throbber
Critical Reviews in Biochemistry and Molecular Biology
`
`ISSN: 1040-9238 (Print) 1549-7798 (Online) Journal homepage: http://www.tandfonline.com/loi/ibmg20
`
`Hypoxia, Clonal Selection, and the Role of HIF-1 in
`Tumor Progression
`
`Gregg L. Semenza
`
`To cite this article: Gregg L. Semenza (2000) Hypoxia, Clonal Selection, and the Role of HIF-1 in
`Tumor Progression, Critical Reviews in Biochemistry and Molecular Biology, 35:2, 71-103
`
`To link to this article: http://dx.doi.org/10.1080/10409230091169186
`
`Published online: 29 Sep 2008.
`
`Submit your article to this journal
`
`Article views: 795
`
`View related articles
`
`Citing articles: 5 View citing articles
`
`Full Terms & Conditions of access and use can be found at
`http://www.tandfonline.com/action/journalInformation?journalCode=ibmg20
`
`Roxane Labs., Inc.
`Exhibit 1017
`Page 001
`
`

`
`Critical Reviews in Biochemistry and Molecular Biology, 35(2):71–103 (2000)
`
`Hypoxia, Clonal Selection, and the Role of
`HIF-1 in Tumor Progression
`
`Gregg L. Semenza
`
`Institute of Genetic Medicine, Departments of Pediatrics and Medicine, The Johns
`Hopkins University School of Medicine, Baltimore, Maryland 21287
`
`* Address for correspondence: Gregg L. Semenza, M.D., Ph.D. Johns Hopkins Hospital,
`CMSC-1004, 600 North Wolfe Street, Baltimore, MD 21287–3914, TEL: 410-955-
`1619 FAX: 410-955-0484, E-mail: gsemenza@jhmi.edu
`
`Referee: Peter M. Glazer, M.D., Ph. D., Yale University School of Medicine
`
`Abstract: Tumor progression occurs as a result of the clonal selection of cells in which
`somatic mutations have activated oncogenes or inactivated tumor suppressor genes leading
`to increased proliferation and/or survival within the hypoxic tumor microenvironment.
`Hypoxia-inducible factor 1 (HIF-1) is a transcription factor that mediates adaptive re-
`sponses to reduced O2 availability, including angiogenesis and glycolysis. Expression of the
`O2-regulated HIF-1α subunit and HIF-1 transcriptional activity are increased dramatically
`in hypoxic cells. Recent studies indicate that many common tumor-specific genetic alter-
`ations also lead to increased HIF-1α expression and/or activity. Thus, genetic and physi-
`ologic alterations within tumors act synergistically to increase HIF-1 transcriptional activ-
`ity, which appears to play a critical role in the development of invasive and metastatic
`properties that define the lethal cancer phenotype.
`
`Key Words: angiogenesis, cancer, glycolysis p53; Ras, vascular endothelial growth factor.
`
`I. INTRODUCTION: GENETIC
`ALTERATIONS IN TUMOR CELLS
`
`A major focus of oncology research over
`the last quarter-century spanning the era of
`molecular biology has been the identifica-
`tion of genetic alterations arising as a result
`of somatic mutation within tumor cells. A
`large body of research data established the
`existence of several groups of genes that,
`when mutated, contribute to the process of
`tumor progression (for review see Vogelstein
`
`1040-9238/00/$.50
`© 2000 by CRC Press LLC
`
`and Kinzler, 1998). The first group that was
`identified consists of oncogenes, in which
`gain-of-function mutations result in in-
`creased expression of a gene product and/or
`expression of a mutant product with
`constitutive activity that is not responsive
`to normal molecular mechanisms of
` downregulation. The second group consists
`of tumor suppressor genes, in which loss-
`of-function mutations eliminate the expres-
`sion or activity of a gene product. A third
`group consists of mutator genes involved in
`DNA repair, in which loss-of-function mu-
`
`71
`
`
`
`Roxane Labs., Inc.
`Exhibit 1017
`Page 002
`
`

`
`tations increase the rate at which mutations
`occur within oncogenes and tumor suppres-
`sor genes.
`The high frequency of mutations involv-
`ing specific genes within specific tumor cell
`types implies that these genetic alterations
`play a key role in the oncogenic transforma-
`tion of that cell type. Tumor-specific muta-
`tions do not result from targeted mutagen-
`esis, but instead represent the effect of
`selection: cells in which a mutation occurs
`by chance within a specific gene that results
`in an increased rate of survival and/or divi-
`sion (that together determine the rate of cell
`proliferation) become overrepresented in the
`tumor relative to cells lacking the mutation.
`Along with the discovery of oncogenes,
`tumor suppressor genes, and mutator genes,
`the concept of clonal selection (Nowell,
`1976) represented a critical advance in our
`understanding of tumor biology.
