You are on page 1of 19

Mutation Research 589 (2005) 47–65

www.elsevier.com/locate/reviewsmr
Community address: www.elsevier.com/locate/mutres
Review

Bile acids as carcinogens in human gastrointestinal cancers


H. Bernsteina,b, C. Bernsteina,*, C.M. Paynea,b, K. Dvorakovaa,b, H. Garewalb,c,d
a
Department of Microbiology and Immunology, College of Medicine, University of Arizona, Tucson AZ 85724, USA
b
Arizona Cancer Center, University of Arizona, Tucson, AZ 85724, USA
c
Department of Internal Medicine, Arizona Cancer Center, College of Medicine, University of Arizona, Tucson AZ 85724, USA
d
Tucson Veterans Affairs Medical Center, Section of Hematology/Oncology, Tucson, AZ 85723, USA
Received 26 January 2004; received in revised form 27 July 2004; accepted 6 August 2004
Available online 22 September 2004

Abstract

Bile acids were first proposed to be carcinogens in 1939 and 1940. On the basis of later work with rodent models, bile acids
came to be regarded as cancer promoters rather than carcinogens. However, considerable indirect evidence, obtained more
recently, supports the view that bile acids are carcinogens in humans. At least 15 reports, from 1980 through 2003, indicate that
bile acids cause DNA damage. The mechanism is probably indirect, involving induction of oxidative stress and production of
reactive oxygen species that then damage DNA. Repeated DNA damage likely increases the mutation rate, including the
mutation rate of tumor suppressor genes and oncogenes. Additional reports, from 1994 through 2002, indicate that bile acids, at
the increased concentrations accompanying a high fat diet, induce frequent apoptosis. Those cells within the exposed population
with reduced apoptosis capability tend to survive and selectively proliferate. That bile acids cause DNA damage and may select
for apoptosis-resistant cells (both leading to increased mutation), indicates that bile acids are likely carcinogens. In humans, an
increased incidence of cancer of the laryngopharyngeal tract, esophagus, stomach, pancreas, the small intestine (near the
Ampulla of Vater) and the colon are associated with high levels of bile acids. The much larger number of cell generations in the
colonic (and, likely, other gastrointestinal) epithelia of humans compared to rodents may allow time for induction and selection
of mutations leading to cancer in humans, although not in rodents.
# 2004 Elsevier B.V. All rights reserved.

Keywords: Bile acids; Carcinogen; Apoptosis; Gastrointestinal cancer; DNA damage; Oxidative stress

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

2. Bile acids in the body . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

3. Potentially more severe consequences of bile acid exposure in humans than in experimental rodent models . . . 50

* Corresponding author. Tel.: +1 520 626 6069; fax: +1 520 626 2100.
E-mail address: bernstein3@earthlink.net (C. Bernstein).

1383-5742/$ – see front matter # 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.mrrev.2004.08.001
48 H. Bernstein et al. / Mutation Research 589 (2005) 47–65

4. Bile acids are implicated in the etiology of human gastrointestinal cancer . . . . . . . . . . . . . . . . . . . . . . . . . . 52


4.1. Upper aerodigestive tract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.2. Esophagus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.3. Stomach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.4. Gallbladder and bile duct . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.5. Pancreas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.6. Small intestine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.7. Colon. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

5. Effects of vitamin D, the vitamin D receptor, and calcium on colon cancer and bile acid regulation . . . . . . . . 53

6. Mechanisms of actions of bile acids: bile acids introduce oxidative/nitrosative stress . . . . . . . . . . . . . . . . . . . 54

7. Mechanisms of actions of bile acids: bile acids induce DNA damage, a likely direct cause of cancer risk . . . . 55

8. Bile acids as likely mutagens in a pathway to gastrointestinal cancer. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

9. Bile acids cause apoptosis, which may select for apoptosis resistance and increased cancer risk . . . . . . . . . . . 58

10. Is it reasonable to hypothesize that bile acids, natural detergents with an important function in digestion,
can also be carcinogens? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

11. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

1. Introduction gens. However, the bile acids, by themselves, were not


carcinogens, failing to induce colon tumors. In
The bile acid, deoxycholic acid, was first proposed to particular, in rats, the bile acids lithocholic, taurodeoxy-
be a carcinogen in 1940 by Cook et al. [1], based on cholic and deoxycholic acids had a promoting effect
induction of tumors in mice when injected, in their on colon carcinogenesis after intrarectal instillation of
flanks, with this bile acid. These authors also cite a 1939 N-methyl-N0 -nitro-N-nitrosoguanidine (MNNG) [3–5].
meeting report by Vittorio Ghiron that ‘‘desoxycholic The bile acid, cholic acid, also had a promoting effect on
acid elicited transplantable subcutaneous fibro-sarco- colon tumor formation in rats after intrarectal instilla-
mas in a high proportion of the mice and rats injected. tion of N-methyl-N-nitrosourea (MNU) [6] or sub-
This is believed to be the first experimental production cutaneous injection of azoxymethane [7]. The promot-
of malignant growths with a compound that exists under ing effect of cholic acid may have been due to the
some conditions in the human body.’’ In a 1999 mouse formation of deoxycholic acid from cholic acid by
experiment, using mice with a germ line mutation in Apc bacterial action in the colon [8]. Further, taurocholate, in
(Min/+) as a model of familial adenomatous polyposis, rats, enhanced the induction by MNNG of hyperplastic
administration of chenodeoxycholate increased duode- and neoplastic lesions in stomach mucosa [9]. Based
nal tumors [2]. This increase occurred without addition largely on these experiments in non-mutated rat model
of a standard carcinogen such as MNNG, MNU or systems, it has been generally assumed that bile acids act
azoxymethane, so that in Min/+ mice, the bile acid as promoters, but not as carcinogens, in humans.
chenodeoxycholate was a carcinogen. Recently, however, considerable indirect evidence
In a number of rat experiments reported between and logical argument supports the view that bile acids
1974 and 1993, several bile acids were shown to be are carcinogens in humans. The evidence and
promoters, increasing tumorigenesis by known carcino- arguments are detailed below, and include evidence
H. Bernstein et al. / Mutation Research 589 (2005) 47–65 49

Fig. 1. Formation of bile acids. Chenodeoxycholic acid and cholic acid are formed in the liver from cholesterol. The bacterial enzyme 7-a
dehydroxylase converts chenodeoxycholic acid into lithocholic acid and converts cholic acid into deoxycholic acid.

that bile acids cause DNA damage, induce frequent Bile acids are avidly re-adsorbed by an active bile
apoptosis, and, after repeated exposure, select for cells salt re-absorptive mechanism in the terminal ileum,
with reduced apoptosis capability. In addition, high or with uptake into ileal columnar epithelium cells,
abnormal bile acid exposure is associated with leaving less than 5% of the bile salt pool to enter the
increased incidence of cancer in the laryngophar- colon [11,14]. After uptake into enterocytes of the
yngeal tract, esophagus, stomach, pancreas, the small ileum, bile salts are shuttled to the basolateral domains
intestine (near the Ampulla of Vater) and the colon. of the cells for efflux into the portal circulation,

2. Bile acids in the body

The primary bile acids, cholic acid and cheno-


deoxycholic acid, are derived from cholesterol,
primarily in the liver, through three main pathways
involving 10–14 different enzymes [10]. The struc-
tures of cholesterol, cholic acid and chenodeoxycholic
acid are shown in Fig. 1. After their synthesis in
hepatocytes, bile acids are excreted as C24 amides
conjugated with glycine or taurine [11]. The cyclical
path of bile acids in the body, after their synthesis in
the liver, is shown schematically in Fig. 2. In
particular, the conjugated bile salts are excreted from
the liver into the gall bladder, at a concentration of
approximately 100 mM [12], and then released into
the intestinal tract when fat enters the proximal portion
of the intestine [13]. Within the small intestine, the
bile acids are critically important for lipid absorption
in the ileum [14]. Fig. 2. The enterohepatic circulation of bile acids.
50 H. Bernstein et al. / Mutation Research 589 (2005) 47–65

Table 1
Bile acid concentrations (mM)a in fecal water of normal individuals on a normal diet [18] and a high fat diet [19]
Bile acid Fecal water (normal diet) Fecal water (high fat diet)
Cholic acid 0.011 (0–0.058) N.D.
Chenodeoxycholic acid 0.004 (0–0.017) N.D.
Deoxycholic acid 0.104 (0.046–0.210) 0.17 (0.07–0.73)
Lithocholic acid 0.013 (0.004–0.043) 0.09 (0.04–0.30)
Ursodeoxycholic acid 0.013 (0–0.039) N.D.
Total bile acids 0.184 (0.088–0.393) 0.36 (0.16–1.96)
a
Median values of bile acids from 25 normal individuals on a normal diet [18] or seven individuals on a high fat diet [19], with the range of
values in parentheses, N.D.: not determined.

