Current Application of Selective COX-2 Inhibitors in Cancer Prevention and Treatment

Publication
Article
OncologyONCOLOGY Vol 16 No 5
Volume 16
Issue 5

The multistep process of carcinogenesis, which can take many years, provides many opportunities for intervention to inhibit disease progression. Effective chemoprevention agents may reduce the risk of cancer by inhibiting the initiation stage of carcinoma through induction of apoptosis or DNA repair in cells harboring mutations, or they may act to prevent promotion of tumor growth. Similarly, chemoprevention may entail blocking cancer progression to an invasive phenotype.

ABSTRACT: The multistep process of carcinogenesis, which can take many years, provides many opportunities for intervention to inhibit disease progression. Effective chemoprevention agents may reduce the risk of cancer by inhibiting the initiation stage of carcinoma through induction of apoptosis or DNA repair in cells harboring mutations, or they may act to prevent promotion of tumor growth. Similarly, chemoprevention may entail blocking cancer progression to an invasive phenotype. Over the past decade, in vitro, preclinical, and clinical data have supported the hypothesis that cyclooxygenase (COX)-2 plays a central role in oncogenesis and that treatment with COX-2 inhibitors offers an effective chemoprevention strategy, as exemplified by the activity of celecoxib (Celebrex) in familial adenomatous polyposis. These COX-2 data have contributed to initiation of clinical trials testing COX-2 inhibitors for the chemoprevention of a wide variety of cancers that overexpress COX-2. [ONCOLOGY 16(Suppl 4):37-51, 2002]

Cyclooxygenases (COXs) are enzymes thatcatalyze the rate-limiting step in the conversion of arachidonic acid toprostaglandins (Figure 1).[1-3] Prostaglandins, along with other arachidonicacid products such as thromboxane and 15-hydroxy-eicosatetraenoic acids, belongto the eicosanoid family of fatty acid molecules, which are known to regulatemany physiologic processes including the inflammatory response and other immuneresponse modulators,[4-6] ovulation,[7,8] and mitogenesis.[9,10] Paradoxically,prostaglandins also have been shown to have anti-inflammatory andimmunosuppressive effects. Studies conducted by Gualde and colleaguesdemonstrated that prostaglandins inhibited T-cell proliferation in vitro.[11]Furthermore, prostaglandins can block production of cytokines by Tlymphocytes.[12]

Synthesis of prostaglandins can be regulated at severaldifferent points in the pathway (Figure 1). In the first step, membranephospholipid is converted to arachidonic acid via phospholipase A2.Subsequently, arachidonic acid is converted to prostaglandin H2 through atwo-step process that involves COX activity to convert arachidonic acid toprostaglandin G2, followed by a peroxidase reaction that is also catalyzed byCOX to produce prostaglandin H2.[13-15]

The COX enzyme family comprises two known isoforms, COX-1 andCOX-2. Cyclooxygenase-1 is a membrane-bound hemoglycoprotein that isconstitutively expressed in the endoplasmic reticulum of cells in most healthytissues and is responsible for local prostaglandin synthesis. In contrast, COX-2is primarily an inducible COX isoform, although low basal expression is apparentin some tissues, including brain and kidney.[16,17] There are a number ofstructural differences between the COX-1 and COX-2 genes, including differencesin the cis elements within the promoter regions and 3´-untranslated domains.

The structure of the COX-2 gene suggests that it is animmediate, early gene product that can be switched on rapidly during theinflammatory response.[18,19] Cyclooxygenase-2 synthesis is inducible by avariety of stimuli, including proinflammatory cytokines such as interleukin-1alpha and -1 beta,[20,21] growth factors such as platelet-derived growthfactor[22,23] and epidermal growth factor,[24,25] and lipopolysaccharide andendothelin.[26,27]

COX-2 Inhibitors and Cancer

Most nonsteroidal anti-inflammatory drugs (NSAIDs) that arecommonly administered to patients inhibit both COX-1 and COX-2. However,inhibition of the inducible isoform, COX-2, is the primary anti-inflammatorymechanism.[5,28,29] Adverse effects associated with long-term use of NSAIDs,including gastritis and gastrointestinal ulceration, in addition to reversibleliver and kidney dysfunction, are thought to be primarily due to inhibition ofthe constitutively expressed COX-1 isoform.[30-32] In recent years, COX-2-specificNSAIDs, including celecoxib (Celebrex) and rofecoxib (Vioxx), have becomeavailable. Selective COX-2 inhibitors are advantageous because they may inhibitpain and the inflammation process in arthritis and oncogenesis. However, they donot inhibit COX-1 enzymes, the products of the "housekeeping genes"required for the maintenance of the gastrointestinal tract and for normal renaland hepatic function.

Rigas and colleagues have demonstrated that colorectal adenomasand adenocarcinomas express elevated levels of prostaglandins.[33] Furthermore,accumulation of prostaglandins is associated with increased expression of COX-2,but not of COX-1.[34] It is also known that prostanoid levels increase duringthe progression from adenoma to adenocarcinoma in patients with familialadenomatous polyposis.[35] In addition, elevated prostanoid expression isassociated with tumor growth, metastatic potential,[36] disease stage,[37]recurrence,[38] and survival[39] in a broad spectrum of tumor types.Furthermore, overexpression of COX-2 in humans has been documented in manycancer types and neoplastic precursor lesions (Table1).[40-81]

These data indicate that selective inhibition of COX-2 may be aneffective strategy for preventing colorectal cancer and also may haveapplication in other cancers. Furthermore, because COX-2 overexpression has beenobserved in both preneoplastic lesions and cancers, chemoprevention interventionis possible at multiple stages of carcinogenesis.

Overexpression of COX-2 may affect a broad range of mechanismsimplicated in the process of carcinogenesis, including angiogenesis, apoptosis,and immune function. Cancer prevention offers more than one opportunity toinhibit disease growth. Effective chemopreventive agents may reduce the risk ofcancer by preventing the initiation stage of carcinoma by inducing apoptosis orDNA repair in cells harboring mutations, or they may act to prevent tumor growthduring the promotion and progression stages of carcinogenesis (Figure2).Ongoing clinical trials evaluating COX-nonspecific and COX-2-specificinhibitors as chemoprevention and therapeutic agents are shown in Table 2[82-84]and are discussed in the following sections.

Colorectal Cancer

Colorectal cancer is a major national health problem. It is thethird leading cause of cancer death in the United States, with 2002 estimates of148,300 new cases and 56,600 deaths.[85] Some (approximately 15%) individualswho develop colorectal carcinoma belong to clinically identifiable high-riskgroups due to familial adenomatous polyposis and hereditary nonpolyposissyndromes.[86] However, the majority of cases of colon carcinoma developsporadically in patients who have no known predisposition for the disease.[87]It is estimated that adherence to the current American Cancer Society andGastrointestinal Society colorectal cancer screening guidelines could lower theannual mortality rate by at least 50% over the next decade.[88]

COX-2: Expression and Preclinical Data

In humans, overexpression of COX-2 has been documented incolorectal adenomas and cancers, but not in normal-appearing mucosa.[89] Forexample, Figure 3 shows immunohistochemical staining for COX-2 in colon adenomatissue. Similar overexpression of COX-2 has been documented in a wide range ofcancers and their precursors (Table 1).[40-81] The chemopreventive effects ofCOX-2 inhibitors on the development of colorectal cancer are the subject ofintense study, and animal models have been useful in investigating colorectalcancer pathogenesis.

A mutation in the adenomatous polyposis coli (APC) gene resultsin spontaneous adenoma formation in the small intestine of APC delta716 knockoutmice. Using this rodent model, Oshima and colleagues demonstrated that there isa cause-effect relationship between COX-2 overexpression and gastrointestinaltumor incidence.[90] It was shown that suppression of one allele of the COX-2gene reduced the number of intestinal polyps by 66%, and suppression of bothalleles resulted in a reduction of 86%.[91] Furthermore, treatment of COX-2-expressingazoxymethane-treated rats with oral celecoxib suppressed formation of colorectaltumors by > 90%, compared with a suppression of 40% to 65% followingadministration of a nonselective COX inhibitor.[91,92]

Reduced Incidence of Colorectal Neoplasia

In prospective cohort studies, long-term administration ofaspirin and other NSAIDs has been associated with a reduction in the incidenceof colorectal adenomas, cancer, and cancer mortality by 40% to 50%.[93-95] Aninverse relationship also has been demonstrated between the use of NSAIDs andthe incidence of colorectal cancer in several case studies.[96,97] Furthermore,clinical trials showed that the administration of sulindac, a commonlyprescribed NSAID, prescribed to familial adenomatous polyposis patients wasassociated with a reduction in the number and size of adenomas.[97-99]

The US Food and Drug Administration recently granted approval ofcelecoxib for treatment of familial adenomatous polyposis. Celecoxib, which wasinitially approved for the relief of signs and symptoms of osteoarthritis andrheumatoid arthritis, is highly selective for COX-2[100] (375-fold greaterselectively compared with COX-1) and has a significantly reduced incidence ofcommon gastrointestinal toxicities, such as bleeding and upper gastrointestinalulcers, associated with NSAIDs.[101]

The pivotal trials of celecoxib for the treatment of familialadenomatous polyposis enrolled 77 patients who were randomized to receive eitherplacebo or celecoxib (100 or 400 mg twice daily) for 6 months.[102] The primaryefficacy end point was the percent change in the number of colorectal adenomas(> 2 mm in size) at 6 months. There was a 4.5% reduction in theplacebo-treated group, a 11.9% reduction in the 100-mg celecoxib-treated group,and a 28.0% reduction in the 400-mg celecoxib-treated group. The decrease inincidence between the 400-mg celecoxib twice-daily group and the placebo groupwas statistically significant (P = .003). The prevalence of adverse events was similar among the treatmentgroups and consisted primarily of diarrhea, dyspepsia, fatigue, upperrespiratory infection, and rash.

The results from the pivotal trial of celecoxib in familialadenomatous polyposis support further investigation of COX-2 inhibitors for anoverall chemoprevention strategy for colorectal tumors in other populations atrisk, including patients with sporadic adenomatous polyps. As shown in Table2,[82-84] there are several recently initiated clinical trials of celecoxib inthe prevention or recurrence of colorectal adenomas. Two are being conductedunder the sponsorship of the Division of Cancer Prevention at the NationalCancer Institute.

One clinical trial is being led by Monica Bertagnolli, MD,Brigham and Women’s Hospital, Dana-Farber Cancer Institute, Boston. This trialis investigating two dose levels of celecoxib compared with placebo with 1- and3-year colonoscopy end points. A second phase III clinical trial (using afactorial design) is studying celecoxib (400 mg/d) vs selenium (in the form ofbaker’s yeast) vs the combination of celecoxib/selenium vs a double placebo.This study is being conducted at the Arizona Cancer Center, Tucson, by Drs.David Alberts and Peter Lance.