`Because it is much easier to manipulate
`tumor cells in plastic dishes than in living
`animals, a great deal of research investigat-
`ing the biological function of oncogenes
`and tumor suppressor genes has been per-
`formed in cultured cells. In this context,
`studying the rate at which cells divide is
`accomplished easily. For this reason, the
`functions attributed to the protein products
`of oncogenes and tumor suppressor genes
`have primarily related to the control of
`cellular proliferation in tissue culture:
`oncogenes promote cell proliferation
`whereas tumor suppressor genes counteract
`this effect. Likewise, the influence of anti-
`and proapoptotic proteins, such as members
`of the BCL2 family, on tumor cell survival
`has been analyzed primarily in the tissue
`culture milieu.
`The goal of this review is to focus on
`adaptive mechanisms by which selected
`tumor cells are able to survive the harsh
`microenvironmental conditions that they
`impose on themselves in vivo. The adapta-
`tions that are described have in common the
`
`72
`
`fact that they are mediated, at least in part,
`by the transcriptional activator hypoxia-in-
`ducible factor 1 (HIF-1). Important proper-
`ties of HIF-1 and common characteristics of
`solid tumors are reviewed first, followed by
`a summary of recent data that have delin-
`eated specific functional relationships be-
`tween genetic alterations, HIF-1, and tumor
`progression. Finally, the clinical implica-
`tions of these results are considered.
`
`II. HIF-1 STRUCTURE AND
`FUNCTION
`
`HIF-1 is a heterodimer composed of HIF-
`1α and HIF-1β subunits (Wang and Semenza,
`1995). Both subunits contain basic-helix-
`loop-helix (bHLH)-PAS domains (Wang et
`al., 1995a). Whereas the bHLH domain de-
`fines a superfamily of eukaryotic transcrip-
`tion factors (reviewed by Semenza, 1998),
`the PAS domain, which was first identified
`in the PER, ARNT, and SIM proteins, de-
`fines a subfamily of bHLH proteins that is
`unique to metazoans (reviewed by Crews
`and Fan, 1999). The HIF-1β subunit is also
`known as the aryl hydrocarbon nuclear
`translocator (ARNT), as it was first shown to
`dimerize with the aryl hydrocarbon receptor
`(Hoffman et al., 1991). Recently, proteins
`with sequence similarity to HIF-1α (HIF-2α
`and HIF-3α) and HIF-1β (ARNT2 and
`ARNT3) have been identified, and all HIF-
`1α-like polypeptides appear able to dimerize
`with all HIF-1β-like polypeptides, at least in
`vitro (reviewed by Semenza, 2000). In knock-
`out mice, homozygosity for a targeted loss-
`o
`f
`-
`f
`u
`n
`c
`t
`i
`o
`n
`mutation in the Hif1a, Epas1, or Arnt gene
`encoding HIF-1α, HIF-2α, or HIF-1β, re-
`spectively, results in embryonic lethality (Iyer
`et al., 1998; Kozak et al., 1998; Maltepe et
`al., 1997; Tian et al., 1998). The develop-
`mental and physiological roles of these fac-
`
`
`
`Roxane Labs., Inc.
`Exhibit 1017
`Page 003
`
`

`
`tors have been reviewed elsewhere recently
`(Crews and Fan, 1999; Semenza, 1999, 2000).
`The HLH and PAS domains of HIF-1α
`and HIF-1β are required for the formation
`of a dimer in which conformation of the
`basic domains allows their recognition of
`specific DNA binding sites containing the
`core sequence 5’-RCGTG-3’ (Jiang et al.,
`1996a; Semenza et al., 1996; Wood et al.,
`1996). HIF-1 functions as a sequence-spe-
`cific transcriptional activator (Forsythe
`et al., 1996; Jiang et al., 1996a; Semenza et
`al. 1996). Both the expression of HIF-1α
`and its ability to activate transcription
`are regulated by the cellular O2 concentra-
`tion (Huang et al., 1996; Jiang et al., 1996b,
`1997b; Pugh et al., 1997; Wang et al.,
`1995a). HIF-1α has a half-life of < 5
`min under posthypoxic conditions, both in
`cultured cells and in vivo (Wang et al., 1995a;
`Yu et al., 1998). Under nonhypoxic condi-
`tions, HIF-1α is subject to ubiquitination
`and proteasomal degradation (Huang et al.,
`1998; Kallio et al., 1999; Salceda and
`Caro, 1997). Under hypoxic conditions,
`ubiquitination of HIF-1α is dramatically re-
`duced (Sutter et al., 2000). The introduction
`of specific missense mutations and dele-
`tions into the HIF-1a sequence results in
`decreased ubiquitination and increased
`expression
`of HIF-1a
`protein
`under nonhypoxic conditions (Sutter et al.,
`2000). Transactivation domain function is
`also under negative regulation in nonhypoxic
`cells and deletions within these domains
`also increase their activity in nonhypoxic
`cells (Jiang et al., 1997b; Pugh et al., 1997).
