You are on page 1of 12

REVIEWS

TARGETING HIF-1 FOR


CANCER THERAPY
Gregg L. Semenza
Hypoxia-inducible factor 1 (HIF-1) activates the transcription of genes that are involved in crucial
aspects of cancer biology, including angiogenesis, cell survival, glucose metabolism and invasion.
Intratumoral hypoxia and genetic alterations can lead to HIF-1α overexpression, which has been
associated with increased patient mortality in several cancer types. In preclinical studies, inhibition
of HIF-1 activity has marked effects on tumour growth. Efforts are underway to identify inhibitors
of HIF-1 and to test their efficacy as anticancer therapeutics.

Interest in the role of hypoxia-inducible factor 1 (HIF-1) interfering RNA that targets HIF1α mRNA8; increased
in cancer biology has grown exponentially in the two expression in von Hippel–Lindau (VHL)-null cells or
decades since its identification1, biochemical purifica- in cells transfected with a HIF-1α expression vector9.
tion2 and molecular characterization3. Much has been The level of proof varies, especially in the present era
learned recently about the cellular and molecular biol- of global analysis by microarrays, in which the quan-
ogy of HIF-1 and its involvement in human cancer tity — but not the quality — of data has increased.
progression, based on the analysis of human cancer Analyses that do not involve the demonstration of a
biopsies and experimental animal models. The con- functional HIF-1 binding site identify both direct and
cept of intratumoral hypoxia as a driving force in can- indirect (secondary) targets of regulation by HIF-1.
cer progression has been covered in a previous review 4, Four groups of direct HIF-1 target genes that are par-
and this review will focus on the potential of HIF-1 as ticularly relevant to cancer encode angiogenic factors,
a therapeutic target. glucose transporters and glycolytic enzymes, survival
factors and invasion factors (FIG. 4). The products of
Master regulator of oxygen homeostasis the genes that HIF-1 regulates act at several steps in
The HIF-1 transcription factor mediates adaptive each of these processes, as shown by recent studies of
responses to changes in tissue oxygenation. HIF-1 is a cancer-cell invasion8,10 (FIG. 5).
heterodimer that consists of a constitutively expressed Expression of several HIF-1 target genes, such as
HIF-1β subunit and a HIF-1α subunit, the expression vascular endothelial growth factor (VEGF), is induced
of which is highly regulated. As for any protein, the level by hypoxia in most cell types. However, for the major-
of HIF-1α expression is determined by the rates of pro- ity of HIF-1 target genes, expression is induced by
tein synthesis and protein degradation. In the case of hypoxia in a cell-type-specific manner. HIF-1 activity
HIF-1α, synthesis is regulated via O2-independent is induced by hypoxia in almost all cell types and
mechanisms (FIG. 1), whereas degradation is regulated therefore HIF-1 alone cannot account for this cell-
primarily via O2-dependent mechanisms (FIG. 2), as type-specific gene expression. Rather, it is the func-
McKusick–Nathans Institute
of Genetic Medicine, described below. tional interaction of HIF-1 with other transcription
The Johns Hopkins More than 60 putative direct HIF-1 target genes factors that determines the subgroup of HIF-1 target
University School of have been identified (FIG. 3) on the basis of one or genes that is activated in any particular hypoxic cell.
Medicine, Baltimore, more of the following: identification of a cis-acting HIF-1 can be viewed as a messenger that is sent to the
Maryland
21287-3914, USA. hypoxia-response element that contains a HIF-1 nucleus to activate transcriptional responses to
e-mail: gsemenza@jhmi.edu binding site1; loss of hypoxia-induced gene expression hypoxia. The details of this response are determined by
doi:10.1038/nrc1187 in HIF1-α-null cells 5–7 or cells treated with small the past (developmental) and present (physiological)

NATURE REVIEWS | C ANCER VOLUME 3 | OCTOBER 2003 | 7 2 1


REVIEWS

Summary prolyl hydroxylases use molecular O2 and 2-oxoglutarate


(α-ketoglutarate) as substrates in a reaction that generates
• Hypoxia-inducible factor 1 (HIF-1) is a heterodimeric protein that consists of two prolyl-hydroxylated HIF-1α and succinate. Under physi-
proteins — HIF-1α and HIF-1β. HIF-1 activates the transcription of many genes that ological conditions, O2 is a limiting substrate19, therefore
code for proteins that are involved in angiogenesis, glucose metabolism, cell providing a mechanism for O2-dependent regulation of
proliferation/survival and invasion/metastasis. HIF-1α expression29. Acetylation of HIF-1α at lysine-532
• HIF-1α protein synthesis is regulated by activation of the phosphatidylinositol by the ARD1 acetyltransferase enhances the interaction
3-kinase (PI3K) and ERK mitogen-activated protein kinase (MAPK) pathways. These of VHL with HIF-1α, promoting its ubiquitylation
pathways can be activated by signalling via receptor tyrosine kinases, non-receptor and degradation30.
tyrosine kinases or G-protein-coupled receptors.
• HIF-1α protein degradation is regulated by O2-dependent prolyl hydroxylation,
which targets the protein for ubiquitylation by E3 ubiquitin-protein ligases. These Growth factor
ligases contain the von Hippel–Lindau tumour-suppressor protein (VHL), which
binds specifically to hydroxylated HIF-1α. Ubiquitylated HIF-1α is rapidly degraded Receptor tyrosine
kinase
by the proteasome.
• HIF-1α is overexpressed in human cancers as a result of intratumoral hypoxia as well
as genetic alterations, such as gain-of-function mutations in oncogenes (for example,
ERBB2) and loss-of-function mutations in tumour-suppressor genes (for example, PI3K RAS
VHL and PTEN). HIF-1α overexpression is associated with treatment failure and
increased mortality.
AKT PTEN MEK
• In xenograft assays, manipulation of HIF-1 activity by genetic or pharmacological
means has marked effects on tumour growth because of effects on angiogenesis,
glucose metabolism and/or cell survival.
mTOR ERK
• Screens are underway to identify small-molecule inhibitors of HIF-1 and to test
their efficacy as anticancer agents. These drugs might represent an important
component of novel combination therapies that are designed to target signalling
molecules in cancer cells. S6K MNK
4E-BPI

programming of each cell. As a result, the total number


of HIF-1 target genes cannot be ascertained by analysis S6 P
of one or a few cell types. Perhaps 1–5% of all human 40S
genes are expressed in response to hypoxia in one or 3′
60S eIF-4E
more cell types in a HIF-1-dependent manner.
Heterogeneity in target-gene expression is observed 5′
even among cell lines that have been derived from can-
HIF-1α protein
cers of the same histopathological type. Similar findings synthesis
have been reported for p53-dependent gene expres-
sion11. An additional complicating factor is the existence HIF-1β
of the related protein HIF-2α, which can also dimerize
with HIF-1β12–15. Heterodimers that contain HIF-1α or
HIF-2α seem to have overlapping but distinct specifici- HIF-1 transcriptional activity
ties, with regard to physiological inducers and target-
gene activation16. A third related protein, HIF-3α, might
function primarily as an inhibitor of HIF-1α17.
Figure 1 | Regulation of HIF-1α protein synthesis.
Growth-factor binding to a cognate receptor tyrosine kinase
Oxygen-dependent regulation of HIF-1 activates the phosphatidylinositol 3-kinase (PI3K) and
Cells transduce decreased O2 concentration into mitogen-activated protein kinase (MAPK) pathways. PI3K
increased HIF-1 activity via a novel O2-dependent post- activates the downstream serine/threonine kinases AKT (also
translational modification (FIG. 2). Three prolyl hydroxy- known as protein kinase B (PKB)) and mammalian target of
lases — known as prolyl hydroxylase-domain protein rapamycin (mTOR). In the MAPK pathway, the extracellular-
signal-regulated kinase (ERK) is activated by the upstream
(PHD) 1–3, or, alternatively, as HIF-1 prolyl hydroxylase MAP/ERK kinase (MEK). ERK, in turn, activates MNK. ERK
(HPH) 1–3 — modify proline(Pro)-402 and -564 of and mTOR phosphorylate p70 S6 kinase (S6K) — which, in
HIF-1α18,19. These proteins were originally designated turn, phosphorylates the ribosomal S6 protein — and the
EGLN1–3 on the basis of sequence homology to EGL9, eukaryotic translation initiation factor 4E (eIF-4E) binding
the HIF-1 prolyl hydroxylase in Caenorhabditis protein (4E-BP1). Binding of 4E-BP1 to eIF-4E inactivates the
elegans19. Hydroxylation of Pro-402 and Pro-564 is latter, inhibiting cap-dependent mRNA translation.
Phosphorylation of 4E-BP1 prevents its binding to eIF-4E.
required for interaction of HIF-1α with the VHL
MNK phosphorylates eIF-4E and stimulates its activity
tumour-suppressor protein20–23. VHL is the recognition directly. The effect of growth-factor signalling is an increase in
component of an E3 ubiquitin-protein ligase that the rate at which a subset of mRNAs within the cell (including
targets HIF-1α for proteasomal degradation24–28. The HIF-1α mRNA) are translated into protein.

