当前位置: 首页 > 期刊 > 《临床肿瘤学》 > 2005年第4期 > 正文
编号:11329580
COX-2: A Molecular Target for Colorectal Cancer Prevention
http://www.100md.com 《临床肿瘤学》
     the Vanderbilt-Ingram Cancer Center, Nashville, TN

    ABSTRACT

    Cyclooxygenase (COX), a key enzyme in the prostanoid biosynthetic pathway, has received considerable attention due to its role in human cancers. Observational and randomized controlled studies in many different population cohorts and settings have demonstrated protective effects of nonsteroidal anti-inflammatory drugs (NSAIDs; the inhibitors of COX activity) for colorectal cancers (CRCs). COX-2, the inducible isoform of cyclooxygenase, is overexpressed in early and advanced CRC tissues, which portends a poor prognosis. Experimental studies have thus identified important mechanisms and pathways by which COX-2 plays an important role in carcinogenesis. Selective COX-2 inhibitors have been approved for use as adjunctive therapy for patients with familial polyposis. The role of COX-2 inhibitors is currently being evaluated for use in wider populations.

    CANCER PREVALENCE

    In 2000, estimates for global statistics on cancer rates indicate that there were 10.1 million new cases of cancer, 6.2 million deaths resulting from cancer, and 22 million people living with the disease.1 Often, 10 to 25 years may pass before a neoplastic cell has gone through enough cell divisions and molecular changes to reach a clinically detectable lesion. Cancer can occur at any age, but risk increases in the elderly. An aging population means that cancer incidence will increase; projected rates for 2050 estimate 57% of those cancer cases detected will be in people 65 years of age or older.1 The impact of cancer in our aging world population illustrates the critical need to identify and apply effective preventive measures.

    Colorectal cancer is ranked the third most common form of cancer worldwide in terms of incidence (estimated to result in 945,000 new cases, 9.4% of the world total) and mortality (492,000 deaths, 7.9% of the total) in 2000.1 High incidence rates are found in Western Europe, North America, and Australia, with the lowest rates found in the sub-Saharan Africa. The National Cancer Institute estimates 146,940 new colorectal cases will be diagnosed in 2004, with an estimated 56,730 deaths in the United States.2 Screening procedures such as colonoscopy can reduce the risk of colorectal cancer mortality by 50%.3 However, routine screening procedures are received by only approximately 50% of Americans age 50 years or older.4 In the United Kingdom, there are 25,000 to 30,000 new cases of colorectal cancer each year.5 Of these new cases, 17,000 result in death, of which 15,000 are in those age 60 years and older. In the United Kingdom, a considerable number of patients present with metastatic disease, and have less than a 50% chance of survival despite aggressive adjuvant therapy.6 Based on data from the United States, average-risk patients account for approximately 75% of all colorectal cancers and include persons older than 50 years with no other known risk factor; moderate-risk patients account for 15% to 20% of all colorectal cancers and are identified as having a diagnosis of colorectal adenomatous polyps, a personal history of resected colorectal cancer, or a positive family history of colorectal adenomatous polyps or cancer; and high-risk patients account for 5% to 15% of all colorectal cancers and include those with familial cancer syndrome—familial adenomatous polyposis (FAP), hereditary nonpolyposis colorectal cancer or long-standing inflammatory bowel disease.7 The majority of colorectal cancers are thus nonhereditary and sporadic, which makes early detection important. Unfortunately, it is not yet economically feasible to identify all individuals in the general population who are at the highest risk for developing colorectal cancer.8 Therefore, most patients present clinically with advanced disease. Advanced metastatic disease remains incurable for the most part, with current treatment regimens having little effect on 5-year survival. Hence, another important approach to reduce the overall morbidity and mortality of colorectal cancer involves its prevention.

    CHEMOPREVENTION AND NSAIDs

    General preventative measures so far can be divided into those that depend on an alteration of lifestyle and those that depend on direct treatment with nutraceutical and/or pharmaceutical agents. Lifestyle alterations have an impact on colorectal cancer; evidence suggests the intake of high fiber and fish oils and a reduction in saturated fats is sufficient to reduce overall risk.9,10 Unfortunately, evidence from cardiovascular disease studies has indicated the difficulty in devising dietary strategies with high compliance that could be applied on a large scale.10

    The International Agency for Research on Cancer (IARC; part of WHO) uses the term chemoprevention to refer to interventions with pharmaceuticals, vitamins, minerals or other chemicals (natural or synthetic) at any of the multiple stages of carcinogenesis to reduce cancer incidence.11 Since colorectal cancer develops over a 10- to 25-year period, a large-scale clinical trial using colon cancer incidence or mortality as an end point would be time consuming, expensive, and difficult to instigate.8 Adenomatous polyps are used as a surrogate end point in some clinical trials in order to determine the efficacy of NSAIDs as chemopreventive agents.7 Prompted by experimental studies, several factors were taken into account when evaluating the antitumor actions of NSAIDs in clinical trials: source of the study population, outcome of interest (adenomatous polyp, cancer incidence, mortality), and reliability of the information regarding NSAID exposure. Prevention of colon cancer by NSAIDs has created considerable interest because of the enormous amount of clinical experience with NSAIDs in the general population.12

    It is estimated that at least 20 to 30 billion aspirin tablets are purchased annually in the United States alone, and that 1% to 2% of the world population consumes at least one aspirin tablet daily.13 At present there are at least 15 NSAIDs on the market, being prescribed at a rate of 70 million NSAID prescriptions per year for individuals suffering from chronic inflammation and pain; for example 10 to 15 million individuals afflicted with rheumatoid arthritis (RA) and osteoarthritis in the United States.13 However, use of aspirin and other NSAIDs accounts for $5 to $10 billion in hospitalization charges and lost work time, and 26,000 deaths per year. The adverse effects attributable to NSAIDs are thought to arise from the inhibition of the constitutive isoform of the target enzyme, cyclooxygenase (COX-1). COX-1 has been hypothesized to function as a housekeeping gene responsible for example, for the production of cytoprotective prostaglandins (PGs) in the gastrointestinal (GI) tract; whereas COX-2 is an immediate-early gene thought to be involved in inflammation, mitogenesis and/or specialized signal transduction mechanisms.14 Hence, the anti-inflammatory, analgesic, and antipyretic properties of NSAIDs are attributed to the inhibition of COX-2, and the GI ulcerations induced by NSAIDs are thought to be due to the inhibition of COX-1. This relationship between NSAIDs, COX-1, and COX-2 has provided the impetus to develop NSAIDs devoid of the adverse effects associated with traditional NSAIDs but which retain COX-2 selectivity to inhibit inflammatory parameters. However, we now know that this original hypothesis was overly simplistic and that even the use of highly selective COX-2 inhibitors is associated with some GI toxicity. Despite these limitations, targeted COX-2 inhibition may have an important role for use in the chemoprevention of colorectal cancer.

    BIOCHEMISTRY OF COX-2

    Arachidonic acid (5,8,11,14-eicosatetraenoic acid) is derived either directly from the diet or from modification of linoleic acid, and resides mostly in an esterified form in cell membranes (Fig 1). Metabolites of the arachidonic acid cascade include products of COX (or PGH synthase [prostaglandin endoperoxide synthase; prostaglandin G/H synthase]; EC1.14.99.1) isoenzymes which utilize membrane-bound arachidonic acid as its substrate. COX has to compete for substrate with lipoxygenase and cytochrome P450 enzymes, the metabolites of which are collectively known as eicosanoids, meaning "twenty" because they are derived from the 20 carbon arachidonic acid. Arachidonic acid is released via the action of cellular phospholipases, liberating free substrate to the active site of COX enzymes.15 COX is a bifunctional (it sequentially catalyzes cyclooxygenase and peroxidase reactions), membrane-bound hemeprotein that catylases the bisoxygenation of arachidonic acid to form PGG2 and the peroxidative reduction of PGG2 to form PGH2.16 COX inserts two molecules of oxygen into the sn-position of arachidonic acid, producing the unstable PGG2, which is rapidly reduced by the hydroperoxidase activity of the COX enzyme to PGH2, the precursor for all prostaglandins. PGH2 is unstable and is subsequently converted to PGs by a family of prostaglandin synthases that are specific for each prostaglandin. The end products of this pathway generated by various cell types reflect the particular cellular abundance of synthase enzymes that utilize PGH2 as a substrate (Fig 1). 17

    NSAIDs are chemically unrelated compounds that share a common therapeutic action. The best known of these are aspirin, ibuprofen, piroxicam, indomethacin, and sulindac. Aspirin is an irreversible inhibitor of the COX active site, covalently modifying the COX protein by acetylating a single serine residue in the substrate-binding channel, blocking the approach of arachidonic acid. Indomethacin, piroxicam, ibuprofen, and sulindac are competitive inhibitors that noncovalently bind to the protein in the substrate channel. The structural differences between COX-1 and COX-2 have been exploited by pharmaceutical companies to develop selective COX-2 inhibitors.18-20 The active site of COX-2 is larger than that of COX-1 with an additional side-pocket18 that can accommodate larger structures. The replacement of isoleucine by valine in the binding site of COX-1, for example, removes constriction in the mouth of this secondary pocket, which results in access by more bulky molecules.21 Hence, compounds specifically designed to bind and fit this additional space are usually potent and selective COX-2 inhibitors.

    Of the selective COX-2 inhibitors, three in the United States and global markets include celecoxib (Searle/Pharmacia/Pfizer, New London, CT), rofecoxib (Merck, Whitehouse, NJ), and valdecoxib (Pharmacia/Pfizer). All have been approved for use in the management of pain and inflammation associated with RA, osteoarthritis, and primary dysmenorrhea. Not only have all three drugs shown similar efficacy to the traditional NSAIDs in treating acute and chronic inflammatory conditions, but these agents have also significantly improved GI tolerability and platelet safety profiles.