An additional randomized, phase III trial is evaluating thepotential for celecoxib to reduce the incidence of sporadic adenomas. Adenomarecurrence rates will be evaluated at a 3-year colonoscopy end point. Theresults of these trials will not be available for several years, but couldestablish COX-2 inhibitors as important components of the management strategyfor colorectal adenomas and the prevention of colon cancer.

Nonmelanoma Skin Cancer

It is well recognized that chronic sun exposure is a majoretiologic agent for skin cancer, contributing to over 1 million new cases ofbasal cell carcinoma and squamous cell carcinoma each year in the United States.Although the majority of skin cancers are basal cell carcinomas, which arerelatively benign, squamous cell carcinomas account for approximately 2,000deaths annually.[103] Ultraviolet (UV)-B exposure is responsible for mostsquamous cell skin cancers.[104-106] It is considered to be the principalcarcinogen in skin cancers and is involved with all stages of carcinogenesis(initiation, promotion, and progression). Ultraviolet-A is also capable ofcausing oxidative stress and is also associated with UV-induced carcinogenesis,particularly melanoma.[106-108] Actinic keratosis is considered a precursorlesion of squamous cell carcinoma, with approximately 60% of squamous cellcarcinomas evolving from actinic keratosis.[109,110]

COX-2: In Vitro and Preclinical Data

Several mechanisms for UV-induced skin carcinogenesis have beendefined, including dysregulation of cell signal transduction pathways[111-113]and upregulation of COX-2 expression.[114,115] In the normal epidermis, thebalance between proliferation of cells in the basal layer, cell differentiationin the suprabasal spinous and granular layers, and apoptosis at the transitionalzone (where the stratum granulosum and stratum corneum meet) is tightlyregulated. Cells lose their proliferative capacity as they undergo terminaldifferentiation and leave the basal cell layer.

When homeostasis is maintained in the epidermis, COX-1 isconstitutively expressed by keratinocytes and COX-2 is abnormallyexpressed.[116] Studies conducted by Athar and colleagues, however, havedemonstrated that COX-2 is expressed in the epidermis in response to tumorpromoting agents,[117] and other colleagues demonstrated that the differentialexpression of COX-2 in mouse skin carcinoma is regulated by cis elements withinthe promoter region of the COX-2 gene.[117,118] Neufang and colleagues showedthat, in a transgenic mouse model, induced expression of COX-2 in the epidermisresulted in abnormal epidermal differentiation.[119] Furthermore, proliferationof epidermal cells was increased and, in some regions of the epidermis, therewas an increase in the number of viable cornified layers.

Prostaglandins are actively synthesized by the human epidermis:COX-2-mediated overexpression of prostaglandin E2 has been shown to increaseepidermal cell proliferation in vitro.[114] Prostaglandins generated by thearachidonic acid cascade have been implicated in various models of skintumorigenesis. Elevated levels of protaglandin E2 have been observed in basalcell carcinoma, squamous cell carcinoma, and actinic keratosis. Data suggestthat prostaglandin E2 overexpression may correlate with the propensity formetastatic and invasive behavior.[120]

In the context of skin tumorigenesis, chemical carcinogenesismodels have implicated COX-2-regulated prostaglandin expression in mediatingtumor promotion.[121,122] Upregulation of COX-2-induced prostaglandinproduction causes an increase in cell growth and, as shown by Tsujii and DuBois,skin carcinoma cells overexpressing COX-2 are resistant to apoptosis.[60,123]

Cyclooxygenase-2 overexpression has also been demonstrated inUV-exposed human skin cells. Buckman and colleagues showed that exposure ofhuman keratinocytes in vitro to UV-B caused a significant increase in expressionof prostaglandin E2.[124] In cultured human keratinocytes, UV-B-induced COX-2activity was blocked completely by indomethacin, a nonspecific COX inhibitor,and by SC58125, a COX-2-specific inhibitor.[125]

Clinical Data and Clinical Trials

Kagoura and colleagues used immunohistochemistry to investigateCOX-2 expression in basal cell carcinoma, Bowen’s disease, squamous cellcarcinoma, and metastatic tumors of the skin.[126] Four of 16 basal cellcarcinoma cases tested positive for COX-2. In Bowen’s disease, the stainingintensity for COX-2 was greater than that observed in basal cell carcinoma.

In addition, 11 of 15 squamous cell carcinoma patients testedpositive for COX-2. The pattern of staining was heterogeneous, with more intensestaining in the centers of the tumor loci. In metastatic tumors, the percentageof COX-2-positive tumor cells and the intensity of staining was low comparedwith Bowen’s disease and basal cell carcinoma, suggesting that COX-2 may playa greater role in the early stages of squamous cell carcinogenesis.

Currently, there are several ongoing clinical trials examiningthe effect of COX-2 inhibition on skin cancer (Table2).[82-84] At theUniversity of California, San Francisco, a study is being conducted to examinewhether celecoxib prevents the development of basal cell carcinoma in patientswith basal cell nevus syndrome. Celecoxib is also being tested in severalongoing, placebo-controlled clinical trials for preventing the development ofnew actinic keratosis lesions.

A phase IIB, double-blind, placebo-controlled trial is ongoingat the University of Alabama with sponsorship from the National CancerInstitute. The primary end point of this study will be inducing actinickeratosis regression. Secondary end points include the effect of celecoxib onpotential surrogate end point biomarkers in areas of actinic keratosis,sun-exposed skin, and non-sun-exposed skin, and the correlation of thesebiomarkers with clinical outcome.

Reports examining COX-2 and skin carcinogenesis illustrate theneed for evaluation of the efficacy of COX-2-selective inhibitors aschemoprevention agents to inhibit the course of UV-induced skin cancer. Futureresearch should combine in vitro techniques, and available transgenic knockoutor overexpression mouse models, in addition to human studies, to more clearlydefine the role of COX-2 in UV-B- and UV-A-induced carcinogenesis of theskin.

Lung Cancer

Lung cancer is the leading cause of cancer deaths in the UnitedStates. Furthermore, lung cancer has a high propensity to spread by hematologicand lymphatic routes. More than 86% of cases occur in current or formersmokers.[127,128] Approximately 75% of lung cancers are classified as non-small-celllung carcinoma. The disease is usually diagnosed at an advanced stage when it istherapeutically intractable.[129,130] Clearly, understanding the biology of thedisease and discovering specific molecular targets to facilitate early detectionand treatment would be valuable.

COX-2: In Vitro and Preclinical Data

Over 30% of non-small-cell lung cancer tumors and cell linesharbor an activating mutation in the K-ras oncogene,[131] which appearsassociated with expression of COX-2 in a panel of non-small-cell lung cancercell lines.[132] Several reports also have shown that NSAIDs block thetransformed growth of non-small-cell lung cancers that express the K-rasmutation.[133-135] It has been hypothesized that the generation of bioactivelipids derived from the arachidonic acid metabolic pathway modulates physiologicand pathologic responses involved in tumor growth and tumor progression in thelung. This theory has been based on data indicating that COX inhibitors inhibitthe growth of human lung cancer cell lines in vitro and in animalmodels,[136,137] and abnormalities in arachidonic acid metabolism are present innon-small-cell lung cancer.[138]

Masferrer and colleagues studied the effect of celecoxib in theLewis lung model.[139] Mice were injected with cancer cells into the foot pad,then fed either a control diet or a diet supplemented with celecoxib. Over aperiod of 30 days, celecoxib caused a dose-dependent decrease in tumor volumeand, at higher doses, celecoxib induced a decrease in the size and number oflung metastases. It was hypothesized that the decrease in tumor size andmetastatic invasion was due to celecoxib-induced inhibition of angiogenicactivity.[140]

Clinical Data and Clinical Trials

Epidemiologic studies have shown that NSAIDs, including aspirin,significantly reduce the risk of lung cancer.[41,42,141] Furthermore, COX-2expression appears elevated in non-small-cell lung cancer,[131,132] andincreased COX-2 expression is associated with poor prognosis.[43] Using in situhybridization with a COX-2 antisense probe, Khuri and colleagues demonstratedthat COX-2 expression in stage I disease was associated with decreased survivalrates.[61]

Cyclooxygenase-2 is diffusely overexpressed in atypicaladenomatous hyperplasia, a possible precursor lesion of adenocarcinoma of thelung.[42] Lau and colleagues demonstrated COX-2 overexpression in 19 of 20 lungcancer specimens using immunohistochemical analyses.[138] In another study,COX-2 overexpression was observed in both squamous cell carcinoma andadenocarcinomas of the lung.[40] Thus, COX-2 may provide a molecular target fortherapy and prevention of lung cancer in smokers and nonsmokers. A pilot studyof COX-2 inhibitors in high-risk tobacco smokers is ongoing at the JonssonCancer Center, University of California, Los Angeles. Other studies evaluatingthe role of celecoxib in heavy smokers are ongoing. In addition, celecoxib isbeing studied in neoadjuvant and therapeutic trials in combination withchemotherapy.

Prostate Cancer

Prostate carcinoma is the most commonly diagnosed cancer and thesecond-leading cause of cancer deaths in men in the United States. It isexpected that approximately 189,000 new cases will be diagnosed and more than30,200 deaths due to prostate cancer will occur in 2002.[85] Treatment foradvanced prostate cancer often involves androgen ablation therapy and eithersurgical removal of the prostate, external beam radiation, or implantation ofradioactive "seeds" into the prostate (brachytherapy). Unfortunately,as prostate carcinoma progresses, it tends to become androgen independent and,therefore, refractory to hormone therapy.[142,143] Patients diagnosed inadvanced stages of the disease have a poor prognosis.[144-146] Agents capable ofinhibiting cell growth and sensitizing prostate carcinoma cells to stimuli thatinduce apoptosis, such as radiation therapy, would enhance the efficacy oftreatment.

COX-2: In Vitro and Preclinical Data

Accumulating evidence suggests that overexpression of COX-2 isassociated with resistance to apoptosis, thereby increasing the tumorigenicpotential of a cancer.[123] Furthermore, selective inhibition of COX-2 has beendemonstrated to induce apoptosis in prostate carcinoma cells in vitro.[76,147]These observations suggest that COX-2 inhibitors could be effectivechemoprevention agents.

Lim and colleagues have examined the potential for COXinhibitors to be used to induce apoptosis in prostate cancer.[77] The COX-1 and-2 inhibitor sulindac was tested in vitro for proapoptotic activity in theprostate cancer cell lines PC3 and LNCaP, and in a normal prostate epithelialcell line PrEC. Apoptosis was quantified following treatment with eithersulindac or sulindac sulfone (exisulind, Aptosyn). Forty-eight hours followingtreatment with either sulindac or exisulind, 50% of PC3 cells and 40% of LNCaPcells underwent apoptosis. However, PrEC cells showed no indication of apoptosisat similar concentrations of drug.[77]

Studies evaluating the effect of COX-2-specific inhibitors onangiogenesis in prostate carcinoma cell lines have been performed. Cell linesLNCaP and PC3, and the control cell line PrEC, were treated with two COX-2-specificinhibitors, etodolac (Lodine) and NS398.[78] Both compounds decreased cellproliferation in the carcinoma cell lines, but not in the normal prostatestromal cell line. A DNA fragmentation assay revealed that both compounds alsoinduced apoptosis in the two carcinoma cell lines, but not in the normal stromalcell line.

