Skip to main content
HAL-INSERM logoLink to HAL-INSERM
. Author manuscript; available in PMC: 2009 Dec 30.
Published in final edited form as: Cancer Epidemiol Biomarkers Prev. 2003 May;12(5):391–400.

Point: From animal models to prevention of colon cancer. Systematic review of chemoprevention in min mice and choice of the model system

Denis E Corpet 1,*, Fabrice Pierre 1
PMCID: PMC2797538  PMID: 12750232

Abstract

The Min (Apc(+/−)) mouse model and the azoxymethane (AOM)-rat model are the main animal models used to study the effect of dietary agents on colorectal cancer. We recently reviewed the potency of chemopreventive agents in the AOM-rat model (Corpet and Taché, 2002). Here we add the results of a systematic review of the effect of diet and agents on the tumor yield in Min mice, based on the results of 179 studies, from 71 articles, and displayed also at the website http://www.inra.fr/reseau-nacre/sci-memb/corpet/indexan.html. The efficacy of agents in the Min mouse model and the AOM-rat model correlated (r=0.66, p<0.001), although some agents that afford strong inhibition in the AOM-rat and the Min mouse increase the tumor yield in the large bowel of mutant mice for reasons not yet understood. Thus, piroxicam, sulindac, celecoxib, difluoromethylornithine, and polyethylene glycol could promote carcinogenesis in the colon of mice. We also compare the results of rodent studies with those from clinical intervention studies of polyp recurrence. We found that the effect of most of the agents tested is consistent across the animal models, except the above-mentioned puzzling mouse colon. Thus our point is that the rodent models can provide guidance in the selection of prevention approaches to colon cancer, in particular suggesting the likely importance of polyethylene glycol, hesperidin, protease inhibitor, sphingomyelin, physical exercise, epidermal growth factor receptor kinase inhibitor, (+)-catechin, resveratrol, fish oil, curcumin, caffeate and thiosulfonate as preventive agents.

Keywords: Animals; Anticarcinogenic Agents; therapeutic use; Azoxymethane; diagnostic use; Chemoprevention; Colonic Neoplasms; chemically induced; prevention & control; Diet; Disease Models, Animal; Humans; Mice; Mice, Mutant Strains; Precancerous Conditions; chemically induced; prevention & control; Randomized Controlled Trials as Topic; Rats

Keywords: animal-model, diet, chemoprevention, colon-carcinogenesis, Min-mice, chemically-induced

Introduction

Puzzling results were presented at recent meetings of the American Association for Cancer Research. The feeding of the nonsteroidal anti-inflammatory drugs (NSAIDs) piroxicam or sulindac to mutant mice with spontaneous tumors, strikingly increases the tumor yield in their colon (13). However, NSAIDs are widely accepted chemopreventive agents against colon cancer in humans (4). These results raise questions about either the animal model, or the NSAIDs protection. We have thus reviewed the results of dietary chemoprevention studies in animal models of colorectal cancer and compared them with the results of clinical intervention studies, looking for consistency among the models.

Two animal models for preclinical testing of chemopreventive agents

Since 1970 investigators have searched for diets or agents that suppress colorectal tumors in rodents. Rodents have almost no spontaneous colon cancer, and a carcinogen is needed to induce the tumors. Most chemopreventive agents were thus tested in rats, usually male Fisher 344, given AOM injections. AOM is a specific colon carcinogen, like its precursor, dimethylhydrazine. The tumors induced are often mutated on K-ras and beta-catenin genes (5), but seldom (15%) on the adenomatous polyposis coli (Apc) gene, and never on the p53 gene (6). Other rodents (mice), and other colon carcinogens were used less frequently, e.g., specific nitrosamines, and heterocyclic amines like 2-amino-1-methyl-6-phenylimidazo [4,5-b]pyridine (PhIP). PhIP induces the Apc mutation frequently (40–60%) (7) and microsatellite instability (8), but no K-ras or p53 mutations (910). AOM is not present in our daily diet, but PhIP is. However, PhIP is almost never used because AOM is less expensive, more potent, and more convenient to use than PhIP. Chemopreventive treatment can be begun before exposure to the carcinogen and during the initiation phase, during the promotion-progression phase, or through both phases. The incidence of colon tumors is the major endpoint in most rats studies. Carcinogen-induced tumors produced in the rat colon share many characteristics with human colorectal cancer, except that they have a lower tendency to metastize.

In 1990, the mutant Min mouse was found with multiple intestinal neoplasia (11). It was shown to have a mutated Apc gene, similar to that in patients with familial adenomatous polyposis, and in many sporadic cancers. This promising animal model mimics the rapid development of adenomatous polyps that affect humans with germline inactivation of one Apc gene. But the K-ras mutations observed in many human tumors were not detected in Min mice polyps (12), and p53 inactivation, frequent in human cancers, does not raise tumor number in Min mice (13). Following the Min mouse discovery with truncated Apc in position 850, other mice have been genetically modified so that one or more oncogenes hold a germline mutation (e.g., truncated Apc in positions 716, 1309, or 1638, and mutated Msh2 or Mlh1). A mutation on Msh2 or Mlh1 genes leads to mismatch repair defect, which makes these mice a model for human hereditary nonpolyposis colorectal cancers (14). These mutant mouse models have increased our understanding of carcinogenesis. They have also provided a model to evaluate the effect of diets and chemopreventive agents. Compared with the AOM-rat model, the use of mutated mice avoids the hazard of carcinogen handling, and leads to shorter studies. Dietary treatments are initiated in mice by the age of 4–5 weeks, when tumors may be already present (except for a few in utero studies). This timing mimics human clinical trials, where dietary treatments are given to adults, likely to bear minute polyps, the visible ones having been removed before randomization. The number of tumors in the small intestine is the primary endpoint in most mouse studies. The major drawback of these mutant models is that, in contrast with the human situation, the mouse tumors occur predominantly in the small intestine and not in the colon.

There are now sufficient results reported from the AOM-rat and Min mouse colon cancer prevention models to make it possible to compare the results of different approaches and to begin to assess which provides the best prediction of response in clinical intervention studies. We consider first the previous review of results of interventions with the rat model, then review results obtained with the Min model.

Review on dietary chemoprevention in the AOM-rat and Min mouse models

Data from our previous systematic review on chemoprevention in the rat model (15) were gathered from 146 articles with the tumor endpoint, and 137 articles with the aberrant crypt focus endpoint, a putative preneoplastic lesion. Tables were built with potency of each agent or diet to reduce the tumor incidence or the number of aberrant crypt foci in the colon of rats. Both tables are available on a website with sorting abilities, http://www.inra.fr/reseau-nacre/sci-memb/corpet. Agents of outstanding potency, that fully suppress colon adenocarcinoma and/or consistently inhibits adenoma and aberrant crypt foci in several independent AOM-rats studies, were (ranked list): polyethylene glycol (PEG), celecoxib, hesperidin, difluoromethylornithine (DFMO) and piroxicam (combined or not), sulindac (sulfone or sulfide), and ursodeoxycholic acid. In addition, treadmill exercise, and S-methyl-methane-thiosulfonate suppressed carcinoma (supported by a single study each). Last, Bowman-Birk protease inhibitor and sphingomyelin were consistently efficient in AOM-initiated mice (15).

All publications relating to the effect of dietary agents tested in Min mice, and other mice with mutations resulting in intestinal tumors, were identified from three databases, ISI Current Contents, Medline, and the American Association for Cancer Research website, for the period from 1990 to May 2002. Data were gathered from 63 articles and eight meeting abstracts, yielding 179 comparisons between a control and a treated group of mice. A primary table (not shown) was built including the following data: mouse strain, mutation, treatment dose duration and vehicle, the number of intestinal adenomas in treated and control groups, the significance of the treatment effect, and, when reported, the size of intestinal adenomas, and the specific number of colonic adenomas. Some papers did not report small and large bowel data separately. In those cases, we reported the total number of adenomas instead of small intestinal adenomas. This primary table was abstracted to give the efficacy of each treatment to reduce the number of adenomas in the small intestine, and in the colon, of mutated mice (Table 1). The results are reported as a percentage of control values. The table is available on a website with sorting abilities, allowing to rank agents by potency, http://www.inra.fr/reseau-nacre/sci-memb/corpet/indexan.html.

Table 1.

Effect of dietary agents on the tumor number in the small intestine and in the colon of Min mice, and other mutant mice.

Category Treatment/mutationa Doseb Durationc Nd Adenoma numbere Treatment effect: treated % control Ref.

