84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide,...

43
1463 Incidence • Prostate cancer is the most commonly diagnosed life-threatening cancer in men (241,740 cases and 28,170 deaths in 2012). • Small prostate cancers are present in 29% of men between ages 30 and 40 and 64% of men between ages 60 and 70. • The lifetime risk of a prostate cancer diagnosis is 1 in 6, and the risk of dying from prostate cancer is 1 in 35. • Age, family history, diet and lifestyle, and ethnicity are risk factors for prostate cancer development. Biological Characteristics • Germline mutations in RNASEL and MSR1, encoding proteins that function in host responses to infection, appear responsible for some cases of hereditary prostate cancer. • An inflammatory lesion, termed proliferative inflammatory atrophy (PIA), is an early precursor to prostate cancer. • Somatic inactivation of GSTP1, encoding a carcinogen-detoxification enzyme, may initiate prostatic carcinogenesis by increasing the vulnerability of prostate cells to damage mediated by oxidant and electrophilic carcinogens. • Gene fusions, involving TMPRSS2 and ETS family transcription factor genes, may contribute to the androgen dependence of prostate cancers. • Defects in the functions of NKX3.1, PTEN, and CDKN1B are common in prostate cancer cells. Screening, Diagnosis, and Staging • Prostate cancer screening using specific antigen (PSA) testing reduces the risk of prostate cancer death but may also lead to overdiagnosis of non-life-threatening disease. • Transrectal ultrasound (TRUS)- guided core needle biopsies are used to diagnose prostate cancer. • Stage, histologic grade (Gleason score), and serum PSA levels are prognostic factors. Primary Therapy • Management options include observational strategies (watchful waiting and active surveillance), anatomic radical prostatectomy (with or without robot-assisted laparoscopic techniques), external beam radiation therapy, and brachytherapy. • A progressive rise in the serum PSA after treatment indicates prostate cancer recurrence. • Depending on the approach used, side effects associated with treatment of localized prostate cancer can include urinary, bowel, and sexual dysfunction. • Salvage therapy for prostate cancer recurrences after initial treatment include external radiation after surgery, or include surgery, brachytherapy, or cryosurgery after external beam radiation. Adjuvant Therapy • Adjuvant androgen suppression can improve survival for some men with prostate cancer treated with external beam radiation therapy. • Adjuvant external radiation improves survival for some men treated with radical prostatectomy. Treatment of Advanced Disease • Androgen suppression, most often accomplished via the use of luteinizing hormone–releasing hormone (LHRH) analogs or antagonists, with or without antiandrogens, is the most commonly used treatment. • Side effects can include loss of libido, hot flashes, gynecomastia, loss of lean muscle mass and bone density, and the development of metabolic syndrome. • Docetaxel and cabazitaxel chemotherapy improves the survival of men with progressive androgen- independent prostate cancer. • Second-line treatments targeting the androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of men previously treated with androgen suppression and taxane chemotherapy. • Bisphosphonates and denosumab antagonize loss of bone density accompanying androgen deprivation, and reduce skeletal complications associated with metastatic prostate cancer progression. • Sipuleucel-T, a dendritic cell vaccine, has shown a survival benefit in men with advanced prostate cancer. Other immunotherapies are under development in clinical trials. SUMMARY OF KEY POINTS 84 Prostate Cancer William G. Nelson, H. Ballentine Carter, Theodore L. DeWeese, Emmanuel S. Antonarakis, and Mario A. Eisenberger

Transcript of 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide,...

Page 1: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1463

Incidence• Prostatecanceristhemost

commonlydiagnosedlife-threateningcancerinmen(241,740casesand28,170deathsin2012).

• Smallprostatecancersarepresentin29%ofmenbetweenages30and40and64%ofmenbetweenages60and70.

• Thelifetimeriskofaprostatecancerdiagnosisis1in6,andtheriskofdyingfromprostatecanceris1in35.

• Age,familyhistory,dietandlifestyle,andethnicityareriskfactorsforprostatecancerdevelopment.

Biological Characteristics• GermlinemutationsinRNASELand

MSR1,encodingproteinsthatfunctioninhostresponsestoinfection,appearresponsibleforsomecasesofhereditaryprostatecancer.

• Aninflammatorylesion,termedproliferativeinflammatoryatrophy(PIA),isanearlyprecursortoprostatecancer.

• SomaticinactivationofGSTP1,encodingacarcinogen-detoxificationenzyme,mayinitiateprostaticcarcinogenesisbyincreasingthevulnerabilityofprostatecellstodamagemediatedbyoxidantandelectrophiliccarcinogens.

• Genefusions,involvingTMPRSS2andETSfamilytranscriptionfactorgenes,maycontributetotheandrogendependenceofprostatecancers.

• DefectsinthefunctionsofNKX3.1,PTEN,andCDKN1Barecommoninprostatecancercells.

Screening, Diagnosis, and Staging• Prostatecancerscreeningusing

specificantigen(PSA)testingreducestheriskofprostatecancerdeathbutmayalsoleadtooverdiagnosisofnon-life-threateningdisease.

• Transrectalultrasound(TRUS)-guidedcoreneedlebiopsiesareusedtodiagnoseprostatecancer.

• Stage,histologicgrade(Gleasonscore),andserumPSAlevelsareprognosticfactors.

Primary Therapy• Managementoptionsinclude

observationalstrategies(watchfulwaitingandactivesurveillance),anatomicradicalprostatectomy(withorwithoutrobot-assistedlaparoscopictechniques),externalbeamradiationtherapy,andbrachytherapy.

• AprogressiveriseintheserumPSAaftertreatmentindicatesprostatecancerrecurrence.

• Dependingontheapproachused,sideeffectsassociatedwithtreatmentoflocalizedprostatecancercanincludeurinary,bowel,andsexualdysfunction.

• Salvagetherapyforprostatecancerrecurrencesafterinitialtreatmentincludeexternalradiationaftersurgery,orincludesurgery,brachytherapy,orcryosurgeryafterexternalbeamradiation.

Adjuvant Therapy• Adjuvantandrogensuppressioncan

improvesurvivalforsomemenwith

prostatecancertreatedwithexternalbeamradiationtherapy.

• Adjuvantexternalradiationimprovessurvivalforsomementreatedwithradicalprostatectomy.

Treatment of Advanced Disease• Androgensuppression,mostoften

accomplishedviatheuseofluteinizinghormone–releasinghormone(LHRH)analogsorantagonists,withorwithoutantiandrogens,isthemostcommonlyusedtreatment.

• Sideeffectscanincludelossoflibido,hotflashes,gynecomastia,lossofleanmusclemassandbonedensity,andthedevelopmentofmetabolicsyndrome.

• Docetaxelandcabazitaxelchemotherapyimprovesthesurvivalofmenwithprogressiveandrogen-independentprostatecancer.

• Second-linetreatmentstargetingtheandrogen-signalingpathway,includingabirateroneacetateandenzalutamide,prolongsurvivalofmenpreviouslytreatedwithandrogensuppressionandtaxanechemotherapy.

• Bisphosphonatesanddenosumabantagonizelossofbonedensityaccompanyingandrogendeprivation,andreduceskeletalcomplicationsassociatedwithmetastaticprostatecancerprogression.

• Sipuleucel-T,adendriticcellvaccine,hasshownasurvivalbenefitinmenwithadvancedprostatecancer.Otherimmunotherapiesareunderdevelopmentinclinicaltrials.

S U M M A R Y O F K E Y P O I N T S

84  Prostate CancerWilliamG.Nelson,H.BallentineCarter,TheodoreL.DeWeese,EmmanuelS.Antonarakis,andMarioA.Eisenberger

Page 2: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1464

Figure 84-1 •  The anatomy of the prostate: rectum, bladder, dorsal vein complex, striated urethral sphincter, pelvic plexus, and neurovascular bundle. 

Dorsal v. complex Seminal vesicle

Vas deferens

Ureter

Urethrallumen

Left neurovascularbundle

Prostate

Bladder

Rectum

Pelvicplexus

Striatedurethralsphincter

Symph

.

INTRODUCTIONIn  2012,  an  estimated  241,740  prostate  cancer  diagnoses  will  be made  in  the  United  States,  accompanied  by  an  estimated  28,170 prostate  cancer  deaths.1  Beginning  around  1994-1996,  with  wide-spread use of serum prostate-specific antigen (PSA) testing and digital rectal  examination  (DRE)  for  prostate  cancer  screening,  and  with increased treatment of clinically localized prostate cancer with surgery or  radiation  therapy,  age-adjusted  prostate  cancer  death  rates  have fallen steadily. Although this trend might indicate a beneficial impact of  prostate  cancer  screening  and/or  early  prostate  cancer  treatment on prostate cancer mortality, mass screening of the general population for prostate cancer remains controversial.2 One challenge for prostate cancer screening is the prevalence of the disease in the United States: autopsy series have revealed small prostate cancers in as many as 29% of men between ages 30 and 40 and 64% of men between ages 60 and 70.3 Obviously, not all of these men are at risk for symptomatic or  life-threatening  prostate  cancer  progression.  In  fact,  many  such men,  if  diagnosed  with  prostate  cancer,  may  be  at  greater  risk  for treatment-associated morbidity.

Currently, for U.S. men, the lifetime risk of a diagnosis of prostate cancer is about 1 in 6, whereas the lifetime risk of death from prostate cancer is on the order of 1 in 35.1 Over the past two decades, treat-ment approaches for men with prostate cancer have changed dramati-cally, with improvement in established prostate cancer treatments and the introduction of new prostate cancer treatment approaches. Now, men diagnosed with prostate cancer often face a bewildering array of treatment  choices.  Clearly,  the  physicians  that  care  for  these  men must weigh the risks of prostate cancer progression against the poten-tial for side effects from treatment, in the context of other health risks and life choices, to best use the current collection of treatments for the greatest benefit.

To  aid  physicians  who  care  for  men  with  prostate  cancer,  this chapter will provide an overview of prostate cancer etiology, biology, screening, detection, diagnosis, prevention, and treatment.

PROSTATE ANATOMY AND FUNCTIONThe prostate is a male sex accessory gland that surrounds the urethra and contributes secretions to the ejaculate (Figure 84-1). Located in 

the  pelvis,  the  prostate  sits  adjacent  to  the  bladder  and  rectum,  is surrounded  incompletely  by  a  thin  capsule  composed  of  collagen, elastin, and smooth muscle, and at the apex of the gland, forms part of the urethral sphincter apparatus.4 Nerves to the corpora cavernosa of the penis, needed for penile erection, travel through fascia along the posterolateral surface of the prostate. These nerves can be recog-nized as a neurovascular bundle by urologists and preserved during radical  prostatectomy  to  minimize  sexual  dysfunction  postopera-tively.5,6 The prostate parenchyma has been divided into three zones that  can  be  seen  by  transrectal  ultrasonography,  and  recognized readily  by  surgical  pathologists  examining  radical  prostatectomy specimens:  a  central  zone,  surrounding  the  ejaculatory  ducts  and accounting for some 25% of the prostate; a transition zone, near the prostatic urethra with 10% of prostate tissue normally; and a periph-eral zone, with the bulk of prostate tissue encompassing the postero-lateral region of the prostate (Figure 84-2).7,8

In addition to prostate cancer, the prostate also frequently mani-fests benign enlargement (benign prostatic hyperplasia or BPH) and chronic or recurrent inflammation (prostatitis). Like prostate cancer, each of these conditions can elevate the PSA, confounding the use of serum PSA testing for prostate cancer screening. When present, BPH tends to be located near the prostatic urethra (in the transition zone), whereas  prostate  cancer  and  the  prostate  cancer  precursor  lesions proliferative inflammatory atrophy (PIA) and prostatic intraepithelial neoplasia (PIN) usually arise in the periphery (the peripheral zone). Prostatic  inflammation,  though  often  prominent  in  the  peripheral zone,  can  be  seen  throughout  the  prostate.  Remarkably,  although prostate cancer, BPH, and prostatitis all commonly afflict U.S. men, and can be simultaneously present in a single prostate gland, mecha-nistic  associations  of  the  three  diseases  have  been  difficult  to demonstrate.

The prostate requires androgenic hormones, and an intact andro-gen receptor,  for normal growth and development.  In  the prostate, the major circulating androgenic hormone, testosterone, produced by Leydig  cells  in  the  testes upon  stimulation by  luteinizing hormone (LH), is converted by 5α-reductase (nicotinamide-adenine dinucleo-tide  phosphate-dependent  δ4-3-ketosteroid  5α-oxidoreductase)  to 5α-dihydrotestosterone (DHT).9 DHT, a more potent androgen than testosterone, binds to intracellular androgen receptors, alters andro-gen receptor conformation to promote dissociation from chaperone 

Page 3: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1465ProstateCancer • CHAPTER84

androgen receptor, PSA, cytokeratins K8 and K18, prostate-specific membrane  antigen  (PSMA),  and prostate-specific  acid phosphatase (PAP); and rare neuroendocrine cells,  that secrete chromogranin A, neuron-specific enolase, and synaptophysin (Figure 84-3).13 The basal epithelial cell compartment likely contains pluripotent prostatic stem cells,  capable of  self-renewal proliferation and of differentiation.  In contrast,  columnar  secretory cells,  specialized  to produce  secretions for the ejaculate, are terminally differentiated, particularly under the influence of androgenic hormones. The prostate epithelium is in turn supported by  a  stroma containing fibroblasts,  smooth muscle  cells, nerves, and blood vessels. Stromal cells, which also express the andro-gen receptor, secrete polypeptide growth factors, such as keratinocyte growth factor (KGF), that contribute to the regulation of epithelial homeostasis  via  a  paracrine  signaling  mechanism.14,15  Abnormal 

proteins, triggers androgen receptor dimerization and transport into the cell nucleus, and activates the expression of selected target genes.10 Stereotypically,  androgen  receptor  target  genes  are  characterized  by the presence of  androgen  response  element  (ARE) DNA sequences within the transcriptional regulatory region, permitting direct binding and trans-activation by the androgen receptor.11 For genes like KLK3 (encoding PSA), which are activated by the androgen receptor selec-tively in prostate cells, and not in cells of other tissues, the transcrip-tional  regulatory  region  also  contains  additional  DNA  sequences (prostate-specific  enhancer  or  PSE)  conferring  prostate-specific expression.12

The  normal  prostate  epithelium  is  composed  of  basal  epithelial cells,  characterized  by  the  expression  of  cytokeratins  K5  and  K14, and  p63,  columnar  secretory  epithelial  cells,  which  express  the 

Figure 84-2 •  Zones of the prostate. The peripheral zone,  accounting  for 70% of  the prostate  gland,  is  the site  of  origin  of ≥70%  of  prostate  cancers;  the  central zone, approximately 25% of the prostate gland, gives rise to only 1% to 5% of prostate cancers; and the transition zone,  ~5%  to  10%  of  the  prostate  gland,  gives  rise  to 20% of prostate cancers and is the site of origin of benign prostatic hyperplasia (BPH). (From Green DR, Shabsign R,  Scardino  PT.  Urological  ultrasonography.  In: Walsh PC,  Rettic  AB,  Stamey  CA,  Vaughan  ED  Jr,  editors. Campbels’s Textbook  of  Urology.  6th  ed.  Philadelphia: Saunders; 1992.)

Transition zone

Central zone

Peripheral zone

Anterior fibromuscular stroma

Figure 84-3 •  The prostate epithelium. 

Smooth muscle cell

Epithelialcompartment

Epithelialcompartment DesmosomeMicrovilli

Golgiaparatus

Secretorygranule

Glandularlumen

Secretorycell

Nerveterminals

Basalcell

Fibroblast Capillary Basementmembrane

Neuroendocrinecell

Page 4: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1466

scavenger  receptor  capable  of  binding  bacterial  lipopolysaccharide and lipoteichoic acid, and oxidized high and low density serum lipo-proteins (oxidized HDL and LDL).29 For both RNASEL and MSR1, not only have mutations been linked to prostate cancer susceptibility in families but variant alleles have been predicted to account for as many  as  13%  and  3%  of  sporadic  prostate  cancer  cases,  respec-tively.27,30,31 The  identification of RNASEL  and MSR1  as  candidate prostate cancer susceptibility genes has intensified interest in the pos-sibility  that  infection and/or  inflammation might contribute  to  the pathogenesis of human prostate cancer. In mice, targeted disruption of  RnaseL  leads  to  diminished  interferon-α  activity  and  increased susceptibility  to  viral  infection,32  whereas  targeted  disruption  of Msr-A leads to increased vulnerability to infection with Listeria mono-cytogenes, Staphylococcus aureus, Escherichia coli, and Herpes simplex virus type 1.29,33-35 Further genetic support for this etiologic mecha-nism  has  come  from  analyses  of  common  variants  of  other  genes encoding  participants  in  host  inflammatory  responses,  including TLR4  and  other  members  of  toll-like  receptor  signaling  pathways, MIC-1, IL1-RN, and COX-2, have also been associated with prostate cancer risk.36-40

An increased risk for prostate cancer has long been known for men in breast cancer families carrying BRCA2 mutations, characterized by disease with an aggressive natural history arising before age 55 years.41 Nonetheless, a role for BRCA2 genotyping in general prostate cancer practice has not been established. However, as evidence has accumu-lated that several additional germline DNA sequence variants may be associated  with  prostate  cancer,  the  possibility  that  genetic  testing might be used to aid in prostate cancer screening, detection, diagnosis, or risk stratification has emerged. In one analysis, five such sequence variants, three single nucleotide polymorphisms (SNPs) at 8q24 and one each at 17q12 and 17q24.3, were found to have a marked asso-ciation with prostate cancer, especially for men with a family history of the disease, showing a 9.46-fold  increased risk (with 95% confi-dence interval of 3.62 to 24.72) for a prostate cancer diagnosis.42

Epidemiology of Prostate CancerAccumulated  epidemiological  evidence  implicates  the  environment as the major contributor to the development of most prostate cancers. Prostate  cancer  incidence  and  mortality  display  wide  geographic variation, with high rates of prostate cancer incidence and mortality in  the United States  and Western Europe,  and  low prostate  cancer risk more  characteristic of Asia.43 African Americans  in  the United States have very high prostate cancer risk.44 The geographic variation in prostate cancer incidence and mortality can best be explained by lifestyle influences, because Asian immigrants to North America typi-cally adopt a higher prostate cancer risk.45-47

The key aspect of lifestyle in the United States most likely respon-sible  for  high  prostate  cancer  incidence  and  mortality  is  the  diet, generally rich in animal fats and meats and poor in fruits and vege-tables.  In  the  Health  Professions  Follow-up  Study,  a  prospective cohort  study  involving  51,529  men,  total  fat  intake,  animal  fat intake, and consumption of red meats were associated with increased risks of prostate cancer development.48 Red meat consumption was similarly correlated with prostate cancer risks in the Physicians Health Study49 and in a large cohort study in Hawaii.50 The cooking of red meats at high temperatures, or on charcoal grills, is known to lead to the  formation  of  both  heterocyclic  aromatic  amine  and  polycyclic aromatic  hydrocarbon  carcinogens.51,52  Ingestion  of  2-amino-1-methyl-6-phenylimidazopyridine  (PhIP),  one  of  the  heterocyclic amine  carcinogens  that  appear  in  “well-done”  red  meats,  leads  to prostate cancer in rats.53 Consumption of dairy products also appears to increase prostate cancer risk, an effect that may be more attribut-able  to  calcium  intake  than  to dietary  fat or protein.54  In contrast, adequate consumption of vegetables and antioxidant micronutrients is  accompanied  by  reduced  prostate  cancer  risk.  Consumption  of tomatoes,  which  contain  lycopene,  and  of  cruciferous  vegetables, 

stromal–epithelial interactions, with disordered regulation of epithe-lial cell proliferation and differentiation, may contribute to the patho-genesis of both prostate cancer and BPH.16

Prostate cancer cells, and PIN cells, arise from the prostatic epi-thelium. Even though transformed, such cells typically retain many of the phenotypic attributes characteristic of differentiated columnar secretory cells,  including the expression of androgen receptor, PSA, PSMA, and PAP. Prostate cancers reminiscent of basal epithelial cells are  exceptionally  rare; prostate  cancers with  features of neuroendo-crine cells  are  somewhat more common.17 However, unlike normal columnar  epithelial  cells,  neoplastic  prostate  epithelial  cells  are capable of proliferation. This has  led to the concept that the target cell  for neoplastic  transformation  in the prostate may be an “inter-mediate” cell, in transit from a basal epithelial stem cell to a differenti-ated columnar secretory epithelial cell, with properties of both stem cells and differentiated cells.18,19 Another feature of neoplastic prostate epithelial cells, as compared with normal basal or columnar secretory cells,  is  that  the  neoplastic  cells  appear  to  use  androgen  receptor signaling  not  only  for  differentiation,  but  also  for  proliferation,  as most  prostate  cancer  cells  tend,  at  least  initially,  to  display  some dependence on androgens  for maintenance of growth and  survival. Somatic fusions between an androgen-regulated gene, TMPRSS2, at chromosome 21q22, and genes encoding members of the ETS family of  transcription  factors,  commonly  found  in  prostate  cancers,  may provide a mechanistic explanation by which androgen signaling can promote prostate cancer cell growth.20,21 Ultimately, in life-threatening prostate cancer, prostate cancer cells escape from the prostate gland, proliferate in lymph nodes, in bones, and in other organs and become less and less dependent on androgenic hormones.

GENETICS AND EPIDEMIOLOGY

Genetic Predisposition to Prostate CancerFamilial clusters of prostate cancer have been recognized since at least 1956, when Morganti et al. reported that men with prostate cancer were  more  likely  to  have  relatives  with  prostate  cancer  than  men without a prostate cancer diagnosis.22 In a study conducted more than three decades later, when detailed family histories were collected from men with prostate cancer and their  spouses,  the men with prostate cancer  were  more  likely  to  have  a  brother  or  father  with  prostate cancer.23 Twin studies, comparing the tendency for concordant pros-tate cancer development between monozygotic twins, sharing all of their genes, and dizygotic twins, sharing half of their genes, have also hinted at a significant contribution of hereditary to prostate cancer: in  a  study  of  44,788  pairs  of  twins  in  Sweden,  Denmark,  and Finland,24 42% of the prostate cancer cases (with a 95% confidence interval  of  29%  to  50%)  were  attributed  to  heredity.  In  principle, familial clustering of prostate cancer cases could be a result of inher-ited susceptibility genes, shared exposure to carcinogenic stresses, or to some sort of detection or diagnosis bias (e.g., the brother of a man diagnosed with prostate cancer may be more likely to pursue screen-ing  for  prostate  cancer).  To  discover  the  genetic  contributions  to prostate cancer in spite of these potential sources of bias, both linkage analyses  of  genetic  loci  in  high-risk  prostate  cancer  families  and genomewide association studies (GWAS) of prostate cancer suscepti-bility in large populations have been conducted. Each approach has generated a number of candidate genes or gene regions (as many as 30 or more), with several  in common, confirming the contribution of heredity to prostate cancer risk but underscoring the complexity of inherited prostate cancer susceptibility.25

Among  the  growing  number  of  prostate  cancer  susceptibility genes  discovered  by  mapping  studies  are  RNASEL  and  MSR1.26,27 RNASEL  encodes  a  latent  endoribonuclease  component  of  an interferon-inducible  2’,5’-oligoadenylate-dependent  RNA  decay pathway that functions to degrade viral and cellular RNA upon viral infection.28 MSR1 encodes subunits of a trimeric class A macrophage 

Page 5: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1467ProstateCancer • CHAPTER84

Areas of the prostate, containing epithelial cells that do not fully dif-ferentiate into columnar secretory cells, have long been recognized as focal atrophy lesions by prostate pathologists.58,67 The term PIA has been used to describe those focal atrophy lesions that contain prolif-erating epithelial cells, are associated with chronic inflammation, and are often  located adjacent  to PIN  lesions and/or prostate cancers.68 The  epithelial  cells  in  PIA  lesions  typically  express  high  levels  of stress-response polypeptides such as GSTP1, GSTA1, and cyclooxy-genase  2  (COX-2).  Loss  of  GSTP1  expression  in  rare  PIA  lesions, attributable to de novo GSTP1 CpG island hypermethylation, may be what leads to the development of PIN and prostate cancer.69

The hypothesis that inflammation might promote prostate cancer development offers new challenges to prostate cancer epidemiology, to  the  search  for  prostate  cancer  susceptibility  genes,  and  to  the molecular pathogenesis of prostate cancer. Although prostatic inflam-mation is common in regions of the world with high prostate cancer risks, whether regions of the world with low prostate cancer risks have less prostatic inflammation has not been determined. Provocative new data hint that PIA lesions appear to be more common in the prostate peripheral zone in men from high-risk prostate cancer regions than in men from low-risk prostate cancer regions. To better test the asso-ciation  of  prostate  inflammation  and  prostate  cancer,  perhaps  new biomarkers  of  prostate  inflammation,  assayable  in  blood,  urine,  or prostate  fluid,  can  be  developed  for  use  in  epidemiology  studies.  In  addition,  polymorphic  genes  encoding  regulators  of  immune responses will likely need to be systematically evaluated for prostate cancer risk associations. Ultimately, if the hypothesis is correct, pros-tate cancer risks may be reduced by therapeutically attenuating pros-tate inflammation.

ETIOLOGICAL AND BIOLOGICAL CHARACTERISTICS

Somatic Genome Alterations in Prostate Cancer CellsProstate cancer cells typically contain a plethora of somatic genome alterations, including gene mutations, gene deletions, gene amplifica-tions, chromosomal rearrangements, and changes in DNA methyla-tion  (Figure  84-5).  In  the  United  States,  prostate  cancer  diagnoses are typically made in men between 60 and 70 years of age, whereas small prostate cancers have been detected at autopsy in nearly 30% of men between 30 and 40 years of age.3 Thus, the somatic genome changes  present  in  prostate  cancers  have  often  accumulated  over many  decades. The  acquisition  of  somatic  genome  changes  in  the prostate may be influenced by lifestyle as well: although small prostate 

which  contain  sulforaphane,  may  protect  against  prostate  cancer development.55,56 Antioxidant micronutrients, such as vitamin E and selenium,  may  reduce  prostate  cancer  risk  only  when  correcting dietary deficiencies: a large trial (SELECT) of supplementation with vitamin E and selenium to prevent prostate cancer failed to show a benefit.57

Prostate Inflammation and Prostate CancerChronic or recurrent inflammation is known to play a causative role in the development of many human cancers, including cancers of the liver, esophagus, stomach, large intestine, and bladder. Inflammatory changes  have  been  recognized  in  prostate  tissues  for  many  years, leading to speculation that  inflammation might contribute in some way  to  prostate  cancer  development.58  However,  over  the  past  few years, evidence has accumulated in support of a more critical role for prostatic inflammation in the pathogenesis of prostate cancer. Inflam-matory changes are present in almost all radical prostatectomy speci-mens  from men with prostate cancer. Because  inflammation  in  the prostate  is not usually associated with symptoms,  the prevalence of prostate inflammation is not known, and the association with pros-tate  cancer  has  been  difficult  to  test.59,60  A  syndrome  of  irritative voiding  symptoms  and pelvic  pain,  perhaps  attributable  to  inflam-mation near the prostatic urethra, is reported by some 9% or more of men between 40 and 79 years of age, with as many as 50% of such men suffering more than one episode by age 80 years.61 Most episodes of  symptomatic  prostatitis  are  not  clearly  attributable  to  specific infectious agents. Even so, sexually transmitted infections do appear to increase prostate cancer risk.62,63 Nonetheless, if prostate infection and  inflammation  lead to prostate cancer,  the mechanism does not appear  likely  to  involve  direct  transformation  of  prostate  epithelial cells  by  microbial  DNA.  Instead,  the  production  of  microbicidal oxidants by inflammatory cells, such as superoxide, nitric oxide, and peroxynitrite, may promote prostate cancer development by trigger-ing cell and genome damage.64,65 Increased production of oxidants by inflammatory  cells  in  the  prostate  may  be  why  decreased  prostate cancer risk has been associated with intake of a variety of antioxidants or  of  nonsteroidal  antiinflammatory  drugs,  and  why  RNASEL  and MSR1, two of the prostate cancer susceptibility genes identified thus far, encode proteins that function in host responses to infections.

Despite  these  provocative  hints,  the  contribution  of  prostate inflammation to prostatic carcinogenesis has been difficult to assess. However,  in  1999,  De  Marzo  et al.  provided  the  most  compelling linkage of prostate inflammation to prostate cancer by proposing that a prostate  lesion,  termed proliferative  inflammatory  atrophy  (PIA), might be a precursor to PIN and to prostate cancer (Figure 84-4).66 

Figure 84-4 •  Proliferative  inflammatory  atrophy (PIA) as a precursor to prostatic intraepithelial neoplasia (PIN) and prostate cancer. (Adapted from Nelson WG, DeMarzo AM, Isaacs WB: Prostate cancer. N Engl J Med 2003;349:366-381.)

Normalprostate

Proliferativeinflammatory

atrophy

Prostaticintraepithelial

neoplasia

Prostatecancer

Columnarcells

Basalcells

Inflammatorycells

Page 6: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1468

dependence  of  prostate  cancer  cells  on  androgenic  hormones  for growth  and  survival,  as  the  expression of ETS  family  transcription factors can be stimulated by androgen action. TMPRSS2-ERG fusions have been detected in ~60% of prostate cancers and in >20% of PIN lesions.81 ERG is highly expressed by many prostate cancers, and not at all by others, though neither ERG expression nor the presence of ERG fusion transcripts appears to have great prognostic significance.81-84 Nonetheless, the presence of TMPRSS2-ERG fusion transcripts does seem to define a molecular subset of prostate cancer. Rarer somatic alterations,  including  SPOP  mutations  (6%  to  15%)  and  SPINK1 overexpression (~10%), are restricted to cases devoid of TMPRSS2-ETS family rearrangement rearrangements.85,86

Hypermethylation  of  CpG  island  sequences  encompassing  the regulatory  region  of  GSTP1,  encoding  the  π–class  glutathione S-transferase (GST) is the most common somatic genome change yet reported for prostate cancer.87,88 GSTs catalyze the detoxification of carcinogens, and of other reactive chemical species, via conjugation with the intracellular scavenger glutathione. In mice, targeted disrup-tion of π-class GST genes leads to increased skin tumors after treat-ment with the carcinogen 7,12-dimethylbenzanthracene (DMBA).90 Similarly, human prostate cancer cells devoid of GSTP1 appear espe-cially vulnerable to genome damage mediated by exposure to N-OH-PhIP, the charred meat carcinogen that causes prostate cancer when fed to rats, and by exposure to oxidant stresses.91 In the normal pros-tate epithelium, GSTP1 is present at high levels in basal cells, and in lower  levels  in columnar  secretory cells,  though the enzyme can be induced in columnar epithelial cells subjected to genome-damaging stresses.  In contrast,  the enzyme is almost never present  in prostate cancer cells. In nearly all cases, the absence of GSTP1 expression in prostate cancer cells can be attributed to hypermethylation of GSTP1 CpG  island  sequences,  a  somatic  genome  change  that  prevents GSTP1  transcription.  Absence  of  GSTP1  expression  and  GSTP1 CpG island hypermethylation may also be characteristic of cells com-prising  PIN  lesions,  thought  to  be  precursors  to  prostate  cancer.92 The  mechanism  by  which  hypermethylated  GSTP1  CpG  island alleles arise during prostatic carcinogenesis remains to be elucidated. Nonetheless, prostate cells carrying inactivated GSTP1 genes appear to  enjoy  some  sort  of  selective  growth  advantage  early  during  the development of prostate cancer.

cancers  have  been  detected  at  autopsy  in  men  from  geographic regions  with  low  prostate  cancer  mortality,  these  small  prostate cancers  are  usually  only  present  in  much  older  men.70-72  Also,  in the  United  States,  prostates  removed  at  radical  prostatectomy  for prostate cancer usually contain more than one prostate cancer lesion (Figure 84-6).

Over  the  years,  several  techniques  have  been  used  to  catalog genome  accidents  in  prostate  cancer  cells,  including  karyotyping, fluorescence  in  situ  hybridization  (FISH),  comparative  genome hybridization,  loss  of  heterozygosity  analyses,  and  genomewide microarray  and/or  sequencing  approaches.  In  one  such  study,  each prostate cancer case exhibited a mean of 3866 base mutations (range 3192  to 5865), 20 nonsilent coding  sequence mutations  (range 13 to 43), and 108 rearrangements (range 43 to 213).73 In another, DNA hypermethylation  was  found  at  5408  regions  of  the  genome,  with 73% of the regions near genes (5′, 3′, or intron–exon junctions), and 27% of the regions at conserved intergenic sites.74 Often, these analy-ses  reveal  different  chromosomal  abnormalities  in  different  cancer cases, in different cancer lesions in the same cancer case, and in dif-ferent areas within the same cancer lesion. The propensity to develop such  a heterogeneous  collection of  somatic  genome  lesions  over  so many years, and in a manner so sensitive to environment and lifestyle, suggests  strongly  that prostate  cancers  likely arise  as  a  consequence of either chronic or recurrent exposure to genome-damaging stresses, defective protection against genome damage, or  some combination of both processes. The resultant genomic instability may be the reason some prostate cancers progress to threaten life.75,76

One  characteristic  somatic  genome  alteration,  a  rearrangement, drives  the  production  of  fusion  transcripts  between  an  androgen-regulated gene, TMPRSS2, at chromosome 21q22, and members of the ETS family of transcription factors (Figure 84-7).20 Fusion part-ners for TMPRSS2 include ERG (also at chromosome 21q22), ETV1 (at  chromosome  7p21),  and  ETV4  (at  chromosome  17q21).20,77,78 The gene rearrangements may result from a mishap during the andro-gen  receptor  transcriptional  trans-activation of TMPRSS2  in which tangling  or  untangling  of  DNA  by TOP2B  triggers  DNA  double-strand  breaks  that  recombine  with  ETS  family  partner  genes  via nonhomologous end-joining (Figure 84-8).79,80 The appearance of the resultant  fusion  transcripts  provides  a  plausible  mechanism  for  the 

Figure 84-5 •  The molecular pathogenesis of prostate cancer. 

Normalprostate

epithelium

Proliferativeinflammatory

atrophy

Prostaticintraepithelial

neoplasia

Localizedprostatecancer

Metastaticprostatecancer

Androgenindependent

cancer

Germ line mutationsRNASEL, MSR1, HOXB13

Chromosome8q gain

Chromosome8p loss

TMPRSS2-ETS familygene fusions

Loss of sequences10q, 13q, 16q

Gains of sequencesat 7p, 7q, Xq

TP53 mutation

hAR gene mutation/amplification

Loss of PTEN

Decrease in NKX3.1

Decrease in p27

GSTP1 CpG islandhypermethylation

Page 7: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1469ProstateCancer • CHAPTER84

NKX3.1 encodes a prostate-specific homeobox gene essential for normal prostate development  that may be a  target  for  somatic  loss on chromosome 8p21.93 NKX3.1 has been shown to bind DNA and to  repress  PSA  expression  via  interactions  with  ETS  transcription factors.94,95 Mice carrying one or two disrupted Nkx3.1 alleles mani-fest prostatic epithelial hyperplasia and dysplasia.96,97 In men, loss of 8p21  DNA  sequences  occurs  early  during  prostatic  carcinogenesis, with 63% of PIN lesions and >90% of prostate cancers, showing loss of  heterozygosity  at  polymorphic  8p21  marker  sequences  in  one report.98  However,  although  mapping  studies  have  indicated  that NKX3.1  lies within  a  common  region of deletion,  encompassing 2 megabases at 8p21, molecular pathology analyses have not yet estab-lished NKX3.1  as  a  somatic  target  for  inactivation during prostatic carcinogenesis  because  somatic  NKX3.1  mutations  have  not  been identified.  Nonetheless,  loss  of  NKX3.1  expression  does  appear  to accompany prostate cancer progression.

PTEN,  a  tumor  suppressor  gene  encoding  a  phosphatase  active against both proteins and lipid substrates, appears to be a common target for somatic alteration during prostate cancer progression.99-106 PTEN  is  an  inhibitor of  the phosphatidylinositol 3′-kinase/protein kinase B (PI3K/Akt) signaling pathway needed for cell cycle progres-sion and cell survival. Although PTEN is expressed by normal pros-tate epithelial cells, and by cells present in PIN lesions, the expression of PTEN is often diminished in prostate cancers, with many prostate cancers containing collections of neoplastic cells with no PTEN.107 

Figure 84-6 •  Multiple  foci of prostate  cancer,  and of prostate  cancer precursor lesions, in the peripheral zone of the prostate. (From Nelson WG, DeMarzo  AM,  Isaacs  WB,  et al.  Prostate  cancer.  N  Engl  J  Med 2003;349:366-381.)

B

A2 cm

C

Transitionzone

Peripheralzone

CarcinomaHigh-grade prostaticintraepithelial neoplasiaAtrophy

Figure 84-7 •  Fusion transcripts and other genetic alterations in prostate cancers.  The  schematic  shows  gene  fusions  involving  ETS  transcription factors and RAF kinase. Also shown are genetic alterations in two other genes (SPINK1 overexpression and SPOP mutations), which are mutually exclusive with ETS rearrangements. The estimated frequencies of each genetic altera-tion  are  also  included.  (From  MacMillan  Publishers  Ltd:  Asian  Journal  of Andrology 2012;14(3):393-9.)

TMPRSS21 4 5-9 10

TMPRSS2

TMPRSS2

TMPRSS2

HERPUD1

SLC45A3

SLC45A3

SLC45A3

SLC45A3

ESRP1 CRAF

CRAFRAF

ETS

SPINK1

SPOP Mutations

NDRG1

ERG

1 4 5–9 10ERG

1 4 5–9 10ERG ERG 45%

ETV1 5%

ETV4 1%

ETV5 1%

RAF 1%

SPINK1 8%SPOP 13%

1 4 5–9 10ERG

1 4 5–11 12ETV1

1 5 6–11 12ETV1

1

1 2–10 10–1611 6 7 178 912 13

1 1 2-12 13ETV4

1 8 9 10 11-17 18

2 3-12 13ETV5

ETV58-12 13

1 2 3

1 2–4 14–155 16

Page 8: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1470

p27 levels appear complex: somatic loss of DNA CDKN1B sequences at 12p12-13 have been reported for only 23% of  localized prostate cancers, 30% of prostate cancer lymph node metastases, and 47% of distant  prostate  cancer  metastases.113  In  place  of  CDKN1B  gene alterations, p27 polypeptide levels may be lowered indirectly by inad-equate PTEN repression of the PI3K/Akt signaling pathway.114-116 In this way,  low p27  levels may be  as much a  result of  loss of PTEN function  as  of  CDKN1B  alterations.  The  critical  contribution  of PTEN  to  epithelial  growth  regulation  in  the prostate  is  evident  in mice, where disruption of Cdkn1b alleles leads to prostatic hyperpla-sia, and Pten+/–Cdkn1b–/– mice develop prostate cancer by 3 months of age.111,117

Metastatic prostate cancer is almost always treated with androgen deprivation, antiandrogens, or a combination of androgen deprivation and antiandrogens.118,119 However, despite such treatment, androgen-independent prostate cancer cells eventually emerge and progress to threaten life. Curiously,  in these cells, androgen receptor expression and androgen receptor signaling remain intact despite the absence of androgens.120,121  Somatic  alterations  of  AR  have  been  reported  for many prostate cancers, especially for androgen-independent prostate 

PTEN defects have been found in a wide variety of cancers and cancer cell lines.101 For prostate cancer, a number of somatic PTEN altera-tions  have  been  reported,  including  homozygous  deletions,  loss  of heterozygosity,  mutations,  and  probable  CpG  island  hypermethyl-ation.  However,  despite  common  losses  of  10q  sequences  near  PTEN in prostate cancers, somatic mutations at the remaining PTEN alleles  are not  as  frequent.  In  a  study of prostate  cancer metastases recovered  at  autopsy,  somatic  PTEN  alterations  were  even  more common than in primary prostate cancers, and a significant hetero-geneity  in  PTEN  defects  in  different  metastatic  deposits  from  the same patient was also evident.106 Haploinsufficiency for PTEN may contribute  to  the  phenotype  of  transformed  cells  in  the  prostate. Pten+/– mice display prostatic hyperplasia and dysplasia, and crosses of Pten+/– mice with Nkx3.1+/– mice have revealed that Pten+/–Nkx3.1+/– mice and Pten+/–Nkx3.1–/– mice develop lesions reminiscent of human PIN.108-110

Defective regulation of p27, a cyclin-dependent kinase inhibitor encoded  by  CDKN1B,  may  also  accompany  prostatic  carcinogene-sis.111,112 In PIN cells and prostate cancer cells, p27 levels are almost always  diminished,  though  the  mechanism(s)  for  the  reduction  in 

Figure 84-8 •  Initiation of  transcription by  androgen  receptor  leads  to DNA strand breaks mediated by TOP2B, which  recombine  to generate  gene rearrangements.  (From  Haffner  MC  et al. Transcription-induced  DNA  double  strand  breaks:  both  oncogenic  force  and  potential  therapeutic  target?  Clin Cancer Res 2011;17:3858-3864.)