Clinical Data and Clinical Trials

Using immunohistochemical analyses, Uotila and colleaguesexamined expression of COX-1 and -2 in prostate carcinoma tumors.[148] Nosignificant difference in COX-1 expression was observed between normal prostateand prostate cancer tissue. However, their data revealed stronger staining ofCOX-2 in the prostate cancer cells compared with normal glandular epithelium ofcontrol prostates. Cyclooxygenase-2 expression was also elevated in theprecursor lesion of prostate carcinoma, prostate intraepithelial neoplasia.[148]

Using similar techniques to quantify COX-2 expression,Kirschenbaum et al showed COX-2 overexpression in 86% of prostateintraepithelial neoplasia lesions and 87% of carcinomas collected duringprostatectomy.[79,149] Furthermore, treatment of prostate carcinoma cells with aselective COX-2 inhibitor induced apoptosis in vivo and in vitro. The in vivoresults revealed that the COX-2 inhibitor decreased microvessel density andangiogenesis. The investigators hypothesized that the decrease in angiogenesiswas caused by inhibition of COX-2-induced expression of vascular endothelialgrowth factor.[79,149]

As a result of the preclinical experimental results showingactivity for exisulind against prostate cancer, this phosphodiesterase/COX-1 and-2 inhibitor currently is being evaluated in a series of phase I/II clinicaltrials. These trials are examining prostate-specific antigen response andmeasurable disease response rate with exisulind as a single agent or incombination with docetaxel (Taxotere) (Table2).[82-84] Secondary objectivesinclude time to disease progression and duration of response in patients withprostate carcinoma.

Celecoxib is also undergoing clinical development forprevention/treatment of prostate cancer (Table2).[82-84] In one phase I/IIstudy, patients will be randomized to receive either celecoxib or placebo priorto radical prostatectomy.[82] The effect of celecoxib on COX-2 expression andangiogenic factors in the prostate will be examined.

These trials of COX-2 inhibitors will help determine theirfuture role in treatment and chemoprevention of prostate cancer. Clearly, themechanisms by which these drugs exert a proapoptotic effect in prostatecarcinoma cells warrant further investigation.

Breast Cancer

The incidence of breast carcinoma varies with age andnationality. In the United States, the incidence increases rapidly to about age45 years, then increases more slowly. At age 25 years, the incidence isapproximately 5 cases in 100,000 women; at age 50 years, the incidence increasesto 150 cases per 100,000 women. At age 75 years, the incidence rises to 200 per100,000 women.[150,151] The development of breast cancer appears to be relatedto ovarian function and hormone production. However, a young age at firstpregnancy, late menarche, and late menopause are factors with a protectiveeffect against breast cancer.[152-154] Furthermore, oophorectomy before age 35years reduces the risk of developing breast cancer by 70%.[155,156]

Therapy usually consists of a combination of surgery, radiation,and chemotherapy. Clearly, molecular target-specific chemoprevention agents withfew or no adverse effects would be valuable to reduce morbidity and mortalityrates from breast cancer.

COX-2: In Vitro and Preclinical Background

Prostaglandins have been implicated in breast carcinogenesis.Breast carcinoma cell lines[53,54] and rodent models[157] express elevatedlevels of COX-2 and prostaglandins compared with the normal tissues from whichthey arise. Inhibition of COX by NSAIDs reduces the development of chemicallyinduced and transplantable mammary cancers.[55,158,159] It is also interestingthat addition of hormones to human breast cancer cell lines[160] or rat mammarycarcinoma cell lines[161] induces expression of prostaglandins.

Studies conducted by Rozic and colleagues examined the role ofprostaglandins in the proliferation, survival, migratory, and invasive behavior,and angiogenic capacity in a highly invasive murine mammary tumor cell line.[55]Northern and Western blot analyses revealed a high expression level of COX-2mRNA and protein, respectively. Their results indicated a high level ofprostaglandin E2 expression, which was completely abrogated by COX-2-specificinhibitors.

Migratory and invasive behavior was measured with an in vitrotranswell migration/invasion assay. Cyclooxygenase-2-specific antagonistsinhibited migration, while a COX-1-specific inhibitor had no effect ontranswell migration. The COX-2-specific inhibitors also blocked angiogenesisin an in vivo assay. These studies suggest that COX-2-specific inhibitors maybe effective for preventing breast tumor development and/or progression.[55]

Other studies also have shown that COX inhibitors blockformation of breast tumors in mouse models.[55,162,163] Some studies suggestthat COX-2 inhibitors may not only prevent mammary carcinogenesis, but may alsoprevent multidrug resistance in breast carcinoma.[164] The molecular mechanismsof prostaglandin-mediated progression of breast cancer are not characterized;however, one mechanism may involve the prostaglandin E2 receptor. Fulton andcolleagues demonstrated that advanced and metastatic breast carcinomas oftenexpress a mutation in the prostaglandin E2 receptor. This mutation may also beinvolved in multidrug resistance.[165]

The potential of celecoxib to prevent mammary cancer was studiedin the dimethylbenzanthracene (DMBA)-induced rat mammary tumor model by Harrisand associates.[166] In this study, animals were fed a control diet or a dietsupplemented with either ibuprofen (a COX-1 and -2 inhibitor) or celecoxib for 1week prior to a single dose of DMBA. Both ibuprofen and celecoxib groups showeda statistically significant reduction in tumor incidence. However, the reductionin tumor incidence observed in the celecoxib group was greater than thatobserved in the ibuprofen-treated group. Furthermore, Alshafie and colleaguesdemonstrated that celecoxib also had a therapeutic effect in the DMBA-inducedrat mammary model.[167]

Clinical Data and Clinical Trials

Hwang and colleagues demonstrated that COX-2 is overexpressed inbreast hyperplasia and atypical hyperplasias.[56] Similarly, these investigatorsshowed that COX-2 is overexpressed in > 56% of breast cancers. Interestingly,COX-2 overexpression was more apparent in ductal carcinomas in situ than inmetastatic breast cancers, suggesting that COX-2 may play an important role inthe early carcinogenesis of the breast.[40] The COX-2-specific inhibitors maybe ideal candidates for chemoprevention of breast cancer for several reasons.The toxicity profile appears to be acceptable for premenopausal women becauseCOX-2 inhibitors do not cause perimenopausal symptoms. However, long-termeffects of the use of celecoxib in these subjects is being evaluated in clinicalstudies. They may be effective in both estrogen-receptor-positive andestrogen-receptor-negative disease.

Several clinical trials with celecoxib are at different pointsof development. One such trial has been initiated examining efficacy ofcelecoxib and trastuzumab (Herceptin) in women with HER2/neu-overexpressingmetastatic breast cancer that is refractory to trastuzumab (Table2).[82-84]Multiple trials are ongoing to evaluate the effect of celecoxib in precancerous,in situ, and invasive breast carcinoma.

Gastric Cancer

The incidence of gastric cancer varies remarkably in differentparts of the world. The incidence is relatively high in Japan and Chile, withapproximately 58.4 cases per 100,000 men and 29.9 cases per 100,000 women.[168]The incidence of gastric cancer is lowest in the Dominican Republic andThailand, where the rate is approximately 5% of that reported in Japan.[168,169]The high rate of gastric carcinoma in Japan is thought to be caused in part bythe high consumption of cured meat and fish. These products contain a highconcentration of N-nitroso compounds that cause formation of potentiallycarcinogenic endogenous nitrosamines.[170,171] In the United States, 21,600cases of gastric cancer are expected in 2002.[85] However, the incidence in theUnited States and Japan has dropped in recent years, perhaps due to an increasedawareness of proper dietary habits.[172,173]

COX-2: Clinical Data

Some studies suggest that a Helicobacter pylori infectionpredisposes a patient to gastric cancer, perhaps owing to increased expressionof COX-2.[174-176] Leung and colleagues examined the association betweenexpression of COX-2 and the presence of a missense mutation in the tumorsuppressor gene p53.[177] Wild-type p53 normally binds to cis elements withinthe promoter region of the COX-2 gene and inhibits its transcription. Of 39patients with gastric cancer, 19 (49%) overexpressed COX-2.[66] Of these, thepatients with the strongest COX-2 expression also had a mutated p53 gene.

In addition, the cancer in these patients with higher COX-2expression and mutated p53 gene was more aggressive, with more lymphaticinvasion and metastases. These data suggest that mutation of the p53 gene mayhave a direct effect on expression of COX-2 in gastric carcinoma, and thatincreased COX-2 expression may be associated with poor prognosis.[66,177]

Other studies revealed a direct relationship between expressionof COX-2 mRNA and increased tumor invasion. Thus, COX-2 inhibitors may provideeffective prevention in patients with H pylori infection and in patients withthis risk factor for gastric cancer. Clinical trials should be developed to testthese agents in this patient population.

Bladder Cancer

Bladder cancer is the fourth leading cause of cancer in men andthe eighth in women. Furthermore, it is the ninth and fourteenth leading causeof US cancer deaths in men and women, respectively.[178] Many etiologic factorshave been identified including occupation (eg, close contact with chemicalcarcinogens),[179,180] diet,[181,182] chronic bladder infections,[183,184] andsmoking.[185,186]

The vast majority of bladder cancers arise from bladderpapilloma precursors, providing significant opportunities for the development ofchemoprevention strategies.

COX-2: In Vitro and Preclinical Background

Recent animal studies indicate that nonspecific COX inhibitorsand specific COX-2 inhibitors reduce the incidence of bladder carcinoma inducedby chemical carcinogens.[187,188] Khan and colleagues examined COX-1 and -2expression in normal dogs and in dogs with transitional cell carcinoma.[49]There was no difference in COX-1 expression between normal and malignant bladdertissue; however, COX-2 was only expressed in carcinoma and in new blood vesselsin the tumor tissue.

Using two rodent models of bladder cancer, Grubbs and colleaguesdemonstrated that the COX-2-specific inhibitor celecoxib was effective ininhibiting chemically induced bladder cancer.[189] Male mice treated withN-butyl-N-(4-hydroxy-butyl)nitrosamine (OH-BBN) compared with no treatmentdeveloped transitional and squamous cell urinary bladder cancer with highmorbidity. Mice pretreated with celecoxib 7 days before initiation of 12 weeklydoses of OH-BBN had a 75% decrease in the incidence of bladder cancer. There wasno decrease in development of preneoplastic lesions, suggesting that COX-2 mayplay a role during the progression stages of bladder carcinoma.

Similar results were observed in a rat model using the samechemical initiator and time course.[189] These data suggest that COX-2-specificinhibitors may not prevent initiation of carcinogenesis in the urinary bladder.However, they may be effective in preventing bladder carcinoma in individualsidentified as high risk and for treatment of early-stage bladder cancer.

