Sie sind auf Seite 1von 35

2012 Collections Booklet

SPECIALSECTION
C
R
E
D
I
T
:

A
N
N
E

W
E
S
T
O
N
/
C
A
N
C
E
R

R
E
S
E
A
R
C
H

U
K
/
V
I
S
U
A
L
S

U
N
L
I
M
I
T
E
D

I
N
C
.

CONTENTS
News
Cancer kesearch and
the :90 8llon Vetaphor
U.S. Cancer Trends
Conbnng Iargeted urugs
to Stop kesstant Iunors
Can Ireatnent Costs 8e Ianed
A lush to lght Cancer
n the uevelopng world
8rothers n Arns Aganst Cancer
Reviews
Lxplorng the uenones of
Cancer Cells: lrogress and lronse
M. R. Stratton
A lerspectve on Cancer
Cell Vetastass
C. L. Chaffer and R. A. Weinberg
Cancer lnnunoedtng:
lntegratng lnnuntys koles n
Cancer Suppresson and lronoton
R. D. Schreiber et al.
See also Editorial p. 1491; Report p. 1612; and Science
Translational Medicine including Research Article by L. Sequist
et al., Science Signaling, Science Careers, Video, and Science
Podcast at www.sciencemag.org/special/cancer2011/.
Cancer Crusade
at 40
I NT RODUCT I ON
Celebratng
an Annversary
In this issue of Science, we commemorate the 40th anniversary of the U.S.
National Cancer Act, which provided a massive stimulus for cancer research.
At the start of this Cancer Crusade, researchers were already tackling some
tough questions, as reected in papers published by Science in 1971. Among
them: How do abnormalities in chromosome number arise in tumor cells? Can
tissue-specic markers be used to determine the epithelial versus mesenchymal
origin of a solid tumor? Can the immune system be manipulated so that it recog-
nizes tumor cells as foreign invaders that must be eliminated from the body? Do
viruses play a role in human cancer?
Skeptics might argue that 40 years later, cancer researchers continue to grap-
ple with the same questions. Perhaps theres some truth in this. But our hope is
that the selection of articles in this special section of Science will explain why
many of these questions have proved so challenging and, more importantly, how
contemporary cancer research is providing a clearer view of the biology that will
lead to answers. Stratton (p.) discusses international efforts to sequence
the complete genomes of a wide range of human tumor types and the impact
that this sequence information is anticipated to have on our understanding of
cancer biology as well as our ability to detect, diagnose, and treat the disease. A
working model for cancer cell metastasis is presented by Chaffer and Weinberg
(p. ), who highlight the important role of cancer stem cells and a develop-
mental process called the epithelial-mesenchymal transition. Schreiber et al. (p.
) describe cancer immunoediting, a conceptual framework that integrates
the immune systems dual roles in inhibiting and promoting cancer growth.
News reports examine other signicant challenges for the eld: Jocelyn
Kaiser (p. ) describes how even the best new drugs eventually seem to fail
and how they might be made more effective. David Malakoff (p. ) outlines a
key social issue: the fast-rising cost of care. Martin Enserink (p. ) reports on
efforts to close the huge disparity between cancer treatment in rich and develop-
ing countries. And Mitch Leslie (p. ) describes how researchers are taking
a new look at the role of p53 and the related proteins p63 and p73 in tumors.
A video report by Robert Frederick appears online at www.sciencemag.
org/special/cancer2011/, along with links to additional reading material.
Science Careers features an article on cancer clinical trials training (http://scim.
ag/cancertrialstraining) and a Q&A with Memorial Sloan-Kettering clinician-
investigator David Solit (http://scim.ag/solitqanda).
It is worth noting that at least one of the questions that concerned cancer
researchers writing in Science back in 1971 has been denitively answered. We
now know that viruses do in fact play a causal role in certain human cancers,
and, thanks to decades of tumor virology research, vaccines against these viruses
have been developed into successful cancer-preventive agents. Thats something
to celebrate.
PAULA KIBERSTIS AND ELIOT MARSHALL
1539 www.sciencemag.org SCIENCE VOL 331 25 MARCH 2011
0325SpecialIntropage.indd 1539 3/18/11 4:42 PM
See also Science Translational Medicine including Research Article
by L. Sequist et al., Science Signaling, Science Careers, Video, and
Science Podcast at www.sciencemag.org/special/cancer2011/.
www.sciencemag.org SCIENCE 1
25 MARCH 2011 VOL 331 SCIENCE www.sciencemag.org 1540
CANCER CCCCCCCCCCAAAAAAAAAANNNNNNNNNNCCCCCCCCCCEEEEEEEEEERRRRRRRRRR
Crusade at 40
U.S. Cancer Trends
SOURCE FOR CANCER STATISTICS: NCI SURVEILLANCE RESEARCH PROGRAM; PHOTOS (LEFT TO RIGHT): LINDA BARTLETT/NCI; BILL BRANSON/NCI; NCI
0
10
20
30
40
50
60
70
80
75 77 79 81 83 85 87 89 91 93 95 97 99 01 03 05 07
Lung & Bronchus
The good news: Lung cancer incidence among
men began to decline in the early 1980s; the death
rate, in the early 1990s. This shows the power of
preventionthrough a campaign to curb tobacco
use. The bad news: The death rate is far higher for
African-American men than for white men, and
deaths among women (of all races) continued to
climb until recent years.
Incidence per
100,000
Cancer Research and
The $90 Billion Metaphor
THERE NEVER WAS AN OFFICIAL WAR ON CANCER. THAT PHRASE FROM NEWS
reports and debates attached itself to the U.S. program that began when Presi-
dent Richard Nixon signed the National Cancer Act in December 1971. The
law made big promises and gave the U.S. National Cancer Institute (NCI) a
token measure of independence. It also encouraged the idea that cancer could
be targeted, like a trip to the moon, and cured. The law was important for
research, RAND medical historian and policy analyst Richard Rettig has writ-
ten: It stopped a decline in NCIs budget.
This reversal began in Congress. Urged on by health activists such as Mary
Lasker, leading Democrats in 1970 adopted curing cancer as their cause. Aware
that it might become a national issue, Nixon embraced it, too. The resulting leg-
islation raised NCIs budget almost overnight by 23% to $233 million; NCIs
funding has continued to climb since then to more than $5 billion per year
although in recent years ination has grown faster. Since the 1971 act, NCI has
spent about $90 billion on science, treatment, and prevention of cancer.
The war metaphor remains part of this legacyand a target for skeptics. In
1986, former NCI biostatistician John Bailar stirred controversy with a bleak
analysis in The New England Journal of Medicine, noting that cancer inci-
dence and mortality rates hadnt changed fundamentally in 15 years. He sug-
gested that the nation was losing the war against cancer. Sweeping declara-
tions continued. In 2003, thenNCI Director Andrew von Eschenbach set a
goal of ending suffering and death from cancer by 2015. Such broad claims
invite balloon-pricking.
But if one sets aside the rhetoric, says Allen Lichter, executive director of
the American Society of Clinical Oncology, its evident that the cancer cam-
paign has changed therapy and saved lives (see indicators for the seven dead-
liest cancers, right). It is true that for certain cancersof the pancreas, brain,
and liver, for examplethe picture has not improved. But the overall U.S. can-
cer mortality rate began to decline in the 1990s. And clinical care looks noth-
ing like it did 40 years ago, says Lichter, who began training as an oncolo-
gist in the 1970s. He speaks of a revolution that took lessons from the 1960s
advances against childhood leukemia to develop adjuvant therapy: the crazy
idea that chemotherapy should be given to patients in remission to treat pre-
sumed microscopic disease that has spread. His list of benets continues with
breast-conserving and microscopic surgery, imaging for diagnosis and disease
management, molecular analysis of tumors and targeted drug therapy, longer
survival times, and much better palliative care. Looking for progress in cancer
can feel like watching the hands of a clock, Lichter admits. But things are
denitely moving in the right direction.
ELIOT MARSHALL
Deaths per 100,000
50
60
70
0
10
20
30
40
75 77 79 81 83 85 87 89 91 93 95 97 99 01 03 05 07
Incidence per 100,000
Deaths per 100,000
Colon & Rectum
The U.S. National Cancer Institute (NCI) spotlighted
declining colorectal cancer trends in its 2010
Annual Report to the Nation on the Status of Can-
cer. Improved diet and screening with colonoscopy,
among other early-detection techniques, are help-
ing to control the second-deadliest cancer. NCIs
modeling predicts that overall mortality could drop
50% by 2020.
40 Years of the War on Cancer
1971
President Richard
Nixon signs the
National Cancer
Act promoting the
National Cancer
Institute.
1973
NCI launches
Surveillance
Epidemiology
and End Results
program to
collect U.S.
cancer data.
1978
Clinical test-
ing begins of
interferon-,
the rst bio-
logical cancer
therapy.
1980
Robert Gallo and others isolate
human T-cell
lymphotro-
pic virus-1,
a cause of
cancer.
2010
Estimated
Deaths 157,300
2010
Estimated
Deaths 51,370
FDA approves
tamoxifen to
prevent breast
cancer recurrence.
1979
Researchers
discover p53,
the mutated
gene most often
seen in tumors.
5-Year
Mortality
Trend
-1.6%
5-Year
Mortality
Trend
-3.0%
0325SpecialNewsSection.indd 1540 3/17/11 3:38 PM
NEWS
2 SCIENCE www.sciencemag.org
25 MARCH 2011 VOL 331 SCIENCE www.sciencemag.org 1540
CANCER CCCCCCCCCCAAAAAAAAAANNNNNNNNNNCCCCCCCCCCEEEEEEEEEERRRRRRRRRR
Crusade at 40
U.S. Cancer Trends
SOURCE FOR CANCER STATISTICS: NCI SURVEILLANCE RESEARCH PROGRAM; PHOTOS (LEFT TO RIGHT): LINDA BARTLETT/NCI; BILL BRANSON/NCI; NCI
0
10
20
30
40
50
60
70
80
75 77 79 81 83 85 87 89 91 93 95 97 99 01 03 05 07
Lung & Bronchus
The good news: Lung cancer incidence among
men began to decline in the early 1980s; the death
rate, in the early 1990s. This shows the power of
preventionthrough a campaign to curb tobacco
use. The bad news: The death rate is far higher for
African-American men than for white men, and
deaths among women (of all races) continued to
climb until recent years.
Incidence per
100,000
Cancer Research and
The $90 Billion Metaphor
THERE NEVER WAS AN OFFICIAL WAR ON CANCER. THAT PHRASE FROM NEWS
reports and debates attached itself to the U.S. program that began when Presi-
dent Richard Nixon signed the National Cancer Act in December 1971. The
law made big promises and gave the U.S. National Cancer Institute (NCI) a
token measure of independence. It also encouraged the idea that cancer could
be targeted, like a trip to the moon, and cured. The law was important for
research, RAND medical historian and policy analyst Richard Rettig has writ-
ten: It stopped a decline in NCIs budget.
This reversal began in Congress. Urged on by health activists such as Mary
Lasker, leading Democrats in 1970 adopted curing cancer as their cause. Aware
that it might become a national issue, Nixon embraced it, too. The resulting leg-
islation raised NCIs budget almost overnight by 23% to $233 million; NCIs
funding has continued to climb since then to more than $5 billion per year
although in recent years ination has grown faster. Since the 1971 act, NCI has
spent about $90 billion on science, treatment, and prevention of cancer.
The war metaphor remains part of this legacyand a target for skeptics. In
1986, former NCI biostatistician John Bailar stirred controversy with a bleak
analysis in The New England Journal of Medicine, noting that cancer inci-
dence and mortality rates hadnt changed fundamentally in 15 years. He sug-
gested that the nation was losing the war against cancer. Sweeping declara-
tions continued. In 2003, thenNCI Director Andrew von Eschenbach set a
goal of ending suffering and death from cancer by 2015. Such broad claims
invite balloon-pricking.
But if one sets aside the rhetoric, says Allen Lichter, executive director of
the American Society of Clinical Oncology, its evident that the cancer cam-
paign has changed therapy and saved lives (see indicators for the seven dead-
liest cancers, right). It is true that for certain cancersof the pancreas, brain,
and liver, for examplethe picture has not improved. But the overall U.S. can-
cer mortality rate began to decline in the 1990s. And clinical care looks noth-
ing like it did 40 years ago, says Lichter, who began training as an oncolo-
gist in the 1970s. He speaks of a revolution that took lessons from the 1960s
advances against childhood leukemia to develop adjuvant therapy: the crazy
idea that chemotherapy should be given to patients in remission to treat pre-
sumed microscopic disease that has spread. His list of benets continues with
breast-conserving and microscopic surgery, imaging for diagnosis and disease
management, molecular analysis of tumors and targeted drug therapy, longer
survival times, and much better palliative care. Looking for progress in cancer
can feel like watching the hands of a clock, Lichter admits. But things are
denitely moving in the right direction.
ELIOT MARSHALL
Deaths per 100,000
50
60
70
0
10
20
30
40
75 77 79 81 83 85 87 89 91 93 95 97 99 01 03 05 07
Incidence per 100,000
Deaths per 100,000
Colon & Rectum
The U.S. National Cancer Institute (NCI) spotlighted
declining colorectal cancer trends in its 2010
Annual Report to the Nation on the Status of Can-
cer. Improved diet and screening with colonoscopy,
among other early-detection techniques, are help-
ing to control the second-deadliest cancer. NCIs
modeling predicts that overall mortality could drop
50% by 2020.
40 Years of the War on Cancer
1971
President Richard
Nixon signs the
National Cancer
Act promoting the
National Cancer
Institute.
1973
NCI launches
Surveillance
Epidemiology
and End Results
program to
collect U.S.
cancer data.
1978
Clinical test-
ing begins of
interferon-,
the rst bio-
logical cancer
therapy.
1980
Robert Gallo and others isolate
human T-cell
lymphotro-
pic virus-1,
a cause of
cancer.
2010
Estimated
Deaths 157,300
2010
Estimated
Deaths 51,370
FDA approves
tamoxifen to
prevent breast
cancer recurrence.
1979
Researchers
discover p53,
the mutated
gene most often
seen in tumors.
5-Year
Mortality
Trend
-1.6%
5-Year
Mortality
Trend
-3.0%
0325SpecialNewsSection.indd 1540 3/17/11 3:38 PM
Breast (female)
Efforts to detect and treat invasive breast cancer partly explain why
incidence rose dramatically in the 1980s, reaching a peak for all
races in 1999. Death rates from breast cancer have been declining
steadily since 198990, although 5-year survival continues to be far
higher for whites than African Americans.
0
30
90
120
150
60
757779 8183 85 878991 93 95 9799 0103 05 07
Incidence per 100,000
Deaths per 100,000
Pancreas
In part because it is difcult to detect early, pancreatic cancer
remains the fourth-deadliest cancer, and incidence and mortality
have hardly budged. The average patient diagnosed with advanced
disease will live only 6 months.
10
12
14
0
2
4
6
8
757779 8183 85 878991 93 95 9799 0103 05 07
Incidence per 100,000
Deaths per 100,000
2010
Estimated
Deaths 36,800
0
40
80
120
160
200
240
757779 8183 85 878991 93 95 9799 0103 05 07
Deaths per 100,000
Incidence
per 100,000
Leukemia
Death rates for the four main types of leukemia have slowly
declined, thanks largely to treatments that combine chemotherapy
drugs. The survival rate is now about 80% for childhood acute
lymphoblastic leukemia.
10
12
14
16
2
0
4
6
8
757779 8183 85 878991 93 95 9799 0103 05 07
Incidence per 100,000
Deaths per 100,000
SOURCE FOR CANCER STATISTICS: NCI SURVEILLANCE RESEARCH PROGRAM; PHOTOS (LEFT TO RIGHT): ISTOCKPHOTO; THINKSTOCK; NCI; PAUL SAKUMA/AP
1983
Researchers create
severe combined
immunodecient
mice, a model for
cancer research.
1989
Nobel Prize for
discovering the rst
proto-oncogene
(Src) awarded to
Harold Varmus and
Michael Bishop.
1981
First cancer-
prevention vaccine
introduced
against human
hepatitis B virus.
2010
Estimated
Deaths 39,840
2010
Estimated
Deaths 32,050
2010
Estimated
Deaths 21,840
1985
Randomized
trial shows that
lumpectomy plus
radiation are
as effective as
mastectomy for
breast cancer.
1986
Biostatistician John
Bailar writes in The
New England Jour-
nal of Medicine, We
are losing the war
against cancer.
5-Year
Mortality
Trend
-2.2%
5-Year
Mortality
Trend
+0.6%
5-Year
Mortality
Trend
-3.3%
5-Year
Mortality
Trend
-1.3%
Liver
Mortality and incidence for liver and bile duct cancers have risen
steadily, linked to infections with hepatitis B and C, which are top risk
factors, along with alcohol abuse. Because tumors usually cannot be
removed with surgery, post-diagnosis survival is brief.
5
6
7
8
0
1
2
3
4
757779 8183 85 878991 93 95 9799 0103 05 07
Incidence per 100,000
Deaths per 100,000
2010
Estimated
Deaths 18,910
5-Year
Mortality
Trend
+2.2%
Prostate
The incidence of prostate cancer, the second most common cancer
in men, spiked in the early 1990s after regulators approved the
prostate-specic antigen (PSA) screening test. Most men treated
after a PSA test had nonlethal tumors.
>
0325SpecialNewsSection.indd 1541 3/17/11 3:38 PM
www.sciencemag.org SCIENCE 3
2012 Cancer Crusade at 40 Collections Booklet
25 MARCH 2011 VOL 331 SCIENCE www.sciencemag.org 1542
CANCER CCCCCCCCCCAAAAAAAAAANNNNNNNNNNCCCCCCCCCCEEEEEEEEEERRRRRRRRRR
Crusade at 40
ONE OF THE BEST HOPES FOR ANTICANCER
drugs in the past decade comes from a simple
idea: Find a weak point in a tumors molec-
ular machinery and throw a well-aimed
wrench into it. The strategy has led to some
dramatic successes, stopping the growth of
certain cancers in their tracks while doing
little or no harm to healthy tissue. But the
pursuit of whats known as targeted ther-
apy has taken many of those involved on
a roller-coaster ride. The new treatments,
after a brilliant debut, tend to lose potency
as tumors develop resistance. After a pause
lasting weeks or months, the cancer may
begin to grow again.
Understanding why this occurs and devis-
ing therapies that will overcome resistance
are the main focus of a growing community
of cancer researchers. A recent example of
their quest involves a drug designed to stop
metastatic melanoma, a highly aggressive
disease for which there is no effective treat-
ment today.
In late 2007, Roche and a biotech com-
pany called Plexxikon began testing a new
targeted drug called PLX4032
for patients with advanced
melanoma. When research-
ers presented the first scans
from treated patients, audi-
ences were stunned: In some
cases, the tumors had almost
disappeared. Eighty percent of
patients got better, a remark-
able response that seemed to
validate the concept of targeted
therapy. There was jubila-
tion about how well it works,
says Charles Sawyers, a can-
cer researcher at Memorial Sloan-Kettering
Cancer Center (MSKCC) in New York City.
Sawyers is a pioneer of the targeted approach
and co-developer of the drug Gleevec that
has been spectacularly effective against
chronic myelogenous leukemia (CML).
But then came the letdown: After about
7 months, most melanoma patients on the
PLX4032 pill saw their tumors begin to
grow again; many died. Last fall, trying to
learn what enabled that puzzling regrowth,
several research teams showed that resistant
tumor cells had found ways to switch back
on the cell pathway that the PLX4032 drug
had jammed. They gave some of the relapsed
patients another new drug that blocks the
pathway at another point, hoping their
tumors would shrink under the dual assault.
The results are still coming in.
In the past few years, researchers have
reported dramatic responses to a hand-
ful of new drugs that are given to patients
with a specic mutation in their tumors. But
the emerging pattern is that although these
drugs can shrink solid tumors and extend
patients lives, they never completely elim-
inate the cancer. For reasons still not well
understood, the initial success is only a foot
in the door, says cancer geneticist Michael
Stratton of the Wellcome Trust Sanger Insti-
tute in Cambridge, U.K., whose team has
found several genetic weak spots in tumors.
Researchers are still working out exactly
what to do next. On the surface, the answer
is straightforward: Identify the ways that
tumors resist the drug, then nd or develop
second-generation drugs that block these
escape routes. With the right drugs on hand,
researchers envision designing a cocktail
perhaps two, three, or more drugsthat, if
given when a patient is rst diagnosed, could
stop tumors from ever evading the block-
ade. This approach has worked for patients
infected with HIV, who usually take three
antiviral drugs. Sawyers and many other
researchers say theres no reason it shouldnt
work for cancer.
But getting a therapy to work and getting
it to endure are two different things. Even if
a combination therapy stops tumor growth,
it may not buy patients more than a tempo-
rary reprieve, researchers admit. To stretch
the benet over years, it might be necessary
to devise one complex cocktail after another,
each tailored to a patients evolving tumors.
Tossing in the wrench
Molecular targeting builds on what research-
ers have learned from 30 years of work on
the genetic changes behind cancer. Uncon-
C
R
E
D
I
T
S

(
T
O
P

T
O

B
O
T
T
O
M
)
:

C
O
P
Y
R
I
G
H
T

N
O
V
A
R
T
I
S

A
G
;

C
O
U
R
T
E
S
Y

A
S
T
R
A
Z
E
N
E
C
A
;

W
I
K
I
M
E
D
I
A

C
O
M
M
O
N
S
;

(
T
A
B
L
E

S
O
U
R
C
E
)

C
H
A
R
L
E
S

S
A
W
Y
E
R
S
/
M
S
K
C
C

A
N
D

J
E
F
F
R
E
Y

E
N
G
E
L
M
A
N
/
M
G
H
;

G
E
O
R
G
E

M
C
G
R
E
G
O
R
/
N
C
I
;

S
C
I
E
N
C
E
Combining Targeted Drugs
To Stop Resistant Tumors
Even the most successful targeted therapies lose potency with time. Researchers hope
to gure out how tumors escape; they aim to turn months of survival into years
NEWS

1991
National
Breast Cancer
Coalition
launched, in
the AIDS
activist style.
1992
FDA
approves
synthetic
yew bark
derivative, Taxol
(paclitaxel), for
breast cancer.
1993
Congress orders
study of environmen-
tal causes of breast
cancer on Long
Island; the 10-year
study will yield no
signicant ndings.
1994
BRCA1 gene,
identied as a
risk for breast
and ovarian can-
cer, is cloned;
BRCA2 cloned
the next year.
Science names
p53 Molecule
of the Year.
1998
FDA approves
Herceptin (trastu-
zumab), a mono-
clonal antibody, for
metastatic breast
tumors that over-
produce HER2.
1996
American Cancer
Society and oth-
ers report the rst
sustained decline in
overall U.S. cancer
deaths, a drop of 2.6%
from 1991 to 1995.
POSSIBLE COCKTAILS OF TARGETED CANCER DRUGS
Median time Possible
Drug Cancer Target to resistance cocktails
Gleevec chronic myelogenous BCR-ABL fusion 5 years Dasatinib or nilotinib
leukemia protein (17% of patients) + T315I inhibitor
Iressa, non-small cell lung EGFR receptor 12 months Tarceva + T790M
Tarceva cancer with inhibitor + (MET inhibitor
EGFR mutation or PI3K inhibitor)
PLX4032 melanoma with V600E BRAF protein 7 months PLX4032 + MEK inhibitor
BRAF mutation
0325SpecialNewsSection.indd 1542 3/17/11 3:38 PM
4 SCIENCE www.sciencemag.org
www.sciencemag.org SCIENCE VOL 331 25 MARCH 2011 1543
SPECIALSECTION
C
R
E
D
I
T
S

(
T
O
P

T
O

B
O
T
T
O
M
)
:

A
D
A
P
T
E
D

F
R
O
M

E
N
G
E
L
M
A
N

&

J

N
N
E
,

C
L
I
N
I
C
A
L

C
A
N
C
E
R

R
E
S
E
A
R
C
H

1
4

(
1
5

M
A
Y

2
0
0
8
)
;

N
C
I
;

G
L
O
G
A
U

P
H
O
T
O
G
R
A
P
H
Y
/
N
C
I
;

N
C
I
trolled cell growth is often driven by an
aberrant protein in the cell membrane
that transmits a spurious signal to the
nucleus, instructing it to divide. Anti-
bodies or small molecules can be used
to block these overactive cell recep-
tors, or a mutated protein farther down
the signaling chain, causing tumors to
shrink dramatically. (By contrast, stan-
dard chemotherapy targets all dividing
cells in the body, which makes it much
more toxic.)
But because tumors are geneti-
cally diverse, resistance seems inev-
itable. Once a targeted drug wipes
out the bulk of a tumor, cells harbor-
ing resistance genes, or alternative
growth instructions, have a chance to
grow. Even a tiny population of resis-
tant cells can expand and take over.
Every single targeted therapy will
select for resistance in this way, says
Carlo Maley, who studies the evolution
of cancer at the University of Califor-
nia, San Francisco.
Gleevec is the classic example. In
95% of patients with CML, the can-
cer is driven by a gene called BCR-ABL that
is formed when two chromosomes swap
pieces, making a fused segment known as
the Philadelphia chromosome. Gleevec,
made by Novartis, is a small molecule that
blocks the BCR-ABL fusion protein. The
drug was approved for CML by regulators
10 years ago this May, and many patients on
it live for at least a decade. But about 17%
of patients develop resistance within 5 years.
Gleevecs developers anticipated this,
they say. Sawyers and others have shown
that in most cases resistance results from
mutated versions of the BCR-ABL pro-
tein that are not affected by the drug and
continue to tell cells to grow. Companies
developed other specific drugs, dasatinib
and nilotinib, that block most forms of the
mutated enzyme and are given to patients
who relapse. Trials published last year show
that the drugs work so well as an alterna-
tive to Gleevec as initial therapy for CML
patients that they may delay resistance for
years, Sawyers says.
However, no cocktail for CML has yet
been tested as an initial therapy in a clinical
trial. One reason, Sawyers says, is that the
cocktail mix is not quite complete. There is
one important mutant version of the BCR-
ABL protein, T315I, that no existing drugs
target. (A promising candidate is in clinical
trials, though.) Furthermore, patients arent
that interested in enrolling in a combina-
tion trial because Gleevec alone works very
well for most patients, says Gleevec co-
developer Brian Druker of Oregon Health
and Science University (OHSU) in Portland.
And as a researcher, Druker says he nds it
hard to ask people to take an experimental
drug as well as Gleevec when only a small
fraction will likely do any better than they
would on Gleevec alone. Would I treat 90
people for the benet of 10? Druker asks.
People are so comfortable with one drug,
Sawyers says.
Gleevec has been used with good
effect to treat another cancer, gastro-
intestinal stromal tumor (GIST), a
rare disease usually caused by muta-
tions in genes called PDFGRA or
KIT, says Michael Heinrich of OHSU.
But resistance mutations can appear
in the KIT protein. Although a drug
called sunitinib targets some of them,
theres no current drug that can patch
up all the holes, Heinrich says. And
because GIST patients live a relatively
long time on Gleevec5 years versus
15 months on chemotherapyits hard
to interest companies in testing a cock-
tail for GIST, he says.
Resistance is a more urgent prob-
lem for lung cancer patients treated
with Iressa and Tarceva (getinib and
erlotinib), the rst big success for tar-
geted therapy after Gleevec. These
nearly identical drugs block a cell
receptor called EGFR that transmits
growth signals. The drugs didnt help
most patients in trials for non-small
cell lung cancer, but researchers real-
ized that they work extremely well on the
roughly 10% of patients who have an EGFR
mutation in their tumors (they tend to be
women, never-smokers, or Asian). Some of
these patients tumors almost vanish when
they receive an EGFR inhibitor. However,
the average patient develops resistance after
about a year.
As with Gleevec for leukemia and GIST,
the trouble is often that tumor cells appear
in which the EGFR receptor has a specic
new mutation (T790M) that prevents the
drugs from binding well. Companies are still
moving toward clinical trials of drugs that
block EGFR proteins with this mutation,
which about half of all patients develop, says
Jeffrey Engelman of Massachusetts General
Hospital (MGH) in Boston. Lung cancer
cells have another way of evading the drug,
moreover: They can make more of a differ-
ent cell receptor, called MET, that can take
over for EGFR and maintain the growth sig-
C
EGFR
Akt
ERBB3 MET
p85
P13K
p110
P P
Akt
ERBB3
p85
P13K
p110
P
A
EGFR
Akt
ERBB3
p85
P13K
p110
P P
B EGFR
T790M
Akt
ERBB3
p85
P13K
p110
P P
p85
P13K
p110
P
MET
P