Persistent DSBsSenescenceApoptosis

Illegitimate repairFormation of structural rearrangements

Geonomic instability

Repair, movement totransaction hubs,

and active transcription

ROSFork collisionDietary carcinogensRepair infidelity

ATM

DNA-PK

POLII

TOP28

TOP28

DNA double strand break

Ku80Ku70

DHT

DHT

DHT

DHT

ARAR

AR

PARP1 Otherrepair

proteins?

AR

Page 9: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1471ProstateCancer • CHAPTER84

both in PIN lesions and in prostate cancer.155-157 At some point, most cancer  cells  activate  the  expression  of  telomerase,  providing  some maintenance of chromosome termini. Telomerase expression has been detected in prostate cancers, but not at high levels in normal prostate tissues or in BPH.155

PROSTATE CANCER SCREENING, EARLY DETECTION, AND PREVENTION

Clinical EvaluationWith widespread opportunistic prostate-specific antigen (PSA)–based screening for prostate cancer, most cases are diagnosed at a stage when symptoms due to the disease are absent. Nevertheless, locally advanced disease can cause lower urinary tract symptoms (obstructive or irrita-tive),  hematospermia,  a  decrease  in  ejaculate  volume,  and  erectile dysfunction if the cancer extends beyond the prostate to involve the bladder neck, urethra, ejaculatory ducts, seminal vesicles, or erecto-genic  nerves,  respectively.  Metastatic  involvement  of  the  skeleton, bone marrow, pelvic lymph nodes, or periureteral lymphatics can lead to bone pain, pancytopenia,  lower extremity edema, or retroperito-neal  fibrosis,  respectively.  Paraneoplastic  syndromes  from  ectopic hormone production by small cell variants of adenocarcinoma, and disseminated  intravascular  coagulation  (DIC)  have  been  associated with metastatic prostate cancer.

Digital Rectal ExaminationIn men with early-stage prostate cancers, physical findings, if present, are usually limited to an abnormal DRE, used for both screening and staging. Palpable areas of induration, or asymmetric firmness of the gland, suggest the presence of prostate cancer, but these findings can also  be  caused  by  prostate  inflammation  (especially  granulomatous prostatitis), by benign prostatic hyperplasia (BPH), and by prostatic stones. DRE has only fair reproducibility in the hands of experienced examiners.158 When used alone for detection of prostate cancer, DRE misses from 23% to 45% of the cancers that are subsequently detected by prostate biopsies done for serum PSA elevations or for transrectal ultrasound (TRUS) abnormalities.159-161 In addition, prostate cancers detected by DRE are at an advanced pathological stage in more than 50% of men.162,163

The positive predictive value of DRE (the  fraction of men who have  prostate  cancer  if  the  DRE  is  abnormal)  ranged  from  4%  to 11%  in  men  with  PSA  levels  from  0.0  to  2.9 ng/mL,  and  from 33%  to  83%  in  men  with  PSA  levels  of  3.0  to  9.9 ng/mL  or more.164 When DRE and PSA are used in prostate cancer screening, detection  rates  are  higher  with  PSA  than  with  DRE  and  highest with both tests  together.165 Furthermore, a DRE abnormality tends to  be  associated  with  the  presence  of  high-grade  cancer.166  Thus DRE  and PSA  are  generally  considered  to be  complementary  tests and  a  prostate  biopsy  is  usually  recommended  for  men  with  an abnormality on DRE that is suspicious for prostate cancer regardless of  the PSA  level.

Serum PSAPSA is a member of the human kallikrein gene family of serine pro-teases encoded by KLK3 located on chromosome 19.167 A component of the ejaculate, PSA is produced by columnar secretory cells in the prostate. PSA expression is regulated by androgens, becoming detect-able  in  serum  at  puberty  accompanying  increases  in  luteinizing hormone and testosterone. In the absence of prostate cancer, serum PSA  levels  increase with age and prostate volume and are generally higher in African Americans. Cross-sectional population data suggest that  the  serum PSA  increases 4% per milliliter of prostate volume, and that 30% and 5% of the variance in PSA can be accounted for by prostate volume and age, respectively.168

cancers. AR  amplification, accompanied by high-level expression of androgen receptors, may promote the growth of androgen-independent prostate cancer cells by  increasing the sensitivity of  the cells  to  low androgen levels.122 AR mutations, encoding androgen receptors with altered  ligand  specificity  have  also  been  detected;  for  some  of  the mutant  androgen  receptors,  even  antiandrogens  can  act  as  agonist ligands.123-125  When  44  mutant  androgen  receptors  from  prostate cancers were evaluated for transcriptional regulatory capabilities, 16% of  the  receptors had  lost  transcriptional  activation  activity, 45% of the receptors had gained some transcriptional regulatory ability, 32% of the receptors maintained some partial transcriptional modulatory activity,  and  the  remaining 7% behaved  like wild-type  receptors.126 In addition to somatic AR gene changes, androgen-independent pros-tate  cancer  cells  with  wild-type  androgen  receptors  may  activate androgen  receptor  signaling  even  in  the  absence  of  androgens,  via posttranslational  modifications  of  the  androgen  receptor  and/or androgen  receptor  coactivators  in  response  to  other  growth  factor signaling pathways.120,127-130

Changes in Gene Expression in Prostate CancersAlterations  in  gene  expression  in  prostate  cancers  have  been  cata-logued using cDNA microarray technologies.131-142 Among the many genes  exhibiting  over-  or  underexpression  in  prostate  cancers,  the products of at least two genes appear consistently increased, and the product of a third gene appears to become elevated during androgen-independent progression. Hepsin,  located at 19q11-13.2,  encodes a transmembrane  serine  protease,  expressed  at  high  levels  in  many normal tissues.143 Hepsin may contribute to prostate cancer progres-sion:  forced  overexpression  of  hepsin  in  mouse  prostates  leads  to disorganization of the epithelial basement membrane and increased metastasis.144 α-Methylacyl-CoA  racemase  (AMACR),  a  mitochon-drial and peroxisomal enzyme that acts on pristanoyl-CoA and C27-bile  acyl-CoA  substrates  to  catalyze  the  conversion  of  R-  to S-stereoisomers  in  order  to  permit  metabolism  by β-oxidation,  has been reported to be overexpressed in almost all prostate cancers.145,146 Germline  AMACR  mutations  lead  to  adult-onset  neuropathy.147 Immunohistochemistry studies, which have revealed that AMACR is occasionally present in normal prostate cells, increased in PIN cells, and further elevated in prostate cancer cells, have prompted the use of antibodies against AMACR as tools for prostate cancer diagnosis by surgical pathologists.146,148 The polycomb protein enhancer of zeste homolog 2 (EZH2), a transcriptional regulatory protein, is elevated in metastatic androgen-independent prostate cancer.149 The mecha-nism by which EZH2 contributes to prostate cancer progression has not been established. However, elevated EZH2 expression in primary prostate cancers portends a poor prognosis.149

Telomere Shortening During Prostatic CarcinogenesisTelomeres, containing repeat DNA sequences at the termini of chro-mosomes,  protect  against  loss  of  chromosome  sequences  during genome replication. DNA ends tend to shorten each generation as a consequence  of  bidirectional  DNA  synthesis  (the  “end-replication” problem);  the  telomere  repeat  sequences  serve  as  templates  for  the enzyme telomerase, which can extend the chromosome termini and maintain chromosome integrity through cell division.150 Growth dys-regulation  accompanying  the  development  of  most  human  cancers tends to  lead to cell proliferation in the absence of telomerase, and to shortened chromosome telomeres.151 Critically shortened telomere sequences may promote genome instability by increasing illegitimate DNA recombination.152,153 Mice carrying disrupted genes needed for a  functioning  telomerase  show  increased numbers of  cancers,  espe-cially when crossed to mice with defective p53 genes.154 In the pros-tate,  short  telomere  repeat  sequences  appear  characteristic  of  cells 

Page 10: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1472

appropriately  thought  of  as  measures  of  a  continuum  of  prostate cancer risk, with the risk of cancer (and of high-grade cancer) rising directly with the PSA level, an observation first made by Gann et al. using  a  cohort  population  study  design.171  In  recent  years,  more emphasis has been placed on integrating several variables associated with the risk of prostate cancer (e.g., age, family history, race, DRE findings, prior biopsy results) to construct risk calculators for estimat-ing  the  need  for  a  prostate  biopsy.177,178 The  future  will  likely  see further integration of genome and other molecular biomarker tools in prostate cancer screening algorithms.

PSA “Density”The major  source of  serum PSA in men without prostate cancer  is the  transition  zone  (TZ)  epithelium,  not  the  epithelium  of  the peripheral zone.179 Because benign prostate enlargement is a growth 

Serum PSA elevations likely occur as a result of disruption of the normal  prostate  architecture,  permitting  PSA  to  diffuse  into  the prostate  parenchyma  and  gain  access  to  the  circulation.  This  can occur in the setting of both benign and malignant prostate diseases (prostatitis,  BPH,  and  prostate  cancer)  and  as  a  result  of  prostate manipulation (prostate massage and prostate biopsy).169 Although the presence  of  some  type  of  prostate  disease  is  the  most  important determinant driving elevation of the serum PSA, an increased serum PSA is not specific for prostate cancer. Furthermore, not all men with prostate disease have elevated serum PSA levels.

Treatments targeting the prostate gland (for BPH or for prostate cancer) can lower serum PSA by decreasing the number of prostatic epithelial  cells  capable  of  producing  PSA,  and  by  decreasing  the amount  of  PSA  produced  by  each  cell.  Modulation  of  sex  steroid hormone  levels  for  treatment  of  BPH  or  prostate  cancer,  radiation therapy  for  prostate  cancer,  and  surgical  ablation  of  prostate  tissue for BPH or prostate cancer can all  lead to decreases  in serum PSA. 5α-Reductase inhibitors, like finasteride and dutasteride, lower PSA levels by 50% after 12 months of treatment.170 Thus for men treated with  these  agents  for  12  months  or  more,  the  serum  PSA  level  should be doubled to estimate  the “true” PSA value.  Interpretation of serum PSA values should always take into account the presence of prostate disease, previous diagnostic procedures, and prostate-targeted treatments.

Serum PSA and Prostate Cancer DetectionThe  serum  PSA  value,  along  with  the  findings  at  DRE,  correlate directly with the risk of prostate cancer at biopsy (Table 84-1). Fur-thermore, the PSA level at a young age anticipates the future risk of being  diagnosed  with  prostate  cancer  decades  later  (Figure  84-9). Gann et al. first showed that the risk of a prostate cancer diagnosis, including the diagnosis of life-threatening disease, incrementally and directly with PSA over the decade after a baseline measurement, even at  low PSA  levels  (below 4.0 ng/mL);  a finding  that has now been confirmed  by  many  others.171-173  These  observations  may  allow  a more targeted approach to prostate cancer screening using a baseline PSA value to direct screening intensity.174,175

In  the early years of PSA testing, most clinicians used PSA as a dichotomous test (i.e., the PSA was “elevated” or not) with elevations triggering a prostate biopsy. However, data from the placebo arm of the Prostate Cancer Prevention Trial (PCPT176), which accrued men with a serum PSA value <3.0 ng/mL, has suggested that there is no PSA cut-off  level with both high sensitivity and high specificity for prostate cancer, and that virtually no level is low enough to exclude the presence of prostate cancer. Instead, serum PSA tests may be more 

Figure 84-9 •  Distribution of bone metastases in men dying of prostate cancer. (Data from Roudier MP et al. Phenotypic heterogeneity of end-stage prostate carcinoma metastatic to bone. Hum Pathol 2003;34:646-653.)

L humerus 64%

Ribs 80%

T8 64%T9 80%T10 80%T11 75%T12 88%

L iliac 92%

L femur 80%

R humerus 75%

Sternum 83%

L1 90%L2 93%L3 85%L4 92%L5 100%

R iliac 100%

R femur 80%

14 AUTOPSY PROSTATECANCER CASES

Table 84-1 Positive Predictive Value of DRE and PSA in a Multicenter Screening Trial

DRE PSA PPV (%)

Abnormal Any 21.4

Any >4 31.54–10 26.1>10 52.9

Normal >4 24.4

Abnormal <4 10.04–10 40.8>10 69.1

Data from Catalona WJ, Richie JP, Ahmann FR, et al. Comparison of digital rectal examination and serum prostate specific antigen in the early detection of prostate cancer: results of a multicenter clinical trial of 6,630 men. J Urol 1994;151:1283.

DRE, Digital rectal examination; PPV, positive predictive value; PSA, prostate-specific antigen.

Page 11: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1473ProstateCancer • CHAPTER84

using one assay for cPSA versus two assays to determine the %fPSA (measuring both tPSA and the %fPSA).

Benign PSA (BPSA), a degraded isoform of fPSA, is preferentially found in nodular tissue involved in benign prostatic enlargement and has  been  shown  to  be  elevated  among  men  with  benign  prostatic hyperplasia.194 A second isoform, proPSA, is an inactive precursor of PSA that contains differing leader sequences of amino acids. When compared to men without prostate cancer, the tissues and serum of prostate cancer patients have an increased proportion of proPSA that is fPSA, a difference that has been used to improve the discrimination of  men  with  and  without  prostate  cancer.195  A  Beckman  Coulter “Prostate  Health  Index”  (PHI)  uses  proPSA,  fPSA,  and  tPSA  in  a formula (proPSA/fPSA) × (tPSA)1/2 that was shown to improve pros-tate cancer detection over total and %fPSA.196 Additionally, there is some  evidence  that  proPSA  is  associated  with  a  more  aggressive prostate cancer phenotype.197

Urine Biomarkers of Prostate CancerPCA-3, a noncoding prostate-specific RNA, was found to be overex-pressed  in prostate cancer  tissue compared with benign  tissue.198,199 Urine assays for PCA3 have been developed that appear to improve the specificity of prostate cancer detection in response to an abnormal serum PSA,200 When compared with %fPSA, a PCA3 score (the ratio of PCA3 RNA to PSA mRNA × 1000) has been shown to be a supe-rior predictor of prostate cancer on a repeat biopsy when an  initial biopsy was negative.201 Although PCA3 may have a higher specificity when  used  for  prostate  cancer  detection,  urine  testing  for  PCA3 exhibits a lower sensitivity when compared with serum PSA testing.

A number of other urinary molecular biomarkers have been evalu-ated for prostate cancer detection, including hypermethylated GSTP1 genes, AMACR proteins,  and TMPRSS2-ERG  fusion mRNA tran-scripts.  In  a  multicenter  evaluation  of  urinary  DNA  methylation markers, including GSTP1, the negative predictive value of the urine test for prostate cancer diagnosis was 87% for men with a PSA of 4.1 to 10.0 ng/mL.202 Although such DNA methylation urine tests might spare men unnecessary  repeat  biopsies,  the  addition of TMPRSS2-ERG mRNA urine testing (expected to be present in half of men with prostate cancer) to PCA3 RNA urine testing may provide a positive predictive value as high as 95%,  steering men who harbor prostate cancer despite a previous negative biopsy toward a definitive diagno-sis.199,203 The discovery and development of new molecular biomark-ers  is  progressing  very  quickly;  in  the  future,  such  biomarkers  (or biomarker panels) will need to be incorporated into algorithms along with  standard  measures  of  risk  (e.g.,  age,  family  history,  race)  to selectively  identify men who should undergo  further evaluation  for the presence of prostate cancer.204

TRUS-Guided Prostate BiopsyTRUS is not an accurate method for localizing early prostate cancer and  is not  recommended  for use  in prostate  cancer  screening. The primary role of TRUS in prostate cancer detection and diagnosis  is to ensure accurate sampling of prostate tissue by prostate biopsies in men suspected of harboring cancer based on serum PSA  levels and DRE.205  This  is  best  accomplished  by  targeting  peripheral  zone lesions that appear hypoechoic by TRUS for biopsy, along with per-forming  systematic  sampling  biopsies  of  areas  without  hypoechoic lesions in the prostate periphery.

TRUS-guided prostate biopsies are performed routinely with an 18-guage  needle  fired  from  a  spring-loaded  gun  through  a  port mounted on the TRUS probe. Most commonly, in preparation for a biopsy procedure, men are administered a fluoroquinolone antibiotic and given a cleansing enema. Injection of a local anesthetic around the  periphery  of  the  prostate  is  used  by  most  urologists  to  reduce discomfort  associated  with  prostate  biopsy.  Major  complications, such as bleeding and/or infection requiring hospitalization, are rare, 

of the transition zone tissue, and because serum PSA levels are largely a reflection of  transition zone volume  in men with benign prostate enlargement,  adjusting  the  serum  PSA  for  either  prostate  volume (PSA density) or more specifically, the transition zone volume (PSA-TZ), has been shown to improve the distinction between those with prostate  cancer  and  those  with  benign  prostate  enlargement.180-183 PSA density is also directly associated with the presence of high-grade cancer on prostate biopsy, and is therefore useful for identifying those men less likely to have high-grade cancer most appropriate for active surveillance.184

PSA “Velocity”Carter et al., using frozen serum samples from an aging study cohort, demonstrated  that  a  PSA  “velocity”  (PSAV;  the  rate  of  change  of serum PSA in ng/mL per year upon repeated testing) of 0.75 ng/mL per year or greater had a specificity of 90% for distinguishing men with prostate cancer in the setting of benign prostatic enlargement, and a specificity of 100% for distinguishing men with prostate cancer in the absence of benign prostatic enlargement when PSA levels were in the range of 4.0 to 10.0 ng/mL (Figure 84-9).185

The PSAV tends to be higher in men with high-grade and high-stage prostate cancer as compared to men with lower grade and stage disease.186,187 Furthermore, men with a PSAV above 2.0 ng/mL per year in the year before a diagnosis of prostate cancer appear to be at an increased risk of prostate cancer death after surgical intervention when compared with men with a PSA velocity of 2.0 ng/mL per year or less.188 For PSAV measured 10 to 15 years before prostate cancer diagnosis (when most men had PSA levels below 4.0 ng/mL), a PSAV less  than  0.35 ng/mL  per  year  has  been  associated  with  a  prostate cancer–specific survival 25 years later of 92%, compared with 54% with  a  PSAV  greater  than  0.35 ng/mL  per  year.189 There  is  also  a direct relationship between the number of times the PSAV exceeds a given threshold (e.g., 0.4 ng/mL per year), a concept known as “risk count,” and the risk of high-grade cancer.190,191 Thus a continuously increasing serum PSA should raise the suspicion of high-grade pros-tate cancer if there is no other explanation for the rise.

Molecular Forms of PSAPSA  in  the  bloodstream  circulates  in  both  bound  and  unbound forms. Most  of  the detectable PSA  in  the  serum  (65%  to 90%)  is bound to α1-antichymotrypsin (“complexed” PSA or cPSA, whereas the  rest  (10%  to  35%)  remains  unbound  (“free”  PSA  or  fPSA).167 The  serum PSA  tests  in  common use  for prostate  cancer detection and  monitoring  effectively  measure  the  sum  of  fPSA  and  cPSA, providing a determination of  the “total”  serum PSA (tPSA). Assays that  can  distinguish  fPSA  and  cPSA  have  been  developed  and approved by the U.S. FDA specifically for use in the early detection of prostate cancer. In general, men with prostate cancer have a greater fraction of serum tPSA bound to α1-antichymotrypsin (cPSA), and a  commensurately  lower  fraction  of  tPSA  that  is  unbound  (fPSA), than men without prostate cancer. This difference  is  thought  to be due  to  the  differential  expression  of  PSA  isoforms  by  cells  in  the transition zone (the zone of origin for BPH) tissue as compared with peripheral zone (the zone where most prostate cancers arise) tissue.

The %fPSA appears most useful in distinguishing men with and without prostate cancer in the setting of tPSA levels between 2 and 10 ng/mL.192  An  fPSA  of  25%  and  a  PSA  density  of  0.078  were shown to have comparable specificity for prostate cancer diagnosis at biopsy (at a sensitivity of 95%), but %fPSA does not require a TRUS for determination.193 Most urologists use %fPSA determinations for decisions about the need for a repeat biopsy in a man with a persis-tently  elevated  serum  PSA  and  previous  negative  prostate  biopsies, where the possibility of a missed prostate cancer may be a concern. New assays for cPSA exhibit comparable specificity and sensitivity for prostate cancer diagnosis as %fPSA, with the potential advantage of 

Page 12: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1474

compelling reasons for prostate cancer prevention. In addition, epi-demiological data, indicating a dominant role for lifestyle factors in prostate cancer development, suggest that prostate cancer risk modi-fication may be feasible, if only through lifestyle modification. Also, because prostatic carcinogenesis takes many decades, there may be a broad window of opportunity to change lifestyle in an effort to retard prostate  cancer development. Clearly,  although  the  specific  lifestyle factors fostering prostate cancer development have not been conclu-sively identified, it is likely that consumption of a diet rich in fruits, vegetables, and antioxidant micronutrients, and poor in saturated fats and “well-done” red meats, may significantly reduce risks of prostate cancer development, and of the development of other diseases char-acteristic of  life  in the developed world. As the etiology of prostate cancer  is  even  better  understood,  new  opportunities  for  prostate cancer prevention may also arise. For example, if prostate inflamma-tion  contributes  to  prostate  cancer  development,  antiinflammatory drugs  might  be  considered  candidate  prostate  cancer  prevention drugs.219  For  drugs  to  be  developed  and  tested  for  prostate  cancer prevention, randomized clinical trials, capable of assessing both drug safety and drug efficacy, will be required.220 Ideally, such trials can be targeted at men with a high risk for prostate cancer development.

Thus far, two classes of agents, 5α-reductase inhibitors and anti-oxidant  micronutrients,  have  been  subjected  to  large  randomized clinical trials; neither class of agents has shown a convincing prostate cancer  prevention  benefit.  In  the  Prostate  Cancer  Prevention Trial (PCPT), the propensity for the 5α-reductase inhibitor finasteride to reduce the prevalence of prostate cancer in healthy men age 55 years and older when given for 7 years was tested.221 For the trial, men (n = 18,882) with a PSA of 3.0 ng/mL or less and a normal DRE were randomized to treatment with finasteride (5 mg/d) or to placebo.221 While  on  study,  men  with  a  PSA  elevation  or  an  abnormal  rectal examination  were  subjected  to  prostate  biopsy;  in  addition,  at  the end of the treatment period, a prostate biopsy was planned for all of the men  in  the  trial. Prostate  cancer was detected  in 18.4% of  the men treated with finasteride versus 24.4% of men receiving placebo (P < 0.001).221 However, high-grade prostate cancers appeared more commonly  associated  with  finasteride  treatment  than  with  placebo (6.4% vs. 5.1%).221 Similarly, in the REDUCE trial, men (n = 6729) with an elevated serum PSA (2.5 to 10.0 ng/mL) and a recent nega-tive  prostate  biopsy  were  randomized  to  receive  the  5α-reductase inhibitor  dutasteride  (0.5 mg/d)  or  placebo.222  For  this  trial,  men underwent prostate biopsies after 2 and 4 years of treatment. With a result  reminiscent  of  PCPT,  overall  prostate  cancer  detection  was reduced  from  25.1%  to  19.9%  for  men  receiving  dutasteride,  but high-grade prostate cancer, which was similar  in the first 2 years of the trial for men treated with dutasteride versus placebo, was increased in dutasteride-treated men in the last 2 years of the trial from <0.1% to  0.5%.  The  findings  of  PCPT  and  REDUCE  may  mean  that 5α-reductase inhibitors prevent or treat low-grade cancers better than high-grade cancers. These mixed results, a reduction in overall pros-tate cancer prevalence but an increase in high-grade prostate cancers, make  prescribing  5α-reductase  inhibitors  to  healthy  men  for  the purpose of preventing prostate cancer very problematic, an opinion shared by the FDA, which has not approved either drug for such an indication.223

Epidemiological  studies have provided compelling evidence  that intake of selenium and of vitamin E might diminish prostate cancer risks, especially in the setting of inadequate dietary consumption.224-230 As  a  result,  a  prospective,  randomized,  placebo-controlled  clinical trial of selenium and vitamin E (SELECT; n = 35,533) was under-taken to test the ability of the antioxidant micronutrients to prevent prostate cancer.231 Selenium (200 µg  selenomethionine), α–tocoph-erol  (400 mg),  the  combination  of  selenium  and α–tocopherol,  or neither, were given  to men  randomized  to  four different  treatment groups using a 2 × 2 factorial design, for 7 to 12 years.231 The men studied  were  age  55  years  or  older  (50  years  or  older  for  African Americans) with an unremarkable DRE and a serum PSA of 

although hematuria and hematospermia are common sequelae of the procedure. More  recently, a  rise  in fluoroquinolone-resistant E. coli strains  have  been  implicated  in  higher  rates  of  hospitalization  for  postbiopsy  sepsis,206  leading  to  more  liberal  use  of  targeted antibiotic  prophylaxis,  based  on  rectal  swab  cultures  done  prior  to  a biopsy.207

The optimal biopsy technique, including the number and place-ment of biopsies for tissue procurement that will minimize the chance of missing a relevant cancer remains controversial. Nonetheless, the best  evidence  available  suggests  that  biopsies  placed  more  laterally within  the  peripheral  zone  of  the  prostate  may  be  important  to exclude prostate cancer in men with elevated serum PSA values and a nonsuspicious DRE. Magnetic resonance imaging (MRI) may have a role in improving the specificity of prostate cancer detection and in identification of sites within the prostate for directing prostate biop-sies.208  At  this  point,  it  is  not  clear  whether  this  use  of  MRI  will improve health outcomes or reduce (vs. increase) the costs of care.

Screening for Prostate CancerOpportunistic  PSA-based  screening  has  been  widespread  in  the United States since the early 1990s. It has been estimated that 45% to 70% of the 30% decline in prostate cancer mortality that occurred in  the 1990s could be attributed to  the  stage migration and earlier treatment of prostate cancer associated with PSA testing.209 Ecological studies of prostate cancer mortality comparing countries with differ-ent rates of PSA uptake, suggest that the detection of more aggressive disease with PSA testing could explain mortality differences.210

Two randomized trials of prostate cancer screening reported dif-ferent results in 2009.211,212 In the Prostate, Lung, Colon, and Ovary (PLCO)  trial,  no  difference  in  mortality  was  found  between  the screening and control groups, whereas in the European Randomized Study of Screening for Prostate Cancer (ERSPC), there was a 21% reduction in prostate cancer deaths in men randomized to screening at a median follow-up of 9 years. However, the absolute reduction in deaths for men screened was only 0.71 per 1000 men, with 48 men diagnosed  with  prostate  cancer  and  treated  to  prevent  one  death, raising a concern about overdiagnosis and overtreatment even despite the mortality reduction. Updated reports from both trials have shown few changes with longer follow-up.213,214 With these conflicting data, as might be expected, champions of prostate cancer screening dismiss the results of the PLCO trial because of high contamination rates in the control arm, high rates of prescreening  in  the population prior to initiation of the trial, and low rates of follow-up for positive screen-ing tests. Nevertheless,  the findings of the PLCO trial do hint that in the setting of ongoing widespread screening as is seen in the United States, further intensification of screening would not likely improve health outcomes.215

The  worry  about  overdiagnosis  and  overtreatment  of  prostate cancer in the United States has led many to reconsider the value of serum PSA testing for prostate cancer screening. A draft recommen-dation on prostate cancer screening from the U.S. Preventive Services Task Force (USPSTF) was released in the fall of 2011 recommending against prostate cancer screening for men of all ages, regardless of race or  family  history.216  These  USPSTF  guidelines  may  prompt  an increase in shared decision making between patients and physicians prior to pursuing prostate cancer screening,217 a more nuanced and targeted approach to screening for those who desire to pursue  it,175 and  an  individualized  approach  to  prostate  cancer  treatment  that includes active surveillance if a diagnosis of prostate cancer is made rather than the rush to curative intervention for all.218

Prostate Cancer PreventionThe high lifetime risks of prostate cancer development, the morbidi-ties associated with treatment of established prostate cancer, and the inability to eradicate life-threatening metastatic prostate cancer offer 

Page 13: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1475ProstateCancer • CHAPTER84

of HGPIN in a prostate biopsy in the apparent absence of prostate cancer is that prostate cancer may have been missed by the prostate sampling strategy used for the biopsy procedure. On repeat biopsy of men  with  HGPIN  on  initial  biopsy,  the  cancer  detection  rate  was 20% among men who underwent standard extended biopsy sampling (14 cores or  less), and 31% among men who underwent a “satura-tion”  biopsy  sampling  (20  biopsy  cores  or  more).242 The  National Comprehensive  Cancer  Network  (NCCN)  recommends  a  repeat biopsy within 1 year for men with HGPIN if the lesion was multifo-cal  (2  cores  or  more  involved)  and/or  the  initial  biopsy  was  not performed using an extended approach (more than 6 cores).243

Of  interest,  unlike  prostate  cancer,  prostate  inflammation,  or BPH, HGPIN lesions are not thought to perturb prostate architec-ture enough to elevate the serum PSA. Recently, increasing attention has  been  afforded  the  notion  that  PIA  (proliferative  inflammatory atrophy) lesions might be precursors to PIN and/or prostate cancer.244 Like  PIN,  PIA  lesions  tend  to  arise  in  the  peripheral  zone  of  the prostate,  where  prostate  cancers  arise,  and  some  PIA  cells  acquire somatic genome alterations reminiscent of prostate cancer cells.66,245 At present, men with PIA  lesions  are not  subjected  to  any kind of treatment,  and presence of PIA on an  initial prostate biopsy  is not thought to predict the detection of prostate cancer on repeat biopsy. The major significance of PIA to the diagnosis of prostate cancer by prostate  biopsy  may  be  propensity  for  such  lesions  to  occasionally exhibit features that mimic prostate cancer.234

The most frequently used approach to histologic grading of pros-tate  cancer  is  the  application  of  Gleason  scoring.246  The  Gleason grade  refers  to  architectural  prostate  cancer  patterns,  numbered  1 (well  differentiated)  to  5  (poorly  differentiated).  Because  prostate cancers are often heterogeneous, Gleason scoring (sometimes referred to as the “combined” Gleason grade) is accomplished by adding the Gleason grade of the most abundant pattern to the Gleason grade of the second most abundant pattern (e.g., a Gleason score of 4 + 3 = 7). The Gleason score, when applied by an expert pathologist, is one of the best tools available for predicting cancer-specific outcomes; the higher  the Gleason  score  the  greater  the  risk  of  cancer  progression with  or  without  treatment.247  Of  note,  because  Gleason  scoring applies pattern grades to the architecture of cancer within the pros-tate,  metastatic  prostate  cancers  detected  by  biopsies  of  metastatic deposits are not assigned a Gleason score. Gleason grading practices have  evolved over  time. As part of  a 2005  International Society of Urological  Pathology  (ISUP)  Consensus  Conference,  a  revised approach was  formulated  and adopted,  subtly  changing  the defini-tions  of  Gleason  patterns  3  and  4  and  improving  the  interrater reproducibility of Gleason grading.248 With the newer ISUP grading strategy, a Gleason score of 3 + 3 = 6 can be expected to have a very low risk of prostate cancer recurrence after primary therapy.

Life-Threatening Prostate Cancer ProgressionProstate  cancer  progression  has  long  been  understood  to  involve metastases  to bones, which are often painful,  and  to  lymph nodes. The  predilection  of  the  disease  for  bones  may  reflect  a  hospitable microenvironment capable of providing growth  factor  support  in a collaboratively malignant milieu (Figure 84-10).249 In turn, prostate cancer  cells  perturb bone homeostasis  in  such  a way  as  to  increase osteoblast activity, which can be readily discerned by bone scanning. This  activity  forms  the  basis  for  bone  scan  imaging,  such  as  with 99mTc-methylene  diphosphonate  (MDP)  and  18F-sodium  fluoride (NaF), in which osteoblastic activity can be visualized throughout the entire  skeleton.250  Reportedly,  NaF  positron  emission  tomography/computed tomography (NaF PET/CT), in which osteoblastic activity is  coregistered  with  bone  radiography,  offers  the  highest  sensitivity and specificity for bone metastasis.251

More recent autopsy studies have provided newer insights into the phenotype(s) of lethal prostate cancer. In addition to expected bone and  lymph  node  metastases,  men  dying  of  prostate  cancer  tend  to 

4.0 ng/mL  or  less.231  Unfortunately,  the  trial  failed  to  show  any reduction in prostate cancer and showed a slight increase in prostate cancer  incidence  among  men  taking α–tocopherol.57,232 When  this result is considered in the context of the findings from epidemiology studies and previous  smaller clinical  trials,  the data  tend  to  suggest that  men  with  low  blood  levels  of  the  antioxidant  micronutrients, who tend to be at the highest risk for prostate cancer development, might be the only men with any chance for benefit from supplemen-tation.225,226,230,233 For this reason, correction of antioxidant micronu-trient  deficiencies  might  ultimately  prove  more  generally  safe  and effective than widespread supplementation (and its  risk of oversup-plementation), which may carry a threat of harm with little benefit.

PATHOLOGY AND PATHWAYS OF SPREAD

Histopathology of Prostate CancerMicroscopic  analysis  of  prostate  tissue  by  a  surgical  pathologist  is needed for the diagnosis of prostate cancer, for determining prostate cancer stage after prostatectomy, and for histologic grading, via  the assignment  of  a  Gleason  score,  to  predict  the  behavior  of  prostate cancer. The majority of prostate cancers are adenocarcinomas, though other types of cancers can appear. Most often, prostate cancer diag-noses  are  made  using  core-needle  biopsy  specimens,  which  sample small  amounts  of  prostate  tissue.  For  many  prostate  cancer  cases, needle biopsies contain only small numbers of prostate cancer cells among  more  plentiful  noncancerous  glands.  Also,  several  prostate conditions,  including  acute,  chronic,  or  granulomatous  prostate inflammation, epithelial atrophy, and PIN, exhibit histologic features that mimic some of those present in prostate cancers.234 Thus, pros-tate cancers can be difficult to recognize in needle biopsy specimens, and difficult to distinguish from other prostate abnormalities. Expe-rienced prostate pathologists use a combination of architectural, cyto-logic,  and ancillary findings  to make a diagnosis of prostate cancer on needle biopsy.234-236 In addition, because normal prostate glands, but  not  glands  present  in  prostate  cancers,  contain  basal  epithelial cells, immunohistochemical staining for basal epithelial cell markers, such as cytokeratins K5 and K14 and the nuclear marker p63, can be used to help distinguish benign from malignant glands in prostate tissue samples. Immunohistochemical staining for AMACR, a pros-tate  cancer  biomarker  discovered  through  cDNA  microarray  tran-scriptome  profiling,  can  aid  in  prostate  cancer  diagnosis.146,148,237 Neither  of  these  immunohistochemistry  reagents  perfectly  distin-guishes prostate cancer: the absence of basal epithelial cell markers is not always diagnostic of prostate cancer, and AMACR expression is absent in some prostate cancers and present in PIN.234 For all of these reasons,  second-opinion  interpretations of prostate biopsy findings, especially when foci of atypical glands suspicious for cancer are identi-fied, are often helpful.

High-grade PIN (HGPIN), a lesion characterized by the prolifera-tion of malignant-appearing prostate epithelial cells within the con-fines of otherwise normal glandular structures, is identified in about 5%  of  men  subjected  to  prostate  biopsies.238,239 The  evidence  that HGPIN is a likely precursor to prostate cancer includes the findings that  (1)  HGPIN  is  more  commonly  present  in  prostates  that  also contain prostate cancer,  (2) HGPIN and prostate cancer both tend to arise in the peripheral zone of the prostate and are often directly contiguous,  and  (iii)  HGPIN  and  prostate  cancer  express  similar biomarkers and share many somatic genome abnormalities.240,241 The notion that HGPIN lesions might be prostate cancer precursors has stimulated interest in possibly treating men with HGPIN to prevent prostate  cancer.241  Unfortunately,  the  natural  history  of  individual HGPIN lesions is not known. Furthermore, HGPIN lesions, which can only be recognized by sampling prostate biopsies, are not easily monitored. These limitations have hindered the use of HGPIN as a response surrogate for cancer prevention drug development. Because HGPIN is not currently treated, the major significance of the finding 

Page 14: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1476

Figure 84-10 •  Modified  Gleason  grading  system  for  prostate  cancer. Gleason patterns 1 to 5 are assigned different patterns of growth for prostatic adenocarcinomas.  A  Gleason  score  is  the  sum  of  the  two  most  common Gleason grades (e.g., Gleason 4 + 3 = 7). (From Epstein JI, An update of the Gleason grading system. J Urol 2010;183:433-440.)

1

2

3

4

5

exhibit  common  metastases  to  liver  (64%),  the  dura  (43%),  and elsewhere.252 The most troubling feature of lethal prostate cancer may be its heterogeneity. In different metastases from the same case, the expression of androgen receptor, and its target PSA, appears markedly variable, present in the majority of cells  in some lesions but mostly absent  for  others.253 This  likely  reflects  ongoing  genome  and  epig-enome  instability.  Genomewide  assessment  of  somatic  genetic  and epigenetic  defects  in  lethal  prostate  cancers  at  autopsy  show  that although disseminated metastatic disease emerges from a single clone, new defects continue to appear in individual metastases.254 Of inter-est,  DNA  hypermethylation  changes  tend  to  be  maintained  in  all metastatic lesions in most cases, at least as much as any of the genetic alterations, whereas DNA hypomethylation, leading to activation of repressed  embryonic  genes,  continues  to  evolve  variably  lesion  to lesion.255 Of course, the greatest worry about ongoing genomic, epi-genomic, and phenotypic heterogeneity in metastatic prostate cancer is  the  likelihood  that  treatment  resistance  will  almost  certainly emerge, regardless of the agent used.

CLINICAL MANIFESTATIONS/PATIENT EVALUATION/STAGING

Evaluation of the Extent of Prostate CancerThe extent of prostate cancer is correlated with clinical tumor stage (Table 84-2), Gleason score (the sum of two Gleason grades; Figure 84-11), and serum PSA level.256 Nomograms, incorporating clinical stage (estimated using DRE findings), the serum PSA level, and the Gleason  score,  have  been  shown  to  be  capable  of  predicting  both pathological prostate cancer extent determined on radical prostatec-tomy  specimens  and  the  long-term  outcome  following  primary tumor  treatment.257-259 The most  commonly used  risk  stratification scheme today is that of D’Amico et al., who suggested that men with prostate cancer can be stratified into low risk (stages T1c to 2a, and serum  PSA  less  than  10 ng/mL,  and  Gleason  score  of  6  or  less), intermediate risk (stage T2b, or serum PSA between 10 and 20 ng/mL, or Gleason score of 7), and high risk (stage T2c, or serum PSA greater than 20 ng/mL, or Gleason score of 8 or greater) groups. The fraction of men free of prostate cancer 10 years after radical prosta-tectomy is significantly different for the risk categories: 83% of men with  low-risk  prostate  cancer,  46%  of  men  with  intermediate-risk prostate cancer, and 29% of men with high-risk prostate cancer.259

A number of pretreatment tools are available for prediction of the pathological  extent  of  tumors  and  the  probability  of  biochemical recurrence.260,261 When  used  before  initiating  prostate  cancer  treat-ment, risk stratification of men with prostate cancer aids in counsel-ing  such  men  about  the  expected  outcome  of  prostate  cancer treatment,  providing  estimates  of  the  chance  that  local  treatments might be curative.

