throbber
Molecular Biology Underlying the Clinical
`Heterogeneity of Prostate Cancer
`An Update
`
`A. Craig Mackinnon, MD, PhD; Benjamin C. Yan, MD, PhD; Loren J. Joseph, MD; Hikmat A. Al-Ahmadie, MD
`
`● Context.—Recent studies have uncovered a number of
`possible mechanisms by which prostate cancers can be-
`come resistant to systemic androgen deprivation, most in-
`volving androgen-independent reactivation of the andro-
`gen receptor. Genome-wide expression analysis with mi-
`croarrays has identified a wide array of genes that are dif-
`ferentially expressed in metastatic prostate cancers
`compared to primary nonrecurrent tumors. Recently, re-
`current gene fusions between TMPRSS2 and ETS family
`genes have been identified and extensively studied for their
`role in prostatic carcinoma.
`Objective.—To review the recent developments in the
`molecular biology of prostate cancer, including those per-
`taining to the androgen receptor and the newly identified
`TMPRSS2-related translocations.
`Data Sources.—Literature review and personal experi-
`ence.
`Conclusions.—Prostatic adenocarcinoma is a heteroge-
`
`Prostate cancer is the most common noncutaneous ma-
`
`lignancy in men and the second most common cause
`of cancer deaths, accounting for 186 320 new annual cases
`and 28 660 deaths in America and an annual incidence of
`679 023 cases and 221 002 deaths worldwide in 2008.1,2 The
`highest incidence of prostate cancer is found in the United
`States, Canada, and Scandinavia, and the lowest in China
`and other parts of Asia.3,4 Risk factors include advancing
`age, African American ethnicity, family history, and diet.5–9
`The increasingly widespread testing for serum levels of
`prostate-specific antigen (PSA) has allowed for the in-
`creasing detection of prostate cancers at earlier stages of
`development. As a result, prostatic adenocarcinoma has
`become a clinically heterogeneous entity, with some early
`carcinomas following an indolent clinical course, remain-
`ing confined to the prostate with little effect on overall
`lifespan, while other cases can lead to the development of
`lethal metastatic disease. Despite the recent advances in
`
`Accepted for publication February 5, 2009.
`From the Department of Pathology, University of Chicago, Chicago,
`Illinois. Dr Al-Ahmadie is now with the Department of Pathology, Me-
`morial Sloan-Kettering Cancer Center, New York, New York.
`The authors have no relevant financial interest in the products or
`companies described in this article.
`Reprints: Hikmat A. Al-Ahmadie, MD, Department of Pathology, Me-
`morial Sloan-Kettering Cancer Center, 1275 York Avenue, New York,
`NY 10065 (e-mail: alahmadh@mskcc.org).
`
`neous group of neoplasms with a broad spectrum of path-
`ologic and molecular characteristics and clinical behav-
`iors. Numerous mechanisms contribute to the develop-
`ment of resistance to androgen ablation therapy, resulting
`in ligand-independent reactivation of the androgen recep-
`tor,
`including amplification, mutation, phosphorylation,
`and activation of coreceptors. Multiple translocations of
`members of the ETS oncogene family are present in ap-
`proximately half of clinically localized prostate cancers.
`TMPRSS2:ERG gene rearrangement appears to be an early
`event in prostate cancer and is not observed in benign or
`hyperplastic
`prostatic
`epithelium. Duplication
`of
`TMPRSS2:ERG appears to predict a worse prognosis. The
`relationship between TMPRSS2:ERG gene rearrangement
`and other morphologic and prognostic parameters of pros-
`tate cancer is still unclear.
`(Arch Pathol Lab Med. 2009;133:1033–1040)
`
`treatment modalities, surgical, radiation, and hormonal
`therapies for prostate cancer are not without complica-
`tions, making the development of methods for distin-
`guishing indolent cancers from their aggressive counter-
`parts necessary to avoid excessive treatment that may lead
`to significant morbidity.10 Furthermore, current methods
`of treating advanced metastatic disease often prove to be
`insufficient in the long-term. Recent research has therefore
`sought to identify new molecular pathways by which in-
`vestigators can distinguish indolent prostate cancers from
`those that go on to pursue a more aggressive clinical
`course, as well as to discover new targets for the treatment
`of metastatic carcinomas.