small int. colon small int. colon
Bile acid Chenodeoxycholic 0.50% 10 w 1 33 0.4 77 150 16
Bile acid Ursodeoxyc. +sulindac 1500–4500 +75 ppm 2 18 15 17
Bile acid Ursodeoxyc. +sulindac 500 +75 ppm 1 18 30 17
Bile acid Ursodeoxycholic. 500–1500 ppm 2 18 73 17
Fat Arachidonic acid 1% 7 w 1 54 101 18
Fat Corn Oil 10% vs 3% 150–300 d 2 23 1.8 137 212 19
Fat DHA/Apc716 3% female 7 w 1 219 31 20
Fat DHA/Apc716 3% male 7 w 1 193 144 20
Fat Fish oil K85 0.4% male 17w/w1 1 108 1.0 33 160 21
Fat Fish oil K85 1.25–2.5% male 17w/w1 2 108 1.0 74 65 21
Fat Fish oil K85 0.4-1.25-2.5% female 17w/w1 3 73 0.8 55 38 21
Fat HighFat LowFiber fat:cellul. 22:0 vs 7:5% 5–6 w 1 35 1.8 98 78 22
Fat Low soybean oil 5% vs 20, no fiber diet 60 d 1 30 0.4 62 57 23
Fat Western diet/Apc1638 Fat+, Ca−, vit.D− 14–34 w 2 1.0 233 24
Fiber Cellulose 5–10% vs 0, in 20%fat 60 d 2 30 0.4 45 107 23
Fiber Fructo-oligosaccharide 5.8% 42 d 1 50 2.1 93 33 25
Fiber Fructo-oligosaccharide 5.8% in T-cell depleted 42 d 1 0.4 200 26
Fiber Guar gum 5–10% vs 0, in 20%fat 60 d 2 30 0.4 59 93 23
Fiber High fiber rodent chow fiber:fat 18:6 vs 0:20% 60 d 1 29 0.4 96 271 23
Fiber Inulin 2.5% 5–6 w 1 35 1.8 140 133 22
Fiber Inulin 10% 9 w 1 55 0.6 123 211 27
Fiber Oat bran 10% 5–6 w 1 35 1.8 133 100 22
Fiber Resistant starch 18.8% 42 d 1 50 2.1 99 143 25
Fiber Rye bran 10% 5–6 w 1 35 1.8 75 78 22
Fiber rye bran 10%/inulin diet 5–6 w 1 40 91 28
Fiber Wheat bran 10% 5–6 w 1 35 1.8 99 117 22
Fiber Wheat bran 7.1% 42 d 1 50 2.1 94 71 25
Fiber Wheat bran 5–10% vs 0, in 20%fat 60 d 2 30 0.4 53 143 23
Fiber Wheat bran, fat-/Apc716 bran:fat 20:5 vs 3:20% 7 w 1 211 2.0 64 36 29
HAA IQ/Apc716 300 ppm 11 w 1 254 84 30
HAA MeIQx/Apc716 400 ppm 11 w 1 254 87 30
HAA N-OH-PhiP/Apc716 50 mg/kg × 5 inj. 5 d/d26 1 17 335 30
HAA PhiP 50 mg/kg × 8 inj. pups 3 w 1 30 0.6 789 519 31
HAA PhiP 25 mg/kg × 8 inj. pups 3 w 1 30 0.6 524 145 31
HAA PhiP 50 mg/kg × 1 inj. 7–11 w 2 47 0.6 346 297 32
HAA PhiP 10 mg/kg × 1 inj. 7–11 w 2 47 0.6 188 157 32
HAA PhiP 50 mg/kg × 4 inj. 4 w 2 72 0.7 108 78 33
HAA PhiP/Apc1638 0.03% male&fem. 182 d 1 3 234 34
HAA PhiP/Apc716 400 ppm 8 w 1 147 114 30
inhib. EGF EKB-569 150 ppm 60 d 1 20 13 35
inhib. EGF EKB−569 +sulindac 150 +37.5 ppm 60 d 1 20 4 35
inhib. EGF EKI-785 300 ppm 60 d 1 17 56 35
inhib. EGF EKI−785 +sulindac 300 +150 ppm 60 d 1 17 5 35
inhib. iNOS Aminoguanidine 1500 ppm in water 10 w 1 73 0.8 69 63 36
inhib. iNOS Arginine deficient diet No arginine diet 10 w 1 102 0.2 67 300 36
inhib. iNOS PBIT 100 ppm/high fat diet 80 d 1 66 10 37
inhib. ODC DFMO 2% 69 d 1 45 3.5 46 92 38
inhib. ODC DFMO 1% 61 d 1 42 57 39
inhib. ODC DFMO +piroxicam 1% +50 ppm 61 d 1 42 27 39
inhib. desat. SC26196 n-6 desaturase 100 mg/kg/d 7 w 1 54 63 18
NSAID 4-ASA or 5-ASA 500 ppm 75 d 2 84 122 40
NSAID 5-ASA balzalazide 250 mg/kg ? 90 d 1 22 28 21 41
NSAID 5-ASA balzalazide 62–125 mg/kg ? 90 d 2 22 49 41
NSAID Aspirine 250–500 ppm 7 w 2 36 1.0 48 65 42
NSAID Aspirine 400 ppm 370 d/d-21 1 34 74 43
NSAID Aspirine 200 ppm 130 d/d21 1 34 115 43
NSAID Piroxicam 0.5 mg/mouse/d 7 d 1 49 5 18
NSAID Piroxicam 25-50-100 ppm 61 d 3 42 66 39
NSAID Piroxicam 50-100-200 ppm 6 w 3 17 0.6 23 150 44
NSAID Piroxicam 50 ppm 25 d/d55 1 23 0.8 35 169 45
NSAID Piroxicam 50 ppm 50 d/d30 1 22 1.5 23 40 45
NSAID Piroxicam 200 ppm 90–180 d 3 53 4 46
NSAID Piroxicam 200 ppm 6–14 d 3 53 7 46
NSAID Piroxicam 200 or 220 ppm 75 d 2 68 18 40
NSAID Piroxicam 200 ppm 2–4 d 2 53 64 46
NSAID Piroxicam/Apc1309 0.05% 10 w 1 31 2.6 52 56 47
NSAID R-flurbiprofen 10 mg/kg/d or/2d gav. 21 d 2 23 35 48
NSAID Sulindac 160 ppm 11.5 w 1 12 0.4 1 25 49
NSAID Sulindac 320 ppm 80 d 1 41 0.8 7 33 50
NSAID Sulindac 0.6 mg/mouse/d 7 d 1 46 48 18
NSAID Sulindac 160 ppm 10 w 1 17 4.5 51 67 51
NSAID Sulindac 75–150 ppm 2 18 22 17
NSAID Sulindac 160 ppm 75 d 1 84 16 40
NSAID Sulindac 300ppm 10 w 1 72 68 52
NSAID Sulindac 150 ppm 60 d 1 17 26 35
NSAID Sulindac/Apc716 150 ppm 8 w 1 201 62 100 53
NSAID Sulindac/Apc716 12 mg/kg/d 8 w 1 424 1.8 84 50 54
NSAID Sulindac/MinMsh2-- 13mg/kg/d 4 w 1 354 13.0 83 108 55
NSAID Sulindac/MinMsh2± 13mg/kg/d 22 w 1 44 4.8 74 104 55
NSAID Sulindac sulfone 50 mg/kg/d gav. 42 d 1 25 77 48
NSAID/2 Celecoxib 1500 ppm 25 d/d55 1 23 0.8 48 253 45
NSAID/2 Celecoxib 500–1500 ppm 50 d/d30 2 22 1.5 29 37 45
NSAID/2 Celecoxib 150–500 ppm 25–50 d 3 23 1.1 73 65 45
NSAID/2 JTE-523 100 ppm 8 w 1 123 68 56
NSAID/2 MF-tricyclic/Apc716 3.5–14 mg/kg/d 8 w 2 424 1.8 44 19 54
NSAID/2 MF-tricyclic/MinMsh2± 13 mg/kg/d 4–22 w 2 200 9.0 46 63 55
NSAID/2 nimesulide 600 ppm/high fat diet 80 d 1 20 100 37
NSAID/2 ONO-AE2-227 300 ppm 7 w 1 61 0.5 69 40 56b
NSAID/2 Rofecoxib/Apc716 25–75 ppm 8 w 2 201 55 53
Other Beef meat 24% 5–6 w 1 35 1.8 150 178 22
Other Bovine lactoferrin 0.2% or 2% 8 w 1 54 1.2 82 83 57
Other Citrobacter rodentium 10E8 cfu one gav. 152 d 1 9 0.8 79 350 58
Other Copper 6 vs 1 ppm 13 w 1 47 1.0 46 27 59
Other DSS 4% in water 4 or 8 d 2 1.7 1421 60
Other Exercise 1.2km/1h/d 7 w 1 37 3.2 89 75 61
Other Food restriction 20% restriction 7 w 1 55 2.3 93 48 62
Other Germ-free status germ-free 85 d 1 32 1.4 88 66 63
Other Methionine 0.7% 4 w 1 26 0.5 96 180 64
Other PEG 3350 10% 10 w 1 42 3.2 53 81 65
Other PEG 8000/MinMsh2+or- 5% in male&fem. 60 d 1 12 0.0 104 3020 66
Other Selenium in broccoli 2.1 ppm Se 10 w 1 67 1.9 71 22 67
Other Selenium p-XSC 10–20 ppm 80 d 2 41 2.8 58 43 68
Other Sphingomyelin -ceramides 0.1% 8 w 3 56 1.4 56 50 69
Other Uroguanylin 26 μg/mouse/d 11 w 2 48 0.7 57 14 70
Other Vegetables & fruits mix. 20% mix in 9%fat diet 110 d/d-20 2 17 1.5 82 152 71
Other Vegetables & fruits mix. 22% mix in 20%fat diet 110 d/d-20 2 17 1.5 162 152 71
Phytochem Acarbose/Apc1309 400 ppm 10 w 1 31 2.6 93 84 47
Phytochem BB protease inhib. 0.1–0.5% 92 d/d-2 2 11 0.6 63 67 72
Phytochem Caffeic CAPE 0.15% 75 d 1 33 37 73
Phytochem Catechin (+) 0.1–1% 75 d 2 26 0.6 27 17 74
Phytochem Curcumin 0.2% 10 w 1 14 94 75
Phytochem Curcumin 0.1% 75 d 1 33 36 73
Phytochem Lignan HMR 200 ppm/inulin diet 5–6 w 1 40 1.3 67 131 28
Phytochem Resveratrol 100 ppm in water 7 w 1 30 4.0 30 0 76
Phytochem Rutin-quercetin 2% 75 d 2 33 85 73
Phytochem Soy isoflavones 475 ppm vs 16 ppm 11 w 2 31 99 77
Phytochem Tea extr. +sulindac 0.1% water +300ppm 10 w 1 72 44 52
Phytochem Tea extract (green) 0.1% iwater 10 w 1 72 78 52
PPAR activ. BRL49,653 20 mg/kg/d 8 w 1 27 0.6 113 525 78
PPAR activ. Troglitazone 150 mg/kg/d 8 w 1 22 0.6 104 283 78
PPAR activ. Troglitazone 0.2% 5 w 1 67 1.0 116 300 79
Vitamin B Folate 8–20 ppm 13 w/w3 2 24 4.6 80 80 80
Vitamin B Folate 2-8-20 ppm 26 w/w3 3 18 2.6 126 120 80
Vitamin B Folate/MinMsh2-- 8 ppm 8 w/w3 1 299 1.7 37 35 81
Vitamin B Folate/MinMsh2-- 8 ppm 5 w/w6 1 70 2.4 422 100 81
Vitamin B Folate + choline 2 ppm+3% vs 0+1.4 70 d/d21 3 29 114 82
Vitamin D 1a,25 (OH)2-D3 3 × 0.01 μg/wk 10 w 1 17 4.5 108 89 51
Vitamin D Ro 26–9114 3 × 5 μg/wk 10 w 1 17 4.5 102 104 51

Notes to table 1:

a

Mutation(s) given when different from the Apc850 mutation (Min mouse)

b

ppm, part per million; % of diet; mg/kg of body weight; inj., injection; gav., gavage.

c

w, weeks; d, days; Treatment start:/w1: from one w after birth;/d-20: from 20 d before birth.

d

Number of similar studies from the same article which were pooled before reporting mean value.

e

Number of adenomas in the small intestine of control mice (some studies report total number of intestinal adenomas), and in their colon.

f

Number of adenomas in treated mice, reported as percent of number in control mice (% smaller than 100 denotes protection). Boldface: significant effect (either decrease or increase).