`Thus, under hypoxic conditions, HIF-1 tran-
`scriptional activity increases rapidly due to
`synergistic effects on HIF-1a protein ex-
`pression and transactivation domain func-
`tion. The HIF-1 transactivation domains
`have been shown to interact with the
`coactivator proteins CBP, p300, SRC-1, and
`TIF2, and these interactions are promoted
`by the redox regulatory proteins thioredoxin
`
`and Ref-1 (Arany et al., 1997; Carrero et al.,
`2000; Ema et al., 1999).
`To date, approximately 30 target genes
`that are transactivated by HIF-1 have been
`identified (Table 1 and data not shown).
`These genes encode proteins that are re-
`quired for angiogenesis, regulation of blood
`vessel tone, and vascular remodeling; cell
`proliferation and viability; erythropoiesis and
`iron metabolism; glucose transport and gly-
`colysis. Remarkably, the vast majority of
`these gene products have been shown to be
`overexpressed in human tumor cells. Each
`HIF-1 target gene contains a hypoxia re-
`sponse element, a cis-acting transcriptional
`regulatory sequence. The presence of a HIF-
`1 binding site is necessary but not sufficient
`to constitute a functional hypoxia response
`element (Forsythe et al., 1996; Semenza et
`al., 1996; Semenza and Wang, 1992).
`
`III. COMMON BIOCHEMICAL
`AND PHYSIOLOGICAL
`CHARACTERISTICS OF SOLID
`TUMORS
`
`A. The Warburg Effect
`
`Seventy years ago, Warburg demon-
`strated that compared with normal cells, tu-
`mor cells are characterized by a marked
`increase in glycolytic metabolism, even
`when cultured in the presence of high O2
`concentrations (Warburg, 1930). Glucose
`transport into tumor cells is also markedly
`increased in order to provide increased
`amounts of the substrate that is ultimately
`converted to lactate by the glycolytic en-
`zymes. Indeed, increased uptake of labeled
`glucose derivatives is utilized in clinical
`diagnostic tests to identify occult tumors in
`patients. Furthermore, glycolytic metabo-
`lism is correlated with disease progression,
`
`73
`
`
`
`Roxane Labs., Inc.
`Exhibit 1017
`Page 004
`
`

`
`Table 1.
`Known HIF-1 Target Genes
`Gene product
`Adenylate kinase 3
`α1B-Adrenergic receptor
`Adrenomedullin
`Aldolase A
`
`Aldolase C
`Endothelin-1 (ET-1)
`Enolase 1 (ENO1)
`Erythropoietin (EPO)
`Glucose transporter 1
`
`Glucose transporter 3
`Glyceraldehyde phosphate dehydrogenase
`Heme oxygenase 1
`Hexokinase 1
`Hexokinase 2
`Insulin-like growth factor 2 (IGF-2)
`IGF binding protein 1
`IGF factor binding protein 3
`Lactate dehydrogenase A
`
`Nitric oxide synthase 2 (NOS2)
`p21
`p35srj
`Phosphofructokinase L
`Phosphoglycerate kinase 1
`
`Plasminogen activator inhibitor-1
`Pyruvate kinase M
`Transferrin
`Transferrin receptor
`Vascular endothelial growth factor (VEGF)
`
`VEGF receptor FLT-1
`
`74
`
`Ref.
`
`Wood et al., 1998
`Eckhart et al., 1997
`Cormier-Regard et al., 1998
`Iyer et al., 1998; Ryan et al., 1998;
`Semenza et al.,1996
`Iyer et al., 1998
`Hu et al., 1998
`Semenza et al., 1996; Iyer et al., 1998
`Semenza and Wang, 1992; Jiang et al.,1996a
`Iyer et al., 1998; Ryan et al., 1998;
`Wood et al., 1998
`Iyer et al., 1998
`Iyer et al., 1998; Ryan et al., 1998
`Lee et al., 1997
`Iyer et al., 1998
`Iyer et al., 1998
`Feldser et al., 1999
`Tazuke et al., 1998
`Feldser et al., 1999
`Iyer et al., 1998; Ryan et al., 1998;
`Semenza et al., 1996
`Melillo et al., 1995; Palmer et al., 1998
`Carmeliet et al., 1998
`Bhattacharya et al., 1999
`Iyer et al., 1998
`Carmeliet et al., 1998; Iyer et al., 1998; Ryan et al.,
`1998
`Kietzmann et al., 1999
`Iyer et al., 1998
`Rolfs et al., 1997
`Lok and Ponka, 1999; Tacchini et al., 1999
`Carmeliet et al., 1998; Forsythe et al., 1996;
`Iyer et al., 1998; Ryan et al., 1998
`Gerber et al., 1997
`
`
`
`Roxane Labs., Inc.