Clinical Data and Clinical Trials

In humans, several studies have shown that COX-2 is expressed ininvasive transitional cell carcinoma, but not in normal bladder tissue.Expression of COX-2 is localized to the carcinoma cells, whereas there appearsto be no COX-2 expression in the stroma.[50-52] It is also interesting that instudies examining the molecular characteristics of bladder carcinoma in patientsin close proximity to Chernobyl following the nuclear accident in 1986,Romanenko and colleagues showed that overexpression of COX-2 was associated withmutations in the p53 gene.[190] These findings are similar to data indicating arelationship between mutant p53 and COX-2 expression in gastric cancer.[177]

Currently, a large clinical study examining inhibition of COX-2and recurrence of bladder cancer is ongoing at The University of Texas M. D.Anderson Cancer Center, Houston (Table 2).[82-84] This study is designed tocompare the time to recurrence following treatment with celecoxib or placebo inpatients with superficial transitional cell carcinoma of the bladder at highrisk for recurrence. This study will also correlate the modulation of one ormore biomarkers with recurrence of bladder cancer and evaluate the quality oflife of patients enrolled on each of the two regimens.

Esophageal Cancer

During the last decade, the incidence of esophagealadenocarcinoma has increased at a disproportionately rapid rate.[191] Themajority of esophageal cancers arise in patients with the premalignant conditionBarrett’s esophagus, which manifests by the replacement of normal esophagealepithelium with a columnar type.[192-195] Barrett’s esophagus is frequentlyassociated with high-grade dysplasia and aneuploidy in esophagealepithelium.[196-199] The condition develops from chronic, severegastroesophageal reflux. Patients afflicted with Barrett’s esophagus have a30- to 40-fold increased risk for development of esophagealadenocarcinoma.[200,201] Postoperative morbidity is high for esophageal cancer,and survival is not favorable (13% to 30% at 5 years).[202-205] Clearly,chemoprevention strategies are needed.

COX-2: In Vitro and Preclinical Background

Research conducted by Li and colleagues demonstrated thataspirin retarded cell growth in an esophageal adenocarcinoma cell line.[206]Growth inhibition was shown to be dose and time dependent.[63] Subsequently,they tested the effect of the COX-2-specific inhibitor NS398 on apoptosis andexpression of genes that regulate apoptosis. In vitro, the COX-2-specificinhibitor induced apoptosis in several esophageal adenocarcinoma cell linesthrough a cytochrome C-dependent pathway. There was a direct relationshipbetween apoptosis and the level of COX-2 expression. Caspase-9 and caspase-3were activated by NS398, and addition of a caspase inhibitor reversed theapoptotic effect of the COX-2 inhibitor.[206] These data suggest that COX-2specific inhibitors may be promising for chemoprevention and treatment ofesophageal adenocarcinoma.

Clinical Data

Kandil and colleagues examined esophageal pinch biopsy specimensfrom patients with Barrett’s esophagus from normal and abnormal tissue.[207]Western analyses for COX-2 protein showed an increase in COX-2 expression in 41%of the Barrett’s esophagus tissue biopsies. No COX-2 expression was observedin adjacent normal tissue. It is of interest that COX-2 was found in Barrett’sesophagus tissue with or without dysplasia, suggesting that COX-2 may play arole in the early stages of development of adenocarcinoma.[45] Additionally,COX-2 is overexpressed in adenocarcinoma arising from Barrett’s esophagus atall stages.[46,208,209]

Currently, a clinical study is ongoing coordinated by the JohnsHopkins Comprehensive Cancer Center, Baltimore, to evaluate the efficacy andsafety of celecoxib in patients with Barrett’s esophagus (Table

Pancreatic Cancer

Pancreatic cancer is a highly invasive, aggressive disease andis the fifth leading cause of cancer deaths in the United States.[210] Althoughthe DNA synthesis inhibitor gemcitabine (Gemzar) can have a modest effect onsurvival, the majority of patients succumb to their disease within 6 monthsfollowing diagnosis.[211,212]

COX-2: In Vitro and Preclinical Data

Studies were conducted to measure the effects of COX-nonspecificand COX-2-specific inhibitors on cell growth and apoptosis in four pancreaticcancer cell lines expressing COX-2.[213] Cell growth, measured by[³H]-thymidineincorporation, showed a dose-dependent decrease in cell proliferation with boththe nonspecific COX inhibitors and the COX-2-specific inhibitor NS298.

Yip-Schneider and colleagues examined the effects of COXinhibitors in combination with gemcitabine in vitro.[74,214] The COX-nonspecificand COX-2-specific inhibitors caused cell cycle arrest primarily in the G1/G0phase through decreased expression of cyclins, whereas gemcitabine caused arrestin the S phase due to its incorporation into DNA. It is of great interest thatCOX inhibitors in combination with gemcitabine had an additive effect oninhibition of cell growth. No significant effect on apoptosis was observed inany of the test groups, suggesting that the drugs alone and in combination mayinduce cell senescence.[74,214]

Clinical Data

Tucker and colleagues examined expression of COX-2 mRNA intissue isolated from pancreatic cancer patients using quantitative reversetranscriptase polymerase chain reaction.[75] These studies revealed a 60-foldincrease in expression of COX-2 mRNA in tissue isolated from 9 of 10 pancreaticadenocarcinomas compared with adjacent normal pancreatic tissue.Immunohistochemical staining indicated that COX-2 expression was localized tothe malignant epithelium. It was also demonstrated that COX-2 was expressed inhuman pancreatic carcinoma cell lines.[75]

Clearly, the effectiveness of COX inhibitors in combination withgemcitabine may have potential for treatment of pancreatic cancer, but continuedclinical development is required. Several studies evaluating the effect ofcelecoxib in combination with chemotherapy for pancreatic cancer have beeninitiated, including a phase II clinical trial of celecoxib and gemcitabine atthe Arizona Cancer Center.

Conclusions

Cyclooxygenase-2 is known to be a mediator of inflammation andother immune processes. It is therefore of interest that several of themalignancies in which COX-2 is overexpressed are associated with a chronicinflammatory condition. For example, esophageal carcinoma is associated withBarrett’s esophagus, and nonmelanoma skin cancers are associated with UVdamage. In addition, gastric carcinoma is associated with an ulcerativecondition caused by infection with H pylori. The data suggest that inhibition ofCOX-2 may inhibit carcinogenesis at a very early stage and may completely blocktumor formation in tumors that arise from nonmalignant inflammatory precursors.As discussed in this review, preclinical and clinical data support thehypothesis that COX-2 plays a role in oncogenesis, and that COX-2 inhibitors mayoffer effective chemoprevention strategies (summarized in Figure4).

Epidemiologic studies revealing an inverse correlation betweenthe incidence of colon cancer and regular use of NSAIDs provided initial cluessuggesting that COX inhibition may be an effective intervention approach topreventing cancer.[215-217] Because NSAIDs are known to block prostaglandinsynthesis by inhibition of the COX enzymes, it was hypothesized that aberrantprostaglandin synthesis may contribute to colorectal neoplasia. Furtherinvestigations documented that COX-2 is barely detectable in normal colonmucosa, but is significantly upregulated in colon carcinoma.[34,47]Cyclooxygenase-2 also is overexpressed in intestinal adenomas in rodent modelsof intestinal tumorigenesis and in a high percentage of colorectaladenomas.[218,219]

More recently, in familial adenomatous polyposis patients, theCOX-2-specific inhibitor celecoxib caused a 28% reduction in colorectaladenomas after 6 months of treatment at a dose of 400 mg twice-daily, whereascontrols given placebo experienced no significant reductions in adenomanumbers.[102] This finding has established celecoxib as a potentially valuablechemoprevention agent in patients at high risk of sporadic colorectal adenomarecurrence.[48,220-222]

Currently, there are three phase III trials under way to testdifferent doses of celecoxib in the prevention of sporadic colorectal polyprecurrence. Clearly, further examination of COX-2 inhibitors in prevention andtreatment of cancer is warranted. Because of the favorable safety profile inhealthy individuals, COX-2 inhibitors may provide an effective and safe optionfor cancer chemoprevention.

References:

1. Topley N, Petersen MM, Mackenzie R, et al: Human peritonealmesothelial cell prostaglandin synthesis: Induction of cyclooxygenase mRNA byperitoneal macrophage-derived cytokines. Kidney Int 46:900-909, 1994.

2. Kulkarni PS: Synthesis of cyclooxygenase products by humananterior uvea from cyclic prostaglandin endoperoxide (PG H2). Exp Eye Res32:197-204, 1981.

3. Holmes DR, Wester W, Thompson RW, et al: Prostaglandin E2synthesis and cyclooxygenase expression in abdominal aortic aneurysms. J VascSurg 25:810-815, 1997.

4. Foster SJ, McCormick ME, Howarth A: The contribution ofcyclooxygenase and lipoxygenase products to acute inflammation in the rat.Agents Actions 17:358-359, 1986.

5. Seibert K, Masferrer JL: Role of inducible cyclooxygenase(COX-2) in inflammation. Receptor 4:17-23, 1994.

6. Vane JR, Mitchell JA, Appleton I, et al: Inducible isoformsof cyclooxygenase and nitric-oxide synthase in inflammation. Proc Natl Acad SciU S A 91:2046-2050, 1994.

7. Joyce IM, Pendola FL, O’Brien M, et al: Regulation ofprostaglandin-endoperoxide synthase 2 messenger ribonucleic acid expression inmouse granulosa cells during ovulation. Endocrinology 142:3187-3197, 2001.

8. Matsumoto H, Ma W, Smalley W, et al: Diversification ofcyclooxygenase-2-derived prostaglandins in ovulation and implantation. BiolReprod 64:1557-1565, 2001.

9. Hwang D, Jang BC, Yu G, et al: Expression ofmitogen-inducible cyclooxygenase induced by lipopolysaccharide: Mediationthrough both mitogen-activated protein kinase and NF-kappa B signaling pathwaysin macrophages. Biochem Pharmacol 54:87-96, 1997.

10. Kujubu DA, Herschman HR: Dexamethasone inhibits mitogeninduction of the TIS10 prostaglandin synthase/cyclooxygenase gene. J Biol Chem267:7991-7994, 1992.

11. Gualde N, Atluru D, Goodwin JS: Effect of lipoxygenasemetabolites of arachidonic acid on proliferation of human T cells and T cellsubsets. J Immunol 134:1125-1129, 1985.

12. Santoli D, Phillips PD, Colt TL, et al: Suppression ofinterleukin 2-dependent human T cell growth in vitro by prostaglandin E (PG E)and their precursor fatty acids. Evidence for a PG E-independent mechanism ofinhibition by the fatty acids. J Clin Invest 85:424-432, 1990.

13. Holtzman MJ: Species-specificity of lipoxygenase andcyclooxygenase activities expressed in pulmonary airway epithelial cells. AdvProstaglandin Thromboxane Leukot Res 17A:177-179, 1987.