The table clearly shows that NSAIDs are, by far, the most potent agents to suppress tumor formation in the small intestine of Min mice. Notably, piroxicam and sulindac decreased the tumor yield by 90% or more in several independent studies. As shown on Table 1 (ranked by potency, on the website), piroxicam or sulindac were used in all of the top-25 studies but one, which involved the epidermal growth factor receptor kinase inhibitor, EKB-569. Specific anti-cyclooxygenase (COX)-2, like celecoxib or MF-tricyclic, were not more potent than non selective NSAIDs. Other agents were less potent than NSAIDs and the best ones decreased the tumor yield by 60–70%: (+)-catechin, resveratrol, fish oil (two studies), curcumin, folic acid and caffeic acid phenetyl ester. Following agents were clearly less potent: cellulose, copper, DFMO, PEG, wheat bran, sphingomyelin, uroguanylin and selenium compound p-XSC.

Comparison of results obtained with the two prevention models

Fig. 1 shows that many agents that suppress tumors in the Min mouse intestine (Table 1) also decrease the incidence of colorectal cancer in AOM-initiated rats (15). A significant correlation was found between the efficacy of agents tested in both models (r=0.66, N=36, p<0.001). It is clear that the most potent chemopreventive agents in the Min mouse small intestine, are also potent in the colon of AOM-initiated rats, and the animal models thus seem consistent.

Figure 1.

Figure 1

Correlation between the effect of various agents on the number of adenomas in Min mice small intestine and on the incidence of colon tumors in AOM-initiated rats.

The correlation coefficient was 0.66 with all the points (p<0.001), and 0.82 after exclusion of two outliers (methionine and folic acid).

Min mice have many tumors in the small intestine (median number 34), but few tumors in the colon (median, 1.0) (Table 1). In contrast, human tumors are rarely found the small intestine, but frequently in the colon. This discrepancy between the mouse model and the human situation led us to examine the effect of diets on the tumor yield in the colon of mutated mice. We thus calculated the ratio between treatment effect in the small intestine and in the colon (two last columns in Table 1). The table was sorted by this ratio, showing that the median ratio was 0.95: on average, the agents have similar efficacy on large and small intestinal tumors. The top and bottom of this ranked table are shown on Table 2, displaying agents with a ratio below 0.4 or above 2.5. Several studies with NSAIDs and peroxisome proliferator-activated receptor (PPAR) agonists show much weaker protection (or specific promotion) to the colon than to the small intestine. In addition high fiber diets, PEG and Citrobacter rodentium can increase specifically the tumor yield in the colon. In contrast resveratrol, folic acid, uroguanylin, selenium in broccoli, and fructo-oligosaccharides afforded a specific protection to the colon (Table 2).

Table 2.

Agents with a very different effect on the tumor number in the small intestine and in the colon of Min mice (ratio below 0.4, or above 2.5, data from Table 1).

Category Treatment Na Treatment effect in the colon: treated % control Protection ratio: effect in small intestine vs colon Ref.
NSAID Sulindac 1 25 0.04 49
NSAID Sulindac 1 33 0.22 50
NSAID Piroxicam 3 150 0.15 44
NSAID Piroxicam 1 169 0.20 45
NSAID/2 Celecoxib 1 253 0.19 45
NSAID/2 Nimesulide 1 100 0.20 37
PPAR activ. BRL49,653 1 525 0.22 78
PPAR activ. Troglitazone 1 283 0.37 78
PPAR activ. Troglitazone 1 300 0.39 79
inhib. iNOS Arginine deficient diet 1 300 0.22 36
Fiber High fiber chow 1 271 0.35 23
Fiber Wheat bran 2 143 0.37 23
Fat Fish oil K85 1 160 0.20 21
Other Citrobacter rodentium 1 350 0.23 58
Other PEG 8000 1 3020 0.03 66

Fiber Fructo-oligosaccharide 1 33 2.8 25
Other Selenium in broccoli 1 22 3.2 67
Other Uroguanylin 2 14 4.0 70
Vitamin Folate/MinMsh2-- 1 100 4.2 81
inhib. iNOS PBIT 1 10 6.6 37
Phytochem Resveratrol 1 0 100 76
a

Note: see abbreviations and notes to Table 1.

Some discrepancies between the small and large bowel are easy to explain, particularly when the effect is modulated by the gut flora. For instance, fructo-oligosaccharides are not digested in the small intestine, but fermented by the microflora in the colon, where they yield butyrate, a possible apoptosis inducer. It may explain why fructo-oligosaccharides can decrease the tumor yield in the colon of Min mice, but not in their small intestine (25). In contrast, the promotion of tumors by Citrobacter rodentium is limited to the colon, where the bacterial density is much higher than in the small intestine (58). The colorectal position of cancers in people is believed to follow, at least in part, the presence of an abundant colonic microflora. From this point of view, the small intestine of Min mice may not be a proper model of the human colon. Also, the tumor promotion by PPAR gamma agonists is much stronger in the colon than in the small intestine. This pattern reflects PPAR gamma expression, high in the colon, low in the rest of the gut, in both mice and humans (83).

In contrast, it is surprising that the most potent chemopreventive agents in the Min mouse small intestine, also potent in the colon of AOM-initiated rats, could sometimes increase the tumor yield in the colon or in the ileum of mutated mice. The NSAIDs piroxicam, sulindac, and celecoxib strongly decrease the number of tumors in the small intestine of mutant mice (Table 1), and strongly decrease the tumor incidence in AOM-initiated rats (84, 85, 86). In contrast, piroxicam increased the number of tumors in the colon of Min mice in four studies out of six (Table 1). In addition, piroxicam caused a ten-fold dose-dependant increase in tumor multiplicity in the distal intestine of Msh2−/− mice (1). In two studies out of three, the sulindac protection in the colon of Min mice is much weaker than in the small intestine (Table 2). In three studies out of four, no protection is seen in the colon of mice with Msh2 or Apc716 mutations (Table 1). In addition, sulindac treatment significantly increased colonic tumors in four mutated mouse models: Apc Min, Apc1638, Apc1638/Mlh1, and Mlh1 mice (2, 3). For instance, in Mlh1+/− mice, sulindac treatment increased the colon tumor incidence from 20% to 91% (3). Also, a late treatment with a high dose of celecoxib increased the tumor yield in the colon of Min mice (45). Thus, very potent chemopreventive NSAIDs could promote tumors in the colon of mutated mice.

Other agents than NSAIDs also yield discrepant results in the colon of rats and mice, for instance DFMO, PEG and inulin. DFMO blocks ornithine decarboxylase, it strongly decreases colon tumor incidence in rats (87), and suppresses polyps in the small intestine of Min mice, but it increased the number of large polyps in the colon of Min mice (38). PEG, a mild laxative, is a strikingly potent chemopreventive agent in rats (88, 89), but in one study out of two (65, 66), it strikingly increased the tumor yield in the colon of Min mice (Table 1). Last, inulin, a natural non digestible oligosaccharide, decreases carcinogenesis in rats, but increases the tumor yield in the colons of Min mice (22, 27).

The reason for these puzzling discrepancies is unclear. Some differences seem spurious: due to the very low number of tumors in the colon of Min mice, an increase may be seen when there is no true effect. However, the promotion by sulindac or piroxicam is seen in several independent studies. Indeed, in Min mice, differences in key enzymes make small and large bowel mucosas react differently to COX inhibitors. Phospholipases A-2 (PLA-2) are key enzymes at the start of the arachidonic acid cascade, that lead to the promoting prostaglandin E2. Arachidonic acid is released from phospholipids by either the secretory sPLA-2 or the cytosolic cPLA-2 (90). cPLA-2 is upregulated in tumors from the colon of humans and rats, and from the mouse small intestine (91, 92). This difference is not seen in the colon of Min mice, where cPLA-2 mRNA is high in both normal and tumor tissues (93). Moreover, mice with a mutated cPLA-2 gene have smaller tumors in the small intestine than wild controls, but the effect is not seen in the colon. In contrast, sPLA-2 does not seem essential for carcinogenesis in humans (94, 95), in PhIP-initiated rats (96), or in Min mice. Indeed, C57Bl6/Min mice with a mutated sPLA-2 gene have more tumors than AKR/Min mice with the intact sPLA-2 gene (97). COX-2 converts arachidonic acid to prostaglandin. It is over-expressed in tumors from the colon of humans and rats, and from the mouse small intestine (92). In Apc-mutated mice, the knocking out of COX-2 dramatically reduces the number and size of small intestinal polyps (54). Conversely, COX-2 upregulation is associated with the development of polyps in the small intestine. In contrast, COX-2 protein is not over-expressed in colonic polyps. COX-2 expression is higher in the large than in the small bowel mucosa (98). Thus, prostaglandin producing enzymes are more expressed, but mice have fewer tumors, in the colon than in the small intestine. This may explain that, in several mice studies, NSAIDs do not prevent, but promote, colon tumors. Curiously, the contrasts observed between tumors and normal mucosa in the colon of humans and rats, are better reproduced in the small than in the large bowel of Min mice.

Polyamines levels are lower in the colon than in the small intestine of Min mice. In spite of a high ornithine decarboxylase expression, a colonic antizyme decreases the polyamine pool (38). This low level of polyamines in the colon may explain why Min mice have few polyps in the colon. Moreover, DFMO treatment reduces polyamine levels in human colon, but not in the colon of Min mice. This may explain why DFMO does not suppress colonic polyps in Min mice. Again, this would suggest that the colon of rats and the small intestine of Min mice are better models than the Min mouse colon.

Comparison of humans data with animal models data

Finally, we would like to know how the results with the two models compare with those obtained to date in clinical trials directed at preventing the recurrence of colonic polyps. How well do the animal models predict what happens in humans? To answer this question we built a table showing the effect of dietary interventions on tumors in rats and mice, and on the recurrence of colonic polyps in humans (Table 3). The mean effect in rats was extracted from a published data base of positive studies (15), to which were added null and negative studies. The mean effect in Min mice was calculated from Table 1. Table 3 is obviously a first approach to such a comparison, since no account was taken of the dose used, and the data presented are not homogeneous across different models.

Table 3. Summary of dietary prevention of colorectal tumors in rats, mice and humans.

Value in treated group percent of control group.