    EVIDENCE OF NSAIDs AS CHEMOPREVENTIVE AGENTS

    During the development of colorectal cancer, PG production is significantly increased in malignant tissue22,23 compared with adjacent normal mucosa, particularly PGs of the E series.24 Sulindac treatment of patients with FAP provided the first insights into the potential chemopreventive actions of NSAIDs. Waddell & Loughry25 showed that sulindac treatment for 1 year could decrease polyp multiplicity and induce regression of polyps in patients with FAP and Gardner’s syndrome. A follow-up report 6 years later demonstrated that cessation of treatment resulted in adenoma recurrence,26 suggesting a cytostatic rather than a cytotoxic effect. Since these studies, a large body of evidence has now emerged revealing a 40% to 50% reduction in colorectal cancer in individuals taking NSAIDs regularly, either in the context of sporadic colorectal cancer or in FAP patients27 (Tables 1 and 2, respectively). These studies comprise observational as well as randomized controlled studies in many different population cohorts and settings demonstrating strong protective properties of NSAIDs; mostly aspirin, but including ibuprofen, indomethacin, and piroxicam against the risk of developing colorectal cancer and reducing mortality related to colorectal cancer (Table 1). Thun et al28 conducted the largest observational study to date on the use of aspirin and colon cancer, in which aspirin use reduced the relative risk of cancer mortality by approximately 40% when consumed more than 16 times per month for at least 1 year. Two recent randomized, double-blind, controlled studies have both confirmed these observational studies, showing aspirin can protect against new adenoma formation thus delaying adenoma development.29,30 With regard to optimal dose regimen, Baron et al30 showed that aspirin only taken at a dose of 81 mg/day and not at the higher dose of 325 mg/day conferred a reduction in advanced adenoma incidence. The reason why no significant antitumor effect in the 325 mg/day group was seen is not certain, because Sandler et al29 showed 325 mg/day was effective. Moreover, patient populations with FAP are sensitive to selective COX-2 inhibition31,32 (Table 2), such that a significant reduction in the number and burden of colorectal polyps provided supporting data for the approval of celecoxib as adjunctive therapy to endoscopic surveillance and surgery for FAP patients. The effect of selective COX-2 inhibition for the chemoprevention of common sporadic colorectal cancers remains to be investigated but is the focus of many international ongoing clinical trials, as well as the effectiveness of other NSAIDs (Table 3). An issue of the long-term safety of selective COX-2 inhibitors also needs to be considered in these future trials, since a potential prothrombotic effect of these inhibitors is of intense debate.33 Randomized controlled studies (VIGOR34 [Vioxx Gastrointestinal Outcomes Research] and CLASS35 [Celecoxib Long-term Arthritis Safety Study]) revealed good upper GI protection compared with other NSAIDs when administered at therapeutic doses. However, the potential inhibition of endothelial cell–derived COX-2 activity and subsequent PGI2 production, which may shift the homeostatic balance toward more thromboxane A2 (TXA2) effects, promoting platelet aggregation and leading to increased cardiovascular events remains controversial.36 These drugs would require clinical and mechanistic safety assessments in patients with known atherosclerotic disease. In this regard, a recent observational study using 54,475 older adults (age 65 years or older) showed a significantly elevated relative risk of acute myocardial infarction (AMI) with rofecoxib > 25 mg/day compared with celecoxib (odds ratio [OR], 1.24; 95% CI, 1.05 to 1.46) or no NSAID treatment (OR, 1.14; 95% CI, 1.00 to 1.31) only during the first 90 days of drug exposure.37 These increased risks were observed in a dose-specific manner, whereas celecoxib was not associated with an increased risk of AMI. Similarly, a large population-based study showed individuals age 66 years or older were at higher risk of admission for congestive heart failure when taking rofecoxib compared with those taking either nonselective NSAIDs or celecoxib.38

    Patients with RA who regularly take NSAIDs also have a reduced incidence of colorectal cancers.39 However, as mentioned in preceding paragraphs, arthritis appears to be a disease that predisposes the patient to the development of ulcers.40 Trujillo et al41 outlined a risk-benefit scenario based on the known effects of aspirin on cancer prevention and intestinal complications. Of 100 individuals older than 50 years, as many as 40 will develop adenomatous polyps, and five will develop colon cancer in their lifetime. If these 40 polyp-bearing individuals were treated with aspirin as a chemopreventive agent, 1.2 patients per year will develop a serious intestinal complication as a result, rising to six individuals within 5 years. Although aspirin could prevent 2.5 colon cancer–related deaths, the risk of a serious ulcer will have exceeded the potential lifetime benefit of preventing five colon cancers. These estimates indicate the need to determine whether long-term usage of selective COX-2 inhibitors would retain their antitumor activities while being devoid of GI complications.

    EXPERIMENTAL EVIDENCE

    In regard to experimental evidence, nonselective NSAIDs have been known to inhibit cancer formation in rodent models of colorectal cancer since the 1980s, and the development of selective COX-2 inhibitors has shown equal promise in rodent models and in models of colonic carcinogenesis, genetically modified tumor development, or human cells grown in nude mice.42-52 For example, selective COX-2 inhibitors can reduce tumor formation in Apc716 (adenomatous polyposis coli716) mice46 and in Min (multiple intestinal neoplasms) mice.47 However, sometimes the doses needed to exhibit an antitumor effect were in excess of anti-inflammatory doses. For example, Reddy et al, 48 using celecoxib in azoxymethane (AOM)-induced mouse carcinogenesis, showed the doses needed to inhibit aberrant crypt foci (ACF; preneoplastic lesions) incidence and multiplicity reached plasma drug concentrations of 3.5 μg/mL (approximately 9 μmol/L; administered in the diet at 1,500 parts per million), whereas the plasma concentration of 0.3 μg/mL (0.8 μmol/L) was sufficient to inhibit adjuvant-induced arthritis. Lower doses, reaching 0.5 μg/mL (1.3 μmol/L) plasma levels did not affect ACF.49 This drug achieves IC50 (inhibitory concentration 50%) values for the inhibition of COX-1 and COX-2 in insect cells in vitro of 13 μmol/L and 0.04 μmol/L, respectively.48 This group (Reddy et al) extended these findings further by demonstrating the antitumor effect of celecoxib in the same model, with reduced tumor multiplicity, incidence and volume,44 even when administered during the promotion or progression stages.50 This finding suggests that celecoxib can retard the growth and/or development of pre-existing neoplastic lesions. Again, the lowest plasma concentration was 2.3 μg/mL (5.8 μmol/L) and the highest was 4.3 μg/mL (11.3 μmol/L). Similarly, nimesulide has shown to dose-dependently inhibit tumor multiplicity, incidence, and size in the AOM-induced colonic cancer model in mice, using doses equivalent to 21 and 39 mg/kg in the diet,51 and inhibit polyp multiplicity in Min mice at the equivalent dose of 39 mg/kg.43 Sawaoka et al52 also showed that COX-2–expressing and non–COX-expressing tumor cells were sensitive to NS-398 through the inhibition of angiogenesis at 10, 30, and 100 mg/kg orally in mice. These doses are sufficiently high to be nonselective, which suggests that mechanisms other than COX-2 inhibition may be involved.53,54 It has also been shown that celecoxib at 1,250 mg/kg of diet fed to nude mice reaches mean plasma levels of 2 to 3 μmol/L concentration, whereas 2 μmol/L in vitro did not affect cell toxicity.55 Interestingly, recent clinical studies indicted that celecoxib caused a statistically significant inhibition of FAP polyps at a dose of 800 mg/day but not at the recommended anti-inflammatory dose of 100 to 200 mg bid.31 The question of COX-independent effects occurring in association with the antitumor properties of selective COX-2 inhibitors at high doses has been reviewed elsewhere56 and includes inhibition of phosphodiesterases and IKK (inhibitor of nuclear factor kappa kinase, beta subunit) kinase by sulindac sulfone; celecoxib can inhibit 3-phosphoinositide-dependent kinase (PDK1) downstream of Akt activity, whereas many NSAIDs can inhibit PPAR (peroxisome proliferator activated receptor delta) activity and matrix metalloproteinases (MMPs), inhibit and/or activate various protein kinases, and downregulate -catenin.

    More compelling evidence for the role of COX-2 in the formation of colorectal cancers has been provided by genetic studies in mice. Genetic inhibition of COX-2 negated the development of colonic polyps in the APC716 murine model.42 Also, the maintenance of COX-2–bearing xenografts requires host COX-2 but not COX-1 expression for the production of angiogenic stimuli.57 In a model of mammary tumorigenesis, overexpression of COX-2 alone in mammary glands was sufficient to induce cellular transformation and resulted in the formation of breast carcinomas.58 Similarly, COX-2 overexpression in basal keratinocytes induces hyperplasia, skin premalignancy59 and skin malignancy.60 These studies provide strong evidence supporting the use of selective COX-2 inhibitors for the chemoprevention of some epithelial-derived cancers.

    The potential role of COX-1 has not been neglected. Recent studies have shown an equal importance of this isoform in colorectal and skin tumorigenesis. A recent comprehensive study identified COX-1 expression detected in human colorectal adenomas as well as COX-2, which highlights the importance of using several analytic techniques to draw conclusions about COX expression.61 Deficiency in either the COX-1 or COX-2 gene was sufficient to inhibit the formation of intestinal polyps in Min mice.62 Selective COX-1 inhibition can reduce ACF in AOM-induced colorectal carcinogenesis.63 Similarly, the genetic deficiency of either COX-1 or COX-2 can prevent terminal differentiation of initiated keratinocytes and reduce skin tumorigenesis.64 It has been postulated that COX-1 may protect cells from the initiated DNA-damaging effects during the early tumorigenesis, whereas COX-2 contributes to tumor promotion, particularly after the loss of the APC gene in colorectal tumorigenesis.65 Cervical66 and ovarian67 cancers are other epithelial-derived malignancies where studies have recently highlighted a role for COX-1 in tumor formation through the production of angiogenic stimuli. Endothelial cell–derived COX-1 has also been shown to be important in angiogenesis,68 such that overexpression leads to malignant transformation in endothelial cells.69 The role of COX-1 in vascular biology was recently extended because selective inhibition of COX-1 prevented malignant cells from primary breast tumors metastasizing to the lung.70 These results highlight the fact that both COX-1 and COX-2 carry out identical reactions leading to the production of PGs such as PGE2.