722 | OCTOBER 2003 | VOLUME 3 www.nature.com/reviews/cancer


REVIEWS

Degradation Ubiquitylation VHL

OH OAc OH OH

Normoxic PHD1–3 ARD1 PHD1–3 FIH-1

P402 K532 P564 N803


bHLH PAS
TAD-N TAD-C

p300/CBP
Hypoxic

Transcriptional
activation

Figure 2 | O2-dependent regulation of HIF-1 activity. O2 regulates the rate at which HIF-1α protein is degraded. In
normoxic conditions, O2-dependent hydroxylation of proline (P) residues 402 and 564 in HIF-1α by the enzymes PHD (prolyl
hydroxylase-domain protein) 1–3 is required for the binding of the von Hippel–Lindau (VHL) tumour-suppressor protein, which
is the recognition component of an E3 ubiquitin-protein ligase. VHL binding is also promoted by acetylation of lysine (K)
residue 532 by the ARD1 acetyltransferase. Ubiquitylation of HIF-1α targets the protein for degradation by the 26S
proteasome. O2 also regulates the interaction of HIF-1α with transcriptional co-activators. O2-dependent hydroxylation of
asparagine (N) residue 803 in HIF-1α by the enzyme FIH-1 (factor inhibiting HIF-1) blocks the binding of p300 and CBP to
HIF-1α and therefore inhibits HIF-1-mediated gene transcription. Under hypoxic conditions, the rate of asparagine and
proline hydroxylation decreases. VHL cannot bind to HIF-1α that is not prolyl-hydroxylated, resulting in a decreased rate of
HIF-1α degradation. By contrast, p300 and CBP can bind to HIF-1α that is not asparaginyl-hydroxylated, allowing
transcriptional activation of HIF-1 target genes. bHLH, basic helix–loop–helix; PAS, Per-Arnt-Sim; TAD-C, carboxy-terminal
transactivation domain; TAD-N, amino-terminal transactivation domain.

HIF-1α transactivation-domain function is also O2- The increase in HIF-1α levels in response to
regulated31,32 via the action of FIH-1 (factor inhibiting growth-factor stimulation differs in two important
HIF-1)33. FIH-1 mediates this effect by hydroxylation of respects from the increase in HIF-1α levels in
asparagine(Asn)-803, which prevents the interaction of response to hypoxia. First, whereas hypoxia increases
HIF-1α with co-activators p300 and CBP34–36. HIF-1α levels in all cell types, growth-factor stimula-
Structural analysis of the HIF-1α domains that interact tion induces HIF-1α expression in a cell-type-specific
with VHL or p300/CBP has shown that hydroxylation manner. Second, whereas hypoxia is associated with
is a molecular switch that markedly alters the affinity of decreased degradation of HIF-1α, growth factors,
these interactions37–40. Hydroxylation is similar to other cytokines and other signalling molecules stimulate
post-translational modifications (for example, phos- HIF-1α synthesis via activation of the phosphatidyli-
phorylation) in that it functions to regulate nositol 3-kinase (PI3K) or mitogen-activated protein
protein–protein interactions, but, unlike other modifi- kinase (MAPK) pathways 44–49 (FIG. 1). Pulse-chase
cations, it is an inherently O2-regulated process. HIF-2α studies of MCF-7 breast cancer cells stimulated with
GLYCOLYTIC METABOLISM
expression and activity are also regulated by proline heregulin showed increased HIF-1α protein synthesis
Two molecules of ATP and and asparagine hydroxylation. that was inhibited by treatment with rapamycin — a
NADH are generated by the macrolide antibiotic that inhibits mammalian target of
conversion of one molecule of Oxygen-independent regulation of HIF-1 rapamycin (mTOR; a kinase that functions down-
glucose to two molecules of
Humans, like other metazoans, are constant and stream of PI3K and AKT)47. The effect of heregulin
pyruvate. The NADH is then
used to reduce pyruvate to obligate consumers of O2. The more cells that are pre- was mediated via the 5′-untranslated region of HIF-1α
lactate. sent in a tissue, the more O2 is consumed. So, when mRNA47. The known targets for phosphorylation by
one cell produces two daughter cells, O2 consump- mTOR are regulators of protein synthesis (FIG. 1).
OXIDATIVE METABOLISM tion increases. It is not surprising that the main path- However, it is not known whether phosphorylation of
Glucose is converted to
pyruvate, which is transported to
ways that transduce proliferative and survival signals these proteins by mTOR is necessary or sufficient for
the mitochondria, converted to from growth-factor receptors also induce HIF-1α increased HIF-1α synthesis. The translation of several
acetyl coenzyme A and oxidized expression (FIG. 1) in what can be viewed as a pre- dozen different mRNAs is known to be regulated by
to CO2 in the citric-acid cycle. emptive strategy for maintaining oxygen homeostasis. this pathway. Specific sequences in the 5′-untranslated
The NADH and FADH2
Proliferating cells express VEGF, which stimulates region could determine the degree to which the trans-
generated in this process provide
electrons to respiratory angiogenesis to provide the additional perfusion that lation of any particular mRNA can be modulated by
cytochromes and, ultimately, to is required to maintain oxygenation of an increased mTOR signalling. HIF-1α protein expression is likely
O2 in the inner mitochondrial number of cells. In addition, proliferating cells prefer- to be particularly sensitive to changes in the rate of
membrane, generating ATP. The entially use GLYCOLYTIC rather than OXIDATIVE METABOLISM synthesis because of its extremely short half-life under
complete oxidation of one
molecule of glucose results in
to generate ATP 41. The concomitant induction of non-hypoxic conditions. In addition to effects on
the production of 36 molecules angiogenesis and glycolysis with cell proliferation is HIF-1α synthesis, activation of the RAF–MEK–ERK
of ATP. mediated partly by activating HIF-1 (REFS 42,43). signalling pathway has also been shown to stimulate

NATURE REVIEWS | C ANCER VOLUME 3 | OCTOBER 2003 | 7 2 3


REVIEWS

Cell proliferation Transcriptional regulation all cell types, the regulation of HIF-1 activity by
Cyclin G2 DEC1 growth-factor signalling is cell-type specific. For exam-
IGF2 DEC2
IGF-BP1 ETS-1
ple, in MCF-7 cells, heregulin induces HIF-1α protein
IGF-BP2 NUR77 synthesis but does not induce transactivation-domain
IGF-BP3 function47, whereas treatment of PC-3 prostate cancer
WAF1 pH regulation
TGF-α Carbonic anhydrase 9 cells with rapamycin inhibits HIF-1α protein stability
TGF-β3 and transactivation-domain function53. Oncogenic
Cell survival Regulation of HIF-1 activity mutations that activate signal-transduction pathways
ADM p35srj
induce HIF-1 activity via various mechanisms (TABLE 1).
EPO
IGF2 Epithelial homeostasis
Loss-of-function mutations in tumour-suppressor
IGF-BP1 Intestinal trefoil factor genes are also associated with increased HIF-1 activity.
IGF-BP2
IGF-BP3 Among these, the most marked effect is observed in
Drug resistance
NOS2
MDR1 clear-cell renal carcinomas (RCCs) and cerebellar
TGF-α
VEGF haemangiomas that have lost VHL function26,54. VHL
Nucleotide metabolism loss of function results in a marked increase in HIF-1
Apoptosis Adenylate kinase 3
NIP3 Ecto-5′-nucleotidase activity in non-hypoxic conditions because of impaired
NIX VHL-dependent ubiquitylation and proteasomal degra-
RTP801 Iron metabolism dation of HIF-1α and HIF-2α. Although the
Ceruloplasmin
Motility Transferrin O2-dependent regulation of transactivation function is
AMF/GPI HIF-1 Transferrin receptor still intact, FIH-1 might become limiting under condi-
c-MET
LRP1 Glucose metabolism
tions of HIF-1α and HIF-2α overexpression, leading to
TGF-α HK1 increased transcriptionally active HIF-1 under non-
HK2 hypoxic conditions in VHL-null cells. For several other
Cytoskeletal structure AMF/GPI
KRT14 ENO1 oncogenes and tumour-suppressor genes, mutation not
KRT18 GLUT1
KRT19 only has a marked effect on cancer progression, but also
GAPDH
VIM LDHA on HIF-1 activity (TABLE 1). Several growth factors, most
PFKBF3 notably insulin-like growth factor-2 (IGF2) and trans-
Cell adhesion PFKL
MIC2 PGK1 forming growth factor-α (TGF-α), are also HIF-1 target
Erythropoiesis
PKM genes8,55. Binding of these factors to their cognate recep-
TPI
EPO tors — the insulin-like growth factor 1 receptor
Angiogenesis Extracellular-matrix metabolism (IGF1R) and epidermal growth-factor receptor (EGFR),
EG-VEGF CATHD respectively — activates signal-transduction pathways
ENG Collagen type V (α1)
FN1 that lead both to HIF-1α expression (as described
LEP
LRP1 MMP2 above) and to cell proliferation/survival. HIF-1 there-
TGF-β3 PAI1
Prolyl-4-hydroxylase α (I) fore contributes to autocrine-signalling pathways that
VEGF
UPAR are crucial for cancer progression (FIG. 6). The mecha-
Vascular tone nisms that lead to increased levels of HIF-1α have been
α1B-adrenergic receptor Energy metabolism
ADM LEP elucidated primarily through experiments in cancer cell
ET1 lines. This work has been complemented by immuno-
Haem oxygenase-1 Amino-acid metabolism
NOS2 Transglutaminase 2 histochemical demonstration of HIF-1α overexpression
in human cancer biopsies.
Figure 3 | Genes that are transcriptionally activated by HIF-1. Genes that are involved in
many processes are transcriptionally activated by HIF-1. ADM, adrenomedullin; ALDA, aldolase HIF-1α expression in human cancers
A; ALDC, aldolase C; AMF, autocrine motility factor; CATHD, cathepsin D; EG-VEGF, endocrine-
gland-derived VEGF; ENG, endoglin; ET1, endothelin-1; ENO1, enolase 1; EPO, erythropoietin;
A common approach for analysing altered expression
FN1, fibronectin 1; GLUT1, glucose transporter 1; GLUT3, glucose transporter 3; GAPDH, of proteins in human cancers is to perform immuno-
glyceraldehyde-3-P-dehydrogenase; HK1, hexokinase 1; HK2, hexokinase 2; IGF2, insulin-like histochemistry on patient biopsy samples. Expression
growth-factor 2; IGF-BP1, IGF-factor-binding-protein 1; IGF-BP2, IGF-factor-binding-protein 2; can be characterized qualitatively (present or absent) or
IGF-BP3, IGF-factor-binding-protein 3; KRT14, keratin 14; KRT18, keratin 18; KRT19, keratin 19; quantitatively (for example, percentage of cells express-
LDHA, lactate dehydrogenase A; LEP, leptin; LRP1, LDL-receptor-related protein 1; MDR1, ing the protein). Statistical analyses are performed to
multidrug resistance 1; MMP2, matrix metalloproteinase 2; NOS2, nitric oxide synthase 2;
determine whether expression of the protein is corre-
PFKBF3, 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase-3; PFKL, phosphofructokinase
L; PGK 1, phosphoglycerate kinase 1; PAI1, plasminogen-activator inhibitor 1; PKM, pyruvate lated with the expression of other biomarkers (which
kinase M; TGF-α, transforming growth factor-α; TGF-β3, transforming growth factor-β3; TPI, could be located either upstream or downstream in a
triosephosphate isomerase; VEGF, vascular endothelial growth factor; UPAR, urokinase common pathway) or with clinical outcome. These
plasminogen activator receptor; VEGFR2, VEGF receptor-2; VIM, vimentin. studies provide associations that can engender mecha-
nistic hypotheses for testing in tissue culture or animal
models. In addition, immunohistochemistry might be
HIF-1α transactivation-domain function50,51. This useful in clinical trials for identifying target populations
effect is due at least in part to phosphorylation by ERK and assessing therapeutic responses.
of the co-activator p300, with which the transactiva- Protein-expression levels in tumours are usually
tion domains interact52. Unlike hypoxia, which induces compared with those of the surrounding normal tissue.
HIF-1α protein stability and transcriptional activity in In general, proteins that promote tumour progression