Agent or Diet a AOM-rats, colon tumor incidenceb Min mice, polyp number, Human trial, polyp recurrence References and notes
Small bowel Large bowel
Selenium 50 (7)c 60 (3) 40 (3) 50 99d
Celecoxib 20 (2) 60 (4) 110 (6±) 70 100e
Aspirin 90 (9±) 70 (4) 70 (2) 80 101f
Sulindac 60 (8) 50 (15) 70 (7±) 80 102g
Calcium 70 (6±) 80 24h, 103
Wheat bran 60 (9) 80 (4) 90 (4±) 90 104
Low fat 80 (10±) 70 (1) 50 (1) 100 105108i
Caloric reduction 50 (3) 90 (1) 50 (1) 100 106108j
Fruits & vegetables 100 (8) 120 (4±) 150 (2) 100 106, 109111
Beta-carotene 80 (3) 110 108, 112
Vit. C + vit. E 100 (11) 110 112114
Psyllium 40 (1) 160 115

Notes to table 3.

a

Tested in clinical trials, ranked by efficacy in humans.

b

Null and negative rats’ studies were added to a previously review of positive studies (15).

c

Data are mean percentages of colorectal tumor incidence in treated group vs. control group, rounded to the nearest ten (100= no effect). Boldface: significant effect; Italics: value not firmly established (single or small size trial, secondary endpoint); Within parentheses: number of pooled studies; ±: discrepant studies in the pool.

d

Colon cancer was a secondary endpoint in the selenium trial, primarily designed to reduce prostate cancer.

e

Polyp reduction shown in FAP patients. No data yet published on sporadic polyps.

f

Significant effect of low dose of aspirin (80 mg/d), no effect of high dose (325 mg/d).

g

Sulindac shows significant protection in FAP patients (3 small-size trials), not on sporadic polyps (2 trials) (4).

h

The low-calcium “Western Diet” increases by +175% the tumor yield in Apc1638 mice, but result was not included in Table 3 because it is also a low-vit.D and high-fat diet.

i

Significant effect in F344 rats, but no effect in SD rats (not shown). In volunteers, the interventions were in part, or led to, a dietary fat intake reduction.

j

The above cited low-fat interventions also led to a reduction of caloric intake, estimated as −18%, −10% and −5% respectively.

Table 3, none-the-less, shows that the effect of most of the diets or agents is consistent across the various models though discrepancies are seen between the effect of agents in humans and in animals as follows.

  • - NSAIDs strongly decrease the tumor yield in the colon of AOM-injected rats, and in the small intestine of mutant mice. This is consistent with epidemiological studies suggesting that, taken collectively, NSAIDs might decrease the colorectal cancer incidence by 45% in humans (4, 116). It is also consistent with the effect of celecoxib and sulindac which decrease the polyp number in familial adenomatous polyposis patients trials. However, as detailed above, several independent studies (but not all) show that some NSAIDs can increase the tumor yield in the colon of mutant mice.

  • - Wheat bran consistently reduces carcinogenesis in animals, but has apparently no significant effect in humans, a discrepancy for insoluble fibers, already pointed out by Giovannucci (117). A soluble fiber, psyllium, decreases carcinogenesis in rats, but increases the tumor recurrence in human volunteers. However, both results are only supported by a single study each. In addition, other soluble fibers similar to psyllium often show promoting properties in AOM-induced rats, an effect that fits the human trial result.

  • - Rats and mice fed a high fat diet have usually more tumors than controls fed a low fat diet. In rodents, the relationship between the colon cancer incidence and the intake of fat remains true when controlled for calorie consumption. Fatty diets with high linoleic acid content, and n-6-polyunsaturated fatty acids, seem particularly consistent promoters in rodents (105). In contrast, neither human trials nor observation studies support fat, or linoleic acid, as tumor promoters in humans (118), a discrepancy already pointed out by Giovannucci (117).

  • - Caloric reduction is a strategy that seems very efficient in animals (Table 3). Overnutrition could be seen as the most potent “carcinogen” in rodents (119). According to Willett, a positive energy balance (caloric intake versus physical activity) is the most powerful and consistent dietary influence on carcinogenesis (120). No published human trial specifically tested the effect of caloric reduction. However, a side effect of interventions with low-fat diet, and with fruits and vegetables, was a modest reduction in caloric intake (106108). The lack of reduction in polyp recurrence seen in these trials (Table 3) suggests the caloric reduction was too small to reduce insuline resistance, a supposed link between overnutrition and carcinogenesis (117, 121).

  • - That fruits and/or vegetables consumption protects against colorectal cancer is a dogma supported by many epidemiological studies (122123), an association that may have been overstated (124). This dogma is challenged by all the experimental studies in rats, mice and humans (Table 3). Indeed, a mixture of fruits and vegetables reproducing the people typical intake marginally increased the tumor yield in most animal studies (71, 109111), but large amounts of black raspberries or of orange juice can inhibit carcinogenesis in rats (125).

Conclusions

There is a close agreement between the many results obtained in the colon of AOM-initiated rats and in the small intestine of Min mice (Fig. 1). There is a reasonable association between these animal studies and the more limited clinical studies (Table 3). However, some results obtained in the colon of Min mice are discrepant from those of the Min mouse small intestine and the AOM-rats, which suggests they should be disregarded until they can be explained. Many promising agents strongly and consistently suppress tumor formation or growth in the small intestine of Min mice, or in the colon of AOM-injected rats. Some of them have already been tested in completed clinical trials: selenium, celecoxib, aspirin, sulindac, calcium, wheat bran, low fat diet, fruits and vegetables diet, beta-carotene, vitamin C and E (Table 3). Others are presently under study in humans: ursodeoxycholate, piroxicam, DFMO, and folic acid. Most published trials show no reduction in polyp recurrence, and the significant protection afforded by celecoxib, aspirin or calcium was modest (Table 3). We thus need new agents or strategies to reduce cancer load.

A conservative approach would be to include an agent or a diet in a human clinical trial only when it shows preventive properties in all available models. We think that this approach might be too conservative, and would have disqualified the testing of celecoxib, piroxicam, sulindac, DFMO, calcium and folic acid in humans, since they have shown promoting properties in some preclinical studies. Our point is that it is appropriate to proceed now with the agents that are particularly potent against carcinogenesis in either rats or mice. The data we have summarized and compared would suggest that these include: PEG, hesperidin, Bowman-Birk protease inhibitor, sphingomyelin, physical exercise, and S-methyl-methane-thiosulfonate (from the AOM-rat model), and EKB-569, (+)-catechin, resveratrol, fish oil, curcumin, and caffeic acid phenetyl ester (from the Min mouse model). These agents showed no toxicity in rodents, and some of them are already used daily by humans on a large scale (PEG, exercise, catechin, fish oil, curcumin). The safety of others, notably EKB-569, still need to be evaluated (126).

Since human studies are extremely long and costly, they require stringent preliminary studies to evaluate side-effects and optimal dosage. In addition, the long term administration of an agent to many people poses ethical problems, as beta-carotene trials in smokers sadly showed. The use of surrogate endpoint biomarkers in step-wise clinical trials might help to decrease both cost and risk. For instance, a short trial on the suppression of aberrant crypt foci in the colon of volunteers (127) could be done before a standard trial on adenoma recurrence. This approach would be particularly appropriate for agents like PEG that clear the aberrant crypt foci quickly from the mucosa (89). This strategy could eventually provide evidence for safe dietary interventions for the prevention of colorectal cancer.

