup to<br />

be the preferred setting. Recent evidence demonstrating that the incremental<br />

benefits associated with contemporary chemotherapy regimens had minimal<br />

impact in high-risk ER-positive disease reinforces this approach (16).<br />

WHICH MARKER?<br />

A number <strong>of</strong> molecular diagnostic tests that provide more accurate prognostic<br />

information that standard clinical features have been developed and externally<br />

validated and are summarized in Table 2, including the Amsterdam 70-gene<br />

pr<strong>of</strong>ile (17,18), Rotterdam 76-gene pr<strong>of</strong>ile (19,20), and a 21-gene pr<strong>of</strong>ile (21,22).<br />

Recently, these assays have been shown to have comparable discriminatory<br />

value in early-stage breast cancer (23).<br />

The 21-gene Oncotype DX TM (Genomic Health, Inc., Redwood City,<br />

California U.S.) includes 16 tumor genes and 5 reference genes, with the result<br />

expressed as a computed “recurrence score” (RS) (Fig. 1) (21). The genes<br />

utilized in the assay were identified in several training set trials (24), followed by<br />

a validation study that was performed in patients with ER-positive, node-negative<br />

breast cancer who received a five-year course <strong>of</strong> tamoxifen (TAM) in<br />

National Surgical Adjuvant <strong>Breast</strong> and Bowel Project (NSABP) trial B-14 (21).<br />

A higher RS was associated with an elevated risk <strong>of</strong> distant recurrence, whether<br />

evaluated as a categorical or a continuous variable (Fig. 2). This assay was<br />

selected for evaluation in the TAILORx trial for several scientific and practical<br />

reasons that are summarized in Table 3 and discussed herein. In addition to the<br />

prospective validation study illustrated by Figure 2, a population-based external<br />

validation study performed in patients treated in the Kaiser Permanente health<br />

care system provided further evidence for the robustness <strong>of</strong> the assay (22).<br />

Moreover, subsequent studies in patients treated with TAM with or without<br />

Table 2<br />

Externally Validated Multigene <strong>Breast</strong> <strong>Cancer</strong> Markers<br />

Number<br />

Population<br />

Commercially<br />

Refs. <strong>of</strong> genes Tissue ER/PR Nodes Adjuvant therapy available<br />

27,28 534 Frozen þ/– þ/– Preoperative chemo Not available<br />

17,18 70 Frozen þ/– þ/– þ/– Chemo; MammaPrint 1<br />

þ/– hormones<br />

21,22 21 Paraffin þ TAM Oncotype DX<br />

embedded<br />

19,20 76 Frozen þ/– No therapy Not available<br />

Abbreviations: ER, estrogen receptor; PR, progesterone receptor; Chemo, chemotherapy; TAM,<br />

tamoxifen.


TAILORx 189<br />

Figure 1<br />

Oncotype DX RS: genes and algorithm. Abbreviation: RS, recurrence score.<br />

Figure 2 Oncotype DX RS predicts distant recurrence as either a categorical variable<br />

(left) or continuous variable (right). Abbreviation: RS, recurrence score.<br />

adjuvant chemotherapy in NSABP trial B20 indicated that only patients with an<br />

elevated RS (RS > 30) derived benefit from adjuvant chemotherapy (25). Other<br />

analyses also indicated that an RS also predicts local recurrence, which may<br />

occur in up to 20% <strong>of</strong> patients with operable breast cancer and which may be<br />

reduced by adjuvant chemotherapy (26). Furthermore, an RS more accurately<br />

predicted recurrence than the Adjuvant! Online model, providing evidence that it<br />

<strong>of</strong>fers information that is complementary to Adjuvant! (12). Finally, although<br />

Oncotoype DX includes a reverse transcriptase polymerase chain reaction<br />

(RT-PCR) assay-based evaluation <strong>of</strong> ER expression, a tumor with high ER<br />

content by RT-PCR may have either a low or a high RS, indicating that it reflects<br />

a measure other than simply ER expression (25).


190 Sparano<br />

Table 3<br />

Scientific Rationale for Selecting Oncotype DX Assay for TAILORx<br />

Rationale<br />

Ref.<br />

Validation study demonstrating that an RS is a prognostic factor for risk <strong>of</strong> distant 21<br />

recurrence in tamoxifen-treated patients, which outperforms clinical criteria in<br />

the multivariate model<br />

External population-based validation study demonstrating an RS predictive <strong>of</strong> 22<br />

breast cancer death<br />

A high RS predictive <strong>of</strong> chemotherapy benefit in tamoxifen-treated patients 25<br />

An RS correlating with risk <strong>of</strong> local recurrence 26<br />

An RS predicting the outcome more accurately than Adjuvant! Online 29<br />

Concordance in the predictive value <strong>of</strong> Oncotype DX compared with other 23<br />

multigene markers<br />

Abbreviation: RS, recurrence score.<br />

There were also several practical reasons for selecting Oncotype DX. First<br />

and foremost is the fact that it involves an RT-PCR-based technique that may be<br />

employed on as little as three 10-mm sections <strong>of</strong> routinely processed and stored<br />

formalin-fixed, paraffin-embedded tissue. In addition, the assay was approved by<br />

Clinical Laboratory Improvement Amendments (CLIA) in the United States and<br />

has received favorable local coverage decisions regarding reimbursement by<br />

Medicare, indicating acceptance <strong>of</strong> the value <strong>of</strong> the assay by third-party payers.<br />

Finally, selecting Oncotype DX for further evaluation by the <strong>Breast</strong> <strong>Cancer</strong><br />

Intergroup represents a logical extension <strong>of</strong> the successful public-private collaboration<br />

between NSABP and Genomic Health.<br />

WHAT TRIAL DESIGN?<br />

In designing a trial integrating a molecular diagnostic test into clinical decision<br />

making, three factors must be considered. First, it is essential to incorporate<br />

features already known about the prognostic and predictive value <strong>of</strong> the test into<br />

the trial design. In this regard, it is clear that patients with a very low RS do very<br />

well with hormonal therapy alone, and patients with a very high RS derive great<br />

benefit from chemotherapy, thereby providing sufficient information to make<br />

definitive treatment recommendations for patients who fall into these groups.<br />

Second, the trial must be designed to address the clinical utility <strong>of</strong> the test in<br />

circumstances where its utility is uncertain or result uninformative. This scenario<br />

applies to patients with a tumor that has a mid-range RS, in which the risk <strong>of</strong><br />

relapse may be sufficiently high to recommend chemotherapy, but in which<br />

chemotherapy may be not be beneficial. Third, it is important to design the trial<br />

in a manner that <strong>of</strong>fers an opportunity to evaluate other clinical cancer tests as<br />

they evolve, obviating the need to repeat a large-scale clinical trial and providing<br />

an opportunity to evaluate the test within a very short time frame.


TAILORx 191<br />

Figure 3<br />

TAILORx schema. Abbreviation: RS, recurrence score.<br />

The factors outlined in the previous paragraph were taken into account<br />

while designing TAILORx. Only patients who meet established clinical criteria<br />

for recommending adjuvant chemotherapy, who are medically suitable candidates<br />

for chemotherapy, and who agree to have their treatment assigned or<br />

randomized on the basis <strong>of</strong> the molecular test are eligible. After informed consent<br />

and “preregistration,” a tumor specimen is forwarded to Genomic Health for the<br />

Oncotype DX assay, with the result reported to the site within 10 to 14 days. After<br />

the Oncotype DX RS is known, patients are then registered and have their treatment<br />

assigned or randomized on the basis <strong>of</strong> the RS (Fig. 3), as described below:<br />

l<br />

l<br />

l<br />

RS is 10 or less: Hormonal therapy alone<br />

RS is 26 or higher: Chemotherapy plus hormonal therapy<br />

RS is 11 to 25: Patients randomized to receive chemotherapy plus hormonal<br />

therapy (the standard treatment arm) versus hormonal therapy alone<br />

(the experimental treatment arm)<br />

Because the study includes only women who meet standard clinical criteria<br />

for adjuvant chemotherapy and are medically suitable candidates for therapy,<br />

chemotherapy plus hormonal therapy is considered the standard treatment arm<br />

for patients with a mid-range RS (i.e., 11–25). Hormonal therapy is considered<br />

the experimental treatment arm for the mid-range group. The primary objective


192 Sparano<br />

<strong>of</strong> the trial is to determine whether hormonal therapy alone is not inferior to<br />

chemotherapy plus hormonal therapy. The trial utilizes a “noninferiority” design<br />

for the mid-range group and is adequately powered to detect a 3% or greater<br />

difference between the two treatment arms.<br />

The chemotherapy regimen and hormonal agents used are prescribed at the<br />

discretion <strong>of</strong> the treating physician but must be consistent with one <strong>of</strong> several<br />

standard options described in the protocol document. Patients may also enroll in<br />

other Intergroup trials as long as the protocol treatment is consistent with the<br />

TAILORx-assigned treatment arm (i.e., hormonal therapy or chemotherapy plus<br />

hormonal therapy). In addition, patients who have already had the Oncotype DX<br />

assay performed are eligible to enroll if the RS is 11 to 25.<br />

At the time <strong>of</strong> registration and treatment assignment, patients are asked to<br />

provide blood samples for banking <strong>of</strong> plasma and peripheral blood mononuclear<br />

cells. Tissue specimens are collected by the Eastern Cooperative Oncology<br />

Group (ECOG) Pathology Coordinating Office, with central testing and confirmation<br />

<strong>of</strong> ER and progesterone receptor (PR) expression, creation <strong>of</strong> tissue<br />

microarrays, and RNA extraction for future studies.<br />

The trial is available through the <strong>Cancer</strong> Trials Support Unit (www.ctsu.org)<br />

and is being coordinated by ECOG. Other participants include NSABP and all<br />

members <strong>of</strong> the North American <strong>Breast</strong> Intergroup, including the Southwest<br />

Oncology Group (SWOG), <strong>Cancer</strong> and Acute Leukemia Group B (CALGB), North<br />

Central <strong>Cancer</strong> Treatment Group (NCCTG), National <strong>Cancer</strong> Institute <strong>of</strong> Canada<br />

(NCIC), and the American College <strong>of</strong> Surgeons Oncology Group (ACOSOG).<br />

RATIONALE FOR THE RS RANGES IN TAILORx<br />

The RS ranges used in TAILORx are different from those originally described<br />

for the low- (30) risk groups as<br />

originally reported for the assay. The range was adjusted in order to minimize<br />

the potential for undertreatment in both the high-risk group (also called “secondary<br />

study group 2”) and the randomized group (also called the “primary study<br />

group”). When the NSABP B20 data were analyzed using the RS ranges used in<br />

TAILORx, the treatment effect <strong>of</strong> chemotherapy was similar to the original<br />

analysis (Table 4). For the analyses described in Table 4, disease-free survival<br />

(DFS) is defined as time to first local, regional, or distant recurrence, second<br />

primary cancer, or death due to any cause, and distant recurrence-free survival<br />

(DRFS) is defined as time to first distant recurrence or death from breast cancer<br />

(contralateral breast cancers, other second primary cancers, and deaths due to<br />

other cancers are censored, and local and regional recurrences are ignored). DFS<br />

has been the primary endpoint evaluated in most breast cancer trials and is the<br />

primary endpoint for TAILORx, whereas DRFS was the primary endpoint in<br />

the NSABP trials evaluating RS. Although a trend favoring the addition <strong>of</strong><br />

chemotherapy becomes evident at an RS <strong>of</strong> approximately 11 when the risk <strong>of</strong><br />

relapse is analyzed in a linear fashion, the 95% confidence intervals completely


TAILORx 193<br />

Table 4 RS and Response to Chemotherapy in B20 Trial (n ¼ 651) by RS<br />

RS<br />

Number <strong>of</strong><br />

patients<br />

TAM 10-yr<br />

DRFS (%)<br />

TAM þ<br />

chemo<br />

10-yr<br />

DRFS (%)<br />

HR (95% CI) for<br />

recurrence by<br />

addition <strong>of</strong><br />

chemo P-value<br />

TAM<br />

10-yr<br />

DFS (%)<br />

TAM þ<br />

chemo<br />

10-yr<br />

DFS (%)<br />

HR (95% CI) for<br />

recurrence by<br />

addition <strong>of</strong> chemo P-value<br />

25 195 (30%) 63 88 0.285 (0.148, 0.551)


194 Sparano<br />

Figure 4 Treatment effect <strong>of</strong> chemotherapy by Oncotype DX RS. Abbreviation: RS,<br />

recurrence score.<br />

overlap in the 11 to 25 RS range (Fig. 4). Finally, a high RS has also been<br />

associated with an elevated risk <strong>of</strong> local relapse, which may also be reduced by<br />

giving adjuvant chemotherapy (26). An RS <strong>of</strong> 11 is associated with a risk <strong>of</strong> both<br />

local and distant relapse <strong>of</strong> about 10%, a threshold that has been typically used<br />

for recommending adjuvant chemotherapy.<br />

CONCLUSION<br />

TAILORx represents a major step forward into the era <strong>of</strong> personalized medicine<br />

for breast cancer. By integrating a molecular diagnostic test into clinical decision<br />

making, patients and clinicians will be able to make more informed decisions<br />

regarding the most appropriate treatment options for a subset <strong>of</strong> patients for<br />

whom the test results in a clear treatment path and refine the utility <strong>of</strong> the assay<br />

for those for whom the result may be uninformative. This trial will also serve as<br />

an important resource for evaluating new molecular signatures and other technologies,<br />

such as proteomics, epigenomics, and pharmacogenomics, as these<br />

technologies evolve.<br />

REFERENCES<br />

1. Hanash S. Integrated global pr<strong>of</strong>iling <strong>of</strong> cancer. Nat Rev <strong>Cancer</strong> 2004; 4:638–644.<br />

2. Simon R. Diagnostic and prognostic prediction using gene expression pr<strong>of</strong>iles in<br />

high-dimensional microarray data. Br J <strong>Cancer</strong> 2003; 89:1599–1604.


TAILORx 195<br />

3. Ransoh<strong>of</strong>f DF. Evaluating discovery-based research: when biologic reasoning cannot<br />

work. Gastroenterology 2004; 127:1028.<br />

4. Sparano JA, Fazzari MJ, Childs G. Clinical application <strong>of</strong> molecular pr<strong>of</strong>iling in<br />

breast cancer. Future Oncol 2005; 1:485–496.<br />

5. Hayes DF, Bast RC, Desch CE, et al. Tumor marker utility grading system: a framework<br />

to evaluate clinical utility <strong>of</strong> tumor markers. J Natl <strong>Cancer</strong> Inst 1996; 88: 1456–1466.<br />

6. McShane LM, Altman DG, Sauerbrei W, et al. REporting recommendations for<br />

tumour MARKer prognostic studies (REMARK). Eur J <strong>Cancer</strong> 2005; 41:1690–1696.<br />

7. Carlson RW, Brown E, Burstein HJ, et al. NCCN Task Force Report: Adjuvant<br />

Therapy for <strong>Breast</strong> <strong>Cancer</strong>. J Natl Compr Canc Netw 2006; 4(suppl 1):S1–S26.<br />

8. Hassett MJ, O’Malley AJ, Pakes JR, et al. Frequency and cost <strong>of</strong> chemotherapyrelated<br />

serious adverse effects in a population sample <strong>of</strong> women with breast cancer.<br />

J Natl <strong>Cancer</strong> Inst 2006; 98:1108–1117.<br />

9. Jemal A, Siegel R, Ward E, et al. <strong>Cancer</strong> statistics, 2006. CA <strong>Cancer</strong> J Clin 2006;<br />

56:106–130.<br />

10. Harlan LC, Clegg LX, Abrams J, et al. Community-based use <strong>of</strong> chemotherapy and<br />

hormonal therapy for early-stage breast cancer: 1987-2000. J Clin Oncol 2006; 24:<br />

872–877.<br />

11. Fisher B, Jeong JH, Dignam J, et al. Findings from recent National Surgical Adjuvant<br />

<strong>Breast</strong> and Bowel Project adjuvant studies in stage I breast cancer. J Natl <strong>Cancer</strong> Inst<br />

Monogr 2001; 30:62–66.<br />

12. Olivotto IA, Bajdik CD, Ravdin PM, et al. Population-based validation <strong>of</strong> the prognostic<br />

model ADJUVANT! for early breast cancer. J Clin Oncol 2005; 23:2716–2725.<br />

13. Whelan T, Levine M, Willan A, et al. Effect <strong>of</strong> a decision aid on knowledge and<br />

treatment decision making for breast cancer surgery: a randomized trial. JAMA<br />

2004; 292:435–441.<br />

14. Whelan TJ, Loprinzi C. <strong>Ph</strong>ysician/patient decision aids for adjuvant therapy. J Clin<br />

Oncol 2005; 23:1627–1630.<br />

15. Simes RJ, Coates AS. Patient preferences for adjuvant chemotherapy <strong>of</strong> early breast<br />

cancer: how much benefit is needed? J Natl <strong>Cancer</strong> Inst Monogr 2001; 30:146–152.<br />

16. Berry DA, Cirrincione C, Henderson IC, et al. Estrogen-receptor status and outcomes<br />

<strong>of</strong> modern chemotherapy for patients with node-positive breast cancer. JAMA 2006;<br />

295:1658–1667.<br />

17. van de Vijver MJ, He YD, van ’t Veer LJ, et al. A gene-expression signature as a<br />

predictor <strong>of</strong> survival in breast cancer. N Engl J Med 2002; 347:1999–2009.<br />

18. Buyse M, Loi S, van ’t Veer LJ, et al. Validation and clinical utility <strong>of</strong> a 70-gene<br />

prognostic signature for women with node-negative breast cancer. J Natl <strong>Cancer</strong> Inst<br />

2006; 98:1183–1192.<br />

19. Wang Y, Klijn JG, Zhang Y, et al. Gene-expression pr<strong>of</strong>iles to predict distant<br />

metastasis <strong>of</strong> lymph-node-negative primary breast cancer. Lancet 2005; 365:671–679.<br />

20. Foekens JA, Atkins D, Zhang Y, et al. Multicenter validation <strong>of</strong> a gene expressionbased<br />

prognostic signature in lymph node-negative primary breast cancer. J Clin<br />

Oncol 2006; 24:1665–1671.<br />

21. Paik S, Shak S, Tang G, et al. A multigene assay to predict recurrence <strong>of</strong> tamoxifentreated,<br />

node-negative breast cancer. N Engl J Med 2004; 351:2817–2826.<br />

22. Habel LA, Shak S, Jacobs MK, et al. A population-based study <strong>of</strong> tumor gene<br />

expression and risk <strong>of</strong> breast cancer death among lymph node-negative patients.<br />

<strong>Breast</strong> <strong>Cancer</strong> Res 2006; 8:R25.


196 Sparano<br />

23. Fan C, Oh DS, Wessels L, et al. Concordance among gene-expression-based<br />

predictors for breast cancer. N Engl J Med 2006; 355:560–569.<br />

24. Cobleigh MA, Tabesh B, Bitterman P, et al. Tumor gene expression and prognosis in<br />

breast cancer patients with 10 or more positive lymph nodes. Clin <strong>Cancer</strong> Res 2005;<br />

11:8623–8631.<br />

25. Paik S, Tang G, Shak S, et al. Gene expression and benefit <strong>of</strong> chemotherapy in<br />

women with node-negative, estrogen receptor-positive breast cancer. J Clin Oncol<br />

2006; 24(23):3726–3734.<br />

26. Mamounas ETG, Bryant J, Paik S, et al. Association between the 21-gene recurrence<br />

score assay (RS) and risk <strong>of</strong> loco-regional failure in node-negative, ER-positive<br />

breast cancer: results from NSABP B-14 and NSABP B-20. <strong>Breast</strong> <strong>Cancer</strong> Res 2005;<br />

94(suppl 1):S16 (abstr 29).<br />

27. Sorlie T, Perou CM, Tibshirani R, et al. Gene expression patterns <strong>of</strong> breast carcinomas<br />

distinguish tumor subclasses with clinical implications. Proc Natl Acad Sci<br />

U S A 2001; 98:10869–10874.<br />

28. Sorlie T, Tibshirani R, Parker J, et al. Repeated observation <strong>of</strong> breast tumor subtypes<br />

in independent gene expression data sets. Proc Natl Acad Sci U S A 2003;<br />

100:8418–8423.<br />

29. Bryant J. Oncotype DX correlates more closely with prognosis than Adjuvant<br />

Online. Paper presented at: 9th International Conference on Primary Therapy <strong>of</strong><br />

Early <strong>Breast</strong> <strong>Cancer</strong> January 26-29, 2005; St. Gallen, Switzerland. Available at:<br />

http://www.oncoconferences.ch/2005/program/overview.htm.


14<br />

Taking Prognostic Signatures to the Clinic<br />

Michail Ignatiadis and Christos Sotiriou<br />

Department <strong>of</strong> Medical Oncology, Translational Research Unit,<br />

Institut Jules Bordet, Universite´ Libre de Bruxelles, Brussels, Belgium<br />

INTRODUCTION<br />

The advent <strong>of</strong> microarray technology and the sequencing <strong>of</strong> the human genome<br />

have enabled scientists to simultaneously interrogate the expression <strong>of</strong> thousands<br />

<strong>of</strong> genes. These gene expression pr<strong>of</strong>iling studies have provided (i) a molecular<br />

classification <strong>of</strong> breast cancer into clinically relevant subtypes, (ii) new tools to<br />

predict disease recurrence and response to different treatments, and (iii) novel<br />

insight into various oncogenic pathways and the process <strong>of</strong> metastatic progression.<br />

<strong>Breast</strong> cancer prognosis has been extensively studied by gene expression<br />

pr<strong>of</strong>iling, resulting in different signatures with little overlap in their constituent<br />

genes. In a large meta-analysis <strong>of</strong> all the above studies, despite the disparity in<br />

their gene lists, molecular subtyping and different prognostic signatures were<br />

shown to carry similar prognostic information. In all these signatures, proliferation<br />

genes were the common driving force with regard to prognostication (1).<br />

Gene expression signatures, despite their current limitations may further<br />

refine breast cancer prognosis and prediction <strong>of</strong> benefit from systemic therapy,<br />

beyond conventional clinicopathological prognostic and predictive factors (2). In<br />

this chapter, we will not refer to gene expression signatures that predict response<br />

to systemic therapy. We will focus only on genomic predictors related to prognosis<br />

and the way these predictors are expected to impact patient management in the<br />

future and accelerate the transition from the “empirical” to “tailored” oncology.<br />

197


198 Ignatiadis and Sotiriou<br />

MOLECULAR CLASSIFICATION OF BREAST CANCER<br />

BY GENE EXPRESSION PROFILING<br />

Clinicians have long recognized that the diagnosis <strong>of</strong> breast cancer includes tumor<br />

types with different natural histories and responses to various treatments. Nevertheless,<br />

traditional histopathological characteristics have been unable to capture the<br />

biological heterogeneity <strong>of</strong> breast tumors. The first studies employing microarray<br />

technology to address this issue were performed by Perou and Sorlie (3–5). They<br />

proposed a molecular classification <strong>of</strong> breast cancer with at least five subgroups on<br />

the basis the expression patterns <strong>of</strong> 500 “intrinsic genes.” Three subgroups were<br />

characterized by low to absent expression <strong>of</strong> the estrogen receptor (ER) and ERrelated<br />

genes, the basal-like, the HER2þ, and the normal breast-like subgroup. The<br />

first subtype was characterized by high expression <strong>of</strong> genes found in myoepithelial<br />

cells in the outer basal layer <strong>of</strong> a normal breast duct, like keratins 5/6 and 17, and<br />

was therefore named basal-like. The HER2þ subgroup was characterized by high<br />

expression <strong>of</strong> several genes <strong>of</strong> the ERBB2 amplicon. The normal breast-like subgroup<br />

showed genes expressed by adipose and nonepithelial tissues and also<br />

revealed a strong expression <strong>of</strong> basal epithelial genes. The other two subtypes<br />

consisted <strong>of</strong> ER-positive tumors and expressed genes <strong>of</strong> the luminal epithelial cells<br />

that are found in the inner layer <strong>of</strong> a normal breast duct, and, as a result, they were<br />

termed luminal. The luminal A subgroup had the highest expression <strong>of</strong> the ERa and<br />

the GATA-binding protein 3 gene, and the second subgroup (luminal B and C)<br />

showed low to moderate expression <strong>of</strong> these genes. To assign each sample to one <strong>of</strong><br />

the five molecular subtypes (basal-like, HER2þ/ER , luminal A, luminal B, normal<br />

breast-like), Hu et al. have created a single sample predictor by calculating the<br />

mean expression pr<strong>of</strong>iles (centroids) <strong>of</strong> a new “intrinsic set <strong>of</strong> 300 genes” for each<br />

subgroup (6). In a multivariate analysis, the basal-like, HER2þ/ER , luminal A,<br />

and luminal B subgroups provided important prognostic information beyond those<br />

provided by traditional predictors like tumor size and grade.<br />

Despite using a different microarray platform, Sotiriou et al. confirmed the<br />

consistency and the prognostic relevance <strong>of</strong> the above molecular classification <strong>of</strong><br />

breast cancer in a different patient population (7). Again, ER status was the most<br />

important discriminator <strong>of</strong> gene expression patterns. Interestingly, although<br />

tumor grade was also strongly reflected in the transcriptome pr<strong>of</strong>ile <strong>of</strong> the primary<br />

tumor, this was not the case for tumor size and axillary nodal invasion.<br />

These results confirmed the important role <strong>of</strong> ER and tumor grade biology in<br />

defining breast cancer molecular subtypes. In conclusion, microarray studies<br />

with different technology platforms and in different patient populations treated<br />

with heterogeneous treatments have been remarkably consistent in reproducing a<br />

similar molecular classification <strong>of</strong> breast cancer.<br />

However, it should be pointed out that the number <strong>of</strong> robust breast cancer<br />

molecular subtypes is not certain, and there is no standardized method for assigning<br />

a molecular subtype to a new case (8). Nevertheless, the breast cancer molecular<br />

classification has begun to alter the way clinical trials are designed. As an example,


Taking Prognostic Signatures to the Clinic 199<br />

new trials have been designed to test the efficacy <strong>of</strong> targeted agents only in the<br />

basal-like breast tumors.<br />

GENE EXPRESSION SIGNATURES AS TOOLS FOR IMPROVING<br />

PROGNOSIS IN BREAST CANCER<br />

Besides the above studies aiming at providing a molecular taxonomy <strong>of</strong> breast<br />

tumors, several independent groups, including our own, have conducted gene<br />

expression pr<strong>of</strong>iling studies with the objective <strong>of</strong> improving upon traditional<br />

prognostic markers used in the clinic for prediction <strong>of</strong> outcome. Two conceptually<br />

different, supervised approaches for prognostic marker discovery on a<br />

genome-wide scale have been applied so far, the “top-down” approach and the<br />

“bottom-up” or “hypothesis-driven” approach. The former derives from a<br />

prognostic model simply by looking for gene expression patterns associated with<br />

the clinical outcome without any a priori biological scenario, whereas the latter<br />

approach first identifies gene expression pr<strong>of</strong>iles linked with a specific biological<br />

phenotype and subsequently correlates these findings to survival.<br />

Using the top-down approach, researchers from the Netherlands <strong>Cancer</strong><br />

Institute (NKI) first identified a 70-gene prognostic signature in a series <strong>of</strong> 78<br />

systemically untreated node-negative breast cancer patients. This signature<br />

(MammaPrint 1 Agendia, Amsterdam, the Netherlands) was then validated on a<br />

larger set <strong>of</strong> 295 young treated and untreated patients, with either node-negative<br />

or -positive breast tumors, from the same institution (9,10). The Rotterdam group<br />

later identified, using a training set <strong>of</strong> 115 breast cancer patients, a 76-gene<br />

signature to predict the development <strong>of</strong> distant metastases in untreated nodenegative<br />

breast cancer patients (11). Interestingly, this group used the ER status<br />

to divide the training set into two subgroups, and the genes selected from each<br />

subgroup (60 and 16 genes for the ER-positive and ER-negative subgroups,<br />

respectively) were combined to form a single signature. The 76-gene signature<br />

(VDX2 array, Veridex LLC, Warren, New Jersey, U.S.) was recently validated in<br />

180 lymph node–negative, untreated patients from multiple institutions (12). The<br />

two groups have used different technology platforms (the NKI group used the<br />

Agilent platform, Santa Clara, California, U.S. and the Rotterdam group used<br />

the Affymetrix platform, Santa Clara, California, U.S.), and although there was<br />

only a three-gene overlap between them, both signatures outperformed the<br />

National Institutes <strong>of</strong> Health (NIH) (13) and St. Gallen criteria (14) in stratifying<br />

patients according to risk <strong>of</strong> relapse. Indeed, while the prognostic signatures<br />

correctly identified those patients who will relapse, they could also identify<br />

many low-risk patients who could be spared from unnecessary chemotherapy.<br />

However, the identification <strong>of</strong> high-risk patients could still be improved, since<br />

half <strong>of</strong> the patients assigned to the poor-signature group, in fact, did not relapse.<br />

A major challenge for gene expression pr<strong>of</strong>iling studies, especially those<br />

with clear clinical implications, is independent validation. Consequently,<br />

TRANSBIG, the translational research network <strong>of</strong> the <strong>Breast</strong> International Group


200 Ignatiadis and Sotiriou<br />

(BIG), conducted an independent validation study <strong>of</strong> these two prognostic signatures<br />

in a series <strong>of</strong> 302 untreated patients from five different centers and across<br />

different statistical facilities (15,16). The two signatures were robust, and their<br />

performance was reproducible and similar in this retrospective series. Interestingly,<br />

both signatures were strong predictors <strong>of</strong> the development <strong>of</strong> metastasis<br />

within five years <strong>of</strong> diagnosis (early metastasis), but their performance decreased<br />

for the detection <strong>of</strong> late metastasis. This could be easily explained, since the<br />

training set in both studies consisted <strong>of</strong> patients who have all relapsed within five<br />

years. However, an intriguing question arises about the potential differences in<br />

the biology and the oncogenic pathways that drive early and late metastasis.<br />

Another group explored the correlation between the expression <strong>of</strong> 250<br />

genes (selected from publicly available microarray data linked with breast cancer<br />

biology) and the clinical outcome in 447 patients from three studies including the<br />

tamoxifen-only group <strong>of</strong> the NASBP trial B-20 to build the Oncotype DX 1<br />

(Genomic Health Inc, Redwood City, California, U.S.) Recurrence Score 1 (RS)<br />

(17). This score (low, intermediate, and high) was based on the expression <strong>of</strong> a<br />

final selected panel <strong>of</strong> 16 cancer-related genes and five reference genes, as<br />

assessed by reverse transcriptase-polymerase chain reaction (RT-PCR) in paraffin-embedded<br />

tumor tissue. Then, the Oncotype DX RS was validated for its<br />

ability to quantify the likelihood <strong>of</strong> distance recurrence in the tamoxifen treated,<br />

ER-positive, node-negative group <strong>of</strong> the NSABP trial B-14 (17). Using the RS,<br />

around 50% <strong>of</strong> the patients in the above study were characterized as low risk<br />

(low RS), with a 6.8% probability <strong>of</strong> distant recurrence at 10 years.<br />

Using the second approach, the hypothesis-driven approach, Chang et al.<br />

tried to identify the molecular similarities between cancer invasion/metastasis<br />

and wound healing (18). They developed a gene signature to characterize the<br />

response <strong>of</strong> fibroblasts to serum. When they applied these genes to primary<br />

gene expression pr<strong>of</strong>iling data from multiple data sets, they observed that the<br />

presence <strong>of</strong> a wound-healing phenotype in the primary tumor was related with<br />

worse clinical outcome in various cancers, including breast cancer. Similar to the<br />

70- and 76-gene signatures, the wound-response signature provided better risk<br />

stratification than the NIH and St. Gallen criteria. Furthermore, it was able to<br />

refine prognosis in the 295 patient data sets <strong>of</strong> van de Vijver et al. (10) beyond<br />

existing clinicopathological risk factors and the 70-gene signature (19). In<br />

conclusion, Chang et al., using a hypothesis-driven approach, apart from identifying<br />

a gene expression prognostic signature, have <strong>of</strong>fered novel insight into<br />

the molecular pathogenesis <strong>of</strong> metastasis by linking this process with wound<br />

healing.<br />

Another example <strong>of</strong> deriving a prognostic gene expression signature using<br />

a hypothesis-driven approach is the study by our group (20) that focused on<br />

histological grade, a well-established pathological parameter rooted in the cell<br />

biology <strong>of</strong> breast cancer. Applying a supervised analysis, we developed a Gene<br />

expression Grade Index (GGI) score based on 97 genes. These genes were<br />

mainly involved in cell cycle regulation and proliferation and were consistently


Taking Prognostic Signatures to the Clinic 201<br />

differentially expressed between low- and high-grade breast carcinomas. Interestingly,<br />

the GGI, which essentially quantifies the degree <strong>of</strong> similarity between<br />

the tumor expression pattern <strong>of</strong> these 97 genes and tumor grade, was able to<br />

reclassify patients with histological grade 2 tumors into two groups with distinct<br />

clinical outcomes similar to those <strong>of</strong> histological grades 1 and 3, respectively<br />

(20). This observation challenges the existence and clinical relevance <strong>of</strong> an<br />

intermediate grade classification. When we explored the implications <strong>of</strong> the joint<br />

distribution <strong>of</strong> ER status and GGI in predicting clinical outcome, we found that<br />

almost all ER-negative tumors with poor clinical outcome were associated with a<br />

high GGI score (high grade), whereas ER-positive tumors were associated with a<br />

heterogeneous mixture <strong>of</strong> gene expression grade values. Recently, our group has<br />

reported that GGI can be used to define two clinically relevant molecular subtypes<br />

in ER-positive breast cancer (21). These two ER-positive molecular subgroups<br />

(high and low genomic grade) were highly comparable to the luminal A<br />

and B classifications and were associated with distinct clinical outcome in<br />

systemically untreated or tamoxifen-only-treated ER-positive breast cancer<br />

populations involving more than 650 patients (21). GGI appeared to be the<br />

strongest predictor <strong>of</strong> clinical outcome, highlighting the prognostic importance<br />

<strong>of</strong> proliferation genes in ER-positive subgroups, as have been already reported<br />

(17). Indeed, proliferation was one <strong>of</strong> the five components associated with the<br />

highest weight in the Oncotype DX RS.<br />

Four additional studies have used the hypothesis-driven approach to build<br />

gene predictors <strong>of</strong> clinical outcome. Oh et al. used estrogen-induced genes,<br />

identified by treating the MCF-7 cell line with 17b-estradiol to build a predictor<br />

<strong>of</strong> clinical outcome on a training set <strong>of</strong> 65 patients with ER-positive breast<br />

tumors (22). A total <strong>of</strong> 822 genes, mainly proliferation and antiapoptosis genes<br />

could separate these tumors into two subgroups with different clinical outcomes.<br />

This gene list was then mapped in three published datasets (4,5,19,23) and an<br />

average expression index or “centroid” was created that could classify the ERpositive<br />

tumors into two clinically relevant subgroups. In another study, Miller et<br />

al. generated a 32-gene expression signature that distinguishes p53 mutant and<br />

wild type tumors. The aim <strong>of</strong> this research was to provide a more accurate<br />

assessment <strong>of</strong> the functional status <strong>of</strong> the p53 pathway than that provided by<br />

mutation analysis using sequencing (24). It should be noted that none <strong>of</strong> the<br />

signature genes were known transcriptional targets <strong>of</strong> p53, nor have they been<br />

related to the p53 pathway. Instead, they were mainly proliferation-related genes,<br />

various transcription factors, and ion transport genes. Once more, the proliferation<br />

genes were a prominent part <strong>of</strong> this signature. The signature outperformed<br />

p53 mutation status, as identified by sequencing in prognosis and prediction <strong>of</strong><br />

response to hormonal therapy and chemotherapy. By combining the hypothesisdriven<br />

approach with comparative cross-species functional genomics, Glinsky<br />

et al. identified an 11-gene “death-from-cancer” signature (25). This signature<br />

can identify a lethal subset <strong>of</strong> human cancers <strong>of</strong> various origins (including breast<br />

cancer) with a high propensity to metastasize. These tumors display the


202 Ignatiadis and Sotiriou<br />

“phenotype” <strong>of</strong> an activated BMI-1 oncogenic pathway that is essential for the<br />

self-renewal <strong>of</strong> normal stem cells. Again, proliferation- and cell cycle–regulating<br />

genes comprise the most essential part <strong>of</strong> this signature. Along the same lines,<br />

Liu et al. isolated and compared at the transcriptional-level tumorigenic CD44þ/<br />

CD24 /low breast cancer cells derived from a small set <strong>of</strong> six breast cancer<br />

patients with normal breast epithilium (26). They developed an “invasiveness”<br />

gene signature (IGS) that was associated with the risk <strong>of</strong> death and metastasis not<br />

only in breast cancer but also in other tumor types. Again, there was a small gene<br />

overlap between IGS and other prognostic signatures, and, not surprisingly, its<br />

prognostic power was significant only in ER-positive and intermediate-grade<br />

tumors, a feature that had already been observed for other prognostic pr<strong>of</strong>iles.<br />

CAN WE INTEGRATE THE DIFFERENT GENE EXPRESSION SIGNATURES?<br />

The above gene expression signatures developed to improve breast cancer<br />

prognostication have little overlap in their constituent genes. A representative<br />

list <strong>of</strong> these signatures is depicted in Table 1. The inevitable question is whether<br />

the “intrinsic subtypes” developed by unsupervised cluster analysis and the<br />

other prognostic signatures have also little overlap in the prognostic information<br />

they convey. Fan et al. recently compared the prognostic ability <strong>of</strong> five gene<br />

expression–based models—intrinsic subtypes (6), 70-gene pr<strong>of</strong>ile (9), wound<br />

response (19), RS (17), and two-gene ratio (23)—in a single dataset <strong>of</strong><br />

295 patients and found that four <strong>of</strong> five predictors had significant agreement in<br />

outcome predictions for individual patients (27). However, this study was limited<br />

to only one dataset <strong>of</strong> 295 patients, examined only five gene expression signatures,<br />

and did not investigate the biology that may drive prognosis among the five<br />

signatures. To address this issue, a large meta-analysis <strong>of</strong> publicly available gene<br />

expression and clinical data from 2833 patients (4–7,9,12,15,20,23,24,28–39) with<br />

breast cancer was performed by the Swiss Institute <strong>of</strong> Bioinformatics in collaboration<br />

with our group (Table 1). In this meta-analysis, the concept <strong>of</strong> “coexpression<br />

modules” (comprehensive list <strong>of</strong> genes with highly correlated<br />

expression) was used to describe important biological processes in breast cancer,<br />

like proliferation, ER and HER2 signaling, tumor invasion, and immune response.<br />

In this study, we sought to depict the connection between these modules, the<br />

previously reported molecular classification, several gene prognostic classifiers,<br />

and the most established clinicopathological variables. A number <strong>of</strong> interesting<br />

conclusions were drawn from this collaborative effort. First, the disparity <strong>of</strong> gene<br />

lists <strong>of</strong> the various gene signatures may be attributed to (i) heterogeneity in patient<br />

populations studied, (ii) the use <strong>of</strong> different microarray platforms with different<br />

probe sets and different methods for data normalization, and (iii) sampling variation<br />

due to small sample size relative to the number <strong>of</strong> genes examined. Second,<br />

using these co-expression modules, instead <strong>of</strong> the five molecular portraits reported<br />

with the intrinsic genes, breast tumors were now grouped into three main subgroups<br />

namely ER /HER , HER2þ, and ERþ/HER2 . Third, the ER /HER ,


Taking Prognostic Signatures to the Clinic 203<br />

Table 1 Gene Expression Signatures Related to <strong>Breast</strong> <strong>Cancer</strong> Prognosis<br />

Gene expression<br />

signatures Biological scenario Microarray platform<br />

Number <strong>of</strong><br />

genes in the<br />

signature<br />

Independent<br />

validation<br />

Prospective<br />

clinical validation Ref.<br />

70-gene signature Clinical outcome Agilent (oligonucleotides) 70 yes Yes (MINDACT<br />

Trial)<br />

76-gene signature Clinical outcome Affymetrix<br />

(oligonucleotides)<br />

(9,10,15)<br />

76 yes No (11,12,16)<br />

Recurrence Score Clinical outcome RT-PCR 21 yes Yes (TAILORX<br />

trial)<br />

Wound-response<br />

signature<br />

Wound healing and tumor<br />

progression<br />

Genomic grade Histological grade and tumor<br />

progression<br />

(17,37)<br />

CDNA (custom made) 512 yes No (18,19)<br />

Affymetrix<br />

(oligonucleotides)<br />

P-53 signature Functional status <strong>of</strong> p-53 Affymetrix<br />

(oligonucleotides)<br />

Death-from-cancer<br />

signature<br />

Invasiveness gene<br />

signature<br />

Bmi-1 oncogenic pathway<br />

self-renewal<br />

Tumorigenic cancer cells<br />

CD44þ/CD24 , low<br />

Affymetrix<br />

(oligonucleotides)<br />

Affymetrix<br />

(oligonucleotides)<br />

97 yes No (20)<br />

32 yes No (24)<br />

11 Yes No (25)<br />

186 Yes No (26)


204 Ignatiadis and Sotiriou<br />

Figure 1 Despite the disparity in their gene lists, different prognostic signatures showed<br />

similar prognostic performance. Proliferation is the common driving force responsible for<br />

their prognostic performance.<br />

HER2þ tumors were characterized by high proliferation, whereas the<br />

ERþ/HER2 was divided in two subgroups, the ERþ/low proliferation and the<br />

ERþ/high proliferation tumors resembling to luminal A and B subtypes,<br />

respectively. Fourth, 10 prognostic signatures [RS (17), 70-gene signature (9),<br />

76-gene signature (11), wound response (18), p53 signature (24), GGI (20),<br />

NCH70 (33), CON52 (38), CCYC (39), ZCOX], despite the disparity in their<br />

gene lists, showed similar prognostic performance (Fig. 1). It is interesting to<br />

note that they all identified as “low risk” the ERþ subtype associated with low<br />

proliferation. Fifth, proliferation was the common driving force responsible for<br />

the performance <strong>of</strong> the various prognostic signatures (Fig. 1). Indeed, to further<br />

investigate the role <strong>of</strong> proliferation genes in relation to the different prognostic<br />

signatures, we divided each signature into two “partial signatures”—one with only<br />

proliferation genes and the other with the complementary nonproliferation genes.<br />

Interestingly, when only proliferation genes were used, the total performance was<br />

not degraded. In contrast, the nonproliferation partial signatures typically show<br />

degraded performance. Sixth, combining the signatures did not improve the performance<br />

as expected from the high concordance in their classifications. Finally,<br />

in the multivariate analysis, including the genomic predictor derived from coexpression<br />

modules, the standard clinicopathological variables, namely tumor size<br />

and nodal status, retained their prognostic value.<br />

PROSPECTIVE VALIDATION OF TWO GENOMIC PREDICTORS<br />

The initial promising results that breast cancer prognostication can be improved<br />

using gene expression signatures need to be validated by well-designed prospective<br />

clinical trials. Two gene expression predictors, the 70-gene expression


Taking Prognostic Signatures to the Clinic 205<br />

signature, (MammaPrint) discussed earlier and the Oncotype DX RS have<br />

reached the final step <strong>of</strong> prospective clinical testing. The BIG and the EORTC<br />

have jointly launched the MINDACT (Microarray for Node-Negative Disease<br />

may Avoid Chemotherapy) trial that directly compares two approaches for<br />

deciding whether to <strong>of</strong>fer adjuvant chemotherapy in node-negative breast cancer<br />

patients, i.e., either using the prognostic information provided by standard<br />

clinicopathological criteria (control arm) or by the 70-gene expression signature<br />

(MammaPrint) (experimental arm). Six thousand women will participate, and<br />

the primary endpoint is to prove that five-year disease-free survival in the<br />

experimental arm will not be inferior to that <strong>of</strong> the control arm; at the same time,<br />

it is estimated that 10% to 15% fewer women will be treated with chemotherapy<br />

in the experimental arm.<br />

A similarly ambitious initiative is taking place in the United States. The<br />

Oncotype DX RS is being evaluated in the TAILORx [Trial Assigning<br />

IndividuaLized Options for Treatment (Rx)] trial, which is part <strong>of</strong> the<br />

National <strong>Cancer</strong> Institute (NCI) Program for the Assessment <strong>of</strong> Clinical<br />

<strong>Cancer</strong> Tests (PACCT). This trial randomizes patients with intermediate RS<br />

to receive either hormonal therapy alone or hormonal therapy plus chemotherapy.<br />

In it, patients with low RS will be treated with hormonal therapy<br />

alone and those with high RS will receive chemotherapy plus hormonal<br />

therapy.<br />

These two trials will provide level I evidence that these gene expression<br />

predictors provide prognostic information beyond standard clinicopathological<br />

factors and are ready for use in clinical practice. Both trials will provide an<br />

excellent opportunity to address issues related to tissue handling and shipping,<br />

reproducibility, quality controls, and standardization <strong>of</strong> the microarray experiments.<br />

Furthermore, they will provide a valuable database for future translational<br />

research.<br />

FUTURE DIRECTION–CONCLUSIONS<br />

The real challenge, once the above trials prospectively validate the two genomic<br />

predictors, will be to integrate these predictors into a comprehensive clinical<br />

decision-making algorithm that will lead toward personalized medicine.<br />

Parameters that need to be included in this algorithm include the following:<br />

Traditional Clinicopathological Parameters<br />

These include patient comorbidities, age, and primary tumor characteristics like<br />

tumor size, tumor grade, nodal status, ER status, HER2/neu status, and lymphovascular<br />

invasion. Many <strong>of</strong> these parameters are incorporated in various<br />

prognostic tools like the Nottingham Prognostic Index (40), Adjuvant! Online<br />

(41), and the NIH (13) and St. Gallen (42) Guidelines.


206 Ignatiadis and Sotiriou<br />

Primary Tumor Gene Expression Signatures<br />

From the above discussion, it is apparent that two genomic predictors are in the<br />

late phase <strong>of</strong> clinical validation. However, apart from the question <strong>of</strong> who can<br />

be spared unnecessary adjuvant chemotherapy (prognosis), a more important<br />

question is which therapy works better in an individual patient (prediction).<br />

Genomic predictors <strong>of</strong> response to multidrug regimens and newer targeted agents<br />

have also been reported (43–49), but their description is beyond the scope <strong>of</strong> this<br />

chapter. Other important technology platforms are being developed to analyze<br />

epigenetic changes in DNA (50), the role <strong>of</strong> microRNAs (51) that will provide us<br />

with a more comprehensive picture <strong>of</strong> the biology <strong>of</strong> the tumor and further refine<br />

prognosis and prediction.<br />

Micrometastatic Cells<br />

The presence <strong>of</strong> cytokeratin positive tumor cells in the bone marrow <strong>of</strong> early breast<br />

cancer patients has been correlated with worse outcome in a meta-analysis<br />

involving more than 4500 patients (52). We have reported that the detection with a<br />

real-time RT-PCR assay <strong>of</strong> Cytokeratin-19 (CK19) mRNA-positive circulating<br />

tumor cells (CTCs) in patients with early breast cancer before (53,54) and after<br />

(55,56) the initiation <strong>of</strong> adjuvant systemic treatment is associated with poor outcome.<br />

However, the additional prognostic information <strong>of</strong> micrometastatic cells to<br />

primary tumor gene expression pr<strong>of</strong>iling data has not yet been examined. Moreover,<br />

there are pilot studies where HER2-positive CTCs and/or disseminated tumor cells<br />

(DTCs) were successfully eliminated with trastuzumab (57), suggesting that CTCs<br />

and/or DTCs may be tested also as tools to predict response to a particular drug.<br />

Molecular Imaging<br />

PET (positron emission tomography) scan using novel PET tracers that target<br />

critical cellular processes like angiogenesis may provide important staging and<br />

functional information and is under investigation in breast cancer (58).<br />

<strong><strong>Ph</strong>armacogenetics</strong> (Host)<br />

The importance <strong>of</strong> pharmacogenetics in predicting clinical outcome in breast<br />

cancer has recently been illustrated in studies focused on the polymorphism <strong>of</strong><br />

enzymes that metabolize tamoxifen (59,60). The study <strong>of</strong> the host is an important<br />

piece <strong>of</strong> the clinical decision-making algorithm.<br />

Proteomics<br />

Proteomics provides the hope <strong>of</strong> discovering novel biological markers that can<br />

be used for early detection, disease diagnosis, prognostication, and prediction <strong>of</strong><br />

response to therapy in early breast cancer (61).


Taking Prognostic Signatures to the Clinic 207<br />

Adjuvant Treatment<br />

It is well documented that adjuvant treatment affects clinical outcome <strong>of</strong> breast<br />

cancer patients (62). From that perspective, the positive adjuvant trials with the<br />

aromatase inhibitors (63) and trastuzumab (64,65), the concept <strong>of</strong> dose-dense<br />

adjuvant chemotherapy (66), and the emerging data about the use <strong>of</strong> taxanes<br />

(67,68) and anthracyclines (69) in the adjuvant setting are rendering both the<br />

treatment decision-making process and prediction <strong>of</strong> outcome increasingly more<br />

complex.<br />

Several efforts have been made to combine at least some <strong>of</strong> the above<br />

parameters to predict the clinical outcome <strong>of</strong> breast cancer patients. For instance,<br />

Pittman et al. have used a modeling approach based on classification trees to<br />

combine clinical and genomic data (70). They nicely showed that by combining<br />

multiple metagenes (defined as summary measures <strong>of</strong> the expression <strong>of</strong> subsets<br />

<strong>of</strong> potentially related genes) with nodal status, they could create integrative<br />

clinicogenomic models. These models provided improved prediction accuracy<br />

compared with either clinical or genomic data alone. Another example <strong>of</strong> such an<br />

effort is the Adjuvant! Online for breast cancer, Genomic Version 7.0, which<br />

combines traditional clinical data, the RS, and information about adjuvant<br />

treatment to provide an estimate <strong>of</strong> clinical outcome in patients with early breast<br />

cancer (71).<br />

However, the accumulation <strong>of</strong> highly complex and multilevel information<br />

will ultimately impose specific data integration requirements. So far these different<br />

types <strong>of</strong> data are not only maintained in a variety <strong>of</strong> different data sources<br />

but are also managed by different complex data management and analytical<br />

systems. Additionally, they involve many institutions and numerous complex<br />

workflows. To address all these critical issues, two initiatives have been<br />

launched that intend to integrate all this information into one comprehensive<br />

biomedical platform—the cancer Biomedical Informatics Grid, “caBIG” (https://<br />

cabig.nci.nih.gov/) and the Advancing Clinico Genomic Trials, “ACGT” (http://<br />

www.eu-acgt.org/) funded by the NCI and European Union, respectively.<br />

In the future, medical oncologists should use a clinically validated treatment<br />

decision-making algorithm that will take into account all this highly<br />

complex and rapidly evolving, multilevel information. In this way, we will move<br />

from the “empirical” to the new exciting era <strong>of</strong> “individual-tailored” oncology.<br />

REFERENCES<br />

1. Sotiriou C, Wirapati P, Kunkel S, et al. Meta-analysis <strong>of</strong> gene-expression pr<strong>of</strong>iles in<br />

breast cancer: towards a unified understanding <strong>of</strong> breast cancer sub-typing and<br />

prognosis signatures. <strong>Breast</strong> <strong>Cancer</strong> Res Treat. 2007; 106(suppl 1):1047(abstr).<br />

2. Ignatiadis M, Desmedt C. Predicting risk <strong>of</strong> breast cancer recurrence using geneexpression<br />

pr<strong>of</strong>iling. <strong>Ph</strong>armacogenomics 2007; 8(1):101–111.<br />

3. Perou CM, Sorlie T, Eisen MB, et al. Molecular portraits <strong>of</strong> human breast tumours.<br />

Nature 2000; 406(6797):747–752.