`Previous cytogenetic and molecular studies11–21 had
`shown that prostatic adenocarcinomas tend to incur fre-
`quent and consistent losses of specific chromosomal loci,
`including chromosomes 8p, 10q, 13q, and 17p, and, less
`commonly, 6q, 7q, 16q, and 18q, as well as gains in chro-
`mosomes 7 and 8q. More recently, investigators compared
`the gene expression profiles of primary nonrecurrent pros-
`tatic adenocarcinomas and metastatic prostate cancers by
`using microarrays.22 Genes that were more highly ex-
`pressed in metastatic carcinomas included those involved
`in DNA synthesis and repair, mitosis, and cell cycle reg-
`ulation, such as RFC5, TOP2A, RFC4, and MAD2L1, which
`have previously been shown to be highly expressed in
`proliferating cells.23 Other differentially expressed genes
`
`Arch Pathol Lab Med—Vol 133, July 2009
`
`Updates in Prostate Cancer—Mackinnon et al 1033
`
`002004
`
`AVENTIS EXHIBIT 2025
`Mylan v. Aventis, IPR2016-00712
`
`

`
`included those involved in signal transduction, transcrip-
`tional regulation, chromatin modification, RNA process-
`ing, protein synthesis, posttranslational protein modifica-
`tion, cell adhesion, cell migration, cytoskeletal regulatory
`elements, extracellular matrix proteins, biosynthetic en-
`zymes, and transport proteins.22 Several unclassified genes
`with unknown functions were also found to be differen-
`tially expressed.22
`
`PROSTATIC ADENOCARCINOMA RESISTANT TO
`ANDROGEN ABLATION THERAPY
`Prostate cancer affects 1 of 9 men older than 65 years.24
`Age correlates with a decrease in the ratio of androgens
`to estrogens in men, suggesting that a physiologic change
`in hormonal status may contribute to the progression of
`preneoplastic lesions to adenocarcinoma.25–27 Androgen
`ablation therapy is the most common systemic treatment
`for metastatic disease; it prevents testosterone production
`by the testes and thereby causes tumor regression during
`the short-term by depleting androgen-dependent carci-
`noma cells.28,29 Androgen deprivation can be achieved sur-
`gically or by medical castration, which can be performed
`by the administration of estrogens and gonadotropin-re-
`leasing agonists and antagonists, and has been shown to
`be effective in treating advanced and metastatic disease in
`several large clinical trials.30–33 However, most prostatic ad-
`enocarcinomas become refractory to androgen ablation.34
`Experiments with animal models of prostatic adenocarci-
`nomas,34,35 such as Dunning R-3327-H rat prostate carci-
`nomas and the transgenic adenocarcinoma mouse prostate
`model, have shown that androgen therapy ultimately se-
`lects for androgen-independent carcinoma cells during the
`long-term, leading to the development of highly aggres-
`sive, androgen-resistant tumors. Despite this progression
`to more aggressive disease, it is still advised that patients
`with hormone-resistant cancers continue to receive andro-
`gen ablation therapy.36
`Aberrant activation of the androgen receptor (AR) can
`result from gene amplification, mutation, phosphorylation,
`activation of coregulators, or androgen-independent acti-
`vation. Most cases of prostatic carcinoma resistant to an-
`drogen ablation therapy demonstrate activation of AR by
`one of these mechanisms.37–39 The AR gene is either mu-
`tated or amplified in 20% to 30% of androgen-resistant
`prostate carcinomas.40–42 Further, 20% of hormone-resis-
`tant carcinomas contain gene amplifications as compared
`to just 2% of hormone-sensitive tumors, suggesting that
`aberrant activation in response to low levels of androgens
`or other ligands may underlie the progression to aggres-
`sive disease that is refractory to androgen ablation thera-
`py.43 Specific mutations, such as the T877A and H874Y
`substitutions, confer increased sensitivity to AR for steroid
`hormones such as progesterone, 17␤-estradiol, or hydroxy-
`flutamide in prostate cancer cell lines and xenografts.44,45
`Mutant AR containing the E231G substitution has also
`been shown to predispose transgenic mice to the devel-
`opment of prostatic intraepithelial neoplasia (PIN), ade-
`nocarcinoma, and metastases.46 Experiments have shown
`that AR hyperactivity results in the formation of prostatic
`neoplasms: overexpression of AR leads to the develop-
`ment of
`focal PIN, whereas AR overexpression in
`LAPC-4 prostate cancer cells and xenografts results in a
`transition from androgen-sensitive disease to androgen-
`resistant cancer.47,48 Phosphorylation of AR at different
`specific serine residues may cause stabilization of the pro-
`
`tein against proteolytic degradation or induce transcrip-
`tional activation of the receptor protein.49
`Inappropriate AR hyperactivity may also be caused by
`activation of coregulators. Recurrent CWR22 tumors were