The abbreviations used are

NSAID

nonsteroidal anti-inflammatory drug

AOM

azoxymethane

Apc

adenomatous polyposis coli

PhIP

2-amino-1-methyl-6-phenylimidazo[4,5-b]pyridine

PEG

polyethylene glycol

DFMO

difluoro-methylornithine

COX

cyclooxygenase

PPAR

peroxisome proliferator-activated receptor

PLA-2

phospholipase A-2

References

  • 1.Jacoby RF, Cole CE, Lubet RA, You M. Effect of the nonspecific Cox1/2 inhibitor piroxicam and the ornithine decarboxylase inhibitor difluoromethylornitine (DFMO) on development of intestinal tumors in mice bearing germline alteration the the Msh2 or APC genes. Proc AACR. 2001;42:1422. [Google Scholar]
  • 2.Yang K, Fan K, Shinozaki H, Newmark H, Edelmann W, Kucherlapati R, Lipkin M. Sulindac increases carcinoma development in the colons of mice with Apc mutations. Proc AACR. 1999;40:3488. [Google Scholar]
  • 3.Yang K, Fan K, Lia M, Edelmann W, Augenlicht LH, Lubet R, Kopelovich L, Kucherlapati R, Lipkin M. Sulindac Increases Tumor Development in the Colon of Mice with Mlh1 +/− Mutation. Proc AACR. 2001;42:1423. [Google Scholar]
  • 4.Thun MJ, Henley J, Patrono C. Nonsteroidal anti-inflammatory drugs as anticancer agents: mechanistic, pharmacologic, and clinical issues. J Natl Cancer Inst. 2002;94:252–266. doi: 10.1093/jnci/94.4.252. [DOI] [PubMed] [Google Scholar]
  • 5.Dashwood RH, Suzui M, Nakagama H, Sugimura T, Nagao M. High frequency of beta-catenin (Ctnnb1) mutations in the colon tumors induced by two heterocyclic amines in the F344 rat. Cancer Res. 1998;58:1127–1129. [PubMed] [Google Scholar]
  • 6.DeFilippo C, Caderni G, Bazzicalupo M, Briani C, Giannini A, Fazi M, Dolara P. Mutations of the Apc gene in experimental colorectal carcinogenesis induced by azoxymethane in F344 rats. Brit J Cancer. 1998;77:2148–2151. doi: 10.1038/bjc.1998.359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Kakiuchi H, Watanabe M, Ushijima T, Toyota M, Imai K, Weisburger JH, Sugimura T, Nagao M. Specific 5′-GGGA-3′ to 5′-GGA-3′ mutation of the Apc gene in rat colon tumors induced by 2-amino-1-methyl--6-phenylimidazo(4,5-b)pyridine. Proc Natl Acad Sci USA. 1995;92:910–914. doi: 10.1073/pnas.92.3.910. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Canzian F, Ushijima T, Serikawa T, Wakabayashi K, Sugimura T, Nagao M. Instability of microsatellites in rat colon tumors induced by heterocyclic amines. Cancer Res. 1994;54:6315–6317. [PubMed] [Google Scholar]
  • 9.Kakiuchi H, Ushijima T, Ochiai M, Imai K, Ito N, Yachi A, Sugimura T, Nagao M. Rare frequency of activation of the ki-ras gene in rat colon tumors induced by heterocyclic amines - possible alternative mechanisms of human colon carcinogenesis. Mol Carcinogenesis. 1993;8:44–48. doi: 10.1002/mc.2940080110. [DOI] [PubMed] [Google Scholar]
  • 10.Makino H, Ushijima T, Kakiuchi H, Onda M, Ito N, Sugimura T, Nagao M. Absence of p53 mutations in rat colon tumors induced by 2-amino-6-methyldipyrido [1,2-a:3′,2′-d]imidazole, 2-amino-3-methylimid azo[4,5-f]quinoline, or 2-amino-1-methyl-6-phenylimidazo[4,5-b]pyridine. Jap J Cancer Res. 1994;85:510–514. doi: 10.1111/j.1349-7006.1994.tb02388.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Moser AR, Pitot HC, Dove WF. A dominant mutation that predisposes to multiple intestinal neoplasia in the mouse. Science. 1990;247:322–324. doi: 10.1126/science.2296722. [DOI] [PubMed] [Google Scholar]
  • 12.Shoemaker AR, Luongo C, Moser AR, Marton LJ, Dove WF. Somatic mutational mechanisms involved in intestinal tumor formation in Min mice. Cancer Res. 1997;57:1999–2006. [PubMed] [Google Scholar]
  • 13.Fazeli A, Steen RG, Dickinson SL, Bautista D, Dietrich WF, Bronson RT, Bresalier RS, Lander ES, Costa J, Weinberg RA. Effects of p53 mutations on apoptosis in mouse intestinal and human colonic adenomas. Proc Natl Acad Sci USA. 1997;94:10199–10204. doi: 10.1073/pnas.94.19.10199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.DeWind N, Dekker M, VanRossum A, VanderValk M, Riele HT. Mouse models for hereditary nonpolyposis colorectal cancer. Cancer Res. 1998;58:248–255. [PubMed] [Google Scholar]
  • 15.Corpet DE, Tache S. Most effective colon cancer chemopreventive agents in rats: A review of aberrant crypt foci and tumor data, ranked by potency. Nutr Cancer. 2002;43 doi: 10.1207/S15327914NC431_1. in the press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Mahmoud NN, Dannenberg AJ, Bilinski RT, Mestre JR, Chadburn A, Churchill M, Martucci C, Bertagnolli MM. Administration of an unconjugated bile acid increases duodenal tumors in a murine model of familial adenomatous polyposis. Carcinogenesis (Lond) 1999;20:299–303. doi: 10.1093/carcin/20.2.299. [DOI] [PubMed] [Google Scholar]
  • 17.Jacoby RF, Cole CE, Hawk ET, Lubet RA. Urodeoxycholate plus low dose sulindac is an effective chemopreventive agent combination that is well tolerated and decreases adenoma multiplicity in the Apc mutan t Min mouse. Proc AACR. 2002;43:3322. [Google Scholar]
  • 18.HansenPetrik MB, McEntee MF, Johnson BT, Obukowicz G, Masferrer J, Zweifel B, Chiu CH, Whelan J. Selective inhibition of delta-6 desaturase impedes intestinal tumorigenesis. Cancer Let. 2002;175:157–163. doi: 10.1016/s0304-3835(01)00715-7. [DOI] [PubMed] [Google Scholar]
  • 19.Wasan HS, Novelli M, Bee J, Bodmer WF. Dietary fat influences on polyp phenotype in multiple intestinal neoplasia mice. Proc Natl Acad Sci USA. 1997;94:3308–3313. doi: 10.1073/pnas.94.7.3308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Oshima M, Takahashi M, Oshima H, Tsutsumi M, Yazawa K, Sugimura T, Nishimura S, Wakabayashi K, Taketo MM. Effects of docosahexaenoic acid (DHA) on intestinal polyp development in apc(delta 716) knockout mice. Carcinogenesis (Lond) 1995;16:2605–2607. doi: 10.1093/carcin/16.11.2605. [DOI] [PubMed] [Google Scholar]
  • 21.Paulsen JE, Elvsaas IKO, Steffensen IL, Alexander J. A fish oil derived concentrate enriched in eicosapentaenoic and docosahexaenoic acid as ethyl ester suppresses the formation and growth of intestinal polyps in the Min mouse. Carcinogenesis (Lond) 1997;18:1905–1910. doi: 10.1093/carcin/18.10.1905. [DOI] [PubMed] [Google Scholar]
  • 22.Mutanen M, Pajari AM, Oikarinen SI. Beef induces and rye bran prevents the formation of intestinal polyps in apc(min) mice: relation to beta-catenin and PKC isozymes. Carcinogenesis (Lond) 2000;21:1167–1173. doi: 10.1093/carcin/21.6.1167. [DOI] [PubMed] [Google Scholar]
  • 23.Yu CF, Whiteley L, Carryl O, Basson MD. Differential dietary effects on colonic and small bowel neoplasia in C57BL/6j Apc Min/+ mice. Dig Dis Sci. 2001;46:1367–1380. doi: 10.1023/a:1010615215528. [DOI] [PubMed] [Google Scholar]
  • 24.Yang K, Edelmann W, Fan KH, Lau K, Leung D, Newmark H, Kucherlapati R, Lipkin M. Dietary modulation of carcinoma development in a mouse model for human familial adenomatous polyposis. Cancer Res. 1998;58:5713–5717. [PubMed] [Google Scholar]
  • 25.Pierre F, Perrin P, Champ M, Bornet F, Meflah K, Menanteau J. Short-chain fructo-oligosaccharides reduce the occurrence of colon tumors and develop gut-associated lymphoid tissue in Min mice. Cancer Res. 1997;57:225–228. [PubMed] [Google Scholar]
  • 26.Pierre F, Perrin P, Bassonga E, Bornet F, Meflah K, Menanteau J. T cell status influences colon tumor occurrence in Min mice fed short chain fructo-oligosaccharides as a diet supplement. Carcinogenesis (Lond) 1999;20:1953–1956. doi: 10.1093/carcin/20.10.1953. [DOI] [PubMed] [Google Scholar]
  • 27.Pajari AM, Rajakangas J, Paivarinta E, Kosam VM, Rafter J, Mutanen M. Inulin modulates intestinal tumor formation partly through an accumulation of cytosolic B-catenin in Min mice. AACR special conference in Cancer Research. Colon cancer: genetics to prevention; Philadelphia, Pennsylvania. March 7–10, 2002; 2002. p. A-16. [Google Scholar]
  • 28.Oikarinen SI, Pajari AM, Mutanen M. Chemopreventative activity of crude hydroxsymatairesinol (HMR) extract in adenomatous polyposis coli multiple intestinal neoplasia (APC) (min) mice (vol 159, pg 183, 2000) Cancer Let. 2000;161:253–258. doi: 10.1016/s0304-3835(00)00662-5. [DOI] [PubMed] [Google Scholar]
  • 29.Hioki K, Shivapurkar N, Oshima H, Alabaster O, Oshima M, Taketo MM. Suppression of intestinal polyp development by low-fat and high-fiber diet in Apc (delta 716) knockout mice. Carcinogenesis (Lond) 1997;18:1863–1865. doi: 10.1093/carcin/18.10.1863. [DOI] [PubMed] [Google Scholar]
  • 30.Oshima M, Oshima H, Tsutsumi M, Nishimura S, Sugimura T, Nagao M, Taketo MM. Effects of 2-amino-1-methyl-6-phenylimidazo [4,5b] pyridine on intestinal polyp development in apc(delta 716) knockout mice. Mol Carcinogenesis. 1996;15:11–17. doi: 10.1002/(SICI)1098-2744(199601)15:1<11::AID-MC3>3.0.CO;2-V. [DOI] [PubMed] [Google Scholar]