2001
FDA approves
Gleevec (imatinib),
a targeted drug, for
chronic myelog-
enous leukemia;
Time calls it a
magic bullet.
Childhood cancer landmark:
nearly 80% of those treated
for acute lymphoblastic
leukemia are free of cancer
events for 5 years or more.
1998
Nobelist James
Watson tells The
New York Times that
blocking the growth
of tumor blood
vessels (antiangiogenesis)
can cure cancer in 2 years.
2003
NCI Director Andrew
von Eschenbach vows
to eliminate
suffering
and death
from cancer
by 2015.
2004
FDA approves
Avastin, an anti-
angiogenesis
drug, for colon
cancer, with
chemotherapy.
2005
NIH launches The
Cancer Genome
Atlas to catalog
genomic changes in
tumors.
Two ways out. (A) The lung
cancer drug Tarceva (blue
spheres) blocks the EGFR
receptor from transmitting a
signal. (B) The T790M muta-
tion prevents the drug from
binding. (C) Tumor cells can
also overexpress the MET
receptor, which takes over
when EGFR is blocked.
>
0325SpecialNewsSection.indd 1543 3/17/11 3:38 PM
2012 Cancer Crusade at 40 Collections Booklet
www.sciencemag.org SCIENCE 5
25 MARCH 2011 VOL 331 SCIENCE www.sciencemag.org 1544
CANCER CCCCCCCCCCAAAAAAAAAANNNNNNNNNNCCCCCCCCCCEEEEEEEEEERRRRRRRRRR
Crusade at 40
C
R
E
D
I
T
S

(
T
O
P

T
O

B
O
T
T
O
M
)
:

C
O
U
R
T
E
S
Y

M
E
M
O
R
I
A
L

S
L
O
A
N
-
K
E
T
T
E
R
I
N
G

C
A
N
C
E
R

C
E
N
T
E
R
;

C
O
U
R
T
E
S
Y

M
A
S
S
A
C
H
U
S
E
T
T
S

G
E
N
E
R
A
L

H
O
S
P
I
T
A
L

(
2
)
;

J
U
P
I
T
E
R

I
M
A
G
E
S
/
T
H
I
N
K
S
T
O
C
K
naling pathway. These escape routes and a
couple of others are problematic but a man-
ageable list, says Jeffrey Settleman, who
left MGH last year for Genentech in South
San Francisco, California.
Hoping to close off two escape routes
at once, a few companies are running tri-
als combining an EGFR inhibitor and a
MET inhibitor. Early results are mixed,
Engelman says. One problem is that the
cocktails didnt target the T790M muta-
tion, says William Pao of Vanderbilt-
Ingram Cancer Center in Nashville. Pao
is optimistic about another early clini-
cal trial, however, that is treating patients
who became resistant to Tarceva or Iressa
with a new, more potent EGFR inhibitor
and an approved EGFR-blocking antibody
called cetuximab; the combination shrank
T790M-carrying tumors in mice. Initial
results are expected this summer.
New escape routes
How melanoma tumors become resistant to
the initially powerful PLX4032 drug is more
complex. Plexxikon began developing the
drug after Strattons team reported 9 years
ago that tumors in half of advanced mela-
noma patients have the same mutation in a
protein called BRAF. This protein is part of
a key growth signaling pathway. Because
PLX4032 blocks mutated BRAF and inhib-
its this pathway only in tumor cells, it can be
given at high doses (Science, 18 December
2009, p. 1619). But nobody was surprised,
says trial co-leader Keith Flaherty of MGH,
when patients on PLX4032 whose tumors
melted away eventually relapsed.
But researchers were surprised to nd
that biopsied tumors from patients who
developed resistance didnt have mutations
in the BRAF protein. Instead, several teams
reported recently that tumor cells appear
to use three different escape routes. This
was disappointing, but some also saw good
news: Two of these forms of resistance func-
tion in the same wayrestoring the growth
pathway by activating a downstream protein
called MEK. This suggests, says David Solit
of MSKCC, that combining a BRAF inhibi-
tor and a drug that blocks MEK could block
both escape routes.
These studies still dont account
for the lions share of resistance cases,
Flaherty says; there is more work to be
done. But they have inspired GlaxoSmith-
Kline, which makes a BRAF inhibitor simi-
lar to PLX4032, to launch a trial combining
its drug with a MEK inhibitor. This will
provide proof of concept of the cocktail
idea, Sawyers says, within a couple of years.
Another cocktail combining PLX4032
with an immunotherapy drug developed
at MSKCC is also under study (Science,
22 October 2010, p. 440).
Cocktails with caveats
Testing combinations of drugs is not new in
cancer research. But these cocktails would
be different because they would be ratio-
nally designed to block tumor escape
routes, researchers say. Although research-
ers are eager to begin, they will need to over-
come some barriers.
One is commercial competition. Many
drugs that target the newly found resistance
pathways are still in development. Compa-
nies are loath to test two unapproved drugs
simultaneously, especially if one comes
from a business rival. They worry that side
effects from one product will bloody a
drug that is safe on its own, Flaherty says.
People in my line of work are all about try-
ing to make this happen because the biology
is so obvious, he says, but drug develop-
ment is another matter.
Flaherty and others are encouraged,
however, by recent examples of compa-
nies teaming up: In 2009, Merck and Astra-
Zeneca agreed in a groundbreaking deci-
sion to test a Merck MEK inhibitor and a
drug blocking another key pathway, PI3K/
Akt. Since then, a couple more companies
have signed such agreements. The motiva-
tion is not just to overcome resistance but
also to explore an exciting possibility: It has
become increasingly clear from cell stud-
ies that pairing two drugs aimed at different
pathways can result in synergistic effects,
says D. Gary Gilliland, senior vice president
and franchise head for oncology at Merck in
North Wales, Pennsylvania. He also praises
new draft U.S. Food and Drug Administra-
tion guidelines that allow for exibility for
testing combinations.
Researchers testing combinations must
also face the fact that tumors are constantly
evolving, Engelman notes. His group pub-
lished a study on non-small cell lung can-
cer this week in Science Translational Medi-
cine (STM) that illustrates this complexity.
Retargeting. Gleevec co-developer Charles Sawyers, lung cancer researcher Jeffrey Engelman, and mela-
noma trial co-leader Keith Flaherty are working on ways to overcome acquired resistance to targeted drugs.
Sawyers Engelman Flaherty
FDA approves
Provenge, an immune
treatment for meta-
static prostate cancer.
It extends life about
4 months and costs
$93,000.
2011
PLX4032, a
targeted cancer
drug, extends
life in patients
with advanced
melanoma.
2010
National
Lung Cancer
Screening
Trial nds
that helical CT screening
can reduce cancer deaths
among smokers.
2009
James Watson writes that
its time to turn from cancer
genetics to understand-
ing the chemical reactions
within cancer cells,
or cell metabolism.
2007-2008
Breast cancer
incidence declines,
attributed to better
screening and
reduced use of hor-
mone replacement
therapy.
2006
FDA approves
Gardasil vac-
cine to prevent
HPV infection,
which can lead
to cervical
cancer.
0325SpecialNewsSection.indd 1544 3/17/11 3:38 PM
6 SCIENCE www.sciencemag.org
www.sciencemag.org SCIENCE VOL 331 25 MARCH 2011 1545
SPECIALSECTION
C
R
E
D
I
T

(
G
R
A
P
H

S
O
U
R
C
E
)
:

A
.

M
A
R
I
O
T
T
O
,

E
T

A
L
.
,

J
O
U
R
N
A
L

O
F

T
H
E

N
A
T
I
O
N
A
L

C
A
N
C
E
R

I
N
S
T
I
T
U
T
E


1
0
3
,

2

(
1
9

J
A
N
U
A
R
Y

2
0
1
1
)
THE AFTERNOON JEFF GUSTAFSON LOST HIS
life to stomach cancer, a rainbow appeared
above his Arlington, Virginia, hospice.
Even if the shimmering arc had led to a
pot of gold, however, its not clear whether
the treasure would have covered the cost
of treating the 48-year-old
builder, who was known for his
wit and compassion. The bills
for just 4 months of dogged,
full-tilt treatment totaled more
than $350,000including
one drug that cost more than
$12,000 per dose.
Gustafsons story isnt that
unusual to experts who track
the costs of treating cancer in
the United Statesand are
increasingly worried about
where they are headed. Over
the past 3 decades, total U.S.
spending on cancer care has
more than quadrupled, reach-
ing $125 billion last year, or
5% of the nations medical bill,
according to a recent estimate. By 2020, it
could grow by as much as 66%, to $207 bil-
lion. Multiple forces are driving the spiral:
a growing and aging population, more peo-
ple living longer with cancer, and new per-
sonalized, or targeted, therapies that can
come with sticker-shock prices of $50,000
or more per patient.
New and more costly, however, havent
necessarily meant better. Although tar-
geted treatments, which attack a molecu-
lar weak spot in the tumors support sys-
tem, have helped improve survival rates for
many cancers, some extend life for just a
few weeks or months (see p. 1542). And the
prices can be sobering: more than $1.2 mil-
lion to extend a lung cancer patients life for
1 year in one scenario involving a costly but
common drug. That example is unusual, but
such numbers have sparked a growingand
sometimes feistydebate
over how best to calculate the
benets of new cancer treat-
ments, whether their use will
lower or raise per-patient
expenses, and who should
decide whether using them is
worth the cost. The question
is, Are we spending too much
for too little? says oncolo-
gist Antonio Tito Fojo of the
U.S. National Cancer Institute
(NCI) in Bethesda, Maryland.
Demographic drivers
Tallying the current and
future costs of treating can-
cer isnt easy. Cancer now
includes more than 100 dif-
His team analyzed 37 biopsy samples from
patients with the EGFR mutation who were
given Iressa or Tarceva and later became
resistant. Although some resistance mutations
were known, others were new, and for 30% of
the samples, his team could not identify the
mechanism. Some tumors even morphed into
a different type of lung cancer that requires
an entirely different treatment. In addition,
biopsies from three patients collected during
the course of treatment showed that tumors
changed: Some that developed resistance
mutations later lost them. Its very, very com-
plex. Its not tting into the simple boxes that
weve made until now, Engelman says.
Cocktail therapy will face another issue
that has not been well explored so far: the risk
that combining two drugsparticularly ones
that target different pathways used by normal
cellscan lead to unacceptable side effects.
Engelman thinks that for this reason, patients
will be able to tolerate a cocktail for only a
short time. He envisions putting them on a
single drug, then intermittently giving them a
pulse of a cocktail for several days. As his
STM study suggests, clinicians would need to
constantly biopsy patients and tailor the cock-
tail to the mutations in their tumor. This could
stretch responses out to years, Engelman says.
A lot of this is going to have to be done
by trial and error, Pao agrees. Sawyers and
others point out, however, that such combi-
nation therapies developed in the 1960s and
70s eventually vanquished most cases of
childhood leukemia, Hodgkin lymphoma,
and testicular cancer.
Major cancer centers are already rou-
tinely genotyping biopsies from patients
throughout their treatment using proce-
dures that are less invasive than surgery,
such as collecting a few tumor cells with a
thick needle. Engelman says this is essen-
tial: We cant be afraid to rebiopsy to see
whats going on.
Cancer researchers acknowledge that
coming up with cocktails to corner can-
cer will be much more difcult than it was
with HIV. Cancer geneticist Bert Vogelstein
of Johns Hopkins University in Baltimore,
Maryland, points out that cancer is different
from AIDS: Among other things, tumors vary
more from patient to patient than HIV does;
the genetic heterogeneity of cancer cells
within a single patient is orders of magni-
tude greater than HIV genotypes in patients,
so tumors have a much larger reservoir of
resistance mechanisms that go beyond those
already uncovered. Unfortunately, weve got
billions of cancer cells ready to become resis-
tant, and it takes 15 years to develop each new
drug, Vogelstein says. Moreover, to wipe out
a tumor completely, he and others say, a cock-
tail might also have to include a drug that tar-
gets stem celllike cells in a tumor that con-
tinuously give rise to tumor cells.
Turning cancer into a manageable dis-
ease, however, would be wonderful, Vogel-
stein says, if it can be done. Advocates of
targeted therapy think they will get there.
Melanoma was completely untreatable
18 months ago, says Neal Rosen of MSKCC.
Give us a chance. JOCELYN KAISER
Can Treatment Costs Be Tamed?
More patients and the rising costs of new cancer treatments spark debate over how
much is too muchand who should decide
NEWS
5% increase in annual costs,
first and last year of care
2% increase in annual costs
Population growth only
Current trends in incidence
and survival, projected
PROJECTED COSTS OF TOTAL U.S. CANCER CARE, 201020
100
2010 2020
160
140
120
200
U
.
S
.

$

(
b
i
l
l
i
o
n
s
)
180
220
Four scenarios. Experts predict total U.S. spending on cancer care could rise by
as much as 66% by 2020, depending on shifts in disease incidence and survival,
and treatment costs.
0325SpecialNewsSection.indd 1545 3/17/11 3:38 PM
2012 Cancer Crusade at 40 Collections Booklet
www.sciencemag.org SCIENCE 7
25 MARCH 2011 VOL 331 SCIENCE www.sciencemag.org 1546
CANCER CCCCCCCCCCAAAAAAAAAANNNNNNNNNNCCCCCCCCCCEEEEEEEEEERRRRRRRRRR
Crusade at 40
C
R
E
D
I
T
S

(
T
O
P

T
O

B
O
T
T
O
M
)
:

C
O
U
R
T
E
S
Y

M
O
L
L
Y

S
I
M
;

(
G
R
A
P
H

S
O
U
R
C
E
)
:

A
.

M
A
R
I
O
T
T
O
,

E
T

A
L
.
,

J
O
U
R
N
A
L

O
F

T
H
E

N
A
T
I
O
N
A
L

C
A
N
C
E
R

I
N
S
T
I
T
U
T
E


1
0
3
,

2

(
1
9

J
A
N
U
A
R
Y

2
0
1
1
)
ferent diseases that affect more than 13 mil-
lion people in the United States each year,
and there is no single source of uniform sta-
tistics. To come up with their recent fore-
casts, a team led by NCIs Angela Mariotto
turned to a massive trove of data kept by
the federal governments Medicare health
insurance program, which now covers some
40 million Americans over age 65. Sieving
the data, they were able to extract a rough
snapshot of the costs associated with treating
nearly 1.8 million Medicare patients diag-
nosed with cancer between 1975 and 2005.
Overall, they examined incidence, survival,
and cost trends for 13 cancers in men and 16
cancers in women, and estimated costs for
patients younger than 65.
The result, published last January in
the Journal of the National Cancer Insti-
tute (JNCI), reveals some of the complex
demographic and technological forces that
are pushing costs down, even as the over-
all total rises. Earlier studies, for instance,
have suggested that average per-patient
costs for many cancers have declined
slightly in recent years, largely because
outpatient treatment has often replaced
more expensive in-hospital stays. The JNCI
study, however, shows that those savings
are being swamped, in part, by the grow-
ing number of older people, who are more
likely to get cancer. Medicare predicts its
rolls will nearly double, to 70 million peo-
ple, by 2020. And the JNCI study forecasts
that some 16% of these older Americans
about 11.4 million peoplewill have can-
cer, up from 8 million today. Another 6.6
million younger people will also be living
with cancer. Even if all other trendssuch
as the cost of individual treatmentdont
change, they estimate that those demo-
graphic changes alone will push national
cancer care costs up 27% by 2020.
Ironically, another factor driving up costs
is that people are now surviving cancers that
might have killed them quickly decades ago.
Some of the largest projected cost increases,
for instance, are linked to providing con-
tinuing care for people living with breast
and prostate cancer, a group expected to
grow by as much as 41%, to nearly 8 mil-
lion in 2020. It will cost some $18 billion
to provide continuing care for these patients,
the team estimates, some 30% to 40% more
than current costs.
The study notes, however, that spending
tends to follow a u-shaped curve over the
period of cancer treatment, with the highest
costs coming just after diagnosis and in the
last year of life, but with relatively less spent
in between. For instance, men under 65 with
stomach cancersuch as Gustafsonspent
an average of $94,000 per year on initial
care in 2010 and $161,000 in the last year of
life, but just $4200 per year for continuing
care. In part, thats because stomach can-
cer can kill quickly, and for those who dont
die early, end-stage costs can be very high.
But the numbers also reect the fact that
younger patients typically get more aggres-
siveand more costlycare than older
ones, perhaps because society perceives a
greater potential benet to extending the
life of a younger person. Patients under 65
now make up about 40% of all cancer cases,
the JNCI team estimated, and typically
receive care that costs about 35% more than
that of older patients.
Worth the cost?
The i mpe ndi ng
demo gr aphi cal l y
driven increases are
i nevi t abl e, t he
authors note. Harder
to pin down, how-
ever, is the future
impact of techno-
logical changes that,
overall, have been
rapidly pushing up
costs. In particular,
chemotherapy prices
have been esca-
lating faster than
other medical costs,
largely because of
the introduction of
targeted drugs that
come with price tags
higher than those of
many older compounds. Although studies
suggest that cancer drugs typically account
for less than 15% of a patients total treat-
ment costs, experts say that is changing.
For instance, an increasingly used drug
called bevacizumabsold under the name
Avastincan cost from $30,000 to $62,000
per patient per course of treatment, depend-
ing on the targeted cancer, according to
estimates compiled by NCIs Fojo and col-
leagues. In contrast, some drugs used at the
beginning of the war on cancer in 1971 cost
just a few hundred dollars per patient.
Part of the problem, Fojo and his col-
leagues have argued in a pair of provocative
papers, is that the new, costlier drugs often
fail to improve survival appreciably. They
dont work for all patients, and when they
do work, the effect is often limited. To high-
light such problems, in both a 2009 JNCI
paper and a 2010 paper published in Clinical
Cancer Research, Fojos team focused on a
breakthrough treatment that was touted
at a major annual cancer research meeting
in 2008. A study had shown that non-small
cell lung cancer patients treated with cetux-
imab (sold as Erbitux) lived, on average,
1.2 months longer than those receiving stan-
dard treatments. One scenario, which called
for using cetuximab for 18 weeks at a cost
of $40,000, translated to a cost of $496,000
for one extra quality-adjusted life year
(QALY). In contrast, they noted, kidney
dialysis, a lifesaving technology Medicare
made universally available in 1972, costs
$129,090 per QALY.
Another scenario, using bevacizumab
Number of people living with cancer
0 2 4 6 8 10 12
Millions
2020
2010
Younger than 65
65 and older
U.S. CANCER PREVALENCE
Surviving. One of the major drivers of increased
costs is that more people are living longer with
cancer.
One life. After standard treatments failed to stop Jeff Gustafsons stomach can-
cer, his insurer approved the experimental use of one expensive drug, but it
failed to extend his life.
0325SpecialNewsSection.indd 1546 3/17/11 3:38 PM
8 SCIENCE www.sciencemag.org
www.sciencemag.org SCIENCE VOL 331 25 MARCH 2011 1547
SPECIALSECTION
C
R
E
D
I
T

(
G
R
A
P
H

S
O
U
R
C
E
)
:

A
.

M
A
R
I
O
T
T
O
,

E
T

A
L
.
,

J
O
U
R
N
A
L

O
F

T
H
E

N
A
T
I
O
N
A
L

C
A
N
C
E
R

I
N
S
T
I
T
U
T
E


1
0
3
,

2

(
1
9

J
A
N
U
A
R
Y

2
0
1
1
)
at a cost of $30,000, translated to a QALY
cost of $1.2 million. And such drugs are not
alone among treatments offering marginal
benet at very high cost, they wrote: More
than 90% of the cancer drugs approved by
the U.S. Food and Drug Administration
(FDA) in the previous 4 years had cost more
than $20,000 for 12-week treatments. We
must stop deluding ourselves into thinking
that prescribing expensive chemothera-
pies and tests are an aberration, a temporary
deviation from an otherwise reasonable cost
trajectory, they concluded.
Along with their critique, Fojo and his col-
leagues offer solutions. They argue that Medi-
care and other U.S. insurers, for instance,
shouldnt pay for drugs that would cost more
than $129,090 per QALY. That would still
make the United States somewhat more gen-
erous than other nations, including the United
Kingdom, Australia, and Canada, whose gov-
ernment-run health services routinely reject
the use of new cancer drugs because they
cant meet certain QALY
costs. Fojos group also
says that drug compa-
nies shouldnt fund tri-
als designed to detect
survival improvements
of fewer than 2 months
unless the drug will cost
less than $20,000 and
should even consider
reducing charges if a
drug doesnt work in a
particular patient. Doc-
tors should use costly
drugs only in the subset
of patients proven to gain
benefit, they say, and
avoid using new drugs
off-label, i.e., to treat
cancers other than those
for which the drug was
approved by FDA. Such
steps, Fojo says, would
focus our attention on using drugs that deliver
proven benefits until researchers can find
better ways to efciently develop and use tar-
geted therapies.
Some disagree sharply. Such ideas are
based on awed premises, Joshua Cohen
and William Looney of Tufts University
School of Medicine in Boston argue in a
November 2010 response published in Nature
Biotechnology. Using QALY calculations
to rule out therapy, for instance, is a biased,
blunt instrument that ends up denying
drugs to patients who need them, the authors
say. They view this as a radically egalitar-
ian approach that forces insurers to choose
between drugs they can afford for all, as
opposed to those they will fund for no one.
They also argue that preventing doctors from
using drugs off-label would hobble the proven
practice of freeing doctors to nd promising
new uses for existing drugs. And it would
stand in stark contrast with clinical practice.
Studies, for instance, suggest that up to 75%
of anticancer drugs are already used off-label.
And price controls would, they argue, ulti-
mately cause investors to reduce funding for
research into new drugs because they couldnt
be sure of recouping their costs. There is one
thing the two sides do agree on: Better, more
organized studies could help steer the right
drugs to the right patients. Better genetic tests,
for instance, could identify patients unlikely
to benet from a particular drug. They could
also help researchers design smaller, less
costly trials that better identify smaller groups
of patients likely to respond; current trials
often wash out these drugs because the ben-
ets seem statistically negligible.
Another idea gaining ground is cover-
age with evidence development (CED),
in which insurers link payments for certain
drugs to efforts to collect data on compara-
tive effectiveness, with the aim of discontinu-
ing coverage for drugs that dont work. Cohen
and Looney even suggest that CED could be
combined with risk-sharing arrangements,
in which insurers and manufacturers agree
to link a drugs price to its performance. U.S.
Medicare managers have already launched
a number of CED experiments, and the new
health care reform law authorizes extensive
new efforts to compare the effectiveness of
different medical treatments, including cancer
drugs. In 2014, it also requires private insur-
ers to cover the routine costs of enrolling can-
cer patients in clinical trials, removing a major
barrier to their participation.
Agonizing decisions
Although such studies didnt help Gustafson,
his treatment did reect both the challenge
and opportunity inherent in efforts to tamp
down the costs of cancer care. His oncolo-
gists, for instance, talked with him and his
wife, Molly Sim, about the costs they were
likely to encounter. Thats something that a
recent statement from the American Society
of Clinical Oncologists says needs to happen
more often, given how many families exhaust
their savings ghting cancer. We were for-
tunate, we had really good insurance, says
Sim, who ended up paying less than $6000
out of pocket.
Gustafsons doctors
also gave him genetic
tests to determine
whether he should be
given one expensive but
promising drug; unfor-
tunately, the tests indi-
cated that he wouldnt
benefit. But when he
failed to respond to
standard treatments,
they proposed a last-
ditch strategy: Sim suc-
cessfully appealed to
her insurer to pay for
off-label use of Avas-
tin at a cost of nearly
$40,000. Gustafson
died in November 2009,
before he could com-
plete the treatment.
Its a story that
Fojowho urges caution in using expensive
drugssays he knows all too well. In the
abstract, you say using these marginal drugs
is crazy, he said recently, shortly after com-
ing off his rounds treating cancer patients.
But when you have the patient in front of
you, and nothing else is working, well, I
might try the crazy thing, too. We used to do
it with cheap drugs, but now we are doing it
with really expensive drugs. The problem is
they really dont work any better.
DAVID MALAKOFF
David Malakoff is a writer in Alexandria, Virginia.
Last year Continuing Initial
THE PRICE OF TREATMENT
0
50,000
100,000
150,000
200,000
250,000
300,000
350,000
(under
65)
(65 and
over)
Female-Breast
A
n
n
u
a
l
i
z
e
d

m
e
a
n

n
e
t

c
o
s
t
s

i
n

2
0
1
0

d
o
l
l
a
r
s
(under
65)
(65 and
over)
Female-Brain
(under
65)
(65 and
over)
Male-Stomach
(under
65)
(65 and
over)
Male-Lung
(under
65)
(65 and
over)
Female-
Pancreas
Male-
Pancreas
Varying costs. The average annual cost of treatment can vary, depending on the patients age and
cancer type. In general, costs are higher just after initial diagnosis and in the last year of life.
0325SpecialNewsSection.indd 1547 3/17/11 3:38 PM
2012 Cancer Crusade at 40 Collections Booklet
www.sciencemag.org SCIENCE 9
25 MARCH 2011 VOL 331 SCIENCE www.sciencemag.org 1548
CANCER CCCCCCCCCCAAAAAAAAAANNNNNNNNNNCCCCCCCCCCEEEEEEEEEERRRRRRRRRR
Crusade at 40
C
R
E
D
I
T
S