Radiographic ImagingAlthough CT scanning is used routinely by radiation oncologists for prostate cancer treatment planning, no imaging technique available today  has  been  proven  to  add  additional  useful  information  when used to evaluate the extent of prostate cancer in men with low- and intermediate-risk disease.262 TRUS and MRI give the most accurate definition of prostatic architecture and anatomy, but current imaging technologies do not provide very precise assessments of cancer extent within  the  prostate  or  the  presence  of  microscopic  foci  of  prostate cancer that have escaped the confines of  the prostate gland. Radio-nuclide bone scans detect metastatic prostate cancer in less than 1% of men with a serum PSA value less than or equal to 20 ng/mL and are not recommended for the initial evaluation of men with low- or intermediate-risk prostate cancer.263 PET has not yet been found to be useful in the evaluation of men with prostate cancer and has no place  in  the  prostate  cancer  staging.264  New  imaging  technologies, including three-dimensional color Doppler, contrast-enhanced color Doppler, magnetic resonance spectroscopy, and high-resolution MRI with magnetic nanoparticles, have great potential for improving the assessment of local and distant prostate cancer extent.262,265,266 Cross-sectional imaging of the pelvis, by CT scan or MRI, for the purpose of detecting lymph node metastases, and radionuclide bone scans for the detection of bony metastases,  should be  reserved  for men with high-risk prostate cancer.

111In-capromab  pendetide,  a  radioimmunoconjugate  featuring  a monoclonal antibody to an intracellular domain of prostate-specific membrane antigen was approved by the FDA for use in the evaluation of  men  with  clinically  localized  prostate  cancer  but  is  rarely  used today  for  assessment  of  prostate  cancer  extent,  even  for  men  with high-risk prostate cancer, in large part because of low scan sensitivity and frequent difficulties in scan interpretation. There may be a role for  111In-capromab pendetide  imaging with  single photon emission computed tomography used in coregistration with computed tomog-raphy (SPECT/CT) for prognostic staging.267 New nuclear medicine agents  capable  of  detecting  PSMA  appear  promising  in  early 

Page 15: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1477ProstateCancer • CHAPTER84

mRNAs or prostate cancer DNAs, polymerase chain reaction (PCR) approaches, capable of astonishing sensitivity, are typically used. Cir-culating  tumor  cells  (CTCs)  appear  early  in  the  course  of  prostate cancer, though simple enumeration of the cells does not provide clear prognostic information for those with localized cancers.276,277 Further, the  presence  of  CTCs  at  diagnosis  does  not  accurately  predict  the development of metastatic disease posttreatment.

PRIMARY THERAPY

Selection of Treatment ApproachMen  thought  to  have  localized  prostate  cancer  face  a  number  of management  options,  including  observational  strategies  (watchful waiting, active surveillance), radical prostatectomy, interstitial brachy-therapy,  and  external  beam  radiation  therapy.  Furthermore,  cryo-therapy, high-intensity focused ultrasound (HIFU), and focal therapy are  also  being  presented  to  some  patients  as  treatment  alternatives despite  limited  long-term  data  to  support  their  use.  A  number  of issues should be considered by patients and physicians before decid-ing on a management strategy. These include the specific side effect profile of different treatments, the possibility that certain comorbid conditions  might  increase  the  severity  of  side  effects  or  the  risk  of complications,  the characteristics of the cancer and the risk of pro-gression  without  treatment,  patient  comorbidities  that  might  limit the need for treatment, and individual patient preferences.278

development.250 Perhaps, one of these tools may overcome the inad-equacies of 111In-capromab.

Serum Biomarker Assays and “Molecular” Classification for PrognosisThe  serum  PSA  level  at  the  time  of  a  prostate  cancer  diagnosis  is correlated with both intermediate and long-term outcomes including tumor  volume,  stage,  grade,  and  freedom  from  relapse  after  treat-ment.268 However,  the  serum PSA cannot be used  alone  to predict disease extent for an individual patient. Instead, PSA values are more typically used as part of multivariable models for predicting the extent of disease and the probability that a cancer will behave in an aggressive manner.269,270 The NCCN recommends the use of the D’Amico clas-sification scheme using serum PSA, clinical stage, and biopsy Gleason score to stratify men into low-, intermediate-, and high-risk groups for selecting management strategies and for counseling patients about prognosis.271 Other PSA  forms used  in  the detection and diagnosis of prostate cancer, such as %fPSA, proPSA, and so forth, are not part of commonly employed risk stratification and prognosis tools.

“Molecular” classification of tumor cells for prognosis refers to the detection  of  circulating  prostate  cancer  cells  and/or  cell  fragments, either  indirectly, by  identifying mRNA species,  like those encoding PSA or PSMA, characteristically expressed by epithelial cells from the prostate,272 or directly, by recovering “circulating” prostate cancer cells by  a  variety  of  isolation  methods.273-275  To  detect  prostate  lineage 

Table 84-2 TNM Staging for Prostate Cancer (from http://www.cancerstaging.org)

PRIMARY TUMOR (T)

CLINICAL

TX Primary tumor cannot be assessed

T0 No evidence of primary tumor

T1 Clinically inapparent tumor neither palpable nor visible by imaging

T1a Tumor incidental histologic finding in 5% or less of tissue resected

T1b Tumor incidental histologic finding in more than 5% of tissue resected

T1c Tumor identified by needle biopsy (e.g., because of elevated PSA)

T2 Tumor confined within prostate1

T2a Tumor involves one-half of one lobe or less

T2b Tumor involves more than one-half of one lobe but not both lobes

T2c Tumor involves both lobes

T3 Tumor extends through the prostate capsule2

T3a Extracapsular extension (unilateral or bilateral)

T3b Tumor invades seminal vesicle(s)

T4 Tumor is fixed or invades adjacent structures other than seminal vesicles, such as external sphincter, rectum, bladder, levator muscles, and/or pelvic wall

PATHOLOGICAL (PT)3

pT2 Organ confined

pT2a Unilateral, one-half of one side or less

1Tumor found in one or both lobes by needle biopsy, but not palpable or reliably visible by imaging, is classified as T1c.

3There is no pathological T1 classification.4Positive surgical margin should be indicated by an R1 descriptor (residual microscopic disease).5When more than one site of metastasis is present, the most advanced category is used. pM1c is most advanced.

pT2b Unilateral, involving more than one-half of side but not both sides

pT2c Bilateral knee

pT3 Extraprostatic extension

pT3a Extraprostatic extension or microscopic invasion of bladder neck4

pT3b Seminal vesicle invasion

REGIONAL LYMPH NODES (N)

CLINICAL

NX Regional lymph nodes were not assessed

N0 No regional lymph node metastasis

N1 Metastases in regional lymph node(s)

PATHOLOGIC

pNX Regional nodes not sampled

pN0 No positive regional nodes

pN1 Metastases in regional node(s)

DISTANT METASTASIS (M)5

M0 No distant metastasis

M1 Distant metastasis

M1a Nonregional lymph node(s)

M1b Bone(s)

M1c Other site(s) with or without bone disease

2Invasion into the prostatic apex or into (but not beyond) the prostatic capsule is classified not as T3 but as T2.

Page 16: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1478

Observational StrategiesWatchful waiting was  a  common  strategy  in  the pre-PSA  era when most cancers were detected at an incurable stage and the morbidity of  primary  treatments  was  high.  The  primary  intent  of  watchful waiting was  to  avoid  treatment until  symptoms  emerged  and non-curative palliative treatment might be necessary.279 In contrast, active surveillance, a more modern approach, involves the selective delayed curative  treatment of men  found  to have disease progression while being carefully monitored. Active surveillance implies a much more diligent monitoring regimen than watchful waiting.

The  Scandinavian  Prostate  Cancer  Group  Study  4  (SPCG-4),  a  trial  conducted  before  the  PSA  screening  had  been  adopted, 

During the past two decades, both surgery and radiation therapy for  prostate  cancer  have  improved  dramatically,  providing  effective local  control of  cancer  in  the prostate while  reducing  the  threat of side  effects. The  absence  of  unbiased  comparative  studies  makes  it difficult for patients to choose an approach that offers a clear advan-tage in terms of disease-free survival or quality of life outcome. Thus patient  preferences  should  play  a  large  role  in  a  shared  decision process between patient and physician.

The  NCCN  guidelines  for  the  management  of  clinically  local-ized  prostate  cancer  reflect  the  needed  consideration  of  both  life expectancy  and  the  risk  profile  of  the  individual  patient’s  cancer in  making  management  recommendations  (http://www.nccn.org; see  Table  84-3).271

Figure 84-11 •  Results of a randomized trial of radical prostatectomy versus watchful waiting. (From Bill-Axelson A et al. Radical prostatectomy versus watchful waiting in early prostate cancer. N Engl J Med 2011;364:1708-1717.)

0.1

Pro

babi

lity

0.6

0.40.5

0.30.2

0.00 3 6 9 12 15

Years

0.1P

roba

bilit

y

0.6

0.40.5

0.30.2

0.00 3 6 9 12 15

Years

0.1

Pro

babi

lity

0.60.7

0.40.5

0.30.2

0.00 3 6 9 12 15

Years

0.1

Pro

babi

lity

0.60.7

0.40.5

0.30.2

0.00 3 6 9 12 15

Years

0.1

Pro

babi

lity

0.6

0.40.5

0.30.2

0.00 3 6 9 12 15

157166

154157

145144

136118

11591

6754

Years

A

Death from any cause, total cohort

P = 0.007 by Gray’s test

B

Death from prostate cancer, total cohort

P = 0.01 by Gray’s test

0.1

C

P = 0.89 by Gray’s test

D

P = 0.41 by Gray’s test

Pro

babi

lity

0.6

0.40.5

0.30.2

0.00 3 6 9 12 15

157166

154157

145144

136118

11591

6754

190182

185177

166162

135133

99101

4242

190182

185177

166162

135133

99101

4242

YearsE

Death from any cause, men <65 yr of age

P < 0.001 by Gray’s test

F

Death from prostate cancer, men <65 yr of age

Death from any cause, men ≥65 yr of age Death from prostate cancer, men ≥65 yr of age

P = 0.008 by Gray’s test

Radicalprostatectomy

Watchful waiting

Watchfulwaiting

No. at riskRadical prostatectomyWatchful waiting

No. at riskRadical prostatectomyWatchful waiting

No. at riskRadical prostatectomyWatchful waiting

No. at riskRadical prostatectomyWatchful waiting

347348

339334

311306

271251

214192

10996

347348

339334

311306

271251

214192

10996

No. at riskRadical prostatectomyWatchful waiting

No. at riskRadical prostatectomyWatchful waiting

Radicalprostatectomy

Watchfulwaiting

Radicalprostatectomy

Watchfulwaiting

WatchfulwaitingRadical

prostatectomy

Radicalprostatectomy

Radicalprostatectomy

Watchfulwaiting

Page 17: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1479ProstateCancer • CHAPTER84

Table 84-3 Management Options for Localized Prostate Cancer by Risk Profile and Life Expectancy

Risk Profile Criteria Management Option

Very low Stage T1c, and Gleason score ≤6, and PSA <10 ng/mL, and <3 biopsy cores with cancer, ≤50% involvement of any core with cancer, PSA density <0.15 ng/mL/cm3

Active surveillance if life expectancy <20 years

Low Stage T1c or T2a, and Gleason score ≤6, and PSA <10 ng/mL

Active surveillance if life expectancy <10 years; radiation (external beam or brachytherapy) or surgery if life expectancy ≥10 years

Intermediate Stage T2b or T2c, or Gleason score 7, or PSA 10–20 ng/mL

Active surveillance or external radiation with/without brachytherapy or surgery if life expectancy <10 yearsSurgery or external radiation with/without ADT with/without brachytherapy if life expectancy ≥10 years

High Stage T3a, or Gleason score 8–10, or PSA >20 ng/mL External radiation + ADT or surgery

Adapted from Carter HB. Management of low (favourable)-risk prostate cancer. BJU Int 2011;108:1684-1695.ADT, androgen deprivation therapy.

Figure 84-12 •  Intensity-modulated radiation therapy (IMRT). A, Representative axial CT slice  from a man with  low-risk prostate cancer  treated by using a seven-field IMRT plan. The isodose distribution is displayed. Dark inner line, prescription isodose curve. B, The IMRT treatment plan shown as a dose/”volume histogram: the curves from left to right represent bladder, rectum, and prostate. 

0 1000 2000 3000 4000 5000 6000 7000 8000

BA Dose (cGy)

Dose-volume histogramN

orm

al v

olum

e

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

randomized men (n = 695) with localized prostate cancer to surgery or watchful waiting (Figure 84-12).280 Although these men would not have  been  considered  optimal  candidates  for  active  surveillance  by current  criteria,  as most of  the men had palpable disease by DRE, surgery  was  accompanied  by  a  40%  reduction  in  prostate  cancer deaths for men with 15 men needing treatment to prevent a single prostate cancer death. Of note, the benefits of surgery were restricted to men under the age of 65 years. The Prostate cancer Intervention Versus Observation Trial (PIVOT), a study conducted in the setting of common PSA screening, randomized men (n = 731) with a mean age of 67 years to surgery or watchful waiting.281 The distribution of prostate  cancer  cases  among  the  study  subjects  included 43%  low-risk, 36% intermediate-risk, and 20% high-risk disease. These data, which  have  been  presented  but  not  yet  published,  suggest  that through 12 years of follow-up there has been no difference in cancer-specific survival attributable to treatment. Thus older men with low-risk,  screen-detected, prostate cancers may not derive great benefits from radical prostatectomy.

Active surveillance seeks to individualize prostate cancer care pri-marily among men thought to harbor low-grade prostate cancers who would  otherwise  be  fit  for  curative  treatment.  The  intent  is  to 

intervene early with curative therapy if the disease should progress,279 while sparing men with more indolent prostate cancer the side effects of  aggressive  primary  therapy.282,283  This  management  approach, which relies on diligent monitoring for signs of disease progression, has been called a variety of names, including “expectant management with curative intent” and others, but active surveillance is the termi-nology most often used today.270,278,284

The  rationale  for  active  surveillance  has  been  derived  from medical  evidence  reflecting  a  variety  of  reported  studies,  including competing  risk  analyses,  surgical  series,  nonrandomized  cohort studies, and randomized trials. Taken together, the findings indicate that  using  a  time horizon of  10  to 15  years,  less  than 3% of men diagnosed with well-differentiated tumors (Gleason score 6 or below) classified as low-risk (based on a serum PSA <10.0 ng/mL and tumor stage  of T2a or  less) will  die  of  prostate  cancer whether  treated or not.281,285-288 Such data clearly call into question the need for treating men with low-risk disease and a life expectancy of less than 10 to 15 years. Unfortunately, curative intervention for favorable-risk prostate cancer appears to be undertaken as commonly as curative interven-tion for higher-risk disease.289-291 This  is of concern because at  least half of newly diagnosed men with prostate cancer have favorable risk 

Page 18: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1480

prostatectomy (RRP) that dramatically decreased operative morbid-ity, to the widespread use of PSA testing that led to an earlier diag-nosis of disease, and to the ease of detection of prostate cancer with TRUS-directed prostate biopsies.6

Radical prostatectomy can be performed under regional or general anesthesia  using  a  perineal,  retropubic,  or  laparoscopic  approach. Radical perineal prostatectomy is not a commonly performed proce-dure  today;  rather,  the  most  common  approach  to  removal  of  the prostate for treatment of prostate cancer is robot-assisted laparoscopic radical prostatectomy  (RALRP). With RALP,  the  robotic  system (a console connected to robotic arms)  translates  the movement of  the surgeon’s fingers into movement of robotic arms attached to the lapa-roscopic instruments used to perform the surgery. Claims of advan-tages  of  RALP  over  other  approaches  have  been  widely  touted  by surgeons,  hospitals,  and  the  device  manufacturer,  resulting  in increased usage of surgery overall, especially in older men.296 However, there is no evidence for improved disease-free or functional outcomes with  RALP  compared  to  RRP.297-299  Instead,  both  disease-free  and quality-of-life outcomes after radical prostatectomy depend more on the experience of the surgeon than on the surgical approach used.300 With  any  surgical  approach,  prior  pelvic  surgery  or  radiation  therapy  tends  to  be  associated  with  an  increased  risk  of  surgical complications.

Regardless of the surgical technique used, in most cases (especially in  the  setting  of  intermediate-risk  or  high-risk  disease),  the  radical prostatectomy proceeds via a staging pelvic lymphadenectomy, focus-ing on the pelvic lymph nodes surrounding the external iliac vein and obturator fossa. The remaining steps of the operation include ligation of the dorsal vein complex anterior to the prostate to control blood loss,  division  of  the  urethra,  identification  and  preservation  of  the neurovascular bundles containing branches of the pelvic nerves inner-vating  the  corpora  cavernosa  necessary  for  penile  erection  (unless wide  excision  of  a  neurovascular  bundle  is  necessary  for  cancer control), division of the bladder neck, resection of the seminal vesi-cles, and construction of a urethrovesical anastomosis.

Hemorrhage and injury to surrounding structures (blood vessels, obturator nerve, ureter, and rectum) are the most common intraop-erative  complications  of  radical  prostatectomy.  In  the  immediate postoperative period, complications can include deep venous throm-bosis and pulmonary emboli, urine anastomotic leak, and postopera-tive  bleeding. The  operative  mortality  rate  (death  within  30  to  60 days) after radical prostatectomy is from 0.4% to 1.6%, depending on age and comorbidity.301 Following surgery, most men spend 1 to 2 nights in the hospital and have an indwelling urinary catheter for 1 to 2 weeks to allow healing of the vesicourethral anastomosis. After catheter  removal,  the  most  common  long-term  complications  of surgery  are  urinary  incontinence  as  a  result  of  intrinsic  sphincter deficiency,  and  erectile  dysfunction  from  injury  to  the  cavernous nerves  innervating  the  corporal bodies of  the penis, both of which negatively impact quality of life.

Urinary Function after Radical ProstatectomyUrinary incontinence rates after radical prostatectomy vary greatly in different reports: from 5% to as high as 50%, with as many as 5% to 10% of men undergoing another procedure for incontinence after radical prostatectomy.297 Some of  the variation  in reported  inconti-nence rates may be attributable to differences in definitions of incon-tinence (stress incontinence vs. more severe difficulties with urinary control),  in  the  time  after  surgery  when  urinary  continence  was assessed (urinary control can improve over as long as a year following surgery), in whether incontinence was reported by treating surgeons in case series or by patients in survey questionnaires, and in the age of the patients undergoing surgery (older age is associated with higher incontinence rates). In the best case series, as many as 95% of men are completely dry 2 years after radical prostatectomy, and as many as 98% of men  report no  significant urinary problems.302,303 Using 

disease,  and  80%  to  90%  of  these  men  undergo  some  form  of  treatment  even when  the  age  at diagnosis  is  above 75 years. Over-treatment  of  nonthreatening  prostate  cancer  probably  reflects  the combination of a fear of harm from cancer, of litigation if aggressive treatment  is  not  recommended,  and  of  misaligned  incentives  that favor treatment in spite of evidence that nonintervention may be the most rational option.

Despite these biases driving overtreatment of favorable-risk pros-tate  cancer,  there has been  a growing  interest  in  active  surveillance approaches as evidenced by the number of academic centers reporting outcomes  from  single-arm  cohort  studies.184,218,270,278,284  As  of  yet, there is no standard approach to selection of men for active surveil-lance, nor are there established guidelines for monitoring regimens, or for criteria to be used as triggers for curative intervention, for men managed by active surveillance. The “ideal” candidate may be a man with favorable-risk prostate cancer (very low-risk to low-risk), a  life expectancy less than 15 years, and a personal desire to avoid the side effects  of  prostate  cancer  treatments. Typical  monitoring  strategies tend to involve DRE and serum PSA determinations at 3- to 6-month intervals, with repeated prostate biopsies at intervals of 1 to 4 years. The most common intervention triggers are a rising serum PSA and an increase in Gleason score of cancer seen on repeat biopsy, a finding that  may  result  from  undersampling  by  the  initial  biopsy  or  from “dedifferentiation” of a low-grade cancer. When managed in this way, approximately 25% to 50% of men will undergo curative interven-tion within 5 to 10 years of surveillance. Of the men proceeding to curative  treatment,  25%  will  do  so  because  of  personal  preference and without any trigger, 30% to 40% will do so because of a higher Gleason score, and 30% to 40% will do so because of an inexorably rising serum PSA.270,284 In one active surveillance series for men with prostate  cancer  (30%  intermediate-risk  disease  and  70%  with favorable-risk disease), the 10-year prostate cancer actuarial survival was 97%.292 An ongoing  trial,  the Prostate Testing  for Cancer  and Treatment (ProtecT) trial, which started in the United Kingdom in 2001, has randomized men with prostate cancer, ages 50 to 69 years to  conformal  radiotherapy,  radical  prostatectomy,  or  active  surveil-lance.293 Results from this trial may further inform the difficult deci-sion that men with localized prostate cancer face in choosing between active surveillance and curative intervention.

Radical ProstatectomyRadical prostatectomy is most often recommended for treatment of men with clinically  localized prostate cancer who have a  life expec-tancy of at  least 10 years or more (Table 84-3). Although there are not  specific or universally  accepted  age  limits  for  radical prostatec-tomy, increasing age is associated with a lower likelihood of receiving surgical treatment for prostate cancer. For example, among men with low-risk or intermediate-risk disease, 3 in 4 at ages 56 to 65 years, 2 in 5 at ages 66 to 75, and 1 to 2 in 10 at age >75 years undergo surgi-cal intervention for prostate cancer.294

Preoperative assessment of men for radical prostatectomy typically includes  a  history  and  physical  examination,  hematology  studies, serum electrolyte studies (with a serum creatinine determination), a urinalysis, coagulation studies, and an electrocardiogram. Many sur-geons recommend a delay of 6 to 8 weeks after prostate needle biopsy to  permit  resolution  of  hematomata/inflammation  caused  by  the biopsy procedure. In anticipation of surgery, men avoid aspirin, non-steroidal  antiinflammatory  agents,  or  high  doses  of  vitamin  E  that might promote excess bleeding.

Radical  prostatectomy  was  not  commonly  recommended  for localized prostate cancer prior to the 1980s because of blood loss, and associated  complications  of  incontinence  and  erectile  dysfunction. But  by  1990,  surgery  was  the  most  commonly  chosen  option  for management  of  prostate  cancer.295  The  increasing  rates  of  radical prostatectomy relative to other management options was in large part due to the description of an anatomic approach to radical retropubic 

Page 19: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1481ProstateCancer • CHAPTER84

in erectile function following radical prostatectomy, and this approach is  routinely  used  today  when  preservation  of  these  nerves  will  not compromise removal of all cancer.

As is the case with urinary continence after radical prostatectomy, reports of return of erectile function vary greatly from 31% to 86% with study differences likely due to definitions of potency used, the methods for assessing return of sexual function, the cohort studied, and  timing  of  assessment  after  surgery.309  Return  of  erections  after surgery  inversely correlates with age, and  is directly associated with the quality of preoperative erections, the frequency of sexual activity preoperatively, and the quality and extent of nerve preservation. In a study using validated questionnaires to assess sexual function, sexual dysfunction was found to remain a big problem one year after treat-ment  in  26%  of  men  who  underwent  radical  prostatectomy,  with distress related to sexual dysfunction described by 44% of sex partners postoperatively.304  Most  of  the  improvement  in  erections  occurs within the first 2 years after surgery.310

Common  approaches  to  managing  erectile  dysfunction  after surgery include PDE-5 inhibitors, vacuum erection devices, vasoac-tive agents placed intraurethrally or injected directly, and placement of a penile prosthesis. Current thinking is  that  injury to the nerves (even when preserved at surgery) supplying corporal tissues results in injury  to  the corporal  tissues  that may be ameliorated by  increased penile blood flow. Thus most sexual medicine experts encourage men to aggressively pursue attempts to achieve erections as soon as possible after surgery. If spontaneous erections do not occur with stimulation and PDE-5 inhibitors, men are encouraged to pursue other options described above, the greatest success achieved being with penile injec-tion therapy.311

Control of Prostate Cancer by Radical ProstatectomyBecause PSA  is  a prostate-specific biomarker,  successful  removal  of all  prostate  tissue  after  radical  prostatectomy  should  result  in  an 

validated  questionnaires  to  assess  urinary  incontinence  and  bother, Sanda  et al.304  found urinary  incontinence  (any pad use)  two years after treatment in 20% of men postsurgery, with a moderate-to-severe urinary problem in 7% (Table 84-4). Most men after radical prosta-tectomy have some degree of stress incontinence for a period of time that usually  resolves within  the first postoperative year,  after which incontinence is not likely to improve further.

Surgical technique has significant consequences for urinary control following radical prostatectomy. Both the striated urinary sphincter musculature  and  smooth  muscle  surrounding  the  urethra  can  be injured  during  surgery.305  Postoperative  strictures  at  the  site  of  the vesicourethral anastomosis can also affect return of urinary control.306 Avoidance of strictures (scars) by performance of a careful, tension-free, vesicourethral anastomosis has led to improved urinary control rates by expert surgeons.307,308 Men with persistent or severe urinary incontinence after radical prostatectomy can be treated with periure-thral injections of bulking agents (silicone) that increase resistance at the bladder outlet, a pelvic sling placed through the perineum that repositions the urethra, and an artificial urinary sphincter—a three-piece  device  made  up  of  a  compressing  urethral  cuff,  a  fluid-filled reservoir placed  in the abdomen that transmits pressure to the ure-thral cuff, and an activating pump placed in the scrotum that transfers fluid from the cuff to the reservoir allowing the cuff to open and the bladder to empty.

Erectile Function after Radical ProstatectomyIn  1982,  Walsh  and  Donker  described  the  anatomy  of  the  nerves traversing the lateral surface of the prostate en route to the corpora cavernosa  of  the  penis,  discerning  the  proximity  of  the  nerves  to vascular structures (the neurovascular bundles) visible at the time of radical prostatectomy.5 This revelation led Walsh et al. to propose a modification of the radical prostatectomy procedure to preserve the neurovascular bundles in an effort to maintain erectile function post-operatively.6 Wide adoption of this modification led to improvements 

Table 84-4 Percentage of Men’s Quality-of-Life Concerns 24 Months after Treatment of Localized Prostate Cancer

Quality-of-Life Domain ProstatectomyExternal Beam Radiation Therapy Brachytherapy

Urinary irritation or obstruction

Dysuria <1 1 5Hematuria 0 <1 1Weak stream 4 10 11Frequency 10 14 20

Urinary incontinence Leaking >1 time per day 14 7 10Frequent dribbling 5 2 3Any pad use 20 5 8Leaking problem 8 5 6

OVERALL URINARY PROBLEM 7 11 16

Bowel function Urgency 2 16 9Frequency <1 10 7Fecal incontinence <1 2 5Bloody stools <1 5 3Rectal pain 2 4 4

OVERALL BOWEL PROBLEM 1 11 8

Sexual function Poor erections 58 60 48Difficulty with orgasm 42 50 38Erections not firm 64 66 54Erections not reliable 51 51 40Poor sexual function 53 58 43

OVERALL SEXUALITY PROBLEM 43 37 29

Adapted from Sanda MG et al. Quality of life and satisfaction with outcome among prostate-cancer survivors. N Engl J Med 2008;358:1250-1261.

Page 20: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1482

These types of low-energy radiation machines were used until cobalt machines became available and provided the first opportunity to treat more deeply seated tumors  in the body. The first  reported series of prostate  cancer  patients  treated  with  60Co  therapy  was  by  George et al. in 1965 and featured men with unresectable disease.318 During this time (beginning in the late 1950s), the megavoltage linear accel-erator was being developed at Stanford University.319 Using this new technology,  the  pioneering  work  of  Bagshaw,  Kaplan,  Del  Regato, and others, ushered in the modern era of radiation therapy for pros-tate cancer by offering the possibility of cure with radiation therapy in this disease.320,321

Now men with prostate cancer have several different radiotherapy options, each of which can be employed with very high precision and with great effectiveness. The integration of computer-based technol-ogy for the design of three-dimensional conformal treatment plans, and  use  of  high-energy  accelerators  with  sophisticated  dynamic shielding,  allows men with prostate  cancer  to be  treated with high doses of radiation, while at the same time sparing surrounding normal tissues.  In  addition,  the  development  of  permanently  implantable radioactive sources, along with the use of real-time imaging and treat-ment planning, has also provided the opportunity for sophisticated prostate  brachytherapy  techniques  resulting  in  safe  and  efficacious treatment of men with prostate cancer.

Traditionally,  men  with  prostate  cancer  referred  for  radiation therapy have tended to be older, to be in poorer health, and to have with higher risk, more advanced tumors than those patients treated surgically. As such, results using less sophisticated radiation therapy techniques  were  less  than  optimal,  raising  concerns  that  radiation therapy might not be as effective as radical prostatectomy. However, with long-term results obtained across a broad range of patients, it is now clear that radiation therapy for prostate cancer provides excellent disease-free survival, comparable to radical prostatectomy for men at similar  risk  of  prostate  cancer  recurrence.  In  this  section,  a  brief description of how men are evaluated and risk-stratified for radiation therapy  will  be  provided,  followed  by  a  description  of  treatment techniques and a review of treatment outcomes for men with low-, intermediate-,  and  high-risk  prostate  cancer.  The  use  of  radiation therapy to treat local prostate cancer recurrences after radical prosta-tectomy will also be reviewed.

External Beam Radiotherapy for Localized Prostate Cancer using 3D-Conformal and Intensity-Modulated ApproachesFor  more  than  four  decades,  external  beam  radiotherapy  has  been widely  used  for  the  definitive  management  of  clinically  localized prostate cancer. More recently, technical advances have permitted the safe  delivery  of  80 Gy  to  the  prostate,  while  minimizing  radiation exposure of nearby normal tissues. Historically, a four-field pelvic box with shielding of the posterior wall of the rectum and anal canal was used to treat the prostate, seminal vesicles, and proximal  lymphatic drainage to a dose of 45 to 50 Gy in 1.8- to 2.0-Gy fractions, with a boost to the prostate to 65 to 70 Gy. Even at these doses, external beam radiation therapy was fairly effective: although overall survival numbers were generally higher  for men treated with radical prosta-tectomy  (often  younger  and  healthier  men),  cause-specific  survival rates were not significantly different.322 Although the four-field box radiation therapy technique is of historic significance only, it is worth noting that side effects (rectal bleeding, dysuria, bowel dysfunction, etc.) associated with this technique were higher than those seen with contemporary technologies and delivery.

To  improve  on  these  outcomes,  new  technical  innovations,  such  as  CT-based  simulation/treatment  planning,  the  introduction  of multileaf collimators in modern linear accelerators, and advances in treatment planning software have allowed for increased precision and accuracy in prostate cancer radiotherapy. For three-dimensional 

undetectable  serum  PSA.  A  detectable  PSA  after  surgery,  and/or  a rising PSA, is an indication of residual prostate cancer that is referred to  as  a  biochemical  recurrence.  Most  series  define  a  biochemical recurrence after surgery as a serum PSA that is above 0.2 ng/mL or rising. Biochemical recurrence rates vary directly with the risk assign-ment before surgery (low-risk, intermediate-risk, high-risk; see Table 84-3), age, the extent and grade of the cancer found on final patho-logical assessment. In modern surgical series, for men with localized prostate cancer, biochemical recurrence-free survival at 5, 10, and 15 years is reported at 84%, 74%, and 66%, respectively.312

Biochemical recurrence is not a surrogate for cancer-specific mor-tality because most men with a biochemical  recurrence as  the only evidence  of  residual  disease  do  not  die  of  prostate  cancer.  A  study  of the natural history of biochemical recurrence in a cohort of men (n =  450)  who  underwent  surgery  for  presumed  localized  prostate cancer, had  rising  serum PSA,  and  received no  treatment until  the development of overt metastases, at a median follow-up of 8 years, revealed that only 30% of the men had developed metastatic disease.313 For these men, the median time from surgery to biochemical recur-rence  was  3  years  and  the  median  metastasis-free  survival  was  10 years. The Gleason score and PSA doubling time were independent predictors of metastasis-free survival, with no evidence of metastases in 94% of men with prostate cancer and a Gleason score of 6 or less versus 19% of those with Gleason scores of 8 to 10, and no evidence of metastases  in 72% of men with  recurrent prostate  cancer  and  a PSA doubling time of >15 months versus 7% of  those with a PSA doubling time of 3 to 9 at 10 years, respectively.

Cancer-specific  survival  after  radical  prostatectomy  was  recently evaluated  in  a  multiinstitutional  study  of  men  (n  =  11,521)  with prostate  cancer  diagnosed  in  the  PSA  era,  revealing  a  cumulative incidence of prostate cancer deaths at 15 years of only 7%.287 Predic-tors  of  prognosis  include  the  Gleason  grade  and  the  presence  of seminal vesicle invasion, with prostate cancer mortality rates of 0.2% to 1.2%, 4.2% to 6.5%, 6.6% to 11%, and 26% to 37% for Gleason scores of 6 or less, 3+4, 4+3, and 8 to 10, respectively, and 0.8% to 1.5%,  2.9%  to  10%,  15%  to  27%,  and  22%  to  30%  for  organ-confined  cancer,  cancer  with  extraprostatic  extension,  cancer  with seminal  vesicle  invasion,  and  cancer  accompanied  by  lymph  node metastasis,  respectively.287 Because  these  results were  from  a  single-arm study without a comparison group, it is not possible to know to what extent the favorable surgical results for screen-detected cancers were a result of surgical intervention for a lethal phenotype versus the intrinsic  indolent biological behavior of many  such cancers. Hope-fully, results from PIVOT, comparing surgery with watchful waiting, and  from  ProtecT,  comparing  conformal  radiotherapy,  prostatec-tomy, and active surveillance will provide more definitive answers to this question.

Radiation TherapyRadiation  therapy  has  been  used  in  the  management  of  prostate cancer  for  nearly  a  century.  Following  Roentgen’s  discovery  of  the x-ray  in  1895,314  and  the  isolation  of  radium  by  Pierre  and  Marie Curie in 1898,315 several pioneering physicians began treating pros-tate  disorders,  including  prostate  cancer,  with  radiation.  In  1910, Paschkis and Tittinger inserted radium into the prostatic urethra with a cystoscope in what may be the use of radiation for prostate cancer. Not long after, Hugh Hampton Young from Johns Hopkins reported the relatively large experience of treating prostate cancer patients with urethral  and  rectal  radium  “applicators.”316  These  and  other  early studies revealed that even when radiation was applied  in this crude manner, it could improve symptoms and kill prostate cancer. However, treatment was technically difficult for the physician and uncomfort-able for the patient. In 1928, the first report on the use of externally delivered  low-energy  kilovoltage  radiation  for  prostate  cancer  was proffered  by  Barringer.317  The  associated  dosimetry  was  not  well worked out and, thus men were treated until their skin turned red. 

Page 21: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1483ProstateCancer • CHAPTER84

Complications of 3D-CRT and IMRTAlthough the use of 3D-CRT was intended to minimize the effects of  high-dose  radiation  on  normal  tissues,  increased  late  complica-tions, such as rectal and urinary toxicity, have been noted with escalat-ing radiation doses used in 3D-CRT.323,324 Rectal toxicity encompasses urgency,  frequency,  pain,  fecal  incontinence,  or  bleeding.  In  a  Fox Chase Cancer Center case series, the 5-year incidence of grade 3 or 4 rectal toxicity at a dose of 75 to 76 Gy was 8%, which upon better shielding of the anterior rectal wall, could be reduced to 2%. IMRT seems  to  be  accompanied  by  less  rectal  complications.  In  one  case series,  men  with  clinically  localized  prostate  cancer  treated  with IMRT to a total dose >81 Gy at Memorial Sloan-Kettering Cancer Center exhibited a  reduction  in  late  rectal  toxicity when compared with 3D-CRT.325 With a median follow-up of 24 months, grade 2 or higher rectal bleeding was seen in 4% of men receiving IMRT, with grade 3 rectal toxicity in only 0.5% (and no grade 4 rectal toxicity) at 3 years. Urinary toxicity, including urethral strictures, tends to be similar for both 3D-CRT and IMRT, with as many as 15% of men suffering late grade 2 urinary complications. Because this may be the result of high-dose radiation to the urethra, decreasing urethral doses for men with prostate  cancer  limited  to  the peripheral  zone of  the prostate may be a means of reducing late urinary toxicity, but better diagnostic  imaging  will  be  necessary  to  identify  men  with  such cancers. Overall, for external beam radiation therapy, 9% to 11% of men treated can expect to have some level of rectal or urinary distress 1 year after treatment (Table 84-4).304

Data on the incidence of erectile dysfunction after external beam radiotherapy  has  been  widely  variable.  Sexual  function  has  many facets  that  are  difficult  to  evaluate  or  quantify.  A  commonly  used qualitative definition of sexual potency is the ability to achieve spon-taneous  erections  sufficient  for  intercourse. Although potency  rates after radical prostatectomy have increased with use of nerve-sparing techniques,  the mechanism of  radiation-related erectile dysfunction appears unrelated to the neurovascular bundles. Zelefsky et al. specifi-cally  addressed  this  topic  by  performing  duplex  ultrasound  studies before and after prostaglandin injection to stimulate erections in men with  radiation-related  erectile  dysfunction.326  A  diminished  peak penile  blood  flow  rate  (<25 mL/min)  was  evident  in  63%  of  such men, with abnormal distensibility of the corpora cavernosa in 32%. Thus, the primary mechanism of radiation therapy-associated impo-tence may be vascular damage rather than nerve damage. Fisch et al. demonstrated a correlation between incidence of erectile dysfunction postradiotherapy  and  radiation  dose  to  the  vascular  penile  bulb.327 Men receiving >70 Gy to >70% of the bulb of the penis are at greatest risk of experiencing radiation therapy–associated erectile dysfunction. A steady decline in potency rates over time is characteristic for men treated with  external  beam  radiation  therapy.  In  a  large  series  (n = 434) of men  treated with  radiation  therapy  at  Stanford University, 86% of the men remained potent at 15 months after treatment, but only 50% were potent 6 years later, and only 30% maintained erectile function for the remainder of their lives.328 Sildenafil administration resulted  in  improvement  of  erectile  function  in  74%  of  men  with radiation therapy–associated erectile dysfunction in a study at Memo-rial Sloan-Kettering Cancer Center.329 Men who do not respond to sildenafil may respond to intracavernosal prostaglandin injections.

Cancer Control by External Beam Radiation TherapyWith the ready availability of serum PSA testing, outcome compari-sons  between  primary  prostate  cancer  treatments  tend  to  focus  on PSA  as  a  marker  of  prostate  cancer  recurrence  (“PSA  relapse”–free survival).  Of  course,  if  successful  in  controlling  cancer,  surgical removal of the prostate eliminates PSA production by both nonneo-plastic and neoplastic cells, whereas radiation therapy may not abolish PSA  production  by  residual  prostate  epithelial  cells  even  as  cancer 

conformal  radiation  therapy  (3D-CRT),  three-dimensional  recon-structions of pelvic CT images are generated to create more accurate target  volumes  (e.g.,  prostate  and  seminal  vesicles)  for  treatments, enabling better identification of critical structures (e.g., bladder and rectum)  to  be  avoided.  An  iterative  process  is  then  used  to  design beam arrangements that will deliver a prescribed dose to the regions of interest and minimize dose to a given volume of an adjacent organ. In  this  way,  the  incident  radiation  volume  “conforms”  to  target volume. The treatment target volumes and normal organs can then be  visualized  in  three  dimensions,  creating  a  “beam’s-eye  view,”  a portrayal of the target area as if looking straight down the path of the radiation beam. During  treatment delivery,  computerized multileaf collimators then shape each individual beam to conform to the shape of  the  target  in  the  beam’s-eye  view.  Intensity-modulated  radiation therapy (IMRT) builds on the advantages of 3D-CRT by using treat-ment  planning  software  capable  of  even  further  optimizing  dose distribution for prostate cancer treatment, by modifying not just the orientation  and  shape  of  the  incident  beams but  also  the  intensity across the designated treatment volume (Figure 84-13).