724 | OCTOBER 2003 | VOLUME 3 www.nature.com/reviews/cancer


REVIEWS

Intratumoral hypoxia Genetic alterations among tumours from all patients, whereas in other can-
cer types the association was observed only in specific
subgroups, such as in early-stage cervical cancer. By con-
* trast, associations between HIF-1α overexpression and
decreased mortality were reported for patients with head
and neck cancer and non-small-cell lung cancer57,58,
although other studies failed to replicate these results59,60.
In a study of ovarian cancer, HIF-1α overexpression
* was correlated with apoptosis in most tumours and the
combination of apoptosis and HIF-1α overexpression
was associated with increased patient survival.
Increased HIF-1α expression and HIF-1 activity However, in ovarian cancers that overexpressed both
HIF-1α and p53 (mutant p53 has an increased half-
life), apoptosis levels were low and were associated with
a statistically significant decrease in overall patient sur-
Metabolic adaptation Apoptosis resistance Angiogenesis Invasion/metastasis vival61. In patients with early-stage oesophageal cancer,
ALDA, ENO1, GADPH, ADM, EPO, ET1, EG-VEGF, ENG, AMF, CATHD, CMET,
GLUT1, GLUT3, GPI, IGF2, NOS2, TGFA LEP, TGF-β3, FN1, KRT14, KRT18,
the combination of HIF-1α overexpression and BCL2
HK1, HK2, LDHA, VEGF, VEGFR2 KRT19, MMP2, UPAR, overexpression was associated with a failure to respond
PFKBF3, PFKL, PGK1, VIM to photodynamic therapy 62. So, the effect of HIF-1α
PGM, TPI
overexpression is dependent on the cancer type and the
Figure 4 | Mechanisms and consequences of HIF-1 activity in cancer cells. presence or absence of genetic alterations that affect the
Immunohistochemical analysis of HIF-1α levels in two separate oropharyngeal cancers. The balance between pro- and anti-apoptotic factors.
biopsy section on the left shows HIF-1α protein (brown staining) in viable cancer cells surrounding
Xenograft studies have provided additional evidence in
areas of necrosis (indicated by asterisks). The cancer cells that express the highest levels of HIF-1α
are at the greatest distance from a blood vessel (indicated by arrows) and are therefore the most support of these concepts.
hypoxic. In the biopsy section on the right, there are no areas of necrosis and HIF-1α is detected
in cancer cells throughout the field, including cells that are immediately adjacent to a blood vessel Manipulating HIF-1 activity in vivo
(arrows), indicating that increased HIF-1α levels are being driven by an O2-independent The most common experimental method for demon-
mechanism, such as through genetic alteration. These two mechanisms are not mutually strating that a particular gene product is involved in
exclusive — genetic alterations can amplify the response to hypoxia. In either case, increased
tumour growth is the xenograft assay, which involves sub-
HIF-1 activity leads to upregulation of genes that are involved in many aspects of cancer
progression, including metabolic adaptation, apoptosis resistance, angiogenesis and metastasis.
cutaneous injection of tumour cells into immunodefi-
See FIG. 3 for gene names. Photomicrographs reprinted with permission from REF. 81 © (2001) cient mice. To show the role of a gene product, the cells
American Association for Cancer Research. are transfected with expression vectors that encode a con-
stitutively active, dominant-negative or wild-type form of
the protein. To test for antitumour effects of a potential
(oncogene products) are usually found to be overex- therapeutic agent, the drug can be administered to mice
pressed, whereas proteins that inhibit progression at various times after injection of the tumour cells. All of
(tumour-suppressor gene products) are underexpressed, these approaches have been used to study HIF-1 (TABLE 3).
or carry missense mutations that result in the accumula- The first cell lines used to investigate the role of Hif-1
tion of a non-functional protein (for example, p53). were derived from the mouse hepatoma cell line
Although immunohistochemical analysis can be useful Hepa1. A Hif-1β-deficient subclone of Hepa1 — desig-
in determining whether a specific protein is present at nated c4 — and spontaneous revertants of c4 were
higher levels in cancer cells compared with surrounding analysed. Compared with wild-type Hepa1 cells, c4
normal tissue, it does not reveal whether the protein car- cells showed markedly reduced xenograft growth and
ries any mutations or has been post-translationally mod- angiogenesis and had increased areas of necrosis.
ified, which could affect its function. Furthermore, the Revertants, on the other hand, regained the ability to
effects of gain or loss of protein function can vary develop into tumours that were histologically similar to
according to the stage of cancer progression. For exam- wild-type Hepa1 cells43,63.
ple, oestrogen-receptor expression is increased in early- Hif1α –/– mouse embryonic stem (ES) cells have also
stage breast cancer and lost in advanced stages. These been analysed. ES cells are among the few non-
provisos notwithstanding, it is difficult to make the case transformed cells that can grow as xenografts. Two inde-
that a particular protein either promotes or inhibits can- pendently derived sets of Hif1α –/– ES cells were
cer progression without corroborating evidence from markedly impaired with respect to xenograft vascular-
human cancer biopsies. ization5,7. In one study, overall growth of xenografts that
Immunohistochemical analyses using monoclonal were derived from Hif1α –/– ES cells was reported to be
antibodies (FIG. 4) revealed that HIF-1α is overexpressed increased, because of reduced apoptosis5, whereas in
in many human cancers54,56. Significant associations another study xenograft growth was reported to be
between HIF-1α overexpression and patient mortality decreased because of reduced angiogenesis7. These data
have been shown in cancers of the brain (oligoden- imply that differences in the genetic background of the
droglioma), breast, cervix, oropharynx, ovary and uterus ES cells or of the xenograft recipients in the two studies
(endometrial) (TABLE 2). In some cancer types, such as had a key role in determining the net effect of HIF-1α
oropharyngeal cancer, the association was observed deficiency on xenograft growth.