  • 31.Andreassen A, Vikse R, Steffensen IL, Paulsen JE, Alexander J. Intestinal tumours induced by the food carcinogen 2-amino-1-methyl-6-phelylimidazo [4,5-b]pyridine in multiple intestinal neoplasia mice have truncation mutations as well as loss of the wild-type Apc+ allele. Mutagenesis. 2001;16:309–315. doi: 10.1093/mutage/16.4.309. [DOI] [PubMed] [Google Scholar]
  • 32.Steffensen IL, Schut HAJ, Paulsen JE, Andreassen A, Alexander J. Intestinal tumorigenesis in multiple intestinal neoplasia mice induced by the food mutagen 2-amino-1-methyl-6- phenylimidazo[4,5-b]pyridine: perinatal susceptibility, regional variation, and correlation with DNA adducts. Cancer Res. 2001;61:8689–8696. [PubMed] [Google Scholar]
  • 33.Steffensen IL, Paulsen JE, Eide TJ, Alexander J. 2-amino-1-methyl-6- phenylimidazo[4,5-b]pyridine increases the numbers of tumors, cystic crypts and aberrant crypt foci in multiple intestinal neoplasia mice. Carcinogenesis (Lond) 1997;18:1049–1054. doi: 10.1093/carcin/18.5.1049. [DOI] [PubMed] [Google Scholar]
  • 34.Sorensen IK, Kristiansen E, Mortensen A, Vankranen H, Vankreijl C, Fodde R, Thorgeirsson SS. Short-term carcinogenicity testing of a potent murine intestinal mutagen, 2-amino-1-methyl-6- phenylimidazo-(4,5-b)pyridine (phIP), in apc1638n transgenic mice. Carcinogenesis (Lond) 1997;18:777–781. doi: 10.1093/carcin/18.4.777. [DOI] [PubMed] [Google Scholar]
  • 35.Torrance CJ, Jackson PE, Montgomery E, Kinzler KW, Vogelstein B, Wissner A, Nunes M, Frost P, Discafani CM. Combinatorial chemoprevention of intestinal neoplasia. Nature Med. 2000;6:1024–1028. doi: 10.1038/79534. [DOI] [PubMed] [Google Scholar]
  • 36.Ahn B, Ohshima H. Suppression of intestinal polyposis in Apc (Min/+) mice by inhibiting nitric oxide production. Cancer Res. 2001;61:8357–8360. [PubMed] [Google Scholar]
  • 37.Rao CV, Malisetty SV, Cooma I, Reddy BS. Chemoprevention of FAP polyps and carcinomas by iNOS and COX2 selective inhibitors administered individually and in combination in the APC Min mice model. Proc AACR. 2002;43:3323. [Google Scholar]
  • 38.Erdman SH, Ignatenko NA, Powell MB, Blohmmangone KA, Holubec H, Guillenrodriguez JM, Gerner EW. APC-dependent changes in expression of genes influencing polyamine metabolism, and consequences for gastrointestinal carcinogenesis, in the Min mouse. Carcinogenesis (Lond) 1999;20:1709–1713. doi: 10.1093/carcin/20.9.1709. [DOI] [PubMed] [Google Scholar]
  • 39.Jacoby RF, Cole CE, Tutsch K, Newton MA, Kelloff G, Hawk ET, Lubet RA. Chemopreventive efficacy of combined piroxicam and difluoromethylornithine treatment of Apc mutant Min mouse adenomas, and selective toxicity against Apc mutant embryos. Cancer Res. 2000;60:1864–1870. [PubMed] [Google Scholar]
  • 40.Ritland SR, Leighton JA, Hirsch RE, Morrow JD, Weaver AL, Gendler SJ. Evaluation of 5-aminosalicylic acid (5-ASA) for cancer chemoprevention: lack of efficacy against nascent adenomatous polyps in the Apc Min mouse. Clin Cancer Res. 1999;5:855–863. [PubMed] [Google Scholar]
  • 41.MacGregor DJ, Kim YS, Siddiki BB, Kwan J, Sleisenger MH, Johnson LK. Induction of colon cancer cell apoptosis in vitro and inhibition of intestinal tumor formation in Min mice by balsalazide and metabolites. Gastrointestinal Oncol. 1996:A553. [Google Scholar]
  • 42.Barnes CJ, Lee M. Chemoprevention of spontaneous intestinal adenomas in the adenomatous polyposis coli Min mouse model with aspirin. Gastroenterology. 1998;114:873–877. doi: 10.1016/s0016-5085(98)70305-1. [DOI] [PubMed] [Google Scholar]
  • 43.Sansom OJ, Stark LA, Dunlop MG, Clarke AR. Suppression of intestinal and mammary neoplasia by lifetime administration of aspirin in Apc(min/+) and Apc(min/+), Msh2(−/−) mice. Cancer Res. 2001;61:7060–7064. [PubMed] [Google Scholar]
  • 44.Jacoby RF, Marshall DJ, Newton MA, Novakovic K, Tutsch K, Cole CE, Lubet RA, Kelloff GJ, Verma A, Moser AR, Dove WF. Chemoprevention of spontaneous intestinal adenomas in the apc (min) mouse model by the nonsteroidal anti-inflammatory drug piroxicam. Cancer Res. 1996;56:710–714. [PubMed] [Google Scholar]
  • 45.Jacoby RF, Seibert K, Cole CE, Kelloff G, Lubet RA. The cyclooxygenase-2 inhibitor celecoxib is a potent preventive and therapeutic agent in the Min mouse model of adenomatous polyposis. Cancer Res. 2000;60:5040–5044. [PubMed] [Google Scholar]
  • 46.Ritland SR, Gendler SJ. Chemoprevention of intestinal adenomas in the apc(min) mouse by piroxicam: kinetics, strain effects and resistance to chemosuppression. Carcinogenesis (Lond) 1999;20:51–58. doi: 10.1093/carcin/20.1.51. [DOI] [PubMed] [Google Scholar]
  • 47.Quesada CF, Kimata H, Mori M, Nishimura M, Tsuneyoshi T, Baba S. Piroxicam and acarbose as chemopreventive agents for spontaneous intestinal adenomas in APC gene 1309 knockout mice. Jap J Cancer Res. 1998;89:392–396. doi: 10.1111/j.1349-7006.1998.tb00576.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Wetcher WJ, Murray DED, Kantoci D, McCracken JD, et al. Treatment and survival study in the C57bl/6J-Apc Min mouse with R-flurbiprofen. Life Sci. 2000;66:745–753. doi: 10.1016/s0024-3205(99)00645-1. [DOI] [PubMed] [Google Scholar]
  • 49.Boolbol SK, Dannenberg AJ, Chadburn A, Martucci C, Guo XJ, Ramonetti JT, Abreugoris M, Newmark HL, Lipkin ML, Decosse JJ, Bertagnolli MM. Cyclooxygenase-2 overexpression and tumor formation are blocked by sulindac in a murine model of familial adenomatous polyposis. Cancer Res. 1996;56:2556–2560. [PubMed] [Google Scholar]
  • 50.Chiu CH, Mcentee MF, Whelan J. Sulindac causes rapid regression of preexisting tumors in min/+ mice independent of prostaglandin biosynthesis. Cancer Res. 1997;57:4267–4273. [PubMed] [Google Scholar]
  • 51.Huerta S, Irwin RW, Heber D, Go VLW, Koeffler HP, Uskokovic MR, Harris DM. 1 alpha,25-(OH)(2)-d-3 and its synthetic analogue decrease tumor load in the apc(min) mouse. Cancer Res. 2002;62:741–746. [PubMed] [Google Scholar]
  • 52.Suganuma M, Ohkura Y, Okabe S, Fujiki H. Combination cancer chemoprevention with green tea extract and sulindac shown in intestinal tumor formation in Min mice. J Cancer Res Clin Oncol. 2001;127:69–72. doi: 10.1007/s004320000189. [DOI] [PubMed] [Google Scholar]
  • 53.Oshima M, Murai N, Kargman S, Arguello M, Luk P, Kwong E, Taketo MM, Evans JF. Chemoprevention of intestinal polyposis in the apc(delta 716) mouse by rofecoxib, a specific cyclooxygenase-2 inhibitor. Cancer Res. 2001;61:1733–1740. [PubMed] [Google Scholar]
  • 54.Oshima M, Dinchuk JE, Kargman SL, Oshima H, Hancock B, Kwong E, Trzaskos JM, Evans JF, Taketo MM. Suppression of intestinal polyposis in apc(delta 716) knockout mice by inhibition of cyclooxygenase 2 (COX-2) Cell. 1996;87:803–809. doi: 10.1016/s0092-8674(00)81988-1. [DOI] [PubMed] [Google Scholar]
  • 55.Lal G, Ash C, Hay K, Redston M, Kwong E, Hancock B, Mak T, Kargman S, Evans JF, Gallinger S. Suppression of intestinal polyps in Msh2-deficient and non-Msh2-deficient multiple intestinal neoplasia mice by a specific cyclooxygenase-2 inhibitor and by a dual cyclooxygenase-1/2 inhibitor. Cancer Res. 2001;61:6131–6136. [PubMed] [Google Scholar]
  • 56.Sasai H, Masaki M, Wakitani K. Suppression of polypogenesis in a new mouse strain with a truncated apc(delta 474) by a novel COX-2 inhibitor, JTE-522. Carcinogenesis (Lond) 2000;21:953–958. doi: 10.1093/carcin/21.5.953. [DOI] [PubMed] [Google Scholar]
  • 56b.Mutoh M, Watanabe K, Kitamura T, Shoji Y, Takahashi M, Kawamori T, Tani K, Kobayashi M, Maruyama T, Kobayashi K, Ohuchida S, Sugimoto Y, Narumiya S, Sugimura T, Wakabayashi K. Involvement of prostaglandin E receptor subtype EP4 in colon carcinogenesis. Cancer Res. 2002;62:28–32. [PubMed] [Google Scholar]
  • 57.Ushida Y, Sekine K, Kuhara T, Tsuda H, et al. Inhibitory effects of bovine lactoferrin on intestinal polyposis in the Apc Min mouse. Cancer letters. 1998;134:141–145. doi: 10.1016/s0304-3835(98)00249-3. [DOI] [PubMed] [Google Scholar]
  • 58.Newman JV, Kosaka T, Sheppard BJ, Fox JG, Schauer DB. Bacterial infection promotes colon tumorigenesis in apc(min/+) mice. J Infect Dis. 2001;184:227–230. doi: 10.1086/321998. [DOI] [PubMed] [Google Scholar]
  • 59.Davis CD, Newman S. Inadequate dietary copper increases tumorigenesis in the Min mouse. Cancer Let. 2000;159:57–62. doi: 10.1016/s0304-3835(00)00531-0. [DOI] [PubMed] [Google Scholar]
  • 60.Cooper HS, Everley L, Chang WC, Pfeiffer G, Lee B, Murthy S, Clapper ML. The role of mutant apc in the development of dysplasia and cancer in the mouse model of dextran sulfate sodium-induced colitis. Gastroenterology. 2001;121:1407–1416. doi: 10.1053/gast.2001.29609. [DOI] [PubMed] [Google Scholar]
  • 61.Colbert LH, Davis JM, Essig DA, Ghaffar A, Mayer EP. Exercise and tumor development in a mouse predisposed to multiple intestinal adenomas. Med Sci Sports Exerc. 2000;32:1704–1708. doi: 10.1097/00005768-200010000-00007. [DOI] [PubMed] [Google Scholar]