`Exhibit 1017
`Page 005
`
`

`
`because lactate concentrations of > 20 µmol/
`g are 6.5-fold more common in primary
`cervical cancers with metastatic spread com-
`pared with non-metastasizing primary tu-
`mors (Schwickert et al., 1995). A correla-
`tion between lactate production and
`metastasis was also reported for head and
`neck tumors (Walenta et al., 1997). In addi-
`tion to aerobic glycolysis (the Warburg ef-
`fect), the existence of intratumoral hypoxia
`provides an additional stimulus for glycoly-
`sis, as described below.
`Even though the Warburg effect is one
`of the most universal characteristics of solid
`tumors, only recently has insight been gained
`into the molecular mechanisms by which
`energy metabolism is shifted from oxida-
`tive to glycolytic pathways. Quiescent thy-
`mocytes cultured in the absence of mito-
`gens have been shown to derive > 80% of
`their ATP from oxidative phosphorylation,
`whereas mitogen-stimulated thymocytes
`derive > 80% of their ATP from glycolysis
`(Brand and Hermfisse, 1997). It has been
`proposed that the switch from oxidative to
`glycolytic metabolism occurs in order to
`reduce the generation of reactive oxygen
`species that might otherwise damage repli-
`cating DNA. If this switch is hard-wired
`into the program of cellular proliferation,
`then it is not surprising that virtually all
`tumor cells would manifest the Warburg
`effect despite their thorough disregard for
`genomic integrity.
`Hypoxia response elements have been
`identified in the promoters of genes encod-
`ing aldolase A (ALDA), enolase 1 (ENO1),
`glucose transporter 3 (GLUT3), lactate de-
`hydrogenase A (LDHA), phosphoglycerate
`kinase 1, and phosphofructokinase L (Ebert
`et al., 1995; Firth et al., 1994, 1995; Semenza
`et al., 1994, 1996) and HIF-1-dependent
`transactivation of the ALDA and ENO1 pro-
`moters has been demonstrated (Semenza et
`al., 1996). A 68-bp fragment of the ENO1
`promoter, which is the most powerful hy-
`
`poxia response element identified to date,
`contains three HIF-1 binding sites (Semenza
`et al., 1996). Analysis of mouse embryonic
`stem cells that are homozygous for a tar-
`geted loss-of-function mutation in the Hif1a
`gene encoding HIF-1α revealed decreased
`expression of 13 different genes encoding
`glucose transporters and glycolytic enzymes
`(Iyer et al., 1998; Ryan et al., 1998). Thus,
`HIF-1 mediates coordinate transcriptional
`activation of the entire pathway from glu-
`cose uptake to lactate production. Further-
`more, expression of HIF-1α is induced when
`tumor cells are cultured under conditions
`associated with cellular proliferation such
`as low cell density and growth factor stimu-
`lation (Feldser et al., 1999; Zhong et al.,
`1998; Zhong et al., 2000). C-MYC, a tran-
`scription factor whose primary role is to
`stimulate cellular proliferation, also
`transactivates several genes encoding gly-
`colytic enzymes, including LDHA (Dang
`and Semenza, 1999; Shim et al., 1997). HIF-
`1 (consensus binding site sequence, 5’-
`RCGTG-3’) and C-MYC (consensus
`binding site sequence, 5’-CACGTG-3’) rec-
`ognize overlapping binding sites in the
`LDHA gene (Dang and Semenza, 1999;
`Semenza et al., 1996; Shim et al., 1997).
`
`B. Angiogenesis
`
`A large body of experimental evidence
`indicates that primary tumors and metastases
`will not grow beyond a volume of several
`mm3 without establishing a blood supply
`(reviewed in Fidler and Ellis, 1994;
`Folkman, 1971, 1990, 1992; Hanahan and
`Folkman, 1996; Zetter, 1998). In the ab-
`sence of vascularization, cell viability is lim-
`ited by the diffusion of O2 and glucose from
`host vessels. Under these circumstances, cell
`division and cell death occur at equal rates
`and no net tumor growth occurs. In
`
`75
`
`
`
`Roxane Labs., Inc.
`Exhibit 1017
`Page 006
`
`

`
`RIP1Tag2 transgenic mice, every pancre-
`atic-islet β cell expresses SV40 T antigen,
`approximately 50% of the islets demonstrate
`β cell hyperplasia, but only a
`subset of hyperplastic islets demonstrate
`neovascularization; an even smaller subset
`of islets progress to carcinoma, but, nota-
`bly, all of these manifest neovascularization
`(Folkman et al., 1989). The onset of angio-
`genesis permits rapid tumor growth, thus
`setting the stage for subsequent develop-
`ment of the invasive and metastatic proper-
`ties that define the lethal cancer phenotype.