14. Seltzer J, Bigby BG, Stulbarg M, et al: O3-induced change inbronchial reactivity to methacholine and airway inflammation in humans. J ApplPhysiol 60:1321-1326, 1986.

15. Hansbrough JR, Atlas AB, Turk J, et al: Arachidonate12-lipoxygenase and cyclooxygenase: PG E isomerase are predominant pathways foroxygenation in bovine tracheal epithelial cells. Am J Respir Cell Mol Biol1:237-244, 1989.

16. Reed DW, Bradshaw WS, Xie W, et al: In vivo and in vitroexpression of a nonmammalian cyclooxygenase-1. Prostaglandins 52:269-284, 1996.

17. Lipsky PE: Role of cyclooxygenase-1 and -2 in health anddisease. Am J Orthop 28:8-12, 1999.

18. Maier JA, Hla T, Maciag T: Cyclooxygenase is animmediate-early gene induced by interleukin-1 in human endothelial cells. J BiolChem 265:10805-10808, 1990.

19. Luckman SM, Dye S, Cox HJ: Induction of members of theFos/Jun family of immediate-early genes in identified hypothalamic neurons: invivo evidence for differential regulation. Neuroscience 73:473-485, 1996.

20. Hinterleitner TA, Saada JI, Berschneider HM, et al: IL-1stimulates intestinal myofibroblast COX gene expression and augments activationof Cl- secretion in T84 cells. Am J Physiol 271:C1262-C1268, 1996.

21. Huang ZF, Massey JB, Via DP: Differential regulation ofcyclooxygenase-2 (COX-2) mRNA stability by interleukin-1 beta (IL-1 beta) andtumor necrosis factor-alpha (TNF-alpha) in human in vitro differentiatedmacrophages. Biochem Pharmacol 59:187-194, 2000.

22. Gilroy DW, Saunders MA, Wu KK: COX-2 expression andcell-cycle progression in human fibroblasts. Am J Physiol Cell Physiol281:C188-C194, 2001.

23. Goppelt-Struebe M, Rehm M, Schaefers HJ: Induction ofcyclooxygenase-2 by platelet-derived growth factor (PDGF) and its inhibition bydexamethasone are independent of NF-kappaB/kappaB transcription factors. NaunynSchmiedebergs Arch Pharmacol 361:636-645, 2000.

24. Goppelt-Struebe M, Wiedemann T, Heusinger-Ribeiro J, et al:Cox-2 and osteopontin in cocultured platelets and mesangial cells: role ofglucocorticoids. Kidney Int 57:2229-2238, 2000.

25. Majima M, Isono M, Ikeda Y, et al: Significant roles ofinducible cyclooxygenase (COX)-2 in angiogenesis in rat sponge implants. Jpn JPharmacol 75:105-114, 1997.

26. Shimada K, Kita T, Yonetani Y, et al: Modulation byendothelin-1 of lipopolysaccharide-induced cyclooxygenase 2 expression in mouseperitoneal macrophages. Eur J Pharmacol 376:285-292, 1999.

27. Shimada K, Kita T, Yonetani Y, et al: The effect ofendothelin-1 on lipopolysaccharide-induced cyclooxygenase 2 expression inassociation with prostaglandin E2. Eur J Pharmacol 388:187-194, 2000.

28. Anderson GD, Hauser SD, McGarity KL, et al: Selectiveinhibition of cyclooxygenase (COX)-2 reverses inflammation and expression ofCOX-2 and interleukin 6 in rat adjuvant arthritis. J Clin Invest 97:2672-2679,1996.

29. Seed MP, Willoughby DA: COX-2, HO NO! Cyclooxygenase-2, hemeoxygenase and nitric oxide synthase: Their role and interactions ininflammation. BIRAs Symposium, Saint Bartholomew’s Hospital, London, 26 April1996. Inflamm Res 46:279-281, 1997.

30. Whelton A: COX-1 sparing and COX-2 inhibitory drugs: Therenal and hepatic safety and tolerability profiles of celecoxib. Am J Ther7:151-152, 2000.

31. Eschwege P, de Ledinghen V, Camilli T, et al:Cyclooxygenases [in French]. Presse Med 30:511-514, 2001.

32. Tanasescu S, Levesque H, Thuillez C: Pharmacology of aspirin[in French]. Rev Med Interne 21(suppl 1):18s-26s, 2000.

33. Rigas B, Goldman IS, Levine L: Altered eicosanoid levels inhuman colon cancer. J Lab Clin Med 122:518-523, 1993.

34. Kargman SL, O’Neill GP, Vickers PJ, et al: Expression ofprostaglandin G/H synthase-1 and -2 protein in human colon cancer. Cancer Res55:2556-2559, 1995.

35. Yang VW, Shields JM, Hamilton SR, et al: Size-dependentincrease in prostanoid levels in adenomas of patients with familial adenomatouspolyposis. Cancer Res 58:1750-1753, 1998.

36. Roller A, Bahr OR, Streffer J, et al: Selective potentiationof drug cytotoxicity by NSAID in human glioma cells: The role of COX-1 and MRP.Biochem Biophys Res Commun 259:600-605, 1999.

37. Karmali RA, Welt S, Thaler HT, et al: Prostaglandins inbreast cancer: Relationship to disease stage and hormone status. Br J Cancer48:689-696, 1983.

38. Fulton AM, Ownby HE, Frederick J, et al: Relationship oftumor prostaglandin levels to early recurrence in women with primary breastcancer: Clinical update. Invasion Metastasis 6:83-94, 1986.

39. Bennett A, Berstock DA, Raja B, et al: Survival time aftersurgery is inversely related to the amounts of prostaglandins extracted fromhuman breast cancers. Br J Pharmacol 66:451P, 1979.

40. Soslow RA, Dannenberg AJ, Rush D, et al: COX-2 is expressedin human pulmonary, colonic, and mammary tumors. Cancer 89:2637-2645, 2000.

41. Ochiai M, Oguri T, Isobe T, et al: Cyclooxygenase-2 (COX-2)mRNA expression levels in normal lung tissues and non-small cell lung cancers.Jpn J Cancer Res 90:1338-1343, 1999.

42. Hosomi Y, Yokose T, Hirose Y, et al: Increasedcyclooxygenase 2 (COX-2) expression occurs frequently in precursor lesions ofhuman adenocarcinoma of the lung. Lung Cancer 30:73-81, 2000.

43. Achiwa H, Yatabe Y, Hida T, et al: Prognostic significanceof elevated cyclooxygenase 2 expression in primary, resected lungadenocarcinomas. Clin Cancer Res 5:1001-1005, 1999.

44. El-Bayoumy K, Iatropoulos M, Amin S, et al: Increasedexpression of cyclooxygenase-2 in rat lung tumors induced by thetobacco-specific nitrosamine 4-(methylnitrosamino)-4-(3-pyridyl)-1-butanone: Theimpact of a high-fat diet. Cancer Res 59:1400-1403, 1999.

45. Wilson KT, Fu S, Ramanujam KS, et al: Increased expressionof inducible nitric oxide synthase and cyclooxygenase-2 in Barrett’s esophagusand associated adenocarcinomas. Cancer Res 58:2929-2934, 1998.

46. Shirvani VN, Ouatu-Lascar R, Kaur BS, et al: Cyclooxygenase2 expression in Barrett’s esophagus and adenocarcinoma: Ex vivo induction bybile salts and acid exposure. Gastroenterology 118:487-496, 2000.

47. Eberhart CE, Coffey RJ, Radhika A, et al: Upregulation ofcyclooxygenase 2 gene expression in human colorectal adenomas andadenocarcinomas. Gastroenterology 107:1183-1188, 1994.

48. Fournier DB, Gordon GB: COX-2 and colon cancer: Potentialtargets for chemoprevention. J Cell Biochem Suppl 34:97-102, 2000.

49. Khan KN, Knapp DW, Denicola DB, et al: Expression ofcyclooxygenase-2 in transitional cell carcinoma of the urinary bladder in dogs.Am J Vet Res 61:478-481, 2000.

50. Mohammed SI, Knapp DW, Bostwick DG, et al: Expression ofcyclooxygenase-2 (COX-2) in human invasive transitional cell carcinoma (TCC) ofthe urinary bladder. Cancer Res 59:5647-5650, 1999.

51. Komhoff M, Guan Y, Shappell HW, et al: Enhanced expressionof cyclooxygenase-2 in high-grade human transitional cell bladder carcinomas. AmJ Pathol 157:29-35, 2000.

52. Ristimaki A, Nieminen O, Saukkonen K, et al: Expression ofcyclooxygenase-2 in human transitional cell carcinoma of the urinary bladder. AmJ Pathol 158:849-853, 2001.

53. Gilhooly EM, Rose DP: The association between a mutated rasgene and cyclooxygenase-2 expression in human breast cancer cell lines. Int JOncol 15:267-270, 1999.

54. Liu XH, Rose DP: Differential expression and regulation ofcyclooxygenase-1 and -2 in two human breast cancer cell lines. Cancer Res56:5125-5127, 1996.

55. Rozic JG, Chakraborty C, Lala PK: Cyclooxygenase inhibitorsretard murine mammary tumor progression by reducing tumor cell migration,invasiveness and angiogenesis. Int J Cancer 93:497-506, 2001.

56. Hwang D, Scollard D, Byrne J, et al: Expression ofcyclooxygenase-1 and cyclooxygenase-2 in human breast cancer. J Natl Cancer Inst90:455-460, 1998.

57. Ryu HS, Chang KH, Yang HW, et al: High cyclooxygenase-2expression in stage IB cervical cancer with lymph node metastasis or parametrialinvasion. Gynecol Oncol 76:320-325, 2000.

58. Formenti S, Felix J, Salonga D, et al: Expression ofmetastases-associated genes in cervical cancers resected in the proliferativeand secretory phases of the menstrual cycle. Clin Cancer Res 6:4653-4657, 2000.

59. Kulkarni S, Rader JS, Zhang F, et al: Cyclooxygenase-2 isoverexpressed in human cervical cancer. Clin Cancer Res 7:429-434, 2001.

60. Tsujii M, Kawano S, DuBois RN: Cyclooxygenase-2 expressionin human colon cancer cells increases metastatic potential [published erratumappears in Cell 94: following 271, 1998]. Proc Natl Acad Sci U S A 94:3336-3340,1997.

61. Khuri FR, Wu H, Lee JJ, et al: Cyclooxygenase-2overexpression is a marker of poor prognosis in stage I non-small-cell lungcancer. Clin Cancer Res 7:861-867, 2001.

62. Sano H, Kawahito Y, Wilder RL, et al: Expression ofcyclooxygenase-1 and -2 in human colorectal cancer. Cancer Res 55:3785-3789,1995.

63. Li M, Lotan R, Levin B, et al: Aspirin induction ofapoptosis in esophageal cancer: A potential for chemoprevention. CancerEpidemiol Biomarkers Prev 9:545-549, 2000.

64. Zimmermann KC, Sarbia M, Weber AA, et al: Cyclooxygenase-2expression in human esophageal carcinoma. Cancer Res 59:198-204, 1999.