208 Ignatiadis and Sotiriou<br />

4. Sorlie T, Perou CM, Tibshirani R, et al. Gene expression patterns <strong>of</strong> breast carcinomas<br />

distinguish tumor subclasses with clinical implications. Proc Natl Acad Sci<br />

U S A 2001; 98(19):10869–10874.<br />

5. Sorlie T, Tibshirani R, Parker J, et al. Repeated observation <strong>of</strong> breast tumor subtypes<br />

in independent gene expression data sets. Proc Natl Acad Sci U S A 2003; 100(14):<br />

8418–8423.<br />

6. Hu Z, Fan C, Oh DS, et al. The molecular portraits <strong>of</strong> breast tumors are conserved<br />

across microarray platforms. BMC Genomics 2006; 7:96.<br />

7. Sotiriou C, Neo SY, McShane LM, et al. <strong>Breast</strong> cancer classification and prognosis<br />

based on gene expression pr<strong>of</strong>iles from a population-based study. Proc Natl Acad Sci<br />

U S A 2003; 100(18):10393–10398.<br />

8. Andre F, Pusztai L. Molecular classification <strong>of</strong> breast cancer: implications for<br />

selection <strong>of</strong> adjuvant chemotherapy. Nat Clin Pract Oncol 2006; 3(11):621–632.<br />

9. van ’t Veer LJ, Dai H, van de Vijver MJ, et al. Gene expression pr<strong>of</strong>iling predicts<br />

clinical outcome <strong>of</strong> breast cancer. Nature 2002; 415(6871):530–536.<br />

10. van de Vijver MJ, He YD, van ’t Veer LJ, et al. A gene-expression signature as a<br />

predictor <strong>of</strong> survival in breast cancer. N Engl J Med 2002; 347(25):1999–2009.<br />

11. Wang Y, Klijn JG, Zhang Y, et al. Gene-expression pr<strong>of</strong>iles to predict distant<br />

metastasis <strong>of</strong> lymph-node-negative primary breast cancer. Lancet 2005; 365(9460):<br />

671–679.<br />

12. Foekens JA, Atkins D, Zhang Y, et al. Multicenter validation <strong>of</strong> a gene expressionbased<br />

prognostic signature in lymph node-negative primary breast cancer. J Clin<br />

Oncol 2006; 24(11):1665–1671.<br />

13. Eifel P, Axelson JA, Costa J, et al. National Institutes <strong>of</strong> Health Consensus Development<br />

Conference Statement: adjuvant therapy for breast cancer, November 1–3,<br />

2000. J Natl <strong>Cancer</strong> Inst 2001; 93(13):979–989.<br />

14. Goldhirsch A, Wood WC, Gelber RD, et al. Meeting highlights: updated international<br />

expert consensus on the primary therapy <strong>of</strong> early breast cancer. J Clin Oncol<br />

2003; 21(17):3357–3365.<br />

15. Buyse M, Loi S, van ’t Veer LJ, et al. Validation and clinical utility <strong>of</strong> a 70-gene<br />

prognostic signature for women with node-negative breast cancer. J Natl <strong>Cancer</strong> Inst<br />

2006; 98(17):1183–1192.<br />

16. Desmedt C, Piette F, Loi S, et al. Strong time-dependency <strong>of</strong> the 76-gene prognostic<br />

signature for node-negative breast cancer patients in the TRANSBIG multi-centre<br />

independent validation series. Clin <strong>Cancer</strong> Res 2007; 13:3207–3214.<br />

17. Paik S, Shak S, Tang G, et al. A multigene assay to predict recurrence <strong>of</strong> tamoxifentreated,<br />

node-negative breast cancer. N Engl J Med 2004; 351(27):2817–2826.<br />

18. Chang HY, Sneddon JB, Alizadeh AA, et al. Gene expression signature <strong>of</strong> fibroblast<br />

serum response predicts human cancer progression: similarities between tumors and<br />

wounds. PLoS Biol 2004; 2(2):E7.<br />

19. Chang HY, Nuyten DS, Sneddon JB, et al. Robustness, scalability, and integration <strong>of</strong><br />

a wound-response gene expression signature in predicting breast cancer survival.<br />

Proc Natl Acad Sci U S A 2005; 102(10):3738–3743.<br />

20. Sotiriou C, Wirapati P, Loi S, et al. Gene expression pr<strong>of</strong>iling in breast cancer:<br />

understanding the molecular basis <strong>of</strong> histologic grade to improve prognosis. J Natl<br />

<strong>Cancer</strong> Inst 2006; 98(4):262–272.


Taking Prognostic Signatures to the Clinic 209<br />

21. Loi S, Haibe-Kains B, Desmedt C, et al. Definition <strong>of</strong> clinically distinct molecular<br />

subtypes in estrogen receptor-positive breast carcinomas through genomic grade.<br />

J Clin Oncol 2007; 25(10):1239–1246.<br />

22. Oh DS, Troester MA, Usary J, et al. Estrogen-regulated genes predict survival in<br />

hormone receptor-positive breast cancers. J Clin Oncol 2006; 24(11):1656–1664.<br />

23. Ma XJ, Wang Z, Ryan PD, et al. A two-gene expression ratio predicts clinical<br />

outcome in breast cancer patients treated with tamoxifen. <strong>Cancer</strong> Cell 2004; 5(6):<br />

607–616.<br />

24. Miller LD, Smeds J, George J, et al. An expression signature for p53 status in human<br />

breast cancer predicts mutation status, transcriptional effects, and patient survival.<br />

Proc Natl Acad Sci U S A 2005; 102(38):13550–13555.<br />

25. Glinsky GV, Berezovska O, Glinskii AB. Microarray analysis identifies a deathfrom-cancer<br />

signature predicting therapy failure in patients with multiple types <strong>of</strong><br />

cancer. J Clin Invest 2005; 115(6):1503–1521.<br />

26. Liu R, Wang X, Chen GY, et al. The prognostic role <strong>of</strong> a gene signature from<br />

tumorigenic breast-cancer cells. N Engl J Med 2007; 356(3):217–226.<br />

27. Fan C, Oh DS, Wessels L, et al. Concordance among gene-expression-based predictors<br />

for breast cancer. N Engl J Med 2006; 355(6):560–569.<br />

28. Calza S, Hall P, Auer G, et al. Intrinsic molecular signature <strong>of</strong> breast cancer in a<br />

population-based cohort <strong>of</strong> 412 patients. <strong>Breast</strong> <strong>Cancer</strong> Res 2006; 8(4):R34.<br />

29. Pawitan Y, Bjohle J, Amler L, et al. Gene expression pr<strong>of</strong>iling spares early breast<br />

cancer patients from adjuvant therapy: derived and validated in two population-based<br />

cohorts. <strong>Breast</strong> <strong>Cancer</strong> Res 2005; 7(6):R953–R964.<br />

30. Huang E, Cheng SH, Dressman H, et al. Gene expression predictors <strong>of</strong> breast cancer<br />

outcomes. Lancet 2003; 361(9369):1590–1596.<br />

31. Huang E, Ishida S, Pittman J, et al. Gene expression phenotypic models that predict<br />

the activity <strong>of</strong> oncogenic pathways. Nat Genet 2003; 34(2):226–230.<br />

32. Korkola JE, DeVries S, Fridlyand J, et al. Differentiation <strong>of</strong> lobular versus ductal breast<br />

carcinomas by expression microarray analysis. <strong>Cancer</strong> Res 2003; 63(21): 7167–7175.<br />

33. Naderi A, Teschendorff AE, Barbosa-Morais NL, et al. A gene-expression signature<br />

to predict survival in breast cancer across independent data sets. Oncogene 2007;<br />

26(10):1507–1516.<br />

34. Farmer P, Bonnefoi H, Becette V, et al. Identification <strong>of</strong> molecular apocrine breast<br />

tumours by microarray analysis. Oncogene 2005; 24(29):4660–4671.<br />

35. Richardson AL, Wang ZC, De NA, et al. X chromosomal abnormalities in basal-like<br />

human breast cancer. <strong>Cancer</strong> Cell 2006; 9(2):121–132.<br />

36. Espinosa E, Vara JA, Redondo A, et al. <strong>Breast</strong> cancer prognosis determined by gene<br />

expression pr<strong>of</strong>iling: a quantitative reverse transcriptase polymerase chain reaction<br />

study. J Clin Oncol 2005; 23(29):7278–7285.<br />

37. Paik S, Tang G, Shak S, et al. Gene expression and benefit <strong>of</strong> chemotherapy in<br />

women with node-negative, estrogen receptor-positive breast cancer. J Clin Oncol<br />

2006; 24(23):3726–3734.<br />

38. Teschendorff AE, Naderi A, Barbosa-Morais NL, et al. A consensus prognostic gene<br />

expression classifier for ER positive breast cancer. Genome Biol 2006; 7(10):R101.<br />

39. Whitfield ML, George LK, Grant GD, et al. Common markers <strong>of</strong> proliferation. Nat<br />

Rev <strong>Cancer</strong> 2006; 6(2):99–106.


210 Ignatiadis and Sotiriou<br />

40. Galea MH, Blamey RW, Elston CE, et al. The Nottingham Prognostic Index in<br />

primary breast cancer. <strong>Breast</strong> <strong>Cancer</strong> Res Treat 1992; 22(3):207–219.<br />

41. Olivotto IA, Bajdik CD, Ravdin PM, et al. Population-based validation <strong>of</strong> the<br />

prognostic model ADJUVANT! for early breast cancer. J Clin Oncol 2005; 23(12):<br />

2716–2725.<br />

42. Goldhirsch A, Glick JH, Gelber RD, et al. Meeting highlights: international expert<br />

consensus on the primary therapy <strong>of</strong> early breast cancer 2005. Ann Oncol 2005;<br />

16(10):1569–1583.<br />

43. Gianni L, Zambetti M, Clark K, et al. Gene expression pr<strong>of</strong>iles in paraffin-embedded<br />

core biopsy tissue predict response to chemotherapy in women with locally advanced<br />

breast cancer. J Clin Oncol 2005; 23(29):7265–7277.<br />

44. Chang JC, Wooten EC, Tsimelzon A, et al. Patterns <strong>of</strong> resistance and incomplete<br />

response to docetaxel by gene expression pr<strong>of</strong>iling in breast cancer patients. J Clin<br />

Oncol 2005; 23(6):1169–1177.<br />

45. Jansen MP, Foekens JA, van Staveren IL, et al. Molecular classification <strong>of</strong> tamoxifen-resistant<br />

breast carcinomas by gene expression pr<strong>of</strong>iling. J Clin Oncol 2005;<br />

23(4):732–740.<br />

46. Hess KR, Anderson K, Symmans WF, et al. <strong>Ph</strong>armacogenomic predictor <strong>of</strong> sensitivity<br />

to preoperative chemotherapy with paclitaxel and fluorouracil, doxorubicin,<br />

and cyclophosphamide in breast cancer. J Clin Oncol 2006; 24(26):4236–4244.<br />

47. Bild AH, Yao G, Chang JT, et al. Oncogenic pathway signatures in human cancers as<br />

a guide to targeted therapies. Nature 2006; 439(7074):353–357.<br />

48. Potti A, Dressman HK, Bild A, et al. Genomic signatures to guide the use <strong>of</strong> chemotherapeutics.<br />

Nat Med 2006; 12(11):1294–1300.<br />

49. Solit DB, Garraway LA, Pratilas CA, et al. BRAF mutation predicts sensitivity to<br />

MEK inhibition. Nature 2006; 439(7074):358–362.<br />

50. Herman JG, Baylin SB. Gene silencing in cancer in association with promoter<br />

hypermethylation. N Engl J Med 2003; 349(21):2042–2054.<br />

51. Iorio MV, Ferracin M, Liu CG, et al. MicroRNA gene expression deregulation in<br />

human breast cancer. <strong>Cancer</strong> Res 2005; 65(16):7065–7070.<br />

52. Braun S, Vogl FD, Naume B, et al. A pooled analysis <strong>of</strong> bone marrow micrometastasis<br />

in breast cancer. N Engl J Med 2005; 353(8):793–802.<br />

53. Xenidis N, Perraki M, Kafousi M, et al. Predictive and prognostic value <strong>of</strong> peripheral<br />

blood cytokeratin-19 mRNA-positive cells detected by real-time polymerase chain<br />

reaction in node-negative breast cancer patients. J Clin Oncol 2006; 24(23):3756–3762.<br />

54. Ignatiadis M, Xenidis N, Perraki M, et al. Different prognostic value <strong>of</strong> Cytokeratin-<br />

19 mRNA-positive Circulating Tumor Cells according to estrogen receptor status in<br />

early breast cancer. J Clin Oncol 2007; 25(33):5194–5202.<br />

55. Xenidis N, Vlachonikolis I, Mavroudis D, et al. Peripheral blood circulating cytokeratin-19<br />

mRNA-positive cells after the completion <strong>of</strong> adjuvant chemotherapy in<br />

patients with operable breast cancer. Ann Oncol 2003; 14(6):849–855.<br />

56. Xenidis N, Mavroudis D, Apostolaki S, et al. Effect <strong>of</strong> adjuvant treatment on the<br />

circulating CK19mRNA positive tumor cells in patients with early stage breast<br />

cancer. Proc Am Soc Clin Oncol 2006; 24(No18S):20012 (abstr).<br />

57. Bozionellou V, Mavroudis D, Perraki M, et al. Trastuzumab administration can<br />

effectively target chemotherapy-resistant cytokeratin-19 messenger RNA-positive<br />

tumor cells in the peripheral blood and bone marrow <strong>of</strong> patients with breast cancer.<br />

Clin <strong>Cancer</strong> Res 2004; 10(24):8185–8194.


Taking Prognostic Signatures to the Clinic 211<br />

58. Quon A, Gambhir SS. FDG-PET and beyond: molecular breast cancer imaging.<br />

J Clin Oncol 2005; 23(8):1664–1673.<br />

59. Goetz MP, Rae JM, Suman VJ, et al. <strong><strong>Ph</strong>armacogenetics</strong> <strong>of</strong> tamoxifen biotransformation<br />

is associated with clinical outcomes <strong>of</strong> efficacy and hot flashes. J Clin<br />

Oncol 2005; 23(36):9312–9318.<br />

60. Nowell S, Sweeney C, Winters M, et al. Association between sulfotransferase 1A1<br />

genotype and survival <strong>of</strong> breast cancer patients receiving tamoxifen therapy. J Natl<br />

<strong>Cancer</strong> Inst 2002; 94(21):1635–1640.<br />

61. Somiari RI, Somiari S, Russell S, Shriver CD. Proteomics <strong>of</strong> breast carcinoma.<br />

J Chromatogr B Analyt Technol Biomed Life Sci 2005; 815(1–2):215–225.<br />

62. Early <strong>Breast</strong> <strong>Cancer</strong> Trialists’ Collaborative Group (EBCTCG). Effects <strong>of</strong> chemotherapy<br />

and hormonal therapy for early breast cancer on recurrence and 15-year<br />

survival: an overview <strong>of</strong> the randomised trials. Lancet 2005; 365(9472):1687–1717.<br />

63. Winer EP, Hudis C, Burstein HJ, et al. American Society <strong>of</strong> Clinical Oncology<br />

technology assessment on the use <strong>of</strong> aromatase inhibitors as adjuvant therapy for<br />

postmenopausal women with hormone receptor-positive breast cancer: status report<br />

2004. J Clin Oncol 2005; 23(3):619–629.<br />

64. Piccart-Gebhart MJ, Procter M, Leyland-Jones B, et al. Trastuzumab after adjuvant<br />

chemotherapy in HER2-positive breast cancer. N Engl J Med 2005; 353(16):1659–1672.<br />

65. Romond EH, Perez EA, Bryant J, et al. Trastuzumab plus adjuvant chemotherapy for<br />

operable HER2-positive breast cancer. N Engl J Med 2005; 353(16):1673–1684.<br />

66. Citron ML, Berry DA, Cirrincione C, et al. Randomized trial <strong>of</strong> dose-dense versus<br />

conventionally scheduled and sequential versus concurrent combination chemotherapy<br />

as postoperative adjuvant treatment <strong>of</strong> node-positive primary breast cancer:<br />

first report <strong>of</strong> Intergroup Trial C9741/<strong>Cancer</strong> and Leukemia Group B Trial 9741.<br />

J Clin Oncol 2003; 21(8):1431–1439.<br />

67. Mamounas EP, Bryant J, Lembersky B, et al. Paclitaxel after doxorubicin plus<br />

cyclophosphamide as adjuvant chemotherapy for node-positive breast cancer: results<br />

from NSABP B-28. J Clin Oncol 2005; 23(16):3686–3696.<br />

68. Martin M, Pienkowski T, Mackey J, et al. Adjuvant docetaxel for node-positive<br />

breast cancer. N Engl J Med 2005; 352(22):2302–2313.<br />

69. Piccart-Gebhart MJ. Anthracyclines and the tailoring <strong>of</strong> treatment for early breast<br />

cancer. N Engl J Med 2006; 354(20):2177–2179.<br />

70. Pittman J, Huang E, Dressman H, et al. Integrated modeling <strong>of</strong> clinical and gene<br />

expression information for personalized prediction <strong>of</strong> disease outcomes. Proc Natl<br />

Acad Sci U S A 2004; 101(22):8431–8436.<br />

71. Adjuvant! For <strong>Breast</strong> <strong>Cancer</strong>, Genomic Version 7.0. http://www.adjuvantonline.com/<br />

breastgenomic.jsp 2006.


15<br />

Tissue is the Issue<br />

Adekunle Raji<br />

Eastern Cooperative Oncology Group Pathology Coordinating Office,<br />

Northwestern University, Evanston, Illinois, U.S.A.<br />

INTRODUCTION<br />

All eyes are on tissue. The biggest challenge facing biomedical discoveries today<br />

is the issue with tissue. The era <strong>of</strong> personalized, individually tailored medicine has<br />

begun; therefore, cells, tissue and its intra- and interenvironment, are the focus <strong>of</strong><br />

pharmacogenetics. The intracellular kinetic and intercellular dynamics are key<br />

factors in the understanding <strong>of</strong> cell behavior before, during, and after therapy.<br />

Undoubtedly the solid type <strong>of</strong> tissue carries the bulk <strong>of</strong> these issues. Other<br />

fluid forms <strong>of</strong> tissue, however, do have their own share <strong>of</strong> these issues, <strong>of</strong> course.<br />

Tissue will be defined here as biological material comprising a network <strong>of</strong><br />

characteristically diverse cells. In essence, blood, urine, surgical resections,<br />

biopsies, marrow, lavage, hair, nail, and other specimen taken from a person will<br />

be considered as tissue. Turn to any biomedical community member and most, if<br />

not all, will have some sort <strong>of</strong> opinion about tissue. Between the ultimate giver<br />

(patients) and ultimate user (laboratorians) lie the challenges. To the patient, care<br />

team, and several entities, the tissue is the issue, and to the laboratories, technologists,<br />

and the interpreters <strong>of</strong> results, the issue is really the tissue.<br />

In the past two years, several working groups were established not only to<br />

understand current issues as well as future anticipated issues but also to address<br />

them. These working groups are usually comprise various clinical and financial<br />

institutions and experts from multiple disciplines, such as medical scientists,<br />

basic scientists, surgeons, laboratories, pathologists, funding agencies, agent<br />

213


214 Raji<br />

manufacturers, patient advocates, statisticians, bioinformatics, research associates,<br />

research administrators, medical staff, and clinical administrators. These<br />

groups guide the interdisciplinary patient care team by addressing the procurement,<br />

collection, preservation, acquisition, processing, transportation, and<br />

storage <strong>of</strong> biological materials <strong>of</strong> high quality and usability that are adequate<br />

for analysis and interpretation. Good news is that specimens with clinical data<br />

are archived as far back as 40 years ago, the bad news is that they were<br />

collected non-uniformaly; today’s good news is that the collection processing<br />

andassociateddataareundergoingharmonization (uniformity). Contingency<br />

plans are also being developed for future anticipated issues. The key areas <strong>of</strong><br />

the tissue issue are obviously procurement, collection, acquisition, processing,<br />

preservation, storage, distribution, and bioinformatics that are successfully<br />

conducted within stringent policies, procedures, rules, and regulations.<br />

PROCUREMENT<br />

A strategy for collecting any biological specimens from patients is needed to<br />

ensure that the behavior <strong>of</strong> cells or tissue after it leaves the body <strong>of</strong> the patient is<br />

consistent with what it was while in the patient’s body. Regardless <strong>of</strong> the<br />

strategy, clear and simple instructions should be written and made available to<br />

the procurement team comprising a preoperation nurse, phlebotomist, surgery<br />

assistant, physician/surgeon, physician assistant, pathology assistant, research<br />

coordinator, patient care nurse, internal specimen transporters, and medical/<br />

research technologist. These instructions are typically included in the local<br />

manual <strong>of</strong> operational procedures and/or clinical protocols and should provide<br />

details <strong>of</strong> the type <strong>of</strong> collection devices such as vacutainers, syringes, punch size,<br />

preservatives, buffers, reagents and the containers needed. The manual must also<br />

include procedures for preparing reagents or buffers such as formalin, RNAlater,<br />

and the optimal cutting temperature (OCT) compound. It must contain information<br />

about the precollection requirements with regard to time, temperature,<br />

and pressure for certain types <strong>of</strong> specimens (Table 1)<br />

COLLECTION<br />

There are several choices <strong>of</strong> collection devices, to the extent that they have<br />

created a great deal <strong>of</strong> confusion in the medical field. It is not surprising to see<br />

several new clinical trial protocols now restricting collection to the use <strong>of</strong><br />

preassembled collection devices, with instructions to assist the procurement<br />

plan and ensure uniform collection, particularly in a multi-institutional study.<br />

While it is possible to obtain serum from redtop, royal blue top, serum separator<br />

top, and plain syringe, it is true that tissue can be fixed in paraformaldhyde,<br />

formalin, acid alcohol, zinc formalin, and many other reagents. Each <strong>of</strong> these<br />

fixatives presents its own advantages and disadvantages toward downstream<br />

applications such as immunohistochemistry (IHC), in situ hybridization (ISH),


Tissue is the Issue 215<br />

Table 1<br />

Roles<br />

Roles and Responsibility<br />

Responsibility<br />

<strong>Ph</strong>ysician/surgeon including<br />

but not limited to pathologist<br />

and intervention radiologist<br />

Nurse<br />

<strong>Ph</strong>lebotomist, nurse, medical<br />

technologist<br />

<strong>Ph</strong>ysician/surgery assistant<br />

Transporter<br />

Research coordinators, project<br />

manager, protocol specialist,<br />

data manager, nurse<br />

coordinator, clinical<br />

research associates<br />

Medical/research technologist<br />

Inform patient, order test, perform surgery/biopsy/<br />

imaging/morphology/assay procedures.<br />

Inform patient, determine timing <strong>of</strong> procedure/draw.<br />

Draw blood via port.<br />

Draw blood into devices. Perform aspiration.<br />

Guide physician and expedite preservation and delivery.<br />

Ensure timely delivery from point <strong>of</strong> collection to next<br />

critical step/location while preserving samples.<br />

Ensure all team members are familiar with their roles<br />

and have adequate information to assist in<br />

performing task. Liaison among team members as<br />

well as facilitate collaboration.<br />

Ensure availability <strong>of</strong> appropriate supplies for<br />

collection and preservation. Effectively and<br />

efficiently process specimens and analyze samples.<br />

comparative genomic hybridization (CGH), polymerase chain reaction (PCR),<br />

blotting (Western, Northern, Southern), enzyme-linked immunosorbent assay<br />

(ELISA), radio immunoassay (RIA), enzyme immunoassay (EIA), and chemiluminescence<br />

assay (CLA). It is also true that plasma can be obtained from all<br />

anticoagulated vacutainers as well as heparin-coated syringes. Protein and<br />

nucleic acids can be obtained from biological samples regardless <strong>of</strong> collection<br />

or preservation buffer. However, experience has shown a great deal <strong>of</strong> variation<br />

in the yield and quality <strong>of</strong> sample products, as well as the analytical results<br />

obtained from the samples. It is therefore important to select one device and one<br />

alternate backup for the collection, preservation, and fixation that will provide<br />

the best representation <strong>of</strong> the cellular content and its content behavior as it was<br />

just before leaving the patient’s body. It is important to select a device and<br />

preservative that are best for yield and quality, as well as most appropriate for<br />

the next step in processing or analysis. Table 2 lists the collection devices,<br />

preservatives, and fixatives that have been selected by large research groups<br />

because <strong>of</strong> their overall performance success.<br />

Fresh tissue. Nonfrozen and nonfixed tissue should be collected directly<br />

into a labeled sterile screw-cap plastic container, with one <strong>of</strong> the<br />

following buffers:<br />

1. <strong>Ph</strong>osphate buffered saline (PBS) for tissue culture and cell viability<br />

studies.<br />

2. Neutral buffered formalin (10%) for gradual fixation.


216 Raji<br />

Table 2 Common Blood Collection Tubes<br />

Closure/stopper color Anticoagulant or additive<br />

Number <strong>of</strong><br />

inversions to mix Examples for use<br />

Red/red in glass tube None None Serum for chemistry, proteomics, immunology, and<br />

banking<br />

Red/red in plastic tube Clot activator 5–10 Serum for chemistry, proteomics, immunology, and<br />

banking<br />

Gold/red marbled SST Clot activator and separation gel 5–10 Serum for chemistry, proteomics, immunology, and<br />

banking. Remote collection.<br />

Lavender/lavender<br />

(purple)<br />

Sodium EDTA (dried), or<br />

potassium EDTA (liquid)<br />

Light blue/light blue Liquid 3.2% sodium citrate, a or<br />

3.8% sodium citrate<br />

Cell preparation<br />

tube CPT<br />

Liquid 3.2% sodium citrate, a or<br />

3.8% sodium citrate. Ficol and<br />

gel separator<br />

Paxgene RNA/DNA Preanalytic/Qiagen nucleic acid<br />

stabilizer<br />

5–10 Whole blood for hematology, certain whole blood<br />

chemistries, e.g., hemoglobin A1c or blood typing.<br />

Proteomics, floating DNA, genomic DNA, cells.<br />

5–10 Plasma for coagulation, Platelet poor plasma. Growth<br />

factors, fragile proteins, viable cells.<br />

5–10 Plasma for coagulation, Platelet poor plasma. Growth<br />

5–10 Genomic expressions.<br />

factors, fragile proteins, viable cells. Remote<br />

collection <strong>of</strong> viable mononuclear cells.<br />

P100 Protein stabilizer 5–10 Small size protein measurement. Proteomic studies.<br />

Tan Sodium heparin (glass), or 5–10 Whole blood for lead, heavy metals, toxicology studies.<br />

potassium EDTA (plastic)


Tissue is the Issue 217<br />

Table 2 Common Blood Collection Tubes (Continued )<br />

Closure/stopper color Anticoagulant or additive<br />

Number <strong>of</strong><br />

inversions to mix Examples for use<br />

Green/green Lithium heparin, or 5–10 Plasma or whole blood for chemistry, PK, and PD.<br />

Sodium heparin 5–10 Plasma for chemistry<br />

Light green/green<br />

marbled<br />

Lithium heparin and separation gel 5–10 Plasma for chemistry<br />

Gray/gray Potassium oxalate/sodium fluoride,<br />

or sodium fluoride, or lithium<br />

iodoacetate, or lithium<br />

iodoacetate/lithium heparin<br />

5–10 Serum or plasma for glucose or alcohol requiring<br />

inhibition <strong>of</strong> glycolysis, C-peptides.<br />

Dark royal blue Sodium heparin, EDTA, or plain 5–10 Plasma or serum for trace elements, toxicology, and<br />

nutrients.<br />

Pale yellow/bright yellow Sodium polyanetholesulfonate 5–10 Blood cultures, microbiology, and infectious diseases<br />

markers.<br />

Yellow Acid citrate dextrose (ACD) 5–10 Blood bank, viable cells alternate to CPT and CellSave<br />

Orange/yellow and<br />

marbled<br />

Clot activator (thrombin) 5–10 Serum for chemistry, ELISA<br />

Pink/pink Potassium EDTA 5–10 Molecular/viral load testing, blood bank, hematology.<br />

a www.bd.com<br />

Abbreviations: SST, serum separator top; PK, pharmacokinetic; PD, pharmacodynamic.


218 Raji<br />

3. RPMI (Roswell Park Memorial Institute) media for tissue culture, cell<br />

viability studies, protein extraction, and multiphoton studies.<br />

4. RNAlater for RNA extraction.<br />

5. OCT bottom and top prior to freezing.<br />

PROCESSING<br />

Generally, processing methods that will ensure optimal quality and yield should be<br />

selected for the processing <strong>of</strong> specimens and samples into next product aliquot.<br />

Blood. It is important to perform all centrifugation in refrigerated and<br />

temperature-controlled chambers. This is to avoid unnecessary heat<br />

incubation <strong>of</strong> samples while spinning. In some instances, it will be<br />

necessarytospinthesampleataslowspeedfollowedbyhighspeedto<br />

eliminate platelets. Selection <strong>of</strong> complex processes to obtain delicate<br />

and fragile by-products is important. Aliquoting into the smallest<br />

useable quantity to avoid freeze-thaw-freeze cycle is also necessary.<br />

For example, serum and plasma can be apportioned into 200 mL, DNA<br />

andRNAcanbeapportionedinto75mg and1mg, respectively.<br />

Fresh frozen tissue. OCT-protected frozen tissue should be cryosectioned<br />

for hematoxylin and eosin staining to assess tumor area and percentage<br />

as part <strong>of</strong> the annotation. The information will guide case selection as<br />

well as steps in macro and micro dissection. Integrity <strong>of</strong> the sample must<br />

be maintained through sterile processes to avoid degradation and/or<br />

contamination. The tissue must be kept frozen until submersed into<br />

extraction buffer. There are many choices <strong>of</strong> methods for the extraction<br />

<strong>of</strong> cellular components such as RNA, DNA, and proteins.<br />

Fixed tissue. Tissue biopsies and/or resection must undergo complete and<br />

adequate fixation in a buffered environment. This is important for the<br />

morphologic review as well as the evaluation <strong>of</strong> reference and novel<br />

markers. Neutral buffered formalin has been widely used and is most<br />

appropriate for current and future translational processes. As previously<br />

mentioned, there are many choices <strong>of</strong> methods for the extraction and<br />

isolation <strong>of</strong> DNA; however, there are only few methods for the<br />

extraction <strong>of</strong> RNA and protein from fixed tissue. Care must be taken<br />

in making selections to avoid loss <strong>of</strong> valuable tissue resource.<br />

Sectioning should be performed with new blades and clean water.<br />

Nonprecipitating formalin should be avoided.<br />

Xylene rather than its substitutes should be used.<br />

Manual rather than automated processes should be used for small,<br />

starting samples.<br />

Washes between steps in any and all processes must be complete to<br />

avoid reagent carry over.


Tissue is the Issue 219<br />

CURRENT NOVEL PROCESSING OF BLOOD AND FIXED TISSUE<br />

Isolation <strong>of</strong> Mononuclear Cells<br />

Monocytes, a type <strong>of</strong> white blood cells, are found both in bone marrow and blood.<br />

The following procedure describes the isolation <strong>of</strong> monocytes in peripheral blood<br />

(peripheral blood mononuclear cells, PBMC).<br />

Steps<br />

1. Centrifuge CPT tubes containing whole blood for 15 minutes at 258C at<br />

2700 rpm.<br />

2. Pipet out the top layer (plasma) and transfer into two labeled tubes and<br />

freeze.<br />

3. Pipet out the lymphocyte layer (the cloudy layer) with a transfer pipette,<br />

placing the layer into a fresh 15-mL tube. Fill tube with room-temperature<br />

PBS.<br />

4. Centrifuge for 10 minutes at 1800 rpm.<br />

5. Discard supernatant and resuspend pellet in 1 mL PBS.<br />

6. Fill tube with PBS, inverting eight times to mix and wash.<br />

7. Centrifuge for 15 minutes at 2000 rpm.<br />

8. Discard supernatant.<br />

9. Proceed to cell sorting or resuspend and freeze pellet in 1 mL <strong>of</strong> RPMI or<br />

DMSO (dimethyl sulfoxide) for later sort.<br />

Note:<br />

For blood collected in any other anticoagulated tube, dilute 1:1 with<br />

RPMI, layer on Histopaque 1077, then proceed with step 1.<br />

Specialized Sorting <strong>of</strong> Circulating Tumor/Endothelial Cells<br />

1. Magnetic labeling (*45 min)<br />

a. Centrifuge sample for 10 minutes at 300g at room temperature.<br />

b. Carefully remove the supernatant completely.<br />

c. Resuspend cell pellet in appropriate amount <strong>of</strong> buffer as specified by<br />

the microbeads package insert. For fewer cells, use the same volume.<br />

d. Add microbeads per package insert.<br />

e. Incubate for 15 minutes at 68Cto128C. For less than 5 10 6 total cells,<br />

use the same volume.<br />

f. Wash cells by adding 10 to 20 times the labeling volume <strong>of</strong> buffer and<br />

centrifuge at 300g for 10 minutes.<br />

g. Remove supernatant completely (save the supernatant in appropriately<br />

labeled 1 mL cryovial) and resuspend cell pellet in appropriate amount<br />

<strong>of</strong> buffer (1 mL <strong>of</strong> buffer/10 8 total cells).<br />

h. Proceed to magnetic separation.


220 Raji<br />

2. Magnetic separation (*60 min)<br />

a. Turn on the manual and/or automated cell sorter.<br />

b. When the machine has finished initializing, perform a column exchange<br />

to remove blank columns and insert magnetic columns. (To save<br />

time, try to perform these functions while the sample and beads<br />

are incubating.)<br />

c. Wipe down probes with alcohol to disinfect.<br />

d. Place the sample in the uptake position, and place empty 15-mL conical<br />

tubes in the rack under the Neg1, Pos1, and Pos2 probes.<br />

e. Choose “Separation” on the main menu.<br />

f. Use the down arrows to select “possel_s” and press “start.”<br />

g. For both the positive and negative cells, separate elution container<br />

must be used (*15 min).<br />

h. Prepare and/or submit samples for next step in processing (DNA extraction,<br />

thin prep slides, etc.).<br />

Extraction <strong>of</strong> RNA from Peripheral Blood in PAXgene<br />

1. Centrifuge the PAXgene Blood RNA Tube for 10 minutes at 3000g to<br />

5000g using a swing-out rotor.<br />

2. Remove the supernatant by decanting or pipetting. Add 4 mL RNase-free<br />

water to the pellet, and close the tube using a fresh secondary Hemogard<br />

closure.<br />

3. Vortex until the pellet is visibly dissolved, and centrifuge for 10 minutes at<br />

3000g to 5000g using a swing-out rotor. Remove and discard the entire<br />

supernatant.<br />

4. Add 350 mL Buffer BR1, and vortex until the pellet is visibly dissolved.<br />

5. Pipet the sample into a 1.5-mL microcentrifuge tube. Add 300 mL Buffer<br />

BR2 and 40 mL proteinase K. Mix by vortexing for 5 seconds, and incubate<br />

for 10 minutes at 558C using a shaker/incubator at 400 to 1400 rpm.<br />

After incubation, set the temperature <strong>of</strong> the shaker/incubator to 658C (for<br />

step 20).<br />

6. Pipet the lysate directly into a PAXgene Shredder spin column (lilac)<br />

placed in a 2-mL processing tube, and centrifuge for 3 minutes at maximum<br />

speed (but not to exceed 20,000g).<br />

7. Carefully transfer the entire supernatant <strong>of</strong> the flow-through fraction to a<br />

fresh 1.5-mL microcentrifuge tube without disturbing the pellet in the<br />

processing tube.<br />

8. Add 350 mL ethanol (96–100%, purity grade p.a.). Mix by vortexing, and<br />

centrifuge briefly (1–2 seconds at 500–1000g) to remove drops from the<br />

inside <strong>of</strong> the tube lid.<br />

9. Pipet 700 mL sample into the PAXgene RNA spin column (red) placed in a<br />

2-mL processing tube and centrifuge for 1 minute at 8000 to 20,000g.


Tissue is the Issue 221<br />

Place the spin column in a new 2-mL processing tube, and discard the old<br />

processing tube containing flow-through.<br />

10. Pipet the remaining sample into the PAXgene RNA spin column, and<br />

centrifuge for 1 minute at 8000 to 20,000g. Place the spin column in a<br />

new 2-mL processing tube, and discard the old processing tube containing<br />

flow-through.<br />

11. Pipet 350 mL Buffer BR3 into the PAXgene RNA spin column. Centrifuge<br />

for 1 minute at 8000 to 20,000g. Place the spin column in a new 2-mL<br />

processing tube and discard the old processing tube containing flow-through.<br />

12. Add 10 mL DNase I stock solution to 70 mL Buffer RDD in a 1.5-mL<br />

microcentrifuge tube. Mix by gently “flicking” the tube and centrifuge<br />

briefly to collect residual liquid from the sides <strong>of</strong> the tube.<br />

13. Pipet the DNase I incubation mix (80 mL) directly onto the PAXgene RNA<br />

spin column membrane, and place on the bench top (20–308C.) for<br />

15 minutes.<br />

14. Pipet 350 mL Buffer BR3 into the PAXgene RNA spin column, and<br />

centrifuge for 1 minute at 8000g to 20,000g. Place the spin column in a<br />

new 2-mL processing tube and discard the old processing tube containing<br />

flow-through.<br />

15. Pipet 500 mL Buffer BR4 to the PAXgene RNA spin column, and<br />

centrifuge for 1 minute at 8000g to 20,000g. Place the spin column in a<br />

new 2-mL processing tube, and discard the old processing tube containing<br />

flow-through.<br />

16. Add another 500 mL Buffer BR4 to the PAXgene RNA spin column.<br />

Centrifuge for 3 minutes at 8000g to 20,000g.<br />

17. Discard the tube containing the flow-through, and place the PAXgene<br />

RNA spin column in a new 2-mL processing tube. Centrifuge for 1 minute<br />

at 8000g to 20,000g.<br />

18. Discard the tube containing the flow-through. Place the PAXgene RNA<br />

spin column in a 1.5-mL microcentrifuge tube, and pipet 20 mL Buffer<br />

RNAse-free water directly onto the PAXgene RNA spin column membrane.<br />

Centrifuge for 1 minute at 8000g to 20,000g to elute the RNA.<br />

19. Repeat the elution step by adding the eluted RNA to the same PAXgene<br />

RNA spin column membrane. Centrifuge for 1 minute at 8000g to 20,000g<br />

to elute the RNA.<br />

20. Incubate the eluate for 5 minutes at 658C. in the shaker/incubator without<br />

shaking. After incubation, chill immediately on ice.<br />

21. If the RNA samples will not be used immediately, store at 208Cor 708C.<br />

Laser Capture Microdissection<br />

Laser capture microdissection (LCM) allows for the microscopic removal <strong>of</strong><br />

specific cells from paraffin or frozen tissue sections mounted on glass slides.


222 Raji<br />

This procedure describes the isolation <strong>of</strong> a pure population <strong>of</strong> cells followed by<br />

its nucleic acid or protein extraction.<br />

1.1. Prepare paraffin section on glass slide.<br />

a. Collect 5 mm paraffin section onto uncoated glass slide, using microtome<br />

and water bath.<br />

b. Allow to air-dry overnight at room temperature or at 428C for up to<br />

8 hours.<br />

c. Deparaffinize and lightly stain slide(s) by immersing in the following:<br />

i. Xylene 2: 5 minutes each<br />

ii. 100% ethanol: 1 minute<br />

iii. 95% ethanol: 1 minute<br />

iv. 70% ethanol: 1 minute<br />

v. Distilled water: 1 minute<br />

vi. Harris hematoxylin: 1 dip (Fisher’s Protocol #245–678)<br />

vii. Tap water: 2 minutes<br />

viii. Distilled water: 1 minute<br />

ix. 95% ethanol: 1 minute<br />

x. Eosin Y: 1 dip (Fisher’s Protocol #314–631)<br />

xi. 95% ethanol: 1 minute<br />

xii. Allow to air-dry<br />

or<br />

1.2. Prepare frozen section on glass slide.<br />

a. Collect 8 mm frozen section on an uncoated glass slide, using cryostat.<br />

b. Store frozen until ready to postfix.<br />

c. Postfix and lightly stain.<br />

d. Remove a maximum <strong>of</strong> four slides from cryostat or freezer, place on<br />

clean paper towel, i.e., KimWipe, and allow to thaw for approximately<br />

30 seconds.<br />

e. Place the slides in 75% ethanol for 30 seconds.<br />

f. Transfer the slides to distilled water for 30 seconds.<br />

g. Place slides on KimWipe, apply 100 mL <strong>of</strong> histogene staining<br />

solution, and stain for approximately 20 seconds.<br />

h. Place the slides in distilled water for 30 seconds.<br />

i. Place the slides for 30 seconds in each <strong>of</strong> the following ethanols in<br />

increasing concentration: 75%, then 95%, and then 100%.<br />

j. Transfer the slides to xylene for 5 minutes.<br />

k. Place the slides on a KimWipe to dry in a ventilation hood for 5 minutes.<br />

2. Dissect area <strong>of</strong> tissue from slide (to be performed same day as steps 1.1<br />

or 1.2).<br />

a. Place slide on Pixcell II LCM [Arcturus (now part <strong>of</strong> Molecular<br />

Devices), Mountain View, California, U.S.] and locate area <strong>of</strong><br />

interest using 4 or 10 objective.


Tissue is the Issue 223<br />

b. Load CapSure HS caps.<br />

c. Turn on laser key, push laser enable button, and locate laser spot on<br />

screen.<br />

d. Focus laser beam to bright, well-defined spot.<br />

e. Test laser, observing melted plastic ring(s).<br />

See following table for power and duration.<br />

Laser spot size (mm) Power (mW) Duration<br />

CapSure HS Small (7.5) 65–75 650–750 ms<br />

Medium (15) 35–45 2.5–3.0 ms<br />

Large (30) 45–55 6.0–7.0 ms<br />

f. View captured cells on cap.<br />

g. Move cap to capping station.<br />

h. Transfer to ExtractSure device from CapSure HS cap.<br />

3. Extract DNA from dissected tissue.<br />

a. Pipet 155 mL <strong>of</strong> reconstitution buffer into one vial <strong>of</strong> proteinase K.<br />

Gently vortex the tube to mix the reagents and immediately place<br />

the tube on ice. Completely dissolve reagent, excessive mixing can<br />

denature proteinase K.<br />

b. Assemble the ExtractSure Sample Extraction Device onto the CapSure<br />

HS LCM Cap.<br />

c. Using clean forceps, remove the cap from the cap insertion tool and<br />

place into the alignment tray (sample facing up).<br />

d. Make sure that the cap snaps securely and lies flat on the bottom <strong>of</strong> the<br />

opening <strong>of</strong> the alignment tray.<br />

e. Position the ExtractSure Device over the cap using clean forceps. The<br />

fill-port should be facing up.<br />

f. Using forceps, push the ExtractSure Device down onto the cap until it<br />

snaps securely into place. Use forceps to ensure that the cap is firmly<br />

attached to the ExtractSure Device.<br />

g. Pipet 10 mL <strong>of</strong> proteinase K extraction solution into the microchamber<br />

formed by the assembled ExtractSure Device and CapSure HS LCM<br />

Cap.<br />

h. Using gloved hands, place a 0.5-mL thin-walled reaction tube over the<br />

ExtractSure device.<br />

i. Cover the assembly in the alignment tray with the incubation block and<br />

incubate it overnight at 658C.<br />

j. After incubation, remove the assembled CapSure HS LCM Cap and<br />

ExtractSure Device from the incubator, place the assembly into a<br />

microcentrifuge and centrifuge for 1 minute at 4000g.


224 Raji<br />

k. Separate the CapSure HS LCM Cap and ExtractSure Device. Close the<br />

microcentrifuge tube containing the extract and heat to 958C for<br />

10 minutes in a heating block to inactivate the proteinase K. Cool<br />

the sample to room temperature.<br />

Isolation <strong>of</strong> DNA, RNA, Protein Using TRIZOL Method<br />

Isolation <strong>of</strong> RNA by TRI Reagent<br />

Homogenization Homogenize fresh tissue samples (that have been stored in<br />

RNAlater) in TRI Reagent (1 mL/50–100 mg tissue) using a glass-Teflon or<br />

Polytron homogenizer. Sample volume should not exceed 10% <strong>of</strong> the volume <strong>of</strong><br />

TRI Reagent used for homogenization.<br />

<strong>Ph</strong>ase separation<br />

1. Store the homogenate for 5 minutes at room temperature to permit the<br />

complete dissociation <strong>of</strong> nucleoprotein complexes.<br />

2. Add 0.2 mL chlor<strong>of</strong>orm to the homogenate, cover the samples tightly, and<br />

shake vigorously for 15 seconds.<br />

3. Store the resulting mixture at room temperature for 2 to 15 minutes and<br />

centrifuge at 12,000g for 15 minutes at 48C.<br />

4. Following centrifugation, the mixture separates into a lower red phenolchlor<strong>of</strong>orm<br />

phase, interphase, and the colorless upper aqueous phase. RNA<br />

remains exclusively in the aqueous phase, whereas DNA and proteins are<br />

in the interphase and organic phase. The volume <strong>of</strong> the aqueous phase is<br />

about 60% <strong>of</strong> the volume <strong>of</strong> TRI Reagent used for homogenization.<br />

RNA precipitation<br />

1. Transfer the aqueous phase to a fresh tube.<br />

2. Precipitate RNA from the aqueous phase by mixing with 0.5 mL <strong>of</strong><br />

isopropanol.<br />

3. Store samples at room temperature for 5 to 10 minutes and centrifuge at<br />

12,000g for eight minutes at 48C to258C. RNA precipitate (<strong>of</strong>ten invisible<br />

before centrifugation) forms a gel-like or white pellet on the side and<br />

bottom <strong>of</strong> the tube.<br />

RNA wash<br />

1. Remove the supernatant and wash the RNA pellet (by vortexing) with 75%<br />

ethanol (1 mL).<br />

2. Centrifuge at 7500g for 5 minutes at 48C to258C.


Tissue is the Issue 225<br />

RNA solubilization<br />

1. Remove the ethanol wash and briefly air-dry the RNA pellet for 3 to 5 minutes.<br />

2. Dissolve RNA water by passing the solution a few times through a pipette<br />

tip and incubating for 10 to 15 minutes at 55 to 608C.<br />

Isolation <strong>of</strong> DNA by TRI Reagent<br />

The DNA is isolated from the interphase and phenol phase separated from the<br />

initial homogenate as described in the RNA isolation protocol.<br />

DNA precipitation<br />

1. Remove the remaining aqueous phase overlying the interphase. Precipitate<br />

DNA from the interphase and organic phase with ethanol.<br />

2. Add 0.3 mL <strong>of</strong> 100% ethanol and mix samples by inversion.<br />

3. Store the samples at room temperature for 2 to 3 minutes and sediment<br />

DNA by centrifugation at 2000g for 5 minutes at 48C to258C.<br />

DNA wash<br />

1. Wash the DNA pellet twice in a solution containing 0.1 M trisodium citrate<br />

in 10% ethanol (1 mL).<br />

2. Store the DNA pellet in the washing solution for 30 minutes at room temperature<br />

with periodic mixing and centrifuge at 2000g for 5 minutes at 48Cto258C.<br />

3. Suspend the DNA pellet in 75% ethanol (1.5–2 mL <strong>of</strong> 75% ethanol). Store<br />

for 10 to 20 minutes at room temperature with periodic mixing and centrifuge<br />

at 2000g for 5 minutes at 48C to258C. This ethanol wash removes<br />

pinkish color from the DNA pellet.<br />

DNA solubilization<br />

1. Remove the ethanol wash and briefly air-dry the DNA pellet by keeping<br />

tubes open for 3 to 5 minutes at room temperature.<br />

2. Dissolve the DNA pellet in 8 mM NaOH by slowly passing through a<br />

pipette. Add (0.3 mL) <strong>of</strong> 8 mM NaOH.<br />

3. Centrifuge at 12,000g for 10 minutes and transfer the resulting supernatant<br />

containing DNA to new tubes.<br />

Isolation <strong>of</strong> Proteins by TRI Reagent<br />

Protein precipitation<br />

1. Aliquot a portion <strong>of</strong> the phenol-ethanol supernatant (0.2–0.5 mL, 1 volume)<br />

into a micr<strong>of</strong>uge tube.


226 Raji<br />

2. Precipitate proteins by adding 3 volumes <strong>of</strong> acetone.<br />

3. Mix by inversion for 10 to 15 seconds to obtain a homogeneous solution.<br />

4. Store samples for 10 minutes at room temperature and sediment the protein<br />

precipitate at 12,000g for 10 minutes at 48C.<br />

Protein wash<br />

1. Decant the phenol-ethanol supernatant and disperse the protein pellet in<br />

0.5 mL <strong>of</strong> 0.3 M guanidine hydrochloride in 95% ethanol þ 2.5 %<br />

glycerol.<br />

2. Disperse the pellet using a pipette tip. After dispersing the pellet, add<br />

another 0.5 mL aliquot <strong>of</strong> the guanidine hydrochloride/ethanol/glycerol<br />

wash solution to the sample and store for 10 minutes at RT.<br />

3. Sediment the protein at 8000g for 5 minutes.<br />

4. Decant the wash solution and perform two more washes in 1 mL each <strong>of</strong><br />

the guanidine/ethanol/glycerol wash solution. Disperse the pellet by vortexing<br />

after each wash to efficiently remove residual phenol.<br />

5. Wash in 1 mL <strong>of</strong> ethanol containing 2.5 % glycerol.<br />

6. Sediment the protein at 8000g for 5 minutes.<br />

7. Decant the alcohol, invert the tube, and dry the pellet for 7 to 10 minutes at<br />

room temperature.<br />

Protein solubilization<br />

1. After briefly air-drying the protein pellet, add 1% SDS (0.2 mL) to the<br />

protein pellet.<br />

2. Gently disperse and solubilize the pellet for 15 to 20 minutes by flicking<br />

the tube or pipetting as required.<br />

Storage <strong>of</strong> RNA, DNA, and/or Protein<br />

Store the solubilized proteins and DNA at 208C or colder, store RNA at 708C<br />

or colder.<br />

Tissue Microarray Creation<br />

This procedure describes the use <strong>of</strong> the MTA-1 (Manual Tissue Arrayer) by<br />

Beecher Instruments, Inc. (Sun Prairie, Wisconsin, U.S.) to create tissue microarray<br />

(TMA) blocks. TMA is a technology which facilitates research by permitting<br />

the collection <strong>of</strong> biological specimens representing a great many individuals<br />

without requiring a great deal <strong>of</strong> storage space. TMA eases comparison <strong>of</strong> the<br />

histological characteristics <strong>of</strong> different patients’ pathologies by placing many<br />

different tissue samples together on the same microscope slide for observation.


Tissue is the Issue 227<br />

Also, TMA helps prevent the unnecessary exhaustion <strong>of</strong> irreplaceable pathological<br />

material, since it permits the same diagnostic and research objectives to be achieved<br />

with far less tissue than was required for such objectives before TMA<br />

existed.<br />

A core <strong>of</strong> paraffin is removed from a “recipient” paraffin block (one<br />

embedded without tissue) and the remaining empty space is filled with a core <strong>of</strong><br />

paraffin-embedded tissue from a “donor” block. This process can be repeated up<br />

to 230 times for one TMA block, resulting in a block containing mapped pieces<br />

from many different blocks, and when sectioned, produces a slide containing<br />

mapped pieces from many different blocks. A corresponding map records each<br />

core’s location in the grid and the information connected with it (patient ID<br />

number, study arm, etc.). The array sections can be used for all histological<br />

staining, including hematoxylin & eosin (H&E), IHC, and ISH.<br />

1. Prepare the donor blocks and corresponding H&E slides.<br />

a. Re-embed tissue, as needed, for more uniform wax consistency and<br />

optimal tissue orientation.<br />

b. Donor tissue should be at least 1.0 mm thick for best results.<br />

c. Prepare a fresh H&E slide from each re-embedded potential donor block.<br />

2. Pathologist characterizes each H&E slide for appropriateness <strong>of</strong> diagnosis<br />

and percentage <strong>of</strong> tumor.<br />

a. Examine each H&E slide microscopically and indicate, with marker,<br />

the area <strong>of</strong> interest (AOI). This area corresponds to the area on the<br />

block from which the core(s) will be taken.<br />

3. Prepare the recipient (TMA) block.<br />

a. Label a blank cassette, using the automated cassette labeler.<br />

b. Determine the size <strong>of</strong> the TMA block and use:<br />

M—the medium mold (30 mm 23 mm) or<br />

L—the large mold (35 mm 23 mm)<br />

c. Make a paraffin block, containing no tissue, using the labeled cassette.<br />

d. Assure that the paraffin block surface is flat and parallel to the<br />

underside <strong>of</strong> plastic cassette and that it contains no air bubbles.<br />

4. Design the array.<br />

a. Determine the diameter <strong>of</strong> the cores to be removed from the donor<br />

blocks. Available core sizes include 0.6, 1.0, 1.5, and 2.0 mm. Core<br />

size is <strong>of</strong>ten indicated by protocol, i.e., breast cancer studies require<br />

the smallest core size—0.6 mm—to conserve tissue while removing<br />

three cores from each donor block during the routine production<br />

<strong>of</strong> three TMA blocks from each group <strong>of</strong> subjects.<br />

b. Allow for enough space (3.0 mm) around the edge <strong>of</strong> the block to avoid<br />

cracking paraffin.