NATURE REVIEWS | C ANCER VOLUME 3 | OCTOBER 2003 | 7 2 5


REVIEWS

a Normal epithelium To evaluate the effect of HIF-1α loss of function in


transformed cells, fibroblasts were isolated from Hif1α –/–
embryos, immortalized by SV40 T-ANTIGEN expression, and
transformed by mutant HRAS expression. In these cells,
Basement HIF-1α deficiency was associated with reduced
membrane tumour mass 16–18 days after injection, although
angiogenesis was not affected64. In tissue culture, gly-
colysis and ATP levels under hypoxic conditions were
markedly reduced42. In xenograft assays, the Hif1α –/–
ECM fibroblasts also manifested increased rates of apoptosis
in response to treatment with carboplatin and etopo-
side — agents that induce DNA double-strand breaks.
b ECM digestion
Cancer cell
Sensitivity to agents that induce single-strand breaks
or inhibit DNA synthesis, however, was unchanged65.
These data indicate that HIF-1 regulates the expression
Cathepsin D of one or more gene products that are required for the
uPAR repair of DNA double-strand breaks induced by
MMP2
chemotherapy or radiation.
HIF-1 activity has also been manipulated in
human cancer cell lines. Expression of VEGF,
xenograft growth and angiogenesis were markedly
increased in HCT116 colon cancer cells that were
Proteases
transfected with an expression vector that encoded
HIF-1α66. HIF-1α overexpression in PCI-10 pancre-
c New ECM production
atic cancer cells was associated with an increased
frequency of xenograft growth after subcutaneous
injection without affecting tumour-microvascular
Fibronectin 1 density 67. HIF-1α-overexpressing PCI-10 cells also
had an increased rate of survival when cultured
under conditions of glucose and oxygen deprivation.
As in the case of HRAS-transformed mouse embryo
fibroblasts described above, metabolic adaptation,
Fibronectin
rather than angiogenesis, seems to be an important
mechanism by which HIF-1α overexpression
promotes PCI-10 xenograft growth.
c Mesenchymal transformation Two strategies have been used to inhibit HIF-1
activity in human cancer cells. The first approach was
based on the demonstration that deletion of the
Keratin 14,
18,19 DNA-binding and transactivation domains results in
Vimentin a dominant-negative form of HIF-1α that can bind
AMF
TGF-α to HIF-1β, resulting in formation of an inactive het-
c-MET erodimer68. Overexpression of this dominant-
negative form of HIF-1α in PCI-43 pancreatic cancer
cells, which constitutively express high levels of HIF-1α,
Figure 5 | HIF-1 target genes that encode invasion
led to an increase in the number of cells undergoing
factors. Invasion of the basement membrane is the defining apoptosis under conditions of glucose and oxygen
characteristic of epithelial cancers. a | Epithelial cells are deprivation and a decreased ability to form tumours
normally constrained by cell–cell contacts and by the in severe combined immunodeficiency (SCID) mice.
basement membrane. b | Cancer cells produce proteases, Tumour vascularization, however, was not affected69.
including the urokinase-type plasminogen-activator receptor The second approach was based on the demonstra-
(uPAR) and matrix metalloproteinase-2 (MMP2), which digest
tion that HIF-1α contains two transactivation domains
the basement membrane/extracellular matrix (ECM).
c | Degraded ECM is replaced by fibronectin and other ECM — TAD-N and TAD-C31,32 (FIG. 2). TAD-C binding to its
proteins that are recognized by integrins that are expressed on co-activators, CBP and p300, is regulated by the
cancer cells. d | An epithelial-to-mesenchymal transformation O2-dependent hydroxylation of Asn-803 by FIH-1 (REFS
SV40 T-ANTIGEN
Large T-antigen — produced in
occurs in which intermediate-filament production is switched 33–36). A fusion protein that consists of GAL4 fused to
the early stage following
from keratin subtypes, which are characteristic of fixed TAD-C inhibits the ability of HIF-1α to interact with
epithelial cells, to keratins and vimentin, which promote the
infection of cells with simian these co-activators, and therefore blocks HIF-1-
virus 40 —promotes fluid structure that is required for motility and which is also
stimulated by expression of secreted factors such as autocrine
dependent transcription. Human breast (MDA-MB-435)
transformation by binding to
and inactivating the host p53 motility factor (AMF) and transforming growth factor-α and colon (HCT116) cancer cells infected with a retro-
and RB (retinoblastoma gene (TGF-α), and surface receptors such as the c-MET tyrosine virus that encodes this fusion protein showed reduced
product) proteins. kinase. HIF-1 target genes that regulate invasion are listed. growth when injected into nude mice70. However, the

726 | OCTOBER 2003 | VOLUME 3 www.nature.com/reviews/cancer


REVIEWS

IGF2 Renal carcinoma has been of particular interest to


Hypoxia
TGF-α investigators in the field, because the defining genetic
IGF1R EGFR lesion in RCC is loss of VHL function, resulting in high
levels of HIF-1α and HIF-2α protein that are not
O2 regulated. As a result, VEGF, GLUT1 and other HIF-1
target genes are constitutively expressed. In 786-O RCC
cells, only HIF-2α is expressed. When an expression vec-
tor encoding wild-type VHL was introduced into these
cells (which are designated 786-O/VHL), HIF-2α expres-
sion became O2 regulated and xenograft growth was
HIF-1 markedly reduced. To show that loss of O2-dependent
degradation of HIF-2α was necessary and sufficient
for the tumorigenic effect of VHL loss of function,
HIF1α 786-O/VHL cells were transfected with an expression
vector that encodes a form of HIF-2α containing a point
IGF2/TGF-α
mutation at the hydroxylation site (Pro531Ala). Ten
weeks after injection, these cells formed even larger
xenografts than 786-O cells, indicating that in this cell
line HIF-2α is a crucial downstream target of VHL71.
Similarly, overexpression in 786-O/VHL cells of a
fusion protein consisting of green fluorescent protein
Figure 6 | Involvement of HIF-1 in autocrine growth- (GFP) fused to the VHL-binding domain of HIF-2α —
factor stimulation of cancer cells. Binding of growth which blocks the interaction between HIF-2α and VHL
factors such as insulin-like growth factor-2 (IGF2) and
transforming growth factor-α (TGF-α) to their cognate
— leads to increased xenograft growth72. By contrast,
receptors — the IGF1 receptor (IGF1R) and the epidermal Vhl –/– ES cells manifest decreased xenograft growth in
growth-factor receptor (EGFR), respectively — stimulates nude mice73.
the expression of HIF-1α. This leads to increased HIF-1 Several important conclusions can be drawn from
transcriptional activity of target genes, which include those these studies. First, in all of the studies in which bona fide
that encode IGF2 and TGF-α. Alternatively, hypoxia can cancer cells were tested, increased levels of HIF-1α or
activate HIF-1, via increased expression of HIF-1α, and
HIF-2α were associated with increased tumour-
initiate autocrine signalling.
xenograft growth, whereas inhibition of HIF-1 activity
markedly impaired tumour growth. However, the spe-
cific consequences of increased HIF-1 activity differed
fusion protein also disrupts the ability of many other according to cell type. In pancreatic cancer cells, both
transcription factors to interact with CBP/p300, and gain- and loss-of-function experiments highlighted the
therefore limits the conclusions that can be drawn from role of HIF-1 activity in regulating glucose metabolism
these studies. and cell survival, whereas in colon cancer cells a correla-
tion between HIF-1α expression, angiogenesis and
tumour growth was observed that was not seen in pan-
Table 1 | Genetic alterations that increase HIF-1 activity creatic cancer cells. Second, HIF-1 activity was inversely
associated with tumour growth only in studies involving
Alteration in tumour Mechanism of HIF-1α induction References
ES cells that lacked the large complement of genetic
VHL loss of function Decreased ubiquitylation 26 alterations that are characteristic of cancer cells. These
p53 loss of function Decreased ubiquitylation 66 results emphasize that the consequences of HIF-1α over-
PTEN loss of function Increased synthesis 48,49 expression are dependent on the cellular context and
PI3K–AKT–mTOR signalling* Increased synthesis 47,48 therefore highlight the importance of using appropriate
MEK–ERK signalling* Increased synthesis 44
models to study the roles of HIF-1 in cancer biology. It
should be noted that all of the studies reported so far
ERBB2 gain of function Increased synthesis 47
involve injection of cells into the subcutaneous space and
EGFR signalling* Increased synthesis 48 therefore fail to replicate the stromal microenvironment
IGF1R signalling* Increased synthesis 44 in which cancer cells normally develop. The observation
PGE2 signalling* Increased synthesis 45,98 that altering HIF-1 activity did not affect the vasculariza-
SRC gain of function Increased synthesis 43 tion of pancreatic cancer xenografts might reflect the fact
ARF loss of function Decreased nucleolar sequestration 99 that those cells were growing in the subcutaneous space
rather than in the pancreas.
BCL2 overexpression Not determined 100
*Increased signalling could be due to genetic alteration in a component of the pathway or an
upstream activator. For example, AKT gain-of-function mutation or PTEN loss-of-function mutation
HIF-1 and clonal selection
induces PI3K–AKT–mTOR signalling; EGFR amplification or TGF-α overexpression induces EGFR Cancer progression involves the selection of cells that
(and PI3K–AKT–mTOR and MEK–ERK) signalling. EGFR, epidermal growth factor receptor; ERK, bear mutations that increase their rate of cell prolifer-
extracellular-signal-regulated kinase; IGFR1, insulin-like growth-factor-1 receptor; MEK, MAP/ERK
kinase; mTOR, mammalian target of rapamycin; PGE2, prostaglandin E2; PI3K, phosphatidylinositol ation and survival. These effects can be mediated by
3-kinase; VHL, von Hippel–Lindau protein. mutations that either directly target components of

NATURE REVIEWS | C ANCER VOLUME 3 | OCTOBER 2003 | 7 2 7


REVIEWS

Table 2 | Increased HIF-1α levels in human cancers* HIF-1 can induce apoptosis — for example, through
the stabilization of p53 (REF. 75) or through transacti-
Tumour type Association References
vation of BNIP3, which encodes a pro-apoptotic
Cervical, early stage Increased mortality 101
BCL2 family member76. In such cells, complementary
Cervical, RTX Increased mortality 102 mutations that inactivate p53 or activate BCL2
Lung, NSCLC Decreased mortality 58 expression could be necessary for the net effect of
Lung, NSCLC Increased mortality (HIF-2α) 59 increased HIF-1 activity to promote cancer-cell sur-
Breast, LN-positive Increased mortality 103 vival. In some cancers (for example, ovarian),
Breast, LN-negative Increased mortality 104
hypoxia-induced HIF-1-mediated apoptosis might
represent an important factor in the positive selection
Oligodendroglioma Increased mortality 105
of p53-null and BCL2-overexpressing cells77. In other
Oropharyngeal SCC Increased mortality, radiation resistance 81 cases, HIF-1α expression could be involved in the
Ovarian Increased mortality (with p53) 61 very early stages of cancer progression, as in the case
Oesophageal, early stage Resistance to PDT (with BCL2) 62 of VHL-null RCC (see later). HIF-1α levels are
Endometrial Increased mortality 106 increased in ductal carcinoma in situ — the pre-
Head and neck SCC, S/P surgery Decreased mortality 57
invasive stage of breast cancer — and are associated
with increased microvascular density 78, indicating
Head and neck SCC Increased mortality (HIF-2α) 60
that HIF-1α expression might have an important role
GI stromal tumour of stomach Increased mortality 107 early in breast cancer progression.
*Protein levels determined by immunohistochemical analysis of biopsy samples. GI, gastrointestinal; At each point in the temporal–spatial progression of
LN, lymph node; NSCLC, non-small-cell lung cancer; PDT, photodynamic therapy; RTX, radiation
therapy; SCC, squamous-cell carcinoma; S/P, status post. a cancer, there is a constant selection for cells bearing
particular (and changing) combinations of genetic and
epigenetic alterations that impart to those cells the
the proliferative/apoptotic machinery or indirectly greatest relative probability of survival and prolifera-
affect these processes; for example, by promoting tion. The ability to identify genetic alterations that are
angiogenesis. Many of these mutations also contribute selected during tumour progression provides a rational
to the invasive and metastatic properties of cancer approach to therapeutic target selection. However, in
cells74. Genetic alterations that activate oncogenes and addition to ‘primary’ targets of clonal selection (onco-
inactivate tumour-suppressor genes also result in genes and tumour-suppressor genes), candidates for
increased HIF-1α expression that might promote the therapeutic target selection should also include ‘sec-
survival of cells that bear these mutations and so con- ondary’ targets, the expression or activity of which are
tribute to their selection. As observed in xenograft affected as a result of genetic alterations involving
assays, the consequences of increased HIF-1 activity several ‘primary’ targets. HIF-1 represents a prime
are cell-type specific. In some cell types, increased example of such a ‘secondary’ target, whereas VEGF