`found to harbor overexpressed transcriptional intermedi-
`ary factor 2 and steroid receptor coactivator 1, which in-
`creased AR transactivation at physiologic androgen con-
`centrations.37 Other coregulators of AR function include
`ARA70, p300, CBP, Tip60, ARA55, ARA54, gelsolin, Stat3,
`and RAC3.50–53 The Foxa1 and Foxa2 proteins are tran-
`scription factors belonging to the forkhead box A (Foxa)
`superfamily (also known as hepatocyte nuclear factor 3
`proteins) that are essential for endodermal development
`and are involved in respiratory, intestinal, and hepatic
`gene expression.54–60 Although Foxa1 was found to be ex-
`pressed in prostatic carcinomas of different grades, Foxa2
`stimulates transactivation of the PSA promoter in an an-
`drogen- and AR-independent manner and has been iden-
`tified in small cell carcinomas and high-grade adenocar-
`cinomas of the prostate, suggesting that Foxa2 regulation
`of gene expression may contribute to progression of pros-
`tatic carcinomas to a more aggressive and androgen-in-
`dependent state.58
`Genome-wide expression analyses61 have identified
`genes that are differentially expressed in prostate cancers
`from patients who had received the gonadotropin-releas-
`ing agonist goserelin and AR antagonist flutamide for 3
`months. Hierarchical clustering algorithms that analyzed
`gene expression profiles classified the specimens accord-
`ing to treatment status, suggesting that distinct transcrip-
`tional programs are activated in prostate carcinomas in
`response to androgen therapy. The genes that were more
`highly expressed in carcinomas treated with androgen ab-
`lation agents included those encoding AR and steroid bio-
`synthetic enzymes, as well as a suite of genes that have
`been previously shown to be targets of AR or have been
`implicated as being regulated by it, including the gene
`encoding PSA (kallikrein-related peptidase 3), KLK3, and
`KLK2, as well as DBI, FASN, IL6, SERPINB5, TGFBR3,
`TMPRSS2, TUBA1, HOXC6, TRG , and FOLH1.61–66 Other
`upregulated genes may represent secondary, indirect ef-
`fects of androgen ablation that occur later than reactiva-
`tion of AR. To identify only those genes that are subject
`to transcriptional regulation by AR, gene expression pro-
`files of LNCaP human prostatic adenocarcinoma cells
`were examined after androgen withdrawal.61 Approxi-
`mately 25% of the genes differentially expressed in carci-
`nomas after chronic androgen ablation therapy also
`showed an altered transcript level in the carcinoma cell
`line. Finally, comparison of the gene expression profiles of
`androgen-resistant cancers to those of cancers that had not
`developed resistance demonstrated that prostatic adeno-
`carcinomas resistant to androgen ablation therapy had
`gene expression profiles more similar to those of untreat-
`ed, androgen-dependent tumors than of cancers under
`conditions of androgen deprivation. This finding suggests
`a reversal in the gene expression profile of androgen-re-
`fractory cancers that is caused by androgen deprivation
`therapy, possibly by ligand-independent reactivation of
`AR, a mechanism that has been proposed by several au-
`thors.67–70 Furthermore, a unique set of genes was ex-
`pressed in androgen-resistant prostatic carcinomas that
`was not expressed in primary androgen-dependent tu-
`mors or in other metastatic carcinomas.61
`Activation of AR can be highlighted by immunohisto-
`
`1034 Arch Pathol Lab Med—Vol 133, July 2009
`
`Updates in Prostate Cancer—Mackinnon et al
`
`

`
`chemistry as a strong nuclear expression in androgen-re-
`sistant prostate cancers.61 Moreover, reactivation of AR
`gene was not due to gene amplification in most cases, as
`it was shown by FISH analysis that only a minority of the
`androgen-resistant carcinomas studied contained ampli-
`fied AR genes. The human prostatic adenocarcinoma xe-
`nograft CWR22, which is propagated in nude mice, reca-
`pitulates the properties of in vivo prostate cancers, with
`an initial period of androgen-dependent proliferation fol-
`lowed by persistent growth several months after androgen
`deprivation.71,72 Androgen receptor protein from a re-
`lapsed CWR22 carcinoma has a half-life that is 2 to 4 times
`that of AR from LNCaP cells, demonstrating that recurrent
`tumors have hyperstabilized AR as compared to andro-
`gen-dependent neoplasms.73 The increased expression,
`greater stability, and nuclear localization of AR in recur-
`rent prostate cancers resistant to androgen deprivation
`correlated with hypersensitivity to low levels of androgens
`in these tumors; androgen ablation–resistant prostate can-
`cers required a significantly much lower concentration of
`dihydrotestosterone than that required by androgen-de-
`pendent tumors for stimulation of proliferative activity.