    REGULATION OF COX-2 EXPRESSION

    COX has a molecular weight of 72 kDa71,72 made up of a dimeric complex of two polypeptides,73,74 each of which requires one molecule of heme for maximal catalytic activity.71 COX-1 was cloned in 1988 by three independent groups.75-77 In 1990, COX gene expression was shown to be increased, producing elevated levels of PGs from mammalian cells in vitro.78,79 In the following year, an inducible isoform of COX was identified using mitogen-stimulated chicken fibroblasts,80 phorbol-ester81 and serum-stimulated82 murine fibroblasts (termed TIS10). This COX was a 4.1 kb mRNA transcript that encoded a protein with 59% sequence homology to the ovine COX. The noninducible or constitutive isoenzyme was renamed COX-1 and the new inducible isoenzyme was designated COX-2.

    The two existing isoforms of COX differ in their protein sequence and regulation of and sensitivity to NSAIDs. The major difference between the two isoenzymes is that COX-1 contains a 17-amino-acid sequence at its amino terminus that is absent in COX-2, whereas COX-2 has an additional 18-amino-acid sequence at its carboxy terminus. However, the amino sequences responsible for catalytic activity are well conserved between the two enzymes. COX-1 and COX-2 isoenzyme regulation differs at the level of both transcription and translation. The COX-1 gene is made up of larger introns compared with the COX-2 gene, accounting for lengths of 22 kb and 8 kb, respectively. The genes are transcribed into mRNA products of 2.8 kb and 4.1 kb for COX-1 and COX-2 respectively. Typical of an immediate-early gene, COX-2 contains several copies of Shaw-Kamen’s sequences in the 3'-untranslated region which confers enhanced mRNA degradation. Thus COX-2 transcripts degrade quickly and their stabilization by pro-inflammatory cytokines contributes to increased COX-2 protein levels during the inflammatory response. In contrast, corticosteroids specifically destabilize COX-2 mRNA and/or interfere with translation without affecting basal COX-1 expression (reviewed in Isakson et al83). This regulation of COX-2 at the transcriptional and translational level implies that fine control of expression is important.

    Oncogenes, growth factors, cytokines, chemotherapeutics, and tumor promoters are among some of the stimuli that induce COX-2 expression. COX-2 induction has been associated with various premalignant and malignant lesions of epithelial origin in organs such as colon, lung, breast, prostate, bladder, stomach, and esophagus (reviewed in Dannenberg et al84). Although the underlying mechanisms of this elevated COX-2 expression in cancer is not known, key cis-acting elements within the start 5' region upstream of the COX-2 gene have been shown to play an important role in the regulation of COX-2. The promoter region of COX-2 consists of many transcription factor binding sites, such as nuclear factor B (NFB), nuclear factor of interleukin-6 (NF-IL-6), cyclic adenosine monophosphate (cAMP) response element (CRE) and hypoxia-inducible factor-1 (HIF-1); most of which are known to be involved in the upregulation of COX-2 by inflammatory stimuli or tumor promoters (reviewed in Dempke et al85). It has been hypothesized that the upregulation of COX-2 prolongs the survival of abnormal cells and thereby favors the accumulation of sequential genetic changes, which increases the risk of tumorigenesis (Fig 2). 86

    However, the over-expression of COX-2 protein in colorectal cancers is likely to occur via several different mechanisms involving complex signaling pathways, since transformed epithelial cells, as well as stromal cells, have been shown to express increased levels of COX-2. For example, the COX-2 gene may be regulated by hypoxia via activation of NFB in human vascular endothelial cells.87 While COX-2 overexpression in cancerous epithelial cells may be induced through the target of normal APC (a member of the Wnt [a term derived from the genes wingless and Int-1] signaling pathway)—the -catenin oncoprotein (reviewed in Bright-Thomas et al88; Fig 3). The principle role of wild-type APC involves the binding and degradation of -catenin. -catenin may exist either bound to membrane or free in the cytosol. Membrane-bound -catenin functions through the transmembrane glycoprotein E-cadherin adhesion molecule, which is responsible for epithelial cell-cell and cell-matrix adhesion and migration. Under normal conditions, cytosolic -catenin is degraded by the action of wild-type APC protein complexed with two accessory proteins (axin and glycogen synthase kinase [GSK]-3). However, mutations in the APC gene, which encodes the vital -catenin-binding regions, renders the APC protein truncated and unable to control cytosolic -catenin. This leads to free -catenin, where it translocates to the nucleus and acts as a transcription factor in concert with the T-cell factor-4 (TCF-4; also known as lymphoid enhancer factor [LEF-1]) complex, a feature associated with progression along the adenoma-carcinoma sequence. The formation of transcriptionally active -catenin and TCF complexes binds to TCF-4 consensus sites in specific target genes, including c-myc, cyclin D1, PPAR, and COX-2. TCF binding sites (TBS) have recently been identified in the COX-2 promoter region,89 such that on modulation, COX-2 can be downregulated by wild-type APC induction and upregulated by nuclear accumulation of -catenin in the presence of mutant APC. Interestingly, Oshima et al42 demonstrated during tumorigenesis in Apc716 mice that induced COX-2 expression occurred either coincidentally or slightly after the loss in wild-type APC allele. This would suggest a direct role of APC loss in COX-2 overexpression. Also, PPAR expression has been recently shown to be upregulated during tumorigenesis in Min mice, and its activation can accelerate tumor growth.90

    COX-2 expression is also regulated by transcriptional and post-transcriptional mechanisms, although transcriptional regulation may play a more important role in the COX-2 expression of human colon cancer cells.91,92 Ras is a member of the small GTPase family and GDP/GTP-regulated signaling molecules. It is found in a mutated form in approximately 50% of large colorectal adenomas and results in the constitutive activation of Ras followed by multiple downstream signaling pathways (Fig 4). The Ras/Rac-1/MEKK/ c-Jun NH2-terminal kinase (JNK) pathway activates COX-2 transcription via the CRE element. Ras/Raf-1/ mitogen-activated protein kinase (MAPK) –extracellular regulated kinase (ERK) kinase (MEK)/ERK operates downstream of growth factor receptor activation (eg, PDGF, serum) to induce the COX-2 promoter and stabilize COX-2 mRNA (reviewed in Dixon93). The p38 MAPK pathway, usually associated with cellular stress and proinflammatory stimuli, has also been found to transform intestinal epithelial cells, in part through COX-2 mRNA stabilization. Finally, sequential activation of phosphatidylinositol 3'-kinase (PI3-K)/ 3-phosphoinositide-dependent kinase (PDK)/Akt/ protein kinase B (PKB) cascade by various growth factors promotes cell survival through the inhibition of apoptosis. This pathway has also been found to be constitutively activated colon cancer cell lines and mediates post-transcriptional stabilization of COX-2 mRNA in intestinal epithelial cells. Recently, evidence has emerged that the Wnt- and Ras-signaling pathways may actually cooperate in the regulation of COX-2 expression in colonic tumor cells.88

    EXPRESSION OF COX-2

    Increased levels of COX-2 mRNA and protein are found in both premalignant and malignant tissues from epithelial and nonepithelial tumors. Gastric, hepatic, esophageal, pancreatic, head and neck, lung, breast, bladder, cervical, endometrial, skin, and colorectal cancers have all shown elevated COX-2 expression when compared with nonmalignant tissue (reviewed in Koki and Masferrer94). COX-2 overexpression in the primary tumor has been associated with a poor clinical outcome. The fact that increased COX-2 was seen in premalignant tissue again illustrates that activation of COX-2 may be an early event during tumorigenesis. Recently, increased COX-2 has also been identified in pituitary tumors95 and in harmartomatous polyps in Peutz-Jeghers syndrome.96 An important issue was recently highlighted when assessing these histologic findings; the treatment status of the patients from which tumor samples are included in these studies needs to be considered because COX expression can be induced by chemotherapeutic agents and hypoxia also. Therefore, it is not known whether expression is related to the treatment regimen, surgery-induced hypoxia, or the primary disease process.97

    Concerning colorectal cancer, Eberhart et al98 were the first to identify significant elevations of COX-2 expression in 85% and 50% of human colorectal carcinomas and adenomas respectively. In normal intestinal tissue, immunolocalization studies have shown the expression of both COX-1 and COX-2 in mucosal epithelial cells, mononuclear cells, vascular endothelial cells and smooth muscle.99 However, in both human and animal models of colorectal cancers, COX-2 expression is dramatically increased in malignancies when compared with adjacent normal mucosa.93 COX-1 expression appears to remain unaltered, or even reduced.100 This has been confirmed in more recent studies of human colonic cancers.93 In vivo models of colorectal cancer have displayed a similar COX expression pattern to that seen in human samples. In the AOM-induced colonic cancer model,101,102 Min mice103 and Apc716 mice,42 increased mRNA and protein levels of COX-2 (but not COX-1) were measured in colorectal tumors compared with normal adjacent tissue. In human colorectal cancer, a positive correlation was made between NFB (cytoplasmic and nuclear) and COX-2 coexpression in neoplastic epithelial cells but not in epithelial cells from normal mucosa, eluding to the possible mechanism of how the COX-2 gene is being activated.104

    However, the source and type of cells responsible for this COX-2 expression has raised many questions. The tumor stroma is now known to contribute to elevated COX-2 expression in colorectal cancers.105 Chapple et al106 found increased levels of COX-2 localized in a subpopulation of macrophages, while neoplastic cells expressed COX-2 in more advanced tumors. Positively stained macrophages for COX-2 in human colorectal adenomas in addition to tumor cells, and in Min mice has also been demonstrated.107,108 Oshima et al42 in their initial study showed localized COX-2 protein to interstitial cells and not malignant cells within small polyps of heterozygote Apc716 mice. In a subsequent study, these authors extended their findings by showing fibroblasts and endothelial cells also express high levels of COX-2 in polyps from FAP patients and Apc716 mice.109 Thus both host and tumor cells may contribute to the production of PGs within the tumor microenvironment and the subsequent development of tumor growth. In keeping with this issue, activated macrophages can transform rat intestinal epithelial cells into a more neoplastic phenotype in a paracrine fashion via a COX-2-dependent mechanism.110 This implies that the increase in COX-2 expression in intestinal interstitial cells is an early event which influences surrounding epithelial cells to transform toward malignancy. Conversely, colonic cancer cells have been shown to induce COX-2 expression in monocytes/macrophages via mucin production.111