Figure 84-13 •  Kaplan-Meier  actuarial  probability  of  relapse-free  sur-vival after  radiation therapy  for prostate cancer,  stratified by radiation dose (78 Gy vs. 70 Gy). A, Men with favorable-prognosis prostate cancer (serum prostate-specific antigen [PSA] ≤10 ng/mL). B, Men with poorer-prognosis prostate cancer (serum PSA, 10 ng/mL). (From Pollack A, Zagars GK, Stark-schall G, et al. Prostate cancer radiation dose response: results of  the M.D. Anderson  phase  III  randomized  trial.  Int  J  Radiat  Oncol  Biol  Phys 2002;53:1097-1105.)

0 20 40 60 80 100

A

0 20 40 60 80 100

B

Fra

ctio

n fr

ee o

f fai

lure

Months after radiotherapy

PSA ≤ 10 ng/mL

78 Gy70 Gy

P=0.46

Fra

ctio

n fr

ee o

f fai

lure

Months after radiotherapy

PSA>10 ng/mL

78 Gy70 Gy

P=0.012

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

Page 22: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1484

Brachytherapy may exhibit  its  greatest benefits  in  the  treatment of men with low-risk, clinically localized, prostate cancer, though in the absence of a prospective randomized trial, this recommendation has been based on retrospective studies with varying follow-up time. One of the largest of these studies reported outcomes for more than 2500  men  with  prostate  cancer  treated  at  11  different  institutions with permanent interstitial brachytherapy.342 The median follow-up for  this  group of men with  stage T1 or T2 prostate  cancer was 63 months and all were treated with either 125I or 103Pd without use of hormonal  therapy. Men with  low-risk,  intermediate-risk, and high-risk prostate cancer had 8-year actuarial PSA relapse-free survivals of 82%, 70%, and 48%, respectively, using ASTRO criteria. The dosi-metric  quality  of  the  implant  was  critical  to  outcome:  for  men  in which 90% of the prostate (D90) received ≥130 Gy, the 8-year PSA relapse-free survival was 93%, whereas for men in which the prostate D90  was  <130 Gy,  the  8-year  PSA  relapse-free  survival  was  76%. There were few, if any, differences attributable to the use of 125I versus 103Pd implants.

cells  are killed. Nonetheless,  a  rising  serum PSA after  radiotherapy for prostate cancer is correlated with the appearance of progressive or metastatic  prostate  cancer  upon  further  follow-up.330  Furthermore, the rate of serum PSA rise may help distinguish between local versus distant treatment failure. Men with slow rates of serum PSA increases are more likely to have local prostate cancer recurrences, whereas men with a rapid serum PSA rise appear more likely to have distant pros-tate cancer metastates.331 In 2006, the American Society for Thera-peutic  Radiology  and  Oncology  (ASTRO)  issued  a  consensus statement  establishing  the  definition  of  recurrence  following  radia-tion therapy as a serum PSA level that  is 2.0 ng/mL over the nadir value.332 This definition is an update to the 1997 ASTRO criteria for biochemical  recurrence  (three  consecutive  rises  of  the  serum  PSA determined at least 3 months apart333) and is now preferred for evalu-ating and comparing radiation treatment strategies. Before these cri-teria were  formulated,  treatment outcomes  in different  studies  and case series were difficult to compare because of the lack of uniformity in defining treatment failure.

External  beam  radiation  therapy  using  3D-CRT  and  IMRT approaches  achieves  effective  prostate  cancer  control.324,334-336  In  a large cohort of men (n = 1100) treated with 3D-CRT at Memorial Sloan-Kettering  Cancer  Center,  the  PSA  relapse-free  survivals  for men with low-risk prostate cancer treated to radiation doses of 64.8 to 70.2 Gy versus 81 Gy, were 77% and 98%, respectively.336 Men with  intermediate-risk  and  high-risk  prostate  cancers  also  showed significant  improvement.  In  a  randomized  trial  of  radiation  dose escalation, higher-dose  treatment was  associated with  improvement in prostate cancer control  rates  for men with prostate cancer, espe-cially men with a pretreatment PSA >10 ng/mL (Figure 84-14).334,337 In  this  study,  men  (n  =  305)  with  stage  T1–T3  prostate  cancer received either radiation therapy to a total dose of 70 Gy using con-ventional radiotherapeutic techniques, or to 78 Gy using a 3D-CRT approach. Results revealed a freedom-from-PSA relapse of 64% and 70% at 6 years  for  the 70-Gy and 78-Gy groups,  respectively  (P = 0.03). Men with a pretreatment serum PSA >10 ng/mL were found to have  the most  significant benefit  from radiation dose escalation, with freedom-from-PSA relapse rates of 62% for the 78-Gy arm and 43% for the 70-Gy arm (P = 0.01). Overall survival was not signifi-cantly different between the two radiation doses, but a trend toward an  improved  freedom-from-distant  metastasis  was  evident  in  men treated on the 78-Gy arm, 98% versus 88% at 6 years (P = 0.056).

BrachytherapyProstate  brachytherapy  involves  the  implantation  of  radioactive sources  into  the  prostate  using  image  guidance  (Figure  84-15).  In principle, brachytherapy offers an attractive means for radiation dose escalation  and  conformality  in  the  treatment  of  clinically  localized prostate cancer. However, the most attractive feature for many men with  prostate  cancer  is  that  this  option  may  be  most  convenient, allowing  a  rapid  return  to  normal  lifestyle  and  activity.  A  typical brachytherapy procedure is performed in 2 hours, often under spinal anesthesia, and does not require an overnight hospital stay. The most common sources in use today are 125I (Iodine) and 103Pd (Palladium) (Table 84-5). Thus far, there have been no compelling data to support the superiority of one isotope or the other. Theoretically, the higher initial dose rate emitted by 103Pd might be advantageous for the treat-ment of cancers,  like prostate cancer, with a  relatively  low αβ  ratio (more radioresistant); however, retrospective data for prostate cancer have been  inconclusive.338 To date, no  long-term prospectively ran-domized data exist directly comparing 125I with 103Pd. Most experi-ence  with  prostate  brachytherapy  has  been  with  permanent  low dose-rate (LDR) implants. Several centers have collected experience with temporary high dose-rate (HDR) implants  for prostate cancer brachytherapy.339-341 Although this treatment approach is not widely used, the available results appear comparable with permanent brachy-therapy for clinically localized prostate cancer.

Figure 84-14 •  Interstitial brachytherapy for prostate cancer. A, Radio-graph obtained after implantation of radioactive seeds. B, CT image showing radioactive  seed  location  within  prostate.  (From  Speight  JL,  Roach  M  III: Imaging  and  radiotherapy  of  the  prostate.  Radiol  Clin  North  Am 2000;38:159-177.)

A

B

Page 23: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1485ProstateCancer • CHAPTER84

The  addition  of  supplemental  external  beam  radiotherapy  to brachytherapy  remains  somewhat  controversial.  Davis  et al.  exam-ined the radial distance of extraprostatic extension of prostate cancer and  found  it  to  be  almost  always  5 mm  or  less,  which  would  be within  a  typical  brachytherapy  dose  distribution.343 Thus  generous periprostatic margins  in brachytherapy planning should obviate the need  for  supplemental  external  beam  radiotherapy.  Proponents  of combinations of brachytherapy and external beam radiation therapy emphasize  both  the  advantage  of  higher  delivered  doses  and  the ability  to  smooth  out  cold  spots  inherent  with  brachytherapy,  the so-called spackle effect. Sylvester et al. have recently reported a ret-rospective review of 232 men with clinically localized prostate cancer treated  with  either  125I  or  103Pd  brachytherapy  and  external  beam radiation  therapy  administered  before  the  implant.344  At  a  median follow-up of 9.4 years,  the biochemical  relapse-free  survival  for  the entire  study group was 74%, with biochemical  relapse-free  survival for  low-risk prostate cancer of 85.8%,  for  intermediate  risk disease 

Figure 84-15 •  Kaplan-Meier estimates of survival for men with prostate cancer from a prospective randomized clinical trial comparing a combination of androgen-deprivation therapy (goserelin acetate for a total of 3 years with cyproterone acetate for 1 month) and radiation therapy with radiation therapy alone. A, Overall survival. B, Relapse-free. O, number of deaths; N, number of subjects. (From Bolla M, Collette L, Blank L, et al. Long-term result with immediate androgen suppression and external irradiation in patients with locally advanced prostate cancer [an EORTC study]: a phase II randomized trial. Lancet 2002;360:103-106.)

0

10

20

30

40

50

60

70

80

90

100

0 1 2 3 4 5 6 7 8

0 1 2 3 4 5 6 7 8

A

B

Ove

rall

surv

ival

(%

)

Time since randomization (yrs)

No. of patients at riskO N81 206 199 177 146 106 70 46 30 1650 207 197 183 166 142 93 71 43 24

Radiotherapy aloneCombined treatment

Log-rank test P<0.0001,hazard ratio 0.51(95% Cl 0.36–0.73)

Bio

chem

ical

ly d

efin

eddi

seas

e-fr

ee s

urvi

val (

%)

Time since randomization (yrs)

No. of patients at riskO N36 66 64 59 50 29 17 9 4 356 170 169 157 138 116 76 50 26 13

Radiotherapy aloneCombined treatment

Log-rank test P<0.0001,hazard ratio 0.42(95% Cl 0.28–0.64)

0

10

20

30

40

50

60

70

80

90

100

Table 84-5 Physical Differences Between 125I and 103Pd Radioactive Seeds

125I 103Pd

Year introduced 1965 1986

Photon energy (keV) 28 21

Half-life (d) 59.4 17

Initial dose rate (for monotherapy) 7 cGy/h 18–20 cGy/h

RBE 1.4 1.9

RBE, relative biological effectiveness.

Page 24: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1486

after brachytherapy tend to emerge within 3 years of implant place-ment, with rectal injury in 5% to 10% of men treated.356,355,357,304

As  popularity  of  prostate  brachytherapy  for  clinically  localized prostate  cancer grew  in  the 1990s,  a  commonly cited advantage of the treatment modality was a lower incidence of treatment-associated erectile  dysfunction  compared  to  external  beam  radiotherapy  or radical  prostatectomy.  Undoubtedly,  this  selling  point  tipped  the scales in favor of brachytherapy for many men faced with selecting a treatment  for  early-stage  prostate  cancer.  For  instance,  Stock  et al. reported a 2-year potency rate of 94% after implant, whereas Wallner et al. reported a 3-year potency rate of 86%.358,359 Studies with longer follow-up,  however,  have  shown  a  continued  decrease  in  sexual potency  over  time.  With  more  long-term  follow-up  in  other  case series, only 57% of men retained potency at 5 years.356 Even Stock et al.  subsequently  reported  a  6-year  potency  rate  of  59%  in  their case  series.360  Notably,  their  study  found  that  70%  of  men  with normal erectile function before implant retained potency at 6 years, whereas  men  with  “erectile  function  sufficient  for  intercourse”  but suboptimal  erections  had  only  a  34%  6-year  potency  rate.  Using postimplant dosimetry studies, Merrick et al. demonstrated that dose to the penile bulb correlated with postimplant erectile dysfunction. In the majority of men who retained potency, the dose delivered to 50%  of  the  penile  bulb  was  less  than  50 Gy.361  Potentially,  this knowledge may result in improved morbidity outcomes with techni-cal  attention  to  this dose  threshold. Erectile dysfunction  is not  the sole  complication  of  prostate  brachytherapy  with  the  potential  to affect sexual quality of life, however, with reports of hematospermia in 28%, orgasmalgia in 15%, and alteration in the intensity of orgasm in 38%.362 These side effects tend to be transient in most men.

Proton Beam RadiotherapyThough only available at a few centers worldwide, there has been an interest  in  proton  beam  therapy  for  prostate  cancer.  The  unique physical properties of protons make them ideal for the treatment of disease in proximity to critical strictures. Specifically, protons deposit the majority of their energy at the very end of their  linear tracks, a phenomenon termed the “Bragg peak.” The dose falls off very rapidly at depths beyond  the Bragg peak. This  is particularly useful  in  the treatment  of  prostate  cancer,  to  minimize  rectal  and  bladder  dose. Investigators from Loma Linda University reported their experience treating men (n = 1255) with T1–T3 prostate cancer.363 Men were treated with protons alone to 74 cobalt Gray equivalents (CGE) or with photons  to 45 Gy  followed by proton boost  to 75 CGE. The median follow-up was 62 months and the 8-year actuarial biochemi-cal disease-free  survival  rate was 73%. In a  recent prospective,  ran-domized trial of 70.2 GyE versus 79.2 GyE (combination of photons plus protons) for men (n = 393) with stage T1b-T2b prostate cancer and  a  serum PSA <15 ng/mL,  at  a median  follow-up of  5.5  years, 61.4%  of  the  men  treated  with  70.2 GyE  versus  80.4%  of  men treated  with  79.2 GyE  were  free  of  biochemical  treatment  failure, supporting  the  concept  that  higher  doses  of  radiation  results  in  a statistically significant reduction in the risk of recurrence of localized prostate cancer.364

Adjuvant Endocrine TherapyADT has been found to improve treatment outcomes in randomized trials  of  men  with  high-risk  prostate  cancer  treated  with  external beam radiation therapy but not as convincingly for men treated with radical prostatectomy (Figure 84-16).365-367 D’Amico et al.  reported results of a large retrospective study (n = 1586) of men treated with 3D-CRT  plus  or  minus  ADT  for  low-risk,  intermediate-risk,  and high-risk prostate cancer.368 In this study, the median radiation dose was  70.2 Gy  and  ADT  was  used  for  2  months  before  radiation therapy,  during  treatment,  and  for  2  months  after  treatment  was completed. With a median follow-up of 51 months, the 5-year PSA 

of  80.3%,  and  for  the  high-risk  group  of  67.8%.  These  results compare favorably with all other surgical and/or radiation series with respect to long-term durable prostate cancer control.

Relative contraindications to the use of prostate brachytherapy are large prostate size, pre-implant obstructive urinary symptoms, history of prior TURP, and the presence of perineural prostate cancer inva-sion on prostate biopsy. Large prostate size has been perceived to be associated with a higher  risk of urinary morbidity postimplant and with  unsuitability  for  implant  because  of  pubic  arch  interference. Men with a prostate volume of greater than 50 mL have been either counseled against brachytherapy or placed on androgen deprivation therapy (ADT) in an attempt to reduce gland size. Nonetheless, the implantation  of  large  prostates  with  radioactive  seeds  has  been described with acceptable morbidity.345,346 In one case series, postim-plant dosimetry quality was found to be independent of prostate size or  use  of  ADT.346  The  correlation  of  preimplantation  obstructive urinary symptoms and postimplantation urinary obstruction has not been  fully  resolved.  Terk  et al.  reported  that  a  high  International Prostate  Symptom  Score  (I-PSS),  a  measure  of  obstructive  urinary symptoms, predicted postimplant urinary retention.347 With the use of α–blockers before and after implant procedures, others have noted no association between preimplant I-PSS and urinary obstruction.348 A  prospective  study  examining  preimplant  urinary  flow  rate  and postvoid  residual,  in  addition  to  I-PSS,  showed  no  association  of obstructive urinary symptoms with postimplant urinary retention or long-term  urinary  function.349  TURP  is  thought  to  be  a  relative contraindication to prostate brachytherapy as it has been associated with  unacceptably  high  rates  of  urinary  incontinence.  This  could possibly be attributable to seed placement approaches that result  in a high central dose to the TURP defect. However, by using a periph-eral  source  loading  approach  to  limit  dose  to  the TURP  defect  to 110% of the prescription dose, the incidence of urinary incontinence may be reduced.350

A phenomenon peculiar  to prostate brachytherapy  that deserves mention is the so-called PSA spike. With a time of onset between 12 and  30  months  postimplantation,  approximately  one-third  of  men with  prostate  cancer  treated  with  brachytherapy  will  experience  a transient increase in serum PSA.351 This spike may be due to radiation-associated prostatitis that compromises prostate architecture, permit-ting  more  PSA  to  appear  in  the  serum.  Notably,  such  PSA  spikes portend no worse long-term outcome.

Toxicity of BrachytherapyThe  short-  and  long-term  sequelae  of  brachytherapy  for  prostate cancer differ  from  those of  external beam  radiotherapy  and  radical prostatectomy. Kleinberg et al. described the morbidity outcomes of the early Memorial Sloan-Kettering Cancer Center experience with permanent  transperineal  brachytherapy,  reporting  that  the  most common  side  effects  were  nocturia  and  dysuria  (80%  and  48%, respectively, 2 months after implantation).352 By 12 months following implantation, these figures had declined (to 45% and 20%). Urinary retention is seen in 3% to 14% of men and usually lasts 1 week or less.  The  most  bothersome  late  complications  of  prostate  cancer brachytherapy are urethral stricture and urinary incontinence. Ragde et al.  reported  a  5.1%  incidence  of  urinary  incontinence  in  men treated with prostate cancer brachytherapy and followed for 7 years; each  of  the  men  with  incontinence  had  a  history  of  TURP.353 Urethral  stricture appeared  in 14.4% of  the men. Others have also seen  an  increased  incidence  of  urinary  incontinence  in  the  setting  of a history of TURP.354 In a cohort study of Medicare beneficiaries (n =  2124)  who  were  treated  with  brachytherapy,  urinary  inconti-nence was noted  in 6.6% and bladder outlet obstruction  requiring intervention was  found in 8.3%.355 In all, as many as 16% of men treated with brachytherapy  for prostate  cancer  can  expect  to  suffer distressful urinary symptoms 2 years after treatment (Table 84-4).304 As is the case for external beam radiation therapy, rectal complications 

Page 25: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1487ProstateCancer • CHAPTER84

improved  PSA  relapse-free  survival  with  the  addition  of  adjuvant radiation therapy following radical prostatectomy in men with these risk factors.378-380 In one study, men (n = 149) with pathological stage T3N0 prostate cancer and an undetectable postoperative serum PSA, adjuvant radiation therapy was given to a median dose of 64.8 Gy to some men (n = 52), whereas the remainder (n = 97) underwent no further treatment.378 In a matched-pair analysis, the 5-year freedom-from-PSA  relapse  rate  was  89%  in  the  adjuvant  radiation  therapy group versus 55% for treatment with surgery alone (P < 0.01).

Three  prospective  randomized  trials  have  been  completed  that tested  the benefits of adjuvant  radiation  therapy versus observation following radical prostatectomy in men with prostate cancer and poor pathological features. The European Organization for Research and Treatment  of  Cancer  (EORTC)  has  published  clinical  trial  results revealing  an  improvement  in  biochemical  relapse-free  survival  and locoregional progression-free survival for the men with pT3 tumors or pT2/T3 prostate cancer and positive surgical margins treated with adjuvant  radiation  therapy.366  In a  second  study  targeting  the  same patient  population  conducted  by  the  Southwest  Oncology  Group (SWOG), men with prostate cancer who received adjuvant radiation not  only  enjoyed  better  biochemical  and  local  control,  but  also appeared  likely  to  have  better  metastasis-free  survival  and  overall survival; these improvements were not yet statistically significant even at a median follow-up of 10 years.381 Nonetheless, men treated with adjuvant radiation therapy were less likely to need hormonal therapy at 5 years (10% vs. 21%, P < 0.001). Avoidance of ADT is likely of significant clinical benefit, reducing or delaying significant morbidity, including hot flashes, diminished bone density,  sexual dysfunction, cognitive dysfunction, and overall  reduced quality of  life. Finally, a third randomized trial comparing adjuvant radiation and observation after  prostatectomy,  which  enrolled  only  men  with  pT3  disease regardless  of  surgical  margin  status,  was  reported  by  the  German Cancer Study Group.382 As with the other two studies, men treated with adjuvant radiation experienced a superior biochemical relapse-free survival with median follow-up of only 3.3 years. Thus, overall, data available today provide a convincing case for adjuvant radiation treatment after prostatectomy. However,  it  is clear that not all men benefit from such treatment. What is needed for the future is a more precise way to stratify men for adjuvant radiation therapy: one such group  may  be  men  with  positive  surgical  margins  after  radical prostatectomy.

relapse-free survival for men with low-risk prostate cancer was 92% with the addition of ADT versus 84% without. Men with intermedi-ate- and high-risk prostate cancer also fared significantly better when given  ADT.  Radiation  Therapy  Oncology  Group  (RTOG)  Trial 94-08, which completed accrual in 2001 and was reported in 2012, was designed to ascertain whether men (n = 1917) with stage T1b–T2 prostate cancer and a serum PSA of 20 ng/mL or  less benefit  from the  addition  of  “complete  androgen  blockade”  given  for  4  months before  and  concomitantly  with  external  beam  radiation  therapy.369 Results of this trial revealed that the addition of androgen suppression to  radiation  significantly  improved  disease-specific  and  overall  sur-vival, reducing overall deaths from 8% to 4% at 10 years.

ADT has also been used with prostate cancer brachytherapy, both to  reduce  the  size  of  the prostate  gland  and  to  improve outcomes. Most prostate glands exhibit some decrease in volume after 3 months of ADT, with an average 30% to 40% reduction, and little  further volume  decreases.370  About  10%  of  prostate  glands  will  show  no volume reduction at all in response to androgen deprivation. Decreas-ing  the  size  of  the  prostate  may  reduce  pubic  arch  interference  in selected  men.  Blank  et al.  reported  that  men  treated  with  ADT tended to have smaller prostates that required fewer seed implants.371 However, at this point there are no data to suggest that smaller pros-tate  volumes  correlate  with  reduced  acute  or  late  morbidity  from brachytherapy. Furthermore, although prospective randomized trials have  demonstrated  improved  survival  in  those  men  with  locally advanced prostate cancer treated with external beam radiotherapy and ADT,  it  is  not  clear  that  these  results  can  be  extrapolated  to  men treated with brachytherapy. A retrospective matched-pair analysis of men  (n  =  60)  with  prostate  cancer  treated  at  Memorial  Sloan-Kettering Cancer Center showed no benefit for the addition of ADT to brachytherapy for men with low-, intermediate-, or high-risk pros-tate cancers.372

Postprostatectomy Adjuvant Radiation TherapyPathological features that portend a higher risk of local recurrence are common  after  radical  prostatectomy.  A  positive  surgical  margin  is associated  with  an  approximately  50%  risk  of  prostate  cancer recurrence.373-377 Other features associated with recurrence are extra-capsular extension, seminal vesicle invasion, and Gleason score of 7 or  greater.  Several  retrospective  series  have  now  demonstrated 

Figure 84-16 •  Treatment of life-threatening prostate cancer. 

Localizedprostatecancer

Recurrentcancer (risingserum PSA)

Metastaticprostatecancer

Castration-resistant

cancer (risingserum PSA)

Metastaticcastration-resistantcancer

Adjuvant hormonal therapy(with radiation)

Androgen deprivationLHRH analogs, LHRHantagonists, anti-androgens,androgen synthesis inhibitors

Chemotherapydocetaxel, cabazitaxel

Immunotherapysipuleucel-T

Bone therapyfor osteoporosis andbone metastases

Bisphosphonateszoledronic acid

RANK ligand inhibitordenosumab

Primary therapysurgery, radiation therapy,active surveillance

Page 26: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1488

androgen deprivation or androgen deprivation  initiated at  the  time of PSA  relapse.391 At 5 years, 65% of men  treated with  immediate androgen deprivation versus 42% of men treated with delayed andro-gen deprivation were  free of PSA relapse. There  are no prospective trials examining the addition of ADT to salvage radiation therapy for local  prostate  cancer  recurrence  after  prostatectomy.  Taylor  et al. reported a benefit to the addition of ADT to salvage radiation therapy in a retrospective case series.392 In this series, adjuvant ADT was given to men who received salvage radiation therapy for a median duration of 24 months. At 5 years, 81% of men receiving ADT (vs. 54% not treated with androgen deprivation) were free of PSA relapse. In con-trast,  Song  et al.  reported  identical median disease-free  survivals  of 26 months for men treated with or without concurrent ADT along with salvage radiation therapy for prostate cancer recurrence.387 The 1999 ASTRO Consensus Panel concluded that there was insufficient evidence  to  support  routine  use  of  ADT  with  postprostatectomy radiation therapy.386

In  general,  men  receiving  radiation  therapy  postprostatectomy experience little in the way of additional morbidity. The incidence of urinary incontinence does not seem to be increased and erectile func-tion  does  not  seem  to  be  worsened  in  men  treated  with  adjuvant radiation  therapy  after  prostatectomy.393,394  Bastasch  et al.  reported that 100% of men who were potent after nerve-sparing radical pros-tatectomy  remained  potent  after  adjuvant  IMRT.395  Despite  these promising reports, side effects remain possible. Katz et al. noted that 19% of men experienced grade 2 or 3 genitourinary toxicity (hema-turia  or  urethral  stricture),  and  12%  of  men  experienced  grade  2 bowel toxicity (with no grade 3 or higher toxicity noted), following salvage radiation with 3D conformal techniques.396

LOCALLY ADVANCED DISEASE AND PALLIATION

Radiation Therapy and Adjuvant Endocrine Therapy for Locally Advanced Prostate CancerCombined modality treatment, using ADT hormonal manipulation in conjunction with external beam radiation therapy, takes advantage of separate and noncompeting modes of cell death such that cells that can survive the insult of one modality cannot survive the other or the additive/synergistic  properties  of  the  two.397  ADT  has  been  used along with radiation therapy for many years in an attempt to modify the outcome of men with stage C (T3) prostate cancer.398 Historically, the rationale for this treatment approach was that these men had an inferior outcome when compared with men with earlier stage prostate cancer treated with radiation therapy. In addition, the tumors were often quite bulky, and it was thought that a course of cytoreductive therapy might provide a more favorable geometry for external irradia-tion. However,  the use of ADT in combination with radiation was not universally accepted throughout the 1970s and 1980s. Radiation therapy  techniques were  improving, and  the  results  from early case series exploring  the benefit of ADT before and/or during radiation therapy were often negative.399,400

In  the  early  1980s,  two  case  series  reported  encouraging  results with  the use of ADT and  external beam  radiation  therapy  to  treat men with  locally  advanced prostate  cancer.401,402 Pilepich  et al.  also found that men with histologically unfavorable prostate cancers who had been treated with ADT and external beam radiation therapy as part of the RTOG 75-06 Trial exhibited similar disease-free survival and overall survival rates as men with more favorable prostate cancers who did not receive ADT along with radiation therapy.403 Since these early  experiences,  phase  3  clinical  trials  have  established  the  local control and survival benefits of ADT given along with external beam radiation therapy for locally advanced prostate cancer.

RTOG 86-10 was a randomized phase 3 clinical trial of external beam radiation therapy alone (standard treatment arm) versus neo-adjuvant  and concomitant  total  androgen  suppression and external 

Salvage Radiotherapy after Radical ProstatectomySalvage radiotherapy refers to the use of radiation therapy postpros-tatectomy in the setting of recognized prostate cancer recurrence. As many as 27% to 53% of men who undergo radical prostatectomy for prostate cancer will have a detectable PSA within 10 years of surgery.383 Subsequently, approximately 25% of men who undergo radical pros-tatectomy will be treated with salvage radiation therapy for recurrent prostate  cancer.384  In  the  setting  of  persistent  or  rising  serum  PSA following  radical  prostatectomy,  it  is  important  to  rule  out  distant metastatic prostate cancer with bone scan, chest radiography, and CT scan  of  the  abdomen  and  pelvis,  prior  to  the  initiation  of  salvage radiation therapy. Partin et al. correlated the rate of serum PSA rise with likelihood of local versus distant relapse after surgery.385 A serum PSA rise of 0.75 ng/mL/year was associated with local recurrence. In 1999, an ASTRO consensus panel concluded that treatment of men with local prostate cancer recurrence after radical prostatectomy and a pre–radiation therapy serum PSA <1.5 ng/mL was more  likely to be successful.386 In addition, doses above 64 Gy were recommended in  the  salvage  setting. Several  studies have demonstrated a Gleason score greater than 7 to be associated with a low likelihood of success-ful  salvage  after  prostate  cancer  recurrence  postprostatectomy.387,388 Caddedu  et al.  reported  no  men  free  of  PSA  relapse  treated  with salvage radiation therapy after prostatectomy for prostate cancers with Gleason scores of 8 or above.388 Similarly, Song et al.  found only 2 of 14 men with Gleason score of 8 or above to be prostate cancer free at the time of analysis.387 These results indicate that men with recur-rent prostate cancer and a Gleason score of 8 or greater are unlikely to benefit from salvage radiation therapy because of the likelihood of microscopic systemic prostate cancer metastases.

However, the largest multiinstitutional retrospective series to date (n = 501 men) has provided a clearer view of the likelihood of benefit for various  subgroups of men with  local prostate  cancer  recurrence after prostatectomy.389 In this study, predictors of a poor response to salvage  radiation  included  Gleason  scores  of  8  to  10,  preradiation PSA level of greater than 2 ng/mL, PSA doubling time after prosta-tectomy of 10 months or less, negative surgical margins and seminal vesicle invasion. Nonetheless, a significant fraction of men with one or  more  of  these  negative  prognostic  features  still  experienced  a durable response to salvage radiation, particularly if the radiation was given prior  to a PSA  level of 2 ng/mL. Even  for  the group of men with the worst combination of prognostic features, a Gleason score of  8  to  10  and  a  preradiation  PSA  of ≥2 ng/mL,  salvage  radiation produced a progression-free  survival of 12% at 4 years.  In  another large retrospective study, Trock et al. reported that men treated with salvage  radiotherapy  resulted  in  a  3-fold  improvement  in  prostate cancer–specific survival compared with men who received no salvage treatment  (HR  =  0.32,  95%  CI  of  0.19  to  0.54;  P  <  0.001).390 Importantly, this benefit was limited to men with a prostate-specific antigen  doubling  time  of  less  than  6  months  and  remained  after adjustment  for  pathological  stage  and  other  established  prognostic factors. Although these findings hint at marked advantage to salvage radiation  therapy  after  prostatectomy,  the  studies  are  retrospective and  the  results  may  be  confounded  by  selection  bias.  Prospective studies  are  needed  to  definitively  determine  whether  patients  with one  or  more  poor  prognostic  features  derive  benefit  from  salvage radiation.  However,  given  the  available  data  and  considering  the limited morbidity of this treatment, especially with IMRT, it is rea-sonable to consider salvage radiation for postprostatectomy patients with PSA recurrence regardless of prognostic factors.

The role of ADT concomitant with radiation therapy given as an adjuvant to surgery as salvage for prostate cancer recurrence following surgery has not been established. In the RTOG 85-31 Trial, a sub-group of men with pathological T3N0 prostate cancer who under-went radical prostatectomy received adjuvant radiation therapy to a dose  of  60  to  65 Gy  and  then  were  randomized  to  immediate 

Page 27: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1489ProstateCancer • CHAPTER84

with those men treated with androgen suppression alone (11.9% vs. 23.9%),  a  benefit  also  manifest  as  a  reduction  in  overall  mortality (29.6% vs. 39.4%). These data firmly establish the role of radiation therapy both  in optimizing  local disease  control  and  in promoting survival of men with high-risk and locally advanced prostate cancer.

TREATMENT OF METASTATIC DISEASE

Natural History of Metastatic Prostate CancerOver the past two and a half decades, widespread and routine clinical use of serum PSA testing has changed not only screening and diag-nosis  of  prostate  cancer,  but  virtually  all  aspects  of  prostate  cancer management.408,409 One consequence of this changing practice pattern is that the conventional stage groupings (TNM staging), developed at a time when many men first presented for care with prostate cancer complicated by symptomatic distant metastases, no longer adequately anticipate the clinical behavior of relapsed prostate cancer seen more commonly today. Before the availability of serum PSA tests, men with newly diagnosed metastatic prostate cancer  faced a median survival of 24 to 30 months. Today, <5% of men with prostate cancer have clinical  evidence  of  distant  metastasis  at  the  time  of  presentation. Now, because most men initially are diagnosed with localized prostate cancer  and  undergo  primary  therapy,  the  natural  history  of  life-threatening  prostate  cancer  tends  to  be  more  protracted,  with  the disease  stereotypically progressing  through  clinical  states of  relapse, hormone naïvety, and castration-resistance (Figure 84-17).

Most prostate cancer relapses after local treatment are diagnosed based on rising serum PSA levels without other clinical or radiological evidence of local or distant recurrence (M0 disease). Often, men with M0 prostate cancer go on to receive ADT before metastases appear, usually resulting in a drop in serum PSA. For such men, a subsequent steady rise in PSA despite adequate hormonal therapy (serum testos-terone <50 ng/dL) has  been  interpreted  as  disease  progression  to  a nonmetastatic castration-resistant  phenotype.  In  this  setting,  further progression to clinically overt metastasis has been labeled metastatic castration-resistant prostate cancer.

Antonarakis et al. reported the experience of a cohort of men (n = 774) with biochemical recurrence after radical prostatectomy.410 At 8 years of follow-up after surgery, some 29% of these men had devel-oped  distant  metastases.  Recurrence  was  defined  as  a  serum  PSA rising to at least 0.2 ng/mL; the median PSA level at the time of the first  metastasis  was  31.4 ng/mL.  For  men  with  recurrent  prostate cancer, the PSA doubling time (PSADT; <3.0 vs. 3.0 to 8.9, vs. 9.0 to 14.9, vs. ≥15.0 months; P < 0.0001), pathological Gleason score (≤7 vs.  8  to 10; P =  0.005),  and  time  from prostatectomy  to PSA recurrence (≤3 vs. >3 years; P < 0.021) all correlated with metastasis-free survival and the 5-year probability of metastatic progression after PSA  recurrence.  Data  from  analyses  of  men  undergoing  radiation therapy appear similar, with PSADT emerging as the strongest pre-dictor of outcome.411

The clinical course of nonmetastatic castration-resistant prostate cancer is not as well described. Presumably, disease parameters such as PSA and PSADT are also likely provide prognostic information in this setting, but the behavior of PSA as a biomarker  in the castrate state may differ from its behavior in the hormone-naïve setting.412,413 The best data available thus far may be the experience of men with nonmetastatic castration-resistant prostate cancer patients enrolled in therapeutic  clinical  trials.  On  the  placebo  arm  of  a  randomized blinded trial of atrasentan for progressive castration-resistant prostate cancer and no radiographic evidence of bone metastases, by 2 years 46% of men had developed bone metastases and 20% had died.414 The median bone metastasis–free survival was 25 months. In a mul-tivariate  analyses,  baseline  PSA  ≥13.1 ng/mL  was  associated  with shorter  overall  survival  (HR =  2.34,  95% CI of  1.71  to 3.21; P < 0.0001), time-to-first bone metastasis (HR = 1.98, 95% CI of 1.43 to 2.74; P < 0.0001), and bone metastasis–free survival (HR = 1.98, 

beam  radiation  therapy  (experimental  treatment  arm).404  Eligible men  had  bulky  (>25 cm2)  locally  advanced  prostate  cancer  (stage T2b-T4, N0-N1, M0). Total androgen suppression involved gosere-lin acetate and flutamide;  radiation therapy was accomplished with 45 Gy delivered  to  the pelvis  followed by a 20-  to 25-Gy boost  to the prostate. Extended fields were used to treat the lymph nodes when they were involved. Findings were a significant advantage to androgen suppression,  and  radiation  exhibited  a  significant  benefit  in  local control, disease-free survival, and cause-specific mortality over radia-tion  alone.  A  second  RTOG  trial  (85-31)  also  targeted  men  with locoregionally advanced prostate cancer, including men with adverse findings at radical prostatectomy.367,405 The aim of this phase 3 trial was to evaluate the role of long-term adjuvant androgen suppression in men with high-risk prostate cancer. Men enrolled on the trial had clinical  stage T3  (>25 cm2)  prostate  cancer,  or T1–T2  disease  and radiographic or histologic evidence of lymph node involvement. Men were  eligible  after  prostatectomy  if  positive  surgical  margins  or seminal vesicle  invasion was  found. ADT featured goserelin acetate alone; radiotherapy was given at a dose of 44 to 46 Gy to the whole pelvis,  with  a  20-  to  25-Gy  boost  to  the  prostate  or  postoperative prostatic fossa for a total dose of 65 to 70 Gy. Again, the addition of androgen  deprivation  to  radiation  therapy  improved  local  disease control,  freedom  from distant metastases,  and disease-free  survival. For men with locoregionally advanced prostate cancers with a Gleason score of 8 to 10, a statistically significant difference in overall survival was also seen.405 The EORTC 22863 trial also tested the contribution of adjuvant ADT, given as goserelin acetate and cyproterone acetate for a month before radiation and goserelin acetate alone for 3 years after  the  radiation  treatment  course,  to  radiation  therapy  for  men with high-stage or high-grade prostate cancer.365,366 Radiation therapy was  delivered  as  50 Gy  to  prostate  and  regional  lymph  nodes,  fol-lowed by a 20-Gy boost to the prostate, for a total prostate dose of 70 Gy.  At  5  years,  the  overall  survival  for  the  men  who  received adjuvant endocrine therapy was 78% versus 62% for men who did not.366

The duration of adjuvant ADT given along with radiation therapy for  locally  advanced  prostate  cancer  has  also  been  tested.  In  the RTOG 92-02 Trial, short-term and long-term androgen suppression were compared.406 All men in the study received 4 months of gosere-lin  acetate  and  flutamide,  2  months  before  and  during  radiation therapy.  Men  were  then  randomized  to  receive  either  no  further therapy  (short-term  androgen  deprivation)  or  to  be  treated  with goserelin acetate  for an additional 24 months  (long-term androgen deprivation). Radiation dose was 65 to 70 Gy to the prostate and 44 to 50 Gy to pelvic nodes. At 5 years, the long-term androgen depriva-tion  group  showed  significant  improvement  in disease-free  survival of 54% versus 34% (P = 0.0001), in clinical local progression of 6.2% versus 13% (P = 0.0001), and in freedom from distant metastasis of 11% versus 17% (P = 0.001). Nonetheless, 5-year overall survival was not significantly different between the two treatment arms (78% vs. 79%),  except  for  the  subset of men with Gleason 8  to 10 prostate cancer, who showed a benefit to longer duration adjuvant endocrine treatment (80% vs. 69%). Despite these results, the optimal sequenc-ing of ADT and radiation has not been fully established. To date, it is not known whether the effects of ADT on prostate cancer control were merely additive to the tumoricidal effects of radiation or were synergistic, providing an enhancement of tumor killing.