`Several of the genes that were found by the microarray
`study to be more highly transcribed in androgen ablation–
`resistant tumors encoded biosynthetic enzymes involved
`in the synthesis of cholesterol, including HMG-CoA syn-
`thase, squalene synthase, lanosterol synthase, and squa-
`lene monooxygenase, the rate-limiting enzyme in sterol
`synthesis.61,74 Androgen ablation–resistant tumors were
`shown to be more strongly immunoreactive for squalene
`monooxygenase than were androgen-dependent tumors.61
`The increased production of steroid biosynthetic enzymes
`in resistant tumors suggests that one mechanism by which
`these carcinomas overcome androgen deprivation is by
`compensatory synthesis of androgens, with consequently
`increased AR activity. Recurrent prostatic carcinomas con-
`sistently exhibit decreased expression of the tumor sup-
`pressor gene PTEN (phosphatase and tensin homolog),
`which carries loss-of-function mutations in advanced pros-
`tate cancers.75 The PTEN protein dephosphorylates phos-
`phatidylinositol-3,4,5-trisphosphate (PIP3), resulting in in-
`hibition of the Akt (protein kinase B) cell survival signal-
`ing pathway.76
`Besides reactivation of AR and loss of PTEN tumor sup-
`pressor activity, other mechanisms for the development of
`hormone-resistant prostate cancers have been proposed.
`Aberrant overexpression or amplification of the HER2/neu
`gene has been identified in prostatic carcinomas77 and el-
`evated serum levels of the HER2/neu extracellular domain
`were found in androgen ablation–refractory prostate can-
`cers.78 Overexpression of the HER2/neu (ERBB2, CD340)
`receptor tyrosine kinase was capable of rescuing LNCaP
`cells from the antiproliferative effect of androgen depri-
`vation and also shortened the latency period for tumor
`formation in castrated mice with severe combined im-
`munodeficiency by 50%.77 Furthermore, HER2/neu can en-
`hance by 15-fold the expression of the AR target PSA in
`the absence of androgens in LAPC-4 cells.77 HER2/neu can
`also activate MAP kinase and PIP3/Akt signaling cas-
`cades, culminating in the phosphorylation of serines 213
`and 791 of AR.79,80 Constitutive Akt activity led to in-
`creased neoplastic growth in LNCaP xenografts.81 Other
`growth factors that may contribute to the development of
`hormone-resistant cancers include insulin-like growth fac-
`
`tor 1, epidermal growth factor, keratinocyte growth factor,
`and factors secreted by neuroendocrine cells.82–84
`The conserved basic helix-loop-helix transcription factor
`TWIST has been shown to be highly expressed in the ma-
`jority (90%) of prostate cancers and only in a minority
`(6.7%) of benign prostatic hyperplasia cases.85 TWIST ex-
`pression levels were also found to be proportional to Glea-
`son grade, and higher levels correlated with metastasis.85
`Experiments with DU145 and PC-3 androgen-resistant
`prostatic adenocarcinoma cell lines85 found that down-reg-
`ulation of TWIST expression by RNA interference led to
`suppression of invasiveness and a reduction in E-cadherin
`expression, as well as loss of the morphologic and molec-
`ular changes that signify the epithelial-mesenchymal tran-
`sition.