  • 62.Kakuni M, Morimura K, Wanibuchi H, Ogawa M, Min W, Hayashi S, Fukushima S. Food restriction inhibits the growth of intestinal polyps in multiple intestinal neoplasia mouse. Jap J Cancer Res. 2002;93:236–241. doi: 10.1111/j.1349-7006.2002.tb02163.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Dove WF, Clipson L, Gould KA, Luongo C, Marshall DJ, Moser AR, Newton MA, Jacoby RF. Intestinal neoplasia in the APC Min pouse: independence from the microbial and natural killer (beige locus) status. Cancer Res. 1997;57:812–814. [PubMed] [Google Scholar]
  • 64.Paulsen JE, Alexander J. Growth stimulation of intestinal tumours in apc(min/+) mice by dietary l-methionine supplementation. Anticancer Res. 2001;21:3281–3284. [PubMed] [Google Scholar]
  • 65.Ansari SH, DiBaise J, Gulizia J, Karolski WJ, Ratashak A, Wali RK, Roy HK. Polyethylene glycol 3350 suppresses intestinal tumorigenesis in the Min mouse. Gastroenterology. 2002;122(A-215) [Google Scholar]
  • 66.Naigamwalla D, Chia MC, Tran TT, Medline A, Hay K, Gallinger S, Bruce WR. Polyethylene glycol 8000 and colon carcinogenesis: inhibition in the F344 rat, promotion in the Min mouse. Cancer Res. 2000;60:6856–6858. [PubMed] [Google Scholar]
  • 67.Davis CD, Zeng HW, Finley JW. Selenium-enriched broccoli decreases intestinal tumorigenesis in multiple intestinal neoplasia mice. J Nutr. 2002;132:307–309. doi: 10.1093/jn/132.2.307. [DOI] [PubMed] [Google Scholar]
  • 68.Rao CV, Cooma I, Rodriguez JGR, Simi B, Elbayoumy K, Reddy BS. Chemoprevention of familial adenomatous polyposis development in the APC(min) mouse model by 1,4-phenylene bis (methylene)selenocyanate. Carcinogenesis (Lond) 2000;21:617–621. doi: 10.1093/carcin/21.4.617. [DOI] [PubMed] [Google Scholar]
  • 69.Schmelz EM, Roberts PC, Kustin EM, Lemonnier LA, Sullards MC, Dillehay DL, Merrill AH. Modulation of intracellular beta-catenin localization and intestinal tumorigenesis in vivo and in vitro by sphingolipids. Cancer Res. 2001;61:6723–6729. [PubMed] [Google Scholar]
  • 70.Shailubhai K, Yu HH, Karunanandaa K, Wang JY, Eber SL, Wang Y, Joo NS, Kim HD, Miedema BW, Abbas SZ, Boddupalli SS, Currie MG, Forte LR. Uroguanylin treatment suppresses polyp formation in the Apc Min/+ mouse and induces apoptosis in human colon adenocarcinoma cells via cyclic GMP. Cancer Res. 2000;60:5151–5157. [PubMed] [Google Scholar]
  • 71.VanKranen HJ, Vaniersel PWC, Rijnkels JM, Beems DB, Alink GM, Vankreijl CF. Effects of dietary fat and a vegetable-fruit mixture on the development of intestinal neoplasia in the apc(min) mouse. Carcinogenesis (Lond) 1998;19:1597–1601. doi: 10.1093/carcin/19.9.1597. [DOI] [PubMed] [Google Scholar]
  • 72.Kennedy AR, BeazerBarclay Y, Kinzler KW, Newberne PM. Suppression of carcinogenesis in the intestines of Min mice by soybean-derived Bowman-Birk inhibitor. Cancer Res. 1996;56:679–682. [PubMed] [Google Scholar]
  • 73.Mahmoud NN, Carothers AM, Grunberger D, Bilinski RT, Churchill MR, Martucci C, Newmark HL, Bertagnolli MM. Plant phenolics decrease intestinal tumors in an animal model of familial adenomatous polyposis. Carcinogenesis (Lond) 2000;21:921–927. doi: 10.1093/carcin/21.5.921. [DOI] [PubMed] [Google Scholar]
  • 74.Weyant MJ, Carothers AM, Dannenberg AJ, Bertagnolli MM. (+)-catechin inhibits intestinal tumor formation and suppresses focal adhesion kinase activation in the min/+ mouse. Cancer Res. 2001;61:118–125. [PubMed] [Google Scholar]
  • 75.Collett GP, Robson CN, Mathers JC, Campbell FC. Curcumin modifies apc(min) apoptosis resistance and inhibits 2-amino 1-methyl-6-phenylimidazo[4,5-b]pyridine (phIP) induced tumour formation in apc(min) mice. Carcinogenesis (Lond) 2001;22:821–825. doi: 10.1093/carcin/22.5.821. [DOI] [PubMed] [Google Scholar]
  • 76.Schneider Y, Duranton B, Gosse F, Schleiffer R, Seiler N, Raul F. Resveratrol inhibits intestinal tumorigenesis and modulates host-defense-related gene expression in an animal model of human familial adenomatous polyposis. Nutr Cancer. 2001;39:102–107. doi: 10.1207/S15327914nc391_14. [DOI] [PubMed] [Google Scholar]
  • 77.Sorensen IK, Kristiansen E, Mortensen A, Nicolaisen GM, Wijnandes JAH, Vankranen HJ, Vankreijl CF. The effect of soy isoflavones on the development of intestinal neoplasia in Apc (min) mouse. Cancer Let. 1998;130:217–225. doi: 10.1016/s0304-3835(98)00139-6. [DOI] [PubMed] [Google Scholar]
  • 78.Lefebvre AM, Chen IH, Desreumaux P, Najib J, Fruchart JC, Geboes K, Briggs M, Heyman R, Auwerx J. Activation of the peroxisome proliferator-activated receptor gamma promotes the development of colon tumors in c57BL/6j-APC(min)/+ mice. Nature Med. 1998;4:1053–1057. doi: 10.1038/2036. [DOI] [PubMed] [Google Scholar]
  • 79.Saez E, Tontonoz P, Nelson MC, Alvarez JGA, TMU, Baird SM, Thomazy VA, Evans RM. Activators of the nuclear receptor PPAR gamma enhance colon polyp formation. Nature Med. 1998;4:1058–1061. doi: 10.1038/2042. [DOI] [PubMed] [Google Scholar]
  • 80.Song J, Medline A, Mason JB, Gallinger S, Kim YI. Effects of dietary folate on intestinal tumorigenesis in the apc(min) mouse. Cancer Res. 2000;60:5434–5440. [PubMed] [Google Scholar]
  • 81.Song J, Sohn KJ, Medline A, Ash C, Gallinger S, Kim YI. Chemopreventive effects of dietary folate on intestinal polyps in apc+/−msh2−/− mice. Cancer Res. 2000;60:3191–3199. [PubMed] [Google Scholar]
  • 82.Sibani S, Melnyk S, Pogribny IP, Wang W, Hioutim F, Deng LY, Trasler J, James SJ, Rozen R. Studies of methionine cycle intermediates (SAM, SAH), DNA methylation and the impact of folate deficiency on tumor numbers in Min mice. Carcinogenesis (Lond) 2002;23:61–65. doi: 10.1093/carcin/23.1.61. [DOI] [PubMed] [Google Scholar]
  • 83.Fajas L, Auboeuf D, Raspe E, Schoonjans K, Lefebvre AM, Saladin R, Najib J, Laville M, Fruchart JC, Deeb S, VidalPuig A, Flier J, Briggs MR, Staels B, Vidals H, Auwerx J. The organization, promoter analysis, and expression of the human PPAR gamma gene. J Biol Chem. 1997;272:18779–18789. doi: 10.1074/jbc.272.30.18779. [DOI] [PubMed] [Google Scholar]
  • 84.Li H, Kramer PM, Lubet RA, Steele VE, Kelloff GJ, Pereira MA. Termination of piroxicam treatment and the occurrence of azoxymethane-induced colon cancer in rats. Cancer Lett. 1999;147:187–193. doi: 10.1016/s0304-3835(99)00296-7. [DOI] [PubMed] [Google Scholar]
  • 85.Rao CV, Rivenson A, Simi B, Zang E, Kelloff G, Steele V, Reddy BS. Chemoprevention of colon carcinogenesis by sulindac, a nonsteroidal anti-inflammatory agent. Cancer Res. 1995;55:1464–1472. [PubMed] [Google Scholar]
  • 86.Kawamori T, Rao CV, Seibert K, Reddy BS. Chemopreventive activity of celecoxib, a specific cyclooxygenase-2 inhibitor, against colon carcinogenesis. Cancer Res. 1998;58:409–412. [PubMed] [Google Scholar]
  • 87.Rao CV, Tokumo K, Rigotty J, Zang E, Kelloff G, Reddy BS. Chemoprevention of Colon Carcinogenesis by Dietary Administration of Piroxicam, alpha-Difluoromethylornithine, 16alpha-Fluoro- 5-Androsten-17-One, and Ellagic Acid Individually and in Combination. Cancer Res. 1991;51:4528–4534. [PubMed] [Google Scholar]
  • 88.Parnaud G, Tache S, Peiffer G, Corpet DE. Polyethylene-glycol suppresses colon cancer and causes dose-dependent regression of azoxymethane-induced aberrant crypt foci in rats. Cancer Res. 1999;59:5143–5147. [PubMed] [Google Scholar]
  • 89.Corpet DE, Parnaud G, Delverdier M, Peiffer G, Tache S. Consistent and fast inhibition of colon carcinogenesis by polyethylene glycol in mice and rats given various carcinogens. Cancer Res. 2000;60:3160–3164. [PubMed] [Google Scholar]
  • 90.Tischfield JA. A reassessment of the low molecular weight phopholipase A2 gene family in mammals. J Biol Chem. 1997;272:17247–17250. doi: 10.1074/jbc.272.28.17247. [DOI] [PubMed] [Google Scholar]
  • 91.Dimberg J, Samuelson A, Hugander A, Soderkvist P. Gene expression of cyclooxygenase-2, group II and cytosolic phospholipase A2 in human colorectal cancer. Anticancer Res. 1998;18:3283–3287. [PubMed] [Google Scholar]
  • 92.Rao CV, Simi B, Wynn TT, Gar K, Reddy BS. Modulating effect of amount and types of dietary fat on colonic mucosal phospholipase A2, phosphatidylinositol-specific phospholipase C activities, and cyclooxygenase metabolite formation furing different stages of colon tumor promotion in male F344 rats. Cancer Res. 1996;56:532–537. [PubMed] [Google Scholar]
  • 93.Takaku K, Sonoshita M, Sasaki N, Uozumi N, Doi Y, Shimizu T, Taketo MM. Suppression of intestinal polyposis in Apc(delta 716) knockout mice by an additional mutation in the cytosolic phospholipase A(2) gene. J Biol Chem. 2000;275:34013–34016. doi: 10.1074/jbc.C000585200. [DOI] [PubMed] [Google Scholar]
  • 94.Dobbie Z, Muller H, Scott RJ. Secretory phospholipase A2 does not appear to be associated with phenotypic variation in familial adenomatous polyposis. Human Genetic. 1996;98:386–390. doi: 10.1007/s004390050226. [DOI] [PubMed] [Google Scholar]