228 Raji<br />

c. Determine the number <strong>of</strong> columns and number <strong>of</strong> rows to use by noting<br />

the following table, which describes the maximum number <strong>of</strong> cores per<br />

TMA block, with respect to core size and mold size.<br />

Maximum Number <strong>of</strong> Cores per TMA Block<br />

Core size (mm)<br />

Spacing between<br />

cores (mm)<br />

Mold size<br />

Maximum no.<br />

<strong>of</strong> cores/spaces<br />

0.6 a yellow 1.5 LM 20 12 ¼ 240<br />

1.0 a green 2.0 LM 15 9 ¼ 135<br />

1.5 a black 2.5 LM 12 7 ¼ 84<br />

2.0 a white 3.0 LM 10 6 ¼ 60<br />

a Punches are color coded.<br />

5. Create a map, identifying each core’s position on the grid its donor block.<br />

a. Enter information into Excel Program (TMA Mapping Tool, Attachment<br />

1) located on Server/Pathcore/TMA. TMA may be mapped using<br />

alternate methods.<br />

b. Include cores from tissue and cell line control blocks, in duplicate.<br />

i. Tissue blocks, i.e., for breast protocols:<br />

Fibroadenoma<br />

Normal endometrium<br />

Normal salivary gland<br />

Normal testis<br />

ii.<br />

Cell line blocks, i.e., for breast protocols:<br />

MCF7<br />

SKBR3<br />

T47D<br />

c. Each TMA block will be asymmetric to assure map orientation, i.e.:<br />

2nd column blank,<br />

4th row blank, and<br />

bottom right space blank<br />

6. Construct the TMA.<br />

a. Install the pair <strong>of</strong> tissue punches, for desired core size, in MTA-1. Install<br />

the slightly larger <strong>of</strong> the pair (identified in blue) on the right (for the<br />

donor block). Install the slightly smaller <strong>of</strong> the pair (identified in red) on<br />

the left for the (for the recipient or TMA block).<br />

b. Insert the recipient (TMA) paraffin block into the MTA-1 block holder.<br />

Secure in place by tightening two screws in block holder with MTA-1<br />

allen wrench.


Tissue is the Issue 229<br />

c. Remove the first core from TMA block.<br />

i. Assure that recipient punch is swung into place.<br />

ii. Align punch over site <strong>of</strong> first core, usually top left corner, by<br />

turning micrometers.<br />

iii. Zero micrometers by pressing the X and Y “ZERO/ABS” buttons.<br />

iv. Gently lower punch into TMA block, using care not to punch too<br />

deeply. Maximum punch depth can be adjusted with large vertical<br />

screw on the left.<br />

v. Raise punch.<br />

vi. Remove paraffin core from punch, using punch’s stylet.<br />

d. Remove core from first donor block.<br />

i. Place the donor block bridge over the TMA block holder.<br />

ii. Swing donor punch into place.<br />

iii. Manually align donor block so that AOI is directly under the<br />

donor punch.<br />

iv. Lower and raise punch.<br />

v. Remove core <strong>of</strong> tissue from punch using punch stylet.<br />

e. Place donor core into TMA block.<br />

i. Remove donor block bridge from over the TMA block holder,<br />

noting location <strong>of</strong> delicate loose core <strong>of</strong> tissue.<br />

ii. Using forceps, align donor core over hole created for it in the<br />

TMA block.<br />

iii. Lower donor core into hole and assure flat level block surface by<br />

pressing evenly with clean glass slide.<br />

f. Adjust the micrometer(s) to move the precision guide to the next X/Y<br />

position. Follow the Spacing Between Cores (mm) Column in the<br />

maximum number <strong>of</strong> cores per TMA block table above to determine<br />

the appropriate increments.<br />

g. Repeat this cycle (steps 6c–6f) to construct the whole array.<br />

h. To punch holes on right half <strong>of</strong> TMA block, rotate it 1808 and replace<br />

in recipient block holder, because the x axis micrometer should not be<br />

extended beyond 9 mm.<br />

i. Smooth and level surface <strong>of</strong> TMA block.<br />

i. Remove the TMA block from the recipient block holder.<br />

ii. Place the TMA block in oven at 378C, maximum temperature, for<br />

30 minutes.<br />

iii. Gently press a glass slide on the top <strong>of</strong> the TMA block, applying<br />

even pressure to push all tissue cores on the block to same level.


230 Raji<br />

7. Section TMA block.<br />

i. Cut a section from the TMA block and stain the slide with H&E for<br />

QA/QC review by the pathologist.<br />

ii. Store TMA block and save TMA map for any future sectioning requests.<br />

Label side <strong>of</strong> block<br />

2nd column left blank.<br />

4th row left blank<br />

Bottom right corner left blank


16<br />

Acquisition and Preservation <strong>of</strong> Tissue<br />

for Microarray Analysis<br />

Brian R. Untch and John A. Olson<br />

Institute for Genome Sciences and Policy and the Department <strong>of</strong> Surgery,<br />

Duke University Medical Center, Durham, North Carolina, U.S.A.<br />

INTRODUCTION<br />

Increasingly, collection <strong>of</strong> normal and neoplastic human tissue is being incorporated<br />

in the design <strong>of</strong> clinical trials and in routine clinical settings under tissue<br />

banking programs. The goal behind these activities is either to facilitate identification<br />

and development <strong>of</strong> robust predictive or prognostic biomarkers or to<br />

enable basic investigation <strong>of</strong> tumor or host biology.<br />

As experience in tissue banking has broadened, the role <strong>of</strong> rigorous collection<br />

protocols, tissue handling, and annotation has become increasingly appreciated.<br />

To this end, several groups have published guiding principles and standard<br />

operating procedures (SOPs) designed to advance the fields <strong>of</strong> tissue banking,<br />

biomarker development, and tumor biology (1,2).<br />

This chapter is designed to summarize key points in the process <strong>of</strong> informed<br />

consent, tissue collection, tissue preservation and storage, as well as extraction <strong>of</strong><br />

macromolecules <strong>of</strong> interest.<br />

INFORMED CONSENT<br />

Tissue banking requires informed consent by patients donating their tissues. The<br />

informed consent document should include standard topics such as study design,<br />

study purpose, and patient risks/benefits. Privacy rights are paramount for<br />

231


232 Untch and Olson<br />

patients when their health information is stored in databases, perhaps indefinitely.<br />

The Health Insurance Portability and Accountability Act and the Common<br />

Rule are federal regulations that require protection <strong>of</strong> patient identification.<br />

Database information and patient identifiers (name, social security number,<br />

institutional medical record ID) should remain separate. An arbitrary code system<br />

should be created to uniquely label specimens and data. The informed<br />

consent document should specify which investigator will keep the anonymous<br />

identifiers and patient identifiers together. In the event that third-party sharing <strong>of</strong><br />

patient identifier data becomes necessary, written authorization must be obtained<br />

from the patient. Also specified in the informed consent document is the<br />

patients’ right to withdraw from the study or have their specimen removed from<br />

the tissue databank. Contact information for these requests is provided to the<br />

patient at the time <strong>of</strong> document signing.<br />

A recent topic <strong>of</strong> debate has been whether giving consent for tissue<br />

banking also allows for indiscriminate investigations related to and unrelated to<br />

the initial study design. Implicit to the potential <strong>of</strong> tissue banking is the use <strong>of</strong><br />

specimens for areas <strong>of</strong> research that are unanticipated or currently unknown.<br />

However, the use <strong>of</strong> specimens for research not specifically noted in the<br />

informed consent document may violate religious and personal beliefs. The<br />

inclusion <strong>of</strong> broad consent for any possible project has been proposed by some as<br />

a solution (3). Others argue that this does not constitute informed consent, as the<br />

possibilities for potential research are unpredictable (4,5). Our approach has been<br />

to specifically describe research to be performed on collected specimens. This<br />

prevents any misunderstanding between patient and researcher. This is not useful<br />

for large repositories that collect tissue for future investigations.<br />

TISSUE ACQUISITION<br />

When tissues are devascularized and removed from the body, cellular metabolism<br />

is altered as hypoxia and acidification occur and nutrient delivery ceases.<br />

At a minimum, macromolecules in ischemic tissues are subject to degradation.<br />

Further, cellular response to ischemia may lead to stress-specific changes in gene<br />

expression. This may ultimately lead to changes in gene and protein expression<br />

or degradation until the cellular machinery stops or tissue preservation occurs.<br />

To accurately and reproducibly analyze gene and protein expression events<br />

within cells, tissue must be preserved as quickly and uniformly as possible.<br />

For all protocols where banking is superimposed upon clinical care, above<br />

all else, banking strategies must not interfere with pathologic assessment <strong>of</strong><br />

either the tumor or specimen resection margins. When designing these protocols,<br />

involvement <strong>of</strong> the pathologist who is custodian <strong>of</strong> the tissue record is critical.<br />

This is not to say that banking cannot occur before specimen processing in<br />

pathology, but that overall specimen integrity must be maintained during the<br />

research specimen collection (e.g., core biopsy <strong>of</strong> lumpectomy specimens).<br />

When a pathologist is immediately available, specimen banking after pathologic


Acquisition and Preservation <strong>of</strong> Tissue for Microarray Analysis 233<br />

gross evaluation is preferred. However, if a pathologist is not immediately<br />

available, then tumor specimens must be collected before pathologic assessment<br />

but without specimen disruption. Regardless <strong>of</strong> when during the processing <strong>of</strong><br />

the pathologic specimen banking occurs, representative samples <strong>of</strong> all banked<br />

tissue should be reviewed by the pathologist to both ensure that critical diagnostic<br />

material is not lost and to document that research tissue samples are<br />

indeed tumor tissues.<br />

To further ensure that research tissue acquisition will not have an impact<br />

on diagnosis, frozen specimens should be held and not analyzed until definitive<br />

pathologic assessment <strong>of</strong> the entire remaining specimen has been completed. At<br />

the time <strong>of</strong> definitive diagnosis <strong>of</strong> the surgical specimen, research specimens can<br />

then be deposited in the tumor bank (Fig. 1). If analysis <strong>of</strong> the surgical specimen<br />

does not demonstrate the abnormal pathology, then research specimens may be<br />

Figure 1 Flow diagram demonstrating the sample handling <strong>of</strong> core biospy specimens in<br />

patients suspected <strong>of</strong> having breast cancer. Research and clinical core biopsies should be<br />

pathologically examined so that the patient’s diagnosis is not affected and that research<br />

specimens contain the tissue <strong>of</strong> interest.


234 Untch and Olson<br />

used for frozen-section analysis. Should the frozen section <strong>of</strong> research specimens<br />

reveal abnormal histology not present in the surgical specimen (e.g., neoplasia<br />

is present in the research specimens and not in the surgical specimen), all<br />

samples can be turned over to pathology for routine processing. While this may<br />

result in a short delay as surgical pathology is reevaluating frozen research<br />

specimens, this can be explained in a preoperative consent form.<br />

In all cases, frozen research specimens should be held frozen until two<br />

weeks after a definitive pathologic report is issued on the diagnostic specimens.<br />

This hold period is desirable, as once DNA and RNA are extracted, no further<br />

diagnostic information relevant to patient care will be available.<br />

For clinical trials and research protocols, specimen collection needs to be<br />

standardized so that tumor samples are treated in a similar manner to minimize<br />

technical variability in gene and protein expression data. Several cooperative<br />

groups have developed banking SOPs for collection <strong>of</strong> formalin-fixed specimens<br />

and blood samples; however, few have until recently focused on frozen-tissue<br />

collection. The American College <strong>of</strong> Surgeons Oncology Group (ACOSOG) has<br />

developed effective procedures for collecting tumor tissue from breast cancer<br />

patients in neoadjuvant studies with good success. The European Human Frozen<br />

Tumor Tissue Bank has also developed similar SPOs that are followed across<br />

institutions and countries in an effort to reduce variability (2). In the former case,<br />

specialized collection kits have been privately developed and are in use by<br />

ACOSOG investigators for multisite tissue and blood collection at the time <strong>of</strong><br />

diagnostic biopsy and again at surgical resection (Fig. 2). Technicians responsible<br />

for tumor collection prepare reagents and containers before excision.<br />

Figure 2 A stereotactic biopsy device to retrieve breast cancer samples from excised<br />

lumpectomy specimens. (A, arrow) Excised lumpectomy specimens are placed in the<br />

device in an oriented fashion, and a specimen radiogram is taken to identify the location <strong>of</strong><br />

the tumor. (B) Coordinates are taken from the radiogram and are used to guide core biopsies<br />

taken from the tumor within an example lumpectomy specimen held in fixed orientation in<br />

the device.


Acquisition and Preservation <strong>of</strong> Tissue for Microarray Analysis 235<br />

Samples are required to be snap frozen within five minutes <strong>of</strong> excision,<br />

regardless <strong>of</strong> the method <strong>of</strong> tissue fixation (described subsequently). Rapid and<br />

uniform handling <strong>of</strong> specimens from patient to preservative is essential.<br />

TECHNIQUES OF CRYOPRESERVATION<br />

RNA degradation by tissue RNases can be detrimental to RNA yield and quality,<br />

with implications on microarray and quantitative PCR experiments. Traditional<br />

practice has been to snap freeze samples in liquid nitrogen followed by permanent<br />

storage at a minimum <strong>of</strong> 808C. This prevents changes to the RNA, and<br />

tumor banks using this method have been highly successful in producing highquality<br />

extracted RNA for microarray analysis. Given the complexity <strong>of</strong> sample<br />

acquisition and multi-institutional collection, recent studies have tried to determine<br />

the importance <strong>of</strong> immediate ex vivo freezing. Micke et al. (6) used ex vivo<br />

tumors at room temperature to determine the amount <strong>of</strong> time until RNA degradation<br />

occurred. Distinct ribosomal peaks were identified up to 16 hours,<br />

indicating that this tissue can remain stable for extended periods <strong>of</strong> time. Other<br />

sample-handling issues, such as freeze-thawing cycles, have also been studied.<br />

Jochumsen et al. (7) established that gene expression analysis is not altered in up<br />

to three freeze-thaw cycles. Despite these reports, it has been our experience that<br />

RNA degradation can be extremely variable, even when SOPs are employed<br />

(Figs. 3,4). To ensure quality RNA, samples should be frozen as soon as possible<br />

after being devascularized.<br />

Figure 3 RNA degradation is variable between two breast tumor samples. Samples are<br />

subjected to the same conditions over 120 minutes. Sample 2 degrades faster than sample 1,<br />

demonstrating the stability variation.


236 Untch and Olson<br />

Figure 4 The effect <strong>of</strong> two-hour extended ischemia time on microarray data. The same<br />

sample had RNA extracted at time 0 and after two hours and hybridized to the U95<br />

Affymetrix platform. Raw expression values were log normalized and graphed. Data<br />

demonstrate that at lower expression values greater variability exists between prolonged<br />

ischemic and nonischemic tissue. There was also variability demonstrated in published<br />

gene signatures describing prognosis. Source: From Refs. 26, 27, 31.<br />

Preservation <strong>of</strong> tissue morphology can be difficult in samples that are snap<br />

frozen. These samples are also subject to desiccation or freezer burn. Samples<br />

can be immersed in optimal cutting temperature (OCT) compound to preserve<br />

tissue morphology and prevent sample damage. This method has been shown to<br />

generate reproducible genomic DNA extraction results (8). Our group has not<br />

observed differences in microarray data when using OCT (unpublished observation).<br />

However, Turbett et al. (9) reported inhibition <strong>of</strong> a polymerase chain<br />

reaction (PCR) assay when compared with non-OCT compound–preserved<br />

specimens. OCT compound can also make specimen thawing and dissection<br />

more difficult. When specimens contain a mix <strong>of</strong> tissue (as in the case <strong>of</strong><br />

undissected tumor specimens with margins), separating whole tumor for RNA<br />

extraction can be difficult, particularly when trying to avoid time-dependant<br />

RNA degradation. Nonetheless, few studies have investigated OCT compound<br />

and its potential to alter microarray results. A recommendation for or against its<br />

use cannot be made at this time.<br />

Because <strong>of</strong> the challenges <strong>of</strong> sample handling (pathology analysis) and the<br />

potential for RNA degradation, RNA stabilization buffers have been developed.<br />

Commercially available RNAlater (Ambion inc.) has been shown to maintain<br />

RNA quality before snap freezing. Several studies have shown consistent<br />

microarray results when storing samples in RNA stabilization buffers (10–12).<br />

One such study found that samples could be stored up to one hour at 378C,<br />

without affecting the RNA or subsequent microarray data (13). These buffers


Acquisition and Preservation <strong>of</strong> Tissue for Microarray Analysis 237<br />

will be important for tissue acquisition protocols, particularly with regard to<br />

multi-institutional trials that will require transfering tissue specimens to a central<br />

tissue bank. This approach allows the samples to be shipped refrigerated rather<br />

than frozen. The principal disadvantages are that proteins are degraded, and<br />

subsequent sectioning for histologic verification <strong>of</strong> tumor identity is problematic.<br />

PARAFFIN-EMBEDED TISSUE FOR MICROARRAY ANALYSIS<br />

The discovery <strong>of</strong> paraffin wax is attributed to the noted hydrocarbon chemist,<br />

Dr. Carl (Karl) Ludwig von Reichenbach (1788–1869). After being developed in<br />

1830, pathology laboratories discovered its potential as a method for storing<br />

and preserving specimens. Throughout the twentieth century, formalin-fixed,<br />

paraffin-embedded tissue became a reliable way to archive tissue. Given the<br />

large quantity <strong>of</strong> specimens available in paraffin, these tissue banks were identified<br />

as a possible tissue source for array-based techniques. Numerous challenges have<br />

been identified that complicate the macromolecular extraction process from these<br />

samples.<br />

Sample handling is an important step that can lead to variability observed<br />

in microarray results. Ischemia time can pr<strong>of</strong>oundly affect protein, DNA, and<br />

RNA extraction because <strong>of</strong> enzymatic degradation. As noted above, sample<br />

collection is ideally performed by snap-freezing tissue in liquid nitrogen<br />

immediately after being devascularized and removed from a patient. Sample<br />

handling <strong>of</strong> formalin-fixed, paraffin-embedded tissue is variable, and this makes<br />

application <strong>of</strong> the array data potentially unreliable.<br />

Many investigators have attempted to develop methods for extraction <strong>of</strong><br />

these samples with mixed results. Given the cross-linking effects <strong>of</strong> formalin, it<br />

was thought that protein extraction for proteomic analysis would be impossible.<br />

However, attempts have been made to isolate proteins from paraffin, but results<br />

have been mixed and more technical development is required (14).<br />

Extraction <strong>of</strong> nucleic acids can be accomplished using commercially<br />

available proteinase digestion buffers (15). This method also eliminates the risk<br />

<strong>of</strong> Rnase activity in the sample as all previously fixed proteins are destroyed.<br />

Formalin fixation does not appear to affect the yield <strong>of</strong> RNA, unless the tissue is<br />

fixed for an extended period <strong>of</strong> time. This can result in cross-linking between<br />

RNA and protein (16). As long as the methylol groups added to the bases during<br />

fixation can be removed, the quality <strong>of</strong> the RNA is not perturbed (17) and results<br />

have been comparable with frozen-tissue controls (18).<br />

The most difficult problem encountered with these specimens has been that<br />

the total DNA or RNA yield from these samples is inadequate and <strong>of</strong>ten<br />

degraded (19). Yield <strong>of</strong> DNA or RNA is dependent on duration <strong>of</strong> protein<br />

digestion, temperature, and pH (20). Most array platforms will require several<br />

micrograms <strong>of</strong> RNA for hybridization. Amplification <strong>of</strong> the RNA has been<br />

suggested as means to increase quantity, but these techniques can result in<br />

considerable variability between samples (21). Further, amplification <strong>of</strong> partially


238 Untch and Olson<br />

degraded samples can also result in unreliable data. Statistical adjustments for<br />

nucleic acid amplification in microarray experiments are being developed and<br />

may eventually prove to be useful.<br />

As extraction, amplification, and statistical techniques improve, formalinfixed,<br />

paraffin-embedded tissue may provide an additional resource for researchers.<br />

Nevertheless, the variability in sample handling history makes application <strong>of</strong> the<br />

array data from these samples less than reliable. At this point in time, use <strong>of</strong> these<br />

samples for application to clinical investigations is discouraged.<br />

FINE NEEDLE ASPIRATION, CORE NEEDLE BIOPSY,<br />

AND SURGICAL SPECIMENS IN MICROARRAY ANALYSIS<br />

Using micorarray analysis, numerous genomic signatures have been developed<br />

that attempt to predict tumor phenotypes, including lymph node metastasis,<br />

tumor growth patterns, postoperative margin positivity, chemosensitivity, and<br />

oncogenic pathway activation. Most <strong>of</strong> these analyses have been based on<br />

samples collected at the time <strong>of</strong> surgery. These large volume samples can yield<br />

adequate RNA for array hybridization and also afford the opportunity to assess<br />

multiple areas <strong>of</strong> the tumor and subcellular compartments via laser capture<br />

microdissection. However, tumor size and growth is known to affect gene<br />

expression, and successful application <strong>of</strong> the developed models to earlier stage,<br />

preoperative biopsy specimens may not be possible.<br />

Nonetheless, application <strong>of</strong> these predictive tests depends on obtaining<br />

adequate tissue for microarray analysis before a definitive surgical intervention.<br />

The preoperative diagnosis <strong>of</strong> breast cancer is typically achieved though two<br />

methods <strong>of</strong> tumor biopsy: core needle biopsy (CNB) and fine needle aspiration<br />

(FNA). For micoarray-based predictive testing, an optimal protocol would allow<br />

these two modalities to collect tissue for pathologic analysis and for molecular<br />

extraction.<br />

FNA is performed by inserting a small gauge needle, typically under image<br />

guidance, into a tumor or lesion. Tissue is removed through a negative-pressure<br />

syringe. Tumor architecture is typically destroyed during this process, but the<br />

resultant sample can be analyzed under the microscope for tumor cell characteristics.<br />

Likewise, this small amount <strong>of</strong> tissue can be harvested for RNA and<br />

microarray analysis. However, concern exists over adequate sample amount (22).<br />

Several studies have shown that FNA can produce tissue in amounts appropriate<br />

for pathology review and array applications (23,24).<br />

CNB is performed under image or visual guidance in a fashion similar to<br />

FNA. However, with CNB, a larger bore needle is used, and the resultant<br />

specimen has maintained tissue architecture, allowing for a more accurate<br />

pathologic analysis. This larger amount <strong>of</strong> tissue can also be used for array-based<br />

studies. Several studies have addressed adequacy <strong>of</strong> tissue acquisition in core


Acquisition and Preservation <strong>of</strong> Tissue for Microarray Analysis 239<br />

biopsy samples (13,24). Studies from our laboratory indicate that core biopsy<br />

specimens can be collected before surgery and can undergo successful RNA<br />

extraction (unpublished observation). The average yield from core biopsy<br />

specimens was 10.4 mg <strong>of</strong> RNA, enough for most array platforms. After<br />

extraction, RNA samples underwent hybridization to the HU95 Affymetrix<br />

microarray platform and then chemosenstivity and oncogenic pathway signatures<br />

were successfully applied. This is consistent with the work <strong>of</strong> Ellis et al. (13),<br />

who was able to extract RNA from core biopsy specimens and distinguish cancer<br />

and noncancer phenotypes.<br />

LASER CAPTURE MICRODISSECTION<br />

Recently, focus has shifted away from whole-tumor microarray analysis to<br />

analysis <strong>of</strong> individual cell types within a tumor. A tumor is composed <strong>of</strong> a<br />

heterogeneous group <strong>of</strong> cells, malignant cells, premalignant cells, normal<br />

parenchymal cells, and supportive cells. This cellular heterogeneity must be<br />

taken into account when applying array-based techniques (25). Numerous successful<br />

expression-pr<strong>of</strong>iling studies have been performed using whole-tumor<br />

macromolecular extraction (26–28). These analyses thus incorporate a variety <strong>of</strong><br />

different cell types. Laser capture microdissection allows one to isolate histologically<br />

distinct cell types and perform subsequent analysis. The technique<br />

employs an infrared laser in conjunction with light microscopy. Prepared histologic<br />

slides are covered with a transfer film. When cells <strong>of</strong> interest are identified,<br />

the laser is activated over the cells. This process causes the cells to attach<br />

to the transfer film. The remainder <strong>of</strong> the sample remains, and the isolated cells<br />

can undergo extraction procedures. Notably, this procedure can result in lower<br />

yields <strong>of</strong> cells requiring amplification <strong>of</strong> nucleic acids for use in microarray<br />

experiments (29,30). Laser capture microdissection has the potential to not only<br />

improve our understanding <strong>of</strong> cellular deregulation but also to help determine<br />

how other cells within the tumor contribute to the overall tumor phenotype.<br />

Immune cells, vascular structures, and normal parenchymal cells within a tumor<br />

may contribute significantly to the growth and metastatic potential. Expression<br />

analysis <strong>of</strong> these individual cell types may help elucidate their roles.<br />

SUMMARY<br />

Acquisition and preservation <strong>of</strong> tissue for microarray experiments requires<br />

understanding <strong>of</strong> informed consent, sample handling, macromolecular extraction,<br />

and sampling methods. Significant advances in these areas have been made<br />

in recent years to improve the quality <strong>of</strong> microarray data. Appreciating these<br />

aspects is paramount if the application <strong>of</strong> microarray techniques is to contribute<br />

meaningfully to the care <strong>of</strong> the cancer patient.


240 Untch and Olson<br />

REFERENCES<br />

1. McShane LM, Altman DG, Sauerbrei W, et al. Reporting recommendations for tumor<br />

marker prognostic studies (REMARK). J Natl <strong>Cancer</strong> Inst 2005; 97(16):1180–1184.<br />

2. Mager SR, Oomen MN, Morente MM, et al. Standard operating procedure for the<br />

collection <strong>of</strong> fresh frozen tissue samples. Eur J <strong>Cancer</strong> 2007; 43(5):828–834.<br />

3. Hansson MG, Dillner J, Bartram CR, et al. Should donors be allowed to give broad<br />

consent to future biobank research? Lancet Oncol 2006; 7(3):266–269.<br />

4. Maschke KJ. Alternative consent approaches for biobank research. Lancet Oncol<br />

2006; 7(3):193–194.<br />

5. Rothstein MA. Expanding the ethical analysis <strong>of</strong> biobanks. J Law Med Ethics 2005;<br />

33(1):89–101.<br />

6. Micke P, Ohshima M, Tahmasebpoor S, et al. Biobanking <strong>of</strong> fresh frozen tissue:<br />

RNA is stable in nonfixed surgical specimens. Lab Invest 2006; 86(2):202–211.<br />

7. Jochumsen KM, Tan Q, Dahlgaard J, et al. RNA quality and gene expression analysis<br />

<strong>of</strong> ovarian tumor tissue undergoing repeated thaw-freezing. Exp Mol Pathol 2007;<br />

82(1):95–102.<br />

8. Sun Y, Hegamyer G, Colburn NH. A simple method using PCR for direct sequencing<br />

<strong>of</strong> genomic DNA from frozen tumor tissue embedded in optimal cutting temperature<br />

compound. Biotechniques 1992; 12(5):639–640.<br />

9. Turbett GR, Sellner LN. The use <strong>of</strong> optimal cutting temperature compound can inhibit<br />

amplification by polymerase chain reaction. Diagn Mol Pathol 1997; 6(5):298–303.<br />

10. Gotzer MA, Patti R, Geoerger G, et al. Biological stability <strong>of</strong> RNA isolated from<br />

RNAlater-treated brain tumor and neuroblastoma xenografts. Med Pediatr Oncol<br />

2000; 34(6):438–442.<br />

11. Florell SR, C<strong>of</strong>fin CM, Holden JA, et al. Preservation <strong>of</strong> RNA for functional<br />

genomic studies: a multidisciplinary tumor bank protocol. Mod Pathol 2001; 14(2):<br />

116–128.<br />

12. Mutter GL, Zahrieh D, Liu C, et al. Comparison <strong>of</strong> frozen and RNALater solid tissue<br />

storage methods for use in RNA expression microarrays. BMC Genomics 2004;<br />

5(1):88.<br />

13. Ellis M, Davis N, Coop A, et al. Development and validation <strong>of</strong> a method for using<br />

breast core needle biopsies for gene expression microarray analyses. Clin <strong>Cancer</strong> Res<br />

2002; 8(5):1155–1166.<br />

14. Becker KF, Schott C, Hipp S, et al. Quantitative protein analysis from formalin-fixed<br />

tissues: implications for translational clinical research and nanoscale molecular<br />

diagnosis. J Pathol 2007; 211(3):370–378.<br />

15. Krafft AE, Duncan BW, Bijwaard KE, et al. Optimization <strong>of</strong> the isolation and<br />

amplification <strong>of</strong> RNA from formalin-fixed, paraffin-embedded tissue: The Armed<br />

Forces Institute <strong>of</strong> Pathology Experience and Literature Review. Mol Diagn 1997;<br />

2(3):217–230.<br />

16. Moller K, Rinke J, Ross A, et al. The use <strong>of</strong> formaldehyde in RNA-protein crosslinking<br />

studies with ribosomal subunits from Escherichia coli. Eur J Biochem 1977;<br />

76(1):175–187.<br />

17. Lewis F, Maughan NJ, Smith V, et al. Unlocking the archive: gene expression in<br />

paraffin-embedded tissue. J Pathol 2001; 195(1):66–71.


Acquisition and Preservation <strong>of</strong> Tissue for Microarray Analysis 241<br />

18. Specht K, Richter T, Muller U, et al. Quantitative gene expression analysis in<br />

microdissected archival formalin-fixed and paraffin-embedded tumor tissue. Am J<br />

Pathol 2001; 158(2):419–429.<br />

19. Jackson DP, Lewis FA, Taylor GR. Tissue extraction <strong>of</strong> DNA and RNA and analysis<br />

by the polymerase chain reaction. J Clin Pathol 1990; 43(6):499–504.<br />

20. Gilbert MT, Haselkorn T, Bunce M, et al. The isolation <strong>of</strong> nucleic acids from fixed,<br />

paraffin-embedded tissues-which methods are useful when? PLoS ONE 2007; 2(6):<br />

E537.<br />

21. Coombs NJ, Gough AC, Primrose JN. Optimisation <strong>of</strong> DNA and RNA extraction<br />

from archival formalin-fixed tissue. Nucleic Acids Res 1999; 27(16):E12.<br />

22. Centeno BA, Enkemann SA, Coppola D, et al. Classification <strong>of</strong> human tumors using<br />

gene expression pr<strong>of</strong>iles obtained after microarray analysis <strong>of</strong> fine-needle aspiration<br />

biopsy samples. <strong>Cancer</strong> 2005; 105(2):101–109.<br />

23. Rapkiewicz A, Expina V, Zujewski JA. The needle in the haystack: application <strong>of</strong><br />

breast fine-needle aspirate samples to quantitative protein microarray technology.<br />

<strong>Cancer</strong> 2007; 111(3):173–184.<br />

24. Symmans WF, Ayers M, Clark EA, et al. Total RNA yield and microarray gene<br />

expression pr<strong>of</strong>iles from fine-needle aspiration biopsy and core-needle biopsy<br />

samples <strong>of</strong> breast carcinoma. <strong>Cancer</strong> 2003; 97(12):2960–2971.<br />

25. Fuller AP, Palmer-Toy D, Erlander MG, et al. Laser capture microdissection and<br />

advanced molecular analysis <strong>of</strong> human breast cancer. J Mammary Gland Biol<br />

Neoplasia 2003; 8(3):335–345.<br />

26. van de Vijver MJ, He YD, van ’t Veer LJ, et al. A gene-expression signature as a<br />

predictor <strong>of</strong> survival in breast cancer. N Engl J Med 2002; 347(25):1999–2009.<br />

27. Sorlie T, Perou CM, Tibshirani R, et al. Gene expression patterns <strong>of</strong> breast carcinomas<br />

distinguish tumor subclasses with clinical implications. Proc Natl Acad Sci<br />

U S A 2001; 98(19):10869–10874.<br />

28. Perou CM, Sorlie T, Eisen MB, et al. Molecular portraits <strong>of</strong> human breast tumours.<br />

Nature 2000; 406(6797):747–752.<br />

29. Caretti E, Devarajan K, Coudry R, et al. Comparison <strong>of</strong> RNA amplification methods<br />

and chip platforms for microarray analysis <strong>of</strong> samples processed by laser capture<br />

microdissection. J Cell Biochem 2008; 103(2):556–563.<br />

30. Schuetz CS, Bonin M, Clare SE, et al. Progression-specific genes identified by<br />

expression pr<strong>of</strong>iling <strong>of</strong> matched ductal carcinomas in situ and invasive breast tumors,<br />

combining laser capture microdissection and oligonucleotide microarray analysis.<br />

<strong>Cancer</strong> Res 2006; 66(10):5278–5286.<br />

31. Ramaswamy S, Ross KN, Lander ES, et al. A molecular signature <strong>of</strong> metastasis in<br />

primary solid tumors. Nat Genet 2003; 33(1):49–54.


17<br />

Tapping into the Formalin-Fixed<br />

Paraffin-Embedded (FFPE) Tissue<br />

Gold Mine for Individualization<br />

<strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong> Treatment<br />

Mark Abramovitz<br />

VM Institute <strong>of</strong> Research, Montreal, Que´bec, Canada<br />

Brian Leyland-Jones<br />

Winship <strong>Cancer</strong> Institute, Emory University School <strong>of</strong> Medicine,<br />

Atlanta, Georgia, U.S.A.<br />

INTRODUCTION<br />

It is now clearly recognized that breast cancer is a heterogeneous disease that<br />

requires personalized targeted therapy in order to have the greatest impact on<br />

survival. This necessitates the development <strong>of</strong> biomarkers that can be used to<br />

predict outcome and response to therapy. Unfortunately, very few biomarkers<br />

have made their way from the research bench to the clinic. The main reason for<br />

this is that in order to validate breast tumor markers, large sample sets from welldefined<br />

patient groups are required, although this could be overcome by tapping<br />

into large numbers <strong>of</strong> well-defined banked tumor specimens. The efforts that are<br />

being made, though, toward the development <strong>of</strong> biomarkers in order to improve<br />

the assessment <strong>of</strong> cancer prognosis and diagnosis are great, given the enormous<br />

potential that they embody.<br />

243


244 Abramovitz and Leyland-Jones<br />

Formalin-fixed, paraffin-embedded (FFPE) tissue samples represent the<br />

largest collection <strong>of</strong> well-annotated, clinical tumor samples that are readily<br />

available for conducting retrospective studies and, therefore, represent a potential<br />

gold mine from which important biomarkers are just waiting to be discovered,<br />

given that the right tools can be brought to bear upon this invaluable resource.<br />

However, until relatively recently, a major caveat regarding the use <strong>of</strong> these FFPE<br />

specimens for genomewide studies was that due to the formalin fixation process,<br />

the standard procedure for handling tumor tissue samples in pathology labs around<br />

the world, RNA degradation occurred, greatly limiting their potential utility. In<br />

fact, with regard to nucleic acid–based assays, FFPE specimens have been mainly<br />

subjected to reverse transcribed-polymerase chain reaction (RT-PCR) analysis.<br />

This chapter will describe the progress that has been made in greatly expanding the<br />

use <strong>of</strong> FFPE tumor specimens for several genomic-based applications that will<br />

greatly increase our ability to tap into this gold mine and extract the biomarker and<br />

genomic data nuggets that will improve cancer prognosis, diagnosis, and treatment.<br />

EXTRACTION OF RNA FROM FFPE TUMOR TISSUE<br />

SAMPLES FOR GENE EXPRESSION ANALYSIS<br />

FFPE tissue samples, a vast archive <strong>of</strong> pathologically well-characterized clinical<br />

samples from randomized trials, are an enormous potential resource that will allow<br />

translational scientists to more fully describe breast cancers at the molecular level<br />

and will ultimately have important etiologic and clinical implications. Even<br />

though the degradation <strong>of</strong> RNA that occurs because <strong>of</strong> the formalin fixation<br />

process results in an RNA species with an average size <strong>of</strong> approximately 200 nt<br />

(1), several groups have illustrated that it is feasible to extract and purify RNA<br />

from such fixed tissue and perform gene expression pr<strong>of</strong>iling (2–9). With the<br />

development <strong>of</strong> real-time quantitative (q) RT-PCR technology that has high<br />

detection sensitivity and high dynamic range, it is now possible to detect even rare<br />

messages in FFPE tissue and to examine the variation <strong>of</strong> expression over quite a<br />

large dynamic range. Amplicons are designed specifically on small segments <strong>of</strong><br />

DNA [


The Formalin-Fixed Paraffin-Embedded (FFPE) Tissue Gold Mine 245<br />

method with multianalyte capability for potential use in clinical research and<br />

diagnostic assays. Two multianalyte assays, one with 48 genes and the other with<br />

92 genes, were designed and tested in different experiments with preserved<br />

human breast cancer tissue. In the first case, all <strong>of</strong> the genes pr<strong>of</strong>iled in FFPE<br />

RNA yielded measurable values, and the overall pr<strong>of</strong>ile was validated by its<br />

similarity to that generated with well-preserved RNA from matched freshly<br />

frozen tissue. In the second experiment using a 92-gene assay, only one <strong>of</strong> the<br />

tested genes failed to yield a signal. Measured levels <strong>of</strong> estrogen receptor (ER),<br />

progesterone receptor (PR), and human epidermal growth factor receptor 2<br />

(HER2) mRNAs were concordant with the levels <strong>of</strong> the respective proteins as<br />

measured by immunohistochemistry (IHC) at an independent clinical reference<br />

laboratory. Approximately 90% concordance was obtained when RT-PCR<br />

expression results for ER and PR were dichotomized into positive and negative<br />

values and compared with ER- and PR-positive and ER- and PR-negative<br />

assignments based on IHC. It is noteworthy that similar levels <strong>of</strong> concordance<br />

were found in other comparative studies <strong>of</strong> IHC versus RT-PCR (15,16).<br />

DEVELOPMENT OF AN RT-PCR-BASED BREAST CANCER<br />

CLINICAL DIAGNOSTIC TEST: ONCOTYPE DX ASSAY<br />

A major development in this area has been the development <strong>of</strong> a multigene qRT-<br />

PCR assay to predict the likelihood <strong>of</strong> breast cancer recurrence in node-negative,<br />

ER-positive, tamoxifen-treated patients (17). The authors studied the likelihood<br />

<strong>of</strong> distant recurrence in breast cancer patients who had involved lymph nodes and<br />

ER-positive tumors that were poorly defined by clinical and histopathologic<br />

measures. They tested whether a prospectively-defined 21-gene RT-PCR assay<br />

<strong>of</strong> FFPE tumor tissue could predict the likelihood <strong>of</strong> distant recurrence in<br />

patients in the National Surgical Adjuvant <strong>Breast</strong> and Bowel Project (NSABP)<br />

B-14 clinical trial. Measured levels <strong>of</strong> gene expression for 16 cancer-related and<br />

5 reference genes were used, by a prospectively defined algorithm, to calculate a<br />

recurrence score (RS) and risk group (low, intermediate, or high) for each<br />

patient. Adequate RT-PCR pr<strong>of</strong>iles were obtained in 668 <strong>of</strong> 675 tumor blocks.<br />

The proportion <strong>of</strong> patients categorized as low, intermediate, or high risk by the<br />

RT-PCR assay was 51%, 23%, and 27%, respectively. The Kaplan–Meier estimates<br />

and 95% confidence intervals for the rates <strong>of</strong> distant recurrence at 10 years<br />

were 6.8% (4.0%, 9.6%), 14.3% (8.3%, 20.3%), and 30.5% (23.6%, 37.4%),<br />

respectively, for the low-, intermediate-, and high-risk groups; the rate for<br />

the low-risk group was significantly lower than the rate for the high-risk group<br />

(p < 0.001). In a multivariate Cox model, RS provides significant (p < 0.001)<br />

predictive power that goes beyond age and tumor size. The RS was also predictive<br />

<strong>of</strong> overall survival (p < 0.001) and could be used as a continuous function<br />

to predict distant recurrence for individual patients. Finally, the RS was prospectively<br />

validated to predict the likelihood <strong>of</strong> distant recurrence in tamoxifentreated<br />

breast cancer patients with node-negative, ER-positive tumors (18,19).


246 Abramovitz and Leyland-Jones<br />

This methodology has been sufficiently successful, so that the Food and<br />

Drug Administration (FDA) approved it in 2004 to become commercially available<br />

[Genomic Health, Inc. (GHI), Redwood City, California, U.S.]. Oncotype DX<br />

is a diagnostic assay that quantifies the likelihood <strong>of</strong> breast cancer recurrence in<br />

women with newly diagnosed stage I or II, node-negative, ER-positive breast<br />

cancer who will be treated with tamoxifen. More importantly, this test will pave<br />

the way for individualization <strong>of</strong> cancer therapy. While the St. Gallen classification<br />

distinguished 7.9% <strong>of</strong> patients as low risk, the GHI RS identified 50.6% <strong>of</strong><br />

patients as low risk. Similarly, St. Gallen overestimated the high-risk patients at<br />

58.8%, while the GHI test classified high-risk patients at only 27.1%. It is obvious<br />

that St. Gallen underestimated the low-risk patients and overestimated the highrisk<br />

ones. The advantage <strong>of</strong> the newly developed multigene RS is, therefore, that<br />

low-risk patients can be saved from potentially toxic, high-dose chemotherapy.<br />

RNA-BASED EXPRESSION PROFILING OF<br />

FFPE BREAST TUMOR SPECIMENS<br />

Expression Pr<strong>of</strong>iling: The DASL Assay<br />

Gene expression pr<strong>of</strong>iling has become an important aspect in prediction, prognosis,<br />

and cancer modeling. Gene signatures resulting from expression pr<strong>of</strong>iling<br />

studies have the potential to define cancer subtypes, predict the clinical outcome<br />

(recurrence <strong>of</strong> disease) and response to specific therapies, and analyze oncogenic<br />

pathways (20–23). Investigations <strong>of</strong> gene pathways and interactions indicated by<br />

gene signatures that are truly predictive <strong>of</strong> the clinical endpoints are necessary to<br />

understand the biology underlying this predictive value. More importantly, when<br />

combined with clinical and demographic factors, multiple forms <strong>of</strong> molecular<br />

(protein- and gene-based) data can provide information that has the potential to<br />

identify unique characteristics <strong>of</strong> individuals and so lead to individualized<br />

treatment strategies (24,25).<br />

Although most array-based assays utilize RNA prepared from frozen<br />

specimens, the newer cDNA-mediated annealing, selection, extension, and<br />

ligation (DASL) assay (Illumina, Inc., San Diego, California, U.S.) platform was<br />

specifically designed to pr<strong>of</strong>ile small transcripts typically found in FFPE tissue<br />

(26–31). The DASL platform is based on massively multiplexed RT-PCR<br />

applied in a microarray format that allows for the determination <strong>of</strong> expression <strong>of</strong><br />

up to 512 genes (502 genes in the standard cancer panel available from Illumina)<br />

using a minimal amount <strong>of</strong> total RNA (100 to *200 ng per assay) isolated<br />

from 96 FFPE tumor tissue samples in a high-throughput format (29,30). In the<br />

procedure, biotinylated random nonamers (biotin-d(N)9) and oligo d(T)18 are<br />

used for cDNA synthesis, and therefore probes are designed so that they can<br />

target any unique region <strong>of</strong> the gene without limiting the selection <strong>of</strong> the optimal<br />

probe to the 3 0 ends <strong>of</strong> transcripts. In addition, because <strong>of</strong> the small size <strong>of</strong> the<br />

targeted gene sequence (50 nucleotides), along with the use <strong>of</strong> random primers in


The Formalin-Fixed Paraffin-Embedded (FFPE) Tissue Gold Mine 247<br />

Figure 1 The DASL platform for pr<strong>of</strong>iling both RNA and miRNA. (A) The DASL assay.<br />

FFPE tumor tissue samples, 5- to 10-mm shaved sections or from microscope slides, are<br />

first subjected to deparaffinization and then RNA extraction. The extracted RNA is<br />

fragmented with an average size <strong>of</strong> approximately 200 nucleotides. Biotinylated cDNA is<br />

made from the fragmented RNA using random primers (9-mers). Complex oligonucleotides<br />

that contain a gene-specific sequence and a unique address sequence are used to<br />

selectively bind to the cDNA. Extension and then ligation follow, resulting in cDNA<br />

products that are subsequently amplified using fluorescently labeled common primers.<br />

Depicted are three gene-specific oligonucleotide probe sets hybridized to cDNAs from a<br />

single gene. The amplified products contain 1536 unique addresses (3 addresses per gene<br />

and therefore 512 genes) and are hybridized to the 50,000-bead array (each unique<br />

complementary address bead at *30-fold redundancy). The address sequence directs<br />

hybridization to their complements on the universal bead arrays. (B) The miRNA array.<br />

cDNA derived from total RNA is recognized as a specific miRNA by MSO hybridization.<br />

Primer extension imparts further target specificity. The product <strong>of</strong> this extension is used as<br />

a template for universal PCR. Abbreviations: DASL, cDNA-mediated annealing, selection,<br />

extension, and ligation; FFPE, formalin-fixed, paraffin-embedded; PCR, polymerase<br />

chain reaction.<br />

the cDNA synthesis, RNAs that are otherwise too degraded for conventional<br />

microarray analysis can be readily detected. Sequence-specific query oligonucleotides<br />

encompassing primer extension, ligation, and universal PCR in highlyplexed<br />

reactions (1536-plex), two-color labeling, and redundant (*30-fold<br />

redundancy <strong>of</strong> each bead type) feature representations are used to probe up to<br />

three different sites on each cDNA (Fig. 1A) and, therefore, lend the assay the<br />

necessary sensitivity and reproducibility for quantitative detection <strong>of</strong> differential<br />

expression using RNA from FFPE tissue (29,30). In an initial study, a 231-gene


248 Abramovitz and Leyland-Jones<br />

cancer panel was used in the DASL assay in order to pr<strong>of</strong>ile both breast and<br />

colon cancer FFPE tumor samples. Cluster analysis was able to separate breast<br />

from colon tissue types and, subsequently, divide each tissue sample set into<br />

cancer versus normal (32). More recently, the DASL assay has been used to<br />

identify RNA signatures in prostate cancer, including a 16-gene set that correlates<br />

with prostate cancer relapse (31,33).<br />

We have designed our own custom 512-gene breast cancer–related gene<br />

panel so that it incorporates previously identified signature genes from various<br />

breast cancer microarray studies that have been used in the classification <strong>of</strong><br />

breast tumors (34,35), in prognosis (MammaPrint) (36) and as predictors <strong>of</strong><br />

outcome to treatment (17,37–40), as well as a host <strong>of</strong> additional genes that have<br />

been implicated as playing a role in a number <strong>of</strong> breast cancer–related processes,<br />

including proliferation, angiogenesis, (41,42), metastasis (43,44), DNA repair,<br />

apoptosis (45), thrombosis (46), and custom-selected genes taken from the most<br />

recently published data on breast cancer (47–49). Additional cancer-related<br />

genes include oncogenes, cell cycle genes, telomerase-related genes, amplified<br />

genes, breast cancer stem cell genes, and senescence-related genes (50–52).<br />

MiRNA Pr<strong>of</strong>iling<br />

MiRNAs (miRNAs) are small noncoding RNA species that regulate the translation<br />

and stability <strong>of</strong> mRNA messages for hundreds <strong>of</strong> downstream target genes<br />

via partial complementarity to short sequences in the 3 0 -untranslated region<br />

(3 0 UTR) <strong>of</strong> mRNAs. The aberrant expression or mutation <strong>of</strong> miRNAs, a distinct<br />

class <strong>of</strong> RNAs that control gene expression via translation repression or degradation,<br />

has been noted in cancers and suggests a key role for miRNAs in<br />

tumorogenesis. MiRNAs can act both as tumor suppressors and as oncogenes,<br />

depending on the sets <strong>of</strong> downstream targets that they regulate (53). Studies have<br />

recently implicated nearly 30 <strong>of</strong> the over 450 known human miRNAs in breast<br />

cancer, as tumor suppressor genes or as oncogenes, with significant deregulation<br />

among only a handful (54). Others are correlated with specific tumor<br />

types as identified by protein expression. For example, the mir-30 cluster is<br />

downregulated in both ER-negative, PR-positive and ER-positive, PR-negative<br />

tumors (54). Differential expression has also been noted by tumor stage or proliferative<br />

indices. The tumor suppressor, p53, was recently shown to regulate<br />

expression <strong>of</strong> mir-34b and mir-34c (55). One key set <strong>of</strong> tumor suppressor miRNAs<br />

includes mir15a-16, which targets the translation <strong>of</strong> the suppressor <strong>of</strong> apoptosis,<br />

Bcl-2 (56). These observations suggest that specific miRNAs are associated with<br />

major breast cancer subgroups, including hormone-positive and hormone-negative<br />

subgroups and the basal or triple negative subtype, and that the deregulated<br />

miRNAs may differ between African-American and Caucasian women.<br />

These low-molecular weight miRNAs can be extracted at the same time as<br />

total RNA is being isolated from FFPE tissue samples. qRT-PCR can be used to<br />

screen tumor samples for miRNA expression. The TaqMan TM array human


The Formalin-Fixed Paraffin-Embedded (FFPE) Tissue Gold Mine 249<br />

miRNA panel from Applied Biosystems allows for the expression pr<strong>of</strong>iling <strong>of</strong><br />

365 miRNAs from one RNA sample. Illumina, Inc., has recently adapted its<br />

DASL assay for high-throughput multiplexed miRNA expression pr<strong>of</strong>iling<br />

(Fig. 1B). It has made a human miRNA panel that can be used to detect 470<br />

miRNAs described in the miRBase database and an additional 265 potential<br />

miRNAs that have been identified in an RNA-primed array-based Klenow<br />

extension (RAKE) analysis study (57,58). With these arrays and tools now<br />

available, our understanding <strong>of</strong> the role that miRNAs play in breast cancer will<br />

be greatly enhanced.<br />

Whole-Genome Expression Pr<strong>of</strong>iling<br />

In order to conduct expression pr<strong>of</strong>iling using microarrays, it has been necessary<br />

to obtain freshly frozen tissue as the starting material for RNA extraction. This<br />

presents limitations with regard to obtaining enough samples, enough tissue per<br />

sample, as well as the logistics <strong>of</strong> storing this type <strong>of</strong> material. Most recently,<br />

new technologies have come into play that enable the use <strong>of</strong> RNA extracted from<br />

FFPE to be used in genomewide gene expression pr<strong>of</strong>iling. The company<br />

NuGEN Technologies, Inc. (San Carlos, California, U.S.) has developed an<br />

amplification system that can take picogram amounts <strong>of</strong> total RNA to yield<br />

micrograms <strong>of</strong> amplified cDNA that can be fragmented and labeled for same-day<br />

hybridization to a variety <strong>of</strong> array formats, including Affymetrix, Agilent, and<br />

Illumina microarrays (59). This type <strong>of</strong> technology will undoubtedly generate<br />

significant microarray data from FFPE breast tumor samples in the near future<br />

further adding to the utility <strong>of</strong> FFPE specimens.<br />

DNA-BASED ARRAY PROFILING THAT EMPLOYS FFPE SAMPLES<br />

Array CGH and SNP Pr<strong>of</strong>iling<br />

DNA copy number alterations (CNAs), which are the result <strong>of</strong> genomic instability,<br />

can give rise to gene amplifications and gene deletions that play important<br />

roles in tumor progression, and it has become apparent that an increased number<br />

<strong>of</strong> CNAs correlate with a poor prognosis for the patient. Several genes that<br />

undergo CNAs have been identified that act as either oncogenes or tumor suppressors,<br />

which are crucial factors in the disease process. With regard to breast<br />

cancer, array comparative genomic hybridization (CGH) has been used principally<br />

to identify novel breast cancer-related genes, genes that can be used to classify<br />

different types <strong>of</strong> breast cancer, genes that yield prognostic or predictive information,<br />

or genes that may represent important therapeutic targets. Single nucleotide<br />

polymorphisms (SNPs) have been linked to individual variations in breast cancer<br />

susceptibility. Genetic variations in genes involved in ER and drug metabolism and<br />

genes involved in cell cycle control as well as variations depending on ethnic<br />

background have been found to contribute to breast cancer susceptibility.