    COX-2–expressing cancer cells induce an angiogenic response in endothelial cells, while endothelial cell-derived COX-1 may also play a significant role in the angiogenic response. Tsujii et al68 showed COX-2 over-expressing cells were sensitive to NS-398 and indomethacin, while COX-null cells were only sensitive to indomethacin. This suggests that the COX-2 from the tumor cells induces the production of proangiogenic factors, whereas COX-1 from the endothelial cells can induce their own tube formation. This was later confirmed in vivo.52,112 Also, fibroblasts are instrumental as a source of angiogenic stimuli for the maintenance to tumor growth57,113 such that an intracrine effect in which COX-2 in stromal cells produce PGs, which increases VEGF within the same cell to have an effect on endothelial cells, may explain why host COX-2 is needed for sustained tumor growth.114 Finally, high microvessel density, IL-8 expression, and numbers of infiltrating macrophages were correlated with poor prognosis and survival time in patients with non–small-cell lung cancer.115 In this study, when macrophages were cocultured with lung tumor cells, tumor cell–derived IL-8 levels were inhibited by aspirin, indomethacin, celecoxib, dexamethasone, and other anti-inflammatory agents. Whether COX-2 (or COX-1)-derived PGs promote tumor growth through the transformed epithelial cells, endothelial cells, fibroblasts or macrophages, or a combination of all these cell types is not yet known. It seems likely that the localization of COX-1 or COX-2 expression changes during tumor progression.

    ANGIOGENESIS AND METASTATIC POTENTIAL

    It is now well established that tumors are dependent on a constant blood supply via neoangiogenesis,116 such that without the induction of angiogenesis, tumors remain only 1 to 2 mm3 in size. During this avascular period, tumors do not increase in size due to the higher numbers of cells undergoing apoptosis compared with those that are proliferating.117 For tumors to grow beyond this size, the balance between naturally occurring angiogenic stimulators and inhibitors is altered in favor of the induction of angiogenesis.116 Tumors acquire this angiogenic switch116 through genetic mutations of various oncogenes and tumor suppressor genes.118-121 The angiogenic component in tumors has considerable potential as a therapeutic target.116

    PGs have been known to contribute to tumor development through their role in angiogenesis.122,123 Peterson124,125 was the first to demonstrate that diclofenac could inhibit the growth of transplantable tumors via the inhibition of angiogenesis. Since the identification of COX-2, its role in angiogenesis has been clearly demonstrated in the progression of colorectal cancer. COX-2–transfected human colon cells possess increased metastatic potential.126 COX-2 has been shown to be involved in models of angiogenesis,127-131 and selective inhibition of COX-2 can block tumor growth via an antiangiogenic mechanism.127,132,133 Selective COX-2 inhibition can also inhibit human colorectal cancer xenografts from metastasizing to the liver.134 Recently, COX-2 has been shown to be involved in the angiogenic switch in Apc716 mice by regulating angiogenesis through the induction of vascular endothelial growth factor (VEGF) in an EP2-receptor–dependent manner.135 However, host-derived PGE2 signaling through the EP3 receptor appears to regulate tumor angiogenesis through the induction of VEGF. 136 In human sporadic colorectal adenomas, interstitial cell–derived COX-2 expression (predominantly from macrophages) is associated with increased angiogenesis.137 A mechanism whereby PGs influence angiogenesis may be through the action of v3 integrin. The integrin v3 is necessary for endothelial cell survival during angiogenesis since its antagonism can induce endothelial cells to undergo apoptosis, leaving pre-existing vascular beds unaffected.138 Similarly, selective COX-2 (and not COX-1) inhibitors can reverse the v3 integrin–mediated angiogenic response.139 In a related study, COX-1 as well as COX-2 induction can modulate angiogenic response through v3 integrin.140 Also, v3 integrin antagonism can reduce liver metastasis and angiogenesis, and improve survival in mice injected with colon cancer cells.141

    IMMUNE MODULATION

    PGs can modulate immune function through a variety of mechanisms,142 however PGE2 in particular appears to have a clearly defined role in the regulation of humoral and cellular immunity.143 Tumor cells are vulnerable to attack by immune effector cells such as lymphokine-activated killer cell, cytotoxic T-lymphocytes (CTLs) and macrophages (reviewed in Young144). However, impairment in function of tumor-infiltrating lymphocytes, circulating T-cells, and macrophages has been demonstrated in cancer patients.145,146 One of the mechanisms attributed to the immune impairment in cancer patients has been increased PGE2 production. Immunosuppression occurs in tissues where PGE2 is high,147,148 and PGE2 can negatively regulate T-lymphocyte proliferation, cytokine production, and cytotoxicity.149 Specifically, PGE2 can mediate the suppression of macrophage-derived, TNF-induced colon cancer cell killing via the suppression of IL-10150 and increased IL-12.151 The tumor microenvironment is predominantly polarized toward TH2-like or immunosuppressive immune responses, and is a common feature of premalignant and malignant diseases (reviewed in O’Byrne et al152). Hence, indomethacin has been shown to increase the number of CTLs in cancer patients153 and stimulate mononuclear cells to increase their tumoricidal capacity.154 More recently, it has been demonstrated that this impaired mononuclear cell function can be restored with treatment with a selective COX-2 inhibitor.155 In fact, COX-2 is known to mediate the imbalance between IL-10 and IL-12 in favor of IL-10 production.156-158 Selective COX-2 inhibition serves to restore the tumor-induced imbalance between IL-10 and IL-12 and promotes antitumor responses in lung cancer159 and metastasis in colorectal cancer.160

    APOPTOSIS

    Tumor growth is significantly influenced by the relative balance between proliferation and apoptosis.117 Decreased cell proliferation alone is not sufficient to inhibit colonic tumor growth.161 Sinicrope et al162 studied the proliferative and apoptotic indices during the adenoma-carcinoma sequence of human colonic tumor samples and found a decrease in apoptosis during tumor progression. Compared with adenomas, lower rates of apoptosis with no change in proliferation were found in colonic carcinomas. This creates an imbalance between apoptosis and proliferation that favors net tumor growth. Colon carcinomas associated with a low apoptotic index have a poor prognosis.163,164 NSAID-induced apoptosis is a phenomenon that has been attributed to the antitumor action of NSAIDs in colon cancer in vivo.45,165-167 COX-2 is known to induce transformation in normal intestinal epithelial cells, resulting in increased bcl-2 expression, increased avidity to extracellular matrix components, reduced transforming growth factor (TGF) receptor expression168 and a prolongation in the cell cycle G1 phase with increased cyclin D expression.169 COX-2 may prevent apoptosis not only by generating the antiapoptotic products PGE2170 and PGI2,171 but also by removing a proapoptotic substrate, arachidonic acid.172 Since PGE2 can activate MAPK activity,130 it is conceivable that Ras signal transduction pathways are involved.85 The overall decrease in apoptosis of these cells may predispose to exhibit an accumulation in sequential genetic changes that can increase the risk of tumorigenesis.

    THE FUTURE IS COMBINATION THERAPY

    COX-2 is now considered as a viable target for chemotherapy,84,85,173,174 especially since selective COX-2 inhibitors have improved safety profiles compared with nonselective NSAIDs and are much less toxic than most chemotherapeutic agents. It is estimated that approximately 26,000 persons die and another 260,000 are hospitalized per year as a consequence of adverse effects of NSAIDs in the United States alone.175 In 2000, approximately 60% of the $4.8 billion total cost for NSAID prescriptions were for selective COX-2 inhibitors (reviewed in Laine176). If long-term use of selective COX-2 inhibitors can prove to be less toxic than traditional NSAIDs, these drugs can have a 50% reduction in adverse effects. Unfortunately, the long-term safety issue with COX-2 inhibitors was recently thrown into question. Clinical trials evaluating the use of selective COX-2 inhibitors for the prevention of polyp recurrence have stopped due to increased cardiovascular and thrombotic adverse effects observed with their long-term use.176a These results dramatically alter the utility of using selective COX-2 inhibitors for chemoprevention in a low-risk population. However, additional short-term studies are underway to evaluate the utility of these inhibitors in adjuvant treatment regimens and in patients with a very high risk of developing colorectal cancer (Table 3).

    For use in colorectal cancer, the role of COX-2 inhibitors in prevention still awaits assessment because their recent release means epidemiologic data will be available only in a few years (Table 3). On the basis of the cumulative experimental and clinical evidence, selective COX-2 inhibitors may be considered as cotherapeutic agents (reviewed in Blanke173). In most preclinical studies, selective COX-2 inhibitors reduce the growth rate of established tumors rather than causing tumor regression.84 Therefore, selective COX-2 inhibition has recently been investigated in various models of colorectal cancer in combination with inducible nitric oxide inhibitors,49 MMP inhibitors,177 ornithine decarboxylase inhibitors (difluoromethylornithine),178 epidermal growth factor receptor kinase inhibitors,179-181 radiotherapy,182 and chemotherapy.160 A small phase II study183 of 10 patients with metastatic colorectal cancer using rofecoxib and chemotherapy did not demonstrate increased efficacy; however, ongoing trials may further clarify the role of these agents for the treatment of colorectal cancer (Table 3). The majority of colorectal cancers over-express COX-2, and this increased expression is thought to inhibit apoptosis, induce angiogenesis, subvert the immune system, and promote tumor invasion. An understanding of the mechanism(s) whereby COX-2 mediates these phenomena awaits further studies.