(
T
O
P

T
O

B
O
T
T
O
M
)
:

C
O
U
R
T
E
S
Y

O
F

P
A
R
T
N
E
R
S

I
N

H
E
A
L
T
H
;

(
G
R
A
P
H

S
O
U
R
C
E
)

G
T
F
.
C
C
C
,

M
E
X
I
C
A
N

H
E
A
L
T
H

F
O
U
N
D
A
T
I
O
N
;

E
S
T
I
M
A
T
E
S

B
A
S
E
D

O
N

I
A
R
C

G
L
O
B
O
C
A
N

D
A
T
A

F
O
R

2
0
0
2

A
N
D

2
0
0
8
BOSTONIn one of his PowerPoint presen-
tations, Lawrence Shulman has a series of
photographs that are hard to forget. One
shows Tushime, an 11-year-old girl in
Rwanda suffering from rhabdomyosarcoma,
a rare cancer of the muscles. A tumor resem-
bling a cauliflower is growing out of her
right cheek.
The good news comes in Shulmans
next three slides. Over a 10-week treatment
course with drugs donated by a U.S. pro-
gram, the mass started to shrink until even-
tually it could be removed surgically. The
last picture shows Tushime, standing with
her happy family anddespite a somewhat
lopsided facelooking healthy.
Shulman is chief medical ofcer at the
Dana-Farber Cancer Institute here, which
is afliated with Harvard Medical School
(HMS), and to him the pictures carry a pow-
erful message: Given that treatments are
readily available, how can you not treat a
child suffering from a very curable cancer?
Its a message that is beginning to be
heard. Global health has gained in promi-
nence on political agendas in recent years, but
attention has been overwhelmingly focused
on infectious diseases. Now, some
argue, its time to start closing an
equally unconscionable gap between
rich and poor nations in cancer pre-
vention, diagnosis, and treatment.
The numbers speak volumes. A
child suffering from leukemia in
Western Europe has an 85% chance
of survival; in the 25 poorest coun-
tries in the world, its just over 10%.
For a man with testicular cancer,
the numbers are about 95% and just
over 40%. Estimates suggest that
less than 5% of the worlds cancer
resources are spent in the develop-
ing world. In many countries, even
painkillers are hard or impossible
to come bya violation of human
rights, Shulman says.
In a bid to change that, oncolo-
gists at topight centers in the United
States and Europe are now taking time out
to help improve cancer care in low- and
middle-income countries. In an article last
month, World Health Organization Director-
General Margaret Chan said that cancer
needs to be acknowledged as a vital part of
the global health agenda. Cancer will also
feature on the agenda at the United Nations
High-level Meeting on Non-Communicable
Diseases in September in New York City.
Shulman and Harvard School of Public
Health Dean Julio Frenk co-chair the new
Global Task Force on Expanded Access
to Cancer Care and Control in Develop-
ing Countries (GTF.CCC), which aims to
move cancer up on the global
priorities list. The group
which combines expertise in
cancer, global health, eco-
nomics, nance, and policy
published a 10-page call to
action last year in The Lancet
and is now working on a col-
lection of papers for the same
journal that details what needs
to be done.
The obstacles are major, and
some people question whether
battling cancer is the wisest use
of scarce global-health money.
Similar doubts were once raised
about infectious diseases,
says HMS physician and GTF.
CCC member Paul Farmer:
A Push to Fight Cancer in the
Developing World
Cancer and other chronic diseases have received little attention from global health
advocates. Thats beginning to change
NEWS
R
a
t
i
o

o
f

m
o
r
t
a
l
i
t
y

t
o

i
n
c
i
d
e
n
c
e
Low
Country income
Lower middle Upper middle High
THE CANCER SURVIVAL GAP
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
Breast
Prostate
Colorectal
Cervix uteri
Non-Hodgkin lymphoma
Thyroid
Testicular
Hodgkin lymphoma
Leukemia (0-14 years of age)
Life and death. For many cancers, the case fatality rate (of which
the ratio of mortality to incidence is a proxy) is much higher in
poor countries than in rich countries.
Prevention. Women wait to be screened
for cervical cancer at a Partners In Health
clinic in Rwanda.
0325SpecialNewsSection.indd 1548 3/17/11 3:38 PM
10 SCIENCE www.sciencemag.org
www.sciencemag.org SCIENCE VOL 331 25 MARCH 2011 1549
SPECIALSECTION
15 years ago people questioned the logistical
and nancial feasibility of treating HIV and
multidrug-resistant tuberculosis (TB) in poor
countries. Yet as Farmer points out, both are
now being addressed on a large scale. Great
strides have been made recently in a range of
other tropical diseases, too. So why cant it be
done for cancer?
A more difcult job
Cancer takes a markedly different toll
depending on the country. Lung cancer, a
major killer in the West, is rarer in Africa,
where fewer people smoke and life expec-
tancy is shorter. Several virus-related can-
cers, on the other hand, are much more
prevalent in poor countries. Cervical can-
cer, caused by the human papillomavirus
(HPV), has become quite rare in rich coun-
tries thanks to screening with Pap smears.
But it is common in the developing world,
where 93% of the estimated 273,000 annual
deaths from this disease occur.
Taken together, however, the burden
of cancer is still lower in the poorer coun-
C
R
E
D
I
T
:

K
E
N
T

D
A
Y
T
O
N
/
H
S
P
H
/

P
R
E
S
I
D
E
N
T

A
N
D

F
E
L
L
O
W
S

O
F

H
A
R
V
A
R
D

C
O
L
L
E
G
E
SUDBURY, MASSACHUSETTSYou can ask me absolutely anything,
says Felicia Knauland shes serious. Knaul, a health economist at
Harvard Medical School in Boston, doesnt mind telling in detail how
breast cancer changed her marriage, her family, and her career. In her
mission to shatter taboos around the disease and improve the lives of
patients in developing countries, her professional and private lives have
become one.
Knaul is one half of a Mexican-Canadian power couple that aims to
end the neglect of cancer as a disease of the poorand will succeed,
if anyone can, say colleagues. The other half is Julio Frenk, the former
health minister of Mexico and a much admired reformer, who is now dean
of the Harvard School of Public Health (HSPH).
Knaul, born and raised in Toronto, says she has always had a keen
interest in poverty and health. She lived in Bogot for 2 years in the early
90s, working on a project for street children and helping the Colom-
bian government reform its health system. After meeting Frenk, who
comes from a family of doctors, she moved to Mexico; she now considers
it home. She joined the Mexican Health Foundation in Mexico City, where
she still leads a research group. When Frenk became minister in Vicente
Foxs Cabinet in 2000, she worked pro bono to help him push through a
major reform that extended basic health coverage to the countrys poor-
est and took effect in 2003.
That didnt prepare her for her own terrifying brush with illness. In
October 2007, a technician in Cuernavaca discovered a lump in Knauls
left breast. It was malignant. After weeks of anguish, she underwent a
mastectomy and started on chemotherapy. Knaul says she was lucky. She
had access to the best doctors in Mexico, andbecause her husband
took a job at the Bill and Melinda Gates Foundation after he left the
government in 2006she even had insurance coverage in the United
States. Knaul used it to seek additional treatment at the Seattle Cancer
Care Alliance. She now rates her 5-year survival chance at between 85%
and 90%.
Thats much more than most Mexican women with breast cancer can
expect. Coverage for breast cancer treatment was added to Frenks insur-
ance plan for the poorest in 2007; in reality, some women still dont get
treatment, for instance, if they live far from a hospital. Early diagnosis is
rare. Many women forgo mammograms even where they are available,
Knaul says, and a culture of machismo often leads men to abandon their
wives or girlfriends if they lose a breast. Knaul recalls a woman recently
diagnosed with breast cancer at one public event saying: A woman with-
out breasts is ugly. I dont want to be ugly.
Knaul decided to start a programit is now a not-for-prot group
called Cncer de Mama: Tmatelo a Pecho (Breast Cancer: Take It to Heart)
that aims to raise awareness of and improve access to prevention, early
detection, and treatment. Sharing her own story might help other women,
she thought, so she wrote
a frank book about her
experience. The resulting
TV interviews with her and
Frenk made waves.
She and Frenk took the
campaign to the next level
after he was appointed dean
of HSPH in 2008 and she
became director of the Har-
vard Global Equity Initiative,
a research program founded
by Nobel laureate
Amartya Sen. She
had the idea to start
a global task force to
expand cancer care in
poor countries. Frenk
now co-chairs it with
Lawrence Shulman,
chief medical ofcer of
the Dana-Farber Cancer
Institute in Boston. Many of their well-placed friends signed on: former
UNAIDS chief Peter Piot, director of the London School of Hygiene and
Tropical Medicine; Columbia University economist and poverty warrior Jef-
frey Sachs; and CNN chief medical correspondent Sanjay Gupta. They have
really brought this to the doorstep of many people at high levels, says Car-
los Rodriguez, a pediatric oncologist at Dana-Farber.
The task force is now collecting evidence on how cancer care in devel-
oping countries can be improved, she says. To do so, she has helped
recruit a series of papers for publication in The Lancet (see main text,
p. 1548). Her own story, she realizes, is atypical for a woman in Mexico. Still,
it reinforces her message, she says: When you say at a meeting, I have
cancer, people listen in a way that happens with very few other diseases.
Her frankness is part of her strategy. Mexican couples thanking her
for her book will sometimes mention chapter 18, she says, in which she
discusses how the chemotherapy-induced menopause shut down her sex
life. At her art-lled house 30 kilometers west of Boston, Knaul also dis-
cussed, matter-of-factly, why, in her case, reconstructive surgery was not
successful. (The implant dropped a couple of inches, she says, because
of a lack of tissue to hold it up.)
Her own marriage didnt suffer, she addson the contrary, the dis-
ease brought the two closer, and for the book, Frenk contributed frag-
ments of four love letters written when she was ill. As Knaul wrote, I had
my boyfriend back. M.E.
In sickness and in health.
Felicia Knaulwith hus-
band, Julio Frenkwrote
about her disease.
k
h
n,
of
cer
I
Making Her Life an Open Book to Promote Expanded Care
0325SpecialNewsSection.indd 1549 3/17/11 3:38 PM
2012 Cancer Crusade at 40 Collections Booklet
www.sciencemag.org SCIENCE 11
25 MARCH 2011 VOL 331 SCIENCE www.sciencemag.org 1550
CANCER CCCCCCCCCCAAAAAAAAAANNNNNNNNNNCCCCCCCCCCEEEEEEEEEERRRRRRRRRR
Crusade at 40
C
R
E
D
I
T
:

T
H
I
N
K
S
T
O
C
K
;

T
A
B
L
E

S
O
U
R
C
E
:

L
.

S
H
U
L
M
A
N
tries, which is one reason why it has failed
to get global health policymakers full atten-
tion. Unfortunately, the developing world is
catching up. In many middle-income coun-
tries, life expectancy is increasing, and obe-
sity and smoking are on the rise, all of which
lead to more cases of cancer. Women are
having their children later in life and breast-
feeding for a shorter time, which increases
their risk of breast cancer.
And treating cancer is much more dif-
cult than treating malaria or TB, Shulman
says. Diagnosis is complex, requiring com-
petently staffed and well-equipped pathol-
ogy labs. Physicians in many countries are
in short supply; oncologists are extremely
scarce. There arent enough surgeons or
operating rooms; 30 countrieshalf of
them in Africadont have a single radia-
tion therapy machine.
Still, in The Lancet paper, Shulman and
colleagues at the task force identied a list
of cancers for which a lot can be done even
in places with poor infrastructure (see table).
These measures include battling tobacco
usewhich increases the risk of various
cancers as well as cardiovascular disease
vaccinating against HPV and hepatitis B,
improving early detection, treating eminently
curable tumors such as Tushimes sarcoma
and childhood leukemia, and improving life-
extending treatment for cancers like Kaposi
sarcoma, an HIV-related cancer of the skin.
The price of drugs is not insurmount-
able, says HMS health economist and breast
cancer advocate Felicia Knaul, who runs
the task forces secretariat (see sidebar,
p. 1549). Of a list of 27 essential cancer
drugs compiled by the task force, 24 are off-
patent, and prices could be brought down
further through negotiations with the phar-
maceutical industry, Knaul says. A recent
study showed that drugs needed to treat a
case of Burkitts lymphomaa cancer that
primarily affects African children and is
associated with the Epstein-Barr virus
cost less than $50, a bargain in terms of life
years saved per dollar.
The biggest challenges in combat-
ing canceras for other diseasesare
to build up expertise and infrastructure
and extend care to the poorest people.
Some private groups are intervening
directly. For the past 18 years, pediat-
ric oncologists at St. Jude Childrens
Research Hospital in Memphis, Ten-
nessee, which sets aside about 1% of its
annual budget for global health, have
pioneered so-called twinning programs
that provide assistance to hospitals in
20 countries to improve cancer care for
children. Several other pediatric hospi-
tals have followed suit. The impact is
real: In El Salvador, where the twin-
ning started, the 5-year survival rate for
children with acute lymphoblastic leu-
kemia went from 10% to 60%, and is
still climbing.
Partners In Health (PIH), best
known for its pioneering work ght-
ing AIDS and TB in Haiti, has long
treated cancer patients as well, says
co-founder and director Farmer: They
often go from one clinic to the next,
seeking care. But PIH is also working with
governments to strengthen their health sys-
temsand expanding cancer care is a key
goal. The Rwandan government is very keen
on it, Farmer says.
Currently, the PIH network still relies on
Bostons medical infrastructure for backup.
Brigham and Womens Hospitals pathol-
ogy department examines specimens to
identify tumors, for instance, an indispens-
able part of diagnosis. (Shulman himself
has returned from Malawi with a suitcase
full of specimens. Fortunately, nobody at
customs checked, he says.) But the plan is
to develop pathology expertise and infra-
structure locally. Brigham and Womens has
donated equipment for Haiti and Rwanda,
and its chief technologist is training staff
members. Were totally committed to doing
this, Shulman says.
Going horizontal
Expanding cancer care will cost money. Take
the new HPV vaccines. So far they have been
introduced in wealthy countries, but can-
cer experts say they would prevent far more
cases of cervical cancer in poor countries,
where Pap smears are seldom done. Their
cost, more than $300 per vaccinated teenage
girl in the West, has been a barrier to their
introduction in the developing world.
But with the economy in a global dip, its
hard to see where the money will come from.
Cancer is competing for attention with the
infectious diseases, which have seen a major
increase in program support over the past
decade but are still underfunded. Meanwhile,
other diseases such as diabetes and mental
illness are vying for attention as well. Men-
tal health advocates are already disappointed
that they wont have a seat at the September
U.N. meeting on noncommunicable diseases,
where the task force is trying to make cancer
well-represented. But its not one or the other,
Farmer argues: We have to get away from the
whole notion of choosing between diseases.
Gene Bukhman, head of Harvards Pro-
gram in Global Non-communicable Disease
and Social Change, advocates building up
health systems broadly so that they can deal
with not just cancer but also diabetes, heart
disease, and a variety of other chronic ail-
ments that each make up a small percentage
of the disease burden. But this approach
sometimes dubbed horizontal as opposed
to vertical disease-specic programsisnt
particularly popular. The Bill and Melinda
Gates Foundation is spending its billions
mostly vertically, focusing on specific dis-
eases. Likewise, many advocacy groups want
to extend their worksay, on breast or prostate
cancerto the worlds poor, but theyre less
interested in helping edgling health systems.
Farmer, too, wants to channel the enthu-
siasm generated by single diseases into help
for a better health system: The breast cancer
advocate needs to see that without a proper
health system were never going to get early
detection or good treatment. Emotional
appeals can help, Bukhman says, and that is
where Tushimes pictures come in. I dont
think weve said this enough. When you see
someone dying needlessly of cancer, it is an
obscene inequalityno different from see-
ing someone die from TB or HIV.
MARTIN ENSERINK
AREAS OF OPPORTUNITY
Here are some cancer types for which experts
say major progress could be made in developing
countries:
Curable with early detection and treatment

Breast cancer

Cervical cancer
Curable with inexpensive chemotherapy

Non-Hodgkin lymphoma

Hodgkin lymphoma

Testicular cancer

Sarcoma in children

Acute lymphoblastic leukemia in children
Palliation with systemic treatment

Advanced breast cancer

Kaposi sarcoma
Preventable

Tobacco-related: lung cancer, head and
neck cancer, bladder cancer

Vaccine-preventable: cervical cancer
(HPV) and liver cancer (hepatitis B)
0325SpecialNewsSection.indd 1550 3/17/11 3:38 PM
12 SCIENCE www.sciencemag.org
www.sciencemag.org SCIENCE VOL 331 25 MARCH 2011 1551
SPECIALSECTION
C
R
E
D
I
T
S

(
T
O
P

T
O

B
O
T
T
O
M
)
:

P
E
T
E
R

A
N
D

M
A
R
I
A

H
O
E
Y
/

W
W
W
.
P
E
T
E
R
H
O
E
Y
.
C
O
M
;

(
T
A
B
L
E

S
O
U
R
C
E
)

P
U
B
M
E
D
YOUVE HEARD OF CHARLES DARWIN, BUT
do you know of his elder brother, Erasmus?
He was a physician, inventor, and philoso-
pher of some repute. Familiar with Cassandra
Austen, the amateur painter and Janes sister?
Probably not. Famous brothers and sisters
often overshadow their siblings.
The same thing happens in molecular
families. Take p53, the tumor-suppressor
protein that was named Sciences Molecule
of the Year in 1993 and has been dubbed
guardian of the genome. Few nonspe-
cialists know that the celebrated p53 is
closely related to two other proteins, p63
and p73. Yet these unheralded siblings are
grabbing the attention of cancer biologists.
New research suggests that p63 and p73 are
erce cancer killers that deserve equal bill-
ing with p53. Instead of a single genome
protector, theres a family of guardians,
says cancer biologist Elsa Flores of M. D.
Anderson Cancer Center in Houston, Texas.
Because efforts to exploit p53 in can-
cer therapies havent yet paid off, some
researchers are now looking to p73 and p63
as alternative tumor treatments. Research-
ers have shrunk or prevented tumors in ani-
mals by targeting p73, and the first clini-
cal trialsattempting to use p73 to combat
a hard-to-treat type of breast cancerhave
already started. It would be very attrac-
tive to nd a way to activate these proteins
in people with tumors, says cancer biologist
Alexander Zaika of Vanderbilt University
Medical Center (VUMC) in Nashville. Strate-
gies that capitalize on the tumor-ghting capa-
bility of the p53 family belong in the arma-
mentarium against cancer, adds molecular
oncologist Wak El-Deiry of the Penn State
Hershey Cancer Institute in Pennsylvania.
p53, the hard target
When a cell suffers DNA damage that can
lead to uncontrolled growth, p53 comes
to the rescue. The p53 protein can trig-
ger DNA repair, stop the dodgy cell from
dividing, or, when the damage is grievous,
prompt it to commit suicide. Because p53
is so potent, cells normally keep levels of
the protein low. Spurring cancer cells to
produce more protein, the argument goes,
could prompt tumors to self-destruct. p53
is a great therapeutic target, says cancer
biologist Kevin Ryan of the Beatson Insti-
tute for Cancer Research in Glasgow, U.K.
Its involvement in tumor suppression is
without question.
Scientists are pursuing a number of
strategies to enlist p53 in the cancer ght
(Science, 2 March 2007, p. 1211). They
have completed or are running several clini-
cal trials for gene therapy approaches, which
involve introducing extra copies of the p53
gene into cancer cells. Researchers at six
institutions in the United States and the
United Kingdom have also begun safety tri-
als on a compound, developed by the phar-
maceutical company Roche, that hikes p53
levels in cells by impeding the proteins nat-
ural recycling.
So far, however, no p53-based treat-
ments have been approved for clinical use
in the United States. The practical difcul-
ties are formidable. For one thing, thanks to
a variety of mutations in p53s gene, more
than half of all tumors
dont carry a working
version of the protein.
Some harbor misshapen
mutants that are inert or
that turn traitor and sub-
vert antitumor defenses.
Even when cancer cells
have a functional form of
p53, they often neutralize
it, for example, by overproducing enzymes
that prompt the proteins destruction.
Attacking many kinds of tumors through
p53 will require pharmacological feats
restoring the gene or reshaping the mal-
formed proteinthat are tougher than the
standard tactic of blocking a molecules
unwanted function. Many of us appreci-
ated it [p53] would be difcult to target,
says cancer biologist Leif Ellisen of Harvard
Medical School (HMS) in Boston. Throw-
ing a wrench into the system is much easier
than xing the system. Yet many research-
ers remain condent that they will eventu-
ally overcome these obstacles. My personal
bet is still on p53, says cancer biologist
Anna Mandinova, also of HMS. But in the
meantime, researchers are looking at other
options, namely, p63 and p73.
Band of molecular brothers
Scientists discovered the p53 protein in
1979 but didnt recognize its importance for
cancer until 10 years later. In 1997, biolo-
gists identied p73 as a molecular relative,
after discovering a DNA sequence closely
resembling the gene for p53. That came
as a bit of a surprise, says cancer biologist
Gerry Melino of the University of Rome
Tor Vergata in Italy. Nobody expected
a protein so close to p53. Yet researchers
reported another relative, p63, the next year.
Although p53 was the rst family member
discovered, p63 and p73 are the older siblings,
evolutionarily speaking. Comparisons of
their genes suggest that p53 evolved from the
ancestral version of p63 and p73 more than
450 million years ago. These elders have a
range of responsibilities. Unlike p53, p63 and
p73 are essential during embryonic develop-
ment. Shaping limbs and giving the skin its
layered structure are among p63s tasks in
an embryo. Formation of
brain regions such as the
hippocampus and cor-
tex depends on p73, as
does maturation of the
immune system. Both
proteins are also neces-
sary for female fertility.
For cancer research-
ers, p63 and p73 have
Brothers in Arms Against Cancer
Cancer researchers are trying to harness siblings of p53, the famous tumor-blocking protein
NEWS
Total PubMed Citations
p53
56,939
p63
2149
p73
1541
0325SpecialNewsSectionR1.indd 1551 3/17/11 5:18 PM
2012 Cancer Crusade at 40 Collections Booklet
www.sciencemag.org SCIENCE 13
25 MARCH 2011 VOL 331 SCIENCE www.sciencemag.org 1552
CANCER CCCCCCCCCCAAAAAAAAAANNNNNNNNNNCCCCCCCCCCEEEEEEEEEERRRRRRRRRR
Crusade at 40
C
R
E
D
I
T

(
G
R
A
P
H

S
O
U
R
C
E
)
:

P
U
B
M
E
D
a big advantage over their sibling: Their
genes are almost never mutated in or lost
from cancer cells. So p63 and p73 could in
theory take over for p53 in almost all can-
cers, and researchers wouldnt have to fret
about restoring a lost gene or reshaping a
distorted protein.
That strategy will work, of course, only if
p73 and p63 are tumor suppressors like their
brother. But that turned out to be surpris-
ingly tough to conrm. A standard experi-
ment for gauging a molecules relevance
to cancerdeleting its gene from mice
didnt clarify the issue. Whereas mice miss-
ing p53 live to adulthood and are beset by
tumors, mice lacking p63 or p73 die young
from other causes, revealing little about their
cancer susceptibility. Complicating the mat-
ter, although some studies found that levels
of p73 or p63 fell in cancer cells, suggesting
that the proteins are antitumor, other work
indicated that their levels soared, implying
that they foster abnormal growth.
The solution to that apparent contra-
diction may lie in the discovery more than
10 years ago that cells dont manufacture
just one kind of each protein. They can fash-
ion at least 12 variants, or isoforms, of the
p73 protein, and at least eight isoforms of
the p63 protein. Scientists divide these vari-
eties into the longer, or TA, isoforms and the
shorter, or N, isoforms. In 2008, Melino,
molecular geneticist Tak Mak of the Univer-
sity of Toronto in Canada, and colleagues
engineered mice that lack all of the TA iso-
forms of p73 but retain the N versions.
The mice were prone to cancer, though they
werent as vulnerable as rodents lacking
p53. The TA isoform-decient mice really
convinced people that these genes are act-
ing as tumor suppressors, says Flores.
Researchers now hypothesize that in can-
cer, TA isoforms are generally good guys,
suppressing unchecked cell growth. The N
isoforms, for the most part, are bad guys.
They can latch on to and disable p53 and
the good TA isoforms, thus aiding cancer.
For example, cancer biologist Alea Mills
of Cold Spring Harbor Laboratory in New
York and colleagues reported in February in
Cell Stem Cell that a common N isoform
of p63 spurs the growth of skin tumors.
Taking on tumors
p63 and p73 battle cancer in several ways.
Like their famous brother, the proteins cull
cells that carry potentially cancer-causing
DNA damage, activating their apoptotic, or
cell suicide, pathways. Fortuitously, some
chemotherapy already takes advantage of
this ability, causing DNA damage that spurs
p63 or p73 to kill tumor cells.
Recent studies suggest that p63 also reins
in metastasis, the migration of tumor cells to
a new location in the body; thats what usu-
ally kills cancer patients. In 2009 in Cell, a
team led by Stefano Piccolo of the Univer-
sity of Padua School of Medicine in Italy
revealed that some mutant forms of p53
found in cancer cells prevent p63 from acti-
vating two genes that curtail metastasis. And
last fall, in a study in Nature, Flores and col-
leagues showed that the TA isoforms of p63
curb metastasis through another mechanism:
boosting levels of microRNAs, RNA snip-
pets that turn down gene activity. The team
discovered that p63s TA isoforms increase
production of Dicer, an enzyme that snips
and activates inert microRNAs. The isoforms
also raise levels of a specific microRNA,
miR-130b, that prevents cells from mov-
ing on. p63 may be the master regulator of
metastasis, says Flores, and activating it
might increase levels of several metastasis-
halting microRNAs by ipping on Dicer.
The big question is whether drug design-
ers can capture p63s and p73s cancer-
quelling talents. The research directed at
this goal isnt as intense or advanced as the
work on p53, but scientists can claim some
encouraging ndings.
One therapeutic strategy attempts to
reduce sibling rivalry in the p53 fam-
ily. The mutant p53s found in cancer cells
often latch onto and neutralize p63 and p73.
Three years ago, a research group from the
Cleveland Clinic in Ohio screened more
than 46,000 compounds and pinpointed
one, named RETRA, that in test tubes
breaks mutant p53s embrace of p73. To
test whether freeing p73 destroys tumors,
the researchers then transplanted human
cancer cells into mice and
administered RETRA. The
treatment cut the number
of tumors that sprouted in
the animals, the scientists
revealed in the Proceed-
ings of the National Acad-
emy of Sciences.
To liberate p73 from a
different inhibitor that is
abundant in cancer cells,
one known as iASPP, Ryan
and colleagues ironically
turned to p53 for inspi-
ration. They developed a
snippet of p53, just 37 of
its nearly 400 amino acids, that in the test
tube separates p73 from iASPP. Dosing can-
cer-ridden mice with the snippet, known as
37AA, shrank the rodents tumors, the team
reported in The Journal of Clinical Investi-
gation in 2007.
RETRAs effects were fairly weak, and
37AA was fragile, so neither is likely to
become a drug. But the importance of stud-
ies like these, says El-Deiry, is that they
show its possible to uncover compounds that
rouse p53s siblings to attack cancer cells.
Scientists are hunting for other molecules
that might spur p73 and p63 into action and
that could make effective drugs. El-Deirys
group, for example, is in the middle of a proj-
ect to screen thousands of small molecules
theyve assessed more than 70,000 so farin
hopes of nding ones that switch on genes in
the p53 pathway. Some of the candidates, he
says, appear to work by activating p73.
At least one group has started clini-
cal trials. Medical oncologist Ingrid Mayer
and cancer biologist Jennifer Pietenpol
of VUMC are targeting a form of breast
cancerknown as triple-negative because
the tumor cells lack three key receptor pro-
teinsthat dees standard treatments such
as tamoxifen and Herceptin. The tumor cells
harbor large quantities of the N isoforms of
p63, which the researchers suspect prevent
p73 from killing the cells. Along with stan-
dard chemotherapy, patients will receive an
existing drug, everolimus, that boosts p73
levels by inhibiting a p73 blocker called
mTOR. The goal is to overcome p63s inter-
ference and enable p73 to kill the tumor cells.
Its too late for Erasmus Darwin to match
his brothers fame. But these clinical trials
and the surge of research on p63 and p73
suggest that the proteins are nally stepping
out of the shadow of their famous sibling.
MITCH LESLIE
On the attack. Researchers are testing antitumor compounds that rely
on p73.
TARGETING P73
Drug candidate Mechanism Results
37AA Disengages p73 Shrinks tumors
from inhibitor in mice
NSC176327 Boosts p73 Kills cancer cells
production in culture
Antisense Curtails production Curbs growth of
gapmers of p73s N isoforms melanoma tumors
in mice
Everolimus Increases p73 levels Trials in progress
RETRA Separates p73 from Impedes tumor
mutant p53 formation in mice
0325SpecialNewsSection.indd 1552 3/17/11 3:38 PM
14 SCIENCE www.sciencemag.org
REVI EW
Exploring the Genomes of Cancer
Cells: Progress and Promise
Michael R. Stratton*
The description and interpretation of genomic abnormalities in cancer cells have been at the
heart of cancer research for more than a century. With exhaustive sequencing of cancer genomes
across a wide range of human tumors well under way, we are now entering the end game of this
mission. In the forthcoming decade, essentially complete catalogs of somatic mutations will be
generated for tens of thousands of human cancers. Here, I provide an overview of what these
efforts have revealed to date about the origin and behavioral features of cancer cells and how
this genomic information is being exploited to improve diagnosis and therapy of the disease.
M
uch of our current understanding of
cancer is based on the central tenet that
it is a genetic disease, arising as a clone
of cells that expands in an unregulated fashion
because of somatically acquired mutations (1).
These somatic mutations include base substitu-
tions, insertions and deletions (indels) of bases,
rearrangements caused by breakage and abnor-
mal rejoining of DNA, and changes in the copy
number of DNAsegments. They also often include
epigenetic changes that are stably inherited over
mitotic DNAreplication, for example, alterations
in methylation of cytosine residues (2).
Whether a mature cancer clone emerges in an
individual person is influenced by environmental
and life-style factors, as well as by the set of ge-
nomic sequence variants present in the fertilized
egg from which the individual develops and that
are therefore found in all somatic cells. These so-
called constitutional or germline mutations can
influence cancer susceptibility in a number of ways,
including directly altering growth of the cancer
clone, altering the mutation rate in somatic cells,
or modulating the metabolism of carcinogens.
Somatic mutations are thought to occur in the
genomes of all normal cells as they proceed
through the rounds of cell division that take place
during development in utero and during replen-
ishment of body tissues in postnatal life. Addi-
tional somatic mutations continue to accumulate
in cancer cells as they divide. The rate of acqui-
sition and the types of somatic mutation that accrue
can be increased by exogenous and endogenous
exposures that cause DNA damage and are miti-
gated by DNA repair processes. Indeed, in the
event that DNArepair fails, the somatic mutation
rate may also increase.
Somatic mutations are more or less randomly
distributed throughout the genome. However, in
the cell that undergoes clonal expansion to become
a cancer, a subsettermed driver mutations
have by chance fallen in a set of key genes, called
cancer genes, and have thus subverted normal
control of cell proliferation, differentiation, death,
and other homeostatic interactions with the tissue
microenvironment (3). Driver mutations confer
growth advantage upon the neoplastic clone, allow-
ing it to expand more than normal cells from the
same tissue, invade into surrounding tissue, and,
in many cases, metastasize. The number of driver
mutations in a cancer cell reflects the number of
mutated cancer genes and thus the deregulation
of cell biological processes required to convert a
normal cell into a symptomatic cancer clone. The
remainingand often the large majority of
mutations are passengers, which, by definition,
do not confer growth advantage. The number of
passenger mutations in a cancer genome primar-
ily reflects the number of mitotic cell divisions
between the fertilized egg and the cancer cell and
the mutation rate at each of these cell divisions.
Thus, the catalog of somatic mutations in the ge-
nome of a cancer cell represents genomic changes
that usually accumulate over several decades. It
includes the mutations responsible for conferring
the various aspects of the neoplastic phenotype
and bears the imprints of the mutational processes
that caused the disease in the first place.
Cataloging Mutations in Human
Cancer Genomes
Over the past half-century a series of technologies
have been deployed to characterize systematical-
ly, at ever-increasing levels of resolution, the state
of cancer genomes across the range of cancer
types (Fig. 1). The earliest, and still one of the
most influential in its impact on cancer science,
was cytogenetic studies of chromosomes from
cancer cells. These revealed abnormalities of chro-
mosome copy number and the presence of somat-
ically acquired rearrangements (chromosomal
translocations). They showed that some cancer
types had very disordered genomes whereas others
displayed few genomic abnormalities. They also
yielded evidence that certain positions in the
genome were recurrently rearranged in particular
cancer types, from which it was inferred that a
cancer gene resided at the rearrangement break-
points. After the widespread adoption of recom-
binant DNA technology in the 1980s, it became
possible to isolate and sequence the genome in
the vicinity of these recurrently rearranged regions,
leading to the identification of many rearranged
cancer genes, particularly in leukemias, lymphomas,
and sarcomas (4).
The next major suite of technologies primar-
ily provided evidence of copy number change in
cancer genomes, but at higher resolution than was
generally possible by cytogenetics. These ap-
proaches confirmed the variation in extent of copy
number change between individual cancer genomes
and highlighted regions showing recurrent in-
creases or reductions in copy number. Subsequent
studies focusing on these recurrently abnormal
regions provided a further harvest of new cancer
genes (5, 6).
These technologies had their limitations. Most
obviously, they could not directly detect base
substitutions or small indels. The emergence of
the draft human genome sequence in 2000 em-
powered the study of cancer genomes in many
ways. In particular, it provided a template for the
design of polymerase chain reaction (PCR) primer
pairs to amplify and sequence (by conventional
sequencing technology) the coding exons of large
numbers of protein-coding genes. This facilitated
more extensive sequencing of cancer genomes,
including whole gene families and subsequently
most coding exons (728). These studies sys-
tematically sampled cancer genomes for somatic
base substitutions and small indels, providing, for
the first time, insights into their prevalence.
However, exploration of noncoding areas of the
genome and larger numbers of cases was still
restricted by high cost and limited sequencing
capacity.
The recent arrival of second-generation DNA
sequencing technologies (29) has further trans-
formed investigation of cancer genomes. These
technologies are being applied in a number of ways.
Because most of the currently known driver mu-
tations change the coding sequences of protein-
coding genes and because protein-coding exons
account for only ~1% of the human genome, se-
quencing is often being thriftily targeted at these
(3032). Use of technologies that extract subsets
of DNA sequences from the whole genome (33),
in combination with second-generation sequenc-
ing, has already allowed sequencing of the
protein-coding exons of roughly 2000 individual
cancers worldwide. This strategy will find base
substitutions and indels in coding exons (and
potentially copy number changes) but will miss
these types of mutation in noncoding regions and
require other analyses of the same genomes to
report most rearrangements. To a similar end, after
extraction of RNA, the transcriptomes of many
SPECIALSECTION
Wellcome Trust Sanger Institute, Hinxton, Cambridge CB10
1SA, UK.
*E-mail: mrs@sanger.ac.uk
www.sciencemag.org SCIENCE VOL 331 25 MARCH 2011 1553
2012 Cancer Crusade at 40 Collections Booklet
www.sciencemag.org SCIENCE 15
hundreds of cancers have been sequenced (3436).
This approach will report substitutions in genes that
have sufficiently high levels of mRNA and can
report rearrangements that are transcribed. Again,
however, abnormalities of noncoding regions will
generally be missed, and protein-truncating muta-
tions maybe difficult tofindif theyactivate nonsense-
mediated RNA decay.
In the longer term, however, the major impact
of these remarkable technology shifts will be to
permit the sequencing of whole cancer genomes
(3743). This strategy, in which genomic DNA
from a cancer (and, in parallel, DNAisolated from
normal tissue of the same person) is randomly
fragmented and hundreds of millions of fragments
are sequenced, can reveal all classes of somatic
change (base substitutions, indels, rearrangements,
copy number changes, and even potentially ep-
igenetic alterations) in all sectors of the genome
(exons, introns, and intergenic regions). It has
paved the way to the generation of almost com-
plete catalogs of somatic mutation for individual
cancers (39, 40), allowing us to set aside our pre-
conceptions of where the important mutations that
cause the disease might lie and, by acquisition of
large numbers of mutations from individual cases,
empowering deeper study of the mutational pro-
cesses that have been operative. A few hundred
whole cancer genomes have already been gener-
ated by sequencing machines worldwide and are
in the process of being analyzed.
The Number of Mutations in Cancer Genomes
As noted above, cytogenetic and copy number
studies revealed that the number of genomic re-
arrangements and copy number changes can dif-
fer markedly between individual cancers. Until
the recent advent of systematic sequencing studies,
however, we had little insight into the numbers of
somatic base substitutions and indels and the
extent of their variation.
We now know that there are usually between
1000 and 10,000 somatic substitutions in the
genomes of most adult cancers, including breast,
ovary, colorectal, pancreas, and glioma (10, 21).
There are cancer types that generally carry relative-
ly few mutationsfor example, medulloblastomas,
testicular germcell tumors, acute leukemias, and car-
cinoids (10, 16)whereas others, suchas lungcan-
cers and melanomas, have many more mutations
(occasionallymore than100,000) (9, 10, 27, 39, 40).
Even within a particular cancer type, individual
tumors often display wide variation in the prev-
alence of base substitutions.
Two major factors account for these differ-
ences in mutation prevalence: differences between
individual cancers in mutation rate at the cell di-
visions that have taken place between the fertilized
egg and the cancer cell and differences in the num-
ber of mitoses in this lineage. The basis for the
high prevalence of somatic substitutions observed
in some cancers is likely to be overwhelming mu-
tagenic exposure such as ultraviolet (UV) light (in
melanoma) or tobacco carcinogens (in lung cancer);
the presence of defective DNA repair mechanisms
(e.g., in colorectal, stomach, and other cancers with
defective DNAmismatch repair); and therapy with
DNA-damaging agents (e.g., in gliomas treated
with the alkylating agent temozolomide) (10, 11).
However, there are individual cancer cases in which
the large number of base substitution mutations
remains unexplained (18).
The reason that some cancer types have rel-
atively few mutations is not completely clear.
Some are tumors of children or young adults, and
therefore it is conceivable that the neoplastic cell
has been through relatively few DNA replica-
tions. Alternatively, it may be that most cancers,
including those with the typical mutation preva-
lence, have experienced an elevated mutation rate
compared to normal cells and that cancers with
low mutation prevalence are the exceptions that
have evolved without it. Because we currently
know little about the prevalence of somatic base
1900 1920 1940 1960 1980 2000 2020 2040
Identification of DNA
as the material of
inheritance
Description of the
double helical
structure of DNA
Second-generation
sequencing
technologies
First sequence of all
exons in a cancer
First complete cancer
genome sequence
Cancer genome
sequences as a
routine diagnostic?
Thousands of complete
cancer genome sequences
400 known cancer genes
First recurrent
chromosomal
rearrangement
in cancer
First somatic
driver mutation
and first cancer
gene identified
Human
genome
sequence
First observations that the
material of inheritance was
abnormal in cancer cells and
consequent proposal that
cancers are clones arising
due to somatic changes
Fig. 1. Time line showing key events in the investigation of the cancer genome.
25 MARCH 2011 VOL 331 SCIENCE www.sciencemag.org 1554
Cancer Crusade at 40
16 SCIENCE www.sciencemag.org
substitutions in normal cells, the importance of an
elevated base substitution mutation rate in cancer
development remains controversial (44, 45).
However, for indels in cancer cases with DNA
mismatch-repair deficiency and for rearrange-
ments and copy number changes in cancers that
have large numbers of these changes, an increased
mutation rate is likely to have been operative.
Systematic sequencing studies have also pro-
vided our first comprehensive insights into the
proportions of driver and passenger mutations.
Thus far, the large majority of base substitutions
in most cancer genomes appear to be passengers
(10, 21). However, these studies also suggest that
there may be many more drivers than can be
unambiguously identified by current approaches.
If the latter interpretation is correct, a substantial
number of cancer genes remain to be discovered,
albeit many contributing infrequently to cancer
development (10, 21).
The Repertoire of Human Cancer Genes
Studies of driver mutations in cancer genes have
yielded many insights into the molecular and
cellular events that convert a healthy cell into a
cancer cell. In recent years, the proteins altered by
driver mutations have become targets for success-
ful anticancer drug development (46). Identifica-
tion of new mutated cancer genes is, therefore,
one of the most important deliverables that
emanates from exploration of cancer genomes.
The primary analytic approach to the identi-
fication of driver mutations and cancer genes has
assumed that passenger mutations are randomly
distributed throughout the genome, whereas driv-
ers (by definition) are clustered in the subset of
genes that are cancer genes. The strategy is thus
to search in a large number of samples of a
specific cancer type, for genes that have a higher
prevalence of somatic mutations than would be
expected by chance alone, followed by verifica-
tion of biological activity in experimental systems.
This basic strategy has been highly effective over
decades and remains the mainstay of cancer gene
discovery today. However, it has become clear
that passenger mutations are not always random-
ly distributed in the genome; their clustering can
mimic that of driver changes, and thus additional
filtering strategies may sometimes be required to
avoid errors in cancer gene identification (47).
The search for cancer genes through system-
atic exploration of cancer genomes by cyto-
genetics, copy number analyses, and sequencing
has been supplemented by targeted somatic
mutational analyses of genes previously identi-
fied as cancer susceptibility genes, by mutational
analyses of biologically plausible candidates, and
by biological assays of transforming activity, no-
tably DNAtransfection through NIH3T3 cells. Col-
lectively, these varied approaches have identified
~400 somatically mutated cancer genes that con-
tribute to neoplastic change in one or more types of
cancer. This number corresponds to roughly 2%of
the protein-coding genes in the human genome
(5, 6) (www.sanger.ac.uk/genetics/CGP/Census/).
Cancer genes are often classified according to
whether they function in a dominant or recessive
manner at the level of the cancer cell. Dominant
cancer genes require only one of the two parental
alleles present in a normal cell to be mutated, and
the encoded protein is usually constitutively ac-
tivated by the mutations. Recessive cancer genes
(also known as tumor suppressor genes) require
mutation of both parental alleles, and these usually
result in inactivation of the encoded protein.
More than 80% of the currently known cancer
genes are dominantly acting; mostly these are
genes that are rearranged in the myriad recurrent
chromosomal translocations particularly found
in leukemias, lymphomas, and sarcomas (www.
sanger.ac.uk/genetics/CGP/Census/). The current
predominance of dominantly acting cancer genes
is in part due to ascertainment bias, and the real
balance remains to be determined.
Most of the known cancer genes were found
through primary cytogenetic analyses, with the
wave of ever higher resolution copy number studies
bringing a further substantial yield. Recent system-
atic sequencing of cancer genomes has provided a
new harvest of cancer genes identified directly
through an elevated prevalence of base substitu-
tions and small indels. These include several
dominant cancer genes, such as BRAF, EGFR,
ERBB2, PIK3CA, IDH1, IDH2, EZH2, FOXL2,
PPP2R1A, and JAK2 (8, 1215, 17, 34, 36, 48, 49)
(some of which were also found by alternative
approaches). Some are on biological pathways pre-
viously implicated in cancer development. Others
for example, IDH1, which encodes isocitrate de-
hydrogenase 1, a component of the Krebs cycle;
or FOXL2, which encodes a tissue-specific tran-
scription factorwould not have featured on many
candidate gene lists.
Several recessive cancer genes (and others for
which the dominant or recessive status is unclear)
have also emerged through systematic sequencing,
including SETD2, KDM6A, KDM5C, PBRM1,
BAP1, ARID1A, DNMT3A, GATA3, DAXX, ATRX,
and MLL2 (7, 12, 16, 19, 20, 3032, 35, 50). Many
of the proteins encoded by this set of genes (and,
in addition, EZH2 and IDH1 among the domi-
nant cancer genes mentioned above) are involved
in chromatin modification and remodeling. For
example, SETD2, EZH2, and MLL2 are histone
H3 methylases, whereas KDM6A and KDM5C
are histone H3 demethylases. ARID1A, PBRM1,
BAP1, ATRX, and DAXX are components of
protein complexes that restructure chromatin, and
DNMT3A is involved in maintenance of cyto-
sine methylation in DNA. Although this sector of
cell biology was previously known to be dis-
rupted through mutation in some cancers, these
discoveries have placed new emphasis on its role
in a range of adult and childhood solid tumor
types and highlight a potentially important link
between somatic mutation and epigenetic changes
that are present in many cancers. This area of
biology promises to be a major focus of activity in
the development of new cancer therapeutics.
Some newly discovered cancer genesfor
example, BRAF, JAK2, ARID1A, EZH2, BAP1,
PBRM1, and DNMT3Aare mutated in a sub-
stantial proportion of cases of a particular cancer
type. Others, such as SETD2 and KDM5C, are
mutated in only a small fraction of cancers of any
class. This appears to be an emerging feature of
the landscape of somatically mutated cancer genes,
of which a relatively limited set are commonly
mutated and a substantial number mutated infre-
quently. From the standpoint of novel drug dis-
covery, the latter presents obvious challenges.
An important perspective on the evolution of
the cancer clone is provided by the number of
mutated cancer genes required to generate an
individual human cancer. It is often speculated
that five mutated cancer genes are necessary (51).
However, higher estimates have been suggested,
and for some hematopoietic neoplasms fewer
may be required. The presence of two to four
driver mutations has been demonstrated in many
cases of various cancer types. In a few years, we
will be able to estimate this core metric of cancer
biology, and the extent to which it varies, more
accurately. Once large numbers of cancer ge-
nomes have been completely sequenced with all
classes of somatic mutation harvested and once
most cancer genes have been identified, robust
direct assessments of the number of mutated
cancer genes in individual cancers will become
achievable.
The Cancer Genome and Drug Discovery
The central role of mutated cancer genes in the
genesis and maintenance of cancer clones renders
them potential Achilles heels to be exploited
for drug discovery. There are now several cele-
brated examples of anticancer drugs that act by
inhibiting the aberrantly activated proteins encoded
by mutated cancer genes (46). A paradigm of
such strategies is the development of imatinib
and subsequent generations of small-molecule
inhibitors of the constitutively activated ABL ki-
nase engendered by the chromosome 9:22 trans-
location in chronic myeloid leukemia (CML) (52).
This advance has transformed the treatment of
CML and, on the way, has helped to revolution-
ize cancer therapeutics. Small-molecule drugs
against mutated versions of EGFR, ERBB2, KIT,
PDGFRA, PML-RARA, MET, and ALK are ei-
ther already in clinical use or being evaluated in
clinical trials (46, 53). Similarly, a therapeutic an-
tibody (trastuzumab) directed against HER2, the
protein encoded by a gene amplified in about
20%of breast cancers, has had a major impact on
treatment of these cancers (54).
An illustrative example of the combined power
of modern genomics, biology, and drug discov-
ery is that of BRAF. Somatic mutations of BRAF
were discovered in an early systematic sequenc-
www.sciencemag.org SCIENCE VOL 331 25 MARCH 2011 1555
SPECIALSECTION
2012 Cancer Crusade at 40 Collections Booklet
www.sciencemag.org SCIENCE 17
ing screen in 2002 (8). BRAF encodes a serine-
threonine kinase and is mutated in 50 to 70% of
malignant melanomas, 10 to 15% of colorectal
cancers, 50% of papillary thyroid cancers, and at
a lower frequency in other cancer types. A single
mutation, V600E (substitution of valine 600 with
glutamic acid), accounts for more than 90% of
mutations and results in constitutive activation of
the BRAF kinase. Being a kinase (with a deep
ATP-binding pocket in which inhibitors can sit)
and being activated by its mutations made mu-
tated BRAF an attractive target for drug de-
velopment. Inhibitors of V600E
mutant BRAF have been tested in
phase 1 trials and have produced
encouraging responses in 80% of
patients with metastatic malignant
melanomas carrying the V600E
mutation (55).
Unfortunately, minor subclones
resistant to BRAF inhibitors ap-
pear to be present in many V600E-
positive malignant melanomas, and
these growout as recurrences. Never-
theless, investigation into the ge-
nomes of recurrences has already
identified some of the mutations that
confer resistance, proffering new
avenues for therapeutic interven-
tion (56, 57). Thus, in the decade since the dis-
covery of BRAF as a mutated cancer gene, the
field has seen small-molecule inhibitors identified
and developed into orally available drugs, the
drugs put through clinical trials and shown to have
anticancer activity, and mechanisms of resistance
to the drugs elucidated. Although we collectively
aspire to even more rapid progress in the future,
this is a remarkable achievement.
In some cancers, direct targeting and inhibi-
tion of constitutively activated proteins encoded
by mutated cancer genes may not be possible.
For example, in clear cell renal cancer, mutated
and activated kinases have not been found.
Indeed, all the operative cancer genes appear to
be recessive (7, 20, 31). Because the proteins
encoded by these genes are already inactivated
by their mutations, other strategiesfor exam-
ple, the development of drugs that exhibit syn-
thetic lethality with particular mutated cancer
genes (58)will have to be adopted.
Genomic Evidence of Mutagenic
and Repair Processes
The patterns of somatic mutation found in a can-
cer genome reflect the DNA damage and muta-
genic processes that have been operative and the
repair mechanisms that have mitigated their
impact. Thus, the cancer genome can be likened
to an archaeological record bearing the imprint of
these processes. The mutational patterns (often
called mutational spectra) incorporate many
types of information, including the numbers of
each class of mutation, the DNAsequences around
each mutated base, and, in transcribed regions,
whether the transcribed or the untranscribed strand
is preferentially mutated.
In the past, mutational spectra have been
assembled with the use of mutations found in
frequently mutated cancer genes, notably the tumor
suppressor gene TP53. In such studies, each in-
formative cancer case usually contributes a single
mutation, and the spectrum is established by
grouping together mutations from multiple cases
of the same cancer type. This approach demon-
strated that lung cancers exhibit many C:G>A:T
transversion mutations, a pattern similar to that
induced in experimental systems by tobacco
carcinogens; that hepatocellular cancers also show
C:G>A:T mutations that are likely to be induced
by aflatoxins, known etiological agents in liver
cancer development; and that skin cancers
predominantly show C:G>T:A mutations of the
pattern known to be caused by UV light (59).
However, a major limitation of these studies is that
mutational spectra generated in this way are
composites of all the mutational spectra present
in a tumor class. Thus, although well powered to
report a strong exposure that dominates a partic-
ular cancer type, they cannot untangle the diverse
mutational processes and patterns that may be
present in some cancer types.
By contrast, partial or complete catalogs of
mutations from individual cancer genomes,
which usually number several thousand somatic
mutations per case, can report with extraordinary
resolution the mutational spectra of individual
cancers, thus revealing the diverse mutational
and repair processes operative within a class of
cancer and even within individual cases. This
type of analysis is in its infancy, but some ex-
amples illustrate its potential. Early systematic
sequencing studies revealed the presence in some
breast cancers of a mutational process charac-
terized by C>T and C>G mutations that occur
almost exclusively at cytosines that follow a
thymine [i.e., at TpC dinucleotides (18)]. The na-
ture of the mutagenic process underlying this pat-
tern of mutations remains mysterious, but future
epidemiological studies correlating its presence
with exogenous exposures and studies replicating
the pattern by examining the effects of chemicals
or DNA repair defects in experimental systems
may elucidate its origin. In the case of melanoma
and lung cancers, the use of essentially complete
catalogs of thousands of mutations from the ge-
nomes revealed the predominant mutational classes
expected of the known exposures underlying these
cancers (39, 40). Although UVexposure accounted
for most mutations in melanoma, there was evi-
dence for at least one additional mutational pro-
cess, the spectrumof which suggested that it may
have been due to reactive oxygen spe-
cies (39). Similarly, by examining the
DNA sequences around somatically
mutated bases in a case of lung can-
cer, it was possible to tease out mul-
tiple distinct mutational processes that
may reflect the complexity of the car-
cinogen mixture present in cigarette
smoke (40).
Traces of DNA repair processes
are also embedded in mutational spec-
tra. For example, in individual mela-
noma and lung cancer cases, evidence
has been found for past activity of
transcription-coupled repair, a subclass
of nucleotide excision repair that is
directed at the transcribed strand of
each gene (3840). These completely sequenced
cancer genomes also revealed that nucleotide ex-
cision repair had been preferentially deployed
to the untranscribed strand of genes, that repair
correlated with the expression level of the target
gene, and that 5 ends of genes had been more
effectively repaired than 3 ends.
Sequencing of cancer genomes has revealed
unexpected features of mutational processes
beyond those that cause base substitutions. For
example, some cancers display many more ge-