Table 3 | Effects of altered HIF-1 activity on tumour growth*


Cell type Manipulation Rate of Tumour Ex vivo‡ References
tumour growth angiogenesis
Decreased HIF-1 activity §
Hepa1 mouse hepatoma Hif-1β LOF – – – 43,63
Mouse ES Hif-1α KO – – – 7
Mouse ES Hif-1α KO + – + 5
MEF (T-Ag, HRASV12) Hif-1α KO – 0 – 42,64
HCT116, MDA-MB-435 HIF1α TAD – – – 70
PCI-43 (pancreatic CA) HIF-1α DN – 0 – 69
MEF (T-Ag, HRASV12) Hif-1α KO, chemo – ND – 65
5 human cancer cell lines YC-1 i.p. – – ND 83
Increased HIF-1 Activity §
HCT116 (colon CA) HIF-1α GOF + + ND 66
PCI-10 (pancreatic CA) HIF-1α GOF + 0 + 67
786-O/VHL HIF-2α (P531A) + + ND 71
786-O/VHL HIF-1α ODD + ND ND 72
Mouse ES Vhl KO – + ND 73
*All effects monitored by xenograft growth in nude mice. ‡Growth or survival of cells cultured under conditions of hypoxia or O2/glucose
deprivation ex vivo. §In this context, HIF-1 refers to both HIF-1α–HIF-1β and HIF-2α–HIF-1β heterodimers. Chemo, chemotherapy; DN,
dominant negative; ES, embryonic stem; GOF, gain of function; i.p., intraperitoneal injection; KO, knockout (homozygous for null allele);
LOF, loss of function; MEF, mouse embryonic fibroblast; ND, not determined; ODD, O2-dependent degradation domain; TAD,
transactivation domain; T-Ag, large T antigen; –, decreased; +, increased; 0, no effect.

728 | OCTOBER 2003 | VOLUME 3 www.nature.com/reviews/cancer


REVIEWS

Table 4 | Novel therapeutic agents that inhibit HIF-1 activity university, government and industry laboratories and
several agents that inhibit HIF-1, angiogenesis and
Agent(s) Molecular target(s) Current status
xenograft growth have been identified (TABLE 4). A
Inhibitors of signal-transduction pathways
preliminary screen of the NCI DIVERSITY SET of small-
BAY 43-9006 RAF kinase Clinical trials molecule chemotherapeutic agents using a cell-based
CCI-779 mTOR Clinical trials assay revealed that TOPOISOMERASE I INHIBITORS block HIF-1α
Celebrex COX2 Clinical trials expression via an undetermined mechanism82. The
PD98059 MEK Not in clinical use small molecule YC-1 (3-(5′-hydroxy-methyl-2′-furyl)-
Trastuzumab (Herceptin) ERBB2 receptor tyrosine kinase Approved agent
1-benzylindazole) was also shown to reduce HIF-1α
levels and xenograft growth83. YC-1 is known to stimu-
ZD-1839 (Iressa), OSI-774 EGFR tyrosine kinase Clinical trials
late soluble guanylate-cyclase activity, but this effect is
Imatinib (Glivec) BCR–ABL, PDGFR tyrosine kinases Approved agent not required for inhibition of HIF-1α levels. The mech-
Small-molecule inhibitors of HIF-1 activity anism by which YC-1 reduces HIF-1α levels has not
2ME2 Microtubule polymerization Clinical trials been established. HIF-1α interacts with the chaperone
17-AAG HSP90 Clinical trials HSP90, and the HSP90 inhibitor 17-allyl-aminogel-
Camptothecin, Topotecan Topoisomerase I Approved agents
danamycin (17-AAG) induces HIF-1α degradation in
a VHL-independent manner 84– 86, indicating that
Pleurotin, 1-methylpropyl Thioredoxin 1 Not in clinical use
2-imidazolyl disulphide HSP90 is required for HIF-1α stability. The redox reg-
ulator thioredoxin-1 also exerts a positive effect on
YC-1 Not determined Not in clinical use
HIF-1α expression. Thioredoxin inhibitors block
BCR–ABL, breakpoint-cluster-region–Abelson-leukaemia; COX2, cyclooxygenase 2; EGFR,
epidermal growth-factor receptor; HSP90, heat-shock protein 90; MEK, MAP/ERK kinase; mTOR, HIF-1α expression and xenograft growth87. Finally,
mammalian target of rapamycin; PDGFR, platelet-derived growth-factor receptor. disruption of microtubule polymerization by
2-methoxyoestradiol (2ME2) has also been shown
to result in decreased HIF-1α levels and decreased
and its receptors represent examples of a ‘tertiary’ tar- VEGF mRNA expression in cultured cells88. In vivo,
get. The hypothesis that increased HIF-1 activity con- 2ME2 decreases tumour-xenograft growth and
tributes to clonal selection and cancer progression is vascularization. YC-1, topoisomerase I inhibitors,
supported by clinical and experimental data. The asso- 17-AAG, thioredoxin inhibitors and 2ME2 share in
ciation between increased HIF-1α expression and common an ability to decrease HIF-1α levels,
increased patient mortality establishes a necessary clini- inhibit the expression of VEGF and other HIF-1 tar-
cal correlation, whereas experimental manipulation of get genes, and impair xenograft growth and vascu-
HIF-1 activity in tumour xenografts provides evidence larization. Therefore, the anticancer effects of these
of causation, as described above. agents might be due, in part, to their inhibition of
HIF-1. On the other hand, it seems that none of
HIF-1 targeted therapeutics these drugs specifically target HIF-1. The lack of
Drugs that specifically target tumour stromal-cell selectivity increases the difficulty in correlating mol-
responses represent an important new class of thera- ecular and clinical responses in patients, but it does
peutic agents. In particular, a large number of drugs not disqualify these drugs as potential anticancer
are in clinical trials at present as anticancer agents agents. Ongoing screens should lead to the identifi-
based on their ability to inhibit angiogenesis79. One cation of more selective HIF-1 inhibitors in the near
concern about this strategy is that inhibition of future. Assuming that this prediction holds, in what
angiogenesis might select for cancer cells that are clinical contexts are HIF-1 inhibitors most likely to
adapted to hypoxic conditions, as these are the cells have therapeutic efficacy?
that are most likely to survive a reduction in perfu-
sion. A large body of data has indicated that hypoxic Candidate patient populations
cancer cells are more likely to be resistant to radiation RCC that is associated with VHL mutations is a good
and chemotherapy, and have increased potential for candidate for HIF-1 targeted therapy. Whereas several
invasion, metastasis and patient mortality 80. Recent RCC cell lines (for example, 786-O) express HIF-2α
studies have provided evidence indicating that HIF-1 but not HIF-1α, immunohistochemical analyses have
mediates resistance to chemotherapy and radia- revealed that overexpression is seen in most primary
NCI DIVERSITY SET tion65,81. Inhibition of HIF-1 activity could therefore cancers54,56. All experimental data indicate that HIF-1 is
A group of approximately 2,000 represent an important component of combination involved in the pathophysiology of this cancer, and the
compounds that is representative anti-angiogenesis therapies. In addition to drugs that upregulation of HIF-1 activity seems to occur in the
of the complete chemical
have been developed specifically as anti-angiogenesis earliest detected neoplastic lesions89. HIF-1 target genes
repository of the National
Cancer Institute’s Developmental agents, it is clear that many novel therapeutic agents that are activated in RCC include platelet-derived
Therapeutics Program. that target signal-transduction pathways have anti- growth factor-β (PDGFB)and TGF-α. These genes
angiogenic effects. This effect seems to be due in part encode proteins that activate the PDGF and EGF recep-
TOPOISOMERASE I INHIBITORS to the fact that inhibition of signal-transduction tors — two of the main receptor tyrosine kinases that
Drugs that inhibit an enzyme
that relaxes supercoiled DNA by
pathways results in decreased levels of HIF-1α. signal via the PI3K and MAPK pathways to stimulate
introducing a transient single- Several examples are provided in TABLE 4. Screens for cell proliferation and survival. In addition, HIF-1
strand break. small-molecule inhibitors of HIF-1 are underway in upregulates VEGF expression, leading to the abundant