    Authors' Disclosures of Potential Conflicts of Interest

    The following authors or their immediate family members have indicated a financial interest. No conflict exists for drugs or devices used in a study if they are not being evaluated as part of the investigation. Consultant/Advisory Role: Raymond N. DuBois, Pharmacia, Novartis. Honoraria: Raymond N. DuBois, Pfizer. For a detailed description of these categories, or for more information about ASCO’s conflict of interest policy, please refer to the Author Disclosure Declaration form and the Disclosures of Potential Conflicts of Interest section of Information for Contributors found in the front of every issue.

    NOTES

    Authors' disclosures of potential conflicts of interest are found at the end of this article.

    REFERENCES

    Parkin DM: Global cancer statistics in the year 2000. Lancet Oncol 2:533-543, 2001

    National Institute for Cancer (NCI), Surveillance and End Results (SEER) Cancer Statistics Review: 1975-2001. http://seer.cancer.gov/csr/1975_2001/

    Walsh JM, Terdiman JP: Colorectal cancer screening: Scientific review. JAMA 289:1288-1296, 2003

    Smith RA, Cokkinides V, Eyre HJ: American Cancer Society guidelines for the early detection of cancer, 2004. CA Cancer J Clin 54:41-52, 2004

    Somerville K, Noble G: Non-steroidal anti-inflammatory drugs: Is the balance shifting Age Ageing 26:417-422, 1997

    Chan TA: Nonsteroidal anti-inflammatory drugs, apoptosis, and colon-cancer chemoprevention. Lancet Oncol 3:166-174, 2002

    O'Shaughnessy JA, Kelloff GJ, Gordon GB, et al: Treatment and prevention of intraepithelial neoplasia: An important target for accelerated new agent development. Clin Cancer Res 8:314-346, 2002

    Dubois RN, Smalley WE: Cyclooxygenase, NSAIDs, and colorectal cancer. J Gastroenterol 31:898-906, 1996

    Reddy BS: Chemoprevention of colon cancer by dietary fatty acids. Cancer Metastasis Rev 13:285-302, 1994

    Langman M, Boyle P: Chemoprevention of colorectal cancer. Gut 43:578-585, 1998

    The International Agency for Research on Cancer (IARC), World Health Organization. http://www.iarc.fr/

    Dubois RN: Nonsteroidal anti-inflammatory drug use and sporadic colorectal adenomas. Gastroenterology 108:1310-1314, 1995

    Lichtenberger LM: Where is the evidence that cyclooxygenase inhibition is the primary cause of nonsteroidal anti-inflammatory drug (NSAID)-induced gastrointestinal injury Topical injury revisited. Biochem Pharmacol 61:631-637, 2001

    Williams CS, Dubois RN: Prostaglandin endoperoxide synthase: Why two isoforms Am J Physiol 270:G393-G400, 1996

    Smith WL, Marnett LJ, DeWitt DL: Prostaglandin and thromboxane biosynthesis. Pharmacol Ther 49:153-179, 1991

    Hamberg M, Samuelsson B: Detection and isolation of an endoperoxide intermediate in prostaglandin biosynthesis. Proc Natl Acad Sci U S A 70:899-903, 1973

    Herschman HR: Regulation of prostaglandin synthase-1 and prostaglandin synthase-2. Cancer Metastasis Rev 13:241-256, 1994

    Luong C, Miller A, Barnett J, et al: Flexibility of the NSAID binding site in the structure of human cyclooxygenase-2. Nat Struct Biol 3:927-933, 1996

    Kurumbail RG, Stevens AM, Gierse JK, et al: Structural basis for selective inhibition of cyclooxygenase-2 by anti-inflammatory agents. Nature 384:644-648, 1996

    Picot D, Loll PJ, Garavito RM: The x-ray crystal structure of the membrane protein prostaglandin H2 synthase-1. Nature 367:243-249, 1994

    Wong E, Bayly C, Waterman HL, et al: Conversion of prostaglandin G/H synthase-1 into an enzyme sensitive to PGHS-2-selective inhibitors by a double His513 –> Arg and Ile523 –> val mutation. J Biol Chem 272:9280-9286, 1997

    Jaffe BM, Parker CW, Philpott GW: Immunochemical measurement of prostaglandin or prostaglandin-like activity from normal and neoplastic cultured tissue. Surg Forum 22:90-92, 1971

    Bennett A, Tacca MD, Stamford IF, et al: Prostaglandins from tumours of human large bowel. Br J Cancer 35:881-884, 1977

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

    Waddell WR, Loughry RW: Sulindac for polyposis of the colon. J Surg Oncol 24:83-87, 1983

    Waddell WR, Ganser GF, Cerise EJ, et al: Sulindac for polyposis of the colon. Am J Surg 157:175-179, 1989

    Smalley WE, DuBois RN: Colorectal cancer and nonsteroidal anti-inflammatory drugs. Adv Pharmacol 39:1-20, 1997

    Thun MJ, Namboodiri MM, Heath CW Jr: Aspirin use and reduced risk of fatal colon cancer. N Engl J Med 325:1593-1596, 1991

    Sandler RS, Halabi S, Baron JA, et al: A randomized trial of aspirin to prevent colorectal adenomas in patients with previous colorectal cancer. N Engl J Med 348:883-890, 2003

    Baron JA, Cole BF, Sandler RS, et al: A randomized trial of aspirin to prevent colorectal adenomas. N Engl J Med 348:891-899, 2003

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

    Phillips RK, Wallace MH, Lynch PM, et al: A randomised, double blind, placebo controlled study of celecoxib, a selective cyclooxygenase 2 inhibitor, on duodenal polyposis in familial adenomatous polyposis. Gut 50:857-860, 2002

    Mukherjee D, Nissen SE, Topol EJ: Risk of cardiovascular events associated with selective COX-2 inhibitors. JAMA 286:954-959, 2001

    Bombardier C, Laine L, Reicin A, et al: Comparison of upper gastrointestinal toxicity of rofecoxib and naproxen in patients with rheumatoid arthritis: VIGOR Study Group. N Engl J Med 343:1520-1528, 2000

    Silverstein FE, Faich G, Goldstein JL, et al: Gastrointestinal toxicity with celecoxib vs nonsteroidal anti-inflammatory drugs for osteoarthritis and rheumatoid arthritis: The CLASS study—A randomized controlled trial: Celecoxib Long-term Arthritis Safety Study. JAMA 284:1247-1255, 2000

    Mukherjee D, Topol EJ: Cox-2: Where are we in 2003 Cardiovascular risk and Cox-2 inhibitors. Arthritis Res Ther 5:8-11, 2003

    Solomon DH, Schneeweiss S, Glynn RJ, et al: Relationship between selective cyclooxygenase-2 inhibitors and acute myocardial infarction in older adults. Circulation 109:2068-2073, 2004

    Mamdani M, Juurlink DN, Lee DS, et al: Cyclo-oxygenase-2 inhibitors versus non-selective non-steroidal anti-inflammatory drugs and congestive heart failure outcomes in elderly patients: A population-based cohort study. Lancet 363:1751-1756, 2004

    Gridley G, McLaughlin JK, Ekbom A, et al: Incidence of cancer among patients with rheumatoid arthritis. J Natl Cancer Inst 85:307-311, 1993

    O'Brien WM: Adverse reactions to nonsteroidal anti-inflammatory drugs: Diclofenac compared with other nonsteroidal anti-inflammatory drugs. Am J Med 80:70-80, 1986

    Trujillo MA, Garewal HS, Sampliner RE: Nonsteroidal antiinflammatory agents in chemoprevention of colorectal cancer: At what cost Dig Dis Sci 39:2260-2266, 1994

    Oshima M, Dinchuk JE, Kargman SL, et al: Suppression of intestinal polyposis in Apc delta716 knockout mice by inhibition of cyclooxygenase 2 (COX-2). Cell 87:803-809, 1996

    Nakatsugi S, Fukutake M, Takahashi M, et al: Suppression of intestinal polyp development by nimesulide, a selective cyclooxygenase-2 inhibitor, in Min mice. Jpn J Cancer Res 88:1117-1120, 1997

    Kawamori T, Rao CV, Seibert K, et al: Chemopreventive activity of celecoxib, a specific cyclooxygenase-2 inhibitor, against colon carcinogenesis. Cancer Res 58:409-412, 1998

    Sheng H, Shao J, Kirkland SC, et al: Inhibition of human colon cancer cell growth by selective inhibition of cyclooxygenase-2. J Clin Invest 99:2254-2259, 1997

    Oshima M, Murai N, Kargman S, et al: Chemoprevention of intestinal polyposis in the Apcdelta716 mouse by rofecoxib, a specific cyclooxygenase-2 inhibitor. Cancer Res 61:1733-1740, 2001

    Jacoby RF, Seibert K, Cole CE, et al: The cyclooxygenase-2 inhibitor celecoxib is a potent preventive and therapeutic agent in the min mouse model of adenomatous polyposis. Cancer Res 60:5040-5044, 2000

    Reddy BS, Rao CV, Seibert K: Evaluation of cyclooxygenase-2 inhibitor for potential chemopreventive properties in colon carcinogenesis. Cancer Res 56:4566-4569, 1996

    Rao CV, Indranie C, Simi B, et al: Chemopreventive properties of a selective inducible nitric oxide synthase inhibitor in colon carcinogenesis, administered alone or in combination with celecoxib, a selective cyclooxygenase-2 inhibitor. Cancer Res 62:165-170, 2002

    Reddy BS, Hirose Y, Lubet R, et al: Chemoprevention of colon cancer by specific cyclooxygenase-2 inhibitor, celecoxib, administered during different stages of carcinogenesis. Cancer Res 60:293-297, 2000

    Fukutake M, Nakatsugi S, Isoi T, et al: Suppressive effects of nimesulide, a selective inhibitor of cyclooxygenase-2, on azoxymethane-induced colon carcinogenesis in mice. Carcinogenesis 19:1939-1942, 1998

    Sawaoka H, Tsuji S, Tsujii M, et al: Cyclooxygenase inhibitors suppress angiogenesis and reduce tumor growth in vivo. Lab Invest 79:1469-1477, 1999