nomic rearrangements than would have been
predicted on the basis of cytogenetic studies (60).
Indeed, different types of rearrangement archi-
tecture predominate in different cancer types. In
some breast cancers, for example, there are frequent
tandem duplications of DNA (60), whereas in
pancreatic cancers this pattern is rare (61). The
genetic defects, or possibly environmental expo-
sures, that underlie these distinctive patterns of
genomic rearrangement are unknown.
Insights have also emerged with respect to the
timing of mutations. In principle, some muta-
tional processes may cause steady accumulation
of mutations over decades, whereas others may
be characterized by a sharp burst over a short
period. A small proportion of cancer genomes
exhibit a distinctive pattern characterized by extra-
ordinary numbers of rearrangements localized
to a small segment of the genome. In such re-
gions, the genome appears to have been shat-
tered and subsequently reassembled by the cell,
albeit in a disordered manner. The structure of
these dense aggregates suggests that the rear-
25 MARCH 2011 VOL 331 SCIENCE www.sciencemag.org 1556
Cancer Crusade at 40
18 SCIENCE www.sciencemag.org
rangements occurred more or less synchronously,
possibly within a single catastrophic cell cycle,
rather than in a serial manner over many cell di-
visions (62).
Revealing the Tree of Clonal Evolution
in Cancer
Although each cancer derives from a single
normal cell, the population of neoplastic cells
constituting the final cancer often has a complex
evolutionary history, which can be visualized in
the following way. Multiple waves of clonal
expansion are thought to be required, each brought
about by an additional driver mutation, to generate
the dominant subclone that manifests as the
symptomatic cancer. Along the way, additional
branching subclones with further drivers may
have been spun off that failed to outcompete
the dominant subclone. The dominant subclone
itself may have spawned a further minor sub-
clone with an additional driver mutation that in
time would dominate. Some of these minor sub-
clones may have been completely extinguished,
but others may persist. Thus, the final population
of cancer cells is composed of the dominant
subclone accompanied by relics of its evolution-
ary past, outstripped rivals, and portents of its
future.
Somatic mutations acquired by cancer cells as
they divide can serve as markers of clonal origin
and thus allowretrospective reconstruction of the
evolutionary tree of individual cancers. These
analyses have revealed the complex subclonal
structure of certain cancers and have demon-
strated that minor subclones at initial presentation
of the cancer are often the source of the major
clone that recurs after treatment (6366).
Metastases have been analyzed by similar
methods, and these studies indicate that they are
usually subclones of the primary cancer. Com-
parison of somatic changes in metastases to those
of the primary tumors fromwhich they originated
has revealed that many likely passed through a
clonal bottleneck, continued to acquire somatic
changes, and diverged from the primary cancer
(37, 41, 61, 67). Comparing genomic changes in
different metastases from the same patient has
yielded provocative insights. For example, in
some cases multiple distinct metastases appar-
ently originated fromthe same minor subclone of
the primary tumor, suggesting that this subclone
possessed enhanced metastatic potential com-
pared to the bulk of the primary tumor (37, 61).
Unexpected relationships between metastases have
also surfaced. For example, in some pancreatic
cancers metastatic to the lung and to the abdomen,
the lung metastases shared a set of somatic changes
with each other, and the abdominal metastases
shared a different set (in addition to the mutations
both groups of metastases shared with the primary
cancer). These results suggest that, rather than
each metastasis being a direct offshoot of the
primary cancer, a single seed of pancreatic cancer
reached the lung and then reseeded further in this
organ to generate the multiple metastases ob-
served, and similarly, a single metastasis seeded
in the abdomen and then reseeded elsewhere in
the peritoneal cavity (61).
The Cancer Genome as a Personalized Diagnostic
As noted above, the successful development of
drugs against proteins encoded by somatically
mutated cancer genes has helped to revolutionize
cancer therapeutics in the past decade. In many
cases, such drugs are only effective against cancers
carrying the relevant mutated cancer gene. Test-
ing for the presence of the mutated gene in biopsies
as a prelude to administering the drug is therefore a
rapidly expanding area of cancer diagnostics that
seems certain to be integrated into future clinical
practice.
There are additional ways in which knowl-
edge of the cancer genome can potentially be
used to improve patient care. Many cancers leak
DNA into the circulation as cells die. Detection
of somatic changes present in the cancer genome
can, in principle, distinguish circulating DNAorig-
inating from the cancer from circulating DNA
derived from normal cells. Such tests would po-
tentially allow monitoring of tumor burden from
measurements on blood samples and might have
utility in a variety of circumstances, including
evaluation of response to treatment and early de-
tection of recurrence. Similar approaches have
been used for many years to monitor disease
burden in several types of leukemia. In these
diseases, they have been applicable because of
the presence of common recurrent driver rear-
rangements (translocations) that allow design of
PCR assays across the rearrangement junction;
these assays serve as sensitive and specific tests
that are straightforward to implement clinically.
This mode of diagnostic has not generally been
employed in solid tumors, in part because there
are relatively fewexamples of common recurrent
rearrangements. However, most solid tumors do
carry multiple passenger rearrangements that are
specific to each individual cancer. The possibil-
ity of sequencing a cancer genome as a real-time
diagnostic to find these rearrangements offers the
potential of developing customized tests of cir-
culating DNA to determine the tumor burden of
most cancer patients. Early proof-of-principle
studies indicate that this approach is technically
feasible (68, 69), and its benefit for patients is
being evaluated.
The leakage of mutated DNA from cancers
into blood or other body fluids also raises the
possibility of early diagnosis by detection of cir-
culating cancer-derived DNA before a cancer be-
comes symptomatic and the tumor burden high.
This is a longer-term vision with additional at-
tendant technical challenges. As with all screening
Drug targets
Early
diagnosis
Drug resistance
Biology of
neoplastic change
Progression
and response
to therapy
Monitoring
cancer burden
Evolution of
the cancer clone
Understanding
metastasis
Mechanisms
of DNA damage
DNA repair
processes
Classification
of cancer The cancer
genome
Fig. 2. Cancer genome analysis is expected to have a far-reaching impact on our understanding of cancer
biology and will likely prompt new approaches to the detection, diagnosis, treatment, and possibly
prevention of the disease.
www.sciencemag.org SCIENCE VOL 331 25 MARCH 2011 1557
SPECIALSECTION
2012 Cancer Crusade at 40 Collections Booklet
www.sciencemag.org SCIENCE 19
approaches, its utility in clinical practice will ul-
timately depend on its sensitivity, false-positive
rate, and impact on mortality.
The Future
The march toward exhaustive sequencing of
cancer genomes across the range of tumor classes
is now under way. This global enterprise is being
conducted under the auspices of the International
Cancer Genome Consortium (70) and currently in-
cludes large-scale sequencing initiatives in Aus-
tralia, Canada, China, France, Germany, India,
Italy, Japan, Mexico, Spain, the United Kingdom,
and the United States. Sequencing of several
hundred cases of each major cancer subtype is
envisaged to provide sufficient statistical power
to detect cancer genes that may be operative in
only 5% of cases. Over the next 5 to 7 years, it is
realistically anticipated that tens of thousands of
cancer genomes will be sequenced. Initially, there
will continue to be a diversity of approaches,
with some studies sequencing the DNA of exons
of protein-coding genes while others continue to
analyze transcriptomes. It is likely, however, that
these large-scale initiatives will ultimately con-
verge on whole-genome sequencing, coupled to
exploration of the transcriptome and epigenome
from the same cases. This convergence will be
encouraged by the falling cost of whole-genome
sequencing; by the convenience of harvesting all
classes of somatic mutation in one experiment;
and by the insight that, albeit large, the human
genome is finite and that we should exploit this
advantageous attribute. The only way of being
sure that nothing important has been missed is to
examine it all.
Collectively, the outcomes of these studies are
expected to have an overarching influence on our
understanding of cancer biology and prompt new
approaches to therapy and potentially prevention
(Fig. 2). They will reveal the full repertoire of
mutated cancer genes that operate across the most
common forms of human cancer, will provide us
with a clear picture of the number and combina-
tions of mutated cancer genes required to gen-
erate each individual cancer, and will shed light
on the mutational and repair processes that have
been operative in generating neoplastic clones in
the first place. Through analysis of samples from
early preinvasive lesions, from metastases, from
recurrences after therapy, and from patients with
known exposures or epidemiological risk factors,
these studies should also provide insights into
disease pathogenesis, progression, and mecha-
nisms of drug resistance.
These studies will also establish a new, com-
prehensive, and biologically rational classification
of human cancer based on genomic abnormalities.
As with any new classification, it will be neces-
sary to evaluate, in a further wave of research, the
ability of this genomic classification scheme to
predict the key features of tumor behavior of most
concern to us, notably progression and response
to therapeutics. We already know that the pres-
ence of certain mutated genes determines re-
sponse to some therapies, and mutational testing
of specific genes has already been introduced into
some clinical trials, particularly when the mutated
gene is the target of the new therapy being eval-
uated. However, with an essentially complete set
of cancer genes to be revealed in a few years, and
the plausibility that some are likely to influence
the clinical behavior of cancer, the ultimate goal
should be to examine the prognostic and predictive
effectiveness of all mutated cancer genes present in
each cancer type, much as similar waves of re-
search in the past have correlated cancer behavior
with clinical parameters, pathology, or specific
biomarkers. In principle, this assessment should be
implemented systematically for both existing and
new patient treatment protocols.
What sort of test design could accomplish
this? Each cancer type is driven by different, al-
though often overlapping, sets of mutated cancer
genes. Thus, customized tests for each cancer
class might be an option. However, a single test
that could be applied to all types of cancer and
access all the relevant information in each type is
an especially attractive prospect. The complete
catalog of somatic mutations provided by the
sequence of the cancer genome fits that descrip-
tion. Although currently expensive for routine
implementation, it is unlikely to remain so for
long, and the costs of performing a cancer ge-
nome sequence in 10 years will be insignificant
compared to other aspects of conducting clinical
trials. Thus, a full cancer genome sequence may
well turn out to be a pragmatic test design for this
next phase of research. One should not, however,
underestimate the technical, scientific, and analytic
challenges intrinsic to this proposal. Moreover,
such a test is unlikely to replace all other intrinsic
predictors of cancer behavior. Nevertheless, given
the rich seam of information that we know is
buried in each cancer genome, the extraordinary
pace of technological advance in sequencing, and
the practical advantage of using a single test in
diverse clinical contexts (including many outside
oncology), it seems reasonable to look forward to
a time in the not-so-distant future when we will
consider a cancer genome sequence as a routine
adjunct in clinical trials and a test we will perform
on many newly diagnosed cancers.
References and Notes
1. M. R. Stratton, P. J. Campbell, P. A. Futreal, Nature 458,
719 (2009).
2. P. W. Laird, Hum. Mol. Genet. 14 (suppl. 1), R65 (2005).
3. D. Hanahan, R. A. Weinberg, Cell 100, 57 (2000).
4. F. Mitelman, B. Johansson, F. Mertens, Nat. Rev. Cancer
7, 233 (2007).
5. P. A. Futreal et al., Nat. Rev. Cancer 4, 177 (2004).
6. T. Santarius, J. Shipley, D. Brewer, M. R. Stratton,
C. S. Cooper, Nat. Rev. Cancer 10, 59 (2010).
7. G. L. Dalgliesh et al., Nature 463, 360 (2010).
8. H. Davies et al., Nature 417, 949 (2002).
9. L. Ding et al., Nature 455, 1069 (2008).
10. C. Greenman et al., Nature 446, 153 (2007).
11. C. Hunter et al., Cancer Res. 66, 3987 (2006).
12. S. Jones et al., Science 330, 228 (2010).
13. R. L. Levine et al., Cancer Cell 7, 387 (2005).
14. J. G. Paez et al., Science 304, 1497 (2004).
15. D. W. Parsons et al., Science 321, 1807 (2008).
16. D. W. Parsons et al., Science 331, 435 (2011).
17. Y. Samuels et al., Science 304, 554 (2004).
18. P. Stephens et al., Nat. Genet. 37, 590 (2005).
19. J. Usary et al., Oncogene 23, 7669 (2004).
20. G. van Haaften et al., Nat. Genet. 41, 521 (2009).
21. L. D. Wood et al., Science 318, 1108 (2007).
22. T. Sjblom et al., Science 314, 268 (2006).
23. R. McLendon et al., Cancer Genome Atlas Research
Network, Nature 455, 1061 (2008).
24. H. Davies et al., Cancer Res. 65, 7591 (2005).
25. D. W. Parsons et al., Nature 436, 792 (2005).
26. A. Bardelli et al., Science 300, 949 (2003).
27. Z. Kan et al., Nature 466, 869 (2010).
28. L. M. Scott et al., N. Engl. J. Med. 356, 459 (2007).
29. J. Zhao, S. F. Grant, Curr. Pharm. Biotechnol. 12, 293
(2010).
30. J. W. Harbour et al., Science 330, 1410 (2010).
31. I. Varela et al., Nature 469, 539 (2011).
32. Y. Jiao et al., Science 331, 1199 (2011).
33. L. Mamanova et al., Nat. Methods 7, 111 (2010).
34. R. D. Morin et al., Nat. Genet. 42, 181 (2010).
35. K. C. Wiegand et al., N. Engl. J. Med. 363, 1532
(2010).
36. S. P. Shah et al., N. Engl. J. Med. 360, 2719 (2009).
37. L. Ding et al., Nature 464, 999 (2010).
38. W. Lee et al., Nature 465, 473 (2010).
39. E. D. Pleasance et al., Nature 463, 191 (2010).
40. E. D. Pleasance et al., Nature 463, 184 (2010).
41. S. P. Shah et al., Nature 461, 809 (2009).
42. E. R. Mardis et al., N. Engl. J. Med. 361, 1058 (2009).
43. T. J. Ley et al., Nature 456, 66 (2008).
44. L. A. Loeb, Semin. Cancer Biol. 20, 279 (2010).
45. W. Bodmer, J. H. Bielas, R. A. Beckman, Cancer Res. 68,
3558, discussion 3560 (2008).
46. D. Stuart, W. R. Sellers, Curr. Opin. Cell Biol. 21, 304
(2009).
47. G. R. Bignell et al., Nature 463, 893 (2010).
48. P. Stephens et al., Nature 431, 525 (2004).
49. H. Yan et al., N. Engl. J. Med. 360, 765 (2009).
50. T. J. Ley et al., N. Engl. J. Med. 363, 2424 (2010).
51. C. Hornsby, K. M. Page, I. P. Tomlinson, Lancet Oncol. 8,
1030 (2007).
52. B. J. Druker, Blood 112, 4808 (2008).
53. E. L. Kwak et al., N. Engl. J. Med. 363, 1693 (2010).
54. F. J. Esteva, D. Yu, M. C. Hung, G. N. Hortobagyi,
Nat. Rev. Clin. Oncol. 7, 98 (2010).
55. K. T. Flaherty et al., N. Engl. J. Med. 363, 809 (2010).
56. R. Nazarian et al., Nature 468, 973 (2010).
57. C. M. Johannessen et al., Nature 468, 968 (2010).
58. F. L. Rehman, C. J. Lord, A. Ashworth, Nat Rev. Clin.
Oncol. 7, 718 (2010).
59. G. P. Pfeifer, Mutat. Res. 450, 1 (2000).
60. P. J. Stephens et al., Nature 462, 1005 (2009).
61. P. J. Campbell et al., Nature 467, 1109 (2010).
62. P. J. Stephens et al., Cell 144, 27 (2011).
63. P. J. Campbell et al., Proc. Natl. Acad. Sci. U.S.A. 105,
13081 (2008).
64. C. G. Mullighan et al., Science 322, 1377 (2008).
65. F. Notta et al., Nature 469, 362 (2011).
66. K. Anderson et al., Nature 469, 356 (2011).
67. S. Yachida et al., Nature 467, 1114 (2010).
68. D. J. McBride et al., Genes Chromosomes Cancer 49, 1062
(2010).
69. R. J. Leary et al., Sci. Transl. Med. 2, 20ra14 (2010).
70. International Cancer Genome Consortium, Nature 464,
993 (2010).
71. M.R.S. thanks A. Futreal, P. Campbell, U. McDermott,
N. Rahman, and many other colleagues for conversations
over the years that have clarified ideas that have found
their way into this Review. Supported by the Wellcome
Trust under grant reference 077012/ Z/ 05/Z.
10.1126/science.1204040
25 MARCH 2011 VOL 331 SCIENCE www.sciencemag.org 1558
Cancer Crusade at 40
20 SCIENCE www.sciencemag.org
REVI EW
A Perspective on Cancer
Cell Metastasis
Christine L. Chaffer
1,3
* and Robert A. Weinberg
1,2,3
*
Metastasis causes most cancer deaths, yet this process remains one of the most enigmatic aspects
of the disease. Building on new mechanistic insights emerging from recent research, we offer our
perspective on the metastatic process and reflect on possible paths of future exploration. We
suggest that metastasis can be portrayed as a two-phase process: The first phase involves the
physical translocation of a cancer cell to a distant organ, whereas the second encompasses the
ability of the cancer cell to develop into a metastatic lesion at that distant site. Although much
remains to be learned about the second phase, we feel that an understanding of the first phase is
now within sight, due in part to a better understanding of how cancer cell behavior can be
modified by a cell-biological program called the epithelial-to-mesenchymal transition.
M
etastasis is responsible for as much as
90% of cancer-associated mortality, yet
it remains the most poorly understood
component of cancer pathogenesis. During meta-
static dissemination, a cancer cell from a primary
tumor executes the following sequence of steps:
It locally invades the surrounding tissue, enters
the microvasculature of the lymph and blood
systems (intravasation), survives and translocates
largely through the bloodstream to microvessels
of distant tissues, exits from the bloodstream
(extravasation), survives in the microenvironment
of distant tissues, and finally adapts to the foreign
microenvironment of these tissues in ways that
facilitate cell proliferation and the formation of a
macroscopic secondary tumor (colonization) (1).
Here we suggest that this complex metastatic
cascade can be conceptually organized and sim-
plified into two major phases: (i) physical trans-
location of a cancer cell from the primary tumor
to the microenvironment of a distant tissue and
then (ii) colonization (Fig. 1). We propose that
an understanding of physical dissemination is in
sight, whereas the second phase, colonization,
involves complex interactions that may still re-
quire several years of research before they come
into clear view. From a therapeutic standpoint,
understanding the mechanisms of physical trans-
location is likely to be important for preventing
metastasis in patients who are diagnosed with
early cancer lesions, whereas understanding the
mechanisms leading to successful colonization
may lead to effective therapies for patients with
already-established metastases.
In the discussion that follows, we provide
our perspective on a selection of issues relevant
to contemporary cancer metastasis research.
Physical Translocation from the Primary Tumor
to the Site of Dissemination
In order for individual or small groups of cancer
cells to break away from the primary tumor and
initiate the metastatic process, these cells must
acquire the ability to migrate and invade. These
traits enable cells to degrade and move through
the extracellular matrix of the surrounding tissue
toward blood and lymphatic vessels, which in
turn provide highways for their passage to dis-
tant secondary sites. The first clinical indication
of metastatic dissemination may come from the
presence of cancer cells in the draining lymph
nodesthose connected directly with the site of
primary tumor formation through lymphatic ves-
sels; more often than not, these draining lymph
nodes seem to represent dead ends rather than
temporary stopping points from which more dis-
tant metastases are launched (2). In fact, spread
to the anatomically distant sites seems to occur
almost entirely through the blood via the process
of hematogenous dissemination (3, 4).
Carcinomas, the tumors on which we focus
in this review, arise in epithelial tissues. Nor-
mally, the cells forming the epithelial sheets in
these tissues are tightly bound to neighboring
cells and to underlying basement membranes by
adherens junctions, tight junctions, desmosomes
and hemi-desmosomes, effectively immobilizing
them in these sheets. These tight physical con-
straints encumber not only normal epithelial
cells, but also those within many benign carcino-
mas. However, as a tumor progresses, carcinoma
cells liberate themselves from these associations
and begin to strike out on their own, first by dis-
solving underlying basement membranes and
then invading adjacent stromal compartments.
This invasiveness seems to empower carcino-
ma cells to both intravasate and subsequently ex-
travasate (5).
A central question, as yet unaddressed, is
whether this acquisition of malignant traits occurs
as an almost-inevitable consequence of primary
tumor progression or as an accidental product
thereof. A widely accepted but still unproven
model of primary tumor formation posits that
cancer cells acquire a sequence of genetic and
epigenetic alterations, each of which confers one
or another form of increased fitness (6). Each of
these alterations can trigger a clonal expansion of
the cells that have acquired it, leading to a suc-
cession of clonal expansions that resembles, at
least formally, a scheme of Darwinian evolution.
This working model raises the question of
whether the development by cancer cells of ag-
gressive traits, such as invasiveness and meta-
static dissemination, reflects the fact that these
traits are locally advantageous for carcinoma
cells within the confines of a primary tumor. The
alternative is more subtle: that multistep clonal
evolution selects for cells that have greatly en-
hanced powers of survival and proliferation with-
in primary tumors. Once formed, such cellsas
an almost-accidental, unselected consequence
of their phenotypic statecan now respond to
contextual signals that induce them to express
highly malignant traits. We do not resolve be-
tween these alternatives here, although obser-
vations described below may help to illuminate
this question.
Cancer stem cells and metastasis. One crit-
ical input into this discussion comes from recent
observations that the neoplastic cells within indi-
vidual tumors are not homogeneous. An im-
portant source of intratumoral heterogeneity has
been revealed by the discovery that populations
of cells within a tumor, like those in the corre-
sponding normal tissues, are organized hierarchi-
cally (7, 8). Thus, the scheme of self-renewing
stem cells (SCs), partially differentiated transit-
amplifying (i.e., progenitor) cells, and fully differ-
entiated end-stage cells seems to be recapitulated
in many carcinomas and other tumor types (9).
The discovery of these cancer stem cells (CSCs)
has forced a major rethinking of tumor biology,
because a variety of cancer-associated traits that
were at one time ascribed to tumor cell popu-
lations as a whole must now be associated with
one or another of these subpopulations: the non-
CSCs and CSCs within each tumor. (For clarifica-
tion, we do not equate the biological properties
of CSCs with normal tissue SCs; instead, we
use the term to define a subpopulation of cancer
cells with greatly enhanced tumor-initiating po-
tential relative to other cancer cells within a tumor.
CSCs should also display self-renewal potential
and the ability to spawn non-CSC progeny.)
These considerations become especially crit-
ical in the present discussion, because many of the
biological traits of high-grade malignancy have
now been traced specifically to the subpopulations
1
Whitehead Institute for Biomedical Research, 9 Cambridge
Center, Cambridge, MA 02142, USA.
2
Department of Biology,
Massachusetts Institute of Technology (MIT), Cambridge, MA
02139, USA.
3
Ludwig MIT Center for Molecular Oncology,
Cambridge, MA 02139, USA.
*To whom correspondence should be addressed. E-mail:
weinberg@wi.mit.edu (R.A.W.); chaffer@wi.mit.edu (C.L.C.)
www.sciencemag.org SCIENCE VOL 331 25 MARCH 2011 1559
SPECIALSECTION
2012 Cancer Crusade at 40 Collections Booklet
www.sciencemag.org SCIENCE 21
of CSCs within carcinomas (1012). Thus, traits
such as motility, invasiveness, and self-renewal,
which are central to malignancy, may in fact be
the ref lection of the actions of the elusive CSC
subpopulations within larger populations of neo-
plastic cells. In many tumors, such cells may
represent a tiny fraction of the total cellularity of
individual tumors, yet these CSCs may be the
critical drivers of their malignant progression.
At a more practical level, CSCs have been
defined by their most central trait: the ability to
seed new tumors when experimentally implanted
into appropriate animal hosts. Though the repre-
sentation of CSCs, assessed in this way, may fluc-
tuate wildly from one tumor to another (13, 14),
it is clear that carcinoma cells can reside in two,
or even three, alternative states of differentiation
within a tumor. Such subpopulations, when sep-
arated by fluorescence-activated cell sorting ac-
cording to distinct cell-surface antigen profiles,
show drastic differences in tumor-initiating abil-
ity. Because these cell fractionations stratify tumor
cell populations that have not previously been
subjected to experimental manipulation, it is clear
that residence in these alternative states pre-
existed in primary tumors before their excision
from hosts.
At one level, the critical role of CSCs in
metastasis is obvious: Tumor initiation by exper-
imentally implanted cells is theoretically anal-
ogous to tumor initiation by disseminated cancer
cells; both processes depend on the ability of cancer
cells to function as founder cells that spawn essen-
tially unlimited numbers of descendants. Hence,
the very traits that are used to define CSCsself-
renewal and tumor-initiating abilitywould seem
to be inextricable elements of successful metas-
tasis formation.
Less intuitive is the observation that CSCs
exhibit yet other traits that are relevant to metas-
tasis: notably motility, invasiveness, and heightened
resistance to apoptosis (1012). This implies that
there is a multifacetedcell-biological programpack-
aged together to empower cancer cells within pri-
mary tumors to execute multiple steps of the
invasion-metastasis cascade.
The epithelial-to-mesenchymal transition and
metastasis. Over the past three decades, develop-
mental biologists have defined a cell-biological
programthe epithelial-to-mesenchymal tran-
sition (EMT)that plays critical roles in early
embryonic morphogenesis (15). This transdiffer-
entiation program, driven by EMT-inducing tran-
scription factors (EMT-TFs), is deployed during
a number of critical steps of morphogenesis, en-
abling cells of epithelial phenotype to generate
mesenchymal derivatives. Importantly, in many
embryonic contexts, the EMT is reversible; thus,
cells that were recently induced to assume a
mesenchymal phenotype may revert back to an
epithelial state via mesenchymal-to-epithelial
transitions.
Recent studies have demonstrated that the
EMT can induce non-CSCs to enter into a CSC-
like state (16, 17). As such, the EMT confers
on epithelial cells precisely the set of traits that
would empower them to disseminate from pri-
mary tumors and seed metastases (15); hence,
it is an attractive solution to understanding the
mechanics of dissemination. Moreover, the height-
ened resistance to apoptosis that is integral to
cells generated by an EMT is surely critical to
the ability of carcinoma cells to survive the rig-
ors of the voyage from primary tumors to sites
of dissemination (18). In addition, the CSC-like
state approached by carcinoma cells that have
passed through an EMT may be critical in their
sites of dissemination for launching new colo-
nies of cancer cells. Merely because the EMT is
an attractive solution does not make it a unique
one, however, and it remains possible that oth-
er, still-undiscovered cell-biological programs op-
erate in certain carcinoma cells as drivers of
malignancy.
Activation of an EMT program during tumor-
igenesis often requires signaling between cancer
cells and neighboring stromal cells (19). Islands
of cancer cells in advanced primary carcinomas
are thought to recruit a variety of cell types into
the surrounding stroma, such as fibroblasts, myo-
fibroblasts, granulocytes, macrophages, mesenchy-
mal stem cells, and lymphocytes; these recruited
cells create a reactive stromain effect, an in-
flammatory microenvironment that appears to
result in the release of EMT-inducing signals.
The carcinoma cells respond to these contextual
signals by activating expression of certain tran-
scription factors (EMT-TFs) that proceed to
orchestrate EMT programs within these cells.
This scenario implies that acti-
vation of an EMT program flows
from two major sources. First, the
biology of the cancer cell-of-origin
before its transformation conspires
with the genetic and epigenetic
changes sustained during primary
tumor formation to generate carcino-
ma cells that are responsive to EMT-
inducing signals. Second, inductive
signals released by the reactive stro-
ma impinge on these responsive
carcinoma cells, causing them to ex-
press various EMT-TFs and thereby
activate previously latent EMT pro-
grams. When portrayed in this way,
the EMT-associated traits are not
direct products of genetic and epi-
genetic evolution in primary tumors;
instead, they represent adaptations to
contextual signals experienced once
primary tumors have formed.
Though it is appealing in con-
cept, the role of the EMT in enabling
metastatic dissemination remains
largely unproven, in part because of
the technical difficulties of capturing
this transitory process in human can-
cer patients. Moreover, the EMT may
only operate in a tiny fraction of
Physical translocation
from primary tumor to distant organ
Colonization
A
Acquisition
of invasive
phenotype
C
CTCs transit to
distant organ
E
Survival at
secondary site
B
Local invasion
cells invade into surrounding
stroma, then intravasate to enter
hematogenous circulation
D
CTCs extravasate
and invade into the
parenchyma of
foreign tissue
F
Adaptation and
proliferation to
form metastases
Differentiated
cancer cell
Transitioning
cancer cell
Cancer
stem cell
Inflammatory cell Stromal cell
Fig. 1. The metastatic cascade. Metastasis can be envisioned as a process that occurs in two major phases: (i) physical
translocation of cancer cells from the primary tumor to a distant organ and (ii) colonization of the translocated cells
within that organ. (A) To begin the metastatic cascade, cancer cells within the primary tumor acquire an invasive
phenotype. (B) Cancer cells can then invade into the surrounding matrix and toward blood vessels, where they
intravasate to enter the circulation, which serves as their primary means of passage to distant organs. (C) Cancer cells
traveling through the circulation are CTCs. They display properties of anchorage-independent survival. (D) At the
distant organ, CTCs exit the circulation and invade into the microenvironment of the foreign tissue. (E) At that foreign
site, cancer cells must be able to evade the innate immune response and also survive as a single cell (or as a small
cluster of cells). (F) To develop into an active macrometastatic deposit, the cancer cell must be able to adapt to the
microenvironment and initiate proliferation.
25 MARCH 2011 VOL 331 SCIENCE www.sciencemag.org 1560
Cancer Crusade at 40
22 SCIENCE www.sciencemag.org
cancer cells that are in intimate contact with
adjacent reactive stroma. This implies that ex-
pression of an EMT program cannot possibly
be discerned by examining the altered genome
of an unfractionated tumor sample.
A clear resolution of the role of EMTs in
high-grade malignancies is complicated by one
further consideration: Although this program is
often depicted as a bi-stable switch that causes
cells to flip from one state into the other, the
biological reality is likely to be more subtle. In
many tumors, epithelial carcinoma cells appear
to advance only partway down the road toward
the mesenchymal state. This yields cancer cells
that concomitantly express epithelial and mesen-
chymal markers and thus persist in a pheno-
typic state that is not encountered
in normal tissues.
Intrinsic versus induced cancer
stem cells. The discussions above
imply that (i) the stem-cell state is
an integral part of the development
of metastases, (ii) in many types of
carcinomas, entrance into this state
is facilitated by passage through an
EMT, and (iii) EMTs can be in-
duced in carcinoma cells by contex-
tual signals received, for example,
from the tumor-associated reactive
stroma. Taken together, these no-
tions imply that CSCs can be formed
de novo through the actions of these
signals. At the same time, these in-
duced CSCs cannot represent the
only source of cells in the CSC pool
within a tumor: Intrinsic CSCs are
likely to exist within tumors from
their very inception, long before re-
active stroma and EMTs become im-
portant (Fig. 2).
We speculate that the presence
of intrinsic and induced subtypes
of CSCs within a tumor may partly
explain the heterogeneity evident
in clinical tumor pathology, where
highly aggressive tumors (such as the claudin-
low and basal-type breast cancers) exhibit a
normal mammary stem cell gene profile and thus
contain high numbers of intrinsic CSCs, whereas
luminal-type breast cancers correlate with a ma-
ture mammary luminal cell phenotype and thus
contain low numbers of intrinsic CSCs (20, 21).
As such, in some subtypes of cancer, intrinsic
preneoplastic SCs are likely to be present al-
ready during the early stages of tumorigenesis.
Such SCs, which already should possess some
of the EMT-associated phenotypes, may play
prominent roles in disseminating carcinoma cells
long before frankly malignant tumors have
developed.
Such thinking confines the induced CSCs to
later stages of tumorigenesis when the extensive