Finally, some have questioned whether the improved local control and survival benefits of combining androgen suppression with radia-tion  for men with high-risk prostate cancer were  simply a  result of the  effects of  long-term androgen  suppression;  that  is, whether  the radiation therapy might not have been needed in this setting. To help answer  this  question, Widmark  et al.  conducted  a prospective  ran-domized  trial  (SPCG-7/SFUO-3) of  long-  term androgen  suppres-sion versus long-term androgen suppression plus pelvic radiation.407 In men (n = 875) receiving combined modality treatment, the 10-year cause-specific  mortality  was  significantly  reduced  when  compared 

Page 28: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1490

median survival is >80 months, with the majority of men manifesting only limited and asymptomatic metastatic disease.419

Treatments Targeting Androgen SignalingBecause androgen signaling contributes to the growth and survival of prostate cancer cells, interfering with the signaling axis tends initially to  benefit  nearly  all  men  with  advanced  prostate  cancer  (Figure 84-18). Typically, the use of ADT for prostate cancer, triggering a fall in serum testosterone to levels <50 ng/mL, results in a decline in the serum PSA,  as  a  consequence of AR  regulation of PSA expression, and  in a diminution  in metastatic  lesions accompanied by  relief of disease-associated  symptoms.  Unfortunately,  despite  continuous androgen deprivation, most men develop castration-resistant prostate cancer (CRPC). Remarkably, CRPC often shows evidence of contin-ued  addiction  to  androgen  signaling  even  though  serum  androgen levels have been therapeutically reduced.420 This phenotype was first demonstrated  in  human  prostate  cancer  xenograft  models,  where progression  to  castration  resistance was  found  to be  attributable  to increased AR abundance, allowing transcriptional trans-activation at lower androgenic hormone levels.421 AR itself  is a known target  for somatic genome alterations in prostate cancer, especially on progres-sion  of  the  disease  in  the  castrate  state.  AR  mutations,  encoding receptors with altered ligand specificity, can result in agonist activity for  antiandrogens,  providing  one  molecular  explanation  for  the curious  “antiandrogen  withdrawal”  syndrome,  in  which  men  with prostate cancer progression, despite treatment with a combination of androgen deprivation and antiandrogens, were found to benefit from discontinuation  of  the  antiandrogen.422  CPRC  cells  may  also  have augmented androgen signaling as a consequence of posttranslational 

95% CI of 1.45 to 2.70; P < 0.0001). The PSAV was also associated with overall  and bone metastasis–free  survival.  In  another placebo-controlled trial (prematurely terminated) of the bisphosphonate zole-dronic acid versus placebo, one-third of men assigned to the placebo arm  developed  bone  metastases  by  2  years,  with  a  median  bone metastasis–free survival of 30 months.415 In this trial, higher baseline PSA and PSAV were also associated with a shorter time-to-first bone metastasis and poorer survival.416

The  outcome  of  men  with  hormone-naïve  metastatic  prostate cancer  has  changed  substantially.  Clinical  trials  employing  similar hormonal  therapies  over  the  past  two  decades  illustrate  the  differ-ences. From 1989 to 1993, the SWOG conducted a trial in men (n = 1387) with newly diagnosed metastatic prostate cancer treated with bilateral orchiectomy with or without the nonsteroidal antiandrogen flutamide, which showed no significant differences between the treat-ment  arms  and  an  overall  median  survival  of  33  months.417  From 1995  to  2009,  the  SWOG  conducted  another  study  in  the  same patient population, testing continuous versus intermittent treatment with GnRH analog and bicalutamide. In this trial, the median sur-vival for continuous hormonal treatment was 49 months, suggesting a  30% decrease  in  the  risk  of  death  for  the  same  treatment  in  the more recent trial versus the older one.418 This apparent improvement in outcome likely reflects the same type of stage migration effect as has been seen for almost all prostate cancer states in the PSA era.

In addition, fewer men contend with major disease-related symp-toms,  such as hemoglobin <10 g/dL, poor performance  status,  and extensive  metastatic  disease  burden,  a  lead-time  effect  almost  cer-tainly attributable to serum PSA monitoring. For men with prostate cancer  recurrence  after  local  treatment  followed  without  the  use  of  androgen deprivation until  radiographic disease progression,  the 

Figure 84-17 •  Sites of action of different treatments that affect androgen action. ACTH, adrenocorticotropic hormone; AR, androgen receptor; DHT, 5α-dihydrotestosterone;  ER,  estrogen  receptor;  GnRH,  gonadotropin-releasing  hormone;  LH,  luteinizing  hormone;  PR,  progesterone  receptor;  T, testosterone. 

Hypothalamus– PR

AR

AR

AR

ER

AR

ER

+

GnRH

GnRH-R

Estrogens

GnRH analogsand antagonists

Antiandrogens Cyproterone acetate Megestrol acetate Flutamide Bicalutamide Nilutamide Enzalutamide

ACTH

Cholesterol

Adrenal androgens

Adrenalglands

Androgen synthesis inhibitorsGlucocorticoidsKetoconazoleAbiraterone acetate

Adrenalandrogens

Gene expression

Prostate

Peripheralorgans

Pituitary

Testis

DHT

T

LHProlactin

Testosterone

T DHT

Cholesterol

Testosterone

AndrogenSynthesisInhibitors

LH-R

5α-reductase inhibitors

Page 29: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1491ProstateCancer • CHAPTER84

Figure 84-18 •  A  meta-analysis  of  survival  from  metastatic  prostate cancer with maximal androgen blockade versus androgen deprivation alone. (From  Prostate  Cancer  Trialists  Cooperative  Group.  Maximum  androgen blockade  in advanced prostate cancer: an overview of  the randomised trial. Lancet 2000;355:1491-1498.)

0 5 10

100

80

60

40

20

0

25.4%

23.6%

6.2%

5.5%

Pro

port

ion

aliv

e (%

)

Time since randomization (yrs)

8000 prostate cancer patients in27 trials of antiandrogen (nilutamide,flutamide, or cyproterone acetate)

Absolute difference1.8% (SE 1.8)

Treatment betterby 0.7% (SE 1.1)logrank 2p>0.1

Androgen suppression onlyAndrogen suppression and antiandrogen

Figure 84-19 •  Responses to enzalutamide in men with castration-resistant prostate cancer who had or had not received previous chemotherapy. (From Scher HI et al. Antitumour activity of MDV3100 in castration-resistant prostate cancer: a phase 1-2 study. Lancet 2010;375:1437-1446.).

PS

A c

hang

e fr

omba

selin

e (%

)

Patient number

–100

–75

–50

–25

0

25

1

No previous chemotherapy (n = 65) Previous chemotherapy (n = 75)

5 9 13 17 21 25 29 33 37 41 45 49 53 57 61 65 69 73 77 81 85 89 93 97 101

105

109

113

117

121

125

129

133

137

modifications of AR or its coactivators driven by growth factor signal-ing pathways, of the expression of AR forms generated from mRNA splice  variants,  and  of  intratumoral  production  of  androgenic hormones.10,420,423,424

Androgen Deprivation TherapyADT for prostate cancer involves reduction of circulating testosterone levels  to <50 ng/mL, most often accomplished via  surgical  removal of the testis (bilateral orchiectomy) or via inhibition of the synthesis and  release  of  pituitary  gonadotropins  with  long-acting  luteinizing hormone–releasing hormone (LHRH) analogs (goserelin, leuprolide, buserelin, histrelin, and triptorelin) and antagonists (abarelix, degare-lix, and cetrorelix) which cause suppression of  testosterone produc-tion  (Figure 84-18). LHRH analogs first  trigger LH  release by  the pituitary, triggering a short-lived elevation of serum testosterone that is occasionally associated with a symptomatic disease “flare” in men with  extensive  bone  metastasis.425  LHRH  antagonists  more  rapidly suppress testosterone levels without such a “flare.” Historically, reduc-tions  in  testosterone were also accomplished via  the administration 

of pharmacological doses of  synthetic estrogens,  such as diethylstil-bestrol, as the first “medical castration” (DES).426 An extensive body of evidence from many prospective randomized clinical trials for men with metastatic prostate cancer  supports  the comparable efficacy of bilateral  orchiectomy,  LHRH  analogs  and  antagonists,  and  DES. However, treatment with DES, in contrast to orchiectomy or to treat-ment with LHRH analogs or antagonists, has been complicated by serious  treatment  adverse  events,  including  congestive  heart  failure and venous thromboemboli. As a consequence, estrogens have been virtually abandoned in favor of LHRH agonists and antagonists for men  with  advanced  prostate  cancer.  Also,  many  men  find  LHRH therapy more acceptable than bilateral orchiectomy.

Antiandrogens and “Complete” Androgen BlockadeAntiandrogens interact directly with the AR to antagonize its trans-activation of  target gene  transcription (Figure 84-18). Many of  the current  antiandrogen  drugs  (flutamide,  nilutamide,  and  bicaluta-mide) were developed initially to be used in combination with LHRH analogs,  with  the  hope  that  such  treatment  combinations  could achieve a “complete” androgen blockade by reducing androgen levels and neutralizing the propensity for adrenal androgen precursors, such as  androstenedione  and  dehydroepiandrosterone,  to  activate  AR function.427 Early reports of the efficacy of this treatment combina-tion prompted the conduct of 27 clinical trials involving 7987 men testing whether “complete” androgen blockade offered any advantage over ADT alone for men with metastatic prostate cancer.428 A com-prehensive review of these trials revealed that only 3 of the 27 studies showed any benefit to the combined approach (Figure 84-19).428 The Agency  for  Health  Care  Policy  and  Research  (AHCPR;  see  report number  99-E012  at  http://www.ahcpr.gov/clinic/index.html)  con-ducted a meta-analysis of all published “complete” androgen blockade clinical trials, finding no difference in 2-year survival rates, and only a  minimal  difference  in  5-year  survival  felt  to  be  of  questionable clinical significance, for combined treatment.

Newer  antiandrogens  promise  to  overcome  some  of  the  mixed agonist/antagonist properties of the first set of available AR-targeting drugs.429  One  of  these  agents,  enzalutamide,  which  reduces  AR ingress into the cell nucleus and impairs its binding to target DNA sequences nuclear translocation, has been subjected to extensive clini-cal trials for men with CRPC, earning FDA approval for this indica-tion in the summer of 2012 (Figure 84-20).430 In one phase 3 trial (the AFFIRM Study), men (n = 1199) with metastatic CRPC that had  progressed  despite  treatment  with  docetaxel  were  randomized 2 : 1 to receive enzalutamide or placebo.431 In this study, enzalutamide therapy was accompanied by a survival benefit (HR = 0.63, 95% CI 

Page 30: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1492

95%. However, in men with prostate cancer, there can be continuous ongoing  production  of  androgens  by  the  adrenal  glands  or  by  the cancer itself.10,433 This nongonadal source of androgens is significant: even in the castrate state, intraprostatic concentrations of testosterone and dihydrotestosterone appear sufficient to stimulate AR. In prostate cancers, such androgens may be able to arise both by local conversion of  adrenal  precursors  and  by  de  novo  synthesis  through  increased expression of steroidogenic enzymes such as CYP17.10,433 Many pros-tate cancer clinicians used high doses of ketoconazole to antagonize this process, often noticing benefits for some men with CRPC. More recently, abiraterone acetate, a pregnenolone analog and much more potent inhibitor of CYP17 than ketoconazole, was approved by the FDA for the treatment of CRPC, based on improvements in survival in randomized trials (Figure 84-21).434 Adverse effects of abiraterone treatment,  though minimal,  included stigmata of  secondary miner-alocorticoid excess,  including fluid retention (30.5%), hypokalemia (17.1%),  and  hypertension  (1.3%).  TAK-700,  another  selective CYP17  inhibitor,  has  reached  phase  3  clinical  trials  for  men  with CRPC.435

Optimal Timing of ADTAlthough there is a general belief that immediate initiation of ADT for men with metastatic prostate  cancer may best  improve disease-related  quality  of  life,  there  is  no  compelling  evidence  of  compro-mised  survival  resulting  from  deferred  treatment  at  the  time  of symptomatic  progression.  The  timing  of  ADT,  and  its  effects  on quality  of  life  and  survival  for  men  with  recurrent  or  metastatic prostate  cancer,  has  become  an  even  more  critical  issue  recently because an increasing number of men are recognized to have recur-rent or metastatic prostate cancer very early because of  increases  in the  serum  PSA.  Randomized  studies  conducted  by  the  Veterans Administration  Cooperative  Urological  Group  (VACURG)  several decades  ago  indicated  that  men  treated  with  placebo  initially  who subsequently  received ADT at  the  time of  clinical  progression  suf-fered no worse a  survival  than men who were  initially  subjected  to ADT.436 More recently,  the Medical Research Council  (MRC) con-ducted a study in men (n = 934) with prostate cancer who did (n = 434)  or  did  not  (n  =  500)  have  metastasis,  randomizing  the  men either  to  immediate  ADT  or  to  ADT  at  the  time  of  symptomatic disease progression.437 Though early results from this study hinted at a survival advantage for men treated with early ADT, a phenomenon 

Figure 84-20 •  Results  of  a  randomized  trial  of  abiraterone  acetate versus  placebo  for  men  with  castration-resistant  prostate  cancer  (CRPC). (Data from De Bono JS et al. Abiraterone and increased survival in metastatic prostate cancer. N Engl J Med 2011;364:1995-2005.)

Placebo

Abirateroneacetate

Overall survival

Sur

viva

l (%

)

Months

No. at riskAbiraterone acetatePlacebo

0

20

40

60

80

100

0 3 6 9 12 15 18 21

797398

736355

657306

520210

282105

6830

23

00

Figure 84-21 •  Results of a  randomized clinical  trial of docetaxel, given at  two different dosing schedules, and pred-nisone  versus  mitoxantrone  and  prednisone  for  men  with androgen-independent metastatic prostate cancer. (Data from Tannock IF, de Wit R, Berry WR, et al. Docetaxel plus pred-nisone or mitoxantrone plus prednisone for advanced prostate cancer. N Engl J Med 2004;351:1502-1512.)

0 3 6 9 12 15 18 21 24 27 30

335334337

296297297

217200192

10410595

372929

543

33

Docetaxelevery 3 wkWeekly

docetaxel

MitoxantronePro

babi

lity

of o

vera

ll su

rviv

al (

%)

Months

No. at RiskDocetaxel every 3 wkWeekly docetaxelMitoxantrone

0

10

20

30

40

50

60

70

80

90

100

of  0.57  to  0.75;  P <  0.0001),  with  an  increase  in  median  survival from  13.6  months  to  18.4  months.  The  second  phase  3  study (PREVAIL) randomized men (n = 1680) with metastatic CRPC who had not yet received docetaxel to enzalutamide or placebo; data from this  study  are  not  yet  available.  A  second  agent,  ARN-509,  which also  appears  to  lack  partial  AR  agonist  activity,  to  prevent  nuclear translocation of AR, and to impair AR binding to DNA, has entered early clinical studies.432 The use of this new generation of antiandro-gens,  as  single  agents  or  in  combination  with  ADT,  for  the  initial treatment of advanced prostate cancer is currently under active evalu-ation in clinical trials.

Inhibitors of Adrenal SteroidogenesisADT,  which  principally  targets  gonadal  production  of  androgenic hormones, decreases total serum testosterone levels by approximately 

Page 31: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1493ProstateCancer • CHAPTER84

mitosis, yet prostate cancers tend to have low growth fractions when compared to many normal cells and most other cancers. An unbiased screen of FDA-approved drugs for toxicity against prostate cancer cell lines unmasked a striking activity of all microtubule-targeted drugs, including colchicine and many others in addition to taxanes, against prostate cancers.447 One possibility is that microtubule function may be needed for other critical cell processes, such as nuclear-cytoplasmic shuttling of  critical  regulatory proteins  essential  for prostate  cancer cell viability, including translocation of AR from the cytoplasm into the nucleus.448

DocetaxelIn early clinical trials, both paclitaxel and docetaxel showed promising activity against prostate cancer, but only docetaxel was subjected to large-scale testing for improving survival of men with mCRPC. Early experiences with docetaxel, given alone in two different dosing sched-ules  and  given  along  with  estramustine,  prompted  the  design  and execution of two definitive randomized trials, TAX-327 and SWOG-9916, which established the survival benefit of docetaxel for mCRPC. The TAX-327 study (n = 1006), considered the pivotal trial for FDA approval  of  docetaxel  in  prostate  cancer,  included  three  treatment arms:  docetaxel  75 mg/m2  every  3  weeks  plus  prednisone  (10 mg daily), weekly docetaxel 30 mg/m2 for 5 of 6 weeks with prednisone, and mitoxantrone 12 mg/m2 every 3 weeks with prednisone.449 The docetaxel regimen featuring every 3-week administration conferred a significantly improved survival (HR = 0.76; 85% CI of 0.62 to 0.92), higher PSA responses (45% vs. 42%), and better pain relief (22% vs. 13%) than mitoxantrone. Weekly docetaxel  showed a trend toward improved survival, a benefit that did not reach statistical significance. Grade  3/4  toxicity  of  docetaxel  was  limited  mostly  to  neutropenia (32%), with an incidence of febrile neutropenia at only 3% or less. Poor  prognostic  factors  for  survival  of  men  with  mCRPC  treated  with  docetaxel  include  liver  metastases,  many  metastatic  sites,  significant pain, poor performance  status,  radiographic  evidence of disease  progression,  high  PSA  level,  low  PSA  doubling  time,  high tumor  grade,  elevated  serum  alkaline  phosphatase,  and  low  blood hemoglobin.450

The  second  study,  SWOG-9916,  evaluated  the  combination  of docetaxel plus estramustine versus mitoxantrone plus prednisone.451 Again,  the  overall  survival  was  superior  in  the  group  receiving docetaxel (HR = 0.80, 95% CI of 0.67 to 0.97). The study subject characteristics were similar to the men enrolled in the TAX-327 study, as were the survival outcomes. The incorporation of estramustine in the docetaxel regimen, however, was characterized by increased gas-trointestinal  and  peripheral  vascular  toxicity,  including  venous thromboembolism (VTE). Because of  the high rate of VTE on the trial, prophylactic low-dose warfarin and aspirin were added, although this  did  not  reduce  the  VTEs  for  men  receiving  estramustine  and docetaxel.  With  the  hazards  associated  with  estramustine  use,  and little evidence for a dramatic benefit for estramustine docetaxel versus docetaxel alone, the estramustine and docetaxel regimen is not cur-rently recommended.

Although the TAX-327 study was designed to include as many as 10 doses of docetaxel given at 3-week intervals, there were not strong data supporting the need for that many treatment cycles. Typically, in  clinical practice,  rather  than delivering 10  consecutive docetaxel treatment  cycles,  the  taxane  is  administered  for  5  to  8  cycles  until disease  stabilization  is  evident,  holding  treatment  for  when  subse-quent  disease  progression  becomes  evident,  whether  manifest  as  a rising serum PSA, clinical worsening, or both. This approach allows for docetaxel retreatment before moving to second-line chemotherapy treatment, occasionally extending disease control without increasing treatment toxicity.452,453

CabazitaxelCabazitaxel  is  a  second-generation  taxane  that  exhibited  cytotoxic activity against a broad range of cancer cell lines and tumor models 

that may have been attributable to underuse of ADT in the delayed therapy arm (54% of men died without endocrine treatment). Sub-sequent  reports  from  the  trial,  with  greater  follow-up  times,  have shown no significant mortality difference  form early versus delayed ADT.

Nonetheless, in a relatively small prospective randomized trial of immediate  ADT  versus  observation  in  men  (n  =  98)  with  node-positive prostate cancer after radical prostatectomy, a benefit to early intervention was evident.438 This study is frequently cited as evidence in support of earlier ADT in men with node-positive prostate cancer, and  is  often  extrapolated  to  rationalize  the  use  of  early  ADT  for almost  recurrent prostate  cancer. Despite  the  limited data  available in this clinical setting, the common use of serial PSA testing to diag-nose  early  disease  relapse  following  primary  treatment  has  pushed practice  patterns  toward  early  ADT.  Against  this  approach  is  the growing body of evidence about the health hazards of ADT, including increases  in  cardiovascular  events,  accelerated  bone  loss,  and  so forth.439 For  these  reasons,  the  relative benefits of  early ADT  for  a rising serum PSA after primary therapy have not been established.

One popular approach for men with asymptomatic recurrent or metastatic prostate cancer has been the use of intermittent ADT. This strategy was suggested from provocative preclinical models suggesting that  intermittent  reductions  in  serum  testosterone  might  delay  the emergence of CRPC.440,441 Unfortunately, available clinical trials have provided little in the way of definitive conclusions regarding the rela-tive  efficacy  of  intermittent  versus  continuous  ADT,  or  definitive guidance  about  the  most  effective  ADT  regimen  for  intermittent administration, the optimal timing or duration of intermittent ADT cycles, how to use serum PSA to aid treatment, and the best definition of  treatment  response  versus  treatment  failure.442  Despite  all  the unanswered questions, intermittent ADT tends to be frequently used for  prostate  cancer  in  the  clinic,  especially  for  men  with  prostate cancer and no evidence of clinical metastasis, or with minimal meta-static disease, who have had complete or near complete responses to ADT, have no active disease-related symptoms, and may be tolerating ADT poorly. Other practical advantages of intermittent ADT include improvement  of  quality  of  life  and  attenuation  of  complications accompanying long-term ADT.443

One  fairly  large  randomized  trial  of  men  with  nonmetastatic recurrent prostate cancer (n = 1386) testing intermittent versus con-tinuous  ADT  showed  a  similar  overall  survival  for  each  treatment schedule (HR = 1.02, 95% CI of 0.86 to 1.21; P for noninferiority = 0.009). Preliminary data from the trial presented in abstract form hinted  that men  treated with  intermittent ADT were  less  likely  to die from prostate cancer, but more likely to die from non–prostate-cancer-related causes.444  In another  large  trial, men with metastatic prostate  cancer  (n = 3040) were  randomized  to  intermittent versus continuous  ADT  after  an  initial  course  of  induction  ADT  over  7 months.445 Responders to the induction regimen (men with a serum PSA ≤4.0 ng/mL at 7 months) went on to receive intermittent ADT or  continuous  ADT  until  the  PSA  rose  to  20 ng/mL  or  clinical disease progression was  seen. Although  the overall  survival was not substantially different for the treatment arms, in a subset of men with minimal prostate cancer (axial skeleton ± lymph node metastasis), a ≥20% inferiority for intermittent ADT could not be excluded.

Cytotoxic Chemotherapy for Metastatic Castration-Resistant Prostate CancerAlthough many systemic chemotherapy drugs have been used for the treatment of metastatic castration-resistant prostate cancer (mCRPC), only  taxanes  have  produced  significant  survival  benefits.  Taxane drugs, including docetaxel and cabazitaxel approved for mCRPC by the  FDA,  disrupt  microtubule  function  and  promote  apoptosis.446 Remarkably, the exact mechanisms by which these drugs act against prostate  cancer  cells  is  not  entirely  clear.  Microtubule  function  is critically  required  for  segregation  of  replicated  chromosomes  at 

Page 32: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1494

against  malignant  cells  while  overcoming  tumor-induced  toler-ance.457  Although  not  traditionally  considered  a  disease  amenable to  immune-directed  approaches,  prostate  cancer  may  be  an  ideal target  for  immunologic attack because of  its  slow growth, allowing a stimulated immune system time to generate an antitumor response, and  its  propensity  to  produce  lineage-specific  proteins  that  may serve as tumor antigens, including PSA and prostatic acid phospha-tase  (PAP).  A  number  of  prostate  cancer  vaccine  approaches  have been  subjected  to  early  clinical  development  to  exploit  this  oppor-tunity.458 One such approach, featuring sipuleucel-T, an autologous PAP-loaded  antigen-presenting  cell  preparation  has  been  advanced to FDA approval.

To  permit  treatment  with  sipuleucel-T,  antigen-presenting  cells (APCs) are collected by leukapheresis from men with prostate cancer and incubated ex vivo with a fusion protein formed from PAP linked to GM-CSF. Subsequently, the primed APCs are reinfused, activating T  cells  to  promote  a  PAP-directed  antitumor  response.459  In  early randomized trials of sipuleucel-T, though only 5 of the 147 men with prostate  cancer  who  received  sipuleucel-T  had  PSA  reductions  of ≥50%,  and  no  objective  tumor  regressions  were  seen,  an  improve-ment in overall survival was evident.460 This prompted a phase 3 trial of men (n = 512) with mCRPC and no visceral metastases or need for narcotics  for pain  (IMPACT).461 The men were  randomized  to sipuleucel-T or a placebo, with an opportunity to receive sipuleucel-T at the time of disease progression (using APCs cryopreserved at the time of placebo preparation); the primary end point for the trial was overall survival. When the trial results were reported, 61% of the men treated with  sipuleucel-T and 70.8% of  the men  receiving placebo had died (HR = 0.78l, P = 0.03). Based on the results of this study, the  FDA  approved  sipuleucel-T  in  April  of  2010  for  men  with asymptomatic  or  minimally  symptomatic  mCRPC.  As  seen  in  the earlier studies, neither time to disease progression nor PSA response rates  were  improved  by  sipuleucel-T  treatment  compared  with placebo.  Only  one  sipuleucel-T–treated  subject  in  the  IMPACT study had partial  radiographic  response, and only 8 of 311 (2.6%) men  experienced  PSA  reductions  of  ≥50%  with  sipuleucel-T. Sipuleucel-T was well tolerated, with adverse events such as fatigue, influenza-like symptoms, chills, low-grade fever, arthralgias/myalgias, and  headaches  appearing  and  resolving  within  24  to  48  hours  of infusion.

Bone-Targeted TreatmentsMen with metastatic prostate cancer are at increased risk of skeletal complications for two main reasons. First, chronic androgen depriva-tion  increases  bone  resorption  and  reduces  bone  mineral  density, resulting  in  osteopenia  and/or  osteoporosis.  Second,  bone  involve-ment  in prostate  cancer  is  extremely  common and occurs  in up  to 75%  of  patients  with  metastatic  disease.  Complications  of  bone metastasis include bone pain, hypercalcemia of malignancy (though rare  for  prostate  cancer),  and  other  skeletal-related  events  (SREs) including  pathological  fractures,  need  for  radiation  or  surgery  to bone, or spinal cord compression.

Zoledronic AcidFor  several  years,  intravenous  bisphosphonates  have  served  as  the mainstay of adjunctive treatment to help maintain skeletal integrity in  patients  with  bone-metastatic  prostate  cancer.  The  benefits  of bisphosphonates were shown definitively in a landmark phase 3 study involving men (n = 643) with mCRPC and asymptomatic or mini-mally  symptomatic  bone  metastases,  in  which  zoledronic  acid  was compared against placebo.462,463 This study demonstrated fewer SREs in  the  zoledronic  acid  group  compared  with  the  placebo  group (33.2% vs. 44.2%; P = 0.021) and an increased median time to first SRE (16.1 vs. 10.6 months; P = 0.009). The results of this study led to the FDA approval of zoledronic acid for men with bone-metastatic CRPC, for the purposes of SRE prevention.

with  greater  potency  than  docetaxel  in  multidrug-resistant  tumor cells, prompting its development as an agent to overcome docetaxel resistance,  perhaps  because  of  its  weaker  binding  affinity  to P-glycoprotein, a drug efflux pump that serves to reduce intracellular concentrations of docetaxel.454 An additional characteristic of cabazi-taxel is its ability to penetrate the blood–brain barrier in vivo, which is not achievable with other taxanes. A phase 1 trial determined that the  principal  dose-limiting  toxicity  of  cabazitaxel  was  neutropenia, with  other  observed  side  effects,  including  nausea,  vomiting,  diar-rhea, and fatigue, appearing generally mild to moderate.455 Two men in  the  trial with mCRPC had partial  responses, with  reductions  in measurable disease deposits and decrements in serum PSA levels.

The phase 3 trial (TROPIC) was a randomized, open-label, mul-tinational study testing whether cabazitaxel plus prednisone improved overall survival compared with mitoxantrone plus prednisone in men (n = 755) with mCRPC that had progressed either during or  after prior docetaxel chemotherapy.456 Men in the trial were given predni-sone 10 mg/d and then either cabazitaxel 25 mg/m2 or mitoxantrone 12 mg/m2 every 3 weeks for 10 cycles. Of note, some 70% of men enrolled  in  the  trial  had  had  progressive  prostate  cancer  within  3 months  of  completing  docetaxel  treatment,  whereas  the  remainder showed  disease  progression  during  docetaxel  therapy.  Trial  results demonstrated  that  cabazitaxel  improved  survival  in  this  clinical setting over mitoxantrone (HR = 0.70, 95% CI of 0.59 to 0.83; P < 0.0001), doubling the median progression-free survival and objective response rates. PSA response rates were also significantly higher  for cabazitaxel treatment.

Men  treated  with  cabazitaxel  received  a  median  of  6  cycles  of treatment,  and  the most common adverse events were hematologic toxicities. For example, 81.7% and 7.5% of men in the cabazitaxel group  experienced  grade  ≥3  neutropenia  and  febrile  neutropenia, respectively, as compared with 58% and 1.3% for men in the mito-xantrone group. Cabazitaxel  treatment  also  increased  the  incidence of nonhematologic  toxicities,  including diarrhea  (46.6% vs. 10.5% for  mitoxantrone  with  grade  ≥3  diarrhea  in  6.2%  vs.  0.3%)  and asthenia (20.5% vs. 4.6% for mitoxantrone with grade ≥3 asthenia in 12.4% vs. 2.4%). Some 5% of men on the cabazitaxel arm died from causes other than disease progression within 30 days of receiving the  drug. This  compares  with  a  1%  drug-related  death  rate  in  the mitoxantrone group. The most common cause of death  in patients who  were  treated  with  cabazitaxel  was  neutropenia,  including  its clinical  consequences  such  as  septicemia. This  problem  was  mini-mized during  the  conduct of  the  trial when TROPIC  investigators were  informed  about  events  and  directed  to  strictly  adhere  to  the study protocol measures  for dose delays and modifications, and for the  management  of  neutropenia  with  granulocyte  macrophage colony-stimulating factor (GM-CSF) according to American Society for Clinical Oncology (ASCO) guidelines.

Based on the results of the TROPIC trial, cabazitaxel was approved by  the  FDA  in  June  2010  and  by  the  EMEA  in  2011  for  use  in  combination  with  prednisone  for  the  treatment  of  docetaxel-pretreated  mCRPC.  Cabazitaxel  is  currently  being  evaluated  as  a first-line treatment of mCRPC in two separate clinical trials. The first study  (FIRSTANA)  is  a  standard multi-institutional, multinational phase 3 comparison between cabazitaxel (in two doses: 25 mg/m2 and 20 mg/m2)  versus  standard  every-3-weeks  docetaxel  (75 mg/m2)  in men with chemotherapy-naïve mCRPC. The second study (TAXYN-ERGY)  is  a  multicenter  randomized  phase  2  trial  to  evaluate  the benefit  of  an  early  switch  from  first-line  docetaxel/prednisone  to cabazitaxel/prednisone  (or  the  opposite  sequence)  in  men  with mCRPC not previously treated with chemotherapy.

Immunotherapy using Sipuleucel-TAn  alternative  strategy  for  the  treatment  of  mCRPC  involves  the use  of  immunotherapeutic  agents.  Cancer  immunotherapy  refers generally to attempts to treat cancer by activating immune responses 

Page 33: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1495ProstateCancer • CHAPTER84

metastases  is  estimated  nearly  40%  to  60%  of  the  administered dose.470 Following encouraging phase 2 findings of  significant anti-tumor activity with a possible suggestion of a survival benefit, 223Ra was  subjected  to  a  randomized,  placebo-controlled,  phase  3  trial (ALSYMPCA).471,472 This study enrolled men (n = 922) with symp-tomatic  bone-metastatic  CRPC,  who  had  previously  received docetaxel or were considered unfit for docetaxel, for treatment with 223Ra (50 kBq/kg) or placebo given in six injections at 4-week inter-vals. At a planned interim analysis, 50% of men treated with 223Ra versus 35% of men receiving placebo had received all six injections, and the overall survival was superior for the 223Ra treatment arm (HR = 0.695, 95% CI of 0.552 to 0.875; P = 0.0019). The overall survival advantage  for  223Ra met  the predetermined  statistical  boundary  for stopping the trial early, and the Independent Data Monitoring Com-mittee (IDMC) recommended halting the trial in June 2011 because of evidence of significant treatment benefit.

In subset analyses, men with a good performance status and men who had not previously been treated with docetaxel showed greater benefits, whereas prior treatment with bisphosphonates did not affect outcome. Among the secondary endpoints, median time to first SRE was significantly improved in the 223Ra treatment arm compared with placebo  (13.6  vs.  8.4  months;  HR  =  0.610,  95%  CI  of  0.461  to 0.807; P = 0.0005), as was the time to alkaline phosphatase progres-sion (HR = 0.163, 95% CI of 0.121 to 0.221; P < 0.00001) and the time to PSA progression (HR = 0.671, 95% CI of 0.546 to 0.826; P  =  0.0002).  These  results  support  a  significant  improvement  in controlling  metastatic  bone  disease  with  223Ra. The  safety  of  223Ra administration was encouraging, with less adverse events in the 223Ra treatment arm and no worrisome toxicities ascribable to 223Ra expo-sure. This safety profile for 223Ra, especially in comparison with the bone-targeting  β–emitters,  may  allow  more  liberal  dosing  and/or extended treatment periods. Also, if 223Ra can control bone metastases without  injuring hematopoiesis  in  the bone marrow,  combinations with  myelosuppressive  chemotherapy,  like  docetaxel  or  cabazitaxel, may be feasible. Each of these expanded uses of 223Ra awaits testing in controlled clinical trials.

SUMMARYAlthough mortality from prostate cancer has declined over the past few years, demographic trends, such as the general aging of the popu-lation,  suggest  that  prostate  cancer  will  remain  one  of  the  most common health threats for men in the developed world. Widespread implementation of prostate cancer screening, using the serum PSA, has  resulted  in a  changing character of prostate  cancer  at  its  initial presentation, with younger men being diagnosed at  earlier prostate cancer  stages  than  ever  before.  The  use  of  serum  PSA  testing  for disease  activity  monitoring  has  changed  the  character  of  prostate cancer throughout the rest of its natural history, with healthier men having  less  prostate  cancer  at  later  stages  of  the  disease  than  ever before. These changes have put new demands on improving prostate cancer treatment, whether offering active surveillance to selected men with indolently progressive disease, minimizing the morbidity of local prostate cancer  treatment, or continuing  to  improve  the efficacy of systemic  prostate  cancer  treatment.  The  androgen  signaling  axis remains  the  major  systemic  treatment  target  for  advanced  prostate cancer,  with  new  agents  introduced  to  restrain  ongoing  androgen receptor activation in castration-resistant disease.

DenosumabMore  recently,  denosumab,  a  monoclonal  antibody  blocking  the receptor  activator  of  nuclear  factor-κB  (RANK)  ligand,  has  also gained FDA approval for the prevention of SREs in bone-metastatic CRPC, based on the finding from a large randomized phase 3 trial.464 RANK ligand is a member of the tumor necrosis factor family that binds to its receptor on immature and mature osteoclasts, promoting differentiation, activation, and survival of osteoclasts.465 By blocking the binding of RANK ligand to RANK, denosumab effectively inhib-its  osteoclastic  activity  and  thus  osteoclast-mediated  bone  resorp-tion.466 Denosumab was developed for treatment of skeletal diseases mediated  by  osteoclasts,  including  metastatic  bone  disease  from  a variety of cancers, multiple myeloma, and hormonal therapy–induced bone loss accompanying cancer treatment.

Denosumab, given as 60 mg subcutaneously every 6 months, was initially studied in men (n = 1468) receiving ADT for nonmetastatic prostate cancer in a randomized, placebo-controlled, phase 3 trial.467 Denosumab therapy was associated with increased bone mineraliza-tion at all sites, with an increase in the bone mineral density of the lumbar  spine  of  5.6%,  as  compared  with  a  loss  1.0%  for  placebo treatment (P = 0.001). Men receiving denosumab along with ADT also had a decreased incidence of new vertebral fractures at 36 months (1.5% vs. 3.9% with placebo; HR = 0.38, 95% CI of 0.19 to 0.78; P = 0.006).

For men with prostate cancer bone metastases, a second phase 3 trial was undertaken, comparing denosumab, given as 120 mg sub-cutaneously every 4 weeks, with zoledronic acid, given as 4 mg intra-venously  every  4  weeks.464 The  primary  end  point  of  the  trial  was time to first SRE, including pathological fracture, need for radiation therapy or surgery, and spinal cord compression. At the time of the efficacy analysis, an improvement in time to first SRE for denosumab was evident  (HR = 0.82, 95% CI of 0.71  to 0.95; P = 0.0002  for noninferiority; P = 0.008 for superiority). There were no differences in overall survival or  in progression-free survival. As a result of this pivotal trial, denosumab gained approval by the FDA in November 2010 for use in men with bone involvement from CRPC. Common toxicities of denosumab include fatigue, nausea, hypophosphatemia, hypocalcemia  (5%  grade  ≥3),  and  osteonecrosis  of  the  jaw  (2%). Therefore denosumab  is  a  reasonable  alternative  to  zoledronate  for the prevention of SREs in patients with metastatic CRPC, and also has the advantage that it does not require dose adjustment or moni-toring for renal impairment. However, men treated with denosumab should  undergo  calcium  supplementation,  and  should  also  have  a careful dental examination prior to initiating therapy.89Strontium, 153Samarium, and 223RadiumIn addition to  the osteoclast-inhibitory maneuvers discussed above, bone-seeking radiopharmaceuticals have also been used in the man-agement  of  patients  with  symptomatic  bone-metastatic  prostate cancer. To date,  two  such agents have been approved by  the FDA: 89Sr  and  153Sm. Both of  these drugs  are  indicated  for  the palliative treatment  of  multifocal  bone  pain  caused  by  widespread  osseous metastases, but neither drug has demonstrated a survival benefit when given to patients with mCRPC.468 89Sr and 153Sm both emit β–par-ticles with track lengths up to a few millimeters resulting in collateral bone marrow toxicity when absorbed in areas of new bone formation. In contrast, α-particles appear to provide more dense ionizing radia-tion (high linear energy transfer radiation) over a narrower range of <100 µm (corresponding to 2 to 10 tumor cell diameters), minimiz-ing myelotoxicity.469

223Ra, an α–emitting calcium mimic, has been subjected to intense scrutiny  for  the  treatment  of  prostate  cancer  bone  metastases. The agent has bone-seeking proclivity without  requiring a  carrier, has  a suitable half-life  (t1/2 = 11.4 days)  allowing convenient dosing,  and has a safe short-lived daughter isotope (219Rn with t1/2 = 4.0 seconds).469 In addition, the total skeletal uptake of 223Ra in men with osteoblastic 

The complete reference list is available online at www.expertconsult.com.

Page 34: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1496

R E F E R E N C E S2.  Barry  MJ.  Screening  for  prostate  cancer—the  

controversy  that  refuses  to  die.  N  Engl  J  Med 2009;360(13):1351–4.

6.  Walsh PC, Lepor H, Eggleston JC. Radical pros-tatectomy  with  preservation  of  sexual  function: anatomical  and  pathological  considerations.  The Prostate 1983;4(5):473–85.

10.  Green SM, Mostaghel EA, Nelson PS. Androgen action and metabolism in prostate cancer. Mol Cell Endocrinol 2012;360(1-2):3–13.

20.  Tomlins SA, Rhodes DR, Perner S, et al. Recurrent fusion of TMPRSS2 and ETS transcription factor genes in prostate cancer. Science 2005;310(5748): 644–8.

21.  Rubin MA, Maher CA, Chinnaiyan AM. Common gene  rearrangements  in  prostate  cancer.  J  Clin Oncol 2011;29(27):3659–68.

23.  Steinberg GD, Carter BS, Beaty TH, et al. Family history and the risk of prostate cancer. The Prostate 1990;17(4):337–47.

24.  Lichtenstein  P,  Holm  NV,  Verkasalo  PK,  et  al. Environmental and heritable factors  in the causa-tion of cancer— analyses of cohorts of twins from Sweden,  Denmark,  and  Finland.  N  Engl  J  Med 2000;343(2):78–85.

42.  Zheng  SL,  Sun  J,  Wiklund  F,  et  al.  Cumulative association  of  five  genetic  variants  with  prostate cancer. N Engl J Med 2008;358(9):910–9.

43.  Hsing  AW,  Tsao  L,  Devesa  SS.  International  trends  and  patterns  of  prostate  cancer  incidence and  mortality.  International  journal  of  cancer 2000;85(1):60–7.

66.  De  Marzo  AM,  Marchi  VL,  Epstein  JI,  Nelson WG.  Proliferative  inflammatory  atrophy  of  the prostate: implications for prostatic carcinogenesis. Am J Pathol 1999;155(6):1985–92.

86.  Barbieri CE, Baca SC, Lawrence MS, et al. Exome sequencing identifies recurrent SPOP, FOXA1 and MED12 mutations in prostate cancer. Nat Genet 2012;44(6):685–9.

87.  Lee WH, Morton RA, Epstein  JI,  et al. Cytidine methylation  of  regulatory  sequences  near  the pi-class glutathione S-transferase gene accompanies human prostatic carcinogenesis. Proceedings of the National Academy of Sciences of the United States of America 1994;91(24):11733–7.