`In summary, a major clinical challenge presented by
`prostate cancer is the treatment of androgen ablation–re-
`sistant carcinomas. Recent experimental evidence suggests
`that there are multiple avenues leading to the development
`of this aggressive form of prostatic carcinoma, which sub-
`vert the molecular mechanisms of the cell to reactivate AR,
`activate its targets, gain inappropriate HER2/neu activity,
`lose PTEN-mediated tumor suppression, or stimulate the
`epithelial-mesenchymal transition via TWIST. Other au-
`thors postulate the existence of androgen-resistant pros-
`tate cancer stem cells that contribute to the growth of ag-
`gressive tumors. Future molecular studies will help fur-
`ther elucidate the diverse signaling pathways underlying
`the pathogenesis of prostate cancers refractory to systemic
`hormonal deprivation and may lead to the development
`of multiple pharmacologic agents and therapeutic modal-
`ities that will halt progression of hormone-naı¨ve tumors.
`
`MOLECULAR BIOLOGY OF TMPRSS2:ERG
`TRANSLOCATION
`A significant role for the ETS gene family, which en-
`codes transcription factors in prostate cancer, was recently
`discovered by using a novel bioinformatics approach
`known as COPA (cancer outlier profile analysis)86 that
`identified the oncogenes ERG (21q22.2), ETV1 (7p21.2),
`ETV4 (17q21), and ETV5 (3q27) as very highly expressed
`in a subset of prostate cancers on the basis of a large set
`of microarray data.86–88 ERG, ETV1, ETV4, and ETV5 are
`members of the ETS family of transcription factors, which
`are characterized by an evolutionarily conserved, 85–ami-
`no acid DNA-binding domain that facilitates binding to
`purine-rich DNA with a GGAA/T core consensus se-
`quence.89 ETS proteins function cooperatively with other
`transcription factors in the regulation of a diversity of cel-
`lular functions including proliferation, differentiation, an-
`giogenesis, hematopoiesis, oncogenic transformation, and
`apoptosis.90 Importantly, translocations involving mem-
`bers of the ETS family have been identified in human leu-
`kemia and solid tumors.89 A mechanism underlying ETS
`overexpression in prostate cancer was established once it
`was recognized that
`the androgen-responsive gene
`TMPRSS2 (see below) is fused to the coding region of an
`ETS family member (for example ERG) as a result of gene
`rearrangement, which was also demonstrated directly by
`studies in vitro.86,91,92 The TMPRRSS2:ERG gene fusion is
`observed in greater than 90% of prostate cancers with
`ETS-family gene rearrangements,93 whereas TMPRSS2:
`ETV1, TMPRSS2:ETV4, and TMPRSS2:ETV5 rearrange-
`ments occur more rarely. Furthermore, ETV1, ETV4, and
`ETV5 have additional
`fusion partners other
`than
`
`Arch Pathol Lab Med—Vol 133, July 2009
`
`Updates in Prostate Cancer—Mackinnon et al 1035
`
`

`
`TMPRSS2, including SLC45A3, HERV-K㛮22q11.3, C15orf21,
`and HNRPA2B1.94
`TMPRSS2 is located at 21q22.2,95 and TMPRSS2 is pre-
`dominantly expressed in luminal epithelial prostate cells,
`with much lower expression in pancreas, kidney, lung, co-
`lon, and liver and no measurable expression in testes, ova-
`ry, placenta, spleen, thymus, circulating leukocytes, brain,
`heart, or skeletal muscle.96,97 TMPRSS2 and ERG are lo-
`cated 3Mb apart on chromosome 21q22.2-22.3 (Figure, A
`through E). The 5⬘ ends of both genes are orientated to-
`ward the telomere, and TMPRSS2 is positioned telomeri-
`cally relative to ERG. Interstitial deletion of the intervening
`intronic genomic DNA is the most common mechanism
`for fusion and is observed in 60% to 90% of TMPRSS2:
`ERG fusion-positive prostate cancers (see below). Regions
`of microhomology exist in the TMPRSS2 and ERG loci,
`suggesting that they might underlie rearrangement events
`during defective homologous recombination.98
`Seventeen different types of TMPRSS2:ERG fusion tran-
`scripts involving various regions of the TMPRSS2 and
`ERG genes have been identified 86,99–102; however, 8 of these
`transcripts are unlikely to result in the translation of func-
`tional ERG proteins due to the introduction of premature
`stop codons. Of the 9 predicted functional TMPRSS2:ERG
`fusion transcripts, 2 code for normal ERG proteins, 6 code
`for amino-terminal–truncated ERG proteins, and 1 codes
`for a bona fide TMPRSS2:ERG fusion protein.101 These
`studies demonstrate that expression of multiple fusion
`mRNAs is common, with TMPRSS2 exon 1 fused to ERG
`exon 4 being the most frequently expressed type of