    Tegeder I, Pfeilschifter J, Geisslinger G: Cyclooxygenase-independent actions of cyclooxygenase inhibitors. FASEB J 15:2057-2072, 2001

    Raz A: Is inhibition of cyclooxygenase required for the anti-tumorigenic effects of nonsteroidal, anti-inflammatory drugs (NSAIDs) In vitro versus in vivo results and the relevance for the prevention and treatment of cancer. Biochem Pharmacol 63:343-347, 2002

    Williams CS, Watson AJ, Sheng H, et al: Celecoxib prevents tumor growth in vivo without toxicity to normal gut: Lack of correlation between in vitro and in vivo models. Cancer Res 60:6045-6051, 2000

    Soh JW, Weinstein IB: Role of COX-independent targets of NSAIDs and related compounds in cancer prevention and treatment, in Dannenberg AJ, DuBois RN (eds): COX-2: A New Target for Cancer Prevention and Treatment. New York, NY, Karger, 2003, pp 259-281

    Williams CS, Tsujii M, Reese J, et al: Host cyclooxygenase-2 modulates carcinoma growth. J Clin Invest 105:1589-1594, 2000

    Liu CH, Chang SH, Narko K, et al: Overexpression of cyclooxygenase-2 is sufficient to induce tumorigenesis in transgenic mice. J Biol Chem 276:18563-18569, 2001

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

    Muller-Decker K, Neufang G, Berger I, et al: Transgenic cyclooxygenase-2 overexpression sensitizes mouse skin for carcinogenesis. Proc Natl Acad Sci U S A 99:12483-12488, 2002

    Chapple KS, Scott N, Guillou PJ, et al: Analysis of cyclooxygenase expression in human colorectal adenomas. Dis Colon Rectum 45:1316-1324, 2002

    Chulada PC, Thompson MB, Mahler JF, et al: Genetic disruption of Ptgs-1, as well as Ptgs-2, reduces intestinal tumorigenesis in Min mice. Cancer Res 60:4705-4708, 2000

    Kitamura T, Kawamori T, Uchiya N, et al: Inhibitory effects of mofezolac, a cyclooxygenase-1 selective inhibitor, on intestinal carcinogenesis. Carcinogenesis 23:1463-1466, 2002

    Tiano HF, Loftin CD, Akunda J, et al: Deficiency of either cyclooxygenase (COX)-1 or COX-2 alters epidermal differentiation and reduces mouse skin tumorigenesis. Cancer Res 62:3395-3401, 2002

    Smith WL, Langenbach R: Why there are two cyclooxygenase isozymes. J Clin Invest 107:1491-1495, 2001

    Sales KJ, Katz AA, Howard B, et al: Cyclooxygenase-1 is up-regulated in cervical carcinomas: Autocrine/paracrine regulation of cyclooxygenase-2, prostaglandin e receptors, and angiogenic factors by cyclooxygenase-1. Cancer Res 62:424-432, 2002

    Gupta RA, Tejada LV, Tong BJ, et al: Cyclooxygenase-1 is overexpressed and promotes angiogenic growth factor production in ovarian cancer. Cancer Res 63:906-911, 2003

    Tsujii M, Kawano S, Tsuji S, et al: Cyclooxygenase regulates angiogenesis induced by colon cancer cells. Cell 93:705-716, 1998

    Narko K, Ristimaki A, MacPhee M, et al: Tumorigenic transformation of immortalized ECV endothelial cells by cyclooxygenase-1 overexpression. J Biol Chem 272:21455-21460, 1997

    Kundu N, Fulton AM: Selective cyclooxygenase (COX)-1 or COX-2 inhibitors control metastatic disease in a murine model of breast cancer. Cancer Res 62:2343-2346, 2002

    Roth GJ, Machuga ET, Strittmatter P: The heme-binding properties of prostaglandin synthetase from sheep vesicular gland. J Biol Chem 256:10018-10022, 1981

    Pagels WR, Sachs RJ, Marnett LJ, et al: Immunochemical evidence for the involvement of prostaglandin H synthase in hydroperoxide-dependent oxidations by ram seminal vesicle microsomes. J Biol Chem 258:6517-6523, 1983

    Roth GJ, Stanford N, Majerus PW: Acetylation of prostaglandin synthase by aspirin. Proc Natl Acad Sci U S A 72:3073-3076, 1975

    Hemler M, Lands WE: Purification of the cyclooxygenase that forms prostaglandins: Demonstration of two forms of iron in the holoenzyme. J Biol Chem 251:5575-5579, 1976

    Merlie JP, Fagan D, Mudd J, et al: Isolation and characterization of the complementary DNA for sheep seminal vesicle prostaglandin endoperoxide synthase (cyclooxygenase). J Biol Chem 263:3550-3553, 1988

    Yokoyama C, Takai T, Tanabe T: Primary structure of sheep prostaglandin endoperoxide synthase deduced from cDNA sequence. FEBS Lett 231:347-351, 1988

    DeWitt DL, Smith WL: Primary structure of prostaglandin G/H synthase from sheep vesicular gland determined from the complementary DNA sequence. Proc Natl Acad Sci U S A 85:1412-1416, 1988

    Fu JY, Masferrer JL, Seibert K, et al: The induction and suppression of prostaglandin H2 synthase (cyclooxygenase) in human monocytes. J Biol Chem 265:16737-16740, 1990

    Masferrer JL, Zweifel BS, Seibert K, et al: Selective regulation of cellular cyclooxygenase by dexamethasone and endotoxin in mice. J Clin Invest 86:1375-1379, 1990

    Xie WL, Chipman JG, Robertson DL, et al: Expression of a mitogen-responsive gene encoding prostaglandin synthase is regulated by mRNA splicing. Proc Natl Acad Sci U S A 88:2692-2696, 1991

    Kujubu DA, Fletcher BS, Varnum BC, et al: TIS10, a phorbol ester tumor promoter-inducible mRNA from Swiss 3T3 cells, encodes a novel prostaglandin synthase/cyclooxygenase homologue. J Biol Chem 266:12866-12872, 1991

    O'Banion MK, Sadowski HB, Winn V, et al: A serum- and glucocorticoid-regulated 4-kilobase mRNA encodes a cyclooxygenase-related protein. J Biol Chem 266:23261-23267, 1991

    Isakson P, Seibert K, Masferrer J, et al: Discovery of a better aspirin. Adv Prostaglandin Thromboxane Leukot Res 23:49-54, 1995

    Dannenberg AJ, Altorki NK, Boyle JO, et al: Cyclo-oxygenase 2: A pharmacological target for the prevention of cancer. Lancet Oncol 2:544-551, 2001

    Dempke W, Rie C, Grothey A, et al: Cyclooxygenase-2: A novel target for cancer chemotherapy J Cancer Res Clin Oncol 127:411-417, 2001

    Fearon ER, Vogelstein B: A genetic model for colorectal tumorigenesis. Cell 61:759-767, 1990

    Schmedtje JF Jr, Ji YS, Liu WL, et al: Hypoxia induces cyclooxygenase-2 via the NF-kappaB p65 transcription factor in human vascular endothelial cells. J Biol Chem 272:601-608, 1997

    Bright-Thomas RM, Hargest R: APC, beta-Catenin and hTCF-4: An unholy trinity in the genesis of colorectal cancer. Eur J Surg Oncol 29:107-117, 2003

    Araki Y, Okamura S, Hussain SP, et al: Regulation of cyclooxygenase-2 expression by the Wnt and ras pathways. Cancer Res 63:728-734, 2003

    Gupta RA, Wang D, Katkuri S, et al: Activation of nuclear hormone receptor peroxisome proliferator-activated receptor-delta accelerates intestinal adenoma growth. Nat Med 10:245-247, 2004

    Kutchera W, Jones DA, Matsunami N, et al: Prostaglandin H synthase 2 is expressed abnormally in human colon cancer: Evidence for a transcriptional effect. Proc Natl Acad Sci U S A 93:4816-4820, 1996

    Shao J, Sheng H, Inoue H, et al: Regulation of constitutive cyclooxygenase-2 expression in colon carcinoma cells. J Biol Chem 275:33951-33956, 2000

    Dixon DA: Regulation of COX-2 expression in human cancer, in Dannenberg AJ, DuBois RN (eds): COX-2: A New Target for Cancer Prevention and Treatment. New York, NY, Karger, 2003, pp 52-71

    Koki AT, Masferrer JL: Celecoxib: A specific COX-2 inhibitor with anticancer properties. Cancer Control 9:28-35, 2002

    Vidal S, Kovacs K, Bell D, et al: Cyclooxygenase-2 expression in human pituitary tumors. Cancer 97:2814-2821, 2003

    McGarrity TJ, Peiffer LP, Amos CI, et al: Overexpression of cyclooxygenase 2 in hamartomatous polyps of Peutz-Jeghers syndrome. Am J Gastroenterol 98:671-678, 2003

    Dubois RN: Evaluation of the whole prostaglandin biosynthetic pathway in lung cancer. Clin Cancer Res 9:1577-1578, 2003

    Eberhart CE, Coffey RJ, Radhika A, et al: Up-regulation of cyclooxygenase 2 gene expression in human colorectal adenomas and adenocarcinomas. Gastroenterology 107:1183-1188, 1994

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

    Kargman SL, O'Neill GP, Vickers PJ, et al: Expression of prostaglandin G/H synthase-1 and -2 protein in human colon cancer. Cancer Res 55:2556-2559, 1995

    Dubois RN, Radhika A, Reddy BS, et al: Increased cyclooxygenase-2 levels in carcinogen-induced rat colonic tumors. Gastroenterology 110:1259-1262, 1996

    Gustafson-Svard C, Lilja I, Hallbook O, et al: Cyclooxygenase-1 and cyclooxygenase-2 gene expression in human colorectal adenocarcinomas and in azoxymethane induced colonic tumours in rats. Gut 38:79-84, 1996

    Williams CS, Luongo C, Radhika A, et al: Elevated cyclooxygenase-2 levels in Min mouse adenomas. Gastroenterology 111:1134-1140, 1996