65. Li Z, Shimada Y, Kawabe A, et al: Suppression of N-nitrosomethylbenzylamine(NMBA)-induced esophageal tumorigenesis in F344 rats by JTE-522, a selectiveCOX-2 inhibitor. Carcinogenesis 22:547-551, 2001.

66. Murata H, Kawano S, Tsuji S, et al: Cyclooxygenase-2overexpression enhances lymphatic invasion and metastasis in human gastriccarcinoma. Am J Gastroenterol 94:451-455, 1999.

67. Akhtar M, Cheng Y, Magno RM, et al: Promoter methylationregulates Helicobacter pylori-stimulated cyclooxygenase-2 expression in gastricepithelial cells. Cancer Res 61:2399-2403, 2001.

68. Saika M, Ueyama T, Senba E: Expression of immediate earlygenes, HSP70, and COX-2 mRNAs in rat stomach following ethanol ingestion. DigDis Sci 45:2455-2462, 2000.

69. Koga H, Sakisaka S, Ohishi M, et al: Expression ofcyclooxygenase-2 in human hepatocellular carcinoma: Relevance to tumordedifferentiation. Hepatology 29:688-696, 1999.

70. Shiota G, Okubo M, Noumi T, et al: Cyclooxygenase-2expression in hepatocellular carcinoma. Hepatogastroenterology 46:407-412, 1999.

71. Bae SH, Jung ES, Park YM, et al: Expression ofcyclooxygenase-2 (COX-2) in hepatocellular carcinoma and growth inhibition ofhepatoma cell lines by a COX-2 inhibitor, NS-398. Clin Cancer Res 7:1410-1418,2001.

72. Wolff H, Saukkonen K, Anttila S, et al: Expression ofcyclooxygenase-2 in human lung carcinoma. Cancer Res 58:4997-5001, 1998.

73. Denkert C, Kobel M, Berger S, et al: Expression ofcyclooxygenase 2 in human malignant melanoma. Cancer Res 61:303-308, 2001.

74. Yip-Schneider MT, Barnard DS, Billings SD, et al:Cyclooxygenase-2 expression in human pancreatic adenocarcinomas. Carcinogenesis21:139-146, 2000.

75. Tucker ON, Dannenberg AJ, Yang EK, et al: Cyclooxygenase-2expression is upregulated in human pancreatic cancer. Cancer Res 59:987-990,1999.

76. Liu XH, Kirschenbaum A, Yao S, et al: Inhibition ofcyclooxygenase-2 suppresses angiogenesis and the growth of prostate cancer invivo. J Urol 164:820-825, 2000.

77. Lim JT, Piazza GA, Han EK, et al: Sulindac derivativesinhibit growth and induce apoptosis in human prostate cancer cell lines. BiochemPharmacol 58:1097-1107, 1999.

78. Kamijo T, Sato T, Nagatomi Y, et al: Induction of apoptosisby cyclooxygenase-2 inhibitors in prostate cancer cell lines. Int J Urol8:S35-S39, 2001.

79. Kirschenbaum A, Liotta DR, Yao S, et al: Immunohistochemicallocalization of cyclooxygenase-1 and cyclooxygenase-2 in the human fetal andadult male reproductive tracts. J Clin Endocrinol Metab 85:3436-3441, 2000.

80. Koki AT, Leahy KM, Masferrer JL: Potential utility of COX-2inhibitors in chemoprevention and chemotherapy. Expert Opin Investig Drugs8:1623-1638, 1999.

81. Chan G, Boyle JO, Yang EK, et al: Cyclooxygenase-2expression is up-regulated in squamous cell carcinoma of the head and neck.Cancer Res 59:991-994, 1999.

82. CenterWatch Clinical Trials Listing Service. Patient andgeneral resources. Available at: http://www.centerwatch.com. Accessed February8, 2002.

83. National Cancer Institute. PDQ® Clinical Trials DatabaseUser’s Guide. Available at: http://www.cancer.gov/cancer_information/doc.aspx?viewid=F2AFAEA4-64BD-4E44-B421-56026E252389.Accessed February 8, 2002.

84. Ryan CW, Stadler WM, Vogelzang NJ: Docetaxel and exisulindin hormone-refractory prostate cancer. Semin Oncol 28:56-61, 2001.

85. Jemal A, Thomas A, Murray T, et al: Cancer statistics, 2002.CA Cancer J Clin 52:23-47, 2002.

86. Dove-Edwin I, Thomas HJ: Review article: The prevention ofcolorectal cancer. Aliment Pharmacol Ther 15:323-336, 2001.

87. Arguedas MR, Heudebert GR, Wilcox CM: Surveillancecolonoscopy or chemoprevention with COX-2 inhibitors in average-riskpost-polypectomy patients: A decision analysis. Aliment Pharmacol Ther15:631-638, 2001.

88. Benson AB III, Desch CE, Flynn PJ, et al: 2000 update ofAmerican Society of Clinical Oncology colorectal cancer surveillance guidelines.J Clin Oncol 18:3586-3588, 2000.

89. Sheehan KM, Sheahan K, O’Donoghue DP, et al: Therelationship between cyclooxygenase-2 expression and colorectal cancer. JAMA282:1254-1257, 1999.

90. Oshima M, Sugiyama H, Kitagawa K, et al: APC gene messengerRNA: Novel isoforms that lack exon 7. Cancer Res 53:5589-5591, 1993.

91. Oshima M, Dinchuk JE, Kargman SL, et al: Suppression ofintestinal polyposis in APC delta716 knockout mice by inhibition ofcyclooxygenase-2 (COX-2). Cell 87:803-809, 1996.

92. Reddy BS: Novel approaches to the prevention of colon cancerby nutritional manipulation and chemoprevention. The Fourth DeWitt S. Goodmanlecture. Cancer Epidemiol Biomarkers Prev 9:239-247, 2000.

93. Thun MJ, Namboodiri MM, Calle EE, et al: Aspirin use andrisk of fatal cancer. Cancer Res 53:1322-1327, 1993.

94. Giovannucci E, Rimm EB, Stampfer MJ, et al: Aspirin use andthe risk for colorectal cancer and adenoma in male health professionals. AnnIntern Med 121:241-246, 1994.

95. Giovannucci E, Egan KM, Hunter DJ, et al: Aspirin and therisk of colorectal cancer in women. N Engl J Med 333:609-614, 1995.

96. Kune GA, Kune S, Watson LF: Colorectal cancer risk, chronicillnesses, operations, and medications: Case control results from the MelbourneColorectal Cancer Study. Cancer Res 48:4399-4404, 1988.

97. Rigau J, Pique JM, Rubio E, et al: Effects of long-termsulindac therapy on colonic polyposis. Ann Intern Med 115:952-954, 1991.

98. Labayle D, Fischer D, Vielh P, et al: Sulindac causesregression of rectal polyps in familial adenomatous polyposis. Gastroenterology101:635-639, 1991.

99. Giardiello FM, Hamilton SR, Krush AJ, et al: Treatment ofcolonic and rectal adenomas with sulindac in familial adenomatous polyposis. NEngl J Med 328:1313-1316, 1993.

100. Lipsky PE, Isakson PC: Outcome of specific COX-2 inhibitionin rheumatoid arthritis. J Rheumatol 24(Suppl 49):9-14, 1997.

101. Geis GS: Update on clinical developments with celecoxib, anew specific COX-2 inhibitor: What can we expect? J Rheumatol 26(Suppl56):31-36, 1999.

102. Steinbach G, Lynch PM, Phillips RK, et al: The effect ofcelecoxib, a cyclooxygenase-2 inhibitor, in familial adenomatous polyposis. NEngl J Med 342:1946-1952, 2000.

103. Marks R: The role of treatment of actinic keratoses in theprevention of morbidity and mortality due to squamous cell carcinoma. ArchDermatol 127:1031-1033, 1991.

104. Das M, Bickers DR, Santella RM, et al: Altered patterns ofcutaneous xenobiotic metabolism in UVB-induced squamous cell carcinoma in SKH-1hairless mice. J Invest Dermatol 84:532-536, 1985.

105. Nomura T, Nakajima H, Hongyo T, et al: Induction of cancer,actinic keratosis, and specific p53 mutations by UVB light in human skinmaintained in severe combined immunodeficient mice. Cancer Res 57:2081-2084,1997.

106. Berg RJ, de Vries A, van Steeg H, et al: Relativesusceptibilities of XPA knockout mice and their heterozygous and wild-typelittermates to UVB-induced skin cancer. Cancer Res 57:581-584, 1997.

107. Godar DE, Thomas DP, Miller SA, et al: Long-wavelength UVAradiation induces oxidative stress, cytoskeletal damage, and hemolysis.Photochem Photobiol 57:1018-1026, 1993.

108. O’Connor I, O’Brien N: Modulation of UVA light-inducedoxidative stress by beta-carotene, lutein, and astaxanthin in culturedfibroblasts. J Dermatol Sci 16:226-230, 1998.

109. Marks R, Rennie G, Selwood TS: Malignant transformation ofsolar keratoses to squamous cell carcinoma. Lancet 1:795-797, 1988.

110. Marks R, Rennie G, Selwood T: The relationship of basalcell carcinomas and squamous cell carcinomas to solar keratoses. Arch Dermatol124:1039-1042, 1988.

111. Radler-Pohl A, Sachsenmaier C, Gebel S, et al: UV-inducedactivation of AP-1 involves obligatory extranuclear steps including Raf-1kinase. EMBO J 12:1005-1012, 1993.

112. Engelberg D, Klein C, Martinetto H, et al: The UV responseinvolving the Ras signaling pathway and AP-1 transcription factors is conservedbetween yeast and mammals. Cell 77:381-390, 1994.

113. Huang C, Ma WY, Dong Z: The extracellular-signal-regulatedprotein kinases (Erks) are required for UV-induced AP-1 activation in JB6 cells.Oncogene 18:2828-2835, 1999.

114. Pentland AP, Schoggins JW, Scott GA, et al: Reduction ofUV-induced skin tumors in hairless mice by selective COX-2 inhibition.Carcinogenesis 20:1939-1944, 1999.

115. Dannhardt G, Ulbrich H: In-vitro test system for theevaluation of cyclooxygenase-1 (COX-1) and cyclooxygenase-2 (COX-2) inhibitorsbased on a single HPLC run with UV detection using bovine aortic coronaryendothelial cells (BAECs). Inflamm Res 50:262-269, 2001.

116. O’Neill GP, Kennedy BP, Mancini JA, et al: Selectiveinhibitors of COX-2. Agents Actions Suppl 46:159-168, 1995.

117. Athar M, An KP, Morel KD, et al: Ultraviolet B(UVB)-induced COX-2 expression in murine skin: An immunohistochemical study.Biochem Biophys Res Commun 280:1042-1047, 2001.

118. Kim Y, Fischer SM: Transcriptional regulation ofcyclooxygenase-2 in mouse skin carcinoma cells. Regulatory role of CCAAT/enhancer-bindingproteins in the differential expression of cyclooxygenase-2 in normal andneoplastic tissues. J Biol Chem 273:27686-27694, 1998.