reactive stroma is first apparent. Accordingly,
induced CSCs may be a feature of tumors able
to recruit a reactive stroma capable of inducing
an EMT. This raises the question of whether the
two types of CSCs play identical roles in tumor
progression and differ only in their origins.
Plasticity between epithelial and mesenchy-
mal states will surely be important as we try to
understand the dynamics of how metastatic col-
onies are initiated after disseminated carcinoma
cells extravasate into tissue parenchyma. Such
carcinoma cells may have previously undergone
a partial or complete EMT within primary tu-
mors, having been induced to do so via hetero-
typic signals originating in the tumor-associated
stroma; components of the EMT program may
then have enabled their physical dissemination.
However, it is plausible that after extravasat-
ing, these cells will not encounter an activated
stroma and, in the absence of associated stro-
mal signaling, may well lapse back to a fully
epithelial state that lacks CSC function. Still,
such cells cannot afford to jettison CSC function
completely if they are to serve as founders of a
metastatic tumor; this suggests that complex
mechanisms operate to maintain the mesenchymal/
CSC state, even in those CSCs whose recent his-
tory suggests that they are induced rather than
intrinsic CSCs.
Circulating tumor cells. The blood of many
patients with advanced primary carcinomas con-
tains circulating cancer cells, at least a subset of
which may be in transit from the primary tumor
to sites of future metastasis. These circulating tu-
mor cells (CTCs) offer the prospect of under-
standing how cells are able to survive in the
circulation. Importantly, the presence of CTCs
and of extravasated, disseminated tumor cells
(DTCs), such as those in the bone marrow of
patients with breast cancer, correlates with in-
creased metastatic burden, aggressive disease,
and a decreased time to relapse (22).
In the longer term, a more thorough under-
standing of CTC and DTC biology may gen-
erate highly useful diagnostic and prognostic
measurements (23). Whether these cells will re-
veal important etiologic mechanisms remains
unclear; for example, the detection of large num-
bers of CTCs may simply reflect a high level of
aggressiveness of a primary tumor, rather than
revealing the particular cells that serve as the
key intermediaries between primary tumors and
metastases.
Our understanding of CTCs and the roles
that they play in metastatic dissemination is still
clouded by some major unresolved biological
issues and by technical issues arising from de-
tection sensitivities of these rare cells. Circulating
carcinoma cells have diameters (20 to 30 mm)
that are far too large to allow them to pass
through the bores of capillaries (~8-mm diameter),
such as those present in the capillary beds of the
lungs (4). By all rights, within minutes of being
released by primary tumors into the venous cir-
culation, CTCs should be trapped in these cap-
illaries during their first pass through the heart.
Yet some persist for far longer periods of time
[with half-lives of 1 to 2.4 hours (24)], which
suggests the possibility that only exceptionally
Intrinsic Induced
CSC with
metastatic potential
Transition to
CSC-like state
EMT
Recruitment of
reactive stroma
Cells poised to
undergo EMT
Differentiated
cancer cell
Transitioning
cancer cell
Responsive
differentiated
cancer cell
Cancer stem cell Inflammatory cell Stromal cell
Fig. 2. Acquisition of the metastatic phenotype. Tumors are heterogeneous populations of cells. CSC sub-
populations are particularly well poised to complete the metastatic cascade. Two alternative means of generating
CSCs are depicted here. Intrinsic CSCs are thought to exist in primary tumors from the very early stages of
tumorigenesis and may be the oncogenic derivatives of normal-tissue stem or progenitor cells. Induced CSCs may
arise as a consequence of the EMT. In this case, carcinoma cells initially recruit a variety of stromal cells, such as
fibroblasts, myofibroblasts, granulocytes, macrophages, mesenchymal stem cells, and lymphocytes. Together these
cells create a reactive microenvironment that releases factors (e.g., Wnt, transforming growth factorb, fibroblast
growth factor) that cause the neighboring cancer cells to undergo the EMT and acquire CSC-like characteristics.
www.sciencemag.org SCIENCE VOL 331 25 MARCH 2011 1561
SPECIALSECTION
2012 Cancer Crusade at 40 Collections Booklet
www.sciencemag.org SCIENCE 23
small or physically plastic CTCs can elude the
sieving action of the pulmonary microvasculature
and thereby accumulate to substantial steady-
state concentrations in the blood. Moreover, if
CTCs occasionally travel in multicell clumps of
far larger diameter, such clumps must lodge al-
most immediately in microvessels and thus have
extremely short lives in the circulation; such
clumps would be strongly underrepresented by
current methods that tally CTC numbers in
patients.
In addition, the tissue factor protein dis-
played on the surface of individual cancer cells
attracts clouds of aggregating platelets (25, 26),
which also increase the effective diameters of
the cancer cells. This suggests that the CTCs
found in the circulation may represent special
subpopulations of cancer cells that, for one rea-
son or another, do not trigger platelet aggrega-
tion. These platelet cloaks may also complicate
the detection of CTCs by occluding the cell-
surface marker antigens that are used to identify
and separate CTCs from the million-fold greater
numbers of nonepithelial cells in the circulation.
Epithelial-specific cell-surface markers are
in widespread use to detect CTCs (2729). This
represents a potential problem, because it is like-
ly that carcinoma cells that have passed through
a partial or complete EMT are no longer detect-
able by epithelial-specific antigens. Enrichment
and detection of CTCs via depletion of hema-
topoietic cells (using antibodies specific for CD45)
may represent one way of circumventing this
issue in the future (23).
Despite these complications, the ability to
isolate CTCs from the circulation may become a
powerful tool to study the biology of migrating
CSCs, especially if those cells differ from CSCs
that reside in the primary tumor site. Moreover,
they offer the prospect of creating a highly useful
(and relatively noninvasive) diagnostic param-
eter: By monitoring longitudinally the concen-
trations of CTCs in a cancer patient, oncologists
may be able to determine, from day to day, how
effectively an applied therapy has been in reduc-
ing the burdens of primary tumors that are pre-
sumably the sources of these CTCs.
Homing. Once lodged in the capillary bed
of a foreign tissue, CTCs may soon extravasate
and invade the foreign parenchyma, or they may
proliferate intraluminally and eventually rupture
the wall of the microvessels in which they are
lodged (3032).
The formation of metastases in certain fa-
vored target organs may be influenced by struc-
tural differences in the capillaries of various
tissues. For example, the sinusoid capillaries in
the bone marrow are formed from single layers
of endothelial cells and are devoid of supporting
mural cells; this design is thought to facilitate
the normal trafficking of hematopoietic cells in
and out of the bone marrow (33). This route
may also present a path of least resistance to
carcinoma cells, and thus may help to explain
why the bone marrow is a favored site of metas-
tasis by cancer cells that originate in diverse
primary tumors (e.g., breast, prostate, lung, and
gastric cancers) (34).
In certain tumor types, the layout of the
circulation may be the strongest determinant
of metastatic tropism. Most frequently cited is
the behavior of colorectal carcinomas (CRCs),
which have a strong preference for generating
liver metastases. In fact, the disseminating CRC
cells may be intrinsically poorly adapted for sur-
vival in the liver microenvironment. However,
because of the portal circulation, which drains
from the mesentery directly into the liver, myriad
carcinoma cells may be dumped over extended
periods of time into the liver microvasculature;
on rare occasion, an otherwise low-probability
event may then generate a liver metastasis. In
these various cases, homing to a particular organ
can be considered to be a passive process that
is determined by circulation patterns and the
physical properties of the vasculature rather than
by particular biological properties of the dis-
seminating cancer cell.
Organ-specific homing may also constitute
an active process, where tissue and cancer cell
specific features determine metastatic dissemi-
nation. The expression by metastasizing cancer
cells of specific proteins (for example, integrins)
seems to play a key role in this process (3537).
We envisage that homing of metastasizing cancer
cells may be a combination of both mechanical
trapping of cancer cells in the microvasculature
of distant organs and cancer cellmediated ad-
hesion to specific luminally displayed compo-
nents of the vasculature (Fig. 3).
Colonization-Adaptation of the Disseminated Cell
to the Microenvironment at the Metastatic Site
Although EMT programs may prove critical to
the physical dissemination of carcinoma cells,
the multiple powers of these programs would
not seem capable of addressing the problem of
colonization. Instead, colonization seems to rep-
resent a far more complex set of phenomena
and a relatively small number of unifying, gen-
eralizable principles. This complexity can be
judged, perhaps simplistically, by listing com-
mon tumors and their known metastatic tropisms.
For example, as noted above, prostate carcino-
mas preferentially metastasize to bone, whereas
CRCs preferentially metastasize to the liver (38).
Nonetheless, these and other primary tumors
can also form metastases at additional, alterna-
tive tissue sites. In each case, the tissue micro-
environment of a primary tumor is likely to differ
markedly from that of the secondary site of dis-
semination, necessitating substantial adaptive
moves by recently arrived cancer cells. The de-
tails of these adaptive programs would seem
to be dictated by the microenvironment of the
starting point (the primary tumor) and the micro-
environment of the landing site (the tissue pa-
renchyma in which a metastasis is founded).
This logic suggests that the number of
distinct adaptive programs can be gauged by a
simple calculation: the product of multiplying
the number of metastasizing primary tumor types
by the number of distinct sites of dissemination.
However, clearly some adaptive moves, such as
the activation of certain ensembles of genes, may
simultaneously confer an ability to colonize mul-
tiple distinct tissues, reducing the diversity of
adaptive programs. Moreover, we know that cer-
tain lung cancers metastasize quickly to multiple
sites, whereas others, such as breast and prostate
carcinomas, often take years to develop meta-
static colonies and then only in a relatively lim-
ited number of sites (38). This suggests that the
differentiation programs of certain normal cells,
such as those in the lungs, position their neo-
plastic descendants to adapt readily to foreign-
tissue microenvironments, whereas other cancer
cell types must laboriously cobble together far
more complex shifts in gene-expression pro-
grams. Moreover, it is unclear how often colo-
nization depends on epigenetic changes versus
genetic mutations in cell genomes.
Nonetheless, solid progress is being made
toward understanding the biochemical adapta-
tions that carcinoma cells must make to thrive in
distant tissues. Gene sets have been defined in
breast tumor xenografts that can predict homing
and colonization of breast cancer cells specifically
to the lung, bone, or brain (3941); conversely,
genomic profiling of metastases has been suc-
cessfully used to predict the sites of primary tu-
mor origin (42). These gene-expression patterns
suggest that carcinoma cells within primary
breast tumors acquire patterns that enable their
subsequent colonization preferentially to specific
target organs. These findings are supported by
recent studies demonstrating that genetically dis-
tinct subpopulations of cells present in primary
tumors are responsible for forming metastases
(43, 44). Precisely how these gene-expression
programs are acquired by the cells within pri-
mary tumors is not yet clear. They may be dic-
tated by the differentiation programs of normal
cells of origin, by somatic genetic and epigenetic
changes that have been selected during multistep
tumor progression, or simply stochastically. In-
terestingly, the discovery of reseeding of primary
tumors by their derived metastases (45) raises a
fourth possibility: Metastatic cells may contami-
nate the gene-expression patterns of the corre-
sponding parental primary tumor by introducing
gene-expression programs that were selected
during earlier metastatic colonization.
In general, colonization is an extremely in-
efficient process, and most cancer cells that suc-
cessfully translocate from the primary tumor to
a secondary site undergo apoptosis within 24 hours
of extravasation (4648). Experimental and clinical
data support the notion that the survivors per-
25 MARCH 2011 VOL 331 SCIENCE www.sciencemag.org 1562
Cancer Crusade at 40
24 SCIENCE www.sciencemag.org
sisting in metastatic sites can exist in at least
three alternative states: (i) as solitary viable can-
cer cells in a quiescent, nonproliferative state
(dormancy); (ii) as micrometastases, which re-
main as small lesions (probably due to a balance
between proliferation and apoptosis); or (iii) as
actively growing macrometastatic lesions (Fig.
3). Our current understanding of the mechanisms
governing the entrance into these three states
and the transitions between them is limited be-
cause of the experimental challenges of studying
dormancy, the extended times required for these
processes to reach completion, the paucity of
appropriate models, and the technical challenges
surrounding the analysis of single cells in vastly
larger living tissues (49).
Successful colonization is presumed to in-
clude the ability to acquire mitogenic stimulation
from growth factors and cytokines that are nat-
urally present in the alien microenvironment
to self-renew and generate a large flock of
descendantsto recruit the necessary supporting
stroma, including an appropriate blood supply
(4, 50). Micrometastases composed of actively
proliferating cells represent attractive venues for
cancer cells to develop complex colonization pro-
grams. Thus, we imagine that active cell divi-
sion is essential for the generation of genetic and
epigenetic alterations; once the resulting variants
arise in these micrometastases, their novel
phenotypes can be tested for an ability to confer
selective advantage in the presence of highly
demanding, otherwise-inhospitable microenvi-
ronments. It is also possible that the tissue micro-
environment at the secondary site may change
over time as a result of aging, disease, or wound-
ing and that these environmental changes function
as triggers that induce the release of disseminated
cells from the dormant state.
Several models of mouse metastasis now
suggest that factors derived from primary tu-
mors can educate distant sites in preparation for,
and before the arrival of, metastasizing cancer
cells (5153). In these studies, factors secreted
by the primary tumors (e.g., VEGF-A, PlGF,
PSAP) are thought to mobilize bone marrow
derived cells that are subsequently attracted to
premetastatic sites. The cells of this premeta-
static niche then release factors (e.g., SDF-1,
S1000A8, S100A9) that can attract disseminat-
ing tumor cells (Fig. 3). This interesting concept
is still in the early stages of investigation.
Consequences of successful colonization. The
outcome of successful colonization is a rapidly
expanding macrometastasis that can now serve
as the focus for disseminating a shower of sec-
ondary metastases. Importantly, many of the
cancer cells that are dispatched from this re-
cently successful metastasis may be invested
with a functional colonization program that may
empower them to colonize either a limited sub-
set of sites throughout the body or, alternative-
ly, multiple distinct tissue types. The throngs of
secondary metastases derived from this shower
will soon eclipse the single initiating metastasis
that spawned them.
Certain microenvironments are almost guar-
anteed to provide hospitable sites for such dis-
seminating, colonization-competent cancer cells:
sites of wound-healing and stroma of already-
established tumors. Strangely enough, one in-
sight has come from dentists who have observed
tumors growing out of extracted tooth sockets
within weeks of oral surgery (54); the tumors
arising in these sites of active wound healing
represented the first clinical manifestations of
previously undetected, widespread metastatic
disease in these patients. Indeed, the similarity
between wound-healing environments and the
hospitable stroma of tumors was encapsulated in
an observation made years ago that tumors are
like wounds that will not heal (55).
This theme has been extended by the recent
observations cited above (45) that the cells from
contralaterally implanted aggressive tumors in
mice can metastasize to one another and at ap-
parently high efficiency. The suggestion here is
that the reactive stroma that arose in each of
these tumors (and contributed to their locally ag-
gressive phenotypes) also generated a hospitable
site for settling and colonization by disseminated
CTCs, including those deriving from the con-
tralaterally implanted tumors and those generated
by the tumor itself. These diverse observations
underscore the notion that, in contrast to normal
tissue stroma, the stromata of aggressive tumors
and sites of wound healing can serve as readily
colonizable microenvironments that make few
adaptive demands on the cells that have dis-
seminated to them.
Conclusions
One of the most revealing findings of the past
decade in cancer research is that cells within any
given carcinoma display a great deal of heter-
ogeneity. This intratumoral heterogeneity is due,
in no small part, to subpopulations of cells that
are phenotypically distinct but genetically iden-
tical; for instance, cancer cells that are less and
more differentiated can share a common set of
genetic alterations. These distinct phenotypic
states, involving CSCs and non-CSCs, could hold
important implications for our understanding of
the biology of tumor progression and clinical
therapy. For example, the biological attributes of
the tumor as a whole may be strongly influenced
by its subpopulation of CSCs. These cells may
Homing Colonization
Differentiated
cancer cell Cancer stem cell Inflammatory cell Extracellular matrix Stromal cell
E
Micrometastasis
A
Mechanical
trapping
C
Pre-metastatic
niche
B
Site-specific
adhesion or
chemoattraction
D
Quiescence
Macrometastasis
F
Fig. 3. Adaptation of metastatic cells to a foreign environment. Homing and colonization of a cancer
cell to a distant organ are complex processes with many questions still unanswered. CTCs transiting
from the primary tumor to a metastatic site can arrive at their destination via a variety of mech-
anisms: (A) CTCs may become lodged in the capillary beds of specific organs due to size. (B) CTCs
may display specific adhesion molecules that enable them to adhere to microvessels in specific
organs, or they may respond to a chemoattractive gradient arising from a particular tissue. (C) CTCs
may preferentially home to organs where a premetastatic niche has prepared a microenvironment
conducive to their survival. (D) Once cancer cells have exited the blood stream (extravasated) they
may first experience a period of quiescence (dormancy) while they adapt to their newfound
microenvironment. (E) Dormant cells may progress to micrometastatic deposits (perhaps in response
to the recruitment of an appropriate stroma or an enhanced ability to respond to proliferative signals
present in the host microenvironment) where their size is kept in check because of a balance in
proliferation, apoptosis, and phagocytosis by the host-tissue immune system. (F) To develop into a
macrometastasis, cancer cells must recruit an adequate blood supply (necessary for growth beyond
1 to 2 mm). The signals or mechanisms responsible for the transition from dormancy to micro-
metastasis to macrometastasis remain largely unknown.
www.sciencemag.org SCIENCE VOL 331 25 MARCH 2011 1563
SPECIALSECTION
2012 Cancer Crusade at 40 Collections Booklet
www.sciencemag.org SCIENCE 25
drive the self-renewal of the tumor cell popula-
tion and, in the context of the present discus-
sion, be responsible for the tumors invasiveness
and metastatic dissemination.
Moreover, the observations that (i) EMT pro-
grams can drive carcinoma cells into states that
approximate the CSC state and (ii) EMTs can be
activated by contextual signals experienced by
carcinoma cells lead to the notion that great
plasticity is likely to exist between the non-CSCs
and CSCs within a tumor. Likewise, these var-
ious observations could hold important implica-
tions for strategies aimed at reducing metastasis
and eradicating minimal residual disease, includ-
ing dormant tumor cells, micrometastases, and
macrometastatic deposits.
Conventional therapeutics, which efficiently
target actively proliferating cells within the pri-
mary tumor, have little impact on quiescent or
slowly proliferating cancer cells and, thus, on
those cells that reside in many micrometastatic
colonies (56, 57). The difficulty of treating this
residual disease may be compounded by the
physiology of certain target organs. For example,
the blood-brain barrier may shield metastases
within the brain from drugs delivered through
the circulatory system (58).
CSCs can erect additional barriers to suc-
cessful treatment. Rapidly accumulating evidence
indicates that CSCs exhibit a heightened resist-
ance to drug-induced death (59). This resistance
may stem from two sources: (i) SCs often retreat
reversibly from the active growth-and-division
cycle into states of quiescence, and (ii) even more
important may be the intrinsic drug resistance
exhibited by the mesenchymal cancer cells that
are the products of the EMT and exhibit traits
associated with CSCs (60). Thus, various obser-
vations of drug-resistant carcinoma cell subpop-
ulations confirm that the surviving cells often
exhibit a more mesenchymal phenotype (61).
Without eradicating carcinoma cells that have
entered into the mesenchymal/CSC state, the on-
cologist is confronted with neoplastic subpopu-
lations that are capable of regrowing primary
tumors and, in addition, dispatching metastatic
travelers to distant organ sites.
The most elusive aspect of the invasion-
metastasis cascade involves the fates of carcinoma
cells after they have disseminated and extrava-
sated into the parenchyma of distant organs. We
presume that the self-renewal ability of recently
disseminated cells is an essential prerequisite to
their successful colonization of such distant tis-
sues, but at present the evidence for this is only
indirect. Interestingly, in the absence of a re-
active stroma and EMT-inducing signals in such
distant tissues, recently disseminated cells that
may have arrived in a quasi-mesenchymal/CSC
state may lapse back to a fully epithelial state and
thereby forfeit the stem-ness that would seem
to be essential for their successful founding of
metastases.
At present, we do not know when and how
carcinoma cells acquire the abilities to colonize
distant organ sites. Clearly, certain cell-biological
programs that have been established and oper-
ated in primary tumors may prove advantageous
when disseminated carcinoma cells initially con-
front the microenvironment of distant tissues.
Accordingly, we do not know whether the adap-
tations required for colonization are largely
carried by these traveling cells to their sites of
metastasis or whether they are instead cobbled
together later in sites of dissemination as carcino-
ma cells struggle to survive and proliferate in for-
eign, potentially hostile tissue microenvironments.
These questions have proven difficult to ad-
dress because the process of colonization in the
great majority of tumors is extraordinarily in-
efficient. It is unclear whether fully dormant
cancer cells are invariably doomed to eventual
elimination or whether they may reawaken after
months and years and suddenly spawn exuber-
ant tumors. We suspect that only those micro-
metastases containing proliferating cells are
capable of exploring multiple alternative pheno-
typic states until they stumble on one that en-
ables them to flourish; at present, this notion is
deduced from first principles rather than being
demonstrated experimentally.
The multiplicity of adaptive programs is an
issue of great interest: they may be shared by
many tumors and at many sites of dissemination.
Alternatively, as argued here, these programs rep-
resent ad hoc solutions that are dictated by the
origins of disseminated cancer cells and the nature
of their newfound homes; if so, we may confront
myriad distinct adaptive programs that are diffi-
cult to rationalize in terms of a common set of
underlying biochemical mechanisms. The thera-
peutic implications of these different scenarios
remain to be determined. Still, on a positive note,
it is clear that the pace of discovery is rapid and
that paths of future exploration are in sight. We
now understand much more about the metastatic
process than we did even a few short years ago.
References and Notes
1. I. J. Fidler, Nat. Rev. Cancer 3, 453 (2003).
2. J. A. Joyce, J. W. Pollard, Nat. Rev. Cancer 9, 239
(2009).
3. D. Hanahan, R. A. Weinberg, Cell 100, 57 (2000).
4. A. F. Chambers, A. C. Groom, I. C. MacDonald, Nat. Rev.
Cancer 2, 563 (2002).
5. J. P. Thiery, J. P. Sleeman, Nat. Rev. Mol. Cell Biol. 7, 131
(2006).
6. L. Foulds, Cancer Res. 14, 327 (1954).
7. D. Bonnet, J. E. Dick, Nat. Med. 3, 730 (1997).
8. M. Al-Hajj, M. S. Wicha, A. Benito-Hernandez,
S. J. Morrison, M. F. Clarke, Proc. Natl. Acad. Sci. U.S.A.
100, 3983 (2003).
9. L. E. Ailles, I. L. Weissman, Curr. Opin. Biotechnol. 18,
460 (2007).
10. E. Charafe-Jauffret et al., Cancer Res. 69, 1302
(2009).
11. R. Pang et al., Cell Stem Cell 6, 603 (2010).
12. P. Marcato et al., Stem Cells 29, 32 (2011).
13. E. Quintana et al., Nature 456, 593 (2008).
14. K. Ishizawa et al., Cell Stem Cell 7, 279 (2010).
15. J. P. Thiery, H. Acloque, R. Y. Huang, M. A. Nieto, Cell
139, 871 (2009).
16. S. A. Mani et al., Cell 133, 704 (2008).
17. A. P. Morel et al., PLoS ONE 3, e2888 (2008).
18. A. Gal et al., Oncogene 27, 1218 (2008).
19. J. Yang, R. A. Weinberg, Dev. Cell 14, 818
(2008).
20. E. Lim et al.; kConFab, Nat. Med. 15, 907
(2009).
21. A. Prat et al., Breast Cancer Res. 12, R68 (2010).
22. S. Braun et al., N. Engl. J. Med. 353, 793 (2005).
23. K. Pantel, C. Alix-Panabires, S. Riethdorf, Nat. Rev.
Clin. Oncol. 6, 339 (2009).
24. S. Meng et al., Clin. Cancer Res. 10, 8152 (2004).
25. E. Camerer et al., Blood 104, 397 (2004).
26. B. Nieswandt, M. Hafner, B. Echtenacher, D. N. Mnnel,
Cancer Res. 59, 1295 (1999).
27. M. Lacroix, Endocr. Relat. Cancer 13, 1033 (2006).
28. K. Pantel, R. H. Brakenhoff, B. Brandt, Nat. Rev. Cancer
8, 329 (2008).
29. M. Mego, S. A. Mani, M. Cristofanilli, Nat. Rev. Clin.
Oncol. 7, 693 (2010).
30. S. Ito et al., Int. J. Cancer 93, 212 (2001).
31. C. W. Wong et al., Am. J. Pathol. 161, 749 (2002).
32. E. Sahai, Nat. Rev. Cancer 7, 737 (2007).
33. H. G. Kopp, S. T. Avecilla, A. T. Hooper, S. Rafii,
Physiology (Bethesda) 20, 349 (2005).
34. C. Alix-Panabires, S. Riethdorf, K. Pantel, Clin. Cancer
Res. 14, 5013 (2008).
35. M. Abdel-Ghany, H. C. Cheng, R. C. Elble, B. U. Pauli,
J. Biol. Chem. 276, 25438 (2001).
36. H. Wang et al., J. Cell Biol. 164, 935 (2004).
37. D. M. Brown, E. Ruoslahti, Cancer Cell 5, 365
(2004).
38. K. R. Hess et al., Cancer 106, 1624 (2006).
39. Y. Kang et al., Cancer Cell 3, 537 (2003).
40. A. J. Minn et al., Nature 436, 518 (2005).
41. P. D. Bos et al., Nature 459, 1005 (2009).
42. G. Bloom et al., Am. J. Pathol. 164, 9 (2004).
43. P. J. Campbell et al., Nature 467, 1109 (2010).
44. S. Yachida et al., Nature 467, 1114 (2010).
45. M. Y. Kim et al., Cell 139, 1315 (2009).
46. I. J. Fidler, J. Natl. Cancer Inst. 45, 773 (1970).
47. J. W. Kim et al., Cancer Lett. 213, 203 (2004).
48. K. J. Luzzi et al., Am. J. Pathol. 153, 865 (1998).
49. P. E. Goss, A. F. Chambers, Nat. Rev. Cancer 10, 871
(2010).
50. J. A. Aguirre-Ghiso, Nat. Rev. Cancer 7, 834
(2007).
51. R. N. Kaplan et al., Nature 438, 820 (2005).
52. S. Hiratsuka, A. Watanabe, H. Aburatani, Y. Maru,
Nat. Cell Biol. 8, 1369 (2006).
53. S. Y. Kang et al., Proc. Natl. Acad. Sci. U.S.A. 106,
12115 (2009).
54. A. Hirshberg, P. Leibovich, I. Horowitz, A. Buchner, J. Oral
Maxillofac. Surg. 51, 1334 (1993).
55. H. F. Dvorak, N. Engl. J. Med. 315, 1650 (1986).
56. G. N. Naumov et al., Breast Cancer Res. Treat. 82, 199
(2003).
57. S. Braun et al., J. Clin. Oncol. 18, 80 (2000).
58. R. J. Weil, D. C. Palmieri, J. L. Bronder, A. M. Stark,
P. S. Steeg, Am. J. Pathol. 167, 913 (2005).
59. X. Li et al., J. Natl. Cancer Inst. 100, 672 (2008).
60. E. Buck et al., Mol. Cancer Ther. 6, 532 (2007).
61. C. J. Creighton et al., Proc. Natl. Acad. Sci. U.S.A. 106,
13820 (2009).
62. C.L.C. is supported by the National Health and
Medical Research Council of Australia and the Advanced
Medical Research Foundation. R.A.W. is supported by
the National Cancer Institute, MIT Ludwig Center for
Molecular Oncology, and Breast Cancer Research Fund.
R.A.W. is a founder and shareholder of Verastem, Inc.,
a biopharmaceutical company focused on discovering
and developing drugs that target cancer stem cells.
10.1126/science.1203543
25 MARCH 2011 VOL 331 SCIENCE www.sciencemag.org 1564
Cancer Crusade at 40
26 SCIENCE www.sciencemag.org
REVI EW
Cancer Immunoediting: Integrating
Immunitys Roles in Cancer
Suppression and Promotion
Robert D. Schreiber,
1
* Lloyd J. Old,
2
Mark J. Smyth
3,4
Understanding how the immune system affects cancer development and progression has been
one of the most challenging questions in immunology. Research over the past two decades has
helped explain why the answer to this question has evaded us for so long. We now appreciate
that the immune system plays a dual role in cancer: It can not only suppress tumor growth by
destroying cancer cells or inhibiting their outgrowth but also promote tumor progression
either by selecting for tumor cells that are more fit to survive in an immunocompetent host or
by establishing conditions within the tumor microenvironment that facilitate tumor outgrowth.
Here, we discuss a unifying conceptual framework called cancer immunoediting, which
integrates the immune systems dual host-protective and tumor-promoting roles.
T
he idea that the immune system can con-
trol cancer has been the subject of debate
for over a century. In the early 1900s, Paul
Ehrlich was perhaps the first to reason that cancer
would be quite common in long-lived organisms
if not for the protective effects of immunity (1).
However, so little was known about the compo-
sition and function of the immune system at the
time that it was simply not possible to assess the
validity of this prediction. It would take nearly 50
years before the idea of immune control of can-
cer resurfaced, stimulated in large part by an en-
hanced understanding of the immune system
combined with the demonstration of the existence
of tumor antigens (2). These advances provided
the foundation upon which Burnet and Thomas
built their cancer immunosurveillance hypoth-
esis, a concept that formally envisaged that
adaptive immunity was responsible for preventing
cancer development in immunocompetent hosts
(3, 4). However, subsequent studies by Stutman
provided little support for this hypothesis. Of
particular note were experiments showing that
the cancer susceptibility of immunocompetent mice
(to both spontaneous and carcinogen-induced tu-
mors) was similar to that of nude mice that had
major but not total immunodeficiency (5, 6). On
the basis of these findings, the cancer immuno-
surveillance hypothesis was largely abandoned,
and soon additional arguments began to surface
as to why cancer immunosurveillance could not
possibly occur. Some investigators argued that tu-
mor cells did not possess the appropriate danger
signals needed to alert the immune system to
the presence of a foreign cell (7), whereas others
suggested that the immune system would ignore
or be tolerant to a developing tumor because
tumor cells were too similar to the normal cells
from which they were derived (8). Still others
showed that persistent activation of the innate,
pro-inflammatory arm of immunity could facil-
itate cellular transformation and promote cancer
outgrowth and argued that this effect of immunity
precluded its capacity to fulfill a protective func-
tion (9, 10).
By the 1990s, improved mouse models of
immunodeficiency on pure genetic backgrounds
became commonplace, permitting a few groups
to reassess the role of immunity in cancer control.
Interest in cancer immunosurveillance was re-
kindled by the discovery of the importance of
interferon-g (IFN-g) in promoting immunologi-
cally induced rejection of transplanted tumor
cells (11) and by the demonstration that mice lack-
ing either IFN-g responsiveness (gene-targeted
mice lacking either the IFN-g receptor or the
STAT1 transcription factor required for IFN re-
ceptor signaling) or adaptive immunity [RAG2
/
mice lacking T cells, B cells, and natural killer
T(NKT) cells] were more susceptible to carcinogen-
induced and spontaneous primary tumor forma-
tion (Fig. 1) (12, 13). Other laboratories soon
began to report similar results, and collectively
these findings documented that the immune sys-
temcan function as an extrinsic tumor suppressor
[(1117), reviewed in (18)].
We now recognize that the immune system
plays at least three distinct roles in preventing
cancer: (i) It protects the host against viral in-
fection and hence suppresses virus-induced tu-
mors; (ii) it prevents the establishment of an
1
Department of Pathology and Immunology, Washington
University School of Medicine, St. Louis, MO63110, USA.
2
New
York Branch of The Ludwig Institute for Cancer Research at
Memorial Sloan-Kettering Cancer Center, NewYork, NY 10021,
USA.
3
Cancer Immunology Program, Peter MacCallum Cancer
Centre, East Melbourne, 3002 Victoria, Australia.
4
Department
of Pathology, University of Melbourne, Parkville, 3010 Victoria,
Australia.
*To whom correspondence should be addressed. E-mail:
schreiber@immunology.wustl.edu
160 80
Time after carcinogen treatment (days)
P
e
r
c
e
n
t
a
g
e