148.  Luo J, Zha S, Gage WR, et al. Alpha-methylacyl-coa racemase: a new molecular marker for prostate cancer. Cancer Res 2002;62(8):2220–6.

165.  Catalona  WJ,  Smith  DS,  Ratliff TL,  et  al.  Mea-surement of prostate-specific antigen in serum as a screening  test  for  prostate  cancer.  N  Engl  J  Med 1991;324(17):1156–61.

181.  Benson MC, Whang IS, Pantuck A, et al. Prostate specific antigen density: a means of distinguishing benign prostatic hypertrophy and prostate cancer. J Urol 1992;147(3 Pt 2):815–6.

184.  Tosoian  JJ,  Trock  BJ,  Landis  P,  et  al.  Active  surveillance program for prostate cancer: an update of  the  Johns  Hopkins  experience.  J  Clin  Oncol 2011;29(16):2185–90.

185.  Carter HB, Pearson JD, Metter EJ, et al. Longitu-dinal evaluation of prostate-specific antigen  levels in men with and without prostate disease. JAMA 1992;267(16):2215–20.

188.  D’Amico AV, Chen MH, Roehl KA, Catalona WJ. Preoperative  PSA  velocity  and  the  risk  of  death 

from  prostate  cancer  after  radical  prostatectomy.  N Engl J Med 2004;351(2):125–35.

211.  Andriole GL, Crawford ED, Grubb 3rd RL, et al. Mortality  results  from  a  randomized  prostate-cancer  screening  trial.  N  Engl  J  Med  2009; 360(13):1310–9.

212.  Schroder  FH,  Hugosson  J,  Roobol  MJ,  et  al. Screening and prostate-cancer mortality  in a  ran-domized  European  study.  N  Engl  J  Med  2009; 360(13):1320–8.

213.  Schroder  FH,  Hugosson  J,  Roobol  MJ,  et  al. Prostate-cancer mortality at 11 years of follow-up. N Engl J Med 2012;366(11):981–90.

244.  De Marzo AM, Platz EA, Sutcliffe S, et al. Inflam-mation  in  prostate  carcinogenesis.  Nature  Rev Cancer 2007;7(4):256–69.

248.  Epstein  JI,  Allsbrook  Jr WC,  Amin  MB,  Egevad LL. The 2005 International Society of Urological Pathology  (ISUP)  Consensus  Conference  on Gleason  Grading  of  Prostatic  Carcinoma.  Am  J Surg Pathol 2005;29(9):1228–42.

250.  Fox JJ, Morris MJ, Larson SM, et al. Developing imaging  strategies  for  castration  resistant prostate cancer. Acta Oncol 2011;50(Suppl 1):39–48.

257.  Partin AW, Kattan MW, Subong EN, et al. Com-bination of prostate-specific antigen, clinical stage, and Gleason score to predict pathological stage of localized  prostate  cancer.  A  multi-institutional update. JAMA 1997;277(18):1445–51.

258.  Han  M,  Partin  AW,  Zahurak  M,  et  al.  Biochemical  (prostate  specific antigen)  recurrence probability  following  radical  prostatectomy  for clinically  localized  prostate  cancer.  J  Urol  2003; 169(2):517–23.

268.  Stamey  TA,  Yang  N,  Hay  AR,  et  al.  Prostate-specific antigen as a serum marker for adenocarci-noma of the prostate. N Engl J Med 1987;317(15): 909–16.

269.  D’Amico AV, Whittington R, Malkowicz SB, et al. Biochemical outcome after  radical prostatectomy, external  beam  radiation  therapy,  or  interstitial radiation  therapy  for  clinically  localized  prostate cancer. JAMA 1998;280(11):969–74.

280.  Bill-Axelson  A,  Holmberg  L,  Ruutu  M,  et  al. Radical  prostatectomy  versus watchful waiting  in early prostate cancer. N Engl J Med 2011;364(18): 1708–17.

304.  Sanda MG, Dunn RL, Michalski J, et al. Quality of  life  and  satisfaction  with  outcome  among prostate-cancer  survivors.  N  Engl  J  Med  2008; 358(12):1250–61.

325.  Zelefsky  MJ,  Fuks  Z,  Hunt  M,  et  al.  High-dose intensity modulated radiation therapy for prostate cancer: early toxicity and biochemical outcome in 772  patients.  Int  J  Radiat  Oncol  Biol  Phys 2002;53(5):1111–6.

332.  Roach 3rd M, Hanks G, Thames Jr H, et al. Defin-ing  biochemical  failure  following  radiotherapy with  or  without  hormonal  therapy  in  men  with clinically  localized  prostate  cancer:  recommenda-tions  of  the  RTOG-ASTRO  Phoenix  Consensus Conference.  Int  J  Radiat  Oncol  Biol  Phys  2006; 65(4):965–74.

338.  Wallner K, Merrick G, True L, et al. I-125 versus Pd-103  for  low-risk  prostate  cancer:  morbidity outcomes  from  a  prospective  randomized  multi-center trial. Cancer J 2002;8(1):67–73.

366.  Bolla  M,  Collette  L,  Blank  L,  et  al.  Long-term results with immediate androgen suppression and external  irradiation  in  patients  with  locally advanced  prostate  cancer  (an  EORTC  study):  a  phase  III  randomised  trial.  Lancet  2002; 360(9327):103–6.

369.  Jones  CU,  Hunt  D,  McGowan  DG,  et  al.  Radiotherapy  and  short-term  androgen  depriva-tion  for  localized  prostate  cancer.  N  Engl  J  Med 2011;365(2):107–18.

390.  Trock  BJ,  Han  M,  Freedland  SJ,  et  al.  Prostate cancer-specific  survival  following  salvage  radio-therapy  vs  observation  in  men  with  biochemical recurrence  after  radical  prostatectomy.  JAMA 2008;299(23):2760–9.

420.  Scher  HI,  Sawyers  CL.  Biology  of  progressive, castration-resistant prostate cancer: directed thera-pies targeting the androgen-receptor signaling axis. J Clin Oncol 2005;23(32):8253–61.

428.  Maximum  androgen  blockade  in  advanced  pros-tate  cancer:  an  overview  of  22  randomised  trials with  3283  deaths  in  5710  patients.  Prostate  Cancer  Trialists’  Collaborative  Group.  Lancet 1995;346(8970):265–9.

430.  Scher  HI,  Beer  TM,  Higano  CS,  et  al.  Antitumour  activity  of  MDV3100  in  castration-resistant prostate cancer: a phase 1-2 study. Lancet 2010;375(9724):1437–46.

434.  de  Bono  JS,  Logothetis  CJ,  Molina  A,  et  al.  Abiraterone  and  increased  survival  in  metastatic prostate  cancer.  N  Engl  J  Med  2011;364(21): 1995–2005.

449.  Tannock IF, de Wit R, Berry WR, et al. Docetaxel plus prednisone or mitoxantrone plus prednisone for  advanced  prostate  cancer.  N  Engl  J  Med 2004;351(15):1502–12.

450.  Armstrong  AJ,  Garrett-Mayer  ES,  Yang  YC,  et  al.  A  contemporary  prognostic  nomogram  for men  with  hormone-refractory  metastatic  prostate cancer: a TAX327 study analysis. Clin Cancer Res 2007;13(21):6396–403.

456.  de  Bono  JS,  Oudard  S,  Ozguroglu  M,  et  al.  Prednisone  plus  cabazitaxel  or  mitoxantrone  for metastatic  castration-resistant  prostate  cancer  progressing  after  docetaxel  treatment:  a  ran-domised open-label trial. Lancet 2010;376(9747): 1147–54.

461.  Kantoff  PW,  Higano  CS,  Shore  ND,  et  al. Sipuleucel-T  immunotherapy  for  castration-resistant  prostate  cancer.  N  Engl  J  Med  2010; 363(5):411–22.

462.  Saad  F,  Gleason  DM,  Murray  R,  et  al.  A  randomized, placebo-controlled trial of zoledronic acid  in  patients  with  hormone-refractory  meta-static  prostate  carcinoma.  J  Natl  Cancer  Inst 2002;94(19):1458–68.

464.  Fizazi K, Carducci M, Smith M, et al. Denosumab versus zoledronic acid for treatment of bone metas-tases  in  men  with  castration-resistant  prostate cancer:  a  randomised, double-blind  study. Lancet 2011;377(9768):813–22.

471.  Nilsson  S,  Franzen  L,  Parker  C,  et  al.  Bone-targeted  radium-223  in  symptomatic, hormone-refractory prostate cancer: a randomised, multicentre,  placebo-controlled  phase  II  study. Lancet Oncol 2007;8(7):587–94.

Page 35: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1496.e1ProstateCancer • CHAPTER84

R E F E R E N C E S1.  Siegel R, Naishadham D,  Jemal A. Cancer  statis-

tics, 2012. CA Cancer J Clin 2012;62(1):10–29.2.  Barry  MJ.  Screening  for  prostate  cancer—the  

controversy  that  refuses  to  die.  N  Engl  J  Med 2009;360(13):1351–4.

3.  Sakr WA, Grignon DJ, Crissman JD, et al. High grade prostatic  intraepithelial neoplasia (HGPIN) and prostatic adenocarcinoma between the ages of 20-69:  an  autopsy  study  of  249  cases.  In  Vivo 1994;8(3):439–43.

4.  Brooks JD. Anatomy of the lower urinary tract and male  genitalia.  In:  Walsh  PC,  editor.  Campbell’s Urology, vol. 1. Philadelphia, PA: Saunders; 2002. p. 41–80.

5.  Walsh  PC,  Donker  PJ.  Impotence  following radical  prostatectomy:  insight  into  etiology  and prevention. J Urol. Sep 1982;128(3):492–7.

6.  Walsh PC, Lepor H, Eggleston JC. Radical pros-tatectomy  with  preservation  of  sexual  function: anatomical  and  pathological  considerations.  The Prostate 1983;4(5):473–85.

7.  McNeal  JE.  The  zonal  anatomy  of  the  prostate. The Prostate 1981;2(1):35–49.

8.  McNeal JE. Normal histology of the prostate. Am J Surg Pathol 1988;12(8):619–33.

9.  Steers  WD.  5alpha-reductase  activity  in  the  prostate.  Urology  2001;58(6  Suppl  1):17–24;  discussion 24.

10.  Green SM, Mostaghel EA, Nelson PS. Androgen action and metabolism in prostate cancer. Mol Cell Endocrinol 2012;360(1-2):3–13.

11.  Roche  PJ,  Hoare  SA,  Parker  MG.  A  consensus DNA-binding site for the androgen receptor. Mol Endocrinol 1992;6(12):2229–35.

12.  Schuur  ER,  Henderson  GA,  Kmetec  LA,  et  al. Prostate-specific antigen expression is regulated by an upstream enhancer. J Biol Chem 1996;271(12): 7043–51.

13.  Goldstein AS, Stoyanova T, Witte ON. Primitive origins  of  prostate  cancer:  in  vivo  evidence  for prostate-regenerating  cells  and  prostate  cancer-initiating cells. Mol Oncol 2010;4(5):385–96.

14.  Peehl DM, Rubin JS. Keratinocyte growth factor: an  androgen-regulated  mediator  of  stromal-epithelial interactions in the prostate. World J Urol 1995;13(5):312–7.

15.  Planz  B,  Wang  Q,  Kirley  SD,  et  al.  Androgen responsiveness of stromal cells of the human pros-tate:  regulation of cell proliferation and keratino-cyte  growth  factor  by  androgen.  J  Urol 1998;160(5):1850–5.

16.  Marker PC, Donjacour AA, Dahiya R, Cunha GR. Hormonal,  cellular,  and  molecular  control  of  prostatic  development.  Dev  Biol  2003;253(2): 165–74.

17.  Oesterling  JE,  Hauzeur  CG,  Farrow  GM.  Small  cell  anaplastic  carcinoma  of  the  prostate:  a  clinical,  pathological  and  immunohistological study  of  27  patients.  J  Urol  1992;147(3  Pt  2): 804–7.

18.  De Marzo AM, Nelson WG, Meeker AK, Coffey DS.  Stem  cell  features  of  benign  and  malignant prostate epithelial cells.  J Urol 1998;160(6 Pt 2): 2381–92.

19.  van Leenders G, Dijkman H, Hulsbergen-van de Kaa C, et al. Demonstration of intermediate cells during human prostate epithelial differentiation in situ  and  in  vitro  using  triple-staining  confocal scanning  microscopy.  Lab  Invest  2000;80(8): 1251–8.

20.  Tomlins SA, Rhodes DR, Perner S, et al. Recurrent fusion of TMPRSS2 and ETS transcription factor genes in prostate cancer. Science 2005;310(5748): 644–8.

21.  Rubin MA, Maher CA, Chinnaiyan AM. Common gene  rearrangements  in  prostate  cancer.  J  Clin Oncol 2011;29(27):3659–68.

22.  Morganti  G,  Gianferrari  L,  Cresseri  A,  et  al. Recherches clinicostastisiques et genetiques sur les neoplasies  de  la  prostate.  Acta  Genet  1956;6: 304–5.

23.  Steinberg GD, Carter BS, Beaty TH, et al. Family history and the risk of prostate cancer. The Prostate 1990;17(4):337–47.

24.  Lichtenstein  P,  Holm  NV,  Verkasalo  PK,  et  al. Environmental and heritable factors  in the causa-tion of cancer—analyses of cohorts of twins from Sweden,  Denmark,  and  Finland.  N  Engl  J  Med 2000;343(2):78–85.

25.  Jin G, Lu L, Cooney KA, et al. Validation of pros-tate  cancer  risk-related  loci  identified  from genome-wide  association  studies  using  family-based association analysis: evidence from the Inter-national Consortium for Prostate Cancer Genetics (ICPCG). Hum Genet 2012;131(7):1095–103.

26.  Carpten J, Nupponen N, Isaacs S, et al. Germline mutations  in  the  ribonuclease  L  gene  in  families showing  linkage  with  HPC1.  Nat  Genet  2002; 30(2):181–4.

27.  Xu J, Zheng SL, Komiya A, et al. Germline muta-tions  and  sequence  variants  of  the  macrophage scavenger receptor 1 gene are associated with pros-tate cancer risk. Nat Genet 2002;32(2):321–5.

28.  Zhou  A,  Hassel  BA,  Silverman  RH.  Expression cloning  of  2-5A-dependent  RNAase:  a  uniquely regulated  mediator  of  interferon  action.  Cell 1993;72(5):753–65.

29.  Platt  N,  Gordon  S.  Is  the  class  A  macrophage  scavenger receptor (SR-A) multifunctional?—The mouse’s tale. J Clin Invest 2001;108(5):649–54.

30.  Casey G, Neville PJ, Plummer SJ, et al. RNASEL Arg462Gln variant is implicated in up to 13% of prostate cancer cases. Nat Genet 2002.

31.  Dejager S, Mietus-Snyder M, Friera A, Pitas RE. Dominant  negative  mutations  of  the  scavenger receptor. Native receptor inactivation by expression of  truncated  variants.  J  Clin  Invest  1993;92(2): 894–902.

32.  Zhou A, Paranjape J, Brown TL, et al. Interferon action and apoptosis are defective  in mice devoid of 2’,5’-oligoadenylate-dependent. RNase L. Embo J 1997;16(21):6355–63.

33.  Suzuki H, Kurihara Y, Takeya M, et al. A role for macrophage  scavenger  receptors  in atherosclerosis and  susceptibility  to  infection.  Nature  1997; 386(6622):292–6.

34.  Peiser L, Gough PJ, Kodama T, Gordon S. Macro-phage class A scavenger receptor-mediated phago-cytosis of Escherichia coli: role of cell heterogeneity, microbial  strain,  and  culture  conditions  in  vitro. Infect Immun 2000;68(4):1953–63.

35.  Thomas  CA,  Li  Y,  Kodama  T,  et  al.  Protection from lethal gram-positive infection by macrophage scavenger receptor-dependent phagocytosis. J Exp Med 2000;191(1):147–56.

36.  Sun  J,  Wiklund  F,  Zheng  SL,  et  al.  Sequence  variants  in Toll-like  receptor gene cluster  (TLR6-TLR1-TLR10)  and  prostate  cancer  risk.  J  Natl Cancer Inst 2005;97(7):525–32.

37.  Lindmark  F,  Zheng  SL,  Wiklund  F,  et  al.  H6D  polymorphism  in  macrophage-inhibitory cytokine-1  gene  associated  with  prostate  cancer.  J Natl Cancer Inst 2004;96(16):1248–54.

38.  Lindmark  F,  Zheng  SL,  Wiklund  F,  et  al. Interleukin-1 receptor antagonist haplotype associ-ated with prostate cancer risk. Br J Cancer 2005; 93(4):493–7.

39.  Xu  J, Lowey  J, Wiklund F,  et  al. The  interaction of four genes in the inflammation pathway signifi-cantly predicts prostate cancer risk. Cancer Epide-miol Biomarkers Prev 2005;14(11 Pt 1):2563–8.

40.  Zheng  SL,  Augustsson-Balter  K,  Chang  B,  et  al. Sequence variants of toll-like receptor 4 are associ-ated  with  prostate  cancer  risk:  results  from  the 

CAncer  Prostate  in  Sweden  Study.  Cancer  Res 2004;64(8):2918–22.

41.  Kote-Jarai Z, Leongamornlert D, Saunders E, et al. BRCA2 is a moderate penetrance gene contribut-ing  to  young-onset  prostate  cancer:  implications for genetic testing in prostate cancer patients. Br J Cancer 2011;105(8):1230–4.

42.  Zheng  SL,  Sun  J,  Wiklund  F,  et  al.  Cumulative association  of  five  genetic  variants  with  prostate cancer. N Engl J Med 2008;358(9):910–9.

43.  Hsing  AW,  Tsao  L,  Devesa  SS.  International  trends  and  patterns  of  prostate  cancer  incidence and  mortality.  International  journal  of  cancer 2000;85(1):60–7.

44.  Reddy S, Shapiro M, Morton Jr R, Brawley OW. Prostate  cancer  in  black  and  white  Americans. Cancer Metastasis Rev 2003;22(1):83–6.

45.  Whittemore  AS,  Kolonel  LN,  Wu  AH,  et  al.  Prostate cancer in relation to diet, physical activity, and body size in blacks, whites, and Asians in the United  States  and  Canada.  J  Natl  Cancer  Inst 1995;87(9):652–61.

46.  Haenszel  W,  Kurihara  M.  Studies  of  Japanese migrants. I. Mortality  from cancer and other dis-eases among Japanese in the United States. J Natl Cancer Inst 1968;40(1):43–68.

47.  Shimizu H, Ross RK, Bernstein L, et al. Cancers of  the  prostate  and  breast  among  Japanese  and white  immigrants  in  Los  Angeles  County.  Br  J Cancer 1991;63(6):963–6.

48.  Giovannucci  E,  Rimm  EB,  Colditz  GA,  et  al.  A prospective study of dietary fat and risk of prostate cancer  [see  comments].  J  Natl  Cancer  Inst 1993;85(19):1571–9.

49.  Gann  PH,  Hennekens  CH,  Sacks  FM,  et  al.  Prospective  study  of  plasma  fatty  acids  and  risk  of prostate cancer. J Natl Cancer Inst 1994;86(4): 281–6.

50.  Le Marchand L, Kolonel LN, Wilkens LR,  et  al. Animal  fat  consumption  and  prostate  cancer:  a prospective  study  in  Hawaii.  Epidemiology 1994;5(3):276–82.

51.  Knize  MG,  Salmon  CP,  Mehta  SS,  Felton  JS. Analysis  of  cooked  muscle  meats  for  heterocyclic aromatic  amine  carcinogens.  Mutat  Res  1997; 376(1-2):129–34.

52.  Lijinsky  W,  Shubik  P.  Benzo(a)pyrene  and  other polynuclear  hydrocarbons  in  charcoal-broiled meat. Science (New York, N.Y) 1964;145:53–5.

53.  Shirai T, Sano M, Tamano S, et al. The prostate:  a  target  for carcinogenicity of 2-amino-1-methyl- 6-  phenylimidazo[4,5-b]pyridine  (PhIP)  derived from  cooked  foods.  Cancer  Res  1997;57(2): 195–8.

54.  Chan  JM,  Stampfer  MJ,  Ma  J,  et  al.  Dairy  products,  calcium,  and  prostate  cancer  risk  in  the  Physicians’  Health  Study.  Am  J  Clin  Nutr 2001;74(4):549–54.

55.  Gann  PH,  Ma  J,  Giovannucci  E,  et  al.  Lower  prostate  cancer  risk  in men with  elevated plasma lycopene  levels:  results  of  a  prospective  analysis. Cancer Res 1999;59(6):1225–30.

56.  Cohen  JH,  Kristal  AR,  Stanford  JL.  Fruit  and  vegetable  intakes  and  prostate  cancer  risk.  J  Natl Cancer Inst 2000;92(1):61–8.

57.  Klein  EA, Thompson  Jr  IM, Tangen  CM,  et  al. Vitamin  E  and  the  risk  of  prostate  cancer:  the Selenium and Vitamin E Cancer Prevention Trial (SELECT). JAMA 2011;306(14):1549–56.

58.  Gardner WA, Bennett BD. The prostate overview: recent insights and speculations. In: Weinstein RS, Garnder WA, editors. Pathology and pathobiology of  the  urinary  bladder  and  prostate.  Baltimore: Williams and Wilkens; 1992. p. 129–48.

59.  Giovannucci  E.  Medical  history  and  etiology  of  prostate  cancer.  Epidemiol  Rev  2001;23(1): 159–62.

Page 36: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1496.e2

60.  Hoekx  L,  Jeuris  W,  Van  Marck  E,  Wyndaele  JJ. Elevated  serum  prostate  specific  antigen  (PSA) related  to  asymptomatic  prostatic  inflammation. Acta Urol Belg 1998;66(3):1–2.

61.  Roberts  RO,  Lieber  MM,  Rhodes  T,  et  al.  Prevalence  of  a  physician-assigned  diagnosis  of prostatitis: the Olmsted County Study of Urinary Symptoms  and  Health  Status  Among  Men. Urology 1998;51(4):578–84.

62.  Hayes RB, Pottern LM, Strickler H, et al. Sexual behaviour, STDs and risks for prostate cancer. Br J Cancer 2000;82(3):718–25.

63.  Dennis  LK,  Dawson  DV.  Meta-analysis  of  measures  of  sexual  activity  and  prostate  cancer. Epidemiology 2002;13(1):72–9.

64.  Xia  Y,  Zweier  JL.  Superoxide  and  peroxynitrite generation  from  inducible  nitric  oxide  synthase  in  macrophages.  Proceedings  of  the  National Academy  of  Sciences  of  the  United  States  of America 1997;94(13):6954–8.

65.  Eiserich JP, Hristova M, Cross CE, et al. Formation of  nitric  oxide-derived  inflammatory  oxidants  by myeloperoxidase  in  neutrophils.  Nature  1998; 391(6665):393–7.

66.  De  Marzo  AM,  Marchi  VL,  Epstein  JI,  Nelson WG.  Proliferative  inflammatory  atrophy  of  the prostate: implications for prostatic carcinogenesis. Am J Pathol 1999;155(6):1985–92.

67.  Franks  LM.  Atrophy  and  hyperplasia  in  the  prostate  proper.  J  Pathol  and  Bacteriol  1954;68: 617–21.

68.  De Marzo AM, et al. A working group classifica-tion of  focal  prostate  atrophy  lesions. Am  J Surg PatholAm J Surg Pathol 2006;In press.

69.  Nakayama  M,  Bennett  CJ,  Hicks  JL,  et  al.  Hypermethylation  of  the  human  glutathione S-transferase-pi  gene  (GSTP1)  CpG  island  is present  in  a  subset  of  proliferative  inflammatory atrophy lesions but not in normal or hyperplastic epithelium of  the prostate: a detailed study using laser-capture  microdissection.  Am  J  Pathol 2003;163(3):923–33.

70.  Shiraishi T, Watanabe M, Matsuura H, et al. The frequency  of  latent  prostatic  carcinoma  in  young males: the Japanese experience. In Vivo 1994;8(3): 445–7.

71.  Yatani R, Shiraishi T, Nakakuki K, et al. Trends in frequency  of  latent  prostate  carcinoma  in  Japan from 1965- 1979 to 1982-1986. J Natl Cancer Inst 1988;80(9):683–7.

72.  Yatani R, Chigusa I, Akazaki K, et al. Geographic pathology  of  latent  prostatic  carcinoma.  Interna-tional journal of cancer 1982;29(6):611–6.

73.  Berger MF, Lawrence MS, Demichelis F, et al. The genomic  complexity  of  primary  human  prostate cancer. Nature 2011;470(7333):214–20.

74.  Yegnasubramanian  S,  Wu  Z,  Haffner  MC,  et  al. Chromosome-wide mapping of DNA methylation patterns  in  normal  and  malignant  prostate  cells reveals  pervasive  methylation  of  gene-associated and conserved intergenic sequences. BMC genom-ics 2011;12:313.

75.  Isaacs  JT,  Coffey  DS.  Adaptation  versus  selection  as  the  mechanism  responsible  for  the relapse  of  prostatic  cancer  to  androgen  ablation therapy  as  studied  in  the  Dunning  R-3327-H adenocarcinoma.  Cancer  Res  1981;41(12  Pt 1):5070–5.

76.  Isaacs  JT,  Wake  N,  Coffey  DS,  Sandberg  AA. Genetic instability coupled to clonal selection as a mechanism for tumor progression in the Dunning R-3327  rat  prostatic  adenocarcinoma  system. Cancer Res 1982;42(6):2353–71.

77.  Perner  S,  Demichelis  F,  Beroukhim  R,  et  al. TMPRSS2:ERG  Fusion-Associated  Deletions Provide Insight into the Heterogeneity of Prostate Cancer. Cancer Res 2006;66(17):8337–41.

78.  Tomlins  SA,  Mehra  R,  Rhodes  DR,  et  al. TMPRSS2:ETV4  gene  fusions  define  a  third 

molecular  subtype of prostate cancer. Cancer Res 2006;66(7):3396–400.

79.  Haffner MC, Aryee MJ, Toubaji A, et al. Androgen-induced  TOP2B-mediated  double-strand  breaks and  prostate  cancer  gene  rearrangements.  Nat Genet 2010;42(8):668–75.

80.  Haffner  MC,  De  Marzo  AM,  Meeker  AK,  et  al. Transcription-induced DNA double strand breaks: both  oncogenic  force  and  potential  therapeutic target? Clin Cancer Res 2011;17(12):3858–64.

81.  Cerveira  N,  Ribeiro  FR,  Peixoto  A,  et  al. TMPRSS2-ERG gene fusion causing ERG overex-pression  precedes  chromosome  copy  number changes in prostate carcinomas and paired HGPIN lesions.  Neoplasia  (New  York,  N.Y.)  2006;8(10): 826–32.

82.  Hermans  KG,  van  Marion  R,  van  Dekken  H,  et  al. TMPRSS2:ERG  fusion  by  translocation  or interstitial deletion is highly relevant in androgen-dependent prostate cancer, but is bypassed in late-stage  androgen  receptor-negative  prostate  cancer. Cancer Res 2006;66(22):10658–63.

83.  Nam RK,  Sugar L, Wang Z,  et  al. Expression of TMPRSS2 ERG Gene Fusion in Prostate Cancer Cells is an Important Prognostic Factor for Cancer Progression. Cancer biology & therapy 2007;6(1).

84.  Petrovics  G,  Liu  A,  Shaheduzzaman  S,  et  al.  Frequent  overexpression  of  ETS-related  gene-1 (ERG1)  in  prostate  cancer  transcriptome.  Onco-gene 2005;24(23):3847–52.

85.  Tomlins SA, Rhodes DR, Yu J, et al. The role of SPINK1  in  ETS  rearrangement-negative  prostate cancers. Cancer cell 2008;13(6):519–28.

86.  Barbieri CE, Baca SC, Lawrence MS, et al. Exome sequencing identifies recurrent SPOP, FOXA1 and MED12 mutations in prostate cancer. Nat Genet 2012;44(6):685–9.

87.  Lee WH, Morton RA, Epstein  JI,  et al. Cytidine methylation  of  regulatory  sequences  near  the pi-class glutathione S-transferase gene accompanies human prostatic carcinogenesis. Proceedings of the National Academy of Sciences of the United States of America 1994;91(24):11733–7.

88.  Nakayama  M,  Gonzalgo  ML,  Yegnasubramanian S, et al. GSTP1 CpG island hypermethylation as a molecular biomarker  for prostate  cancer.  J Cell Biochem 2004;91(3):540–52.

89.  The effect of vitamin E and beta carotene on the incidence of lung cancer and other cancers in male smokers.  The  Alpha-Tocopherol,  Beta  Carotene Cancer  Prevention  Study  Group.  N  Engl  J  Med 1994;330(15):1029–35.

90.  Henderson CJ, Smith AG, Ure J, et al.  Increased skin  tumorigenesis  in mice  lacking pi  class gluta-thione S- transferases. Proceedings of the National Academy  of  Sciences  of  the  United  States  of America 1998;95(9):5275–80.

91.  Nelson  CP,  Kidd  LC,  Sauvageot  J,  et  al.  Protection  against  2-hydroxyamino-1-methyl-6-phenylimidazo[4,5-  b]pyridine  cytotoxicity  and DNA  adduct  formation  in  human  prostate  by  glutathione  S-transferase  P1.  Cancer  Res  2001; 61(1):103–9.

92.  Brooks JD, Weinstein M, Lin X, et al. CG island methylation changes near the GSTP1 gene in pros-tatic  intraepithelial  neoplasia.  Cancer  Epidemiol Biomarkers Prev 1998;7(6):531–6.

93.  Bieberich  CJ,  Fujita  K,  He  WW,  Jay  G.  Prostate-specific  and  androgen-dependent  expres-sion  of  a  novel  homeobox  gene.  J  Biol  Chem 1996;271(50):31779–82.

94.  Steadman  DJ,  Giuffrida  D,  Gelmann  EP.  DNA-binding  sequence  of  the  human  prostate-specific homeodomain  protein  NKX3.1.  Nucleic  acids research 2000;28(12):2389–95.

95.  Chen H, Nandi AK, Li X, Bieberich CJ. NKX-3.1 interacts with prostate-derived Ets factor and regu-lates the activity of the PSA promoter. Cancer Res 2002;62(2):338–40.

96.  Bhatia-Gaur  R,  Donjacour  AA,  Sciavolino  PJ,  et  al.  Roles  for  Nkx3.1  in  prostate  development and cancer. Genes Dev 1999;13(8):966–77.

97.  Abdulkadir  SA,  Magee  JA,  Peters  TJ,  et  al.  Conditional loss of Nkx3.1 in adult mice induces prostatic  intraepithelial  neoplasia.  Molecular  and cellular biology 2002;22(5):1495–503.

98.  Emmert-Buck MR, Vocke CD, Pozzatti RO, et al. Allelic loss on chromosome 8p12-21 in microdis-sected  prostatic  intraepithelial  neoplasia.  Cancer Res 1995;55(14):2959–62.

99.  Li  J,  Yen  C,  Liaw  D,  et  al.  PTEN,  a  putative protein  tyrosine  phosphatase  gene  mutated  in human brain, breast, and prostate cancer. Science (New York, N.Y.) 1997;275(5308):1943–7.

100.  Steck  PA,  Pershouse  MA,  Jasser  SA,  et  al.  Identification  of  a  candidate  tumour  suppressor gene,  MMAC1,  at  chromosome  10q23.3  that  is mutated in multiple advanced cancers. Nat Genet 1997;15(4):356–62.

101.  Teng  DH,  Hu  R,  Lin  H,  et  al.  MMAC1/PTEN mutations in primary tumor specimens and tumor cell lines. Cancer Res 1997;57(23):5221–5.

102.  Myers MP, Pass I, Batty IH, et al. The lipid phos-phatase  activity of PTEN  is  critical  for  its  tumor supressor  function.  Proceedings  of  the  National Academy  of  Sciences  of  the  United  States  of America 1998;95(23):13513–8.

103.  Myers  MP,  Stolarov  JP,  Eng  C,  et  al.  P-TEN,  the  tumor  suppressor  from  human  chromosome 10q23, is a dual- specificity phosphatase. Proceed-ings  of  the  National  Academy  of  Sciences  of  the United States of America 1997;94(17):9052–7.

104.  Maehama  T,  Dixon  JE.  The  tumor  suppressor, PTEN/MMAC1,  dephosphorylates  the  lipid second  messenger,  phosphatidylinositol  3,4,5- trisphosphate.  J  Biol  Chem  1998;273(22): 13375–8.

105.  Cairns  P,  Okami  K,  Halachmi  S,  et  al.  Frequent inactivation of PTEN/MMAC1  in primary pros-tate cancer. Cancer Res 1997;57(22):4997–5000.

106.  Suzuki H, Freije D, Nusskern DR, et al. Interfocal heterogeneity of PTEN/MMAC1 gene alterations in  multiple  metastatic  prostate  cancer  tissues. Cancer Res 1998;58(2):204–9.

107.  McMenamin ME, Soung P, Perera S, et al. Loss of PTEN  expression  in  paraffin-embedded  primary prostate cancer correlates with high Gleason score and  advanced  stage.  Cancer  Res  1999;59(17): 4291–6.

108.  Podsypanina  K,  Ellenson  LH,  Nemes  A,  et  al. Mutation of Pten/Mmac1 in mice causes neoplasia in  multiple  organ  systems.  Proceedings  of  the National Academy of Sciences of the United States of America 1999;96(4):1563–8.

109.  Di  Cristofano  A,  Pesce  B,  Cordon-Cardo  C,  Pandolfi  PP.  Pten  is  essential  for  embryonic  development and tumour suppression. Nat Genet 1998;19(4):348–55.

110.  Kim MJ, Cardiff RD, Desai N, et al. Cooperativity of  Nkx3.1  and  Pten  loss  of  function  in  a  mouse model  of  prostate  carcinogenesis.  Proceedings  of the  National  Academy  of  Sciences  of  the  United States of America 2002;99(5):2884–9.

111.  Cordon-Cardo C, Koff A, Drobnjak M, et al. Dis-tinct altered patterns of p27KIP1 gene expression in benign prostatic hyperplasia and prostatic carci-noma. J Natl Cancer Inst 1998;90(17):1284–91.

112.  Guo  Y,  Sklar  GN,  Borkowski  A,  Kyprianou  N. Loss  of  the  cyclin-dependent  kinase  inhibitor p27(Kip1) protein  in human prostate cancer cor-relates  with  tumor  grade.  Clin  Cancer  Res 1997;3(12 Pt 1):2269–74.

113.  Kibel  AS,  Faith  DA,  Bova  GS,  Isaacs  WB.  Loss  of  heterozygosity  at  12P12-13  in  primary  and metastatic  prostate  adenocarcinoma.  J  Urol 2000;164(1):192–6.

114.  Graff  JR,  Konicek  BW,  McNulty  AM,  et  al. Increased  AKT  activity  contributes  to  prostate 

Page 37: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1496.e3ProstateCancer • CHAPTER84

cancer  progression  by  dramatically  accelerating prostate  tumor growth and diminishing p27Kip1 expression. J Biol Chem 2000;275(32):24500–5.

115.  Gottschalk AR, Basila D, Wong M, et al. p27Kip1 is  required  for  PTEN-induced  G1  growth  arrest. Cancer Res 2001;61(5):2105–11.

116.  Nakamura  N,  Ramaswamy  S,  Vazquez  F,  et  al. Forkhead transcription factors are critical effectors of  cell  death  and  cell  cycle  arrest  downstream  of PTEN.  Molecular  and  cellular  biology  2000; 20(23):8969–82.

117.  Di  Cristofano  A,  De  Acetis  M,  Koff  A,  et  al.  Pten  and  p27KIP1  cooperate  in  prostate  cancer tumor  suppression  in  the  mouse.  Nat  Genet 2001;27(2):222–4.

118.  Eisenberger MA, Blumenstein BA, Crawford ED, et al. Bilateral orchiectomy with or without fluta-mide for metastatic prostate cancer. N Engl J Med 1998;339(15):1036–42.

119.  Crawford  ED,  Eisenberger  MA,  McLeod  DG,  et  al.  A  controlled  trial  of  leuprolide  with  and without flutamide in prostatic carcinoma. N Engl J Med 1989;321(7):419–24.

120.  Feldman  BJ,  Feldman  D.  The  development  of androgen-independent  prostate  cancer.  Nature reviews. Cancer 2001;1(1):34–45.

121.  van  der  Kwast  TH,  Schalken  J,  Ruizeveld  de Winter JA, et al. Androgen receptors in endocrine-therapy-resistant  human  prostate  cancer.  Interna-tional journal of cancer 1991;48(2):189–93.

122.  Koivisto  P,  Kononen  J,  Palmberg  C,  et  al.  Androgen  receptor  gene  amplification:  a  possible molecular  mechanism  for  androgen  deprivation therapy  failure  in  prostate  cancer.  Cancer  Res 1997;57(2):314–9.

123.  Tan  J,  Sharief  Y,  Hamil  KG,  et  al.  Dehydroepi-androsterone activates mutant androgen receptors expressed in the androgen-dependent human pros-tate  cancer  xenograft  CWR22  and  LNCaP  cells. Mol Endocrinol 1997;11(4):450–9.

124.  Veldscholte J, Voorhorst-Ogink MM, Bolt-de Vries J, et al. Unusual specificity of the androgen recep-tor in the human prostate tumor cell line LNCaP: high  affinity  for  progestagenic  and  estrogenic  steroids.  Biochim  Biophys  Acta  1990;1052(1): 187–94.

125.  Culig Z, Hobisch A, Cronauer MV, et al. Mutant androgen  receptor  detected  in  an  advanced-stage prostatic carcinoma is activated by adrenal andro-gens  and  progesterone.  Mol  Endocrinol  1993; 7(12):1541–50.

126.  Shi XB, Ma AH, Xia L, et al. Functional analysis of 44 mutant androgen receptors from human pros-tate cancer. Cancer Res 2002;62(5):1496–502.

127.  Sadar MD, Gleave ME. Ligand-independent acti-vation of the androgen receptor by the differentia-tion agent butyrate in human prostate cancer cells. Cancer Res 2000;60(20):5825–31.

128.  Craft  N,  Shostak  Y,  Carey  M,  Sawyers  CL.  A mechanism  for  hormone-independent  prostate cancer  through  modulation of  androgen  receptor signaling  by  the  HER-2/neu  tyrosine  kinase. Nature medicine 1999;5(3):280–5.

129.  Hobisch  A,  Eder  IE,  Putz T,  et  al.  Interleukin-6 regulates  prostate-specific  protein  expression  in prostate carcinoma cells by activation of the andro-gen receptor. Cancer Res 1998;58(20):4640–5.

130.  Nazareth LV, Weigel NL. Activation of the human androgen  receptor  through  a  protein  kinase  A  signaling  pathway.  J  Biol  Chem  1996;271(33): 19900–7.

131.  Dhanasekaran SM, Barrette TR, Ghosh D,  et  al. Delineation  of  prognostic  biomarkers  in  prostate cancer. Nature 2001;412(6849):822–6.

132.  Luo  JH, Yu YP, Cieply K,  et  al. Gene  expression analysis  of  prostate  cancers.  Mol  Carcinog  2002; 33(1):25–35.

133.  Stamey TA,  Warrington  JA,  Caldwell  MC,  et  al. Molecular  genetic  profiling of Gleason  grade 4/5 

prostate  cancers  compared  to  benign  prostatic hyperplasia. J Urol 2001;166(6):2171–7.

134.  Welsh JB, Sapinoso LM, Su AI, et al. Analysis of gene  expression  identifies  candidate  markers  and pharmacological targets in prostate cancer. Cancer Res 2001;61(16):5974–8.

135.  Magee JA, Araki T, Patil S, et al. Expression profil-ing  reveals  hepsin  overexpression  in  prostate cancer. Cancer Res 2001;61(15):5692–6.