NATURE REVIEWS | C ANCER VOLUME 3 | OCTOBER 2003 | 7 2 9


REVIEWS

vascularization that is a characteristic of RCCs. So, second key limitation of the xenograft model relates to
increased HIF-1 activity in RCC has direct and marked the crucial importance of the interactions that take place
effects on both cancer and stromal cells. between cancer cells and the stromal microenvironment.
Administration of a HIF-1 inhibitor, together with a The factors required for angiogenesis, invasion and
tyrosine kinase inhibitor that targets VEGF/PDGF metastasis in one tissue differ significantly from those
and/or EGFR, is one potential therapeutic regimen in required in another site. The subcutaneous space into
this disease. which xenografts are transplanted does not provide the
Another candidate is glioblastoma multiforme same microenvironment as in common sites of human
(GBM). Like RCC, GBM is an intractable disease, and cancer such as the breast, colon and lung. Also, most
patients typically survive for less than 1 year. Among patients die of metastatic disease, which cannot be mod-
gliomas, there is strong correlation between HIF-1α elled in encapsulated subcutaneous xenografts. Finally,
expression, tumour grade and tumour vasculariza- xenograft assays are performed in immunodeficient
tion90. PTEN loss of function and EGFR gain of func- (nude or SCID) mice, so the inflammatory cell popula-
tion are commonly observed in primary GBM, and tions that are involved in human cancer progression are
are known to increase HIF-1α levels48,49. GBMs are not represented in this model.
highly vascularized tumours that are characterized by An incremental but significant advance that is rele-
areas of necrosis surrounded by PSEUDOPALISADING CELLS vant to the issue of the tumour microenvironment is
that express high levels of HIF-1α protein and VEGF the use of orthotopic transplantation. Glioblastoma
mRNA. So, in GBM, hypoxia-induced HIF-1α cells injected into the rodent brain form tumours that
expression seems to also be involved in the disease share remarkable histological similarity to invasive
pathophysiology. GBM is the most highly invasive human glioblastomas90. Certain human breast cancer
cancer known and high levels of HIF-1α expression cell lines, when injected into the mammary fat pad,
have been detected at the leading edge of invading show a high incidence of metastasis91. These models
cancer cells in both human biopsies and following provide an opportunity to study important aspects of
ORTHOTOPIC TRANSPLANTATION of labelled glioma cells into cancer biology that cannot be addressed in subcuta-
the mouse brain 90. In patients with GBM, a HIF-1 neous xenografts, such as further analysis of the role
inhibitor might be effective in combination with a of HIF-1 in invasion and metastasis.
rapamycin derivative (that targets increased More elegant models of cancer have been developed
PI3K–AKT–mTOR signalling induced by PTEN loss by engineering loss-of-function mutations in tumour-
of function) anti-angiogenic agents and/or tyrosine suppressor genes or gain-of-function mutations in
kinase inhibitors. oncogenes into the mouse genome. This approach can
A number of other tumour types have also been model hereditary tumour-predisposition syndromes
found to upregulate HIF-1, as determined by immuno- in which a heterozygous mutation in a tumour-
histochemical studies (TABLE 2). In patients with suppressor gene such as TP53 is present in the
oropharyngeal cancers, HIF-1α overexpression has germline. However, for most oncogenes, mutations
been associated with radiation resistance and increased have not been detected in the human germline, proba-
mortality, regardless of tumour grade, stage or other bly because they would result in developmental defects.
biomarkers81. Importantly, HIF-1α overexpression in Novel approaches, such as the use of retroviral infec-
the primary tumour was shown to predict radiation tion of the target tissue to induce oncogene expression,
resistance in lymph-node metastases. These tumours might provide better mouse models of cancer92,93. A key
might therefore be susceptible to HIF-1-targeted thera- advantage of these models is that they reproduce the
pies. The combination of HIF-1α overexpression and heterogeneity of cancer progression. A drawback (espe-
(mutant) p53 overexpression also seems to define a cially as models for drug testing) is that they result in
subset of ovarian cancers that might be susceptible to only a limited number of animals that bear tumours of
treatment with HIF-1 inhibitors61. a particular stage at any given time. The development
of in vivo tumour-imaging techniques94 could provide a
Better predictors of clinical efficacy solution to this problem as tumour-bearing mice can
The tumour-xenograft model represents the current be identified by periodic screening and then started on
standard for preclinical testing of anticancer agents. This therapy. Testing of potential therapeutic agents (target-
PSEUDOPALISADING CELLS
Rows of viable cells surrounding
model, however, has too many limitations — only a few ing HIF-1 or any other molecule) for efficacy in at least
areas of necrosis that are a of which will be discussed here — to remain an accept- one such model should be considered as a necessary
histopathological characteristic able gateway to clinical trials. First, human cancers prerequisite to clinical trials.
of glioblastoma multiforme. develop by a stepwise process of mutation and selection. Novel imaging techniques also provide a means to
Although clonal selection occurs, cancers are character- monitor the response to therapy in patients 95–97.
ORTHOTOPIC
TRANSPLANTATION
ized by a tremendous degree of genetic heterogeneity. In vivo imaging can be used to monitor tumour-cell
The introduction of foreign Among the millions of cells in a tumour, there are likely apoptosis and proliferation, angiogenesis, blood flow,
tumour cells into another to exist subpopulations that are resistant to any single oxygenation and glucose metabolism. Developments
species at the site from which therapeutic agent that might be administered and, for this in magnetic resonance imaging and positron emis-
they were derived. For example,
injection of human breast cancer
reason, multidrug regimens are likely to be necessary to sion tomography should provide the means to moni-
cells into the mouse mammary successfully treat advanced cancers. This degree of genetic tor any of these variables in patients in the near
fat pad. heterogeneity does not exist in a tumour xenograft. A future. Imaging can also be used to monitor the

730 | OCTOBER 2003 | VOLUME 3 www.nature.com/reviews/cancer


REVIEWS

response of a drug target to therapy. For example, outcome. This will be particularly crucial for
imaging techniques could be used to show inhibition of inhibitors that target kinases or transcription factors
receptor-tyrosine-kinase activity or gene transcription, such as HIF-1 that regulate many key physiological
and to correlate this molecular effect with patient pathways in cancer cells.