119. Neufang G, Furstenberger G, Heidt M, et al: Abnormaldifferentiation of epidermis in transgenic mice constitutively expressingcyclooxygenase-2 in skin. Proc Natl Acad Sci U S A 98:7629-7634, 2001.

120. Vanderveen EE, Grekin RC, Swanson NA, et al: Arachidonicacid metabolites in cutaneous carcinomas. Evidence suggesting that elevatedlevels of prostaglandins in basal cell carcinomas are associated with anaggressive growth pattern. Arch Dermatol 122:407-412, 1986.

121. Chariyalertsak S, Sirikulchayanonta V, Mayer D, et al:Aberrant cyclooxygenase isozyme expression in human intrahepaticcholangiocarcinoma. Gut 48:80-86, 2001.

122. Marks F, Muller-Decker K, Furstenberger G: A causalrelationship between unscheduled eicosanoid signaling and tumor development:Cancer chemoprevention by inhibitors of arachidonic acid metabolism. Toxicology153:11-26, 2000.

123. Tsujii M, DuBois RN: Alterations in cellular adhesion andapoptosis in epithelial cells overexpressing prostaglandin endoperoxide synthase2. Cell 83:493-501, 1995.

124. Buckman SY, Gresham A, Hale P, et al: COX-2 expression isinduced by UVB exposure in human skin: Implications for the development of skincancer. Carcinogenesis 19:723-729, 1998.

125. Muller-Decker K, Kopp-Schneider A, Marks F, et al:Localization of prostaglandin H synthase isoenzymes in murine epidermal tumors:Suppression of skin tumor promotion by inhibition of prostaglandin H synthase-2.Mol Carcinog 23:36-44, 1998.

126. Kagoura M, Toyoda M, Matsui C, et al: Immunohistochemicalexpression of cyclooxygenase-2 in skin cancers. J Cutan Pathol 28:298-302, 2001.

127. Schottenfeld D: Patient risk factors and the detection ofearly cancer. Prev Med 1:335-351, 1972.

128. Dominioni L, Strauss GM, Imperatori A, et al: Screening forlung cancer. Chest Surg Clin N Am 10:729-736, 2000.

129. Strauss GM: Measuring effectiveness of lung cancerscreening: From consensus to controversy and back. Chest 112:216S-228S, 1997.

130. Strauss GM, Gleason RE, Sugarbaker DJ: Screening for lungcancer. Another look; a different view. Chest 111:754-768, 1997.

131. Heasley LE, Thaler S, Nicks M, et al: Induction ofcytosolic phospholipase A2 by oncogenic Ras in human non-small cell lung cancer.J Biol Chem 272:14501-14504, 1997.

132. Blaine SA, Wick M, Dessev C, et al: Induction of cPLA2 inlung epithelial cells and non-small-cell lung cancer is mediated by Sp1 andc-Jun. J Biol Chem 276:42737-42743, 2001.

133. Avis IM, Jett M, Boyle T, et al: Growth control of lungcancer by interruption of 5-lipoxygenase-mediated growth factor signaling. JClin Invest 97:806-813, 1996.

134. Castonguay A, Rioux N, Duperron C, et al: Inhibition oflung tumorigenesis by NSAIDs: A working hypothesis. Exp Lung Res 24:605-615,1998.

135. Bouchard L, Castonguay A: Inhibitory effects ofnonsteroidal anti-inflammatory drugs (NSAIDs) on the metabolism of4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone (NNK) in mouse lung explants.Drug Metab Dispos 21:293-298, 1993.

136. Moody TW, Leyton J, Zakowicz H, et al: Indomethacin reduceslung adenoma number in A/J mice. Anticancer Res 21:1749-1755, 2001.

137. Rioux N, Castonguay A: Prevention of NNK-induced lungtumorigenesis in A/J mice by acetylsalicylic acid and NS-398. Cancer Res58:5354-5360, 1998.

138. Lau SS, McMahon JB, McMenamin MG, et al: Metabolism ofarachidonic acid in human lung cancer cell lines. Cancer Res 47:3757-3762, 1987.

139. Masferrer JL, Leahy KM, Koki AT, et al: Antiangiogenic andantitumor activities of cyclooxygenase-2 inhibitors. Cancer Res 60:1306-1311,2000.

140. Leahy KM, Koki AT, Masferrer JL: Role of cyclooxygenases inangiogenesis. Curr Med Chem 7:1163-1170, 2000.

141. Bauer AK, Dwyer-Nield LD, Malkinson AM: High cyclooxygenase1 (COX-1) and cyclooxygenase 2 (COX-2) contents in mouse lung tumors.Carcinogenesis 21:543-550, 2000.

142. Sella A, Konichezky M, Flex D, et al: Low PSA metastaticandrogen-independent prostate cancer. Eur Urol 38:250-254, 2000.

143. Lara PN Jr, Meyers FJ: Treatment options inandrogen-independent prostate cancer. Cancer Invest 17:137-144, 1999.

144. Surya BV, Provet JA: Manifestations of advanced prostatecancer: Prognosis and treatment. J Urol 142:921-928, 1989.

145. da Fonseca FP, Lopes A, Melarato Junior WA, et al:Evaluation of prostate specific antigen in the prognosis of patients withadvanced prostate cancer. Rev Paul Med 116:1798-1802, 1998.

146. Hovsepian JA, Byar DP: Quantitative radiology for stagingand prognosis of patients with advanced prostatic carcinoma. Correlations withother pretreatment characteristics. Urology 14:145-150, 1979.

147. Liu XH, Yao S, Kirschenbaum A, et al: NS398, a selectivecyclooxygenase-2 inhibitor, induces apoptosis and downregulates bcl-2 expressionin LNCaP cells. Cancer Res 58:4245-4249, 1998.

148. Uotila P, Valve E, Martikainen P, et al: Increasedexpression of cyclooxygenase-2 and nitric oxide synthase-2 in human prostatecancer. Urol Res 29:23-28, 2001.

149. Kirschenbaum A, Klausner AP, Lee R, et al: Expression ofcyclooxygenase-1 and cyclooxygenase-2 in the human prostate. Urology 56:671-676,2000.

150. Yonemoto RH: Breast cancer in Japan and United States:Epidemiology, hormone receptors, pathology, and survival. Arch Surg115:1056-1062, 1980.

151. Warren JL, Feuer E, Potosky AL, et al: Use of Medicarehospital and physician data to assess breast cancer incidence. Med Care37:445-456, 1999.

152. MacMahon B, Trichopoulos D, Brown J, et al: Age atmenarche, probability of ovulation, and breast cancer risk. Int J Cancer29:13-16, 1982.

153. Newcomb PA, Storer BE, Longnecker MP, et al: Lactation anda reduced risk of premenopausal breast cancer. N Engl J Med 330:81-87, 1994.

154. Chie WC, Hsieh C, Newcomb PA, et al: Age at any full-termpregnancy and breast cancer risk. Am J Epidemiol 151:715-722, 2000.

155. Meiser B, Butow P, Barratt A, et al: Attitudes towardprophylactic oophorectomy and screening utilization in women at increased riskof developing hereditary breast/ovarian cancer. Gynecol Oncol 75:122-129, 1999.

156. Helzlsouer KJ: Bad news/good news: Information about breastcancer risk following prophylactic oophorectomy. J Natl Cancer Inst91:1442-1443, 1999.

157. Kundu N, Yang Q, Dorsey R, et al: Increasedcyclooxygenase-2 (COX-2) expression and activity in a murine model of metastaticbreast cancer. Int J Cancer 93:681-686, 2001.

158. Giercksky KE: COX-2 inhibition and prevention of cancer.Best Pract Res Clin Gastroenterol 15:821-833, 2001.

159. Subbaramaiah K, Chung WJ, Michaluart P, et al: Resveratrolinhibits cyclooxygenase-2 transcription and activity in phorbol ester-treatedhuman mammary epithelial cells. J Biol Chem 273:21875-21882, 1998.

160. Valentin-Opran A, Eilon G, Saez S, et al: Estrogens andantiestrogens stimulate release of bone resorbing activity by cultured humanbreast cancer cells. J Clin Invest 75:726-731, 1985.

161. Rudland PS: Stem cells and the development of mammarycancers in experimental rats and in humans. Cancer Metastasis Rev 6:55-83, 1987.

162. Morecki S, Yacovlev E, Gelfand Y, et al: Induction ofantitumor immunity by indomethacin. Cancer Immunol Immunother 48:613-620, 2000.

163. Jang M, Pezzuto JM: Cancer chemopreventive activity ofresveratrol. Drugs Exp Clin Res 25:65-77, 1999.

164. Ratnasinghe D, Phang JM, Yeh GC: Differential expressionand activity of phosphatases and protein kinases in adriamycin sensitive andresistant human breast cancer MCF-7 cells. Int J Oncol 13:79-84, 1998.

165. Fulton AM, Zhang SZ, Chong YC: Role of the prostaglandin E2receptor in mammary tumor metastasis. Cancer Res 51:2047-2050, 1991.

166. Harris RE, Alshafie GA, Abou-Issa H, et al: Chemopreventionof breast cancer in rats by celecoxib, a cyclooxygenase-2 inhibitor. Cancer Res60:2101-2103, 2000.

167. Alshafie GA, Abou-Issa HM, Seibert K, et al:Chemotherapeutic evaluation of celecoxib, a cyclooxygenase-2 inhibitor, in a ratmammary tumor model. Oncol Rep 7:1377-1381, 2000.

168. Pausawasdi A, Suntharabha S, Tanwatcharabhan P, et al:Clinical study of gastric cancer. J Med Assoc Thai 63:655-661, 1980.

169. Pongchairerks P, Chalermsanyakorn P, Tanjapatkul M:Occurrence of intestinal metaplasia of the stomach in Thai patients withgastritis, benign ulcer, and gastric cancer. J Surg Oncol 43:101-105, 1990.

170. Guadagni S, Walters CL, Smith PL, et al: N-nitrosocompounds in the gastric juice of normal controls, patients with partialgastrectomies, and gastric cancer patients. J Surg Oncol 63:226-233, 1996.

171. Mirvish SS: Role of N-nitroso compounds (NOC) andN-nitrosation in etiology of gastric, esophageal, nasopharyngeal, and bladdercancer and contribution to cancer of known exposures to NOC. Cancer Lett93:17-48, 1995.

172. Kaneko S, Yoshimura T: Time trend analysis of gastriccancer incidence in Japan by histological types, 1975-1989. Br J Cancer84:400-405, 2001.

173. Sihvo EI, Salminen JT, Ramo OJ, et al: The epidemiology ofoesophageal adenocarcinoma: Has the cancer of gastric cardia an influence on therising incidence of oesophageal adenocarcinoma? Scand J Gastroenterol35:1082-1086, 2000.