136.  Luo J, Duggan DJ, Chen Y, et al. Human prostate cancer and benign prostatic hyperplasia: molecular dissection by gene expression profiling. Cancer Res 2001;61(12):4683–8.

137.  Waghray A, Schober M, Feroze F, et al. Identifica-tion  of  differentially  expressed  genes  by  serial analysis  of  gene  expression  in  human  prostate cancer. Cancer Res 2001;61(10):4283–6.

138.  Nelson PS, Han D, Rochon Y, et al. Comprehen-sive  analyses  of  prostate  gene  expression:  conver-gence  of  expressed  sequence  tag  databases, transcript profiling and proteomics. Electrophore-sis 2000;21(9):1823–31. [pii].

139.  Xu  J,  Stolk  JA,  Zhang  X,  et  al.  Identification  of differentially  expressed  genes  in  human  prostate cancer  using  subtraction  and  microarray.  Cancer Res 2000;60(6):1677–82.

140.  Walker  MG,  Volkmuth  W,  Sprinzak  E,  et  al.  Prediction  of  gene  function  by  genome-scale expression  analysis:  prostate  cancer-associated genes. Genome research 1999;9(12):1198–203.

141.  Huang  GM,  Ng  WL,  Farkas  J,  et  al.  Prostate cancer  expression  profiling  by  cDNA  sequencing analysis. Genomics 1999;59(2):178–86.

142.  Rhodes  DR,  Barrette  TR,  Rubin  MA,  et  al.  Meta-analysis of microarrays: interstudy validation of  gene  expression  profiles  reveals  pathway  dys-regulation  in  prostate  cancer.  Cancer  Res  2002; 62(15):4427–33.

143.  Tsuji A, Torres-Rosado A, Arai T, et al. Hepsin, a cell  membrane-associated  protease.  Characteriza-tion,  tissue  distribution,  and  gene  localization.  J Biol Chem 1991;266(25):16948–53.

144.  Klezovitch  O,  Chevillet  J,  Mirosevich  J,  et  al. Hepsin promotes prostate cancer progression and metastasis. Cancer cell 2004;6(2):185–95.

145.  Schmitz W,  Albers  C,  Fingerhut  R,  Conzelmann E.  Purification  and  characterization  of  an  alpha-methylacyl-CoA racemase from human liver. Eur J Biochem 1995;231(3):815–22.

146.  Jiang Z, Woda BA, Rock KL, et al. P504S: a new molecular marker for the detection of prostate car-cinoma.  Am  J  Surg  PatholAm  J  Surg  Pathol 2001;25(11):1397–404.

147.  Ferdinandusse  S,  Denis  S,  Clayton  PT,  et  al.  Mutations  in  the  gene  encoding  peroxisomal alpha-methylacyl-CoA racemase cause adult-onset sensory  motor  neuropathy.  Nat  Genet  2000; 24(2):188–91.

148.  Luo J, Zha S, Gage WR, et al. Alpha-methylacyl-coa racemase: a new molecular marker for prostate cancer. Cancer Res 2002;62(8):2220–6.

149.  Varambally S, Dhanasekaran SM, Zhou M, et al. The  polycomb  group  protein  EZH2  is  involved  in  progression  of  prostate  cancer.  Nature  2002; 419(6907):624–9.

150.  Greider  CW,  Blackburn  EH.  Identification  of  a  specific  telomere  terminal  transferase  activity  in  Tetrahymena  extracts.  Cell  1985;43(2  Pt  1): 405–13.

151.  Maser  RS,  DePinho  RA.  Connecting  chromo-somes, crisis, and cancer. Science 2002;297(5581): 565–9.

152.  O’Hagan RC, Chang S, Maser RS, et al. Telomere dysfunction  provokes  regional  amplification  and deletion  in  cancer  genomes.  Cancer  cell  2002; 2(2):149–55.

153.  Hackett  JA,  Feldser  DM,  Greider  CW. Telomere dysfunction  increases mutation  rate  and genomic instability. Cell 2001;106(3):275–86.

154.  Chin L, Artandi SE, Shen Q, et al. p53 deficiency rescues  the  adverse  effects  of  telomere  loss  and cooperates with telomere dysfunction to accelerate carcinogenesis. Cell 1999;97(4):527–38.

155.  Sommerfeld HJ, Meeker AK, Piatyszek MA, et al. Telomerase  activity:  a  prevalent marker  of malig-nant  human  prostate  tissue.  Cancer  Res  1996; 56(1):218–22.

156.  Meeker AK, Gage WR, Hicks JL, et al. Telomere length assessment in human archival tissues: com-bined  telomere  fluorescence  in  situ  hybridization and  immunostaining.  Am  J  Pathol  2002; 160(4):1259–68.

157.  Meeker  AK,  Hicks  JL,  Platz  EA,  et  al. Telomere shortening  is  an  early  somatic  DNA  alteration  in  human  prostate  tumorigenesis.  Cancer  Res 2002;62(22):6405–9.

158.  Smith DS, Catalona WJ. Interexaminer variability of digital  rectal examination  in detecting prostate cancer. Urology 1995;45(1):70–4.

159.  Ellis WJ,  Chetner  MP,  Preston  SD,  Brawer  MK. Diagnosis  of  prostatic  carcinoma:  the  yield  of serum prostate specific antigen, digital rectal exam-ination  and  transrectal  ultrasonography.  J  Urol 1994;152(5 Pt 1):1520–5.

160.  Cooner WH, Mosley BR, Rutherford Jr CL, et al. Prostate  cancer  detection  in  a  clinical  urological practice by ultrasonography, digital rectal examina-tion  and  prostate  specific  antigen.  J  Urol.  Jun 1990;143(6):1146–52; discussion 1152–44.

161.  Catalona WJ, Richie JP, Ahmann FR, et al. Com-parison  of  digital  rectal  examination  and  serum prostate  specific  antigen  in  the  early detection of prostate  cancer:  results  of  a  multicenter  clinical trial of 6,630 men. J Urol 1994;151(5):1283–90.

162.  Thompson  IM,  Rounder  JB,  Teague  JL,  et  al. Impact  of  routine  screening  for  adenocarcinoma  of  the  prostate  on  stage  distribution.  J  Urol 1987;137(3):424–6.

163.  Epstein  JI,  Walsh  PC,  Carmichael  M,  Brendler CB.  Pathologic  and  clinical  findings  to  predict tumor extent of nonpalpable  (stage T1c) prostate cancer. JAMA 1994;271(5):368–74.

164.  Schroder  FH,  van  der  Maas  P,  Beemsterboer  P,  et  al. Evaluation of  the digital  rectal  examination as a  screening  test  for prostate cancer. Rotterdam section  of  the  European  Randomized  Study  of Screening for Prostate Cancer. J Natl Cancer Inst 1998;90(23):1817–23.

165.  Catalona  WJ,  Smith  DS,  Ratliff TL,  et  al.  Mea-surement of prostate-specific antigen in serum as a screening  test  for  prostate  cancer.  N  Engl  J  Med 1991;324(17):1156–61.

166.  Gosselaar  C,  Roobol  MJ,  Roemeling  S,  Schroder  FH.  The  role  of  the  digital  rectal  examination  in  subsequent  screening visits  in  the European  randomized  study  of  screening  for  prostate  cancer  (ERSPC),  Rotterdam.  Eur  Urol 2008;54(3):581–8.

167.  Lilja  H.  Biology  of  prostate-specific  antigen. Urology 2003;62(5 Suppl 1):27–33.

168.  Oesterling JE, Jacobsen SJ, Chute CG, et al. Serum prostate-specific  antigen  in  a  community-based population of healthy men. Establishment of age-specific  reference  ranges.  JAMA  1993;270(7): 860–4.

169.  Klein LT, Lowe FC. The effects of prostatic manip-ulation  on  prostate-specific  antigen  levels.  Urol Clin North Am 1997;24(2):293–7.

170.  Guess  HA,  Heyse  JF,  Gormley  GJ.  The  effect  of finasteride on prostate-specific  antigen  in men with  benign  prostatic  hyperplasia.  The  Prostate 1993;22(1):31–7.

171.  Gann  PH,  Hennekens  CH,  Stampfer  MJ.  A  prospective  evaluation  of  plasma  prostate-specific antigen  for  detection  of  prostatic  cancer.  JAMA 1995;273(4):289–94.

172.  Fang  J,  Metter  EJ,  Landis  P,  et  al.  Low  levels  of prostate-specific antigen predict  long-term risk of 

Page 38: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1496.e4

prostate cancer: results from the Baltimore Longi-tudinal  Study  of  Aging.  Urology  2001;58(3): 411–6.

173.  Lilja H, Cronin AM, Dahlin A, et al. Prediction of significant prostate cancer diagnosed 20 to 30 years later  with  a  single  measure  of  prostate-specific antigen at or before age 50. Cancer 2011;117(6): 1210–9.

174.  Schroder FH. Stratifying risk—the U.S. Preventive Services Task Force and prostate-cancer screening. N Engl J Med 2011;365(21):1953–5.

175.  Zhu  X,  Albertsen  PC,  Andriole  GL,  et  al.  Risk-based  prostate  cancer  screening.  Eur  Urol 2012;61(4):652–61.

176.  Thompson IM, Ankerst DP, Chi C, et al. Operat-ing  characteristics  of  prostate-specific  antigen  in men  with  an  initial  PSA  level  of  3.0 ng/ml  or lower. JAMA 2005;294(1):66–70.

177.  Thompson IM, Ankerst DP, Chi C, et al. Assessing prostate  cancer  risk:  results  from  the  Prostate Cancer  Prevention  Trial.  J  Natl  Cancer  Inst 2006;98(8):529–34.

178.  van  Vugt  HA,  Roobol  MJ,  Kranse  R,  et  al.  Prediction of prostate  cancer  in unscreened men: external validation of a risk calculator. Eur J Cancer 2011;47(6):903–9.

179.  Lepor  H,  Wang  B,  Shapiro  E.  Relationship between  prostatic  epithelial  volume  and  serum prostate-specific  antigen  levels.  Urology  1994; 44(2):199–205.

180.  Benson MC, Whang IS, Olsson CA, et al. The use of prostate specific antigen density to enhance the predictive  value  of  intermediate  levels  of  serum prostate specific antigen. J Urol 1992;147(3 Pt 2): 817–21.

181.  Benson MC, Whang IS, Pantuck A, et al. Prostate specific antigen density: a means of distinguishing benign prostatic hypertrophy and prostate cancer. J Urol 1992;147(3 Pt 2):815–6.

182.  Djavan  B,  Zlotta  A,  Kratzik  C,  et  al.  PSA,  PSA density, PSA density of  transition zone,  free/total PSA ratio, and PSA velocity for early detection of prostate  cancer  in  men  with  serum  PSA  2.5  to 4.0 ng/mL. Urology 1999;54(3):517–22.

183.  Djavan B, Zlotta AR, Remzi M,  et  al. Total  and transition zone prostate volume and age: how do they  affect  the  utility  of  PSA-based  diagnostic parameters  for  early  prostate  cancer  detection? Urology 1999;54(5):846–52.

184.  Tosoian  JJ,  Trock  BJ,  Landis  P,  et  al.  Active  surveillance program for prostate cancer: an update of  the  Johns  Hopkins  experience.  J  Clin  Oncol 2011;29(16):2185–90.

185.  Carter HB, Pearson JD, Metter EJ, et al. Longitu-dinal evaluation of prostate-specific antigen  levels in men with and without prostate disease. JAMA 1992;267(16):2215–20.

186.  Berger  AP,  Deibl  M,  Strasak  A,  et  al.  Large-scale study of clinical impact of PSA velocity: long-term PSA kinetics as method of differentiating men with from  those  without  prostate  cancer.  Urology 2007;69(1):134–8.

187.  Etzioni  RD,  Howlader  N,  Shaw  PA,  et  al.  Long-term  effects  of  finasteride  on  prostate  specific  antigen  levels:  results  from  the  prostate cancer  prevention  trial.  J  Urol  2005;174(3): 877–81.

188.  D’Amico AV, Chen MH, Roehl KA, Catalona WJ. Preoperative  PSA  velocity  and  the  risk  of  death from  prostate  cancer  after  radical  prostatectomy.  N Engl J Med 2004;351(2):125–35.

189.  Carter  HB,  Ferrucci  L,  Kettermann  A,  et  al.  Detection of  life-threatening prostate cancer with prostate-specific antigen velocity during a window of  curability.  J  Natl  Cancer  Inst  2006;98(21): 1521–7.

190.  Carter  HB,  Kettermann  A,  Ferrucci  L,  et  al. Prostate-specific  antigen  velocity  risk  count  assessment:  a  new  concept  for  detection  of  

life-threatening prostate cancer during window of curability. Urology 2007;70(4):685–90.

191.  Loeb S, Metter EJ, Kan D, et al. Prostate-specific antigen  velocity  (PSAV)  risk  count  improves  the specificity  of  screening  for  clinically  significant prostate  cancer.  BJU  Int  2012;109(4):508–13;  discussion 513–504.

192.  Roddam AW, Duffy MJ, Hamdy FC, et al. Use of prostate-specific  antigen  (PSA)  isoforms  for  the detection  of  prostate  cancer  in  men  with  a  PSA level  of  2-10 ng/ml:  systematic  review  and meta-analysis. Eur Urol 2005;48(3):386–99; discussion 398–389.

193.  Catalona  WJ,  Southwick  PC,  Slawin  KM,  et  al. Comparison of percent free PSA, PSA density, and age-specific PSA cutoffs for prostate cancer detec-tion and staging. Urology 2000;56(2):255–60.

194.  Mikolajczyk  SD,  Marker  KM,  Millar  LS,  et  al.  A  truncated  precursor  form  of  prostate-specific antigen is a more specific serum marker of prostate cancer. Cancer Res 2001;61(18):6958–63.

195.  Mikolajczyk  SD,  Catalona  WJ,  Evans  CL,  et  al. Proenzyme  forms  of  prostate-specific  antigen  in serum  improve  the  detection  of  prostate  cancer. Clinical chemistry 2004;50(6):1017–25.

196.  Le BV, Griffin CR, Loeb S,  et  al.  [-2]Proenzyme prostate specific antigen is more accurate than total and free prostate specific antigen in differentiating prostate cancer  from benign disease  in a prospec-tive  prostate  cancer  screening  study.  J  Urol 2010;183(4):1355–9.

197.  Sokoll LJ, Sanda MG, Feng Z, et al. A prospective, multicenter,  National  Cancer  Institute  Early Detection Research Network study of [-2]proPSA: improving prostate cancer detection and correlat-ing with cancer aggressiveness. Cancer Epidemiol Biomarkers Prev 2010;19(5):1193–200.

198.  Bussemakers MJ, van Bokhoven A, Verhaegh GW, et  al.  DD3:  a  new  prostate-specific  gene,  highly overexpressed  in  prostate  cancer.  Cancer  Res 1999;59(23):5975–9.

199.  Hessels D, Schalken JA. The use of PCA3 in the diagnosis  of  prostate  cancer.  Nature  reviews. Urology 2009;6(5):255–61.

200.  Auprich M, Bjartell A, Chun FK, et al. Contem-porary  role  of  prostate  cancer  antigen  3  in  the management  of  prostate  cancer.  Eur  Urol  2011; 60(5):1045–54.

201.  Haese A, de la Taille A, van Poppel H, et al. Clini-cal  utility  of  the  PCA3  urine  assay  in  European men  scheduled  for  repeat  biopsy.  Eur  Urol 2008;54(5):1081–8.

202.  Baden  J,  Green  G,  Painter  J,  et  al.  Multicenter evaluation  of  an  investigational  prostate  cancer methylation assay. J Urol 2009;182(3):1186–93.

203.  Tomlins  SA,  Aubin  SM,  Siddiqui  J,  et  al.  Urine TMPRSS2:ERG  fusion  transcript  stratifies prostate  cancer  risk  in  men  with  elevated  serum PSA.  Science  translational  medicine  2011;3(94): 94ra72.

204.  Etzioni  R,  Kooperberg  C,  Pepe  M,  et  al.  Combining  biomarkers  to  detect  disease  with application  to  prostate  cancer.  Biostatistics  2003; 4(4):523–38.

205.  Boczko J, Messing E, Dogra V. Transrectal sonog-raphy  in  prostate  evaluation.  Radiol  Clin  North Am 2006;44(5):679–87, viii.

206.  Loeb S, Carter HB, Berndt SI, et al. Complications after  prostate  biopsy:  data  from  SEER-Medicare.  J Urol 2011;186(5):1830–4.

207.  Taylor  AK,  Zembower  TR,  Nadler  RB,  et  al.  Targeted  antimicrobial  prophylaxis  using  rectal swab cultures in men undergoing transrectal ultra-sound  guided  prostate  biopsy  is  associated  with reduced incidence of postoperative infectious com-plications  and  cost  of  care.  J  Urol  2012;187(4): 1275–9.

208.  Kumar V, Jagannathan NR, Thulkar S, Kumar R. Prebiopsy  magnetic  resonance  spectroscopy  and 

imaging  in  the diagnosis of prostate  cancer.  Int  J Urol 2012;19(7):602–13.

209.  Etzioni R, Tsodikov A, Mariotto A, et al. Quantify-ing  the  role of PSA  screening  in  the US prostate cancer  mortality  decline.  Cancer  Causes  Control 2008;19(2):175–81.

210.  Collin  SM,  Martin  RM,  Metcalfe  C,  et  al.  Prostate-cancer mortality  in  the USA and UK  in 1975-2004:  an  ecological  study.  Lancet  Oncol 2008;9(5):445–52.

211.  Andriole GL, Crawford ED, Grubb 3rd RL, et al. Mortality  results  from  a  randomized  prostate-cancer  screening  trial.  N  Engl  J  Med  2009; 360(13):1310–9.

212.  Schroder  FH,  Hugosson  J,  Roobol  MJ,  et  al. Screening and prostate-cancer mortality  in a  ran-domized  European  study.  N  Engl  J  Med  2009; 360(13):1320–8.

213.  Schroder  FH,  Hugosson  J,  Roobol  MJ,  et  al. Prostate-cancer mortality at 11 years of follow-up. N Engl J Med 2012;366(11):981–90.

214.  Andriole GL, Crawford ED, Grubb 3rd RL, et al. Prostate  cancer  screening  in  the  randomized  Prostate,  Lung,  Colorectal,  and  Ovarian  Cancer Screening Trial: mortality results after 13 years of follow-up.  J  Natl  Cancer  Inst  2012;104(2): 125–32.

215.  Miller AB. New data on prostate-cancer mortality after PSA screening. N Engl J Med 2012;366(11): 1047–8.

216.  Chou R, LeFevre ML. Prostate cancer screening—the evidence, the recommendations, and the clini-cal implications. JAMA 2011;306(24):2721–2.

217.  McNaughton-Collins MF, Barry MJ. One man at a  time—resolving  the PSA controversy. N Engl  J Med 2011;365(21):1951–3.

218.  Ganz PA, Barry JM, Burke W, et al. National Insti-tutes  of  Health  State-of-the-Science  Conference: role  of  active  surveillance  in  the  management  of men  with  localized  prostate  cancer.  Ann  Intern Med 2012;156(8):591–5.

219.  Bardia  A,  Platz  EA,  Yegnasubramanian  S,  et  al. Anti-inflammatory drugs,  antioxidants,  and pros-tate  cancer  prevention.  Current  opinion  in  phar-macology 2009;9(4):419–26.

220.  Prevention  of  cancer  in  the  next  millennium: Report  of  the Chemoprevention Working Group to the American Association for Cancer Research. Cancer Res 1999;59(19):4743–58.

221.  Thompson IM, Goodman PJ, Tangen CM, et al. The influence of finasteride on the development of prostate  cancer.  N  Engl  J  Med  2003;349(3): 215–24.

222.  Andriole  GL,  Bostwick  DG,  Brawley  OW,  et  al. Effect of dutasteride on the risk of prostate cancer. N Engl J Med 2010;362(13):1192–202.

223.  Theoret  MR,  Ning  YM,  Zhang  JJ,  et  al.  The  risks  and  benefits  of  5alpha-reductase  inhibitors  for  prostate-cancer  prevention.  N  Engl  J  Med 2011;365(2):97–9.

224.  Hardell  L,  Degerman  A,  Tomic  R,  et  al.  Levels  of selenium in plasma and glutathione peroxidase in erythrocytes in patients with prostate cancer or benign hyperplasia. Eur J Cancer Prev 1995;4(1): 91–5.

225.  Chan JM, Stampfer MJ, Ma J, et al. Supplemental vitamin  E  intake  and  prostate  cancer  risk  in  a  large cohort of men in the United States. Cancer Epidemiol Biomarkers Prev 1999;8(10):893–9.

226.  Brooks  JD,  Metter  EJ,  Chan  DW,  et  al.  Plasma selenium  level  before  diagnosis  and  the  risk  of  prostate cancer development. J Urol 2001;166(6): 2034–8.

227.  Kristal AR, Stanford JL, Cohen JH, et al. Vitamin and  mineral  supplement  use  is  associated  with reduced risk of prostate cancer. Cancer Epidemiol Biomarkers Prev 1999;8(10):887–92.

228.  Helzlsouer  KJ,  Huang  HY,  Alberg  AJ,  et  al.  Association  between  alpha-tocopherol, 

Page 39: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1496.e5ProstateCancer • CHAPTER84

gamma-tocopherol,  selenium,  and  subsequent prostate  cancer.  J  Natl  Cancer  Inst  2000; 92(24):2018–23.

229.  Nomura AM, Lee  J,  Stemmermann GN, Combs Jr GF. Serum selenium and subsequent risk of pros-tate  cancer.  Cancer  Epidemiol  Biomarkers  Prev 2000;9(9):883–7.

230.  Yoshizawa K, Willett WC, Morris SJ, et al. Study of prediagnostic selenium level in toenails and the risk of advanced prostate cancer. J Natl Cancer Inst 1998;90(16):1219–24.

231.  Klein  EA,  Thompson  IM,  Lippman  SM,  et  al. SELECT: the next prostate cancer prevention trial. Selenum and Vitamin E Cancer Prevention Trial. J Urol 2001;166(4):1311–5.

232.  Lippman SM, Klein EA, Goodman PJ, et al. Effect of  selenium  and  vitamin  E  on  risk  of  prostate cancer  and  other  cancers:  the  Selenium  and Vitamin  E  Cancer  Prevention  Trial  (SELECT). JAMA 2009;301(1):39–51.

233.  Duffield-Lillico  AJ,  Dalkin  BL,  Reid  ME,  et  al. Selenium  supplementation,  baseline  plasma  sele-nium  status  and  incidence  of  prostate  cancer:  an analysis  of  the  complete  treatment  period  of  the Nutritional  Prevention  of  Cancer  Trial.  BJU  Int 2003;91(7):608–12.

234.  DeMarzo AM, Nelson WG, Isaacs WB, Epstein JI. Pathological  and  molecular  aspects  of  prostate cancer. Lancet 2003;361(9361):955–64.

235.  Epstein JI. Diagnostic criteria of limited adenocar-cinoma  of  the  prostate  on  needle  biopsy.  Hum Pathol 1995;26(2):223–9.

236.  Baisden BL, Kahane H, Epstein JI. Perineural inva-sion,  mucinous  fibroplasia,  and  glomerulations: diagnostic  features  of  limited  cancer  on  prostate needle  biopsy.  Am  J  Surg  Pathol  1999;23(8): 918–24.

237.  Rubin  MA,  Zhou  M,  Dhanasekaran  SM,  et  al. alpha-Methylacyl coenzyme A racemase as a tissue biomarker  for  prostate  cancer.  JAMAJAMA 2002;287(13):1662–70.

238.  McNeal JE, Bostwick DG. Intraductal dysplasia: a premalignant  lesion  of  the  prostate.  Hum  Pathol 1986;17(1):64–71.

239.  Wills  ML,  Hamper  UM,  Partin  AW,  Epstein  JI. Incidence  of  high-grade  prostatic  intraepithelial neoplasia  in  sextant  needle  biopsy  specimens. Urology 1997;49(3):367–73.

240.  Haggman MJ, Macoska JA, Wojno KJ, Oesterling JE. The relationship between prostatic intraepithe-lial  neoplasia  and  prostate  cancer:  critical  issues.  J Urol 1997;158(1):12–22.

241.  O’Shaughnessy JA, Kelloff GJ, Gordon GB, et al. Treatment  and  prevention  of  intraepithelial  neo-plasia:  an  important  target  for  accelerated  new agent  development.  Clin  Cancer  Res  2002;8(2): 314–46.

242.  Lee  MC,  Moussa  AS,  Zaytoun  O,  et  al.  Using  a  saturation  biopsy  scheme  increases  cancer  detection during repeat biopsy in men with high-grade prostatic  intra-epithelial  neoplasia. Urology 2011;78(5):1115–9.

243.  Kawachi  MH,  Bahnson  RR,  Barry  M,  et  al. NCCN  clinical  practice  guidelines  in  oncology: prostate cancer early detection. J Natl Compr Canc Netw 2010;8(2):240–62.

244.  De Marzo AM, Platz EA, Sutcliffe S, et al. Inflam-mation  in  prostate  carcinogenesis.  Nature  Rev Cancer 2007;7(4):256–69.

245.  Shah  R,  Mucci  NR,  Amin  A,  et  al.  Postatrophic hyperplasia  of  the  prostate  gland:  neoplastic  pre-cursor  or  innocent  bystander?  Am  J  Pathol 2001;158(5):1767–73.

246.  Gleason DF, Mellinger GT. Prediction of prognosis for prostatic adenocarcinoma by combined histo-logical  grading  and  clinical  staging.  J Urol  1974; 111(1):58–64.

247.  Fine SW, Amin MB, Berney DM, et al. A contem-porary update on pathology reporting for prostate 

cancer:  biopsy  and  radical  prostatectomy  speci-mens. Eur Urol 2012;62(1):20–39.

248.  Epstein  JI,  Allsbrook  Jr WC,  Amin  MB,  Egevad LL. The 2005 International Society of Urological Pathology  (ISUP)  Consensus  Conference  on Gleason  Grading  of  Prostatic  Carcinoma.  Am  J Surg Pathol 2005;29(9):1228–42.

249.  Casimiro S, Guise TA, Chirgwin J. The critical role of the bone microenvironment in cancer metasta-ses. Mol Cell Endocrinol 2009;310(1-2):71–81.

250.  Fox JJ, Morris MJ, Larson SM, et al. Developing imaging  strategies  for  castration  resistant prostate cancer. Acta Oncol 2011;50 Suppl 1:39–48.

251.  Even-Sapir  E,  Metser  U,  Mishani  E,  et  al.  The detection of bone metastases in patients with high-risk  prostate  cancer:  99mTc-MDP  Planar  bone scintigraphy,  single-  and  multi-field-of-view SPECT, 18F-fluoride PET, and 18F-fluoride PET/CT. J Nucl Med 2006;47(2):287–97.

252.  Rubin  MA,  Putzi  M,  Mucci  N,  et  al.  Rapid (“warm”)  autopsy  study  for  procurement  of  metastatic prostate cancer. Clin Cancer Res 2000; 6(3):1038–45.

253.  Shah  RB,  Mehra  R,  Chinnaiyan  AM,  et  al. Androgen-independent prostate cancer is a hetero-geneous  group  of  diseases:  lessons  from  a  rapid autopsy  program.  Cancer  Res  2004;64(24): 9209–16.

254.  Liu W,  Laitinen  S,  Khan  S,  et  al.  Copy  number analysis  indicates  monoclonal  origin  of  lethal  metastatic  prostate  cancer.  Nat  Med  2009;15(5): 559–65.

255.  Yegnasubramanian  S,  Haffner  MC,  Zhang  Y,  et  al.  DNA  hypomethylation  arises  later  in  prostate  cancer  progression  than  CpG  island hypermethylation  and  contributes  to  metastatic tumor  heterogeneity.  Cancer  Res  2008;68(21): 8954–67.

256.  Partin  AW,  Yoo  J,  Carter  HB,  et  al.  The  use  of  prostate  specific  antigen,  clinical  stage  and Gleason score to predict pathological stage in men with  localized  prostate  cancer.  J  Urol  1993; 150(1):110–4.

257.  Partin AW, Kattan MW, Subong EN, et al. Com-bination of prostate-specific antigen, clinical stage, and Gleason score to predict pathological stage of localized  prostate  cancer.  A  multi-institutional update. JAMA 1997;277(18):1445–51.

258.  Han  M,  Partin  AW,  Zahurak  M,  et  al.  Biochemical  (prostate  specific antigen)  recurrence probability  following  radical  prostatectomy  for clinically  localized  prostate  cancer.  J  Urol  2003; 169(2):517–23.

259.  D’Amico AV, Whittington R, Malkowicz SB, et al. Predicting  prostate  specific  antigen  outcome  preoperatively in the prostate specific antigen era. J Urol 2001;166(6):2185–8.

260.  Lughezzani  G,  Briganti  A,  Karakiewicz  PI,  et  al. Predictive and prognostic models in radical prosta-tectomy candidates: a critical analysis of the litera-ture. Eur Urol 2010;58(5):687–700.

261.  Shariat SF, Kattan MW, Vickers AJ, et al. Critical review  of  prostate  cancer  predictive  tools.  Future Oncol 2009;5(10):1555–84.

262.  Purohit RS, Shinohara K, Meng MV, Carroll PR. Imaging  clinically  localized  prostate  cancer.  Urol Clin North Am 2003;30(2):279–93.

263.  Oesterling JE. Using PSA to eliminate the staging radionuclide  bone  scan.  Significant  economic implications.  Urol  Clin  North  Am  1993;20(4): 705–11.

264.  Hofer C, Kubler H, Hartung R,  et  al. Diagnosis and  monitoring  of  urological  tumors  using  posi-tron emission  tomography. Eur Urol 2001;40(5): 481–7.

265.  Harisinghani  MG,  Barentsz  J,  Hahn  PF,  et  al. Noninvasive detection of clinically occult  lymph-node metastases in prostate cancer. N Engl J Med 2003;348(25):2491–9.

266.  Pinto  F, Totaro  A,  Palermo  G,  et  al.  Imaging  in prostate  cancer  staging:  present  role  and  future perspectives. Urol Int 2012;88(2):125–36.

267.  Ellis  RJ,  Kaminsky  DA,  Zhou  EH,  et  al.  Ten-year  outcomes:  the  clinical  utility  of  single photon emission computed tomography/computed tomography capromab pendetide (Prostascint) in a cohort  diagnosed  with  localized  prostate  cancer. Int J Radiat Oncol Biol Phys 2011;81(1):29–34.

268.  Stamey  TA,  Yang  N,  Hay  AR,  et  al.  Prostate-specific  antigen  as  a  serum  marker  for adenocarcinoma  of  the  prostate.  N  Engl  J  Med 1987;317(15):909–16.

269.  D’Amico AV, Whittington R, Malkowicz SB, et al. Biochemical outcome after  radical prostatectomy, external  beam  radiation  therapy,  or  interstitial radiation  therapy  for  clinically  localized  prostate cancer. JAMA 1998;280(11):969–74.

270.  Cooperberg MR, Carroll PR, Klotz L. Active sur-veillance for prostate cancer: progress and promise. J Clin Oncol 2011;29(27):3669–76.

271.  Mohler  JL.  The  2010  NCCN  clinical  practice guidelines  in oncology on prostate  cancer.  J Natl Compr Canc Netw 2010;8(2):145.

272.  Moreno JG, Croce CM, Fischer R, et al. Detection of hematogenous micrometastasis in patients with prostate cancer. Cancer Res 1992;52(21):6110–2.

273.  Ts’o PO, Pannek J, Wang ZP, et al. Detection of intact  prostate  cancer  cells  in  the  blood  of  men with prostate cancer. Urology 1997;49(6):881–5.

274.  Bastian PJ, Palapattu GS, Lin X, et al. Preoperative serum  DNA  GSTP1  CpG  island  hypermethyl-ation and the risk of early prostate-specific antigen recurrence  following  radical  prostatectomy.  Clin Cancer Res 2005;11(11):4037–43.

275.  Allard  WJ,  Matera  J,  Miller  MC,  et  al.  Tumor  cells circulate in the peripheral blood of all major carcinomas but not in healthy subjects or patients with  nonmalignant  diseases.  Clin  Cancer  Res 2004;10(20):6897–904.

276.  Danila  DC,  Fleisher  M,  Scher  HI.  Circulating tumor  cells  as  biomarkers  in  prostate  cancer.  Clin Cancer Res 2011;17(12):3903–12.

277.  Economos C, Morrissey C, Vessella RL. Circulat-ing  tumor  cells  as  a marker of  response:  implica-tions  for  determining  treatment  efficacy  and evaluating  new  agents.  Current  Opin  Urol.  May 2012;22(3):190–6.

278.  Carter HB. Management of  low (favourable)-risk prostate cancer. BJU Int 2011.

279.  Parker C. Active surveillance: towards a new para-digm in the management of early prostate cancer. Lancet Oncol 2004;5(2):101–6.

280.  Bill-Axelson  A,  Holmberg  L,  Ruutu  M,  et  al. Radical  prostatectomy  versus  watchful  waiting  in  early  prostate  cancer.  N  Engl  J  Med  2011; 364(18):1708–17.

281.  Wilt TJ, Brawer MK, Barry MJ, et al. The Prostate cancer  Intervention Versus Observation Trial:VA/NCI/AHRQ  Cooperative  Studies  Program  #407 (PIVOT): design and baseline results of a random-ized  controlled  trial  comparing  radical  prostatec-tomy  to watchful waiting  for men with clinically localized  prostate  cancer.  Contemp  Clin  Trials 2009;30(1):81–7.

282.  Carter  HB,  Walsh  PC,  Landis  P,  Epstein  JI. Expectant  management  of  nonpalpable  prostate cancer  with  curative  intent:  preliminary  results.  J Urol 2002;167(3):1231–4.

283.  Choo  R,  Klotz  L,  Danjoux  C,  et  al.  Feasibility study: watchful waiting for localized low to inter-mediate  grade  prostate  carcinoma  with  selective delayed  intervention  based  on  prostate  specific antigen,  histological  and/or  clinical  progression.  J Urol 2002;167(4):1664–9.

284.  Dahabreh  IJ,  Chung  M,  Balk  EM,  et  al.  Active surveillance in men with localized prostate cancer:  a  systematic  review.  Ann  Intern  Med 2012;156(8):582–90.

Page 40: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1496.e6

285.  Stattin  P,  Holmberg  E,  Johansson  JE,  et  al.  Outcomes  in  localized  prostate  cancer:  National Prostate  Cancer  Register  of  Sweden  follow-up study. J Natl Cancer Inst 2010;102(13):950–8.

286.  Parker C, Muston D, Melia J, et al. A model of the natural history of screen-detected prostate cancer, and the effect of radical treatment on overall sur-vival. Br J Cancer 2006;94(10):1361–8.

287.  Eggener SE, Scardino PT, Walsh PC, et al. Predict-ing  15-year  prostate  cancer  specific  mortality  after  radical  prostatectomy.  J  Urol  2011;185(3): 869–75.

288.  Shappley  3rd  WV,  Kenfield  SA,  Kasperzyk  JL,  et  al.  Prospective  study  of  determinants  and  outcomes  of  deferred  treatment  or  watchful waiting  among  men  with  prostate  cancer  in  a nationwide  cohort.  J  Clin  Oncol  2009; 27(30):4980–5.

289.  Cooperberg  MR,  Broering  JM,  Carroll  PR.  Time  trends  and  local  variation  in primary  treat-ment  of  localized  prostate  cancer.  J  Clin  Oncol 2010;28(7):1117–23.

290.  Cooperberg  MR,  Broering  JM,  Kantoff  PW, Carroll PR. Contemporary trends in low risk pros-tate cancer: risk assessment and treatment. J Urol 2007;178(3 Pt 2):S14–19.

291.  Hamilton  AS,  Albertsen  PC,  Johnson TK,  et  al. Trends in the treatment of localized prostate cancer using supplemented cancer registry data. BJU Int 2011;107(4):576–84.

292.  Klotz L, Zhang L, Lam A, et al. Clinical results of long-term follow-up of a large, active surveillance cohort with localized prostate cancer. J Clin Oncol 2010;28(1):126–31.

293.  Lane  JA,  Hamdy  FC,  Martin  RM,  et  al.  Latest results  from  the  UK  trials  evaluating  prostate cancer  screening  and  treatment:  the  CAP  and ProtecT  studies.  Eur  J  Cancer  2010;46(17): 3095–101.

294.  Bechis  SK,  Carroll  PR,  Cooperberg  MR.  Impact of  age  at  diagnosis  on  prostate  cancer  treatment and survival. J Clin Oncol 2011;29(2):235–41.

295.  Mettlin CJ, Murphy GP, Ho R, Menck HR. The National Cancer Data Base report on longitudinal observations  on  prostate  cancer.  Cancer  1996; 77(10):2162–6.

296.  Barbash GI, Glied SA. New technology and health care  costs—the  case  of  robot-assisted  surgery.  N Engl J Med 2010;363(8):701–4.

297.  Boorjian  SA,  Eastham  JA,  Graefen  M,  et  al.  A critical analysis of the long-term impact of radical prostatectomy on cancer control and function out-comes. Eur Urol 2012;61(4):664–75.

298.  Barry MJ, Gallagher PM, Skinner JS, Fowler Jr FJ. Adverse  effects  of  robotic-assisted  laparoscopic versus  open  retropubic  radical  prostatectomy among a nationwide random sample of medicare-age men. J Clin Oncol 2012;30(5):513–8.

299.  Duffey B, Varda B, Konety B. Quality of evidence to  compare  outcomes  of  open  and  robot-assisted laparoscopic  prostatectomy.  Current  urology reports 2011;12(3):229–36.

300.  Vickers  AJ.  Great  meaningless  questions  in urology:  which  is  better,  open,  laparoscopic,  or robotic  radical  prostatectomy?  Urology  2011; 77(5):1025–6.

301.  Begg CB, Riedel ER, Bach PB, et al. Variations in morbidity  after  radical  prostatectomy.  N  Engl  J Med 2002;346(15):1138–44.

302.  Walsh  PC,  Marschke  P,  Ricker  D,  Burnett  AL. Patient-reported  urinary  continence  and  sexual function  after  anatomic  radical  prostatectomy. Urology 2000;55(1):58–61.

303.  Walsh  PC.  Radical  prostatectomy  for  localized prostate  cancer  provides  durable  cancer  control with excellent quality of  life: a  structured debate.  J Urol 2000;163(6):1802–7.

304.  Sanda MG, Dunn RL, Michalski J, et al. Quality of  life  and  satisfaction  with  outcome  among 

prostate-cancer  survivors.  N  Engl  J  Med  2008; 358(12):1250–61.

305.  Walsh PC. Anatomic radical retropubic prostatec-tomy.  In:  Walsh  PC,  editor.  Campbell’s  Urology, vol.  4.  Philadelphia,  PA:  Saunders;  2002.  p. 3107–30.

306.  Geary ES, Dendinger TE, Freiha FS, Stamey TA. Incontinence and vesical neck strictures following radical  retropubic  prostatectomy.  Urology  1995; 45(6):1000–6.

307.  Steiner  MS.  Continence-preserving  anatomic radical  retropubic  prostatectomy.  Urology 2000;55(3):427–35.

308.  Walsh  PC,  Marschke  PL.  Intussusception  of  the  reconstructed  bladder  neck  leads  to  earlier  continence  after  radical  prostatectomy.  Urology 2002;59(6):934–8.

309.  Dubbelman  YD,  Dohle  GR,  Schroder  FH.  Sexual function before and after radical retropubic prostatectomy:  A  systematic  review  of  prognostic indicators  for  a  successful  outcome.  Eur  Urol 2006;50(4):711–8; discussion 718–20.

310.  Huang GJ, Sadetsky N, Penson DF. Health related quality of life for men treated for localized prostate cancer  with  long-term  followup.  J  Urol  2010; 183(6):2206–12.

311.  Mulhall  JP,  Bella  AJ,  Briganti  A,  et  al.  Erectile function rehabilitation in the radical prostatectomy patient. J Sex Med 2010;7(4 Pt 2):1687–98.