`Indeed, for many tumor types there is a
`statistically significant correlation between
`tumor angiogenesis (as measured by blood
`vessel density) and patient survival (re-
`viewed in Fox, 1997; Jekunen and Kairemo,
`1997; Zetter, 1998).
`Tumor angiogenesis occurs as a result
`of increased expression of angiogenic
`factors and decreased expression of
`antiangiogenic factors. The angiogenic fac-
`tor that has been shown to have the most
`important role in mediating blood vessel
`formation in response to developmental,
`physiological, or oncogenic stimuli is vas-
`cular endothelial growth factor (VEGF) (re-
`viewed in Ferrara and Davis-Smyth, 1997).
`There is a strong correlation between VEGF
`expression and blood vessel density and
`clinical outcome in many tumor types
`(Ferrara and Davis-Smyth, 1997). Inhibi-
`tion of VEGF expression or binding to its
`receptor on endothelial cells has a dramatic
`effect on tumor growth, invasion, and
`metastasis in animal models (Benjamin et
`al., 1999; Benjamin and Keshet, 1997;
`Borgstrom et al., 1996; Cheng et al., 1996;
`Grunstein et al., 1999; Im et al., 1999; Jain
`et al., 1998; Kim et al., 1993; Millauer et al.,
`1994, 1996; Saleh et al., 1996; Shaheen et
`al., 1999; Skobe et al., 1997; Yuan et al.,
`1996). Somatic mutations resulting in
`oncogene activation (e.g., H-Ras and v-Src)
`and tumor suppressor gene inactivation (e.g.,
`
`76
`
`p53 and VHL) are associated with increased
`VEGF gene expression (reviewed in Ferrara
`and Davis-Smyth, 1997). In addition, in sev-
`eral tumor types, most notably glioblastoma
`multiforme (Shweiki et al., 1992), VEGF
`expression is correlated with intratumoral
`hypoxia (Ferrara and Davis-Smyth, 1997).
`The role of HIF-1 in activating hypoxia-
`induced transcription of the VEGF gene is
`well established. A hypoxia response ele-
`ment is located approximately 1 kb 5’ to the
`transcription start site of the human VEGF
`gene (Forsythe et al., 1996; Liu et al., 1995)
`and is also present in the mouse and rat Vegf
`genes (Levy et al., 1995; Shima et al., 1996).
`A reporter gene containing a 48-bp VEGF
`hypoxia response element upstream of an
`SV40 promoter and luciferase coding se-
`quences is expressed in a hypoxia-inducible
`manner. Forced overexpression of HIF-1α
`and HIF-1β results in reporter gene expres-
`sion under nonhypoxic conditions and a su-
`perinduction in response to hypoxia
`(Forsythe et al., 1996). A 3-bp mutation that
`disrupts the HIF-1 binding site in the hy-
`poxia response element resulted in a loss of
`reporter gene expression in response to hy-
`poxia or coexpressed HIF-1α and HIF-1β
`(Forsythe et al., 1996). In mouse embryonic
`stem cells homozygous for a targeted muta-
`tion in the Hif1a gene encoding HIF-1α
`there is no induction of VEGF mRNA ex-
`pression in response to hypoxia (Carmeliet
`et al., 1998; Iyer et al., 1998; Ryan et al.,
`1998), although VEGF expression is still
`induced by glucose deprivation (Iyer et al.,
`1998).
`In a series of mouse hepatoma subclones,
`there is a strong correlation between the level
`of hypoxia-induced HIF-1 DNA-binding ac-
`tivity and VEGF mRNA expression (Forsythe
`et al., 1996; Salceda et al., 1996; Wood et al.,
`1996). Furthermore, when these subclones are
`injected into nude mice, there is a strong cor-
`relation between the levels of hypoxia-induced
`HIF-1 and VEGF expression ex vivo and tu-
`
`
`
`Roxane Labs., Inc.
`Exhibit 1017
`Page 007
`
`

`
`mor xenograft growth and angiogenesis in
`vivo (Jiang et al., 1997a; Maxwell et al., 1997).
`In situ hybridization analysis of tumors de-
`rived from HIF-1expressing subclones has
`demonstrated that viable cells surrounding
`areas of necrosis express high levels of VEGF
`and GLUT3 mRNA, whereas such expression
`is not seen at similar locations in tumors de-
`rived from HIF-1 nonexpressing subclones
`(Maxwell et al., 1997), suggesting that physi-
`ologic induction of hypoxia-inducible genes
`mediated by HIF-1 is important for tumor
`growth and vascularization. Vascularization
`of teratomas derived from mouse embryonic
`stem cells is also dramatically affected by loss
`of HIF-1α expression (Carmeliet et al., 1998;
`Ryan et al. 1998).