250 Abramovitz and Leyland-Jones<br />

Array CGH has been used in the classification <strong>of</strong> breast tumors and has<br />

shown that subtypes can be discerned on the basis <strong>of</strong> genomic instability patterns.<br />

In one study, array CGH, made up <strong>of</strong> 2464 genomic clones, was applied to<br />

the analysis <strong>of</strong> CNAs in 62 sporadic invasive ductal carcinomas, and three breast<br />

tumor subtypes were delineated based on genomic DNA CNAs (60). In another<br />

study, 89 breast cancer tumors <strong>of</strong> the luminal A, luminal B, HER2, and basal-like<br />

subtypes were also found to yield distinct CNA pr<strong>of</strong>iles, suggesting that genomic<br />

instability mechanisms differ between subtypes (61). Most array CGH studies<br />

have made use <strong>of</strong> DNA prepared from fresh or frozen tumor samples; however,<br />

the ability to tap into archival FFPE clinical samples would greatly accelerate the<br />

rate at which disease-related genomic change, including translocation breakpoints,<br />

could be more easily identified. Recent studies have addressed this<br />

important issue and shown that DNA can be prepared from FFPE tissue samples<br />

and used successfully in array CGH (62–64). Van Beers and colleagues (2006)<br />

have also developed a multiplex PCR predictor that can be used to assess the<br />

quality <strong>of</strong> the DNA extracted from FFPE samples (65) prior to use in array CGH.<br />

Recently, Oosting and co-workers (2007) (66) have compared the Affymetrix<br />

GeneChip and Bac array CGH with the Illumina BeadArray platform, which is a<br />

high-density SNP microarray that allows for the measurement <strong>of</strong> both genomic<br />

copy number and loss <strong>of</strong> heterozygosity (LOH). The BeadArray platform (Fig. 2)<br />

is amenable for use with fragmented DNA from FFPE tissue because <strong>of</strong> the fact<br />

that only short, intact genomic segments <strong>of</strong> approximately 40 bp flank each SNP<br />

on the array. The authors found that when paired comparisons <strong>of</strong> DNA prepared<br />

from freshly frozen and FFPE tissue were conducted, the patterns <strong>of</strong> genomic<br />

aberrations were basically identical.<br />

The use <strong>of</strong> array CGH and SNP pr<strong>of</strong>iling in conjunction with the vast<br />

archive <strong>of</strong> FFPE clinical breast cancer samples will allow for the identification <strong>of</strong><br />

chromosomal aberrations as well as novel variants that contribute to breast<br />

cancer, and this will lead to uncovering novel breast cancer-related genes that<br />

will be key to understanding the molecular mechanisms underlying the disease.<br />

Methylation Pr<strong>of</strong>iling<br />

Epigenetic changes, which involve very subtle molecular modifications, can<br />

cause activation or suppression <strong>of</strong> genes that can lead to abnormalities resulting<br />

in tumorigenesis (67), and these epigenetic changes have been found to occur in<br />

many cancers, including breast, colon, prostate, and blood cancer. The epigenotype<br />

is affected by such factors as the environment, age, epigenotype <strong>of</strong> the<br />

parents, and the genotype at the loci that regulate DNA methylation and chromatin,<br />

and these factors can cause aberrant methylation <strong>of</strong> the cytosine residue <strong>of</strong><br />

DNA, resulting in gene regulation.<br />

Illumina has developed an array-based platform (Fig. 3) that can be used in<br />

conjunction with its standard <strong>Cancer</strong> Methylation Panel, which spans 1505 CpG<br />

loci selected from 807 genes with 71% <strong>of</strong> those selected genes containing at least


The Formalin-Fixed Paraffin-Embedded (FFPE) Tissue Gold Mine 251<br />

Figure 2 Whole-genome expression pr<strong>of</strong>iling on the BeadChip array. (A) Gene-specific<br />

probes are attached to beads, which are then assembled into the arrays. For simplicity,<br />

this diagram shows only one oligomer attached to the bead; actual beads have hundreds<br />

<strong>of</strong> thousands <strong>of</strong> copies <strong>of</strong> the same sequence attached. (B) Sentrix Human-6 and<br />

HumanRef-8 Expression BeadChips. More than 48,000 different bead types, each with<br />

a unique sequence, are represented on the Human-6 Expression BeadChip, greater<br />

than 24,000 on the HumanRef-8 Expression BeadChip. Each bead contains hundreds<br />

<strong>of</strong> thousands <strong>of</strong> copies <strong>of</strong> covalently attached oligonucleotide probes. Beads are<br />

assembled into greater than 1.6 million pits, each measuring 3 mm in diameter,<br />

generating an average 30-fold redundancy for each sequence represented on the array.<br />

2 CpG sites. The <strong>Cancer</strong> Methylation Panel includes genes selected from various<br />

functional classes, including oncogenes and tumor suppressor genes; genes<br />

involved in cell cycle control, DNA repair, differentiation, and apoptosis; and<br />

X-linked and imprinted genes (68). Although DNA prepared from FFPE tumor<br />

samples has not yet been specifically validated for use in the assay, it would be<br />

useful to be able to scan large numbers <strong>of</strong> FFPE samples for methylation patterns,<br />

which could provide important clues regarding tumor type.<br />

THE WAY FORWARD<br />

The goal is to obtain as complete a molecular picture <strong>of</strong> a breast tumor as possible,<br />

which can now be realistically undertaken using FFPE tissue specimens, truly<br />

untapped gold now ready to be brought to the surface and subjected to a thorough<br />

nucleic acid characterization. Pr<strong>of</strong>iling platforms can now be brought to<br />

bear on these invaluable samples in order to build a comprehensive picture <strong>of</strong> the<br />

tumor. In the scheme shown in Figure 4, FFPE samples, either sections or cores,<br />

are used to extract DNA, RNA, and miRNA. These nucleic acids can now be


252 Abramovitz and Leyland-Jones<br />

Figure 3 The DNA methylation assay. (A) In order to perform a DNA methylation assay,<br />

genomic DNA must first undergo bisulfite conversion. Cytosines are converted to uracils,<br />

but methyl-cytosine is unreactive. (B) In order to take advantage <strong>of</strong> this differential<br />

reactivity to bisulfite treatment, two pairs <strong>of</strong> complex oligonucleotide probe sets are<br />

designed for each CpG site. The first probe pair is made up <strong>of</strong> an ASO and a LSO to<br />

interrogate the methylated state <strong>of</strong> the CpG site and a corresponding ASO-LSO pair for the<br />

unmethylated state. The ASO is used to determine whether a site is methylated or not and<br />

also incorporates a universal PCR primer sequence, P1 or P2. P1 and P2 are fluorescently<br />

labeled, each with a different dye, and associated with the “T” (unmethylated) allele or<br />

the “C” (methylated) allele, respectively. The LSO is made up <strong>of</strong> three parts: a CpG<br />

(Continued)


The Formalin-Fixed Paraffin-Embedded (FFPE) Tissue Gold Mine 253<br />

Figure 4 Analysis <strong>of</strong> FFPE specimens on multiple array platforms. FFPE tissue samples,<br />

as sections on either slides, sections, or cores are available for extraction <strong>of</strong> nucleic acids.<br />

DNA, RNA, and miRNA can be isolated and applied to various array platforms in order to<br />

maximize the molecular information that can be obtained from a single tumor sample.<br />

DNA can be subjected to array CGH, SNP analysis, LOH pr<strong>of</strong>iling, and methlylation<br />

analysis. RNA can be used in conjunction with expression pr<strong>of</strong>iling platforms that include<br />

the DASL assay and whole-genome expression pr<strong>of</strong>iling on different microarray platforms.<br />

MiRNA can be used to conduct miRNA pr<strong>of</strong>iling. This ability to utilize FFPE<br />

tumor tissues for detailed molecular characterization will undoubtedly have a pr<strong>of</strong>ound<br />

impact on our understanding <strong>of</strong> tumor biology. Abbreviations: FFPE, formalin-fixed,<br />

paraffin-embedded; CGH, comparative genomic hybridization; SNP, single nucleotide<br />

polymorphisms; LOH, loss <strong>of</strong> heterozygosity; DASL, cDNA-mediated annealing, selection,<br />

extension, and ligation.<br />

<<br />

locus-specific sequence, an address sequence in the middle corresponding to a complementary<br />

address sequence on the bead array, and a universal PCR priming site (P3) at the<br />

3 0 -end. The pooled assay oligonucleotides are first annealed to bisulfite-converted<br />

genomic DNA. An allele-specific primer extension step is then carried out using ASOs<br />

that are extended only if their 3 0 -base is complementary to their cognate CpG site in the<br />

genomic DNA template. Allele-specific extension is followed by ligation <strong>of</strong> the extended<br />

ASOs to their corresponding LSOs to create PCR templates. Common primers P1, P2, and<br />

P3 are then used to amplify the ligated products, which are subsequently hybridized to a<br />

microarray bearing the complementary address sequences. Abbreviations: ASO, allelespecific<br />

oligonucleotide; LSO, locus-specific oligonucleotide; PCR, polymerase chain<br />

reaction.


254 Abramovitz and Leyland-Jones<br />

subjected to a panoply <strong>of</strong> analysis platforms that can be used to detect CNAs,<br />

analyze SNPs, and investigate methylation aberrations, regulated gene expression,<br />

and signaling pathway changes, as well as miRNA alterations. Data derived from<br />

tumor tissue analyses can be used to further our understanding <strong>of</strong> the different<br />

subtypes, which could lead to new insights into breast cancer etiology, prediction<br />

<strong>of</strong> disease outcome, response to therapy, and ultimately to better and personalized<br />

patient management.<br />

REFERENCES<br />

1. Masuda N, Ohnishi T, Kawamoto S, et al. Analysis <strong>of</strong> chemical modification <strong>of</strong> RNA<br />

from formalin-fixed samples and optimization <strong>of</strong> molecular biology applications for<br />

such samples. Nucleic Acids Res 1999; 27:4436–4443.<br />

2. Abrahamsen HN, Steiniche T, Nexo E, et al. Towards quantitative mRNA analysis in<br />

paraffin-embedded tissues using real-time reverse transcriptase-polymerase chain<br />

reaction: a methodological study on lymph nodes from melanoma patients. J Mol<br />

Diagn 2003; 5:34–41.<br />

3. Finke J, Fritzen R, Ternes P, et al. An improved strategy and a useful housekeeping<br />

gene for RNA analysis from formalin-fixed, paraffin-embedded tissues by PCR.<br />

Biotechniques 1993; 14:448–453.<br />

4. Godfrey TE, Kim SH, Chavira M, et al. Quantitative mRNA expression analysis<br />

from formalin-fixed, paraffin-embedded tissues using 5 0 nuclease quantitative<br />

reverse transcription-polymerase chain reaction. J Mol Diagn 2000; 2:84–91.<br />

5. Reichmuth C, Markus MA, Hillemanns M, et al. The diagnostic potential <strong>of</strong> the<br />

chromosome translocation t(2;13) in rhabdomyosarcoma: a Pcr study <strong>of</strong> fresh-frozen<br />

and paraffin-embedded tumour samples. J Pathol 1996; 180:50–57.<br />

6. Rupp GM, Locker J. Purification and analysis <strong>of</strong> RNA from paraffin-embedded<br />

tissues. Biotechniques 1988; 6:56–60.<br />

7. Sheils OM, Sweeney EC. TSH receptor status <strong>of</strong> thyroid neoplasms–TaqMan RT-<br />

PCR analysis <strong>of</strong> archival material. J Pathol 1999; 188:87–92.<br />

8. Specht K, Richter T, Muller U, et al. Quantitative gene expression analysis in<br />

microdissected archival formalin-fixed and paraffin-embedded tumor tissue. Am<br />

J Pathol 2001; 158:419–429.<br />

9. Stanta G, Bonin S. RNA quantitative analysis from fixed and paraffin-embedded<br />

tissues: membrane hybridization and capillary electrophoresis. Biotechniques 1998;<br />

24:271–276.<br />

10. Abramovitz M, Leyland-Jones B. Application <strong>of</strong> array-based genomic and epigenomic<br />

technologies to unraveling the heterogeneous nature <strong>of</strong> breast tumors: on<br />

the road to individualized treatment. <strong>Cancer</strong> Genomics Proteomics 2007; 4:135–145.<br />

11. Chen J, Byrne GEJr., Lossos IS. Optimization <strong>of</strong> RNA extraction from formalinfixed,<br />

paraffin-embedded lymphoid tissues. Diagn Mol Pathol 2007; 16:61–72.<br />

12. Chung JY, Braunschweig T, Hewitt SM. Optimization <strong>of</strong> recovery <strong>of</strong> RNA from<br />

formalin-fixed, paraffin-embedded tissue. Diagn Mol Pathol 2006; 15:229–236.<br />

13. Penland SK, Keku TO, Torrice C, et al. RNA expression analysis <strong>of</strong> formalin-fixed<br />

paraffin-embedded tumors. Lab Invest 2007; 87:383–391.


The Formalin-Fixed Paraffin-Embedded (FFPE) Tissue Gold Mine 255<br />

14. Cronin M, <strong>Ph</strong>o M, Dutta D, et al. Measurement <strong>of</strong> gene expression in archival<br />

paraffin-embedded tissues: development and performance <strong>of</strong> a 92-gene reverse<br />

transcriptase-polymerase chain reaction assay. Am J Pathol 2004; 164:35–42.<br />

15. Chevillard S, Muller A, Levalois C, et al. Reverse transcription-polymerase chain<br />

reaction (RT-PCR) assays <strong>of</strong> estrogen and progesterone receptors in breast cancer.<br />

<strong>Breast</strong> <strong>Cancer</strong> Res Treat 1996; 41:81–89.<br />

16. Knowlden JM, Gee JM, Bryant S, et al. Use <strong>of</strong> reverse transcription-polymerase<br />

chain reaction methodology to detect estrogen-regulated gene expression in small<br />

breast cancer specimens. Clin <strong>Cancer</strong> Res 1997; 3:2165–2172.<br />

17. Paik S, Shak S, Tang G, et al. A multigene assay to predict recurrence <strong>of</strong> tamoxifentreated,<br />

node-negative breast cancer. N Engl J Med 2004; 351:2817–2826.<br />

18. Paik S, Tang G, Shak S, et al. Gene expression and benefit <strong>of</strong> chemotherapy in<br />

women with node-negative, estrogen receptor-positive breast cancer. J Clin Oncol<br />

2006; 24:3726–3734.<br />

19. Paik S. Development and clinical utility <strong>of</strong> a 21-gene recurrence score prognostic<br />

assay in patients with early breast cancer treated with tamoxifen. Oncologist 2007;<br />

12:631–635.<br />

20. Bild AH, Yao G, Chang JT, et al. Oncogenic pathway signatures in human cancers as<br />

a guide to targeted therapies. Nature 2006; 439:353–357.<br />

21. Huang E, Ishida S, Pittman J, et al. Gene expression phenotypic models that predict<br />

the activity <strong>of</strong> oncogenic pathways. Nat Genet 2003; 34:226–230.<br />

22. Huang ES, Black EP, Dressman H, et al. Gene expression phenotypes <strong>of</strong> oncogenic<br />

signaling pathways. Cell Cycle 2003; 2:415–417.<br />

23. Ramaswamy S, Tamayo P, Rifkin R, et al. Multiclass cancer diagnosis using tumor<br />

gene expression signatures. Proc Natl Acad Sci U S A 2001; 98:15149–15154.<br />

24. Nevins JR, Huang ES, Dressman H, et al. Towards integrated clinico-genomic models<br />

for personalized medicine: combining gene expression signatures and clinical factors<br />

in breast cancer outcomes prediction. Hum Mol Genet 2003; 12(2):R153–R157.<br />

25. van de Vijver MJ, He YD, van ’t Veer LJ, et al. A gene-expression signature as a<br />

predictor <strong>of</strong> survival in breast cancer. N Engl J Med 2002; 347:1999–2009.<br />

26. Abramovitz M, Ordanic-Kodani M, Wang Y, et al. Optimization <strong>of</strong> RNA extraction<br />

from FFPE tissues for expression pr<strong>of</strong>iling in the DASL assay. Biotechniques 2007<br />

(in press).<br />

27. Abramovitz M, Leyland-Jones B. A systems approach to clinical oncology: focus on<br />

breast cancer. Proteome Sci 2006; 4:5.<br />

28. Bibikova M, Yeakley JM, Chudin E, et al. Gene expression pr<strong>of</strong>iles in formalinfixed,<br />

paraffin-embedded tissues obtained with a novel assay for microarray analysis.<br />

Clin Chem 2004; 50:2384–2386.<br />

29. Bibikova M, Talantov D, Chudin E, et al. Quantitative gene expression pr<strong>of</strong>iling in<br />

formalin-fixed, paraffin-embedded tissues using universal bead arrays. Am J Pathol<br />

2004; 165:1799–1807.<br />

30. Fan JB, Yeakley JM, Bibikova M, et al. A versatile assay for high-throughput gene<br />

expression pr<strong>of</strong>iling on universal array matrices. Genome Res 2004; 14:878–885.<br />

31. Li HR, Wang-Rodriguez J, Nair TM, et al. Two-dimensional transcriptome pr<strong>of</strong>iling:<br />

identification <strong>of</strong> messenger RNA is<strong>of</strong>orm signatures in prostate cancer from archived<br />

paraffin-embedded cancer specimens. <strong>Cancer</strong> Res 2006; 66:4079–4088.


256 Abramovitz and Leyland-Jones<br />

32. Bibikova M, Talantov D, Chudin E, et al. Quantitative gene expression pr<strong>of</strong>iling in<br />

formalin-fixed, paraffin-embedded tissues using universal bead arrays. Am J Pathol<br />

2004; 165:1799–1807.<br />

33. Bibikova M, Chudin E, Arsanjani A, et al. Expression signatures that correlated with<br />

Gleason score and relapse in prostate cancer. Genomics 2007; 89:666–672.<br />

34. Sorlie T, Perou CM, Tibshirani R, et al. Gene expression patterns <strong>of</strong> breast carcinomas<br />

distinguish tumor subclasses with clinical implications. Proc Natl Acad Sci<br />

U S A 2001; 98:10869–10874.<br />

35. Sotiriou C, Neo SY, McShane LM, et al. <strong>Breast</strong> cancer classification and prognosis<br />

based on gene expression pr<strong>of</strong>iles from a population-based study. Proc Natl Acad Sci<br />

U S A 2003; 100:10393–10398.<br />

36. van ’t Veer LJ, Dai H, van de Vijver MJ, et al. Gene expression pr<strong>of</strong>iling predicts<br />

clinical outcome <strong>of</strong> breast cancer. Nature 2002; 415:530–536.<br />

37. Ayers M, Symmans WF, Stec J, et al. Gene expression pr<strong>of</strong>iles predict complete<br />

pathologic response to neoadjuvant paclitaxel and fluorouracil, doxorubicin, and<br />

cyclophosphamide chemotherapy in breast cancer. J Clin Oncol 2004; 22:2284–2293.<br />

38. Chang JC, Wooten EC, Tsimelzon A, et al. Patterns <strong>of</strong> resistance and incomplete<br />

response to docetaxel by gene expression pr<strong>of</strong>iling in breast cancer patients. J Clin<br />

Oncol 2005; 23:1169–1177.<br />

39. Jansen MP, Foekens JA, van Staveren IL, et al. Molecular classification <strong>of</strong> tamoxifenresistant<br />

breast carcinomas by gene expression pr<strong>of</strong>iling. J Clin Oncol 2005; 23:<br />

732–740.<br />

40. Pawitan Y, Bjohle J, Amler L, et al. Gene expression pr<strong>of</strong>iling spares early breast<br />

cancer patients from adjuvant therapy: derived and validated in two population-based<br />

cohorts. <strong>Breast</strong> <strong>Cancer</strong> Res 2005; 7:R953–R964.<br />

41. Heuze-Vourc’h N, Liu M, Dalwadi H, et al. IL-20, an anti-angiogenic cytokine that<br />

inhibits COX-2 expression. Biochem Biophys Res Commun 2005; 333:470–475.<br />

42. Zhong H, Chiles K, Feldser D, et al. Modulation <strong>of</strong> hypoxia-inducible factor 1alpha<br />

expression by the epidermal growth factor/phosphatidylinositol 3-kinase/PTEN/<br />

AKT/FRAP pathway in human prostate cancer cells: implications for tumor angiogenesis<br />

and therapeutics. <strong>Cancer</strong> Res 2000; 60:1541–1545.<br />

43. Kang Y, Siegel PM, Shu W, et al. A multigenic program mediating breast cancer<br />

metastasis to bone. <strong>Cancer</strong> Cell 2003; 3:537–549.<br />

44. Minn AJ, Gupta GP, Siegel PM, et al. Genes that mediate breast cancer metastasis to<br />

lung. Nature 2005; 436:518–524.<br />

45. Miller LD, Smeds J, George J, et al. An expression signature for p53 status in human<br />

breast cancer predicts mutation status, transcriptional effects, and patient survival.<br />

Proc Natl Acad Sci U S A 2005; 102:13550–13555.<br />

46. Kwaan HC, Parmar S, Wang J. Pathogenesis <strong>of</strong> increased risk <strong>of</strong> thrombosis in<br />

cancer. Semin Thromb Hemost 2003; 29:283–290.<br />

47. Gonzalez MA, Pinder SE, Callagy G, et al. Minichromosome maintenance protein 2<br />

is a strong independent prognostic marker in breast cancer. J Clin Oncol 2003;<br />

21:4306–4313.<br />

48. McLean GW, Carragher NO, Avizienyte E, et al. The role <strong>of</strong> focal-adhesion kinase<br />

in cancer: a new therapeutic opportunity. Nat Rev <strong>Cancer</strong> 2005; 5:505–515.<br />

49. Zhao C, Yasui K, Lee CJ, et al. Elevated expression levels <strong>of</strong> NCOA3, TOP1, and<br />

TFAP2C in breast tumors as predictors <strong>of</strong> poor prognosis. <strong>Cancer</strong> 2003; 98:18–23.


The Formalin-Fixed Paraffin-Embedded (FFPE) Tissue Gold Mine 257<br />

50. Brabletz T, Jung A, Spaderna S, et al. Opinion: migrating cancer stem cells - an<br />

integrated concept <strong>of</strong> malignant tumour progression. Nat Rev <strong>Cancer</strong> 2005; 5:744–749.<br />

51. Collado M, Gil J, Efeyan A, et al. Tumour biology: senescence in premalignant<br />

tumours. Nature 2005; 436:642.<br />

52. Dikmen ZG, Gellert GC, Jackson S, et al. In vivo inhibition <strong>of</strong> lung cancer by<br />

GRN163L: a novel human telomerase inhibitor. <strong>Cancer</strong> Res 2005; 65:7866–7873.<br />

53. Hammond SM. MicroRNAs as oncogenes. Curr Opin Genet Dev 2006; 16:4–9.<br />

54. Iorio MV, Ferracin M, Liu CG, et al. MicroRNA gene expression deregulation in<br />

human breast cancer. <strong>Cancer</strong> Res 2005; 65:7065–7070.<br />

55. Corney DC, Flesken-Nikitin A, Godwin AK, et al. MicroRNA-34b and microRNA-<br />

34c are targets <strong>of</strong> p53 and cooperate in control <strong>of</strong> cell proliferation and adhesionindependent<br />

growth. <strong>Cancer</strong> Res 2007; 67:8433–8438.<br />

56. Cimmino A, Calin GA, Fabbri M, et al. miR-15 and miR-16 induce apoptosis by<br />

targeting BCL2. Proc Natl Acad Sci U S A 2005; 102:13944–13949.<br />

57. Berezikov E, van Tetering G, Verheul M, et al. Many novel mammalian microRNA<br />

candidates identified by extensive cloning and RAKE analysis. Genome Res 2006;<br />

16:1289–1298.<br />

58. Griffiths-Jones S, Grocock RJ, van Dongen S, et al. miRBase: microRNA sequences,<br />

targets and gene nomenclature. Nucleic Acids Res 2006; 34:D140–D144.<br />

59. Turner, L, de Feo, G, Heath, JD, et al. WT-Ovation TM FFPE system: RNA amplification<br />

for gene expression pr<strong>of</strong>iling using RNA from formalin fixed, paraffin<br />

embedded (FFPE) tissue. Los Angeles, CA: American Association <strong>of</strong> <strong>Cancer</strong><br />

Research (AACR) 20073 (abstr).<br />

60. Fridlyand J, Snijders AM, Ylstra B, et al. <strong>Breast</strong> tumor copy number aberration<br />

phenotypes and genomic instability. BMC <strong>Cancer</strong> 2006; 6:96.<br />

61. Bergamaschi A, Kim YH, Wang P, et al. Distinct patterns <strong>of</strong> DNA copy number alteration<br />

are associated with different clinicopathological features and gene-expression subtypes<br />

<strong>of</strong> breast cancer. Genes Chromosomes <strong>Cancer</strong> 2006; 45:1033–1040.<br />

62. Johnson NA, Hamoudi RA, Ichimura K, et al. Application <strong>of</strong> array CGH on archival<br />

formalin-fixed paraffin-embedded tissues including small numbers <strong>of</strong> microdissected<br />

cells. Lab Invest 2006; 86:968–978.<br />

63. Joosse SA, van Beers EH, Nederl<strong>of</strong> PM. Automated array-CGH optimized for archival<br />

formalin-fixed, paraffin-embedded tumor material. BMC <strong>Cancer</strong> 2007; 7:43.<br />

64. Little SE, Vuononvirta R, Reis-Filho JS, et al. Array CGH using whole genome<br />

amplification <strong>of</strong> fresh-frozen and formalin-fixed, paraffin-embedded tumor DNA.<br />

Genomics 2006; 87:298–306.<br />

65. van Beers EH, Joosse SA, Ligtenberg MJ, et al. A multiplex PCR predictor for aCGH<br />

success <strong>of</strong> FFPE samples. Br J <strong>Cancer</strong> 2006; 94:333–337.<br />

66. Oosting J, Lips EH, van Eijk R, et al. High-resolution copy number analysis <strong>of</strong> paraffinembedded<br />

archival tissue using SNP BeadArrays. Genome Res 2007; 17:368–376.<br />

67. Esteller M. <strong>Cancer</strong> epigenomics: DNA methylomes and histone-modification maps.<br />

Nat Rev Genet 2007; 8:286–298.<br />

68. Bibikova M, Lin Z, Zhou L, et al. High-throughput DNA methylation pr<strong>of</strong>iling using<br />

universal bead arrays. Genome Res 2006; 16:383–393.


18<br />

<strong>Breast</strong> <strong>Cancer</strong> Stem Cells and Their<br />

Niche: Lethal Seeds in Lethal Soil<br />

Danuta Balicki<br />

Department <strong>of</strong> Medicine and Research Centre, Centre Hospitalier de<br />

l’Universite´ de Montre´al; Department <strong>of</strong> Medicine, Universite´ de Montre´al;<br />

and <strong>Breast</strong> <strong>Cancer</strong> Prevention and Genetics and <strong>Breast</strong> <strong>Cancer</strong> Risk<br />

Assessment Clinic, Ville Marie Multidisciplinary <strong>Breast</strong> <strong>Cancer</strong> Centre,<br />

Montre´al, Que´bec, Canada<br />

Brian Leyland-Jones<br />

Winship <strong>Cancer</strong> Institute, Emory University School <strong>of</strong> Medicine,<br />

Atlanta, Georgia, U.S.A.<br />

Max S. Wicha<br />

Comprehensive <strong>Cancer</strong> Centre, Department <strong>of</strong> Internal Medicine, Division <strong>of</strong><br />

Hematology-Oncology, University <strong>of</strong> Michigan, Ann Arbor, Michigan, U.S.A.<br />

INTRODUCTION<br />

The quest for optimal breast cancer therapy has been thwarted by recurrent<br />

treatment failures. As a result, the necessity has emerged to concomitantly rethink<br />

the fundamental basis <strong>of</strong> breast cancer, and to ensure that it is adequately and<br />

optimally targeted by therapeutic agents, and combinations <strong>of</strong> these agents.<br />

Adding to the complexity <strong>of</strong> this disease is the heterogeneity <strong>of</strong> breast cancer, and<br />

the realization that breast cancer is not a single disease, but a family <strong>of</strong> diseases.<br />

The daunting challenge that we face with a newly diagnosed breast cancer patient<br />

is one-to-one communication that tailors an individualized treatment strategy<br />

259


260 Balicki et al.<br />

based on the corresponding breast cancer pathophysiology. At the same time, we<br />

are mandated to address the urgency <strong>of</strong> this therapy on the basis <strong>of</strong> the underlying<br />

clinical assessment, the risks <strong>of</strong> metastases and death, as well as the benefits and<br />

adverse effects associated with therapeutic regimens to ensure that an enlightened<br />

decision among available therapeutic options can be mutually reached with an<br />

individual patient. As our understanding <strong>of</strong> the pathophysiology <strong>of</strong> breast cancer is<br />

incomplete, it is critical to channel the information that we glean from observations<br />

that deviate from our current models into a unified paradigm shift where therapeutic<br />

targets are validated, and subsequently attacked and destroyed. Given the<br />

complexity <strong>of</strong> breast cancer, the highest level <strong>of</strong> therapeutic success is most<br />

likely to be achieved when the uniqueness <strong>of</strong> an individual patient’s breast<br />

cancer is matched with individualized, combined, and targeted therapy.<br />

CANCER STEM CELL HYPOTHESIS<br />

The cancer stem cell hypothesis is a major player in the current paradigm shift in<br />

oncology (1,2). <strong>Cancer</strong> is, in essence, a genetic disease where three groups <strong>of</strong> genes<br />

are responsible for tumor progression, namely oncogenes, tumor suppressor genes,<br />

and stability genes (3,4). <strong>Cancer</strong> is a multistage process involving the accumulation<br />

<strong>of</strong> genetic and epigenetic changes in genes involved in the regulation <strong>of</strong> cell<br />

growth, DNA repair, and metastasis. In the traditional stochastic model, the vast<br />

majority <strong>of</strong> cancer cells can proliferate extensively to form new tumors. In contrast,<br />

the cancer stem cell model suggests that only rare cancer stem cells have the<br />

intrinsic ability to proliferate extensively to form new tumors (Fig. 1) (1). Clinical<br />

observations indicate that breast cancer responses to therapy do not necessarily<br />

correlate with patient survival. Tumor-initiating cells with the concomitant stem<br />

cell properties <strong>of</strong> self-renewal and successive maturation <strong>of</strong> progenitor cells into<br />

highly differentiated cells are referred to as “cancer stem cells” (Fig. 2).<br />

The cancer stem cell hypothesis provides an explanation for our clinical<br />

observations, including those which deviate from the stochastic model, by suggesting<br />

that therapies that kill cancer cells but not cancer stem cells may lead to tumor<br />

regression, but that the disease will subsequently recur as the cancer stem cells act as<br />

lethal seeds to repopulate the tumor (5). Thus, tumor regression in itself may represent<br />

an incomplete clinical end point, as it reflects cell death limited to differentiated tumor<br />

cells, while sparing cancer stem cells (Fig. 3). The “dandelion hypothesis” has been<br />

used to illustrate this pattern <strong>of</strong> clinical activity as it is analogous to cutting <strong>of</strong>f a<br />

dandelion (or other weed) at ground level, which may initially appear to produce the<br />

desired effect, but only the elimination <strong>of</strong> the root will prevent the weed from<br />

regrowing (6). Therefore, effective therapies should target cancer stem cells, with<br />

limited effects on normal cells. In the presence <strong>of</strong> therapies that kill cancer stem<br />

cells, tumors are predicted to lose their ability to generate new cancer cells,<br />

subsequently degenerate, leading to complete tumor eradication.<br />

Stem cells are multipotent with high proliferative potential and with<br />

concomitant properties <strong>of</strong> self-renewal and differentiation into multilineage


<strong>Breast</strong> <strong>Cancer</strong> Stem Cells and Their Niche: Lethal Seeds in Lethal Soil 261<br />

Figure 1 A comparison <strong>of</strong> the stochastic model where most cancer cells can proliferate<br />

extensively to give rise to new tumors with the cancer stem cell model where only a few<br />

cancer stem cells can give rise to new tumors. Abbreviation: CSC, cancer stem cell.<br />

cells. Recent findings suggest that stem cell properties are integral and fundamental<br />

to the formation and perpetuation <strong>of</strong> human cancers, including breast<br />

cancer (7). The quiescence, self-renewal, and differentiation <strong>of</strong> stem cells are<br />

regulated by both intrinsic and extrinsic mechanisms. Intrinsic mechanisms<br />

include those governing the epigenetic state <strong>of</strong> stem cells, for example, hematopoietic<br />

stem cells are controlled by chromatin remodelers such as polycomb<br />

group proteins (8,9). The extrinsic mechanisms include changes in stem cell<br />

fate that are dictated by the environment, i.e., the stem cell niche. As our tissues<br />

are organized into stem cell hierarchies ranging from stem cells with extensive<br />

proliferative and self-renewal capacity to mature cells with little or no capacity<br />

for cell division, it is likely that cancer cell populations might also be organized<br />

in stem cell hierarchies, ranging from a small number <strong>of</strong> cells that are responsible<br />

for fueling the uncontrolled growth <strong>of</strong> the tumor cells to a large array <strong>of</strong> differentiated<br />

daughter cells (10). While the vast majority <strong>of</strong> cells in a tumor are<br />

merely the nontumorigenic daughter cells <strong>of</strong> cancer stem cells, the highly proliferative<br />

capacities <strong>of</strong> stem cells may drive the continued expansion <strong>of</strong> malignant<br />

cells. As cellular proliferation decreases with differentiation, the long-lived<br />

and slowly dividing stem and progenitor cells are likely candidates for the<br />

accumulation <strong>of</strong> mutations associated with carcinogenesis (11,12).


262 Balicki et al.<br />

Figure 2 NSCs and CSCs have the capacity <strong>of</strong> self-renewal and differentiation into<br />

progenitor cells. NSCs divide asymmetrically giving rise to a stem cell and a differentiated<br />

progenitor cell. CSCs may lose asymmetric division, leading to selectively selfrenewal,<br />

thereby increasing the number <strong>of</strong> CSCs and the subsequent tumor burden.<br />

Abbreviations: NSCs, normal stem cells; CSCs, cancer stem cells.<br />

The multilineage differentiation <strong>of</strong> cancer stem cells may contribute to the<br />

heterogeneity <strong>of</strong> certain tumors, including breast cancer. The parallels between<br />

normal stem cells and cancer stem cells suggest that some cancer stem cells<br />

arise from mutated normal stem cells, and are responsible for the proliferative<br />

capacity <strong>of</strong> cancers (Fig. 2) (7). The observation that stem and cancer cells share<br />

certain signaling pathways that regulate their self-renewal suggests that normal<br />

stem cells can, in fact, give rise to cancer cells (8). Mutated progenitor cells<br />

which acquire self-renewal properties may also become cancer stem cells<br />

(Fig. 2) (7,13). Epigenetic gene silencing and associated promoter CpG island<br />

DNA hypermethylation may precede genetic changes in premalignant cells and<br />

foster the accumulation <strong>of</strong> additional genetic and epigenetic hits that contribute


<strong>Breast</strong> <strong>Cancer</strong> Stem Cells and Their Niche: Lethal Seeds in Lethal Soil 263<br />

Figure 3 Therapies which kill cancer cells, but not CSCs lead to an initial tumor<br />

regression, but as the CSCs are not eliminated, the tumor will regenerate. In contrast,<br />

therapies that target CSCs also eliminate the genesis <strong>of</strong> new tumor cells and result in<br />

tumor regression. Abbreviation: CSCs, cancer stem cells.<br />

to the generation <strong>of</strong> cancer stem cells from stem or early progenitor cells (14).<br />

The accumulation <strong>of</strong> mutations in cancer stem cells may disrupt the tight control<br />

<strong>of</strong> normal stem cells, leading to the dysregulation <strong>of</strong> self-renewal, tumorigenesis,<br />

loss <strong>of</strong> asymmetric division, and aberrant differentiation (Fig. 2) (12,15). The<br />

fate <strong>of</strong> stem cells, including self-renewal and differentiation, may also be programmed<br />

by the stem cell niche, i.e., the cellular microenvironment providing<br />

necessary support and stimuli to sustain self-renewal and other environmental<br />

factors (11,13,16–19). In normal breast development, mammary stem cells differentiate<br />

into myoepithelial, alveolar epithelial, and ductal epithelial cells (20).<br />

<strong>Cancer</strong> stem cells may acquire additional features associated with tumor progression,<br />

metastases, and therapeutic failure, including genetic instability and<br />

drug resistance (11). Mechanisms <strong>of</strong> therapeutic resistance include cellular<br />

quiescence and the acquisition <strong>of</strong> protein expression responsible for the cellular<br />

efflux <strong>of</strong> chemotherapeutic agents. The elimination <strong>of</strong> the cancer stem cell<br />

compartment <strong>of</strong> a tumor may be necessary to achieve stable, long-lasting cancer<br />

remissions. Therapeutic strategies designed to eradicate the cancer stem


264 Balicki et al.<br />

compartment should selectively, or at least preferentially, target the self-renewal,<br />

survival, and proliferative pathways specific to cancer stem cells, or induce<br />

differentiation. Thus, advances in our knowledge <strong>of</strong> stem cell properties are<br />

pivotal to the specific and effective targeting <strong>of</strong> these cells in therapeutic<br />

strategies, including cancer therapies.<br />

CURRENT STATUS OF BREAST STEM CELL RESEARCH<br />

The human mammary gland is organized during development as a hierarchy <strong>of</strong><br />

stem and progenitor cells that are successively limited in their multilineage<br />

potential and proliferative capacities. Stem cells represent approximately 1 in<br />

250 epithelial cells <strong>of</strong> the normal human breast (17). The transplantation <strong>of</strong><br />

mammary cells into cleared murine fat pads is a functional assay used to identify<br />

mammary epithelial cells with stem cell properties based on their ability to<br />

generate tumors (21,22). The existence <strong>of</strong> a stem cell origin <strong>of</strong> murine mammary<br />

development was established using transplantation <strong>of</strong> cells at limiting dilutions.<br />

These experiments confirmed the clonal nature <strong>of</strong> the regenerated alveolar,<br />

ductal, lobular, and more complex structures (22–24). The unequivocal demonstration<br />

<strong>of</strong> the existence <strong>of</strong> a murine mammary stem cell occurred with the<br />

reconstitution <strong>of</strong> a complete and functional mammary gland from a single lineage<br />

negative (Lin )/cluster <strong>of</strong> differentiation (CD)24 þ /CD29 þ cell, marked<br />

with a b-galactosidase (LacZ) transgene (25).<br />

Three types <strong>of</strong> murine mammary epithelial cell progenitors have been<br />

identified thus far (22–24). The first is a bipotent progenitor with features <strong>of</strong><br />

both luminal and myoepithelial characteristics, the second has luminal features,<br />

and the third has myoepithelial features. Analogous human studies <strong>of</strong> stem and<br />

progenitor cells have been limited by the challenges associated with the development<br />

<strong>of</strong> suitable in vivo xenotransplantation assays. Alternative strategies<br />

have been undertaken to assay human breast stem cells and to optimize conditions<br />

that support the in vitro growth and differentiation <strong>of</strong> primitive human epithelial<br />

cells. These include the proliferation <strong>of</strong> nonadherent mammospheres in suspension<br />

from human breast cancer tissues (17,26) and murine mammary fat pads<br />

(27). These mammospheres are enriched in stem and progenitor cells, as illustrated<br />

by their ability to differentiate along all three mammary epithelial lineages<br />

and to clonally generate complex functional structures in reconstituted threedimensional<br />

culture systems (17). Additional studies to characterize stem and<br />

progenitor cells include the identification <strong>of</strong> side populations using flow cytometry,<br />

i.e., low Hoechst 33342 dye-retaining populations (28–30). Side populations<br />

are enriched 30-fold in mammospheres derived from human tissues (17).<br />

The use <strong>of</strong> mammosphere assays subsequently lead to the discovery that<br />

the Notch signaling pathway plays a pivotal role in normal human mammary<br />

development through its activity in both stem and progenitor cells, concurrently<br />

affecting self-renewal and lineage-specific differentiation. As a result, it was<br />

proposed that abnormal Notch signaling contributes to mammary carcinogenesis


<strong>Breast</strong> <strong>Cancer</strong> Stem Cells and Their Niche: Lethal Seeds in Lethal Soil 265<br />

by dysregulating the self-renewal <strong>of</strong> normal mammary stem cells (26). Additional<br />

pathways thought to be involved in stem cell self-renewal include<br />

hedgehog (HH), Bmi-1, wingless (Wnt/b-catenin) and phosphatase and tensin<br />

homolog deleted on chromosome ten (PTEN) (11,20,31).<br />

Further studies in human breast stem/progenitor cells identified a subset <strong>of</strong><br />

human breast cancer cells using flow cytometry with the phenotype CD44 þ /<br />

CD24 /Lin that possess highly tumorigenic properties in a nonobese diabetic/<br />

severe combined immunodeficiency (NOD/SCID) xenograft model (21). As few<br />

as 200 cells with the epithelial-specific antigen (ESA) þ , CD44 þ /CD24 /Lin<br />

phenotype consistently generated tumors in mice, whereas even 20,000 cells<br />

with alternative phenotypes were ineffective. The isolation and serial passage <strong>of</strong><br />

the tumorigenic CD44 þ /CD24 /Lin cell population in NOD/SCID mice gave<br />

rise to additional tumorigenic CD44 þ /CD24 /Lin cells, as well as phenotypically<br />

diverse nontumorigenic cells, that composed the bulk <strong>of</strong> the tumors,<br />

thereby confirming the self-renewal and differentiating capacities <strong>of</strong> these cells.<br />

While the unequivocal demonstration <strong>of</strong> stem cell characteristics necessitates<br />

model systems capable <strong>of</strong> tumor generation from a single cell, a population <strong>of</strong><br />

CD44 þ /CD24 /Lin cells appears to exhibit properties <strong>of</strong> human breast stem/<br />

progenitor cells in NOD/SCID mice.<br />

GENETIC SIGNATURES AND POLYMORPHISMS WHICH<br />

MAY INFLUENCE BREAST CANCER DEVELOPMENT<br />

AND TREATMENT OUTCOMES<br />

The study <strong>of</strong> molecular breast cancer portraits identifies patterns <strong>of</strong> molecular<br />

breast cancer markers using gene expression microarrays, which subsequently<br />

are useful in the selection <strong>of</strong> therapy. These portraits also seek to overcome<br />

potential differences in drug sensitivity or target expression between tumorinitiating<br />

cells and the more frequent nontumorigenic cells (32). The rationale for<br />

the pursuit <strong>of</strong> new drug discovery based on molecular targets results from the<br />

significant increment <strong>of</strong> the effectiveness <strong>of</strong> these drugs as compared with<br />

conventional chemotherapy. A recent systematic analysis <strong>of</strong> the consensus <strong>of</strong><br />

human breast and colorectal cancers included the analysis <strong>of</strong> 13,023 genes in<br />

11 breast and 11 colorectal cancers, and revealed that individual tumors accumulate<br />

an average <strong>of</strong> approximately 90 mutated genes, but that only a subset <strong>of</strong><br />

these contribute to the neoplastic process (33). Using stringent criteria, 189 genes<br />

(average <strong>of</strong> 11 per tumor) were found to be mutated at significant frequencies.<br />

The vast majority <strong>of</strong> these genes were not known to be genetically altered in<br />

tumors and are predicted to affect a wide range <strong>of</strong> cellular functions, including<br />

transcription, adhesion, and invasion. The genetic landscape <strong>of</strong> breast and colorectal<br />

cancer defined in this study provides new targets for diagnostic and<br />

therapeutic intervention.<br />

Gene signatures <strong>of</strong> tumor invasiveness in tumorigenic CD44 þ CD24 /low<br />

breast cancer cell populations correlate with overall and metastasis-free survival,


266 Balicki et al.<br />

independently <strong>of</strong> established clinical and pathological variables (34).<br />

CD44 þ CD24 /low breast cancer cells were also shown to increase from a median<br />

<strong>of</strong> 4.8% to 14.8% after chemotherapy in paired breast cancer biopsies from<br />

35 patients obtained before and after 12 weeks <strong>of</strong> neoadjuvant chemotherapy.<br />

Microarray analyses <strong>of</strong> CD44 þ CD24 /low breast cancer cells using the Affymetrix<br />

U133 GeneChip demonstrated an increased expression <strong>of</strong> self-renewal pathways.<br />

An increase in mammosphere formation <strong>of</strong> residual tumors was also observed<br />

postchemotherapy. Taken together, these results suggest that therapy resistant<br />

cancer stem cells may exist in the neoadjuvant setting <strong>of</strong> breast cancer therapy, and<br />

that breast cancer patients may benefit from the therapeutic targeting designed to<br />

eradicate these cells (35).<br />

In addition to chemoresistance, progenitor cells may also be resistant to<br />

radiation. This radioresistance has been shown in an animal model to be<br />

mediated at least in part by Wnt signaling, which has been implicated in stem<br />

cell survival (36). Radioresistance was investigated by treating primary BALB/c<br />

mouse mammary epithelial cells with clinically relevant doses <strong>of</strong> radiation. An<br />

enrichment in normal progenitor cells [stem cell antigen (SCA) 1-positive and<br />

side population progenitors] was identified.<br />

Specifically, radiation selectively enriched for progenitors in mammary<br />

epithelial cells isolated from transgenic mice with activated Wnt/b-catenin<br />

signaling but not for background-matched controls. Irradiated stem cell antigen<br />

1-positive cells had a selective increase in active b-catenin and survivin<br />

expression compared with SCA 1-negative cells. Radiation also induced enrichment<br />

<strong>of</strong> side population progenitors cells <strong>of</strong> the MCF-7 human breast adenocarcinoma<br />

cell line. In comparison to differentiated cells, progenitor cells have<br />

different cell survival properties that may facilitate the development <strong>of</strong> targeted<br />

antiprogenitor cell therapies.<br />

Closely intertwined with response prediction are an individual patient’s<br />

genetic polymorphisms, which may increase breast cancer susceptibility and<br />

influence the efficacy or safety <strong>of</strong> breast cancer treatments. The accumulation <strong>of</strong><br />

genetic factors that underlie breast cancer susceptibility or therapeutic escape<br />

may trigger the transformation <strong>of</strong> a normal stem or progenitor cell into a cancer<br />

stem cell. The variation <strong>of</strong> treatment responses between individuals also lies in<br />

differences in DNA sequences that alter the expression or function <strong>of</strong> proteins<br />

that are drug targets. These proteins are encoded by genes identified in early<br />

studies that are linked to highly penetrant, single-gene traits. Germline polymorphisms<br />

which confer susceptibility to breast cancer are associated with the<br />

highly penetrant breast cancer genes BRCA1 and BRCA2, p53, PTEN, ataxia<br />

telangiectasia mutated (ATM), and candidate BRCA3 genes (37). In addition,<br />

BRCA1 has been suggested to function as a breast stem cell regulator (38). Thus,<br />

BRCA1 may contribute to a conceptual link between hereditary and sporadic<br />

breast cancer. The loss <strong>of</strong> BRCA1 function may produce a block in cell differentiation<br />

and an increase in the self-renewal <strong>of</strong> mammary stem cells, resulting<br />

in the expansion <strong>of</strong> the stem cell pool. Taken together with its known function in


<strong>Breast</strong> <strong>Cancer</strong> Stem Cells and Their Niche: Lethal Seeds in Lethal Soil 267<br />

DNA repair, BRCA1 may generate genetically unstable stem cells which are targets<br />

for additional carcinogenic events and the aggressive triple negative “basal” stem<br />

cell–like phenotype (39).<br />

Low-penetrance genes that may influence breast cancer susceptibility<br />

alone or in combination with environmental factors include genes in DNA repair,<br />

metabolic pathways, and estrogen-related pathways including HRAS1 minisatellite,<br />

CHEK2, XRCC1, XRCC2, RAD51, XRCC3, and LIG4 (37). Metabolic<br />

enzymes such as N-acetyltransferases (NAT) and gluthathione-S-transferases<br />

(GST) are involved in the metabolic inactivation <strong>of</strong> carcinogens. Mutations in<br />

these enzymes can result in increased toxicity and exposure to mutagenic substances,<br />

thereby increasing mutations, which contribute to carcinogenesis.<br />

Superoxide dismutases (SODs) are metalloenzymes that protect cells from free<br />

radicals, and their mutations can lead to increased breast cancer susceptibility<br />

(37). Cytochrome p450s (CYP) metabolically activate carcinogens into chemically<br />

reactive electrophiles. CYP1B1 is the most common form in the breast,<br />

and its alanine and serine (ALA/SER) polymorphism significantly increases<br />

breast cancer susceptibility (37). Additional targets <strong>of</strong> polymorphisms that<br />

increase breast cancer susceptibility include COMT, CYP17, CYP19, PPAR-g,<br />

TGF-b, HSP70-1, HSP70-2, and HSP70-HOM (37). Germline polymorphisms<br />

may also influence an individual’s susceptibility to the effects <strong>of</strong> breast cancer<br />

treatments (37). For example, dihydropyrimidine dehydrogenase (DPD) and<br />

thymidylate synthase (TS) are pivotal to the metabolism <strong>of</strong> 5-FU, and specific<br />

polymorphisms <strong>of</strong> these enzymes lead to 5-FU toxicity. Paclitaxel is metabolized<br />

by cytochrome P450, and specific CYP polymorphisms lead to defective metabolism.<br />

Additional breast cancer treatments with known polymorphisms, which<br />

may effect drug efficacy or safety include tamoxifen (ER, CYP2D6, SULT1A1),<br />

aromatase inhibitors (CYP19, CYP1A2, CYP2C9, CYP3A), cyclophosphamide<br />

(GST), methotrexate (MTHFR), doxorubicin (GST), and epirubicin (UGT2B7)<br />

(40). Future advances hinge on the more difficult challenge <strong>of</strong> elucidating multigene<br />

determinants <strong>of</strong> breast cancer susceptibility and drug responses through an<br />

intersection <strong>of</strong> genomics and medicine, which has the potential to yield a new set<br />

<strong>of</strong> molecular diagnostic tools that can be used to individualize and optimize drug<br />

therapy (41). Additional refinements can be accomplished through coupling in<br />

vitro drug sensitivity data with microarray data, to develop gene expression signatures<br />

that predict sensitivity to individual chemotherapeutic drugs (42). Thus, as<br />

we continue to fine-tune our stratification, we also approach our ultimate goal <strong>of</strong><br />

defining individual molecular breast cancer portraits.<br />

STEM CELL SANCTUARIES AS A MECHANISM<br />

OF THERAPEUTIC RESISTANCE<br />

Targeting <strong>of</strong> stem cells is particularly complex since this cellular population is<br />

largely quiescent. Thus, therapeutic agents directed against cycling cells are<br />

predictably ineffective in this population, which is essentially shielded from


268 Balicki et al.<br />

these drugs. <strong>Cancer</strong>s that respond to therapy initially may acquire drug resistance<br />

during the course <strong>of</strong> treatment, while other cancers appear to be intrinsically<br />

resistant. The cancer stem cell hypothesis suggests that in both <strong>of</strong> these cases, the<br />

resting cancer stem cell, which is both the cancer-initiating cell and its source<br />

<strong>of</strong> replenishment under selective pressure, has innate drug resistance by virtue <strong>of</strong><br />

its quiescent stem cell phenotype. Acquired drug resistance in more differentiated<br />

cancer cells may occur through gene amplification or rearrangement,<br />

thereby contributing to an aggressive phenotype. Through the accumulation <strong>of</strong><br />

mutations that distinguish cancer stem cells from their normal counterparts, these<br />

cells may acquire additional features associated with tumor progression,<br />

metastases, and therapeutic failure, including genetic instability, radioresistance,<br />

and chemoresistance (11,36). In the case <strong>of</strong> gliomas, cancer stem cells contribute<br />

to radioresistance through the preferential activation <strong>of</strong> the DNA damage<br />

checkpoint response and an increase in DNA repair capacity (43). Radioresistance<br />

may also result from a selective increase <strong>of</strong> cell survival pathways in<br />

progenitor cells, including b-catenin and survivin (36).<br />

Stem cell niches regulate stem cell proliferation and cell fate decisions, and<br />

they also play a protective role by shielding stem cells from environmental<br />

insults. In this manner, vascular niches might protect brain cancer stem cells<br />

from chemo- and radiotherapies, and the opportunity for these cells to reform a<br />

tumor mass in spite <strong>of</strong> an initial clinical response (44). With the acquisition <strong>of</strong><br />

properties <strong>of</strong> tumor progression and metastases, cancer stem cells may travel to<br />

sanctuary sites, such as across the blood-brain barrier, where it may be difficult<br />

to eradicate them, even with targeted therapies.<br />

The drug resistance <strong>of</strong> normal tissue stem cells is constitutively mediated<br />

by multidrug resistance (MDR) transporters and detoxifying enzymes (45). DNA<br />

repair mechanisms, resistance to apoptosis, and telomerase activity also contribute<br />

to the stability <strong>of</strong> normal tissue stem cells. Among the mechanisms<br />

responsible for stem cell drug resistance, there is mounting speculation that ABC<br />

transporters repress the maturation and differentiation <strong>of</strong> stem cells (46,47).<br />

However, since breast cancer resistance protein 1 (BCRP1) knockout mice<br />

display normal hematopoiesis instead <strong>of</strong> accelerated hematopoietic maturation,<br />

it is clear that multiple factors determine the maturation and differentiation <strong>of</strong><br />

stem cells (48). In addition, protection from hypoxia appears to be another<br />

function <strong>of</strong> ABC transporters. As a result, it has been postulated that Bcrp<br />

expression protects hematopoietic stem cells from hypoxic environments by<br />

preventing porphyrin accumulation that causes mitochondrial death (49).<br />

Thus, leading edge-targeted breast cancer therapies must overcome all <strong>of</strong><br />

the obstacles associated with cancer stem cell biology, including both the<br />

intrinsic and extrinsic properties <strong>of</strong> resistance, which define these enigmatic cells<br />

and the cloisters that keep them safe. From the cancer stem cell hypothesis, it<br />

follows that systemic administration <strong>of</strong> an efficacious MDR reversal agent would<br />

render tumor and normal tissue stem cells equally susceptible to the chemotherapeutic<br />

agents, <strong>of</strong>fering no net gain in therapeutic index (45). Thus, targeting