312.  Han M, Partin AW, Pound CR, et al. Long-term biochemical  disease-free  and  cancer-specific  sur-vival following anatomic radical retropubic prosta-tectomy. The  15-year  Johns  Hopkins  experience. Urol Clin North Am 2001;28(3):555–65.

313.  Antonarakis  ES,  Feng  Z,  Trock  BJ,  et  al.  The natural  history  of  metastatic  progression  in  men with  prostate-specific  antigen  recurrence  after radical  prostatectomy:  long-term  follow-up.  BJU Int 2012;109(1):32–9.

314.  Roentgen  WC.  Proceedings  of  the  Wurzburg Phisico-Medical Society. December 28 1895.

315.  Curie MS. Recherches  sur  les  substances  radioac-tives. In: Faculte des Sciences de Paris pour obtenir le grade de docteur es  science physiques. 2nd ed. Paris: Gauthier-Villas; 1904.

316.  Young HH, Frontz WA. Some new methods in the treatment of carcinoma of the lower genito-urinary tract with radium. J Urol 1917;1:505–41.

317.  Barringer  BS.  Phases  of  the  pathology,  diagnosis and  treatment  of  cancer  of  the  prostate.  J  Urol 1928:407–11.

318.  George  FW,  Carlton  CE,  Dykhuizen  RF,  Dillon JR.  Cobalt-60  telecurietherapy  in  the  definitive treatment of carcinoma of the prostate: A prelimi-nary report. J Urol 1965;93:102–9.

319.  Kaplan HS, Bagshaw MA. The Stanford medical linear accelerator III. Application to clinical prob-lems  of  Radiation  Therapy.  Stanford  Med  Bull 1957;15:141–51.

320.  Bagshaw MA, Kaplan HS, Sagerman RH. Linear Accelerator Supervoltage Radiotherapy VII. Carci-noma of the Prostate. Radiology 1965;85:121–9.

321.  Del  Regato  JA.  Radiotherapy  in  the  conservative treatment of operable and locally inoperable carci-noma of the prostate. Radiology 1967;88:761–6.

322.  Fowler JE Jr, Braswell NT, Pandey P, et al. Experi-ence  with  radical  prostatectomy  and  radiation therapy for localized prostate cancer at a Veterans Affairs  Medical  Center.  J  Urol  1995;153:1026– 31.

323.  Zelefsky MJ, Cowen D, Fuks Z, et al. Long term tolerance  of  high  dose  three-dimensional  confor-mal radiotherapy in patients with localized prostate carcinoma. Cancer 1999;85(11):2460–8.

324.  Hanks  GE,  Hanlon  AL,  Schultheiss  TE,  et  al.  Dose  escalation  with  3D  conformal  treatment:  five  year  outcomes,  treatment  optimization,  and future  directions.  Int  J  Radiat  Oncol  Biol  Phys 1998;41(3):501–10.

325.  Zelefsky  MJ,  Fuks  Z,  Hunt  M,  et  al.  High-dose intensity modulated radiation therapy for prostate cancer: early toxicity and biochemical outcome in 772  patients.  Int  J  Radiat  Oncol  Biol  Phys 2002;53(5):1111–6.

326.  Zelefsky  MJ,  Eid  JF.  Elucidating  the  etiology  of erectile  dysfunction  after  definitive  therapy  for prostatic  cancer.  Int  J  Radiat  Oncol  Biol  Phys 1998;40(1):129–33.

327.  Fisch BM, Pickett B, Weinberg V, Roach M. Dose of radiation received by the bulb of the penis cor-relates  with  risk  of  impotence  after  three-dimensional  conformal  radiotherapy  for  prostate cancer. Urology 2001;57(5):955–9.

328.  Bagshaw MA, Cox RS, Ray GR. Status of radiation treatment of prostate cancer at Stanford University. NCI Monogr 1988;7:47–60.

329.  Zelefsky MJ, McKee AB, Lee H, Leibel SA. Effi-cacy of oral sildenafil in patients with erectile dys-function  after  radiotherapy  for  carcinoma  of  the prostate. Urology 1999;53(4):775–8.

330.  Zagars GK. The prognostic significance of a single serum  prostate-specific  antigen  value  beyond  six months after radiation therapy for adenocarcinoma of  the  prostate.  Int  J  Radiat  Oncol  Biol  Phys 1993;27(1):39–45.

331.  Sartor CI, Strawderman MH, Lin XH, et al. Rate of PSA rise predicts metastatic versus  local  recur-rence  after  definitive  radiotherapy.  Int  J  Radiat Oncol Biol Phys 1997;38(5):941–7.

332.  Roach 3rd M, Hanks G, Thames Jr H, et al. Defin-ing  biochemical  failure  following  radiotherapy with  or  without  hormonal  therapy  in  men  with clinically  localized  prostate  cancer:  recommenda-tions  of  the  RTOG-ASTRO  Phoenix  Consensus Conference.  Int  J  Radiat  Oncol  Biol  Phys 2006;65(4):965–74.

333.  Consensus statement: guidelines for PSA following radiation therapy. American Society for Therapeu-tic Radiology and Oncology Consensus Panel. Int J Radiat Oncol Biol Phys 1997;37(5):1035–41.

334.  Pollack A, Zagars GK, Starkschall G, et al. Prostate cancer  radiation  dose  response:  results  of  the  M. D.  Anderson  phase  III  randomized  trial.  Int  J Radiat Oncol Biol Phys 2002;53(5):1097–105.

335.  Zelefsky  MJ,  Leibel  SA,  Gaudin  PB,  et  al.  Dose escalation with three-dimensional conformal radia-tion therapy affects the outcome in prostate cancer. Int  J  Radiat  Oncol  Biol  Phys  1998;41(3): 491–500.

336.  Zelefsky  MJ,  Fuks  Z,  Hunt  M,  et  al.  High  dose  radiation  delivered  by  intensity  modulated conformal  radiotherapy  improves  the outcome of localized  prostate  cancer.  J  Urol  2001;166(3): 876–81.

337.  Pollack A, Smith LG, von Eschenbach AC. Exter-nal beam radiotherapy dose response characteristics of  1127  men  with  prostate  cancer  treated  in  the  PSA  era.  Int  J  Radiat  Oncol  Biol  Phys 2000;48(2):507–12.

338.  Wallner K, Merrick G, True L, et al. I-125 versus Pd-103  for  low-risk  prostate  cancer:  morbidity outcomes  from  a  prospective  randomized  multi-center trial. Cancer J 2002;8(1):67–73.

339.  Martinez AA, Pataki I, Edmundson G, et al. Phase II prospective study of the use of conformal high-dose-rate  brachytherapy  as  monotherapy  for  the treatment of favorable stage prostate cancer: a fea-sibility  report.  Int  J  Radiat  Oncol  Biol  Phys 2001;49(1):61–9.

340.  Yoshioka Y, Nose T, Yoshida K, et al. High-dose-rate interstitial brachytherapy as a monotherapy for localized  prostate  cancer:  treatment  description and preliminary results of a phase I/II clinical trial. Int  J  Radiat  Oncol  Biol  Phys  2000;48(3): 675–81.

341.  Deger  S,  Boehmer  D,  Turk  I,  et  al.  High  dose  rate brachytherapy of localized prostate cancer. Eur Urol 2002;41(4):420–6.

Page 41: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1496.e7ProstateCancer • CHAPTER84

342.  Zelefsky  MJ,  Kuban  DA,  Levy  LB,  et  al.  Multi-institutional  analysis  of  long-term  outcome  for stages T1-T2 prostate cancer  treated with perma-nent  seed  implantation.  Int  J  Radiat  Oncol  Biol Phys 2007;67(2):327–33.

343.  Davis  BJ,  Pisansky TM,  Wilson TM,  et  al.  The radial distance of extraprostatic extension of pros-tate  carcinoma:  implications  for  prostate  brachy-therapy. Cancer 1999;85(12):2630–7.

344.  Sylvester JE, Grimm PD, Blasko JC, et al. 15-Year biochemical  relapse  free  survival  in  clinical  Stage T1-T3 prostate cancer following combined exter-nal beam radiotherapy and brachytherapy; Seattle experience.  Int  J  Radiat  Oncol  Biol  Phys 2007;67(1):57–64.

345.  Stone  NN,  Stock  RG.  Prostate  brachytherapy  in patients with prostate  volumes >/= 50 cm3: dosi-metric  analysis  of  implant  quality.  Int  J  Radiat Oncol Biol Phys 2000;46(5):1199–204.

346.  Merrick  GS,  Butler  WM,  Dorsey  AT,  Lief  JH. Effect of prostate size and isotope selection on dosi-metric quality following permanent seed implanta-tion. Tech Urol 2001;7(3):233–40.

347.  Terk  MD,  Stock  RG,  Stone  NN.  Identification  of patients at increased risk for prolonged urinary retention  following  radioactive  seed  implantation of the prostate. J Urol 1998;160(4):1379–82.

348.  Merrick  GS,  Butler  WM,  Lief  JH,  Dorsey  AT. Temporal  resolution of urinary morbidity  follow-ing prostate brachytherapy. Int J Radiat Oncol Biol Phys 2000;47(1):121–8.

349.  Landis D, Wallner K, Locke J, et al. Late Urinary Function  After  Prostate  Brachytherapy.  Brachy-therapy 2002;1:21.

350.  Wallner K, Lee H, Wasserman S, Dattoli M. Low risk  of  urinary  incontinence  following  prostate brachytherapy in patients with a prior transurethral prostate  resection.  Int  J  Radiat  Oncol  Biol  Phys 1997;37(3):565–9.

351.  Merrick  GS,  Butler  WM,  Wallner  KE,  et  al. Prostate-specific  antigen  spikes  after  permanent prostate  brachytherapy.  Int  J  Radiat  Oncol  Biol Phys 2002;54(2):450–6.

352.  Kleinberg  L,  Wallner  K,  Roy  J,  et  al.  Treatment-related  symptoms during  the first  year following transperineal  125I prostate  implantation. Int  J  Radiat  Oncol  Biol  Phys  1994;28(4): 985–90.

353.  Ragde H, Blasko JC, Grimm PD, et al. Interstitial iodine-125 radiation without adjuvant therapy  in the treatment of clinically localized prostate carci-noma. Cancer 1997;80(3):442–53.

354.  Talcott  JA,  Clark  JA,  Stark  PC,  Mitchell  SP.  Long-term  treatment  related  complications  of brachytherapy for early prostate cancer: a survey of patients  previously  treated.  J  Urol  2001;166(2): 494–9.

355.  Benoit  RM,  Naslund  MJ,  Cohen  JK.  Complica-tions after prostate brachytherapy in the Medicare population. Urology 2000;55(1):91–6.

356.  Zelefsky MJ, Hollister T, Raben A, et al. Five-year biochemical outcome and toxicity with transperi-neal  CT-  planned  permanent  I-125  prostate implantation  for  patients  with  localized  prostate cancer.  Int  J  Radiat  Oncol  Biol  Phys  2000; 47(5):1261–6.

357.  Lawton CA, DeSilvio M, Lee WR, et al. Results of a  phase  II  trial  of  transrectal  ultrasound-guided permanent radioactive implantation of the prostate for definitive management of localized adenocarci-noma of the prostate (radiation therapy oncology group  98-05).  Int  J  Radiat  Oncol  Biol  Phys 2007;67(1):39–47.

358.  Stock RG, Stone NN, Iannuzzi C. Sexual potency following  interactive  ultrasound-guided  brachy-therapy for prostate cancer. Int J Radiat Oncol Biol Phys 1996;35(2):267–72.

359.  Wallner K, Roy J, Harrison L. Tumor control and morbidity  following  transperineal  iodine  125 

implantation for stage T1/T2 prostatic carcinoma. J Clin Oncol 1996;14(2):449–53.

360.  Stock RG, Kao J, Stone NN. Penile erectile func-tion after permanent radioactive seed implantation for  treatment  of  prostate  cancer.  J  Urol  2001; 165(2):436–9.

361.  Merrick GS, Butler WM, Wallner KE, et al. The importance  of  radiation  doses  to  the  penile  bulb vs. crura in the development of postbrachytherapy erectile dysfunction. Int J Radiat Oncol Biol Phys 2002;54(4):1055–62.

362.  Merrick  GS,  Wallner  KE,  Butler  WM.  Manage-ment of  sexual dysfunction after prostate brachy-therapy. Oncology (Huntingt) 2003;17(1):52–62; discussion 62, 67–70, 73.

363.  Slater JD, Rossi Jr CJ, Yonemoto LT, et al. Proton therapy for prostate cancer: the initial Loma Linda University  experience.  Int  J  Radiat  Oncol  Biol Phys 2004;59(2):348–52.

364.  Zietman AL, DeSilvio ML, Slater JD, et al. Com-parison of conventional-dose vs high-dose confor-mal  radiation  therapy  in  clinically  localized adenocarcinoma  of  the  prostate:  a  randomized controlled trial. JAMA 2005;294(10):1233–9.

365.  Bolla  M,  Gonzalez  D,  Warde  P,  et  al.  Improved survival in patients with locally advanced prostate cancer treated with radiotherapy and goserelin. N Engl J MedN Engl J Med 1997;337(5):295–300.

366.  Bolla  M,  Collette  L,  Blank  L,  et  al.  Long-term results with immediate androgen suppression and external  irradiation  in  patients  with  locally advanced  prostate  cancer  (an  EORTC  study):  a  phase  III  randomised  trial.  Lancet  2002; 360(9327):103–6.

367.  Pilepich MV, Caplan R, Byhardt RW, et al. Phase III trial of androgen suppression using goserelin in unfavorable-  prognosis  carcinoma  of  the  prostate treated  with  definitive  radiotherapy:  report  of Radiation  Therapy  Oncology  Group  Protocol 85-31. J Clin Oncol 1997;15(3):1013–21.

368.  D’Amico AV, Schultz D, Loffredo M,  et  al. Bio-chemical outcome following external beam radia-tion therapy with or without androgen suppression therapy  for  clinically  localized  prostate  cancer. JAMA 2000;284(10):1280–3.

369.  Jones CU, Hunt D, McGowan DG, et al. Radio-therapy  and  short-term  androgen  deprivation  for localized  prostate  cancer.  N  Engl  J  Med 2011;365(2):107–18.

370.  Gleave ME, Goldenberg SL, Chin JL, et al. Ran-domized  comparative  study  of  3  versus  8-month neoadjuvant hormonal therapy before radical pros-tatectomy:  biochemical  and  pathological  effects.  J Urol 2001;166(2):500–6; discussion 506–7.

371.  Blank  KR,  Whittington  R,  Arjomandy  B,  et  al. Neoadjuvant androgen deprivation prior to trans-perineal  prostate brachytherapy:  smaller  volumes, less  morbidity.  Cancer  J  Sci  Am  1999;5(6): 370–3.

372.  Potters L, Torre T, Ashley R, Leibel S. Examining the  role  of  neoadjuvant  androgen  deprivation  in patients undergoing prostate brachytherapy. J Clin Oncol 2000;18(6):1187–92.

373.  Kupelian PA, Katcher J, Levin HS, Klein EA. Stage T1-2  prostate  cancer:  a  multivariate  analysis  of factors  affecting  biochemical  and  clinical  failures after radical prostatectomy. Int J Radiat Oncol Biol Phys 1997;37(5):1043–52.

374.  Grossfeld  GD,  Tigrani  VS,  Nudell  D,  et  al.  Management  of  a  positive  surgical  margin  after radical  prostatectomy:  decision  analysis.  J  Urol 2000;164(1):93–9; discussion 100.

375.  Anscher MS, Prosnitz LR. Multivariate analysis of factors  predicting  local  relapse  after  radical prostatectomy—possible indications for postoper-ative  radiotherapy.  Int  J  Radiat  Oncol  Biol  Phys 1991;21(4):941–7.

376.  Zietman AL, Shipley WU, Coen JJ. Radical pros-tatectomy and radical radiation therapy for clinical 

stages  T1  to  2  adenocarcinoma  of  the  prostate:  new  insights  into  outcome  from  repeat  biopsy  and  prostate  specific  antigen  followup.  J  Urol 1994;152(5 Pt 2):1806–12.

377.  Paulson DF. Impact of radical prostatectomy in the management of clinically  localized disease. J Urol 1994;152(5 Pt 2):1826–30.

378.  Valicenti  RK,  Gomella  LG,  Ismail  M,  et  al.  The  efficacy  of  early  adjuvant  radiation  therapy  for pT3N0 prostate cancer: a matched-pair analy-sis.  Int  J  Radiat  Oncol  Biol  Phys  1999;45(1): 53–8.

379.  Petrovich  Z,  Lieskovsky  G,  Stein  JP,  et  al.  Com-parison of surgery alone with surgery and adjuvant radiotherapy for pT3N0 prostate cancer. BJU Int 2002;89(6):604–11.

380.  Taylor  NKJ,  Kuban  DA,  et  al.  Adjuvant  and Salvage Radiotherapy After Radical Prostatectomy For Prostate Cancer. Int J Radiat Oncol Biol Phys 2003;56(3):755–63.

381.  Thompson  Jr  IM, Tangen CM, Paradelo  J,  et  al. Adjuvant radiotherapy for pathologically advanced prostate cancer: a randomized clinical trial. JAMA 2006;296(19):2329–35.

382.  Bottke  D, Wiegel T.  Adjuvant  radiotherapy  after radical prostatectomy: indications, results and side effects. Urol Int 2007;78(3):193–7.

383.  Pound  CR,  Partin  AW,  Eisenberger  MA,  et  al. Natural history of progression after PSA elevation following  radical  prostatectomy.  JAMA  1999; 281(17):1591–7.

384.  Lu-Yao  GL,  Potosky  AL,  Albertsen  PC,  et  al. Follow-up prostate cancer treatments after radical prostatectomy:  a  population-based  study.  J  Natl Cancer Inst 1996;88(3-4):166–73.

385.  Partin AW, Pearson JD, Landis PK, et al. Evalua-tion  of  serum  prostate-specific  antigen  velocity after  radical  prostatectomy  to  distinguish  local recurrence  from  distant  metastases.  Urology 1994;43(5):649–59.

386.  Cox JD, Gallagher MJ, Hammond EH, et al. Con-sensus statements on radiation therapy of prostate cancer: guidelines for prostate re-biopsy after radia-tion and for radiation therapy with rising prostate-specific  antigen  levels  after  radical  prostatectomy. American  Society  for Therapeutic  Radiology  and Oncology  Consensus  Panel.  J  Clin  Oncol 1999;17(4):1155.

387.  Song DY, Thompson TL, Ramakrishnan V, et al. Salvage  radiotherapy  for  rising  or  persistent  PSA after  radical  prostatectomy.  Urology  2002;60(2): 281–7.

388.  Cadeddu JA, Partin AW, DeWeese TL, Walsh PC. Long-term results of radiation therapy for prostate cancer recurrence following radical prostatectomy. J Urol 1998;159(1):173–7; discussion 177–8.

389.  Stephenson  AJ,  Shariat  SF,  Zelefsky  MJ,  et  al. Salvage radiotherapy for recurrent prostate cancer after  radical prostatectomy.  JAMA 2004;291(11): 1325–32.

390.  Trock  BJ,  Han  M,  Freedland  SJ,  et  al.  Prostate cancer-specific  survival  following  salvage  radio-therapy  vs  observation  in  men  with  biochemical recurrence  after  radical  prostatectomy.  JAMA 2008;299(23):2760–9.

391.  Corn BW, Winter K, Pilepich MV. Does androgen suppression  enhance  the  efficacy  of  postoperative irradiation? A secondary analysis of RTOG 85-31. Radiation  Therapy  Oncology  Group.  Urology 1999;54(3):495–502.

392.  Taylor  CD,  Elson  P,  Trump  DL.  Importance  of continued  testicular  suppression  in  hormone-refractory  prostate  cancer.  J  Clin  Oncol 1993;11(11):2167–72.

393.  Van  Cangh  PJ,  Richard  F,  Lorge  F,  et  al.  Adjuvant radiation therapy does not cause urinary incontinence  after  radical  prostatectomy:  results  of  a  prospective  randomized  study.  J  Urol  1998; 159(1):164–6.

Page 42: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

PartIII:SpecificMalignancies1496.e8

394.  Formenti  SC,  Lieskovsky  G,  Skinner  D,  et  al. Update  on  impact  of  moderate  dose  of  adjuvant radiation on urinary continence and sexual potency in  prostate  cancer  patients  treated  with  nerve-sparing  prostatectomy.  Urology  2000;56(3): 453–8.

395.  Bastasch MD, Teh BS, Mai WY, et al. Post-nerve-sparing  prostatectomy,  dose-escalated  intensity-modulated radiotherapy: effect on erectile function. Int J Radiat Oncol Biol Phys 2002;54(1):101–6.

396.  Katz  MS,  Zelefsky  MJ,  Venkatraman  ES,  et  al. Predictors  of  biochemical  outcome  with  salvage conformal radiotherapy after radical prostatectomy for  prostate  cancer.  J  Clin  Oncol  2003;21(3): 483–9.

397.  DeWeese  TL,  Song  DY.  Current  evidence  for  the  role  of  combined  androgen  suppression  and radiation  in  the  treatment  of  adenocarcinoma  of the prostate. Urology 2000;55(2):169–74.

398.  Del Regato  J. Radiotherapy  for  carcinoma of  the prostate.  A  report  from  the  Committee  for  the Cooperative Study of Radiotherapy for Carcinoma of the Prostate 1968.

399.  Neglia WJ, Hussey DH, Johnson DE. Megavoltage radiation therapy for carcinoma of the prostate. Int J Radiat Oncol Biol Phys 1977;2(9-10):873–83.

400.  van der Werf-Messing B, Sourek-Zikova V, Blonk DI. Localized advanced carcinoma of the prostate: radiation  therapy  versus  hormonal  therapy.  Int  J Radiat Oncol Biol Phys 1976;1(11-12):1043–8.

401.  Green  N,  Bodner  H,  Broth  E,  et  al.  Improved control of bulky prostate carcinoma with sequen-tial  estrogen  and  radiation  therapy.  Int  J  Radiat Oncol Biol Phys 1984;10(7):971–6.

402.  Mukamel  E,  Servadio  C,  Lurie  H.  Combined external  radiotherapy  and  hormonal  therapy  for localized  carcinoma  of  the  prostate. The  Prostate 1983;4(3):283–7.

403.  Pilepich MV, Krall JM, Sause WT, et al. Prognostic factors  in  carcinoma  of  the  prostate—analysis  of RTOG study 75-06. Int J Radiat Oncol Biol Phys 1987;13(3):339–49.

404.  Pilepich MV, Winter K, John MJ, et al. Phase III radiation  therapy  oncology  group  (RTOG)  trial 86-10 of androgen deprivation adjuvant to defini-tive radiotherapy in locally advanced carcinoma of the  prostate.  Int  J  Radiat  Oncol  Biol  Phys 2001;50(5):1243–52.

405.  Lawton CA, Winter K, Murray K, et al. Updated results of the phase III Radiation Therapy Oncol-ogy  Group  (RTOG)  trial  85-31  evaluating  the potential benefit of  androgen  suppression  follow-ing  standard  radiation  therapy  for  unfavorable prognosis  carcinoma of  the prostate.  Int  J Radiat Oncol Biol Phys 2001;49(4):937–46.

406.  Hanks  GE,  Lu  JD,  Machtay  M,  et  al.  RTOG  protocol 92-02: a phase III trial of the use of long term androgen suppression following neoadjuvant hormonal  cytoreduction  and  radiotherapy  in locally  advanced  carcinoma  of  the  prostate. (Abstract).  Int  J  Radiat  Oncol  Biol  Phys 2000;48(Suppl):112.

407.  Widmark A, Klepp O, Solberg A, et al. Endocrine treatment, with or without radiotherapy, in locally advanced  prostate  cancer  (SPCG-7/SFUO-3):  an open  randomised  phase  III  trial.  Lancet  2009; 373(9660):301–8.

408.  Partin AW, Hanks GE, Klein EA,  et  al. Prostate-specific  antigen  as  a  marker  of  disease  activity  in prostate cancer. Oncology (Huntingt) 2002;16(9): 1218–24; discussion 1224, 1227-18 passim.

409.  Partin AW, Hanks GE, Klein EA,  et  al. Prostate-specific  antigen  as  a  marker  of  disease  activity  in prostate  cancer.  Oncology  (Huntingt) 2002;16(8):1024–38,  1042;  discussion  1042, 1047–1028, 1051.

410.  Antonarakis  ES,  Trock  BJ,  Feng  Z,  et  al.  The natural  history  of  metastatic  progression  in  men with  PSA-recurrent  prostate  cancer  after  radical 

prostatectomy:  25-year  follow-up.  J  Clin  Oncol 2009;27(15S).

411.  D’Amico AV, Moul JW, Carroll PR, et al. Surro-gate end point for prostate cancer-specific mortal-ity after radical prostatectomy or radiation therapy. J Natl Cancer Inst 2003;95(18):1376–83.

412.  Hussain M, Goldman B, Tangen C, et al. Prostate-specific  antigen  progression  predicts  overall  sur-vival  in  patients  with  metastatic  prostate  cancer: data from Southwest Oncology Group Trials 9346 (Intergroup Study 0162) and 9916. J Clin Oncol 2009;27(15):2450–6.

413.  Scher  HI,  Eisenberger  M,  D’Amico  AV,  et  al.  Eligibility  and  outcomes  reporting  guidelines  for clinical  trials  for  patients  in  the  state  of  a  rising prostate-specific  antigen:  recommendations  from the  Prostate-Specific  Antigen  Working  Group.  J Clin Oncol 2004;22(3):537–56.

414.  Nelson JB, Love W, Chin JL, et al. Phase 3, ran-domized, controlled trial of atrasentan in patients with  nonmetastatic,  hormone-refractory  prostate cancer. Cancer 2008;113(9):2478–87.

415.  Smith  MR,  Kabbinavar  F,  Saad  F,  et  al.  Natural history of rising serum prostate-specific antigen in men  with  castrate  nonmetastatic  prostate  cancer.  J Clin Oncol 2005;23(13):2918–25.

416.  Smith MR, Cook R, Lee KA, Nelson JB. Disease and host characteristics as predictors of time to first bone metastasis and death in men with progressive castration-resistant  nonmetastatic  prostate  cancer. Cancer 2011;117(10):2077–85.

417.  Eisenberger MA, Blumenstein BA, Crawford ED, et al. Bilateral orchiectomy with or without fluta-mide for metastatic prostate cancer. N Engl J Med 1998;339(15):1036–42.

418.  Hussain M, Tangen CM, Higano C, et al. Absolute prostate-specific  antigen  value  after  androgen deprivation  is  a  strong  independent  predictor  of survival  in  new  metastatic  prostate  cancer:  Data from Southwest Oncology Group trial 9346 (INT-0162). J Clin Oncol 2006;24(24):3984–90.

419.  Makarov DV, Humphreys EB, Mangold LA, et al. The natural history of men  treated with deferred androgen deprivation therapy in whom metastatic prostate cancer developed following radical prosta-tectomy.  J  Urol  2008;179(1):156–61;  discussion 161-152.

420.  Scher  HI,  Sawyers  CL.  Biology  of  progressive, castration-resistant prostate cancer: directed thera-pies targeting the androgen-receptor signaling axis. J Clin Oncol 2005;23(32):8253–61.

421.  Chen CD, Welsbie DS, Tran C,  et  al. Molecular determinants of resistance to antiandrogen therapy. Nat Med 2004;10(1):33–9.

422.  Kelly  WK,  Scher  HI.  Prostate  specific  antigen decline  after  antiandrogen  withdrawal:  the  fluta-mide  withdrawal  syndrome.  J  Urol.  Mar  1993; 149(3):607–9.

423.  Hu R, Lu C, Mostaghel EA, et  al. Distinct  tran-scriptional  programs  mediated  by  the  ligand-dependent  full-length  androgen  receptor  and  its splice  variants  in  castration-resistant  prostate cancer. Cancer Res 2012;72(14):3457–62.

424.  Montgomery RB, Mostaghel EA, Vessella R, et al. Maintenance of  intratumoral  androgens  in meta-static prostate cancer: a mechanism for castration-resistant tumor growth. Cancer Res 2008;68(11): 4447–54.

425.  Thompson  IM,  Zeidman  EJ,  Rodriguez  FR. Sudden death due to disease flare with luteinizing hormone-releasing  hormone  agonist  therapy  for carcinoma  of  the  prostate.  J  Urol  1990;144(6): 1479–80.

426.  Cox RL, Crawford ED. Estrogens in the treatment of prostate cancer. J Urol 1995;154(6):1991–8.

427.  Labrie F, Dupont A, Giguere M, et al. Advantages of the combination therapy in previously untreated and treated patients with advanced prostate cancer. J Steroid Biochem 1986;25(5B):877–83.

428.  Maximum  androgen  blockade  in  advanced  pros-tate  cancer:  an  overview  of  22  randomised  trials with 3283 deaths in 5710 patients. Prostate Cancer Trialists’  Collaborative  Group.  Lancet  1995; 346(8970):265–9.

429.  Chen Y, Clegg NJ, Scher HI. Anti-androgens and androgen-depleting  therapies  in  prostate  cancer: new agents for an established target. Lancet Oncol 2009;10(10):981–91.

430.  Scher HI, Beer TM, Higano CS, et al. Antitumour activity of MDV3100 in castration-resistant pros-tate  cancer:  a  phase  1-2  study.  Lancet  2010; 375(9724):1437–46.

431.  De  Bono  JS,  Fizazi  K,  Saad  F,  et  al.  Primary,  secondary,  and  quality-of-life  endpoint  results from the phase III AFFIRM study of MDV3100, an  androgen  receptor  signaling  inhibitor.  J  Clin Oncol 2012;30:(suppl; abstr 4519).

432.  Rathkopf DE, Danila DC, Slovin SF, et al. A first-in-human, open-label, phase I/II safety, pharmaco-kinetic,  and  proof-of-concept  study  of  ARN-509 in  patients  with  progressive  advanced  castration-resistant  prostate  cancer  (CRPC).  J  Clin  Oncol 2011;29:(suppl; abstr TPS190).

433.  Mostaghel EA, Page ST, Lin DW, et al. Intrapros-tatic  androgens  and  androgen-regulated  gene expression  persist  after  testosterone  suppression: therapeutic  implications  for  castration-resistant prostate  cancer.  Cancer  Res  2007;67(10): 5033–41.

434.  de  Bono  JS,  Logothetis  CJ,  Molina  A,  et  al.  Abiraterone  and  increased  survival  in  metastatic prostate  cancer.  N  Engl  J  Med  2011;364(21): 1995–2005.

435.  Vasaitis TS, Bruno RD, Njar VC. CYP17 inhibi-tors for prostate cancer therapy. J Steroid Biochem Mol Biol 2011;125(1-2):23–31.

436.  Byar DP, Corle DK. Hormone therapy for prostate cancer:  results  of  the  Veterans  Administration Cooperative  Urological  Research  Group  studies. NCI Monogr 1988;(7):165–70.

437.  Immediate versus deferred treatment for advanced prostatic  cancer:  initial  results  of  the  Medical Research  Council  Trial.  The  Medical  Research Council Prostate Cancer Working Party Investiga-tors  Group.  British  journal  of  urology  1997; 79(2):235–46.

438.  Messing EM, Manola J, Sarosdy M, et al. Immedi-ate hormonal therapy compared with observation after  radical  prostatectomy  and  pelvic  lymphade-nectomy  in  men  with  node-positive  prostate cancer. N Engl J Med 1999;341(24):1781–8.

439.  Saylor PJ, Smith MR. Adverse effects of androgen deprivation therapy: defining the problem and pro-moting  health  among  men  with  prostate  cancer.  J Natl Compr Canc Netw 2010;8(2):211–23.

440.  Akakura K, Bruchovsky N, Goldenberg SL, et al. Effects  of  intermittent  androgen  suppression  on  androgen-dependent  tumors.  Apoptosis  and serum  prostate-specific  antigen.  Cancer  1993; 71(9):2782–90.

441.  Russo P, Liguori G, Heston WD, et al. Effects of intermittent diethylstilbestrol diphosphate admin-istration  on  the  R3327  rat  prostatic  carcinoma. Cancer Res 1987;47(22):5967–70.

442.  Conti PD, Atallah AN, Arruda H, et al. Intermit-tent  versus  continuous  androgen  suppression  for prostatic  cancer.  Cochrane  Database  Syst  Rev 2007(4):CD005009.

443.  Oefelein MG. Health related quality of  life using serum  testosterone  as  the  trigger  to  re-dose  long acting  depot  luteinizing  hormone-releasing hormone agonists in patients with prostate cancer. J Urol 2003;169(1):251–5.

444.  Klotz  L,  O’Callaghan  CJ,  Ding  K,  et  al.  A  phase III randomized trial comparing intermittent versus  continuous  androgen  suppression  for patients with PSA progression after radical therapy: NCIC  CTG  PR.7/SWOG  JPR.7/CTSU  JPR.7/

Page 43: 84 - Prostate Cancer · androgen-signaling pathway, including abiraterone acetate and enzalutamide, prolong survival of ... theoduction intr of new prostate cancer treatment approaches.

1496.e9ProstateCancer • CHAPTER84

UK Intercontinental Trial CRUKE/01/013. J Clin Oncol 2011;29: (suppl 7; abstr 3).

445.  Hussain M, Tangen CM, Higano CS, et al. Inter-mittent (IAD) versus continuous androgen depri-vation  (CAD)  in  hormone  sensitive  metastatic prostate cancer (HSM1PC) patients (pts): Results of  S9346  (INT-0162),  an  international  phase  III trial. J Clin Oncol 2012;30:(suppl; abstr 4).

446.  Jordan MA, Wilson L. Microtubules as a target for anticancer  drugs.  Nature  reviews.  Cancer  2004; 4(4):253–65.

447.  Platz  EA,  Yegnasubramanian  S,  Liu  JO,  et  al.  A  novel  two-stage,  transdisciplinary  study  identi-fies digoxin as a possible drug for prostate cancer treatment. Cancer Discov 2011;1(1):68–77.

448.  Darshan  MS,  Loftus  MS,  Thadani-Mulero  M,  et al. Taxane-induced blockade to nuclear accumu-lation  of  the  androgen  receptor  predicts  clinical responses in metastatic prostate cancer. Cancer Res 2011;71(18):6019–29.

449.  Tannock IF, de Wit R, Berry WR, et al. Docetaxel plus prednisone or mitoxantrone plus prednisone for  advanced  prostate  cancer.  N  Engl  J  Med 2004;351(15):1502–12.

450.  Armstrong AJ, Garrett-Mayer ES, Yang YC, et al. A  contemporary  prognostic  nomogram  for  men with  hormone-refractory  metastatic  prostate cancer: a TAX327 study analysis. Clin Cancer Res 2007;13(21):6396–403.

451.  Petrylak  DP,  Tangen  CM,  Hussain  MH,  et  al. Docetaxel and estramustine compared with mito-xantrone  and  prednisone  for  advanced  refractory prostate  cancer.  N  Engl  J  Med  2004;351(15): 1513–20.

452.  Beer TM, Garzotto M, Henner WD, et al. Inter-mittent  chemotherapy  in  metastatic  androgen-independent  prostate  cancer.  Br  J  Cancer  2003; 89(6):968–70.

453.  Meulenbeld  HJ,  Hamberg  P,  de Wit  R.  Chemo-therapy  in patients with  castration-resistant  pros-tate  cancer.  Eur  J  Cancer  2009;45(Suppl  1): 161–71.

454.  Paller  CJ,  Antonarakis  ES.  Cabazitaxel:  a  novel second-line  treatment  for  metastatic  castration-resistant prostate cancer. Drug Design Devel Ther 2011;5:117–24.

455.  Mita AC, Denis LJ, Rowinsky EK,  et  al. Phase  I and  pharmacokinetic  study  of  XRP6258  (RPR 116258A),  a  novel  taxane,  administered  as  a 1-hour  infusion  every  3  weeks  in  patients  with advanced  solid  tumors.  Clin  Cancer  Res  2009; 15(2):723–30.

456.  de  Bono  JS,  Oudard  S,  Ozguroglu  M,  et  al.  Prednisone  plus  cabazitaxel  or  mitoxantrone  for  metastatic  castration-resistant  prostate  cancer progressing  after  docetaxel  treatment:  a  ran-domised open-label trial. Lancet 2010;376(9747): 1147–54.

457.  Drake CG. Prostate cancer as a model for tumour immunotherapy.  Nat  Rev  Immunol  2010;10(8): 580–93.

458.  Antonarakis  ES,  Drake  CG.  Current  status  of immunological therapies for prostate cancer. Curr Opin Urol 2010;20(3):241–6.

459.  Small  EJ,  Fratesi  P,  Reese  DM,  et  al.  Immuno-therapy of hormone-refractory prostate cancer with antigen-loaded  dendritic  cells.  J  Clin  Oncol 2000;18(23):3894–903.

460.  Higano CS, Schellhammer PF, Small EJ, et al. Inte-grated  data  from  2  randomized,  double-blind, placebo-controlled, phase 3 trials of active cellular immunotherapy  with  sipuleucel-T  in  advanced prostate cancer. Cancer 2009;115(16):3670–9.

461.  Kantoff  PW,  Higano  CS,  Shore  ND,  et  al. Sipuleucel-T  immunotherapy  for  castration-resistant  prostate  cancer.  N  Engl  J  Med  2010; 363(5):411–22.

462.  Saad F, Gleason DM, Murray R, et al. A random-ized, placebo-controlled trial of zoledronic acid in patients with hormone-refractory metastatic pros-tate  carcinoma.  J  Natl  Cancer  Inst  2002;94(19): 1458–68.

463.  Saad F, Gleason DM, Murray R, et al. Long-term efficacy  of  zoledronic  acid  for  the  prevention  of 

skeletal  complications  in patients with metastatic hormone-refractory prostate cancer. J Natl Cancer Inst 2004;96(11):879–82.

464.  Fizazi K, Carducci M, Smith M, et al. Denosumab versus zoledronic acid for treatment of bone metas-tases  in  men  with  castration-resistant  prostate cancer:  a  randomised, double-blind  study. Lancet 2011;377(9768):813–22.

465.  Vallet S, Smith MR, Raje N. Novel bone-targeted strategies  in  oncology.  Clin  Cancer  Res 2010;16(16):4084–93.

466.  Brown  JM, Corey E, Lee ZD,  et  al. Osteoprote-gerin and rank ligand expression in prostate cancer. Urology 2001;57(4):611–6.

467.  Smith  MR,  Egerdie  B,  Hernandez  Toriz  N,  et  al.  Denosumab  in  men  receiving  androgen-deprivation  therapy  for prostate cancer. N Engl  J Med 2009;361(8):745–55.

468.  Goyal  J,  Antonarakis  ES.  Bone-targeting  radio-pharmaceuticals  for  the  treatment  of  prostate cancer  with  bone  metastases.  Cancer  letters 2012;323(2):135–46.

469.  Nilsson S, Larsen RH, Fossa SD, et al. First clinical experience with alpha-emitting radium-223 in the treatment of  skeletal metastases. Clin Cancer Res 2005;11(12):4451–9.

470.  Bruland  OS,  Nilsson  S,  Fisher  DR,  Larsen  RH. High-linear energy transfer irradiation targeted to skeletal  metastases  by  the  alpha-emitter  223Ra: adjuvant or alternative to conventional modalities? Clin Cancer Res 2006;12(20 Pt 2):6250s–7s.

471.  Nilsson  S,  Franzen  L,  Parker  C,  et  al.  Bone-targeted  radium-223  in  symptomatic,  hormone-refractory  prostate  cancer:  a  randomised, multicentre,  placebo-controlled  phase  II  study. Lancet Oncol 2007;8(7):587–94.

472.  Parker  P,  Nilsson  S,  Heinrich  D,  et  al.  Updated analysis  of  the  phase  III,  double-blind,  random-ized, multinational  study of  radium-223 chloride in  castration-resistant  prostate  cancer  (CRPC) patients  with  bone  metastases  (ALSYMPCA).  J Clin Oncol 2012;30:(suppl; abstr LBA4512)