`In addition to hypoxia, VEGF expres-
`sion is also induced by exposure of both
`transformed and nontransformed cells to a
`variety of growth factors, including epider-
`mal growth factor, basic fibroblast growth
`factor (FGF-2), insulin-like growth factor 1
`(IGF-1), interleukin 1β, platelet-derived
`growth factor, and tumor necrosis factor α
`(TNF-α) (Jackson et al., 1997; Ryuto et al.,
`1996; Stavri et al., 1995a, 1995b; Warren et
`al., 1996). Remarkably, all of these
`cytokines/growth factors also induce HIF-
`1α protein expression and/or HIF-1 DNA-
`binding activity (Feldser et al., 1999;
`Hellwig-Burgel et al., 1999; Zelzer et al.,
`1998; Zhong et al., 2000).
`
`C. Hypoxia
`
`Given the strong correlation between
`angiogenesis and tumor progression de-
`scribed above, one might predict that that
`there is a positive correlation between tu-
`mor oxygenation and tumor progression.
`Remarkably, the opposite is true: first, in
`the majority of human cancers that have
`been analyzed, the mean pO2 is markedly
`
`lower than that of normal tissue in the same
`organ (reviewed in Brown and Giaccia,
`1998; Vaupel, 1996; Vaupel et al., 1989).
`Second, clinical studies in which oxygen-
`ation of cervical or head and neck tumors
`has been measured in situ have demonstrated
`that an intratumoral pO2 of < 10 mmHg is
`associated with a significantly increased
`frequency of tumor invasion and metastasis
`and of patient death (Brizel et al., 1996;
`Hockel et al., 1996). In both cervical and
`head and neck cancers, a correlation be-
`tween elevated lactate levels and metastasis
`has also been reported (Schwickert et al.,
`1995; Walenta et al., 1997).
`What is the basis for the apparent para-
`dox regarding the effects of tumor hypoxia
`and angiogenesis on clinical outcome? The
`blood vessels induced by tumors are struc-
`turally and functionally abnormal, resulting
`in marked regional heterogeneity in tumor
`perfusion (reviewed in Brown and Giaccia,
`1998; Gillies et al., 1999; Vaupel, 1996;
`Vaupel et al., 1989). Analysis of tumor xe-
`nografts has revealed that although there is
`an inverse correlation between mean pO2
`and distance of a tumor cell from the nearest
`capillary (reaching a mean pO2 of 0 at > 200
`µm), individual tumor cells with a pO2 of 0
`can be identified immediately adjacent to a
`blood vessel, indicating an absence of blood
`flow (Helmlinger et al., 1997). These data
`suggest that the induction of angiogenesis is
`necessary but not sufficient to ensure tumor
`progression.
`Intratumoral hypoxia is also associated
`with genetic instability and resistance to
`chemotherapy and radiation, thus providing
`a basis for tumor recurrence (reviewed by
`Brown and Giaccia, 1994, 1998; Yuan and
`Glazer, 1998). For chemotherapy, it is has
`been hypothesized that hypoperfused tumor
`tissue would be deprived both of O2 and any
`chemotherapeutic agent administered
`parenterally; alternatively, cell cycle arrest
`under hypoxic conditions may provide pro-
`
`77
`
`
`
`Roxane Labs., Inc.
`Exhibit 1017
`Page 008
`
`

`
`tection against agents that induce apoptosis
`in proliferating cells. For radiation, it has
`been hypothesized that decreased O2 con-
`centrations would result in decreased for-
`mation of reactive oxygen species, which
`are believed to trigger radiation-induced cell
`death. Whereas these proposed mechanisms
`may indeed contribute to the survival of
`hypoxic tumor cells following cancer
`therapy, it is also possible that hypoxic tu-
`mor cells express survival factors that me-
`diate protection against any apoptosis-in-
`ducing stimulus. Exposure of cultured
`glioma cells to hypoxia results in increased
`survival when the cells are subsequently
`exposed to chemotherapeutic agents (Liang,
`1996), an effect that is obviously indepen-
`dent of vascularization/perfusion. Insulin-
`like growth factor 2 (IGF-2) represents an
`example of a hypoxia-induced survival fac-
`tor (Kim et al., 1998), and its expression
`appears to be regulated by HIF-1 (Feldser et
`al., 1999).