174. Uemura N, Okamoto S, Yamamoto S, et al: Helicobacter pyloriinfecion and the development of gastric cancer. N Engl J Med 345:784-789, 2001.

175. Chung AY, Chow PK, Yu WK, et al: Prevalence of Helicobacterpylori in gastric cancer in a South-East Asian population by 14C-urea breathtest. ANZ J Surg 71:574-576, 2001.

176. Moshkovitz M, Niv Y: Helicobacter pylori and gastric cancer[in Hebrew]. Harefuah 140:738-743, 2001.

177. Leung WK, To KF, Ng YP, et al: Association betweencyclo-oxygenase-2 overexpression and missense p53 mutations in gastric cancer.Br J Cancer 84:335-339, 2001.

178. Mungan NA, Aben KK, Schoenberg MP, et al: Genderdifferences in stage-adjusted bladder cancer survival. Urology 55:876-880, 2000.

179. Shaham J, Melzer A, Kaufman Z, et al: Occupation andbladder cancer [in Hebrew]. Harefuah 131:382-386, 456, 1996.

180. Barbone F, Franceschi S, Talamini R, et al: Occupation andbladder cancer in Pordenone (north-east Italy): A case-control study. Int JEpidemiol 23:58-65, 1994.

181. Steinmaus CM, Nunez S, Smith AH: Diet and bladder cancer: Ameta-analysis of six dietary variables. Am J Epidemiol 151:693-702, 2000.

182. Hebert JR, Miller DR: A cross-national investigation ofdiet and bladder cancer. Eur J Cancer 30A:778-784, 1994.

183. Vodvarka P, Jancova J: Bacterial infection of urine inpatients with bladder cancer. Neoplasma 35:243-250, 1988.

184. Kantor AF, Hartge P, Hoover RN, et al: Urinary tractinfection and risk of bladder cancer. Am J Epidemiol 119:510-515, 1984.

185. Castelao JE, Yuan JM, Gago-Dominguez M, et al:Non-steroidal anti-inflammatory drugs and bladder cancer prevention. Br J Cancer82:1364-1369, 2000.

186. Mannetje A, Kogevinas M, Chang-Claude J, et al: Smoking asa confounder in case-control studies of occupational bladder cancer in women. AmJ Ind Med 36:75-82, 1999.

187. Piazza GA, Thompson WJ, Pamukcu R, et al: Exisulind, anovel proapoptotic drug, inhibits rat urinary bladder tumorigenesis. Cancer Res61:3961-3968, 2001.

188. Tramontana M, Catalioto RM, Lecci A, et al: Role ofprostanoids in the contraction induced by a tachykinin NK2 receptor agonist inthe hamster urinary bladder. Naunyn Schmiedebergs Arch Pharmacol 361:452-459,2000.

189. Grubbs CJ, Lubet RA, Koki AT, et al: Celecoxib inhibitsN-butyl-N-(4-hydroxybutyl)-nitrosamine-induced urinary bladder cancers in maleB6D2F1 mice and female Fischer-344 rats. Cancer Res 60:5599-5602, 2000.

190. Romanenko A, Morimura K, Wanibuchi H, et al: Increasedoxidative stress with gene alteration in urinary bladder urothelium after theChernobyl accident. Int J Cancer 86:790-798, 2000.

191. Flora-Filho R, Camara-Lopes LH, Zilberstein B: Histologicalcriteria of esophagitis in the gastroesophageal reflux disease. Reevaluation ofthe sensitivity of the esophageal 24-hours pHmetry [in Portuguese]. ArqGastroenterol 37:197-202, 2000.

192. Dent J, Bremner CG, Collen MJ, et al: Barrett’soesophagus. J Gastroenterol Hepatol 6:1-22, 1991.

193. Spechler SJ, Lee E, Ahnen D, et al: Long-term outcome ofmedical and surgical therapies for gastroesophageal reflux disease: Follow-up ofa randomized controlled trial. JAMA 285:2331-2338, 2001.

194. Spechler SJ: Pathogenesis and epidemiology of Barrettesophagus [in German]. Chirurg 65:84-87, 1994.

195. Reid BJ, Levine DS, Longton G, et al: Predictors ofprogression to cancer in Barrett’s esophagus: Baseline histology and flowcytometry identify low- and high-risk patient subsets. Am J Gastroenterol95:1669-1676, 2000.

196. Pech O, Gossner L, May A, et al: Management of Barrett’soesophagus, dysplasia, and early adenocarcinoma. Best Pract Res ClinGastroenterol 15:267-284, 2001.

197. Sharma P: Controversies in Barrett’s esophagus:Management of high-grade dysplasia. Semin Gastrointest Dis 12:26-32, 2001.

198. Sharma P, Morales TG, Bhattacharyya A, et al: Dysplasia inshort-segment Barrett’s esophagus: A prospective 3-year follow-up. Am JGastroenterol 92:2012-2016, 1997.

199. Panjehpour M, Overholt BF, Haydek JM, et al: Results ofphotodynamic therapy for ablation of dysplasia and early cancer in Barrett’sesophagus and effect of oral steroids on stricture formation. Am J Gastroenterol95:2177-2184, 2000.

200. Terry P, Lagergren J, Wolk A, et al: Reflux-inducingdietary factors and risk of adenocarcinoma of the esophagus and gastric cardia.Nutr Cancer 38:186-191, 2000.

201. Eckardt VF, Kanzler G, Bernhard G: Life expectancy andcancer risk in patients with Barrett’s esophagus: A prospective controlledinvestigation. Am J Med 111:33-37, 2001.

202. Eloubeidi MA, Wallace MB, Hoffman BJ, et al: Predictors ofsurvival for esophageal cancer patients with and without celiac axislymphadenopathy: Impact of staging endosonography. Ann Thorac Surg 72:212-220 [incldiscussion], 2001.

203. Salazar JD, Doty JR, Lin JW, et al: Does cell typeinfluence post-esophagectomy survival in patients with esophageal cancer? DisEsophagus 11:168-171, 1998.

204. Swisher SG, Putnam JB Jr: Influence of lymph nodedissection on survival in esophageal cancer. J Surg Oncol 69:117-118, 1998.

205. Kawahara K, Maekawa T, Okabayashi K, et al: The number oflymph node metastases influences survival in esophageal cancer. J Surg Oncol67:160-163, 1998.

206. Li M, Wu X, Xu XC: Induction of apoptosis bycyclo-oxygenase-2 inhibitor NS398 through a cytochrome C-dependent pathway inesophageal cancer cells. Int J Cancer 93:218-223, 2001.

207. Kandil HM, Tanner G, Smalley W, et al: Cyclooxygenase-2expression in Barrett’s esophagus. Dig Dis Sci 46:785-789, 2001.

208. Lord RV, Danenberg KD, Danenberg PV: Cyclooxygenase-2 inBarrett’s esophagus, Barrett’s adenocarcinomas, and esophageal SCC: Readyfor clinical trials. Am J Gastroenterol 94:2313-2315, 1999.

209. Wallner B, Sylvan A, Janunger KG, et al:Immunohistochemical markers for Barrett’s esophagus and associations toesophageal Z-line appearance. Scand J Gastroenterol 36:910-915, 2001.

210. Longnecker DS, Karagas MR, Tosteson TD, et al: Racialdifferences in pancreatic cancer: comparison of survival and histologic types ofpancreatic carcinoma in Asians, blacks, and whites in the United States.Pancreas 21:338-343, 2000.

211. Burris H, Storniolo AM: Assessing clinical benefit in thetreatment of pancreas cancer: Gemcitabine compared to 5-fluorouracil. Eur JCancer 33 (suppl) 1:18S-22S, 1997.

212. Rothenberg ML, Moore MJ, Cripps MC, et al: A phase II trialof gemcitabine in patients with 5-FU-refractory pancreas cancer. Ann Oncol7:347-353, 1996.

213. Ding XZ, Tong WG, Adrian TE: Blockade of cyclooxygenase-2inhibits proliferation and induces apoptosis in human pancreatic cancer cells.Anticancer Res 20:2625-2631, 2000.

214. Yip-Schneider MT, Sweeney CJ, Jung SH, et al: Cell-cycleeffects of nonsteroidal anti-inflammatory drugs and enhanced growth inhibitionin combination with gemcitabine in pancreatic carcinoma cells. J Pharmacol ExpTher 298:976-985, 2001.

215. Thun MJ: NSAID use and decreased risk of gastrointestinalcancers. Gastroenterol Clin North Am 25:333-348, 1996.

216. Greenberg HE, Gottesdiener K, Huntington M, et al: A newcyclooxygenase-2 inhibitor, rofecoxib (Vioxx), did not alter the antiplateleteffects of low-dose aspirin in healthy volunteers. J Clin Pharmacol40:1509-1515, 2000.

217. Logan RF, Little J, Hawtin PG, et al: Effect of aspirin andnonsteroidal anti-inflammatory drugs on colorectal adenomas: Case-control studyof subjects participating in the Nottingham faecal occult blood screeningprogramme. BMJ 307:285-289, 1993.

218. Boolbol SK, Dannenberg AJ, Chadburn A, et al:Cyclooxygenase-2 overexpression and tumor formation are blocked by sulindac in amurine model of familial adenomatous polyposis. Cancer Res 56:2556-2560, 1996.

219. DuBois RN: COX-2 in large bowel cancer: A one-sided story.Gut 45:636-637, 1999.

220. DuBois RN, Giardiello FM, Smalley WE: Nonsteroidalanti-inflammatory drugs, eicosanoids, and colorectal cancer prevention.Gastroenterol Clin North Am 25:773-791, 1996.

221. Ishikawa H: Chemoprevention of familial cancer [inJapanese]. Gan To Kagaku Ryoho 24:951-957, 1997.

222. Vadlamudi R, Mandal M, Adam L, et al: Regulation ofcyclooxygenase-2 pathway by HER2 receptor. Oncogene 18:305-314, 1999.

Related Videos
Increasing screening for younger individuals who are at risk of colorectal cancer may help mitigate the rising early incidence of this disease.
Laparoscopy may reduce the degree of pain or length of hospital stay compared with open surgery for patients with colorectal cancer.
Rahul Gosain, MD; Sam Klempner, MD; and Rohit Gosain, MD, presenting slides
Rahul Gosain, MD; Sam Klempner, MD; and Rohit Gosain, MD, presenting slides
Rahul Gosain, MD; Sam Klempner, MD; and Rohit Gosain, MD, presenting slides
Rahul Gosain, MD; Sam Klempner, MD; and Rohit Gosain, MD, presenting slides
Rahul Gosain, MD; Sam Klempner, MD; and Rohit Gosain, MD, presenting slides
Tailoring neoadjuvant therapy regimens for patients with mismatch repair deficient gastroesophageal cancer represents a future step in terms of research.
Not much is currently known about the factors that may predict pathologic responses to neoadjuvant immunotherapy in this population, says Adrienne Bruce Shannon, MD.
The toxicity profile of tislelizumab also appears to look better compared with chemotherapy in metastatic esophageal squamous cell carcinoma.