<strong>Breast</strong> <strong>Cancer</strong> Stem Cells and Their Niche: Lethal Seeds in Lethal Soil 269<br />

the properties that distinguish cancer stem cells from their normal stem cell<br />

counterparts is necessary.<br />

CELLULAR BREAST CANCER STEM CELL TARGETS<br />

The identification <strong>of</strong> cancer stem cell markers is necessary for the isolation<br />

and purification <strong>of</strong> cancer stem cells as they are potential targets <strong>of</strong> novel drug<br />

discovery and development. The functional targets that define stem cells are<br />

their abilities to self-renew and differentiate using serial passages in xenotransplantation<br />

models. The current breast cancer stem cell hypothesis suggests that<br />

the most primitive human breast stem cell does not express the estrogen receptor<br />

(ER) while more mature progenitors may or may not express ER depending on<br />

their location in the differentiation pathway (12). A recent study using reverse<br />

transcriptase polymerase chain reaction (RT-PCR) and immunohistochemical<br />

analyses established that mouse mammary stem cells were negative for both ERa<br />

and the progesterone receptor (50). An independent study further supported this<br />

result by showing that ER-positive cells retained very little stem cell activity (51).<br />

Throughout much <strong>of</strong> mammary development during embryogenesis, the developing<br />

gland consists <strong>of</strong> a simple duct system, with a primitive mammary bud, the<br />

growing terminal end <strong>of</strong> the duct, with no ERa expression until the 30th week <strong>of</strong><br />

gestation. These structures are thought to be made <strong>of</strong> primitive ductal and<br />

myoepithelial progenitors. Later in gestation, ERa-positive cells are present in<br />

the growing terminal ends <strong>of</strong> these ducts, followed by a surge <strong>of</strong> ERa-positive<br />

cells in this region after birth (12). In the developing mammary gland, insulin<br />

growth factor-1 (IGF-1) is required to form ducts and terminal end buds, indicating<br />

its critical role in stimulating the proliferation and differentiation <strong>of</strong><br />

primitive breast epithelium (12).<br />

ER-positive progenitors arise after 30 weeks <strong>of</strong> gestation. They respond to<br />

estrogen stimulation by proliferating and generating ductal epithelial cells and<br />

alveolar cells that form the adult mammary tree. Maintenance <strong>of</strong> the myoepithelial<br />

progenitor pool and the stem cell compartment may rely on paracrine<br />

signaling initiated by the ER-positive progenitor cells. <strong>Breast</strong> carcinogenesis<br />

results from mutations affecting stem and progenitor cells. Thus, the phenotype<br />

<strong>of</strong> different subtypes <strong>of</strong> breast cancer is determined both by the cell <strong>of</strong> origin and<br />

the particular mutations driving carcinogenesis. Thus, ER-negative tumors arise<br />

from stem cells or early ER-negative progenitor cells, while ER-positive tumors<br />

arise from either ER-negative or ER-positive progenitors (12).<br />

On the basis <strong>of</strong> these observations, a breast tumor model has been proposed<br />

consisting <strong>of</strong> three types <strong>of</strong> tumors. The first is an ER-negative tumor termed<br />

type 1, which arises from the most primitive ER-negative stem or early progenitor<br />

cells. Using classic histopathology and molecular pr<strong>of</strong>iling studies, these<br />

tumors are poorly differentiated and display a basal phenotype with expression<br />

markers <strong>of</strong> both luminal epithelial and myoepithelial cells. Clinically, these<br />

tumors are highly aggressive and are associated with a poor prognosis. Type 2


270 Balicki et al.<br />

tumors also arise from early ER-negative stem or early progenitor cells. However,<br />

this group includes stem cell mutations, which allow for the differentiation<br />

<strong>of</strong> a specific subset <strong>of</strong> tumor cells into ER-positive cells, which may range from a<br />

low percentage (6–10%) similar to that seen in normal tissue to a high percentage<br />

(70–100%). Thus, these tumors are heterogeneous, with intermediate differentiation<br />

and prognosis. Type 3 tumors arise from the transformation <strong>of</strong> ER-positive<br />

progenitors. They are predicted to exclusively express luminal markers and have<br />

the best prognosis as they contain differentiated cells and respond to antiestrogen<br />

therapy (12). Additional breast stem cell markers include CD44, CD24, aldehyde<br />

dehydrogenase (ALDH), Oct-4, Bmi-1, and Wnt/b-catenin.<br />

As discussed above, a small population <strong>of</strong> cancer cells characterized by<br />

CD44 expression but low or undetectable levels <strong>of</strong> CD24 (CD44 þ CD24 /low )<br />

have a high tumorigenic capacity when injected into immunodeficient mice (52).<br />

In contrast, the rest <strong>of</strong> the cancer cells, called nontumorigenic breast cancer cells,<br />

have little or no tumorigenic ability. Tumors in mice that originate from purified<br />

tumorigenic breast cancer cells contain a mixture <strong>of</strong> both tumorigenic and<br />

nontumorigenic breast cancer cells, which confirms that the CD44 þ CD24 /low<br />

population shares the capacity for self-renewal with normal stem cells. Gene<br />

expression pr<strong>of</strong>iling <strong>of</strong> tumorigenic breast cancer cells was utilized to develop<br />

prognostic tools to assess survival in patients with breast cancer and to predict<br />

clinical outcomes in breast cancer (34). An invasiveness gene signature <strong>of</strong> 186<br />

genes was developed to predict the likelihood <strong>of</strong> a tumor to metastasize. This<br />

invasiveness gene signature is associated with the risk <strong>of</strong> death and metastasis<br />

not only in breast cancer but also in lung cancer, prostate cancer, and medulloblastoma.<br />

These results suggest the critical relevance <strong>of</strong> the CD44 þ CD24 /low<br />

tumorigenic subclass <strong>of</strong> breast cancer cells as a pivotal contributor to the association<br />

<strong>of</strong> the invasiveness signature to clinical outcome. In addition, CD44, a<br />

cell-surface receptor that mediates cell–cell contacts, promotes the engraftment<br />

<strong>of</strong> leukemic stem cells in the bone marrow and may thus have an additional role<br />

as a critical component <strong>of</strong> the cancer stem cell niche (53).<br />

As a result <strong>of</strong> the growing interest <strong>of</strong> identifying new markers for identifying<br />

and isolating stem cells, assays based on ALDH activity have been proposed<br />

to be a promising alternative in both murine and human mammary stem<br />

cells. ALDH is a family <strong>of</strong> cytosolic enzyme is<strong>of</strong>orms responsible for oxidizing<br />

intracellular aldehydes, including vitamin A, to carboxylic acids. It is highly<br />

expressed in hematopoietic progenitors, in intestinal crypt cells, as well as in<br />

breast tumor cells (54–56). ALDH1 or (ALDH isoenzyme 1 or ALDH1A1) has<br />

been shown to be responsible for the resistance observed in hematopoietic stem<br />

cells and breast cancer tumor cells to the alkylating agent, cyclophosphamide<br />

(56–58). A novel system has been designed to detect ALDH using a visible light<br />

excitable fluorochrome named Aldefluor 1 (BODIPY-conjugated aminoacetaldehyde,<br />

BAAA) metabolized by both murine and human cells (59).<br />

Fluorescence-activated cell sorter (FACS) <strong>of</strong> the normal breast cell population<br />

revealed that 6% <strong>of</strong> this population is ALDH-positive (60). In breast cancer cell


<strong>Breast</strong> <strong>Cancer</strong> Stem Cells and Their Niche: Lethal Seeds in Lethal Soil 271<br />

lines, ALDH-positive cells were more frequently observed in basal-like than in<br />

luminal cell lines. ALDH-positive cells form mammospheres at a high efficiency<br />

and are tumorigenic in NOD/SCID mice (61).<br />

A recent study showed that when acute myeloid leukemic (AML) patient<br />

cells were sorted on the basis <strong>of</strong> Aldefluor staining, three distinct possibilities<br />

were observed (62). High levels <strong>of</strong> aldehyde dehydrogenase (ALDH hi ) cells from<br />

a first group <strong>of</strong> patients were identified as normal nonleukemic stem cells, were<br />

present in low number, and were able to reconstitute a complete hematopoietic<br />

system when transplanted into NOD/SCID mice. However, a second group <strong>of</strong><br />

patients had ALDH hi cells displaying a higher side scattering pattern than normal<br />

stem cells and these cells were also significantly more abundant. These cells<br />

correspond to leukemic stem cells. Finally, a third group <strong>of</strong> patients showed no<br />

ALDH hi cells. This study demonstrates that FACS analyses based on the first<br />

ALDH isozyme (ALDH1 or ALDHA1) activity can be used to identify and<br />

isolate leukemic stem cells (58). In contrast to the toxicity associated with<br />

Hoechst 33342, ALDH is relatively mild and allows for cell sorting using viable<br />

cell populations, enriched in stem cells. The availability <strong>of</strong> antibodies derived<br />

against ALDH1 allows for the study <strong>of</strong> their expression in fixed tissue samples<br />

using immunohistochemistry (Fig. 4). Immunohistochemistry <strong>of</strong> 577 human<br />

breast carcinoma specimens from two different sources detected 24% and 30%<br />

ALDH-positive cells, respectively. In this study, ALDH expression was strongly<br />

correlated with tumor size, grade, ER status, human epidermal growth factor<br />

receptor 2 (HER2) overexpression, and poor prognosis (60).<br />

OCT-4 is an additional cancer stem cell marker. The OCT-4 oncogene<br />

encodes a nuclear protein that belongs to a family <strong>of</strong> transcription factors containing<br />

the POU DNA–binding domain that plays an important role in maintaining<br />

the pluripotent state <strong>of</strong> embryonic stem cells (63). It may prevent expression <strong>of</strong><br />

genes activated during differentiation, and its expression is downregulated during<br />

differentiation. OCT-4 is normally found in the pluripotent stem cells <strong>of</strong> pregastrulation<br />

embryos, including oocytes, early cleavage-stage embryos, and the inner<br />

cell mass <strong>of</strong> the blastocyst, and knockout <strong>of</strong> OCT-4 causes early lethality in mice<br />

because <strong>of</strong> the absence <strong>of</strong> an inner cell mass (63). Thus, these activities associated<br />

with OCT-4 suggest that it plays a pivotal role in mammalian development and in<br />

the self-renewal <strong>of</strong> embryonic stem cells. The expression <strong>of</strong> genes such as OCT-4<br />

is potentially correlated with tumorigenesis and may affect some aspects <strong>of</strong> tumor<br />

behavior, such as tumor recurrence or resistance to therapies. OCT-4 is expressed<br />

in bladder cancer, as well as in CD44 þ /CD24 breast cancer–initiating cells<br />

(64,65). It may serve as a multifunctional factor involved in major biological<br />

processes such as embryonic development, control <strong>of</strong> differentiation, and stem<br />

cell–based carcinogenesis (64).<br />

Bmi-1 is a cancer stem cell marker that belongs to the polycomb group<br />

family <strong>of</strong> proteins, and was first identified as a c-myc cooperating oncogene in<br />

murine lymphomagenesis. Bmi-1 is involved in the regulation <strong>of</strong> diverse biological<br />

processes involved in X chromosome inactivation, repression <strong>of</strong> cellular


272 Balicki et al.<br />

Figure 4 ALDH1 immunostaining. Immunohistochemistry using anti-ALDH1 on tissue<br />

microarray samples from two patients (A, B) with infiltrating ductal carcinoma <strong>of</strong> the breast,<br />

both magnified 200, with a scale bar <strong>of</strong> 100 mm. Abbreviation: ALDH1, aldehyde dehydrogenase<br />

isoenzyme 1.<br />

senescence, carcinogenesis, and stem cell renewal (66). Bmi-1 oncoprotein<br />

overexpression correlates with axillary lymph node metastases in invasive ductal<br />

breast cancer (67). In addition, the Bmi-1 oncogene induces telomerase activity<br />

and immortalizes human mammary epithelial cells (68). Bmi-1 regulates selfrenewal<br />

<strong>of</strong> normal and malignant human mammary stem cells, stimulates<br />

mammosphere formation and colony size in vitro, and increases ductal/alveolar<br />

development in humanized NOD/SCID mammary fatpads (20).<br />

The final stem cell marker in this chapter is wingless (Wnt)/b-catenin. The<br />

canonical Wnt/b-catenin signaling pathway modulates the delicate balance<br />

between stemness, differentiation, and tumorigenesis in several adult stem cell


<strong>Breast</strong> <strong>Cancer</strong> Stem Cells and Their Niche: Lethal Seeds in Lethal Soil 273<br />

niches such as the hair follicles in the skin, the mammary gland, hematopoiesis,<br />

and the intestinal crypt (8,69). Differences in the levels <strong>of</strong> Wnt signaling activity<br />

reflect tumor heterogeneity and are likely to account for distinct cellular activities<br />

such as proliferation and epithelial-mesenchymal transitions, which prompt<br />

tumor growth and malignant behavior, respectively. This pivotal pathway is<br />

highly conserved in evolution and is known to regulate cell-fate decisions, cell<br />

proliferation, morphology, migration, apoptosis, differentiation, and stem cell<br />

self-renewal. Recent evidence which suggests that Wnt signaling plays a role in<br />

human breast cancer include the detection <strong>of</strong> elevated levels <strong>of</strong> nuclear and/or<br />

cytoplasmic b-catenin using immunohistochemistry, and the overexpression or<br />

downregulation <strong>of</strong> specific Wnt proteins. The Wnt pathway has also been<br />

implicated in normal stem cell self-renewal in vivo, and there is evidence that<br />

dysregulation <strong>of</strong> this pathway in the mammary gland and other organs may play<br />

a key role in carcinogenesis (70).<br />

The Wnt family <strong>of</strong> secreted proteins includes the well-characterized<br />

canonical Wnt signaling pathway, in which Wnt ligands signal through the<br />

stabilization <strong>of</strong> b-catenin. In this pathway, Wnt proteins bind to a family <strong>of</strong><br />

frizzled receptors in a complex with the low-density lipoprotein receptor–related<br />

proteins 5 and 6 (LRP5/6) coreceptors to activate Dishevelled (Dsh). Subsequently,<br />

Dsh inhibits the activity <strong>of</strong> the b-catenin destruction complex [adenomatous<br />

polyposis coli (APC), axin, and glycogen synthase kinase-3b (GSK-<br />

3b)], which phosphorylates b-catenin in the absence <strong>of</strong> the ligands. As a result,<br />

b-catenin is stabilized and translocated to the nucleus, where it recruits transactivators<br />

to high mobility group (HMG)–box DNA-binding proteins <strong>of</strong> the<br />

lymphoid-enhancer factor/T-cell factor (LEF/TCF) family. In the absence <strong>of</strong><br />

Wnt signaling, b-catenin remains in the cytoplasm, where it forms the b-catenin<br />

destruction complex. GSK-3b phosphorylates b-catenin, which targets the protein<br />

for ubiquitin-mediated degradation. When the Wnt pathway is activated,<br />

GSK-3b is inhibited, blocking b-catenin phosphorylation and its subsequent<br />

degradation (71). In addition, several b-catenin-independent Wnt signaling<br />

pathways, known as noncanonical, have been shown to be crucial for different<br />

aspects <strong>of</strong> vertebrate embryo development (71).<br />

MOLECULAR BREAST CANCER STEM CELL TARGETS<br />

Molecular signatures <strong>of</strong> stem cells are currently being identified. The mutations that<br />

accumulate in cancer stem cells could inappropriately activate Bmi-1, Wnt/b-catenin,<br />

Notch, sonic hedgehog (SHH), and PTEN pathways that are involved in the regulation<br />

<strong>of</strong> self-renewal <strong>of</strong> normal stem cells and could potentially lead to cancer<br />

(Figs. 5 and 6) (26,72–74). These molecular markers can be exploited to purify<br />

cancer stem cells, which can then be characterized in the most intimate detail,<br />

by the modern tools <strong>of</strong> gene expression pr<strong>of</strong>iling and informatics. The failure to<br />

shut <strong>of</strong>f the self-renewal pathways could confer self-renewal on cells and microenvironments<br />

capable <strong>of</strong> evading homeostatic regulation, leading to tumorigenesis.


274 Balicki et al.<br />

Figure 5 Stem cell self-renewal pathways include the HH, Notch, Bmi-1, Wnt, and<br />

PTEN pathways. Dysregulation <strong>of</strong> the self-renewal pathways <strong>of</strong> NSCs and progenitor<br />

cells may lead to CSCs and tumor formation. NSCs are ER negative and they differentiate<br />

into ER-positive progenitor cells, which further differentiate into myoepithelial, alveolar<br />

epithelial, and ductal epithelial cells. CSCs in their niche may lose asymmetric division,<br />

leading to the clonal expansion <strong>of</strong> cancerous cells and tumor formation. Abbreviations:<br />

NSCs, normal stem cells; CSCs, cancer stem cells; ER, estrogen receptor.<br />

The intimate interrelationship <strong>of</strong> multiple pathways critical to tumorigenesis and<br />

to normal stem cell self-renewal raises the specter that cancer therapies targeting<br />

such pathways might ablate resident stem cells (75). Drugs that selectively target<br />

the self-renewal pathways <strong>of</strong> tumors will be attractive candidates for cancer<br />

therapies if they can effectively spare the normal bystander stem cells that use<br />

self-renewal as part <strong>of</strong> their regular functional repertoire. Alternatively,


<strong>Breast</strong> <strong>Cancer</strong> Stem Cells and Their Niche: Lethal Seeds in Lethal Soil 275<br />

Figure 6 A model <strong>of</strong> the interrelationship <strong>of</strong> stem cell self-renewal pathways including a<br />

conceptual link between sporadic and hereditary breast cancers. BRCA1 may inhibit stem<br />

cell self-renewal, while HER2 overexpression may increase the stem cell pool through<br />

regulation <strong>of</strong> self-renewal. In this model, there is bidirectional cross talk between the<br />

Notch and HH pathways with subsequent regulation <strong>of</strong> the polycomb gene Bmi-1 and the<br />

Wnt/b-catenin pathway. SHH, IHH, and DHH are HH ligands, while Gli1 and 2 are two<br />

HH transcription factors that are positive mediator <strong>of</strong> HH signaling. Cylcopamine inhibits<br />

the HH pathway. The DSL peptide activates Notch signaling, while the GSI blocks the<br />

intramembrane cleavage <strong>of</strong> Notch required for signaling. PTEN affects Wnt signaling<br />

through GSK-3b. Exisulind, bromoindirubin-3 0 -oxime, and imatinib inhibit the Wnt/bcatenin<br />

pathway. Abbreviations: BRCA, breast cancer; HER, human epidermal growth<br />

factor receptor 2; HH, hedgehog; SHH, sonic hedgehog; IHH, Indian hedgehog; DHH,<br />

desert hedgehog; PTEN, phosphatase and tensin homolog deleted on chromosome ten;<br />

GSI, gamma secretase inhibitor.


276 Balicki et al.<br />

differentiating therapies may also be effective if they can efficiently shift the<br />

balance <strong>of</strong> stem cell divisions toward differentiation rather than self-renewal and<br />

stem cell expansion. In addition to Bmi-1, and Wnt/b-catenin discussed above,<br />

Notch, SHH pathways, and PTEN will be reviewed below.<br />

The Notch signaling pathway plays an important role in regulating proliferation,<br />

survival, and differentiation <strong>of</strong> many cell types, including the regulation<br />

<strong>of</strong> self-renewal <strong>of</strong> adult stem cells and differentiation <strong>of</strong> progenitor cells<br />

along a particular lineage. These outcomes include increased survival or death,<br />

proliferation or growth arrest, and commitment to or blockage <strong>of</strong> differentiation.<br />

There are four mammalian Notch homologues, Notch 1 to 4, which interact with<br />

the DSL ligands: Delta, Delta-like, Jagged1, and Jagged2 in vertebrates (71).<br />

Notch 1 and Notch 4 homologues are involved in the normal development<br />

<strong>of</strong> the mammary gland, and mouse mammary tumors are associated with their<br />

mutated forms. Notch signaling is critical to normal human mammary development<br />

where it acts on both stem cells and progenitor cells, affecting selfrenewal<br />

and lineage-specific differentiation. Notch signaling may also contribute<br />

to mammary carcinogenesis by dysregulating the self-renewal pathways <strong>of</strong><br />

normal mammary stem cells. There is bidirectional cross talk between the Notch<br />

and HH pathways with subsequent regulation <strong>of</strong> the polycomb gene Bmi-1, as<br />

well as the Wnt signaling network (71). The DSL ligands activate Notch signaling,<br />

while the gamma secretase inhibitor (GSI) blocks the intramembrane<br />

cleavage <strong>of</strong> Notch required for signaling (71). The abnormal expression <strong>of</strong> Notch<br />

receptors in epithelial metaplastic lesions and neoplastic lesions suggests that<br />

Notch may act as a protooncogene (26). In spite <strong>of</strong> the simplicity <strong>of</strong> the Notch<br />

signaling pathway, it yields varied outcomes in different contexts (76). These<br />

different outcomes are mediated through a novel signaling pathway in which<br />

Notch receptors on the cell surface are processed to give rise to a nuclear<br />

transcriptional activation complex. Activated Notch modulates stem and progenitor<br />

cell renewal through the upregulation <strong>of</strong> the cyclin-dependent kinase<br />

inhibitors p21 cip1 and p27 kip1 (77). Notch signaling is also required for the<br />

survival <strong>of</strong> niche cells (78). Notch receptor activation promotes the survival <strong>of</strong><br />

neural stem cells by inducing the expression <strong>of</strong> the specific target genes hairy and<br />

enhancer <strong>of</strong> split 3 (Hes3) and SHH. The underlying mechanism occurs through<br />

the rapid activation <strong>of</strong> cytoplasmic signals, including the serine/threonine kinase<br />

AKT, the transcription factor STAT3, and mammalian targets <strong>of</strong> rapamycin<br />

(mTOR) molecular targets <strong>of</strong> rapamycin (79).<br />

The HH signaling pathway has important roles in embryonic development<br />

and tumorigenesis (80). In a number <strong>of</strong> tumors, including prostatic and pancreatic<br />

cancers, HH signaling is abnormally upregulated, and effective signaling<br />

inhibition can repress tumor growth and induce apoptosis (81–83). The core<br />

components <strong>of</strong> the HH pathway are evolutionarily conserved from insects to<br />

mammals. They include the secreted ligand HH, the membrane receptor Patched<br />

(PTCH), the membrane signal transductor Smoothened (SMO), and the transcription<br />

factor Ci/glioma (GLI). However, while in Drosophila, there is only


<strong>Breast</strong> <strong>Cancer</strong> Stem Cells and Their Niche: Lethal Seeds in Lethal Soil 277<br />

one secreted ligand (HH), one receptor (Ptc), and one transcription factor (Ci),<br />

there are three secreted ligands, sonic HH (SHH )/desert HH (DHH)/Indian HH<br />

(IHH), two receptors (PTCH1/2), and three transcription factors (GLI1/2/3) in<br />

mammals (81). The HH signaling components PTCH1 and GLI1/2 are highly<br />

expressed in normal mammary stem and progenitor cells, and activated HH signaling<br />

is mediated by Bmi-1 (30). Addition <strong>of</strong> Hh ligands increases the expression<br />

<strong>of</strong> the transcription factors Gli1 and Gli2, which are positive mediators <strong>of</strong> Hh<br />

signaling (20). SHH can potentiate tumor cell proliferation. Components <strong>of</strong> the<br />

SHH–GLI pathway are expressed in adult human prostate cancer, <strong>of</strong>ten with<br />

enhanced levels in tumors versus normal prostatic epithelia. Cyclopamine and<br />

anti-SHH antibodies inhibit the proliferation <strong>of</strong> GLI1–PSA primary prostate tumor<br />

cultures (83). The SHH signaling pathway also regulates adult neuronal selfrenewal,<br />

and may play a role in brain tumor biology (4).<br />

The final pathway discussed herein is PTEN, which is a phosphatase that<br />

inhibits cellular proliferation by negative regulation <strong>of</strong> signaling through the<br />

phosphatidylinositol-3-OH kinase (PI3K) pathway. The activation <strong>of</strong> PI3K leads<br />

to phosphorylation and activation <strong>of</strong> the Akt protein, which in turn can potentially<br />

phosphorylate a multitude <strong>of</strong> proteins (84). The PTEN–Akt pathway regulates<br />

stem cell activation by helping control nuclear localization <strong>of</strong> the Wnt<br />

pathway effector b-catenin. PTEN inhibits, whereas Akt enhances, the nuclear<br />

localization <strong>of</strong> b-catenin. Akt is thought to coordinate with Wnt signaling to<br />

assist in the activation <strong>of</strong> b-catenin in intestinal stem cells, either through<br />

phosphorylation <strong>of</strong> GSK-3b or by phosphorylation <strong>of</strong> b-catenin itself. The<br />

phosphorylation by Akt inactivates GSK-3b, helping the Wnt signal to permit<br />

b-catenin to escape proteasome-mediated degradation and facilitating its nuclear<br />

translocation and activity in transcriptional regulation (85).<br />

In addition to other downstream effectors, the highly branched PI3K<br />

pathway activates the mammalian target <strong>of</strong> rapamycin (mTOR). Molecules such<br />

as p21 Cip1/Waf1 are found downstream to PTEN, and they influence stem cell<br />

proliferation, without necessarily implicating self-renewal. FOXO regulatory<br />

proteins may have the effects <strong>of</strong> PTEN on p21 Cip1/Waf1 by influencing the<br />

expression <strong>of</strong> a range <strong>of</strong> crucial genes for stem cell function, including cell<br />

division, and tolerance <strong>of</strong> oxidative stress. Thus, the perturbation <strong>of</strong> PTEN may<br />

thus result in nonrenewing cell divisions or death <strong>of</strong> stem cells (84). PTEN has<br />

also been reported to be a guardian <strong>of</strong> genome integrity and <strong>of</strong> the maintenance<br />

<strong>of</strong> chromosomal stability through the physical interaction with centromeres and<br />

control <strong>of</strong> DNA repair. PTEN acts on chromatin and regulates the expression <strong>of</strong><br />

Rad51, which reduces the incidence <strong>of</strong> spontaneous double strand breaks (86).<br />

PTEN is commonly deleted or otherwise inactivated in diverse cancers,<br />

including hematopoietic malignancies, glioblastomas, endometrial carcinomas,<br />

breast carcinomas, and prostate carcinomas (86). PTEN deletion or HER2<br />

overexpression may modulate stem cell function leading to increased invasion<br />

and tumorigenicity (87). PTEN deletions are implicated in the generation <strong>of</strong><br />

transplantable leukemia-initiating cells, and the depletion <strong>of</strong> normal hematopoietic


278 Balicki et al.<br />

stem cells (88). PTEN normally maintains hematopoietic stem cells in a quiescent<br />

state, and in the absence <strong>of</strong> PTEN, hematopoietic stem cells are driven into the cell<br />

cycle, eventually leading to depletion <strong>of</strong> hematopoietic stem cell reserves. Thus,<br />

PTEN activity identifies a mechanistic difference between the maintenance <strong>of</strong><br />

normal stem cells and cancer stem cells. These effects were mostly mediated by<br />

mTOR as they are inhibited by rapamycin, which depletes leukemia-initiating<br />

cells but also restores normal hematopoietic stem cell function (88). This mechanistic<br />

distinction shall enable the design <strong>of</strong> therapies that target the PTEN<br />

pathway to combat leukemia with the goal <strong>of</strong> killing leukemic stem cells, without<br />

damaging the normal stem cell pool (75). In addition to leukemia, this pathway<br />

may be an attractive therapeutic target for treating a broad spectrum <strong>of</strong> tumors<br />

carrying inactivated PTEN because mTOR activity can be suppressed by a number<br />

<strong>of</strong> compounds including rapamycin (75). Genes that promote stem cell quiescence<br />

might be particularly amenable to such targeted strategies because although normal<br />

stem cells thrive in a quiescent world, forcing cancer stem cells to adopt<br />

quiescence could lead to their eradication (75).<br />

TARGETS OF THE STEM CELL NICHE<br />

While cancer stem cells may indeed be lethal seeds in themselves, they may<br />

whither and die unless they are supported by environmental factors, which<br />

perpetuate their survival and self-renewal. The stem cell niche provides this<br />

microenvironment for cancer stem cells, and the combination <strong>of</strong> lethal seeds in<br />

lethal soil is likely to lead to the most aggressive forms <strong>of</strong> breast cancer, with the<br />

most fatal course. The stem cell niche determines cell fate decisions based on<br />

surrounding cells, extracellular matrix components, and secreted local and systemic<br />

factors (11). The 1889 proposal by Stephen Paget that metastasis depends<br />

on cross talk between selected cancer cells (the “seeds”) and specific organ<br />

microenvironments (the “soil”) still holds forth today, and may also apply to the<br />

interrelationship between cancer stem cells and their niche (89–91). <strong>Cancer</strong><br />

metastasis requires the seeding, homing, and successful colonization <strong>of</strong> specialized<br />

cancer stem cells at distant sites (92). In this context, it is not surprising<br />

that the inhibition <strong>of</strong> the homing factor CXCR4 has been shown to effectively<br />

prevent both primary tumor formation as well as metastasis in animal models<br />

(92). Stem cells <strong>of</strong> various tissues exist within protective niches that are composed<br />

<strong>of</strong> differentiated cells which provide direct cell contacts and secreted<br />

factors that maintain stem cells primarily in a quiescent state by providing cues<br />

for the inhibition <strong>of</strong> proliferation (8,44). The mutations that distinguish cancer<br />

stem cells from normal stem cells may enable them to escape niche control.<br />

Alternatively, the niche provides proliferation or differentiation-inducing signals<br />

at times when high numbers <strong>of</strong> progenitors are needed and can quickly give rise<br />

to all committed cell lineages (8). As a result, the uncontrolled proliferation <strong>of</strong><br />

stem cells and tumorigenesis may be a consequence <strong>of</strong> the dysregulation <strong>of</strong><br />

extrinsic factors within the niche. In this case, the stem cell niche might represent


<strong>Breast</strong> <strong>Cancer</strong> Stem Cells and Their Niche: Lethal Seeds in Lethal Soil 279<br />

targets for cancer therapy (44). Given that postnatal estrogen exposure is related to<br />

breast cancer risk, this effect could also apply in utero, where the developing fetus<br />

is exposed to estrogen levels up to 10-fold higher than at any other time in the<br />

individual’s lifetime. Under these circumstances, estrogens or other hormones may<br />

create the fertile soil that is responsible for an increase in the responsiveness <strong>of</strong><br />

estrogen-responsive tissue to undergo oncogenesis later in life (93).<br />

Epigenetic modulation <strong>of</strong> gene expression is highly abnormal in cancers,<br />

which <strong>of</strong>ten show aberrant promoter CpG island hypermethylation and transcriptional<br />

silencing <strong>of</strong> tumor suppressor genes and prodifferentiation factors.<br />

The aberrant epigenetic landscape <strong>of</strong> the cancer cell is also characterized by a<br />

massive genomic hypomethylation, an altered histone code for critical genes, and<br />

a global loss <strong>of</strong> monoacetylated and trimethylated histone H4 (94). In embryonic<br />

stem cells, these genes are maintained in a state that is ready for transcription<br />

mediated by a bivalent promoter chromatin pattern consisting <strong>of</strong> the repressor,<br />

histone H3 methylated at Lys27 (H3K27) by polycomb group proteins, plus the<br />

activator, methylated H3K4. However, in contrast to embryonic stem cells,<br />

embryonic carcinoma cells add two key repressors, dimethylated H3K9 and<br />

trimethylated H3K9, which are both associated with DNA hypermethylation in<br />

adult cancers (14). Polycomb group proteins reversibly repress genes required<br />

for differentiation in embryonic stem cells. Stem cell polycomb group targets are<br />

up to 12-fold more likely to have cancer-specific promoter DNA hypermethylation<br />

than nontargets, supporting a stem cell origin <strong>of</strong> cancer where<br />

reversible gene repression is replaced by permanent silencing, locking the cell<br />

into a perpetual state <strong>of</strong> self-renewal, thereby predisposing to subsequent<br />

malignant transformation (95). Thus, cell chromatin patterns and transient<br />

silencing <strong>of</strong> these important regulatory genes in stem or progenitor cells may<br />

leave these genes vulnerable to aberrant DNA hypermethylation and heritable<br />

gene silencing during tumor initiation and progression.<br />

Vascular niches have been described in the case <strong>of</strong> neural stem cells, which<br />

were found to be concentrated in regions that are rich in blood vessels (18).<br />

These vessels shelter neural stem cells from apoptotic stimuli and maintain a<br />

proper balance between self-renewal and differentiation. The endothelial cells<br />

that line blood vessels are key components <strong>of</strong> the vascular niche as they secrete<br />

factors that promote stem cell survival and self-renewal. As the cotransplantation<br />

<strong>of</strong> tumor cells with endothelial cells leads to more rapid tumor formation, the<br />

vascular niche may contribute to tumor initiation and development (44). In an<br />

established tumor, cancer stem cells that find a vascular niche may continue to<br />

self-renew, while the remaining cells may differentiate to form the bulk <strong>of</strong> the<br />

tumor, but not to its long-term maintenance (18). Thus, the identification <strong>of</strong><br />

endothelial cell signals, which regulate the self-renewal <strong>of</strong> cancer stem cells,<br />

may also contribute to our understanding <strong>of</strong> the interrelationship <strong>of</strong> cancer stem<br />

cells and their niche. There is some evidence that there is a bidirectional relationship<br />

between cancer stem cells and their niche. For example, gliomas secrete<br />

elevated levels <strong>of</strong> vascular endothelial growth factor (VEGF), which increases


280 Balicki et al.<br />

endothelial cell migration and tube formation (43). In addition, VEGF promotes<br />

neural stem cell survival (96). Thus, the effectiveness <strong>of</strong> antiangiogenic agents in<br />

cancer therapy may relate to the prevention <strong>of</strong> new blood vessels that inhibit<br />

tumor growth. In addition, antiangiogenic therapy may disrupt a vascular niche,<br />

which is necessary for cancer stem cell renewal (18).<br />

In the case <strong>of</strong> hematopoietic stem cell niches, studies have focused on the<br />

endosteum, the inner surface <strong>of</strong> the bone that interfaces with the bone marrow.<br />

Regulation <strong>of</strong> hematopoietic stem cell function is governed by factors secreted<br />

by osteoblasts, osteoclasts, and stromal fibroblasts within the endosteum (8).<br />

Imaging data suggests that hematopoietic stem cells reside near the endosteum,<br />

and that this localization and maintenance is also usually adjacent to CXCL12-<br />

secreting reticular cells, irrespective <strong>of</strong> whether the hematopoietic stem cells<br />

are in vascular or endosteal locations (97,98). The stimulation via parathyroid<br />

hormone–related protein receptors (PPR) or an increase in number <strong>of</strong> osteoblastic<br />

cells in the periosteum with high levels <strong>of</strong> the Notch ligand Jagged1,<br />

enlarged the stem cell niche with a resultant increase in the number <strong>of</strong> hematopoietic<br />

stem cells with evidence <strong>of</strong> Notch1 activation in vivo (99). The<br />

activation <strong>of</strong> Notch1 has been shown to result in enhanced self-renewal <strong>of</strong><br />

hematopoietic stem cells possibly by inducing self-renewal genes including<br />

Hes1 (8). In addition, the bone morphogenetic protein signal has an essential<br />

role in inducing haematopoietic tissue during embryogenesis within the hematopoietic<br />

stem cell niche (100). Mesenchymal stem cells have been shown to<br />

improve engraftment <strong>of</strong> hematopoietic stem cells, and suppress graft-versushost<br />

disease after allogeneic hematopoietic stem cell transplantation. In addition,<br />

when tumor cells are injected into NOD/SCID mice in conjunction with<br />

mesenchymal stem cells, their growth is much faster as compared to the group<br />

receiving only tumor cells. To explain this observation, it was suggested that<br />

mesenchymal stem cells have the ability to form a cancer stem cell niche in<br />

which tumor cells can preserve their proliferative potential to sustain the<br />

malignant process (101).<br />

Activated, irradiated, mammary fibroblasts are critical to the microenvironment<br />

<strong>of</strong> mammary stem cells as they are required for the successful<br />

repopulation <strong>of</strong> cleared mammary fat pads in immunocompromised mice (102).<br />

These epithelial-stromal interactions are also important in malignant transformation<br />

and likely to be important with cancer stem cells (11). Fibroblasts secrete<br />

matrix metalloproteinases and protease inhibitors and mobilize and activate<br />

growth factors such as tumor growth factor-b (TGF-b), which has been shown to<br />

promote tumor progression and invasion (103). Carcinoma-associated fibroblasts<br />

promote angiogenesis, as well as tumor growth, by secreting stromal cell-derived<br />

factor-1 (SDF-1), which interacts with the chemokine receptor CXCR4 on tumor<br />

cells and endothelial cells (104). As radiographic breast density correlates with<br />

increased breast cancer risk, as well as increased fibroblasts and collagen<br />

deposition, this modified niche may promote the transformation <strong>of</strong> initiated stem<br />

cells into cancer stem cells (11,105).


<strong>Breast</strong> <strong>Cancer</strong> Stem Cells and Their Niche: Lethal Seeds in Lethal Soil 281<br />

FUTURE CHALLENGES: INDIVIDUALIZED THERAPY BASED<br />

ON BREAST CANCER STEM CELLS AND THEIR NICHE<br />

In the last few years, there has been a significant evolution in breast cancer care, if<br />

not a revolution (106). Whereas, we previously focused on tumor size, nodal<br />

status, and certain prognostic features based on tumor pathology, the paradigm has<br />

now shifted to include the underlying molecular pathways that drive cancerigenesis,<br />

as well as the cellular targets <strong>of</strong> breast cancer, including cancer stem cells<br />

and their niche. The intricate circuitry <strong>of</strong> pathways that define the hallmarks <strong>of</strong><br />

cancer and the corresponding complex molecular interactions are being translated<br />

into therapeutic benefit (107,108). Individualized drivers are gradually being<br />

factored into patient assessments in order to astutely and accurately determine the<br />

clinical course <strong>of</strong> breast cancer patients on the basis <strong>of</strong> their unique and complex<br />

cellular and molecular breast cancer portraits.<br />

<strong>Breast</strong> cancer stem cell markers and the critical molecules in the pathways<br />

that drive their self-renewal are necessary to complete the molecular portrait <strong>of</strong><br />

breast cancer. Our failure to eradicate most cancers may be as fundamental as a<br />

misidentification <strong>of</strong> the target cancer stem cell (Fig. 3) (6). The fact that stem cells<br />

rarely divide and have unique cellular properties, coupled with the observation that<br />

they may have high levels <strong>of</strong> drug transporters to pump chemotherapy agents out<br />

<strong>of</strong> the cell, has led many to believe that traditional chemo- or radiation therapies<br />

are insufficient to clear these tumor-initiating cells from the body (92). Clinical<br />

observations indicate that breast cancer responses to therapy do not necessarily<br />

correlate with patient survival and are incomplete clinical endpoints. Instead, the<br />

effectiveness <strong>of</strong> these treatments should be evaluated by decreased cancer recurrence<br />

and metastases as measures <strong>of</strong> therapeutic efficacy (92). As the molecular<br />

pathways that govern normal and cancer stem cells overlap, the targeting <strong>of</strong> these<br />

pathways could have harmful effects on the homeostatic function <strong>of</strong> normal stem<br />

cells. Targeted therapies that specifically eliminate cancer stem cells could further<br />

revolutionize the ways by which we treat cancer.<br />

As the proliferation <strong>of</strong> cancer stem cells is important for the growth <strong>of</strong> the<br />

primary tumor, as well as progression to metastatic disease, inhibition <strong>of</strong> the selfrenewal<br />

capacity <strong>of</strong> these cells is an obvious target for their elimination. Therapies<br />

such as cyclopamine (targeting the HH signaling) and exisulind, bromoindirubin-3 0 -<br />

oxime, and imatinib (targeting the Wnt/b-catenin pathways) were designed to<br />

inhibit self-renewal pathways critical to cancer stem cells, and have achieved<br />

varying levels <strong>of</strong> success (92). Another strategy to eliminate self-renewal is to force<br />

cancer stem cells to differentiate (Fig. 6).<br />

As cancer stem cells are currently being identified in a broad spectrum <strong>of</strong><br />

cancers (44,109–112), significant advances in their fundamental properties are<br />

forthcoming, and will serve as a scaffold to expedite our understanding and identification<br />

<strong>of</strong> cancer stem cells within specific types <strong>of</strong> cancers, including breast<br />

cancer. The existence <strong>of</strong> overlapping cancer stem cell pathways within different<br />

cancer types will also stimulate the discovery and development <strong>of</strong> new drugs to


282 Balicki et al.<br />

target these pathways, as they are likely to be effective and marketable in more than<br />

merely a small subset <strong>of</strong> cancer patients.<br />

The addition <strong>of</strong> molecular markers in our patient assessments acknowledges<br />

the pivotal role that the corresponding molecular pathways play in fueling<br />

tumorigenesis. This fine-tuning <strong>of</strong> patient prognostic factors will ensure that the<br />

oncogenic drivers <strong>of</strong> each individual patient’s tumor pathophysiology are both<br />

identified and tailored into the definition and optimization <strong>of</strong> their specific<br />

therapeutic strategy, which will translate into their best possible clinical outcome<br />

and overall survival. As a result, it is critical that our therapeutic armamentarium<br />

target the specific fingerprint <strong>of</strong> cellular and molecular pathways defined for an<br />

individual patient. Therapeutic resistance via quiescence and the sequestration <strong>of</strong><br />

cancer stem cells into sanctuaries that are impermeable to most therapeutic<br />

agents must be taken into consideration in the design <strong>of</strong> individualized breast<br />

cancer therapies. Therapeutic resistance may also result from an inadequate<br />

depletion <strong>of</strong> key cellular and molecular elements within the cancer stem cell<br />

niche. The elucidation <strong>of</strong> the secreted factors that set up the premetastasis and<br />

metastatic niche could provide both diagnostic as well as therapeutic benefits.<br />

Drugs such as Bevacizumab (Avastin 1 ) that reduce intratumoral pressure<br />

associated with leaky tumoral vasculature and thereby increase the introduction<br />

<strong>of</strong> intravenous agents within the tumor bed may have the critical synergy necessary<br />

in combinations <strong>of</strong> targeted agents to permeate cancer stem cell sanctuaries<br />

and cancer stem cell niches (18,44,113,114). In addition, they may target<br />

both the vascular niche and the associated cancer stem cells (44). Thus, it is not<br />

surprising that clinical trials <strong>of</strong> Bevacizumab combined with the chemotherapeutic<br />

drug CPT-11 suggest that this combination is one <strong>of</strong> the most<br />

effective treatments <strong>of</strong> glioblastoma (115). Other agents that stimulate cycling <strong>of</strong><br />

quiescent cancer stem cells may also be beneficial in combined therapies, which<br />

include agents that eradicate cycling cells. In addition, individualized therapy<br />

must also have the long-term objectives <strong>of</strong> predicting and preparing to eradicate<br />

the molecular pathways that may circumvent those pathways that drove the<br />

initial breast cancer pathophysiology, and were targeted by first-line targeted<br />

therapies. This strategy also underlines the critical objective <strong>of</strong> eradicating every<br />

single cancer stem cell and its niche in newly diagnosed breast cancer patients,<br />

thereby striving to eliminate all possibilities that these lethal seeds in lethal soil<br />

may ignite tumor recurrence in the future.<br />

Therefore, the genesis <strong>of</strong> individualized breast cancer therapy including<br />

the eradication <strong>of</strong> cancer stem cells has become a necessity to achieve optimal<br />

breast cancer patient care. It is imperative to understand the biology <strong>of</strong> cancer<br />

stem cells and their niche to optimize our cancer treatment strategies to prevent<br />

the longevity <strong>of</strong> cancer stem cells from interfering with our own (18). The<br />

associated costs <strong>of</strong> its future reality precipitate a call to rethink our healthcare<br />

strategies with a focus on cancer prevention and therapeutic schedules which<br />

deliver targeted agents when they are both necessary and most effective in our<br />

quest to optimally treat an individual breast cancer patient with the ultimate goal


<strong>Breast</strong> <strong>Cancer</strong> Stem Cells and Their Niche: Lethal Seeds in Lethal Soil 283<br />

<strong>of</strong> a stable, long-term remission. In this setting, the identification and targeting <strong>of</strong><br />

breast cancer stem cells and their niche hold great promise in the development <strong>of</strong><br />

new therapeutic translational strategies using state-<strong>of</strong>-the-art individualized<br />

therapies for breast cancer patients.<br />

ACKNOWLEDGMENTS<br />

Rhyna Salinas (CHUM) is gratefully acknowledged for her expert assistance in<br />

the preparation <strong>of</strong> this chapter. The contributions made by Drs. Francine<br />

Jolicoeur, Isabel Garcia de la Fuente, and Louis Gaboury (CHUM) to Figure 4<br />

during their collaboration with D.B. are greatly appreciated. Drs. Connie Eaves,<br />

Gabriela Dontu, Suling Liu, Christophe Ginestier, Hasan Korkaya, and Julie<br />

Dutcher are thankfully acknowledged for insightful and inspirational discussions.<br />

Dr. Connie Eaves is also gratefully recognized for her invaluable role<br />

as D.B.’s first stem cell research mentor and role model, who initially planted the<br />

seed that she continues to nourish. D.B. is a scholar <strong>of</strong> Fonds de la recherche en<br />

santé du Québec (FRSQ) and principal investigator <strong>of</strong> the National <strong>Cancer</strong><br />

Institute <strong>of</strong> Canada’s Clinical Trials Group (NCIC CTG). This chapter was also<br />

made possible through a Canadian Institutes for Health Research (CIHR)<br />

operating grant to D.B..<br />

REFERENCES<br />

1. Wicha MS, Liu S, Dontu G. <strong>Cancer</strong> stem cells: an old idea—a paradigm shift.<br />

<strong>Cancer</strong> Res 2006; 66(4):1883–1890.<br />

2. Balicki D. <strong>Cancer</strong> stem cells and resistance to treatment. Oncol Exchange 2007;<br />

6(1): 37–40.<br />

3. Vogelstein B, Kinzler KW. <strong>Cancer</strong> genes and the pathways they control. Nat Med<br />

2004; 10(8):789–799.<br />

4. Massard C, Deutsch E, Soria JC. Tumour stem cell-targeted treatment: elimination<br />

or differentiation. Ann Oncol 2006; 17(11):1620–1624.<br />

5. Wicha MS. <strong>Cancer</strong> stem cells and metastasis: lethal seeds. Clin <strong>Cancer</strong> Res 2006;<br />

12(19):5606–5607.<br />

6. Jones RJ, Matsui WH, Smith BD. <strong>Cancer</strong> stem cells: are we missing the target?<br />

J Natl <strong>Cancer</strong> Inst 2004; 96(8):583–585.<br />

7. Jordan CT, Guzman ML, Noble M. <strong>Cancer</strong> stem cells. N Engl J Med 2006; 355(12):<br />

1253–1261.<br />

8. Rizo A, Vellenga E, de Haan G, et al. Signaling pathways in self-renewing hematopoietic<br />

and leukemic stem cells: do all stem cells need a niche? Hum Mol Genet<br />

2006; 15(spec no 2):R210–R219.<br />

9. Schwartz YB, Pirrotta V. Polycomb silencing mechanisms and the management <strong>of</strong><br />

genomic programmes. Nat Rev Genet 2007; 8(1):9–22.<br />

10. Bernstein A. <strong>Cancer</strong> stem cells: the centrality <strong>of</strong> translational research to cancer<br />

control. CMAJ 2007; 176(1):29–30.


284 Balicki et al.<br />

11. Dontu G, Liu S, Wicha MS. Stem cells in mammary development and carcinogenesis:<br />

implications for prevention and treatment. Stem Cell Rev 2005; 1(3):207–213.<br />

12. Dontu G, El Ashry D, Wicha MS. <strong>Breast</strong> cancer, stem/progenitor cells and the<br />

estrogen receptor. Trends Endocrinol Metab 2004; 15(5):193–197.<br />

13. Krivtsov AV, Twomey D, Feng Z, et al. Transformation from committed progenitor<br />

to leukaemia stem cell initiated by MLL-AF9. Nature 2006; 442(7104):818–822.<br />

14. Ohm JE, McGarvey KM, Yu X, et al. A stem cell-like chromatin pattern may predispose<br />

tumor suppressor genes to DNA hypermethylation and heritable silencing.<br />

Nat Genet 2007; 39(2):237–242.<br />

15. Clevers H. Stem cells, asymmetric division and cancer. Nat Genet 2005; 37(10):<br />

1027–1028.<br />

16. Scadden DT. The stem-cell niche as an entity <strong>of</strong> action. Nature 2006; 441(7097):<br />

1075–1079.<br />

17. Dontu G, Abdallah WM, Foley JM, et al. In vitro propagation and transcriptional<br />

pr<strong>of</strong>iling <strong>of</strong> human mammary stem/progenitor cells. Genes Dev 2003; 17(10):<br />

1253–1270.<br />

18. Yang ZJ, Wechsler-Reya RJ. Hit ’em where they live: targeting the cancer stem cell<br />

niche. <strong>Cancer</strong> Cell 2007; 11(1):3–5.<br />

19. Moore KA, Lemischka IR. Stem cells and their niches. Science 2006; 311(5769):<br />

1880–1885.<br />

20. Liu S, Dontu G, Mantle ID, et al. Hedgehog signaling and Bmi-1 regulate selfrenewal<br />

<strong>of</strong> normal and malignant human mammary stem cells. <strong>Cancer</strong> Res 2006;<br />

66(12):6063–6071.<br />

21. Al-Hajj M, Wicha MS, ito-Hernandez A, et al. Prospective identification <strong>of</strong><br />

tumorigenic breast cancer cells. Proc Natl Acad Sci U S A 2003; 100(7):3983–3988.<br />

22. Stingl J, Eirew P, Ricketson I, et al. Purification and unique properties <strong>of</strong> mammary<br />

epithelial stem cells. Nature 2006; 439(7079):993–997.<br />

23. Smith GH. Experimental mammary epithelial morphogenesis in an in vivo model:<br />

evidence for distinct cellular progenitors <strong>of</strong> the ductal and lobular phenotype. <strong>Breast</strong><br />

<strong>Cancer</strong> Res Treat 1996; 39(1):21–31.<br />

24. Kordon EC, Smith GH. An entire functional mammary gland may comprise the<br />

progeny from a single cell. Development 1998; 125(10):1921–1930.<br />

25. Shackleton M, Vaillant F, Simpson KJ, et al. Generation <strong>of</strong> a functional mammary<br />

gland from a single stem cell. Nature 2006; 439(7072):84–88.<br />

26. Dontu G, Jackson KW, McNicholas E, et al. Role <strong>of</strong> Notch signaling in cell-fate<br />

determination <strong>of</strong> human mammary stem/progenitor cells. <strong>Breast</strong> <strong>Cancer</strong> Res 2004;<br />

6(6):R605–R615.<br />

27. Youn BS, Sen A, Behie LA, et al. Scale-up <strong>of</strong> breast cancer stem cell aggregate<br />

cultures to suspension bioreactors. Biotechnol Prog 2006; 22(3):801–810.<br />

28. Hadnagy A, Gaboury L, Beaulieu R, et al. SP analysis may be used to identify<br />

cancer stem cell populations. Exp Cell Res 2006; 312(19):3701–3710.<br />

29. Goodell MA, Brose K, Paradis G, et al. Isolation and functional properties <strong>of</strong><br />

murine hematopoietic stem cells that are replicating in vivo. J Exp Med 1996;<br />

183(4):1797–1806.<br />

30. Balicki D, Beaulieu B. <strong>Cancer</strong> stem cell side populations. In: Bapat SA, ed. <strong>Cancer</strong><br />

Stem Cells. Hoboken: John Wiley and sons, Inc., 2008 (in press).<br />

31. Dontu G, Al-Hajj M, Abdallah WM, et al. Stem cells in normal breast development<br />

and breast cancer. Cell Prolif 2003; 36(suppl 1):59–72.