`
`IV. IMMUNOHISTOCHEMICAL
`ANALYSIS OF HIF-1α
`EXPRESSION IN TUMORS
`
`A. Increased Expression of HIF-
`1ααααα in Common Human Cancers
`and Their Metastases
`
`The analysis of tumor-derived cell lines
`either in tissue culture or xenograft assays
`provides a means to investigate molecular
`mechanisms underlying tumor biology. Yet,
`such investigations must rest on a firm foun-
`dation of histopathological studies utilizing
`clinical material that provide the basis for
`generating hypotheses to be tested in model
`systems. Although clinical investigation is
`often limited by the lack of optimal con-
`trols, in the case of cancer, the patient’s
`
`78
`
`normal tissue often can be utilized effec-
`tively for this purpose. With these principles
`in mind, a mouse monoclonal antibody spe-
`cific for HIF-1α was generated, and a sen-
`sitive immunohistochemical assay was de-
`veloped to detect HIF-1α expression in
`formalin-fixed and paraffin-imbedded hu-
`man tumor biopsy specimens (Zhong et al.,
`1999). Whereas HIF-1α expression was not
`detected in the corresponding normal tissue
`(either adjacent to the tumor in the biopsy or
`in separate autopsy specimens), HIF-1α ex-
`pression was detected in two-thirds of the
`primary tumors of the brain, breast, colon,
`lung, ovary, and prostate that were analyzed
`(Table 2). Although HIF-1α expression was
`frequently detected in many, but not all,
`types of malignant tumors, it was not de-
`tected in benign tumors such as breast fi-
`broadenoma and uterine leiomyoma (Zhong
`et al., 1999). In contrast, HIF-1α expression
`was detected in early neoplastic lesions
`(breast comedo-type ductal carcinoma
`in situ, colon adenoma, and prostatic
`intraepithelial neoplasia) detected inciden-
`tally in tumor specimens, suggesting that
`increased HIF-1α expression can occur early
`in tumor progression (Zhong et al., 1999).
`In animal models, induction of angiogen-
`esis has been demonstrated during the tran-
`sition from hyperplasia to neoplasia
`(Folkman et al., 1989). HIF-1α expression
`in early neoplastic lesions may promote
`vascularization mediated by VEGF and
`possibly other HIF-1-regulated angiogenic
`factors.
`
`B. HIF-1ααααα Expression in Brain
`Tumors
`
`Immunohistochemical analysis of brain
`tumors was particularly instructive (Zagzag
`et al., 2000). Several studies have shown
`that compared with low-grade gliomas, the
`
`
`
`Roxane Labs., Inc.
`Exhibit 1017
`Page 009
`
`

`
`Table 2
`Expression of HIF-1a in Human Cancers Compared with Corresponding Normal Tissuea
`
`Organ
`
`Normal Tissue
`
`Benign Tumors
`
`Primary Cancers
`
`Metastases
`
`Brain
`Breast
`Colon
`Lung
`Ovary
`Prostate
`Uterus
`
`0/10
`0/18
`0/24
`0/10
`0/10
`0/12
`0/3
`
`—
`0/10b
`—
`—
`—
`—
`0/2e
`
`5/9
`25/52
`22/22
`3/3
`2/2
`9/11
`—
`
`—
`9/13c
`9/10c
`—
`—
`5/10c,d
`—
`
`aData from Zhong et al., 1999.
`bFibroadenomas
`cLymph node metastases
`dBone metastases
`eLeiomyomas
`
`high-grade glioblastoma multiforme (GBM)
`has
`significantly
`higher levels of VEGF expression and
`neovascularization (Leung et al., 1997;
`Pietsch et al., 1997; Plate et al., 1992; Plate
`and Mennel, 1995; Plate and Warnke, 1997;
`Takano et al., 1996; Takekawa and Sawada,
`1998). Even in low-grade astrocytomas, pa-
`tient survival is inversely correlated with
`VEGF expression and vascular density
`(Abdulrauf et al., 1998). GBM is also char-
`acterized by extensive areas of necrosis, rep-
`resenting tumor cells located beyond the
`effective diffusion distance of O2 from the
`nearest blood vessel. VEGF mRNA is highly
`
`expressed in the pseudopalisading cells sur-
`rounding necrotic areas, suggesting that hy-
`poxia is a stimulus for VEGF expression by
`these cells (Plate et al., 1992; Shweiki et al.,
`1992). Remarkably, the expression of HIF-
`1α is correlated with tumor grade and, in
`particular, with vascularity (Table 3 and Fig-
`ure 1). Expression of HIF-1β was also ana-
`lyzed by immunohistochemistry, and its ex-
`pression is also correlated with tumor grade,
`although it is more widely expressed than
`HIF-1α and is not as strongly correlated
`with tumor vascularity. Most striking is the
`pattern of HIF-1α protein expression in
`GBM which are identical to that described
`
`Table 3
`Vascularity and Expression of HIF-1ααααα and HIF-1βββββ in Human Brain Tumorsa
`
`Tumor
`Low grade astrocytoma
`Low grade mixed glioma
`Anaplastic astrocytoma
`Anaplastic oligodendroglioma
`Anaplastic mixed gliomad
`Glioblastoma multiforme
`Hemangioblastoma
`
`n
`2
`2
`2
`3
`9
`14
`10
`
`Vascularityb
`1.0
`2.5
`3.5
`3.7
`2.4
`3.9
`3.7
`
`HIF-1αααααc
`1.5
`2.5
`3.5
`3.0
`2.8
`4.0
`3.2
`
`HIF-1βββββc
`1.0
`0.5
`2.5
`2.3
`2.4
`3.2
`3.2
`
`aData from Zagzag et al., 2000.