transported by portal vein blood to the liver, and and after a high fat diet [19] are shown in Table 1. Bile
extracted from portal blood plasma by hepatocytes acids in the gallbladder [20] (later ejected into the
[11]. Fasting and postprandial levels of bile acids in small intestine, and subsequently diluted with food),
the portal venous blood are approximately 0.014 and or in esophageal aspirates [21] (from duodenal reflux)
0.043 mM, respectively [15]. The liver is efficient in are conjugated with taurine or glycine, and may be
removing bile acids from the circulation, so that sulfated in addition, as shown in Table 2. For persons
peripheral arterial blood serum of healthy individuals on a high fat diet, both deoxycholic acid and
has an approximate bile acid concentration of lithocholic acid may be substantially raised in the
0.003 mM [15]. fecal water in the colon (Table 1). Conjugated bile
There are approximately 6–12 enterohepatic acids are at much higher levels in the gallbladder, and
circulations per day, so that there is an average entry upon delivery, near the Ampulla of Vater, than bile
of 20% of the bile acid pool per day into the colon acids in any other region of the body (Tables 1 and 2).
[16]. Bile acids entering the colon are metabolized by Reflux of duodenal contents into the esophagus of
the anaerobic bacterial flora. Bacteria of the large patients can yield peak (short term) concentrations of
intestine carry out two major and a number of minor conjugated bile acids at levels as high as the maximum
reactions on the bile salts [17]. The first is bile acid concentration reported in the fecal water of
deconjugation to release free bile acids, which are the colon after a high fat meal (Tables 1 and 2). Based
only poorly ionized and are lipophilic. The second in part on DNA damaging effects, evidence of muta-
major reaction is 7-a dehydroxylation to yield genicity, and higher frequency in serum of adenoma
deoxycholic and lithocholic acids, respectively, from patients, the secondary bile acids, deoxycholic acid and
cholic and chenodeoxycholic acids (Fig. 1) [17]. lithocholic acid are considered the ones most likely to
Deoxycholic acid is partly absorbed in the colon and be implicated in colon carcinogenesis [16,17].
enters the enterohepatic circulation, where it is
conjugated in the liver and secreted in the bile [16].
Lithocholic acid is fairly insoluble and little of it is 3. Potentially more severe consequences of bile
reabsorbed [17]. Thus, the circulating bile acid pool acid exposure in humans than in experimental
(conjugated when it leaves the gallbladder, and then rodent models
de-conjugated by action of bacterial enzymes after it
enters the colon) is composed of about 30–40% each On the basis of the observation that bile acids by
of cholic acid and chenodeoxycholic acid, about 20– themselves do not induce colonic tumors in most
30% of deoxycholic acid, and less than 5% of rodent model systems, but do enhance the induction of
lithocholic acid [16]. Further bacterial degradation in tumors by carcinogens such as MNNG, MNU and
the colon and alterations in the liver produce tertiary azoxymethane, it is ordinarily assumed that in
bile acids, such as ursodeoxycholic acid, which also humans, like rodents, bile acids are promoters, not
enter the circulating bile acid pool. Bile acid carcinogens. However, there is a fundamental differ-
compositions, reported for the fecal water (which ence between the rodent models and human gastro-
bathes the colon wall), both with a normal diet [18] intestinal cancer. The rodent models and humans have
H. Bernstein et al. / Mutation Research 589 (2005) 47–65 51

Table 2
Mean bile acid concentrations in mM in the gallbladder [20] (126 normal individuals) and in esophageal aspirates [21] (24 patients)
Gallbladder (S.D.)ab Esophagus (S.E.M.)
Glycocholic acid 29.96 (22.56) 0.098 (0.034)
Taurocholic acid 24.62 (27.61) 0.013 (0.006)
Glycochenodeoxychlic acid 29.86 (38.25) 0.069 (0.025)
Taurochenodeoxycholic acid 18.46 (18.49) 0.013 (0.008)
Glycodeoxycholic acid 17.10 (16.25) 0.030 (0.011)
Taurodeoxycholic acid 30.44 (37.91) 0.005 (0.003)
Glycolithocholic acid 0.48 (0.88) 0.002 (0.002)
Taurolithocholic acid 0.40 (0.81) 0.0
Sulfoglycolithocholic acid 11.26 (19.67) N.D.
Sulfotaurolithocholic acid 0.56 (0.55) N.D.
Glycoursodeoxycholic acid 2.48 (1.28) 0.0
Tauroursodeoxycholic acid 0.52 (0.42) 0.0
Total bile acids 166.14 (70.94) N.D.
Maximum peak total bile acids (during reflux of patient) N.D. 2.015
a
Millimolar concentrations of individual bile acids were calculated from values given in ref. [20].
b
Abbreviations: S.D., standard deviation; S.E.M., standard error of the mean; N.D., not determined.