    Charalambous MP, Maihofner C, Bhambra U, et al: Upregulation of cyclooxygenase-2 is accompanied by increased expression of nuclear factor-kappaB and IkappaB kinase-alpha in human colorectal cancer epithelial cells. Br J Cancer 88:1598-1604, 2003

    Shattuck-Brandt RL, Varilek GW, Radhika A, et al: Cyclooxygenase 2 expression is increased in the stroma of colon carcinomas from IL-10(-/-) mice. Gastroenterology 118:337-345, 2000

    Chapple KS, Cartwright EJ, Hawcroft G, et al: Localization of cyclooxygenase-2 in human sporadic colorectal adenomas. Am J Pathol 156:545-553, 2000

    Bamba H, Ota S, Kato A, et al: High expression of cyclooxygenase-2 in macrophages of human colonic adenoma. Int J Cancer 83:470-475, 1999

    Hull MA, Booth JK, Tisbury A, et al: Cyclooxygenase 2 is up-regulated and localized to macrophages in the intestine of Min mice. Br J Cancer 79:1399-1405, 1999

    Sonoshita M, Takaku K, Oshima M, et al: Cyclooxygenase-2 expression in fibroblasts and endothelial cells of intestinal polyps. Cancer Res 62:6846-6849, 2002

    Ko SC, Chapple KS, Hawcroft G, et al: Paracrine cyclooxygenase-2-mediated signalling by macrophages promotes tumorigenic progression of intestinal epithelial cells. Oncogene 21:7175-7186, 2002

    Inaba T, Sano H, Kawahito Y, et al: Induction of cyclooxygenase-2 in monocyte/macrophage by mucins secreted from colon cancer cells. Proc Natl Acad Sci U S A 100:2736-2741, 2003

    Brown JR, Seed MP, Willoughby DA: Relationship between apoptosis, angiogenesis and colon-26 tumour growth after oral NSAID-treatment. Adv Exp Med Biol 507:409-414, 2002

    Adegboyega PA, Mifflin RC, DiMari JF, et al: Immunohistochemical study of myofibroblasts in normal colonic mucosa, hyperplastic polyps, and adenomatous colorectal polyps. Arch Pathol Lab Med 126:829-836, 2002

    Prescott SM: Is cyclooxygenase-2 the alpha and the omega in cancer J Clin Invest 105:1511-1513, 2000

    Chen JJ, Yao PL, Yuan A, et al: Up-regulation of tumor interleukin-8 expression by infiltrating macrophages: Its correlation with tumor angiogenesis and patient survival in non-small cell lung cancer. Clin Cancer Res 9:729-737, 2003

    Folkman J: Tumor angiogenesis: Therapeutic implications. N Engl J Med 285:1182-1186, 1971

    Holmgren L, O'Reilly MS, Folkman J: Dormancy of micrometastases: Balanced proliferation and apoptosis in the presence of angiogenesis suppression. Nat Med 1:149-153, 1995

    Hanahan D, Folkman J: Patterns and emerging mechanisms of the angiogenic switch during tumorigenesis. Cell 86:353-364, 1996

    Rastinejad F, Polverini PJ, Bouck NP: Regulation of the activity of a new inhibitor of angiogenesis by a cancer suppressor gene. Cell 56:345-355, 1989

    Dameron KM, Volpert OV, Tainsky MA, et al: Control of angiogenesis in fibroblasts by p53 regulation of thrombospondin-1. Science 265:1582-1584, 1994

    Rak J, Mitsuhashi Y, Bayko L, et al: Mutant ras oncogenes upregulate VEGF/VPF expression: Implications for induction and inhibition of tumor angiogenesis. Cancer Res 55:4575-4580, 1995

    Ziche M, Jones J, Gullino PM: Role of prostaglandin E1 and copper in angiogenesis. J Natl Cancer Inst 69:475-482, 1982

    Form DM, Auerbach R: PGE2 and angiogenesis. Proc Soc Exp Biol Med 172:214-218, 1983

    Peterson HI: Effects of prostaglandin synthesis inhibitors on tumor growth and vascularization: Experimental studies in the rat. Invasion Metastasis 3:151-159, 1983

    Peterson HI: Tumor angiogenesis inhibition by prostaglandin synthetase inhibitors. Anticancer Res 6:251-253, 1986

    Tsujii M, Kawano S, Dubois RN: Cyclooxygenase-2 expression in human colon cancer cells increases metastatic potential. Proc Natl Acad Sci U S A 94:3336-3340, 1997

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

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

    Chiarugi V, Magnelli L, Gallo O: Cox-2, iNOS and p53 as play-makers of tumor angiogenesis (review). Int J Mol Med 2:715-719, 1998

    Jones MK, Wang H, Peskar BM, et al: Inhibition of angiogenesis by nonsteroidal anti-inflammatory drugs: Insight into mechanisms and implications for cancer growth and ulcer healing. Nat Med 5:1418-1423, 1999

    Daniel TO, Liu H, Morrow JD, et al: Thromboxane A2 is a mediator of cyclooxygenase-2-dependent endothelial migration and angiogenesis. Cancer Res 59:4574-4577, 1999

    Tomozawa S, Nagawa H, Tsuno N, et al: Inhibition of haematogenous metastasis of colon cancer in mice by a selective COX-2 inhibitor, JTE-522. Br J Cancer 81:1274-1279, 1999

    Leahy KM, Ornberg RL, Wang Y, et al: Cyclooxygenase-2 inhibition by celecoxib reduces proliferation and induces apoptosis in angiogenic endothelial cells in vivo. Cancer Res 62:625-631, 2002

    Yamauchi T, Watanabe M, Hasegawa H, et al: The potential for a selective cyclooxygenase-2 inhibitor in the prevention of liver metastasis in human colorectal cancer. Anticancer Res 23:245-249, 2003

    Seno H, Oshima M, Ishikawa TO, et al: Cyclooxygenase 2- and prostaglandin E(2) receptor EP(2)-dependent angiogenesis in Apc(Delta716) mouse intestinal polyps. Cancer Res 62:506-511, 2002

    Amano H, Hayashi I, Endo H, et al: Host prostaglandin E(2)-EP3 signaling regulates tumor-associated angiogenesis and tumor growth. J Exp Med 197:221-232, 2003

    Chapple KS, Scott N, Guillou PJ, et al: Interstitial cell cyclooxygenase-2 expression is associated with increased angiogenesis in human sporadic colorectal adenomas. J Pathol 198:435-441, 2002

    Brooks PC, Montgomery AM, Rosenfeld M, et al: Integrin alpha v beta 3 antagonists promote tumor regression by inducing apoptosis of angiogenic blood vessels. Cell 79:1157-1164, 1994

    Dormond O, Foletti A, Paroz C, et al: NSAIDs inhibit alpha V beta 3 integrin-mediated and Cdc42/Rac-dependent endothelial-cell spreading, migration and angiogenesis. Nat Med 7:1041-1047, 2001

    Murphy JF, Steele C, Belton O, et al: Induction of cyclooxygenase-1 and -2 modulates angiogenic responses to engagement of alphavbeta3. Br J Haematol 121:157-164, 2003

    Reinmuth N, Liu W, Ahmad SA, et al: Alphavbeta3 integrin antagonist S247 decreases colon cancer metastasis and angiogenesis and improves survival in mice. Cancer Res 63:2079-2087, 2003

    Goodwin JS: Immunologic effects of nonsteroidal anti-inflammatory drugs. Am J Med 77:7-15, 1984

    Goodwin JS: Prostaglandins and host defense in cancer. Med Clin North Am 65:829-844, 1981

    Young MR: Eicosanoids and the immunology of cancer. Cancer Metastasis Rev 13:337-348, 1994

    Farinas MC, Rodriguez-Valverde V, Zarrabeitia MT, et al: Contribution of monocytes to the decreased lymphoproliferative response to phytohemagglutinin in patients with lung cancer. Cancer 68:1279-1284, 1991

    Yoshino I, Yano T, Yoshikai Y, et al: Oligoclonal T lymphocytes infiltrating human lung cancer tissues. Int J Cancer 47:654-658, 1991

    Brunda MJ, Herberman RB, Holden HT: Inhibition of murine natural killer cell activity by prostaglandins. J Immunol 124:2682-2687, 1980

    Kubota Y, Sunouchi K, Ono M, et al: Local immunity and metastasis of colorectal carcinoma. Dis Colon Rectum 35:645-650, 1992

    Goodwin JS, Ceuppens J: Regulation of the immune response by prostaglandins. J Clin Immunol 3:295-315, 1983

    Kambayashi T, Alexander HR, Fong M, et al: Potential involvement of IL-10 in suppressing tumor-associated macrophages: Colon-26-derived prostaglandin E2 inhibits TNF-alpha release via a mechanism involving IL-10. J Immunol 154:3383-3390, 1995

    Kuroda E, Yamashita U: Mechanisms of enhanced macrophage-mediated prostaglandin E2 production and its suppressive role in Th1 activation in Th2-dominant BALB/c mice. J Immunol 170:757-764, 2003

    O'Byrne KJ, Dalgleish AG, Browning MJ, et al: The relationship between angiogenesis and the immune response in carcinogenesis and the progression of malignant disease. Eur J Cancer 36:151-169, 2000

    Cross DS, Platt JL, Juhn SK, et al: Administration of a prostaglandin synthetase inhibitor associated with an increased immune cell infiltrate in squamous cell carcinoma of the head and neck. Arch Otolaryngol Head Neck Surg 118:526-528, 1992

    Braun DP, Ahn MC, Harris JE, et al: Sensitivity of tumoricidal function in macrophages from different anatomical sites of cancer patients to modulation of arachidonic acid metabolism. Cancer Res 53:3362-3368, 1993

    Lang S, Lauffer L, Clausen C, et al: Impaired monocyte function in cancer patients: Restoration with a cyclooxygenase-2 inhibitor. FASEB J 17:286-288, 2003