<strong>Breast</strong> <strong>Cancer</strong> Stem Cells and Their Niche: Lethal Seeds in Lethal Soil 285<br />

32. Al-Hajj M, Becker MW, Wicha M, et al. Therapeutic implications <strong>of</strong> cancer stem<br />

cells. Curr Opin Genet Dev 2004; 14(1):43–47.<br />

33. Sjoblom T, Jones S, Wood LD, et al. The consensus coding sequences <strong>of</strong> human<br />

breast and colorectal cancers. Science 2006; 314(5797):268–274.<br />

34. Liu R, Wang X, Chen GY, et al. The prognostic role <strong>of</strong> a gene signature from<br />

tumorigenic breast-cancer cells. N Engl J Med 2007; 356(3):217–226.<br />

35. Chang JC, Li X, Wong H, et al. Journal <strong>of</strong> Clinical Oncology, 2007 ASCO Annual<br />

Meeting Proceedings Part I. Vol. 25, No. 18S (June 20 Supplement), 2007:9S.<br />

36. Woodward WA, Chen MS, Behbod F, et al. WNT/beta-catenin mediates radiation<br />

resistance <strong>of</strong> mouse mammary progenitor cells. Proc Natl Acad Sci U S A 2007;<br />

104(2):618–623.<br />

37. Lymberis SC, Parhar PK, Katsoulakis E, et al. <strong>Ph</strong>armacogenomics and breast<br />

cancer. <strong>Ph</strong>armacogenomics 2004; 5(1):31–55.<br />

38. Foulkes WD. BRCA1 functions as a breast stem cell regulator. J Med Genet 2004;<br />

41(1):1–5.<br />

39. Liu S, Ginestier S, Dontu G, et al. BRCA1 regulates human mammary stem cell<br />

self-renewal and differentiation. In: Proceedings <strong>of</strong> the 98th Annual Meeting <strong>of</strong> the<br />

American Association for <strong>Cancer</strong> Research, Los Angeles, CA, 14–18 April 2007.<br />

<strong>Ph</strong>iladelphia, PA: AACR; 2007:1346–1347. Abstract nr 5700.<br />

40. Stearns V, Davidson NE, Flockhart DA. <strong><strong>Ph</strong>armacogenetics</strong> in the treatment <strong>of</strong><br />

breast cancer. <strong>Ph</strong>armacogenomics J 2004; 4(3):143–153.<br />

41. Evans WE, Relling MV. Moving towards individualized medicine with pharmacogenomics.<br />

Nature 2004; 429(6990):464–468.<br />

42. Potti A, Dressman HK, Bild A, et al. Genomic signatures to guide the use <strong>of</strong><br />

chemotherapeutics. Nat Med 2006; 12(11):1294–1300.<br />

43. Bao S, Wu Q, McLendon RE, et al. Glioma stem cells promote radioresistance<br />

by preferential activation <strong>of</strong> the DNA damage response. Nature 2006; 444(7120):<br />

756–760.<br />

44. Calabrese C, Poppleton H, Kocak M, et al. A perivascular niche for brain tumor<br />

stem cells. <strong>Cancer</strong> Cell 2007; 11(1):69–82.<br />

45. Donnenberg VS, Donnenberg AD. Multiple drug resistance in cancer revisited: the<br />

cancer stem cell hypothesis. J Clin <strong>Ph</strong>armacol 2005; 45(8):872–877.<br />

46. Kim M, Turnquist H, Jackson J, et al. The multidrug resistance transporter ABCG2<br />

(breast cancer resistance protein 1) effluxes Hoechst 33342 and is overexpressed in<br />

hematopoietic stem cells. Clin <strong>Cancer</strong> Res 2002; 8(1):22–28.<br />

47. Zhou S, Schuetz JD, Bunting KD, et al. The ABC transporter Bcrp1/ABCG2 is<br />

expressed in a wide variety <strong>of</strong> stem cells and is a molecular determinant <strong>of</strong> the sidepopulation<br />

phenotype. Nat Med 2001; 7(9):1028–1034.<br />

48. Zhou S, Morris JJ, Barnes Y, et al. Bcrp1 gene expression is required for normal<br />

numbers <strong>of</strong> side population stem cells in mice, and confers relative protection to<br />

mitoxantrone in hematopoietic cells in vivo. Proc Natl Acad Sci U S A 2002;<br />

99(19):12339–12344.<br />

49. Krishnamurthy P, Ross DD, Nakanishi T, et al. The stem cell marker Bcrp/ABCG2<br />

enhances hypoxic cell survival through interactions with heme. J Biol Chem 2004;<br />

279(23):24218–24225.<br />

50. Asselin-Labat ML, Shackleton M, Stingl J, et al. Steroid hormone receptor status <strong>of</strong><br />

mouse mammary stem cells. J Natl <strong>Cancer</strong> Inst 2006; 98(14):1011–1014.


286 Balicki et al.<br />

51. Sleeman KE, Kendrick H, Robertson D, et al. Dissociation <strong>of</strong> estrogen receptor<br />

expression and in vivo stem cell activity in the mammary gland. J Cell Biol 2007;<br />

176(1):19–26.<br />

52. Al Hajj M, Wicha MS, Benito-Hernandez A, et al. Prospective identification <strong>of</strong><br />

tumorigenic breast cancer cells. Proc Natl Acad Sci U S A 2003; 100(7):3983–3988.<br />

53. Williams DA, Cancelas JA. Leukaemia: niche retreats for stem cells. Nature 2006;<br />

444(7121):827–828.<br />

54. Kastan MB, Schlaffer E, Russo JE, et al. Direct demonstration <strong>of</strong> elevated aldehyde<br />

dehydrogenase in human hematopoietic progenitor cells. Blood 1990; 75(10):<br />

1947–1950.<br />

55. Russo JE, Hilton J, Colvin OM. The role <strong>of</strong> aldehyde dehydrogenase isozymes in<br />

cellular resistance to the alkylating agent cyclophosphamide. Prog Clin Biol Res<br />

1989; 290:65–79.<br />

56. Sladek NE, Kollander R, Sreerama L, et al. Cellular levels <strong>of</strong> aldehyde dehydrogenases<br />

(ALDH1A1 and ALDH3A1) as predictors <strong>of</strong> therapeutic responses to cyclophosphamidebased<br />

chemotherapy <strong>of</strong> breast cancer: a retrospective study. Rational individualization<br />

<strong>of</strong> oxazaphosphorine-based cancer chemotherapeutic regimens. <strong>Cancer</strong> Chemother<br />

<strong>Ph</strong>armacol 2002; 49(4):309–321.<br />

57. Gordon MY, Goldman JM, Gordon-Smith EC. 4-Hydroperoxycyclophosphamide<br />

inhibits proliferation by human granulocyte-macrophage colony-forming cells (GM-<br />

CFC) but spares more primitive progenitor cells. Leuk Res 1985; 9(8):1017–1021.<br />

58. Sladek NE. Human aldehyde dehydrogenases: potential pathological, pharmacological,<br />

and toxicological impact. J Biochem Mol Toxicol 2003; 17(1):7–23.<br />

59. Storms RW, Trujillo AP, Springer JB, et al. Isolation <strong>of</strong> primitive human hematopoietic<br />

progenitors on the basis <strong>of</strong> aldehyde dehydrogenase activity. Proc Natl<br />

Acad Sci U S A 1999; 96(16):9118–9123.<br />

60. Ginestier C, Hur M, Charafe-Jauffret E, et al. ALDH1 is a marker <strong>of</strong> normal and<br />

breast cancer stem cells and predicts poor clinical outcome. Targeted Oncol 2007;<br />

2(suppl 1):S9.<br />

61. Charafe-Jauffret E, Ginestier C, Hur MH, et al. Aldefluor-positive populations in<br />

breast cell lines: evidence for cancer stem cells properties and gene expression pr<strong>of</strong>ile<br />

(GEP). In: Proceedings <strong>of</strong> the 98th Annual Meeting <strong>of</strong> the American Association for<br />

<strong>Cancer</strong> Research, Los Angeles, CA, 14–18 April 2007. <strong>Ph</strong>iladelphia, PA: AACR,<br />

2007:310. Abstract 1307.<br />

62. Pearce DJ, Taussig D, Simpson C, et al. Characterization <strong>of</strong> cells with a high<br />

aldehyde dehydrogenase activity from cord blood and acute myeloid leukemia<br />

samples. Stem Cells 2005; 23(6):752–760.<br />

63. Lee J, Kim HK, Rho JY, et al. The human OCT-4 is<strong>of</strong>orms differ in their ability to<br />

confer self-renewal. J Biol Chem 2006; 281(44):33554–33565.<br />

64. Atlasi Y, Mowla SJ, Ziaee SA, et al. OCT-4, an embryonic stem cell marker, is<br />

highly expressed in bladder cancer. Int J <strong>Cancer</strong> 2007; 120(7):1598–1602.<br />

65. Ponti D, Costa A, Zaffaroni N, et al. Isolation and in vitro propagation <strong>of</strong> tumorigenic<br />

breast cancer cells with stem/progenitor cell properties. <strong>Cancer</strong> Res 2005;<br />

65(13):5506–5511.<br />

66. Kang MK, Kim RH, Kim SJ, et al. Elevated Bmi-1 expression is associated with<br />

dysplastic cell transformation during oral carcinogenesis and is required for cancer<br />

cell replication and survival. Br J <strong>Cancer</strong> 2007; 96(1):126–133.


<strong>Breast</strong> <strong>Cancer</strong> Stem Cells and Their Niche: Lethal Seeds in Lethal Soil 287<br />

67. Kim JH, Yoon SY, Jeong SH, et al. Overexpression <strong>of</strong> Bmi-1 oncoprotein correlates<br />

with axillary lymph node metastases in invasive ductal breast cancer. <strong>Breast</strong> 2004;<br />

13(5):383–388.<br />

68. Dimri GP, Martinez JL, Jacobs JJ, et al. The Bmi-1 oncogene induces telomerase<br />

activity and immortalizes human mammary epithelial cells. <strong>Cancer</strong> Res 2002;<br />

62(16):4736–4745.<br />

69. Fodde R, Brabletz T. Wnt/beta-catenin signaling in cancer stemness and malignant<br />

behavior. Curr Opin Cell Biol 2007; 19(2):150–8.<br />

70. Dontu G, Wicha MS. Survival <strong>of</strong> mammary stem cells in suspension culture:<br />

implications for stem cell biology and neoplasia. J Mammary Gland Biol Neoplasia<br />

2005; 10(1):75–86.<br />

71. Liu S, Dontu G, Wicha MS. Mammary stem cells, self-renewal pathways, and<br />

carcinogenesis. <strong>Breast</strong> <strong>Cancer</strong> Res 2005; 7(3):86–95.<br />

72. Warner JK, Wang JC, Takenaka K, et al. Direct evidence for cooperating genetic<br />

events in the leukemic transformation <strong>of</strong> normal human hematopoietic cells. Leukemia<br />

2005; 19(10):1794–1805.<br />

73. He XC, Zhang J, Tong WG, et al. BMP signaling inhibits intestinal stem cell selfrenewal<br />

through suppression <strong>of</strong> Wnt-beta-catenin signaling. Nat Genet 2004;<br />

36(10):1117–1121.<br />

74. Adolphe C, Narang M, Ellis T, et al. An in vivo comparative study <strong>of</strong> sonic, desert<br />

and Indian hedgehog reveals that hedgehog pathway activity regulates epidermal<br />

stem cell homeostasis. Development 2004; 131(20):5009–5019.<br />

75. Rossi DJ, Weissman IL. Pten, tumorigenesis, and stem cell self-renewal. Cell 2006;<br />

125(2):229–231.<br />

76. Roy M, Pear WS, Aster JC. The multifaceted role <strong>of</strong> Notch in cancer. Curr Opin<br />

Genet Dev 2007; 17(1):52–59.<br />

77. Bullock TE, Wen B, Marley SB, et al. Potential <strong>of</strong> CD34 in the regulation <strong>of</strong><br />

symmetrical and asymmetrical divisions by hematopoietic progenitor cells. Stem<br />

Cells 2006; 25(4):844–851.<br />

78. Song X, Call GB, Kirilly D, et al. Notch signaling controls germline stem cell niche<br />

formation in the Drosophila ovary. Development 2007; 134(6):1071–1080.<br />

79. Androutsellis-Theotokis A, Leker RR, Soldner F, et al. Notch signalling regulates<br />

stem cell numbers in vitro and in vivo. Nature 2006; 442(7104):823–826.<br />

80. Kalderon D. The mechanism <strong>of</strong> hedgehog signal transduction. Biochem Soc Trans<br />

2005; 33(pt 6):1509–1512.<br />

81. Shimokawa T, Svard J, Heby-Henricson K, et al. Distinct roles <strong>of</strong> first exon variants<br />

<strong>of</strong> the tumor-suppressor Patched1 in Hedgehog signaling. Oncogene 2007; 26(34):<br />

4889–4896; [Epub 2007 February 19].<br />

82. Sanchez P, Hernandez AM, Stecca B, et al. Inhibition <strong>of</strong> prostate cancer proliferation<br />

by interference with Sonic Hedgehog-GLI1 signaling. Proc Natl Acad Sci U S A<br />

2004; 101(34):12561–12566.<br />

83. Ohuchida K, Mizumoto K, Fujita H, et al. Sonic hedgehog is an early developmental<br />

marker <strong>of</strong> intraductal papillary mucinous neoplasms: clinical implications<br />

<strong>of</strong> mRNA levels in pancreatic juice. J Pathol 2006; 210(1):42–48.<br />

84. Janzen V, Scadden DT. Stem cells: good, bad and reformable. Nature 2006;<br />

441(7092):418–419.<br />

85. He XC, Yin T, Grindley JC, et al. PTEN-deficient intestinal stem cells initiate<br />

intestinal polyposis. Nat Genet 2007; 39(2):189–198.


288 Balicki et al.<br />

86. Shen WH, Balajee AS, Wang J, et al. Essential role for nuclear PTEN in maintaining<br />

chromosomal integrity. Cell 2007; 128(1):157–170.<br />

87. Korkaya H, Iovino F, Wicha MS. PTEN and HER2 regulate self-renewal and<br />

invasion <strong>of</strong> human mammary stem cells. In: Proceedings <strong>of</strong> the 98th Annual<br />

Meeting <strong>of</strong> the American Association for <strong>Cancer</strong> Research, Los Angeles, CA, 14–18<br />

April 2007. <strong>Ph</strong>iladelphia, PA: AACR, 2007:305. Abstract nr 1287.<br />

88. Yilmaz OH, Valdez R, Theisen BK, et al. Pten dependence distinguishes haematopoietic<br />

stem cells from leukaemia-initiating cells. Nature 2006; 441(7092):475–482.<br />

89. Fidler IJ. The pathogenesis <strong>of</strong> cancer metastasis: the ‘seed and soil’ hypothesis<br />

revisited. Nat Rev <strong>Cancer</strong> 2003; 3(6):453–458.<br />

90. Fidler IJ, Yano S, Zhang RD, et al. The seed and soil hypothesis: vascularisation<br />

and brain metastases. Lancet Oncol 2002; 3(1):53–57.<br />

91. Fidler IJ. Seed and soil revisited: contribution <strong>of</strong> the organ microenvironment to<br />

cancer metastasis. Surg Oncol Clin N Am 2001; 10(2):257–269, vii–viiii.<br />

92. Li F, Tiede B, Massague J, et al. Beyond tumorigenesis: cancer stem cells in<br />

metastasis. Cell Res 2007; 17(1):3–14.<br />

93. Savarese TM, Low HP, Baik I, et al. Normal breast stem cells, malignant breast<br />

stem cells, and the perinatal origin <strong>of</strong> breast cancer. Stem Cell Rev 2006; 2(2):<br />

103–110.<br />

94. Esteller M. The necessity <strong>of</strong> a human epigenome project. Carcinogenesis 2006;<br />

27(6):1121–1125.<br />

95. Widschwendter M, Fiegl H, Egle D, et al. Epigenetic stem cell signature in cancer.<br />

Nat Genet 2007; 39(2):157–158.<br />

96. Wada T, Haigh JJ, Ema M, et al. Vascular endothelial growth factor directly inhibits<br />

primitive neural stem cell survival but promotes definitive neural stem cell survival.<br />

J Neurosci 2006; 26(25):6803–6812.<br />

97. Sugiyama T, Kohara H, Noda M, et al. Maintenance <strong>of</strong> the hematopoietic stem cell<br />

pool by CXCL12-CXCR4 chemokine signaling in bone marrow stromal cell niches.<br />

Immunity 2006; 25(6):977–988.<br />

98. Kiel MJ, Morrison SJ. Maintaining hematopoietic stem cells in the vascular niche.<br />

Immunity 2006; 25(6):862–864.<br />

99. Calvi LM, Adams GB, Weibrecht KW, et al. Osteoblastic cells regulate the<br />

haematopoietic stem cell niche. Nature 2003; 425(6960):841–846.<br />

100. Zhang J, Niu C, Ye L, et al. Identification <strong>of</strong> the haematopoietic stem cell niche and<br />

control <strong>of</strong> the niche size. Nature 2003; 425(6960):836–841.<br />

101. Ramasamy R, Lam EW, Soeiro I, et al. Mesenchymal stem cells inhibit proliferation<br />

and apoptosis <strong>of</strong> tumor cells: impact on in vivo tumor growth. Leukemia 2007;<br />

21(2):304–310.<br />

102. Kuperwasser C, Chavarria T, Wu M, et al. Reconstruction <strong>of</strong> functionally normal<br />

and malignant human breast tissues in mice. Proc Natl Acad Sci U S A 2004;<br />

101(14):4966–4971.<br />

103. Bhowmick NA, Chytil A, Plieth D, et al. TGF-beta signaling in fibroblasts<br />

modulates the oncogenic potential <strong>of</strong> adjacent epithelia. Science 2004; 303(5659):<br />

848–851.<br />

104. Orimo A, Gupta PB, Sgroi DC, et al. Stromal fibroblasts present in invasive human<br />

breast carcinomas promote tumor growth and angiogenesis through elevated SDF-1/<br />

CXCL12 secretion. Cell 2005; 121(3):335–348.


<strong>Breast</strong> <strong>Cancer</strong> Stem Cells and Their Niche: Lethal Seeds in Lethal Soil 289<br />

105. Boyd NF, Guo H, Martin LJ, et al. Mammographic density and the risk and<br />

detection <strong>of</strong> breast cancer. N Engl J Med 2007; 356(3):227–236.<br />

106. Hortobagyi GN. Trastuzumab in the treatment <strong>of</strong> breast cancer. N Engl J Med 2005;<br />

353(16):1734–1736.<br />

107. Chabner BA. In defense <strong>of</strong> cell-line screening. J Natl <strong>Cancer</strong> Inst 1990; 82(13):<br />

1083–1085.<br />

108. Hanahan D, Weinberg RA. The hallmarks <strong>of</strong> cancer. Cell 2000; 100(1):57–70.<br />

109. Li C, Heidt DG, Dalerba P, et al. Identification <strong>of</strong> pancreatic cancer stem cells.<br />

<strong>Cancer</strong> Res 2007; 67(3):1030–1037.<br />

110. Dirks PB. Brain tumour stem cells: the undercurrents <strong>of</strong> human brain cancer and<br />

their relationship to neural stem cells. <strong>Ph</strong>ilos Trans R Soc Lond B Biol Sci 2008;<br />

363(1489):139–52.<br />

111. Piccirillo SG, Reynolds BA, Zanetti N, et al. Bone morphogenetic proteins inhibit<br />

the tumorigenic potential <strong>of</strong> human brain tumour-initiating cells. Nature 2006;<br />

444(7120):761–765.<br />

112. Prince ME, Sivanandan R, Kaczorowski A, et al. Identification <strong>of</strong> a subpopulation<br />

<strong>of</strong> cells with cancer stem cell properties in head and neck squamous cell carcinoma.<br />

Proc Natl Acad Sci U S A 2007; 104(3):973–978.<br />

113. Jain RK, Forbes NS. Can engineered bacteria help control cancer? Proc Natl Acad<br />

Sci U S A 2001; 98(26):14748–14750.<br />

114. Heldin CH, Rubin K, Pietras K, et al. High interstitial fluid pressure—an obstacle in<br />

cancer therapy. Nat Rev <strong>Cancer</strong> 2004; 4(10):806–813.<br />

115. Reardon DA, Wen PY. Therapeutic advances in the treatment <strong>of</strong> glioblastoma:<br />

rationale and potential role <strong>of</strong> targeted agents. Oncologist 2006; 11(2):152–164.


19<br />

Molecular Imaging in Individualized<br />

<strong>Cancer</strong> Management<br />

David M. Schuster<br />

Division <strong>of</strong> Nuclear Medicine and Molecular Imaging, Department<br />

<strong>of</strong> Radiology, Emory University, Atlanta, Georgia, U.S.A.<br />

Diego R. Martin<br />

Department <strong>of</strong> Radiology, Emory University, Atlanta, Georgia, U.S.A.<br />

INTRODUCTION<br />

X-rays were discovered by Karl Roentgen in 1895, and with that discovery came<br />

the ability to see structures within the living human body without having to directly<br />

sample the tissue. Imaging, particularly cross-sectional techniques, has revolutionized<br />

the practice <strong>of</strong> medicine. These techniques employ ionizing radiation, such<br />

as computed tomography (CT) scanning, single-photon emission computed<br />

tomography (SPECT), and positron emission tomography (PET), as well as nonionizing<br />

energy, such as ultrasound and magnetic resonance imaging (MRI). Some<br />

<strong>of</strong> the most exciting and practical major advances are unfolding within the MRI and<br />

PET technologies, specifically with regard to molecular imaging.<br />

Molecular imaging is defined as the in vivo characterization and measurement<br />

<strong>of</strong> biological processes at cellular and molecular levels. By definition,<br />

molecular imaging is not restricted to any one tool; it may involve optical,<br />

ultrasound, gamma ray, positron, or magnetic imaging (1). Since the 1960s, bone<br />

scanning has played a major role in the management <strong>of</strong> breast cancer. However,<br />

in the past decade, the role <strong>of</strong> radionuclide molecular imaging has expanded<br />

291


292 Schuster and Martin<br />

significantly in the clinical management <strong>of</strong> breast cancer because <strong>of</strong> PET,<br />

mammoscintigraphy, and sentinel lymph node techniques. Molecular imaging is<br />

also instrumental in drug development, gene therapy, and basic science research<br />

<strong>of</strong> breast cancer. Mammography and ultrasound have been and continue to be<br />

very important tools in the diagnosis and management <strong>of</strong> breast cancer, but other<br />

modalities, such as MRI, have enhanced our ability to detect occult disease,<br />

especially in those with high genetic risk, and to follow breast carcinoma during<br />

and after therapy (2,3).<br />

In this chapter, we will examine MRI and PET with regard to molecular<br />

imaging as applied to individualized breast cancer management; this chapter is<br />

not intended to be a comprehensive review <strong>of</strong> molecular imaging methods. We<br />

have specifically chosen MRI and PET because we believe the integration <strong>of</strong><br />

both technologies will lead to further progress in our ability to interrogate disease<br />

processes in a manner analogous to the classical pathologist, but in a<br />

noninvasive fashion. This ability will continue to expand as technology progresses.<br />

But in essence, the future is already here.<br />

RADIONUCLIDE TECHNIQUES:<br />

CURRENT CONCEPTS AND APPLICATIONS<br />

The advantage <strong>of</strong> radionuclide techniques over other imaging approaches is their<br />

ability to label any chemical species with an isotope <strong>of</strong> choice and to detect that<br />

radiotracer with high sensitivity. When certain radionuclides decay, gamma rays<br />

(photons) are given <strong>of</strong>f, which are detected with special crystal detectors. The<br />

gamma emissions <strong>of</strong> SPECT or PET radiotracers are energetic enough to travel<br />

through tissues and thus enable imaging <strong>of</strong> the whole body. Carbon, nitrogen,<br />

hydrogen, and oxygen have positron-emitting isotopes and are ideally suited for<br />

designing probes to image natural biochemical and molecular process in vivo.<br />

Probes can either home in on targets on cell surfaces, such as antigens and<br />

receptors, or intracellular structures such as internalized receptors.<br />

Single-photon imaging is the basis <strong>of</strong> simple techniques such as planar<br />

bone scanning or more sophisticated cross-sectional approaches such as SPECT.<br />

High-energy dual-photon imaging is utilized for PET, permitting better resolution<br />

and sensitivity than single-photon techniques. Yet, each method has its<br />

place. For example, sentinel lymph node procedures, employing single-photon<br />

radiotracers, are rapidly becoming the standard <strong>of</strong> care in breast cancer lymph<br />

node staging (4). For all practical purposes modern PET scanning is now<br />

accomplished as part <strong>of</strong> a PET-CT device, which fuses the functional imaging <strong>of</strong><br />

PET with the anatomic detail <strong>of</strong> CT.<br />

In 1984, Beaney and coworkers reported (5) on the use <strong>of</strong> [ 15 O]H 2 O PET<br />

to study blood flow, the oxygen extraction ratio, and oxygen utilization in breast<br />

cancer. The earliest studies on the use <strong>of</strong> 18 F-fluorodeoxyglucose (FDG) in the<br />

diagnosis <strong>of</strong> breast cancer were conducted in 1989 (6,7). Figure 1 is an example<br />

<strong>of</strong> 18 F-FDG PET uptake in a primary breast mass and axillary lymph nodes.


Molecular Imaging in Individualized <strong>Cancer</strong> Management 293<br />

Figure 1 (A) Axial CT, (B) attenuation corrected PET, and (C) fused image in a 50-year-old<br />

female with a 4-cm primary right breast cancer (arrow) and malignant axillary lymph<br />

nodes (arrowhead ). Abbreviations: CT, computed tomography; PET, positron emission<br />

tomography.<br />

Fluorodeoxyglucose is actively transported into the cell, where it is irreversibly<br />

phosphorylated by hexokinase, trapping it within the cell, where it does<br />

not proceed further down the glycolytic pathway. There is a correlation between<br />

FDG uptake and glucose transporter-1 (Glut-1) expression, the mitotic activity<br />

index, percent necrosis (likely reflecting the proliferation rate), tumor cells/<br />

volume, hexokinase-1 expression, the number <strong>of</strong> lymphocytes (but not macrophages),<br />

microvessel density, and possibly the expression <strong>of</strong> the p53 gene (8–12).<br />

PET is currently <strong>of</strong> limited value in routine screening for breast cancer (13)<br />

and cannot take the place <strong>of</strong> fine sectioning and immunohistochemical lymph<br />

node evaluation, but it has an important role to play in a select group <strong>of</strong> problematic<br />

diagnostic situations and in staging patients with high-risk disease (14–17).<br />

PET can provide important information by detecting locoregional lymph node<br />

involvement as well as distant metastases. Because <strong>of</strong> its high positive predictive<br />

value, FDG PET may obviate a sentinel lymph node procedure and axillary<br />

dissection.<br />

The use <strong>of</strong> neoadjuvant chemotherapy has increased the rate <strong>of</strong> breast<br />

conservation surgery (18). Clinical response does not necessarily correlate with<br />

pathologic response. FDG PET has been shown to aid in monitoring response to<br />

chemotherapy and has prognostic benefit (19). Nonresponders and those<br />

developing progressive disease or distant metastases can be identified earlier,<br />

and this information may prove useful in changing therapies and avoiding side<br />

effects <strong>of</strong> chemotherapy that is not effective. Baseline PET combined with PET<br />

after the first course <strong>of</strong> chemotherapy is highly accurate in this regard (16). A<br />

meta-analysis <strong>of</strong> the FDG PET literature (20) reports an overall 81% sensitivity,<br />

96% specificity, and 92% accuracy for monitoring response to therapy. Elevated<br />

FDG uptake was found to be an independent predictor <strong>of</strong> worse prognosis and<br />

decreased relapse-free survival in breast cancer (10). Figure 2 illustrates a dramatic<br />

response <strong>of</strong> metastatic breast cancer to fulvestrant (Faslodex) therapy.<br />

FDG PET has been shown to locate unsuspected metastases with high<br />

sensitivity and specificity (21,22). Detection <strong>of</strong> early recurrence may have


294 Schuster and Martin<br />

Figure 2 (A) Anterior view from a maximal intensity projection <strong>of</strong> an FDG-PET scan<br />

showing extensive breast cancer pleural implants in the left chest and (B) after one dose <strong>of</strong><br />

fulvestrant (Faslodex) after which the implants resolved. Note normal cardiac (arrow) and<br />

renal (arrowhead) uptake is better seen in B than in A. The patient was status post<br />

unrelated remote right nephrectomy. Abbreviations: FDG,<br />

18 F-fluorodeoxyglucose;<br />

PET, positron emission tomography.<br />

important survival benefits, prompting the use <strong>of</strong> new therapies as well as curative<br />

or palliative surgery. FDG PET is considered <strong>of</strong> great efficacy in the evaluation <strong>of</strong><br />

patients with suspected recurrent breast cancer, especially to differentiate true<br />

recurrence from postsurgical and radiation sequelae. The ability <strong>of</strong> PET to detect<br />

suspected clinical recurrence and in patients with elevated tumor markers has been<br />

demonstrated in a number <strong>of</strong> studies (23–32).<br />

MRI: CURRENT CONCEPTS AND APPLICATIONS<br />

In 1973, Paul Lauterbur discovered that introduction <strong>of</strong> magnetic gradients<br />

superimposed over a magnetic field makes it possible to create two-dimensional<br />

images <strong>of</strong> structures, with the first experiments showing the ability to differentiate<br />

between tubes filled with regular or heavy water (33). No other imaging<br />

method can achieve this differentiation. This work was the foundation for a<br />

revolutionary tissue-imaging technique that generates images on the basis <strong>of</strong><br />

molecular differences between or within tissues.<br />

To understand the added value <strong>of</strong> MRI to clinical sciences, it is useful to<br />

compare it with CT. CT can produce images <strong>of</strong> high spatial resolution, which are<br />

interpreted on the basis <strong>of</strong> anatomy. A distinct feature <strong>of</strong> MRI is that anatomically<br />

correct images are produced, but these images are based on molecular<br />

events that produce much higher s<strong>of</strong>t tissue contrast between tissues and between


Molecular Imaging in Individualized <strong>Cancer</strong> Management 295<br />

Figure 3 Liver cross-sectional imaging showing multiple hypervascular breast metastases.<br />

(A) On contrast-enhanced CT, the liver lesions are not visualized. (B) On contrastenhanced<br />

MRI, the lesions are conspicuous with innumerable tumors scattered throughout<br />

the liver (6 are shown: B, black arrows). Abbreviations: CT, computed tomography; MRI,<br />

magnetic resonance imaging.<br />

abnormal and normal tissues. It has been shown for most clinically important<br />

applications, including the brain and the solid organs <strong>of</strong> the body, that s<strong>of</strong>t tissue<br />

contrast resolution is more important than spatial resolution. Even on contrastenhanced<br />

CT, the ability to generate s<strong>of</strong>t tissue contrast resolution is less than<br />

that <strong>of</strong> MRI, and this difference may contribute to the relative advantage that<br />

MRI provides for lesion detection (Fig. 3).<br />

Unlike CT, which relies on the degree <strong>of</strong> X-ray absorption to differentiate<br />

tissues, MRI contrast differences are based on molecular interactions between<br />

protons located in water molecules within and between cells, neighboring<br />

molecules that influence mobility, such as charged glycoproteins and membranes,<br />

and species that can alter the magnetic field at the molecular level, such<br />

as iron and lipid. There are essentially an infinite variety <strong>of</strong> formulations <strong>of</strong> MRI<br />

techniques that can exploit these interactions to generate different contrast patterns.<br />

Similar to the concept <strong>of</strong> a pathologist using different histochemical stains<br />

to visualize molecular species, the expert MR imager will utilize an optimized<br />

set <strong>of</strong> MRI sequences to visualize the characteristics that may be unique to<br />

specific pathologies. Though the initial applications <strong>of</strong> MRI had been for<br />

imaging diseases <strong>of</strong> the brain, MRI is now able to image faster and with concomitant<br />

improvement in image quality. This has been critical for improving the<br />

capacity <strong>of</strong> MRI for rapidly imaging the chest, abdomen, and pelvis, where<br />

motion from respiration or cardiac contractions may deteriorate images (34,35).<br />

<strong>Breast</strong> MRI (Fig. 4) is a recent additional imaging method for the evaluation<br />

<strong>of</strong> primary breast carcinoma and is useful for early detection and staging<br />

(36–40). MRI has been found to provide maximal sensitivity, as compared to<br />

traditional mammography and ultrasound, but with comparatively lower specificity.<br />

The fundamental components to the current state <strong>of</strong> the art <strong>of</strong> breast MRI<br />

is to use the combination <strong>of</strong> time-resolved gadolinium uptake analysis in combination<br />

with the evaluation <strong>of</strong> the contours <strong>of</strong> a mass using high-contrast and


296 Schuster and Martin<br />

Figure 4 (A) Right breast carcinoma shown on sagittal plane dynamically enhanced<br />

gadolinium MRI at the indicated time intervals. Subsequently, (B) a region <strong>of</strong> interest is<br />

applied within the tumor, and a time-intensity curve is generated. (C) A lymph node<br />

showing a similar enhancement pattern (arrow) is shown on a more lateral slice through<br />

the right axilla on a representative 3-min sagittal image. (D) The enhancing primary breast<br />

carcinoma and the metastatic lymph node in the right axilla can be seen on an axial MRI<br />

image <strong>of</strong> both breasts (arrows), which correlates with findings on (E) PET-CT (arrows),<br />

confirming the presence <strong>of</strong> metastasis in the node. Note that the position <strong>of</strong> the breast is<br />

different on the MRI compared with the PET-CT due to prone MRI versus supine PET<br />

positioning. Abbreviations: MRI, magnetic resonance imaging; PET, positron emission<br />

tomography; CT, computed tomography.<br />

high-resolution imaging. MRI techniques have benefited from improvements in<br />

surface coils specifically dedicated to imaging both breasts simultaneously.<br />

The most important sequence used for routine breast MRI is the threedimensional<br />

gradient echo (3D GRE) technique. Uniform and complete suppression<br />

<strong>of</strong> the breast lipid signal has been a key area <strong>of</strong> focus for improvement<br />

<strong>of</strong> these sequences. With time-resolved imaging after an MRI contrast injection<br />

(Fig. 4), malignant focal breast carcinomas typically show enhancement that


Molecular Imaging in Individualized <strong>Cancer</strong> Management 297<br />

remains constant over several minutes or may diminish. Benign tumors such as<br />

fibromas may show differential contrast uptake, with progressive increase in<br />

contrast accumulation within the benign mass. However, the specificity for these<br />

differentiating features is relatively moderate. A more recent approach has<br />

incorporated lessons learned previously from conventional X-ray mammography;<br />

the margins <strong>of</strong> the tumor may show patterns specific to malignancy. Data<br />

are developing to support that the analysis <strong>of</strong> tumor margin morphology will<br />

improve MRI specificity for the diagnosis <strong>of</strong> breast carcinoma and separate this<br />

condition from benign tumors. Here, further development <strong>of</strong> newer MRI strategies,<br />

such as perfusion studies or spectroscopy, or a possible combination with<br />

metabolic imaging provided by PET, may help improve this apparent current<br />

limitation.<br />

Lymphatic spread may result in local lymph node involvement. Findings<br />

on MRI may be relatively specific when the lymph nodes are found to be over<br />

1 cm in diameter and show contrast enhancement that is similar in time course<br />

and intensity to those seen in the primary breast mass. However, for smaller<br />

lymph nodes the specificity may be significantly lower. Combined imaging with<br />

FDG-PET and MRI may help aid the specificity <strong>of</strong> the MRI study and the<br />

identification <strong>of</strong> the most concerning lymph nodes (Fig. 4).<br />

Currently accepted applications for MRI include screening for high-risk<br />

women, evaluation <strong>of</strong> dense fibroglandular breast tissue, and surveillance after<br />

surgical alteration or breast implant placement (36–40). Generalized use <strong>of</strong> MRI<br />

for evaluation <strong>of</strong> all women in the screening age group has not been widely<br />

accepted.<br />

With regard to hematogenous tumor spread, the liver is the organ most<br />

commonly involved in metastatic diseases, usually from breast, colon, pancreas,<br />

kidney, and carcinoid cancers, or melanoma. However, most liver tumors are<br />

benign, and benign liver lesions are common, including cysts, bile duct hamartomas,<br />

hemangiomas, adenomas, and focal nodular hyperplasia. The specificity <strong>of</strong><br />

a noninvasive test is the litmus test for utility in tumor evaluation. We have shown<br />

that although MRI is slightly more sensitive than CT for tumor detection, the most<br />

important distinction is that MRI is highly specific, approaching 98%, while CT is<br />

significantly less specific, approaching 80% (41–44). The ability to image rapidly<br />

and include the evaluation <strong>of</strong> the entire body is undergoing development. The<br />

potential for assessing metastatic burden will be a primary potential role for this<br />

capability.<br />

FUTURE DEVELOPMENTS<br />

The resolution <strong>of</strong> PET scanners is currently limited in comparison with anatomic<br />

imaging, but continued technological improvements are expected<br />

(45,46). Sensitivity for initial diagnosis will likely improve as device resolution<br />

advances and as special focused field PET-mammography imaging systems are<br />

developed (47–49). Small field-<strong>of</strong>-view cameras not only <strong>of</strong>fer better sentinel


298 Schuster and Martin<br />

lymph node and primary tumor detection because <strong>of</strong> increased spatial resolution<br />

but can also be mounted on maneuverable arms and used in an intraoperative<br />

setting, helping the surgeon realize a tumor-free margin (50).<br />

MR spectroscopic analysis has the potential to elucidate pathophysiological<br />

processes and may become useful for the assessment <strong>of</strong> primary or<br />

metastatic neoplasms, or the assessment <strong>of</strong> tumor response to therapy. MRI<br />

strategies are providing new diagnostic methods <strong>of</strong> evaluating dynamic pathophysiological<br />

processes using time-resolved imaging. These methods include the<br />

use <strong>of</strong> an injected contrast agent that can be visualized as it perfuses and then<br />

redistributes within the tissues, much as a radiotracer is sometimes used for<br />

nuclear imaging techniques (51,52). The advantages <strong>of</strong> MRI for this application<br />

include the capacity for high temporal and spatial resolution and the ability to<br />

combine this component <strong>of</strong> an examination with other elements <strong>of</strong> MRI to<br />

provide comprehensive evaluation during a single study. This approach is being<br />

used to detect and characterize diseases, including neoplasms, inflammation and<br />

infection, and fibrotic processes (53–57).<br />

Apoptosis, or programmed cell death, is an elementary process <strong>of</strong> life that<br />

maintains homeostasis. Apoptosis is physiologic and is different from necrosis,<br />

which is pathologic, <strong>of</strong>ten eliciting an inflammatory reaction. Successful chemotherapy<br />

increases apoptosis and enhances the uptake <strong>of</strong> annexin in the tumor.<br />

As the tumor becomes necrotic, the uptake <strong>of</strong> annexin reduces. PET and SPECT<br />

imaging with radiolabeled annexin have been used to assess apoptosis (58–61).<br />

Molecular process that occur within breast cancer cells can be imaged by<br />

using a radiotracer based on amino acid metabolism, such as methionine, which<br />

can serve as an indirect measure <strong>of</strong> protein synthesis, or an amine, such as<br />

choline, which can reflect membrane biosynthesis.<br />

11 C-methionine and<br />

11 C-choline have been used in assessing breast cancer and metastases (62,63).<br />

DNA synthesis within breast cancer has been examined with thymidine analogues<br />

such as 3 0 -deoxy-3 0 -[ 18 F]fluorothymidine (FLT), which has been<br />

employed to image intratumoral proliferation (64,65). [ 15 O]H 2 O PET can also be<br />

useful for tumor perfusion (65). These radiotracers can serve to characterize<br />

tumor as well as measure therapeutic response.<br />

Estrogen and progestin receptors <strong>of</strong>fer another avenue to image breast<br />

cancer. Estrogen receptors (ERs) have been successfully targeted using 16a-[ 18 F]<br />

fluoro-17b-estradiol (FES), and early clinical studies have shown them to be as<br />

sensitive as FDG imaging and probably more specific (66). In one study <strong>of</strong><br />

patients with ER-positive breast cancer, the patients also underwent FES PET.<br />

PET was able to predict global response to hormonal therapy in human epidermal<br />

growth factor receptor 2 (HER2)-negative patients better than pathology<br />

(24% vs. 47%) (67,68). Radiolabeled chemotherapeutic agents such as [ 18 F]<br />

fluorotamoxifen may be useful to detect ER-positive metastases and can also<br />

help in following response to treatment (69). 18 F-labeled paclitaxel has also been<br />

shown in a mouse model with a breast carcinoma xenograft to predict tumor<br />

response to paclitaxel chemotherapy (70).


Molecular Imaging in Individualized <strong>Cancer</strong> Management 299<br />

Peptides that localize to receptors are small-sized molecules and clear<br />

rapidly via the kidney, making these peptides attractive imaging probes. Peptides,<br />

such as bombesin (BN), labeled with single-photon and dual-photon<br />

radionuclides are being studied both for imaging and for therapeutic purposes<br />

(71). Gastrin-releasing peptide (GRP) receptors, a subtype <strong>of</strong> BN receptors, are<br />

overexpressed in breast carcinoma. Octreoscan, a single-photon radiotracer<br />

somatostatin analogue, has also been used to study breast cancer (72).<br />

Several gene therapy approaches have been considered for the management<br />

<strong>of</strong> breast cancer, such as ablation <strong>of</strong> oncogenic products, restoration <strong>of</strong> ER<br />

expression, alteration <strong>of</strong> genes that are involved in apoptosis, and activation <strong>of</strong><br />

tumor suppressor genes. Reporter genes have been used to study promoter or<br />

regulatory elements involved in gene expression conventionally studied by tissue<br />

biopsy and immunohistochemistry. Now noninvasive in vivo methods are<br />

available in the form <strong>of</strong> radionuclide techniques or optical imaging. The ability<br />

to image deeper structures in larger animal models and in humans is the main<br />

advantage <strong>of</strong> radionuclide technique over optical imaging. For example, 8-[ 18 F]<br />

Fluoropenciclovir has been studied as a reporter probe for herpes simplex virus<br />

type 1 thymidine kinase (HSV1-tk) gene expression (73). Other possibilities still<br />

in very early development include specially designed, small radiolabelled antisense<br />

oligodeoxynucleotides that may help image gene expression (20). PET<br />

imaging with an appropriate PET reporter gene and probe combination may<br />

become a powerful tool for molecular imaging <strong>of</strong> human gene expression (74).<br />

HER2/neu overexpression in breast cancer is associated with poor response<br />

to chemotherapy and an unfavorable prognosis. <strong>Ph</strong>ase I clinical trials <strong>of</strong> E1A<br />

gene therapy targeting HER2/neu-overexpressed breast cancer has shown<br />

increased apoptosis (75). A novel PET radiotracer, (68)Ga-DOTA-F(ab 0 )(2)-<br />

herceptin, has been developed, which has the potential to follow tumor<br />

response to the heat shock protein 90 (Hsp90) inhibitor 17-allylamino-17-<br />

demethoxygeldanamycin (17AAG) (76). In an animal model, downregulation <strong>of</strong><br />

HER2 occurs independently <strong>of</strong> glycolytic response as measured by 18 F-FDG-<br />

PET. This method could be used to rapidly screen tumor response to a variety <strong>of</strong><br />

HER2 inhibitors and choose the correct agent within hours rather than wait for<br />

weeks or months to determine if there is a response. Multidrug resistance as<br />

manifested by MDR1 P-glycoprotein transport activity also has the potential to<br />

be studied with SPECT and PET radiotracers (77).<br />

Angiogenesis, the outgrowth <strong>of</strong> new capillaries, is an important component<br />

in tumor formation and spread. Imaging <strong>of</strong> angiogenesis for diagnosis and following<br />

therapy can be performed with PET and SPECT radiotracers targeting<br />

integrin, VEGF, and other biomarkers (78). For example, (64)Cu-DOTA-VEGF<br />

(121) PET has been successful in small animals to specifically image in vivo<br />

expression <strong>of</strong> VEGF (79).<br />

Tumor hypoxia may result in resistance to therapy. [ 18 F]Fluoromisonidazole-3-fluoro-1-(2<br />

0 -nitro-1 0 -imidazolyl)-2-propanol ([ 18 F]MISO) and Cu-diacetylbis(N4-methylthiosemicarbazone)<br />

(Cu-ATSM) radiotracers for PET and blood


300 Schuster and Martin<br />

oxygen level–dependent (BOLD) MRI are some <strong>of</strong> the techniques being utilized<br />

for in vivo imaging but require further study (80,81).<br />

Finally, nanoparticles technology may be utilized to achieve diagnostic<br />

imaging with both PET and MRI and enable combined modality radionuclide<br />

therapy with intracellular drug delivery (82). The feasibility <strong>of</strong> magnetic<br />

frequency–activated thermal necrosis with antibody-linked iron oxide nanoparticles<br />

conjugated with the single-photon radiotracer Indium-111 has been<br />

recently demonstrated in breast cancer xenografts (83).<br />

An understanding <strong>of</strong> pathology, together with the capabilities <strong>of</strong> MRI,<br />

facilitates the use <strong>of</strong> a particular MRI feature for the evaluation <strong>of</strong> a range <strong>of</strong><br />

diseases that affect different organ systems. The application <strong>of</strong> MRI timeresolved<br />

perfusion imaging for tumor ranges from primary diagnosis to the<br />

determination <strong>of</strong> the degree <strong>of</strong> tumor neoangiogenesis. The ability to quantify<br />

tumor angiogenesis may provide a useful surrogate marker <strong>of</strong> tumor responsiveness<br />

prior to initiating therapy or indicate tumor response even prior to any<br />

change in the overall volume <strong>of</strong> tumor tissue (Fig. 4). MRI may be useful for the<br />

specific evaluation <strong>of</strong> chemotherapeutic antiangiogenic agents such as anti-<br />

VEGF (84,85). The possibility <strong>of</strong> using MRI molecular imaging capabilities for<br />

directing therapy is being explored, including the detection <strong>of</strong> reporter gene<br />

transcription, the use <strong>of</strong> MRI tissue thermometry to guide therapeutic ultrasound<br />

for noninvasive tumor ablation, and the use <strong>of</strong> functional brain MRI to plan the<br />

surgical resection <strong>of</strong> tumors, including metastases to the brain (51,52,86–91).<br />

Most importantly, in a manner analogous to the fusion <strong>of</strong> PET and CT<br />

scanning, the fusion <strong>of</strong> PET and MRI scanning holds even greater promise in<br />

providing a one-stop shop for molecular function and exquisite anatomic detail<br />

but with a markedly reduced radiation dose. There have been a number <strong>of</strong> studies<br />

examining the utility <strong>of</strong> MRI and FDG PET. For example, Walter et al. studied<br />

42 lesions in 40 patients preoperatively (92). The sensitivity <strong>of</strong> FDG PET was<br />

63%, and the specificity was 91%; the sensitivity <strong>of</strong> MRI was 89%, and the<br />

specificity was 74%. Thus, both modalities can play complementary roles, with<br />

overall improved accuracy when used in conjunction. Other papers examining<br />

the complementary strengths and weaknesses <strong>of</strong> both modalities have emerged<br />

(93–95). A recent intriguing study in a small group <strong>of</strong> patients examines the<br />

utility <strong>of</strong> MRI with ultrasmall superparamagnetic iron oxide particles in conjunction<br />

with PET, achieving 100% sensitivity and specificity on a per-patient<br />

basis for local lymph node staging when both modalities are employed (96).<br />

Further technical development is anticipated to facilitate combining MRI<br />

and PET (97). We have already started accumulating evidence <strong>of</strong> the advantages<br />

in this pairing <strong>of</strong> technologies, as compared with using the combination <strong>of</strong><br />

CT and PET. Figure 5 is an example from our institution <strong>of</strong> how PET and MRI<br />

data can be combined utilizing special s<strong>of</strong>tware registration. While the above<br />

PET-MRI studies have been performed on separate devices, necessitating<br />

moving and scheduling the patients between two physically separate units, a<br />

combined PET-MRI scanner in a single gantry is now a reality (45,46,97,98).


Molecular Imaging in Individualized <strong>Cancer</strong> Management 301<br />

Figure 5 <strong>Breast</strong> cancer patient (from Fig. 4) with axial MRI, PET, and CT images through<br />

the pelvis. (A) Shows the MRI and PET images before and after deformable registration <strong>of</strong><br />

the PET to the MRI. (B) The original CT (upper left) andPET-CT(lower left), compared<br />

with the MRI (upper right) and deformable registered PET-MRI (lower right). Increased<br />

FDG activity in the perirectal space on PET-CT (arrow) is <strong>of</strong> uncertain origin; however, the<br />

activity from the PET scan clearly corresponds to a benign fibroid on the MRI, which is not<br />

visible on CT. Benign fibroids may occasionally take up FDG in premenopausal women.<br />

The fusion <strong>of</strong> PET and MRI increased diagnostic accuracy in this patient. Abbreviations:<br />

MRI, magnetic resonance imaging; PET, positron emission tomography; CT, computed<br />

tomography; FDG, 18 F-fluorodeoxyglucose.<br />

OUR VISION<br />

Molecular imaging will continue to have an evolving role in diagnosis, staging,<br />

personalization <strong>of</strong> therapy, monitoring response to therapy, and detecting<br />

recurrence <strong>of</strong> breast carcinoma, both local and widespread. A patient will arrive<br />

in the imaging suite and have the entire disease process interrogated on a specially<br />

modified PET-MRI system, which will utilize dynamic MRI techniques<br />

and MR spectroscopy as well as novel PET radiotracers. Complex processes<br />

such as glycolysis, tumor perfusion, cell proliferation, protein synthesis, genetic<br />

expression, angiogenesis, and various other pr<strong>of</strong>iling will be examined. There is<br />

potential for additional information and improvements over current invasive<br />

tissue sampling and histopathology, which, for example, may not be truly representative<br />

<strong>of</strong> the heterogeneity present within a primary tumor or metastases.<br />

This process will not simply serve as a baseline to monitor therapy but<br />

will help individualize therapy since all the parameters can be entered into a<br />

matrix and therapy chosen on the basis <strong>of</strong> data derived from a cumulative<br />

outcomes database, including surgical, chemotherapeutic, radiotherapy, and<br />

genetic therapy options. Drug-target matching and optimal dosing can be<br />

assessed utilizing the powerful molecular techniques described above. These<br />

techniques will also facilitate targeted biopsies, which will improve genomic<br />

and proteomic pr<strong>of</strong>iling and permit a more tailored radiotherapy. Designer<br />

therapeutic and imaging agents may also be synthesized on the basis <strong>of</strong> a<br />

patient’s particular pr<strong>of</strong>ile.


302 Schuster and Martin<br />

These same methods may then be used to quickly determine treatment<br />

response as measured by glycolysis, amino acid transport, tumor perfusion,<br />

proliferation, and apoptosis, among other metabolic indices, and also monitor for<br />

resistant or mutating clones and for recurrence. Chemotherapeutic and genetic<br />

therapies may then be altered or doses optimized on the basis <strong>of</strong> rapid molecular<br />

imaging feedback rather than maximizing therapy based only on toxicity effects.<br />

The possibility <strong>of</strong> therapy using nanoparticle delivery <strong>of</strong> killing doses <strong>of</strong><br />

radiotherapy combined with chemotherapy may also be possible. The same<br />

nanoparticles could also carry MR and PET imaging tracers.<br />

The concept <strong>of</strong> the “one magic bullet” for the generic enemy <strong>of</strong> cancer has<br />

fallen by the wayside. A host <strong>of</strong> ordinary tools optimized for a particular patient<br />

and his or her own cancer can be combined in a unique fashion with the help <strong>of</strong><br />

the “magic window” afforded by molecular imaging. Though defeating the<br />

cancer is a l<strong>of</strong>ty goal, these tools may also enable patients to live with cancer and<br />

manage the entire disease process holistically in cooperation with their medical<br />

caregivers. The true cost-benefit ratios <strong>of</strong> therapies can then be ascertained. We<br />

believe the technology already exists, but further research, social, and financial<br />

investment, as well as societal vision is required for this dream to flower into<br />

everyday reality.<br />

REFERENCES<br />

1. Basilion JP. Current and future technologies for breast cancer imaging. <strong>Breast</strong><br />

<strong>Cancer</strong> Res 2001; 3(1):14–16.<br />

2. Trecate G, Vergnaghi D, Manoukian S, et al. MRI in the early detection <strong>of</strong> breast<br />

cancer in women with high genetic risk. Tumori 2006; 92(6):517–523.<br />

3. Bartella L, Smith CS, Dershaw DD, et al. Imaging breast cancer. Radiol Clin North<br />

Am 2007; 45(1):45–67.<br />

4. Krynyckyi BR, Kim CK, Goyenechea MR, et al. Clinical breast lymphoscintigraphy:<br />

optimal techniques for performing studies, image atlas, and analysis <strong>of</strong> images.<br />

Radiographics 2004; 24(1):121–145 (discussion 139–145).<br />

5. Beaney RP, Lammertsma AA, Jones T, et al. Positron emission tomography for invivo<br />

measurement <strong>of</strong> regional blood flow, oxygen utilisation, and blood volume in<br />

patients with breast carcinoma. Lancet 1984; 1(8369):131–134.<br />

6. Minn H, Soini I. [18F]fluorodeoxyglucose scintigraphy in diagnosis and follow up <strong>of</strong><br />

treatment in advanced breast cancer. Eur J Nucl Med 1989; 15(2):61–66.<br />

7. Kubota K, Matsuzawa T, Amemiya A, et al. Imaging <strong>of</strong> breast cancer with [18F]<br />

fluorodeoxyglucose and positron emission tomography. J Comput Assist Tomogr<br />

1989; 13(6):1097–1098.<br />

8. Bos R, van Der Hoeven JJ, van Der Wall E, et al. Biologic correlates <strong>of</strong> (18)<br />

fluorodeoxyglucose uptake in human breast cancer measured by positron emission<br />

tomography. J Clin Oncol 2002; 20(2):379–387.<br />

9. Brown RS, Wahl RL. Overexpression <strong>of</strong> Glut-1 glucose transporter in human breast<br />

cancer. An immunohistochemical study. <strong>Cancer</strong> 1993; 72(10):2979–2985.