  • 95.Riggins GJ, Markowitz S, Wilson JK, Vogelstein B, Kinzler KW. Absence of secretory phospholipase A2 gene alterations in human colorectal cancer. Cancer Res. 1995;55:5184–5186. [PubMed] [Google Scholar]
  • 96.Ishiguro Y, Ochiai M, Sugimura T, Nagao M, Nakagama H. Strain differences of rats in the susceptibility to aberrant crypt foci formation by 2-amino-1-methyl-6-phenylimidazo-[4,5-b]pyridine: no implication of apc and pla2g2a genetic polymorphisms in differential susceptibility. Carcinogenesis (Lond) 1999;20:1063–1068. doi: 10.1093/carcin/20.6.1063. [DOI] [PubMed] [Google Scholar]
  • 97.Gould KA, Detrish WF, Borenstein N, Lander ES, Dove WF. Mom1 is a semi-dominant modifier of intestinal adenoma size and multiplicity in Min/+ mice. Genetics. 1996;144:1769–1776. doi: 10.1093/genetics/144.4.1769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Kawajiri H, Hsi LC, Kamitani H, Ikawa H, Geller M, Ward T, Eling TE, Glasgow WC. Arachidonic and linoleic acid metabolism in mouse intestinal tissue: evidence to novel lipoxygenase activity. Arch Biochem Biophys. 2002;398:51–60. doi: 10.1006/abbi.2001.2685. [DOI] [PubMed] [Google Scholar]
  • 99.Clark LC, Dalkin B, Krongrad A, Combs GF, Jr, Turnbull BW, Slate EH, Witherington R, Herlong JH, Janosko E, Carpenter D, Borosso C, Falk S, Rounder J. Decreased incidence of prostate cancer with selenium supplementation: results of a double-blind cancer prevention trial. Brit J Urol. 1998;81:730–734. doi: 10.1046/j.1464-410x.1998.00630.x. [DOI] [PubMed] [Google Scholar]
  • 100.Steinbach G, Lynch PM, Phillips RK, Wallace MH, Hawk E, Gordon GB, Wakabayashi N, Saunders B, Shen Y, Fujimura T, Su LK, Levin B. The effect of celecoxib, a cyclooxydenase-2 inhibitor, in familial adenomatous polyposis. New England J Med. 2000;342:1946–1952. doi: 10.1056/NEJM200006293422603. [DOI] [PubMed] [Google Scholar]
  • 101.Baron JA, Cole BF, Mott LA, PPSG Aspirin chemoprevention of colorectal cancer. Proc AACR. 2002;43:3319. [Google Scholar]
  • 102.Giardiello FM, Yang VW, Hylind LM, et al. Primary chemoprevention of familial adenomatous polyposis with sulindac. New Engl J Med. 2002;346:1054–1059. doi: 10.1056/NEJMoa012015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Baron JA, Beach M, Mandel JS, Vanstolk RU, Haile RW, Sandler RS, Rothstein R, Summers RW, Snover DC, Beck GJ, Bond JH, Greenberg ER. Calcium supplements for the prevention of colorectal adenomas. New Engl J Med. 1999;340:101–107. doi: 10.1056/NEJM199901143400204. [DOI] [PubMed] [Google Scholar]
  • 104.Alberts DS, Martinez ME, Roe DJ, Guillenrodriguez JM, Marshall JR, Vanleeuwen JB, Reid ME, Ritenbaugh C, Vargas PA, Bhattacharyya AB, Earnest DL, Sampliner RE, Parish D, Koonce K, Fales L. Lack of effect of a high-fiber cereal supplement on the recurrence of colorectal adenomas. New Engl J Med. 2000;342:1156–1162. doi: 10.1056/NEJM200004203421602. [DOI] [PubMed] [Google Scholar]
  • 105.Zhao LP, Kushi LH, Klein RD, Prentice RL. Quantitative Review of Studies of Dietary Fat and Rat Colon Carcinoma. Nutr Cancer. 1991;15:169–177. doi: 10.1080/01635589109514124. [DOI] [PubMed] [Google Scholar]
  • 106.Schatzkin A, Lanza E, Corle D, Lance P, Iber F, Caan B, Shike M, Weissfeld J, Burt R, Cooper MR, Kikendall JW, Cahill J, Freedman L, Marshall J, Schoen RE, Slattery M. Lack of effect of a low-fat, high-fiber diet on the recurrence of colorectal adenomas. New Engl J Med. 2000;342:1149–1155. doi: 10.1056/NEJM200004203421601. [DOI] [PubMed] [Google Scholar]
  • 107.McKeown-Eyssen GE, Bright-See E, Bruce WR, Jazmaji V Toronto-Polyp-Prevention-Group. A randomized trial of a low fat high fibre diet in the recurrence of colorectal polyps. J Clin Epidemiol. 1994;47:525–536. doi: 10.1016/0895-4356(94)90299-2. [DOI] [PubMed] [Google Scholar]
  • 108.MacLennan R, Macrae F, Bain C, Battistutta D, Chapuis P, Gratten H, Lambert J, Newland RC, Ngu M, Russell A, Ward M, Wahlqvist ML. Randomized trial of intake of fat, fiber, and beta carotene to prevent colorectal adenomas. J Natl Cancer Inst. 1995;87:1760–1766. doi: 10.1093/jnci/87.23.1760. [DOI] [PubMed] [Google Scholar]
  • 109.Alink GM, Kuiper HA, Hollanders VMH, Koeman JH. Effect of heat processing and of vegetables and fruit in human diets on 1,2-dimethylhydrazine-induced colon carcinogenesis in rats. Carcinogenesis (Lond) 1993;14:519–524. doi: 10.1093/carcin/14.3.519. [DOI] [PubMed] [Google Scholar]
  • 110.Rijnkels JM, Hollanders VMH, Woutersen RA, Koeman JH, Alink GM. Absence of an inhibitory effect of a vegetables-fruit mixture on the initiation and promotion phases of azoxymethane-induced colorectal carcinogenesis in rats fed low- or high-fat diets. Nutr Cancer. 1998;30:124–129. doi: 10.1080/01635589809514651. [DOI] [PubMed] [Google Scholar]
  • 111.Rijnkels JM, Hollanders VMH, Woutersen RA, Koeman JH, Alink GM. Interaction of dietary fat and of a vegetables/fruit mixture on 1,2-dimethylhydrazine-induced colorectal cancer in rats. Nutr Cancer. 1997;27:261–266. doi: 10.1080/01635589709514536. [DOI] [PubMed] [Google Scholar]
  • 112.Greenberg ER, Baron JA, Tosteson TD, Freeman DH, Beck GJ, Bond JH, Colacchio TA, Coller JA, Frankl HD, Haile RW, Mandel JS, Nierenberg DW, Rothstein R, Snover DC, Stevens MM, Summers RW, Vanstolk RU. Clinical trial of antioxidant vitamins to prevent colorectal adenoma. New Engl J Med. 1994;331:141–147. doi: 10.1056/NEJM199407213310301. [DOI] [PubMed] [Google Scholar]
  • 113.Angres G, Beth M. In: Effects of dietary constituents on carcinogenesis in different tumor models: an overview from 1975 to 1988 in Cancer and Nutrition. Alfin-Slater RB, Kritchevsky D, editors. Plenum Press; NY: 1991. pp. 337–485. [Google Scholar]
  • 114.McKeown-Eyssen G, Holloway C, Jazmaji V, Bright-See E, Dion P, Bruce WR. A randomized trial of vitamins C and E in the prevention of recurrence of colorectal polyps. Cancer Res. 1988;48:4701–4705. [PubMed] [Google Scholar]
  • 115.BonithonKopp C, Kronborg O, Giacosa A, Rath U, Faivre J. Calcium and fibre supplementation in prevention of colorectal adenoma recurrence: a randomised intervention trial. Lancet. 2000;356:1300–1306. doi: 10.1016/s0140-6736(00)02813-0. [DOI] [PubMed] [Google Scholar]
  • 116.Giovannucci E. The prevention of colorectal cancer by aspirin use. Biomed Pharm. 1999;53:303–308. doi: 10.1016/S0753-3322(00)88500-5. [DOI] [PubMed] [Google Scholar]
  • 117.Giovannucci E. Insulin and colon cancer. Cancer Causes Control. 1995;6:164–179. doi: 10.1007/BF00052777. [DOI] [PubMed] [Google Scholar]
  • 118.Willett WC. Polyunsaturated fat and the risk of cancer. Brit Medical J. 1995;311:1239–1240. doi: 10.1136/bmj.311.7015.1239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Lutz WK, Schlatter J. Chemical carcinogens and overnutrition in diet-related cancer. Carcinogenesis (Lond) 1992;13:2211–2216. doi: 10.1093/carcin/13.12.2211. [DOI] [PubMed] [Google Scholar]
  • 120.Willett WC. Diet and cancer: one view at the start of the millenium. Cancer Epidemiol Biomark Prev. 2001;10:3–8. [PubMed] [Google Scholar]
  • 121.McKeown-Eyssen G. Epidemiology of colorectal cancer revisited: are serum triglycerides and/or plasma glucose associated with risk ? Cancer Epidemiol Biomark Prev. 1994;3:687–695. [PubMed] [Google Scholar]
  • 122.Gerber M, BoutronRuault MC, Hercberg S, Riboli E, Scalbert A, Siess MH. Actualités en cancérologie: fruits, légumes et cancers, Une synthèse du réseau Nacre. Bull Cancer. 2002;89:293–312. [PubMed] [Google Scholar]
  • 123.WCRF. Nutrition and the prevention of cancer: a global perspective. Ed. World Cancer Research Fund and American Institute for Cancer Research. Menasha, USA: Banta Book Group; 1997. [Google Scholar]
  • 124.Michels KB, Giovannucci E, Joshipura KJ, Rosner BA, Stampfer MJ, Fuchs CS, Colditz GA, Speizer FE, Willett WC. Prospective study of fruit and vegetable consumption and incidence of colon and rectal cancers. J Natl Cancer Inst. 2000;92:1740–1752. doi: 10.1093/jnci/92.21.1740. [DOI] [PubMed] [Google Scholar]
  • 125.Harris GK, Gupta A, Nines RG, Kresty LA, Habib SG, Frankel WL, Laperle K, Gallaher DD, Schwartz SJ, Stoner GD. Effects of lyophilized black raspberries on azoxymethane-induced colon cancer and 8-hydroxy-2′-deoxyguanosine levels in the Fischer 344 rat. Nutr Cancer. 2001;40:125–133. doi: 10.1207/S15327914NC402_8. [DOI] [PubMed] [Google Scholar]
  • 126.Morris K. Combination rules in colorectal-cancer chemoprevention. The Lancet Oncology. 2000;1:66. [Google Scholar]
  • 127.Takayama T, Katsuki S, Takahashi Y, Ohi M, Nojiri S, Sakamaki S, Kato J, Kogawa K, Miyake H, Niitsu Y. Aberrant crypt foci of the colon as precursors of adenoma and cancer. New Engl J Med. 1998;339:1277–1284. doi: 10.1056/NEJM199810293391803. [DOI] [PubMed] [Google Scholar]

RESOURCES