Prostate Cancer

Published on March 2017 | Categories: Documents | Downloads: 84 | Comments: 0 | Views: 1231
of 118
Download PDF   Embed   Report

Comments

Content

Chapter 10. PROSTATE CANCER
Acknowledgments: Hema Samaratunga, Chris Schmidt, Mark Frydenberg, John Yaxley for their guidance and advice in preparation of this document and to the authors of the previous edition Mitchell H Sokoloff, William B Isaacs, and Leland WK Chung.

I. INTRODUCTION
Prostate Cancer is an increasingly common diagnosis in Western societies and in those emulating Western lifestyles and diets. In the year 2001 there were estimated to be 198,100 new cases and almost 31,900 deaths attributable to this condition in the United States (1). Approximately one in seven American men will be diagnosed with prostate cancer during their lifetime, making it the most common solid tissue cancer in the United States. Despite advances in prevention and early detection, refinements in surgical technique and improvements in adjuvant radio-therapy and chemotherapy, the ability to cure many patients with prostate cancer remains elusive. However, mortality rates are changing. Baade et al recently reviewed international trends in prostate cancer mortality and reported significant reductions in prostate-cancer mortality in the UK, USA, Austria, Canada, Italy, France, Germany, Australia and Spain with downward trends in the Netherlands, Ireland and Sweden (2). Detection of this disease earlier, as a consequence of introduction of the prostate specific antigen (PSA) blood test, has been acknowledged by the NCI as one factor contributing to lowering the mortality rate over the past few years (3-6). The use of PSA testing has been estimated to provide a diagnostic lead-time of up to 10 years (7-11). In the mid to late 1980s only one third of prostate cancers were diagnosed at curable stages compared with today when 80% are staged clinically as organ-confined and potentially curable (12-14). Unfortunately, however, even when the tumour is thought to be localized, up to 25% of men have non-localised disease which declares itself subsequently (15). Since curative therapies are directed to localised tumours (3,4,7,16), extending effective but non-invasive treatments to include both primary and secondary lesions remains a major goal and challenge. Once prostate cancer metastasizes, apart from causing loss of life, its toll is often considerable with regard to morbidity from both the disease itself and administered therapies. As a result of increasing numbers of men having their prostate cancers diagnosed earlier, more patients are now eligible for treatment with curative intent. Improved surgical and radiation-based treatments have been developed so that the prognosis of a man diagnosed today with prostate cancer is better than ever before.

II. ANATOMY AND PHYSIOLOGY
The term "prostate", originally derived from the Greek word prohistani which means "to stand in front of," has been attributed to Herophilus of Alexandria who used the term in 355 B.C.E. to describe the small organ located in front of the bladder (17). The prostate gland is a small firm organ, about the size of a chestnut, located below the bladder and in front of the rectum. The urethra, the channel through which urine is voided, passes from the bladder and through the prostate and penis. The primary function of the prostate gland, which contracts with ejaculation, is to provide enzymes to maintain the fluid nature of seminal fluid and to nourish sperm as they pass through the the prostatic and penile urethra to outside the body.

1

Chapter 10. PROSTATE CANCER

III. NATURAL HISTORY OF PROSTATE DISEASE
Traditionally, prostate cancer was considered a disease of "older men." As such, it was generally accepted that "men never died from prostate cancer, they died of other conditions with prostate cancer." Consequently, treatment was conservative and directed toward palliation and prevention of the emergence of debilitating and painful sequelae. In addition, diagnosis was generally made on the basis of palpating a rockhard and nodular prostate on digital rectal exam [DRE] or by symptoms and signs of primary or secondary tumors, such as urinary obstruction, back pain, nerve root or, less commonly, spinal cord compression. Almost invariably, most tumors had already disseminated at the time of diagnosis and, therefore, were incurable. It was in the mid-1980s, with the introduction of the PSA blood test that prostate cancer began to be diagnosed earlier and in younger men. Although methods of diagnosis and treatment of localized disease have become wellestablished, the management of prostate cancer remains controversial due to its variable biologic course, imprecision with clinical staging and limitations in prediction of the clinical outcome of patients with both organ-confined and locally-invasive disease - not to mention the morbidity associated with all currently established treatments. It is sobering to muse that, were the unwanted effects of diagnosis and treatment insignificant, the dilemma of whether or not to treat would not be an issue.

IV. BASIC SCIENCE OF PROSTATE CANCER
(i) Introduction
As described in the previous section the prostate consists of two major cellular types, the stroma and the epithelium. The stromal component is comprised of smooth muscle cells, fibroblasts and endothelial cells; 5 cell sub-types, basal epithelium, secretory epithelium, transit amplifying cells, neuroendocrine cells and stem cells, constitute the epithelium Prostate cancer is non-regulated proliferation primarily of the epithelium. A number of factors may contribute to dedifferentiation and proliferation of epithelial cells, including aberrations in surrounding stromal cells via cytokines and growth factors secreted by both stromal and epithelial cell types (18,19). To date, the exact triggers for development of prostatic adenocarcinoma have not been elucidated, although epidemiological studies have shown links to both environmental (discussed subsequently) and genetic contributors. A variety of growth factors and cytokines have been shown to influence the growth rate and development of prostatic cancers, and a number of proteins and genes have been identified whose regulation and/or function are altered between malignant and benign states. Several of these proteins and genes have been flagged as potential biomarkers and therapeutic targets for prostate cancer. This chapter will provide an overview of the current literature, and will discuss factors that contribute to the development and progression of PCa, as well as the potential of new biomarkers for diagnosing and treating this disorder.

A Genetics of Prostate Cancer
Genetic aberrations, whether inherited or acquired, are universally associated with cancer development, growth, and progression. Several cancer types, such as breast, renal and colorectal, have been shown to have a familial component (20). In an effort to identify important genetic determinants associated with hereditary prostate cancer, a number of research groups have employed different methodologies involving large-scale populations (epidemiology), families (with one or more affected individuals) and twins. 2

Chapter 10. PROSTATE CANCER

(i) Epidemiology:
Major genetic epidemiologic studies published in the last two decades support the notion that prostate cancer may exist as clusters in families. In the 1980s, a Utah Mormon genealogy study found that prostate cancer exhibited the fourth strongest degree of familial clustering after lip, melanoma, and ovarian cancers . Prostate cancer, interestingly, had a higher familial association than either colon or breast carcinoma, which are known to be predisposed by genetic or familial components. A later study, determined cancer pedigrees in 691 men with prostate cancer and 640 spouse controls, and found a positive familial history of prostate cancer . They concluded that men with an affected father or brother were twice as likely to develop prostate cancer as men with no affected relatives. Although these studies strongly suggest that familial clustering of prostate cancer risk does exist, they did not address the underlying aetiological mechanisms. Indeed, familial clustering can reflect either shared environmental and lifestyle risk factors, or a genetic mechanism, or indeed both. In order to address the inherent difficulties in separating the inheritable and environmental causes of prostate cancer, the International Consortium for Prostate Cancer Genetics (ICPCG) was formed. In early 2005 they published a world-wide study of 1,233 families located in North America, Australia, and Europe. While the study included some families with Asian, Hispanic, Native American and African-American backgrounds, 1,166 of the 1,233 families were Caucasian. The researchers found 5 regions (5q12, 8p21, 15q11, 17q21 and 22q12) in human chromosomes that potentially could harbour cancer susceptibility genes. In addition, the authors flagged chromosomal regions which linked to families with 5 or more affected members (22q12, 1q25, 8q13, 13q14, 16p13 and 17q21) and those which corresponded with early onset of disease (less than 65 years; 3p24, 5q35, 11q22, and Xq12) . While, to date, there are no known cancer susceptibility genes in these loci, work is ongoing to refine the mapping of these regions to the gene level.

(ii) Twin Studies:
Twin studies compare the similarities (concordance rate) of a trait or traits, traced to either monozygotic (MZ) or dizygotic (DZ) twins to dissect the relative genetic and environmental influences of a disease. Several twin studies (although with limited subject numbers) of prostate cancer reported higher concordance rates in MZ twins compared to DZ twins, implicating a genetic contribution for prostate cancer. Lichtenstein et al (2000) (24) analyzed Swedish, Finnish and Danish twin registries and concluded that the recurrent risk in the twin of an affected man was 21.1% for MZ and 6.4% for DZ. More recent mathematical modeling of genetic risk of a number of cancers, by the same research group (25), found that the contribution of genetic susceptibility to prostate, breast, and colorectal cancers was small to moderate. Although these aforementioned studies clearly demonstrated a potential genetic basis of prostate cancer, albeit low, for a portion of familial prostate cancers, this rate is much lower than coeliac disease where strongly heritable components were found in 75% of the MZ with 11% prevalence in first-degree relatives . The significant differences in the contribution by the heritable element on the development of prostate cancer, as opposed to coeliac disease, suggest the importance of gene-environment interaction, which may be pivotal in prostate carcinogenesis.

(iii) Gene Association Studies:
In addition to hereditary prostate cancer-associated genes, a number of studies have focused on genetic components that may be associated with prostate cancer development. Association studies are generally case-control studies based on the comparison of frequencies of an allele in unrelated cases and normal controls. A significant difference in allele frequency between cases and controls can be expected if a particular allele harbours causal variations or if that allele is closely linked with (or in close proximity to) a disease-causing mutation. These types of association studies are sus3

Chapter 10. PROSTATE CANCER ceptible to aetiological (genetic versus environmental), inheritance (dominant, recessive, or X-linked), and locus heterogeneity, as well as allelic (same gene but different mutations) and founder (same mutation in different genetic background) heterogeneity. Despite these difficulties, association studies have identified genes that may be associated with, or contribute to, prostate cancer development and progression. 1. BRCA1 and BRCA2 BRCA1 and BRCA2 genes, linked to breast cancer development, have also been examined in prostate cancer due to links between these cancers’ aetiology, observed through epidemiologic, biologic, and molecular studies (27, 28). Recent association studies demonstrated an increase in the frequency of BRCA1 and BRCA2 mutations in prostate cancer carriers . The most recent studies have shown that the rate of prostate cancer diagnosis after age 60 in cases with a deleterious mutation in BRCA1 or BRCA2 was 5.2% compared with 1.9% of controls. . In addition, data reported thus far show that mutations of BRCA1 and BRCA2 may increase the risk of prostate cancer, particularly in the case of BRCA2 for early-onset disease. The contribution of germ line mutations in these genes to familial clusters of prostate cancer needs to be more fully assessed.

2. Androgen Receptor Pathway Androgens serve as a most important ligand for prostate cancer, stimulating growth via the androgen receptor action pathway. The androgen receptor action pathway, broadly defined, includes the androgen receptor, the 5-alpha-reductase gene and the various members of the cytochrome p450 family. Mutations within the AR gene are detected in 10 – 20% of prostate cancer specimens, with mutations more common in hormone escape than hormone sensitive specimens . The androgen receptor (AR) is comprised of three important domains: N-terminal transactivation domain, DNA binding domain, and C-terminal steroid binding domain. Mutations in the N-terminal transactivation domain include variations in the number of polymorphic triplet repeats located within this area of the gene . Several studies have suggested that AR genes with shorter repeat lengths may increase the risk of developing more aggressive prostate cancer, as the number of polyglycine repeats is inversely related to the ability of the AR to activate target genes . Mutations at the steroid-binding domain have been associated the promiscuous nature of AR which recognizes non-androgen ligands, such as oestrogen, progesterone, and antiandrogens as agonists that confer AR-induced gene transactivation and growth stimulatory activities (36). Overall, the increase in incidence of mutations and the variety of mutations observed in the AR in metastatic cancers, suggest that these regions of AR heterogeneity may be acquired through gene-environment interactions. Amplification or duplication of the AR gene has been associated with the transition from hormone sensitive to hormone refractory tumours. Edwards et al, (2003) (37) found that less than 5% of hormone sensitive prostate cancers contained more than one copy of the AR gene, compared with 20-30% of hormone resistant tumours. In 80% of these cancers there was a corresponding increase in AR protein concentrations. However, it was the presence of the gene duplication and not just higher AR protein levels which significantly predicted poorer survival rates for patients in their study . Alterations in the 3-beta-hydroxysteriod dehydrogenase gene (also a member of the AR pathway) have also been strongly associated with hereditary prostate cancer . 3-beta-hydroxysteroid dehydrogenase metabolises dihydrotestosterone to 3-alphaand 3-beta-diol. Recently Guerini et al (2005) have found that 3-beta-diol does not bind androgen receptors but efficiently binds the estrogen receptor (ER-beta), and exerts an inhibitory effect on prostate cancer cells through the activation of ER-beta signaling. Subsequently, deletiorous mutations in the 3-beta-hydroxysteriod dehydrogenase gene are thought to confer a growth advantage to cancer cells. 4

Chapter 10. PROSTATE CANCER Although the androgen receptor pathway dominates in hormonal-tumour interactions in prostate cancer, other relationships are also operative. As well as oestrogens having an increasingly recognised role, the growth hormone secretagogue ghrelin and a preproghrelin isoform have been shown to be highly expressed in prostate cancer with activation via the protein kinase pathway (39). Both inhibin and activin subunits are known to be expressed in the normal prostate with the inhibin alpha subunit gene down regulated in prostate cancer with a loss of heterozygosity at the gene locus and methylation of the promoter (40). Overexpression of the activin betaCsubunit has been demonstrated to increase activin AC heterodimer levels, concomitantly reducing activin A levels and decreasing activin signaling in PC3 cells (41).

(iv) Clinical and pathological characterization of hereditary Prostate Cancer
To determine what differences might distinguish hereditary prostate cancer from its sporadic counterparts, a number of clinical features of prostate cancer were examined by Carter, et al. . Clinical stage at presentation, pre-operative PSA, final pathologic stage, and prostate weight were examined in a series of approximately 650 patients divided among three categories. Individuals were classified as having hereditary disease if 3 or more relatives were affected in a single generation, prostate cancer occurred in each of 3 successive generations in either paternal or maternal lineages, or 2 relatives were affected under the age of 65 years. For the other groups, either no other family members were affected (sporadic disease), or other family members were affected but not to the extent found in families classified as hereditary. In summary, no unique clinical or pathological characteristics distinguished hereditary prostate cancer in this group of patients. This parallel between hereditary and sporadic prostate cancer extends to the incidence of multifocality found in both of these categories.

(v) Other molecular changes in Prostate Cancer
While mutations or gene duplications have been associated with familial prostate cancers, other molecular changes have been documented in the majority of prostate tumours. Although the mechanisms behind these changes have not always been elucidated, it is believed that changes in gene methylation patterns, gene expression profiles and expression of non-coding mRNA’s play important roles in cancer development and progression. 1. Silencing of gene expression by CpG methylation One of the mechanisms used by cells to alter gene expression is by conjugation of a methyl group on certain cytosine residues in the gene promoter. In a normal cell this may function to maintain cellular differentiation, but recent reports have described several genes which are abnormally silenced, or activated, in many tumour types including prostate cancer, by promoter methylation. . Lodygin et al, (2005) (43) described many genes, including cell cycle regulating genes, angiogenesis inhibitors, apoptosis inducing genes, growth factor receptors and transcriptional control genes that are “switched off” in prostate cancer. The “switch” could be thrown into reverse by treatment with 5-aza-2’ deoxycytidine, a chemical that removes methyl groups from cytosine residues. A study of 41 primary tumours from prostate cancer patients found that frequencies of CpG methylation detected in the promoter region of a selection of the above genes were as follows, GPX3 (Glutathione peroxidase – detoxification), 93%; SRFP1 (Secreted frizzled-related protein 1 – signaling), 83%; COX2 (cyclooxygenase 2 –signaling) 78%; DKK3 (Dickkopf homologue 3 – signaling) 68%; GSTM1(Glutathione S-transferase M1 - detoxification), 58% and KIP2/p57 (cell cycle control), 56%.

5

Chapter 10. PROSTATE CANCER The authors also found that methylation in some gene promoters did not always equate to reduced transcription rates, and that some genes apparently “switched on” after 5-aza-2’ deoxycytidine treatment did not show detectable CpG methylation. They concluded that silencing of genes by CpG methylation occurs at an early stage of prostate cancer development. This may have implications for the use of certain methylation events as diagnostic markers for prostatic disease (44). Recently, global methylation in the chromosomes of prostate cancer cells has been studied with specific antibodies that detect all CpG sites, both in gene promoters and elsewhere in the chromosomes (45). Interestingly, the overall amount of methylation in tumour cells was decreased when compared with benign cells from the same individuals. It was also noted that men with recurrent disease and/or positive surgical margins showed overall higher levels of methylation in their tumour cells compared with men with early stage, surgically curable disease (45). Brothman et al, (2005) (45) suggest that genomic hypermethlyation is associated with condensed DNA. Hypomethylation of these regions permits a reduction of DNA condensation and may result in chromosomal rearrangement and genetic instability. In addition they suggest that global DNA hypomethlyation precedes hypermethylation of specific gene promoters (such as GSTP1) in cancer development, and the balance between the two methylation states may be critical in determination of aggressive vs non-aggressive disease (45). However, whether CpG methylation is a cause or consequence of cancer remains to be shown.

2. Non-coding mRNA’s: With the rise of microarray technologies, analyses of changes in multiple gene transcription rates have been described. Interestingly, some changes in the transcriptional output of the human genome points to the existence of a significant number of noncoding RNA transcripts, derived from intronic genomic regions, with some of these being oriented antisense relative to the protein-coding mRNA of the gene. To investigate the expression level of intronic messages in human tissues, a subset of approximately 2,000 totally intronic Expressed Sequence Tags (EST) clusters and 2,000 clusters from exonic segments of known genes, was selected for intronic microarray analysis. Hybridization of these intronic microarrays with 27 prostate tumours and corresponding adjacent normal tissue revealed that in prostate cancer, the fraction of expressed messages arising from exonic or intronic transcripts were similar . Moreover, the expression levels of 23 intronic non-coding transcripts correlated (p value - 0.001) with the degree of prostate tumour differentiation. It has not been determined whether the expression of intronic antisense RNAs in tumours is a true mechanism of cancer disregulation or just reflects broad errors in promoter recognition/transcription initiation. However, regulatory elements have been identified for a number of non-coding RNA’s, including DD3/PCA3 and PCGEM1 – see below.

B. Stromal Epithelial Interaction in Prostate Cancer
While a variety of genetic and environmental determinants may contribute to the cause of malignancy, cancer cells display similar phenotypes and modes of development such as changes in cell adhesion, attachment, migration and invasion. These phenotypes and the genes controlling them are in turn regulated by epigenetic factors, such as soluble proteins, steroids and growth factors released by immune cells and adjacent stromal cells (e.g. fibroblasts, smooth muscle cells, endothelial cells, and neuroendocrine cells) and alterations in the extracellular matrix (ECM) surrounding tumor epithelium . Reciprocal cellular interaction between stroma and epithelium is involved in foetal prostate development, post-natal prostate growth and maturation, maintenance of differentiation status, hormonal responsiveness, and the aging and senescence 6

Chapter 10. PROSTATE CANCER of prostate gland in adulthood ). During neoplastic progression, key phenotypic changes in prostate cancer cells are also modulated by a dynamic, two-way communication between tumour epithelium and various stromal cells, including fibroblasts, smooth muscle cells, vascular endothelium and bone-derived cells including osteoblasts . All of these cell types and the factors they secrete form the cell’s microenvironment that can either enhance or repress cancer development and progression. One of the challenges in understanding carcinogenesis is determining how the tumour microenviroment interacts with genetic changes in the cells themselves.

(i) AR-mediated cell communication
Experimental evidence using co-cultured human benign prostatic hyperplasia (BPH)-derived stromal and epithelial cells indicate that the expression of AR, PSA, and 5 alpha-reductase in epithelial cells relies on the inductive influence of neighbouring stromal cells . Similarly, in co-cultured rat prostatic epithelial and fibroblast cells, the androgen responsiveness of prostate epithelial cells can be conferred by the presence of fibroblastic cells . Therefore the action of AR in prostate development is mediated by the local stroma. In addition, oestrogens have been shown to have a role in cancer development, yet the oestrogen receptor is predominantly localized in stromal cells. . Animal models have demonstrated prostatic cell proliferation occurs in androgen deficient mice due to exogenous oestrogen. . In addition, elevated androgen levels in oestrogen deficient mice induced prostatic hyperplasia . While neither of these hormones alone induced prostate cancer, combined androgen and oestrogen therapy did evoke prostatic dysplasia and adenocarcinoma. It is proposed, therefore, that both androgens and oestrogens influence the process of prostate carcinogenesis . The function of androgen ablation therapy (ABT) is to prevent activation of AR regulated genes. Failure of ABT, as in hormonally resistant prostate cancer, was logically linked to genetic modification of the AR, as described above. More recently however, there has been recognition of the effect of post-translational modifications on the AR . In 1998, Blok et al. demonstrated phosphylation of the AR in response to binding by the ligand DHT (dihydrotestosterone). In addition to stabilizing the active AR homodimers, this phosphorylation can influence AR mediated gene activation . In vivo, phosphorylation of AR by MAPK (serine-threonine kinase) and AKT (protein kinase B) sensitizes AR to low levels of DHT, allowing low levels of androgens, or alternative steroids to induce translocation of the AR to the nucleus, facilitating gene activation . MAPK has been shown to be activated in cell lines derived from hormone-refractory tumours, and is correlated with advanced stage and grade in prostate cancer . While the exact mechanism is not fully known, transfection of cells with c-Ras (cell growth regulator; activated in many cancer types), leads to increased activation of MAPK and development of hormone escape . AKT can be activated via PI3 kinase pathway (involved in cell growth, adhesion and migration in many tumours; (69), and specifically phosphorylates AR at Ser210 and Ser790 (70)– see Figure 1 The control of AR function also involves interaction of the receptor with a number of co-factors or proteins that bind either as part of a complex or directly to AR and increase (co-activators) or inhibit (co-repressors) the transcriptional activity of the AR. Co-activators include CBP (CREB-binding protein), beta-catenin, ARA55 and ARA70 which all act to alter ligand specificity of the AR . These mechanisms include allowing the antiandrogens such as hydroxyflutamide (CBP, ARA70) and bicalutamide (ARA70) to act as agonists and/or by permitting low concentrations of adrenal androgens or oestradial to activate AR (ARA55, beta-catenin, ARA70; . Some co-activators activate AR in the absence of ligands, such as SRC-1, SRC-3, p300, Tip60, and c-Jun. C-Jun binds to the AR, promoting homodimerization and subsequently activating AR dependent transcription initiation of downstream genes (such as PSA) . SRC-1 and SRC-3 are members of the steroid receptor cofactor family, which is commonly overexpressed in hormone-refractory prostate cancer. This 7

Chapter 10. PROSTATE CANCER family of proteins normally facilitates AR transcriptional activity in the presence of androgens, however, phosphorylation of SRC-1 by MAPK (activated as discussed above), may be one mechanism by which SRC-1 activates AR in the absence of androgens. Tip60 is linked to the transcription activation activity of AR by inducing changes to AR acetylation. In cell line models, in the absence of androgens, p300 is required for Il-6 (see below) stimulated growth, and it has been proposed that p300 plays an important role in the development of hormone refractory tumours .

Figure 1. Phosphorylation of the androgen receptor (AR) by MAPK and AKT. Numerous signalling factors (intracellular and extracellular), involved in inducing cell growth and proliferation, stimulate the Ras pathway to activate MAPK. Subsequently, MAPK phosphorylates AR, enabling AR to form dimers, enhancing ARE (androgen response element) dependent gene expression. Similarly, intraand extra-cellular factors inducing cell growth and migration and inhibiting cell adhesion activate the PI3 kinase pathway resulting in AKT dependent phosphorylation, dimerisation and activation of AR. A less well understood mechanism of AR co-activators involves movement of AR to the nucleus. In LNCaP cells, STAT3 binds ligand-free AR and facilitates its translocation to the nucleus, and it is via this mechanism that the STAT3/AR complex activates AR dependent genes (in the absence of androgens) in response to IL-6 stimulation .

(ii) Soluble Factors in Cell signalling
The mediation of stromal-epithelial interactions in the normal and malignant prostatic environment involves a number of soluble factors that can serve paracrine, autocrine or intracrine functions (72). Several soluble factors have been identified performing a variety of functions from angiogenesis, growth enhancement, and dedifferentiation. These are listed in Table 1 (72). Table 1. Most commonly cited soluble factor signaling pathways regulating prostate growth and differentiation (from Chung et al., 2005 [19]) 8

Chapter 10. PROSTATE CANCER Soluble Growth Factor References Name Source Receptor Receptor Location Function Regulation at Androgen Independent Progression Disease prognosis neg correlation Disease prognosis pos correlation van Moorselaar RJ, 2002) van Moorselaar RJ, 2002) Knudsen BS, 2004; LailTrecker M, 1998

VEGF

Epithelium, stroma

VEGFR-1, Epithe2 lium, stroma

Angiogenic factor

bFGF (FGF-2)

Stroma

FGF-2R

Epithelium, stroma

Angiogenic factor

HGF/SF

Stroma

c-met

Disease Epithelium Stimulates progression pos cell correlation growth Stroma Induces apoptosis, increases angiogenesis stimulates stromal but inhibits epithelial cell growth

TGF-beta

TGF-beta Epithelium I-III receptors

AugBachman, mented 2004 and expres2005; sion at Yingling, androgen 2004 withdrawal

IGF-I

Stroma

IGF-IR

Epithelium, stroma

UpDjavan B, Stimulates regulation 2001; cell at disease Rubin, growth, 2003 blocks Progression apoptosis Promotes differentiation and apoptosis inhibition Increasing IL-6 signaling during disease progression Edwards, 2005a and b; Hideshima., 2005; Royuela M, 2004

IL-6

Epithelium, stroma

IL-6R, sIL-6R

Epithelium, stroma

9

Chapter 10. PROSTATE CANCER Soluble Growth Factor References (KGF) (FGF-7) Stroma Gp130 KGF-R StimuEpithelium lates cell Growth Stromal Planz, KGF ex1999 and pression 2004 responded to androgen

1. VEGF Recruitment of neovascular endothelial cells to proliferating cancer cells is thought to be required for the maintenance and stimulation of tumour growth, and is mediated by vascular endothelial growth factor (VEGF) and its receptors. VEGF has been shown to be secreted by both glandular and surrounding stromal cells, and VEGF expression can be modulated by a number of treatments including androgen ablation, finasteride, and thalidomide .

2. IL-6 High levels of IL-6 secretion from prostate fibroblasts, smooth muscle cells, and tumour cells themselves, are thought to be a mechanism of ligand independent activation of AR activation via the PI3K-Akt, STAT3 and MAPK pathways in PCa . Interference with IL-6 signaling is a potential means of modulating the growth of advanced prostate cancer. Importantly, IL-6 is secreted by bone marrow stromal cells (BMSCs), and this secretion is further augmented by direct interaction between tumour cells and BMSCs . Studies using an anti-IL-6 monoclonal antibody have shown tumoricidal effects in a murine model

3. IGF-1 The insulin-like growth factor-I (IGF-I) pathway is involved with malignant transformation in various tissues. In prostate cancer, it has been proposed that IGF-1 induces ligand-independent activation of the androgen receptor and enhances the expression of matrix metalloproteinase-2 and urokinase plasminogen activator (see next section, insoluble cell signalling). Progression to androgen independence has also been linked to deregulation of the IGF-1-IGF-1-receptor axis (76). Manipulations of the IGF axis have shown therapeutic potential, for example, antisense RNA to IGF-I receptor inhibits prostate cancer proliferation and invasion, while increasing IGF binding protein 3 expression induces cell death .

4. HGS/SF, EGF and bFGF HGS/SF, EGF and bFGF are all members of a large family of heparin bound growth factors. Hepatocyte growth factor/scatter factor (HGF/SF) and its receptor, the c-met proto-oncogene, were shown to be predominantly expressed by localized and metastatic prostate cancer. Experimental evidence suggests that HGF/SF and c-met downstream signaling may regulate prostate cancer growth and metastasis through enhanced IL-6, androgen receptor, extracellular matrix and integrin interaction . Epidermal growth factor (EGF) its receptor (EGFR), and other family members, erbB2/neu, erbB3 and erbB4, are known to have a role in prostate cancer progression through their interactions with a broad spectrum of soluble factors and their downstream converging signaling pathways (81, 82). 10

Chapter 10. PROSTATE CANCER 5. TGF-beta Transforming growth factor beta type 1 is a ubiquitous cytokine originally named for its ability to transform fibrolasts in culture (83). TGF-beta both inhibits the growth of normal epithelial cells, but paradoxically can induce cancer cell proliferation and promote an Epithelial to Mesenchymal transition (EMT). The mammalian TGF-beta family has 3 subtypes, and can bind to three cell surface receptors (type I, II and III). The type II receptor functions as a tumor repressor gene, and absence of this receptor results in resistance to the growth inhibitory effects of TGF-beta. This growth factor also stimulates angiogenesis, extracellular matrix turnover and host immune surveillance, although the receptors mediating this function have not been well defined (84). TGF-beta specific inhibitors have been developed and have been proposed to have utility in treatment of progressive prostate cancer .

(ii) Insoluble factors and the Extracellular Matrix in Cell signalling
At the stromal /tumor cell interface of adenocarcinomas, there is a noticeable derangement of Extracellular Matrix (ECM), due mainly to the activity of enzymes derived from the host stroma. (72). However, the mechanisms of this breakdown of ECM are unknown, and whether cancer cell invasion develops before or after interacting with host stroma, or whether stroma response is subject to reciprocal regulation by cancer cells remains undetermined. De Wever, and Mareel (2003)(86) suggest that the maintenance of epithelial homeostasis requires the participation of stroma, and therefore it is likely that stromal changes are subsequent to epithelial aberrations. ECM and its degradative products could signal to cancer cells through their cell surface integrin or non-integrin associated receptors. Subsequently, cell behaviours, such as cell polarity, secretion, adhesion, motility and invasion, and integrated cell functions, such as proliferation, differentiation and survival could be affected. 1 Integrins Integrins are a family of heterodimeric, transmembrane receptors that mediate the attachment of cells to the surrounding ECM and function as sensors of the environment. Alterations in integrin expression and signaling have been implicated in many aspects of tumorigenesis and metastasis including cell survival, migration, and invasion. In prostate cancer, the progression from normal to metastatic cells is accompanied by changes in the repertoire of integrins expressed and up-regulation of key adhesion-dependent signaling pathways . Important mediators of Integrin signaling include the adhesion kinases FAK (focal adhesion kinase) PAK (p21 activated kinase) and MAPK (a serine/threonine kinase) . Potentially, Integrins may have more complex roles by coordinating their actions with metalloproteinases and serine proteases, which together may increase cancer cell invasion and migration into secondary sites of cancer growth (72).

2. FAK Overexpression and increased phosphorylation of FAK has been shown in a variety of tumours and these changes correspond to a malignant and metastatic phenotype . FAK has been shown to promote cell surface expression of metalloproteinases, particularly MT1-MMP. This results in activation of other MMP’s – particularly MMP2 enabling ECM degradation and tumour cell invasion .

3.MMP’s: Matrix metalloproteinases (MMPs) are involved in tumour invasion and metastasis in various malignancies. MMP-2 and MMP-9 are capable of digesting collagen type IV, a significant component of basement membranes, and have been implicated in 11

Chapter 10. PROSTATE CANCER prostate cancer progression and metastases ). Preliminary studies show the stimulation of protease induced receptors produced increased levels of MMP-2 and MMP-9 activity in prostate cancer cell lines, indicating their potential role in the metastasis of prostate cancer cells . A crucial role for MMP-9 has been demonstrated in the colonization of bone by prostate cancer cells . Net MMP-9 activity in bone tissues peaked 2 weeks after injection of prostate cancer cells (PC-3), coinciding with a wave of osteoclast recruitment. In vitro, co-culture of PC3 cells with bone tissue led to activation of pro-MMP-9 and increased in secreted MMP-9 activity. Activation of pro-MMP-9 was prevented by metalloprotease inhibitors but not by inhibitors of other classes of proteases. The authors concluded that their data suggested that osteoclast-derived MMP-9 may represent a potential therapeutic target in bone metastasis . MMP-2 has been implicated in initiation of metastases, that is in promoting the movement of tumour cells from the primary lesion into the lymphatic or circulatory systems , but does not appear to have a role in bone metasteses colonisation . Recently inhibition of MMP-2 by genistin, a form of dietary soy, via MAPK and TGF-beta, was demonstrated. This study suggested a physiological role for genestin and confirmed epidemiological studies which demonstrated that dietary intake of genisten was associated with lower rates of metastatic prostate cancer .

4. Urokinase Plasminogen activator The plasminogen plasmin proteolytic cascade is a multi-functional pathway which facilitates a spectrum of biological processes including ECM remodelling during wound healing and tumour invasion, and metastasis. The urokinase-type plasminogen activator (uPA) and its receptor (uPAR), initiate this cascade by converting plasminogen to plasmin. Plasmin subsequently degrades a range of ECM components and activates MMPs (see above) (95-98). Over expression of uPA or uPAR is a feature of a number of malignancies, including prostate, and is correlated with tumour progression and metastasis. In contrast, inhibition of expression of uPA or uPAR or inhibition of uPA and uPAR interaction leads to a reduction in the invasive and metastatic capacity of many tumors (96, 98-100).

5. Kallikreins Kallikreins (KLKs) are highly conserved serine proteases that play key roles in a variety of physiological and pathological processes . Possibly the best known KLK with respect to prostate cancer is KLK3 or PSA. Recently, two other androgen regulated KLK’s have been shown to have altered expression in PCas, KLK2 and KLK4 . The protein product of KLK2, (glandular kallikrein; hGK2) is secreted in ejaculate, and like PSA has been shown to have a number of substrates (105). In particular, KLK2 can proteolyse several IGFB’s (IGFB-2, -3, -4 and -5) more efficiently and at lower concentrations than PSA, and potentially may stimulate the IGF axis. KLK4 may also play a role in the IGF axis by activating pro-PSA/KLK3 which induces IGF activity by cleaving IGFB-3. PSA (KLK3) also cleaves latent pro-transforming growth factor-beta (TGF-beta) to active TGF, thereby regulating prostatic cell growth and bone homeostatis. In addition, it has also been suggested that PSA also plays a role in degradation of laminin (the ECM glycoprotein), stimulating cell invasion through the ECM. Both KLK2 and KLK4 can activate pro-uPA to active uPA, initiating the plasminogen plasmin proteolytic cascade (see above). KLK2, can also inactivate plasminogen activator inhibitor-1, repressing regulatory control of the plasminogen plasmin proteolytic cascade. (105) Overexpression studies have shown that KLK4 induces transcriptional repression of E-cadherin, with associated increase in vimentin . Veveris-Lowe et al (2005) (106) concluded that the loss of E-cadherin and associated increase in vimentin are indicative 12

Chapter 10. PROSTATE CANCER of EMT and that KLK4, may have a functional role in the progression of prostate cancer through promotion of tumour cell migration.

C. Molecular Markers of Prostate Cancer.
One of the challenges in diagnosing prostate cancer is the heterogeneity of the disease phenotypically. As described above, alterations occur in cancer cells at many levels, both pre and post transcriptionally (in the genome and transcribed RNA), and post translational protein modifications (such as phosphorylations). The quest for reliable prostate cancer markers to identify cancer cells in blood, bone marrow, urine, prostatic tissue itself and semen, has embraced a number of candidate biomarkers in addition to PSA. Many methods have been used to identify potential markers for prostate cancer, including analysis of gene expression (microarrays and quantitative PCR) and protein marker (histopathology, Mass-spectrophotometry) changes between disease and non-disease states. Interestingly it has been noted that the genes, their transcription rates and subsequent proteins altered in prostate cancer can be grouped into a number of cellular pathways. These include cell adhesion, cell-cycle regulation, cell signaling, angiogenesis and apoptosis. Some of these proteins and genes have been discussed above, and therefore only their applicability as diagnostic/prognostic or therapeutic targets will be discussed in this section. Table 2. Summary of molecular aberrations in prostate Cancer (Modified from Quinn et al., European Journal of Cancer 41 (2005) 858–887 [107]) Process Apoptosis Key molecules/markers p53, Bcl-2, Clusterin References Downing SR, 2003 Augustin, 2003; Rubio, 2005; Scaltriti, 2004a; Scaltriti, 2004b Chung, 2005; Williams, 2005; Yang; 2004 Edwards, 2005a and b Quinn, 2005; Epstein 2005; Verambally, 2002 Quinn et al, 2005; Burger, 2002; Haese A, 2005; Kurek R, 2004; Lintula S, 2005; Steuber T, 2005; Stephenson et al, 1999; Ang, 2004 Quinn et al, 2005; Burger, 2002; Chang, 2004 Bussemakers, 1999; Landers, 2005; ; Schalken et al., 2005; Kumar-Sinha, 2004; Jiang, 2004; Srikantan V, 2002; ; Zhigang, 2005; 13

Signal transduction

TGF-beta, KGF, EGF, Caveolins AR c-Myc, p16INK4A, p27KIP1, pRb, apoptotic index, Ki67 p53, EZH2 E-cadherin, alpha-catenin, delta-catenin, metalloproteinases, kallikreins, CD151

Cell cycle regulation

Cell adhesion and cohesion

Angiogenesis Other Molecular Markers

VEGF, VEGF receptors, nitric oxide, PSMA AMACR; DD3/PCA3; pCGEM-1; Hepsin; PSCA

Chapter 10. PROSTATE CANCER 1. Cell Signalling markers As previously discussed, cell-cell signaling is mediated by a number of soluble and insoluble proteins and factors. These include growth factors, such as TGF-beta, EGF and IGF, and cytokines such as Il-6. Several of these factors have utility as biomarkers, such as TGF-beta and, as mentioned above, inhibitors to several of these compounds have been developed (72). Caveolins are major structural proteins of Caveolae, specialized plasma membrane invaginations that are abundant in smooth muscle cells, adipocytes, and endothelium, and act as regulators of signal transduction (108). In a number of tumour models (including prostate) caveolin-1 has been implicated in oncogenic cell transformation and subsequent metastasis. Studies on knock-out mice and in breast cancer models, indicate that caveolin-1 normally functions as a negative regulator of cell transformation and tumorigenesis. However, in prostate cancer caveolin-1 may function as a tumour promoter, potentially via both genetic and post-translational modifications (108). In addition, a recent study showed that c-Myc and caveolin-1 immunopositivity correlated positively with Gleason score (P = 0.0253) and positive surgical margin (P = 0.0006). Yang (109) found that the combination of positive c-Myc and caveolin-1 in patients with clinically confined prostate carcinoma was a significant prognostic marker for disease progression after surgery. As described above, changes in the androgen receptor have been long associated with prostate cancer development. While alterations in the function of this gene appear to be critical for cancer progression, these alterations occur at many levels, both in the genome (such as point mutations) and also post transcriptional and translational (ie at the RNA and protein levels (62, 63). Therefore using detection of any one of these changes as a screening tool for prostate cancer has not yet been shown to be practicable.

2. Cell Adhesion markers Aberrant expression of cell-cell adhesion molecules (C-CAMs) is often associated with the development of tumours. Decreased expression of many C-CAMs including E-cadherin, have been associated with the progression of prostate cancer and several other types of neoplasm (107). While CD44, another cell adhesion molecule, has been associated with prostate cancer, there have been inconsistencies across the few studies as to the prognostic value of this marker . While the loss of these molecules may be useful in immuno-histopathologically- based diagnoses, molecular based profiling may depend more on overexpressed molecules. Therefore, the activity of MMP’s 2 and 9, uPA and uPAR, and KLK2, 3 (PSA) and 4 (see above) in interfering with cell adhesion, may prove to be more relevant Both KLK2 mRNA and its protein product, hK2, have been examined for efficacy as stand alone biomarkers or in combinations with other known biomarkers such as PSA . Thus far none of the published studies has found KLK2 to be useful prognostically and it appears to have limited diagnostic value; only one study has found that KLK2 (measured by quantitative PCR) in combination with PSA distinguishes cancers from BPH . A recent study from China found that a functional C748T polymorphism in KLK2 may be associated with increased risk for developing prostate cancer. The frequency of the CC, CT and TT genotypes was 65.7%, 32.7% and 1.6% in patients with prostate cancer and 56.0%, 37.5% and 6.5%, respectively, in controls (p = 0.010). Therefore, C allele carriers (CC and CT genotypes) were at significantly higher risk for prostate cancer than TT homozygous subjects (p = 0.002) . As noted previously, KLK4, has also been associated with prostate cancer (114). The KLK4 protein product, hK4, is the first member of the KLK family that is intracellularly localized , and KLK4 expression is regulated by androgens, oestrogen and progesterone in prostate cancer cells. In situ hybridization on normal and hyperplastic 14

Chapter 10. PROSTATE CANCER prostate samples indicated that KLK4 is predominantly expressed in the basal cells of the normal prostate gland and overexpressed in prostate cancer . Some cell adhesion associated molecules such as CD151, a member of the tetraspanin family which interacts with integrins (see above), may prove to be relevant prognostically. CD151 plays a role as a link between extracellular matrix and intracellular structures, and increased protein levels of CD151 in well and moderately differentiated prostate cancers correlate with disease relapse subsequent to radical prostatectomy (115)

3. Markers of apoptosis Apoptosis or programmed cell death is part of cell growth and cycling in normal and benign cells. In cancers, key regulators of apoptosis can be pro-apoptotic or antiapoptotic such as p53 and Bcl-2 respectively, show abnormal function and expression. While there are many genes and proteins which fall into this category, many are also cell cycle regulators (such as p21 and p16) and will be discussed below. The pro-apoptotic protein p53 regulates transcription of genes required for G1-phase growth arrest of cells in response to DNA damage. Mutant p53 protein accumulation in malignant cell nuclei has been shown to be a poor prognostic indicator in several human carcinomas including breast, lung and colorectal (107). Mutations in p53 have been shown to be a common event in early stage, organ-confined prostate cancer and the loss of p53 function via expression of viral or cellular oncoproteins also seems common (116). A number of studies have reported that p53 nuclear accumulation in ≥ 20% of tumour cells is adversely prognostic (117). Although using p53 as a diagnostic marker has been debated due to the heterogeneity of expression within tumours, Quinn et al (2000 & 2005) (107, 117), propose that p53 has great potential as a prognostic marker as metastatic, recurrent and androgen resistant cancers show higher number of cells with p53 immunoreactivity compared with primary tumours. Bcl-2 was initially identified as an apoptosis-inhibiting proto-oncogene in B-cell lymphomas. Its value as a diagnostic marker is unclear with some groups showing that only limited numbers of prostatic tumours express Bcl-2 (118, 119). However, other researchers have shown correlation with Bcl-2 and poor prognostic outcomes, with increased numbers of high-grade and metastatic tumours having Bcl-2 immunoreactivity . Also Bcl-2 overexpression in tumours has been associated with resistance to radiotherapy (123). Subsequently many scientists are looking at Bcl-2 inhibitors (such as transgenes, and antisense RNA oligonucleotides) as a means of sensitizing Bcl-2 expressing tumours to chemo and radiotherapies (125). Another study showed that testosterone-repressed prostate message-2 (TRPM-2), also known as clusterin or sulfated glycoprotein-2, was elevated following androgen withdrawal in both normal and malignant tissues . In prostate adenocarcinoma, TRPM-2/clusterin expression may be useful as both a diagnostic and prognostic marker, with increased TRPM-2/clusterin protein expression evident in prostate cancer (96%) compared with BPH (73%) and normal prostate epithelium (17%) . Pins et al (2004) (129) indicated that TRPM-2/clusterin immunoreactivity in stromal cells surrounding the tumour epithelium predicted PSA relapse but staining within the primary tumour epithelium was not prognostic. The anti-apoptotic function of TRPM-2/clusterin is well documented, both by overexpression studies and through the activities of TRPM-2/clusterin specific inhibitors (124). However, some studies have shown pro-apoptotic functions of TRPM-2/clusterin in PC-3 androgen-independent prostate cancer cells. Cells overexpressing an intracellular, non secreted form of TRPM-2/clusterin showed signal-independent nuclear localization of the protein - leading to G2-M phase blockade followed by caspase-dependent apoptosis (127, 131). While TRPM-2/clusterin is an attractive target for theraputics, caution is warranted as it seems to have a number of functions 15

Chapter 10. PROSTATE CANCER in cell cycle and apoptosis, with the nuclear form (nClu) being proapoptotic while the secreted form (sClu) has prosurvival effects (131).

4. Cell cycle Regulators as markers Genetic aberrations in the control of progression in the cell cycle are present in most human cancers (107). There are many molecules involved in this process whose expression is altered in prostate cancer, including the cyclin family, RBp (retinoblastoma protein), p16, p21, p27, p53, Smad4, FHIT, and PTEN/MMAC1. Increased expression of p16 and p21 has been associated with poorer prognostic outcomes, and while loss of RBp and p27 may also have some prognostic value, further studies are warranted (107, 110). Strong expression of EZH2, a catalytic subunit of the polycomb repressor complex 2, in clinically localized prostate cancer is related to poor prognosis . EZH2 is also overexpressed in hormone refractory prostate cancers (133). Regulation of EZH2 is controlled by the E2F3 transcription factor and recently Foster (2004) (134) showed that nuclear expression of E2F3 in 20% or more of prostate epithelial cells is also an indicator of an unfavourable clinical outcome.

5. Markers of Angiogenesis As discussed above, angiogenesis or blood vessel growth is an essential factor in cancer growth and progression (135). A key component in angiogenesis, VEGF, is highly expressed in most prostate cancers and has value prognostically (107). A number of VEGF and VEGF-R antibodies and peptide antagonists have been developed with the specific goal of targeting the neovasculature and growing cancer cells. This targeting pathway may become highly important since hypoxia is a known important factor that induces VEGF production . Another molecule whose expression has been associated with angiogenesis is PSMA (prostate-specific membrane antigen) which is expressed by both tumor epithelium and tumour associated endothelial cells and neovasculature (137). A number of researchers, (138-140)have made a series of antibodies against the external domain of this protein, which have been used both pre-clinically and clinically for the diagnosis and therapy of prostate cancer. Chromosomal localization of PSMA gene, however, is controversial. It has been mapped to two regions, chromosome 11p11-12 and 11q14 and it has been proposed that a PSMA-like gene may exist in one of the two chromosomal regions through the process of gene duplication. Our research group has shown that a specific PSMA transcript has applicability as a biomarker for prostate cancer, particularly when used in combination with other gene transcripts (such as DD3/PCA3) (143, 144). PSMA may function as a ligand internalising receptor, an enzyme playing a role in nutrient uptake, and a peptidase involved in signal transduction in prostate epithelial cells (145).

6 Other molecular markers A number of other gene transcripts have been identified which may have utility as diagnostic and prognostic markers that do not fit into the catergories discussed above. These include Hepsin a transmembrane serine protease, AMACR, (alpha16

Chapter 10. PROSTATE CANCER methylacyl-CoA racemase), PSCA (prostate Stem cell antigen) and the non-coding RNA’s DD3 and PCGEM-1. Hepsin, a type II transmembrane serine protease is differentially expressed in prostate cancer compared with normal and BPH affected prostate tissue (144, 146). Interestingly, in-vitro studies in prostate cell lines found that Hepsin overexpression had growth inhibitory effects . More recently a model for Hepsin and tumour progression was found when the soluble form of Hepsin was found to activate HGF (hepatocyte Growth Factor) . Potentially, antagonists to Hepsin and subsequently HGF activation could be useful therapeutically. Alpha-methylacyl-CoA racemase (AMACR) overexpression in prostate and other cancer tissues has been well characterized at both the mRNA and protein levels (151-153). This enzyme functions in the peroxisomal beta oxidation of branched-chain fatty acid molecules and has been implicated in the link between high meat high fat diets and the increased incidence of prostate cancer observed by many epidemiological studies (154). AMACR immunohistochemistry is being used in conjunction with normal Haemotoxylin and Eosin staining to aid in the diagnosis of prostate cancer histologically (128, 155, 156). As with many of the markers already mentioned clinicians are finding that using AMACR in conjunction with other prostatic markers gives better diagnostic results than AMACR alone (J153, 157). Molinie (2004)(157) showed that basal cells of normal prostatic glands stained with p63 in 100% of cases, while carcinomas had a p63-/AMACR+ profile, PIN were p63+/AMACR+, and benign lesions were p63+/AMACR-. Recently a number of splice variants of the AMACR transcript have been described which may be relevant for strategies targeting AMACR expression such as RNA antisense oligonucleotides (158). PSCA was identified as a cell surface antigen expressed by prostate cancer cells and is regulated by the androgen receptor ). It is central to the development of the prostate gland and could provide a new diagnostic and therapeutic target for PCa (161). PSCA overexpression has more recently been shown in pancreatic and urothelial tumor models (162-164) and has been proposed as a target antigen with immuno-based therapeutics (165, 166). DD3/PCA3 was identified by differential display technology in 1999 as a non-coding RNA highly specific to prostate cancer . Subsequently a number of researchers have confirmed the over-expression of DD3 (144,168) in a number of different cohorts. More recently the uPM3 (Bostwick) test has utilized this detection of this RNA in a PCR based assay in urines of prostate cancer patients. Two independent studies showed significant improvements in detecting cancer compared with the use of PSA alone (169, 170). Tinzl (2004)(169) reported 82% sensitivity, 76% specificity for the uPM3 assay compared to 98% sensitivity, 5% specificity, for tPSA (at a cutoff of 2.5 ng/ml). In the tPSA categories <4, 4-10 and >10 ng/ml sensitivity was 73%, 84% and 84% and specificity was 61%, 80% and 70%, respectively (169). The Canadian-based study (170) sampled 517 patients undergoing biopsy at five centres, for which 86% had an assessable sample. The overall uPM3 sensitivity and specificity in this sample group was 66% and 89%, respectively. Once again, in the tPSA categories <4, 4-10 and >10 ng/ml, the sensitivity of the uPM3 assay was 74%, 58% and 79% with specificity of 91%, 91% and 80% respectively. The positive predictive value of uPM3 was 75% compared with 38% for total PSA. (170). PCGEM1, also a putative non-coding RNA was also identified by differential display analysis of paired normal and prostate cancer tissues, with subsequent Northern blot analysis of tissues showing that PCGEM1 was expressed exclusively in the human prostate. In-situ analysis showed tumor associated overexpression in 84% of prostate cancer patients while reverse transcription PCR assays revealed tumour-associated overexpression in 56% of patients . Interestingly PCGEM1 over-expression has been shown to be significantly higher in prostate cancer cells of African-American men than in Caucasian-American men (P=0.0002). In addition, ‘normal’ prostate epithelial cells from prostate cancer patients with a family history of prostate cancer also displayed increased PCGEM1 expression (P=0.0400). A physiological role for PCGEM1 17

Chapter 10. PROSTATE CANCER in cell growth regulation has been suggested, with cells transfected with PCGEM1 displaying cell proliferation and an increase in colony formation .

D Future Directions
Recently the focus of some studies has changed from identification of individual markers to utilising combinations of known prostate cancer-specific markers as predictors of disease recurrence after treatment with curative intent (172, 173). In our laboratory we have investigated the utility of using combinations of biomarkers in a PCR based diagnostic assay for prostate cancer. We found that using a combination of Hepsin, DD3 and PSMA allowed us to distinguish 100% of prostate tumour from BPH tissues (144). As with diagnosis, it has been proposed that a number of genes can be used to build a “fingerprint” of an aggressive tumour . The challenge remaining is how to apply this information in biologically relevant samples such as serum, ejaculate and urine sediments. With the development of methods that allow large scale gene expression profiling, such as microarrays and quantitative RT-PCR, the list of genetic alterations in prostate cancer cells has increased dramatically. Now the same technology is being applied to proteins (proteomics). These advances in technology should allow the scientist to determine that genetic changes translate into the proteome and identify post-translational modifications which are biologically significant in cancers. Although this chapter has attempted to summarise the current literature, this area of science is ever-expanding and subsequently there are many other biomarkers for prostate canceer, and other cancers which may be relevant in PCa, that were not discussed. Over the past decade there has been an explosion of research into the basic science of prostate cancer. Consequently, we are developing new paradigms for understanding the natural history of the disease as well as creating novel approaches to therapy. Later in this chapter, some of these advances will be discussed in further detail. Currently, surgery and radiation therapy are the conventionally accepted primary treatment modalities for localized prostate cancer. However, with the advent of new molecular-based therapies, combination therapy using surgery and/or radiation with these novel agents has been proposed as a method of improving therapeutic outcomes. More importantly, molecular-based therapies are being studied in phase I, II, and II trials in men with advanced and hormone-refractory prostate cancer, for which no curative therapies currently exist.

V. PREVENTION OF PROSTATE CANCER
Introduction
Predisposition to any cancer can be considered in terms of environmental and extrinsic factors, such as foods and interventions, and intrinsic or inherent susceptibility related to the person’s genes apportioned at conception. For prostate cancer in particular, many people hold strong opinions regarding various strategies to avoid contracting it and for treatment of the condition, very often without any clear evidence of beneficial effect. It should be borne in mind that many advocated supplements and unconventional therapies are expensive and, although some ultimately may be proven to be of benefit, many will not and some may be harmful.

18

Chapter 10. PROSTATE CANCER

Dietary & Natural Therapies
(i) Dietary fat and animal meat
A body of epidemiological data links high meat consumption and a diet rich in saturated fats with an increased risk of prostate cancer. (175-177). As stated in the previous section, alpha-methylacyl-CoA racemase (AMCR), an enzyme involved in oxidation of fatty acid molecules, has been implicated in the link between high fat diets and prostate cancer (154). Fatty acids in dietary fats can be divided into essential and non-essential with the former consisting of omega-6 (linoleic acid derived) and omega-3 fatty acids (which are linolenic acid in origin). Current evidence suggests that omega-3 fatty acids offer a protective role. (178, 179). Mutagens in cooked meats and fat, which include heterocyclic amines and polycyclic aromatic hydrocarbons, are known to be activated by cytochrome p450 and N-acetyl transferase enzymes in the prostate to produce an increase in markers of dietary stress. (180). These, in turn, generate reactive oxygen species (ROS) leading to mutation of DNA base pairs and oncogene expression. Another mode of action invoked is degradation of omega-6 fatty acids via the arachidonic acid pathway production of prostaglandin and lipoxygenase products which, in addition to resulting in inflammation, are known to decrease apoptosis, stimulate proliferation and induce angiogenesis. (181-184).

(ii) Lycopene
The relevance of the anti-oxidant lycopene, an isomer of β -carotene, has been reported in relation to various cancers, including prostate, in a number of epidemiological studies. Lycopene, which is the carotenoid responsible for the red colour in tomatoes is thought to have a protective effect as an inverse relationship has been observed between tomato intake or blood lycopene level and the risk of cancer. (180, 185, 186)

(iii) Selenium & vitamin E
The essential micronutrient selenium is present in water and food, especially seafood, meats and Brazil nuts. In countries with selenium-poor soils such as New Zealand, locally produced foodstuffs may be low in selenium content. A number of epidemiological studies have reported an increased risk of prostate and other cancers with selenium deficiency and, conversely, a reduced incidence of prostate cancer with selenium supplementation. (187, 188). Selenium is a component of many proteins, such as glutathione peroxidase which is a key enzyme in surveillance against oxidative stress, involved in maintenance of genomic stability. (189). In addition to an effect in reducing tumour initiation, selenium is considered to have an inhibitory effect in relation to progression by blocking cellcycle arrest through down-regulation of CDK1, CDK2 and cyclin A and activation of p19/INK4d and p21/WAF1. (190). Gianduzzo et al (2003) (191) demonstrated in a randomised, controlled trial that oral selenium supplementation resulted in significantly higher levels of this compound in prostatic tissue removed at transurethral resection of the prostate (TURP). Importantly, these investigators also found that blood and prostate levels correlated poorly, suggesting that peripheral blood measurements are a poor indicator of prostatic selenium content (191). Vitamin E is present in a wide range of foods which include vegetables, vegetable oils, nuts and egg yolks. Through its most active form, α-tocopherol, vitamin E causes G1 cell-cycle arrest via cell regulatory proteins D1, D3 and E and cdk2 and cdk4. (190) In the Alpha-Tocopherol, Beta-Carotene study in the United States, a one-third reduction in prostate cancer incidence and a 40% reduction in prostate cancer deaths 19

Chapter 10. PROSTATE CANCER were reported in men taking vitamin E (The Alpha-Tocopherol, Beta Carotene Cancer Prevention Study Group, 1994) Currently, 2 large prevention studies are in progress to evaluate selenium and vitamin E supplementation. SELECT (Selenium and Vitamin E Cancer Prevention Trial) is American based and APPOSE (Australian Prostate cancer Prevention Trial Using Selenium) are expected to produce results in the next few years (180, 192).

(iv) Soy products/isoflavones/green tea
Dietary differences between people living in Asian countries and those in the West correlate with marked differences in prostate cancer incidence. However, with adoption of western dietary patterns, there is an increase incidence of diseases of affluence, such as diabetes mellitus, breast, colorectal and prostate cancer (193). These traditional differences have been attributed to the high content of soy, tofu and isoflavinoids in many traditional asian diets. (194, 195). The major isoflavone components of soy, including genistein, daidzein and their metabolites, have been shown to inhibit benign and malignant prostatic epithelial cell growth, down-regulate androgen-related genes and reduce tumour growth in some animal models (190). Huang et al (2005) (94) reported inhibition of MMP-2 by genistein via MAPK and TGF-beta. This study suggested a physiological role for genestin and confirming epidemiological studies which demonstrated that dietary intake of genisten was associated with lower rates of metastatic prostate cancer . deVere-White et al (2005)(196) very recently presented their experience with GCP, a proprietary nutritional supplement, the major constituents of which are genestine and diadzin. The preparation functions as a tyrosine kinase inhibitor and facilitates entry of prostate cancer cells into apoptosis. For13 men on active surveillance for their prostate cancer, six months of GCP resulted in a lowering of PSA; the trial is ongoing (196). Dalais et al (2004)(197) studied 29 men diagnosed with prostate cancer and scheduled to undergo RRP who were randomized to one of three groups; soy (high phytoestrogen), soy and linseed (high phytoestrogen), or wheat (low phytoestrogen) in the form of specially prepared breads. Their findings indicated that a daily diet containing four slices of a bread rich in soy grits favourably influenced the PSA level and the free/total PSA ratio in patients with prostate cancer, providing supportive evidence for epidemiological studies claiming that men who consume high phytoestrogen diets have a reduced risk of prostate cancer development and progression (197). Schröder et al (2005) (198) reported on 49 patients with rising serum PSA levels after radical prostatectomy (n = 34) or radiotherapy (n = 15) who participated in a randomised, double-blind, placebo-controlled crossover study of a dietary supplement of soy, isoflavones, lycopene, silymarin and antioxidants as main ingredients. The soy-based dietary supplement delayed PSA progression in a significant fashion with a 2.6 fold increase in the PSA doubling time from 445 to 1150 days for the supplement and placebo periods (198). An integral part of most asian diets, green tea has had many benefits accorded to it, often through association with other dietary components. In terms of its affect on inhibiting prostate cancer development, it is thought to provide a benefit through the presence of polyphenols (also present in red wine) which induce apoptosis and inhibit proliferation via WAF1/p21 cell- cycle effects (190)

Pharmacological
(i) COX-2 Inhibition
Continuous use of Nonsteroidal anti-inflammatory drugs (NSAIDs), in particular as20

Chapter 10. PROSTATE CANCER pirin, has been reported to reduce the likelihood of development of several malignancies, including prostate cancer (199). Indeed, aspirin and other NSAIDs have been considered to decrease the risk of developing prostate cancer by 24-66% (200-2002) Non-selective NSAIDS such as aspirin inhibit both cyclooxygenase (COX-1 & COX-2) enzymes with the COX-2 isoform believed to be the relevant one in terms of prostate cancer development. COX-2 converts arachidonic acid to prostaglandin so both nonspecific and specific COX-2 inhibitors act at this site, although some specific COX-2 inhibitors may have other actions as well. Specific COX-2 inhibition by celecoxib and nimesulide has been reported to reduce expression of several androgen-inducible genes, repress androgen-receptor mediated activation of PSA and hK2 promoter activity and repress androgen receptor protein expression. (203). Cost implications aside, the recent reporting of previously unappreciated serious side-effects with two prominent COX-2 inhibitors, rofecoxib and celecoxib, have resulted in some reluctance to change from using NSAIDS to selective COX-2 inhibitors for prostate cancer prophylaxis.

(ii) Finasteride
Based on the observations that androgens are necessary for the development of prostate cancer and that men with a congenital deficiency of the 5-α-reductase type 2 enzyme do not develop prostate cancer, a 7-year randomised, controlled trial with the drug finasteride was undertaken with 18 882 men who were >55 years, had a normal prostate on digital rectal examination and a PSA of <3 ng/ml. The study was ended 15 months prematurely because the end-point had been reached and continuing the trial would not have changed the outcome. Men who received finasteride, which inhibits conversion of testosterone to dihydrotestosterone (DHT) by targeting the 5-α-reductase type 2 enzyme, had a prostate cancer prevalence reduction of 24.8% (24.4% to 18.4%. However, the prevalence of Gleason 7-10 tumours was higher in the finasteride arm (6.4% versus 5.1%) although 98% of the tumours were clinically localised. (204). Klein et al (2005) (205) have questioned the validity of the conclusion in relation to the rate of clinically-significant prostate cancer detection in this trial. They present a model of risk and benefit that estimates the potential influence of histological artefact (due to finasteride-induced effect on prostatic epithelial appearances) in the assignment of excess risk for high-grade disease and possible overdetection bias introduced by finasteride-induced volume reduction in prostates for the treated patients (205). A further industry-sponsored study, the REDUCE (Reduction by DUtasteride in prostate cancer Events) trial is in progress. In this randomised, controlled study involving 8 000 men with PSA values between 2 and 10 ng/ml dutasteride, which inhibits both isoforms of the 5-α-reductase enzyme, is being used in the treatment arm. All patients in this trial are being biopsied at least twice during the study unlike the finasteride study in which prostatic biopsies were recommended if the annual PSA (adjusted for the effect of finasteride) exceeded 4 ng/ml or if the digital rectal examination was abnormal. Recent research suggests that inhibition of the irreversible action of 5α-reductase to convert testosterone to the more transcriptionally active dihydrotestosterone may have untoward effects in relation to prostate cancer. Dihydrotestosterone in turn is hydroxylated to 3α-diol and 3β -Adiol which do not bind to the androgen receptor but have a strong affinity for oestrogen receptors, the result of which is thought to have a direct effect on prostate development and homeostasis (206). The binding of 3β -Adiol to oestrogen receptor beta (ERβ ) induces expression of the cell adhesion molecule E-cadherin, loss of which is associated with a more aggressive phenotype in prostate cancer cells (38).

21

Chapter 10. PROSTATE CANCER

Conclusion
The genesis of prostate cancer is multifactorial, with changes arising over a considerable period of time. Consequently, with respect to preventative or ameliorating measures, it is axiomatic that these should be implemented early and sustained for longer rather than shorter duration if their beneficial effects are to be maximised.

VI. DIAGNOSING PROSTATE CANCER
A. Introduction
Indisputably, early detection and treatment of prostate cancer are more effective than ever before with a consistent decline in mortality rates in many Western countries (2). However, this statement does not accomodate the fact that many men continue to die with prostate cancer and that the consequences of this tumour and its treatments may contribute significantly to the demise of many of these men, both directly and indirectly. Albertsen et al (207, 208) estimated that, for conservative treatment, men with moderately and poorly differentiated tumours lost approximately 4-5 years and 6-8 years of life, respectively, compared with those with well-differentiated tumours whose life-expectancies were little different to age-matched controls (207, 208). This problem of identifying patients with the more aggressive cancers is relevant to men of all ages, consistent with the findings of Parker et al (2001) (209) who concluded in their meta-analysis that age is not a significant prognostic factor in contemporary clinical practice (209). Using the SEER database, Lu-Yao and Yao (1997) (210) found that the disease-specific 5 year survival for poorly differentiated tumours was 63-69% (210). Albertsen et al (2005) (211), reporting on 767 men diagnosed between 1971 and 1984 and treated without curative intent, found that, for low-grade prostate cancers (Gleason 2-4), men have a minimal risk of dying from prostate cancer during 20 years of follow-up. Those with Gleason scores of 5 or 6 have an intermediate risk of prostate cancer being the cause for their demise but patients with high-grade prostate cancers (Gleason 8-10) have a high probability of dying from prostate cancer within 10 years of diagnosis (211). The zeal to diagnose and treat must be tempered by realisation that neither detection nor therapeutic intervention may be in the best interests of many men. Establishing the right balance between the benefits and disadvantages of diagnosis and treatment is problematical and, consistent with the variation in the natural history of prostate cancer, it is not difficult for the unwanted effects of intervention to be worse than the disease being treated. Of necessity, it is important to recognise that the risks of competing mortalities are very different for men of different ages. Based on Queensland data, Baade et al (2005) estimated that about 60% of men diagnosed in their 50s could be expected to die a premature death (ie, before reaching 80 years of age) from prostate cancer compared with corresponding rates of 50% and 38% for men diagnosed at 60 and 70 years. Thus, for a 50 year-old man, a diagnosis of prostate cancer is more likely to result in a premature death from prostate cancer than for men diagnosed in their 70s (212). Hence, whether or not to diagnose the condition should be a carefully considered decision.

B. Making Decisions about whether or not to diagnose Prostate cancer
While prostate cancer is the most common male malignancy in the developed world and the second most common cause of cancer deaths, uncertainties remain about management practices at several points in the illness continuum. For example, owing to a lack of definitive data confirming that widespread screening for prostate cancer will reduce the death rate from this disease, population-based screening for 22

Chapter 10. PROSTATE CANCER prostate cancer in asymptomatic men is not currently recommended in most countries . Rather, it is suggested that men should be able to access PSA testing as long as they are fully informed of the pros and cons of testing. For those diagnosed with localised prostate cancer, further decisions present with three possible treatment options at a minimum: watchful waiting, radiation therapy, or RRP - extending more recently to include in some settings brachytherapy and laparoscopic and robotic surgery . Men who are diagnosed with advanced disease will also face difficult treatment decisions such as when to commence treatment and what method of hormonal ablation to select, each with various quality of life ‘tradeoffs’, to accept . In the setting in which no one treatment approach is clearly superior with regards to cancer cure and where quality of life outcomes differ markedly, the quality of patients’ decision making about medical treatments is critical. As a result, strategies to assist in meaningfully considering prostate cancer treatment options, and the risks and benefits of these options in order to achieve high quality patient decisions, are essential . The approach that is considered to be optimal for achieving high quality patient decisions is shared decision making . Shared decision making is defined as a process carried out between a patient and his health care professional where both parties share information and the patient understands the risks and benefits of each treatment option, participates in the decision to the extent that he desires and makes a decision consistent with his preferences and values, or defers the decision to another time. Shared decision making may not be easy to achieve for all patients . For example, although many patients with cancer indicate a preference for sharing decision making with their clinicians, some, in the case of prostate cancer between 8% to 58% of men, prefer a passive decision making role where clinicians make treatment decisions on their behalf . However, clinicians still need to understand patients’ preferences to ensure that they are making quality decisions on behalf of their patients. As well, there is often a gap between the clinical ideal of shared decision making and actual clinical practice where decision complexity and time constraints may make this approach difficult for both parties to achieve . There are, however, defined strategies and decision aids that can facilitate this process .

Supporting Patient Choice about Testing for Prostate cancer
Many groups advocate an informed decision-making process as an evidence-based approach and necessary precursor to screening for early prostate cancer . Others have suggested that informed decision-making on this health topic is also necessary as a medico-legal risk management strategy . While some researchers have suggested a set of information that needs to be communicated to men about this health decision , there are few explicit guidelines on this subject . Problematically, patients and clinicians do not agree on core content . It has been advised that, for any screening test, patients need to understand the purpose of the test, the likelihood of false-negatives and false-positives, the uncertainties and risks associated with testing, significant medical, social or financial implications of testing and any possible sequelae and follow up care plans (www.ipdas.ohri.ca). Such information needs to be communicated to patients in a logical and balanced sequence in order to promote better understanding and increased decisional control by men. One approach that has been widely tested in primary care in Australia is the use of six decision steps (see Table 1). Each decision step logically follows to prompt the clinician to overview important health information, with tailoring suggested in Step 1 to ensure the discussion is consistent with the patient’s concerns. For example, for a man with a significant family history of prostate cancer, this factor is likely to be central to the patient discussion . Men who experience uncomplicated lower urinary tract symptoms (LUTS) often worry about prostate cancer, so addressing this concern first may be priority . In this regard, resources for patients that explain about male reproductive health problems such as urinary symptoms and sexual dysfunction are available at www.andrologyaustralia.org1. As well, National Health and 23

Chapter 10. PROSTATE CANCER Medical Research Council guidelines are available about the management of LUTS (http://www.health.gov.au/nhmrc/publications/synopses/cp42syn.htm). Table. Box 1: Six Decision Steps Six Decision Steps for Informed Choice about PSA Testing in Asymptomatic Men 1. Identify the patient’s main concern 2. Explain where the prostate is and tests available to detect prostate cancer 3. Discuss prostate cancer risk and risk factors 4. Explain the pros and cons of early detection of prostate cancer 5. Identify patient’s personal preferences 6. Support the patient’s choice, and if requested implement a prostate cancer risk management plan

Source: Steginga S, Pinnock C, Baade P. "The early detection of prostate cancer in general practice: supporting patient choice ", practice resource in “Supporting patients’ choice about PSA testing in general practice” A collaborative project of the Queensland Cancer Fund. Brisbane, 2005 From this point, checking to ensure the patient has a basic understanding of both the prostate and possible tests is needed and, given many men may be unaware of the location and function of the prostate gland, an anatomical diagram may be a useful teaching tool here. Next, a consideration of individual risk with regard to both the incidence and mortality of prostate cancer is needed. Communicating health risks effectively is a challenge in the provision of effective decision support. In general people find probabilities hard to understand, often estimate their level of risk incorrectly, and tend not to weigh up pros and cons in a systematic way when deciding about treatments . As well, population based statistics provide data about populations, not individuals, so risk communication needs to acknowledge this as a limitation and, where possible, refer to age-based risk estimates and relevant individual factors such as family history ). There are a number of communication strategies that have been suggested to help patients understand risk. These include 1. using numbers as well as words to explain risk 2. where possible providing the absolute risk or benefit 3. using frequencies rather than single event probabilities 4. using consistent denominators 5. putting the risk into context by comparing it to other life events 6. offering both the possible negative and positive outcomes to balance the message frame . However, a quality health decision goes beyond the simple transfer of information and includes consideration and incorporation of each patient’s values and personal preferences . Thus, Step 5 in Box 1 prompts the clinician to discuss each man’s individual preferences. A number of strategies can be used to do this, most commonly the use of a pros and cons exercise in which patients are encouraged to explicitly consider the factors that matter most to them personally in this decision, and the 24

Chapter 10. PROSTATE CANCER direction and leaning of their preferences either for or against each possible option. One approach to support this process for this health topic is the inclusion of a values table within a decision card (see Table 1). A decision aid that incorporates both the six decision steps and this values clarification exercise can be found on http://www.ncci.org.au/services/PSA_decision_card.pdf . Table 3. What is most important to you? FOR: Is this like you? AGAINST: Is this like you?

I’m concerned that I might get prostate cancer I want the best chance of finding it early, if I do get it I’m not interested in waiting for all the proof to be in

I think my chance of getting prostate cancer is low I am not convinced about the effectiveness of testing I am more concerned about avoiding treatment side effects, if there’s no guarantee I’d be reducing my risk of dying from prostate cancer

I want to do everything possible to reduce my risk of dying from prostate cancer Decision aids are also effective in supporting patients to make informed choices. With regards to PSA testing, patient focussed decision aids have been found to be effective in increasing men’s knowledge about PSA testing and decreasing decision-related distress , with a variable effect on actual testing behaviour. A range of aids is freely available from the web ( www.prostatehealth.org.au2; www.cdc.gov/cancer/prostate3; www.cancerbacup.org.uk4). Cancer helplines also often provide such information, for example, The Cancer Council Australia Cancer Helpline on 13 11 20; the UK helpline on 0808 800 1234; the USA Cancer Helpline on 1800 227 2345.

C. Digital Rectal Examination
Traditionally, palpation of the prostate by digital rectal examination (DRE) was the manner by which a diagnosis of prostate cancer was suspected. In historical series, up to 50% of palpable masses were attributable to prostate cancer (243, 244). Although DRE by itself is a poor method for diagnosing this malignancy (245, 246), it does still have an important diagnostic role as 25% of tumours are detected in men with normal PSA levels (247). Unfortunately, when a prostate cancer is diagnosed based on a palpable tumour, the risk of the patient already harbouring metastatic or locally advanced malignancy is considerable (248-250). However, a PSA-based prostate cancer detection strategy which omits DRE runs the low risk of missing some curable cancers (251).

D. Prostate-specific Antigen (PSA)
PSA testing has revolutionized the detection of prostate cancer and monitoring of its treatment. However, its application in early detection is contentious with attitudes to population-based screening for prostate cancer in asymptomatic men varying considerably. (252, 253) In the US, PSA screening is currently recommended for all men older than 50 years and advocated to begin at 45 years in those with first-degree relatives with prostate 25

Chapter 10. PROSTATE CANCER cancer and African-American men because of their higher risk of contracting this tumour.( 247, 254) Population screening is not currently recommended in most Western countries (213) although case selection is widely practised with peer organisations endorsing the importance of informed patient participation in decision-making for diagnostic testing, as outlined above.

Background to PSA as a screening test
The FDA initially approved PSA testing in 1986 for monitoring the disease status of prostate cancer patients and, subsequently in 1994, it was endorsed as a screening method for prostate cancer (255). However, the benefits for screening in improving survival are yet to be proven (256). Fitzpatrick (255) summarised the state of the evidence for screening recently. Although the results from the Quebec trial published in 1999 (16) and the experience in the Tyrol compared with the rest of Austria (3) are encouraging, the results of two large randomised trials are awaited with considerable interest. The Prostate, Lung, Colorectal and Ovarian Cancer Study (PLCO) (n = 74 000) in the US and the European Randomised Study for Screening for Prostate Cancer (ERSPCC) ( n = 239 000) from Europe are both on-going with results from the PLCO study due in 2006 and those from the ERSPC trial in 2008. Whether these 2 studies demonstrate a benefit for screening in reducing mortality, and from prostate cancer in particular, it is unlikely that early detection will disappear which, currently, remains PSA-based. The topic is reviewed in BJU Int, volume 95 supplement 3, 2005.

Strengths, limitations and adaptations of PSA
Although emergence of an abnormal PSA due to prostate cancer can precede the appearance of a palpable mass by as many as ten years (8, 10, 11), tumour-induced symptoms by 5-10 years and, on average, death by prostate cancer by 17 years (257), it must be remembered that PSA is not a test for prostate cancer (258, 259). An abnormal serum PSA merely indicates that something is likely to be awry in the prostate that includes cancer as one possibility. Contemporarily, the large majority of patients diagnosed with prostate cancer present with elevated serum PSA levels, mostly between 4-10 ng/ml, (260) in the absence of any other discernable abnormality. Although the prevalence of prostate cancer depends on the population examined, most men undergoing prostatic biopsies – usually the next step in elucidating the cause of an elevated PSA - have negative histology for cancer, even when >8 transrectal ultrasound (TRUS)-guided biopsies are taken,(261-263) with biopsies repeated if suspicion of an undetected malignancy is high.(260). In order to increase the likelihood of detecting prostate cancer earlier, a serum PSA ‘cut-off’ of 3 ng/ml has been advocated. However, Lodding et al (1998) (251) reported that approximately 15% of prostate cancers detected by investigating a serum PSA level between 3 and 4 ng/ml had extraprostatic growth (251). Furthermore, for a PSA threshold of 3 ng/ml, the negative biopsy rate is increased which Hessels et al (2004) (264) considered to be 70-80% (264). Thompson et al (2005) (265) for the Prostate Cancer Prevention Trial, reported on a randomised, prostpective study of 18, 882 healthy men >55 years, PSA levels <3.0 ng/ml and normal DREs followed for 7 years with annual DREs and serum PSA measurements conducted from 1993 to 2003 at 221 US centres. Biopsies were recommended for either a PSA >4.0 ng/ml or an abnormal DRE and for all participants at the end of 7 years. They concluded from their findings that there is no cutpoint of PSA with simultaneous high sensitivity and high specificity for monitoring healthy men for prostate cancer but rather a continuum of prostate cancer risk at all values of PSA (265). In their analysis of 3,446 consecutive volunteers with a PSA level of 2.0-10.0 ng/ml (free PSA 18% or less) screened in the Tyrol project who underwent 10 systematic prostate biopsies and an additional five Doppler-enhanced targeted biopsies on the 26

Chapter 10. PROSTATE CANCER basis of age-specific PSA reference ranges, Pelzer et al (2005) (266) focused on findings of those patients with PSA levels <4 ng/ml. They found that more than one third of prostate cancers were detected in 313 men with a PSA value of 2-3.9 ng/mL. Of these 313 patients, 24% had a Gleason score of 7 or greater compared with 33% of 560 patients with a PSA value of 4.0-10.0 ng/mL (P = 0.004). In addition, the prostate cancer cases with a low PSA level occurred in younger patients and at lower stages with smaller prostate volumes (266).Aus et al (2005) (267) also found similarly in their study of 5,855 men, 539 cases of whom had prostate cancer detected after a median follow-up of 7.6 years. However, based on their finding that there was not a single case of prostate cancer detected within 3 years in 2950 men (50.4% of the screened population) with an initial PSA level <1 ng/ml, they concluded that retesting intervals should be individualized on the basis of the PSA level, and that the large group of men with PSA levels of <1 ng/ml can be safely scheduled for a 3-year testing interval (267). On the basis of an evaluation of their data from the Rotterdam section of the ERSPC, Roobol et al concur that patients with a PSA of <1 ng/ml are at low risk of developing prostate cancer. They concluded that a strategy of PSA screening every 8 years for men with a PSA level of <1.0 ng/ml would result in a minimal risk of missing an aggressive cancer at a curable stage (268). Table 4a. Age-based Ranges for PSA Age range 40-49 50-59 60-69 70-79 50th percentile median 0.65 0.85 1.39 1.64 95th percentile upper limit of normal 2.0 3.0 4.0 5.0

Between 50th & 95th percentile, higher long-term risk of cancer PSA increases at ~3.3% pa – if rate of increase is greater, the risk of cancer is greater References: 11, 269-271

Table 4b. Non-prostate cancer contributors to increases in PSA

1. Benign prostatic hyperplasia (BPH) 2. Ejaculation (both free & total) up to 48 hours 3. Bacterial infection of prostate 4. Prostatic massage 5. Instrumentation (including catheterisation) of prostatic urethra 6. Prostatic biopsy #Finasteride lowers PSA levels by ~50% http://ncci.org.au/services/prostate_GPresources.htm

As indicated in table 4(b), non-malignant causes can produce an elevated PSA (>4.0 27

Chapter 10. PROSTATE CANCER ng/ml): these include infection, benign prostatic hyperplasia [BPH] and ageing (endorsing age-based reference ranges) (272). Instrumentation of the prostate and urinary tract can also raise PSA levels (273). Certain drugs, such as finasteride can lower PSA values by approximately 50% (274, 275). Physical examination and symptomatology can help differentiate BPH and prostatitis from cancer. The combination of a serum PSA test and digital rectal exam remains the most sensitive combination for diagnosing a prostatic malignancy (245, 247). Because of these limitations, adaptations to enhance the diagnostic utility of PSA for diagnosing prostate cancer have been advocated. It is important to recognise that these serve only as a guide in helping to decide whether or not to proceed to biopsies, with limited utility for extrapolation to individual patients. 1. Free/total PSA This test measures the percentage of free (or unbound) PSA in the blood, and compares it with the percentage bound to proteins (α1 antichymotrypsin and α2 macroglobulin) with its application most useful in younger men, as alluded to, above. In prostate cancer, most of the PSA in blood is bound so the lower the ratio of free to total PSA or the percentage of free PSA, the higher the likelihood that the patient has prostate cancer. The proportion of free PSA in seminal fluid is much higher than in serum, consistent with its physiological role in liquefaction (276). Levels of free-PSA but not complex-PSA in blood significantly correlate with PSA in semen in young men, with blood levels of complex-PSA, but not free-PSA, increasing with age (277). The free/total PSA test can help to discriminate between patients with indeterminate PSA levels (4-10.0 ng/ml) who are at the greatest risk of having prostate cancer, in particular aggressive disease (257, 278).

2. PSA velocity PSA velocity measures the speed at which a series of PSA values increases in value over a period of time (279). Any change in PSA >0.75 ng/ml in a year, is concerning for cancer although bacterial prostatic infection may be responsible for this degree of change.

3. PSA density PSA density is a measure of the concentration of PSA in a man’s prostate. It compares the value of his PSA and the size of his prostate (280). Most neoplastic prostate glands produce higher serum PSA levels than do non-malignant glands. Consequently, a serum PSA of 5.0 ng/ml in a patient with a 20 gram prostate is more worrisome for cancer than that a PSA of 5.0 ng/ml in a man with a 60 gram prostate, especially if there is a predominance of transitional zone tissue (BPH) in the latter. To determine the PSA density, a PSA level is obtained and is divided by the volume of the prostate, as estimated by TRUS. A value >0.15 ng/ml per gram of prostate tissue is considered worrisome for prostate cancer. PSA density has been extended to include transition zone measurements in relation to the overall size of the prostate as the transition zone is the site in which BPH develops with ~25% of prostate cancers also arising in this zone. The larger the transition zone in relation to the overall size of the gland, the lower the likelihood of prostate cancer, other things being equal.

E. Prostatic Fluid
Deficiencies in the use of PSA in the diagnosis of prostate cancer have led to research into examining the cellular contents of the prostate more directly. 28

Chapter 10. PROSTATE CANCER Bostwick Laboratories provide a test that assays for PCA3/DD3 RNA ( http://www.bostwicklaboratories.com/) from prostatic cells in urine immediately following DRE (168, 169). Tinzl et al (2004) (169) reported that detection of the non-coding PCA/DD3 RNA, which is highly overexpressed in most prostate cancers (167), provided sensitivities of 73%, 84% and 84% with specificities of 61%, 80% and 70% for serum PSA values of <4 ng/ml, 4-10 ng/ml and >10 ng/ml, respectively, in the detection of prostate cancer (169). Fradet et al (2004) (170) reported on a multi-centre evaluation of this marker in a paper titled ‘uPM3, a new molecular urine test for the detection of prostate cancer’. 443 of 517 (86%) samples provided by patients undergoing prostatic biopsies at 5 centres were assessable by this molecular assay. The overall uPM3 sensitivity and specificity were 66% and 89%, respectively. In men with a PSA level <4 ng/ml, the sensitivity was 74% and specificity 91%; for PSA levels 4-10 ng/ml, the sensitivity was 58% and specificity 91%; for PSA levels >10 ng/ml, the sensitivity and specificity was 79% and 80%, respectively. The overall accuracy was 81% compared with 43% and 47% for total PSA at a cutoff of 2.5 and 4.0 ng/ml, respectively.

F. Prostate Needle Biopsies
Once the possibility of a prostatic cancer is raised, whether by rectal examination, PSA parameters, or a combination of both, the second part of the contemporary two-step early-diagnostic approach, TRUS-guided prostate biopsies, is usually performed. TRUS imaging permits spatial positioning of spring-loaded biopsy needles to provide a methodical approach for obtaining tissue cores for standard histopathology. With few exceptions, TRUS imaging by itself is non-diagnostic as only gross changes register as an abnormal appearance on the monitor. The number of biopsy cores taken is important with the chance of missing a cancer by standard sextant biopsy estimated to be approximately 25% (281) so that, more recently, the numbers of cores recommended are at least 8 and preferably a minimum of 10. In addition, it is advocated that biopsies should be directed laterally and that they should include the anterior horns of the peripheral zone (263, 282-287). Many urologists routinely take 12 biopsy cores now to minimise the likelihood of missing cancer. Guidelines established by the American Urologic Association recommend prostate needle biopsies for any man with a PSA value greater than 4.0 ng/ml, or an abnormal prostate on digital rectal examination (247). The issue of repeat biopsies was addressed by Djavan et al (2001) particularly in relation to when it is reasonable to stop repeating the biopsies. Cancer-detection rates in 1051 men biopsied were 22%, 10%, 5% and 4% with 1-4 TRUS biopsy sessions with 58%, 60.9%, 86.3% and 100%, respectively, having organ-confined disease. Recently, Yanke et al (2005) extended experience with the Kattan Nomogram to predict the likelihood of a positive finding at a subsequent biopsy session. Predictor variables studied in the nomogram were patient age, family history of prostate cancer, prostate specific antigen slope, months from initial negative biopsy session, months from previous negative biopsy session, cumulative number of negative cores previously taken and previously detected high grade PIN or atypical small acinar proliferation. The authors evaluated a total of 356 repeat biopsy procedures for 230 patients. The mean number of total cores per patient was 17.9 with 78 men having biopsies positive for cancer. The area under the ROC curve was 0.71, which was greater than any single risk factor (288). Routine practice involves peri-operative antibiotic prophylaxis with a pre-procedural enema and not proceeding if any faeces at all is present in the rectum, as determined by DRE. Since TRUS biopsies are unpleasant and uncomfortable, many urologists use anaesthesia (local or general) as a routine. Minor morbidity is common with this procedure with well over 50% of patients experiencing at least one complication. Fortunately the dreaded complication of life-threatening sepsis is uncommon, generally <1%, even though rates of bacteraemia vary greatly; blood in the urine, ejaculate and 29

Chapter 10. PROSTATE CANCER faeces are not infrequent sequelae with some men having difficulty voiding immediately following the procedure. (289-291).

G. Histologic Analysis
The Gleason Grading System
The biopsy result provides important information for the patient and clinician on which to base management decisions (292, 293). In addition to the pre-biopsy PSA level, important prognostic factors include tumour volume (percentage of the core involved and the number of positive cores) and the histological grade of the tumour. Increasing tumour burden and poor histologic differentiation are associated with a higher risk of metastatic disease, an increased chance of post-treatment failure, and a worse overall prognosis (275, 294, 295). Histological analysis is based on the Gleason grading system that is regarded as the ‘gold standard’ for classifying prostatic adenocarcinoma (296). Based on architectural patterns, tumour is assigned a rating between 1 and 5, with higher numbers representing less differentiated, more aggressive tumours (see Table 1 and Figure 1). A single prostate can harbour multiple foci of different histologic patterns of adenocarcinoma, and it is possible to have Gleason grade 3, 4 and 5 patterns in the same specimen: 85% of prostate tumours are multifocal. The Gleason score (or Gleason sum) is generated by combining the values of the first and second most common (dominant and subdominant). grades (i.e.: in a tumour with mostly Gleason grade 3 and some Gleason grade 4 disease, the Gleason score will be 3+4 = 7), assessed by the uropathologist using low-power light microscopy. The Gleason score provides important prognostic information. Table 5. Gleason grading system Grade 1&2 3 Histology closely-packed glands forming a nodule small infiltrating glands, completelumen formation fused glands, incomplete lumen formation solid sheet or single cells, no lumen formation Biologic Behaviour Indolent disease, rarely progressive most common pattern; less aggressive than pattern 4 indicates tumour progression Very aggressive, late stage

4 5

30

Chapter 10. PROSTATE CANCER

Figure 2a. The presence of Gleason grade 4 or greater histology carries a significantly poorer prognosis (297, 298). Stamey demonstrated that Gleason score 7 tumours can be stratified, based on the amount of grade 4 disease (299). Those with <50% grade 4 behave similarly to Gleason score 6 (more favourable), while those with >50% grade 4 act like Gleason score 8 (unfavourable) cancers. The transition from Gleason 3 to Gleason 4 appears to be a common event and represents a critical juncture in which the tumor acquires a significantly more aggressive phenotype.

31

Chapter 10. PROSTATE CANCER

Prostatic Intra-epithelial Neoplasia

Figure 2b. Prostatic Intra-epithelial Neoplasia Prostatic intraepithelial neoplasia [PIN] is believed to be a precursor of prostate cancer, given the strong association between high grade PIN and prostatic adenocarcinoma (300-303). The presence of high grade PIN is often indicative of the presence of prostate cancer. It has been shown that more than 80 percent of prostates with adenocarcinoma also contain high-grade PIN (PIN-11 & III). High-grade PIN has cytologic features resembling cancer and carries many of the genetic alterations of prostate cancer. The finding of high-grade PIN alone in a biopsy has been cited as an indication to proceed with repeat biopsies given the high co-frequency between high-grade PIN and carcinoma. However, in current practice, the predictive value of PIN in finding cancer on subsequent biopsies has declined, probably due to the extended biopsy techniques yielding higher rates of initial cancer detection (304). A diagnosis of PIN by itself is certainly insufficient for a patient to undergo either radical prostatectomy or radiotherapy.

Atypical prostatic glandular proliferations
Foci of atypical glands, also labeled atypical small acinar proliferation of uncertain significance, have features suspicious for but not diagnostic of cancer. These encompass a variety of lesions including benign mimickers of cancer, high-grade prostatic intraepithelial neoplasia (PIN), and small foci of carcinoma which, for a variety of reasons, cannot be accurately diagnosed. The reported incidence of these lesions on prostate needle biopsies is 1.5% to 5.3% (304). Patients with atypical glands on needle biopsy have a high risk of harbouring cancer. The reported incidence of prostate cancer from repeat biopsies has ranged from 34 to 60%. (304-306). Following an atypical diagnosis, biopsies need to be repeated (307).

H Staging system
Once a diagnosis of prostate cancer is made, it must be determined whether the pa32

Chapter 10. PROSTATE CANCER tient is a candidate for potentially curative treatment (surgery or radiation). This depends upon several factors, including general health and projected longevity in conjunction with the likelihood that the cancer is still localized within the prostate and has not yet metastasized. The most important factor, however, is the patient’s decision after he has considered the ‘pros and cons’ of the various choices as they relate to him (see below). Currently, the TNM system is used for staging, and prostate cancers can be assigned both a clinical stage and, should the prostate be removed surgically, a pathologic stage. This differentiation is important with the clinical and pathological stage designated by the letters ‘c’ and ‘p’, respectively, preceding the stage denotation (e.g. cT2a = clinically, tumour is palpably involving one lobe of the prostate or less). Table 6. TNM staging classifications Primary Tumour Tx T0 T1 T1a T1b T1c T2 T2a T2b T2c T3 T3a T3b T4 Lymph Nodes Nx N0 N1 Distant Metastases Mx M0 M1 M1a M1b Regional nodes not assessed No Metastases No distant Non-regional lymph nodes Bone(s) 33 Regional nodes were not assessed No regional nodes Regional node metastases Primary tumour cannot be assessed No evidence of primary tumour Clinically inapparent tumour not palpable not visible by imaging Incidental tumour in < 5% of TUR tissue Incidental tumour in > 5% of TUR tissue Needle biopsy prompted by elevated PSA Organ confined Tumour involves one half of one lobe or less Tumour involves more than half of one lobe but not both lobes Tumour involves both lobes Tumour extends beyond the prostatic capsule Extracapsular, unilateral and bilateral Tumour invades seminal vesicles (s) Tumour invades bladder neck, sphincter, rectum, pelvic side wall

Chapter 10. PROSTATE CANCER Primary Tumour M1c Other site(s) with or without bone disease

VII. PROSTATE CANCER TREATMENT OPTIONS
A. Introduction
Traditionally, for a man with clinically-localized prostate cancer, there have been three treatment options: watchful waiting, radical RRP, and radiotherapy. For patients with advanced prostate cancer or for those men with sufficiently serious comorbidities such that surgery and radiotherapy are contraindicated, treatment options have generally been limited to hormonal ablation or watchful waiting (308, 309). Over the past few years, chemotherapy and other experimental approaches also have been championed for those with advanced disease (310). A recent paper by Bill-Axelson et al (2005) (311) is timely in relation to whether treatment with curative intent, in particular by RRP, makes a difference to survival. From October 1989 to February 1999, these authors recruited 695 men who were randomised to RRP or watchful waiting. The updated data from this Scandinavian study demonstrates a significant difference in the overall (and not just cancer deemed) death rates in favour of RRP with a median of 8.2 years of follow-up; 83/347 men in the surgery group and 106/348 men assigned to watchful-waiting died (P=0.04). (For 30/347 men randomised to surgery (8.6%) and 50/348 men assigned to watchful waiting (14.4%), death was attributed to prostate cancer). In their interim report, Holmberg et al (2002) (312) revealed the difference between their patients and those presenting contemporarily. For those men randomised to watchful waiting and RRP, respectively 1. only 10.9% (38/348) and 12.4% (43/347) had their prostate cancers diagnosed through investigations of an abnormal PSA level 2. 45.4% (158/348) and 50.7% (176/347) of men had serum PSA levels >10.1 ng/ml at baseline 3. 29.6% (103/348) and 26.2% (91/347) had Gleason scores of >7 4. for 9.5% (33/348) and 13.3% (46/347) the diagnosis was made cytologically or the biopsy specimen could not be retrieved. Nevertheless, the updated 2005 results of this randomised, controlled trial are notable in that RRP reduced disease-specific mortality, overall mortality, and the risks of metastases and local progression. Although the absolute reduction in the risk of death after 10 years was small, the reductions in the risks of metastasis and local tumour progression were substantial (311). Both RRP and radiotherapy (including brachytherapy) have undergone substantial modifications during the past 1-2 decades. Unfortunately, however, both these forms of potentially curative treatment continue to have risks of significant unwanted effects in subgroups of men (see below). As a result, patients have to choose and, in so doing, ‘trade off’ the risks of side-effect risks from one form of treatment with those from others. Other options, such as cryotherapy and High-Intensity Focused Ultrasound (HIFU) do not have sufficiently-established track records at this time to be recommended as routine options for localised prostate cancer, except in research settings. Unwanted consequences of HIFU vary considerably with impotence rates 44%61%, grade 2-3 incontinence 0%-14%, and rectal fistulae 0.7%-3.2% (313). Although there have been a significant advancements in understanding the mechanisms of can34

Chapter 10. PROSTATE CANCER cer cryotherapy for tumour destruction with improved delivery methods resulting in more effective local prostate cancer control, its most appropriate application is for patients with bulky local disease and local recurrence after radiation therapy. These topics have been reviewed recently in an excellent supplement of Urology (314, 315). http://www.nhmrc.gov.au/publications/_files/cp88.pdf http://www.nelh.nhs.uk/guidelinesdb/html/Prostate-ft.htm http://www.uroweb.nl/files/uploaded_files/2005ProstateCancer.pdf http://www.cancer.gov/cancertopics/understanding-prostate-cancertreatment/page5

B. Pre-treatment decision-making
Making decisions about prostate cancer treatments
Men who are diagnosed with prostate cancer often find themselves in a situation of choosing between unfamiliar and often complex treatment choices, while facing the psychological distress of a cancer diagnosis. In this context decision-related distress is common and persistent . Population-based studies suggest that the medical treatments received by men with localised prostate cancer are influenced by age and by socio-economic factors that affect access to medical services. For example, assessed 3,073 North American men six months after treatment for localised prostate cancer. Conservative management was predicted by later stage disease, physical comorbidity, older age, and being unmarried. As well, radical prostatectomy was received more frequently by Hispanics compared with non-Hispanic Caucasians and less frequently by men with lower education and income levels. utilised cancer registry data about 2,941 patients diagnosed with prostate cancer in the Netherlands and found that the presence of co-morbidities such as cardiovascular disease or diabetes had little effect on what medical treatments men received. Rather, men were more likely to receive radical prostatectomy if they were younger, had a small clinically localised tumour that was moderately differentiated and when they had been diagnosed in a hospital with a high clinical case load. These researchers concluded that treatment was mostly determined by the patient’s age and the extent of the urologist’s surgical experience. assessed clinical and socio-demographic factors to identify predictors of treatment choices amongst 1,809 North American men, predominantly white middle class and affluent, who were diagnosed with prostate cancer as a result of a screening program. Younger age, a higher cancer stage and PSA level and being of non African-American race predicted receiving curative treatments. Men with normal sexual function were more likely to receive watchful waiting whereas men with normal urinary function were more likely to receive radical prostatectomy. Several studies have found that the clinician’s recommendation strongly influences men’s treatment choice . As well, lay health beliefs, such as the view that surgery is the best way to cure a cancer are a strong influence, with the use of such beliefs acting as a short cut to more effortful systematic processing . As one example, found only 13% of men made their decision about treatment for localised prostate cancer by weighing up the risks and benefits of each different medical treatment. Thus, there is a need for care to be taken in supporting informed and patient oriented choice. In order to support men’s decision making, as a minimum men should be provided with evidence-based patient decision support materials to provide them with an opportunity to become well informed about their treatment options. A recent Cochrane review provides guidance on acceptable attributes of evidence based decision aids and some of these can be web accessed ( www.prostatehealth.org.au10; www.ohri.ca/decisionaid11). As well, a generic decision aid, the Ottawa Personal Decision Guide is available that can be adapted for use for most patient populations (www.ohri.ca/decisionaid). Two of the present authors (SKS and RAG) are currently 35

Chapter 10. PROSTATE CANCER validating a revised version of the Ottawa Personal Decision Guide within a randomised control trial of a multi-component intervention targeted to the specific challenges men experience in the early diagnostic and treatment phase of prostate cancer. This approach integrates psychoeducation and decision support in a novel approach . The psycho-educational component is informed by the stress and coping model and problem solving therapy . Structured counselling protocols and patient education materials underpin a telephone based nurse delivered support intervention that commences at diagnosis and extends 6 to 8 weeks after treatment. Peer support may also be helpful to men at the time of deciding about treatments. Peers can provide support from the perspective of shared personal experience . In this regard, men who have been previously diagnosed with prostate cancer can provide first hand advice about what it is like to live with the effects of treatments, practical advice about ways to cope, and ongoing social support. A range of peer support programs are available world wide, with research suggesting they are positively received by men . Peer support programs work well when they are integrated into a broader support framework and are linked to clinicians, and many are available that do this (table 7). Table 7. Examples of prostate cancer Specific Peer support programmes Group Example Group Example Canadian prostate cancer Network (CAN) Us Too (US) Website Website http://www.cpcn.org/

http://www.ustoo.com/

PSA: Prostate cancer Support Association (UK)

http://www.prostatecancersupport.info/

Prostate cancer Foundation of Australia (AUST)

http://www.prostate.org.au/support.htm

Prostate Awareness and Support Society (NZ) The Scottish Association of prostate cancer Support Groups Irish Cancer Society: Men against cancer

http://www.prostate.org.nz/index.html http://www.prostatescot.co.uk/

http://www.cancer.ie/support/mac.php

C. Clinical staging
(i) Clinically insignificant tumours
Because there was a pronounced increase in the number of men diagnosed with 36

Chapter 10. PROSTATE CANCER prostate cancer following the introduction and widespread use of the PSA blood test, concern arose that a proportion of these patients had insignificant disease which did not warrant treatment. As a result, Epstein et al (1994) (334) developed criteria to identify insignificant prostate cancer. These included a PSA density <0.15, a biopsy Gleason score <6, the presence of disease in fewer than 3/6 biopsy cores and <50% prostate cancer involvement in each of these cores. Bastian et al (2004) (335) reported on the analysis of these criteria in relation to 237 patients who had RRPs between December 2002 and August 2003. The large majority (67%) had only one TRUS biopsy core positive and most (89.9% had Gleason 6 in the RRP specimen. However, 9.7% had Gleason 7 or 8 tumours and 8.4% had non-organ confined disease, illustrating the problem of ‘undercalling’ when relying on biopsy information for stratifying patients into a good prognostic group.

(ii) Determining whether the tumour is localised
The presence of any cancer cells outside the prostate locally or at distant sites following treatment with curative intent, is regarded as a portent for ultimate treatment failure. The likelihood of cancer becoming evident after prostatectomy and radiotherapy increases with pathologic stage (336, 337), both these therapies being potentially curative for men with localized (T1 and T2; N0; M0) prostate cancer (336-338). Because these treatments are considerable undertakings and may cause problems in terms of their unwanted effects, it is imperative that, as far as possible, pre-treatment staging excludes any evidence of locally-advanced or metastatic disease. However, the clinical staging of prostate cancer is inexact and this imprecision continues to be a serious limitation in the overall management of patients with this malignancy. PSA serology is a poor predictor of pathologic stage and needle biopsies can misrepresent - usually understate - the volume, histology, and thus the expected behaviour of a patient’s tumour (339, 340). In addition, the presence of extraprostatic extension is easily underappreciated by digital palpation, and conventional imaging studies can fail to detect metastatic spread (341). Approximately 25% of men with clinically localised disease experience an early relapse despite successful treatment of the primary lesion, and up to 50 percent of men with clinically organ-confined lesions are discovered to be understaged at time of surgery (15, 342-346). Cancer cells, particularly in the case of higher grade tumours, are apt to change phenotype and be motile. Facilitated by enzymic breakdown of the extracellular matrix, infiltration of motile cancer cells proceeds into extracellular tissues with intravasation into vascular and lymphatic channels, occurring much earlier than was appreciated previously. However, the rate-limiting steps in terms of metastasising are extravasation from the circulation into tissues remote from the primary lesion and the ability to grow and thrive in these new environments which, for prostate cancer, are particularly the bone of the axial skeleton and pelvic and retroperitoneal lymp nodes. A very recent report indicates that bone marrow-derived haemopoietic progenitor cells expressing vascular endothelial growth factor 1 (VEGFR1) home to tumour-specific pre-metastatic sites and form clusters, preceding the arrival of tumour cells. These VEGFR1+ cells also express VLA-4 which binds to fibronectin, upregulated in resident fibroblasts by tumour-specific growth factors, to provide a permissive niche for incoming tumour cells. Thus a favourable environment for secondary tumour cells to become established is pre-arranged even before arrival or the tumour cells themselves (347). One important though easily neglected role of clinicians is to inform patients of the potential that RRP or radiotherapy, including brachytherapy, may not result in cure. To help with this task, the Partin Tables may be used (248, 260). Initially published in 1993, these were subsequently updated to accommodate an earlier stage of presentation at diagnosis for the majority of patients (348). The Partin Tables have been validated by others in multicentre studies ( 349 ) although not all reports have confirmed an improvement from earlier versions ( 350 ). In addition, Steuber at al ( 2005 ) found that the Partin Tables were less predictive for predominantly transition zone 37

Chapter 10. PROSTATE CANCER tumours. They attributed this variance to the different biological tumour characteristics of transitional zone lesions ( 351 ). First published in 1993, subsequently modified (352) and validated in a number of centres (353), nomograms have been created to summarize a multivariate logistic regression analysis for the prediction of pathologic stage using the combination of DRE, serum PSA and Gleason score. While these estimations do not predict whether or not a given patient will be cured with surgery, they do provide an indication of likelihood of disease-free recurrence at 5 years. Stratifying patients is often a helpful exercise in this regard. The use of nomograms based on age have been extended recently to include the use of percentage free PSA for determining the presence of prostate cancer ( 354 ). Problems with the application of tables and nomogram findings to individual patients are that PSA is a labile enzyme with serum levels affected by factors other than prostate cancer, TRUS biopsies tend to understate cancer both in terms of the extent of tumour and its Gleason score and clinical staging is subjective and relatively insensitive. Nevertheless, tables and nomograms based on pre-operative findings are useful to a degree. Post-operatively, however, incorporation of more definitive and comprehensive parameters obtained from pathology of the RRP specimen itself, enable formulation of more predictive nomograms (355). A number of studies have stratified patients with respect to likelihood of tumour progression with recurrent disease, if it does become evident, likely to do so for the first time within 5 years after RRP. Partin et al were the first to develop a simple biostatistical model equation which allowed categorisation of cases after RRP into 3 risk groups reflecting a low, intermediate and high likelihood of PSA recurrence (356). After an extensive multifactorial regression analysis, they identified only 3 variables viz. a sigmoidal transformation of PSA, Gleason sum of the RRP specimen and margin status or tumour confinement within the prostate. Other studies have confirmed the importance of these parameters and the prognostic value of the 3-group categorisation approach (357, 358). Table. Box 2. Stratifying patients: Factors indicating a high likelihood of recurrence following treatment with curative intent PSA >10 ng/ml Gleason grade 4 or 5 pattern on biopsy Stage 2b (and higher stage) tumour

(iii) Methods for Clinical Staging of Prostate cancer
A combination of the following parameters is used routinely to estimate whether a prostate cancer is localized: 1. clinical stage of the primary tumour (based on digital rectal examination) 2. PSA serology 3. the histologic grade of the prostate biopsy/ies (Uropathology expertise is essential in this interpretation) 4. The number of positive cores and percentage involvement by tumour 5. Imaging studies to identify the presence of metastases

38

Chapter 10. PROSTATE CANCER 1. Digital rectal examination (DRE) Digital rectal examination understages organ-confined disease (245, 342-346). In a published series of 601 men undergoing RRP, of the 565 men with cT2 disease, only 52% had organ-confined tumours (248). By comparison, of 36 men with cT3 disease, 19% had organ-confined lesions. Although there are reports of cure by RRP for cT3 disease, the presence of a bulky extra-prostatic tumour generally indicates a poor outcome, given the high associated risk of metastatic disease (359, 360). Of particular importance is palpable disease at the prostatic apex. The presence of a nodule on digital rectal examination at the apex is often indicative of extraprostatic extension at that location and tends to foreshadows a positive surgical margin or poorer result post-radiotherapy (361-363).

2. PSA Serology As PSA levels increase, the chance of non-localised disease increases accordingly (247). Partin reported that, for men with PSA <10 ng/ml, 70-80% will have organconfined disease. This decreases to approximately 50% for men with PSA >10 ng/ml, and to approximately 25% for men with PSA >50 ng/ml (248). Sanwick et al (1998) (295) showed that 37.5% of patients with PSA values between 10-15 ng/ml and all patients with PSA values >15 ng/ml had evidence of extracapsular extension (295). Because PSA values vary widely within a given stage and overlap between different stages, the predictive value of PSA in determining pathologic stage is weak (340). Even with the combined use of DRE, serum PSA, and TRUS, it is not feasible to reliably estimate the stage of an individual tumour prior to treatment. However, after studying the PSA velocity of 1095 men with localized prostate cancer during the year before diagnosis, D’Amico et al (2004) (364) concluded that patients whose PSA levels increase >2.0 ng per ml are particularly at risk of dying from prostate cancer despite undergoing RRP. Despite initial enthusiasm for RT-PCR molecular staging assays, at this time there is no indication for their routine use in preoperative staging (365-367).

3. Histology Oesterling et al. demonstrated the importance of Gleason grade on prostate needle biopsies as a predictor of final pathology (368). Gleason score 6 tumours were associated with a 24% risk of capsular penetration and a 29% probability of positive surgical margins. This increased to a 62% risk of capsular penetration and a 48% probability of positive surgical margins for Gleason score 7 cancers, and 85% and 59%, respectively, for Gleason score 8-10. Of 72 men with Gleason sore 8-10 tumors, Partin et al. noted extraprostatic disease in 92% (248). Furthermore, some circumspection should be exercised in using biopsy histology for prognosticating since there is a tendency to under-represent the Gleason score and tumour volume relative to pathology findings in the whole gland.

4. Tumour Volume from TRUS-guided biopsies McNeal and associates have demonstrated that tumour volume on prostatectomy specimens correlates with extracapsular extension (369). However, they found that an accurate measurement of tumour volume based on biopsy findings is very difficult technically. McNeal and Stamey have further proposed that the biological aggressiveness of prostate cancer is a direct function of the tumour volume (370). Hence, it can be implied that the larger the volume with biopsies, the greater the likelihood of extracapsular disease. This relationship has been confirmed in several studies which suggest that the tumour extent with prostate biopsies can predict T3 disease. Using multivariate analysis models, Goto et al., determined that the length of cancer in needle biopsies could predict extraprostatic disease, and Sebo et al. demonstrated that the percent of needle biopsy cores and the surface area positive for cancer were strong 39

Chapter 10. PROSTATE CANCER predictors of pathologic stage and tumour volume of the pathologic specimen (294, 371).

5. Imaging Clinical staging is performed routinely with technetium labelled phosphate radioisotope bone scanning. Sites of increased radio-tracer uptake are those with greater metabolic activity, so called ‘hot-spots’, and a malignant cause has to be differentiated from non-malignant diagnoses such as arthritis and Pagets Disease. The predilection for prostate cancer metastases to strongly favour the axial skeleton with limb involvement (apart from upper femora) much less common, helps in differentiating malignant prostatic from other causes of increased radio-isotope activity. Routine bone scanning is usual prior to treatment, irrespective of the likelihood or lack of likelihood of metastases being demonstrable, as a bone scan at this time serves as a baseline reference for subsequent monitoring. CT scanning is not always performed at baseline in men with a prostate cancer diagnosis. For detecting soft-tissue metastases, CT is the usual first-line investigation. However lymph node deposits need to be >1cm before they are usually regarded as pathological with the enlargement presumed to be due to prostate cancer until proven otherwise, especially if sited in pelvic and para-aortic regions. Magnetic Resonance Imaging (MRI) involving the use of an endorectal coil has been promoted to provide an improvement in detection of extracapsular disease but, in the overall context of patient assessment, this investigation adds little to clinical staging; however, MRI can be helpful in evaluating spinal secondaries, especially in relation to possible neural compression. The roles of positron emission tomography (PET) and ProstaScint scanning are not well established. A 70-80% sensitivity has been cited for ProstaScint in detecting lymph node metastases (372).

D. Management Options
Although routine PSA-based detection has increased the percentage of newly-diagnosed cases potentially curable by surgery or radiotherapy, a considerable proportion of patients are not candidates for such intervention (1). Common exclusionary factors include unresectable (locally-advanced) disease, ‘poor tumour pathology’, the presence of demonstrable metastases and advancing age in association with a limited life-expectancy, this last factor begging the question why a diagnosis of prostate cancer was sought in the first place. In addition, there are some men who refuse therapy altogether for fear of anaesthesia or concerns over risks of incontinence and impotence. Because prostate cancer is relatively slowly growing, only those men with at least a 10-year life expectancy are usually considered for aggressive, potentially-curative treatment. Although RRP is not offered to men with cT3 or cT4 disease, radiotherapy, in combination with androgen ablation therapy, does have a more recognised role in locally advanced disease, especially with neo-adjunctive or adjunctive androgen suppressive treatment (373-377). Justification may be based on a need to achieve local control of the cancer, however, a proportion of patients with cT1 or cT2 tumors that are upstaged to pT3 at time of surgery have been reported to have durable cancer remissions (336, 378). For clinically localised disease, there are 3 management options viz. RRP, radiotherapy in the form of either external beam radiotherapy or brachytherapy and watchful waiting. The last of these is not synonymous with patient disregard as, apart from a need for ongoing support and counselling, these men need to be monitored for disease progression and for potential future therapeutic intervention, which may include treatment with curative intent. 40

Chapter 10. PROSTATE CANCER

E. Sexuality and prostate cancer
While most men appear to adjust well psychologically to the experience of prostate cancer, sexual difficulties are frequently described as a bothersome problem experienced by men following both surgery and radiation therapy for prostate cancer . The nature of these difficulties goes beyond the symptom of erectile dysfunction, although this is the effect most often described in the research literature. Physical effects men may experience include: 1. Erectile dysfunction 2. penile shortening 3. loss of sexual desire 4. less satisfying orgasms, and for a small percentage of men, 5. painful orgasms . The inability to achieve a spontaneous erection sufficient for penetrative sex has an obvious direct impact upon intimate relations. However, this is more than just a physical consequence, with the domains of quality of life that may be affected by erectile dysfunction including impaired sexual performance; and changes in relationships with women, sexual imaginings and masculinity . For example, men who experience erectile dysfunction after treatment for prostate cancer describe fears about intimate contact with partners, embarrassment about the failure or potential failure to obtain an erection, and awkwardness about needing to use mechanical devices or penile injections within an intimate encounter . As well, men describe changes in the way they relate to women generally, with regards to losing the enjoyment of sexual feelings in response to women they find attractive. This loss of desire and arousal is broader than changes in the men’s actual relationships and extends to sexual fantasies and imaginings. Men may experience a loss of feelings of masculinity and of self worth, as well as embarrassment and loss of self esteem about other related physical changes such as penile shortening. Studies suggest that an early return to sexual activity using intracavernous injections, specifically by three months after surgery, may increase the recovery rate of spontaneous erections after prostatectomy and improve men’s responses to erectile dysfunction treatments . However, problematically, after prostate cancer men may be reluctant to seek help for sexual difficulties, with studies finding that up to five years after treatment only about half of men seek medical treatment for erectile dysfunction . Further, for those men who do seek treatment, improvements in erectile function are modest. Thus, sexual dysfunction is an ongoing quality of life problem for many men after being treated for prostate cancer. Negative attitudes to help seeking for sexual problems are a barrier to men’s psychosexual adjustment, suggesting a potential role for a counselling intervention . In particular, normalising help seeking for sexual problems; setting realistic goals and expectations about sexual function after treatment; reinforcing the need for flexibility, patience and persistence in managing erectile dysfunction; and encouraging and teaching a problem solving approach may be helpful. With regards to sources of help for sexual problems, men report a preference for help from their urologist or cancer specialist; written information about erectile dysfunction before and after treatment and one to one support . Including the man’s partner in medical consultations about sexual problems is also likely to be helpful. Table 8 lists internet sites and self help books that may be helpful to patients and clinicians. As well, prostate cancer support groups may also be helpful for some men. A further approach is the integration of elements of sex therapy, such as sexual communication and stimulation, with medical treatments for erectile dysfunction. In this regard, a recent pilot of a four session counselling program to enhance sexual rehabilitation after treatment for localised prostate cancer produced improved sexual satisfaction in both men and their partners at three month follow up, and increased 41

Chapter 10. PROSTATE CANCER utilisation of medical treatments for erectile dysfunction at three and six months . Improvements in sexual satisfaction were not maintained at six months. Contrary to expectations, the presence of the man’s partner during the intervention did not affect study outcomes. These authors are currently trialling an internet based version of this program in order to improve the acceptability of the program to men and study generalisability. Table 8. Examples of Internet Sites Relating to Prostate cancer and Sexuality Internet Sites Oncolink has links to chatlines for sexuality and cancer or fertility and cancer, and links to other useful sites. http://www.oncolink.upenn.edu/psychosocial/sexuality Cancer Source offers interactive tools and community resources. http://www.cancersource.com Association of Cancer Online Resources is a cancer online information system that offers access to electronic mailing lists and links to other sites. http://www.acor.org Andrology Australia has resources about sexual dysfunction. http://www.andrologyaustralia.org The Lions Prostate cancer Website has information on prostate cancertreatment and support groups, links t other sites, as well as an email advisory service. http://www.prostatehealth.org.au Books Sexuality and Fertility after Cancer by Leslie R. Schover, Ph.D., John Wiley & Sons, 1997. A how-to and educational book for all types of cancer. Sexuality & Cancer: For the Man Who Has Cancer and His Partner, prepared for the American Cancer Society by Leslie R. Schover, Ph.D. Large booklet available free of charge from ACS website and offices. His Prostate and Me: A Couple Deals with Prostate cancer by Desiree Lyon Howe, Winedale Publishing Company, 2002. A woman’s perspective of coping as a couple with prostate cancer. Making Love Again:Hope for Couples Facing Loss of Sexual Intimacy, by Virginia and Keith Laken, Ant Hill Press, 2002. An unusually open, emotional account of one couple’s struggle to get their sex life back to normal after radical prostatectomy. The Lovin’ Ain’t Over, The Couples guide to Better Sex after Prostate Disease by Ralph and Barbara Alterowitz, Health Education Literary Publisher, Westbury, NY.

Acknowledgment: We are grateful to Professor Lesley Schover for assistance in compiling this table

VIII. RADICAL RETROPUBIC PROSTATECTOMY (RRP)
There are two traditional approaches to the prostate for its total removal, retropubic and perineal. Throughout the world, the large majority of total (or so-called radical) prostatectomies are undertaken by the retropubic route with perineal prostatectomy performed by only a small number of urologists. The procedure of perineal prostatectomy is suited particularly to the removal of smaller prostates. 42

Chapter 10. PROSTATE CANCER Contemporary surgical management of prostate cancer has been shaped by two seminal events, development of the anatomic RRP by Walsh (393) and application of serum PSA as a means for early detection, resulting in many more young men being diagnosed with this condition than ever before. Consequently, there has been a concerted effort to improve long-term cancer control together with preservation of urinary and sexual function. Description of the anatomic RRP has improved understanding of prostatic and pelvic anatomy and of the structures involved in maintaining urinary continence and erectile function. As a result, more patients are cured of their cancers today and are satisfied with their quality of life post-prostatectomy, than ever before (394, 395). The modern RRP evolved from a succession of anatomic dissections and modifications to historical techniques (396, 397). The legacy of the anatomic RRP which permits preservation of the neurovascular bundle with maintenance of erections for many, better sphincter preservation with improved continence rates and more effective management of the dorsal vein complex resulting in less blood loss, has paved the way for not only faster and less morbid recovery (338) but also the newer adaptions of laparoscopic and robotic forms of RRP.

A. Preoperative Consultation and Care
In addition to having pathology and staging details reviewed together with their overall medical status, it is usual to provide candidates for RRP with detailed printed information and, as a routine, have concerns and questions addressed pre-operatively. Blood may be taken in advance to be used as needed peri-operatively as an autologous transfusion. It is standard procedure for compression stockings to be fitted immediately before surgery with intermittent calf compression applied throughout and after the operation, as part of the prophylaxis against deep venous thrombosis. Since the small risk of peri-operative mortality (usually <0.5%) is mostly due to cardiovascular/respiratory complications, some surgeons insist on having all candidates for RRP evaluated with stress testing by cardiovascular physicians.

Delayed commencement of treatment
In terms of timing, it is not uncommon for surgery to be delayed 6 weeks after TRUS biopsies to permit resolution of biopsy-induced inflammation. Patients often are concerned that a delay in treatment is deleterious to the likelihood of the surgery being curative. Graefen et al (2005) (398) analysed data from 795 patients with clinically localized prostate cancer who underwent RRP between January 1992 and June 2000 in relation to the time from biopsy to the date of RRP as a potential prognostic factor. For a mean follow-up of 33 months (1-116 months), 25% of the patients failed on the basis of a postoperative PSA level >0.1 ng/ml. They concluded that a treatment delay in the investigated time span of a few months did not adversely affect recurrence free survival rates and recommended that patients can be reassured that they can evaluate their management options without compromising efficacy due to a delay in initiation of treatment (398). Boorjian et al (2005 (399) reviewed 3,969 consecutive patients who underwent RRP for clinically localised cancer within one year of diagnosis. They reported that the time from biopsies to surgery did not influence the probability of biochemical recurrence, even for those considered to be at high risk of biochemical recurrence. The clinical and pathological features and not the delay were the important factors for estimating risk of biochemical recurrence (399).

43

Chapter 10. PROSTATE CANCER

B. Operative Technique
Most RRPs are performed under regional (epidural) or general anesthesia. Advantages of the epidural approach include diminished blood loss, decreased incidence of pulmonary embolus and improved post-operative analgesia. (400-402). If general anaesthesia is used, post-operative analgesia is usually delivered parenterally for at least 24 hours, often via patient controlled analgesia. Subsequently, oral analgesics such as NSAIDs are used as indicated. Commonly, a Pfannenstiel or a midline incision is made from the umbilicus to the pubis and continued through to the space of Retzius. The peritoneum is mobilized superiorly, enabling the surgery to be undertaken extraperitoneally. A lymphadenectomy is performed, if indicated (see below). The principles and techniques of the anatomic RRP (393) constitute the basis for the operative procedure with various modifications such as bladder neck preservation having been incorporated and used by some urologists. The entire operation is performed under complete visualization with meticulous dissection in a relatively bloodless field, to facilitate optimal cancer control, maximize post-operative continence and, if nerve-sparing is attempted, retain potency. When the specimen is removed, it is examined intraoperatively to ensure completeness of the resection and exclude any evidence of grossly positive margins or violation of the prostatic capsule. An indwelling urethral catheter is left for 1-2 weeks to maximize optimal healing of the vesicourethral anastomosis. A drain is placed near the anastomosis site: it is usually left in position for several days.

C Nerve-sparing prostatectomy
Sexual function is important in middle aged and older men (403,404) as well as their partners. Helgason et al (1996) (403) reported physiological potency for 435 randomly selected Swedish men aged 50-59, 60-69 and 70-80 years to be 97%, 76% and 51%, respectively. Blanker et al (2001) (404) confirmed this finding in data collected from 1688 men with the prevalence of significant dysfunction ranging from 3% in 50-54 year olds to 26% in males between 70-78 years. Consequently, it is not surprising that potency problems rate highly as an important consideration not only in deciding whether or not to proceed to active treatment but in determining which treatment to pursue. The importance of sexual function as a consideration is illustrated by Singer et al (2001)(405) who reported that men undergoing treatment for prostate cancer were willing to exchange an approximate 20% chance of being cured of their cancers for an increased prospect of remaining potent after treatment (405). Consequently for many patients, the emphasis of RRP is no longer focused solely on cancer control, but extends to include lifestyle issues (394, 395, 406,407). This is not a dissimilar paradigm shift to that which has occurred with breast cancer treatment, as sexuality issues have taken a prominent role influencing treatment decisions with radical and partial mastectomies replaced by combination lumpectomy and adjuvant therapy. Potency and continence rates associated with nerve-sparing surgery vary among surgeons and academic centres (see Table 9). It is fair to say that a degree of scepticism is expressed regarding some of the claims made in respect to post-operative potency, in particular. However, explanations for inconsistencies in cited results include differences in surgical skill, patient selection and outcome measurement methodology. Among these, patient selection ranks particularly highly. Nonetheless, it is undisputed that increasing numbers of patients are sexually potent after surgery, albeit with the use of the phosphodiesterase inhibiting drugs (Viagra™, Cialis™ and Levitra™), as a result of the modifications incorporated into the technique of anatomic nerve-sparing RRP. Table 9. Potency rates after nerve-sparing radical prostatectomy Study 44 Potency Rate

Chapter 10. PROSTATE CANCER Study Jonler, et al (1994) (408) Fowler, et al (1993) (409) Heathcote, et al (1998) (410) Ojdeby, et al (1996) (411) Gaylis, et al (1998) (412) Talcott, et al (1997) (413) Sole-Balcells, et al (1992) (414) Davidson, et al (1996) (415) Ritchie, et al (1989) (416) Catalona, et al (1999) (417) Quinlan, et al (1991) (418) Leandri, et al (1992) (419) Noh, et al (2005) (420) Walsh, (2000) (421) Potency Rate 9% 11% 12% 14% 18% 21% 39% 43% 45% 67% 68% 71% 72% 86%

Although definitions for above parameters vary, quality of life studies indicate that the majority of patients are satisfied with the outcomes of their surgery (394, 395, 406, 407). Link et al (2005) (422) reported their experience with laparoscopic RRP using the validated EPIC and a 5-item International Index of Erectile Function questionnaires; 78.9% of men who were previously potent were having sexual intercourse, albeit with the use of phosphodiesterase inhibitors in most instances, 12 months following surgery. Sexual bother decreased to 64% of baseline at 3 months and did not show any significant improvement subsequently. By contrast, the sexual function sub-domain, which refects erectile performance better, decreased to a mean of 36% at 3 months but showed improvement subsequently to 51% of baseline at 6 months and 64% of baseline at 12 months. Recovery of sexual function was not significantly affected by age or pre-operative potency status, although the extent of nerve sparing was a significant predictor of outcome (422).

(i) Anatomical Considerations of Nerve-sparing Surgery
Normal post-prostatectomy erectile function depends upon preservation of the autonomic cavernous nerves, located within the neurovascular bundles of the penis. These nerves are located immediately posterolateral to the prostatic capsule, within the periprostatic fascia at the junction of the lateral and posterior portions of the endopelvic fascia. At this position, the lateral pelvic fascia is comprised of two layers, the levator and the prostatic fascia. The neurovascular bundles pass between these layers bilaterally and can be preserved by entering the periprostatic fascia laterally and gently and meticulously dissecting them off the prostatic fascia. If both neurovascular bundles are preserved, potency (defined as the ability to sustain sufficient erections for sexual intercourse without any aids other than phosphodiesterase inhibitors) can be as high as 68- 86% (338, 417). When only one bundle is saved, potency rates diminish substantially (423). Given the importance of sexual function to many men with prostate cancer and the differences when only one bundle is spared instead of two, it is common practice for every effort to be made to preserve both neurovascular bundles whenever possible, as long as the potential of cure is not considered to be jeopardized – consistent with the wishes of the patient.

(ii) Effect of Nerve-Sparing Surgery on Cancer Control
The neurovascular bundles are mostly located outside of the prostatic capsule so, 45

Chapter 10. PROSTATE CANCER unless dissection violates the capsule, nerve-sparing radical prostatectomy should not compromise surgical cancer control if the tumour is organ-confined. Epstein et al followed patients with positive surgical margins present only in the posterolateral region and determined that post-prostatectomy relapse was uncommon (424). In this study, of 507 men with cT1 and cT2 tumours, the most common sites of positive margins were distal (22%), posterior (17%), and posterolateral (14%). Rosen confirmed these findings in 144 men undergoing radical prostatectomy (425). They demonstrated that, when positive margins were present, <10% had involvement in the region of the neurovascular bundles. Hence, in cases of extraprostatic extension, involvement of the posterolateral margin is less common than for other locations. Epstein also examined paired specimens from men with prostate tumors highly suspicious for posterolateral involvement who underwent nerve-sparing prostatectomy followed by excision of the neurovascular bundles on the suspect side (423). Cancer was found in the bundles in only 17.5% of the men. In men with clinically-suspicious but pathologically negative posterolateral margins, no tumour was found in the resected bundles.

(iii) Sural nerve grafting
In order to attempt to maximise preservation of the neurovascular bundles, technical modifications to RRP have been reported; many of these involve methods which improve visualisaton of the neurovascular bundles (420, 426, 427, 428). The potential problem of impotence, when preservation of the neurovascular bundle is not considered appropriate, has been addressed by sural nerve grafting, employing techniques established in the management of facial and peripheral nerve injuries (429, 430). Although this approach has met with a mixed reception by urologists, claimed success rates vary with one study having reported return of erectile activity in 75% of men with maximum return of function 14-18 months post-RRP (431). However, most Urologists cite much lower success rates.

(iv) Laparoscopic and Robotic prostatectomy
As expected, the recently introduced techniques of laparoscopic and robotic prostatectomy differ from open radical prostatectomy in that they have incorporated some adaptations for practical reasons. However, the principles of the prostatic cancer surgery remain the same and results to date are comparable (426, 432). Keyhole approaches are attractive to patients because of perceptions of reduced surgical trauma and morbidity with, by extrapolation, less post-operative pain - even though this is uncommonly problematic with the open approach. A recently reported prospective study of robot assisted laparoscopic prostatectomy (n = 159) is illustrative in that it did not provide a clinically meaningful decrease in post-operative pain compared with RRP (n = 154), which the authors ascribed primarily to the low pain scores reported in both groups. They concluded that outcomes other than pain will ultimately determine the role of laparoscopic radical prostatectomy and Robot assisted laparoscopic prostatectomy (433). Conceptually, better vision is afforded the surgeon with forms of endoscopic surgery to permit easier identification of structures, particularly apically (434, 435), but there is the potential problem of delay if (the infrequent) conversion to an open procedure is required in the event of bleeding. Unlike laparoscopic surgery, roboticallyperformed procedures involve hand-eye coordination and hand positions which are more akin to those used in open surgery; in addition, the learning period appears to be shorter (436). Although the capital costs associated with robotic surgery are, at present, very large, these are likely to decrease with time.

46

Chapter 10. PROSTATE CANCER

D. Pelvic Lymphadenectomy
Following bone, the pelvic lymph nodes are the second most common site of prostate cancer metastases (437). While radionuclide bone scans provide a sensitive method for evaluating skeletal lesions, the sensitivity of CT scans to delineate lymph node metastases is poor (48, 49, 438). Pelvic lymphadenectomy prior to RRP to detect metastatic lymphatic involvement is limited to patients at high risk for tumour cell dissemination, then only proceeding to RRP if the lymph nodes are free of tumour. In the PSA-era of the past 15 years, the rate of positive lymph nodes in patients undergoing RRP has plummeted from approximately 25-30% to under 5% (439-442). Consequently, there has been considerable debate regarding which patients should have lymph node dissections. It is generally accepted that men with Gleason scores < 7 and PSA values <10 ng/ml do not require pelvic lymphadenectomy (424). Some investigators have increased these limits to include PSA values from 10-20 ng/ml and Gleason 7 tumours (443). The presence of a positive lymph node is a predictor of post-operative recurrence. Most studies demonstrate that such foci of cancer herald subsequent curative treatment failure: 10-year biochemical disease-free survival rates are low (361, 444). In a study designed to determine if early versus delayed hormonal ablation therapy in men with lymph node positive disease was advantageous, Messing et al (1999) (445) reported on 51 men with lymph node positive disease who were randomized to delayed therapy. At an average follow-up of 7 years, 16 percent had no biochemical or clinical evidence of disease and were never started on androgen ablation (445). Pelvic lymphadenopathy can be performed as a separate procedure through a "minilaparotomy" or by laparoscopy, or can be combined with RRP (446). All nodal tissue medial to the external iliac artery, from the junction of the external iliac and hypogastric arteries to the obturator foramen is removed with this procedure. It is then sent for frozen section histologic analysis. If lymphadenectomy is performed prior to a planned RRP and the frozen section report indicates that there is no cancer evident, it is usual to proceed with the prostatectomy. If there are tumour cells present in the specimen, the planned RRP is almost always aborted. Complications of pelvic lymphadenectomy are infrequent (415, 447-450). Injury to the obturator nerve can occur in 1% of patients causing an inability to adduct the thigh. If complete transection occurs, primary anastomosis can be performed. Bleeding from obturator and iliac vessels is uncommon as are ureteric injuries. A lymphocele may form in the pelvis which can become symptomatic and require sclerosis or formal drainage into the peritoneal cavity. (451).

IX. RESULTS OF RADICAL RETROPUBIC PROSTATECTOMY
A. Disease-free Recurrence and Survival
(i) Pre-operative PSA
The results of published series vary but, depending very much on the cases selected, survival rates at 10 years have been reported to exceed 90% (338, 452, 453) with a comparable percentage of continent and many men potent post-operatively (338). Lower preoperative PSA, pathologic stage, and Gleason grade are associated with high rates of cure after RRP. For patients with preoperative PSA values <10 ng/ml, Catalona demonstrated that 7-year disease-free survival was 76-93% (417). If PSA was >10 ng/ml, disease-free survival decreased to 51%. Similar results have been shown by others (247, 454, 455).

47

Chapter 10. PROSTATE CANCER

(ii) Stage & Grade
Not unexpectedly, the probability of cure is considerably greater for patients with localized, non-palpable, and well to moderately differentiated tumors. The risk of treatment failure rises in relation to increases in clinical and pathological stage and correlates with increasingly aggressive histologic patterns. Survival rates has been reported to decrease with increasing pathologic stage, from 81% for pT2 lesions, 76% for pT3 tumors and 19% for N(+) disease. As indicated above, survival rates also vary with grade. In a multi-institutional study, Gerber et al (1996) (456) demonstrated that 10 year progression-free rates correlated with grade with 87% of low-grade tumors, 68% of moderate and 52% of high-grade cancers progressing, with other authors reporting comparable results (339, 340, 455-457).

(iii) Positive margins
A positive surgical margins is present if the tumour extends to the edge of the specimen. Patients with positive surgical margins, whether surgically induced or not, have approximately a 2 fold higher recurrence rate compared with patients with negative margins. The progression-free probability at 5 years for men with negative margins at RRP has been reported to be ~80% compared with 42-64% for men with positive margins (297, 458)

(iv) Molecular markers
A number of molecular markers have been evaluated for their prognostic value but none has been adopted for routine use in practice. Amongst others, the potential of CD151, c-myc and calveolin-1 overexpression in predicting outcome has been discussed earlier in this chapter. Overexpression of p21 protein has been reported to be a predictor of response to salvage radiotherapy after RRP (459) and Fizazi et al (2002) (460) correlated p21 expression with progression to androgen independence. In addition, expression of p53 and the combined loss of PTEN and p27 have been shown to be predictive in the identification of the likelihood of disease recurrence (461, 462). Fradet et al (2004) (463) demonstrated that expression of NF-κB in prostate cancer cells at the sites of positive margins was highly indicative of biochemical recurrence (463).

(v) Lymph node & seminal vesicle involvement
Although clearly in the minority, some men with microscopic N(+) disease have undetectable PSA levels at 5-10 years. Progression in N(+) disease appears to be related to the volume of tumour in the lymph nodes. As well, not all men with seminal vesicle involvement have poor outcomes (464). Despite the observation that seminal vesicle and lymph node involvement are associated with poor results post-prostatectomy, in the Catalona series, at seven years the disease-free survival rates were 26% for men with seminal vesicle involvement and 9% for men with N(+) disease (337). There also appears to be a difference if seminal vesicle involvement is a direct extension from the primary tumour, in which case it behaves more like an extracapsular extension rather than a separate metastatic focus which implies systemic involvement.

(vi) PSA monitoring
PSA serology is very useful in monitoring for disease recurrence post-RRP. If all cancer tissue is removed, PSA values should drop to the undetectable range (between 0-0.02 ng/ml). Understandably, patients become strongly focussed on their postoperative PSA values with raised levels following RRP indicating persisting prostatic 48

Chapter 10. PROSTATE CANCER tissue at a distant and/or local site (444, 452, 465). Most PSA recurrences occur in the first few years after surgery and precede clinical evidence of disease by as much as 5 years. It is rare for the PSA to become detectable after levels have been undetectable for 5 years. The rate of PSA doubling time is an important predictor of poor prognosis (452, 466, 467). Biochemical Recurrence (BCR)/Biochemical(PSA) Failure (BF) Using the definition of PSA failure of the American Society of Therapeutic Radiation and Oncology’s (ASTRO’s) 1996 consensus statement (ie 3 consecutively rising PSA values, each obtained at least 3 months apart), D’Amico et al analysed 888 patients followed for a median time of 38 months (8-100) after RRP. Based on the medical literature, they established 3 groups using known prognostic factors (viz. PSA level, biopsy Gleason score and the American Joint Commission on Cancer [AJCC] staging system). Over 95% of the PSA failures were evident for the intermediate and highrisk groups by 4 years. Questioning the ASTRO guidelines, Amling et al reviewed 2,782 RPP patients followed for a median of 6.3 years. After examining four points of “PSA cut-off”, these authors concluded that a PSA >0.4ng/ml was most appropriate since a significant number of patients with a lower PSA do not continue to increase or proceed to demonstrable metastases (468). Freedland et al (2003) (469) undertook a retrospective review of 358 men who underwent RRP between 1991 and 2001 to determine the most relevant cutpoint for determining biochemical recurrence. For patients with a detectable postoperative PSA value from 0.11 to 0.2 ng/ml, the 1 and 3-year risk of PSA progression was 64% and 93%, respectively. For men with a PSA value from 0.21-0.3 ng/ml, the 1 and 3-year risk of PSA progression was 86% and 100%, respectively. The 5-year risk of PSA recurrence using a greater than 0.1 ng/ml cutpoint resulted in a 43% risk of recurrence compared with only 23% for a greater than 0.5 ng/mL cutpoint (469). Contemporarily, it has become common practice to regard a PSA of >0.2ng/ml as the cutpoint to indicate BCR. Reporting on 213 men who had biochemical failure on the basis of a PSA of >0.2 ng/ml following RRP for a mean follow-up of 56 months (range 1 to 125), 99 of whom were treated with androgen ablation and/or radiation therapy at the time of detection of BF, Jhaveri et al (1999) (470) cited 10-year overall survival rates of 88% compared with 93% for those who did not have BF. They found no difference in survival rates in relation to age (>65 years), preoperative PSA >10 ng/ml, biopsy or specimen Gleason score >7, clinical Stage including the presence of extracapsular extension, positive surgical margins, and seminal vesicle invasion. (470).

(vii) Prognosis
Pound et al (1999) (452) reported in a retrospective study on the fate of 304 men with biochemical escape (PSA >0.2 ng/ml) following RRP, 103 of whom progressed to metastatic disease, not having had androgen suppressive therapy. The median actuarial time to metastases was 8 years from the time of PSA elevation and once patients developed metastases, the median actuarial time to death was 5 years. The time interval from surgery to the appearance of metastatic disease was predictive of the time until death (452). However, it is clear that BCR in itself is not a predictor of imminent demise from prostate cancer and, as alluded to above, PSA doubling time (PSADT) is becoming accepted increasingly as a preferred prognostic indicator (471). Albertsen et al (2004) reported that men who had a PSADT of <1 year had a high risk of dying of prostate cancer within 10 years of diagnosis, and conversely (472). D’Amico et al (2003) (473) examined the results of men who experienced BCR subsequent to RRP and radiation therapy. From their analyses of 5,918 men following RRP and 2,751 patients after radiation therapy, they concluded that a PSADT of <3months met the criteria of sur49

Chapter 10. PROSTATE CANCER rogacy in predicting death from prostate cancer at a median survival time of 6 years (473).

X. COMPLICATIONS OF RADICAL RETROPUBIC PROSTATECTOMY
A. Incontinence
Table 10. Continence rates after radical retropubic prostatectomy Study Percent Continent Average Post-operative Length of Time to Regain Continence 6 months 12 months 6 months 18 months 2 months 18 months 6 months 12 months

Shelfo et al (474) Lowe (475) Goluboff et al (476) Catalona et al (417) Kaye et al (477) Walsh et al (421) Seaman et al (478) Poore et al (479)

88% 85 to 100% 92% 92% 93% 93% 97% 100%

Continence rates from several institutions are represented in the table (above). In general, >90% of patients are continent by their reckoning at 12 months. Most men report little bother from urinary symptoms after RRP with socially acceptable continence achieved usually within 3-6 months of surgery, although, for some patients, regaining control can take up to 18 to 24 months. Factors influencing continence include maintaining the integrity of periurethral support, the precision of the vesicourethral anastomosis, patient age and co-morbidities, including pre-prostatectomy detrusor dysfunction which may persist post-operatively. Return of urinary control is considered to take longer in men over 65 years. Patients with atherosclerotic and diabetic vasculopathy are more prone to prolonged incontinence, probably the result of poor vascularization and healing of the periurethral and sphincteric tissues. Biofeedback and Kegel exercises have been reported to hasten the return of urinary control (480). If urinary control does not return, and the incontinence is demonstrated to be sphincter failure by urodynamics investigations, an artificial urinary sphincter may resolve the problem (481). Bladder neck contractures occur in less than 10% of patients and can cause symptoms ranging from poor urinary flow to complete incontinence (482, 483). Bladder neck contractures can result from non-mucosa-to-mucosal anastomosis, following heavy intraoperative bleeding and after prolonged urinary extravasation, with a previous transurethral resection of the prostate possibly predisposing to this problem. Treatment can consist of simple dilatation, although surgical incision of the scarred tissue is usually performed. Urethral stricturing can also require dilatation and/or incision with formal urethroplasty rarely warranted.

B. Impotence (& sexual dysfunction – see section VII. E)
Spontaneous post-prostatectomy erectile function depends upon preservation of at 50

Chapter 10. PROSTATE CANCER least one neurovascular bundle. While surgical skill is an important factor in the successful outcome of nerve-sparing prostatectomy, patient selection is paramount. Younger men (<60-65 years old) with small, non-palpable, and low-grade tumours have the best outcomes (421). Phosphodiesterase inhibitors (e.g.sildenafil/Viagra™, tadalafil/Cialis™, vardenafil/Levitra™), intracorporeal injection of vasoactive drugs such as prostaglandin E1 or Alprostadil, transurethral vasodilators (medicated urethral system for erection [MUSE]) and vacuum constriction devices (VCD) all have a role in erectile rehabilitation (484, 485). Although patients clearly prefer oral medications (391), these may be less effective in the short term, particularly if temporary neurovascular damage (neuropraxia) is present following RRP (485) as well as following external beam radiotherapy, (389). The results of the various treatment approaches were reviewed recently by Raina et al (2005) (485). They reported that VCD efficacy rates range from 60-80% with compliance at 12 months between 50% and 70%. Tightness or pain from the constriction ring and decreased sensation of the penis, in particular the glans, were common causes for noncomliance. Approximately 50% of men were afforded benefit from MUSE using Alprostadil with some non-nerve sparing RRP patients also able to achieve tumescence satisfactory for vaginal penetration. In addition to not achieving satisfactory penile tumescence, urethral pain and burning were cited causes for discontinuation. Intracorporeal penile injection therapy with PGE1 or Alprostadil is claimed to provide adequate rigidity in >75% of patients but ~50% of men do not continue with the treatment long-term. In addition to a lack of success with the technique, physical and emotional problems in addition to pain with with the injection were promoted as reasons for discontinuation. Penile fibrosis is reported to occur in up to 15% of men (485). Avoidance of penile hypoxia through regular tissue oxygenation via erections is considered to lessen the likelihood of lacunar fibrosis and, ultimately, erectile incapacity. Thus, there is a vogue currently to advocate a return to penile tumesence with the early commencement of one or a combination of the above treatments following RRP. Not unexpectedly, the use of oral phospodiesterase inhibiting drugs is most popular in this regard but, especially in the early post-operative period, other methods may be more effective. Despite high (>75%) efficacy and satisfaction rates, penile implant surgery is restricted to only those men with persisting erectile failure after having exhausted all less invasive options and all hope of spontaneous erectile recovery (485, 486).

C. Bleeding
The anatomic approach to RRP has improved awareness of, and abilities to control, the venous plexuses surrounding the prostate. Consequently, massive blood loss is rare. nevertheless, during nerve-sparing surgery it is not uncommon for 600-1200 mL of blood to be lost. Many patients are encouraged to donate autologous blood preoperatively in case a transfusion is required intraoperatively or in the immediate post-operative period. The need for transfusion may be lessened by the use of a ‘cellsaver’ during the operative procedure.

D. Infection
Despite using an iodine or alcohol-based skin preparation preoperatively, any time a skin incision is made there is the risk of both local and systemic infection. For this reason, it is usual to administer antibacterial drugs prophylactically. During RRP, the bladder is opened and, inevitably, urine spills into the surgical field. Several strategies can be used to diminish the risk of infection. Preoperative urinalysis and urine cultures identify patients with potential urinary tract infections which can be treated. After the vesicourethral anastomosis is completed, lavaging the pelvis with saline may help and post-operative drainage as a routine decreases the likelihood of localised sepsis occurring. 51

Chapter 10. PROSTATE CANCER

E. Visceral Injuries
Rectal injuries occur in under 1% of RRPs and may be more likely subsequent to previous rectal surgery or pelvic irradiation (415, 448, 482). If a rectal injury does occur, it can be repaired easily in most instances by oversewing the rectal wall. It is for this reason that some surgeons suggest that all patients are given a bowel preparation prior to surgery, to minimise wound contamination and lessen the need for a covering colostomy should an injury occur. Although extremely rare, ureteric injuries can happen during transection of the bladder neck and dissection of the seminal vesicles. Correction usually consists of a ureteroneocystotomy with insertion of a ureteric stent.

F. Deep vein thrombosis, pulmonary embolus & other problems
Deep vein thrombosis and pulmonary embolus are diagnosed in approximately 1% of men having a RRP (415, 448, 482, 487). Their probability can be reduced by ensuring adequate hydration peri-operatively, using elastic compression stockings with sequential compression applied during and after surrgery and encouraging early ambulation. Subcutaneous heparin or clexane is also commonly used. All peri-operative complications are increased in patients with conditions characterised by impaired healing or reduced pelvic tissue vascularity (482). Thus, prior pelvic radiation, previous rectal surgery, inflammatory bowel disease, past procedures on the prostate, urethra, or bladder, diabetes mellitus, and vasculopathies can contribute to prolonged incontinence, impotence, hemorrhage, infection, and visceral injuries. In addition, co-morbid cardiac and respiratory conditions can contribute to perioperative morbidity and mortality.

XI. TREATMENT OF POST-PROSTATECTOMY RECURRENCES
Overall, approximately 35% of men undergoing RRP will have a recurrence of their prostate cancer within 5 to 10 years (342-346). The following preoperative findings are associated with a high risk of surgical failure: cT3 lesions, cT1 or cT2 tumors with >Gleason grade 4 number on prostate biopsy, and serum PSA >10 ng/ml. Additional risk factors for recurrence after attempted definitive therapy include such additional pathologic determinants as tumor volume, vascular invasion, seminal vesicle or lymph node involvement, and positive surgical margins. The location of the positive margin can influence the likelihood of tumor recurrence and must be taken into consideration (292). The use of post-operative radiotherapy has been the focus of extensive investigation in recent years, including both retrospective and randomised prospective studies. The use of radiotherapy post-RRP can be with either adjuvant or salvage intent. Adjuvant radiotherapy is that used immediately post-operatively, in the aim of eliminating potential but as yet unidentifiable residual disease. Salvage therapy is conducted when recurrence is suspected - typically when a PSA rise is seen - although the immediate use of radiotherapy when an undetectable post-RRP PSA is not returned may also be regarded in this category.

Adjuvant radiotherapy
A number of retrospective comparative series have examined the potential benefit to adjuvant radiotherapy. Unselected series have suggested biochemical control rates in the combined surgery plus adjuvant radiotherapy cases of 52-93% compared to surgery only rates of 25-74% (488-492). More concordant results were seen in two published matched pair analyses which showed surgery only freedom from bio52

Chapter 10. PROSTATE CANCER chemical failure (FFbF) of 55-59% which was increased to 88-89% with the addition of radiotherapy (493, 494). Three randomised studies of adjuvant therapy have now been announced in abstract or publication form. These trials have focused on a subset of men at high risk of having residual disease with inclusion requiring either pathologically involved surgical margins or pathological stage T3 disease. The EORTC 22911 trial (495) randomised 1005 such patients to observation or post-RRP radiotherapy (60Gy) from 1992-2001. 30% of these had a detectable post-operative PSA. At a median follow-up of 5 years, biochemical progression-free survival was significantly improved in the irradiated arm (74.0% versus 52.6%; hazard ratio 0.48 [95% CI 0.37-0.62]). Clinical progression and loco-regional failure were also significantly improved. Too few death events have been recorded as yet to show a difference, although approximately twice as many deaths from prostate cancer had occurred in the observation arm compared with the irradiated. Grade 3 toxicity was uncommon, and not significantly different between the treatment arms (4.2% and 2.6% cumulative incidence rate at 5 years for the radiotherapy and observed arms respectively). In a detailed companion study, urinary incontinence (as defined by patient reported questionnaires and pad weight measurements) was found to not be increased by the addition of radiotherapy (496). Two additional similar randomised studies have been presented in abstract form to date. The German Cancer Society trial (ARO 96-02 / AUO AP 0995) was designed to asses to impact of adjuvant radiotherapy in those with pT3 disease or positive margins who achieved an undetectable PSA post-RRP (497). With a median follow-up of over 3 years, the results from 261 patients showed a better than 20% absolute improvement in the biochemical failure rate in the radiotherapy arm (hazard ratio 0.40 for the addition of radiotherapy when analysed by treatment received, p<0.0001). Grade 3 rectal toxicity was not observed in either group. Similarly, the Southwest Oncology Group study (SWOG 8794) randomised 419 pT3 and margin positive men to have either observation or radiotherapy (498). 55% had a detectable post-RRP PSA. The risk of biochemical recurrence was reduced by 56% by adding adjuvant radiotherapy, and prospective quality of life assessment suggested no significant differences in the gastrointestinal, urinary and sexual domains at 5 years. The risk of needing future salvage androgen suppression therapy was reduced by 56% also and clinically apparent relapse by 38% (both statistically significant). Distant metastasis and survival events were too infrequent in both studies to make meaningful assessment. This combination of randomised trial data (amounting to level 1 evidence) suggest that there is an unequivocal capacity for post-RRP adjuvant radiotherapy to approximately halve the chance of having a future PSA-detectable tumour recurrence in pT3 / margin positive patients, while maintaining a low level of toxicity. More maturity to the data is awaited to determine the overall impact this has on distant metastases or survival.

Salvage radiotherapy
The role of salvage radiotherapy is far less clearly defined, with no prospective studies of efficacy or toxicity to guide decisions. Retrospective analyses suggest that the PSA level prior to initiating the salvage radiotherapy is strongly predictive of outcome (499-501), with some series suggesting that treatment at PSA levels below 1 ng/mL do substantially better than those above this level (502). This PSA effect has been shown to be independent of the PSA doubling time (503), potentially indicating that these patients benefit from early referral for treatment regardless of PSA dynamics. Indeed, one series demonstrated an independent benefit to the use of immediate adjuvant therapy rather than waiting until requiring salvage (504). Other prognostic factors which have described, although not consistently, have included seminal vesicle invasion, margin positivity, post-operative PSA nadir level, high gleason grade, and radiation dose (501-503). Larger series typically show a 5 year biochemical control rate in the order of 50% typically (501, 505), and over 70% for those with a pretreatment PSA of less than 1 ng/mL and an operative Gleason score of 7 or lower 53

Chapter 10. PROSTATE CANCER (502). Despite the apparent ability to provide a substantial chance for long-term control in relapsed men, there is concerning data that post-RRP radiotherapy may be under-utilised. Only 55% and the 38% of the biochemical failures in the observation arms of the EORTC and SWOG randomised trials respectively underwent potentially curative salvage radiotherapy, with the remaining being treated palliatively with observation or androgen suppression therapy. Prospective efficacy and toxicity data will be required to fully appreciate the therapeutic ratio of radiotherapy in this group of men.

XII. RADIOTHERAPY
A. Introduction
Contemporary radiation therapy can result in cure of prostate cancer in a substantial proportion of cases. Deliberation continues regarding when it is best used rather than other radical treatments, in particular, RRP. Direct randomised evaluation of them would appear unlikely to ever happen, given that a highly anticipated and well funded international trial comparing radical surgery with seed brachytherapy recently closed due to poor accrual (ACOSOG Z0070 – the “SPIRIT” trial). Without such evidence, non-randomised data describing both tumour control and quality of life issues must be balanced against each patient’s concerns and overall health status. Multidisciplinary clinics should be regarded as the standard forum for helping patients balance these issues. Irrespective of the perceived tumour biology, men under 60 years of age are offered surgical therapy in many centres based on, amongst other issues, the practical advantages of ascertainment of pathological tumour stage (discussed above), more accurate grading and a more dependable early PSA endpoint of success in most cases. Many patients, however, will be offered radiotherapy preferentially. This may be on the basis of advancing age, a high risk of not achieving surgical complete clearance or patient preference. Evidence to help determine appropriate therapy is now available for many clinical scenarios within this group of men.

B. Stratification of risk of progression & metastases
In many cases, radiotherapy treatment decisions are now driven by the patient’s risk profile tumour containment. These profiles were typically developed using historical data, and either externally validated or confirmed as clinically relevant by stratification levels in prospective trials. The most simple of these is a three tiered system with (506): 1. Low risk: PSA≤10 ng/mL and Gleason Score 2-6 and stage cT1-cT2a 2. High risk: either cT3 or cT4 or PSA>20 ng/mL or Gleason Score 8-10 3. Intermediate risk falls between these levels While nomogram models (507) have been shown to have more reliable discrimination of outcome, the three level model serves well in relation to known therapeutic option categories. These may involve combinations of external beam radiotherapy (EBRT), brachytherapy (BT) or androgen suppression (AS) therapy.

C. External Beam Radiotherapy
For those with low risk prostate cancer, modern EBRT techniques appear to be highly efficacious (508, 509). PSA control would be expected in over 85% of such cases 54

Chapter 10. PROSTATE CANCER presently. These treatments are typically delivered using conformal 3D techniques, or intensity modulated radiotherapy (IMRT) where these more complex treatments may aid in controlling toxicity, especially at higher doses. These excellent results are yet to have an additional demonstrated benefit when combined with therapies such as AS. High risk prostate cancer traditionally has been a disease with a poor outcome when treated with radiotherapy alone (as with surgery). In the days before PSA detection of tumours, biochemical recurrence rates were typically over 80% using radiotherapy as primary therapy (510). With the risk of both local and metastatic progression being of concern, randomised trials investigated the role of adjuvant AS in what are now known to be typically high risk patients. Published trials of locally advanced disease showed AS to have a beneficial effect when given after radiotherapy for a duration of either 2 years (511), 3 years (375) or indefinitely (512). An overall survival benefit of 16% at 5 years was seen in the Bolla et al trial when compared with radiotherapy alone. Hence, most men today with advanced cancers will have this style of treatment offered as a minimum. Fit men in this group may also be offered therapy on a clinical trial, as the improved results in this group are continuing to be optimised. Intermediate risk prostate cancer, however, appears to have a lesser reliance on treatment of sub-clinical metastatic disease present at diagnosis, as borne out in trials evaluating the value of local therapy intensity. Responses to increasing radiation doses (“dose escalation”) have now been shown in a randomised trial. For the subset of PSA 10-20 ng/mL, the freedom from failure rate at 6 years increased from 43% to 62% when the radiation dose was increased from 70Gy to 78Gy (513). This supports the results of other dose escalation observational series (509, 511, 514). Additionally, short-term use of AS combined with EBRT has shown to be of benefit to some patients. The precise explanation for this improvement in results is less clear though. Observational studies suggested a benefit predominately to intermediate and possibly high risk men (515). A number of randomised trials also now exist which, to some extent, cover this group of men. The mature RTOG 86-10 trial examined the role of 2 months of complete AS before as well as during radiotherapy in patients with bulky primary tumours and was conducted prior to the wide availability of PSA. The subanalysis by grade showed significant gains in loco-regional control, cancer-specific and overall survival in the Gleason 2-6 tumours, also further reinforcing that bulky high-grade cancers should be approached as high risk disease. Another trial has recently reported a freedom from failure together with a survival advantage to having 6 months of neoadjuvant AS (NAAS) prior to 70Gy of EBRT in intermediate and high risk men (516). Somewhat disparate though are the results from the Trans-Tasman Radiation Oncology Group (TROG) 96.01 trial which compared radiotherapy of 66Gy alone to the same dose in combination with either 3 or 6 months of neoadjuvant maximum androgen blockade. Looking at the risk subsets of men from this study shows that the benefit shown for the use of 6 months NAAS (particularly in terms of freedom from BF and cancer-specific mortality) was primarily limited to the high risk men. The power to detect a difference, however, was limited by the small proportion of intermediate risk patients (517). Also somewhat confounding matters is a large 2x2 phase III trial of men with a calculated risk of nodal positivity >15% showing that 4 months of AS was a significant benefit only when given prior to and during radiotherapy (rather than adjuvantly), and furthermore, only when combined with whole pelvic radiation fields (rather than prostate only) (518). As most of these trials showing a benefit to AS in combination with radiotherapy have employed lower doses of radiation than used presently, the precise indications for combination therapy in this group are unclear, it does appear that a subgroup of men, who usually fit within the typical intermediate risk profile, will benefit from short term neoadjuvant therapy. Toxicity: The major concerns with EBRT are damage to normal tissues at risk viz. the rectum, bladder/urethra as well as the neurovascularity involved with erectile function. Modern planning systems have capabilities to produce detailed descriptions of 55

Chapter 10. PROSTATE CANCER the dose to structures of interest (dose-volume histograms). Applying accurate constraints to these parameters is partly the explanation for the observation that serious toxicity from modern 3D EBRT is uncommon, with evidence suggesting that it is less common now than with previous EBRT planning and treatment techniques, even those previously using much lower doses (519). During a typical radical course of EBRT, symptomatic urinary toxicity occurs in most men which can be a combination of so-called irritative or obstructive symptoms. Grade I-II toxicity typically not requiring any or only minor medical intervention happens in approximately 50%, while serious problems potentially requiring invasive intervention happen in less than 1%. Urethral stricturing is a major concern after radiotherapy in high doses, and occur in approximately <1% presently. Towards the end of a course of radiotherapy, most men will be aware of some rectal urgency associated with frequency. Rectal bleeding during or soon after therapy occurs in 0-2% of men in modern series (519) with the late bleeding rate being approximately 1% also. A measurable, but not clinically problematic, change in bowel habit will be observed in 10-20% (519). Changes in erectile function related to increases in radiation dose are yet to be adequately documented in long-term data from large scale randomised dose trials. Observational series do suggest that in those with adequate erections prior to radiotherapy, 50% will maintain erectile function for more than 2 years (520). The control arm (radiation alone to 66Gy) of a large Australian randomised trial showed that 1 year after radiotherapy, approximately half of the patients sexually active prior to treatment remained active (376). This trial also showed that the addition of AS to radiotherapy does not appear to increase the toxicity of radiotherapy beyond that seen with radiation alone, in agreement with other published results (521).

D. Brachytherapy
Brachytherapy (BT) for prostate cancer can involve 2 forms of interstitial implant therapy, either permanent low dose-rate (LDR) radioactive seed implant or a temporary high dose-rate (HDR) implant.

(i) LDR seed implant
For men with early stage prostate cancer, the use of seed BT has increased dramatically over the past decade with potentially a quarter of all early cancers now being treated this way in the USA (522). The modern technique uses accurate imaging and guidance systems to place metal seeds, containing either Iodine125 or Palladium103 as the radioactive source, in a planned array within the prostate to deliver a conformal dose of radiotherapy. Typically a TRUS unit is used to image the prostate in real-time (with CT or MRI used in some centres) and a template utilised to guide needles loaded with the seeds into the prostate transperineally. The procedure is usually relatively simple and brief from the patient’s viewpoint, often administered in an outpatient setting, making it attractive to many. A number of series have now reported both PSA and toxicity outcomes for many thousands of men treated this way (523, 524). Although the longterm results published for this procedure amount to beyond 10 years at only a few centres (525, 526), the tumour control rates appear to be similar to those of surgery. Biochemical control rates for those in the low risk group are presently greater than 85% (524, 525, 527), and shown to be equivalent or better than results of similar EBRT treated cohorts. Moving beyond the low risk category, the use of seed BT as the sole treatment modality almost universally shows an increase in the biochemical failure (BF) rates. In these cases, recommendations range from using seed BT in combination with EBRT, in combination with AS, or using EBRT only. No randomised evidence exists to help with this decision, however, there is considerable experience using seed BT in combination with EBRT indicating potential for good outcomes in this group 56

Chapter 10. PROSTATE CANCER (528). High risk men typically do poorly with seed BT alone and are rarely considered suitable for this therapy. Toxicity Following a seed implant, the dose of radiation is delivered over some months, depending on the half-life of the isotope. The predominant toxicity of the implant is that of urinary irritation, and typically the symptoms peak at 2-10 weeks (isotoperelated) and the majority will have returned to normal at 12 months (529). These initial symptoms can be troublesome, with most series reporting rises in the International Prostate Symptom Score of 7-12 units at their worst, indicating moderate bother (373, 530, 531). Rectal toxicity is uncommon, with bleeding or fistula formation occurring in <1% (532). Erectile function maintenance has been reported in a number of prospective brachytherapy studies. The highly conformal dose distribution of seed BT theoretically may reduce the dose of radiation to the structures important for maintaining erectile function, and this seems to be demonstrated in the relatively early figures. The loss of adequate erectile function rates appears to occur in 30-50% of patients by 3 years (533, 534) and longer term results are awaited.

(ii) HDR temporary implant
HDR brachytherapy of the prostate has been utilised in a number of centres now for well over a decade (535-541). Proponents of the technique are in favour of the highly conformal nature of the radiation dosimetry able to be achieved with HDR, as well as a theoretical biological advantage in prostate cancer (542). It is performed by placing a series of fine BT catheters in the prostate transperineally, using a technique similar to the guidance of the needles for LDR seed BT. These are subsequently loaded with a radioactive source which sits at various positions in the catheters for various durations to build up a dose of radiation as predetermined by a computer planning system. It is typically used in combination with a truncated course of EBRT, although there are experimental protocols in place for using it as monotherapy. Results using this technique are encouraging, and would appear to justify the theoretical basis of its use. Series routinely report freedom from biochemical failure results of better than 90% and better than 80% for low and intermediate risk patients respectively (537, 539, 540). Late urinary complications (grade 3) appear to occur in about 5%, and are typically ‘obstructive’ symptoms. Significant rectal toxicity is uncommon; the maintenance of erectile function is poorly reported to date.

E. Defining biochemical failure (BF) after radiotherapy
The definition of therapeutic failure based on the post-radiotherapy PSA profile (“biochemical failure”) is a more difficult prospect than following radical prostatectomy. After surgery, essentially all the PSA-secreting tissue is removed, and hence an undetectable level is considered desirable. Naturally, this situation does not exist following radiotherapy; with the prostate remaining in situ with variable residual function. Early after the inception of PSA testing, the publication of BF results was based on a variety of interpretations which were felt by individuals to describe a high likelihood of recurrent disease. In response to concerns regarding the potential lack comparability of such published results, an ASTRO consensus panel was convened in 1996 to discuss future BF publication criteria. The panel considered data from a number of clinical databases, all treated with external beam therapy. The subsequently published definition of BF, known as the ASTRO consensus definition (ACD), stipulated BF to have occurred when three consecutive PSA rises are seen and the date of failure to be retrospectively backdated to the 57

Chapter 10. PROSTATE CANCER midpoint of the PSA nadir and first rise. This was largely based on the finding that three consecutive PSA rises was followed by a further rise in 89% of cases, plus the clinically plausible argument that the recurrence actually commenced well prior to the confirmatory third rise. Relatively little was known at the time regarding the relationship between the ACD BF status and clinical endpoints (543). The ACD has been widely embraced for the analysis of post-radiotherapy results. A large number of concerns about its performance have been raised however. Although BF would appear to occur rarely after 5 years using the ACD, the reliability of the results are known to be highly dependent on the duration of follow-up. Practically this has shown to potentially worsen the derived freedom from BF (FFBF) rate by approximately 25% when the same data is analysed at a median follow-up of 6 years against that obtained at 2 years (544). This creates difficulty when examining the results of more recent series (for example, a new treatment technique) with historical data; the contemporary data will be favourably biased. This phenomenon is known to be an artefact of the backdating in the ACD. Furthermore, the known independent prognostic value of ACD BF and overall survival (relative risk 1.27) is substantially weakened by backdating (545). The ACD has also been criticised for having unpredictable results when applied to treatment modalities other than EBRT, for which it was not planned to be utilised. False positive results can be problematic with seed brachytherapy, where benign PSA fluctuations (“bounces”) are seen in approximately one-third of cases (546). Similarly, when the testosterone recovers after combined AS and radiotherapy there may be a concurrent PSA rise of a level sufficient to trigger ACD BF in up to approximately 20% of cases (547, 548). There is poor applicability to surgical data, with backdating particularly creating differences in the late risk of failure (468). A further surgical analysis suggested that the ACD may underestimate the risk of BF by as much as 30%, however this study illustrates the propensity for the ACD to be misinterpreted (549). By stipulating the nadir date to be the date of surgery, rather than the defined last non-rising PSA date, the backdating effect was exaggerated to the theoretical extreme, further reinforcing the finding that the there are multiple areas of potential misinterpretation in the ACD (550), and that original publication guidelines are poorly adhered to (551). A number of analyses of very large and mature patient cohorts have been published which suggest alternative BF definitions may be superior to the ACD. These have focused on the correlation of PSA dynamics and subsequent clinical events in the aim of finding a BF definition which is both sensitive in detecting recurrence early, yet specific to those factors which are likely to be clinically relevant in the future – data which was unknown at the time of the original consensus meeting. The most ambitious undertaking was that of Thames et al (2003) (552). This study assimilated PSA and clinical outcomes of 4839 T1-T2 prostate cancer patients from nine USA institutions treated with EBRT without planned AS. A total of 102 different definitions of BF were assessed using a variety of quantitative measures of prediction of clinical failure. A number of definitions were found to have substantially better performance than the ACD. In particular, nadir (the lowest level to date) plus either 2 or 3 ng/mL showed good sensitivity, specificity and predictive indices in relation to clinical failure without resorting to backdating and prospective in nature (that is, it used only the PSA data available prior to failure, as would happen in the clinic). The nadir + 2 ng/mL definition, for example, showed a sensitivity, specificity and relative risk of clinical failure of 0.74, 0.84 and 58.4 respectively, while the ACD had respective values of 0.61, 0.80 and 7.0. Similarly, Kestin et al (2004) (553) used the mature data of 727 men to show the nadir + 2 ng/mL definition had a 73% versus 3% rate of clinical failure at 10 years for those who were deemed as BF or not respectively, while the ACD had 64% against 14%. Further analysis of the sensitivity and specificity characteristics using receiver operating characteristics show that the optimum balance appears with the nadir + 2 ng/ml or an absolute level of 3 ng/ml definitions (554). Analysis of the interaction of various BF definitions and AS therapy have been per58

Chapter 10. PROSTATE CANCER formed by Pickles et al (2003) (548). The nadir + 2 ng/mL performed better in terms of predicting clinical failure in their cohort of 1490 patients, and especially so in those previously receiving AS, and this was also confirmed by Zietman et al (2005) (547). It would also appear that the definition of the nadir + 2 ng/mL style will be relatively unlikely to be influenced by PSA bounces related to seed brachytherapy, given the median bounce level is 0.4-0.7 ng/mL (546) and less than 3% are of sufficient level to trigger a false positive BF with this definition (555). In view of the considerable advances in the understanding of PSA biology following radiotherapy, another consensus meeting to discuss an update to the ACD was convened in January 2005. Results from this meeting are expected soon.

F. Management of local failure after radiotherapy
A detectable and rising PSA above the nadir value following radiation therapy may represent local or distant failure, or both. Determining the site or sites of tumour, although highly desirable, is notoroiously difficult with currently available diagnostic techniques with BF often preceding clinically evidence of metastases by several years. Consequently, a diagnosis of localized disease without metastases, can be only inferred. However, selecting the most appropriate therapeutic approach may be aided by reference to initial clinical staging, PSA and Gleason score from TRUS biopsies. In addition to pre-treatment clinical and post-treatment pathological factors, PSA kinetics and, specifically, a pretreatment PSA velocity >2 ng/ml/year, an interval to PSA failure <3 years and a post-treatment PSA doubling time <3 months place a man at increased risk for metastases and subsequent prostate cancer-specific mortality, making these patients poor candidates for local-only salvage therapy. Men with any Gleason 8-10 tumour in their TRUS biopsy cores or those whose cores indicated seminal vesicle involvement are also very likely to have non-localised disease. As a result, Lee and D’Amico (2005) (556) identified an optimal candidate for local-only salvage therapy as one whose pretreatment PSA velocity was <2 ng/ml/year, interval to PSA failure exceeded 3 years, post-treatment PSA doubling time was at least 12 months and whose TRUS biopsy cores contained a Gleason score of less than 8 (556). Touma et al (2005) (557) reviewed the current status of local salvage therapies following radiation failure for prostate cancer, reporting that salvage RRP offered 5-year biochemical relapse-free rates of 55 to 69%. They identified a life expectancy of at least 10 years, pre-radiation and pre-operative PSA <10 ng/ml and localized preoperative stage as desirable factors for RRP after radiotherapy failure. However, patients need to be aware that higher complication rates are to be expected compared with primary RRP, which include rectal injuries, bladder neck contracture and urinary incontinence (557). As discussed above in Section VII, the most appropriate application of cryotherapy is for patients with bulky local disease and local recurrence after radiation therapy (315). Touma et al (2005) (557) noted that urinary incontinence, impotence, pelvic pain and urinary retention were the major side effects of salvage cryotherapy cited and considered salvage cryotherapy to be a valid option in hormonally naïve patients with preoperative PSA <10 ng/ml, Gleason score <8 and clinical stage <T3, being especially suitable for older men with some comorbidities but who are still considered to be reasonable anesthetic risks (557). HIFU (also addressed briefly in Section VII) may have a role in these types of patients but relevant data to make this assertion are not yet available (313). Apart from local treatments, androgen suppression treatments may be considered (detailed in Section XIII, below).

G. Quality of life issues in prostate cancer radiotherapy
Considerable interest exists regarding the treatment modality in early prostate cancer, and the impact it has on the patient’s quality of life (QOL). Once more, randomised 59

Chapter 10. PROSTATE CANCER data are lacking, although at least three large prospective series have now been published comparing treatment modalities (413, 558, 559). EBRT was typically associated with the most bowel dysfunction, which settles over 1-2 years, and then tends to be stable. Acute urinary bother is mild, but significant worsening of symptoms at later stages (beyond 5 years) has been reported and requires further investigation (559). Sexual function consistently deteriorated over time in all series after EBRT, although notably, at the same rate as the untreated controls of one series. Seed BT has initial elevation of urinary bother that settles to a large degree in the first 2 years. There is however, the suggestion that there may be subsequent deterioration of continence over the next 4 years. This has not been corroborated in other studies to date (560). Rectal morbidity, although typically minor, is documented to steadily settle with time. Erectile function deteriorates in the initial 1-2 years typically, but the longer term results of one series suggest little change in the QOL impact of this beyond that time. Each of these series had a RRP control arm for comparison, which yielded consistent findings. RRP always led to a high QOL impact in terms of urinary continence and sexual dysfunction within the first 6 months post-therapy. These both improve over the first 2 years, but should be considered likely to remain stable beyond that time. Long-term follow-up comparisons in well designed trials will be required to ultimately decide the overall QOL trade-offs with each particular therapy.

XIII. SYSTEMIC THERAPIES
Although much attention is focussed on early diagnosis and localised disease, a considerable proportion of patients continues to present with extra-prostatic disease (561). In addition, approximately 25% of men treated initially with curative intent experience prostate specific antigen (PSA) failure or develop clinically detectable metastases within 5 years following therapy (562). The cornerstone of treatment for patients with prostate cancer metastases is hormonal manipulation: approximately 80% of patients have a durable clinical regression with androgen suppression therapy.

A. Hormonal Therapy
It is now well over 60 years since Huggins and Hodges reported their observation (308, 309) that prostate cancer was an androgen dependent tumour which regressed following bilateral orchidectomy. In the interim, a number of alternative methods have become available for achieving castrate testosterone levels. These include bilateral orchidectomy/orchiectomy (regarded as the reference treatment), Luteinising Hormone Releasing Hormone (LHRH) agonists, antiandrogens and oestrogens, all of which have significant adverse or unwanted effects.

(i) Bilateral orchidectomy/orchiectomy, LHRH agonists & antiandrogens:
Historically, bilateral orchidectomy has been the reference treatment for advanced prostate cancer which, by targeting the major source of androgen production, dramatically diminishes the number of tumour epithelial cells and accompanying neovasculature. LHRH agonists, after initially stimulating LH and hence testosterone synthesis by the testicles, occupy LHRH receptors to prevent subsequent production of both these hormones. In addition, LHRH agonists can have a a further action by targeting LHRH receptors on prostate tumour cells (563, 564). The unwanted effects of bilateral orchidectomy and LHRH agonists are generally considered to be comparable apart from the fact that the former requires an opera60

Chapter 10. PROSTATE CANCER tion, with physical and possible psychological consequences, and the latter necessitates regular interval injections and commencement of therapy has an accompanying risk of an initial surge in testosterone (the so-called flare reaction) due the drug initially stimulating production of luteinising hormone before blocking production. Reduced libido and impotence are to be expected following surgical or medical castration together with a loss of bone substance and muscle mass. As a condition, osteoporosis is underdiagnosed in men and it is only relatively recently that its relationship with bilateral orchidectomy and LHRH agonists has been appreciated (565567). The high prevalence of osteoporosis together with its debilitating consequence of bone fractures has prompted advocacy for the use of bisphosphonates in prostate cancer patients, especially those committed to longstanding castration (see below). Hot flushes can be problematic for many men and tiredness from anaemia can compound debility. A minority of men receiving LHRH analogues develop gynaecomastia. The American Society of Clinical Oncology recommends bilateral orchidectomy or LHRH agonists as initial androgen suppression treatments (568). Non-steroidal antiandrogens may be considered alternatives but the steroidal anti-androgen cyproterone acetate should not be offered as monotherapy (568). The UK Committee on the Safety of Medicines recommends that, because of the risk of hepatotoxicity, cyproterone use in prostate cancer should be restricted to short courses unless patients are unresponsive to or intolerant of other treatments. Unwanted effects with nonsteroidal antiandrogens are common with gynaecomastia and breast pain troublesome for many: a cessation rate for these medications has been reported to be 4-10%, in particular with flutamide (568, 569). Hepatotoxicity is a potential problem with all antiandrogens but especially cyproterone acetate.( 568). Green et al (2004) (570) reported the results of our study of 82 men randomised to leuprorelin (Lucrin™), goserelin (Zoladex™), cyproterone acetate (Androcur™) or watchful waiting, 62 of whom completed 12 months of follow-up. In addition to a non-treatment (watchful waiting) control group, a non-cancer community reference group was also evaluated. Findings were compared in relation to cognition and quality of life. Most patients had serum PSA levels between 30 and 60 ng/ml at baseline. Using well-validated and established instruments, they found that ~50% of men in all treatment groups, but none of the controls, had significant cognitive deterioration at 12 months, in particular in relation to complex information processing. Although the cognitive defects were of a magnitude comparable with sleep-deprivation or mild inebriation, there was no consistent association between subjective cognitive changes and objective deficits. In addition, there were more instances of a deterioration of quality of life for the men on hormonal treatments at 12 months, in particular in relation to sexual function (570). Unlike other studies addressing cognition and quality of life, Green et al’s paper is important as all patients were randomised to management regimens which included a non-drug (control) arm and this study was completely independent of any industry-sponsorship. Consequently, these findings are much more compelling than those reported from less robustly designed studies and those trials sponsored or supported by industry (571-577)

(ii) Oestrogens
Oral oestrogen therapy, the most common form of androgen suppressive medication for many years, is now rarely used as first line hormonal treatment because of associated cardiovascular complications with oral administration, although this route of delivery is employed not uncommonly in Japan for a short period to offset the flare effect of LHRH agonists (578). An increased cardiovascular morbidity was reported to be present with the oral form of oestrogens even in patients without overt cardiovascular disease affecting one quarter of such patients during their first year of treatment (579). However, a dose-response relationship is said to be present in terms of cardiovascular morbidity and mortality with one mg daily of diethylstilboestrol 61

Chapter 10. PROSTATE CANCER (DES) (with or without aspirin) stated to be comparable with bilateral orchidectomy in the treatment of advanced disease but without increased cardiovascular complications. Klotz et al (1999) (580) found that venous thrombosis was not prevented when DES was prescribed together with low-dose warfarin (580). The increased susceptibility to cardiovascular complications in patients taking oestrogens appears to be critically related to the route of administration. This predeliction is reported to be significantly reduced (581) or avoided by parenteral delivery. An increased synthesis of coagulation factors, in particular Factor VII, results from oral oestrogen therapy and this is thought to be responsible for the increased rate of cardiovascular problems in these patients (582, 583). Consequently, many investigators have advocated a re-evaluation of oestrogen treatment (584-587), especially since all forms of oestrogen, including parenteral and transdermal patch preparations, are cheap (588) and, unlike bilateral orchidectomy and LHRH agonists in particular, this medication is not considered to induce osteoporosis ( 581). Furthermore, oestrogens may have a role in ameliorating agitation in some men receiving LHRH agonists. Ockrim et al (2003) (589) reported their experience with 20 men with newly diagnosed locally advanced or metastatic prostate cancer treated with transdermal estradiol patches in particular in relation to bone mineral density. They found that at 1 year that the mean bone mineral density had increased by 3.6% and concluded that transdermal estradiol protects against bone loss in men with prostate cancer and may improve bone density in those at risk for osteoporotic fracture. Local experience with this form of delivery has been that patch displacement can occur with sweating, especially in active men: serial serum testosterone levels may be used to optimise the frequency and duration of patch application.

(iii) Commencement of hormone therapy
Despite many attempts to demonstrate otherwise, there is no clear evidence that commencing androgen suppression therapy early improves survival (568, 590). Early commencement does, however, increase the likelihood and duration of unwanted effects. Since in terms of lifestyle effects, commencing androgen suppression is likely to adversely affect libido, potency, physical mobility and strength, body habitus, cognition and liver function (with anti-androgens), a body of clinicians advocates carefully delaying commencement by balancing the unwanted effects of treatment with those of the disease being treated. In conjunction with patients’ wishes, PSA doubling times and development of lesions on bone scans can be helpful in deciding when to commence androgen suppression. By upsetting homeostasis at a molecular level, androgen suppressive therapies may in fact contribute to tumour progression in those remaining prostate cancer cells after commencement of treatment (38). Unlike androgen receptors which are sited in prostate epithelial cells, oestrogen receptors are present in both epithelial and stromal cells. Compartmentalisation of the two ERs is reported with ERα exclusively in stroma and ERβ predominantly in the epithelial compartment (38, 591). ERβ is considered to have a role against prostate cancer dedifferentiation in contrast to the proliferative effect of ERα and androgens (592). In prostatic epithelial cells, testosterone is reduced by 5α-reductase to dihydrotestosterone which, in turn, is converted into 3α-diol and 3β -Adiol. Unlike testosterone and dihydrotestosterone, these two metabolites do not bind to the androgen receptor but possess high affinity for oestrogen receptors. By binding to estrogen receptor β (ERβ ), 3β -Adiol induces expression of the cell adhesion molecule E-cadherin, the presence of which in prostate cancer cell membranes is associated with a less aggressive phenotype. Decreased expression of many C-CAMs including E-cadherin, have been associated with the progression of prostate cancer (38,107). However, there are many clinicians and patients who become pre-occupied with lowering the serum PSA levels at all costs. A survey of American Urologists indicated that 68% recommended hormone suppression therapy for an elevated PSA after radical prostatectomy (562, 593). Thus, an increasing PSA often serves as the trigger for 62

Chapter 10. PROSTATE CANCER commencement of what translates into long-term androgen suppressive therapy for a large proportion of patients in spite of a lack of clear evidence of a survival benefit with early treatment (594, 595).

(iv) Combined androgen Blockade (CAB)/Maximum Adrogen Blockade (MAD)
Since bilateral orchidectomy and LHRH analogue treatment address testicular androgen production exclusively and a small contribution to overall androgen levels is made by adrenal androgens, CAB was initiated. However, despite extensive trialling, only a modest survival benefit has been demonstrated but at the cost of a higher sideeffect profile for patients (596). The limited survival benefit appears to be associated with the use of non-steroidal antiandrogens and only becomes evident after 5 years of therapy. This topic has been reviewed recently by Loblaw et al (2004) (568).

(v) Intermittent Androgen Blockade (IAB)
Intermittent hormone therapy was instituted to lessen the duration of unwanted effects from androgen suppression therapy, usually being limited to those patients who demonstrate a pronounced PSA response and find the unwanted effects of androgen suppression problematical. However, this approach to androgen delivery begs the question, why was the treatment started when it was since, most often, IAB seems to be used for men who commenced their androgen suppression very early. Other issues with IAB are that not all the adverse affects of androgen suppression are reversible and recovery of the hypothalamic-pituitary-testicular axis is variable, especially after prolonged LHRH administration. Indeed, castrate levels of testosterone and LH may persist for up to 1 year (or even longer) after discontinuing LHRH agonist treatment (597).

(vi) Hormone escape
Once prostate cancer metastasizes following effective androgen suppression therapy, its toll, with regard to pain, suffering, and disability, can be considerable due to incapacitating sequelae of disseminated and hormone-refractory disease, for which no curative treatments currently exist. PSA evidence of hormone escape usually precedes other barometers of disease progression and metastasis. Newling et al (1993) (598) reported that, for men with newly diagnosed metastatic prostate cancer who were randomised in the EORTC study 30853 to receive goserelin and flutamide or bilateral orchidectomy, the median time for survival following PSA progression was 52 weeks compared with 41 weeks for bone metastases and 28 weeks and 33 weeks for progression of regional and distant lymph nodes, respectively (598). However, as suggested by a reducing mortality rate for this condition in many countries, these estimates may no longer be accurate for a variety of reasons addressed throughout this chapter. As outlined above, prostate cancer has a particular predilection for androgens that serve as this malignancy’s preferred ligand. When androgen suppression is invoked initially, the tumour regression effect can be dramatic with both epithelial cells and tumour neovasculature being affected significantly (599). However, tumour repression is not permanent with a median time to relapse of 18 months (600). Nevertheless, maintaining suppression of circulating androgens remains important as the androgen receptor pathway continues to be very active in ‘hormone escape’ patients. Amplification and overexpression of the androgen receptor gene, as well as posttranslational modifications to the AR occur, resulting in the cancer cells becoming ‘super-sensitive’ to androgens (601-603). In conjunction with these changes is activation of other genes that, through signal transduction pathways, facilitate receptivity to a variety of ligands including other 63

Chapter 10. PROSTATE CANCER hormones and drugs, especially antiandrogens (604). Consequently stopping an antiandrogen may afford a temporary respite to tumour regression manifested by a reduction of serum PSA in some patients (605). Paradoxically, for those patients receiving monotherapy in the form of LHRH agonists or a previous bilateral orchidectomy, the addition of an antiandrogen may cause a clinical regression for a short time and this is often recommended before proceeding to chemotherapy or radiotherapy in the form of radio-isotopes or local external beam treatment to isolated troublesome secondary deposits. Recent research has provided support for modulation of the oestrogen receptor axis in disease no longer responding to androgen suppression by LHRH agonists, bilateral orchidectomy or antiandrogens by targeting oestrogen receptors in metastatic disease (606). Hormone-refractory prostate cancer is said to be present when there is evidence of progression despite the use of first and second line hormonal manipulation. Although the administration of bisphosphonate and chemotherapeutic interventions may provide benefit in the short term (see below) these palliating approaches merely serve to temporise the situation in this subterminal/terminal phase of the disease. Clarke (2003) (607) divided the urological issues to be considered into: 1. Lower urinary tract dysfunction 2. Ureteric obstruction 3. Skeletal, dysfunction 4. Bone mrrow insufficiency 5. Lymphoedema 6. Rectal obstruction/infiltration 7. Pain 8. Psychological dysfunction/impaired quality of life. It is this last point in particular which needs to be considered paramount so that unreasonable attempts to prolong life are not undertaken, especially when these are not in concert with individual patient’s wishes. For example, it may be preferable not to treat ureteric obstruction and allow the patient to die painlessly from uraemia than protract his demise for a short but miserable period by instigating various interventions. A number of other disciplines are often involved at this stage as indicated, which include medical and radiation oncology, interventional radiology, pain management specialists and palliative care clinicians amongst others. It is important that pastoral care support is available as appropriate. A further web-site is: http://www.cancer.gov/cancertopics/understanding-prostate-cancertreatment/page6

B. Chemotherapy
In reports published to 1991, the rates of objective clinical response of prostate cancer to available cytotoxic agents were disappointingly poor. A summary overall response rate of just 8.7% (with a 95% confidence interval of 6.4-9.0%) was responsible for Yagoda and Petrylak (608) concluding hormone-refractory prostate cancer to be unresponsive to conventional chemotherapy of the time. Initial randomised data for the use of mitoxantrone chemotherapy in prostate cancer (thought to be the most effective agent of the period) became available in 1996 with the publication by Tannock et al (609). By randomising 161 men with symptomatic hormone-refractory disease, they were able to show a significant reduction in the need for analgesia (the primary end-point) in those who received mitoxantrone plus 64

Chapter 10. PROSTATE CANCER prednisone compared with the control arm of prednisone only. Analgesic responses were also substantially longer in the chemotherapy arm. In a subsequent CALGB trial looking at survival end-points, no survival benefit was shown when mitoxantrone was added to hydrocortisone, although there was an overall improvement in the quality of life and duration of analgesic response (8 months on average) in those on the chemotherapy arm (610). Further preclinical activity had also been suggested when prostate cancer was exposed to taxanes, a class of microtubule stabilising agents. Predominately, these agents work by blocking the ability of cells to depolymerize the microtubule cytoskeleton during normal mitosis, thus inducing a lethal cell cycle arrest at the G2M phase. In particular, docetaxel appeared the most active form and was taken to clinical testing. A number of phase I/II trials were conducted using either single agent docetaxel or in combination with estramustine, an agent thought to possibly potentiate the taxane effect by acting on a different microtubule pathway. Used as single agent therapy, docetaxel showed significant PSA responses (a PSA level decline of greater than 50%) in 38-48%, and up to 68% in the combination therapy trials (611). Responses in measurable soft tissue disease ranged from 20-55%. Neutropenia was a feature in over two-thirds of the patients treated on a three weekly single agent docetaxel schedule, and less common when given weekly. Similar toxicity was seen with the combination therapy trials, except for the addition of thrombo-embolic complications in up to 10% related to the estramustine. Large randomised trials to comprehensively test the efficacy of docetaxel in advanced hormone-refractory prostate cancer, in comparison with mitoxantrone, have now been published. The TAX 327 trial of Tannock et al (612) randomly allocated 1006 patients to one of three treatment arms, the ‘control reference’ being three weekly mitoxantrone plus prednisone (M/P): the docetaxel arms were given either weekly for 5 of 6 weeks or three weekly with prednisone (D/P). The three weekly schedule of D/P showed a significantly improved overall survival compared with M/P, with the risk of death reduced by 24% (95% CI 0.62-0.94, p=0.009) leading to a median survival prolongation of 2.5 months (18.9 against 16.5 months). Other significant benefits were seen in the decrease of PSA levels, analgesic responses and the patientreported quality of life assessment in the D/P group. Toxicity was lowest in the M/P arm, while the highest in the weekly D/P group. As this group did not, however, show a survival benefit, it was concluded that the three weekly schedule of D/P was optimal. The other phase III trial was conducted by the SouthWest Oncology Group (SWOG) and accrued 674 eligible patients to either receive M/P or docetaxel and estramustine (D/E) three weekly (613). Using the primary overall survival endpoint, a significant benefit was found for the D/E arm, with the median survival increasing from 15.6 to 17.5 months associated with a hazard ratio of 0.80. PSA declines of >50% were seen in 50% of D/E men, and 27% of those on M/P (p<0.001). Pain relief was not significantly different between the arms and there was substantially more toxicity in those having D/E, although the neutropenia rates were comparable. The consistency of the results between these large trials has demonstrated that alteration of the natural history of hormone-refractory prostate cancer can be made using cytotoxic agents, and survival advantages, albeit small, can be made. Therefore, in the context of a patient with a rising PSA on hormonal therapy, the first line use of docetaxel-based chemotherapy is considered standard of care by many. For those not medically suitable for this therapy, mitoxantrone-based therapy should still be considered for its known palliative capacity in those with symptomatic disease, along with lesser toxicity.

C Bisphosphonates
Bisphosphonates were initially examined for the prophylaxis and treatment of os65

Chapter 10. PROSTATE CANCER teoporosis, particularly in women (614) but, more recently, for osteoporosis in men following androgen deprivation +/- external beam radiotherapy (614, 615). Bisphosphonates inhibit osteoclast activity so, in addition to their potentially protective effect with respect to the development of osteoporosis there is a possible role in metastatic bone disease. Osteoclast activity is an integral part of the metastatic process for both osteolytic and, more commonly, osteoblastic bone metastases in prostate cancer (616). Most experience with bisphosphonates in cancer has been with multiple myeloma and breast cancer patients but results from these studies cannot be extrapolated to men with prostate cancer. In addition, not all bisphosphonates are equal with studies with the first generation compound Clodronate failing to show a clear advantage compared with placebo (617). However, the third generation bisphosphonate zoledronic acid did demonstrate increased apoptosis in prostate cancer cell lines in vitro and inhibited growth of osteoblastic and osteolytic metastases in vivo (618). Furthermore, Zoledronate has been shown recently to expand γδ T cells which exhibit cytolytic activity independent of MHC (see next section; (619). Saad et al (2002) (620) reported their experience with 4 mg and 8 mg of zoledronic acid given intravenously 3-weekly over 5 minutes initially, but subsequently 15 minutes to increase renal safety, in a double-blind randomised controlled trial for 15 months. A total of 643 patients with documented bone metastases were randomised to one of the 3 groups. Only 98/214 (45.8%) and 77/221 (35.3%) of the patients who initially received 4 mg and 8 mg of zoledronic acid, respectively, received at least 12 months of study drug compared with 77/208 (37%) randomised to placebo. The 8 mg dose was reduced to 4 mg during the study because of renal toxicity. The reasons for discontinuation were withdrawal of consent, adverse events and death, most common in the 8/4mg zoledronic acid group, and unsatisfactory therapeutic effect, especially in the placebo group. During the study, at least one skeletal-related event occurred in 71 (33.2%) compared with 92 (44.2%) of patients randomised to 4 mg zoledronic acid and placebo, respectively. Pain and analgesic scores increased more in patients who received placebo than zoledronic acid but there were no differences in disease progression, performance status or quality-of-life scores among the groups (620). Saad et al (2004) (621) subsequently reported the results from 122 men who completed a total of 24 months on study. Fewer patients in the 4-mg zoledronic acid group than in the placebo group had at least one SRE (38% versus 49%). The median time to the first skeletal related event was 488 days for the 4-mg zoledronic acid group versus 321 days for the placebo group (P =.009). Compared with placebo, 4 mg of zoledronic acid reduced the ongoing risk of SREs by 36%. These authors concluded that long-term treatment with 4 mg of zoledronic acid is safe and provides sustained clinical benefits for men with metastatic hormone-refractory prostate cancer. Since the optimal timing for commencing administration of zoledronate may be at an earlier phase in the disease, trials are underway with patients with hormone sensitive rather than refractory disease.

D Bone-seeking radio-isotopes
Widespread skeletal metastases may not be easily amenable to relatively localised external beam radiotherapy. The systemic use of bone-seeking radio-isotopes can be useful in this situation, with proven efficacy for two agents – Strontium89 and Samarium153-ethylenediaminetetramethylene (EDTMP). Strontium89 (Sr89) is an agent which behaves biologically in a manner analogous to calcium (with which it shares a relationship on the periodic table of elements), and hence is incorporated into sclerotic bone metastases avidly. It is a pure beta emitter, with a half-life of approximately 50 days, although half is typically excreted from the body in 2 weeks (mainly in the urine). Samarium153 is chelated to EDTMP to enable preferential binding to bone. Physically it has a much shorter half-life (46 hours) than strontium89 and has gamma as well as beta emissions, enabling imaging on 66

Chapter 10. PROSTATE CANCER gamma cameras. Producing radiation damage for only a short distance in tissues, these agents concentrate effect in bone metastases with little deposition of dose in soft tissues. Following intravenous administration, maximal effect on pain is usually seen in 24 weeks. Randomised trials comparing Sr89 against or in combination with EBRT showed no significant difference in analgesic efficacy, although the occurrence of new areas of pain and analgesic requirements was significantly reduced in two randomised series (622, 623), although one other suggested inferiority to EBRT (624). Similar efficacy is reported for Samarium153 (625, 626), with both typically reducing pain to some degree in 70% of patients, an effect which lasts for 3-4 months on average. A small percentage of patients will experience a flare (a temporary increase) in pain in the first week. The predominate toxicity of these agents is that of bone marrow suppression, and in particular, most patients will have a measurable decrease in the platelet count (on average a 30% drop) or white cell count. Repeat doses must therefore be given with caution (especially within 3 months) and close monitoring of the blood count will be required. This will be a prime concern in patients being considered for chemotherapy. These agents are presently being further investigated for their efficacy in combination with chemotherapy (627, 628).

E. Emerging Therapies: Vaccines in Prostate cancer
(i) Introduction
Of the various treatment approaches being examined for prostate cancer, nothing seems to have captured the public’s imagination quite as vividly as have vaccines. Although considerable advances have been made in understanding the processes involved with different vaccine approaches, overall clinical results remain modest so, with changes being implemented continually this form of therapy must still be regarded as experimental. Historically, the prostate was considered to be an “immunologically privileged” site (629, 630) and was regarded as lacking a network of intraprostatic (631) and afferent lymphatics (632) . Although neither McCullough nor Gittes was able to demonstrate extraprostatic lymphatic drainage following intra-glandular injections of iodinated emulsified oils and India ink carbon particles, respectively (629), Gardiner et al in 1979 described lymphoscintigraphic evidence of lymphatic drainage following intraglandular injections of technetium labelled antimony sulphide colloid into clinically benign prostates of volunteer patients (633, 634) That intraprostatic lymphatics do exist was elegantly demonstrated by Zeng et al (2004, 2005) (635, 636) who showed recently that, not only does the prostate contain lymphatics, but peritumoural lymhangiogenesis is demonstrable within prostates harbouring cancer. Furthermore, they found a relationship between peri-tumoural lymphatic vessel density and the presence of lymph node metastases (635, 636). Further endorsement of Zeng et al’s findings (635, 636) in relation to finally destroying the myth of the prostate being an ‘immunologically privileged’ site, is Vesulainen et al’s report that lymphocytic infiltrates correlate with an improved 10-year survival for patients with primary prostate cancer (637). This observation by Vesalainen et al (1994) (637) also serves to encourage approaches to enhance lymphocytic infiltration with enhancement of its cytotoxic activity through prostate cancer immunotherapy.

(ii) Effector arms available for recruitment for cell killing in tumour vaccines
Active vaccination strategies can exploit a number of candidate immunological cells as effectors or mediators for immunological therapies (638). These include: 67

Chapter 10. PROSTATE CANCER 1. B cells with production of antibodies 2. Monocytes/macrophages 3. Natural killer (NK) cells 4. NK T-cells 5. γδ T-cells 6. αβ Cytotoxic CD8+ T-cells 7. αβ CD4+ Helper T-cells A primitive level of tumour target recognition can be employed by monocytes, involving interactions with overabundant or aberrant cell surface molecules on cancers. NK cells, which provide the earliest effector mechanism against disseminated blood-borne metastases, identify absence of self on the basis of aberrant or absent expression of major histocompatibility (MHC) class I antigens. NKT and some γδ T-cells subsets recognize lipid antigens presented on CD1c, which is related in evolution to the major histocompatibility complex (MHC) molecules (see below). The balance between interactions with numerous inhibitory and activating receptors for MHC and other MHC-related molecules governs the outcome of NK-mediated recognition of tumour cells. However, the major roles of these more broadly reactive (and therefore immediately responsive) cell types in tumour immunity may be in directing the initial phase of activation of more specific effector arms of adaptive immunity, namely αβ CD4+ and CD8+ T-cells.

(iii) Methods of cell killing
1. Antibody-directed cell killing The interleukins (IL) 4, 5, 6 & 10 activate or influence B lymphocyte differentiation to antibody-making plasma cells. However, cell killing via antibodies is dependent upon more than just attachment of the binding Fab fragment of the immunoglobulin molecule to its specific target epitope. Additional requirements include activation of complement to induce phagocytosis, or interaction with NK-cells, polymorphonuclear leucocytes, or monocytes to provide antibody dependent cell-mediated cytotoxicity (ADCC). These mechanisms may also contribute to the success of clinically available immunotherapies based on the passive administration of monoclonal antibodies to destroy targeted cancer cells, such as Herceptin®, used in cancer of the breast. Unlike in breast cancer, Her-2neu does not appear to be overexpressed frequently in prostate cancer (639) so that the drug has little role here. Specific antibodyrelated mechanisms which trigger cell death in prostate cancer are under development, but are not yet available clinically. Despite correlations between survival and levels of anti-tumour antibodies, there is little evidence that antibodies actively induced by immunization can induce tumour regression.

2. Cell-directed killing Macrophages, which are the archetypical phagocyte, destroy cells by releasing reactive oxygen intermediates and tumour necrosis factor (TNF). By contrast, NK, NKT and cytotoxic (αβ and γδ ) T-lymphocytes (CTL) effect cell killing chiefly through the release of perforin and granzymes, or by ligating Fas on the tumour cell surface to cause apoptosis of targets. As well as malignant epithelial cells, tumour stroma, including endothelial cells, fibroblasts and infiltrating cells, may also be an important target for cell-directed killing (640). Further, interferon γ (IFNγ ) production by T and NK cells may be crucial in tumour immunity, either by inhibiting stromal functions (e.g., angiogenesis) or enhancing tumour cell recognition by CTL, for example by upregulating MHC antigens. 68

Chapter 10. PROSTATE CANCER

(iv) αβ Cytotoxic T-cells, Helper T-cells and Dendritic Cells:
Both αβ CD8+ Cytotoxic T-cells and CD4+ Helper T-cells are unable to recognise naked antigen, requiring it to be presented to them on a platter that, for the former is the MHC class I receptor and for CD4+ Helper T-cells is the MHC class II receptor. The antigen is recognised by CD8+ and CD4+ T cells in the form of short, 8-10 or 12-20 amino acid fragments of proteins, respectively. Critically, the antigen may not be cell-surface associated, but may be derived from any intracellular compartment. Indeed, defective ribosomal products (including mistranslated proteins) are preferential targets for CD8+ T cells (641), while membrane-associated proteins are frequent targets for CD4+ T cells.

Figure 3. Production of peptides for presentation on MHC receptors for T-cell recogntion. The 9-11 amino-acid peptides displayed by MHC class I receptor molecules on the cell surface are generally, but not exclusively, derived from endogenous proteins made by the antigen-presenting cell. A process called cross priming may also allow external proteins to be presented via the class I pathway. The 12-20 amino-acid peptides displayed by the MHC class II receptor molecules on the cell surface are typically derived from extracellular proteins taken up by the antigen presenting cell. CD4 and CD8 T cells are referred to as the adaptive arm of the cellular immune response because the relevant antigen-specific, and naïve yet-to-be-primed, T cells must expand considerably in number before sufficient are present to be effective. This expansion occurs in the specialised environment of lymph nodes, in response to antigen presented by antigen presenting cells (APCs). One obvious possibility is that tumour cells infiltrating lymph nodes might act as APC, and cause this expansion (642). However, most reports support mature dendritic cells (DC) as the only APCs with the ability to prime naïve T-cells, and from the perspective of therapeutic potential, most recent attention has concentrated on directly or indirectly targeting antigen to these. Nonetheless, a recent report indicates that γδ T-cells may also have this capacity (643) being able to simulate mature DC function by processing and displaying 69

Chapter 10. PROSTATE CANCER antigens as well as providing co-stimulatory signals sufficient for strong induction of naïve αβ T-cell proliferation and differentiation. DCs constitute only ~0.2% of the circulating white blood cell population (644) but are distributed throughout tissues where they act as sentinels. They function as biological vacuum-cleaners by pinocytosing, endocytosing and phagocytosing extracellular antigens, processing up to 4 times their own volume of extracellular fluid (ECF) in 1 hour, and converting proteins into peptide (645). A more important source of antigens may be apoptotic or necrotic cells; their display of aberrant surface molecules (such as phosphatidyl serine, usually confined to the inner plasma membrane leaflet), marks them as targets for phagocytosis by DC. DC have the unique property of efficiently cross-presenting antigens: that is, they are capable of processing antigens derived from other cells, and presenting derived peptide fragments on their MHC class I molecules, for presentation to CD8 T cells. Like other “professional” APC such as monocytes, they can also present extracellular antigens via MHC class II, for presentation to CD4 T cells. Highly motile DCs with their captured antigens migrate via lymphatic channels to lymph nodes where the antigen is presented to prime naïve T-cells or re-activate resting memory T-cells. Signalling through pattern-recognition receptors on DC, such as Toll-Like Receptors (TLR), facilitates their migration, and subsequent secretion of IL-2 & IL-12, which appear to be crucial cytokines for the development of a successful cellular response. Early interactions with activated NK cells and CD4 T cells in the lymph node may further contribute to the effectiveness of the priming process (646649).

70

Chapter 10. PROSTATE CANCER

Figure 4. Precise alignment of co-stimulatory & adhesion molecules required for T-cell recognition of presented antigen in MHC receptor

(v) Ignorance, Tolerance to self, and autoimmunity
Autoimmune disorders tell us that the immune system can effectively target self antigens. Some self-reactive T cells escape deletion in the thymus (i.e., central tolerance) and inactivation in the periphery (i.e., peripheral tolerance), for example via regulatory T cells. Insufficient presentation of antigen or access to target tissues could account for “ignorance”, in which potentially self-reactive T cells remain in a resting state. For immunological purposes, most tumour antigens (with the possible exception of mutated or aberrantly translated proteins) fall into these categories. The challenge for tumour vaccine strategies is to convince the immune system that these antigens are legitimate targets for attack.

(vi) Some strategies employed by tumours to thwart cytotoxicity
Successful prostate tumours ‘evolve’ genetically, epigenetically, or environmentally to evade detection and destruction by immune defence mechanisms. Strategies include: 1. Reduced MHC class I expression 71

Chapter 10. PROSTATE CANCER 1. Loss of co-stimulatory molecules 1. Antigen negative variants 1. Production of mucins to disguise antigens 1. Tumour production of Fasl 1. Significantly reduced DC numbers 2. Expression of activation markers by only a small subset of DCs (650) 3. Suppression of T-cell proliferation by PSA in a dose-dependent manner (651)

(vii) Immunotherapeutic strategies
The current status of vaccine studies in prostate cancer was recently reviewed by McNeel and Malkovsky (2005) (652): a brief outline of the various approaches available is provided below. A major limitation with immunotherapy studies is a lack of known antigens recognized by T cells and expressed by prostate cancers. This deficiency is related to the difficulty of generating cell lines from patients’ tumours. Antigens validated thus far (such as hTERT, survivin, PSMA, PSA) have been first proposed on the basis of selective overexpression in tumours, rather than discovered ab initio using T cells recognizing tumour cells. We thus have no knowledge of their relative importance in T cell mediated immunity. Nonetheless, the prostate is a highly specialized organ, and therefore (if tolerance can be overcome) potentially expresses many putative specific targets for an immune response. (a) Passive immunisation Although passive immunization with in vitro-activated anti-tumour T cells is used against other tumours, in particular for lymphomas as a result of Epstein-Barr viral infections following organ transplantation, the relative difficulty of generating and validating anti-tumour T cells has discouraged this approach in prostate cancer.

(b) Non-specific immune stimulants & adjuvants 1. Cytokines The use of cytokines such as GM-CSF and IL-2 therapeutically was initially in the form of unphysiological doses delivered intravenously and, as such, was associated with significant systemic side-effects. More recently, cytokine activities have been harnessed more discretely as part of other treatments rather than as therapies in their own right. GM-CSF, in particular, is an important component in many vaccines, regulating growth and differentiation of haemopoietic cells and acting at several sites in the generation of the immune response. These include activation of ADCC of neutrophils, chemo-attraction of eosinophils and induction of differentiation of DCs (652).

2. Non-specific Adjuvants Flt3 ligand, is a growth and differentiation factor for DCs. In a transgenic murine model, Flt3 ligand on its own was able to result in prostate cancer regression (653) In support of this finding, McNeel et al (2003) (654) reported a marked increase in DCs in the blood with vaccine preparations containing flt3 ligand (654). 72

Chapter 10. PROSTATE CANCER An attenuated form of the tuberculosis bacilli, bacillus Calmete-Guérin (BCG), is a non-specific immune stimulant which was initially developed to vaccinate patients against tuberculosis. The major role for BCG oncologically is as an intravesical agent in superficial bladder cancer where it is administered to minimize tumour recurrences. However, it has potential utility in DC-based vaccines as a non-specific immune stimulant. Another mycobacterium, mycobacterium vaccae was administered in combination with irradiated cell lines by Eaton et al (2002) (655) in 60 men with prostate cancer: although there was absence of a clinical response, some increases in specific antibodies were present in association with T cell proliferation (655). A further non-specific adjuvant we have used in one of our studies is keyhole limpet haemocyanin (KLH). However, one concern with the use of nonspecific adjuvants such as Flt3 ligand, BCG and KLH is that, although they may invoke proliferation and expansion of dendritic or T-cells, these may not induce/include those particularly relevant lymphocytes to produce the desired response for effective tumour cell killing.

3. Specific Adjuvants Ligands for the Toll-Like Receptors (TLRs), pattern recognition receptors of the immune system, have received much attention recently, in particular deoxycytidyl-deoxyguanosin oligodeoxynucleorides (CpG). CpG which mediates its activities via TLR-9 induces DC activation and proliferation, increased co-stimulatory molecule expression and secretion of IFN, Il-1, 6 &12 and TNFα (656). In addition to adjuvants specifically included in vaccine preparations to enhance efficacy, other therapies may interact with patients’ immunological responses. Certain nitrogen-containing bisphosphonates such as paidronate and zoledronic acid are potent stimulators of Vγ 9Vδ 2 cells which constitute the majority of the small population of γδ T-cells in the peripheral circulation. Vγ 9Vδ 2 cells recognise non-peptide antigens and, as they are not reliant upon antigen presentation via the MHC receptor, are able to effect target cell killing rapidly (657).

(c) Active immunization The recent expansion of interest in vaccines has been largely focused on DCs together with αβ Cytotoxic CD8+ T-cells and Helper CD4+ T-lymphocyte recruitment and activation. Potential sources of DC cells for vaccine production are umbilical cord blood, bone marrow and peripheral blood. The discovery that myeloid DCs can be generated readily from monocytes or very early (CD34+) precursors has served as a great boost with most studies using monocyte-derived DCs (MoDCs) which are loaded with antigen in-vitro. This topic has been reviewed recently in an excellent article by Figdor et al (2004) (644). To date, the large majority of prostate cancer vaccine studies have been phase I trials undertaken on patients with advanced disease. Large tumour burdens and the heterogeneity of these patients’ cancers together with potential sub-optimal immunocompetence of the vaccine recipients may serve to understate the real potential of this therapeutic approach. In contrast to cytotoxic chemotherapy regimens, a striking feature has been just how well these vaccines have been tolerated with a virtual absence of serious adverse events. Heiser et al (2002) (658) examined escalating doses of PSA mRNA-transfected DCs in their vaccines without any evidence of dose-related toxicity or adverse effects and Ridgway et al (2003) (659) reported that, in 100 trials which involved >1000 patients, there was a complete absence of severe adverse events (659) Certainly, the concern of inducing serious auto-immune reactions has not been seen, at least in the short-term.

73

Chapter 10. PROSTATE CANCER 1. Whole cell vaccines As clinical studies with re-injection of irradiated autologous tumour cells alone did not demonstrate a significant benefit, immune stimulants were introduced and used concomitantly. Simons reported injection of autologous cancer cells transfected with retroviral construct to express GM-CSF into patients with advanced prostate cancer (660) as a prelude to the use of the androgen dependent and PSA producing LNCaP and androgen independent and non-PSA secreting PC3 cell lines transfected to express GM-CSF (661). Because of the limited number of prostate cancer cell lines available for use as a source of antigens for presentation to T cells, one strategy to increase the number and variety of antigens has involved culturing freshly obtained prostatic tumours which are then transfected after several passages to confer immortality (Onyvax™). Although this approach promises to provide a greater range of readily available and suitable antigens, that the cells are transfected virally is likely to limit their utility to patients with advanced disease, at least initially. In order to overcome the limitations imposed by cell lines, we have been harvesting soft tissue metastases from men with progressing hormone-escape cancer. The tumour is then processed, which includes a sub-lethal dose of radiation, prior to being made available to DCs derived from cells of the monocyte lineage (MoDCs) in vitro. In this ongoing trial, we have had one man who has demonstrated a complete response at 12 months with total resolution of large lymph node metastases radiologically and a PSA which has decreased from 150 ng/ml at baseline to 4.7 ng/ml. Other patients have demonstrated partial responses - approximately 20% overall – but most have not. Amongst other explanations for the variability in response, this may indicate that harvested tumours did not reflect a sufficiently comprehensive representation of antigenically relevant molecular changes common to all metastases which could be processed by DCs and presented to αβ T cells to result in killing of all tumour cells.

2. Protein-based vaccines A more defined approach than using whole cells containing a variety of proteins and other factors, is to use flagged proteins, in particular those which are predominantly prostate specific (lineage markers) as the source of antigens for priming DCs. Unlike whole cell preparations, these have the potential advantage of being recombinant with the added possibility of subtle modifications being incorporated to enhance their immunogenicity with APC processing. Those examined have included prostate acid phosphatase (PAP) (662), and prostate specific antigen (PSA) (663-665) in particular.

3. Peptide-based vaccines As αβ cytotoxic CD8+ T-cells and αβ helper CD4+ T-cells recognize antigens as processed 8-10 or 12-20 amino acids, respectively, by APCs through their appropriate MHC receptors, so vaccines have been produced with specific peptides. Among the many molecular targets to which peptide sequences can be constructed, prostate specific membrane antigen (PSMA) peptides were used most widely especially in early phase I-II trials, with reported response rates of 20-30% (666, 667) Unlike PSA which, at a cellular level, is expressed more strongly in non-cancerous cells, PSMA is overexpressed in most prostate cancers as well as in tumour neovasculature (138). Although peptides are very attractive in the sense that they can be synthesized and made available as ‘off-the shelf’ preparations, they have the huge disadvantage that patient responses are MHC restricted. Consequently, people who do not have the appropriate HLA phenotype, which for PSMA peptides is HLA-A2+ (constituting ~40% of populations in many western countries), are unlikely to mount a TH1 response if they are administered vaccines based on PSMA peptides. 74

Chapter 10. PROSTATE CANCER In order to overcome this problem, Noguchi et al (2003) (668) tested 10 men with hormone escape prostate cancer to determine whether or not cytotoxic T cell precursors were detectable for 14 peptides. As these peptides were HLA-A24 restricted, all ten were shown to express the HLA-A24+ phenotype before commencement of the study. Patients were then vaccinated subcutaneously with up to 4 types of peptides to which their pre-vaccination peripheral blood mononuclear cells reacted. Four of the 10 men developed increased cytotoxic T-cell responses to peptides and cancer cells with anti-peptide IgG antibodies identified in 7 patients. One man developed a partial response with an 89% decrease in PSA. Stable disease was demonstrated in 5 of the 10 patients for a median duration of 2 months (668).

4. RNA Recently, use of RNA has attracted much attention for a number of reasons, not least of which is the need for only minute amounts of tumour tissue or recombinant material required. Tumour RNA potentially encodes multiple epitopes for many HLA alleles and, consequently, extends the scope of vaccination to cancers in which potent T cell epitopes have not been identified (669) In order to permit this approach, techniques have been developed recently for highly efficient RNA transfection of DCs by electroporation (670-672). RNA taken up by DCs is translated and the resulting polypeptides have the potential to bind to MHC molecules for presentation to T cells. Cytoplasmic antigens in somatic cells, as a result of intracellular degrading processes, are channelled preferentially via the Class-I presentation pathway, thereby activating primarily antigen-specific CD8+ T cells (673). By contrast, membrane-associated antigens are more likely to be shunted through the endosomal pathway and peptides presented on MHC-Class II receptor for CD4+ T-cell recognition (Figure 3). The importance of mounting both a CD4+ and a CD8+ response for effective tumour killing has been highlighted recently. Furthermore, the importance of activating specific CD4+ T cells concurrently to facilitate initiation, potentiation & maintenance of an effective anti-tumour immune response may be harnessed more precisely via RNA (674-676). Heiser et al (2001) (677) reported induction of a polyclonal CaP-specific cytotoxic Tcell lymphocyte (CTL) response with DCs transfected with amplified tumour RNA (677). Subsequently the same group published their experience with a vaccine consisting of DCs transfected with mRNA encoding PSA. They found a potent T-cell mediated anti-tumour response in-vitro prompting a phase I study in which this finding was replicated in all patients, 6/7 of whom had a significant decrease in the log slope of PSA (644). More recently, a strategy to curtail immortality of tumour cells was reported by Su et al (2005) (669) who targeted telomerase, the enzyme that prevents telomeres on the ends of chromosomes from shortening (669). By priming DCs with the human teleomerase reverse transcriptase (hTERT), αβ CD8+ cytotoxic T-cells can be primed to target tumours. Although this strategy, which can be engineered to obtain both CD8 and CD4 responses, is likely to be comprehensive in targeting most tumour cells, a theoretical concern is that other immortal cells such as non-cancer stem cells also may be affected by vaccines targeting telomerase (669).

5. Carbohydrates Tumours expressing high levels of certain carbohydrate antigens exhibit a greater propensity for progressing and metastasising, which is manifested in a reduced survival rate. In prostate cancer, globo-H, MUC1,GM2 and Thompson-Friedenreich antigen have been found to be preferentially expressed (678,679) in this regard and this has led to vaccine trials with combination carbohydrate-KLH conjugates (680, 681). 75

Chapter 10. PROSTATE CANCER 6. Viruses Based on the observation that a number of viral pathogens are able to elicit potent cytotoxic T-cell responses, viral constructs which express specific epitopes have been used. These include recombinant vaccinia and fowlpox viruses expressing PSA and B7.1 co-stimulatory genes (682, 683) In addition, DNA plasmids which encode for selelcted genes, have been used to produce immune responses to the gene products (684).

(viii) Summary
Although considerable advances are being made in vaccine therapies, they remain experimental and, as such, need to be integrated into clinical management in the form of research trials. However, it is hoped that the promise they bring will be realised increasingly so that they become an established form of therapy in the foreseeable future.

XVII. CONCLUSIONS
The contemporary management of prostate cancer has been shaped the utilization of PSA serology for prostate cancer detection and monitoring. One result of PSA early detection programmes is that younger patients are being diagnosed with and treated for prostate cancer. Consequently, there has been a concerted effort to improve long-term cancer control while preserving urinary and sexual function. This has been facilitated by the anatomic radical prostatectomy with, more recently, endoscopic modifications, as well as by advances in radiotherapy and other treatments. Consequently, increasing numbers of men are being cured of their prostate cancers and, by tailoring therapy to accomodate patients’ wishes, more and more are able to be satisfied with their quality of life throughout what is often a long period of management. The past 10 years has witnessed a surge in basic, clinical, and translational research in prostate cancer. Fortunately, we are in the midst of a period of unprecedented collaboration between scientists and clinicians placing patients in a most favourable position to benefit from advances in basic science and experimental therapeutics. Even if a cure is not imminent, the welfare and quality of life of those diagnosed with this condition promises to continue to improve at an exponential rate through the careful and critical application of research findings.

References
1. Greenlee RT, Hill-Harmon MB, Murray T, Thun M. (2001) Cancer Statistics, 2001. CA Cancer Journal for Clinicians, 51(1):15. 2. Baade PD, Coory MD, Aitken JF. (2004) International trends in prostate-cancer mortality: the decrease is continuing and spreading. Cancer Causes Control, 15(3): 237-41. 3. Bartsch G, Horninger W, Klocker H, Reissigl A, Oberaigner W, Schonitzer D, Severi G, Robertson C, Boyle P; Tyrol Prostate Cancer Screening Group.(2000) Decrease in prostate cancer mortality following introduction of prostate specific antigen screening in the federal state of Tyrol, Austria. Journal of Urology 2000, 163(4):88. 4. Etzioni, R, Legler, JM, Feuer, EJ, Merrill, RM, Cronin, KA, Hankey, BF. Cancer surveillance series: interpreting trends in prostate cancer--part III: Quantifying the 76

Chapter 10. PROSTATE CANCER link between population prostate-specific antigen testing and recent declines in prostate cancer mortality. Journal of the National Cancer Institute 1999, 91(12):1033. 5. Hankey BF, Fuer EJ, Clegg LX, Hayes RB, Legler JM, Prorok PC, Ries LA, Merrill RM, Kaplan RS. Cancer Surveillance Series: Interpreting trends in prostate cancer – Part I: Evidence of the effects of screening in recent prostate cancer incidence, mortality and survival rates. J Natl Cancer Inst, 1999, 91:1017-1024. 6. Schroder FH, Kranse R. Verification bias and the prostate-specific antigen test--is there a case for a lower threshold for biopsy? N Engl J Med,2003; 349(4):393-5. 7. Partin AW, Stutzman RE. Elevated prostate-specific antigen, abnormal prostate evaluation on digital rectal examination, and transrectal ultrasound and prostate biopsy. Urol Clin North Am,1998; 25(4):581-9. 8. Jhaveri FM, Klein EA, Kupelian PA, Levin HS. Declining rates of extracapsular extension after radical prostatectomy: evidence for continued stage migration. J Clinical Oncology, 1999; 17(10):3167-72. 9. Brawer, MK. Prostate-specific antigen: critical issues. Urology, 1994; 44(6A Suppl):9. 10. Carter, HB, Pearson, JD. Prostate-specific antigen testing for early diagnosis of prostate cancer: formulation of guidelines. Urology 1999, 54(5):780. 11. Gann PH, Hennekens CH, Sampfer MJ. A prospective evaluation of plasma prostate specific antigen for detection of prostatic cancer. JAMA, 1995, 273:289-94. 12. Smith DS, Catalona WJ, The nature of prostate cancer detected through prostate specific antigen based screening. J Urol, 1994, 152:1732-6. 13. Hoedemaeker RF, Rietbergen JB, Kranse R, Schröder FH, van derKwast TH. Histopathological prostate cancer characteristics at radical prostatectomy after population based screening. J Urol, 2000, 164:411-5. 14. Luboldt HJ, Bex A, Swoboda A, Husing J, Rubben H. Early detection of prostate cancer in Germany: a study using digital rectal examination and 4.0 ng/ml prostatespecific antigen as cutoff. Eur Urol, 2001, 39:131-7. 15. Freedland SJ, Presti JC Jr, Amling CL, Kane CJ, Aronson WJ, Dorey F, Terris, MK. SEARCH Database Study Group. Time trends in biochemical recurrence after radical prostatectomy: results of the SEARCH database. Urology, 2003, 61:736-741. 16. Labrie F, Candas B, Dupont A, Cusan L, Gomez JL, Suburu RE, Diamond P, Levesque J, Belanger A.. Screening decreases prostate cancer death: first analysis of the 1988 Quebec prospective randomized controlled trial. Prostate 1999, 38(2):83-91. 17. Kirby, RS, Christmas, TJ, Brawer, MK. Prostate cancer. London: Times Mirror International Publishers. 1996. 18. Peehl DM, Feldman D. (2004). "Interaction of nuclear receptor ligands with the Vitamin D signaling pathway in prostate cancer." J Steroid Biochem Mol Biol. 92(4): 307-315. 19. Chung LW, Baseman A, Assikis V, Zhau HE. (2005). "Molecular insights into prostate cancer progression: the missing link of tumor microenvironment." J Urol 173(1): 10-20. 20. Fearon ER. (1997). "Human cancer syndromes: clues to the origin and nature of cancer." Science 278(5340): 1043-1050. 21. Cannon LBD, Skolnick M, Hunt S, Lyon J, Smart C. " Genetic epidemiology of prostate cancer in the Utah Mormon genealogy." Cancer Surv, 1982; 1(47): 47-69. 22. Steinberg GD, Carter B, Beaty T, Childs B, Walsh P. (1990). "Family history and the risk of prostate cancer." Prostate 17: 337-347. 23. Xu JDL, Chang BL, Adams TS, Turner AR, Meyers DA, Eeles RA, Easton DF, Foulkes WD, Simard J, Giles GG, Hopper JL, Mahle L, Moller P, Bishop T, Evans 77

Chapter 10. PROSTATE CANCER C, Edwards S, Meitz J, Bullock S, Hope Q, Hsieh CL, Halpern J, Balise RN, OakleyGirvan I, Whittemore AS, Ewing CM, Gielzak M, Isaacs SD, Walsh PC, Wiley KE, Isaacs WB, Thibodeau SN, McDonnell SK, Cunningham JM, Zarfas KE, Hebbring S, Schaid DJ, Friedrichsen DM, Deutsch K, Kolb S, Badzioch M, Jarvik GP, Janer M, Hood L, Ostrander EA, Stanford JL, Lange EM, Beebe-Dimmer JL, Mohai CE, Cooney KA, Ikonen T, Baffoe-Bonnie A, Fredriksson H, Matikainen MP, Tammela TLj, Bailey-Wilson J, Schleutker J, Maier C, Herkommer K, Hoegel JJ, Vogel W, Paiss T, Wiklund F, Emanuelsson M, Stenman E, Jonsson BA, Gronberg H, Camp NJ, Farnham J, Cannon-Albright LA, Seminara D; The ACTANE Consortium. "A combined genomewide linkage scan of 1,233 families for prostate cancer-susceptibility genes conducted by the international consortium for prostate cancer genetics." Am J Hum Gene, 2005; 77(2): 219-229. 24. Lichtenstein P, Holm NV, Verkasalo PK, Iliadou A, Kaprio J, Koskenvuo M, Pukkala E, Skytthe A, Hemminki K. "Environmental and heritable factors in the causation of cancer--analyses of cohorts of twins from Sweden, Denmark, and Finland." N Engl J Med, 2000; 343(2): 78-85. 25. Baker SG, Lichtenstein P, Kaprio J, Holm N. (2005). "Genetic susceptibility to prostate, breast, and colorectal cancer among Nordic twins." Biometrics, 2005; 61(1): 55-63. 26. Greco LRR, Coto I, Di Cosmo N, Percopo S, Maglio M, Paparo F, Gasperi V, Limongelli MG, Cotichini R, D’Agate C, Tinto N, Sacchetti L, Tosi R, Stazi MA. (2002). "The first large population based twin study of coeliac disease." Gut. 50(5): 624-628. 27. Lorenzo Bermejo J, and Hemminki K. "Risk of cancer at sites other than the breast in Swedish families eligible for BRCA1 or BRCA2 mutation testing." Ann Oncol, 2004; 15(12): 1834-41. 28. Lorenzo Bermejo J and Hemminki K. "A population-based assessment of the clustering of breast cancer in families eligible for testing of BRCA1 and BRCA2 mutations." Ann Oncol, 2005; 16(2): 322-9. 29. van Asperen C J, Brohet RM, Meijers-Heijboer EJ, Hoogerbrugge N, Verhoef S, Vasen HF, Ausems MG, Menko FH, Gomez Garcia EB, Klijn JG, Hogervorst FB, van Houwelingen JC, van’t Veer LJ, Rookus MA, van Leeuwen FE, Netherlands Collaborative Group on Hereditary Breast Cancer (HEBON). "Cancer risks in BRCA2 families: estimates for sites other than breast and ovary." J Med Genet, 2005; 42(9): 711-719. 30. Kirchhoff TKN, Mitra N, Nafa K, Huang H, Palmer C, Gulati T, Wadsworth E, Donat S, Robson ME, Ellis NA, Offit K. "BRCA mutations and risk of prostate cancer in Ashkenazi Jews." Clin Cancer Res, 2004; 10(9): 2918-2921. 31. Suzuki H, Ueda T, Ichikawa T, Ito H. " Androgen receptor involvement in the progression of prostate cancer." Endocrine-related cancer, 2003; 10: 209-216. 32. Taplin ME, Bubley GJ, Ko YJ, Small EJ, Upton M, Rajeshkumar B, Balk SP. Cancer and Leukemia Group B Study 9663. "Androgen receptor mutations in androgen-in dependent prostate cancer. Cancer and Leukemia group B study, 9663." J Clinical oncology, 2003; 21: 2673-2678. 33. Febbo P, and Kantoff PW. "Androgen receptor polymorphisms and prostate cancer risk." Prostate Cancer: Biology, Genetics, and the New Therapeutics. (Chung, LWK, Isaacs, WB, Simmons, JW, editors). Totowa: Humana Press Inc, 2001: 95-109. 34. Montgomery JS, Price DK, Figg WD. "The androgen receptor gene and its influence on the development and progression of prostate cancer." Journal of Pathology, 2001; 195(2): 138-146. 35. Binnie MC, Alexander FE, Heald C, Habib FK. "Polymorphic forms of prostate specific antigen and their interaction with androgen receptor trinucleotide repeats in prostate cancer." Prostate, 2005; 63(4): 309-315. 36. Han G, Foster BA, Mistry S, Buchanan G, Harris JM, Tilley WD, Greenberg NM. Hormone status selects for spontaneous somatic androgen receptor variants that 78

Chapter 10. PROSTATE CANCER demonstrate specific ligand and cofactor dependent activities in autochthonous prostate cancer. J Biol Chem, 2001; 276(14):11204-13. 37. Edwards J, Krishna NS, Grigor KM, Bartlett JM. "Androgen receptor gene amplification and protein expression in hormone refractory prostate cancer." Br J Cancer, 2003; 89(3): 552-6. 38. Guerini V, Sau D, Scaccianoce E, Rusmini P, Ciana P, Maggi A, Martini PGV, Katzenellenbogen BS, Martini L, Motta M, Poletti A. The androgen derivative 5aAndrostane-3 ß,17 ß-diol inhibits prostate cancer cell migration through activation of the estrogen receptor ß subtype. Cancer Res,2005; 65: 5445-53. 39. Yeh AH, Jeffery PL, Duncan RP, Herington AC, Chopin LK. Ghrelin and a novel preproghrelin isoform are highly expressed in prostate cancer and ghrelin activates mitogen-activated protein kinase in prostate cancer. Clin Cancer Res,2005;11(23):8295-303. 40. Risbridger GP, Mellor SL, McPherson SJ, Schmitt JF. The contribution of inhibins and activins to malignant prostate disease. Mol Cell Endocrinol„ 2001;180(1-2):149-53 41. Mellor SL, Ball EM, O’Connor AE, Ethier JF, Cranfield M, Schmitt JF, Phillips DJ, Groome NP, Risbridger GP. Activin betaC-subunit heterodimers provide a new mechanism of regulating activin levels in the prostate. Endocrinology,2003;144(10):4410-9. 42. Carter BS, Bova GS, Beaty TH, Steinberg GD, Childs B, Isaacs WB, Walsh PC. "Hereditary prostate cancer: epidemiologic and clinical features (review)." J Urol, 1993; 150: 797-802. 43. Lodygin D, Epanchintsev A, Menssen A, Diebold J, Hermeking H. "Functional epigenomics identifies genes frequently silenced in prostate cancer." Cancer Res, 2005; 65(10): 4218-4227. 44. Lodygin D, Hermeking H. (2005). "The role of epigenetic inactivation of 14-33sigma in human cancer." Cell Res, 2005; 15(4): 237-246. 45. Brothman AR, Swanson G, Maxwell TM, Cui J, Murphy KJ, Herrick J, Speights VO, Isaac J, Rohr LR. "Global hypomethylation is common in prostate cancer cells: a quantitative predictor for clinical outcome?" Cancer Genet Cytogenet, 2005; 156(1): 31-6. 46. Reis M, Pina F, Dias M, Perez M, Barros H. "Ganglionar metastization on radical prostatectomy." Acta Med Port, 2004; 17(1): 1-7. 47. Chung, L. "Stromal-epithelial interactions dictate prostate cancer androgen independent metastatic progression." Cancer Detection and Prevention, 1998; 22(Supp 1): 57. 48. Chung, LWK, Anderson NG, Neuauer BL, Cunha GR, Thompson TC, Rocco AK. Tissue interaction in prostate development: roles of sex steroids. In Prostatic Cells: Structure and Function, 1981. Alan R Liss. New York: 177-203. 49. Cunha GR & Chung LWK. Stromal-epithelial interactions. I. Introduction of prostatic phenotype of urothelium of testicular feminized (Tfm/y) mice. J Steroid Biochem, 1981; 14:1317-1321. 50. Chung, L. W. (2003). "Prostate carcinoma bone-stroma interaction and its biologic and therapeutic implications." Cancer 97(3 Suppl): 772-8. 51. Hsieh CL, Gardner TA, Miao L, Balian G, Chung LW. "Cotargeting tumor and stroma in a novel chimeric tumor model involving the growth of both human prostate cancer and bone stromal cells." Cancer Gene Ther, 2004; 11(2): 148-155. 52. Ruiz M, Pettaway C, Song R, Stoeltzing O, Ellis L, Bar-Eli M. "Activator protein 2alpha inhibits tumorigenicity and represses vascular endothelial growth factor transcription in prostate cancer cells." Cancer Res, 2004; 64(2): 631-638. 53. Wong YC, Tam NN. "Differentiation of stromal smooth muscle as a factor in prostate carcinogenesis." Differentiation, 2002; 70(9-10): 633. 79

Chapter 10. PROSTATE CANCER 54. Bayne CW, Donnelly F, Chapman K, Bollina P, Buck C, Habib F. (1998). "A novel co-culture model for benign prostatic hyperplasia expressing both isoforms of 5reductase." J Clin Endocrinol Metab 83(1): 206-213. 55. Grant ES, Batchelor KW, Habib FK. "Androgen-independence of primary epithelial cell cultures of the prostate is associated with a down-regulation of androgen receptor gene expression." Prostate, 1996; 29: 339-349. 56. Chang SM, Chung LW. "Interaction between prostatic fibroblast and epithelial cells in culture: role of androgen." Endocrinology, 1989; 125(5): 2719-2727. 57. Condon (2005). "Seminars in Cancer." Biology 15: 132-137. 58. Bianco JJ, Handelsman DJ, Pedersen JS, Risbridger GP. "Direct response of the murine prostate gland and seminal vesicles to estradiol." Endocrinology, 2002; 143(12): 4922-4933. 59. McPherson SJ, Wang H, Jones ME, Pedersen J, Iismaa TP, Wreford N, Simpson ER, 39. 39. Risbridger GP. (2001). "Elevated androgens and prolactin in aromatasedeficient mice cause enlargement, but not malignancy, of the prostate gland." Endocrinology 142(6): 2458-4267. 60. Cunha GR, Ricke W, Thomson A, Marker PC, Risbridger G, Hayward SW, Wang YZ, Donjacour AA, Kurita T. "Hormonal, cellular, and molecular regulation of normal and neoplastic prostatic development." J Steroid Biochem Mol Biol, 2004;. 92(4): 221236. 61. Risbridger GP, Ball EM, Wang H, Mellor SL, Peehl DM. "Re-evaluation of inhibin alpha subunit as a tumour suppressor in prostate cancer." Mol Cell Endocrinol, 2004;. 15(225 1-2): 73-76. 62. Edwards J, Bartlett JM. "The androgen receptor and signal-transduction pathways in hormone-refractory prostate cancer." BJU Int, 2005; 95(Part 1): 1319-1326. 63. Edwards J and Bartlett JM. "The androgen receptor and signal-transduction pathways in hormone-refractory prostate cancer. Part 2: Androgen-receptor cofactors and bypass pathways." BJU Int, 2005; 95(9): 1327-35. 64. Blok LJ, de Ruiter PE, Brinkmann AO. "Forskolin-induced dephosphorylation of the AR receptor impairs ligand binding." Biochemistry, 1998; 37: 3850-3857. 65. Lee HJ, Chang C. "Recent advances in androgen receptor action." Cellular Mol Life Sci, 2003; 60: 1613-1622. 66. Uzgare AR, Kaplan PJ, Greenberg NM. "Differential expression and/or activation of P38MAPK, erk1/2, and jnk during the initiation and progression of prostate cancer." Prostate, 2003; 55(2): 128-139. 67. Bakin RE, Gioeli D, Bissonette EA, Weber MJ. (2003). "Attenuation of Ras signaling restores androgen sensitivity to hormone-refractory C4-2 prostate cancer cells." Cancer Res. 63(8): 1975-1980. 68. Bakin RE, Gioeli D, Sikes RA, Bissonette EA, Weber MJ. (2003). "Constitutive activation of the Ras/mitogen-activated protein kinase signaling pathway promotes androgen hypersensitivity in LNCaP prostate cancer cells." Cancer Res. 63(8): 19811989. 69. Lin HK, Hu YC, Yang L, Altuwaijri S, Chen YT, Kang HY, Chang C. "Suppression versus induction of androgen receptor functions by the phosphatidylinositol 3-kinase/Akt pathway in prostate cancer LNCaP cells with different passage numbers." J Biol Chem, 2003; 278(51): 50902-7. 70. Ghosh PM, Malik S, Bedolla R, Kreisberg JI. "Akt in prostate cancer: possible role in androgen-independence." Curr Drug Metab, 2003; 4(6): 487-96. 71. Debes JD, Sebo TJ, Lohse CM, Murphy LM, Haugen de AL, Tindall DJ. "p300 in prostate cancer proliferation and progression." Cancer Res, 2003; 63(22): 7638-7640.

80

Chapter 10. PROSTATE CANCER 72. Chung LW, Baseman A, Assikis V, Zhau HE. "Molecular insights into prostate cancer progression: the missing link of tumor microenvironment." J Urol, 2005; 173(1): 10-20. 73. van Moorselaar RJ, Voest EE. "Angiogenesis in prostate cancer: its role in disease progression and possible therapeutic approaches." Mol Cell Endocrinol, 2002; 197(12): 239-250. 74. Hideshima T, Podar, K, Chauhan D, Anderson KC. " Cytokines and signal transduction." Best practice and research Clinical Haematology, 2005; 18(4): 509-524. 75. Royuela M, Ricote M, Parsons MS, Garcia-Tunon I, Paniagua R, de Miguel MP. "Immunohistochemical analysis of the IL-6 family of cytokines and their receptors in benign, hyperplasic, and malignant human prostate." J. Pathology, 2004; 202(1): 41-49. 76. Papatsoris AG, Karamouzis MV, Papavassiliou AG. "Novel insights into the implication of the IGF-1 network in prostate cancer." Trends Mol Med, 2005; 11(2): 52-5. 77. Djavan B, Waldert M, Seitz C, Marberger M. "Insulin-like growth factors and prostate cancer." World J Urology, 2001; 19: 225 78. Rubin J, Chung LW, Fan X, Zhu L, Murphy T.C,Nanes M.S, Rosen C.J "Prostate carcinoma cells that have resided in bone have an upregulated IGF-I axis." Prostate, 2003; 58: 41-49. 79. Knudsen BS, Edlund M. (2004). "Prostate cancer and the met hepatocyte growth factor receptor." Adv Cancer Research, 2004; 91: 31-67. 80. Lail-Trecker M, Gulati R, Peluso JJ. "A role for hepatocyte growth factor/scatter factor in regulating normal and neoplastic cells of reproductive tissues." J soc Gyn Invest, 1998; 5(3): 114-121. 81. Gregory CW, Fei X, Ponguta LA, He B, Bill HM, French FS, Wilson EM. "Epidermal growth factor increases coactivation of the androgen receptor in recurrent prostate cancer." J Biol Chem, 2004;279(8): 7119-30. 82. Ratan HL, Gescher A, Steward WP, Mellon JK. "ErbB receptors: possible therapeutic targets in prostate cancer?" BJU Int, 2003; 92(9): 890-5. 83. Bachman, KE and Park BH. Duel nature of TGF-beta signaling: tumor suppressor vs. tumor promoter. Curr Opin Oncol, 2005; 17(1): 49-54. 84. Bachman KE, Blair BG, Brenner K, Bardelli A Arena S, Zhou S, Hicks J, De Marzo AM, Argani P, Park BH. p21(WAF1/CIP1) mediates the growth response to TGF-beta in human epithelial cells. Cancer Biol Ther, 2004; 3(2): 221-5. 85. Yingling JM, Blanchard KM, and Sawyer JS. "Development of TGF-beta signaling inhibitors for cancer therapy." Nature Reviews, 2004; 3(12): 1011-1022. 86. De Wever O and Mareel M. "Role of tissue stroma in cancer cell invasion." J Pathol, 2003; 200(4): 429-47. 87. Slack-Davis JK, Parsons JT. "Emerging Views of Integrin Signaling:Implications for Prostate Cancer." Journal of Cellular Biochemistry, 2004; 91(1): 41-46. 88. Rovin JD, Frierson HF Jr, Ledinh W, Parsons JT, Adams RB. "Expression of Focal Adhesion kinase in Normal and Pathologyic Human prostate tissues." The Prostate, 2002; 53(2): 124-132. 89. Wu X, Gan B, YooY, Guan J-L. "FAK-mediated Src phosphorylation of Endophilin A2 inhibits endocytosis of MT1-MMP and promotes ECM degradation." Developmental Cell, 2005; 9(2): 185-196. 90. Morgia G, Falsaperla M, Malaponte G, Madonia M, Indelicato M, Travali S, Mazzarino MC. "Matrix metalloproteinases as diagnostic (MMP-13) and prognostic (MMP-2, MMP-9) markers of prostate cancer." Urol Res, 2005;. 33(1): 44-50. 91. Sauer CG, R. A., Grobholz R. (2005). "Genetic analysis of neuroendocrine tumor cells in prostatic carcinoma." Prostate. 19. 81

Chapter 10. PROSTATE CANCER 92. Wilson SR, Gallagher S, Warpeha K, Hawthorne SJ. "Amplification of MMP-2 and MMP-9 production by prostate cancer cell lines via activation of protease-activated receptors." Prostate, 2004; 60(2): 168-174. 93. Dong Z, Bonfil RD, Chinni S, Deng X, Trindade Filho JC, Bernardo M, Vaishampayan U, Che M, Sloane BF, Sheng S, Fridman R, Cher ML. (2005). "Matrix metalloproteinase activity and osteoclasts in experimental prostate cancer bone metastasis tissue." Am J Pathol, 2005; 166(4): 1173-86. 94. Huang XCS, Xu L, Liu Y, Deb DK, Platanias LC, Bergan RC. "Genistein inhibits p38 map kinase activation, matrix metalloproteinase type 2, and cell invasion in human prostate epithelial cells." Cancer Res, 2005; 65(8): 3470-3480. 95. Choong PF and Nadesapillai AP. "Urokinase plasminogen activator system: a multifunctional role in tumor progression and metastasis." Clin Orthop Relat Res, 2003;(415 Suppl): S46-58. 96. Festuccia C, Dolo V, Guerra F, Violini S, Muzi P, Pavan A, Bologna M. "Plasminogen activator system modulates invasive capacity and proliferation in prostatic tumor cells." Clin Exp Metastasis, 1998; 16(6): 513-28. 97. Gonzalez-Gronow M, Grenett HE, Weber MR, Gawdi G, Pizzo SV. "Interaction of plasminogen with dipeptidyl peptidase IV initiates a signal transduction mechanism which regulates expression of matrix metalloproteinase-9 by prostate cancer cells." Biochem J, 2001; 355(Pt 2): 397-407. 98. Gonzalez-Gronow M, Grenett HE, Gawdi G, Pizzo SV. "Angiostatin directly inhibits human prostate tumor cell invasion by blocking plasminogen binding to its cellular receptor, CD26." Exp Cell Res, 2005; 303(1): 22-31. 99. Sidenius N and Blasi F. "The urokinase plasminogen activator system in cancer: recent advances and implication for prognosis and therapy." Cancer Metastasis Rev, 2003; 22(2-3): 205-22. 100. Margheri F, D’Alessio S, Serrati S, Pucci M, Del Rosso A, Benelli R, Ferrari N, Noonan DM, Albini A, Fibbi G, Del Rosso M. "The urokinase-type plasminogen activator, its receptor and u-PA inhibitor type-1 affect in vitro growth and invasion of Kaposi’s sarcoma and capillary endothelial cells: role of HIV-Tat protein." Int J Oncol, 2005; 27(1): 223-35. 101. Xi Z, Kaern J, Davidson B, Klokk TI, Risberg B, Trope C, Saatcioglu F. "Kallikrein 4 is associated with paclitaxel resistance in ovarian cancer." Gynecol Oncol, 2004; 94(1): 80-85. 102. Haese AVV, Lilja H, Kattan MW, Rittenhouse HG, Pettersson K, Chan DW, Huland H, Sokoll LJ, Partin AW. "Comparison of predictive accuracy for pathologically organ confined clinical stage T1c prostate cancer using human glandular kallikrein 2 and prostate specific antigen combined with clinical stage and Gleason grade." J Urol, 2005; 173(3): 752-756. 103. Stephan C, Jung K, Soosaipillai A, Yousef GM, Cammann H, Meyer H, Xu C, Diamandis EP. "Clinical utility of human glandular kallikrein 2 within a neural network for prostate cancer detection." BJU Int, 2005; 96(4): 521-527. 104. Steuber T, Vickers AJ, Haese A, Becker C, Pettersson K, Chun FK, Kattan MW, Eastham JA, Scardino PT, Huland H, Lilja H. (2005). "Risk assessment for biochemical recurrence prior to radical prostatectomy: Significant enhancement contributed by human glandular kallikrein 2 (hK2) and free prostate specific antigen (PSA) in men with moderate PSA-elevation in serum." Int J Cancer, 2006; 118(5): 1234-40. 105. Clements JA, Willemsen NM, Myers SA, Dong Y. "The tissue kallikrein family of serine proteases: functional roles in human disease and potential as clinical biomarkers." Crit Rev Clin Lab Sci, 2004; 41(3): 265-312. 106. Veveris-Lowe TL, Lawrence MG, Collard RL, Bui L, Herington AC, Nicol DL, Clements JA. "Kallikrein 4 (hK4) and prostate-specific antigen (PSA) are associated 82

Chapter 10. PROSTATE CANCER with the loss of E-cadherin and an epithelial-mesenchymal transition (EMT)-like effect in prostate cancer cells." Endocr Relat Cancer, 2005; 12(3): 631-643. 107. Quinn DI, Henshall SM, Sutherland RL. Molecular markers of prostate cancer outcome. Eur J Cancer, 2005; 41(6): 858-87. 108. Williams TM and Lisanti MP. "Caveolin-1 in oncogenic transformation, cancer, and metastasis." Am J Physiol Cell Physiol, 2005; 288(3): C494-506. 109. Yang G, Timme TL, Frolov A, Wheeler TM, Thompson TC. "Combined c-Myc and caveolin-1 expression in human prostate carcinoma predicts prostate carcinoma progression." Cancer, 2005; 103(6): 1186-94. 110. Epstein JI, Amin M, Boccon-Gibod L, Egevad L, Humphrey PA, Mikuz G, Newling D, Nilsson S, Sakr W, Srigley JR, Wheeler TM, Montironi R. "Prognostic factors and reporting of prostate carcinoma in radical prostatectomy and pelvic lymphadenectomy specimens." Scand J Urol Nephrol Suppl, 2005; 216: 34-63. 111. Kurek RNG, Tselis N, Konrad L, Martin T, Roeddiger S, Aumuller G, Zamboglou N, Lin DW, Tunn UW, Renneberg H. "Prognostic value of combined "triple"reverse transcription-PCR analysis for prostate-specific antigen, human kallikrein 2, and prostate-specific membrane antigen mRNA in peripheral blood and lymph nodes of prostate cancer patients." Clin Cancer Res, 2004; 10(17): 5808-5814. 112. Lintula S, Stenman J, Bjartell A, Nordling S, Stenman UH. "Relative concentrations of hK2/PSA mRNA in benign and malignant prostatic tissue." Prostate, 2005; 63(4): 324-329. 113. Chiang CH, Hong CJ, Chang YH, Chang LS, Chen KK. "Human kallikrein-2 gene polymorphism is associated with the occurrence of prostate cancer." J Urol, 2005; 173(2): 429-432. 114. Stephenson SA, Verity K, Ashworth LK, Clements JA. "Localization of a new prostate-specific antigen-related serine protease gene, KLK4, is evidence for an expanded human kallikrein gene family cluster on chromosome 19q13.3-13.4." J Biol Chem, 1999; 274(33): 23210-4. 115. Ang J, Lijovic M, Ashman LK, Kan K, Frauman AG. “CD151 protein expression predicts the clinical outcome of low-grade primary prostate cancer better than histologic grading: a new prognostic indicator?” Cancer Epidemiol Biomarkers Prev, 2004; 13(11 Pt 1): 1717-21. 116. Downing SR, Russell PJ, Jackson P. "Alterations of p53 are common in early stage prostate cancer." Can J Urol, 2005; 10(4): 1924-1933. 117. Quinn DI, Henshall SM, Head DR, Golovsky D, Wilson JD, Brenner PC, Turner JJ, Delprado W, Finlayson JF, Stricker PD, Grygiel JJ, Sutherland RL. "Prognostic significance of p53 nuclear accumulation in localized prostate cancer treated with radical prostatectomy." Cancer Res, 2000; 60(6): 1585-94. 118. Augustin H, Hammerer PG, Graefen M, Palisaar J, Daghofer F, Huland H, Erbersdobler A. “Characterisation of biomolecular profiles in primary high-grade prostate cancer treated by radical prostatectomy.” J Cancer Res Clin Oncol, 2003; 129(11): 662-8. 119. Rubio J, Ramos D, Lopez-Guerrero JA, Iborra I, Collado A, Solsona E, Almenar S, Llombart-Bosch A. "Immunohistochemical expression of ki-67 antigen, cox-2 and bax/bcl-2 in prostate cancer; prognostic value in biopsies and radical prostatectomy specimens." Eur Urol, 2005; 48(5): 745-51. 120. Fonseca GN, Srougi M, Leite KR, Nesrallah LJ, Ortiz V."The role of HER2/neu, BCL2, p53 genes and proliferating cell nuclear protein as molecular prognostic parameters in localized prostate carcinoma." Sao Paulo Med J, 2004; 122(3): 124-7 122(3): 124-127. 121. Kaur P, Kallakury BS, Sheehan CE, Fisher HA, Kaufman RP Jr, Ross JS. "Survivin and Bcl-2 expression in prostatic adenocarcinomas." Arch Pathol Lab Med, 2004; 128(1): 39-43. 83

Chapter 10. PROSTATE CANCER 122. Tulunay O, Orhan D, Baltaci S, Gogus C, Muftuoglu YZ. “Prostatic ductal adenocarcinoma showing Bcl-2 expression.” Int J Urol.2004; Sep 11(9):805-8. 123. Rosser CJ, Reyes AO, Vakar-Lopez F, Levy LB, Kuban DA, Hoover DC, Lee AK, Pisters LL. "Bcl-2 is significantly overexpressed in localized radio-recurrent prostate carcinoma, compared with localized radio-naive prostate carcinoma." Int J Radiat Oncol Biol Phys, 2003; 56(1): 1-6. 124. Gleave M and Jansen B. "Clusterin and IGFBPs as antisense targets in prostate cancer." Ann N Y Acad Sci, 2003; 1002: 95-104. 125. Rosser CJ, Tanaka M, Pisters LL, Tanaka N, Levy LB, Hoover DC, Grossman HB, McDonnell TJ, Kuban DA, Meyn RE. (2004). "Adenoviral-mediated PTEN transgene expression sensitizes Bcl-2-expressing prostate cancer cells to radiation." Cancer Gene Ther, 2004; 11(4): 273-9. 126. Scaltriti M, Brausi M, Amorosi A, Caporali A, D’Arca D, Astancolle S, Corti A, Bettuzzi S. ""Clusterin (SGP-2, ApoJ) expression is downregulated in low- and highgrade human prostate cancer."" Int J Cancer, 2004; 108(1): 23-30. 127. Scaltriti M, Brausi M, Amorosi A, Caporali A, D’Arca D, Astancolle S, Corti A, Bettuzzi S. "Clusterin overexpression in both malignant and nonmalignant prostate epithelial cells induces cell cycle arrest and apoptosis." Br J Cancer, 2004; 91(10): 18421850. 128. Zhou M, Aydin H, Kanane H, Epstein JI. "How often does alpha-methylacyl-CoA-racemase contribute to resolving an atypical diagnosis on prostate needle biopsy beyond that provided by basal cell markers?" Am J Surg Pathol, 2004; 28(2): 239-243. 129. Pins MR, Fiadjoe JE, Korley F, Wong M, Rademaker AW, Jovanovic B, Yoo TK, Kozlowski JM, Raji A, Yang XJ, Lee C. "Clusterin as a possible predictor for biochemical recurrence of prostate cancer following radical prostatectomy with intermediate Gleason scores: a preliminary report." Prostate Cancer Prostatic Dis, 2004; 7(3): 243248. 130. Zellweger T, Kiyama S, Chi K, Miyake H, Adomat H, Skov K, Gleave ME. "Overexpression of the cytoprotective protein clusterin decreases radiosensitivity in the human LNCaP prostate tumour model." BJU Int, 2003; 92(4): 463-469. 131. Caccamo AE, Scaltriti M, Caporali A, D’Arca D, Corti A, Corvetta D, Sala A, Bettuzzi S. Ca2+ depletion induces nuclear clusterin, a novel effector of apoptosis in immortalized human prostate cells. Cell Deat Differ. 2005 12(1):101-4. 132. Bracken AP, Pasini D, Capra M, Prosperini E, Colli E, Helin K. "EZH2 is downstream of the pRB-E2F pathway, essential for proliferation and amplified in cancer." Embo J, 2003; 22(20): 5323-5335. 133. Varambally S, Dhanasekaran SM, Zhou M, Barrette TR, Kumar-Sinha C, Sanda MG, Ghosh D, Pienta KJ, Sewalt RG, Otte AP, Rubin MA, Chinnaiyan AM. "The polycomb group protein EZH2 is involved in progression of prostate cancer." Nature, 2002; 419(6907): 624-9. 134. Foster CS, Falconer A, Dodson AR, Norman AR, Dennis N, Fletcher A, Southgate C, Dowe A, Dearnaley D, Jhavar S, Eeles R, Feber A, Cooper CS. (2004). "Transcription factor E2F3 overexpressed in prostate cancer independently predicts clinical outcome." Oncogene, 2004; 23(35): 5871-9. 135. Hrouda D, Nicol DL, Gardiner RA. "The role of angiogenesis in prostate development and the pathogenesis of prostate cancer." Urol Res, 2003; 30(6): 347-55. 136. Blancher C, Harris AL. "The molecular basis of the hypoxia response pathway: Tumor hypoxia as a therapy target." Cancer Metastasis Rev, 1998; 17: 187-194. 137. Chang SS. Monoclonal antibodies and prostate-specific membrane antigen. Curr Opin Investig Drugs, 2004; 5(6): 611-5.

84

Chapter 10. PROSTATE CANCER 138. Chang SS, Reuter VE, Heston WD, Bander NH, Grauer LS, Gaudin PB. "Five different anti-prostate-specific membrane antigen (PSMA) antibodies confirm PSMA expression in tumor-associated neovasculature." Cancer Res, 1999; 59(13): 3192-8. 139. Israeli RS, Powell CT, Fair WR, Heston WD. "Molecular cloning of a complementary DNA encoding a prostate-specific membrane antigen." Cancer Res, 1993; 53(2): 227-30. 140. Belldegrun A, Bander NH, Lerner SP, Wood Dp, Pantuck AJ. "Society of Urologic Oncology Biotechnology Forum: new approaches and targets for advanced prostate cancer." Journal of Urology, 2001; 166: 1316. 141. Leek J, Lench N, Maraj B, Bailey A, Carr IM, Andersen S, Cross J, Whelan P, MacLennan KA, Meredith DM, et al."Prostate-specific membrane antigen: evidence for the existence of a second related human gene." Br J Cancer, 1995; 72: 583-588. 142. Rinker-Schaeffer CW, Hawkins AL, Su SL, Israeli RS, Griffin CA, Isaacs JT, Heston WD. "Localization and physical mapping of the prostate-specific antigen (PSM) gene to human chromosome 11." Genomics, 1995; 30: 105-108. 143. Burger MJ, Tebay MA, Keith PA, Samaratunga HM, Clements J, Lavin MF, Gardiner RA. "Expression analysis of delta-catenin and prostate-specific membrane antigen: their potential as diagnostic markers for prostate cancer." Int J Cancer, 2002; 100(2): 228-37. 144. Landers, K. A., M. J. Burger, Tebay MA, Purdie DM, Scells B, Samaratunga H, Lavin MF, Gardiner RA.. "Use of multiple biomarkers for a molecular diagnosis of prostate cancer." Int J Cancer, 2005; 114(6): 950-6. 145. Rajasekaran AK, Anilkumar G, Christiansen JJ. "Is prostate-specific membrane antigen a multifunctional protein?" Am J Physiol Cell Physiol, 2005; 288(5): C975-81. 146. Dhanasekaran SM, Barrette TR, Ghosh D, Shah R, Varambally S, Kurachi K, Pienta KJ, Rubin MA, Chinnaiyan AM. "Delineation of prognostic biomarkers in prostate cancer." Nature, 2001; 412(4689): 822-826. 147. Srikantan V, Valladares M, Rhim JS, Moul JW, Srivastava S. "HEPSIN inhibits cell growth/invasion in prostate cancer cells." Cancer Res, 2002; 62(23): 6812-6812. 148. Herter SPD, Aaron W, Gabriele T, Cutler G, Cao P, Bhatt AS, Choe Y, Craik CS, Walker N, Meininger D, Hoey T, Austin RJ. "Hepatocyte growth factor is a preferred in vitro substrate for human hepsin, a membrane-anchored serine protease implicated in prostate and ovarian cancers." Biochem J, 2005; 390(Pt 1): 125-136. 149. Kirchhofer DMP, Bullens S, Peale F, Bunting S. "A monoclonal antibody that inhibits mouse tissue factor function." J Thromb Haemost, 2005; May;3(5):1098-9. 3(5): 1089-1090. 150. Kirchhofer DPM, Lipari MT, Billeci K, Fan B, Moran P. (2005). "Hepsin activates pro-hepatocyte growth factor and is inhibited by hepatocyte growth factor activator inhibitor-1B (HAI-1B) and HAI-2." FEBS Lett, 2005; 259(9): 1945-1950. 151. Schalken JA. Validation of molecular targets in prostate cancer. BJU Int.2005; 96 Suppl 2:23-9. 152. Kumar-Sinha CSR, Laxman B, Tomlins SA, Harwood J, Schmitz W, Conzelmann E, Sanda MG, Wei JT, Rubin MA, Chinnaiyan AM. "Elevated alpha-methylacyl-CoA racemase enzymatic activity in prostate cancer." Am J Pathol, 2004; 164(3): 787-793. 153. Jiang Z, Li C, Fischer A, Dresser K, Woda BA. "Using an AMACR (P504S)/34betaE12/p63 cocktail for the detection of small focal prostate carcinoma in needle biopsy specimens." Am J Clin Pathol, 2005; 123(2): 231-6. 154. Mobley J A., Leav I, Zielie P, Wotkowitz C, Evans J, Lam YW, L’Esperance BS, Jiang Z, Ho SM. “Branched fatty acids in dairy and beef products markedly enhance alpha-methylacyl-CoA racemase expression in prostate cancer cells in vitro.” Cancer Epidemiol Biomarkers Prev, 2003; 12(8): 775-83. 85

Chapter 10. PROSTATE CANCER 155. Varma M, Jasani B. "Diagnostic utility of immunohistochemistry in morphologically difficult prostate cancer: review of current literature." Histopathology, 2005; 47(1): 1-16. 47(1): 1-16. 156. Zhou M, Jiang Z, Epstein JI. "Expression and diagnostic utility of alpha-methylacyl-CoA-racemase (P504S) in foamy gland and pseudohyperplastic prostate cancer." Am J Surg Pathol, 2003; 27(6): 772-8. 157. Molinie V, Fromont G, Sibony M, Vieillefond A, Vassiliu V, Cochand-Priollet B, Herve JM, Lebret T, Baglin AC. "Diagnostic utility of a p63/alpha-methyl-CoAracemase (p504s) cocktail in atypical foci in the prostate." Mod Pathol, 2004; 17(10): 1180-90. 158. Mubiru JN, Shen-Ong GL, Valente AJ, Troyer DA. "Alternative spliced variants of the alpha-methylacyl-CoA racemase gene and their expression in prostate cancer." Gene, 2004; 327(1): 89-98. 159. Reiter RE, Gu Z, Watabe T, Thomas G, Szigeti K, Davis E, Wahl M, Nisitani S, Yamashiro J, Le Beau MM, Loda M, Witte ON. "Prostate stem cell antigen: a cell surface marker overexpressed in prostate cancer." Proc Natl Acad Sci USA, 1998; 95(11): 1735-1740. 160. Zhigang Z, Wenlu S. "Complete androgen ablation suppresses prostate stem cell antigen (PSCA) mRNA expression in human prostate carcinoma." Prostate, 2005; 65(4): 299-305. 161. Han KR, Seligson DB, Liu X, Horvath S, Shintaku PI, Thomas GV, Said JW, Reiter RE. "Prostate stem cell antigen expression is associated with gleason score, seminal vesicle invasion and capsular invasion in prostate cancer." J Urol, 2004; 171(3): 111721. 162. Wente MN, Jain A, Kono E, Berberat PO, Giese T, Reber HA, Friess H, Buchler MW, Reiter RE, Hines OJ. "Prostate stem cell antigen is a putative target for immunotherapy in pancreatic cancer." Pancreas, 2005; 31(2): 119-25. 163. Rodriguez JA, Li M, Yao Q, Chen C, Fisher WE. "Gene overexpression in pancreatic adenocarcinoma: diagnostic and therapeutic implications." World J Surg, 2005; 29(3): 297-305. 164. Cheng L, Reiter RE, Jin Y, Sharon H, Wieder J, Lane TF, Rao J. "Immunocytochemical analysis of prostate stem cell antigen as adjunct marker for detection of urothelial transitional cell carcinoma in voided urine specimens." J Urol, 2003; 169(6): 2094-100. 165. Matsueda S, Kobayashi K, Nonaka Y, Noguchi M, Itoh K, Harada M. "Identification of new prostate stem cell antigen-derived peptides immunogenic in HLA-A2(+) patients with hormone-refractory prostate cancer." Cancer Immunol Immunother, 2004; 53(6): 479-89. 166. Matsueda S, Yao A, Ishihara Y, Ogata R, Noguchi M, Itoh K, Harada M. "A prostate stem cell antigen-derived peptide immunogenic in HLA-A24- prostate cancer patients." Prostate, 2004; 60(3): 205-13. 167. Bussemakers MJ, van Bokhoven A, Verhaegh GW, Smit FP, Karthaus HF, Schalken JA, Devruyne FM, Ru N, Isaacs WB. “DD3: a new prostate-specific gene, highly overexpressed in prostate cancer.” Cancer Res, 1999; 59(23): 5975-9. 168. Hessels D, Klein Gunnewiek JM, van Oort I, Karthaus HF, van Leenders GJ, van Balken B, Kiemeney LA, Witjes JA, Schalken JA. “(DD3(PCA3)-based molecular urine analysis for the diagnosis of prostate cancer." Eur Urol, 2003; 44(1): 8-15. 169. Tinzl M, Marberger M, Horvath S, Chypre C. DD3 PCA3 analysis in urine – a new perspective for detecting prostate cancer. Eur Urol, 2004; 46: 182-7. 170. Fradet Y, Saad F, Aprikian A, Dessureault J, Elhilali M, Trudel C, Masse B, Piche L, Chypre C. “uPM3, a new molecular urine test for the detection of prostate cancer.” Urology, 2004; 64(2): 311-5.

86

Chapter 10. PROSTATE CANCER 171. Petrovics G, Liu A, Shaheduzzaman S, Furasato B, Sun C, Chen Y, Nau M, Ravindranath L, Chen Y, Dobi A, Srikantan V, Sesterhenn IA, McLeod DG, Vahey M, Moul JW, Srivastava S. "Frequent overexpression of ETS-related gene-1 (ERG1) in prostate cancer transcriptome." Oncogene, 2005; 24(23): 3847-3852. 172. Di Lorenzo G, De Placido S, Autorino R, De Laurentiis M, Mignogna C, D’Armiento M, Tortora G, De Rosa G, D’Armiento M, De Sio M, Bianco AR, D’Armiento FP. "Expression of biomarkers modulating prostate cancer progression: implications in the treatment of the disease." Prostate Cancer Prostatic Dis, 2005; 8(1): 54-9. 173. Ekici SCW, Duncan R, Gomez P, Civantos F, Soloway MS, Lokeshwar VB. "Comparison of the prognostic potential of hyaluronic acid, hyaluronidase (HYAL-1), CD44v6 and microvessel density for prostate cancer." Int J Cancer, 2004; 112(1): 121-129. 174. Glinsky GV, Glinskii AB, Stephenson AJ, Hoffman RM, Gerald WL. "Gene expression profiling predicts clinical outcome of prostate cancer." Journal of Clinical Investigation, 2004; 113(6): 913-923. 175. Statland BE. “Nutrition and cancer.” Clin Chem, 1992; 38(8B Pt 2): 1587-94 176. Giovannucci E, Rimm EB, Colditz GA, Stampfer MJ, Ascherio A, Chute CC, Willett WC. “A prospective study of dietary fat and risk of prostate cancer.” J Natl Cancer Inst, 1993; 85(19): 1571-9. 177. Fleshner N, Bagnell PS, Klotz L, Venkateswaran V. “Dietary fat and prostate cancer.” J Urol, 2004;171(2 Pt 2): S19-24. 178. Schuurman AG, van-den-Brandt PA, Dorant E, Brants HA, Goldbohm RA. Association of energy and fat intake with prostate carcinoma risk: results from The Netherlands Cohort Study.” Cancer, 1999; 86(6) 1019-27. 179. Zock PL, Katan MB. “Linoleic acid intake and cancer risk: a review and metaanalysis.” Am J Clin Nutr, 1998; 68(1): 142-53. 180. Kumar V. “Nutritional aspects of prostate cancer.” Urol News, 2005; 9(2): 10-13. 181. Shackelford RE, Kaufmann WK, Paules RS. “Cell cycle control, checkpoint mechanisms and genotoxic stress.” Environ Health Perspect, 1999; 107 (Suppl 1): 5-24. 182. Esposito F, RussoL, Russo T, Cimino F. “Retinoblastoma protein dephosphorylation is an early event of cellular reponse to prooxidant conditions.” FEBS Lett, 2000; 470(2): 211-5. 183. Chetcuti A, Margan SH, Russell P, Mann S, Millar DS, Clark SJ, Rogers J, Handelsman DJ, Dong Q. “Loss of annexin II heavy and light chains in rost and its precursors.” Cancer Res, 2001; 61(17): 6331-4. 184. Dannenberg AJ, Altorki NK, Boyle JO, Dang C, Howe LR, Weksler BB, Subbaramaiah K. “Cyclo-oxygenase 2: a pharmacological target for the prevention of cancer.” Lancet Oncol, 2001; 2(9): 544-51. 185. Giovannucci E, Rimm EB, Liu Y, Stampfer MJ, Willett WC. “ A prospective study of tomato products, lycopene, and prostate cancer risk.” J Natl Cancer Inst, 1999; 91(4):317-31. 186. Giovannucci E, Rimm EB, LiuY, Stampfer MJ, Willett WC. “ A prospective study of tomato products, lycopene, and prostate cancer risk.” J Natl Cancer Inst, 2002; 94(5): 391-8. 187. Giovannucci E. “Selenium and risk of prostate cancer.” Lancet, 1998; 352(9130): 755-6. 188. Clark LC, Dalkin B, Krongrad A, Combs GF Jr, Turnbull BW, Slate EH, Witherington R, Herlong JH, Janosko E, Carpenter D, Borosso C, Falk S, Rounder J. “Decreased incidence of prostate cancer with selenium supplementation: results of a double-blind cancer prevention trial.” Br J Urol, 1998; 81(5):730-4. 87

Chapter 10. PROSTATE CANCER 189. Karunasinghe N, Ryan J, Tuckey J, Masters J, Jamieson M, Clarke L-C, Marshall JR, Ferguson LR. “DNA stability and serum selenium levels in a high-risk group for prostate cancer.” Cancer Epidemiol Biomarkers Prev, 2004; 13(3): 391-7. 190. Klein EA, Thompson IM. “Update on chemoprevention of prostate cancer.” Curr Opin Urol, 2004; 14(3):143-9. 191. Gianduzzo TR, Holmes EG, Tinggi U, Shahin M, Mactaggart P, Nicol D. “Prostatic and peripheral blood selenium levels after oral supplementation.” J Urol, 2003; 170:870-3. 192. Klein EA, Lippman SM, Thompson IM, Goodman PJ, Albanes D, Taylor PR, Coltman C. “The selenium and vitamin E cancer prevention trial.” World J Urol, 2003; 21(1): 21-7. 193. Evans S, Gillatt DA, Persad R. “Prostate cancer and race.” Urology News, 2003; 7(6):12-7. 194. Hebert JR, Hurley TG, Olendzki BC, Teas J, Ma Y, Hampl JS. “Nutritional and socioeconomic factors in relation to prostate cancer mortality: a cross-national study.” J Natl Cancer Inst, 1998; 90(21): 1637-47. 195. Lee MM, Gomez SL, Chang JS, Wey M, Wang RT, Hsing AW. “Soy and isoflavone consumption in relation to prostate cancer risk in China.” Cancer Epidemiol Biomarkers Prev, 2003; 12(7): 665-8. 196. De Vere White R, Hackman R, Soares S. “Use of a genestine-enriched compound for secondary chemoprevention of prostate cancer.” Urological Research Society abstract, 2005. 197. Dalais FS, Meliala A, Wattanapenpaiboon N, Frydenberg M, Suter DA, Thomson WK, Wahlqvist ML. “Effects of a diet rich in phytoestrogens on prostate-specific antigen and sex hormones in men diagnosed with prostate cancer.” Urology, 2004; 64(3): 510-5. 198. Schröder FH, Roobol MJ, Boeve ER, de Mutsert R, Zuijdgeest-van Leeuwen SD, Kersten I, Wildhagen MF, van Helvoort A. “Randomized, double-blind, placebocontrolled crossover study in men with prostate cancer and rising PSA: effectiveness of a dietary supplement.” Eur Urol,2005; 48(6):922-30 199. Jacobs EJ, Rodriguez C, Mondul AM, Connell CJ, Henley SJ, Calle EE, Thun MJ. “A large cohort study of aspirin and other nonsteroidal anti-inflammatory drugs and prostate cancer incidence.” J Natl Cancer Inst, 2005; 97(13): 975-80. 200. Basler JW, Piazza GA. “Nonsteroidal anti-inflammatory drugs and cyclooxygenase-2 selective inhibitors for prostate cancer chemoprevention.” J Urol, 2004; 171(2 Pt 2): S59-62. 201. Pruthi RS, Derksen E, Gaston K. “Cyclooxygenase-2 as a potential target in the prevention and treatment of genitourinary tumors: a review.” J Urol, 2003; 169(6): 2352-9. 202. Hussain T, Gupta S, Mukhtar H. “Cyclooxygenase-2 and prostate carcinogenesis.” Cancer Lett, 2003; 191(2): 125-35. 203. Pan Y, Zhang JS, Gazi MH, Young CY. “The cyclooxygenase 2-specific nonsteroidal anti-inflammatory drugs celecoxib and nimesulide inhibit androgen receptor activity via induction of c-Jun in prostate cancer cells.” Cancer Epidemiol Biomarkers Prev, 2003; 12(8): 769-74. 204. Thompson IM, Goodman PJ, Tangen CM, Lucia MS, Miller GJ, Ford LG, Lieber MM, Cespedes RD, Atkins JN, Lippman SM, Carlin SM, Ryan A, Szczepanek CM, Crowley JJ, Coltman CA Jr. “The influence of finasteride on the development of prostate cancer.” N Engl J Med, 2003; 349(3): 215-24.

88

Chapter 10. PROSTATE CANCER 205. Klein EA, Tangen CM, Goodman PJ, Lippman SM, Thompson IM. “Assessing benefit and risk in the prevention of prostate cancer: the prostate cancer prevention trial revisited.” J Clin Oncol, 2005; 23: 7460-6. 206. Imamov O, Lopatkin NA, Gustafsson J-A. “Estrogen receptor ß in prostate cancer.” N Engl J Med, 2004; 351(26): 2773-4. 207. Albertsen PC, Fryback DG, Storer BE, Kolon TF, Fine J. “Long-term survival among men with conservatively treated localized prostate cancer.” JAMA, 1995; 274(8): 626-31. 208. Albertsen P-C, Fryback DG, Storer BE, Kolon TF, Fine J. “The impact of comorbidity on life expectancy among men with localized prostate cancer.” J Urol, 1996; 156(1): 127-32. 209. Parker CC, Gospodarowicz M, Warde P. “Does age influence the behaviour of localized prostate cancer?” BJU Int, 201; 87(7): 629-37. 210. Lu-Yao GL, Yao SL. “Population-based study of long-term survival in patients with clinically localised prostate cancer.” Lancet, 1997; 349(9056): 906-10. 211. Albertsen PC, Hanley JA, Fine J. “20 year outcomes following conservative management of clinically localized prostate cancer.” JAMA, 2005; 293: 2095-2101. 212. Baade PD, Steginga SK, Pinnock C & Aitken JF. “Communicating prostate cancer risk: What should we be telling our patients.” Medical Journal of Australia, 2005; 182, 472-475. 213. Melia J, Hewitson P, Austoker J. “Introduction: Review of screening for Prostate cancer.” British Journal of Urology International, 2005; 95(Supplement 3), 1-3. 214. Australian Cancer Network Working Party on Management of Localised Prostate cancer. (2002). Clinical practice guidelines: Evidence-based information and recommendations for the management of localised Prostate cancer. Canberra: National Health and Medical Research Council. 215. Green HJ, Pakenham KI, Headley BC & Gardiner RA. “Coping and healthrelated quality of life in men with prostate cancer randomly assigned to hormonal medication or close monitoring.” Psycho-Oncolgy, 11, 401-414. 216. O’Connor AM, Rostom A, Fiset V, Tetroe J, Entwistle V, Llewellyn-Thomas H, et al. “Decision aids for patients facing health treatment or screening decisions: Systematic review.” British Medical Journal, 1999; 319, 731-734. 217. Frosch DL, Kaplan RM. “Shared decision making in clinical medicine: Past research and future directions.” American journal of Preventative Medicine, 1999; 17(4), 285-294. 218. Elwyn G, Edwards A, Kinnersley P. “Shared decision-making in primary care: The neglected second half of the consultation.” British Journal of Primary Practice, 1999; 49, 477-482. 219. Davison BJ, Degner LF & Morgan TR. “Information and decision-making preferences of men with prostate cancer.” Oncology Nurses Forum, 1995; 22(9), 1404-1408. 220. Davison BJ, Gleave ME, Goldenberg SL, Degner LF, Hoffart D & Berkowitz J. “Assessing information and decision preferences of men with prostate cancer and their partners.” Cancer Nursing, 2002; 25(1), 42-49. 221. Barry MJ. “Involving patients in medical decisions: How can physicians do better?” Journal of the American Medical Association, 1999; 282(24), 2356-2357. 222. Braddock CH, Edwards KA, Hasenberg NM, Laidley TL & Levinson W. “Informed decison making in outpatient practice: Time to get back to basics.” Journal of the American Medical Association, 1999; 282(24), 2313-2320. 223. O’Connor AM, Stacey D, Rovner D, Holmes-Rovner M, Tetroe J, LlewellynThomas H. “Decision aids for people facing health treatment or screening decisions” (Cochrane review) (Vol. Issue 3): Oxford: 2001; Update Software. 89

Chapter 10. PROSTATE CANCER 224. Australian Health Technology Advisory Committee. “Prostate cancer screening.” Canberra: Australian Government Publishing Service, 1996. 225. The Cancer Council Australia. “National cancer prevention policy 2004-2006.” NSW: The Cancer Council Australia, 2004. 226. Wilkinson S, Chodak G. “Informed consent for prostate-specific antigen screening.” Urology, 2003; 61(1), 2-4. 227. Bird S. “Discussing benefits and risks with patients psa testing.” Australian Family Physician, 2004; 33(4), 266-267. 228. Dunn IB, Kirk D. “Legal pitfalls in the diagnosis of prostate cancer.” British Journal of Urology International, 2000; 86, 304-307. 229. Carter HB. “Informed consent for prostate-specific antigen screening.” Urology, 2003; 61, 13-14. 230. Ito K, Schroder FH. “Informed consent for prostate-specific antigen-based screening - European view.” Urology, 2003; 61, 20-22. 231. Hewitson P, Austoker J. “Part 2: Patient information, informed decision making and the psycho-social impact of prostate-specific antigen testing.” British Journal of Urology International, 2005; 95(Supplement 3), 16-32. 232. Chan E, Sulmasy D. “What men should know about prostate-specfic antigen screening before giving consent.” American Journal of Medicine, 1998; 105, 266-274. 233. General Medical Council. “Seeking patients’ considerations.” London: 1998; General Medical Council. consent: The ethical

234. Bratt O, Damber JE, Emanualsson M, KristoffersonU, Lundgren R, Olsson H, Gronberg H. “Risk perception, screening practice and interest in genetic testing among unaffected men in families with hereditary prostate cancer.” European Journal of Cancer, 2000; 36, 235-241. 235. Browne CT, Van der Meulen J, Newman S, Mundy AR & Emberton M. “The fear of prostate cancer in men with lower urinary tract symptoms: Should symptomatic men be screened?” British Journal of Urology, 2003; 91, 30-32. 236. Roberts RO, Rhodes T, Panser LA, Girman CJ, Chute CG, Oesterling JE, Lieber MM, Jacobsen SJ. “Natural history of prostatism: Worry and embarrassment from urinary symptoms and health care seeking behavior.” Urology, 1994; 43, 621-628. 237. Lloyd A, Hayes P, Bell PR, & Naylor AR. “The role of risk and benefit perception in informed consent for surgery.” Medical Decision Making, 2001; 21, 141-149. 238. Redelmeier DA, Rozin P & Kahneman D. “Understanding patients’ decisions: Cognitive and emotional perspectives.” Journal of the American Medical Association, 1993; 270(1), 72-76. 239. Steginga SK, Occhipinti S, Gardiner RA, Yaxley J & Heathcote P. “Making decisions about treatment for localised prostate cancer.” British Journal of Urology International, 2002; 89(3), 255-260. 240. Gigerenzer G, Edwards A. “Simple tools for understanding risks: From innumeracy to insight.” British Medical Journal, 2003; 327, 741-744. 241. Gurmankin AD, Baron J, & Armstrong K. “The effect of numerical statements of risk on trust and comfort with hypothetical physician risk communication.” Medical Decision Making, 2004; 24, 265-271. 242. Paling J. “Strategies to help patients understand risks.” British Medical Journal, 2003; 327, 745-748. 243. Brawer MK. “The diagnosis of prostatic carcinoma.” Cancer, 1993; 71 (3 Suppl):899. 244. Jewitt, HJ. “Significance of the palpable prostatic nodule.” JAMA, 1956; 160:838. 90

Chapter 10. PROSTATE CANCER 245. Catalona WJ, Richie JP, Ahmann FR, Hudson MA, Scardino PT, Flanigan RC, deKernion JB, Ratliff TL, Kavoussi LR, Dalkin BL. “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(5):1283. 246. Ellis WJ, Chetner MP, Prestion SD, Brawer MK. “Diagnosis of prostatic carcinoma: the yield of serum prostate specific antigen, digital rectal examination, and transrectal ultrasonography.” Journal of Urology, 1994; 152(5):1520. 247. AUA Commentary. “Prostate-specific antigen best practice policy.” Oncology, 2000; 14(2):267. 248. Partin AW, Yoo J, Carter HB, Pearson JD, Chan DW, Epstein JI, Walsh PC. “The use of prostate specific antigen, clinical stage and Gleason score to predict pathological stage in men with localized prostate cancer.” 1993; Journal of Urology, 150:110. 249. Thompson IM, Rounder JB, Teague JL, Peek M, Spence CR. “Impact of direct screening for adenocarcinoma of the prostate on stage distribution.” Journal of Urology, 1987; 137(3):424. 250. McLaughlin AP, Saltzstein SL, McCullough DL, Gittes RF. “Prostatic carcinoma: incidence and location of unsuspected lymphatic metastases.” Journal of Urology, 1976; 115(1):89. 251. Lodding P, Aus G, Bergdahl S, Frösing R, Lilja H, Pihl CG, Hugosson J. “Characteristics of screening detected prostate cancer in men 50 to 66 years old with 3 to 4 ng/ml Prostate Specific Antigen.” J Urol, 1998; 159:899-903. 252. Catalona WJ. “Prostate cancer screening.” BJU Int, 1996; 94(7):964-6. 253. Kirby R, Fitzpatrick J. “Prostate-specific antigen testing for the early detection of prostate cancer.” BJU Int, 2004; 94(7):966-7 254. Smith RA, Cokkinides V, Eyre HJ. “American Cancer Society guidelines for the early detection of cancer.” CA Cancer J Clin, 2003; 53: 27-43. 255. Fitzpatrick J. “PSA screening for prostate cancer.” Urol News, 2004; 9 (1): 6-9. 256. Donovan JL, Hamdy FC, Neal DE. “Screening for prostate cancer. The case against.” Prostate-Testing for Cancer and Treatment study group. Ann R Coll Surg Engl, 2005; 87(2): 90-1. 257. Stenman UH, Abrahamsson PA, Aus G, Lilja H, Bangma C, Hamdy FC, BocconGibod L, Ekman P. “Prognostic value of serum markers for prostate cancer.” Scand J Urol Nephrol Suppl.2005; May;(216):64-81. 258. Brawer MK. “Prostate-specific antigen: current status.” CA Cancer Journal for Clinicians, 1999; 49(5):264. 259. Catalona WJ, Smith DS, Ornstein DK. “Prostate cancer detection in men with serum PSA concentrations of 2.6 to 4.0 ng/ml and benign prostate examination: enhancement of specificity of PSA measurement.” JAMA, 1997; 277(18):1452. 260. Partin AW, Kattan MW, Subong EN, Walsh PC, Wojno KJ, Oesterling JE, Scardino PT, Pearson JD. “Combination 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. 261. Brossner C, Bayer G, Madersbacher S, Kuber W, Klingler C, Pycha A. “Twelve prostate biopsies detect significant cancer volumes (> 0.5 mL).” BJU Int, 2000; 85:7057. 262. Naughton CK, Miller DC, Mager DE, Ornstein DK, Catalona WJ. “A prospective randomized trial comparing 6 versus 12 prostate biopsy cores: impact on cancer detection.” J Urol, 2000; 164:388-92. 263. Presti JC Jr, Chang JJ, Bhargava V, Shinohara K. “The optimal systematic prostate biopsy scheme should include 8 rather than 6 biopsies: results of a prospective clinical trial.” J Urol, 2000; 163:163-7. 91

Chapter 10. PROSTATE CANCER 264. Hessels D, Verhaegh GW, Schalken JA, Witjes JA. “Applicability of biomarkers in the early diagnosis of prostate cancer.” Expert Rev Mol Diagn,2004; 4:513-26. 265. Thompson IM, Ankerst DP, Chi C, Lucia MS, Goodman PJ, Crowley JJ, Parnes HL, Coltman CA Jr. “Operating 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. 266. Pelzer AE, Tewari A, Bektic J, Berger AP, Frauscher F, Bartsch G, Horninger W. “Detection rates and biologic significance of prostate cancer with PSA less than 4.0 ng/mL: observation and clinical implications from Tyrol screening project.” Urology, 2005; 66(5):1029-33. 267. Aus G, Damber JE, Khatami A, Lilja H, Stranne J, Hugosson J. “Individualized screening interval for prostate cancer based on prostate-specific antigen level: results of a prospective, randomized, population-based study.” Arch Intern Med,2005; 165(16):1857-61. 268. Roobol MJ, Roobol DW, Schroder FH. “Is additional testing necessary in men with prostate-specific antigen levels of 1.0 ng/mL or less in a population-based screening setting?” (ERSPC, section Rotterdam). Urology, 2005; 65(2):343-6. 269. Oesterling JE, Jacobsen SJ, Klee GG, Pettersson K, Piironen T, Abrahamsson PA, Stenman UH, Dowel B, Lovgren T, Lilja H. “Free complexed and total serum prostate antigen: the establishment of appropriate reference ranges for their concentrations and ratios.” J Urol, 1995; 154:1090-5. 270. Fang J, Metter EJ, Landis P, Chan DW, Morrell CH, Carter HB. “Low levels of prostate-specific antigen predict long-term risk of prostate cancer: Results from the Baltimore longitudinal study of aging.” Urology, 2001; 58:411-6. 271. Carter HB, Pearson JD, Metter EJ, Brant LJ, Chan DW, Andres R, Fozard JL, Walsh PC. “Longitudinal evaluation of prostate –specific antigen levels in men and without prostate disease.” JAMA, 1992; 267:2215-20. 272. Ornstein DK, Smith DS, Humphrey PA, Catalona WJ. “The effect of prostate volume, age, total prostate specific antigen level and acute inflammation on the percentage of free serum prostate specific antigen levels in men without clinically detectable Prostate cancer.” Journal of Urology, 1998; 159(4):1234. 273. Yuan JJ, Coplen DE, Petros JA, Figenshau RS, Ratliff TL, Smith DS, Catalona WJ. “Effects of rectal examination, prostatic massage, ultrasonography, and needle biopsy of prostate-specific antigen levels.” Journal of Urology, 1992; 147(6):810. 274. Guess HA, Heyse JF, Gormley GJ, Stoner E, Oesterling JE. Effect of finasteride on serum PSA concentration in men with benign prostatic hyperplasia: results from the North American Phase III clinical trial. Urologic Clinics of North America, 1993; 20(4):627. 275. de la Taille A, Katz A, Bagiella E, Olsson CA. O’Toole KM. Rubin MA. Perineural invasion on prostate needle biopsy: an independent predictor of final pathologic stage. Urology, 1999; 54(6):1039. 276. Clements JA, Merritt T, Devoss K, Swanson C, Hamlyn L, Scells B, Rohde P, Lavin MF, Yaxley J, Gardiner RA. “Inactive free : total prostate specific antigen ratios in ejaculate from men with suspected and known prostate cancer, compared with young control men.” BJU Int.2000; Sep;86(4):453-8. 277. Savblom C, Malm J, Giwercman A, Nilsson JA, Berglund G, Lilja H. “Blood levels of free-PSA but not complex-PSA significantly correlates to prostate release of PSA in semen in young men, while blood levels of complex-PSA, but not free-PSA increase with age.” Prostate.2005; Sep 15;65(1):66-72. 278. Lilja H. “Significance of different molecular forms of serum PSA.” Urologic Clinics of North America, 1993; 20(4):681. 279. Carter, HB, Pearson, JD. “PSA velocity for the diagnosis of early prostate cancer.” Urologic Clinics of North America, 1993; 20(4):665. 92

Chapter 10. PROSTATE CANCER 280. Seaman E, Whang M, Olsson CA, Katz A, Cooner WH, Benson MC. “PSA density: role in patient evaluation and management.” Urologic Clinics of North America, 1993; 20(4):653. 281. Daneshagari F, Taylor GD, Miller GJ, Crawford ED. “Computer simulation of the probability of detecting low volume carcinoma of the prostate with six random systematic core biopsies.” Urology, 1995; 45:604-9. 282. Gore JL, Shariat SF, Miles BJ, Kadmon D, Jiang N, Wheeler TM, Slawin KM. “Optimal combinations of systematic sextant and laterally directed biopsies for the detection of prostate cancer.” J Urol, 2001; 165:1554-9. 283. Djavan B, Ravery V, Zlotta A, Dobronski P, Dobrovits M, Fakhari M, Seitz C, Susani M, Borkowski A, Boccon-Gibod L, Schulman CC, Marberger M. “Prospective evaluation of prostate cancer detected on biopsies 1, 2, 3 and 4: when should we stop?” J Urol, 2001; 166:1679-83. 284. Schröder FH, Albertsen P, Boyle P, Candas B, Catalon AW, Cheng C, DeKoning HJ, Fourcade R, Hugosson J, Moul J, Perrin P, Roehrborn C, Rübben H, Stephenson R, Yamanaka H. “Early Detection and Screening for prostate cancer,” In: Prostate cancer, 3rd International Consultation Eds: L Denis, G Bartsch, S Khoury, M Murai, A Partin. Editions 21, Paris, France. 2003. 285. Vashi AR, Wojno KJ, Gillespie B, Oesterling JE. “A model for the number of cores per prostate biopsy based on patient age and prostate gland volume.” Journal of Urology, 1998; 159(3):920. 286. Bostwick DG. “Evaluating prostate needle biopsy: therapeutic and prognostic importance.” Ca: A Cancer Journal for Clinicians 1993; 47(5):297. 287. Stricker HJ, Ruddock LJ, Wan J, Belville WD. “Detection of non-palpable prostate cancer. A mathematical and laboratory model.” British Journal of Urology, 1993; 71(1):43. 288. Yanke BV, Gonen M, Scardino PT, Kattan MW. “Validation of a nomogram for predicting positive repeat biopsy for prostate cancer.” J Urol, 2005; 173 (2):421-4. 289. Enlund AL, Varenhorst E. “Morbidity of ultrasound-guided transrectal core biopsy of the prostate without prophylactic antibiotic therapy. A prospective study in 415 cases.” Br J Urol, 1997; 997; 79(5): 777-80. 290. Rodriguez, LV, Terris, MK. “Risks and complications of transrectal ultrasound guided prostate needle biopsy: a prospective study and review of the literature.” Journal of Urology, 1998; 160(6):2115. 291. Rodriguez LV, Terris MK. “Risks and complications of transrectal ultrasound.” Curr Opin Urol, 2000; 111-116. 292. Byar DP, Mostofi FK, Veteran’s Administrative Cooperative Urologic Research Group. “Carcinoma of the prostate: prognostic evaluation of certain pathologic features in 208 radical prostatectomies.” Cancer 1972; 30(1):5. 293. Mostofi FK, Davis CJ, Sesterhenn IA. “Histopathology of prostate cancer.” In: Prostate Diseases. H Lepor and RK Lawson, editors. Philadelphia: WB Saunders Co. 1993; p 229. 294. Sebo TJ, Bock BJ, Cheville JC, Lohse C, Wollan P, Zincke H. “The percent of cores positive for cancer in prostate needle biopsy specimens is strongly predictive of tumor stage and volume at radical prostatectomy.” Journal of Urology, 2000; 163(1):174. 295. Sanwick JM, Dalkin BL, Nagle RB. “Accuracy of prostate needle biopsy in predicting extracapsular tumor extension at radical retropubic prostatectomy: application in selecting patients for nerve-sparing surgery.” Urology, 1998; 52(5):814. 296. Gleason DF. “Histologic grading of prostate cancer: a perspective.” Human Pathology, 1992; 23(3):273.

93

Chapter 10. PROSTATE CANCER 297. Epstein JI, Partin AW, Sauvegout J, Walsh PC. “Prediction of progression following radical prostatectomy: a multivariate analysis of 721 men with long-term followup.” American Journal of Surgical Pathology, 1996; 20(2):286. 298. McNeal JE, Villers AA, Redwine EA, Freiha FS, Stamey TA. “Histologic differentiation, cancer volume, and pelvic lymph node metastasis in adenocarcinoma of the prostate.” Cancer, 1990; 66(6):1225. 299. Stamey TA, McNeal JE, Yemoto CM, Sigal BM, Johnstone IM. “Biological determinants of cancer progression in men with prostate cancer.” JAMA, 1990; 281(15):1395. 300. Haggman MJ, Macoska JA, Wojno KJ, Oesterling JE. “The relationship between prostatic intraepithelial neoplasia and prostate cancer: critical issues.” Journal of Urology, 1997; 158(1):12. 301. Bostwick DG, Pacelli A, Lopez-Beltran A. “Molecular biology of prostatic intraepithelial neoplasia.” Prostate, 1997; 29(1):117-134. 302. Bostwick DG, Brawer MK. “Prostatic intra-epithelial neoplasia and early invasion in Prostate cancer.” Cancer, 1987; 59(4):788. 303. McNeal JE, Bostwick DG. “Intraductal dysplasia: a premalignant lesion of the p Human Pathology”, 1986; 17(1):64. 304. Schlesinger C, Bostwick DG, Iczkowski KA. “High-grade prostatic intraepithelial neoplasia and atypical small acinar proliferation: predictive value for cancer in current practice” Am J Surg Pathol, 2005; 29 (9): 1201-1207. 305. Iczkowski KA, Basseler TJ, Schwob VS, Basseler IC, Kunnel BS, Orozco RE, Bostwick DG. “Diagnosis of "suspicious for malignancy" in prostate biopsies: predictive value for cancer.” Urology, 1998; 51(5):749-757. 306. Chan TY, Epstein JI. “Follow-up of atypical prostate needle biopsies suspicious for cancer.” Urology. 1999; 53(2):351-5. 307. Allen EA, Kahane H, Epstein JI. “Repeat biopsy strategies for men with atypical diagnoses on initial prostate needle biopsy.” Urology, 1998; 52:803-807. 308. Huggins C & Hodges CV. “Studies on prostatic cancer. The effect of castration, of estrogen, and of androgen injection on serum phosphates in metastatic carcinoma of the prostate.” Cancer Research, 1941; 1(1):293. 309. Huggins C, Hodges CV. “Studies of prostate cancer: II The effects of castration on advanced carcinoma.” Arch Surg, 1941; 43:209-28. 310. Small EJ, Reese DM, Vogelzang NJ. “Hormone-refractory prostate cancer: an evolving standard of care.” Seminars in Oncology, 1999; 26(5):61. 311. Bill-Axelson A, Holmberg L, Ruutu M, Haggman M, Andersson SO, Bratell S, Spangberg A, Busch C, Nordling S, Garmo H, Palmgren J, Adami HO, Norlen BJ, Johansson JE. “Radical prostatectomy versus watchful waiting in early prostate cancer.” Scandinavian-Prostate-Cancer-Group-Study-No.-4. N Engl J Med, 2005; 352(19): 1977-84. 312. Holmberg L, Bill-Axelson A, Helgesen F, Salo JO, Folmerz P, Haggman M, Andersson SO, Spangberg A, Busch C, Nordling S, Palmgren J, Adami HO, Johansson JE, Norlen BJ. “A randomized trial comparing radical prostatectomy with watchful waiting in early prostate cancer.” Scandinavian-Prostatic-Cancer-Group-StudyNumber-4 N Engl J Med, 2002; 347(11): 781-9. 313. Pickles T, Goldenberg SL, Steinhoff G. “Technology review: High-Intensity Focused Ultrasound for prostate cancer.” Canadian J Urol, 2005; 12:2593-7. 314. Theodorescu D. “Cancer cryotherapy: evolution and biology.” Urology, 2004; 6(suppl 4): S9-19. 315. Cohen JK. “Cryosurgery of the prostate: techniques and indications.” Urology, 2004; 6 (suppl 4): S20-6. 94

Chapter 10. PROSTATE CANCER 316. Steginga SK, Occhipinti S. “The application of the heuristic systematic processing model to treatment decision making about prostate cancer.” Medical Decision Making, 2004; 24, 573-583. 317. Harlan LC, Potosky A, Gilliland FD, Hoffman R, Albertsen PC, Hamilton AS, Eley JW, Stanford JL, Stephenson RA. “Factors associated with initial therapy for clinically localised prostate cancer: prostate cancer outcomes study.” Journal of the National Cancer Institute, 2001; 93(24), 1864-1871. 318. Post PN, Kil PJM, Hendrik AJM., Janssen-Heijnen MLG, Crommelin MA & Coebergh JWW. “Comorbidity in patients with prostate cancer and its relevance to treatment choice.” British Journal of Urology International, 1999; 84, 652-656. 319. Yan Y, Carvalhal GF, Catalona WJ & Young JD. “Primary treatment choices for men with clinically localised prostate cancer detected by screening.” Cancer, 2000; 88, 1122-1130. 320. Boehmer U, Clark JA. “Married couples perspectives on prostate cancer diagnosis and treatment decision making.” Pscho-Oncology, 2001; 10, 147-155. 321. Demark-Wahnefried W, Schildkraut JM, Iselin CE, Conlisk E, Kavee A, Aldrich TE, Lengerich EJ, Walther PJ, Paulson DF. “Treatment options, selection, and satisfaction among african american and white men with prostate carcinoma in North Carolina.” Cancer, 1998; 83, 320-330. 322. Diefenbach MA, Dorsey J, Uzzo RG, Hanks GE, Greenberg RE, Horwitz E, Newton F, Engstrom PF. “Decision-making strategies for patients with localized prostate cancer.” Seminars in Urologic Oncology, 1998; 20(1), 55-62. 323. Holmboe ES, Concato J. “Treatment decisions for localized prostate cancer.” Archives of General Internal Medicine, 2000; 15, 694-701. 324. O’Rourke ME, Germino BB. “Prostate cancer treatment decisions: A focus group exploration.” Oncology Nurses Forum, 1998; 25(1), 97-104. 325. Mazur DJ, Hickman DH. “Patient preferences for management of localized prostate cancer.” Western Journal of Medicine, 1996; 165, 26-30. 326. O’Rourke ME. “Narrowing the options: The process of deciding on prostate cancer treatment.” Cancer Investigation, 1999; 17(5), 349-359. 327. Steginga SK, Occhipinti S. “Dispositional optimism as a predictor of men’s decision-related distress after localized prostate cancer”. Health Psychology (In press). 328. Folkman, S, Lazarus RS. “Coping as a mediator of emotion.” Journal of Personality and Social Psychology, 1998; 54(3), 466-475. 329. Nezu AM, Nezu CM, Friedman SH, FaddisS & Houts PS. “A problem-solving approach helping cancer patients cope.” Washington: American Psychological Association, 1998. 330. Dunn J, Steginga SK, Millichap D & Rosoman N. “A review of peer support in the context of cancer.” Journal of Psychosocial Oncology, 2003; 21(2), 55-67. 331. Coreil J, Behal R. “Man to man prostate cancer support groups.” 1999; Cancer Practice, 7(3), 122-129. 332. Manne S. “Prostate cancer support and advocacy groups: Their role for patients and family members.” Seminars in Urologic Oncology, 2002; 20(1), 45-54. 333. Steginga SK, Pinnock C, Gardner M, Gardiner RA & Dunn J. “Evaluating peer support for prostate cancer: The prostate cancer peer support inventory.” British Journal of Urology International, 2005; 95, 46-50. 334. Epstein JI, Walsh PC, Carmichael M, Brender CB. “Pathologic and clinical findings to predict tumor extent of nonpalpable (Stage T1c) prostate cancer.” JAMA, 1994; 271:368-74. 95

Chapter 10. PROSTATE CANCER 335. Bastian PJ, Mangold LA, Epstein JI, Partin AW. “Characteristics of insignificant clinical T1c prostate tumors: a contemporary analysis.” Cancer, 2004; 101:2001-5. 336. Catalona WJ, Smith DS. “Five-year tumor recurrence rates after anatomic radical retropubic prostatectomy for prostate cancer.” Journal of Urology, 1994; 152(5):1837. 337. Catalona WJ, Smith DS. “Cancer recurrence and survival rates after anatomic radical retropubic prostatectomy for prostate cancer: intermediate-term results.” Journal of Urology 1998; 160(6):2428. 338. Walsh PC. “Radical prostatectomy for localized prostate cancer provides durable cancer control with excellent quality of life: a structured debate.” Journal of Urology, 2000, 163(6):1802. 339. Epstein JI. “Diagnostic criteria of limited adenocarcinoma of the prostate on needle biopsy.” Human Pathology, 1995; 26(2):223. 340. Oesterling JE, Martin SK, Bergstralh EJ, Lowe FC. “The use of prostate-specific antigen in staging patients with newly diagnosed prostate cancer.” JAMA, 1993; 269(1):57. 341. Walsh PC, Jewitt HJ. “Radical surgery for prostate cancer.” Cancer, 1980; 45(Suppl):1906. 342. Lu-Yao G, McLarran D, Wasson J, Wennberg JE. “An assessment of radical prostatectomy.” JAMA, 1993; 269:2633. 343. Voges GE, McNeal JE, Redwine EA, Freiha FS, Stamey TA. “Morphologic analysis of surgical margins with positive findings in prostatectomy for adenocarcinoma of the prostate.” Cancer 1992; 69:520. 344. Lerner SP, Seale-Hawkins C, Carlton CE, Scardino PT. “The risk of dying of prostate cancer in patients with clinically localized disease” Journal of Urology, 1991; 146:1040. 345. Schellhammer PF. “Radical prostatectomy: patterns of local failure and survival.” Urology, 1988; 31:191. 346. Catalona WJ, Dresner SM. “Nerve-sparing radical prostatectomy: extraprostatic tumor extension and preservation of erectile function.” Journal of Urology, 1985; 134:1149. 347. Kaplan RN, Riba RD, Zacharoulis S, Bramley AH, Vincent L, Costa C, MacDonald D, Jin D, Shido K, Kerns S, Zhu Z, Hicklin D, Wu Y, Port JL, Altorki N, Port ER, Ruggero D, Shmelkov SV, Jensen KK, Rafii S, Lyden D. “VEGFR1-positive haematopoietic bone marrow progenitors initiate the pre-metastatic niche.” Nature, 2005; 438/8:820-7. 348. Khan MA, Partin AW. “Partin tables: past and present.” BJU Int, 2003; 92: 7-11. 349. Eskicorapci SY, Karabulut E, Turkeri L, Baltaci S, Cal C, Toktas G, Akpinar H, Ozer G, Sozen S, Tokuc R, Lekili M, Soylu A, Albayrak S, Sahin H, Alpar R, Ozen H. “Validation of 2001 Partin tables in Turkey: a multicenter study.” Eur Urol,2005; 47(2):185-9. 350. Augustin H, Eggert T, Wenske S, Karakiewicz PI, Palisaar J, Daghofer F, Huland H, Graefen M. “Comparison of accuracy between the Partin tables of 1997 and 2001 to predict final pathological stage in clinically localized prostate cancer.” J Urol, 2004; 171(1):177-81. 351. Steuber T, Karakiewicz PI, Augustin H, Erbersdobler A, Lange I, Haese A, Chun KH, Walz J, Graefen M, Huland H. “Transition zone cancers undermine the predictive accuracy of Partin table stage predictions.” J Urol,2005; 173(3):737-41. 352. Kattan MW, Eastham JA, Stapleton AM, Wheeler TM, Scardino PT. “A preoperative nomogram for disease recurrence following radical prostatectomy for prostate cancer.” Journal of the National Cancer Institute, 1998; 90(10):766-71.

96

Chapter 10. PROSTATE CANCER 353. Graefen M, Karakiewicz PI, Cagiannos I, Klein E, Kupelian PA, Quinn DI, Henshall SM, Grygiel JJ, Sutherland RL, Stricker PD, de Kernion J, Cangiano T, Schroder FH, Wildhagen MF, Scardino PT, Kattan MW. “Validation study of the accuracy of a postoperative nomogram for recurrence after radical prostatectomy for localized prostate cancer.” J Clin Oncol.2002; Feb 15;20(4):951-6. 354. Karakiewicz PI, Benayoun S, Kattan MW, Perrotte P, Valiquette L, Scardino PT, Cagiannos I, Heinzer H, Tanguay S, Aprikian AG, Huland H, Graefen M. J Urol.2005; Jun;173(6):1930-4. 355. Stephenson AJ, Scardino PT, Eastham JA, Bianco FJ Jr, Dotan ZA, DiBlasio CJ, Reuther A, Klein EA, Kattan MW. “Postoperative nomogram predicting the 10-year probability of prostate cancer recurrence after radical prostatectomy.” J Clin Oncol, 2005;23:7005-12. 356. Partin AW, Piantadosi S, Sanda MG, Epstein JI, Marshall FF, Mohler JL, Brendler CB, Walsh PC, Simons JW. “Selection of men at high risk for disease recurrence for experimental adjuvant therapy following radical prostatectomy.” Urology, 1995;45:831838. 357. Moul JW Connelly RR, Lubeck DP, Bauer JJ, Sun L, Flanders SC, Grossfeld GD, Carroll P. “Predicting risk of prostate specific antigen recurrence after radical prostatectomy with the Center for Prostate Disease Research and Cancer of the Prostate Strategic Urologic Research Endeavor databases.” J Urol, 2001; 166:1322-1327. 358. Vollmer RT, Humphrey PA. “The relative importance of anatomic and PSA factors to outcomes after radical prostatectomy for prostate cancer.” Am J Clin Pathol, 2001; 116:864-870. 359. Ohori M, Wheeler TM, Dunn JK, Stamey TA, Scardino PT. “The pathological features and prognosis of prostate cancer detectable with current diagnostic tests.” Journal of Urology, 1994; 152(5):1712. 360. Gervasi LA, Mata J, Easley JD, Wilbanks JH, Seale-Hawkins C, Carlton CE Jr, Scardino PT. “Prognostic significance of lymph node metastases in prostate cancer.” Journal of Urology 1989, 142(2):332. 361. Malavaud B, Villers A, Ravery V, Tollon C, Rischmann P, Charlet JP, Plante P, Boccon-Gibod L, Sarramon JP. “Role of preoperative positive apical biopsies in the prediction of specimen-confined prostate cancer after radical retropubic prostatectomy: a multi-institutional study.” European Urology, 2000; 37(3):281. 362. Alsikafi NF, Brendler CB. “Surgical modifications of radical retropubic prostatectomy to decrease incidence of positive surgical margins.” Journal of Urology, 1998; 159(4):1281. 363. Stamey TA, Villers AA, McNeal JE, Link PC, Freiha FS. “Positive surgical margins at radical prostatectomy: importance of the apical dissection.” Journal of Urology 1990; 143(3):1143. 364. 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. 365. Gao CL, Maheshwari S, Dean RC, Tatum L, Mooneyhan R, Connelly RR, McLeod DG, Srivastava S, Moul JW. “Blinded evaluation of reverse transcriptase-polymerase chain reaction prostate-specific antigen peripheral blood assay for molecular staging of prostate cancer.” Urology 1999; 53(4):714. 366. Ellis WJ, Vessella RL, Corey E, Arfman EW, Oswin MM, Melchior S, Lange PH. “The value of a reverse transcriptase polymerase chain reaction assay in preoperative staging and follow-up of patients with prostate cancer.” Journal of Urology, 1998; 159(4):1134. 367. Sokoloff MH, Tso CL, Kaboo R, Nelson S, Ko J, Dorey F, Figlin RA, Pang S, deKernion J, Belldegrun A. “Quantitative polymerase chain reaction does not im97

Chapter 10. PROSTATE CANCER prove prostate cancer staging: a clinical-pathologic-molecular analysis of 121 patients.” Journal of Urology, 1996; 156(5):1560. 368. Oesterling JE, Brendler CB, Epstein JI, Kimball AW Jr, Walsh PC. “Correlation of clinical stage, serum prostatic acid phosphatase, and preoperative Gleason grade with final pathological stage in 275 patients with clinically localized adenocarcinoma of the prostate.” Journal of Urology, 1987; 138(1):92. 369. McNeal JE. “Cancer volume and site of origin of adenocarcinoma of the prostate: relationship to local and distant spread.” Human Pathology, 1992; 23(2):258. 370. Stamey TA, Freiha FS, McNeal JE, Redwine EA, Whittemore AS, Schmid HP. “Localized prostate cancer. Relationship of tumor volume to clinical significance for treatment of prostate cancer.” Cancer 1993; 71(3 Suppl):933. 371. Goto Y, Ohari M, Scardino PT. “Use of systemic biopsy results to predict pathologic stage in patients with clinically localized prostate cancer: a preliminary report.” International Journal of Urology 1998; 5(3):337. 372. Brassell SA, Rosner IL, McLeod DG. “Update on magnetic resonance imaging, ProstaScint and novel imaging in prostate cancer.” Curr Opin Urol, 2005; 15: 163-6. 373. Crook J, McLean M, Catton C, Yeung I, Tsihlias J, Pintilie M. “Factors influencing risk of acute urinary retention after TRUS-guided permanent prostate seed implantation.” Int J Radiat Oncol Biol Phys 2002; 52, 453-60. 374. Bolla M, Gonzalez D, Warde P, Dubois JB, Mirimanoff RO, Storme G, Bernier J, Kuten A, Sternberg C, Gil T, Collette L, Pierart M. “Improved survival in patients with locally advanced prostate cancer treated with radiotherapy and goserelin.” N Engl J Med, 1997; 337(5): 295-300. 375. Bolla M, Collette L, Blank L, Warde P, Dubois J B, Mirimanoff RO, Storme G, Bernier J, Kuten A, Sternberg C, Mattelaer J, Lopez-Torecilla J, Pfeffer JR, Lino-Cutajar C, Zurlo A, Pierart M. 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. 376. Lamb DS, Denham JW, Mameghan H, Joseph D, Turner S, Matthews J, Franklin I, Atkinson C, North J, Poulsen M, Kovacev O, Robertson R, Francis L, Christie D, Spry NA, Tai KH, Wynne C, Duchesne G. Acceptability of short term neo-adjuvant androgen deprivation in patients with locally advanced prostate cancer. Radiother Oncol, 2003; 68(3): 255-67. 377. Pilepich MV, Winter K, Lawton C A, Krisch RE, Wolkov HB, Movsas B, Hug EB, Asbell SO, Grignon D. “Androgen suppression adjuvant to definitive radiotherapy in prostate carcinoma--long-term results of phase III RTOG 85-31.” Int J Radiat Oncol Biol Phys, 2005; 61(5): 1285-90. 378. Van den Ouden D, Hop WC, Schroder FH. “Progression in a survival of patients with locally advanced prostate cancer (T3) treated with radical prostatectomy as monotherapy.” Journal of Urology, 1998; 160(4):1392-7. 379. Potosky A, Davis WW, Hoffman RM, Stanford JL, Stephenson RA, Penson DF., et al. “Five-year outcomes after prostatectomy or radiotherapy for prostate cancer: The prostate cancer outcomes study.” Journal of the National Cancer Institute, 2004; 96(18), 1358-1367. 380. Schover LR, Fouladi RT, Warneke CL, Neese L, Klein EA, Zippe C, Kupelian PA. “Defining sexual outcomes after treatment for localized prostate cancer.” Cancer, 2002; 95, 1773-1785. 381. Steginga SK, Occhipinti S, Dunn J, Gardiner RA, Heathcote P, Yaxley J. “The supportive care needs of men with prostate cancer.” Psycho-Oncology, 2001; 10, 6675. 382. Steginga SK, Occhipinti S, Gardiner RA, Yaxley J, & Heathcote P. “A prospective study of men’s psychological and decision-related adjustment after treatment for localized prostate cancer.” Urology, 2004; 63, 751-756. 98

Chapter 10. PROSTATE CANCER 383. Savoie M, Kim SS & Soloway MS. “A prospective study measuring penile length in men treated with radical prostatectomy for prostate cancer.” The Journal of Urology, 2002; 169, 1462-1464. 384. Barnas JL, Pierpaoli S, Ladd P, Valenzuela R, Aviv N, Parker M, Waters WB, Flanigan RC, Mulhall JP. “The prevalence and nature of orgasmic dysfunction after radical prostatectomy.” British Journal of Urology International, 2004; 94, 603-605. 385. Bokhour BG, Clark JA, Thomas S, Silliman RA, Talcott JA. “Sexuality after treatment for early prostate cancer. Exploring the meanings of "erectile dysfunction". Journal of General Internal 2001; Medicine, 16, 649-655. 386. Gontero P, Fontana F, Bagnasacco A, Panella M, Kocjancic E, Pretti G, Frea B. “Is there an optimal time for intracavernous prostaglandin e1 rehabilitation following nonnerve sparing radical prostatectomy? Results from a hemodynamic prospective study.” The Journal of Urology, 2003; 169, 2166-2169. 387. Montorsi F, Guazzoni G, Strambi LF, Da Pozzo LF, Nava L, Barbieri L, Rigatti P, Pizzini G, Miani A. “Recovery of spontaneous erectile function after nerve sparing radical retropubic prostatectomy with and without early intracavernous injection sof alprostadil: Results of a prospective, randomised trial.” The Journal of Urology, 1997; 158(4), 1408-1410. 388. Schover LR, Fouladi RT, Warneke CL, Neese L, Klein EA, Zippe C, Kupelian PA. “Seeking help for erectile dysfunction after treatment for prostate cancer.” Archives of Sexual Behavior, 2004; 33(5), 443-454. 389. Stephenson RA, Mori M, Hsieh YC, Beer TM, Stanford JL, Gilliland FD, Hoffman RM, Potosky AL. “Treatment of erectile dysfunction following therapy for clinically localized and End Results Prostate Cancer Outcomes Study.” J Urol, 2005; 174: 646-50. 390. Davison BJ, Keyes M, Elliott S, Berkowitz J, Goldenberg SL. “Preferences for sexual information resources in patients treated for early stage prostate cancer with either radical prostatectomy or brachytherapy.” British Journal of Urology International, 2004; 93, 965-969. 391. Neese L, Schover LR, Klein EA, Zippe C, Kupelian PA. “Finding help for sexual problems after prostate cancer treatment: A phone survey of men’s and women’s perspectives.” Psycho-Oncology, 2003; 12, 463-473. 392. Canada AL, Neese L, Sui D & Schover LR. “A pilot intervention to enhance sexual rehabilitation for couples after treatment for localized prostate cancer”. Cancer. (In press). 393. Walsh, PC. “Radical retropubic prostatectomy with reduced morbidity: an anatomic approach.” NCI Monographs, 1988; 7:133. 394. Litwin MS, Hays RD, Fink A, Ganz PA, Leake B, Leach GE, Brook RH. “Qualityof-life outcomes in men treated for localized prostate cancer” JAMA, 1995; 273(2):129. 395. Litwin MS, McGuigan KA, Shpall AI, Dhanani N. “Recovery of health related quality of life in the year after radical prostatectomy: early experience.” Journal of Urology, 1999; 161(2):515. 396. Reiner WG, Walsh PC. “An anatomical approach to the surgical management of the dorsal vein and Santorinin’s plexus during radical retropubic surgery.” 1979; Journal of Urology, 121(1):198. 397. Walsh PC, Lepor H, Eggleston JC. “Radical prostatectomy with preservation of sexual function: anatomical and pathological considerations.” Prostate, 1983; 4:473. 398. Graefen M, Walz J, Chun KH, Schlomm T, Haese A, Huland H. „Reasonable delay of surgical treatment in men with localized prostate cancer--impact on prognosis?” Eur Urol.2005; Jun;47(6):756-60. 399. Boorjian SA, Bianco FJ Jr, Scardino PT, Eastham JA. “Does the time from biopsy to surgery affect biochemical recurrence after radical prostatectomy?” BJU Int, 2005; 96:773-6. 99

Chapter 10. PROSTATE CANCER 400. Worwag E, Chodak GW. “Overnight hospitalization after radical prostatectomy: the impact of two clinical pathways on patient satisfaction, length of hospitalization, and morbidity.” Anesthesia & Analgesia, 1998; 87(1):62. 401. Shir Y, Frank SM, Brendler CB, Raja SN. “Postoperative morbidity is similar in patients anesthetized with epidural and general anesthesia for radical prostatectomy.” Urology, 1994; 44(2):232. 402. Shir Y, Raja SN, Frank SM, Brendler CB. “Intraoperative blood loss during radical retropubic prostatectomy: epidural versus general anesthesia.” Urology, 1995; 45(6):993. 403. Helgason AR, Adolfsson J, Dickman P, Arver S, Fredrikson M, Gothberg M, Steineck G. “Sexual desire, erection, orgasm and ejaculatory functions and their importance to elderly Swedish men: a population-based study.” Age Ageing, 1996; 25(4): 285-91. 404. Blanker MH, Bosch JL, Groeneveld FP, Bohnen AM, Prins A, Thomas S, Hop WC. “Erectile and ejaculatory dysfunction in a community-based sample of men 50 to 78 years old: prevalence, concern, and relation to sexual activity.” Urology, 2001; 57(4): 763-8. 405. Singer PA, Tasch ES, Stocking C, Rubin S, Siegler M, Weichselbaum R.„Sex or survival: trade-offs between quality and quantity of life.” Journal of Clinical Oncology 1991; 9(2):328. 406. Mantz CA, Nautiyal J, Awan A, Kopnick M, Ray P, Kandel G, Niederberger C, Ignacio L, Dawson E, Fields R, Weichselbaum R, Vijayakumar S.« Potency preservation following conformal radiotherapy for localized prostate cancer: impact of neoadjuvant androgen blockade, treatment technique, and patient-related factors.” Cancer Journal From Scientific American 1999; 5(4):230. 407. Krupski T, Petroni GR, Bissonette EA, Theodorescu D. “Quality-of-life comparison of radical prostatectomy and interstitial brachytherapy in the treatment of clinically localized prostate cancer.” Urology, 2000; 55(5);736. 408. Jonler M, Messing EM, Rhodes PR, Bruskewitz RC. “Sequelae of radical prostatectomy.” British Journal of Urology, 1994; 74:352. 409. Fowler FJ Jr, Barry MJ, Lu-Yao G, Roman A, Wasson J, Wennberg JE. “Patientreported complications and follow-up treatment after radical prostatectomy. The National Medicare Experience: 1988-1990” (updated June 1993). Urology, 1993; 42:622. 410. Heathcote PS, Mactaggart PN, Boston RJ, James AN, Thompson LC, Nicol DL. “Health-related quality of life in Australian men remaining disease-free after radical prostatectomy.” Medical J of Australia, 1998; 168:483. 411. Ojdeby G, Claezon A, Brekkan E, Haggman M, Norlen BJ. “Urinary incontinence and sexual impotence after radical prostatectomy.” Scandinavian Journal of Urology and Nephrology, 1996; 30:473. 412. Gaylis FD, Friedel WE, Armas OA. “Radical retropubic prostatectomy outcomes at a community hospital.” Journal of Urology, 1998; 159:167. 413. Talcott JA, Manola J, Clark JA, Kaplan I, Beard CJ, Mitchell SP, Chen RC, O’Leary MP, Kantoff PW, D’Amico AV. “Time course and predictors of symptoms after primary prostate cancer therapy.” J Clin Oncol 2003; 21, 3979-86. 414. Sole-Balcells F, Villavicencio H, Ortiz A. “Postsurgical management of the patient undergoing radical prostatectomy.” British Journal of Urology, 1992; 70(1 Suppl):43. 415. Davidson PJ, van den Ouden D, Schroeder FH. “Radical prostatectomy: prospective assessment of mortality and morbidity.” European Urology, 1996; 29(2):168. 416. Ritchie AW, James K, deKernion JB. “Early post-operative morbidity of total prostatectomy.” British Journal of Urology, 1989; 64(3):511.

100

Chapter 10. PROSTATE CANCER 417. Catalona WJ, Carvalhal GF, Mager DE, Smith DS. “Potency, continence and complication rates in 1,870 consecutive radical retropubic prostatectomies.” Journal of Urology, 1999; 162(2):433. 418 Quinlan DM, Epstein JI, Carter BS, Walsh PC. “Sexual function following radical prostatectomy: influence of preservation of neurovascular bundles.” Journal of Urology, 1991; 145(3):998. 419. Leandri P, Rossignol G, Gautier JR, Ramon J. “Radical retropubic prostatectomy: morbidity and quality of life. Experience with 620 consecutive cases.” Journal of Urology, 1992; 147(3):883. 420. Noh C, Kshirsagar A, Mohler JL. “Outcomes after radical retropubic prostatectomy.” Urology, 2003; 61(2): 412-6. 421. Walsh PC, Marschke P, Ricker D, Burnett AL. “Patient-reported urinary continence and sexual function after anatomic radical prostatectomy.” Urology, 2000; 55(1):58. 422. Link RE, Su LM, Sullivan W, Bhayani SB, Pavlovich CP. “Health related quality of life before and after laparoscopic radical prostatectomy.” J Urol, 2005; 173(1): 175-9; discussion 179. 423. Epstein JI. “Evaluation of radical prostatectomy capsular margins of resection. The significance of margins designated as negative, closely approaching, and positive.” American Journal of Surgical Pathology, 1995; 14(7):626. 424. Epstein JI, Pizov G, Walsh PC. “Correlation of pathologic findings with progression after radical retropubic prostatectomy.” Cancer, 1993; 71(11):3582. 425. Rosen MA, Goldstone L, Lapin S, Wheeler T, Scardino PT. “Frequency and location of extracapsular extension of positive surgical margins in radical prostatectomy specimens.” Journal of Urology, 1992; 142(2):331. 426. Roumeguere T, Bollens R, Vanden-Bossche M, Rochet D, Bialek D, Hoffman P, Quackels T, Damoun A, Wespes E, Schulman CC, Zlotta AR. “Radical prostatectomy: a prospective comparison of oncological and functional results between open and laparoscopic approaches.” World J Urol, 2003; 20(6): 360-6. 427. Chuang MS, O’Connor RC, Laven BA, Orvieto MA, Brendler CB. “Early release of the neurovascular bundles and optical loupe magnification lead to improved and earlier return of potency following radical retropubic prostatectomy.” J Urol, 2005; 173(2): 537-9. 428. Porpiglia F, Ragni F, Terrone C, Renard J, Musso F, Grande S, Cracco C, Ghignone G, Scarpa RM. “Is laparoscopic unilateral sural nerve grafting during radical prostatectomy effective in retaining sexual potency?” BJU Int, 2005; 95(9): 1267-71. 429. Kim ED, Scardino PT, Hampel O, Mills NL, Wheeler TM, Nath RK. “Interposition of sural nerve restores function of cavernous nerves resected during radical prostatectomy.” J Urol, 1999; 61(1): 188-92. 430. Scardino PT, Kim ED. “Rationale for and results of nerve grafting during radical prostatectomy.” Urology, 2001; 57(6): 1016-9. 431. Kim ED, Nath R, Slawin KM, Kadmon D, Miles BJ, Scardino PT. “Bilateral nerve grafting during radical retropubic prostatectomy: extended follow-up.” Urology, 2001; 58(6): 983-7. 432. Joseph JV, Vincente I, Madeb R, Erturk E, Patel, HRH. “Robot-assisted vs laparoscopic radical proatatectomy: are there any differences?” BJU Int, 2005; 96:39-42 433. Webster TM, Herrell SD, Chang SS, Cookson MS, Baumgartner RG, Anderson LW, Smith JA Jr. “Robotic assisted laparoscopic radical prostatectomy versus retropubic radical prostatectomy: a prospective assessment of postoperative pain.” J Urol.2005; Sep, 174(3):912-4; discussion 914.

101

Chapter 10. PROSTATE CANCER 434. Ahlering TE, Eichel L, Edwards RA, Lee DI, Skarecky DW. “Robotic radical prostatectomy: a technique to reduce pT2 positive margins.” Urology, 2004;64(6): 1224-8. 435. Touijer K, Kuroiwa K, Saranchuk JW, Hassen WA, Trabulsi EJ, Reuter VE, Guillonneau B. “Quality improvement in laparoscopic radical prostatectomy for pT2 prostate cancer: impact of video documentation review on positive surgical margin.” J Urol, 2005; 173(3): 765-8. 436. Costello AJ, Haxhimolla H, Crowe H, Peters, JS. “Installation of telerobotic surgery and initial experience with telerobotic radical prostatectomy.” BJU Int, 2005; 96:34-8. 437. Pienta K, Redman B, Hussain M. “Phase II study of estramustine and oral etoposide in hormone-refractory adenocarcinoma of the prostate.” Journal of Clinical Oncology, 1994; 12(10):2005. 438. Chung LWK. “Implications of stromal-epithelial interaction in human prostate cancer growth, progression and differentiation.” Semin Cancer Biol, 1993; 4: 183-192. 439. Labrie F, Candas B, Cusan L, Gomez JL, Diamond P, Suburu R, Lemay M. “Diagnosis of advanced or noncurable prostate cancer can be practically eliminated by prostate-specific antigen.” Urology, 1996; 47(2):212. 440. Arai Y, Kanamaru H, Yoshimura K, Okubo K, Kamoto T, Yoshida O. “Incidence of lymph node metastasis and its impact on long-term prognosis in clinically localized prostate cancer.” International Journal of Urology, 1998; 5(5):459-65. 441. Ekman P. “Predicting pelvic lymph node involvement in patients with localized prostate cancer.” European Urology, 1997; 32(Suppl 3):60-4. 442. Danella JF, deKernion JB, Smith RB, Steckel J. “The contemporary incidence of lymph node metastases in prostate cancer: implications for laparoscopic lymph node dissection.” Journal of Urology, 1993; 149(6):1488-91. 443. Alagiri M, Colton MD, Seidmon EJ, Greenberg RE, Hanno PM. “The staging of pelvic lymphadenectomy: implications as an adjunctive procedure for clinically localized prostate cancer.” British Journal of Urology, 1997; 80(2):243-6. 444. Partin AW, Pearson JD, Landis PK, Carter HB, Pound CR, Clemens JQ, Epstein JI, Walsh PC. “Evaluation of serum prostate-specific antigen velocity after radical prostatectomy to distinguish local recurrence from distant metastasis.” Urology, 1994; 43(3):649. 445. Messing EM, Manola J, Sarodsy M, Wilding G, Crawford ED, Truno D. “Immediate hormonal therapy compared with observation after radical prostatectomy and pelvic lymphadenopathy in men with node positive prostate cancer.” New Engl J Med, 1999; 341(24):1781. 446. Steiner MS, Marshall FF. “Mini-laparotomy staging pelvic lymphadenectomy (minilap). Alternative to standard and laparoscopic pelvic lymphadenectomy.” Urology, 1993; 41(3):201. 447. Wagner JR, Russo P. “Urologic complications of major pelvic surgery.” Seminars in Surgical Oncology, 2000; 18(3):216. 448. Heinzer H, Graefen M, Noldus J, Hammerer P, Huland H. “Early complication of anatomical radical retropubic prostatectomy: lessons from a single-center experience.” Urologia Internationalis, 1997; 59(1):30. 449. Paul DP, Loening SA, Narayana AS, Culp DA. “Morbidity from pelvic lymphadenectomy in staging carcinoma of the prostate.” Journal of Urology, 1983; 129(3):1141. 450. McCullough DL, McLaughlin AP, Gittes RF. “Morbidity of pelvic lymphadenectomy and radical prostatectomy for prostatic cancer.” Journal of Urology, 1977; 117(2):206. 102

Chapter 10. PROSTATE CANCER 451. Sibert L, Descargues G, Scotte M, Planet M, Thoumas D, Annoot M, Boillot B, Grise P. “Laparoscopic drainage of a lymphocele after radical prostatectomy.” Annales d’Urologie, 1994; 28(4):202. 452. Pound CR, Partin AW, Eisenberger MA, Chan DW, Pearson JD, Walsh PC. “Natural history of progression after PSA elevation following radical prostatectomy.” JAMA, 1999; 281(17): 1591-7. 453. Krongrad A, Lai H, Lai S. “Survival after radical prostatectomy.” JAMA, 1997; 278(1):44-46. 454. Tefilli MV, Gheiler EL, Tiguert R, Banerjee M, Sakr W, Grignon D, Wood DP Jr, Pontes JE. “Role of radical prostatectomy in patients with prostate cancer of high Gleason score.” Prostate, 1999; 39(1):60. 455. Kupelian P, Katcher J, Levin H, Zippe C, Klein E. “Correlation of clinical and pathologic factors with rising prostate-specific antigen profiles after radical prostatectomy alone for clinically localized prostate cancer.” Urology, 1996; 48(2):249. 456. Gerber GS, Thisted RA, Scardino PT, Frohmuller HG, Schroeder FH, Paulson DF, Middleton AW Jr, Rukstalis DB, Smith JA Jr, Schellhammer PF, Ohori M, Chodak GW. “Results of radical prostatectomy in men with clinically localized prostate cancer.” JAMA 1996; 276(8):615. 457. Blute ML, Bostwick DG, Bergstralh EJ, Slezak JM, Martin SK, Amling CL, Zincke H. “Anatomic site-specific positive margins in organ-confined prostate cancer and its impact on outcome after radical prostatectomy.” Urology, 1997; 50(5):733. 458. Ohori M, Wheeler TM, Kattan MW, Goto Y, Scardino PT. “Prognostic significance of positive surgical margins in radical prostatectomy specimens.” J Urol, 1995; 154:1818-24. 459. Rigaud J, Tiguert R, Decobert M, Hovington H, Latulippe E, Laverdiere J, Larue H, Lacombe L, Fradet Y. “Expression of p21 cell cycle protein is an independent predictor of response to salvage radiotherapy after radical prostatectomy.” Prostate, 2004; 58:269-76. 460. Fizazi K, Martinez LA, Sikes CR, Johnston DA, Stephens LC, McDonnell TJ, Logothetis CJ, Trapman J, Pisters LL, Ordonez NG, Troncoso P, Navone NM. “The association of p21 (WAF-1/CIP1)) with progression to androgen-independent prostate cancer.” Clin Cancer Res, 2002; 8:775-81. 461. Bauer JJ, Sesterhenn IA, Mostofi KF, McLeod DG, Srivastava S, Moul JW. “p53 nuclear protein expression is an independent prognostic marker in clinically localized prostate cancer patients undergoing radical prostatectomy.” Clin Cancer Res,1995; 1:1295-300. 462. Halvorsen OJ, Haukaas SA, Akslen LA. “Combined loss of PTEN and p27 expression is associated with tumor cell proliferation by Ki-67 and increased risk of recurrent disease in localized prostate cancer.” Clin Cancer Res, 2003; 9:1474-9. 463. Fradet V, Lessard L, Bégin LR, Karakeiwicz P, Masson A-M M, Saad F. “Nuclear Factor-?B nuclear localization is predictive of biochemical recurrence in patients with positive margin prostate cancer.” Clin Cancer Res, 2004; 10: 8460-4. 464. Walsh PC, Partin AW, Epstein JI. “Cancer control and quality of life following anatomical radical retropubic prostatectomy: results at 10 years.” Journal of Urology, 1994; 152(5 Pt 2):1831. 465. Pound CR, Partin AW, Epstein JI. “Prostate-specific antigen after anatomic radical retropubic prostatectomy: patterns of recurrence and cancer control.” Urologic Clinics of North America, 1997; 242(2):395. 466. Partin AW, Pound CR, Clemens JQ, Epstein JI, Walsh PC. “Serum PSA after anatomic radical prostatectomy.” “The Johns Hopkins experience after 10 years.” Urol Clin North Am, 1993; 20(4): 713-25.

103

Chapter 10. PROSTATE CANCER 467. Paulson DF. “Impact of radical prostatectomy in the management of clinically localized disease.” J Urol 1994; 152:1826-1830. 468. Amling CL, Bergstralh EJ, Blute ML, Slezak JM, Zincke H. “Defining prostate specific antigen progression after radical prostatectomy: what is the most appropriate cut point?” J Urol, 2001; 165:1146-1151. 469. Freedland SJ, Sutter ME, Dorey F, Aronson WJ. “Defining the ideal cutpoint for determining PSA recurrence after radical prostatectomy. Prostate-specific antigen.” Urology, 2003; 61(2): 365-9. 470. Jhaveri FM, Zippe CD, Klein EA, Kupelian PA. “Biochemical failure does not predict overall survival after radical prostatectomy for localized prostate cancer: 10year results.” Urology, 1999; 54(5): 884-90. 471. Prostate Cancer End Points Workshop Presentations, 2004. 472. Albertsen PC, Hanley JA, Penson DF, Fine J. “Validation of increasing prostate specific antigen as a predictor of prostate cancer death after treatment of localized prostate cancer with surgery or radiation.” J Urol, 2004; 171:2221-5 473. D’Amico AV, Moul JW, Carroll PR, Sun L, Lubeck D, Chen MH. “Surrogate end point for prostate cancer-specific mortality after radical prostatectomy or radiation therapy.” J Natl Cancer Inst, 2003; 95(18): 1376-83. 474. Shelfo SW, Obek C, Soloway MS. “Update on bladder neck preservation during radical retropubic prostatectomy: impact on pathologic outcome, anastomotic sutures, and continence.” Urology, 1998; 51(1):73. 475. Lowe BA. “Preservation of the anterior urethral ligamentous attachments in maintaining post-prostatectomy urinary continence.” Journal of Urology 1997; 158(6):2137. 476. Goluboff ET, Saidi JA, Mazer S. “Urinary continence after radical prostatectomy: the Columbia experience.” Journal of Urology, 1998; 159(4):1276. 477. Kaye KW, Creed KE, Wilson GJ, D’Antuono M, Dawkins HJ. “Urinary continence after radical retropubic prostatectomy. Analysis and synthesis of contributing factors: a unified concept.” British Journal of Urology, 1997; 80(3):444. 478. Seaman EK & Benson MC. “Improved continence with tubularized bladder neck reconstruction following radical retropubic prostatectomy.” Urology 1996; 47(4):532. 479. Poore RE, McCullough DL, Jarow JP. “Puboprostatic ligament sparing improves urinary continence after radical retropubic prostatectomy.” Urology, 1998; 51(1):67. 480. Moul JW. “Pelvic muscle rehabilitation in males following prostatectomy.” Urologic Nursing, 1998; 18(4):296. 481. Montague DK, Angermeier KW. “Post-prostatectomy urinary incontinence: the case for artificial urinary sphincter implantation.” Urology, 2000; 55(1):2. 482. Bahnson RR, Catalona WJ. “Complications of radical retropubic prostatectomy.” In: Complications of Urologic Surgery: Prevention and Management (second edition); RB Smith and RM Ehrlich, editors. Philadelphia: WB Saunders Co, 1990; p.386. 483. Meulen T, Zambon JV, Janknegt RA. “Treatment of anastomotic strictures and urinary incontinence after radical prostatectomy with urolume wallstent and AMS 800 artificial sphincter.” Journal of Endourology, 1999; 13(7):517. 484. Mulcahy JJ. “Erectile function after radical prostatectomy.” Seminars in Urologic Oncology, 2000; 18(1):71. 485. Raina R, Agarwal A, Zippe CD. “Management of erectile dysfunction after radical prostatectomy.” Urology, 2005; 66:923-9. 486. Ohebshalom M, Parker M, Guhring P, Mulhall JP. “The efficacy of sildenafil citrate following radiation therpay for prostate cancer: temporal considerations.” J Urol, 2005; 174: 258-62. 104

Chapter 10. PROSTATE CANCER 487. Leibovitch I, Foster RS, Wass JL, Rowland RG, Bihrle R, Little JS Jr, Kopecky KK, Donohue JP.“Color Doppler flow imaging for deep vein thrombosis screening in patients undergoing pelvic lymphadenopathy and radical retropubic prostatectomy for prostatic cancer.” Journal of Urology 1995; 153(6):1866. 488. Choo, R., G. Hruby, J. Hong, E. Hong, G. DeBoer, C. Danjoux, G. Morton, L. Klotz, E. Bhak, and A. Flavin. Positive resection margin and/or pathologic T3 adenocarcinoma of prostate with undetectable postoperative prostate-specific antigen after radical prostatectomy: to irradiate or not? Int J Radiat Oncol Biol Phys, 2002 52:67480. 489. Schild, S. E., W. W. Wong, G. L. Grado, M. Y. Halyard, D. E. Novicki, S. K. Swanson, T. R. Larson, and R. G. Ferrigni. The result of radical retropubic prostatectomy and adjuvant therapy for pathologic stage C prostate cancer. Int J Radiat Oncol Biol Phys, 1996 34:535-41. 490. Syndikus, I., T. Pickles, E. Kostashuk, and L. D. Sullivan. Postoperative radiotherapy for stage pT3 carcinoma of the prostate: improved local control. J Urol, 1996 155:1983-6. 491. Vargas, C., L. L. Kestin, D. W. Weed, D. Krauss, F. A. Vicini, and A. A. Martinez. Improved biochemical outcome with adjuvant radiotherapy after radical prostatectomy for prostate cancer with poor pathologic features. Int J Radiat Oncol Biol Phys, 2005 61:714-24. 492. Zietman, A. L., J. J. Coen, W. U. Shipley, and A. F. Althausen. Adjuvant irradiation after radical prostatectomy for adenocarcinoma of prostate: analysis of freedom from PSA failure. Urology, 1993 42:292-8; discussion 298-9. 493. Leibovich, B. C., D. E. Engen, D. E. Patterson, T. M. Pisansky, E. E. Alexander, M. L. Blute, E. J. Bergstralh, and H. Zincke. Benefit of adjuvant radiation therapy for localized prostate cancer with a positive surgical margin. J Urol, 2000 163:1178-82. 494. Valicenti, R. K., L. G. Gomella, M. Ismail, S. E. Strup, S. G. Mulholland, A. P. Dicker, R. O. Petersen, and C. J. Newschaffer. The efficacy of early adjuvant radiation therapy for pT3N0 prostate cancer: a matched-pair analysis. Int J Radiat Oncol Biol Phys, 1999 45:53-8. 495. Bolla, M., H. van Poppel, L. Collette, P. van Cangh, K. Vekemans, L. Da Pozzo, T. M. de Reijke, A. Verbaeys, J. F. Bosset, R. van Velthoven, J. M. Marechal, P. Scalliet, K. Haustermans, and M. Pierart. Postoperative radiotherapy after radical prostatectomy: a randomised controlled trial (EORTC trial 22911). Lancet, 2005 366:572-8. 496. Van Cangh, P. J., F. Richard, F. Lorge, Y. Castille, A. Moxhon, R. Opsomer, L. De Visscher, F. X. Wese, and P. Scaillet. Adjuvant radiation therapy does not cause urinary incontinence after radical prostatectomy: results of a prospective randomized study. J Urol, 1998 159:164-6. 497. ASCO 2005 abstract 4513. 498. ASTRO 2005 abstract 0001. 499. Do, L. V., T. M. Do, R. Smith, and R. G. Parker. Postoperative radiotherapy for carcinoma of the prostate: impact on both local control and distant disease-free survival. Am J Clin Oncol, 2002 25:1-8. 500. Garg, M. K., S. Tekyi-Mensah, S. Bolton, J. Velasco, E. Pontes, D. P. Wood, Jr., A. T. Porter, and J. D. Forman. Impact of postprostatectomy prostate-specific antigen nadir on outcomes following salvage radiotherapy. Urology, 1998 51:998-1002. 501. Pisansky, T. M., T. F. Kozelsky, R. P. Myers, D. W. Hillman, M. L. Blute, S. J. Buskirk, J. C. Cheville, R. G. Ferrigni, and S. E. Schild. Radiotherapy for isolated serum prostate specific antigen elevation after prostatectomy for prostate cancer. J Urol, 2000 163:845-50. 502. Duchesne GM, Dowling C, Frydenberg M, Joseph D, Gogna NK, Neerhut G, Roberts S, Spry N, Turner S, WooH. Outcome, morbidity, and prognostic factors 105

Chapter 10. PROSTATE CANCER in post-prostatectomy radiotherapy: an Australian multicenter study. Urology, 2003 61(1): 179-83. 503. Leventis, A. K., S. F. Shariat, M. W. Kattan, E. B. Butler, T. M. Wheeler, and K. M. Slawin. Prediction of response to salvage radiation therapy in patients with prostate cancer recurrence after radical prostatectomy. J Clin Oncol, 2001 19:1030-9. 504. Morris, M. M., K. C. Dallow, A. L. Zietman, J. Park, A. Althausen, N. M. Heney, and W. U. Shipley. Adjuvant and salvage irradiation following radical prostatectomy for prostate cancer. Int J Radiat Oncol Biol Phys, 1997 38:731-6. 505. Anscher, M. S., R. Clough, and R. Dodge. Radiotherapy for a rising prostatespecific antigen after radical prostatectomy: the first 10 years. Int J Radiat Oncol Biol Phys, 2000 48:369-75. 506. Bahnson RR, Hanks GE, Huben RP, Kantoff P, Kozlowski JM, Kuettel M, Lange P H, Logothetis C, Pow-Sang JM, Roach M, 3rd Sandler H, Scardino PT, Taylor RJ, Urban DA, Walsh PC, Wilson TG. “NCCN Practice Guidelines for Prostate Cancer.” Oncology 2000; (Williston Park) 14, 111-9. 507. Kattan MW, Zelefsky M.J, Kupelian PA, Scardino PT, Fuks Z, Leibel SA. “Pretreatment nomogram for predicting the outcome of three-dimensional conformal radiotherapy in prostate cancer.” J Clin Oncol 2000; 18, 3352-9. 508. Kupelian PA, Kuban D, Thames H, Levy L, Horwitz E, Martinez A, Michalski J, Pisansky T, Sandler H, Shipley W, Zelefsky M, Zietman A. “Improved biochemical relapse-free survival with increased external radiation doses in patients with localized prostate cancer: the combined experience of nine institutions in patients treated in 1994 and 1995.” Int J Radiat Oncol Biol Phys 2003; 57, S271-2. 509. Zelefsky MJ, Fuks Z, Hunt M, Lee HJ, Lombardi D, Ling CC, Reuter VE, Venkatraman ES, Leibel SA. “High dose radiation delivered by intensity modulated conformal radiotherapy improves the outcome of localized prostate cancer.” J Urol 2001; 166, 876-81. 510. Pilepich MV, Krall JM, al-Sarraf M, John MJ, Doggett RL, Sause WT, Lawton CA, Abrams RA, Rotman M, Rubin P. “Androgen deprivation with radiation therapy compared with radiation therapy alone for locally advanced prostatic carcinoma: a randomized comparative trial of the Radiation Therapy Oncology Group.” Urology 45, 1995; 616-23. 511. Hanks GE, Hanlon AL, Epstein B, Horwitz EM. “Dose response in prostate cancer with 8-12 years’ follow-up.” Int J Radiat Oncol Biol Phys 2002; 54, 427-35. 512. Lawton CA, Winter K, Murray K, Machtay M, Mesic JB, Hanks GE, Coughlin CT, Pilepich MV. “Updated results of the phase III Radiation Therapy Oncology Group (RTOG) trial 85-31 evaluating the potential benefit of androgen suppression following standard radiation therapy for unfavorable prognosis carcinoma of the prostate.” Int J Radiat Oncol Biol Phys 2001; 49, 937-46. 513. Pollack A, Zagars GK, Starkschall G, Antolak JA, Lee JJ, Huang E, von Eschenbach, AC, Kuban DA, Rosen I. “Prostate cancer radiation dose response: results of the M. D. Anderson phase III randomized trial.” Int J Radiat Oncol Biol Phys 2002; 53, 1097-105. 514. Kupelian P, Kuban D, Thames H, Levy L, Horwitz E, Martinez A, Michalski J, Pisansky T, Sandler H, Shipley W, Zelefsky M, Zietman A. “Improved biochemical relapse-free survival with increased external radiation doses in patients with localized prostate cancer: The combined experience of nine institutions in patients treated in 1994 and 1995.” Int J Radiat Oncol Biol Phys 2005; 61, 415-9. 515. D’Amico AV, Schultz D, Loffredo M, Dugal, R, Hurwitz M, Kaplan I, Beard CJ, Renshaw AA, Kantoff PW. “Biochemical outcome following external beam radiation therapy with or without androgen suppression therapy for clinically localized prostate cancer.” Jama 2000; 284, 1280-3.

106

Chapter 10. PROSTATE CANCER 516. D’Amico AV, Manola J, Loffredo M, Renshaw AA, DellaCroce A, Kantoff PW. “6-month androgen suppression plus radiation therapy vs radiation therapy alone for patients with clinically localized prostate cancer: a randomized controlled trial.” Jama 2004; 292, 821-7. 517. Denham JW, Steigler A, Lamb DS, Joseph D, Mameghan H, Turner S, Matthews J, Franklin I, Atkinson C, North J, Poulsen M,Christie D, Spry NA, Tai KH, Wynne C, Duchesne G, Kovacev O, D’Este C. “Short-term androgen deprivation and radiotherapy for locally advanced prostate cancer: results from the Trans-Tasman Radiation Oncology Group 96.01 randomised controlled trial.” Lancet Oncol 2005; 6, 841-50. 518. Roach M, 3rd DeSilvio M, Lawton C, Uhl V, Machtay M, Seider MJ, Rotman M, Jones C, Asbell SO, Valicenti RK, Han S, Thomas CR Jr, Shipley WS. “Phase III trial comparing whole-pelvic versus prostate-only radiotherapy and neoadjuvant versus adjuvant combined androgen suppression: Radiation Therapy Oncology Group 9413.” J Clin Oncol 2003; 21, 1904-11. 519. Morris DE, Emami B, Mauch PM, Konski AA, Tao ML, Ng AK, Klein EA, Mohideen N, Hurwitz MD, Fraas BA, Roach M, 3rd Gore EM, Tepper JE. “Evidence-based review of three-dimensional conformal radiotherapy for localized prostate cancer: an ASTRO outcomes initiative.” Int J Radiat Oncol Biol Phys 2005; 62, 3-19. 520. Robinson JW, Moritz S, Fung T. “Meta-analysis of rates of erectile function after treatment of localized prostate carcinoma.” International Journal of Radiation Oncology*Biology*Physics 2002; 54, 1063-1068. 521. Zurlo A, Collette L, van Tienhoven G, Blank L, Warde P, Dubois J, Jeanneret W, Storme G, Bernier J, Kuten A, Pierart M, Bolla M. “Acute toxicity of conventional radiation therapy for high-risk prostate cancer in EORTC trial 22863.” Eur Urol 2002; 42, 125-32. 522. Cooperberg MR, Lubeck DP, Meng MV, Mehta SS, Carroll PR. “The changing face of low-risk prostate cancer: trends in clinical presentation and primary management.” J Clin Oncol 2004; 22, 2141-9. 523. Beyer DC, McKeough T, Thomas T. “Impact of short course hormonal therapy on overall and cancer specific survival after permanent prostate brachytherapy.” Int J Radiat Oncol Biol Phys 2005; 61, 1299-305. 524. Blasko JC, Grimm PD, Sylsvester JE, Cavanagh W. “The role of external beam radiotherapy with I-125/Pd-103 brachytherapy for prostate carcinoma.” Radiother Oncol 2000; 57, 273-8. 525. Potters L, Morgenstern C, Calugaru E, Fearn P, Jassal A, Presser J, Mullen E. “12year outcomes following permanent prostate brachytherapy in patients with clinically localized prostate cancer.” J Urol 2005; 173, 1562-6. 526. Sylvester JE, Blasko JC, Grimm PD, Meier R, Malmgren JA. “Ten-year biochemical relapse-free survival after external beam radiation and brachytherapy for localized prostate cancer: the Seattle experience.” Int J Radiat Oncol Biol Phys 2003; 57, 944-52. 527. Beyer DC, Brachman DG. “Failure free survival following brachytherapy alone for prostate cancer: comparison with external beam radiotherapy.” Radiother Oncol 2000; 57, 263-7. 528. Critz FA, Levinson AK, Williams WH, Holladay CT, Griffin VD, Holladay DA. “Simultaneous radiotherapy for prostate cancer: 125I prostate implant followed by external-beam radiation.” Cancer J Sci Am 1998; 4, 359-63. 529. Merrick GS, Butler WM, Wallner KE, Galbreath RW, Murray B, Zeroski D, Lief JH. “Dysuria after permanent prostate brachytherapy.” Int J Radiat Oncol Biol Phys 2003; 55, 979-85. 530. Lee WR, Hall MC, McQuellon RP, Case LD, McCullough DL. “A prospective quality-of-life study in men with clinically localized prostate carcinoma treated with 107

Chapter 10. PROSTATE CANCER radical prostatectomy, external beam radiotherapy, or interstitial brachytherapy.” Int J Radiat Oncol Biol Phys 2001;51, 614-23. 531. Williams SG, Millar JL, Duchesne GM, Dally MJ, Royce PL, Snow RM. “Factors predicting for urinary morbidity following 125iodine transperineal prostate brachytherapy.” Radiother Oncol 2004; 73, 33-8. 532. Mueller A, Wallner K, Merrick G, Ford E, Sutlief S, Cavanagh W, Butler W. “Perirectal seeds as a risk factor for prostate brachytherapy-related rectal bleeding.” Int J Radiat Oncol Biol Phys 2004; 59, 1047-52. 533. Macdonald AG, Keyes M, Kruk A, Duncan G, Moravan V, Morris W J. “Predictive factors for erectile dysfunction in men with prostate cancer after brachytherapy: Is dose to the penile bulb important?” International Journal of Radiation Oncology*Biology*Physics 2005; 63, 155-163. 534. Merrick GS, Butler WM, Galbreath RW, Stipetich RL, Abel LJ, Lief JH. “Erectile function after permanent prostate brachytherapy.” Int J Radiat Oncol Biol Phys 2002; 52, 893-902. 535. Astrom L, Pedersen D, Mercke C, Holmang S, Johansson KA. “Long-term outcome of high dose rate brachytherapy in radiotherapy of localised prostate cancer.” Radiother Oncol 2005; 74, 157-61. 536. Deger S, Boehmer D, Turk I, Roigas J,Wernecke KD,Wiegel T, Hinkelbein W, Dinges S, Budach V, Loening SA. “High dose rate brachytherapy of localized prostate cancer.” Eur Urol 2002; 41, 420-6. 537. Demanes D J, Rodriguez RR, Schour L, Brandt D, Altieri G. “High-dose-rate intensity-modulated brachytherapy with external beam radiotherapy for prostate cancer: California endocurietherapy’s 10-year results.” International Journal of Radiation Oncology*Biology*Physics 2005; 61, 1306-1316. 538. Duchesne GM, Das R, Toye W, See A. “Dose distribution and morbidity after high dose rate brachytherapy for prostate cancer: influence of V150 and V200 parameters.” Australas Radiol 2002; 46, 384-9. 539. Galalae RM, Martinez A, Mate T, Mitchell C, Edmundson G, Nuernberg N, Eulau S, Gustafson G, Gribble M, Kovacs G. “Long-term outcome by risk factors using conformal high-dose-rate brachytherapy (HDR-BT) boost with or without neoadjuvant androgen suppression for localized prostate cancer.” Int J Radiat Oncol Biol Phys 2004; 58, 1048-55. 540. Martinez AA, Gustafson G, Gonzalez J, Armour E, Mitchell C, Edmundson G, Spencer W, Stromberg J, Huang R, Vicini F. “Dose escalation using conformal highdose-rate brachytherapy improves outcome in unfavorable prostate cancer.” Int J Radiat Oncol Biol Phys 2002; 53, 316-27. 541. Mate TP, Gottesman JE, Hatton J, Gribble M, Van Hollebeke L. “High dose-rate afterloading 192Iridium prostate brachytherapy: feasibility report.” Int J Radiat Oncol Biol Phys 1998; 41, 525-33. 542. Duchesne GM & Peters LJ. “What is the alpha/beta ratio for prostate cancer? Rationale for hypofractionated high-dose-rate brachytherapy.” Int J Radiat Oncol Biol Phys 1999; 44, 747-8. 543. Consensus statement: guidelines for PSA following radiation therapy. “American Society for Therapeutic Radiology and Oncology Consensus Panel.” Int J Radiat Oncol Biol Phys. 1997; 37, 1035-41. 544. Vicini FA, Ziaja EL, Kestin LL, Brabbins DS, Stromberg JS, Gonzalez JA, Martinez AA. “Treatment outcome with adjuvant and salvage irradiation after radical prostatectomy for prostate cancer.” Urology, 1999; 54(1):111. 545. Williams SG. “Ambiguities within the ASTRO consensus definition of biochemical failure: never assume all is equal.” Int J Radiat Oncol Biol Phys. 2004; 58(4):108392. 108

Chapter 10. PROSTATE CANCER 546. Stock RG, Stone NN, Cesaretti JA. “Prostate-specific antigen bounce after prostate seed implantation for localized prostate cancer: descriptions and implications.” Int J Radiat Oncol Biol Phys 2003; 56, 448-53. 547. Zietman AL, Christodouleas JP, Shipley WU. “PSA bounces after neoadjuvant androgen deprivation and external beam radiation: impact on definitions of failure.” Int J Radiat Oncol Biol Phys. 2005; 62(3):714-8. 548. Pickles T, Kim-Sing C, Morris WJ, Tyldesley S, Paltiel C. “Evaluation of the Houston biochemical relapse definition in men treated with prolonged neoadjuvant and adjuvant androgen ablation and assessment of follow-up lead-time bias.” Int J Radiat Oncol Biol Phys 2003; 57, 11-8. 549. Gretzer MB, Trock BJ, Han M, Walsh PC. “A critical analysis of the interpretation of biochemical failure in surgically treated patients using the American Society for Therapeutic Radiation and Oncology criteria.” J Urol 2002; 168, 1419-22. 550. Williams SG, Duchesne GM, Millar JL, Pratt GR. “Both pretreatment prostatespecific antigen level and posttreatment biochemical failure are independent predictors of overall survival after radiotherapy for prostate cancer.” Int J Radiat Oncol Biol Phys 2004; 60, 1082-7. 551. McMullen KP, Lee WR. “A structured literature review to determine the use of the American Society for Therapeutic Radiology and Oncology consensus definition of biochemical failure.” Urology 2003; 61, 391-6. 552. Thames H, Kuban D, Levy L, Horwitz EM, Kupelian P, Martinez A, Michalski J, Pisansky T, Sandler H, Shipley W, Zelefsky M, Zietman A. “Comparison of alternative biochemical failure definitions based on clinical outcome in 4839 prostate cancer patients treated by external beam radiotherapy between 1986 and 1995.” Int J Radiat Oncol Biol Phys. 2003; 57(4):929-43. 553. Kestin LL, Vicini FA, Martinez AA. “Practical application of biochemical failure definitions: what to do and when to do it.” Int J Radiat Oncol Biol Phys. 2002; 53(2):304-15. 554. Kestin L. Personal communication (2005). 555. Pickles T, ASTRO abstract 05-0058 (2005). 556. Lee AK, D’Amico AV. “Utility of prostate-specific antigen kinetics in addition to clinical factors in the selection of patients for salvage local therapy.” J Clin Oncol,2005; 10;23(32):8192-7. 557. Touma NJ, Izawa JI, Chin JL. “Current status of local salvage therapies following radiation failure for prostate cancer.” J Urol, 2005; 173(2):373-9. 558. Korfage IJ, Essink-Bot ML, Borsboom GJ, Madalinska JB, Kirkels WJ, Habbema JD, Schroder FH, de Koning HJ. “Five-year follow-up of health-related quality of life after primary treatment of localized prostate cancer.” Int J Cancer 2005;116, 291-6. 559. Miller DC, Sanda MG, Dunn RL, Montie JE, Pimentel H, Sandler HM, McLaughlin WP, Wei JT. “Long-term outcomes among localized prostate cancer survivors: health-related quality-of-life changes after radical prostatectomy, external radiation, and brachytherapy.” J Clin Oncol 2005; 23, 2772-80. 560. Merrick GS, Wallner KE, Butler WM. “Minimizing brachytherapy-related morbidity.” Urology 2003; 62, 786-92. prostate

561. Frydenberg M, Giles GG, Mameghan H, Thursfield VJ, Millar J, Wheelahan JB, Bolton DM, Syme, RR. “Prostate cancer in Victoria in 1993: patterns of reported management.” Med J Aust, 2000; 172(6): 270-4. 562. Holtgrewe HL. “The economics of urologic practice in the twenty-first century.” Urol Clin North Am, 1998; 25(1): 1-13. 563. Halmos G, Arencibia JM, Schally AV, Davis R, Bostwick DG. “High incidence of receptors for luteinizing hormone-releasing hormone (LHRH) and LHRH receptor gene expression in human prostate cancers.” J Urol, 2000; 163(2): 623-9. 109

Chapter 10. PROSTATE CANCER 564. Schally AV, Nagy A. “Chemotherapy targeted to cancers through tumoral hormone receptors.” Trends Endocrinol Metab, 2004; 15(7): 300-10. 565. Townsend MF, Sanders WH, Northway RO, Graham SD Jr. “Bone fractures associated with leutinizing hormone-releasing hormone agonists used in the treatment of prostate carcinoma.” Cancer, 1997; 79:545-50. 566. Daniell HW. “Osteoporosis after orchiectomy for prostate cancer.” J Urol, 1997; 157: 439-44. 567. Shahinian VB, Yong-Fang K, Freeman JL, Goodwin JS. “Risk of fracture after androgen deprivation for prostate cancer.” N Eng J Med, 2005; 352:154-64. 568. Loblaw DA, Mendelson DS, Talcott JA, Virgo KS, Somerfield MR, Ben-Josef E, Middleton R, Porterfield H, Sharp SA, Smith TJ, Taplin ME, Vogelzang NJ, Wade JL, Bennett CL Jr, Scher HI. “American Society of Clinical Oncology recommendations for the initial hormonal management of androgen-sensitive metastatic, recurrent, or progressive prostate cancer.” American-Society-of-Clinical-Oncology, 2004; J Clin Oncol, 22(14): 2927-41. 569. Kaisary AV, Tyrrell CJ, Beacock C, Lunglmayr G, Debruyne F. “A randomised comparison of monotherapy with Casodex 50 mg daily and castration in the treatment of metastatic prostate carcinoma.” Casodex Study Group. Eur Urol, 1995; 28(3): 215-22. 570. Green HJ, Pakenham KI, Headley BC, Yaxley J, Nicol DL, Mactaggart PN, Swanson CE, Watson RB, Gardiner RA. “Quality of life compared during pharmacological treatments and clinical monitoring for non-localized prostate cancer: a randomized controlled trial.” BJU Int, 2004; 93(7): 975-9. 571. Cassileth BR, Soloway MS, Vogelzang NJ, Chou JM, Schellhammer PD, Seidmon EJ, Kennealey GT. “Quality of life and psychosocial status in stage D prostate cancer. Zoladex Prostate Cancer Study Group.” Qual Life Res. 1992; Oct, 1(5): 323-9. 572. Da-Silva FC, Fossa SD, Aaronson NK, Serbouti S, Denis L, Casselman J, Whelan P, Hetherington J, Fava C, Richards B, Robinson MR. “The quality of life of patients with newly diagnosed M1 prostate cancer: experience with EORTC clinical trial 30853.” Eur J Cancer, 1996; 32A(1): 72-7. 573. Bhayani SB, Andriole GL. “Hormonal manipulation for rising PSA after radical prostatectomy.” Semin Urol Oncol, 1999; 17(3): 148-53. 574. Herr HW, O’Sullivan M. “Quality of life of asymptomatic men with nonmetastatic prostate cancer on androgen deprivation therapy.” J Urol, 2000; 163(6): 1743-6. 575. Klotz L. Hormone therapy for patients with prostate carcinoma. Cancer. 2000; 88(12 Suppl): 3009-14. 576. Salminen E, Portin R, Korpela J, Backman H, Parvinen LM, Helenius H, Nurmi M. “Androgen deprivation and cognition in prostate cancer.” Br J Cancer, 2003; 89(6): 971-6. 577. Cherrier MM, Rose AL, Higano C. “The effects of combined androgen blockade on cognitive function during the first cycle of intermittent androgen suppression in patients with prostate cancer.” J Urol, 2003; 170(5): 1808-11. 578. Usui S, Suzuki K, Yamanaka H, Nakano T, Nakajima K, Hara Y, Okazaki M. “Estrogen treatment of prostate cancer increases triglycerides in lipoproteins as demonstrated by HPLC and immunoseparation techniques.” Clin Chim Acta, 2002; 317(1-2): 133-43. 579. Henriksson P, Edhag O. (1986) Orchidectomy versus oestrogen for prostatic cancer: cardiovascular effects. Br Med J (Clin-Res-Ed), 293(6544): 413-5. 580. Klotz L, McNeill I, Fleshner N. “A phase 1-2 trial of diethylstilbestrol plus low dose warfarin in advanced prostate carcinoma.” J Urol, 1999; 161(1): 169-72. 110

Chapter 10. PROSTATE CANCER 581. Mikkola A, Ruutu M, Aro J, Rannikko S, Salo J. “The role of parenteral polyestradiol phosphate in the treatment of advanced prostatic cancer on the threshold of the new millennium.” Ann Chir Gynaecol, 1999; 88(1): 18-21. 582. Iversen P. “Orchidectomy and oestrogen therapy revisited.” Eur Urol. 1998; 34 Suppl 3: 7-11. 583. Lundgren R, Nordle O, Josefsson K. “Immediate estrogen or estramustine phosphate therapy versus deferred endocrine treatment in nonmetastatic prostate cancer: a randomized multicenter study with 15 years of followup.” The South Sweden Prostate Cancer Study Group, J Urol, 1995; 153(5): 1580-6. 584. Cox RL, Crawford ED. “Estrogens in the treatment of prostate cancer.” J Urol, 1995; 154(6): 1991-8. 585. Bishop MC. “Experience with low-dose oestrogen in the treatment of advanced prostate cancer: a personal view.” Br J Urol, 1996; 78(6): 921-7; discussion 927-8. 586. Kato M, Kase H, Okada K, Tohma T, Ishiwata D. “Reevaluation and potential therapeutic effects of estrogen therapy to prostate cancer patients.” Nippon Rinsho. 1998;56(8): 2140-4. 587. Henriksson P, Carlstrom K, Pousette A, Gunnarsson PO, Johansson CJ, Eriksson B, Altersgard-Brorsson AK, Nordle O, Stege R. “Time for revival of estrogens in the treatment of advanced prostatic carcinoma? Pharmacokinetics, and endocrine and clinical effects, of a parenteral estrogen regimen.” Prostate, 1999; 40(2): 76-82. 588. Bland LB, Garzotto M, DeLoughery TG, Ryan CW, Schuff KG, Wersinger EM, Lemmon D, Beer TM. “Phase II study of transdermal estradiol in androgen-independent prostate carcinoma.” Cancer, 2005; 15; 103(4): 717-23. 589. Ockrim JL, Lalani EN, Laniado ME, Carter SS, Abel PD. “Transdermal estradiol therapy for advanced prostate cancer--forward to the past?” J Urol, 2003; 169(5): 17357. 590. Nair B, Wilt T, MacDonald R, Rutks I. “Early versus deferred androgen suppression in the treatment of advanced prostatic cancer.” Cochrane Database Syst Rev, 2002; (1):CD003506. 591. Abate-Shen C, Shen MM. “Mouse models of prostate carcinogenesis.” Trends Genet, 2002; 18(5): S1-5. 592. Lopatkin NA, Gustafsson J-A. “Estrogen receptor ß in prostate cancer.” N Engl J Med, 2004; 351:2773-4. 593. Mariani AJ, Glover M, Arita S. “Medical versus surgical androgen suppression therapy for prostate cancer: a 10-year longitudinal cost study.” J Urol, 2001; 165(1): 104-7. 594. Eisenberger MA, Walsh PC. “Early androgen deprivation for prostate cancer?” N Engl J Med, 1999; 341(24): 1837-8. 595. Walsh PC, DeWeese TL, Eisenberger MA. “A structured debate: immediate versus deferred androgen suppression in prostate cancer-evidence for deferred treatment.” J Urol, 2001; 166(2): 508-15; discussion 515-6. 596. Samson DJ, Seidenfeld J, Schmitt B, Hasselblad V, Albertsen PC, Bennett CL, Wilt, TJ, Aronson N. “Systematic review and meta-analysis of monotherapy compared with combined androgen blockade for patients with advanced prostate carcinoma.” Cancer. 2002; 95(2): 361-76. 597. Hall MC, Fritzsch RJ, Sagalowsky AI, Ahrens A, Petty B, Roehrborn CG. “Prospective determination of the hormonal response after cessation of luteinizing hormone-releasing hormone agonist treatment in patients with prostate cancer.” Urology, 1999; 53(5): 898-902; discussion 902-3. 598. Newling DW, Denis L, Vermeylen K. “Orchiectomy versus goserelin and flutamide in the treatment of newly diagnosed metastatic prostate cancer. Analysis of 111

Chapter 10. PROSTATE CANCER the criteria of evaluation used in the European Organization for Research and Treatment of Cancer - Genitourinary Group Study 30853.” Cancer, 1993; 72(12 Suppl): 3793-8. 599. Buttyan R, Ghafar MA, Shabsigh A. “The effects of androgen deprivation on the prostate gland: cell death mediated by vascular regression.” Curr Opin Urol 2000; 10(5): 415-20. 600. Sciarra A, Casale P, Colella D, Di-Chiro C, Di-Silverio F. “Hormone-refractory prostate cancer? Anti-androgen withdrawal and intermittent hormone therapy.” Scand J Urol Nephrol, 1999; 33(4): 211-6. 601. Visakorpi T, Hyytinen E, Koivisto P, Tanner M, Keinanen R, Palmberg C, Palotie A, Tammela T, Isola J, Kallioniemi OP. “In vivo amplification of the androgen receptor gene and progression of human prostate cancer.” Nat Genet, 1995; 9(4): 401-6. 602. Linja MJ, Savinainen KJ, Saramaki OR, Tammela TL, Vessella RL, Visakorpi T. “Amplification and over expression of androgen receptor gene in hormone-refractory prostate cancer.” Cancer Res, 2001; 61(9): 3550-5. 603. Gregory CW, Johnson RT Jr, Mohler JL, French FS, Wilson EM. “Androgen receptor stabilization in recurrent prostate cancer is associated with hypersensitivity to low androgen.” Cancer Res, 2001; 61(7): 2892-8. 604. Gottlieb B, Lehvaslaiho H, Beitel LK, Lumbroso R, Pinsky L, Trifiro M. “The Androgen Receptor Gene Mutations Database.” Nucleic Acids Res, 1998; 26(1): 2348. 605. Taplin ME, Bubley GJ, Ko YJ, Small EJ, Upton M, Rajeshkumar B, Balk SP. “Selection for androgen receptor mutations in prostate cancers treated with androgen antagonist.” Cancer Res, 1999; 59(11): 2511-5. 606. Lai JS, Brown LG, True LD, Hawley SJ, Etzioni RB, Higano CS, Ho S-M, Vessella RL, Corey E. “Metastases of prostate cancer express estrogen receptor-beta.” Urology, 2004; 64:814-20. 607. Clarke NW. “The management of hormone-relapsed prostate cancer.” BJU Int, 2003; 92(8): 860-8. 608. Yagoda A & Petrylak D. “Cytotoxic chemotherapy hormone-resistant prostate cancer.” Cancer 1993; 71, 1098-109. for advanced

609. Tannock IF, Osoba D, Stockler MR, Ernst DS, Neville AJ, Moore MJ, Armitage G R, Wilson JJ, Venner PM, Coppin CM, Murphy KC. “Chemotherapy with mitoxantrone plus prednisone or prednisone alone for symptomatic hormone-resistant prostate cancer: a Canadian randomized trial with palliative end points.” J Clin Oncol 1996; 14, 1756-64. 610. Kantoff PW, Halabi S, Conaway M, Picus J, Kirshner J, Hars V, Trump D, Winer EP, Vogelzang NJ. “Hydrocortisone with or without mitoxantrone in men with hormone-refractory prostate cancer: results of the cancer and leukemia group B 9182 study.” J Clin Oncol 1999; 17, 2506-13. 611. Pienta KJ & Smith DC. “Advances in prostate cancer chemotherapy: a new era begins.” CA Cancer J Clin 2005; 55, 300-18. 612. Tannock IF, de Wit R, Berry WR, Horti J, Pluzanska A, Chi KN, Oudard S, Theodore C, James ND, Turesson I, Rosenthal MA, Eisenberger MA. “Docetaxel plus prednisone or mitoxantrone plus prednisone for advanced prostate cancer.” N Engl J Med 2004; 351, 1502-12. 613. Petrylak DP, Tangen CM, Hussain MH, Lara PN Jr, Jones JA, Taplin ME, Burch PA, Berry D, Moinpour C, Kohli M, Benson MC, Small EJ, Raghavan D, Crawford ED. “Docetaxel and estramustine compared with mitoxantrone and prednisone for advanced refractory prostate cancer.” N Engl J Med 2004; 351, 1513-20. 614. Cranney A, Wells G, Wilan A. “Meta-analyses of therapies for postmenopausal osteoporosis.” Endocr Rev, 2002; 23(4):508-16. 112

Chapter 10. PROSTATE CANCER 615. Smith MR, Eastham J, Gleason DM, Shasha D, Tchekmedyian S, Zinner N. “Randomized controlled trial of zoledronic acid to prevent bone loss in men receiving androgen deprivation therapy for nonmetastatic prostate cancer.” J Urol, 2003; 169: 2008-12. 616. Lipton A, Small E, Saad F, Gleason D, Gordon D, Smith M, Rosen L, Kowalski MO, Reitsma D, Seaman J. “The new bisphosphonate, Zometa (zoledronic acid) decreases skeletal complications in both osteolytic and osteoblastic lesions: a comparison to pamidronate.” Cancer Invest, 2002; 20(suppl 2):45-54. 617. Kasamon KM, Dawson NA. “Update on hormone-refractory prostate cancer.” Curr Opin Urol, 2004: 14: 185-98. 618. Corey E, Brown LG, Quinn JE, Poot M, Roudier MP, Higano CS, Vessella RL. “Zoedronic acid exhibits inhibitory effects on osteoblastic and osteolytic metastases of prostate cancer.” Clin Cancer Res, 2003: 9:295-308. 619. Sato K, Kimura S, Segawa H, Yokota A, Matsumoto S, Kuroda J, Nogawa M, Yuasa T, Kiyono Y, Wada H, Maekawa T. “Cytotoxic effects of gammadelta T cells expanded ex vivo by a third generation bisphosphonate for cancer immunotherapy.” Int J Cancer, 2005; 116(1): 94-9. 620. Saad F, Gleason DM, Murray R, Tchekmedyian S, Venne P, Lacombe L, Chin JL, Vinholes JJ. “A randomized, placebo-controlled trial of zoledronic acid in patients with normone-refractory metastatic prostate carcinoma”. JNCI, 2002; 94(19):1458-68. 621. Saad F, Gleason DM, Murray R, Tchekmedyian S, Venner P, Lacombe L, Chin JL, Vinholes JJ, Goas JA, Zheng M. “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. 622. Quilty PM, Kirk D, Bolger JJ, Dearnaley DP, Lewington VJ, Mason MD, Reed NS, Russell JM, Yardley J. “A comparison of the palliative effects of strontium-89 and external beam radiotherapy in metastatic prostate cancer.” Radiother Oncol 1994; 31(1): 33-40. 623. Porter AT, McEwan AJ, Powe JE, Reid R, McGowan DG, Lukka H, Sathyanarayana JR, Yakemchuk VN, Thomas GM, Erlich LE. “Results of a randomized phase-III trial to evaluate the efficacy of strontium-89 adjuvant to local field external beam irradiation in the management of endocrine resistant metastatic prostate cancer.” Int J Radiat Oncol Biol Phys 1993; 25(5): 805-13. 624. Oosterhof GO, Roberts JT, de Reijke TM, Engelholm SA, Horenblas S, von der Maase H, Neymark N, Debois M, Collette L. "Strontium(89) chloride versus palliative local field radiotherapy in patients with hormonal escaped prostate cancer: a phase III study of the European Organisation for Research and Treatment of Cancer, Genitourinary Group." Eur Urol 2003; 44(5): 519-26. 625. Collins C, Eary JF, Donaldson G, Vernon C, Bush NE, Petersdorf S, Livingston RB, Gordon EE, Chapman CR, Appelbaum FR. “Samarium-153-EDTMP in bone metastases of hormone refractory prostate carcinoma: a phase I/II trial.” J Nucl Med 1993; 34(11): 1839-44. 626. Serafini AN, Houston SJ, Resche I, Quick DP, Grund FM, Ell PJ, Bertrand A, Ahmann FR, Orihuela E, Reid RH, Lerski RA, Collier BD, McKillop JH, Purnell GL, Pecking AP, Thomas FD, Harrison KA. “Palliation of pain associated with metastatic bone cancer using samarium-153 lexidronam: a double-blind placebo-controlled clinical trial.” J Clin Oncol 1998; 16(4): 1574-81. 627. Sciuto R, Festa A, Rea S, Pasqualoni R, Bergomi S, Petrilli G, Maini CL. "Effects of low-dose cisplatin on 89Sr therapy for painful bone metastases from prostate cancer: a randomized clinical trial." J Nucl Med 2002; 43(1): 79-86. 628. Nilsson S, Strang P, Ginman C, Zimmermann R, Edgren M, Nordstrom B, Ryberg M, Kalkner KM, Westlin JE. “Palliation of bone pain in prostate cancer using chemotherapy and strontium-89. A randomized phase II study.” J Pain Symptom Manage 2005; 29(4): 352-7. 113

Chapter 10. PROSTATE CANCER 629. Gittes RF, McCullough DL. “Occult carcinoma of the prostate: an oversight of immune surveillance--a working hypothesis.” J Urol, 1974: 112(2): 241-4. 630. Vieweg J, Rosenthal FM, Bannerji R, Heston WD, Fair WR, Gansbacher B, Gilboa E. “Immunotherapy of prostate cancer in the Dunning rat model: use of cytokine gene modified tumor vaccines.” Cancer-Res, 1994; 54(7): 1760-5. 631. Rodin AE, Larson DL, Roberts DK. “Nature of the perineural space invaded by prostatic carcinoma.” Cancer, 1967; 20: 1772-9. 632. Smith M. “The lymphatics of the prostate.” Invest Urol, 1996; 3: 439-4). 633. Gardiner RA, Fitzpatrick JM, Constable AR, Cranage RW, O’Donoghue EPN, Wickham JEA. “Human prostatic lymphoscintigraphy. A preliminary report.” Br J Urol, 1979; 51: 300 3. 634. Gardiner RA, Fitzpatrick JM, Constable AR, Cranage RW, O’Donoghue, EPN and Wickham JEA. “Improved techniques in radionuclide imaging of prostatic lymph nodes.” Br J Urol, 1979; 51: 561 4. 635. Zeng Y, Opeskin K, Horvath LG, Sutherland RL, Williams ED. “Lymphatic vessel density and lymph node metastasis in prostate cancer.” Prostate. 2005; 65 (3): 222 -30. 636. Zeng Y, Opeskin K, Baldwin ME, Horvath LG, Achen MG, Stacker SA, Sutherland RL, Williams ED. “Expression of vascular endothelial growth factor receptor-3 by lymphatic endothelial cells is associated with lymph node metastasis in prostate cancer.” Clin Cancer Res, 2004; 10(15): 5137-44. 637. Vesalainen S, Lipponen P, Talja M, Syrjanen K. “Histological grade, perineural infiltration, tumour-infiltrating lymphocytes and apoptosis as determinants of longterm prognosis in prostatic adenocarcinoma.” Eur J Cancer, 2004; 30A(12): 1797-1803. 638. Fong L, Small EJ. “Immunotherapy for prostate cancer.” Semin-Oncol 2003; 30: 649-58. 639. Reese DM, Small EJ, Magrane G, Waldman FM, Chew K, Sudilovsky D. “HER2 protein expression and gene amplification in androgen-independent prostate cancer.” Am-J-Clin-Pathol. 2001; 116(2): 234-9. 640. Blankenstein T. “The role of tumor stroma in the interaction between tumor and immune system.” Curr Opin Immunol. 2005; 17(2): 180-6. 641. Yewdell JW, Schubert U, Bennink JR. “At the crossroads of cell biology and immunology: DRiPs and other sources of peptide ligands for MHC class I molecules.” J Cell Sci, 2001; 114(Pt 5): 845-51. 642. Zinkernagel RM, Ehl S, Aichele P, Oehen S, Kundig T, Hengartner H. “Antigen localisation regulates immune responses in a dose- and time-dependent fashion: a geographical view of immune reactivity.” Immunol Rev, 1997; 156: 199-209. 643. Brandes M, Willimann K, Moser B. „Professional antigen-presentation function by human gammadelta T Cells.” Science, 2005; 309(5732): 264-8. 644. Figdor CG, deVries IJ, Lesterhuis WJ, Melief CJ. “Dendritic cell immunotherapy: mapping the way.” Nat Med, 2004; 10(5): 475-80. 645. Yao V, Platell C, Hall JC. “Dendritic cells.” ANZ J Surg, 2002; 72(7): 501-6. 646. Martin-Fontecha A, Thomsen LL, Brett S, Gerard C, Lipp M, Lanzavecchia A, Sallusto F. “Induced recruitment of NK cells to lymph nodes provides IFN-gamma for T(H)1 priming.” Nat Immunol, 2004; 5(12): 1260-5. 647. Ridge JP, Di-Rosa F, Matzinger P. “A conditioned dendritic cell can be a temporal bridge between a CD4+ T-helper and a T-killer cell.” Nature, 1998; 393(6684): 474-8. 648. Bennett SR, Carbone FR, Karamalis F, Flavell RA, Miller JF, Heath WR. “Help for cytotoxic-T-cell responses is mediated by CD40 signalling.” Nature, 1998; 393(6684): 478-80.

114

Chapter 10. PROSTATE CANCER 649. Schoenberger SP, Toes RE, van der Voort EI, Offringa R, Melief CJ. “T-cell help for cytotoxic T lymphocytes is mediated by CD40-CD40L interactions.” Nature, 1998; 393(6684): 480-3. 650. Troy A, Davidson P, Atkinson C, Hart D. “Phenotypic characterisation of the dendritic cell infiltrate in prostate cancer.” J Urol, 1998; 160(1): 214-9. 651. Kennedy-Smith, AG, McKenzie JL, Owen MC, Davidson PJ, Vuckovic S, Hart DN. “Prostate specific antigen inhibits immune responses in vitro: a potential role in prostate cancer.” J Urol, 2002; 168(2): 741-7. 652. McNeel DG, Malkovsky M. “Immune-based therapies for prostate cancer.” Immunol Lett, 2005; 96(1): 3-9. 653. Ciavarra RP, Somers KD, Brown RR, Glass WF, Consolvo PJ, Wright GL, Schellhammer PF. “Flt3-ligand induces transient tumor regression in an ectopic treatment model of major histocompatibility complex-negative prostate cancer.” Cancer Res, 2000; 60(8): 2081-4. 654. McNeel DG, Knutson KL, Schiffman K, Davis DR, Caron D, Disis ML. “Pilot study of an HLA-A2 peptide vaccine using flt3 ligand as a systemic vaccine adjuvant.” J Clin Immunol, 2003; 23(1): 62-72. 655. Eaton JD, Perry MJ, Nicholson S, Guckian M, Russell N, Whelan M, Kirby RS. “Allogeneic whole-cell vaccine: a phase I/II study in men with hormone-refractory prostate cancer.” BJU Int, 2002;89(1): 19-26. 656. Heckelsmiller K, Rall K, Beck S, Schlamp A, Seiderer J, Jahrsdorfer B, Krug A, Rothenfusser S Endres S, Hartmann G. “Peritumoral CpG DNA elicits a coordinated response of CD8 T cells and innate effectors to cure established tumors in a murine colon carcinoma model.” J Immunol, 2002; 169(7): 3892-9. 657. Dieli F, Gebbia N, Poccia F, Caccamo N, Montesano C, Fulfaro F, Arcara C, Valerio MR, Meraviglia S, DiSano C, Sireci G, Salerno A. “Induction of gammadelta Tlymphocyte effector functions by bisphosphonate zoledronic acid in cancer patients in vivo.” Blood, 2003; 102(6): 2310-1. 658. Heiser A, Coleman D, Dannull J, Yancey D, Maurice MA, Lallas CD, Dahm P, Niedzwiecki D, Gilboa E, Vieweg J. “Autologous dendritic cells transfected with prostate-specific antigen RNA stimulate CTL responses against metastatic prostate tumors.” J Clin Invest, 2002; 109(3): 409-17. 659. Ridgway D. “The first 1000 dendritic cell vaccinees.” Cancer Invest, 2003; 21(6): 873-86. 660. Simons JW, Mikhak B, Chang JF, DeMarzo AM, Carducci MA, Lim M, Weber CE, Baccala AA, Goemann MA, Clift SM, Ando DG, Levitsky HI, Cohen LK, Sanda MG, Mulligan RC, Partin AW, Carter HB, Piantadosi S, Marshall FF, Nelson WG. “Induction of immunity to prostate cancer antigens: results of a clinical trial of vaccination with irradiated autologous prostate tumor cells engineered to secrete granulocytemacrophage colony-stimulating factor using ex vivo gene transfer.” Cancer Res, 1999; 59(20): 5160-8. 661. Parkinson RJ, Mian S, Bishop MC, Gray T, Li G, McArdle SE, Ali S, Rees RC. “Disabled infectious single cycle herpes simplex virus (DISC-HSV) is a candidate vector system for gene delivery/expression of GM-CSF in human prostate cancer therapy.” Prostate, 2003; 56(1): 65-73. 662. Fong L, Brockstedt D, Benike C, Breen JK, Strang G, Ruegg CL, Engleman EG. “Dendritic cell-based xenoantigen vaccination for prostate cancer immunotherapy.” J Immunol, 2001; 167(12): 7150-6. 663. Harris DT, Matyas GR, Gomella LG, Talor E, Winship MD, Spitler LE, Mastrangelo MJ. “Immunologic approaches to the treatment of prostate cancer.” Semin Oncol, 1999; 26(4): 439-47.

115

Chapter 10. PROSTATE CANCER 664. Meidenbauer N, Harris DT, Spitler LE, Whiteside TL. “Generation of PSA-reactive effector cells after vaccination with a PSA-based vaccine in patients with prostate cancer.” Prostate, 2000; 43(2): 88-100. 665. Barrou B, Benoit G, Ouldkaci M, Cussenot O, Salcedo M, Agrawal S, Massicard S, Bercovici N, Ericson ML, Thiounn N. “Vaccination of prostatectomized prostate cancer patients in biochemical relapse, with autologous dendritic cells pulsed with recombinant human PSA.” Cancer Immunol Immunother, 2004; 53(5): 453-60. 666. Tjoa BA, Simmons SJ, Bowes VA, Ragde H, Rogers M, Elgamal A, Kenny GM, Cobb OE, Ireton RC, Troychak MJ, Salgaller ML, Boynton AL, Murphy GP. “Evaluation of phase I/II clinical trials in prostate cancer with dendritic cells and PSMA peptides.” Prostate, (1998) 36(1): 39-44. 667. Murphy GP, Tjoa BA, Simmons SJ, Jarisch J, Bowes VA, Ragde H, Rogers M, Elgamal A, Kenny GM, Cobb OE, Ireton RC, Troychak MJ, Salgaller ML, Boynton AL. “nfusion of dendritic cells pulsed with HLA-A2-specific prostate-specific membrane antigen peptides: a phase II prostate cancer vaccine trial involving patients with hormone-refractory metastatic disease.” Prostate, 1999; 38(1): 73-8. 668. Noguchi M, Kobayashi K, Suetsugu N, Tomiyasu K, Suekane S, Yamada A, Itoh K, Noda S. “Induction of cellular and humoral immune responses to tumor cells and peptides in HLA-A24 positive hormone-refractory prostate cancer patients by peptide vaccination.” Prostate, 2003; 57(1): 80-92. 669. Su Z, Dannull J, Yang BJ, Dahm P, Coleman D, Yancey D, Sichi S, Niedzwiecki D, Boczkowski D, Gilboa E, Vieweg J. “Telomerase mRNA-transfected Dendritic Cells stimulate antigen-specific CD8+ and CD4+ T cell responses in patients with metastatic prostate cancer.” J Immunol, 2005; 174:3798-3807. 670. Strobel I, Berchtold S, Gotze A, Schulze U, Schuler G, Steinkasserer A. “Human dendritic cells transfected with either RNA or DNA encoding influenza matrix protein M1 differ in their ability to stimulate cytotoxic T lymphocytes.” Gene Ther, 2000; 7(23): 2028-35. 671. VanTendeloo VF, Ponsaerts P, Lardon F, Nijs G, Lenjou M, VanBroeckhoven C, VanBockstaele DR, Berneman ZN. “Highly efficient gene delivery by mRNA electroporation in human hematopoietic cells: superiority to lipofection and passive pulsing of mRNA and to electroporation of plasmid cDNA for tumor antigen loading of dendritic cells.” Blood, 2001; 98(1): 49-56. 672. Grunebach F, Muller MR, Nencioni A, Brossart P. “Delivery of tumor-derived RNA for the induction of cytotoxic T-lymphocytes.” Gene Ther,2003; 10(5): 367-74. 673. Rock KL, Goldberg AL. “Degradation of cell proteins and the generation of MHC class I-presented peptides.” Annu Rev Immunol, 1999; 17: 739-79. 674. Rock KL, Clark K. “Analysis of the role of MHC class II presentation in the stimulation of cytotoxic T lymphocytes by antigens targeted into the exogenous antigenMHC class I presentation pathway.” J Immunol, 1996; 156(10): 3721-6. 675. Pardoll DM, Topalian, SL. “The role of CD4+ T cell responses in antitumor immunity.” Curr Opin Immunol, 1998; 10(5): 588-94. 676. Toes RE, Ossendorp F, Offringa R, Melief CJ. “CD4 T cells and their role in antitumor immune responses.” J Exp Med, 1999; 189(5): 753-6. 677. Heiser A, Maurice MA, Yancey DR, Wu NZ, Dahm P, Pruitt SK, Boczkowski D, Nair SK, Ballo MS, Gilboa E, Vieweg J. “Induction of polyclonal prostate cancerspecific CTL using dendritic cells transfected with amplified tumor RNA.” J Immunol, 2001; 166(5): 2953-60. 678. Livingston P. “Ganglioside vaccines with emphasis on GM2.” Semin Oncol, (1998) 25(6): 636-45. 679. Zhang S, Zhang HS, Reuter VE, Slovin SF, Scher HI, Livingston PO. “Expression of potential target antigens for immunotherapy on primary and metastatic prostate cancers.” Clin Cancer Res, 1998; 4(2): 295-302. 116

Chapter 10. PROSTATE CANCER 680. Musselli C, Livingston PO, Ragupathi G. “Keyhole limpet hemocyanin conjugate vaccines against cancer: the Memorial Sloan Kettering experience.” J Cancer Res Clin Oncol, 2001: 127 Suppl 2: R20-6. 681. Slovin SF, Ragupathi G, Musselli C, Olkiewicz K, Verbel D, Kuduk SD, Schwarz JB, Sames D, Danishefsky S, Livingston PO, Scher HI. “Fully synthetic carbohydratebased vaccines in biochemically relapsed prostate cancer: clinical trial results with alpha-N-acetylgalactosamine-O-serine/threonine conjugate vaccine.” J Clin Oncol, 2003; 21(23): 4292-8. 682. Sanda MG, Smith DC, Charles LG, Hwang C, Pienta KJ, Schlom J, Milenic D, Panicali D, Montie JE. “Recombinant vaccinia-PSA (PROSTVAC) can induce a prostatespecific immune response in androgen-modulated human prostate cancer.” Urology, 1999; 53(2): 260-6. 683. Cavacini LA, Duval M, Eder JP, Posner MR. “Evidence of determinant spreading in the antibody responses to prostate cell surface antigens in patients immunized with prostate-specific antigen.” Clin Cancer Res, 2002; 8(2): 368-73. 684. Raz E, Carson DA, Parker SE, Parr TB, Abai AM, Aichinger G, Gromkowski SH, Singh M, Lew D, Yankauckas MA. “Intradermal gene immunization: the possible role of DNA uptake in the induction of cellular immunity to viruses.” Proc Natl Acad Sci USA, 1994; 91(20): 9519-23.

Notes
1. http://www.andrologyaustralia.org/ 2. http://www.prostatehealth.org.au/ 3. http://www.cdc.gov/cancer/prostate 4. http://www.cancerbacup.org.uk/ 5. http://www.bostwicklaboratories.com/ 6. http://www.nhmrc.gov.au/publications/_files/cp88.pdf 7. http://www.nelh.nhs.uk/guidelinesdb/html/Prostate-ft.htm 8. http://www.uroweb.nl/files/uploaded_files/2005ProstateCancer.pdf 9. http://www.cancer.gov/cancertopics/understanding-prostate-cancertreatment/page5 10. http://www.prostatehealth.org.au/ 11. http://www.ohri.ca/decisionaid

117

Chapter 10. PROSTATE CANCER

118

Sponsor Documents

Or use your account on DocShare.tips

Hide

Forgot your password?

Or register your new account on DocShare.tips

Hide

Lost your password? Please enter your email address. You will receive a link to create a new password.

Back to log-in

Close