Molecular Imaging in Individualized <strong>Cancer</strong> Management 303<br />

10. Oshida M, Uno K, Suzuki M, et al. Predicting the prognoses <strong>of</strong> breast carcinoma<br />

patients with positron emission tomography using 2-deoxy-2-fluoro[18F]-D-glucose.<br />

<strong>Cancer</strong> 1998; 82(11):2227–2234.<br />

11. Crippa F, Seregni E, Agresti R, et al. Association between [18F]fluorodeoxyglucose<br />

uptake and postoperative histopathology, hormone receptor status, thymidine labelling<br />

index and p53 in primary breast cancer: a preliminary observation. Eur J Nucl<br />

Med 1998; 25(10):1429–1434.<br />

12. Smith TA, Sharma RI, Thompson AM, et al. Tumor 18F-FDG incorporation is<br />

enhanced by attenuation <strong>of</strong> P53 function in breast cancer cells in vitro. J Nucl Med<br />

2006; 47(9):1525–1530.<br />

13. Samson DJ, Flamm CR, Pisano ED, et al. Should FDG PET be used to decide<br />

whether a patient with an abnormal mammogram or breast finding at physical<br />

examination should undergo biopsy? Acad Radiol 2002; 9(7):773–783.<br />

14. Noh DY, Yun IJ, Kang HS, et al. Detection <strong>of</strong> cancer in augmented breasts by<br />

positron emission tomography. Eur J Surg 1999; 165(9):847–851.<br />

15. Wahl RL, Helvie MA, Chang AE, et al. Detection <strong>of</strong> breast cancer in women after<br />

augmentation mammoplasty using fluorine-18-fluorodeoxyglucose-PET. J Nucl Med<br />

1994; 35(5):872–875.<br />

16. Zimny M, Siggelkow W. Positron emission tomography scanning in gynecologic and<br />

breast cancers. Curr Opin Obstet Gynecol 2003; 15(1):69–75.<br />

17. Czernin J. FDG-PET in breast cancer: a different view <strong>of</strong> its clinical usefulness. Mol<br />

Imaging Biol 2002; 4(1):35–45.<br />

18. Rose C, Dose J, Avril N. Positron emission tomography for the diagnosis <strong>of</strong> breast<br />

cancer. Nucl Med Commun 2002; 23(7):613–618.<br />

19. Kostakoglu L, Goldsmith SJ. 18F-FDG PET evaluation <strong>of</strong> the response to therapy for<br />

lymphoma and for breast, lung, and colorectal carcinoma. J Nucl Med 2003; 44(2):<br />

224–239.<br />

20. Gambhir SS, Czernin J, Schwimmer J, et al. A tabulated summary <strong>of</strong> the FDG PET<br />

literature. J Nucl Med 2001; 42(suppl 5):S1–S93.<br />

21. Avril N, Dose J, Janicke F, et al. Assessment <strong>of</strong> axillary lymph node involvement in<br />

breast cancer patients with positron emission tomography using radiolabeled 2-<br />

(fluorine-18)-fluoro-2-deoxy-D-glucose. J Natl <strong>Cancer</strong> Inst 1996; 88(17):1204–1209.<br />

22. Dose J, Bleckmann C, Bachmann S, et al. Comparison <strong>of</strong> fluorodeoxyglucose positron<br />

emission tomography and “conventional diagnostic procedures” for the detection <strong>of</strong><br />

distant metastases in breast cancer patients. Nucl Med Commun 2002; 23(9):857–864.<br />

23. Kim TS, Moon WK, Lee DS, et al. Fluorodeoxyglucose positron emission tomography<br />

for detection <strong>of</strong> recurrent or metastatic breast cancer. World J Surg 2001; 25(7):<br />

829–834.<br />

24. Eubank WB, Mank<strong>of</strong>f DA, Vesselle HJ, et al. Detection <strong>of</strong> locoregional and distant<br />

recurrences in breast cancer patients by using FDG PET. Radiographics 2002; 22(1):<br />

5–17.<br />

25. Lonneux M, Borbath II, Berliere M, et al. The place <strong>of</strong> whole-body PET FDG for the<br />

diagnosis <strong>of</strong> distant recurrence <strong>of</strong> breast cancer. Clin Positron Imaging 2000; 3(2):<br />

45–49.<br />

26. Siggelkow W, Zimny M, Faridi A, et al. The value <strong>of</strong> positron emission tomography<br />

in the follow-up for breast cancer. Anticancer Res 2003; 23(2C):1859–1867.


304 Schuster and Martin<br />

27. Moon DH, Maddahi J, Silverman DH, et al. Accuracy <strong>of</strong> whole-body fluorine-<br />

18-FDG PET for the detection <strong>of</strong> recurrent or metastatic breast carcinoma. J Nucl<br />

Med 1998; 39(3):431–435.<br />

28. Grahek D, Montravers F, Kerrou K, et al. [18F]FDG in recurrent breast cancer:<br />

diagnostic performances, clinical impact and relevance <strong>of</strong> induced changes in<br />

management. Eur J Nucl Med Mol Imaging 2004; 31(2):179–188.<br />

29. Suarez M, Perez-Castejon MJ, Jimenez A, et al. Early diagnosis <strong>of</strong> recurrent breast<br />

cancer with FDG-PET in patients with progressive elevation <strong>of</strong> serum tumor<br />

markers. Q J Nucl Med 2002; 46(2):113–121.<br />

30. Liu CS, Shen YY, Lin CC, et al. Clinical impact <strong>of</strong> [(18)F]FDG-PET in patients with<br />

suspected recurrent breast cancer based on asymptomatically elevated tumor marker<br />

serum levels: a preliminary report. Jpn J Clin Oncol 2002; 32(7):244–247.<br />

31. Pecking AP, Mechelany-Corone C, Bertrand-Kermorgant F, et al. Detection <strong>of</strong> occult<br />

disease in breast cancer using fluorodeoxyglucose camera-based positron emission<br />

tomography. Clin <strong>Breast</strong> <strong>Cancer</strong> 2001; 2(3):229–234.<br />

32. Maldonado A, Sancho F, Calvo FA, et al. The role <strong>of</strong> FDG-PET in diagnosing<br />

suspected tumor recurrence based on increased levels <strong>of</strong> plasma markers. Clin<br />

Positron Imaging 1998; 1(4):246.<br />

33. Pearson H. Magnetic pioneers net Nobel for putting medicine in the picture. Nature<br />

2003; 425(6958):547.<br />

34. Martin DR, Danrad R, Hussain SM. MR imaging <strong>of</strong> the liver. Radiol Clin North Am<br />

2005; 43(5):861–886, viii.<br />

35. Martin DR, Friel HT, Danrad R, et al. Approach to abdominal imaging at 1.5 Tesla and<br />

optimization at 3 Tesla. Magn Reson Imaging Clin N Am 2005; 13(2):241–254, v–vi.<br />

36. Deurloo EE, Peterse JL, Rutgers EJ, et al. Additional breast lesions in patients eligible<br />

for breast-conserving therapy by MRI: impact on preoperative management and<br />

potential benefit <strong>of</strong> computerised analysis. Eur J <strong>Cancer</strong> 2005; 41(10):1393–1401.<br />

37. Lalonde L, David J, Trop I. Magnetic resonance imaging <strong>of</strong> the breast: current<br />

indications. Can Assoc Radiol J 2005; 56(5):301–308.<br />

38. Lehman CD, Blume JD, Weatherall P, et al. Screening women at high risk for breast<br />

cancer with mammography and magnetic resonance imaging. <strong>Cancer</strong> 2005; 103(9):<br />

1898–1905.<br />

39. Liberman L. <strong>Breast</strong> MR imaging in assessing extent <strong>of</strong> disease. Magn Reson Imaging<br />

Clin N Am 2006; 14(3):339–349, vi.<br />

40. Warner E, Causer PA. MRI surveillance for hereditary breast-cancer risk. Lancet<br />

2005; 365(9473):1747–1749.<br />

41. Eisenberg LB, Warshauer DM, Woosley JT, et al. CT and MRI <strong>of</strong> hepatic focal<br />

nodular hyperplasia with peripheral steatosis. J Comput Assist Tomogr 1995; 19(3):<br />

498–500.<br />

42. Martin DR, Semelka RC. Imaging <strong>of</strong> benign and malignant focal liver lesions. Magn<br />

Reson Imaging Clin N Am 2001; 9(4):785–802, vi–vii.<br />

43. Semelka RC, Martin DR, Balci C, et al. Focal liver lesions: comparison <strong>of</strong> dualphase<br />

CT and multisequence multiplanar MR imaging including dynamic gadolinium<br />

enhancement. J Magn Reson Imaging 2001; 13(3):397–401.<br />

44. Lee VS, Lavelle MT, Krinsky GA, et al. Volumetric MR imaging <strong>of</strong> the liver and<br />

applications. Magn Reson Imaging ClinN Am 2001; 9(4):697–716, v–vi.<br />

45. Lucignani G. Time-<strong>of</strong>-flight PET and PET/MRI: recurrent dreams or actual realities?<br />

Eur J Nucl Med Mol Imaging 2006; 33(8):969–971.


Molecular Imaging in Individualized <strong>Cancer</strong> Management 305<br />

46. Cherry SR. The 2006 Henry N. Wagner Lecture: Of mice and men (and positrons)–<br />

advances in PET imaging technology. J Nucl Med 2006; 47(11):1735–1745.<br />

47. Raylman RR, Majewski S, Weisenberger AG, et al. Positron emission mammography-guided<br />

breast biopsy. J Nucl Med 2001; 42(6):960–966.<br />

48. Murthy K, Aznar M, Thompson CJ, et al. Results <strong>of</strong> preliminary clinical trials <strong>of</strong> the<br />

positron emission mammography system PEM-I: a dedicated breast imaging system<br />

producing glucose metabolic images using FDG. J Nucl Med 2000; 41(11):1851–1858.<br />

49. Adler LP, Weinberg IN, Bradbury MS, et al. Method for combined FDG-PET and<br />

radiographic imaging <strong>of</strong> primary breast cancers. <strong>Breast</strong> J 2003; 9(3):163–166.<br />

50. Aarsvold JN, Mintzer RA, Yasillo NJ, et al. A miniature gamma camera. Ann N Y<br />

Acad Sci 1994; 720:192–205.<br />

51. Curran SD, Muellner AU, Schwartz LH. Imaging response assessment in oncology.<br />

<strong>Cancer</strong> Imaging 2006; 6:S126–S130.<br />

52. Kiessling F, Morgenstern B, Zhang C. Contrast agents and applications to assess tumor<br />

angiogenesis in vivo by magnetic resonance imaging. Curr Med Chem 2007; 14(1):<br />

77–91.<br />

53. Ansari M, Araoz PA, Gerard SK, et al. Comparison <strong>of</strong> late enhancement cardiovascular<br />

magnetic resonance and thallium SPECT in patients with coronary disease<br />

and left ventricular dysfunction. J Cardiovasc Magn Reson 2004; 6(2):549–556.<br />

54. Goldfarb JW, Shinnar M. Free-breathing delayed hyperenhanced imaging <strong>of</strong> the<br />

myocardium: a clinical application <strong>of</strong> real-time navigator echo imaging. J Magn<br />

Reson Imaging 2006; 24(1):66–71.<br />

55. Saeed M, Weber O, Lee R, et al. Discrimination <strong>of</strong> myocardial acute and chronic<br />

(scar) infarctions on delayed contrast enhanced magnetic resonance imaging with<br />

intravascular magnetic resonance contrast media. J Am Coll Cardiol 21 2006; 48(10):<br />

1961–1968.<br />

56. Schneider G, Ahlhelm F, Seidel R, et al. Contrast-enhanced cardiovascular magnetic<br />

resonance imaging. Top Magn Reson Imaging 2003; 14(5):386–402.<br />

57. Martin DR, Semelka RC. Magnetic resonance imaging <strong>of</strong> the liver: review <strong>of</strong><br />

techniques and approach to common diseases. Semin Ultrasound CT MR 2005; 26(3):<br />

116–131.<br />

58. Yagle KJ, Eary JF, Tait JF, et al. Evaluation <strong>of</strong> 18F-annexin V as a PET imaging<br />

agent in an animal model <strong>of</strong> apoptosis. J Nucl Med 2005; 46(4):658–666.<br />

59. Murakami Y, Takamatsu H, Taki J, et al. 18F-labelled annexin V: a PET tracer for<br />

apoptosis imaging. Eur J Nucl Med Mol Imaging 2003.<br />

60. Dekker B, Keen H, Shaw D, et al. Functional comparison <strong>of</strong> annexin V analogues<br />

labeled indirectly and directly with iodine-124. Nucl Med Biol 2005; 32(4):403–413.<br />

61. Keen HG, Dekker BA, Disley L, et al. Imaging apoptosis in vivo using 124I-annexin<br />

V and PET. Nucl Med Biol 2005; 32(4):395–402.<br />

62. Inoue T, Kim EE, Wong FC, et al. Comparison <strong>of</strong> fluorine-18-fluorodeoxyglucose<br />

and carbon-11-methionine PET in detection <strong>of</strong> malignant tumors. J Nucl Med 1996;<br />

37(9):1472–1476.<br />

63. Leskinen-Kallio S, Nagren K, Lehikoinen P, et al. Carbon-11-methionine and PET is<br />

an effective method to image head and neck cancer. J Nucl Med 1992; 33(5):691–695.<br />

64. Smyczek-Gargya B, Fersis N, Dittmann H, et al. PET with [18F]fluorothymidine for<br />

imaging <strong>of</strong> primary breast cancer: a pilot study. Eur J Nucl Med Mol Imaging 2004;<br />

31(5):720–724.


306 Schuster and Martin<br />

65. Buck AK, Schirrmeister H, Mattfeldt T, et al. Biological characterisation <strong>of</strong> breast<br />

cancer by means <strong>of</strong> PET. Eur J Nucl Med Mol Imaging 2004; 31(suppl 1):S80–S87.<br />

66. Dehdashti F, Mortimer JE, Siegel BA, et al. Positron tomographic assessment <strong>of</strong><br />

estrogen receptors in breast cancer: comparison with FDG-PET and in vitro receptor<br />

assays. J Nucl Med 1995; 36(10):1766–1774.<br />

67. Linden HM, Stekhova S, Link JM, et al. HER2 expression and uptake <strong>of</strong> 18F-fluoroestradiol<br />

predict response <strong>of</strong> breast cancer to hormonal therapy. J Nucl Med 2004;<br />

45:85–86.<br />

68. Shields AF. Positron emission tomography measurement <strong>of</strong> tumor metabolism and<br />

growth: its expanding role in oncology. Mol Imaging Biol 2006; 8(3):141–150.<br />

69. Inoue T, Kim EE, Wallace S, et al. Preliminary study <strong>of</strong> cardiac accumulation <strong>of</strong> F-18<br />

fluorotamoxifen in patients with breast cancer. Clin Imaging 1997; 21(5):332–336.<br />

70. Hsueh WA, Kesner AL, Gangl<strong>of</strong>f A, et al. Predicting chemotherapy response to<br />

paclitaxel with 18F-Fluoropaclitaxel and PET. J Nucl Med 2006; 47(12):1995–1999.<br />

71. Zhang H, Schuhmacher J, Waser B, et al. DOTA-PESIN, a DOTA-conjugated<br />

bombesin derivative designed for the imaging and targeted radionuclide treatment <strong>of</strong><br />

bombesin receptor-positive tumours. Eur J Nucl Med Mol Imaging 2007; 34(8):<br />

1198–1208.<br />

72. van Eijck CH, Kwekkeboom DJ, Krenning EP. Somatostatin receptors and breast<br />

cancer. Q J Nucl Med 1998; 42(1):18–25.<br />

73. Iyer M, Barrio JR, Namavari M, et al. 8-[18F]Fluoropenciclovir: an improved<br />

reporter probe for imaging HSV1-tk reporter gene expression in vivo using PET.<br />

J Nucl Med 2001; 42(1):96–105.<br />

74. Penuelas I, Boan J, Marti-Climent JM, et al. Positron emission tomography and gene<br />

therapy: basic concepts and experimental approaches for in vivo gene expression<br />

imaging. Mol Imaging Biol 2004; 6(4):225–238.<br />

75. Hung MC, Hortobagyi GN, Ueno NT. Development <strong>of</strong> clinical trial <strong>of</strong> E1A gene<br />

therapy targeting HER-2/neu-overexpressing breast and ovarian cancer. Adv Exp<br />

Med Biol 2000; 465:171–180.<br />

76. Smith-Jones PM, Solit D, Afroze F, et al. Early tumor response to Hsp90 therapy using<br />

HER2 PET: comparison with 18F-FDG PET. J Nucl Med 2006; 47(5):793–796.<br />

77. Sharma V, Prior JL, Belinsky MG, et al. Characterization <strong>of</strong> a 67Ga/68Ga radiopharmaceutical<br />

for SPECT and PET <strong>of</strong> MDR1 P-glycoprotein transport activity in<br />

vivo: validation in multidrug-resistant tumors and at the blood-brain barrier. J Nucl<br />

Med 2005; 46(2):354–364.<br />

78. Choe YS, Lee KH. Targeted in vivo imaging <strong>of</strong> angiogenesis: present status and<br />

perspectives. Curr <strong>Ph</strong>arm Des 2007; 13(1):17–31.<br />

79. Cai W, Chen K, Mohamedali KA, et al. PET <strong>of</strong> vascular endothelial growth factor<br />

receptor expression. J Nucl Med 2006; 47(12):2048–2056.<br />

80. Padhani AR, Krohn KA, Lewis JS, et al. Imaging oxygenation <strong>of</strong> human tumours.<br />

Eur Radiol 2007; 17(4):861–872.<br />

81. O’Donoghue JA, Zanzonico P, Pugachev A, et al. Assessment <strong>of</strong> regional tumor<br />

hypoxia using 18F-fluoromisonidazole and 64Cu(II)-diacetyl-bis(N4-methylthiosemicarbazone)<br />

positron emission tomography: comparative study featuring microPET<br />

imaging, Po2 probe measurement, autoradiography, and fluorescent microscopy in<br />

the R3327-AT and FaDu rat tumor models. Int J Radiat Oncol Biol <strong>Ph</strong>ys 2005; 61(5):<br />

1493–1502.


Molecular Imaging in Individualized <strong>Cancer</strong> Management 307<br />

82. DeNardo SJ, Denardo GL. Targeted radionuclide therapy for solid tumors: an<br />

overview. Int J Radiat Oncol Biol <strong>Ph</strong>ys 2006; 66(suppl 2):S89–S95.<br />

83. Denardo SJ, Denardo GL, Natarajan A, et al. Thermal dosimetry predictive <strong>of</strong><br />

efficacy <strong>of</strong> 111In-ChL6 nanoparticle AMF-induced thermoablative therapy for<br />

human breast cancer in mice. J Nucl Med 2007; 48(3):437–444.<br />

84. Kanematsu M, Semelka RC, Leonardou P, et al. Angiogenesis in hepatocellular<br />

nodules: correlation <strong>of</strong> MR imaging and vascular endothelial growth factor. J Magn<br />

Reson Imaging 2004; 20(3):426–434.<br />

85. Kanematsu M, Semelka RC, Osada S, et al. Magnetic resonance imaging and<br />

expression <strong>of</strong> vascular endothelial growth factor in hepatocellular nodules in cirrhosis<br />

and hepatocellular carcinomas. Top Magn Reson Imaging 2005; 16(1):67–75.<br />

86. Andrew ER. Nuclear magnetic resonance and the brain. Brain Topogr 1992; 5(2):<br />

129–133.<br />

87. Cai W, Rao J, Gambhir SS, et al. How molecular imaging is speeding up antiangiogenic<br />

drug development. Mol <strong>Cancer</strong> Ther 2006; 5(11):2624–2633.<br />

88. Charles-Edwards EM, deSouza NM. Diffusion-weighted magnetic resonance imaging<br />

and its application to cancer. <strong>Cancer</strong> Imaging 2006; 6:135–143.<br />

89. Chenevert TL, Sundgren PC, Ross BD. Diffusion imaging: insight to cell status and<br />

cytoarchitecture. Neuroimaging Clin N Am 2006; 16(4):619–632, viii–ix.<br />

90. Lee JH, Huh YM, Jun YW, et al. Artificially engineered magnetic nanoparticles for<br />

ultra-sensitive molecular imaging. Nat Med 2007; 13(1):95–99.<br />

91. Lee JH, Jun YW, Yeon SI, et al. Dual-mode nanoparticle probes for high-performance<br />

magnetic resonance and fluorescence imaging <strong>of</strong> neuroblastoma. Angew<br />

Chem Int Ed Engl 2006; 45(48):8160–8162.<br />

92. Walter C, Scheidhauer K, Scharl A, et al. Clinical and diagnostic value <strong>of</strong> preoperative<br />

MR mammography and FDG-PET in suspicious breast lesions. Eur Radiol<br />

2003; 13(7):1651–1656.<br />

93. Antoch G, Vogt FM, Freudenberg LS, et al. Whole-body dual-modality PET/CT and<br />

whole-body MRI for tumor staging in oncology. JAMA 2003; 290(24):3199–3206.<br />

94. Schmidt GP, Schoenberg SO, Schmid R, et al. Screening for bone metastases: wholebody<br />

MRI using a 32-channel system versus dual-modality PET-CT. Eur Radiol<br />

2006; 17(4):939–949.<br />

95. Brix G, Henze M, Knopp MV, et al. Comparison <strong>of</strong> pharmacokinetic MRI and [18F]<br />

fluorodeoxyglucose PET in the diagnosis <strong>of</strong> breast cancer: initial experience. Eur<br />

Radiol 2001; 11(10):2058–2070.<br />

96. Stadnik TW, Everaert H, Makkat S, et al. <strong>Breast</strong> imaging. Preoperative breast cancer<br />

staging: comparison <strong>of</strong> USPIO-enhanced MR imaging and 18F-fluorodeoxyglucose<br />

(FDC) positron emission tomography (PET) imaging for axillary lymph node staging:<br />

initial findings. Eur Radiol 2006; 16(10):2153–2160.<br />

97. Schmidt GP, Haug AR, Schoenberg SO, et al. Whole-body MRI and PET-CT in the<br />

management <strong>of</strong> cancer patients. Eur Radiol 2006; 16(6):1216–1225.<br />

98. Catana C, Wu Y, Judenh<strong>of</strong>er MS, et al. Simultaneous acquisition <strong>of</strong> multislice PET<br />

and MR images: initial results with a MR-compatible PET scanner. J Nucl Med<br />

2006; 47(12):1968–1976.


20<br />

<strong>Cancer</strong> Nanotechnology for Molecular<br />

Pr<strong>of</strong>iling and Individualized Therapy<br />

May Dongmei Wang<br />

Departments <strong>of</strong> Biomedical Engineering and Electrical<br />

and Computer Engineering, Georgia Institute <strong>of</strong> Technology<br />

and Emory University, Atlanta, Georgia, U.S.A.<br />

Jonathan W. Simons<br />

The Prostate <strong>Cancer</strong> Foundation, Santa Monica, California, U.S.A.<br />

Shuming Nie<br />

Departments <strong>of</strong> Biomedical Engineering and Chemistry<br />

and the Winship <strong>Cancer</strong> Institute, Emory University and<br />

Georgia Institute <strong>of</strong> Technology, Atlanta, Georgia, U.S.A.<br />

INTRODUCTION<br />

<strong>Cancer</strong> nanotechnology is an interdisciplinary area <strong>of</strong> research in science, engineering,<br />

and medicine with broad applications in molecular pr<strong>of</strong>iling and individualized<br />

therapy (1–4). The basic rationale is that nanometer-sized particles such as<br />

biodegradable micelles, semiconductor quantum dots (QDs), and iron oxide nanocrystals<br />

have functional or structural properties that are not available from either<br />

molecular or bulk materials (5–15). When linked with biotargeting ligands such as<br />

monoclonal antibodies, peptides, or small molecules, these nanoparticles are used to<br />

target malignant tumors with high affinity and specificity. In the “mesoscopic” size<br />

range <strong>of</strong> 5 to 100 nm diameter, nanoparticles also have large surface areas and<br />

functional groups for conjugating to multiple diagnostic (e.g., optical, radioisotopic,<br />

309


310 Wang et al.<br />

or magnetic) and therapeutic (e.g., anticancer) agents. The emergence <strong>of</strong> nanotechnology<br />

has opened new opportunities for personalized oncology in which cancer<br />

detection, diagnosis, and therapy are tailored to each individual’s molecular pr<strong>of</strong>ile,<br />

and also for predictive oncology in which genetic or molecular information is used to<br />

predict tumor development, progression, and clinical outcome.<br />

Recent advances have led to the development <strong>of</strong> functional nanoparticles<br />

that are covalently linked to biological molecules such as peptides, proteins,<br />

nucleic acids, or small-molecule ligands (1–4). Medical applications have also<br />

appeared, such as the use <strong>of</strong> superparamagnetic iron oxide nanoparticles as a<br />

contrast agent for lymph node prostate cancer detection (16), and the use <strong>of</strong><br />

polymeric nanoparticles for targeted gene delivery to tumor vasculatures (17). In<br />

particular, the U.S. Food and Drug Administration (FDA) recently approved<br />

Abraxane TM , an albumin-paclitaxel (Taxol TM ) nanoparticle for the treatment <strong>of</strong><br />

metastatic breast cancer (18). A phase I clinical trial determined that the maximum<br />

tolerated dose (MTD) <strong>of</strong> single-agent albumin-bound paclitaxel every<br />

three weeks was 300 mg/m 2 in patients with solid tumors (breast cancer and<br />

melanoma). A second phase I trial demonstrated five patient responses among<br />

39 pretreated patients with advanced solid tumors, including one response in a<br />

patient with non–small cell lung cancer (NSCLC), three responses in patients<br />

with ovarian cancer, and one in breast cancer. The dose-limiting toxicity was<br />

myelosuppression, the MTD was 270 mg/m 2 , and premedication was not<br />

required. Subsequent use <strong>of</strong> Abraxane in both phase II and phase III trials proved<br />

that this new formulation was considerably superior to Taxol. The use <strong>of</strong><br />

nanoparticles for drug delivery and targeting is one <strong>of</strong> the most exciting and<br />

clinically important applications <strong>of</strong> cancer nanotechnology.<br />

CANCER BIOMARKERS AND NANOTECHNOLOGY<br />

<strong>Cancer</strong> markers are broadly defined as altered or mutant genes, RNA, proteins,<br />

lipids, carbohydrates, small metabolite molecules, and altered expression <strong>of</strong><br />

those that are correlated with a biological behavior or a clinical outcome (19–22).<br />

Most cancer biomarkers are discovered by molecular pr<strong>of</strong>iling studies on the<br />

basis <strong>of</strong> an association or correlation between a molecular signature and cancer<br />

behavior. In the cases <strong>of</strong> both breast and prostate cancers, a major progression<br />

step is the appearance <strong>of</strong> so-called “lethal phenotypes” (causing patient death)<br />

such as bone metastatic, hormone-independent, and radiation- and chemotherapyresistant<br />

phenotypes. It has been hypothesized that each <strong>of</strong> these aggressive<br />

behaviors or phenotypes could be understood and predicted by a defining set <strong>of</strong><br />

biomarkers (20). By critically defining the inter-relationships between these<br />

biomarkers, it could be possible to diagnose and determine the prognosis <strong>of</strong> a<br />

cancer on the basis <strong>of</strong> a patient’s molecular pr<strong>of</strong>ile, leading to personalized and<br />

predictive medicine.<br />

One example is the drug trastuzumab (Herceptin TM ; Genentech/Roche,<br />

California, U.S./Basel, Switzerland), a monoclonal antibody designed to target


<strong>Cancer</strong> Nanotechnology 311<br />

amplified and overexpressed ERBB2 (also known as HER2) tyrosine kinase<br />

receptor found in only approximately 25% to 30% <strong>of</strong> breast cancers. FDA<br />

approval <strong>of</strong> trastuzumab was predicated on the availability <strong>of</strong> a test to detect<br />

ERBB2 overexpression. Both an immunohistochemistry (IHC) assay for the<br />

expressed protein (HercepTest TM ; Dako, Glostrup, Denmark) and a nucleic acid–<br />

based fluorescence in situ hybridization (FISH) test (PathVysion; Abbott Laboratories,<br />

Illinois, U.S.) have been approved as in vitro diagnostics to guide<br />

trastuzumab treatment decisions. In another example, the clinical response <strong>of</strong><br />

lung cancer patients to the epidermal growth factor receptor (EGFR) tyrosine<br />

kinase inhibitor gefitinib (Iressa TM ; AstraZeneca, London, U.K.) is associated<br />

with a small number <strong>of</strong> genetic mutations (23,24). Thus, a molecular diagnostic<br />

test could be used to identify patients that are most likely to respond to this drug.<br />

Despite these advances, critical studies that can clearly link biomarkers<br />

with cancer behavior remain a significant challenge. One difficulty is that most<br />

cancer tumors (especially prostate and breast cancer) are highly heterogeneous,<br />

containing a mixture <strong>of</strong> benign, cancerous, and stromal cells. Current technologies<br />

for molecular pr<strong>of</strong>iling including reverse-transcriptase polymerase<br />

chain reaction (RT-PCR), gene chips, protein chips, two-dimensional gel electrophoresis,<br />

biomolecular mass spectrometry (e.g., MALDI-MS, i.e., matrix<br />

assisted laser desorption ionization mass spectrometry, ES-MS, i.e., electro spray<br />

mass spectroscopy, and SELDI-MS, i.e., surface enhanced laser desorption/<br />

ionization) are not designed to handle this type <strong>of</strong> heterogeneous samples<br />

(25,26). Furthermore, a limitation shared by all these technologies is that they<br />

require destructive preparation <strong>of</strong> cells or tissue specimens into a homogeneous<br />

solution, leading to a loss <strong>of</strong> valuable three-dimensional cellular and tissue<br />

morphological information associated with the original tumor. The development<br />

<strong>of</strong> nanotechnology, especially bioconjugated nanoparticles, provides an essential<br />

link by which biomarkers could be functionally correlated with cancer behavior<br />

(Fig. 1).<br />

NANOTYPING<br />

Semiconductor QDs are tiny light-emitting particles on the nanometer scale, and<br />

are emerging as a new class <strong>of</strong> fluorescent labels for biology and medicine<br />

(8–15). In comparison with organic dyes and fluorescent proteins, QDs have<br />

unique optical and electronic properties such as size-tunable light emission,<br />

superior signal brightness, resistance to photobleaching, and simultaneous excitation<br />

<strong>of</strong> multiple fluorescence colors (Fig. 2). These properties are most promising<br />

for improving the sensitivity and multiplexing capabilities <strong>of</strong> molecular histopathology<br />

and disease diagnosis. Recent advances have led to highly bright and<br />

stable QD probes that are well suited for pr<strong>of</strong>iling genetic and protein biomarkers<br />

in intact cells and clinical tissue specimens (2–4). In contrast to in vivo imaging<br />

applications where the potential toxicity <strong>of</strong> cadmium-containing QDs is a major<br />

concern, immunohistological staining is performed on in vitro or ex vivo clinical


312 Wang et al.<br />

Figure 1 Size-tunable optical properties <strong>of</strong> semiconductor QDs for multiplexed molecular<br />

pr<strong>of</strong>iling <strong>of</strong> cancer. (A) Fluorescence emission spectra <strong>of</strong> different sizes <strong>of</strong> CdSe QDs.<br />

(B) Relative sizes <strong>of</strong> the five sizes <strong>of</strong> CdSe QDs corresponding to the above spectrum.<br />

From left to right the diameters are: 2.1 nm, 2.5 nm, 2.9 nm, 4.7 nm, and 7.5 nm.<br />

Abbreviations: CdSe, cadium selenide; QDs, quantum dots.<br />

patient samples. As a result, the use <strong>of</strong> multicolor QD probes in IHC is likely one<br />

<strong>of</strong> the most important and clinically relevant applications in the near term (27).<br />

Indeed, significant opportunities exist at the interface between QD nanotechnology<br />

and signature biomarkers for cancer diagnosis and individualized<br />

therapy. In particular, QD nanoparticle probes have been used to quantify a panel<br />

<strong>of</strong> biomarkers on intact cancer cells and tissue specimens, allowing a correlation<br />

<strong>of</strong> traditional histopathology and molecular signatures for the same material<br />

(2–4,27). A single nanoparticle is large enough for conjugation to multiple<br />

ligands, leading to enhanced binding affinity and exquisite specificity through a<br />

“multivalency” effect (28). These features are especially important for the<br />

analysis <strong>of</strong> cancer biomarkers that are present at low concentrations or in small<br />

numbers <strong>of</strong> cells.<br />

Quantitative biomarker information can be obtained by using a spectrometer<br />

attached to the fluorescence microscope. It is however very important to<br />

use a common protein such as b actin or glyceraldehyde 3-phosphate dehydrogenase<br />

(GAPDH) as an “internal control.” That is, one <strong>of</strong> the QD–Ab (antibody)<br />

conjugates should be designed to measure the product <strong>of</strong> a housekeeping gene<br />

that is expressed at relatively constant levels in all cells. The use <strong>of</strong> an internal<br />

control holds great promise for overcoming a number <strong>of</strong> major problems in<br />

biomarker quantification, such as differences in the probe brightness, variations<br />

in probe binding efficiency, uneven light illumination, and detector responses.


<strong>Cancer</strong> Nanotechnology 313<br />

Figure 2 Nanotechnology linking biomarkers with cancer behavior and clinical outcome.<br />

The schematic diagram shows multiplexed detection and quantification <strong>of</strong> cancer<br />

biomarkers on intact cells or tissues with multicolor nanoparticle probes. (Left) The<br />

images show cancer cells labeled with quantum dots, and (right) the drawings suggest<br />

how surface and intracellular biomarkers could be quantified by wavelength-resolved<br />

spectroscopy or spectral imaging.<br />

The majority <strong>of</strong> available tumor specimens are archived, formalin-fixed paraffinembedded<br />

(FFPE) tissues that might be several decades old. As the clinical<br />

outcomes <strong>of</strong> these tissues are already known, these specimens are well suited for<br />

examining the relationship between molecular pr<strong>of</strong>ile and clinical outcome in<br />

retrospective studies.<br />

It is also critically important to validate the QD staining data with other<br />

available techniques. For this purpose, O’Regan and coworkers (2) have<br />

obtained QD molecular pr<strong>of</strong>iling data from standard human breast cell specimens<br />

and have compared the corresponding biomarker data with traditional IHC<br />

and FISH techniques. Briefly, slides from formaldehyde-fixed paraffin cell<br />

blocks were stained in accordance with standard pathological protocols for three<br />

breast cancer biomarkers—estrogen receptor (ER), progesterone receptor (PR),<br />

and HER2. This panel <strong>of</strong> protein biomarkers was selected because <strong>of</strong> its clinical<br />

significance in human breast cancer diagnosis and treatment (29–32). The traditional<br />

IHC results were analyzed by two independent observers and scored<br />

with a standard scale from 0 (no visible staining in the nucleus or membrane) to


314 Wang et al.<br />

3þ (strong and complete membrane or nuclear staining in > 10% <strong>of</strong> malignant<br />

stained cells). For a comparative analysis <strong>of</strong> QD pr<strong>of</strong>iling with traditional IHC, it<br />

is necessary to normalize the absolute fluorescence intensities <strong>of</strong> QD–Ab signals,<br />

so that relative percentage values are calculated from the maximum signal<br />

strength.<br />

The results reveal that a 3þ score for ER, PR, or HER2 by traditional IHC<br />

corresponds to 85% to 100% relative expression <strong>of</strong> the antigen by QD–Ab measurement,<br />

and that 1þ or 2þ scores by traditional IHC correspond to 11% to 48%<br />

expression as determined by QD quantification. It should be noted that classification<br />

<strong>of</strong> antigens expressed at low levels (1þ or 2þ) is subjective, requiring<br />

experience and <strong>of</strong>ten resulting in considerable interobserver variations. In contrast,<br />

quantitative QD measurements allow accurate determination <strong>of</strong> tumor antigens at<br />

low levels. For example, PR expression in MCF-7 cells and ER expression in BT-<br />

474 cells are both classified as 1þ by traditional IHC, but quantitative QD<br />

measurements indicate major differences in PR expression (16.8%) and ER<br />

expression (47.7%) in these two cell lines. This indicates that the quantitative<br />

nature <strong>of</strong> QD-based molecular pr<strong>of</strong>iling could simplify and standardize categorization<br />

<strong>of</strong> antigens that are expressed at low levels. This is <strong>of</strong> fundamental<br />

importance in the management <strong>of</strong> breast cancer, as the likely benefit <strong>of</strong> hormonal<br />

therapies and trastuzumab depends directly on not just the presence but also the<br />

quantity <strong>of</strong> hormone or HER2 receptors (33–35) (Fig. 3).<br />

Wang and coworkers (27) have developed an integrated image processing<br />

and bioinformatics s<strong>of</strong>tware tool called Q-IHC for quantitative analysis <strong>of</strong> biomarker<br />

expression and distribution in IHC images. In comparison with previous<br />

image processing s<strong>of</strong>tware for automated feature extraction and quantitative<br />

Figure 3 Targeting the HER 2/Neu antigen on breast cancer cells by using antibodyconjugated<br />

quantum dots. The image shows individual breast cancer cells with<br />

nuclear counterstaining.


<strong>Cancer</strong> Nanotechnology 315<br />

analysis (36,37), this s<strong>of</strong>tware system is capable <strong>of</strong> handling imaging data from<br />

both traditional and QD-based IHC. To measure the distribution <strong>of</strong> labeled<br />

antigens, multiple slides <strong>of</strong> IHC imaging data are acquired to capture selected<br />

tissue structures. After image acquisition, an image-processing module carries<br />

out automatic boundary identification, semiautomatic image segmentation, and<br />

color-based tissue classification on the basis <strong>of</strong> biomarker staining. Then, an<br />

image analysis module quantifies the various biomarker features into numerical<br />

values. These values become distinct features and are used for comparison with<br />

clinical diagnosis. After validation by a physician, the quantitative data and rules<br />

describing biomarker features are stored in a database. This semiautomatic<br />

image processing and quantification system is designed to provide molecular<br />

pr<strong>of</strong>iling data that are more objective, more consistent, and more reproducible<br />

than completely manual or automated quantification methods. The s<strong>of</strong>tware tools<br />

process image files from slide scanners in Matlab, which is a collection <strong>of</strong><br />

various engineering processing tools. In addition, a user-friendly graphical<br />

interface allows users to give input and feedback to improve the system quality.<br />

NANOTHERAPEUTICS<br />

Most current anticancer agents do not greatly differentiate between cancerous<br />

and normal cells, leading to systemic toxicity and adverse effects. Consequently,<br />

systemic applications <strong>of</strong> these drugs <strong>of</strong>ten cause severe side effects in other tissues<br />

(such as bone marrow suppression, cardiomyopathy, and neurotoxicity), which<br />

greatly limit the maximal allowable dose <strong>of</strong> the drug. In addition, rapid elimination<br />

and widespread distribution into nontargeted organs and tissues require administration<br />

<strong>of</strong> a drug in large quantities, which is not economical and <strong>of</strong>ten complicated<br />

because <strong>of</strong> nonspecific toxicity. Nanotechnology <strong>of</strong>fers a more targeted approach<br />

and could thus provide significant benefits to cancer patients.<br />

Rapid vascularization in fast-growing cancerous tissues is known to result<br />

in leaky, defective architecture and impaired lymphatic drainage. This structure<br />

allows an enhanced permeation and retention (EPR) effect (38–42), as a result <strong>of</strong><br />

which nanoparticles accumulate at the tumor site. For such passive targeting<br />

mechanism to work, the size and surface properties <strong>of</strong> drug delivery nanoparticles<br />

must be controlled to avoid uptake by the reticuloendothelial system<br />

(RES) (43). To maximize circulation time and targeting ability, the optimal size<br />

should be less than 100 nm in diameter and the surface should be hydrophilic to<br />

circumvent clearance by macrophages. A hydrophilic surface <strong>of</strong> the nanoparticles<br />

safeguards against plasma protein adsorption, and it can be achieved<br />

through hydrophilic polymer coatings such as polyethylene glycol (PEG),<br />

poloxamines, poloxamers, polysaccharides or through the use <strong>of</strong> branched or<br />

block amphiphilic copolymers (44–47). The covalent linkage <strong>of</strong> amphiphilic<br />

copolymers (polylactic acid, polycaprolactone, and polycyanonacrylate chemically<br />

coupled to PEG) is generally preferred, as it avoids aggregation and ligand<br />

desorption when in contact with blood components.


316 Wang et al.<br />

An alternative passive targeting strategy is to utilize the unique tumor<br />

environment in a scheme called tumor-activated prodrug therapy. The drug is<br />

conjugated to a tumor-specific molecule and remains inactive until it reaches the<br />

target (48). Overexpression <strong>of</strong> the matrix metalloproteinase (MMP), MMP-2, in<br />

melanoma has been shown in a number <strong>of</strong> preclinical as well as clinical<br />

investigations. Mansour et al. (49) reported a water-soluble maleimide derivative<br />

<strong>of</strong> doxorubicin (DOX) incorporating an MMP-2 specific peptide sequence (Gly-<br />

Pro-Leu-Gly-Ile-Ala-Gly-Gln) that rapidly and selectively binds to the cysteine-<br />

34 position <strong>of</strong> circulating albumin. The albumin-DOX conjugate is efficiently<br />

and specifically cleaved by MMP-2, releasing a DOX tetrapeptide (Ile-Ala-Gly-<br />

Gln-DOX) and subsequently DOX. pH and redox potential have been also<br />

explored as drug-release triggers at the tumor site (50). Another passive targeting<br />

method is the direct local delivery <strong>of</strong> anticancer agents to tumors. This approach<br />

has the obvious advantage <strong>of</strong> excluding the drug from the systemic circulation.<br />

However, administration can be highly invasive, as it involves injections or<br />

surgical procedures. For some tumors such as lung cancers that are difficult to<br />

access, the technique is nearly impossible to use.<br />

Active targeting is usually achieved by conjugating to the nanoparticle a<br />

targeting component that provides preferential accumulation <strong>of</strong> nanoparticles in<br />

the tumor-bearing organ, in the tumor itself, individual cancer cells, or intracellular<br />

organelles inside cancer cells. This approach is based on specific<br />

interactions such as lectin-carbohydrate, ligand-receptor, and antibody-antigen<br />

(51). Lectin-carbohydrate is one <strong>of</strong> the classic examples <strong>of</strong> targeted drug delivery<br />

(52). Lectins are proteins <strong>of</strong> nonimmunological origin, which are capable <strong>of</strong><br />

recognizing and binding to glycoproteins expressed on cell surfaces. Lectin<br />

interactions with certain carbohydrates are very specific. Carbohydrate moieties<br />

can be used to target drug delivery systems to lectins (direct lectin targeting), and<br />

lectins can be used as targeting moieties to target cell surface carbohydrates<br />

(reverse lectin targeting). However, drug delivery systems based on lectincarbohydrate<br />

have mainly been developed to target whole organs (53), which can<br />

pose harm to normal cells. Therefore, in most cases, the targeting moiety is<br />

directed toward specific receptors or antigens expressed on the plasma membrane<br />

or elsewhere at the tumor site.<br />

The cell surface receptor for folate is inaccessible from the circulation to<br />

healthy cells because <strong>of</strong> its location on the apical membrane <strong>of</strong> polarized epithelia,<br />

but it is overexpressed on the surface <strong>of</strong> various cancers like ovary, brain,<br />

kidney, breast, and lung malignancies (54,55). Surface plasmon resonance<br />

studies revealed that folate-conjugated PEGylated cyanoacrylate nanoparticles<br />

had a 10-fold higher affinity for the folate receptor than the free folate ones did<br />

(56). Folate receptors are <strong>of</strong>ten organized in clusters and bind preferably to the<br />

multivalent forms <strong>of</strong> the ligand. Furthermore, confocal microscopy demonstrated<br />

selective uptake and endocytosis <strong>of</strong> folate-conjugated nanoparticles by tumor<br />

cells bearing folate receptors. Interest in exploiting folate receptor targeting<br />

in cancer therapy and diagnosis has rapidly increased, as attested by many


<strong>Cancer</strong> Nanotechnology 317<br />

conjugated systems including proteins, liposomes, imaging agents, and neutron<br />

activation compounds (54,55).<br />

For enhanced tumor-specific targeting, the differences between cancerous<br />

cells and normal cells may be exploited. By virtue <strong>of</strong> their small size, nanoparticles<br />

entail a high surface area that not only paves a way for more efficient<br />

drug release but also a better strategy for functionalization. There is a growing<br />

body <strong>of</strong> knowledge on unique cancer markers due to recent advances in proteomics<br />

and genomics. They form the basis <strong>of</strong> complex interactions between<br />

bioconjugated nanoparticles and cancer cells. Carrier design and targeting<br />

strategies may vary according to the type, developmental stage, and location <strong>of</strong><br />

cancer (57). There is much synergy between imaging and nanotechnology in<br />

biomedical applications. Many <strong>of</strong> the principles used to target delivery <strong>of</strong> drugs<br />

to cancer may also be applied to target imaging and diagnostic agents to enhance<br />

detection sensitivity in medical imaging. With engineered multifunctional<br />

nanoparticles, the full in vivo potential <strong>of</strong> cancer nanotechnology in targeted<br />

drug delivery and imaging can be realized.<br />

BIONANOINFORMATICS<br />

For the long-term goal <strong>of</strong> personalized and predictive oncology, it is important to<br />

integrate biotechnology and nanotechnology with information technology<br />

(Fig. 4). A major step in this direction is the development <strong>of</strong> the <strong>Cancer</strong> Bioinformatics<br />

Grid (caBIG), the “World Wide Web for <strong>Cancer</strong> Research” spearheaded<br />

by the U.S. National <strong>Cancer</strong> Institute (NCI) (58). Through data sharing<br />

and standardization, this national initiative aims at improving the infrastructures<br />

<strong>of</strong> health care information technology, to facilitate the process from clinical data<br />

Figure 4 Schematic diagram showing the integration <strong>of</strong> biotechnology, nanotechnology,<br />

and information technology for personalized oncology.


318 Wang et al.<br />

collection to computational data mining, and to accelerate the use <strong>of</strong> biomolecular<br />

markers for personalized treatment. The caBIG organizational structure<br />

is divided into working groups, each focused on different aspects <strong>of</strong> the<br />

interoperability problem specific to cancer research. Two overarching work<br />

groups, “Architecture” and “Vocabularies and Common Data Elements” have<br />

begun to produce detailed specifications that will guide the design <strong>of</strong> interoperable<br />

bioinformatics tools. In practice, caBIG provides a common platform<br />

from which to launch even more ambitious integrated solutions. Bioinformatics<br />

laboratories can adopt the basic infrastructure, databases, and functionality<br />

provided by caBIG working groups to fill gaps in knowledge and to carry out<br />

information management responsibilities. In addition to these shared tools and<br />

architectures, caBIG is the perfect test bed for a suite <strong>of</strong> standard tools that can<br />

be understood, verified, and used by clinicians around the world. Another major<br />

effort is <strong>Cancer</strong> Bioinformatics Infrastructure Objects (caBIO), an application<br />

programming interface (API) that was first developed as Java objects and was<br />

later extended to support various programming platforms. These objects represent<br />

the most fundamental concepts in bioinformatics research such as genes,<br />

ontology, and sequences. Also under development are network programs such as<br />

caBIONet based on Micros<strong>of</strong>t’s .NET technology for data analysis and integration<br />

(59).<br />

OUTLOOK<br />

Looking into the future, there are a number <strong>of</strong> research themes or directions that<br />

are particularly promising but require concerted effort for success. The first<br />

direction is the design and development <strong>of</strong> nanoparticles with mon<strong>of</strong>unctions,<br />

dual functions, three functions, or multiple functions. For cancer and other<br />

medical applications, important functions include imaging (single or dualmodality),<br />

therapy (single drug or combination <strong>of</strong> two or more drugs), and<br />

targeting (one or more ligands). With each added function, nanoparticles could<br />

be designed to have novel properties and applications. For example, binary<br />

nanoparticles with two functions could be developed for molecular imaging,<br />

targeted therapy, or for simultaneous imaging and therapy (but without targeting).<br />

Bioconjugated QDs with both targeting and imaging functions will be used<br />

for targeted tumor imaging and molecular pr<strong>of</strong>iling applications. Conversely,<br />

ternary nanoparticles with three functions could be designed for simultaneous<br />

imaging and therapy with targeting, targeted dual-modality imaging, or for targeted<br />

dual-drug therapy. Quaternary nanoparticles with four functions can be<br />

conceptualized in future to have the abilities <strong>of</strong> tumor targeting, dual-drug<br />

therapy, and imaging. The second direction is nanoparticle molecular pr<strong>of</strong>iling<br />

(nanotyping) for clinical oncology; i.e., the use <strong>of</strong> bioconjugated nanoparticle<br />

probes to predict cancer behavior, clinical outcome, treatment response, and<br />

individualize therapy. This should start with retrospective studies <strong>of</strong> archived<br />

specimens because the patient outcome is already known for these specimens.


<strong>Cancer</strong> Nanotechnology 319<br />

The key hypotheses to be tested are that nanotyping a panel <strong>of</strong> tumor markers<br />

will allow more accurate correlations than single tumor markers, and that the<br />

combination <strong>of</strong> nanotyping tumor gene expression and host stroma are both<br />

important in defining the aggressive phenotypes <strong>of</strong> cancer as well as determining<br />

the response <strong>of</strong> early-stage disease to treatment (chemotherapy, radiation, or<br />

surgery). The third important direction is to study nanoparticle distribution,<br />

excretion, metabolism, and pharmacodynamics in in vivo animal models. These<br />

investigations will be very important in the development <strong>of</strong> nanoparticles for<br />

clinical applications in cancer imaging or therapy.<br />

In conclusion, nanotechnology is emerging as an enabling technology for<br />

personalized oncology in which cancer detection, diagnosis, and therapy are<br />

tailored to each individual’s tumor molecular pr<strong>of</strong>ile and also for predictive<br />

oncology in which genetic or molecular markers are used to predict disease<br />

development, progression, and clinical outcomes (60,61).<br />

ACKNOWLEDGEMENTS<br />

We are grateful to Dr. Leland Chung, Dr. Ruth O’Regan, and Dr. Dong Shin for<br />

insightful discussions. Our research in cancer nanotechnology was supported by<br />

National Institutes <strong>of</strong> Health (NIH) grants (R01 CA108468-01 and U54CA119338)<br />

and the Georgia <strong>Cancer</strong> Coalition Distinguished <strong>Cancer</strong> Scholars Program (to May<br />

Dongmei Wang and Shuming Nie).<br />

REFERENCES<br />

1. Ferrari M. <strong>Cancer</strong> nanotechnology: Opportunities and challenges. Nat Rev <strong>Cancer</strong><br />

2005; 5:161–171.<br />

2. Yezhelyev M, Gao XH, Xing Y, et al. Emerging use <strong>of</strong> nanoparticles in diagnosis<br />

and treatment <strong>of</strong> breast cancer. Lancet Oncol 2006; 7:657–767.<br />

3. Sinha R, Kim GK, Nie SM, et al. Nanotechnology in cancer therapeutics: bioconjugated<br />

nanoparticles for drug delivery. Mol <strong>Cancer</strong> Ther 2006; 5:1909–1917.<br />

4. Xing Y, Kim GJ, Nie SM, et al. Nanotechnology Applications in <strong>Cancer</strong>. Annu Rev<br />

Biomed Eng 2007; 9:257–288 (review).<br />

5. Henglein A. Small-particle research - physicochemical properties <strong>of</strong> extremely small<br />

colloidal metal and semiconductor particles. Chem Rev 1989; 89:1861–1873.<br />

6. Schmid G. Large clusters and colloids - metals in the embryonic state. Chem Rev<br />

1992; 92:1709–1727.<br />

7. Niemeyer CM. Nanoparticles, proteins, and nucleic acids: biotechnology meets<br />

materials science. Angew Chem Int Ed Engl 2001; 40:4128–4158.<br />

8. Alivisatos AP. The use <strong>of</strong> nanocrystals in biological detection. Nat Biotechnol 2004;<br />

22:47–52.<br />

9. Alivisatos AP. Semiconductor clusters, nanocrystals, and quantum dots. Science<br />

1996; 271:933–937.<br />

10. Alivisatos AP, Gu WW, Larabell C. Quantum dots as cellular probes. Annu Rev<br />

Biomed Eng 2005; 7:55–76.