`bVascularity: 0, vascular hyperplasia not detected; 1, minimal; 2, mild; 3, moderate; 4, marked
`increase in tumor vascularity compared with surrounding normal brain tissue
`cHIF-1α or HIF-1β: 0, nuclear staining not detected by immunohistochemistry; 1, staining
`in less than 1% of nuclei; 2, 1–10% of nuclei; 3, 10–50%; 4, greater than 50%.
`dastrocytoma/oligodendroglioma
`
`79
`
`
`
`Roxane Labs., Inc.
`Exhibit 1017
`Page 010
`
`

`
`Figure 1. Correlation between brain tumor grade, vascularity, and HIF-1a expression. Based on an
`analysis of 42 human brain tumors (Zagzag et al., 2000). See Table 3 for definitions.
`
`high
`for VEGF mRNA, with
`of HIF-1α
`in
`levels
`detected
`pseudopalisading cells surrounding areas of
`necrosis (Zagzag et al., 2000). Taken to-
`gether, these data are consistent with the
`hypothesis that HIF-1 mediates hypoxia-
`induced VEGF expression in GBM.
`Hemangioblastoma is another extremely
`vascular brain tumor. The vascularity of
`hemangioblastomas is so great that, as the
`name implies, they were originally believed
`to be tumors of hemangioblasts, the stem
`cells from which hematopoietic and endo-
`thelial cells are derived. However, it is now
`apparent that the “stromal” cells (as op-
`posed to the blood vessels) are in fact tumor
`cells that produce extremely high levels of
`VEGF (Flamme et al., 1998; Krieg et al.,
`1998; Stratmann et al., 1997; Wizigman-
`Voos et al., 1995). Unlike GBM, hemangio-
`blastomas are so richly vascularized that
`areas of necrosis are unusual. Yet, immuno-
`histochemistry revealed that in those he-
`mangioblastomas with the greatest degree
`
`80
`
`of vascularization, HIF-1α was expressed
`in the majority of tumor cells, whereas ex-
`pression was not detected in vascular cells
`(Zagzag et al., 2000; Zhong et al., 1999).
`The detection of high levels of HIF-1α in
`tumor cells immediately adjacent to most
`blood vessels suggests that such expression
`is not hypoxia induced.
`A similar pattern of expression was de-
`tected in another extremely vascularized
`tumor, clear cell renal carcinoma (Zhong et
`al., 1999). Hemangioblastoma and clear cell
`renal carcinoma share in common inactiva-
`tion of the VHL gene encoding the von
`Hippel-Lindau tumor suppressor protein
`(Gnarra et al., 1994; Herman et al., 1994;
`Kanno et al., 1994; Shuin et al., 1994). In
`renal carcinoma cell lines lacking VHL ex-
`pression, the normal O2-regulated expres-
`sion of HIF-1α and HIF-2α is lost; these
`proteins are constitutively expressed at high
`levels under nonhypoxic conditions and
`activate transcription of downstream target
`genes, including VEGF (Maxwell et al.,
`
`
`
`Roxane Labs., Inc.
`Exhibit 1017
`Page 011
`
`

`
`1999). Analysis of multiple renal carcinoma
`cell lines with VHL loss-of-function revealed
`that HIF-2α was overexpressed in all,
`whereas in several lines HIF-1α expression
`was completely absent. In contrast, HIF-1α
`was highly expressed in all of the heman-
`gioblastoma (n = 10) and clear cell renal
`carcinoma (n = 1) biopsies that were ana-
`lyzed by immunohistochemistry (Zagzag et
`al., 2000; Zhong et al., 1999), suggesting
`that there may be a selection against HIF-
`1α overexpression in cultured renal carci-
`noma cells that is not relevant to their bio-
`logical behavior in vivo. However, primary
`clear cell renal carcinomas express an
`antisense HIF-1α RNA species (Thrash-
`Bingham and Tartof, 1999). Thus, the role
`of HIF-1α expression in these tumors re-
`quires further analysis.
`
`C. HIF-1ααααα and Tumor Invasion
`
`GBM is one of the most highly inva-
`sive forms of human cancer and patients
`with this tumor have a life expectan

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket