Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Mol Nutr Food Res. Author manuscript; available in PMC 2012 Jun 1.
Published in final edited form as:
PMCID: PMC3163457
NIHMSID: NIHMS318192
PMID: 21538852

Chemopreventive Effects of Tea in Prostate Cancer: Green Tea vs. Black Tea

Abstract

The polyphenol compositions of green tea (GT) and black tea (BT) are very different due to post-harvest processing. GT contains higher concentrations of monomeric polyphenols, which affect numerous intracellular signaling pathways involved in prostate cancer (CaP) development. BT polymers, on the other hand, are poorly absorbed and are converted to phenolic acids by the colonic microflora. Therefore, after consumption of GT higher concentrations of polyphenols are found in the circulation while after BT consumption the phenolic acid levels in the circulation are higher. The majority of in vitro cell culture, in vivo animal, and clinical intervention studies examine the effects of extracts of GT or purified (-)-epigallocatechin-3-gallate (EGCG) on prostate carcinogenesis. These studies provide strong evidence supporting a chemopreventive effect of GT, but results from epidemiological studies of GT consumption are mixed. While the evidence for a chemopreventive effect of BT is much weaker than the body of evidence with regard to GT, there are several animal BT intervention studies demonstrating inhibition of CaP growth. This article will review in detail the available epidemiological and human clinical studies, as well as animal and basic mechanistic studies on GT and BT supporting a chemopreventive role in CaP.

1 Introduction

The evidence for chemoprevention by any bioactive substance is drawn from a combination of epidemiological, animal and basic mechanistic studies as well as limited amounts of evidence from human intervention studies. This is particularly true for tea which is the second most commonly consumed beverage in the world after water. The vast majority of the tea consumed in the world is black tea (BT), comprising approximately eighty percent of all tea consumed [1]. Nonetheless, the scientific evidence for the chemopreventive activities of green tea (GT) is far more extensive than that available for BT. The key chemical differences between BT and GT result from post-harvest processing. During GT tea production the endogenous oxidase enzymes in tea leaves are inactivated by heating so that the green tea polyphenols (GTPs), including epigallocatechin gallate (EGCG), are preserved. In BT, natural oxidation after harvesting results in polymerization and formation of theaflavins and thearubigins with small amounts of EGCG remaining in BT [1]. After ingestion, both BT and GT undergo fermentation by normal colonic flora also known as the microbiome. This metabolism results in the formation of phenolic acids, which may be active in chemoprevention. Therefore, the key differences between GT and BT are that GTPs are found in higher concentration in the circulation after consumption of GT compared to BT. On the other hand, the phenolic acid levels in the circulation are higher after consumption of BT by comparison to GT. This article will review in detail the available epidemiological and human clinical studies, as well as animal and basic mechanistic studies on GT and BT supporting a chemopreventive role in prostate cancer (CaP). This review will also highlight the importance of investigating the bioavailability and metabolism of polyphenols in CaP chemoprevention.

2 Prostate cancer

Prostate cancer is the second most frequently diagnosed cancer and the sixth leading cause of cancer death among men worldwide, with 914,000 new cases and 258,000 deaths in 2008 [2]. More than half of these cases and deaths are expected to occur in more developed countries [2]. The highest country-specific age-standardized incidence rate (per 100000 population) in 2002 was 124.8 in the USA, and the lowest in Bangladesh (0.3), in China (1.6) and in India (4.4) [3]. Much of the observed geographic variation in CaP incidence may be due to differences in prostate-specific antigen (PSA) testing, and the ability to detect latent CaP [2]. However, differences in the use of PSA testing cannot explain all of the international variation, since there was already more than a 50-fold difference in international CaP incidence rates across countries in 1980 before the PSA test was introduced [3]. This indicates that environmental factors have a substantial role in determining cancer risk. In support of an environmental influence it was observed that men born in Japan and immigrating to Hawaii assimilated the host country's cancer rate both in their lifetime and in succeeding generations [4]. Diet is probably one of the most important environmental risk factors for CaP. The CaP incidence is low in Asian countries, and one possible explanation is the high intake of soy, tea, fish, fruits and vegetables and the reduced intake of red meat and fatty foods by comparison to the Western Diet [5]. However, the incidence of CaP is increasing rapidly in Asian countries due to the rapid introduction of a Western lifestyle, especially in urban centers [5]. CaP is typically diagnosed in men over 50 years of age and the rate of growth and progression is typically slow, which makes CaP an ideal disease in which to study the utility of preventive strategies, including those resulting from changes in nutrition and lifestyle [6]. Due to the fact that most CaP treatments carry the risk of side effects, there is an increasing trend for men diagnosed with less advanced CaP to choose expectant management, also called active surveillance, which provides additional opportunities to study non-toxic chemopreventive strategies such as GT.

3 Green and black tea

Tea is the most commonly consumed beverage second only to water in the world. All different varieties of tea including white, green, oolong, black and pu-erh teas are manufactured from the leaves of Camellia sinensis. For the production of GT the leaves are heat-treated to retain the typical polyphenols, also known as flavan-3-ols, including (-)-epigallocatechin (EGC), (-)-epigallocatechin-3-gallate (EGCG), (-)-epicatechin (EC), and (-)-epicatechin-3-gallate (ECG) [1]. For the production of BT the leaves undergo fermentation in moist and warm conditions [1]. Endogenous enzymes enhance the formation of polymers such as theaflavin and thearubigen. Therefore, BT contains a relatively small amount of monomeric polyphenols (Table 1). Other differences are the higher content of theanine and lower content of caffeine in GT [7]. Worldwide, 78 percent of tea produced is BT. By contrast GT comprises 20% of tea production and is the preferred form in China and Japan, as well as a few countries in North Africa and the Middle East. On the other hand, BT is consumed in European countries, the US, India, and many Arabic countries. The main tea producing countries are China, India, Kenya, Sri Lanka, Turkey, Indonesia, and Japan (Food and Agriculture Organization of the United Nations—Production FAOSTAT).

Table 1

Composition of green and black tea solids (percent) [1].

ConstituentsGreen Tea (%)Black Tea (%)
Flavan-3-ols30-423-10
Flavonols5-106-8
Other Flavonoids2-4--
Theogallin2-3--
Gallic acid0.5--
Quinic acid2.0--
Theanine4-60-2
Methylxanthines7-98-11
Theaflavin--3-6
Thearubigin--12-18

GT has been studied extensively for its chemopreventive activities with regard to CaP [8-10]. After evaluating all the evidence it becomes clear that there is a strong role for GT in chemoprevention of CaP, but much more research is needed on the chemopreventive potential of BT. More clinical research is needed for both GT and BT.

4 Animal studies

Convincing evidence exists showing that GT extracts (GTE) decrease tumor growth in a variety of animal models. However, there are very few animal studies using brewed GT, and BT has not been studied as often as GT. However, several investigations using BT extracts demonstrated an inhibition of CaP growth in mouse xenograft and a rat models [1;11;12]. The majority of mouse studies are summarized in the review by Johnson et al [8]. Mouse xenograft models used both androgen-dependent (LNCaP and CWR22Rv1) and androgen-independent (PC-3, LNCaP-R104, CWR22R) cell lines. Both demonstrated a decrease in tumor growth with administration of EGCG or GTPs (Table 2). We will highlight a few important studies. Two studies investigated whether inhibition of tumor growth by GT intervention depends on the stage of CaP development [13;14]. In the study by Harper et al. 0.06% EGCG in drinking water decreased tumor growth at 12 weeks but not at 28 weeks of age in transgenic adenocarcinoma of the mouse prostate (TRAMP) model by inhibiting proliferation, inducing apoptosis, decreasing androgen receptor, insulin-like growth factor-1 (IGF-1), IGF-1 receptor, phosphor-extracellular signal-regulated kinases (ERK) 1 and 2, cyclooxygenase-2 and inducible nitric oxide synthase (iNOS) [14]. The study by Adhami et al. demonstrated that treatment of TRAMP mice with 0.1% of GTE extended tumor-free survival significantly when the intervention was initiated prior to week 28. In addition, a decrease in IGF-1, IGFBP-3 and phosphatidylinositol 3-kinase (PI3K)/AKT/ERK was only observed when treatment was started prior to week 28 [13]. Another important mouse study explored the question of the GTE concentration necessary for chemoprevention in an androgen dependent xenograft model tumor model in nude mice [11]. The investigators determined that the administration of 0.05% of GTE in drinking water was associated with a significant extension in days to reach 1200 mm tumor volume but not at 0.01%. This was associated with a decrease in PSA, an increase in apoptosis (PARP, caspase, Bax, Bcl2) and a decrease in vascular endothelial growth factor (VEGF) [11]. However in a few mouse studies GT administration was not effective [15;16]. For example in the study by Zhou et al. an intraprostatic inoculation of male severe combined immune deficient (SCID) mice with the androgen-sensitive human LNCaP CaP cell line was used. Oral administration of brewed BT but not GT in drinking water demonstrated a significant decrease in tumor weight associated with decreased serum PSA [15]. China green and black leaf tea imported from Shanghai Tea Import was used at a concentration of 15 g of tea leaves per liter of water [15]. On a weight basis this is a much higher intake of BT than typical for humans. Another interesting xenograft mouse model included the subcutaneous injection of TRAMP-C1 cells with or without lipopolysaccharide (LPS) in C57/Bl mice [16]. 0.6% GTE (59% EGCG) was administered in drinking water starting three days prior to tumor injection. GTPs did not inhibit tumor growth. However, LPS-induced recruitment of inflammatory polymorphonuclear phagocytes (PMNs) significantly decreased tumor growth, which was inhibited by GTP administration [16].

Table 2

Studies of prostate cancer prevention and treatment in animal models.

Animal ModelChemopreventive AgentOutcomeReference
Androgen-sensitive LNCaP intraprostatic in SCID miceOral brewed GT and BT: 15g leaves/L of water↓tumor growth, ↓DHT with GT and BT, ↓PSA with BT not GTZhou JR et al. 2003 [15]
Androgen-sensitive CWR22Rv1 in nude miceOral 0.05 and 0.01% of GTE (62% EGCG) when tumor was 400mm3↓tumor growth, ↑apoptosis, ↓VEGF, ↓PSASidiqui IA et al. 2006 [11]
Androgen-sensitive CWR22Rv1 in nude miceOral 1.25% BT extract↓tumor growth, ↑apoptosis, ↓VEGF, ↓PSASidiqui IA et al. 2006 [11]
CWR22R androgen-independent in nude miceEGCG and EGCG-P↓tumor growth, ↑apoptosis, ↓angiogenesis↓PSALee SC et al. 2008 [100]
s.c.TRAMP-C1 cells into C57/Bl male miceOral 3d prior to tumor 0.6% of GTE (59% EGCG)No effect on tumor growthSartor L et al. 2004 [16]
TRAMP C57BL/6 miceOral 0.1% GT extract (62% EGCG) starting at 8 wks of ageAt 20 wks 44% inhibition of tumor growth, 30 wks 42%, ↓serum IGF-1, ↑IGFBP-3, ↑apoptosisGutpa S et al. 2001 [101]
TRAMP C57BL/6 miceOral 0.1% GTE (62% EGCG) starting at 8 wks of age↓tumor growth, ↓serum IGF-1, ↑IGF BP-3, ↓PI3K, ↓Akt, ↓ERK1/2, ↓VEGF, ↓urokinase plasminogen activator, ↓matrix metalloproteinasesAdhami VM et al. 2004 [74]
TRAMP C57BL/6 miceOral 0.3% GTE (51.9% EGCG)8-genes differentiated between prostate of wt from transgenic mice and TRAMP+GTP responsive from non-responsive miceScaltriti M et al. 2006 [102]
TRAMP C57BL/6 miceOral 0.3% GTE (51.9% EGCG)At 24 wks only 20% developed tumors in GT group, ↓MCM7McCarthy S et al. 2007 [103]
TRAMP C57BL/6 miceOral 0.06% EGCG starting at 5 weeksAt 12 weeks ↓tumor growth, but not at 28 wks. ↑apoptosis, ↓AR, ↓serum IGF-1,↓ERK1/2, ↓COX-2,↓iNOSHarper CE et al. 2007 [14]
TRAMP C57BL/6 miceOral 0.1% GTE (62% EGCG) starting at 4 wks of age↓NFkB, IKKa, IKKb, RANK, NIK, STAT-3 with a trend to decrease over time (8-32 wks)Siddiqui IA 2008 [18]
TRAMP C57BL/6 miceOral 0.1% GTE (62% EGCG) starting at 6, 12, and 18 wks of ageTumor-free survival of 38 wks (GTP at 8 wks), 31 wks (GTP at 12 wks) and 24 wks (GTP at 18 wks) vs 19wks in controlAdhami VM 2009 [13]
TRAMP C57BL/6 miceOral 0.05% GTE at 4 wks of ageNo effect on tumor weight and urine 8-OHdGTeichert F 2008 [19]
TRAMP C57Bl/6:FVB 50:50Oral 0.1% of GTE (35% EGCG) at 4 or 6 wks of age and oral 0.1-0.6% GTPsNo effect on tumor progression. At 12 weeks of 0.1-0.6% GTPs no effect on DNA methylation statusMorey Kinney SR et al. 2009 [20]
Lobund Wistar ratsOral 0.2% of GT (freeze dried – 10% EGCG)50% less tumors compared to water cntr, no effect on 8-OHdG in prostate, ↑MnSODO'Sullivan J et al. 2008 [21]
Nobel rats CaP induced with testosteroneOral 2% GT and 200 g soy protein/kg dietOnly GT+soy ↓prostate hyperplasia, ↓NFκB p50 binding activity, ↓TNF-a, ↓IL-6, ↑apoptosisHsu A et al. 2010 [104]
Wistar rats s.c. injection of testosterone (5mg/kg bwt)Oral 0.5, 1, 1.5% (w/v) BT extract↓oxidative stress induced by testosterone, ↓SOD, ↓GST, ↓GR, ↓Cat, ↓lipid peroxidationSiddiqui IA et al. 2005 [12]

Most mouse studies investigating the chemopreventive and therapeutic effects of GT were performed using the TRAMP model and using the chemical EGCG or GTEs. TRAMP mice express the rat probasin (PB)-SV40 early gene (T/t antigen) construct under prostate specific control of the minimal rat probasin promoter and display mild to severe hyperplasia of the prostate epithelium, resembling prostatic intraepithelial neoplasia (PIN) by 6-12 weeks of age. Between 10 to 16 weeks well-differentiated neoplasia is generally observed, between 18-24 weeks of age all of the mice will display primary tumors, and by week 30 will display metastases to distant sites [17]. Table 2 gives an overview of the TRAMP tea intervention studies. Most of the studies used GTE enriched in EGCG. The lowest effective concentration was 0.1% GTE in drinking water. In most studies GTPs were administered starting at 4 to 8 weeks of age. When GT was started at later points in time, the effect on tumor progression was decreased [13;18]. This could indicate that GT has more of a preventive effect earlier in transgenic tumor progression and less of a therapeutic effect once tumors are established. Moreover, not all investigations using TRAMP mice demonstrated protection. A study by Teichert et al. used 0.05% GTPs with unknown EGCG content and did not show a decrease in tumor weight [19]. Another study by Morey Kinney et al. administered orally 0.1% GTPs with 35% of EGCG and did not show a decrease in tumor progression [20]. In the same study no effect on DNA methylation in the prostate of TRAMP mice was found after administration of 0.1-0.6% GTPs in drinking water.

Few rat studies have been performed to investigate the chemopreventive effect of GT. One study by O'Sullivan et al. showed a decrease in tumor progression of 50% compared to water control but no decrease in tissue oxidative DNA damage marker, 8-hydroxydeoxyguanosine (8-OHdG). However manganese superoxide dismutase (MnSOD) protein expression was increased by the tea treatment [21].

Animal models clearly provide rapid answers by comparison to clinical studies, and serve a useful purpose in highlighting potential actions and interactions of candidate chemopreventive substances such as tea. However, there is some question as to how these studies relate to human benefit. Many of the studies are done with subcutaneous injection of tumor cells into an immune-impaired host and do not replicate the complex microenvironment of the human prostate gland. Nonetheless, the xenograft tumors are invaded by host immune and stromal cells and demonstrate tumor angiogenesis and elements of metastasis in appropriate model systems. Therefore, they do recapitulate elements of human CaP. Ultimately human intervention studies provide the best quality of evidence for the chemopreventive potential of tea.

5 Human epidemiological studies

As summarized in Table 3, twelve population studies, including five case-control and seven cohort studies, have been published evaluating the association of tea consumption with CaP. Among these population studies four evaluated the use of GT alone, two evaluated GT and BT, one evaluated only BT and five did not define the type of tea. The studies analyzing “tea” from Western countries presumably examine BT since this is the most common tea consumed in Western countries. Among the cohort studies, one GT study showed a protective effect in advanced CaP [22] and one BT study showed a protective effect in localized CaP associated with “tea” consumption [23]. Two GT cohort studies in Japan showed no association, one GT and BT cohort study from Hawaii in men with Japanese ancestry showed no effect and two cohort studies not defining the type of tea (Canada, USA) showed no effect (Table 3).

Table 3

Population studies investigating the prevention and treatment of prostate cancer using green and black tea.

LocationType of StudyDaily DoseOdds RatioReference
Green tea, JapanCohort Study, advanced CaP N=49,920, cases=404≥ 5 cups, protective0.6 (p=0.01)Kurahashi N et al. 2008 [22]
Black tea, HawaiiCohort of men of Japanese decent n=7833, cases=149 >58 yrsOnce per day, > 10 years, protective0.6 (p=0.04)Heilbrun LK et al. 1986 [23]
Green tea, JapanCohort Study, n=19561, cases=110≥5 cups, no association0.85 (P=0.81)Kikuchi N et al. 2006 [28]
Green tea, JapanCohort Study N=18115, cases=196No associationP=0.16Allen NE et al. 2004 [30]
Green tea and black tea, USA (Hawaii)Cohort Study n=7999, cases=174 Japanese ancestryNo associationGT ever vs never 1.47 BT ever vs never 0.83Severson et al. 1989 [29]
Tea, CanadaCohort Study, n=3400, cases=145>500 mL, No association1.02Ellison et al. 2000 [105]
Tea, UK (London)Prospective Cohort n=14085, cases=185>10 compared to <4, no association0.8 (p=0.3)Kinlen LJ et al. 1988 [106]
Green tea, ChinaCase control cases=130, controls=274Protective effect0.28Jian L et al. 2004 [24]
Green and black tea, JapanCase control cases=140, controls=140≥10 cups protective trend, not significantGT 0.67 (p=0.3) BT 1.5Sonoda T et al. 2004 [25]
Tea, CanadaCase control cases=617, controls=637>500 ml, protective0.7 (p=0.05)Jain MG et al. 1998 [31]
Tea, ItalyCase control cases=107, Controls=6147>1 cup no association0.9LaVecchia C et al. 1992 [26]
Tea, CanadaCase Control Study (cases=1623, controls=1623)> 4 cups/day, No associationVilleneuve et al. 1999 [27]

Among the case control studies one GT study from China showed a protective effect [24]. One tea study from Canada, most likely investigating BT, showed a protective effect and another study from Japan investigating GT and BT showed a trend to protective effect [25]. However there were two case control studies from Italy and Canada with “tea” not showing an association [26;27].

The following cohort studies provide evidence for a positive association with the intake of GT and a reduced incidence of CaP

The cohort study by Kurahashi et al. used the Japan Public Health Center-based cohort with 49,920 men aged 40-69 years assessed GT consumption habit at baseline and followed participants for 11 years [22]. GT was not significantly associated with localized CaP, but demonstrated a protective effect for advanced CaP [odds ratio (OR)=0.52, ptrend=0.01] in comparing men drinking 5 or more cups/day compared with those who consumed less than 1 cup per day. [22]. An earlier cohort study by Heilbrun et al. involving 7833 men of Japanese ancestry living in Hawaii observed a weak but significant negative association of BT consumption and CaP incidence, with relative risk (RR) being 0.6 for those consuming more than one cup of BT daily versus almost never [23].

The following cohort studies fail to show any association of GT intake and prostate cancer incidence

The study by Kikuchi et al. using the Ohsaki cohort with 19 561 men aged 40-79 followed a Japanese population for 7 years and found 110 incident cases of CaP. There was no protective effect of GT (OR=0.85, ptrend=0.81) [28].

Another cohort of 7999 men of Japanese ancestry living in Hawaii showed a borderline significant increase in risk for GT consumption, OR = 1.47 (95% CI: 0.99–2.19), but no association for BT [29]. Another cohort study by Allen et al. of 18 115 men including 196 cases of CaP demonstrated that men who drank 5 or more cups of GT had a 29% non-significant increase in CaP risk compared to those who drank tea less than once per day, (OR=1.29, ptrend=0.16) [30].

The following case control studies provide evidence for a positive association with the intake of GT and a reduced incidence of CaP

The case control study from Southern China included 130 cases and 274 controls and provided information on duration, quantity and frequency of usual tea consumption [24]. Among the cases 55.4% were tea consumers compared to 79.9% of the controls. The CaP risk declined with increasing frequency, duration and quantity of GT consumption. The adjusted OR was 0.28 for tea drinking vs. non-tea drinking [24]. A slight reduction in CaP risk was also reported by a larger Canadian case-control study involving 617 cases and 637 population controls [31]. An OR of 0.70 (95% CI: 0.50–0.99) was reported with consumption of more than 500 g (approximately two cups) of tea per day [31].

Other case control studies fail to show any association of GT intake and prostate cancer incidence

The Japanese case control study by Sonoda et al. included 140 cases and controls and evaluated dietary habits based on a semi-quantitative food frequency questionnaire [25]. A modest non-significant reduction in risk (OR=0.67, p=0.3) of the fourth vs. first quartile >10 cups vs. <1 cup of GT was found. However in this study only the intake of fish and natto were associated with a significant decrease in risk of CaP while the effect for tofu and all soy products were not significant [25]. Two more case control studies from Canada and Italy with 107 cases and 1623 cases, respectively, showed no association between tea consumption and CaP [26;27]. In summary evidence from population studies of the protective effect of tea and CaP is not conclusive. The variability between population studies is most likely based on the fact that the majority of publications seeking an association between tea consumption and the risk of CaP did not consider the type of tea consumed, the method of tea preparation and the tea polyphenol content, which is dependent on the method of preparation. Our own study of commercially available green teas in the United States showed a large variation in GTP content [32]. Most of the studies showing a decrease in CaP risk were performed in Asian countries (Japan and China) or in Hawaii with men of Japanese descent. It should be noted that the failure to demonstrate an association of CaP incidence with the intake of tea should not be taken as proof of a lack of any association. As already indicated, population studies depend on numerous lifestyle variables which often cannot be adequately accounted for in analysis and can lead to false associations or failure to demonstrate associations. Intervention studies in humans also have numerous limitations discussed below and may also not yield conclusive evidence.

6 Human intervention studies

Data from two intervention studies in patients with localized CaP and two in patients with hormone refractory CaP have been published [33-36]. The first intervention study of localized CaP, performed in Italy, showed that GTPs are effective in delaying progression of premalignant lesions to CaP was [33]. 60 men with high-grade prostate intraepithelial neoplasia (HG-PIN) in the baseline biopsy were randomized to either 600 mg of a decaffeinated GTE containing 51.8% of EGCG or placebo daily. At the end of one year repeat biopsy samples were taken and men in the placebo group had a 30 percent incidence of CaP compared to only 3 percent in the group receiving the GTE [33]. Serum PSA showed a non-significant trend to decrease in GTE-treated men at 9 and 12 months. A significant improvement was observed for the international prostate symptoms score. GTE administration also reduced lower urinary tract symptoms, suggesting that GTPs might be helpful in treating benign prostate hyperplasia (BPH). A two year follow up assessment was published in 2008 in a letter to the editor of European Urology [37]. Only 9 participants from the placebo-arm and 13 from the GTE-arm underwent a third biopsy sample collection. Despite the high drop-out rate the two arms remained balanced and large enough for statistical analysis. Two further cancer diagnoses appeared in the placebo arm and one in the GTE-arm. This indicates that overall, even after suspension of the GTE treatment, the GT-arm experienced an almost 80% reduction in CaP diagnosis compared to the placebo group [37]. In the second smaller intervention trial with localized CaP 25 men diagnosed with stage I, II, or III CaP, who were scheduled for prostatectomy consumed 4 capsules of Polyphenon E daily containing a total daily dose of 800 mg of EGCG [34]. The average intervention time was 4-6 weeks. Only one patient reported mild nausea related to the Polyphenon E intake. No adverse effects on liver function were observed. A significant decrease in serum PSA, hepatic growth factor (HGF), vascular endothelial growth factor (VEGF), insulin growth factor 1 (IGF-1) and ratio of IGF-1 to insulin growth factor binding protein 3 (IGFBP-3) was found [34]. In addition two intervention studies from Canada and the US investigating hormone refractory CaP showed a very limited protective effect on CaP by consumption of GTE [35;36]. In one study 6g of a pulverized GT powder which contained sugar, citric acid and flavoring was administered [36]. However no information was provided about the GTP content of this GT product. Only one participant of 42 showed a decrease in PSA level. The second intervention study with hormone refractory CaP administered a dose of 250 mg of GTE containing 75% GTPs twice daily [35]. Nine patient of 15 consuming the GTE for more than 2 months had progressive disease and six showed an apparent slowing of disease progression with a slow rise in PSA serum level. However, there were no “responders” as per the traditional definition of a PSA drop of greater than 50%. There were no control groups in either of the two studies. Unfortunately no clear conclusion can be drawn from the studies of hormone refractory disease due to lack of information on the GTP content in the first study [36] and the use of a relatively low dose in the second study [35]. The results of the intervention studies support the findings from animal studies that GT is more effective in prevention and early stages of CaP and less effective in hormone refractory disease.

7 Bioavailability and metabolism of tea polyphenols

The limited bioavailability and extensive metabolism of tea polyphenols after absorption provides a challenge to their utilization for chemoprevention [38]. Tea polyphenols are mainly absorbed from the small intestine. The absorption is regulated by multidrug resistance-associated protein transporter (MRP) and monocarboxylate transporter (MCT) [38]. MRP1 is located at the basal membrane and MRP2 at the apical membrane. GTPs are taken up into the epithelial cells and metabolized to glucuronides, sulfates and methyl metabolites [39]. Both GTPs and metabolites can be transported to the intestinal lumen or into the vascular system [38]. In humans and mice the rate of conjugation depends on the chemical structure of the GTPs. Gallated polyphenols such as EGCG and ECG are found in the circulation in the free form whereas nongallated polyphenols circulate mostly in conjugated form [40]. In addition all GTPs undergo methylation catalyzed by catechol-O-methyltransferase (COMT) [41]. Non-gallated polyphenols are excreted through the kidney into the urine, whereas gallated polyphenols are not found in the urine and have been demonstrated to be excreted through the enterohepatic circulation [40].

Human prostate tissue concentrations have been determined by our laboratory in men who participated in a phase II tea intervention study at the University of California Los Angeles. Men consumed 6 cups of GT or BT 3-8 weeks prior to prostatectomy. The 6 cups of GT provided 571 mg of EGCG, 291 mg of EGC, 75 mg of EC and 89 mg of ECG daily. Tissue aliquots collected after prostatectomy from men consuming GT contained EGCG, 4″-O-MeEGCG and ECG (Table 5) [42]. No GTPs were found in tissue from men in the BT or control groups. Participants collected urine samples during the intervention. Urine samples from men consuming GT contained EGC, 4′-O-MeEGC and EC (Table 6). Urine samples from men consuming BT contained 100 fold lower amounts of the same GTPs. No GTPs were found in the control group consuming only water [42]. In a prior pilot study, also designed as a preprostatectomy trial, a Darjeeling BT was used, which contained higher amounts of GTPs [43]. In this human study no theaflavins were found in the human prostate after BT consumption. However, when administering a very high concentration of decaf. BT extract (5%) mixed into the diet to C57BL/6 mice we demonstrated that theaflavins can be absorbed into liver, prostate, small intestine and colon [43].

Table 5

Concentration of GTPs in human prostate tissue and the percent occurring in free, glucuronidated or sulfated form from men consuming 6 cups of GT daily for 3-6 weeks (mean±std; n=8) [42].

GTPTotal (pmol/g prostate)Glucuronide (%)Sulfate (%)Free form (%)
EGCG42.1±32.423.210.666.2
4″-O-methyl EGCG38.9±19.513.310.176.6
ECG17.8±10.111.39.878.9

Table 6

Concentration of GTPs in human urine from men consuming either 6 cups of GT or BT for 3-6 weeks (μmol/g creatinine; mean±std, n=8) [42].

EGCEC4'-MeEGC
Baseline0.0 ± 0.00.0 ± 0.00.0 ± 0.0
Week 3 GT intervention7.8 ± 3.96.9 ± 3.412.4 ± 5.9
Week 3 BT intervention0.016 ± 0.020.011± 0.010.023 ± 0.03

Due to the high rate of EGCG-methylation in human prostate tissue, we investigated the bioactivity of the methyl metabolite of EGCG. We demonstrated that methylation of EGCG decreased its pro-apoptotic and nuclear factor kappa B (NFκB) inhibitory activities [42]. Similar observations were made by other investigators, who found that methylation of EGCG and ECG decreased their proteasome-inhibitory activity [44]. As shown in Table 5 and and66 about 50% or more of EGCG and EGC found in human prostate and urine are in the methylated and less active form. Similar methylation rates were found in tissues of mice drinking brewed GT as drinking water (Figure 2) (unpublished data). GTP concentrations were similar in mouse and human tissues. However in human tissue only EGCG, 4″-MeEGCG and ECG were found, whereas the mouse tissue contained EC, EGC, 4′-MeECG, EGCG and 4″-MeEGCG. After the intragastric administration of 164 μmol/kg (equivalent to about 2.2 mg per 30 g mouse) EGCG to mice similar EGCG tissue concentration were found with (Cmax) for EGCG (nmol/g) 0.09-0.2 in prostate, 0.03-0.1 in liver, 0.002-0.01 in lung, 45±14 in small intestine and 7.9±2.4 in colon [45]. A study by Meng et al. found EGCG and 4′4″-DiMeEGCG in liver, kidney and small intestine, plasma, urine and feces [46]. As substrates of methyltransferases, GTPs are able to inactivate methylating enzymes such as COMT and DNA methyltransferase (DNMT1). There is the potential that GT may inhibit carcinogenesis by limiting methylation of DNA and affecting the expression of proteins that stimulate proliferation [47;48]. The inhibition of DNMT1 activity by GTPs in cell culture has been demonstrated to lead to demethylation of the CpG islands in the promoter regions and the reactivation of methylation-silenced genes such as p16INK4a, retinoic acid receptor beta, O6-methylguanine methyltransferase, human mutL homolog 1, and glutathione S-transferase-pi (GSTp1) [48]. Since CaP is commonly associated with hypermethylation and silencing of GSTp1 it could be possible that GT intervention may assist in reactivation of GSTp1 [49;50] thus increasing its antioxidant activity and inhibiting tumor growth.

An external file that holds a picture, illustration, etc.
Object name is nihms-318192-f0002.jpg

Tissue concentrations of GTPs were determined in SCID mice inoculated s.c. with LAPC4 androgen-dependent prostate cancer cells. Drinking water was replaced with brewed green tea containing 0.07% GTPs freshly prepared every 2 days. 10 weeks after tumor cell inoculation mice were sacrificed and tumor, lung, liver and kidney tissue was collected and analyzed by high performance liquid chromatography with Coularray electrochemical detection (mean±std; n=5)[107].

Due to the limited absorption of GTPs and BT theaflavins, the GTPs remaining in the small intestine are transformed in the colon by the microflora to phenolic acids. In a study designed to examine the effects of GT and BT digestion products, an artificial colon containing human colonic microflora (TNO, Holland) was incubated with GT and BT concentrates [51]. The major phenolic acids found in the artificial colon content after GT and BT digestion were 3-methoxy-4-hydroxyphenylacetic acid, 4-hydroxyphenyl acetic acid, 3,4- dihydroxyphenylacetic acid, 3,3-hydroxyphenylpropionic acid, 2,4,6-trihydroxybenzoic acid [51].

The composition of BT varies strongly depending on the growing region and manufacturing conditions [32]. For example, Darjeeling tea contains higher amounts of EGCG compared to regular BTs [32]. Most in vivo BT intervention studies have analyzed plasma GTP concentrations. A study by Mulder et al. demonstrated that theaflavins are only minimally bioavailable [52]. 800 mg of theaflavin, equivalent to 30 cups of BT, was administered to two volunteers leading to a plasma concentration of 1 μg/L at 2 hours [52]. It is unlikely that theaflavin and thearubigins directly contribute to the chemopreventive effects of BT. Biological effects of BT consumption most likely are due to the GTPs and phenolic acids resulting from colonic metabolism. Therefore, in vitro evidence of potential molecular targets of BT theaflavins will not be discussed as this research is unlikely to be relevant to CaP chemoprevention. However, the bioactivity of the microflora products may be more relevant to the chemopreventive effects of BT. Further research is needed to determine whether the tissue concentrations of phenolic acids are sufficient to exhibit a chemopreventive effect. Preliminary research indicated that the IC50 concentration to inhibit cell proliferation in HCT116 colon cancer cells at 75 μmol/L is much higher than the achievable tissue concentrations [51].

BT frequently is consumed in combination with a small amount of milk. The majority of investigations demonstrated that the addition of milk did not decrease the bioavailability of GTPs from BT [53;54]. However, one study by Reddy et al demonstrated a decrease of plasma GTPs when BT was consumed with milk [55]. The same study also demonstrated that the plasma antioxidant activity was not affected by milk [55]. Other studies showed mixed results on the effect of BT and milk combination on the antioxidant activity. The study by Kyle et al. 2007 demonstrated no decrease of the plasma antioxidant activity but the study by Ryan et al. demonstrated a decrease in the in vitro antioxidant of BT when combined with milk [54;56].

8 Molecular targets of GTPs in prostate cancer

Polyphenolic botanical extracts such as GTPs exert their effects on tumor growth through multiple mechanisms reviewed below.

8.1 Antioxidant/prooxidant activity

Most molecular targets have been investigated in mechanistic studies conducted in cell culture. The stability of tea polyphenols decreases with increasing pH above pH 7. Extensive studies by Neilson et al. demonstrated that in alkaline conditions EGCG undergoes dimerization and autoxidation involving the B-ring to form homodimers or heterodimers with other tea polyphenols such as EGC, forming theasinensins [57]. This process leads to the concurrent formation of hydrogen peroxide [58] and results in prooxidant actions of EGCG in cell culture [59]. Hydrogen peroxide in cell culture exhibits many activities similar to EGCG which makes the interpretation of results from studies of EGCG complex. Therefore many results from cell culture experiments will need to be confirmed in animal studies. The antioxidant activity of tea polyphenols has been summarized in a review by Lambert and Elias (2010) [60] and is also described in a chapter by Lambert et al. in this special edition.

To date, no human studies have been published demonstrating antioxidant activity of GTPs in CaP. Unpublished data from our laboratory demonstrated a decrease in oxidative DNA damage (ratio of 8-OHdG/guanosine) and oxidative protein damage (carbonyl protein) in LAPC4 prostate xenograft tumors in SCID mice consuming brewed GT in place of drinking water (manuscript in preparation). The GT contained 700 mg/L (0.07%) of total GTPs including 634 mg (0.063%) of EGCG.

Since tissue concentrations of GTPs are very low, a direct chemical antioxidant effect is unlikely. However, EGCG may act indirectly through stimulating the transcription factor, erythroid 2p45 (NF-E2)-related factor 2 (Nrf2) [61]. Nrf2 mediates the expression of key antioxidant enzymes through the antioxidant-response element (ARE) [62]. These antioxidant enzymes include glutathione reductase (GSR), glutathione peroxidase (GPX), glutathione-S-transferase (GST), glutamate-cysteine ligase (GCL), manganese superoxide dismutase (MnSOD), NAD(P)H:quinone oxidoreductase (NQO1), heme oxygenase-1 (HO-1), thioredoxin reductase1 (TRX1) and peroxiredoxin (PRX1) [63]. It has been demonstrated that Nrf2 and members of the glutathinone-S-transferase (GST) mu family are extensively suppressed by gene methylation in human CaP [64;65]. Using the TRAMP transgene and Nrf2 knockout murine models it was demonstrated that the loss of Nrf2 initiates a detrimental cascade of reduced GST expression, elevated ROS levels and ultimately DNA damage associated with tumorigenesis [65;66]. It has been demonstrated that treatment with EGCG was able to stimulate nuclear accumulation, ARE binding and transcriptional activity of Nrf2 in MCF10A breast cancer and Caco-2 colon cancer cells [61;67].

8.2 Apoptosis and cell cycle arrest

GTPs have been demonstrated to induce cell cycle arrest at G1 phase and induce apoptosis in androgen dependent LNCaP and androgen-independent PC-3 and DU145 cells [68-70]. However different mechanisms have been held responsible. For example EGCG-induced apoptosis in human prostate carcinoma LNCaP cells (p53 wildtype) was mediated via modulation of two related pathways: (a) stabilization of p53 by phosphorylation on critical serine residues and p14ARF-mediated downregulation of murine double minute 2(MDM2) protein, and (b) negative regulation of NFκB activity, thereby decreasing the expression of the proapoptotic protein Bcl-2 [68]. The altered expression of Bcl-2 family members triggered the activation of initiator caspases 9 and 8 followed by activation of effector caspase 3 followed by poly (ADP-ribose) polymerase cleavage and induction of apoptosis [68]. In DU145 (p53 mutant) and PC-3 (p53 null), apoptosis was induced through different mechanisms such as through an increase in reactive oxygen species (ROS) formation and mitochondrial depolarization [69]. These observations were further supported by the fact that the rank order of the effects of different GTPs in growth suppression, apoptosis induction, ROS formation and mitochondrial depolarization were similar (i.e. ECG > EGCG > EGC > EC) [69]. In another study EGCG treatment of LNCaP and DU145 cells resulted in significant dose- and time-dependent (i) upregulation of the protein expression of WAF1/p21, KIP1/p27, INK4a/p16, and INK4c/p18, (ii) down-modulation of the protein expression of cyclin D1, cyclin E, cdk2, cdk4, and cdk6, but not of cyclin D2, (iii) increase in the binding of cyclin D1 toward WAF1/p21 and KIP1/p27, and (iv) decrease in the binding of cyclin E toward cdk2 [70].

8.3 Insulin-like growth factors and binding proteins

The insulin-like growth factors (IGF-1 and IGF-2) and their binding proteins (IGFBP) play central roles in cell growth, differentiation, survival, transformation and metastasis. The biological effects of the IGFs are mediated by the IGF-1 receptor (IGF-1R), a receptor tyrosine kinase homologous to the insulin receptor (IR) [71]. IGF-1 interacts with IGF-1R to induce a series of ligand-mediated receptor activation and mitogenic responses, including PI3K/AKT and RAS/RAF/MAPK cascade, controlling cell survival and cell proliferation, respectively. Activated AKT also stimulates NFκB transcriptional activity and the mammalian target of rapamycin (mTOR) [72]. Epidemiological observations indicate that circulating IGF-1 levels are positively associated with increased risk of CaP [72]. A study by Li et al demonstrated that EGCG is a highly potent inhibitor (IC50=14 μmol/L) of IGF-1R tyrosine kinase activity and malignant cell growth [73]. Furthermore, it was found that IGF-1R autophosphorylation in the presence of increasing ATP concentrations was unaltered by EGCG treatment [73]. Mouse studies using the TRAMP model demonstrated the inhibition of IGF-1 signaling [14;74]. In addition, the above-mentioned human Polyphenon E intervention study by McLarty et al confirmed a decrease in IGF-1 and the ratio of IGF-1 to IGFBP-3 [34].

8.4 Inflammation

Inflammation is implicated as a major risk factor of CaP. Population studies have found an increased relative risk of CaP in men with prior histories of prostatitis [75]. Benign prostatic hyperplasia (BPH), a condition which often precedes and coexists with CaP, demonstrates signs of inflammatory response. Specifically, almost all human BPH specimens showed inflammatory infiltration and high expression of pro-inflammatory cytokines, including interleukin-17 (IL-17) that promotes stromal growth and chronic inflammation [76]. Interestingly, inflammatory pathways, including the cyclooxygenase-2 (COX-2) and NFκB, are over-expressed in human prostate adenocarcinomas compared with normal prostate tissues and targeting these inflammatory pathways have shown promise as an intervention strategy for CaP [75;77]. In vitro EGCG treatment of LNCaP cells with 20-80 μmol/L was associated with a decrease in DNA binding activity of NFκB. Furthermore, EGCG decreased TNFα (a known inducer of NFκB)-induced NFκB activity. In addition, a decrease in the protein levels of the p65 subunit of NFκB in nuclear lysates of LNCaP cells treated with EGCG was observed. This observation indicated that reduced availability of NFκB subunits in the nucleus may be responsible for the decreased transcriptional activity [68]. NFκB can modulate the transcriptional activation of genes associated with cell proliferation, angiogenesis, metastasis, tumor promotion, inflammation and suppression of apoptosis via the regulation of gene expression of genes such as bcl-2, bcl-xl, cIAP, survivin, TRAF, COX-2, MMP-9, iNOS and cell cycle-regulatory components. [78]. NFκB also amplifies inflammatory signals, including COX-2 and pro-inflammatory cytokines, such as IL-6, by acting as a transcriptional activator [77]. Cyclooxygenase is an enzyme involved in the synthesis of prostaglandins from arachidonic acid. Overexpression of COX-2 has been implicated in many pathologic conditions, including cancer. It has been demonstrated that EGCG inhibits COX-2 without affecting COX-1 expression at both the mRNA and protein levels, in androgen-sensitive LNCaP and androgen-insensitive PC-3 human prostate carcinoma cells [79].

8.5 Angiogenesis

Solid tumors cannot grow beyond 2-3 mm in diameter before the diffusion limit of oxygen (100-200 μm) is reached and hypoxia develops. Therefore both the growth and metastasis of tumors are dependent on the formation of new blood vessels (angiogenesis) [80]. A switch to the angiogenic phenotype requires a local change in balance between proangiogenic factors and angiogenic inhibitors [81]. Proangiogenic factors such as VEGF can be stimulated by hypoxia via the hypoxia-responsive transcription factor HIF-1α [82]. Immunohistochemical studies have confirmed the upregulation of HIF-1α and VEGF in areas of human CaP compared with normal prostate and benign prostatic hyperplasia (BPH) [83]. The expression and activation of HIF-1α is tightly regulated by cellular oxygen concentration. Administration of GTPs and EGCG in drinking water to TRAMP mice was associated with tumor inhibition, decreased VEGF and angiogenesis [74]. In cell culture experiments the situation is more complicated. For example Thomas et al. demonstrated that 20-40 μmol of EGCG inhibited prolyl hydroxylation of HIF-1α, thus preventing HIF-1α and pVHL interaction in PC-3 cells. However they also demonstrated a stimulation of Hif-1α protein levels at normoxic condition [84]. This most likely is related to the proteasome inhibitor activity of EGCG. The effect of EGCG on VEGF secretion has been demonstrated in multiple cancer models [85;86]. In stromal fibroblasts derived from primary prostate cancers treatment with 10 μmol/L of EGCG decreased VEGF secretion into the medium [34]. In TRAMP mice consuming 0.1% GTE in drinking water, serum VEGF was decreased significantly at 24 weeks [74]. An effect on serum VEGF was also observed in the human Polyphenon E intervention study of McLarty et al. [34].

8.6 Cancer metastasis

Multiple cellular signaling pathways have been involved in the processes of cancer cell invasion and metastasis. Matrix metalloproteinases (MMP) play a crucial role in the development and metastatic spread of cancer. One of the earliest events in the metastatic spread of cancer is the invasion through the basement membrane and proteolytic degradation of the extracellular matrix proteins, such as collagens, laminin, elastin and fibronectin etc, and non-matrix proteins. MMPs are the important regulators of tumor growth, both at the primary site and in distant metastases [87]. Treatment of DU145 cells with 5-40 μg/L of EGCG resulted in dose-dependent inhibition of induced pro-MMP-2 and pro and active forms of MMP-9 concomitant with marked inhibition of phosphorylation of ERK1/2 and p38 [87]. The HGF/c-Met pathway is another important regulator of signaling pathways implicated in the processes of invasion and metastasis of most human cancers, including CaP. Exposure of DU145 prostate tumor cells to hepatic growth factor (HGF) stimulates the activation of c-Met and downstream PI3-kinase and MAPK pathways, leading to increased scattering, motility, and invasion, which was prevented by the addition of 5 μmol/L of EGCG [88]. The polyphenol epigallocatechin-3-gallate also affects lipid rafts in the cell membrane to block activation of the c-Met receptor in CaP cells. The c-Met inhibitory activity of EGCG may be facilitated through altering lipid raft structures [88].

8.7 Cancer stem cells

Epithelial-mesenchymal transition (EMT) induction in cancer cells results in the acquisition of invasive and metastatic properties [89]. Recent reports indicate that the emergence of cancer stem cells (CSCs) occurs in part as a result of EMT through cues from tumor stromal components [90;91]. Cancer stem cells undergoing metastasis usually express EMT markers (vimentin, slug, snail and β-catenin). A study by Tang et al. in 2010 indicated that human CaP cell lines contain a small population of CD44+CD133+ cancer stem cells and their self-renewal capacity was inhibited by EGCG [92]. Furthermore, EGCG inhibited the self-renewal capacity of CD44+a2b1+CD133+ CSCs isolated from human primary prostate tumors, as measured by spheroid formation in suspension. EGCG induced apoptosis by activating caspase-3/7 and inhibiting the expression of Bcl-2, survivin and XIAP in CSCs. Furthermore, EGCG inhibited EMT by inhibiting the expression of vimentin, slug, snail and nuclear β-catenin, and the activity of the LEF-1/TCF responsive reporter, and also retarded CSC's migration and invasion, suggesting the blockade of signaling involved in early metastasis.

8.8 Androgen receptor

Androgens not only play an important role in the development and function of the prostate but they are also intimately involved in the development and progression of CaP. Within the prostate, testosterone is converted to the more potent androgen dihydrotestosterone (DHT) via the action of the 5α-reductase enzyme. DHT is the primary prostatic androgen and promotes the growth and survival of normal, hyperplastic and malignant prostate tissues. Throughout the different stages of CaP (PIN, localised, recurrent, and metastatic) there is an increase in expression of 5α-reductase particularly in localised high-grade carcinoma [93]. In in vitro experiments it has been demonstrated that EGCG inhibited 5α-reductase with an IC50 of 15 μmol/L [94]. Prostate tumor growth is primarily regulated by androgen binding and transcription signals of the androgen receptor (AR). Androgen-deprivation therapy (ADT), which suppresses the binding of androgens to the androgen receptor (AR), has been the mainstay of treatment for recurrent CaP after primary treatment. Despite suppression of prostate tumor growth, ADT eventually fails, leading to hormone-refractory tumor growth, even though functional AR is often present and even overexpressed in hormone-refractory CaP cells [95]. EGCG suppresses cell proliferation, PSA expression, and AR mRNA and transcriptional activity of AR in androgen dependent and independent LNCaP sublines [96-98]. In addition, in cell culture studies it has been demonstrated that GTE and EGCG, but not EC, inhibited both basal and kinase-stimulated testosterone production in rat Leydig cells [99]. Further in vivo experiments with oral supplementation with GT are needed to confirm the effects of EGCG on hormone production.

9. Conclusion

Strong evidence from in vitro and in vivo animal studies supports the role of GT in CaP prevention. However, prior to assuming that these benefits translate to humans, several points need to be considered. First, as pointed out in this review, most of the in vitro studies do not take the pro-oxidant activity of EGCG at alkaline pH into consideration. With the elimination of the “hydrogen peroxide” effect, much higher concentrations of GTPs would be necessary to induce the same effects. Second, it appears that mouse tissue bioavailability of GTPs is considerably different from human tissue. Therefore the bioavailability in human tissue may be more limited and the potential of GTPs more limited.

Nonetheless, human population studies provide some supportive evidence for a decrease in risk of CaP associated with increased consumption of tea. Evidence is stronger for GT compared to BT. However, for the evidence to be convincing, population studies showing a beneficial effect of tea need to be replicated. The evidence from human clinical trials demonstrating a decrease in the rate of tumor progression from PIN to adenocarcinoma together with evidence of a decrease in serum markers of tumor progression provide strong support for the preventive actions of GT in localized CaP [33;34]. Further intervention studies are needed to demonstrate the effect of brewed tea, which is consumed in amounts relevant to the demonstrated intakes in epidemiological studies. Currently our group is conducting a phase II clinical trial investigating the effects of the consumption of 6 cups of GT or BT prior to prostatectomy on biomarkers of CaP. This trial will be completed in 2011 (Clinical Trial ID: NCI-2010-00973).

Overall, epidemiological and human clinical studies, as well as animal and basic mechanistic studies on GT and BT support a chemopreventive role in CaP with an emphasis on GT, but more research efforts at many levels is needed.

An external file that holds a picture, illustration, etc.
Object name is nihms-318192-f0001.jpg

Chemical structure of methylated and non-methylated EGCG and EGC.

Table 4

Human intervention studies using green tea extracts in localized and advanced prostate cancer.

LocationType of StudyEffectDaily DoseReference
Green tea extract, ItalyIntervention Study (n=60) 1 yearDecreased progression from PIN to CaP600 mg GTE (310.8 mg EGCG)Bettuzzi S et al. 2006 [33]
Green tea extract, Italy1 year follow upSignificantly lower rate of CaPNo interventionBrausi M et al. 2008 [37]
Polyphenon E, USIntervention Study (n=25)Decreased serum PSA, HGF, VEGF1300 mg Polyphenon E (800 mg EGCG)McLarty J et al. 2009 [34]
Green tea extract, CanadaIntervention study (n=19) in hormone refractory CaPNo effect500 mg GTE (112.5 mg EGCG)Choan E et al. 2005 [35]
Green tea powder, USAIntervention study (n=42) in hormone refractory CaPPSA decreased by 50% in one out of 42 subjects, no effect, Grade 1 and 2 toxicity in 69% and no side effects in 31% of participants6 g instant green tea powder in 6x 1g doses in hot waterJatoi A et al. 2003 [36]

Footnotes

10 Conflict of Interest

There are no conflicts of interest to disclose for any of the authors.

9 References

1. Sharma V, Rao LJ. A thought on the biological activities of black tea. Crit Rev.Food Sci.Nutr. 2009;49:379–404. [PubMed] [Google Scholar]
2. Jemal A, Center MM, DeSantis C, Ward EM. Global patterns of cancer incidence and mortality rates and trends. Cancer Epidemiol.Biomarkers Prev. 2010;19:1893–1907. [PubMed] [Google Scholar]
3. Baade PD, Youlden DR, Krnjacki LJ. International epidemiology of prostate cancer: geographical distribution and secular trends. Mol.Nutr.Food Res. 2009;53:171–184. [PubMed] [Google Scholar]
4. Kolonel LN, Altshuler D, Henderson BE. The multiethnic cohort study: exploring genes, lifestyle and cancer risk. Nat.Rev.Cancer. 2004;4:519–527. [PubMed] [Google Scholar]
5. Namiki M, Akaza H, Lee SE, Song JM, Umbas R, Zhou L, Lee BC, Cheng C, Chung MK, Fukagai T, Hinotsu S, Horie S. Prostate Cancer Working Group report. Jpn.J.Clin.Oncol. 2010;40(Suppl 1):i70–i75. [PubMed] [Google Scholar]
6. Jayachandran J, Freedland SJ. Prevention of prostate cancer: what we know and where we are going. Am.J.Mens.Health. 2008;2:178–189. [PubMed] [Google Scholar]
7. Siddiqui IA, Saleem M, Adhami VM, Asim M, Mukhtar H. Tea beverage in chemoprevention and chemotherapy of prostate cancer. Acta Pharmacol.Sin. 2007;28:1392–1408. [PubMed] [Google Scholar]
8. Johnson JJ, Bailey HH, Mukhtar H. Green tea polyphenols for prostate cancer chemoprevention: a translational perspective. Phytomedicine. 2010;17:3–13. [PMC free article] [PubMed] [Google Scholar]
9. Lee AH, Fraser ML, Binns CW. Tea, coffee and prostate cancer. Mol.Nutr.Food Res. 2009;53:256–265. [PubMed] [Google Scholar]
10. Adhami VM, Ahmad N, Mukhtar H. Molecular targets for green tea in prostate cancer prevention. J Nutr. 2003;133:2417S–2424S. [PubMed] [Google Scholar]
11. Siddiqui IA, Zaman N, Aziz MH, Reagan-Shaw SR, Sarfaraz S, Adhami VM, Ahmad N, Raisuddin S, Mukhtar H. Inhibition of CWR22Rnu1 tumor growth and PSA secretion in athymic nude mice by green and black teas. Carcinogenesis. 2006;27:833–839. [PubMed] [Google Scholar]
12. Siddiqui IA, Raisuddin S, Shukla Y. Protective effects of black tea extract on testosterone induced oxidative damage in prostate. Cancer Lett. 2005;227:125–132. [PubMed] [Google Scholar]
13. Adhami VM, Siddiqui IA, Sarfaraz S, Khwaja SI, Hafeez BB, Ahmad N, Mukhtar H. Effective prostate cancer chemopreventive intervention with green tea polyphenols in the TRAMP model depends on the stage of the disease. Clin.Cancer Res. 2009;15:1947–1953. [PMC free article] [PubMed] [Google Scholar]
14. Harper CE, Patel BB, Wang J, Eltoum IA, Lamartiniere CA. Epigallocatechin-3-Gallate suppresses early stage, but not late stage prostate cancer in TRAMP mice: mechanisms of action. Prostate. 2007;67:1576–1589. [PubMed] [Google Scholar]
15. Zhou JR, Yu L, Zhong Y, Blackburn GL. Soy phytochemicals and tea bioactive components synergistically inhibit androgen-sensitive human prostate tumors in mice. J Nutr. 2003;133:516–521. [PMC free article] [PubMed] [Google Scholar]
16. Sartor L, Pezzato E, Dona M, Dell'Aica I, Calabrese F, Morini M, Albini A, Garbisa S. Prostate carcinoma and green tea: (-)epigallocatechin-3-gallate inhibits inflammation-triggered MMP-2 activation and invasion in murine TRAMP model. Int.J Cancer. 2004;112:823–829. [PubMed] [Google Scholar]
17. Greenberg NM, DeMayo F, Finegold MJ, Medina D, Tilley WD, Aspinall JO, Cunha GR, Donjacour AA, Matusik RJ, Rosen JM. Prostate cancer in a transgenic mouse. Proc.Natl.Acad.Sci.U.S.A. 1995;92:3439–3443. [PMC free article] [PubMed] [Google Scholar]
18. Siddiqui IA, Shukla Y, Adhami VM, Sarfaraz S, Asim M, Hafeez BB, Mukhtar H. Suppression of NFkappaB and its regulated gene products by oral administration of green tea polyphenols in an autochthonous mouse prostate cancer model. Pharm.Res. 2008;25:2135–2142. [PMC free article] [PubMed] [Google Scholar]
19. Teichert F, Verschoyle RD, Greaves P, Thorpe JF, Mellon JK, Steward WP, Farmer PB, Gescher AJ, Singh R. Determination of 8-oxo-2'-deoxyguanosine and creatinine in murine and human urine by liquid chromatography/tandem mass spectrometry: application to chemoprevention studies. Rapid Commun.Mass Spectrom. 2009;23:258–266. [PubMed] [Google Scholar]
20. Morey K, Sr., Zhang W, Pascual M, Greally JM, Gillard BM, Karasik E, Foster BA, Karpf AR. Lack of evidence for green tea polyphenols as DNA methylation inhibitors in murine prostate. Cancer Prev.Res.(Phila) 2009;2:1065–1075. [PMC free article] [PubMed] [Google Scholar]
21. O'Sullivan J, Sheridan J, Mulcahy H, Tenniswood M, Morrissey C. The effect of green tea on oxidative damage and tumour formation in Lobund-Wistar rats. Eur.J.Cancer Prev. 2008;17:489–501. [PMC free article] [PubMed] [Google Scholar]
22. Kurahashi N, Sasazuki S, Iwasaki M, Inoue M, Tsugane S. Green tea consumption and prostate cancer risk in Japanese men: a prospective study. Am.J.Epidemiol. 2008;167:71–77. [PubMed] [Google Scholar]
23. Heilbrun LK, Nomura A, Stemmermann GN. Black tea consumption and cancer risk: a prospective study. Br.J.Cancer. 1986;54:677–683. [PMC free article] [PubMed] [Google Scholar]
24. Jian L, Xie LP, Lee AH, Binns CW. Protective effect of green tea against prostate cancer: a case-control study in southeast China. Int.J.Cancer. 2004;108:130–135. [PubMed] [Google Scholar]
25. Sonoda T, Nagata Y, Mori M, Miyanaga N, Takashima N, Okumura K, Goto K, Naito S, Fujimoto K, Hirao Y, Takahashi A, Tsukamoto T, Fujioka T, Akaza H. A case-control study of diet and prostate cancer in Japan: possible protective effect of traditional Japanese diet. Cancer Sci. 2004;95:238–242. [PMC free article] [PubMed] [Google Scholar]
26. La Vecchia C, Negri E, Franceschi S, D'Avanzo B, Boyle P. Tea consumption and cancer risk. Nutr.Cancer. 1992;17:27–31. [PubMed] [Google Scholar]
27. Villeneuve PJ, Johnson KC, Kreiger N, Mao Y. Risk factors for prostate cancer: results from the Canadian National Enhanced Cancer Surveillance System. The Canadian Cancer Registries Epidemiology Research Group. Cancer Causes Control. 1999;10:355–367. [PubMed] [Google Scholar]
28. Kikuchi N, Ohmori K, Shimazu T, Nakaya N, Kuriyama S, Nishino Y, Tsubono Y, Tsuji I. No association between green tea and prostate cancer risk in Japanese men: the Ohsaki Cohort Study. Br.J.Cancer. 2006;95:371–373. [PMC free article] [PubMed] [Google Scholar]
29. Severson RK, Nomura AM, Grove JS, Stemmermann GN. A prospective study of demographics, diet, and prostate cancer among men of Japanese ancestry in Hawaii. Cancer Res. 1989;49:1857–1860. [PubMed] [Google Scholar]
30. Allen NE, Sauvaget C, Roddam AW, Appleby P, Nagano J, Suzuki G, Key TJ, Koyama K. A prospective study of diet and prostate cancer in Japanese men. Cancer Causes Control. 2004;15:911–920. [PubMed] [Google Scholar]
31. Jain MG, Hislop GT, Howe GR, Burch JD, Ghadirian P. Alcohol and other beverage use and prostate cancer risk among Canadian men. Int.J.Cancer. 1998;78:707–711. [PubMed] [Google Scholar]
32. Henning SM, Fajardo-Lira C, Lee HW, Youssefian AA, Go VLW, Heber D. Catechin content of 18 teas and a green tea extract supplement correlated with the antioxidant capacity. Nutr.Cancer. 2003;45:226–235. [PubMed] [Google Scholar]
33. Bettuzzi S, Brausi M, Rizzi F, Castagnetti G, Peracchia G, Corti A. Chemoprevention of human prostate cancer by oral administration of green tea catechins in volunteers with high-grade prostate intraepithelial neoplasia: a preliminary report from a one-year proof-of-principle study. Cancer Res. 2006;66:1234–1240. [PubMed] [Google Scholar]
34. McLarty J, Bigelow RL, Smith M, Elmajian D, Ankem M, Cardelli JA. Tea Polyphenols Decrease Serum Levels of Prostate-Specific Antigen, Hepatocyte Growth Factor, and Vascular Endothelial Growth Factor in Prostate Cancer Patients and Inhibit Production of Hepatocyte Growth Factor and Vascular Endothelial Growth Factor In vitro. Cancer Prev.Res.(Phila Pa) 2009;2:673–682. [PubMed] [Google Scholar]
35. Choan E, Segal R, Jonker D, Malone S, Reaume N, Eapen L, Gallant V. A prospective clinical trial of green tea for hormone refractory prostate cancer: an evaluation of the complementary/alternative therapy approach. Urol.Oncol. 2005;23:108–113. [PubMed] [Google Scholar]
36. Jatoi A, Ellison N, Burch PA, Sloan JA, Dakhil SR, Novotny P, Tan W, Fitch TR, Rowland KM, Young CY, Flynn PJ. A phase II trial of green tea in the treatment of patients with androgen independent metastatic prostate carcinoma. Cancer. 2003;97:1442–1446. [PubMed] [Google Scholar]
37. Brausi M, Rizzi F, Bettuzzi S. Chemoprevention of human prostate cancer by green tea catechins: two years later. A follow-up update. Eur.Urol. 2008;54:472–473. [PubMed] [Google Scholar]
38. Lambert JD, Sang S, Lu AY, Yang CS. Metabolism of dietary polyphenols and possible interactions with drugs. Curr.Drug Metab. 2007;8:499–507. [PubMed] [Google Scholar]
39. Lu H, Meng X, Yang CS. Enzymology of methylation of tea catechins and inhibition of catechol-O-methyltransferase by (-)-epigallocatechin gallate. Drug Metab Dispos. 2003;31:572–579. [PubMed] [Google Scholar]
40. Henning SM, Choo JJ, Heber D. Nongallated compared with gallated flavan-3-ols in green and black tea are more bioavailable. J.Nutr. 2008;138:1529S–1534S. [PMC free article] [PubMed] [Google Scholar]
41. Meng X, Lee MJ, Li C, Sheng S, Zhu N, Sang S, Ho CT, Yang CS. Formation and identification of 4'-O-methyl-(-)-epigallocatechin in humans. Drug Metab Dispos. 2001;29:789–793. [PubMed] [Google Scholar]
42. Wang P, Aronson WJ, Huang M, Zhang Y, Lee RP, Heber D, Henning SM. Green tea polyphenols and metabolites in prostatectomy tissue: implications for cancer prevention. Cancer Prev.Res.(Phila) 2010;3:985–993. [PMC free article] [PubMed] [Google Scholar]
43. Henning SM, Aronson W, Niu Y, Conde F, Lee NH, Seeram NP, Lee RP, Lu J, Harris DM, Moro A, Hong J, Leun PS, Barnard RJ, Ziaee HG, Csathy G, Go VLW, Wang H, Heber D. Tea polyphenols and theaflavins are found in prostate tissue of humans and mice after green and black tea consumption. J Nutr. 2006;136:1839–143. [PubMed] [Google Scholar]
44. Landis-Piwowar KR, Wan SB, Wiegand RA, Kuhn DJ, Chan TH, Dou QP. Methylation suppresses the proteasome-inhibitory function of green tea polyphenols. J.Cell Physiol. 2007;213:252–260. [PubMed] [Google Scholar]
45. Lambert JD, Lee MJ, Lu H, Meng X, Hong JJ, Seril DN, Sturgill MG, Yang CS. Epigallocatechin-3-gallate is absorbed but extensively glucuronidated following oral administration to mice. J Nutr. 2003;133:4172–4177. [PubMed] [Google Scholar]
46. Meng X, Sang S, Zhu N, Lu H, Sheng S, Lee MJ, Ho CT, Yang CS. Identification and Characterization of Methylated and Ring-Fission Metabolites of Tea Catechins Formed in Humans, Mice, and Rats. Chem.Res.Toxicol. 2002;15:1042–1050. [PubMed] [Google Scholar]
47. Chen D, Wang CY, Lambert JD, Ai N, Welsh WJ, Yang CS. Inhibition of human liver catechol-O-methyltransferase by tea catechins and their metabolites: structure-activity relationship and molecular-modeling studies. Biochem.Pharmacol. 2005;69:1523–1531. [PubMed] [Google Scholar]
48. Fang M, Chen D, Yang CS. Dietary polyphenols may affect DNA methylation. J.Nutr. 2007;137:223S–228S. [PubMed] [Google Scholar]
49. Yegnasubramanian S, Kowalski J, Gonzalgo ML, Zahurak M, Piantadosi S, Walsh PC, Bova GS, De Marzo AM, Isaacs WB, Nelson WG. Hypermethylation of CpG islands in primary and metastatic human prostate cancer. Cancer Res. 2004;64:1975–1986. [PubMed] [Google Scholar]
50. Nakayama M, Gonzalgo ML, Yegnasubramanian S, Lin X, De Marzo AM, Nelson WG. GSTP1 CpG island hypermethylation as a molecular biomarker for prostate cancer. J.Cell Biochem. 2004;91:540–552. [PubMed] [Google Scholar]
51. Gao K, Xu A, Krul C, Venema K, Liu Y, Niu Y, Lu J, Bensoussan L, Seeram NP, Heber D, Henning SM. Of the major phenolic acids formed during human microbial fermentation of tea, citrus, and soy flavonoid supplements, only 3,4-dihydroxyphenylacetic acid has antiproliferative activity. J Nutr. 2006;136:52–57. [PubMed] [Google Scholar]
52. Mulder TP, van Platerink CJ, Wijnand Schuyl PJ, Van Amelsvoort JM. Analysis of theaflavins in biological fluids using liquid chromatography-electrospray mass spectrometry. J.Chromatogr.B Biomed.Sci.Appl. 2001;760:271–279. [PubMed] [Google Scholar]
53. van het Hof KH, Kivits GA, Weststrate JA, Tijburg LB. Bioavailability of catechins from tea: the effect of milk. Eur.J.Clin.Nutr. 1998;52:356–359. [PubMed] [Google Scholar]
54. Kyle JA, Morrice PC, McNeill G, Duthie GG. Effects of infusion time and addition of milk on content and absorption of polyphenols from black tea. J.Agric.Food Chem. 2007;55:4889–4894. [PubMed] [Google Scholar]
55. Reddy VC, Vidya Sagar GV, Sreeramulu D, Venu L, Raghunath M. Addition of milk does not alter the antioxidant activity of black tea. Ann.Nutr.Metab. 2005;49:189–195. [PubMed] [Google Scholar]
56. Ryan L, Petit S. Addition of whole, semiskimmed, and skimmed bovine milk reduces the total antioxidant capacity of black tea. Nutr.Res. 2010;30:14–20. [PubMed] [Google Scholar]
57. Neilson AP, Song BJ, Sapper TN, Bomser JA, Ferruzzi MG. Tea catechin auto-oxidation dimers are accumulated and retained by Caco-2 human intestinal cells. Nutr.Res. 2010;30:327–340. [PMC free article] [PubMed] [Google Scholar]
58. Neilson AP, Hopf AS, Cooper BR, Pereira MA, Bomser JA, Ferruzzi MG. Catechin Degradation with Concurrent Formation of Homo- and Heterocatechin Dimers during in Vitro Digestion. J.Agric.Food Chem. 2007;55:8941–8949. [PubMed] [Google Scholar]
59. Hong J, Lu H, Meng X, Ryu JH, Hara Y, Yang CS. Stability, cellular uptake, biotransformation, and efflux of tea polyphenol (-)-epigallocatechin-3-gallate in HT-29 human colon adenocarcinoma cells. Cancer Res. 2002;62:7241–7246. [PubMed] [Google Scholar]
60. Lambert JD, Elias RJ. The antioxidant and pro-oxidant activities of green tea polyphenols: a role in cancer prevention. Arch.Biochem.Biophys. 2010;501:65–72. [PMC free article] [PubMed] [Google Scholar]
61. Na HK, Kim EH, Jung JH, Lee HH, Hyun JW, Surh YJ. (-)-Epigallocatechin gallate induces Nrf2-mediated antioxidant enzyme expression via activation of PI3K and ERK in human mammary epithelial cells. Arch.Biochem.Biophys. 2008;476:171–177. [PubMed] [Google Scholar]
62. Kobayashi A, Kang MI, Watai Y, Tong KI, Shibata T, Uchida K, Yamamoto M. Oxidative and electrophilic stresses activate Nrf2 through inhibition of ubiquitination activity of Keap1. Mol.Cell Biol. 2006;26:221–229. [PMC free article] [PubMed] [Google Scholar]
63. Surh YJ, Na HK. NF-kappaB and Nrf2 as prime molecular targets for chemoprevention and cytoprotection with anti-inflammatory and antioxidant phytochemicals. Genes Nutr. 2008;2:313–317. [PMC free article] [PubMed] [Google Scholar]
64. Bostwick DG, Meiers I, Shanks JH. Glutathione S-transferase: differential expression of alpha, mu, and pi isoenzymes in benign prostate, prostatic intraepithelial neoplasia, and prostatic adenocarcinoma. Hum.Pathol. 2007;38:1394–1401. [PubMed] [Google Scholar]
65. Yu S, Khor TO, Cheung KL, Li W, Wu TY, Huang Y, Foster BA, Kan YW, Kong AN. Nrf2 expression is regulated by epigenetic mechanisms in prostate cancer of TRAMP mice. PLoS.One. 2010;5:e8579. [PMC free article] [PubMed] [Google Scholar]
66. Frohlich DA, McCabe MT, Arnold RS, Day ML. The role of Nrf2 in increased reactive oxygen species and DNA damage in prostate tumorigenesis. Oncogene. 2008;27:4353–4362. [PubMed] [Google Scholar]
67. Zhang ZM, Yang XY, Yuan JH, Sun ZY, Li YQ. Modulation of NRF2 and UGT1A expression by epigallocatechin-3-gallate in colon cancer cells and BALB/c mice. Chin Med.J.(Engl.) 2009;122:1660–1665. [PubMed] [Google Scholar]
68. Hastak K, Gupta S, Ahmad N, Agarwal MK, Agarwal ML, Mukhtar H. Role of p53 and NF-kappaB in epigallocatechin-3-gallate-induced apoptosis of LNCaP cells. Oncogene. 2003;22:4851–4859. [PubMed] [Google Scholar]
69. Chung LY, Cheung TC, Kong SK, Fung KP, Choy YM, Chan ZY, Kwok TT. Induction of apoptosis by green tea catechins in human prostate cancer DU145 cells. Life Sci. 2001;68:1207–1214. [PubMed] [Google Scholar]
70. Gupta S, Ahmad N, Nieminen AL, Mukhtar H. Growth inhibition, cell-cycle dysregulation, and induction of apoptosis by green tea constituent (-)-epigallocatechin-3-gallate in androgen-sensitive and androgen-insensitive human prostate carcinoma cells. Toxicol.Appl.Pharmacol. 2000;164:82–90. [PubMed] [Google Scholar]
71. Rosenzweig SA, Atreya HS. Defining the pathway to insulin-like growth factor system targeting in cancer. Biochem.Pharmacol. 2010;80:1115–1124. [PMC free article] [PubMed] [Google Scholar]
72. Kojima S, Inahara M, Suzuki H, Ichikawa T, Furuya Y. Implications of insulin-like growth factor-I for prostate cancer therapies. Int.J.Urol. 2009;16:161–167. [PubMed] [Google Scholar]
73. Li M, He Z, Ermakova S, Zheng D, Tang F, Cho YY, Zhu F, Ma WY, Sham Y, Rogozin EA, Bode AM, Cao Y, Dong Z. Direct inhibition of insulin-like growth factor-I receptor kinase activity by (-)-epigallocatechin-3-gallate regulates cell transformation. Cancer Epidemiol.Biomarkers Prev. 2007;16:598–605. [PubMed] [Google Scholar]
74. Adhami VM, Siddiqui IA, Ahmad N, Gupta S, Mukhtar H. Oral consumption of green tea polyphenols inhibits insulin-like growth factor-I-induced signaling in an autochthonous mouse model of prostate cancer. Cancer Res. 2004;64:8715–8722. [PubMed] [Google Scholar]
75. Palapattu GS, Sutcliffe S, Bastian PJ, Platz EA, De Marzo AM, Isaacs WB, Nelson WG. Prostate carcinogenesis and inflammation: emerging insights. Carcinogenesis. 2005;26:1170–1181. [PubMed] [Google Scholar]
76. Kramer G, Mitteregger D, Marberger M. Is benign prostatic hyperplasia (BPH) an immune inflammatory disease? Eur.Urol. 2007;51:1202–1216. [PubMed] [Google Scholar]
77. Hsu A, Bray TM, Ho E. Anti-inflammatory activity of soy and tea in prostate cancer prevention. Exp.Biol.Med.(Maywood.) 2010;235:659–667. [PMC free article] [PubMed] [Google Scholar]
78. Luqman S, Pezzuto JM. NFkappaB: a promising target for natural products in cancer chemoprevention. Phytother.Res. 2010;24:949–963. [PubMed] [Google Scholar]
79. Hussain T, Gupta S, Adhami VM, Mukhtar H. Green tea constituent epigallocatechin-3-gallate selectively inhibits COX-2 without affecting COX-1 expression in human prostate carcinoma cells. Int.J.Cancer. 2005;113:660–669. [PubMed] [Google Scholar]
80. Daniel KG, Kuhn DJ, Kazi A, Dou QP. Anti-angiogenic and anti-tumor properties of proteasome inhibitors. Curr.Cancer Drug Targets. 2005;5:529–541. [PubMed] [Google Scholar]
81. Huss WJ, Hanrahan CF, Barrios RJ, Simons JW, Greenberg NM. Angiogenesis and prostate cancer: identification of a molecular progression switch. Cancer Res. 2001;61:2736–2743. [PubMed] [Google Scholar]
82. Zhang Q, Tang X, Lu Q, Zhang Z, Rao J, Le AD. Green tea extract and (-)-epigallocatechin-3-gallate inhibit hypoxia- and serum-induced HIF-1alpha protein accumulation and VEGF expression in human cervical carcinoma and hepatoma cells. Mol.Cancer Ther. 2006;5:1227–1238. [PubMed] [Google Scholar]
83. Boddy JL, Fox SB, Han C, Campo L, Turley H, Kanga S, Malone PR, Harris AL. The androgen receptor is significantly associated with vascular endothelial growth factor and hypoxia sensing via hypoxia-inducible factors HIF-1a, HIF-2a, and the prolyl hydroxylases in human prostate cancer. Clin.Cancer Res. 2005;11:7658–7663. [PubMed] [Google Scholar]
84. Thomas R, Kim MH. Epigallocatechin gallate inhibits HIF-1alpha degradation in prostate cancer cells. Biochem.Biophys.Res.Commun. 2005;334:543–548. [PubMed] [Google Scholar]
85. Sartippour MR, Shao ZM, Heber D, Beatty P, Zhang L, Liu C, Ellis L, Liu W, Go VL, Brooks MN. Green tea inhibits vascular endothelial growth factor (VEGF) induction in human breast cancer cells. J Nutr. 2002;132:2307–2311. [PubMed] [Google Scholar]
86. Shimizu M, Shirakami Y, Sakai H, Yasuda Y, Kubota M, Adachi S, Tsurumi H, Hara Y, Moriwaki H. (-)-Epigallocatechin gallate inhibits growth and activation of the VEGF/VEGFR axis in human colorectal cancer cells. Chem.Biol.Interact. 2010;185:247–252. [PubMed] [Google Scholar]
87. Katiyar SK. Matrix metalloproteinases in cancer metastasis: molecular targets for prostate cancer prevention by green tea polyphenols and grape seed proanthocyanidins. Endocr.Metab Immune.Disord.Drug Targets. 2006;6:17–24. [PubMed] [Google Scholar]
88. Duhon D, Bigelow RL, Coleman DT, Steffan JJ, Yu C, Langston W, Kevil CG, Cardelli JA. The polyphenol epigallocatechin-3-gallate affects lipid rafts to block activation of the c-Met receptor in prostate cancer cells. Mol.Carcinog. 2010;49:739–749. [PubMed] [Google Scholar]
89. Singh A, Settleman J. EMT, cancer stem cells and drug resistance: an emerging axis of evil in the war on cancer. Oncogene. 2010;29:4741–4751. [PMC free article] [PubMed] [Google Scholar]
90. Gupta PB, Chaffer CL, Weinberg RA. Cancer stem cells: mirage or reality? Nat.Med. 2009;15:1010–1012. [PubMed] [Google Scholar]
91. Lawson DA, Witte ON. Stem cells in prostate cancer initiation and progression. J.Clin.Invest. 2007;117:2044–2050. [PMC free article] [PubMed] [Google Scholar]
92. Tang SN, Singh C, Nall D, Meeker D, Shankar S, Srivastava RK. The dietary bioflavonoid quercetin synergizes with epigallocathechin gallate (EGCG) to inhibit prostate cancer stem cell characteristics, invasion, migration and epithelialmesenchymal transition. J.Mol.Signal. 2010;5:14. [PMC free article] [PubMed] [Google Scholar]
93. Rosenberg MT, Froehner M, Albala D, Miner MM. Biology and natural history of prostate cancer and the role of chemoprevention. Int.J.Clin.Pract. 2010;64:1746–1753. [PubMed] [Google Scholar]
94. Hiipakka RA, Zhang HZ, Dai W, Dai Q, Liao S. Structure-activity relationships for inhibition of human 5alpha-reductases by polyphenols. Biochem.Pharmacol. 2002;63:1165–1176. [PubMed] [Google Scholar]
95. Niu Y, Chang TM, Yeh S, Ma WL, Wang YZ, Chang C. Differential androgen receptor signals in different cells explain why androgen-deprivation therapy of prostate cancer fails. Oncogene. 2010;29:3593–3604. [PubMed] [Google Scholar]
96. Chuu CP, Chen RY, Kokontis JM, Hiipakka RA, Liao S. Suppression of androgen receptor signaling and prostate specific antigen expression by (-)-epigallocatechin-3-gallate in different progression stages of LNCaP prostate cancer cells. Cancer Lett. 2009;275:86–92. [PMC free article] [PubMed] [Google Scholar]
97. Pezzato E, Sartor L, Dell'Aica I, Dittadi R, Gion M, Belluco C, Lise M, Garbisa S. Prostate carcinoma and green tea: PSA-triggered basement membrane degradation and MMP-2 activation are inhibited by (-)epigallocatechin-3-gallate. Int.J Cancer. 2004;112:787–792. [PubMed] [Google Scholar]
98. Ren F, Zhang S, Mitchell SH, Butler R, Young CY. Tea polyphenols down-regulate the expression of the androgen receptor in LNCaP prostate cancer cells. Oncogene. 2000;19:1924–1932. [PubMed] [Google Scholar]
99. Figueiroa MS, Cesar Vieira JS, Leite DS, Filho RC, Ferreira F, Gouveia PS, Udrisar DP, Wanderley MI. Green tea polyphenols inhibit testosterone production in rat Leydig cells. Asian J.Androl. 2009;11:362–370. [PMC free article] [PubMed] [Google Scholar]
100. Lee SC, Chan WK, Lee TW, Lam WH, Wang X, Chan TH, Wong YC. Effect of a prodrug of the green tea polyphenol (-)-epigallocatechin-3-gallate on the growth of androgen-independent prostate cancer in vivo. Nutr.Cancer. 2008;60:483–491. [PubMed] [Google Scholar]
101. Gupta S, Hastak K, Ahmad N, Lewin JS, Mukhtar H. Inhibition of prostate carcinogenesis in TRAMP mice by oral infusion of green tea polyphenols. Proc.Natl.Acad.Sci.U.S.A. 2001;98:10350–10355. [PMC free article] [PubMed] [Google Scholar]
102. Scaltriti M, Belloni L, Caporali A, Davalli P, Remondini D, Rizzi F, Astancolle S, Corti A, Bettuzzi S. Molecular classification of green tea catechin-sensitive and green tea catechin-resistant prostate cancer in the TRAMP mice model by quantitative real-time PCR gene profiling. Carcinogenesis. 2006;27:1047–1053. [PubMed] [Google Scholar]
103. McCarthy S, Caporali A, Enkemann S, Scaltriti M, Eschrich S, Davalli P, Corti A, Lee A, Sung J, Yeatman TJ, Bettuzzi S. Green tea catechins suppress the DNA synthesis marker MCM7 in the TRAMP model of prostate cancer. Mol.Oncol. 2007;1:196–204. [PMC free article] [PubMed] [Google Scholar]
104. Hsu A, Bruno RS, Lohr CV, Taylor AW, Dashwood RH, Bray TM, Ho E. Dietary soy and tea mitigate chronic inflammation and prostate cancer via NFkappaB pathway in the Noble rat model. J.Nutr.Biochem. 2010 [Epub ahead of print] [PMC free article] [PubMed] [Google Scholar]
105. Ellison LF. Tea and other beverage consumption and prostate cancer risk: a Canadian retrospective cohort study. Eur.J.Cancer Prev. 2000;9:125–130. [PubMed] [Google Scholar]
106. Kinlen LJ, Willows AN, Goldblatt P, Yudkin J. Tea consumption and cancer. Br.J.Cancer. 1988;58:397–401. [PMC free article] [PubMed] [Google Scholar]
107. Henning SM, Niu Y, Lee NH, Thames GD, Minutti RR, Wang H, Go VL, Heber D. Bioavailability and antioxidant activity of tea flavanols after consumption of green tea, black tea, or a green tea extract supplement. Am.J Clin Nutr. 2004;80:1558–1564. [PubMed] [Google Scholar]
-