`TMPRSS2:ERG fusion. Alternative splicing of
`the
`TMPRSS2:ERG gene is proposed as the most likely basis
`for the multiple different types of fusion mRNAs ob-
`served.99,101
`
`DETECTION AND PREVALENCE OF TMPRSS2:ERG GENE
`FUSION IN PROSTATE CANCER
`Most TMPRSS2:ERG gene fusion events in patients with
`clinically localized prostate cancer (ie, patients identified
`through PSA screening who have potentially curable dis-
`ease by surgical resection) are characterized by using ei-
`ther fluorescence in situ hybridization (FISH) or quanti-
`tative reverse transcription–polymerase chain reaction
`(RT-PCR). One common FISH method uses break-apart
`probes that bind the 5⬘ (ie, green) and 3⬘ (ie, red) ends of
`the ERG gene (Figure). In normal prostate tissue, both of
`these probes hybridize to the ERG locus and generate ad-
`jacent green and red fluorescent signals in the nucleus. In
`contrast, prostate cancer cells harboring rearranged ERG
`demonstrate distinct, separate green and red fluorescent
`signals, as the probes are split because of change in chro-
`mosome structure. One advantage of this method is that
`it also reveals the mechanism underlying the rearrange-
`ment by detection of the commonly occurring 3-Mb inter-
`stitial deletion between TMPRSS2 and ERG. In such cases,
`this deletion manifests by the loss of the 5⬘ (green) signal
`in nuclei of malignant cells.
`The initial report86 identifying the TMPRSS2:ERG gene
`rearrangement demonstrated that 47% of clinically local-
`ized prostate cancers contain the TMPRSS2:ERG fusion,
`and two-thirds of these translocations are formed second-
`ary to interstitial deletion between the TMPRSS2 and ERG
`genes
`on
`chromosome
`21q22.2-22.3.
`Subsequent
`work101,103–106 confirms that the TMPRSS2:ERG rearrange-
`ment is present in approximately 50% of primary prostate
`
`cancer samples with interstitial deletion of the 5⬘ region
`of ERG occurring in 60% of the TMPRSS2:ERG-positive
`primary prostate cancers. Gene rearrangements involving
`ETV1 and ETV4 are rare, accounting for approximately
`2% of all observed gene alterations.
`
`MORPHOLOGY OF TMPRSS2:ERG PROSTATE CANCER
`Initially, morphologic analysis of prostate cancer cases107
`identified 5 features strongly associated with the presence
`of the TMPRSS2:ERG fusion, which included blue-tinged
`mucin, cribriform growth pattern, macronucleoli, intra-
`ductal tumor spread, and signet ring cell features. Of the
`cases demonstrating none of these features, 24% were
`TMPRSS2:ERG fusion-positive. Conversely, 55%, 86%, and
`93% of cases with 1, 2, or 3⫹ features, respectively, were
`TMPRSS2:ERG fusion-positive.
`The association of the TMPRSS2:ERG fusion with high-
`grade prostatic intraepithelial neoplasia (HGPIN) was ob-
`served in up to 21% of cases in several studies.94,108,109 Un-
`like localized prostate cancer, in which TMPRSS2:ERG fu-
`sion-positive prostate cancers typically demonstrate over-
`expression of ERG, overexpression of ERG is less constant
`in TMPRSS2:ERG fusion-positive HGPIN.109 TMPRSS2:
`ERG fusions were not present in benign prostatic epithe-
`lium or other lesions not associated with prostate cancer,
`specifically, benign prostatic hyperplasia and proliferative
`inflammatory atrophy. Furthermore, all TMPRSS2:ERG-
`positive HGPIN cases identified by FISH showed the same
`fusion pattern as the matching prostate cancer from the
`same patient, and no fusion-positive HGPIN cases asso-
`ciated with fusion-negative prostate cancer were identified
`by FISH,108,110 suggesting TMPRSS2:ERG may play a role
`in the progression from HGPIN to adenocarcinoma.92
`Based upon (1) the homogenous distribution of the fu-
`sion gene throughout the cancer, (2) the absence of de-
`tectable TMPRSS2:ERG fusion events in normal and hy-
`perplastic prostate tissue, (3) the finding that TMPRSS2:
`ERG-positive HGPIN lesions show the same fusion pat-
`tern with the matching prostate cancer, and (4) the fact
`that TMPRSS2:ERG fusion-positive HGPIN is never ob-
`served with fusion-negative matching prostate cancer, it
`can be suggested that TMPRSS2:ERG fusion is an early
`event in the development of prostate adenocarcinoma. In
`support of these findings, transgenic mice expressing the
`equivalent
`truncated ERG gene coded by human
`TMPRSS2:ERG develop mouse PIN (mPIN), along with
`loss of the p63-positive basal layer adjacent to mPIN foci.92
`These findings strongly suggest that TMPRSS2:ERG fu-
`sion HGPIN is a true precursor for TMPRSS2:ERG-posi-
`tive prostate cancer.
`
`TMPRSS2:ERG AND CLINICAL PROGNOSIS
`As a wide array of distinct TMPRSS2:ERG fusions have
`been identified, several studies have explored whether
`specific gene fusion isoforms correlate with aggressive
`clinical behavior. One group101 observed that expression of
`TMPRSS2:ERG fusion variants consisting of exons 1 and
`2 of TMPRSS2 fused to exon 4 of ERG (T2E4) and, to a
`lesser extent, exon 1 of TMPRSS2 juxtaposed to exons 2
`or 3 of ERG (T1E2/3) are associated with early recurrence
`and seminal vesicle invasion. Interestingly, these fusion
`isoforms all contain the native ATG translation initiation
`codon from either TMPRSS2 or ERG, raising the possibil-
`ity that increased efficiency of translation from native
`translation start codons may underlie the correlation be-
`
`1036 Arch Pathol Lab Med—Vol 133, July 2009
`
`Updates in Prostate Cancer—Mackinnon et al
`
`

`
`Model of TMPRSS2:ERG gene rearrangements and fluorescence in situ hybridization (FISH) patterns observed in prostate cancer. A, Physical map
`of the TMPRSS2 and ERG loci on 21q22.2-22.3. T and C orientate toward the telomeric and centromeric regions, respectively. 5⬘ERG (green) and
`3⬘ ERG (red) FISH break-apart probes are positioned above the chromosome relative to where they hybridize. The TMPRSS2 and ERG loci are
`separated by approximately 3 Mb. B, Chromosome structure (upper) and nuclear FISH pattern (lower) observed in a normal ERG locus. This pattern
`is observed in approximately 50% of all clinically localized prostate cancers. C, Separated 5⬘ ERG and 3⬘ ERG loci due to gene rearrangement.
`This pattern has been referred to as Esplit (see text). Note the retention of both the 5⬘ ERG and 3⬘ ERG FISH probe signals. D, Chromosome
`structure and nuclear FISH pattern observed with an interstitial deletion of the 5⬘ERG locus. Note the loss of the 5⬘FISH probe signal. This is the
`most common TMPRSS2:ERG gene rearrangement pattern observed in prostate cancer and has been referred to as 1Edel. E, Interstitial deletion of
`the 5⬘ ERG locus accompanied by a duplication of 3⬘ ERG sequence. Note the loss of the 5⬘ FISH probe and duplication of the 3⬘ FISH probe
`signals. This pattern of TMPRSS2:ERG gene rearrangement is associated with a worse clinical prognosis and has been referred to as 2⫹Edel.
`
`tween TMPRSS2:ERG fusion type and aggressive clinical
`course. Alternatively, it may represent altered biochemical
`properties of the TMPRSS2:ERG fusion protein.
`A comprehensive FISH analysis of TMPRSS2:ERG re-
`arrangement in 445 cases111 has demonstrated that pros-
`tate cancer in which the 5⬘ portion of ERG is deleted has
`significantly worse cause-specific and overall survival
`than prostate cancer in which ERG is either not disrupted
`(ie, normal EGR) or contains a balanced ERG translocation
`(ie, split EGR). When ERG-deleted prostate cancers were
`further analyzed, the authors observed that prostate can-
`cer with 2 or more copies of the 3⬘ ERG region showed
`much worse clinical behavior: the survival rate of patients
`with duplication of the TMPRSS2:ERG gene rearrange-
`ment was 25% at 8 years compared to 90% for patients
`with prostate cancer without ERG rearrangement. A sep-
`arate study examining 521 patients112 reported similar re-
`sults in which duplication of TMPRSS2:ERG was associ-
`
`Arch Pathol Lab Med—Vol 133, July 2009
`
`ated with higher clinical stage and aggressive disease.
`Furthermore, analysis of 214 patients with prostate cancer
`suggested that multiple copies of TMPRSS2:ERG were as-
`sociated with greater prostate cancer–specific mortality, al-
`though this study113 was not statistically significant. Taken
`together, these results demonstrate that ERG gene copy
`number may provide useful prognostic information for
`patients with prostate cancer.
`The association of TMPRSS2:ERG gene rearrangement
`with Gleason score, aggressive disease, and prognosis is
`unclear, as multiple studies with conflicting findings have
`been described. A population-based study114 of 252 men
`followed up expectantly with low-stage (T1a-bNXM0)
`prostate cancer explored the risk of metastasis or prostate
`cancer–specific death based upon the presence of the
`TMPRSS2:ERG fusion. These authors determined that
`TMPRSS2:ERG fusion-positive prostate cancer is associ-
`ated with higher Gleason score (⬎7) than fusion-negative
`Updates in Prostate Cancer—Mackinnon et al 1037
`
`

`
`prostate cancer. Furthermore, cumulative incidence regres-
`sion analysis demonstrated a significant association be-
`tween TMPRSS2:ERG fusion-positive prostate cancer and
`metastases or disease-specific death. Another study115 re-
`ported similar findings. However, several other stud-
`ies98,110,116–118 failed to establish any correlation between
`TMPRSS2:ERG gene fusion status and Gleason score, tu-
`mor stage, or clinical outcomes. Lastly, using a xenograft
`model system, Hermans et al119 demonstrated that ad-
`vanced, AR-negative tumors did not express
`the
`TMPRSS2:ERG fusion transcript despite its presence in
`the genomic DNA, indicating that TMPRSS2:ERG is not
`involved in androgen-independent growth of these xeno-
`grafts. This finding suggests that the TMPRSS2:ERG fu-
`sion is important during the early, androgen-sensitive
`stage of
`tumor growth, but androgen-dependent
`TMPRSS2:ERG expression is bypassed and down-regulat-
`ed as tumor growth progresses and becomes androgen
`resistant.119
`In summary, the discovery of TMPRSS2 gene rearrange-
`ments helped broaden our understanding of the molecular
`pathology of prostate cancer, and the numerous studies
`that followed the initial report confirmed the high preva-
`lence of TMPRSS2:ETS gene alterations in prostate cancer,
`as well as advanced our understanding of androgen sig-
`naling during prostate cancer progression. Mouse and hu-
`man studies clearly demonstrate the occurrence of
`TMPRSS2:ETS gene fusion events in early HGPIN lesions;
`however, no evidence to date demonstrates a direct role
`for TMPRSS2:ETS fusion genes in the progression to ad-
`enocarcinoma, a fact suggesting a requirement for addi-
`tional genetic mutations in the course of prostate cancer
`development.
`
`References
`1. Jemal A, Siegel R, Ward E, et al. Cancer statistics, 2008. CA Cancer J Clin.
`2008;58:71–96.
`2. Parkin DM, Bray F, Ferlay J, Pisani P. Global cancer statistics, 2002. CA
`Cancer J Clin. 2005;55:74–108.
`3. Gronberg H. Prostate cancer epidemiology. Lancet. 2003;361:859–864.
`4. Quinn M, Babb P. Patterns and trends in prostate cancer incidence, survival,
`prevalence and mortality—part I: international comparisons. BJU Int. 2002;90:
`162–173.
`5. American Cancer Society. Cancer Facts and Figures, 2003. Atlanta, GA:
`American Cancer Society; 2003.
`6. Bratt O. Hereditary prostate cancer: clinical aspects. J Urol. 2002;168:906–
`913.
`7. Carter BS, Beaty TH, Steinberg GD, Childs B, Walsh PC. Mendelian inher-
`itance of familial prostate cancer. Proc Natl Acad Sci U S A. 1992;89:3367–3371.
`8. Haas GP, Sakr WA. Epidemiology of prostate cancer. CA Cancer J Clin.
`1997;47:273–287.
`9. Steinberg GD, Carter BS, Beaty TH, Childs B, Walsh PC. Family history and
`the risk of prostate cancer. Prostate. 1990;17:337–347.
`10. Michaelson MD, Cotter SE, Gargollo PC, Zietman AL, Dahl DM, Smith
`MR. Management of complications of prostate

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