o
f

m
i
c
e

w
i
t
h

t
u
m
o
r
s
0
0
20
40
60
80
100
Immunodecient
mice
Immunodecient
mice
Immunocompetent
mice
Immunocompetent
mice
Carcinogen
Carcinogen
200
days
200
days
Fig. 1. The immune status of mice is a critical determinant of their susceptibility to tumors induced by
chemical carcinogens. Over the past two decades, numerous studies have established that immu-
nodeficient mice are more tumor prone than are immunocompetent mice after treatment with car-
cinogens such as MCA. The immunodeficient mice tested in such experiments include gene-targeted mice
on pure genetic backgrounds with deficits of innate or adaptive immunity as well as wild-type mice
rendered immunodeficient by chronic administration of monoclonal antibodies that, for example, deplete
CD4
+
and CD8
+
T cells or interferon-g. Immunodeficiency has also been found to increase the sus-
ceptibility of untreated mice to spontaneously arising tumors and to increase the incidence of tumor
formation in mouse genetic models of cancer. Schematic is based on experiments described in (13).
www.sciencemag.org SCIENCE VOL 331 25 MARCH 2011 1565
SPECIALSECTION
2012 Cancer Crusade at 40 Collections Booklet
www.sciencemag.org SCIENCE 27
inflammatory environment that facilitates tumor-
igenesis by eliminating pathogens and by prompt
resolution of inflammation; and (iii) it eliminates
tumor cells in certain tissues because nascent
transformed cells often co-express ligands for
activating receptors on innate immune cells and
tumor antigens (see below) that are recognized
by immune receptors on lymphocytes of the
adaptive immune system. This third role is most
pertinent to our discussion.
Tumor Antigens and Cancer Immunosurveillance
A fundamental tenet of tumor immunology in
general and of cancer immunosurveillance in par-
ticular is that cancer cells express antigens that
differentiate them from their nontransformed
counterparts. The existence of tumor antigens was
first demonstrated by the finding that mice immu-
nized with chemically induced tumors were pro-
tected against subsequent rechallenge with the same
tumor [reviewed in (2)]. These types of tumor anti-
gens became known as transplantation rejection
antigens, and similar antigens have since been
demonstrated in a wide variety of experimentally
induced tumors [such as those induced by dif-
ferent carcinogens, viruses, or ultraviolet (UV)
irradiation] and even in spontaneous tumors. Sub-
sequent molecular studies revealed that these anti-
gens were often products of mutated cellular genes,
aberrantly expressed normal genes, or genes en-
coding viral proteins. In the case of human cancer,
identification of tumor antigens required the devel-
opment of novel in vitro detection and cloning
methods that used as probes antibodies and cyto-
lytic T lymphocytes (CD8
+
Tcells) derived from
cancer patients that were specific for the autolo-
gous tumor (1922). The human tumor antigens
discovered in these and other ways include differ-
entiation antigens (such as melanocyte differen-
tiation antigens), mutational antigens (such as p53),
overexpressed cellular antigens (such as HER-2),
viral antigens (such as human papillomavirus pro-
teins), and cancer/testis (CT) antigens that are ex-
pressed in germ cells of testis and ovary but silent
in normal somatic cells (such as MAGE and NY-
ESO-1) (23). Thus, the identification of this large
array of immunogenic mouse and human tumor
antigens puts to rest the long-held viewthat tumor
antigens are overexpressed normal proteins and
therefore were subject to immunological tolerance.
The Cancer Immunoediting Hypothesis
The discovery in 2001 that the immune system
controls not only tumor quantity but also tumor
quality (immunogenicity) (13, 24) prompted a
major revision of the cancer immunosurveillance
hypothesis. This study revealed that tumors formed
in mice that lacked an intact immune system
were, as a group, more immunogenic (and hence
were classified as unedited) than similar tumors
derived fromimmunocompetent mice (and hence
were termed edited) (Fig. 2). The notion that
the immune system not only protects the host
against tumor formation but also shapes tumor
immunogenicity is the basis of the cancer im-
munoediting hypothesis, which stresses the dual
host-protective and tumor-promoting actions of
immunity on developing tumors.
We postulate that the cancer immunoediting
process, in its most complex embodiment, pro-
ceeds sequentially through three distinct phases
that we have termed elimination, equilibri-
um, and escape (Fig. 3) (18, 2429). Howev-
er, in some cases tumor cells may directly enter
into either the equilibrium or escape phases with-
out passing through an earlier phase. In addition,
external factors may influence the directionality
of the flow. The latter consideration may help
explain the influences of environmental stress, im-
mune system deterioration accompanying aging,
and even immunotherapeutic intervention on
tumor cell outgrowth in human cancer patients.
Elimination. The elimination phase is best
described as an updated version of cancer im-
munosurveillance, in which the innate and adapt-
ive immune systems work together to detect the
presence of a developing tumor and destroy it
before it becomes clinically apparent. The mecha-
nisms by which the immune system is alerted to
the presence of a developing tumor are not fully
understood. Among the possibilities are the clas-
sical danger signals such as Type I IFNs as
originally described by Matzinger (7), which we
now know are induced early during tumor devel-
opment. These cytokines activate dendritic cells
and promote induction of adaptive anti-tumor
immune responses. However, roles for different
damage-associated molecular pattern molecules
(DAMPs) need also to be considered because
they are released either directly fromdying tumor
cells [suchas highmobilitygroupbox1(HMGB1)]
Day
T
u
m
o
r

s
i
z
e

Day
T
u
m
o
r

s
i
z
e

30 0 Day
T
u
m
o
r

s
i
z
e

20
30 0
20
30 0
20
Unedited
progressor
Edited
progressor
Unedited
regressor
Primary
tumor
growing in:
Tumor
cell
line
Inject
into naive WT
recipients
Monitor
tumor
growth
Growth
profile
Tumor
phenotype
Immunodeficient
Mouse
Immunodeficient
Mouse
Immunocompetent
mouse
Fig. 2. Tumors in immunocompetent mice are qualitatively different from tumors in immunodeficient
mice. This observation, which led to the formulation of the cancer immunoediting hypothesis, is based on
comparative analyses of carcinogen-induced tumors harvested from immunocompetent and immuno-
deficient mice. In these experiments, tumor cell lines were established from tumors arising in each group
of mice, and these cells were then injected into immunodeficient recipient mice or immunocompetent
wild-type (WT) recipient mice. Tumor cells from carcinogen-treated WT mice formed progressively growing
tumors in both immunodeficient mice (not shown) and nave syngeneic immunocompetent mice (blue)
100% of the time. In contrast, although tumor cells from carcinogen-treated immunodeficient mice grew
progressively when transplanted into immunodeficient mice (not shown), only half of the tumor cell lines
were capable of forming progressively growing tumors in nave syngeneic immunocompetent recipients
(purple), whereas the other half of the cell lines were rejected by the recipients (red). Thus, tumors from
immunodeficient mice are termed unedited and further designated as progressor or regressor to
denote their growth phenotypes after injection into nave WT recipients. Carcinogen-induced tumors from
immunocompetent mice are termed edited because they are less immunogenic and show only a
progressor growth phenotype. Schematic is based on experiments described in (13).
25 MARCH 2011 VOL 331 SCIENCE www.sciencemag.org 1566
Cancer Crusade at 40
28 SCIENCE www.sciencemag.org
or from damaged tissues (such as hyaluronan
fragments) as solid tumors begin to grow in-
vasively (30). A third potential mechanism may
involve stress ligands such as RAE-1 and H60
(mouse) or MICA/B (human) that are frequently
expressed on the surface of tumor cells. Such lig-
ands bind to activating receptors on innate im-
mune cells, leading to release of pro-inflammatory
and immunomodulatory cytokines, which in turn
establish a microenvironment that facilitates the
development of a tumor-specific adaptive im-
mune response (31). Although in some experi-
mental systems, activation of innate immunity
can protect against tumor development, in most
systems effective cancer immunosurveillance re-
sponses require the additional expression of tu-
mor antigens capable of propagating the expansion
of effector CD4
+
and CD8
+
Tcells. Thus, coordi-
nated and balanced activation of both innate and
adaptive immunity is needed to protect the host
against a developing tumor. If tumor cell destruc-
tion goes to completion, the elimination phase
represents an endpoint of the cancer immunoedit-
ing process.
The elimination phase has not yet been di-
rectly observed in vivo, but its existence has been
inferred from the earlier onset or greater pene-
trance of neoplasia in mice lacking certain im-
mune cell subsets, recognition molecules, effector
pathways, or cytokines and by studies comparing
tumor initiation, growth, and metastases in wild-
type versus immunodeficient mice [reviewed in
(18)]. These studies have revealed that the im-
mune components required for effective elimina-
tion of any given tumor are dependent on specific
characteristics of the tumor, such as how it orig-
inated (spontaneous versus carcinogen-induced),
its anatomic location, and its rate of growth.
Equilibrium. Rare tumor cell variants may
survive the elimination phase and enter the equi-
librium phase, in which the adaptive immune
system prevents tumor cell outgrowth and also
sculpts the immunogenicity of the tumor cells.
We envisage equilibrium to be the longest phase
of the cancer immunoediting processperhaps
extending throughout the life of the host. As
such, it may represent a second stable endpoint
of cancer immunoediting. In equilibrium, the im-
mune system maintains residual tumor cells in
a functional state of dormancy, a term used to
describe latent tumor cells that may reside in
patients for decades before eventually resuming
growth as either recurrent primary tumors or dis-
tant metastases (32). Equilibrium thus represents
a type of tumor dormancy in which outgrowth
of occult tumors is specifically controlled by
immunity.
An early suggestion that the immune system
could maintain tumor cells in a dormant/equilibrium
state came from tumor transplantation experi-
ments in which mice were primed with a trans-
plantable tumor and then rechallenged with the
same tumor in order to induce tumor latency (33).
However, stronger evidence for the existence of
an immunologically mediated equilibrium phase
came from primary tumorigenesis experiments
showing that immunocompetent mice treated
with low-dose carcinogen [3-methylcholanthrene
(MCA)] harbored occult cancer cells for an ex-
tended time period even when the mice did not
develop any apparent tumors (34). When the
immune system of these mice was ablated [by
administering monoclonal antibodies (mAbs) that

T
Transformed
cells
Normal
tissue
Elimination Equilibrium Escape
Cancer Immunoediting
Extrinsic tumor
suppression
Tumor growth
promotion
Tumor dormancy
and editing
Danger
signals
Tumor
antigens
NKR
ligands
Intrinsic tumor suppression
(senescence, repair,
and/or apoptosis)
Carcinogens
Radiation
Viral infections
Chronic inflammation
Inherited genetic mutations
Antigen loss
MHC loss
Innate &
adaptive
immunity
Normal cell
Highly immunogenic
transformed cell
Poorly immunogenic
and immunoevasive
transformed cells
IFN-/
IFN-