    Huang M, Stolina M, Sharma S, et al: Non-small cell lung cancer cyclooxygenase-2-dependent regulation of cytokine balance in lymphocytes and macrophages: Up-regulation of interleukin 10 and down-regulation of interleukin 12 production. Cancer Res 58:1208-1216, 1998

    Berg DJ, Zhang J, Lauricella DM, et al: IL-10 is a central regulator of cyclooxygenase-2 expression and prostaglandin production. J Immunol 166:2674-2680, 2001

    Sharma S, Stolina M, Yang SC, et al: Tumor cyclooxygenase 2-dependent suppression of dendritic cell function. Clin Cancer Res 9:961-968, 2003

    Stolina M, Sharma S, Lin Y, et al: Specific inhibition of cyclooxygenase 2 restores antitumor reactivity by altering the balance of IL-10 and IL-12 synthesis. J Immunol 164:361-370, 2000

    Yao M, Kargman S, Lam EC, et al: Inhibition of cyclooxygenase-2 by rofecoxib attenuates the growth and metastatic potential of colorectal carcinoma in mice. Cancer Res 63:586-592, 2003

    Kulkarni N, Zang E, Kelloff G, et al: Effect of the chemopreventive agents piroxicam and D,L-alpha-difluoromethylornithine on intermediate biomarkers of colon carcinogenesis. Carcinogenesis 13:995-1000, 1992

    Sinicrope FA, Roddey G, McDonnell TJ, et al: Increased apoptosis accompanies neoplastic development in the human colorectum. Clin Cancer Res 2:1999-2006, 1996

    Langlois NE, Lamb J, Eremin O, et al: Apoptosis in colorectal carcinoma occurring in patients aged 45 years and under: Relationship to prognosis, mitosis, and immunohistochemical demonstration of p53, c-myc and bcl-2 protein products. J Pathol 182:392-397, 1997

    Sugamura K, Makino M, Kaibara N: Apoptosis as a prognostic factor in colorectal carcinoma. Surg Today 28:145-150, 1998

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

    Samaha HS, Kelloff GJ, Steele V, et al: Modulation of apoptosis by sulindac, curcumin, phenylethyl-3-methylcaffeate, and 6-phenylhexyl isothiocyanate: Apoptotic index as a biomarker in colon cancer chemoprevention and promotion. Cancer Res 57:1301-1305, 1997

    Barnes CJ, Lee M: Chemoprevention of spontaneous intestinal adenomas in the adenomatous polyposis coli Min mouse model with aspirin. Gastroenterology 114:873-877, 1998

    Tsujii M, Dubois RN: Alterations in cellular adhesion and apoptosis in epithelial cells overexpressing prostaglandin endoperoxide synthase 2. Cell 83:493-501, 1995

    DuBois RN, Shao J, Tsujii M, et al: G1 delay in cells overexpressing prostaglandin endoperoxide synthase-2. Cancer Res 56:733-737, 1996

    Sheng H, Shao J, Morrow JD, et al: Modulation of apoptosis and Bcl-2 expression by prostaglandin E2 in human colon cancer cells. Cancer Res 58:362-366, 1998

    Cutler NS, Graves-Deal R, LaFleur BJ, et al: Stromal production of prostacyclin confers an antiapoptotic effect to colonic epithelial cells. Cancer Res 63:1748-1751, 2003

    Cao Y, Prescott SM: Many actions of cyclooxygenase-2 in cellular dynamics and in cancer. J Cell Physiol 190:279-286, 2002

    Blanke CD: Celecoxib with chemotherapy in colorectal cancer. Oncology (Huntingt) 16:17-21, 2002

    Ricchi P, Zarrilli R, Di Palma A, et al: Nonsteroidal anti-inflammatory drugs in colorectal cancer: From prevention to therapy. Br J Cancer 88:803-807, 2003

    Grover JK, Yadav S, Vats V, et al: Cyclo-oxygenase 2 inhibitors: Emerging roles in the gut. Int J Colorectal Dis 18:279-291, 2003

    Laine L: Approaches to nonsteroidal anti-inflammatory drug use in the high-risk patient. Gastroenterology 120:594-606, 2001

    Drazen JM: COX-2 inhibitors: A lesson in unexpected problems. N Engl J Med 2005

    Wagenaar-Miller RA, Hanley G, Shattuck-Brandt R, et al: Cooperative effects of matrix metalloproteinase and cyclooxygenase-2 inhibition on intestinal adenoma reduction. Br J Cancer 88:1445-1452, 2003

    Fischer SM, Conti CJ, Viner J, et al: Celecoxib and difluoromethylornithine in combination have strong therapeutic activity against UV-induced skin tumors in mice. Carcinogenesis 24:945-952, 2003

    Torrance CJ, Jackson PE, Montgomery E, et al: Combinatorial chemoprevention of intestinal neoplasia. Nat Med 6:1024-1028, 2000

    Mann M, Sheng H, Shao J, et al: Targeting cyclooxygenase 2 and HER-2/neu pathways inhibits colorectal carcinoma growth. Gastroenterology 120:1713-1719, 2001

    Tortora G, Caputo R, Damiano V, et al: Combination of a selective cyclooxygenase-2 inhibitor with epidermal growth factor receptor tyrosine kinase inhibitor ZD1839 and protein kinase A antisense causes cooperative antitumor and antiangiogenic effect. Clin Cancer Res 9:1566-1572, 2003

    Dicker AP, Williams TL, Grant DS: Targeting angiogenic processes by combination rofecoxib and ionizing radiation. Am J Clin Oncol 24:438-442, 2001

    Becerra CR, Frenkel EP, Ashfaq R, et al: Increased toxicity and lack of efficacy of Rofecoxib in combination with chemotherapy for treatment of metastatic colorectal cancer: A phase II study. Int J Cancer 105:868-872, 2003

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

    Rosenberg L, Palmer JR, Zauber AG, et al: A hypothesis: Nonsteroidal anti-inflammatory drugs reduce the incidence of large-bowel cancer. J Natl Cancer Inst 83:355-358, 1991

    Logan RF, Little J, Hawtin PG, et al: Effect of aspirin and non-steroidal anti-inflammatory drugs on colorectal adenomas: Case-control study of subjects participating in the Nottingham faecal occult blood screening programme. BMJ 307:285-289, 1993

    Gann PH, Manson JE, Glynn RJ, et al: Low-dose aspirin and incidence of colorectal tumors in a randomized trial. J Natl Cancer Inst 85:1220-1224, 1993

    Suh O, Mettlin C, Petrelli NJ: Aspirin use, cancer, and polyps of the large bowel. Cancer 72:1171-1177, 1993

    Muscat JE, Stellman SD, Wynder EL: Nonsteroidal antiinflammatory drugs and colorectal cancer. Cancer 74:1847-1854, 1994

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

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

    Reeves MJ, Newcomb PA, Trentham-Dietz A, et al: Nonsteroidal anti-inflammatory drug use and protection against colorectal cancer in women. Cancer Epidemiol Biomarkers Prev 5:955-960, 1996

    Bansal P, Sonnenberg A: Risk factors of colorectal cancer in inflammatory bowel disease. Am J Gastroenterol 91:44-48, 1996

    La Vecchia C, Negri E, Franceschi S, et al: Aspirin and colorectal cancer. Br J Cancer 76:675-677, 1997

    Rosenberg L, Louik C, Shapiro S: Nonsteroidal antiinflammatory drug use and reduced risk of large bowel carcinoma. Cancer 82:2326-2333, 1998

    Freedman AN, Michalek AM, Weiss HA, et al: Aspirin use and p53 expression in colorectal cancer. Cancer Detect Prev 22:213-218, 1998

    Coogan PF, Rosenberg L, Louik C, et al: NSAIDs and risk of colorectal cancer according to presence or absence of family history of the disease. Cancer Causes Control 11:249-255, 2000

    Garcia Rodriguez LA, Huerta-Alvarez C: Reduced incidence of colorectal adenoma among long-term users of nonsteroidal antiinflammatory drugs: A pooled analysis of published studies and a new population-based study. Epidemiology 11:376-381, 2000

    Breuer-Katschinski B, Nemes K, Rump B, et al: Long-term use of nonsteroidal antiinflammatory drugs and the risk of colorectal adenomas: The Colorectal Adenoma Study Group. Digestion 61:129-134, 2000

    Garcia-Rodriguez LA, Huerta-Alvarez C: Reduced risk of colorectal cancer among long-term users of aspirin and nonaspirin nonsteroidal antiinflammatory drugs. Epidemiology 12:88-93, 2001

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

    Labayle D, Fischer D, Vielh P, et al: Sulindac causes regression of rectal polyps in familial adenomatous polyposis. Gastroenterology 101:635-639, 1991

    Tonelli F, Valanzano R: Sulindac in familial adenomatous polyposis. Lancet 342:1120, 1993

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

    Nugent KP, Farmer KC, Spigelman AD, et al: Randomized controlled trial of the effect of sulindac on duodenal and rectal polyposis and cell proliferation in patients with familial adenomatous polyposis. Br J Surg 80:1618-1619, 1993

    Pasricha PJ, Bedi A, O'Connor K, et al: The effects of sulindac on colorectal proliferation and apoptosis in familial adenomatous polyposis. Gastroenterology 109:994-998, 1995

    Hirota C, Iida M, Aoyagi K, et al: Effect of indomethacin suppositories on rectal polyposis in patients with familial adenomatous polyposis. Cancer 78:1660-1665, 1996

    Giardiello FM, Offerhaus JA, Tersmette AC, et al: Sulindac induced regression of colorectal adenomas in familial adenomatous polyposis: Evaluation of predictive factors. Gut 38:578-581, 1996

    Winde G, Schmid KW, Brandt B, et al: Clinical and genomic influence of sulindac on rectal mucosa in familial adenomatous polyposis. Dis Colon Rectum 40:1156-1168, 1997

    van Stolk R, Stoner G, Hayton WL, et al: Phase I trial of exisulind (sulindac sulfone, FGN-1) as a chemopreventive agent in patients with familial adenomatous polyposis. Clin Cancer Res 6:78-89, 2000(Joanne R. Brown, Raymond )