1. Semenza, G. L. & Wang, G. L. A nuclear factor induced by 25. Kamura, T. et al. Activation of HIF1α ubiquitination by a mechanism for HIF-1-mediated vascular endothelial
hypoxia via de novo protein synthesis binds to the human reconstituted von Hippel-Lindau (VHL) tumor suppressor growth factor expression. Mol. Cell. Biol. 21, 3995–4004
erythropoietin gene enhancer at a site required for complex. Proc. Natl Acad. Sci. USA 97, 10430–10435 (2001).
transcriptional activation. Mol. Cell. Biol. 12, 5447–5454 (2000). 48. Zhong, H. et al. Modulation of HIF-1α expression by the
(1992). 26. Maxwell, P. H. et al. The tumour suppressor protein VHL epidermal growth factor/phosphatidylinositol
2. Wang, G. L. & Semenza, G. L. Purification and targets hypoxia-inducible factors for oxygen-dependent 3-kinase/PTEN/AKT/FRAP pathway in human prostate
characterization of hypoxia-inducible factor 1. J. Biol. Chem. proteolysis. Nature 399, 271–275 (1999). cancer cells: implications for tumor angiogenesis and
270, 1230–1237 (1995). 27. Ohh, M. et al. Ubiquitination of hypoxia-inducible factor therapeutics. Cancer Res. 60, 1541–1545 (2000).
3. Wang, G. L., Jiang, B.-H., Rue, E. A. & Semenza, G. L. requires direct binding to the β-domain of the von Hippel- 49. Zundel, W. et al. Loss of PTEN facilitates HIF-1-mediated
Hypoxia-inducible factor 1 is a basic-helix-loop-helix-PAS Lindau protein. Nature Cell Biol. 2, 423–427 (2000). gene expression. Genes Dev. 14, 391–396 (2000).
heterodimer regulated by cellular O2 tension. Proc. Natl 28. Tanimoto, K., Makino, Y., Pereira, T. & Poellinger, L. 50. Richard, D. E. et al. p42/p44 mitogen-activated protein
Acad. Sci. USA 92, 5510–5514 (1995). Mechanism of regulation of the hypoxia-inducible factor-1α kinases phosphorylate hypoxia-inducible factor 1α (HIF-1α)
4. Harris, A. L. Hypoxia — a key regulatory factor in tumor by the von Hippel-Lindau tumor protein. EMBO J. 19, and enhance the transcriptional activity of HIF-1. J. Biol.
growth. Nature Rev. Cancer 2, 38–46 (2001). 4298–4309 (2000). Chem. 274, 32631–32637 (1999).
5. Carmeliet, P. et al. Role of HIF-1α in hypoxia-mediated 29. Jiang, B.-H., Semenza, G. L., Bauer, C. & Marti, H. H. 51. Sodhi, A. et al. The Kaposi’s sarcoma-associated herpes
apoptosis, cell proliferation and tumour angiogenesis. Hypoxia-inducible factor 1 levels vary exponentially over a virus G protein-coupled receptor up-regulates vascular
Nature 394, 485–490 (1998). physiologically relevant range of O2 tension. Am. J. Physiol. endothelial growth factor expression and secretion through
6. Iyer, N. V. et al. Cellular and developmental control of O2 271, C1172–C1180 (1996). mitogen-activated protein kinase and p38 pathways acting
homeostasis by hypoxia-inducible factor 1α. Genes Dev. 30. Jeong, J. W. et al. Regulation and destabilization of HIF-1α on hypoxia-inducible factor 1α. Cancer Res. 60, 4873–4880
12, 149–162 (1998). by ARD1-mediated acetylation. Cell 111, 709–720 (2002). (2000).
7. Ryan, H. E., Lo, J. & Johnson, R. S. HIF-1α is required for 31. Jiang, B.-H. et al. Transactivation and inhibitory domains of 52. Sang, N. et al. MAPK signaling up-regulates the activity of
solid tumor formation and embryonic vascularization. hypoxia-inducible factor 1α: modulation of transcriptional hypoxia-inducible factors by its effects on p300. J. Biol.
EMBO J. 17, 3005–3015 (1998). activity by oxygen tension. J Biol. Chem. 272, 19253–19260 Chem. 278, 14013–14019 (2003).
8. Krishnamachary, B. et al. Regulation of colon carcinoma cell (1997). 53. Hudson, C. C. et al. Regulation of hypoxia-inducible
invasion by hypoxia-inducible factor 1. Cancer Res. 63, 32. Pugh, C. W. et al. Activation of hypoxia-inducible factor-1; factor 1alpha expression and function by the mammalian
1138–1143 (2003). definition of regulatory domains within the α subunit. J Biol. target of rapamycin. Mol. Cell. Biol. 22, 7004–7014
9. Wykoff, C. C. et al. Identification of novel hypoxia Chem. 272, 11205–11214 (1997). (2002).
dependent and independent target genes of the von Hippel- 33. Mahon, P. C., Hirota, K. & Semenza, G. L. FIH-1: a novel 54. Zhong, H. et al. Overexpression of hypoxia-inducible factor
Lindau (VHL) tumour suppressor by mRNA differential protein that interacts with HIF-1α and VHL to mediate 1α in common human cancers and their metastases.
expression profiling. Oncogene 19, 6297–6305 (2000). repression of HIF-1 transcriptional activity. Genes Dev. 15, Cancer Res. 59, 5830–5835 (1999).
10. Pennacchietti, S. et al. Hypoxia promotes invasive growth by 2675–2686 (2001). 55. Feldser, D. et al. Reciprocal positive regulation of hypoxia-
transcriptional activation of the met protooncogene. Cancer 34. Hewitson, K. S. et al. Hypoxia-inducible factor (HIF) inducible factor 1α and insulin-like growth factor 2. Cancer
Cell 3, 347–361 (2003). asparagine hydroxylase is identical to factor inhibiting HIF Res. 59, 3915–3918 (1999).
11. Yu, J. et al. Identification and classification of p53-regulated (FIH) and is related to the cupin structural family. J. Biol. 56. Talks, K. L. et al. The expression and distribution of the
genes. Proc. Natl Acad. Sci. USA 96, 14517–14522 (1999). Chem. 277, 26351–26355 (2002). hypoxia-inducible factors HIF-1α and HIF-2α in normal
12. Ema, M. et al. A novel bHLH-PAS factor with close 35. Lando, D. et al. Asparagine hydroxylation of the HIF human tissues, cancers, and tumor-associated
sequence similarity to hypoxia-inducible factor 1α regulates transactivation domain a hypoxic switch. Science 295, macrophages. Am. J. Pathol. 157, 411–421 (2000).
the VEGF expression and is potentially involved in lung and 858–861 (2002). 57. Beasley, N. J. et al. Hypoxia-inducible factors HIF-1α and
vascular development. Proc. Natl Acad. Sci. USA 94, 36. Lando, D. et al. FIH-1 is an asparaginyl hydroxylase enzyme HIF-2α in head and neck cancer: relationship to tumor
4273–4278 (1997). that regulates the transcriptional activity of hypoxia-inducible biology and treatment outcome in surgically resected
13. Flamme, I. et al. HRF, a putative basic helix-loop-helix-PAS- factor. Genes Dev. 16, 1466–1471 (2002). patients. Cancer Res. 62, 2493–2497 (2002).
domain transcription factor is closely related to hypoxia- 37. Dames, S. A. et al. Structural basis for HIF-1α/CBP 58. Volm, M. & Koomagi, R. Hypoxia-inducible factor (HIF-1)
inducible factor-1α and developmentally expressed in blood recognition in the cellular hypoxic response. Proc. Natl and its relationship to apoptosis and proliferation in lung
vessels. Mech. Dev. 63, 51–60 (1997). Acad. Sci. USA 99, 5271–5276 (2002). cancer. Anticancer Res. 20, 1527–1533 (2000).
14. Hogenesch, J. B. et al. Characterization of a subset of the 38. Freedman, S. J. et al. Structural basis for recruitment of 59. Giatromanolaki, A. et al. Relation of hypoxia inducible factor
basic-helix-loop-helix-PAS superfamily that interacts with CBP/p300 by hypoxia-inducible factor-1α. Proc. Natl Acad. 1α and 2α in operable non-small cell lung cancer to
components of the dioxin signaling pathway. J. Biol. Chem. Sci. USA 99, 5367–5372 (2002). angiogenic/molecular profile of tumours and survival. Br. J.
272, 8581–8593 (1997). 39. Hon, W. C. et al. Structural basis for the recognition of Cancer 85, 881–890 (2001).
15. Tian, H., McKnight, S. L. & Russell, D. W. Endothelial PAS hydroxyproline in HIF-1α by pVHL. Nature 417, 975–978 60. Koukourakis, M. I. et al. Hypoxia-inducible factor (HIF1A and
domain protein 1 (EPAS1), a transcription factor selectively (2002). HIF2A), angiogenesis, and chemoradiotherapy outcome of
expressed in endothelial cells. Genes Dev. 11, 72–82 (1997). 40. Min, J. H. et al. Structure of an HIF-1α-pVHL complex: squamous cell head-and-neck cancer. Int. J. Radiat. Oncol.
16. Brusselmans, K. et al. Hypoxia-inducible factor-2α (HIF- hydroxyproline recognition in signaling. Science 296, Biol. Phys. 53, 1192–1202 (2002).
2α) is involved in the apoptotic response to hypoglycemia 1886–1889 (2002). 61. Birner, P. et al. Expression of hypoxia-inducible factor 1α in
but not to hypoxia. J. Biol. Chem. 276, 39192–39196 41. Brand, K. A. & Hermfisse, U. Aerobic glycolysis by epithelial ovarian tumors: its impact on prognosis and on
(2001). proliferating cells: a protective strategy against reactive response to chemotherapy. Clin. Cancer Res. 7, 1661–1668
17. Makino, Y. et al. Inhibitory PAS domain protein (IPAS) is a oxygen species. FASEB J. 11, 388–395 (1997). (2001).
hypoxia-inducible splicing variant of thehypoxia-inducible 42. Seagroves, T. N. et al. Transcription factor HIF-1 is a 62. Koukourakis, M. I. et al. Hypoxia inducible factor (HIF-1α
factor-3α locus. J. Biol. Chem. 277, 32405–32408 (2002). necessary mediator of the pasteur effect in mammalian and HIF-2α) expression in early esophageal cancer and
18. Bruick, R. K. & McKnight, S. L. A conserved family of prolyl- cells. Mol. Cell. Biol. 21, 3436–3444 (2001), response to photodynamic therapy and radiotherapy.
4-hydroxylases that modify HIF. Science 294, 1337–1340 43. Jiang, B. H., Agani, F., Passaniti, A. & Semenza, G. L. Cancer Res. 61, 1830–1832 (2001).
(2001). V-SRC induces expression of hypoxia-inducible factor 1 63. Maxwell, P. H. et al. Hypoxia-inducible factor-1 modulates
19. Epstein, A. C. et al. C. elegans EGL-9 and mammalian (HIF-1) and transcription of genes encoding vascular gene expression in solid tumors and influences both
homologs define a family of dioxygenases that regulate HIF endothelial growth factor and enolase 1: involvement of HIF-1 angiogenesis and tumor growth. Proc. Natl Acad. Sci. USA
by prolyl hydroxylation. Cell 107, 43–54 (2001). in tumor progression. Cancer Res. 57, 5328–5335 (1997). 94, 8104–8109 (1997).
20. Ivan, M. et al. HIFα targeted for VHL-mediated destruction 44. Fukuda, R. et al. Insulin-like growth factor 1 induces 64. Ryan, H. E. et al. Hypoxia-inducible factor-1α is a positive
by proline hydroxylation: implications for O2 sensing. hypoxia-inducible factor 1-mediated vascular endothelial factor in solid tumor growth. Cancer Res. 60, 4010–4015
Science 292, 464–468 (2001). growth factor expression, which is dependent on MAP (2000).
21. Jaakkola, P. et al. Targeting of HIF-α to the von Hippel- kinase and phosphatidylinositol 3-kinase signaling in 65. Unruh, A. et al. The hypoxia-inducible factor-1α is a
Lindau ubiquitylation complex by O2-regulated prolyl colon cancer cells. J. Biol. Chem. 277, 38205–38211 negative factor for tumor therapy. Oncogene 22,
hydroxylation. Science 292, 468–472 (2001). (2002). 3213–3220 (2003).
22. Masson, N. et al. Independent function of two destruction 45. Fukuda, R., Kelly, B. & Semenza, G. L. Vascular endothelial 66. Ravi, R. et al. Regulation of tumor angiogenesis by p53-
domains in hypoxia-inducible factor-alpha chains activated growth factor gene expression in colon cancer cells induced degradation of hypoxia-inducible factor 1α. Genes
by prolyl hydroxylation. EMBO J. 20, 5197–5206 (2001). exposed to prostaglandin E2 is mediated by hypoxia- Dev. 14, 34–44 (2000).
23. Yu, F., White, S. B., Zhao, Q. & Lee, F. S. HIF-1α binding to inducible factor 1. Cancer Res. 63, 2330–2334 (2003). 67. Akakura, N. et al. Constitutive expression of hypoxia-
VHL is regulated by stimulus-sensitive proline hydroxylation. 46. Hellwig-Burgel, T., Stiehl, D. P. & Jelkmann, W. in Oxygen inducible factor 1α renders pancreatic cancer cells resistant
Proc. Natl Acad. Sci. USA 98, 9630–9635 (2001). Sensing: Responses and Adaptation to Hypoxia (eds Lahiri, to apoptosis induced by hypoxia and nutrient deprivation.
24. Cockman, M. E. et al. Hypoxia inducible factor-α binding S., Semenza, G. L. & Prabhakar, N. R.) 95–108 (Marcel Cancer Res. 61, 6548–6554 (2001).
and ubiquitylation by the von Hippel-Lindau tumor Dekker, Inc., New York, 2003). 68. Jiang, B. H. et al. Dimerization, DNA binding, and
suppressor protein. J. Biol. Chem. 275, 25733–25741 47. Laughner, E. et al. HER2 (neu) signaling increases the rate transactivation properties of hypoxia-inducible factor 1.
(2000). of hypoxia-inducible factor 1α (HIF-1α) synthesis: novel J. Biol. Chem. 271, 17771–17778 (1996).