TRAIL
NKG2D
Perforin
TNF
IL-12 CTLA-4
PD-1
CTLA-4
PD-1 MDSC
reg
CD8+
T cell
CD8+
T cell
TGF-
IDO
IL-6, IL-10
Galectin-1
CD8+
T cell
NK
M
CD8+
T cell
IFN-
IL-12
CD4+
cell
PD-L1
CD8+
T cell
CD8+
T cell
NK
NKT
cell
M
DC
CD4+
T cell
CD4+
T cell
T
cell
T
Fig. 3. The cancer immunoediting concept. Cancer immunoediting is an extrinsic tumor suppressor
mechanism that engages only after cellular transformation has occurred and intrinsic tumor suppressor
mechanisms have failed. In its most complex form, cancer immunoediting consists of three sequential
phases: elimination, equilibrium, and escape. In the elimination phase, innate and adaptive immunity
work together to destroy developing tumors long before they become clinically apparent. Many of the
immune molecules and cells that participate in the elimination phase have been identified, but more work
is needed to determine their exact sequence of action. If this phase goes to completion, then the host
remains free of cancer, and elimination thus represents the full extent of the process. If, however, a rare
cancer cell variant is not destroyed in the elimination phase, it may then enter the equilibrium phase, in
which its outgrowth is prevented by immunologic mechanisms. T cells, IL-12, and IFN-g are required to
maintain tumor cells in a state of functional dormancy, whereas NK cells and molecules that participate in
the recognition or effector function of cells of innate immunity are not required; this indicates that
equilibrium is a function of adaptive immunity only. Editing of tumor immunogenicity occurs in the
equilibrium phase. Equilibrium may also represent an end stage of the cancer immunoediting process and
may restrain outgrowth of occult cancers for the lifetime of the host. However, as a consequence of
constant immune selection pressure placed on genetically unstable tumor cells held in equilibrium, tumor
cell variants may emerge that (i) are no longer recognized by adaptive immunity (antigen loss variants or
tumors cells that develop defects in antigen processing or presentation), (ii) become insensitive to
immune effector mechanisms, or (iii) induce an immunosuppressive state within the tumor microenvi-
ronment. These tumor cells may then enter the escape phase, in which their outgrowth is no longer blocked
by immunity. These tumor cells emerge to cause clinically apparent disease. [Figure adapted from (18)]
www.sciencemag.org SCIENCE VOL 331 25 MARCH 2011 1567
SPECIALSECTION
2012 Cancer Crusade at 40 Collections Booklet
www.sciencemag.org SCIENCE 29
deplete T cells and IFN-g], tumors rapidly ap-
peared at the original MCA injection site in half
of the mice. Tumor cells isolated fromthese lesions
were highly immunogenic and thus resembled un-
edited sarcoma cells derived from MCA-treated
immunodeficient RAG2
/
mice. Further analy-
ses revealed that adaptive immunityspecifically,
interleukin-12 (IL-12), IFN-g, CD4
+
, and CD8
+
Tcellsbut not innate immunity was responsible
for maintaining the occult tumor cells in equi-
librium. This observation mechanistically distin-
guishes equilibriumfromelimination because the
latter displays an obligate requirement for both
innate and adaptive immunity. Additional studies
with different mouse tumor models have con-
firmed the capacity of the immune system to
control the outgrowth of occult primary carcino-
mas and metastases for extended periods of time
(35, 36). In the low-dose MCA system, equilib-
rium appears to be the result of both the growth
inhibitory and cytocidal actions of immunity on
the residual tumor cells (34). Conceivably, the
same immune functions also provide the selec-
tive pressure that promote outgrowth of tumor
cells that have acquired the most immunoevasive
mutations.
Escape. In the escape phase, tumor cells that
have acquired the ability to circumvent immune
recognition and/or destruction emerge as pro-
gressively growing, visible tumors. Progression
from equilibrium to the escape phase can occur
because the tumor cell population changes in re-
sponse to the immune systems editing functions
and/or because the host immune system changes
in response to increased cancer-induced immu-
nosuppression or immune system deterioration.
Tumor cell escape can occur through many
different mechanisms [reviewed in (18, 2426,
28, 37, 38)]. At the tumor cell level, alterations
leading to reduced immune recognition (such as a
loss of antigens) or increased resistance to the
cytotoxic effects of immunity (for example, through
induction of anti-apoptotic mechanisms involv-
ing persistent activation of pro-oncogenic tran-
scription factors such as STAT3 or expression of
anti-apoptotic effector molecules such as BCL-2)
promote tumor outgrowth. Loss of tumor antigen
expression is one of the best-studied escape mech-
anisms, and it can occur in at least three ways: (i)
through emergence of tumor cells that lack ex-
pression of strong rejection antigens, (ii) through
loss of major histocompatibility complex (MHC)
class I proteins that present these antigens to
tumor-specific T cells, or (iii) through loss of an-
tigen processing function within the tumor cell
that is needed to produce the antigenic peptide
epitope and load it onto the MHC class I mol-
ecule. All of these alterations are probably driven
by a combination of genetic instability inherent in
all tumor cells and the process of immunoselec-
tion (24, 38). The end result is the generation via a
Darwinian selection process of poorly immuno-
genic tumor cell variants that become invisible
to the immune system and thus acquire the ca-
pacity to grow progressively.
Alternatively, escape may result from the es-
tablishment of an immunosuppressive state with-
in the tumor microenvironment (39). Tumor cells
can promote the development of such a state by
producing immunosuppressive cytokines such
as vascular endothelial growth factor (VEGF),
transforming growth factorb (TGF-b), galectin,
or indoleamine 2,3-dioxygenase (IDO) and/or
by recruiting regulatory immune cells that func-
tion as the effectors of immunosuppression [re-
viewed in (18)]. Regulatory Tcells (T
reg
cells) and
myeloid-derived suppressor cells (MDSCs) are
two major types of immunosuppressive leuko-
cyte populations that play key roles in inhib-
iting host-protective antitumor responses. T
reg
cells are CD4
+
T cells that constitutively express
CD25 and the transcription factor Foxp3. When
stimulated, they inhibit the function of tumor-
specific T lymphocytes by producing the immu-
nosuppressive cytokines IL-10 and TGF-b; by
expressing the negative co-stimulatory molecules
CTLA-4, PD-1, and PD-L1; and by consuming
IL-2, a cytokine that is critical for maintaining
CTL function. MDSCs are a heterogeneous group
of myeloid progenitor cells and immature mye-
loid cells that inhibit lymphocyte function by in-
ducing T
reg
cells; producing TGF-b; depleting or
sequestering the amino acids arginine, trypto-
phan, or cysteine required for T cell function; or
nitrating T cell receptors or chemokine receptors
on tumor-specific T cells.
Cancer Immunoediting Versus Inflammation
Inflammation is a complex physiological process
that normally functions to maintain tissue ho-
meostasis in response to tissue stressors such as
infection or tissue damage (40). Acute inflam-
mation (innate immunity) frequently precedes the
development of protective adaptive immune re-
sponses to pathogens and cancer. Chronic in-
flammation, on the other hand, has been shown
to contribute to tumorigenesis at all stages. It
contributes to cancer initiation by generating geno-
toxic stress, to cancer promotion by inducing cel-
lular proliferation, and to cancer progression by
enhancing angiogenesis and tissue invasion (41).
On the basis of these observations, it has been
proposed that inflammation and tumor immunity
are mutually exclusive processes (9, 10).
In our view, a more likely interpretation is that
tumor-promoting inflammation and protective tu-
mor immunity are dynamically interconnected
processes that vie for dominance as tumor cells
develop and transit through cancer immunoedit-
ing (42). This scenario is supported by data from
several different experimental systems. First, al-
though tumor induction in MCA-treated mice
requires the participation of pro-inflammatory
cytokines/signaling (such as IL-1b, IL-23, or
MyD88) the tumors, once formed, became sus-
ceptible to control by other components of im-
munity (such as IFN-g, IFN-a/b, IL-12, or T
cells) (43). Thus, tumor-promoting inflammation
and cancer immunosurveillance/immunoediting
can coexist within the same tumor model. Sec-
ond, immune components with pro-oncogenic
activity can also promote induction of tumor im-
munity, depending on when they are recruited
into the cancer development process. For ex-
ample, whereas MyD88 and IL-1b clearly pro-
mote carcinogen-induced tumorigenesis in mouse
models (4447), the same proteins have the op-
posite effect at later stages of tumorigenesis
that is, they promote development of protective
immune responses against established tumors by
facilitating recognition of tumor cells undergoing
immunogenic death (4850). This paradoxical
role of inflammatory cytokines and the immune
response in cancer is also illustrated by the ob-
servation that tumor necrosis factora (TNF-a)
has both tumor-promoting and antitumor activ-
ities in mouse and Drosophila tumor models (51)
and by more recent work showing that in a
mouse melanoma model, IFN-g is required both
for UVB-induced tumor formation and for im-
mune rejection of these tumors (52). Lastly, in-
flammation can play an important role during
tumor escape, when inflammatory cells are re-
cruited to the site of a progressively growing tu-
mor, undergo activation by cancer-derived products
(such as VEGF), and suppress protective tumor
immunity (41).
Cancer Immunoediting in Humans
Although studies of tumor development in mice
served as the main driver for the formulation of
the cancer immunoediting hypothesis, evidence
has since been obtained indicating that immu-
noediting also occurs in humans and can alter the
course of tumor development in cancer patients.
We discuss three key types of evidence support-
ing this conclusion; more comprehensive sum-
maries can be found in (18, 24).
Intratumoral immune responses predict pa-
tient prognosis. The strongest evidence of cancer
immunoediting in humans comes from reports
that correlate the quantity, quality, and spatial dis-
tribution of tumor-infiltrating lymphocytes (TILs)
with patient survival. Tumor infiltration by IFN-g
producing Th1 CD4
+
T cells and CD8
+
T cells,
and the presence of cytokines such as IFN-g and
TNF-a that promote tumor control, has been as-
sociated with an improved prognosis for patients
with many different cancers. A study of melano-
ma patients provided an early indication that TILs
are associated with a favorable patient prognosis
(53, 54). A subsequent landmark study by Naito
et al. demonstrated that the presence and location
of one particular type of TIL, CD8
+
T cells, in
colon cancers had a particularly important influ-
ence on clinical outcome; specifically, accumu-
lation of CD8
+
Tcells within the tumor predicted
improved patient survival, whereas accumulation
of the same cells at the tumor margin had no effect
25 MARCH 2011 VOL 331 SCIENCE www.sciencemag.org 1568
Cancer Crusade at 40
30 SCIENCE www.sciencemag.org
on survival (55). Subsequent studies in ovarian
cancer, melanoma, and colon cancer confirmed this
observation and further showed that the ratio and
distribution patterns of intratumoral CD8
+
Tcells
and T
reg
cells were critical determinants of prog-
nosis (5659). Recent exciting studies of human
colon and lung cancers have not only confirmed
these observations but have provided quantitative
insights into the key variables involved (58, 59).
Remarkably, the type and density of lymphocytes
infiltrating these cancers was found to be a more
powerful prognostic indicator than previous path-
ological criteria for tumor staging and was even
more predictive than correlating disease progres-
sion with oncogene expression.
Spontaneous immune responses in cancer
patients. A major advance to the field of tumor
immunology came from the demonstration that
cancer patients can develop high levels of anti-
body and T cell responses to antigens expressed
in their tumors [reviewed in (60)]. These immune
responses are generally observed in patients with
progressively growing tumors, indicating that
immune recognition of cancer does not always
result in immune protection. However, there is
presently no way to know whether such immune
responses influence the rate or pattern of tumor
growth in these patients and whether these re-
sponses represent the footprint of incomplete or
ongoing elimination or equilibrium phases of can-
cer immunoediting. An example of the latter
comes from the analysis of individuals with para-
neoplastic neurologic disorders (PNDs). PNDs
arise as a consequence of antibody and T cell
responses against certain autologous tumors that
ectopically express proteins normally expressed
only in cells of the nervous system (61, 62). This
antitumor response develops into an autoimmune
response as it attacks normal neurons that express
the tumor-associated antigens. The neurologic
dysfunctions observed in PND usually become
evident before the tumor is discovered.
Immunodeficiency is associated with a higher
risk of cancer. Immunodeficiency has been linked
to increased cancer risk in patients with AIDS
and in transplant recipients maintained on im-
munosuppressants [reviewed in (18, 24)]. Al-
though the cancers arising in these patients are
typically those with a viral etiology such as lym-
phomas (Epstein-Barr virus), Kaposis sarcoma
(herpesviruses), and cervical cancer (human pap-
illoma viruses), there is at least some evidence
that these patients are at greater risk for malignan-
cies of the colon, lung, pancreas, kidney, head
and neck, and endocrine system as well as non-
melanoma skin cancers. Melanoma incidence
rates are also 2 to 10 times higher than average in
renal transplant patients. Interestingly, increased
incidences of other cancersincluding breast,
prostate, ovarian, brain, and testeshave not been
observed in immunosuppressed transplant patients.
Insights into the role of the immune systemin
human cancer have also come from anecdotal
reports of cancer being transferred from an organ
donor to the immunosuppressed recipient (63, 64).
In one study, two individuals received kidney
transplants from the same cadaver donor, and
both recipients later succumbed to malignant
melanoma that was shown by tissue typing to be
of donor origin. Medical records revealed that
16 years before her death, the donor had been
diagnosed with malignant melanoma and suc-
cessfully treated. One interpretation of these find-
ings is that the donors kidneys contained dormant
melanoma cells held in equilibrium by the do-
nors immune system. The transfer of the kidney
to nave and immunosuppressed recipients may
have removed the immune pressure holding the
tumor cells in equilibrium and thus allowed the
occult melanoma cells to grow out into clinically
apparent cancer. Together, these clinical obser-
vations are consistent with the hypothesis that de
novo malignancies arise only in certain permissive
microenvironments created by immunosuppres-
sive regimens that suspend or severely compro-
mise the elimination and/or equilibriumphases of
cancer immunoediting.
Cancer Immunoediting in Immunotherapy
With our newfound knowledge of the immune
systems capacity to not only recognize and de-
stroy cancer but also to shape cancer immunoge-
nicity, more informed attempts to control cancer
via immunological means can nowbe pursued. It
is now well accepted that progressively growing,
clinically apparent tumors in cancer patients have
developed successful strategies to circumvent the
natural, extrinsic tumor-suppressor mechanisms
of immunity. Thus to be effective, immunothera-
pies will have to increase the quality or quantity
of immune effector cells, reveal additional pro-
tective tumor antigens, and/or eliminate cancer-
induced immunosuppressive mechanisms. Multiple
forms of immunotherapy are being explored to
achieve these objectives. These include (i) vac-
cine approaches to elicit strong specific immune
responses to tumor antigens such as MAGE-3
and NY-ESO-1; (ii) approaches involving adop-
tive transfer of in vitro expanded, naturally aris-
ing, or genetically engineered tumor-specific
lymphocytes; (iii) therapeutic administration of
monoclonal antibodies such as Rituximab (di-
rected against CD20 on leukemia and lymphoma
cells) and Herceptin (directed against HER2 on
breast cancer cells) to target and eliminate tumor
cells; and (iv) approaches that inhibit or destroy
the molecular or cellular mediators of cancer-
induced immunosuppression such as CTLA-4,
PD-1, or T
reg
cells.
Quantitative analyses performed on patients
undergoing various forms of cancer immuno-
therapy have revealed that the cancer immunoedit-
ing process reoccurs either in part or in its entirety
during therapy. Specifically, whereas some treated
patients display responses that recapitulate the
elimination phase of cancer immunoediting (for
example, they develop increased numbers of
tumor-specific T cells with intact effector func-
tion, or they show destruction of some or all
tumor cells) others show evidence for estab-
lishment of a therapeutically induced equilibrium
phase, and still others display evidence for devel-
opment of additional escape mechanisms, such
as outgrowth of antigen-loss variants. Recently,
the occurrence of all three phases of cancer im-
munoediting has been documented in a melanoma
patient with preexisting NY-ESO-1 immunity un-
dergoing CTLA-4 blockade monotherapy (65).
When observed over a 28-month period after
initiation of immunotherapy, melanoma lesions
could be identified that disappeared (elimination),
were held in a protracted state of growth dor-
mancy (equilibrium), or continued to grow (es-
cape). Thus, cancer immunoediting can occur not
only when the unmanipulated immune system
encounters a developing tumor but also when an
established tumor is subjected to immunotherapy.
Future Directions
We envision that future work on cancer immu-
noediting will address five major questions:
(i) What immune effector processes mediate
cancer elimination, equilibrium, and escape? T
cells play a critical role in mediating both natu-
ral and therapeutically induced cancer immuno-
editing responses. However, it remains unclear
whether they represent the ultimate effectors of
these processes. Although activated T cells and
other lymphocytes can certainly kill tumor cells,
they also elaborate a variety of cytokines such as
IFN-g and TNF-a that can exert profound cy-
tostatic and cytocidal effects on tumor cells, ac-
tivate tumor cytotoxicity in other cell types (such
as macrophages) present in the tumor micro-
environment, and block tumor angiogenesis. Iden-
tifying the molecular mechanisms and targets
responsible for cancer elimination, equilibrium,
and escape will determine whether the three
phases of cancer immunoediting are manifest by
similar or distinct effector processes.
(ii) How do the antigens of nascent tumors
differ from the antigens of established, clinically
apparent tumors? Almost all of our knowledge of
tumor antigens is based on analyses of advanced
cancers in imunocompetent hosts. It is important
to identify the antigens expressed in early devel-
oping tumors because these are the initial targets
of the elimination phase of cancer immunoedit-
ing. In addition, it would be interesting to define
the antigens of tumors from immunosuppressed
individuals because these antigens may not have
undergone extensive editing and thus may be
similar to the antigens of nascent tumors. Are the
major antigens of unedited tumors more likely to
be associated with driver or passenger mutations?
An answer to this question may provide insights
into the capacity of immunity to eliminate devel-
oping tumor cells, hold them in an equilibrium
state, or facilitate their outgrowth. Can we use
www.sciencemag.org SCIENCE VOL 331 25 MARCH 2011 1569
SPECIALSECTION
2012 Cancer Crusade at 40 Collections Booklet
www.sciencemag.org SCIENCE 31
information from high-throughput screening of
cancer genomes and proteomes or other cutting-
edge techniques to rapidly identify the mutations
and epigenetic changes in unedited and edited
cancer cells that result in formation of function-
ally relevant tumor antigens?
(iii) What is the link between the types of
antigens expressed in a tumor and the mechanism
of cellular transformation? With the exception of
viral-mediated oncogenesis, little is known about
whether and how the mechanisms leading to cell
transformation affect the quality or quantity of tu-
mor antigens. The possibility should be considered
that experimental tumors, in which transformation
is driven by strong oncogenes, may harbor fewer
passenger mutations than do spontaneous tumors
(66). Because passenger mutations can produce tu-
mor antigens, oncogene-driven cancer models may
therefore not always be optimal models for explor-
ing the immunology of naturally developing tu-
mors. However, a recent study revealed a previously
unknown capacity of the immune system to
sustain tumor regression upon oncogene inacti-
vation (67). These considerations emphasize the
need for further work on defining the relation-
ships between cellular transformation mechanisms
and tumor immunogenicity. In the future, careful
consideration should be given to the use of cancer
models that most closely recapitulate both the
biology and immunology of human cancers.
(iv) Is a durable state of equilibrium a desir-
able and attainable endpoint for cancer immuno-
therapy? We currently know very little about the
effector mechanisms that operate in the equilib-
rium phase. To date, mouse studies reveal that T
cells and IFN-g contribute to equilibrium, but the
recognition pathways and immune network and
mechanisms remain unclear. If these can be iden-
tified, it might be possible to develop cancer ther-
apies aimed at recapitulating immunological
tumor dormancy. It will also be important to un-
derstand what effect conventional interventions,
such as surgery, radiotherapy, and chemotherapy
have on the equilibrium phase.
(v) Howcan we most effectively inhibit cancer-
induced immunosuppressive mechanisms at the
tumor site so as to boost the host-protective anti-
tumor effects of preexisting or therapeutically
induced immunity without concomitantly induc-
ing life-threatening autoimmunity? Arguably, this
may be the most pressing question in all of tumor
immunology. Unlike other therapies that target
cancer cells, therapies aimed at inhibiting immu-
nosuppression target the immune system itself.
An exciting approach being evaluated in clinical
trials involves the use of monoclonal antibodies to
blockade immunosuppressive molecules such as
CTLA-4 or PD-1 expressed by Tcells. In a related
approach, the effectiveness of monoclonal anti-
bodies that block the PD-1 ligand, PD-L1, which
can be expressed on both tumor cells and normal
host cells, is also being explored. These types of
therapies have been designatedcheckpoint block-
ade (68). In the case of CTLA-4 blockade, a
recent phase III clinical trial reported that therapy
with CTLA-4blocking antibodies imparted a
significant survival benefit in approximately one-
third of patients with metastatic melanoma, mak-
ing this drug a promising treatment for cancer
(69). The success of the current CTLA-4 blockade
clinical trials has stimulated interest in blocking
other potential effectors of immunosuppression,
including the soluble (such as IDO and TGF-b)
and cellular (such as T
reg
cells and MDSCs) me-
diators of the process. Clearly, there is much to
be learned about the benefits and risks of inhib-
iting the different immunosuppressive mecha-
nisms that may be concurrently operating in the
cancer patient.
Conclusion
The cancer immunoediting concept attempts to
integrate the diverse effects of the immune sys-
tem on tumor development and outgrowth. With
elucidation of the molecular and cellular mech-
anisms that underlie the elimination, equilibrium,
and escape phases of this process, it should be
possible to develop newcancer immunotherapies
that are safer and more efficacious in a substan-
tial percentage of cancer patients. Given the
well-established effects of immunity on cancer
development and outgrowth, escape from im-
mune control can now be viewed as one of the
Hallmarks of Cancer (70).
References and Notes
1. P. Ehrlich, Ned. Tijdschr. Geneeskd. 5, 273 (1909).
2. L. J. Old, E. A. Boyse, Annu. Rev. Med. 15, 167 (1964).
3. M. Burnet, BMJ 1, 841 (1957).
4. L. Thomas, Cellular and Humoral Aspects of the
Hypersensitive States, H. Lawrence, Ed. (Hoeber-Harper,
New York, 1959).
5. O. Stutman, Science 183, 534 (1974).
6. O. Stutman, Adv. Cancer Res. 22, 261 (1975).
7. P. Matzinger, Annu. Rev. Immunol. 12, 991 (1994).
8. D. Pardoll, Annu. Rev. Immunol. 21, 807 (2003).
9. F. Balkwill, A. Mantovani, Lancet 357, 539 (2001).
10. M. Karin, Y. Cao, F. R. Greten, Z. W. Li, Nat. Rev.
Cancer 2, 301 (2002).
11. A. S. Dighe, E. Richards, L. J. Old, R. D. Schreiber,
Immunity 1, 447 (1994).
12. D. H. Kaplan et al., Proc. Natl. Acad. Sci. U.S.A. 95,
7556 (1998).
13. V. Shankaran et al., Nature 410, 1107 (2001).
14. M. J. Smyth et al., J. Exp. Med. 192, 755 (2000).
15. M. J. Smyth et al., J. Exp. Med. 191, 661 (2000).
16. S. E. Street, J. A. Trapani, D. MacGregor, M. J. Smyth,
J. Exp. Med. 196, 129 (2002).
17. M. Girardi et al., Science 294, 605 (2001).
18. M. D. Vesely, M. H. Kershaw, R. D. Schreiber, M. J. Smyth,
Annu. Rev. Immunol. 10.1146/annurev-immunol-
031210-101324 (2011).
19. L. J. Old, Cancer Res. 41, 361 (1981).
20. A. Knuth, B. Danowski, H. F. Oettgen, L. J. Old,
Proc. Natl. Acad. Sci. U.S.A. 81, 3511 (1984).
21. P. van der Bruggen et al., Science 254, 1643 (1991).
22. U. Sahin et al., Proc. Natl. Acad. Sci. U.S.A. 92,
11810 (1995).
23. M. A. Cheever et al., Clin. Cancer Res. 15, 5323 (2009).
24. G. P. Dunn, A. T. Bruce, H. Ikeda, L. J. Old,
R. D. Schreiber, Nat. Immunol. 3, 991 (2002).
25. G. P. Dunn, L. J. Old, R. D. Schreiber, Immunity 21, 137
(2004).
26. G. P. Dunn, L. J. Old, R. D. Schreiber, Annu. Rev.
Immunol. 22, 329 (2004).
27. G. P. Dunn, C. M. Koebel, R. D. Schreiber, Nat. Rev.
Immunol. 6, 836 (2006).
28. M. J. Smyth, G. P. Dunn, R. D. Schreiber, Adv. Immunol.
90, 1 (2006).
29. J. B. Swann, M. J. Smyth, J. Clin. Invest. 117, 1137 (2007).
30. G. P. Sims, D. C. Rowe, S. T. Rietdijk, R. Herbst,
A. J. Coyle, Annu. Rev. Immunol. 28, 367 (2010).
31. N. Guerra et al., Immunity 28, 571 (2008).
32. J. A. Aguirre-Ghiso, Nat. Rev. Cancer 7, 834 (2007).
33. J. D. Farrar et al., J. Immunol. 162, 2842 (1999).
34. C. M. Koebel et al., Nature 450, 903 (2007).
35. S. Loeser et al., J. Exp. Med. 204, 879 (2007).
36. J. Eyles et al., J. Clin. Invest. 120, 2030 (2010).
37. L. Zitvogel, A. Tesniere, G. Kroemer, Nat. Rev. Immunol.
6, 715 (2006).
38. H. T. Khong, N. P. Restifo, Nat. Immunol. 3, 999 (2002).
39. S. Radoja, T. D. Rao, D. Hillman, A. B. Frey, J. Immunol.
164, 2619 (2000).
40. R. Medzhitov, Nature 454, 428 (2008).
41. S. I. Grivennikov, F. R. Greten, M. Karin, Cell 140, 883
(2010).
42. J. D. Bui, R. D. Schreiber, Curr. Opin. Immunol. 19, 203
(2007).
43. M. W. Teng et al., Proc. Natl. Acad. Sci. U.S.A. 107,
8328 (2010).
44. W. E. Naugler et al., Science 317, 121 (2007).
45. S. Rakoff-Nahoum, R. Medzhitov, Science 317, 124 (2007).
46. J. B. Swann et al., Proc. Natl. Acad. Sci. U.S.A. 105,
652 (2008).
47. Y. Krelin et al., Cancer Res. 67, 1062 (2007).
48. L. Apetoh et al., Nat. Med. 13, 1050 (2007).
49. M. Obeid et al., Nat. Med. 13, 54 (2007).
50. F. Ghiringhelli et al., Nat. Med. 15, 1170 (2009).
51. J. B. Cordero et al., Dev. Cell 18, 999 (2010).
52. M. R. Zaidi et al., Nature 469, 548 (2011).
53. W. H. Clark Jr. et al., J. Natl. Cancer Inst. 81, 1893 (1989).
54. C. G. Clemente et al., Cancer 77, 1303 (1996).
55. Y. Naito et al., Cancer Res. 58, 3491 (1998).
56. E. Sato et al., Proc. Natl. Acad. Sci. U.S.A. 102, 18538
(2005).
57. I. S. van Houdt et al., Int. J. Cancer 123, 609 (2008).
58. F. Pags et al., N. Engl. J. Med. 353, 2654 (2005).
59. J. Galon et al., Science 313, 1960 (2006).
60. M. Dougan, G. Dranoff, Annu. Rev. Immunol. 27, 83 (2009).
61. M. L. Albert, R. B. Darnell, Nat. Rev. Cancer 4, 36 (2004).
62. M. L. Albert et al., Nat. Med. 4, 1321 (1998).
63. H. Myron Kauffman et al., Transplantation 74, 358 (2002).
64. R. M. MacKie, R. Reid, B. Junor, N. Engl. J. Med. 348,
567 (2003).
65. J. Yuan et al., Cancer Immun. 10, 1 (2010).
66. M. DuPage et al., Cancer Cell 19, 72 (2011).
67. K. Rakhra et al., Cancer Cell 18, 485 (2010).
68. A. J. Korman, K. S. Peggs, J. P. Allison, Adv. Immunol. 90,
297 (2006).
69. F. S. Hodi et al., N. Engl. J. Med. 363, 711 (2010).
70. D. Hanahan, R. A. Weinberg, Cell 144, 646 (2011).
71. We thank M. D. Vesely and M. H. Kershaw for
invaluable advice and contributions to this review and
E. R. Unanue and P. M. Allen for constructive
suggestions. R.D.S. is supported by grants from the
National Cancer Institute, the Cancer Research Institute,
and the Ludwig Institute for Cancer Research. L.J.O. is
supported by The Ludwig Institute for Cancer Research
and grants from the Cancer Research Institute. M.J.S.
is supported by a National Health and Medical Research
Council of Australia (NH&MRC) Australia Fellowship
and Program Grant and a grant from the Association
for International Cancer Research. We apologize to all
the investigators whose research could not be
appropriately cited because of the journals space
limitations. R.D.S. is a co-founder and member of the
Board of Directors of and is a paid scientific advisor
for Igenica, a biopharmaceutical company dedicated
to the discovery and development of antibody-based
therapeutics for the treatment of cancer.
10.1126/science.1203486
25 MARCH 2011 VOL 331 SCIENCE www.sciencemag.org 1570
Cancer Crusade at 40
32 SCIENCE www.sciencemag.org
To learn more, visit aaas.org/plusyou/project2061
AAAS is here promoting universal science literacy.
In 1985, AAAS founded Project 2061 with the goal of helping all Americans become literate in science, mathematics, and
technology. With its landmark publications Science for All Americans and Benchmarks for Science Literacy, Project 2061 set out
recommendations for what all students should know and be able to do in science, mathematics, and technology by the time they
graduate from high school. Today, many of the state standards in the United States have drawn their content from Project 2061.
Every day Project 2061 staff use their expertise as teachers, researchers, and scientists to evaluate textbooks and assessments,
create conceptual strand maps for educators, produce groundbreaking research and innovative books, CD-ROMs, and profes-
sional development workshops for educators, all in the service of achieving our goal of universal science literacy.
As a AAAS member, your dues help support Project 2061 as it works to improve science education. If you are not yet a AAAS
member, join us. Together we can make a difference.
To learn more, visit aaas.org/plusyou/project2061
AAAS is here promoting universal science literacy.
In 1985, AAAS founded Project 2061 with the goal of helping all Americans become literate in science, mathematics, and
technology. With its landmark publications Science for All Americans and Benchmarks for Science Literacy, Project 2061 set out
recommendations for what all students should know and be able to do in science, mathematics, and technology by the time they
graduate from high school. Today, many of the state standards in the United States have drawn their content from Project 2061.
Every day Project 2061 staff use their expertise as teachers, researchers, and scientists to evaluate textbooks and assessments,
create conceptual strand maps for educators, produce groundbreaking research and innovative books, CD-ROMs, and profes-
sional development workshops for educators, all in the service of achieving our goal of universal science literacy.
As a AAAS member, your dues help support Project 2061 as it works to improve science education. If you are not yet a AAAS
member, join us. Together we can make a difference.
www.aaas.org

Das könnte Ihnen auch gefallen