320 Wang et al.<br />

11. Pinaud F, Michalet X, Bentolila LA, et al. Advances in fluorescence imaging with<br />

quantum dot bio-probes. Biomaterials 2006; 27:1679–1687.<br />

12. Michalet X, Pinaud FF, Bentolila LA, et al. Quantum dots for live cells, in vivo<br />

imaging, and diagnostics. Science 2005; 307:538–544.<br />

13. Gao XH, Yang LL, Petros JA, et al. In vivo molecular and cellular imaging with<br />

quantum dots. Curr Opin Biotechnol 2005; 16: 63–72.<br />

14. Smith AM, Gao X, Nie S. Quantum dot nanocrystals for in vivo molecular and<br />

cellular imaging. <strong>Ph</strong>otochem <strong>Ph</strong>otobiol 2004; 80:377–3785.<br />

15. Chan WC, Maxwell DJ, Gao X, et al. Luminescent quantum dots for multiplexed<br />

biological detection and imaging. Curr Opin Biotechnol 2002; 13:40–46.<br />

16. Harisinghani MG, Barentsz J, Hahn PF, et al. Noninvasive detection <strong>of</strong> clinically<br />

occult lymph-node metastases in prostate cancer. N Engl J Med 2003; 348:2491–2495.<br />

17. Hood JD, Bednarski M, Frausto R, et al. Tumor regression by targeted gene delivery<br />

to the neovasculature. Science 2002; 296:2404–2407.<br />

18. Robinson DM, Keating GM. Albumin-bound paclitaxel in metastatic breast cancer.<br />

Drugs 2006; 66:941–948.<br />

19. Hahn WC, Weinberg RA. Modelling the molecular circuitry <strong>of</strong> cancer. Nat Rev<br />

<strong>Cancer</strong> 2002; 2:331–341.<br />

20. Liotta L, Petricoin E. Molecular pr<strong>of</strong>iling <strong>of</strong> human cancer. Nat Rev Genet 2000;<br />

1:48–56.<br />

21. Petricoin EF, Zoon KC, Kohn EC, et al. Clinical proteomics: translating benchside<br />

promise into bedside reality. Nat Rev Drug Discov 2002; 1:683–695.<br />

22. Menon U, Jacobs IJ. Recent developments in ovarian cancer screening. Curr Opin<br />

Obstet Gynecol 2000; 12:39–42.<br />

23. Lynch TJ, Bell DW, Sordella R, et al. Activating mutations in the epidermal growth<br />

factor receptor underlying responsiveness <strong>of</strong> non-small-cell lung cancer to gefitinib.<br />

N Engl J Med 2004; 350:2129–2139.<br />

24. Paez JG, Janne PA, Lee JC, et al. EGFR mutations in lung cancer: correlation with<br />

clinical response to gefitinib therapy. Science 2004; 304:1497–1500.<br />

25. Michener CM, Ardekani AM, Petricoin EF, et al. Genomics and proteomics:<br />

application <strong>of</strong> novel technology to early detection and prevention <strong>of</strong> cancer. <strong>Cancer</strong><br />

Detect Prev 2002; 26:249–255.<br />

26. Srinivas PR, Verma M, Zhao Y, et al. Proteomics for cancer biomarker discovery.<br />

Clin Chem 2002; 48:1160–1169.<br />

27. Xing Y, Chaudry Q, Shen C, et al. Bioconjugated quantum dots for multiplexed and<br />

quantitative immunohistochemistry. Nat Protoc 2007; 2:1152–1165.<br />

28. Akerman ME, Chan WC, Laakkonen P, et al. Nanocrystal targeting in vivo. Proc<br />

Natl Acad Sci U S A 2002; 99:12617–12621.<br />

29. Umemura S, Osamura RY. Utility <strong>of</strong> immunohistochemistry in breast cancer practice.<br />

<strong>Breast</strong> <strong>Cancer</strong> 2004; 11:334–338.<br />

30. Thomson TA, Hayes MM, Spinelli JJ, et al. HER-2/neu in breast cancer: interobserver<br />

variability and performance <strong>of</strong> immunohistochemistry with 4 antibodies compared<br />

with fluorescent in situ hybridization. Mod Pathol 2001; 14:1079–1086.<br />

31. Vogel CL, Cobleigh MA, Tripathy D, et al. Efficacy and safety <strong>of</strong> trastuzumab as a<br />

single agent in first-line treatment <strong>of</strong> HER2-overexpressing metastatic breast cancer.<br />

J Clin Oncol 2002; 20:719–726.<br />

32. Elledge RM, Green S, Pugh R, et al. Estrogen receptor (ER) and progesterone<br />

receptor (PR), by ligand-binding assay compared with ER, PR and pS2, by


<strong>Cancer</strong> Nanotechnology 321<br />

immunohistochemistry in predicting response to tamoxifen in metastatic breast<br />

cancer: a Southwest Oncology Group Study. Int J <strong>Cancer</strong> 2000; 89:111–117.<br />

33. Marquez DC, Pietras RJ. Membrane-associated binding sites for estrogen contribute<br />

to growth regulation <strong>of</strong> human breast cancer cells. Oncogene 2001; 20:5420–5430.<br />

34. Hicks DG, Tubbs RR. Assessment <strong>of</strong> the HER2 status in breast cancer by fluorescence<br />

in situ hybridization: a technical review with interpretive guidelines. Hum<br />

Pathol 2005; 36:250–261.<br />

35. Konecny G, Pauletti G, Pegram M, et al. Quantitative association between HER-2/<br />

neu and steroid hormone receptors in hormone receptor-positive primary breast<br />

cancer. J Natl <strong>Cancer</strong> Inst 2003; 95:142–153.<br />

36. Matkowskyj KA, Schonfeld D, Benya RV. Quantitative immunohistochemistry by<br />

measuring cumulative signal strength using commercially available s<strong>of</strong>tware <strong>Ph</strong>otoshop<br />

and Matlab. J Histochem Cytochem 2000; 48:303–311.<br />

37. Matkowskyj KA, Schonfeld D, Benya RV. Quantitative immunohistochemistry by<br />

measuring cumulative signal strength accurately measures receptor number. J Histochem<br />

Cytochem 2003; 51:205–214.<br />

38. Matsumura Y, Maeda H. A new concept for macromolecular therapeutics in cancer<br />

chemotherapy: mechanism <strong>of</strong> tumoritropic accumulation <strong>of</strong> proteins and the antitumor<br />

agent smancs. <strong>Cancer</strong> Res 1986; 46:6387–6392.<br />

39. Duncan R. The dawning era <strong>of</strong> polymer therapeutics. Nat Rev Drug Discov 2003;<br />

2:347–360.<br />

40. Jain RK. Transport <strong>of</strong> molecules, particles, and cells in solid tumors. Annu Rev<br />

Biomed Eng 1999; 1:241–263.<br />

41. Jain RK. Delivery <strong>of</strong> molecular medicine to solid tumors: lessons from in vivo<br />

imaging <strong>of</strong> gene expression and function. J Control Release 2001; 74:7–25.<br />

42. Jain RK. Understanding barriers to drug delivery: High resolution in vivo imaging is<br />

key. Clin <strong>Cancer</strong> Res 1999; 5:1605–1606.<br />

43. Gref R, Minamitake Y, Peracchia MT, et al. Biodegradable Long-Circulating<br />

Polymeric Nanospheres. Science 1994; 263:1600–1603.<br />

44. Ringsdorf H. Structure and properties <strong>of</strong> pharmacologically active polymers. J Polymer<br />

Sci Polymer Symp 1975; 51:135–153.<br />

45. Davis FF. The origin <strong>of</strong> pegnology. Adv Drug Deliv Rev 2002; 54:457–458.<br />

46. Moghimi SM, Hunter AC. Poloxamers and poloxamines in nanoparticle engineering<br />

and experimental medicine. Trends Biotechnol 2000; 18:412–420.<br />

47. Park EK, Lee SB, Lee YM. Preparation and characterization <strong>of</strong> methoxy poly(ethylene<br />

glycol)/poly(epsilon-caprolactone) amphiphilic block copolymeric nanospheres<br />

for tumor-specific folate-mediated targeting <strong>of</strong> anticancer drugs. Biomaterials 2005; 26:<br />

1053–1061.<br />

48. Chari RV. Targeted delivery <strong>of</strong> chemotherapeutics: tumor-activated prodrug therapy.<br />

Adv Drug Deliv Rev 1998; 31:89–104.<br />

49. Mansour AM, Drevs J, Esser N, et al. A new approach for the treatment <strong>of</strong> malignant<br />

melanoma: Enhanced antitumor efficacy <strong>of</strong> an albumin-binding doxorubicin prodrug<br />

that is cleaved by matrix metalloproteinase 2. <strong>Cancer</strong> Res 2003; 63:4062–4066.<br />

50. Guo X, Szoka FC. Chemical approaches to triggerable lipid vesicles for drug and<br />

gene delivery. Acc Chem Res 2003; 36:335–341.<br />

51. Allen TM. Ligand-targeted therapeutics in anticancer therapy. Nat Rev <strong>Cancer</strong> 2002;<br />

2:750–763.


322 Wang et al.<br />

52. Kannagi R, Izawa M, Koike T, et al. Carbohydrate-mediated cell adhesion in cancer<br />

metastasis and angiogenesis. <strong>Cancer</strong> Sci 2004; 95:377–384.<br />

53. Yamazaki N, Kojima S, Bovin NV, et al. Endogenous lectins as targets for drug<br />

delivery. Adv Drug Deliv Rev 2000; 43:225–244.<br />

54. Leamon CP, Reddy JA. Folate-targeted chemotherapy. Adv Drug Deliv Rev 2004;<br />

56:1127–1141.<br />

55. Leamon CP, Low PS. Folate-mediated targeting: from diagnostics to drug and gene<br />

delivery. Drug Discov Today 2001; 6: 44–51.<br />

56. Stella B, Arpicco S, Peracchia MT, et al. Design <strong>of</strong> folic acid-conjugated nanoparticles<br />

for drug targeting. J <strong>Ph</strong>arm Sci 2000; 89:1452–1464.<br />

57. Vicent MJ, Duncan R. Polymer conjugates: nanosized medicines for treating cancer.<br />

Trends Biotechnol 2006; 24:39–47.<br />

58. Covitz PA, Hartel F, Schaefer C, et al. caCORE: a common infrastructure for cancer<br />

informatics. Bioinformatics 2003; 19:2404–2412.<br />

59. Kraj P, McIndoe RA. caBIONet: a .NET wrapper to access and process genomic data<br />

stored at the National <strong>Cancer</strong> Institute’s Center for Bioinformatics databases. Bioinformatics<br />

2005; 21:3456–3458.<br />

60. Weinshilboum R, Wang LW. <strong>Ph</strong>armacogenomics: Bench to bedside. Nat Rev Drug<br />

Discov 2004; 3:739–748.<br />

61. Evans WE, Relling MV. Moving towards individualized medicine with pharmacogenomics.<br />

Nature 2004; 429:464–468.


Index<br />

ABCB1 polymorphisms, 19–20<br />

Abraxane TM , 310<br />

Acute myeloid leukemic (AML) cells, 271<br />

Adenosine triphosphate (ATP), 52<br />

Adjuvant chemotherapy, 100–101, 207<br />

Adjuvant! Online, 187, 189, 205, 207<br />

Adjuvant on Line (AOL), vs. RS assay,<br />

103–104<br />

AdnaGen system, 157<br />

Adriamycin. See Doxorubicin<br />

Advancing Clinico Genomic Trials<br />

(ACGT), 207<br />

16a-[ 18 F] fluoro-17b-estradiol (FES), 298<br />

Affymetrix GeneChip, 250, 266<br />

Affymetrix microarray, 239<br />

AIB1, 171<br />

Akt, 173, 277<br />

Aldefluor 1 , 270<br />

Aldehyde dehydrogenase (ALDH), 270–271<br />

ALDH. See Aldehyde dehydrogenase<br />

(ALDH)<br />

17-allylamino-17-demethoxygeldanamycin<br />

(17AAG), 299<br />

American College <strong>of</strong> Surgeons Oncology<br />

Group (ACOSOG), 234<br />

American Joint Committee on <strong>Cancer</strong><br />

(AJCC) classification system,<br />

159–160<br />

Aminoglutethimide, 174<br />

Amonafide, 17<br />

Amphiphilic copolymers, covalent linkage<br />

<strong>of</strong>, 315<br />

Amplicons, 244<br />

Anastrozole, 32, 174<br />

chemical structure <strong>of</strong>, 175<br />

vs. tamoxifen, 177<br />

Angiogenesis<br />

PET and SPECT imaging <strong>of</strong>, 299<br />

Animal model, 299<br />

Anthracyclines<br />

CBR and, 28<br />

oxidative stress and, 28–29<br />

Antiapoptosis genes, 201, 202<br />

Anticytokeratin antibodies, 153, 154<br />

Antiestrogens, 168–174<br />

Anti-VEGF, 300<br />

API. See Application programming<br />

interface (API)<br />

Apoptosis, 298<br />

Application programming interface<br />

(API), 318<br />

Aromatase inhibitors (AI), 32, 137<br />

analysis <strong>of</strong> gene expression with,<br />

177–178<br />

chemical structures <strong>of</strong>, 175<br />

mechanisms <strong>of</strong> resistance, 175–176<br />

steroidal and nonsteroidal, 174<br />

vs. tamoxifen, 176–177<br />

323


324 Index<br />

Array CGH, 250<br />

ATAC trial, 175–177<br />

AutoMACS TM system, 153–155<br />

Automatic magnetic cell sorting system.<br />

See AutoMACS TM system<br />

Avidin-biotin complex peroxidase<br />

technique, 154<br />

BAC. See Bacterial artificial<br />

chromosomes (BAC)<br />

Bacterial artificial chromosomes (BAC),<br />

64–65<br />

b-catenin, 273<br />

Bcl-2 tumor supressor, 248<br />

BCRP1 (breast cancer resistance<br />

protein 1), 268<br />

17b-estradiol (E2), 168, 201<br />

Bioinformatics laboratories, 318<br />

Biomarkers, 51. See also <strong>Cancer</strong><br />

biomarkers<br />

Bionanoinformatics, 317–318<br />

Biopsies, gene expression pr<strong>of</strong>iling in,<br />

91–92<br />

Biotinylated random nonamers, 246<br />

Blood collection tubes, 216–217<br />

Blood tisssue, 218<br />

Bmi-1, 271–272<br />

BMI-1 oncogenic pathway, 202<br />

BOLD (blood oxygen level-dependent)<br />

MRI, 300<br />

Bone marrow<br />

cancer stem cells in, 152<br />

cytokeratin positive tumor cells in, 206<br />

micrometastasis, 154<br />

tumor cells, 154<br />

Bone mineral density, 178<br />

Bone scanning, 291, 292<br />

BRCA1, 5, 50–51, 266, 267<br />

BRCA2, 5<br />

<strong>Breast</strong> cancer<br />

genetic susceptibility, 5<br />

luminal, 198<br />

molecular classification <strong>of</strong>, 3, 198–199<br />

patients information<br />

storage and protection <strong>of</strong>, 231–232<br />

treatment with genetic<br />

polymorphism, 267<br />

<strong>Breast</strong> cancer cells<br />

circulating. See Circulating tumor cells<br />

subgroups <strong>of</strong>, 202–204<br />

<strong>Breast</strong> <strong>Cancer</strong> Intergroup, 186, 188, 190<br />

<strong>Breast</strong> International Group (BIG) 1-98<br />

study, 175<br />

<strong>Breast</strong> MRI, 295–296<br />

<strong>Breast</strong> tumor model, 269<br />

caBIG. See <strong>Cancer</strong> Bioinformatics Grid<br />

(caBIG)<br />

caBIONet, 318<br />

<strong>Cancer</strong> Bioinformatics Infrastructure<br />

Objects (caBIO) API, 318<br />

<strong>Cancer</strong> biomarkers, 310–311<br />

<strong>Cancer</strong> Biomedical Informatics Grid<br />

(caBIG), 207<br />

development <strong>of</strong>, 317<br />

organizational structure <strong>of</strong>, 318<br />

<strong>Cancer</strong> cells vs. normal cells, 317<br />

<strong>Cancer</strong> metastasis, 278<br />

<strong>Cancer</strong> Methylation Panel, 250, 251<br />

<strong>Cancer</strong>ous tissues, vascularization in, 315<br />

<strong>Cancer</strong> stem cells. See Stem cells<br />

<strong>Cancer</strong> Trials Support Unit, 192<br />

CapSure HS LCM Cap, 223, 224<br />

Carbohydrate moieties, 316<br />

Carbonyl reductase (CBR), 28<br />

Carcinogenesis, 269<br />

CAT. See Catalase<br />

Catalase (CAT), 29<br />

11 C-choline, 298<br />

CD24 breast cancer cells, 202<br />

CD44+ breast cancer cells, 202<br />

CD44 + CD24 /loo breast cancer cells,<br />

265, 266, 270<br />

CD44 expression, 270<br />

cDNA synthesis, 246, 247<br />

CellPr<strong>of</strong>ile TM kit, 153<br />

CellSearch TM system, 153<br />

for prognostic and predictive value <strong>of</strong><br />

CTCs, 158<br />

CellSpotter TM Analyzer, 153<br />

c-erbB-2, 161<br />

CGH. See Comparative genomic<br />

hybridization (CGH)<br />

Chemotherapeutic agents, 263


Index 325<br />

Chemotherapy, 159<br />

adjuvant. See Adjuvant chemotherapy<br />

cytotoxic, 161<br />

evaluation <strong>of</strong> marker, 186<br />

neoadjuvant. See Neoadjuvant<br />

chemotherapy<br />

residual clinical risk and, 102–103<br />

Chromatin immunoprecipitation (ChIP), 53<br />

Chromogenic in situ hybridization<br />

(CISH), 139<br />

Circulating female hormone, 168<br />

Circulating tumor cells (CTCs)<br />

baseline and follow-up, 159<br />

CD326-positive, 154, 155<br />

clinical application <strong>of</strong>, 160–161<br />

detection, 153–157<br />

clinical benefits <strong>of</strong>, 161<br />

future uses <strong>of</strong>, 162<br />

immunostains <strong>of</strong>, 154<br />

magnetic labeling and separation <strong>of</strong>,<br />

219–220<br />

prognostic value <strong>of</strong>, 158<br />

for staging <strong>of</strong> MBC, 159–160<br />

CISH. See Chromogenic in situ<br />

hybridization (CISH)<br />

Classification <strong>of</strong> breast tumors, 250<br />

Clinical decision making, 190<br />

Clinical Laboratory Improvement<br />

Amendments (CLIA), 190<br />

Clinical trials<br />

for antiestrogens, 168<br />

designing, 190–192<br />

protocols, 214, 228<br />

for treatment with MAPK-specific<br />

inhibitors, 172<br />

Clinicogenomic models, 207<br />

Cloned probe arrays, in CGH, 65–66<br />

Cluster analysis, 248<br />

11 C-methionine, 298<br />

CNA, Copy number abnormalities (CNA)<br />

CNB. See Core needle biopsy (CNB)<br />

Coexpression modules, 202<br />

Combined imaging, with FDG-PET and<br />

MRI, 297<br />

Common Rule, 232<br />

Comparative genomic hybridization<br />

(CGH), 62, 249<br />

BAC, 64–65<br />

[Comparative genomic hybridization<br />

(CGH)]<br />

cloned probe arrays in, 65–66<br />

DNA in, 69–71<br />

metaphase chromosomes in, 65<br />

MIP, 66–69<br />

oligonucleotide-based arrays, 65–66<br />

Complete pathological response (pCR),<br />

101–102<br />

Computed tomography (CT) scanning, 294<br />

Contingency plans, 214<br />

Copy number abnormalities (CNA), 61, 249<br />

assessment by CGH. See Comparative<br />

genomic hybridization (CGH)<br />

gene deregulation by, 62–63<br />

Coregulators, 171<br />

Core needle biopsy (CNB), 83, 238–239<br />

Cox model, 245<br />

Cox regression, 159<br />

CpG dinucleotides, 45–46<br />

Cryopreservation techniques, 235–237<br />

CTCs. See Circulating tumor cells (CTCs)<br />

Cu-ATSM, 299<br />

Cu-DOTA-VEGF, 299<br />

CXCR4 chemokine receptor, 278, 280<br />

Cycling cells, 267, 282<br />

Cyclophosphamide<br />

CYP3A4 in, 27<br />

CYP2B6 in, 26–27<br />

and GST, 27–28<br />

CYP. See Cytochrome p450s (CYP)<br />

CYP3A4, in cyclophosphamide, 27<br />

CYP1B1, 267<br />

CYP2B6, in cyclophosphamide, 26–27<br />

CYP2C8<br />

in paclitaxel, 31<br />

CYP2D6<br />

genotyping, 170<br />

in TAM, 32<br />

Cytochrome p450s (CYP), 267<br />

Cytology, 51–52<br />

Cytotoxic agents, in neoadjuvant<br />

chemotherapy, 114–115<br />

Cytotoxic therapies, 109<br />

Dandelion hypothesis, 260<br />

DASL assay, 246


326 Index<br />

Database information, and patient<br />

identifiers, 232<br />

Daunorubicin, 28<br />

DCIS. See Ductal carcinoma in situ<br />

(DCIS)<br />

Deacetylation, <strong>of</strong> histones, 47<br />

Decision aids, 187<br />

Decision boards, 187<br />

DES. See Diethylstilbestrol (DES)<br />

DFS. See Disease-free survival<br />

3D GRE technique, 296–297<br />

Diethylstilbestrol (DES), vs.<br />

tamoxifen, 168<br />

Dihydropyrimidine dehydrogenase<br />

(DPD), 18–19<br />

Discovery-based research, 185<br />

Disease-free survival (DFS), 32–36, 109<br />

defined, 192<br />

GST in, 33<br />

MnSOD, 36<br />

MTHFR in, 33<br />

SULT1A1, 36<br />

UGT2B15, 36<br />

Disseminated tumor cells (DTC), 206<br />

Distant recurrence-free survival (DRFS)<br />

defined, 192<br />

DNA, in CGH, 69–71<br />

DNA hypermethylation, 279<br />

DNA isolation, 225–226<br />

DNA methylation, 45<br />

DNMT in, 46–47<br />

inhibitors, 53–54, 55<br />

DNA methyltransferases (DNMT), 46–47<br />

DNA microarray analysis, 107–108<br />

DNA synthesis, 298<br />

4 0 ,6-doamidino-2-phenylindole<br />

(DAPI), 153<br />

Docetaxel, 19–20<br />

Doxorubicin (DOX), 28, 316<br />

DPD. See Dihydropyrimidine<br />

dehydrogenase (DPD)<br />

Drosophila, 276<br />

Drugs<br />

chemotherapy, 159<br />

development<br />

chemotherapy, 159<br />

for CTC, 161<br />

predictive markers in, 141–143<br />

DSL ligands, 276<br />

Ductal carcinoma in situ (DCIS), 174<br />

E2. See 17b-estradiol (E2)<br />

Eastern Cooperative Oncology Group<br />

(ECOG), 192<br />

EasySep TM system, 155–157<br />

EGFR. See Epidermal growth factor<br />

receptor (EGFR)<br />

Embryogenesis, 269<br />

Embryonic stem cells, 271, 279<br />

Endocrine therapy, 167–168<br />

with antiestrogens, 168–174<br />

with aromatase inhibitors,<br />

174–178<br />

Endothelial cells, 279, 280<br />

magnetic labeling and separation <strong>of</strong>,<br />

219–220<br />

Endoxifen. See 4-hydroxy-N-desmethyl<br />

tamoxifen<br />

Enzymes, metabolic, 267<br />

EpCAM. See Epithelial cell adhesion<br />

molecule-1 (EpCAM)<br />

Epidermal growth factor receptor (EGFR),<br />

140–141, 171, 172<br />

Epigenetics<br />

DNA methylation. See DNA<br />

methylation<br />

histones. See Histones<br />

Epithelial cell adhesion molecule-1<br />

(EpCAM), 153<br />

Epithelial stem cells, 263<br />

ER. See Estrogen receptors (ER)<br />

ERa, 136–138, 168<br />

gene transcription, 169<br />

ERa-positive breast cancer, 168, 173<br />

ERa-positive cells, 269<br />

ERBB2, 62, 311<br />

ER-negative breast cancer, 168, 170<br />

ER-negative tumor cells, 269–270<br />

ER-positive breast cancer, 187, 246<br />

ER-positive tumors, 198, 245<br />

Estrogen receptors (ER), 51, 108, 298<br />

ERa, 136–138, 168<br />

gene transcription, 169<br />

gene expression <strong>of</strong>, 55–56<br />

mutations, 170


Index 327<br />

European Human Frozen Tumor Tissue<br />

Bank, 234<br />

European Union, 207<br />

Exemestane, 174<br />

chemical structure <strong>of</strong>, 175<br />

ExtractSure device, 223, 224<br />

FDG PET, 293–294<br />

FES. See 16a-[ 18 F] fluoro-17b-estradiol<br />

(FES)<br />

FES PET, 298<br />

18 F-fluorodeoxyglucose (FDG), 292<br />

[ 18 F] fluorotamoxifen, 298<br />

FFPE. See Formalin-fixed<br />

paraffin-embedded (FFPE)<br />

FFPE tissues, 313<br />

Fibroblasts, 280<br />

Fibromas, 297<br />

Ficoll-Paque TM , density gradient<br />

procedure, 156<br />

Fine needle aspiration (FNA), 83, 238<br />

Fixed tissue, 218<br />

18 F-labeled paclitaxel, 298<br />

Flow cytometry, 264<br />

Fluorescence-activated cell sorter<br />

(FACS), 270, 271<br />

Fluorescence microscope, 312<br />

Fluorescent in situ hybridization<br />

(FISH), 311<br />

for HER2/neu testing, 138–139<br />

Fluorodeoxyglucose, 293<br />

5-Fluorouracil (5FU)<br />

and DPD, 18–19<br />

DPD and, 30<br />

MTHFR and, 30, 31<br />

TS and, 31<br />

TS inhibition by, 19<br />

[ 18 F]MISO ([ 18 F]fluoromisonidazole-<br />

3-fluoro-1-(2 0 -nitro-1 0 -imidazolyl)-<br />

2-propanol), 299<br />

FNA. See Fine needle aspiration (FNA)<br />

Folate receptors, 316<br />

Food and Drug Administration<br />

(FDA), 153, 174, 246,<br />

311, 3111<br />

Formalin, 215, 218<br />

cross-linking effects <strong>of</strong>, 237<br />

Formalin-fixed paraffin-embedded<br />

(FFPE) tissue<br />

extraction <strong>of</strong> RNA from tissue samples<br />

<strong>of</strong>, 98–99, 244–245<br />

RNA-based expression pr<strong>of</strong>iling <strong>of</strong>, 246<br />

Forward genetics approach, 16<br />

FOXO regulatory proteins, 277<br />

Freeze-thawing cycles, 235<br />

Fulvestrant, 174<br />

Ga-DOTA-F(ab 0 )(2)-herceptin, 299<br />

Gamma rays, 292<br />

Gamma secretase inhibitor (GSI), 276<br />

Gefitinib, 172, 311<br />

Gene expression<br />

analysis, 177–178, 244–245<br />

pr<strong>of</strong>iling. See Gene expression pr<strong>of</strong>iling<br />

70-gene expression, 108, 124–126, 205.<br />

See also Gene expression signature<br />

in MINDACT trial, 130–132<br />

prognostic value <strong>of</strong>, 126–127<br />

in RASTER trial, 130<br />

in TRANSBIG study, 127–129<br />

Gene expression analysis, 177–178,<br />

244–245<br />

Gene expression Grade Index (GGI),<br />

200–201<br />

Gene expression pr<strong>of</strong>iling, 79–81, 97–98,<br />

198–199, 246, 270<br />

and chemotherapy, 100–101, 111–114<br />

in neoadjuvant setting, 101–102<br />

for classification <strong>of</strong> breast cancer,<br />

198–199<br />

and docetaxel, 112–113<br />

oncogenic pathway analysis by, 116–118<br />

pCR prediction, 101–102<br />

RNA expression, 98–100<br />

RS in. See Recurrence score (RS) assay<br />

RT-PCR for, 99<br />

Gene expression signatures, 265–267. See<br />

also Gene expression pr<strong>of</strong>iling<br />

for improving prognosis, 199–202<br />

integration with, 202–204<br />

Gene methylation<br />

BRCA1, 50–51<br />

estrogen receptors (ER), 51<br />

RASSF1A, 49–50


328 Index<br />

21-gene pr<strong>of</strong>ile. See Oncotype DX TM<br />

76-gene pr<strong>of</strong>ile, 129<br />

Genetic polymorphisms, 266–267<br />

Genetic susceptibility, 5<br />

Gene transcription, 19<br />

Genomic alterations<br />

prognostic value <strong>of</strong>, 158<br />

Genomic Health, 190, 191<br />

Genomic Health, Inc. (GHI), 246<br />

Genomic hypomethylation, 279<br />

Genomic predictors, 204–205<br />

Germline polymorphisms, 267<br />

Gestation, 269<br />

GHI. See Genomic Health, Inc. (GHI)<br />

Gliomas, 268<br />

Glutathione-S-transferases (GST), 16, 267<br />

and cyclophosphamide, 27–28<br />

in DFS, 33<br />

Glycophorin A, 156<br />

Graphical interface, user-friendly, 315<br />

GRP (gastrin-releasing peptide)<br />

receptors, 299<br />

GSK-3b, 273<br />

Harmonization, 214<br />

HAT. See Histone acetyltransferases<br />

(HAT)<br />

Health Insurance Portability and<br />

Accountability Act, 232<br />

Hedgehog (HH), 265<br />

ligands, 276–277<br />

signaling pathway, 276–277<br />

transcription factors, Gli1 and<br />

Gli2, 277<br />

H & E (hematoxylin & eosin), 227<br />

Hematopoietic cells, human, 156<br />

Heparin-coated syringes, 215<br />

HER2. See Human epidermal growth<br />

factor receptor 2 (HER2)<br />

HER2+, 198<br />

HER3, 171<br />

HER4, 171<br />

HER2/neu, 138–139, 171–172<br />

Heterochromatin, 47<br />

HH. See Hedgehog (HH)<br />

Histone acetyltransferases (HAT), 47<br />

Histone deacetylases (HDAC), 55<br />

Histones<br />

deacetylation, 47<br />

HAT, 47<br />

HDAC. See Histone deacetylases<br />

(HDAC)<br />

methylation <strong>of</strong>, 278<br />

modifiers, 48<br />

posttranslational modification <strong>of</strong>, 47<br />

ChIP for, 53<br />

H3K9 repressors, 278<br />

Homogenization, 224<br />

Hormonal therapy, 191<br />

adjuvant, 187<br />

Hormone receptors, 161<br />

Hormone replacement therapy<br />

(HRT), 173<br />

Human epidermal growth factor receptor<br />

2 (HER2). See also HER2/neu<br />

gene amplification in CTCs, 158<br />

Hybridization. See Comparative genomic<br />

hybridization (CGH)<br />

4-hydroxy-N-desmethyl tamoxifen,<br />

20, 170<br />

Hypothesis-based research, 185<br />

IHC. See Immunohistochemistry (IHC)<br />

Illumina, Inc., 249<br />

Illumina BeadArray, 250<br />

Illumina microarrays, 249<br />

Image file processing, 315<br />

Immunohistochemistry (IHC), 135–136,<br />

227, 311, 313<br />

for ERa, 136–138<br />

for HER2/neu, 138<br />

for PR, 136–138<br />

Immunohistological staining, 311–312<br />

Indium-111, 300<br />

In situ hybridization (ISH), 227<br />

International Stage IV Stratification<br />

Study, 160<br />

International Union Against <strong>Cancer</strong><br />

(UICC), 159<br />

Invasiveness gene signature (IGS), 202<br />

Iressa. See Gefitinib<br />

Isoniazid, 16–17


Index 329<br />

Java objects, 318<br />

Kaiser Permanente health care system,<br />

188<br />

Kaplan-Meier estimator, 245<br />

Krishnamurthy, Savitri, 154<br />

Laser capture microdissection (LCM),<br />

2–3, 221–224, 239<br />

Lauterbur, Paul, 294<br />

LCM. See Laser capture microdissection<br />

(LCM)<br />

Lectin-carbohydrate, 316<br />

Lectins, 316<br />

Lethal phenotypes, 310<br />

Letrozole, 32, 174<br />

chemical structure <strong>of</strong>, 175<br />

vs. tamoxifen, 176, 177<br />

Ligand-binding domain (LBD), 170<br />

Liver tumors, 297<br />

Lobular carcinoma in situ (LCIS), 174<br />

Loss <strong>of</strong> heterozygosity (LOH), 66, 250<br />

Luminal epithelial cells, genes <strong>of</strong>, 198<br />

Lymph nodes, 297<br />

microarrays for, 3–5<br />

recurrence <strong>of</strong> breast cancer in, 245<br />

Macromolecular extraction, 237, 239<br />

Magnetic labeling and separation, <strong>of</strong><br />

tumor tissue, 219–220<br />

MammaPrint. See 70-gene expression<br />

Mammary gland, 264<br />

Mammography, 292<br />

Manganese superoxide dismutase<br />

(MnSOD), 29<br />

in DFS, 36<br />

MAPK. See Mitogen-activated protein<br />

kinase (MAPK)<br />

MassARRAY, 52<br />

Matlab slide scanners, 315<br />

Matrix metalloproteinase (MMP), 316<br />

MBC. See Metastatic breast cancer<br />

(MBC)<br />

MCF-7 breast cancer cells, 170, 172<br />

Medicare, 190<br />

Mercaptopurine, 17<br />

Metaphase chromosomes, in CGH, 65<br />

Metastatic breast cancer (MBC)<br />

development <strong>of</strong>, 151<br />

diagnosis <strong>of</strong>, 158–159<br />

stage IV, 160<br />

HER2-positive, 177, 178<br />

staging, 159–160<br />

Methotrexate (MTX), 18, 30, 31<br />

Methylation<br />

detection <strong>of</strong>, 51–53<br />

<strong>of</strong> DNA. See DNA methylation<br />

<strong>of</strong> genes. See Gene methylation<br />

pr<strong>of</strong>iling, 250–251<br />

therapy, 53–55<br />

Methyltetrahydr<strong>of</strong>olate reductase<br />

(MTHFR), 18<br />

in DFS, 33<br />

and 5-Fluorouracil (5FU), 30, 31<br />

Microarray for Node-Negative Disease<br />

may Avoid Chemotherapy<br />

(MINDACT) trial, 130–132, 205<br />

Microarray Quality Control Project<br />

(MAQC), 82<br />

Microarrays<br />

for lymph node status, 3–5<br />

with OCT, 236<br />

with paraffin-embeded tissue, 236–237<br />

RNA expression, 2, 4<br />

with RNA stabilization buffers,<br />

237–238<br />

technical reproducibility <strong>of</strong>, 82–83<br />

using CNB, 238–239<br />

using FNA, 238<br />

using LCM, 239<br />

Micrometastatic carcinoma, 154<br />

Micrometastatic cells, 206<br />

microRNAs, 206<br />

Micros<strong>of</strong>t Excel, 228<br />

MIP. See Molecular inversion probe (MIP)<br />

miRNA pr<strong>of</strong>iling, 248–249<br />

Mitogen-activated protein kinase<br />

(MAPK), 172<br />

MMP. See Matrix metalloproteinase<br />

(MMP)<br />

MMP-2, 316


330 Index<br />

MnSOD. See Manganese superoxide<br />

dismutase (MnSOD)<br />

Molecular classification, <strong>of</strong> breast<br />

cancer, 3<br />

Molecular diagnostic test, 186, 188,<br />

190, 311<br />

Molecular epidemiologic studies, 32–36<br />

Molecular imaging, 291–292. See also<br />

Positron emission tomography<br />

(PET)<br />

Molecular inversion probe (MIP), 66–69<br />

PCR in, 67<br />

Molecular markers, 79, 160, 188–190<br />

defined, 185–186<br />

development process <strong>of</strong>, 81–82<br />

in stem cells, 265, 273<br />

Molecular pr<strong>of</strong>iling<br />

defined, 185<br />

developmental technologies for, 311<br />

Molecular signatures, 186<br />

Monoclonal antibodies, 153, 156<br />

Monocytes, 219<br />

MPO. See Myeloperoxidase (MPO)<br />

MRI, 294–297<br />

applications for, 297<br />

mRNAs, 245<br />

MTA-1 (Manual Tissue Arrayer), 226<br />

MTHFR. See Methyltetrahydr<strong>of</strong>olate<br />

reductase (MTHFR)<br />

mTOR (mammalian targets <strong>of</strong><br />

rapamycin), 276<br />

MTX. See Methotrexate<br />

Mutations<br />

estrogen receptors, 170<br />

Myeloperoxidase (MPO), 29<br />

N-acetyltransferases, 16, 267<br />

Nanoparticles. See also Nanotechnology<br />

bioconjugated, 311, 317<br />

design and development <strong>of</strong>, 318<br />

molecular pr<strong>of</strong>iling. See Nanotyping<br />

superparamagnetic iron oxide, 310<br />

Nanotechnology, 311<br />

for cancer detection, diagnosis, and<br />

therapy, 319<br />

QD, 312<br />

Nanotherapeutics, 315–317<br />

Nanotyping, 311–315, 318–319<br />

National <strong>Cancer</strong> Institute (NCI), 186,<br />

207, 317<br />

National Institutes <strong>of</strong> Health (NIH),<br />

199, 200<br />

cancer guidelines, 205<br />

National Surgical Adjuvant <strong>Breast</strong> and<br />

Bowel Project (NSABP), 99–100,<br />

188, 190, 192, 245<br />

N-CoR, 171<br />

Neoadjuvant chemotherapy, 5–7, 78,<br />

109–111<br />

cytotoxic agents in, 114–115<br />

predictive markers <strong>of</strong> response to,<br />

101–102<br />

response predictors <strong>of</strong>, 78–79<br />

use <strong>of</strong>, 293<br />

Netherlands <strong>Cancer</strong> Institute (NKI), 199<br />

.NET technology, 318<br />

Noncanonical pathways. See Wnt<br />

signaling<br />

Nonnucleoside inhibitors, 53, 54<br />

Non-small cell lung cancer (NSCLC), 310<br />

Nontumorigenic cells, 265<br />

North American <strong>Breast</strong> Intergroup, 192<br />

Notch receptors, 276<br />

Notch signaling pathway, 264, 276<br />

Nottingham Prognostic Index, 205<br />

NSABP. See National Surgical Adjuvant<br />

<strong>Breast</strong> and Bowel Project<br />

(NSABP)<br />

Nucleic acids<br />

amplification, 238<br />

collection, 215<br />

extraction, 237<br />

pr<strong>of</strong>ilation, 251<br />

Nucleoside inhibitors, 53, 54<br />

NuGEN Technologies, Inc., 249<br />

OCT. See Optimal cutting temperature<br />

(OCT)<br />

OCT-4, 271<br />

Octreoscan, 299<br />

[ 15 O]H 2 O PET, 292<br />

4OHT (4-hydroxytamoxifen), 171<br />

Oligonucleotide-based arrays,<br />

in CGH, 65–66


Index 331<br />

Oligonucleotides, 247<br />

Oncogenes, 62, 251<br />

Oncogenic pathways, deregulation <strong>of</strong>,<br />

116–118<br />

cytotoxic agents, 118<br />

Oncologists, 207<br />

Oncotype DX assay, 188, 190, 192, 194,<br />

200, 205, 246<br />

Optimal cutting temperature (OCT),<br />

218, 236<br />

p53<br />

signature genes, 201<br />

tumor suppressor, 248<br />

PACCT. See Program for the assessment<br />

<strong>of</strong> clinical cancer tests (PACCT)<br />

Paclitaxel, 19–20, 267<br />

CYP2C8 in, 31<br />

Paget, Stephen, 278<br />

Papanicolaou (Pap)-stained cytospin, 154<br />

Paraffin, 222, 227<br />

Paraffin-embedded tissue, 237–238<br />

Paraffin wax, discovery <strong>of</strong>, 237<br />

Partial signatures, 204<br />

Pathologic report, 234<br />

Pathologist, 232, 233<br />

Patient identifiers, 232<br />

PAXgene, 220–221<br />

p21 Cip1/Waf1 , 277<br />

PCR. See Polymerase chain reaction<br />

(PCR); Polymerase chain reaction<br />

(PCR), in MIP<br />

pCR. See Complete pathological response<br />

(pCR)<br />

Peptides, 299<br />

Peripheral blood<br />

cancer stem cells in, 152<br />

extraction <strong>of</strong> RNA from, 220–221<br />

isolation <strong>of</strong> monocytes in, 219<br />

separation from mononuclear cells, 156<br />

Peripheral blood mononuclear cells<br />

(PBMC), 219<br />

PET. See Positron emission tomography<br />

(PET)<br />

PET-MRI scanner, 300<br />

PET scanners, 297<br />

<strong>Ph</strong>armacodynamics, 13<br />

<strong>Ph</strong>armacoepidemiology, 14<br />

<strong><strong>Ph</strong>armacogenetics</strong>, 14, 206<br />

molecular epidemiologic studies, 32–36<br />

<strong>Ph</strong>armacogenomic predictors, 79–80<br />

published studies, 86–90<br />

validation studies <strong>of</strong>, 84–85<br />

<strong>Ph</strong>armacokinetics, 13<br />

<strong>Ph</strong>armacology, 13–14<br />

<strong>Ph</strong>osphatase and tensin homolog deleted<br />

on chromosome ten (PTEN),<br />

265, 277<br />

<strong>Ph</strong>osphate buffered saline (PBS), 215<br />

<strong>Ph</strong>osphatidylinositol-3-OH kinase (PI3K),<br />

173, 277<br />

Polycomb group (PcG) proteins, 46,<br />

261, 279<br />

Polymerase chain reaction (PCR), 236<br />

detection <strong>of</strong> tumor cells, 158<br />

in MIP, 67<br />

Positron emission tomography (PET),<br />

206, 292<br />

Posttranslational modification,<br />

<strong>of</strong> histones, 47<br />

PR. See Progesterone receptor (PR)<br />

Predictive markers<br />

defined, 136<br />

in drug development, 141–143<br />

ERa, 136–138<br />

HER2/neu, 138–139<br />

PR, 136–138<br />

Preoperative chemotherapy. See<br />

Neoadjuvant chemotherapy<br />

Preoperative systemic chemotherapy<br />

(PST). See Neoadjuvant<br />

chemotherapy<br />

Preregistration, 191<br />

Preservatives, 214, 215<br />

Primary study group, 192<br />

Privacy rights, 231<br />

Progenitor cells, mutated, 262, 266<br />

Progesterone receptor (PR), 136–138,<br />

176–177<br />

Prognosis, in breast cancer<br />

improving, 199–202<br />

Prognostic marker, 199<br />

Program for the assessment <strong>of</strong> clinical<br />

cancer tests (PACCT), 186<br />

Proliferation genes, 201, 202, 204


332 Index<br />

Prostate cancer, RNA signatures in, 248<br />

Proteinase digestion buffers, 237<br />

Protein biomarkers, 313. See also <strong>Cancer</strong><br />

biomarkers<br />

Protein collection, 215<br />

Protein isolation, 225–226<br />

Proteomics, 206<br />

PTEN. See <strong>Ph</strong>osphatase and tensin<br />

homolog deleted on chromosome<br />

ten (PTEN)<br />

PTEN pathway, 277–278<br />

Pyrosequencing, 52–53<br />

QD-Ab, 312<br />

QDs. See Quantum dots (QDs)<br />

Q-IHC tools, 314<br />

qRT-PCR (realtime quantitative-PCR)<br />

for screening tumor samples for<br />

miRNA expression, 248<br />

Quantification system, 315<br />

Quantum dots (QDs)<br />

multicolor, 312<br />

nanotechnology, 312<br />

semiconductor, 311<br />

Radiation, resistance to<br />

(Radioresistance), 266<br />

Radionuclide techniques, 292–294<br />

vs. optical imaging, 299<br />

RAKE (RNA-primed array-based Klenow<br />

extension) analysis study, 249<br />

Randomization, 192<br />

Rapamycin, molecular targets <strong>of</strong>. See<br />

MTOR (mammalian targets <strong>of</strong><br />

rapamycin)<br />

RASSF1A, 49–50<br />

RASTER trial, 130<br />

Rationally designed inhibitors, 53,<br />

54, 55<br />

Reactive oxygen species (ROS), 27<br />

and mitochondrial permeability, 29<br />

Recurrence score (RS), 100, 200, 245<br />

and adjuvant chemotherapy, 100–101<br />

clinical utility <strong>of</strong>, 103–104<br />

GHI, 246<br />

for node positive patients, 104<br />

[Recurrence score (RS)]<br />

and pCR, 101<br />

to predict recurrence <strong>of</strong> breast<br />

cancer, 245<br />

vs. AOL, 103–104<br />

REMARK guidelines, 186<br />

Reporter genes, 299<br />

Response predictors, <strong>of</strong> neoadjuvant<br />

chemotherapy, 5–7, 78–79<br />

molecular markers as, 79<br />

Reticuloendothelial system (RES), 315<br />

Reverse transcriptase-polymerase chain<br />

reaction (RT-PCR), 99, 244<br />

assay-based evaluation <strong>of</strong> ER<br />

expression, 189<br />

and gene signature, 200<br />

RNA<br />

amplification, 237–238<br />

degradation, 235, 247<br />

expression microarray, 2, 4<br />

hybridization, 238, 239<br />

isolation, 224–225<br />

RNAlater, 218, 236<br />

RNA stabilization buffers, 236–237<br />

RosetteSep TM method, 156<br />

Rotterdam group, 199<br />

RPMI (Roswell Park Memorial Institute)<br />

media, 218<br />

RS. See Recurrence score (RS) assay<br />

RT-PCR assay, 245<br />

Sample handling, 237<br />

Selective ER downregulators (SERDs),<br />

173, 174<br />

Selective ER modulators (SERMs), 173<br />

Semiautomatic image processing, 315<br />

Single nucleotide polymorphism (SNP),<br />

14–15, 249<br />

pr<strong>of</strong>iling, 250<br />

Single-photon imaging, 292<br />

SMRT, 171<br />

SNP. See Single nucleotide polymorphism<br />

(SNP)<br />

Sonic hedgehog (SHH), 276, 277<br />

SRC-1, 171<br />

St. Gallen classification, 246<br />

St. Gallen criteria, 199, 200


Index 333<br />

St. Gallen guidelines, 205<br />

Standard operating procedures (SOPs)<br />

for collection <strong>of</strong> tissue samples, 234<br />

for RNA degradation, 235<br />

STAR (Study <strong>of</strong> Tamoxifen and<br />

Raloxifene) trial, 173<br />

STAT3 transcription factor, 276<br />

Stem cell lines, derivation <strong>of</strong>, 266<br />

Stem cell niche, 261, 263, 268<br />

molecular targets <strong>of</strong>, 278–280<br />

Stem cell regulation, 280<br />

Stem cells<br />

cellular proliferation, 261<br />

clinical research on, 264–265<br />

differentiation <strong>of</strong>, 261, 262<br />

division <strong>of</strong>, 276<br />

drug resistance <strong>of</strong>, 268<br />

embryonic, 271, 279<br />

functional targets, 269–273<br />

gene expression pr<strong>of</strong>iling <strong>of</strong>, 270<br />

hematopoietic, 278, 280<br />

hypothesis, 260–264, 268<br />

intestinal, 277<br />

leukemic, 278<br />

mesenchymal, 280<br />

molecular markers, 265, 273<br />

mutations, 263, 270, 278<br />

nontumorigenic daughter cells <strong>of</strong>, 261<br />

Stem cell sanctuaries, 267–269<br />

Stem cell survival, neural, 279, 280<br />

Stochastic model, 260, 261<br />

SULT1A1<br />

in DFS, 36<br />

in TAM, 32<br />

Superoxide dismutases (SODs), 267<br />

Swiss Institute <strong>of</strong> Bioinformatics, 202<br />

TAM. See Tamoxifen (TAM)<br />

Tamoxifen (TAM), 20, 31–32, 56, 137<br />

adjuvant, 169<br />

CYP2D6 in, 32<br />

metabolizers <strong>of</strong>, 170<br />

SULT1A1 in, 32<br />

UGT2B15 in, 32<br />

vs. aromatase inhibitors, 176–177<br />

vs. DES, 168<br />

TaqMan TM array, 248–249<br />

Taxanes, 19–20<br />

CYP2C8 in, 31<br />

Thiopurine methyltransferase (TPMT),<br />

17–18<br />

Thymidylate synthase (TS), 19<br />

Tissue<br />

acquisition <strong>of</strong>, 232–235<br />

biopsies, 218<br />

collection, 214–218<br />

control blocks, 228<br />

defined, 213<br />

processing, 218<br />

procurement, 214<br />

Tissue banking, 231–232<br />

Tissue microarray (TMA), 226–227<br />

blocks, 228, 229, 230<br />

mapping, 227, 228<br />

TMA. See Tissue microarray (TMA)<br />

TRANSBIG study, 127–129, 199–200<br />

Transcriptional regulation, 277<br />

Trastuzumab, 138, 172, 310–311<br />

Trial Assigning Individualized Options for<br />

Treatment (TAILORx) trial, 178, 186<br />

designing <strong>of</strong>, 191<br />

recurrence score ranges in, 192–194<br />

TRI reagent<br />

for isolation <strong>of</strong> DNA, 225–226<br />

for isolation <strong>of</strong> RNA, 224–225<br />

for tissue microarray creation, 226–230<br />

TRIZOL method, 224–230<br />

TS. See Thymidylate synthase (TS)<br />

Tumor-activated prodrug therapy, 316<br />

Tumor growth factor-b (TGF-b), 280<br />

Tumor hypoxia, 299<br />

Tumorigenesis, 250, 273, 274<br />

Tumors<br />

detection <strong>of</strong>, 158<br />

invasion, 202<br />

regression, 260<br />

size and growth, 238, 239<br />

Tumor suppressor genes, 251<br />

Tumor tissue. See Tissue<br />

UGT2B15<br />

in DFS, 36<br />

in TAM, 32<br />

Ultrasound, 292


334 Index<br />

Vacutainers, use <strong>of</strong>, 215<br />

Vascular endothelial growth factor<br />

(VEGF), 279, 280<br />

Vascularization, 315<br />

VEGF. See Vascular endothelial growth<br />

factor (VEGF)<br />

White blood cells. See Monocytes<br />

Whole-genome expression pr<strong>of</strong>iling, 249<br />

Wnt/b-catenin, 272<br />

Wnt signaling, 266, 273<br />

World Wide Web for <strong>Cancer</strong> Research,<br />

317<br />

Wound-healing, gene phenotype, 200<br />

WWOX expression, 56<br />

Xenograft model, 265<br />

Xenotransplantation, 264, 269<br />

X-ray mammography, 297<br />

Xylene, 218


Oncology<br />

about the book…<br />

<strong><strong>Ph</strong>armacogenetics</strong> is becoming increasingly relevant in the diagnosis, treatment,<br />

and recovery <strong>of</strong> cancer patients. A major problem facing oncologists is the outstanding<br />

varied efficacy <strong>of</strong> treatment. Promising advances in pharmacogenetics<br />

have allowed the development <strong>of</strong> effective agents which will enable personalized<br />

cancer chemotherapy to become routine for the clinical practice.<br />

Written by experts in the field and combining information that is unable to be<br />

found in a single source volume, <strong>Ph</strong>armocogenetics <strong>of</strong> <strong>Breast</strong> <strong>Cancer</strong>:<br />

• combines a complete overview <strong>of</strong> pharmacogenetics and how it relates to<br />

breast oncology for diagnosis, treatment, and the recovery <strong>of</strong> patients using<br />

an individualized therapy model<br />

• is the first single source reference demonstrating the application <strong>of</strong><br />

pharmacogenetics in the care <strong>of</strong> breast cancer patients<br />

• enables physicians a coherent interpretation <strong>of</strong> the emerging science <strong>of</strong><br />

pharmacogenetics, aiding them to incorporate individual therapies in their<br />

own practice<br />

• gives practical guidance on various forms <strong>of</strong> specimen collection, tissue<br />

selection, and handling procedures<br />

about the editor...<br />

BRIAN LEYLAND-JONES is Associate Vice President and Director, Emory Winship<br />

<strong>Cancer</strong> Institute, Emory University, Atlanta, Georgia. Dr. Leyland-Jones’ main<br />

research interests are pharmacodynamics, pharmacokinetics, and pharmacogenetics<br />

in oncological clinical trials; translation <strong>of</strong> preclinical models into the clinic;<br />

biomarker endpoints in <strong>Ph</strong>ase I/II clinical trials; and screening and mechanistic<br />

studies <strong>of</strong> novel targeted and chemotherapeutic anti-cancer agents. He has<br />

authored more than 125 peer-reviewed articles and book contributions, 150<br />

abstracts, and 29 patents.<br />

Printed in the United States <strong>of</strong> America<br />

H8637

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!