NATURE REVIEWS | C ANCER VOLUME 3 | OCTOBER 2003 | 7 3 1


REVIEWS

69. Chen, J. et al. Dominant-negative hypoxia-inducible factor pathway in prostate cancer cells. Cancer Res. 62, 99. Fatyol, K. & Szalay, A. A. The p14ARF tumor suppressor
1α reduces tumorigenicity of pancreatic cancer cells 2478–2482 (2002). protein facilitates nucleolar sequestration of HIF-1α and
through the suppression of glucose metabolism. Am. 86. Zagzag, D. et al. Geldanamycin inhibits migration of glioma inhibits HIF-1 mediated transcription. J. Biol. Chem. 276,
J. Pathol. 162, 1283–1291 (2003). cells in vitro: a potential role for hypoxia-inducible factor 28421–28429 (2001).
70. Kung, A. L. et al. Suppression of tumor growth through (HIF-1α) in glioma cell invasion. J. Cell. Physiol. 196, 100. Iervolino, A. et al. Bcl-2 overexpression in human melanoma
disruption of hypoxia-inducible transcription. Nature Med. 6, 394–402 (2003). cells increases angiogenesis through VEGF mRNA
1335–1340 (2000). 87. Welsh, S. J. et al. The thioredoxin redox inhibitors stabilization and HIF-1-mediated transcriptional activity.
71. Kondo, K. et al. Inhibition of HIF is necessary for tumor 1-methylpropyl 2-imidazolyl disulfide and pleurotin inhibit FASEB J. 16, 1453–1455.
suppression by the von Hippel-Lindau protein. Cancer Cell hypoxia-induced factor 1α and vascular endothelial 101. Birner, P. et al. Overexpression of hypoxia-inducible factor 1α
1, 237–246 (2002). growth factor formation. Mol. Cancer Ther. 2, 235–243 is a marker for an unfavorable prognosis in early-stage
72. Maranchie, J. K. et al. The contribution of VHL substrate (2003). invasive cervical cancer. Cancer Res. 60, 4693–4696 (2000).
binding and HIF-1α to the phenotype of VHL loss in renal 88. Mabjeesh, N. J. et al. 2ME2 inhibits tumor growth and 102. Burri, P. et al. Significant correlation of hypoxia-inducible
cell carcinoma. Cancer Cell 1, 247–255 (2002). angiogenesis by disrupting microtubules and dysregulating factor-1α with treatment outcome in cervical cancer treated
73. Mack, F. A. et al. Loss of pVHL is sufficient to cause HIF HIF. Cancer Cell 3, 363–375 (2003). with radical radiotherapy. Int. J. Radiat. Oncol. Biol. Phys.
dysregulation in primary cells but does not promote tumor 89. Mandriota, S. J. et al. HIF activation identifies early lesions 56, 494–501 (2003).
growth. Cancer Cell 3, 75–88 (2003). in VHL kidneys: evidence for site-specific tumor 103. Schindl, M. et al. Overexpression of hypoxia-inducible factor
74. Bernards, R. & Weinberg, R. A. Metastasis genes: a suppressor function in the nephron. Cancer Cell 1, 1α is associated with an unfavorable prognosis in lymph
progression puzzle. Nature 418, 823 (2002). 459–468 (2002). node-positive breast cancer. Clin. Cancer Res. 8,
75. An, W. G. et al. Stabilization of wild-type p53 by hypoxia- 90. Zagzag, D. et al. Expression of hypoxia-inducible factor 1α 1831–1837 (2002).
inducible factor 1α. Nature 392, 405–408 (1998). in brain tumors: association with angiogenesis, invasion, and 104. Bos, R. et al. Levels of hypoxia-inducible factor-1α
76. Bruick, R. K. Expression of the gene encoding the progression. Cancer 88, 2606–2618 (2000). independently predict prognosis in patients with lymph node
proapoptotic Nip3 protein is induced by hypoxia. Proc. Natl 91. Price, J. E., Polyzos, A., Zhang, R. D. & Daniels, L. M. negative breast carcinoma. Cancer 97, 1573–1581 (2003).
Acad. Sci. USA 97, 9082–9087 (2000). Tumorigenicity and metastasis of human breast 105. Birner, P. et al. Expression of hypoxia-inducible factor-1α in
77. Graeber, T. G. et al. Hypoxia-mediated selection of cells with carcinoma cell lines in nude mice. Cancer Res. 50, oligodendrogliomas: its impact on prognosis and on
diminished apoptotic potential in solid tumours. Nature 379, 717–721 (1990). neoangiogenesis. Cancer 92, 165–171 (2001).
88–91 (1996). 92. Holland, E. C. Gliomagenesis: genetic alterations and 106. Sivridis, E. et al. Association of hypoxia-inducible factors 1α
78. Bos, R. et al. Levels of hypoxia-inducible factor 1α during mouse models. Nature Rev. Genet. 2, 120–129 and 2α with activated angiogenic pathways and prognosis
breast carcinogenesis. J. Natl Cancer Inst. 93, 309–314 (2001). in patients with endometrial carcinoma. Cancer 95,
(2001). 93. Van Dyke, T. & Jacks, T. Cancer modeling in the modern 1055–1063 (2002).
79. Semenza, G. L. Angiogenesis in ischemic and neoplastic era: progress and challenges. Cell 108, 135–144 107. Takahashi, R. et al. Hypoxia-inducible factor-1α expression
disorders. Annu. Rev. Med. 54, 17–28 (2003). (2002). and angiogenesis in gastrointestinal stromal tumor of the
80. Hockel, M. & Vaupel, P. Tumor hypoxia: definitions and 94. Pomper, M. G. Can small animal imaging accelerate drug stomach. Oncol. Rep. 10, 797–802 (2003).
current clinical, biologic, and molecular aspects. J. Natl development? J. Cell. Biochem. 39 (Suppl.), 211–220
Cancer Inst. 93, 266–276 (2001). (2002).
81. Aebersold, D. M. et al. Expression of hypoxia-inducible 95. Artemov, D., Mori, N., Ravi, R. & Bhujwalla, Z. M. Magnetic Online links
factor 1α: a novel predictive and prognostic parameter in the resonance molecular imaging of the her-2/neu receptor.
radiotherapy of oropharyngeal cancer. Cancer Res. 61, Cancer Res. 63, 2723–2727 (2003). DATABASES
2911–2916 (2001). 96. Bhujwalla, Z. M. et al. Reduction of vascular and permeable The following terms in this article are linked online to:
82. Rapisarda, A. et al. Identification of small molecule inhibitors regions in solid tumors detected by macromolecular Cancer.gov: http://cancer.gov/
of hypoxia-inducible factor 1 transcriptional activation contrast magnetic resonance imaging after treatment with brain cancer | breast cancer | cervical cancer | endometrial cancer |
pathway. Cancer Res. 62, 4316–4324 (2002). antiangiogenic agent TNP-470. Clin. Cancer Res. 9, head and neck cancer | oesophageal cancer | oropharyngeal
83. Yeo, E. J. et al. YC-1: a potential anticancer drug targeting 355–362 (2003). cancer | ovarian cancer | pancreatic cancer | prostate cancer
hypoxia-inducible factor 1. J. Natl Cancer Inst. 95, 516–525 97. Mankoff, D. A. et al. Blood flow and metabolism in locally LocusLink: http://www.ncbi.nlm.nih.gov/LocusLink/
(2003). advanced breast cancer: relationship to response to AKT | ARD1 | BCL2 | BNIP3 | CBP | FIH-1 | GLUT1 | HIF-1α |
84. Isaacs, J. S. et al. Hsp90 regulates a von Hippel-Lindau- therapy. J. Nucl. Med. 43, 500–509 (2002). HIF-2α | HIF-3α | HSP90 | IGF2 | mTOR | p53 | p300 | PDGF-β |
independent hypoxia-inducible factor-1α-degradative 98. Liu, X. H. et al. Prostaglandin E2 induces hypoxia-inducible PI3K | RAF | TGF-α | VEGF | VHL |
pathway. J. Biol. Chem. 277, 29936–29944 (2002). factor-1α stabilization and nuclear localization in a human OMIM: http://www.ncbi.nlm.nih.gov/omim/
85. Mabjeesh, N. J. et al. Geldanamycin induces degradation of prostate cancer cell line. J. Biol. Chem. 277, 50081–50086 glioblastoma multiforme
hypoxia-inducible factor 1α protein via the proteosome (2002). Access to this interactive links box is free online.

732 | OCTOBER 2003 | VOLUME 3 www.nature.com/reviews/cancer

You might also like