very different time periods of exposure to bile acids. rodent models does not rule out the possibility of a
The rodent experiments were carried out for one year carcinogenic effect of bile acids in humans.
or less (e.g. [1–5]). In humans, cancers of the digestive That MNNG, MNU and azoxymethane are
tract generally develop after many years of exposure. effective carcinogens over the time span of these
In humans, the mean age of occurrence of colon rodent experiments implies that they are potent
cancer is 60 years for men, and 59 years for women mutagens in rat colon. In contrast, bile acids may
[22]. For esophageal adenocarcinoma, the mean age is be much weaker mutagens, whose cumulative effect
64 years [23]. would only result in carcinogenesis after many years
The cell turnover times (defined as the time of exposure, as appears to be the case in humans.
necessary to replace the number of cells present in the Another factor making these rodent models of colon
entire cell population) of rapidly renewing cell carcinogenesis less than ideal for drawing the conclu-
populations are about the same regardless of life span sion that bile acids are non-carcinogens in humans is the
of mice (2.5 years), rats (3.5 years) and man (70 years) different spectrum of bile acids to which the GI tract of
[24]. Thus, rapidly renewing cell populations are rodents is ordinarily exposed compared to humans. For
replaced many more times during the life span of man instance, in rat feces, the predominant bile acids are
than mice or rats. Colonic epithelium renews itself hyodeoxycholic acid (23%), muricholic acid (b and v
roughly 365 times during the life span of mice and rats combined) (21%) and deoxycholate (17%) (see con-
and 5110 times in man [24]. At the mean age that ventional rats in [25]), whereas in human fecal water the
humans develop colon cancer (59–60 years) there predominant bile acids are deoxycholate (58%),
would have been about a 30-fold greater cell turnover lithocholate (12%) and cholic acid (4%) (calculated
in the human colon than in rodent models where the from individuals on a low fiber diet [26]) [see also,
animals were exposed to bile acids for a year or less. Table 1 (normal diet)]. In the rodent models where bile
Since carcinogenic mutations arise as the result of acids are administered in the diet [6,7] cholic acid is the
errors of DNA replication, the opportunity for such bile acid usually used. When bile acids are administered
mutations is at least 30-fold greater in humans than in by repeated intrarectal instillation, various bile acids
the rodent models. Thus, repeated exposure of the have been used; lithocholate and taurodeoxycholate [3],
colon to high concentrations of bile acids over a 59–60 deoxycholate [4], and cholate and chenodeoxycholate
year period in humans may be expected to have more [5]. Since the rat GI tract presumably is adapted to cope
severe consequences than a 1 year exposure in rodents. with a substantially different spectrum of bile acids than
The lack of carcinogenic effect in 1 year (or less) in that of humans, the rat response to these exposures
52 H. Bernstein et al. / Mutation Research 589 (2005) 47–65

probably does not closely mimic the response of the refluxate than normal individuals or those with
human GI tract to these bile acids. minimal esophagitis [29,30]. Furthermore, individuals
with early (T1 tumor stage) adenocarcinoma have an
even higher exposure to bile than individuals with
4. Bile acids are implicated in the etiology of uncomplicated Barrett’s metaplasia [31]. The concept
human gastrointestinal cancer that bile acids have a significant role in esophageal
adenocarcinoma has been suggested by numerous
Although bile acids have been predominantly investigators (e.g. [32–35]). Animal models and
studied in relation to promotion of colon cancer, it human studies suggest that esophageal mucosal injury
is now evident that bile acids are implicated as playing can arise from a synergistic damaging effect of
a role in carcinogenesis throughout the human gastro- conjugated bile acids (e.g. taurocholate and tauro-
intestinal (GI) tract. We consider it likely that bile deoxycholate) and HCl, as well as a synergistic
acids may act by a common underlying mechanism at interaction of unconjugated bile acids and trypsin at
various sites within the GI tract. However, conditions more neutral pH values [36].
vary widely from site to site within the GI tract. It is
certainly possible that at any given site some factor other 4.3. Stomach
than bile acid exposure, or in combination with bile acid
exposure, is more important in carcinogenesis, or that In the remnant stomach of rats after gastrectomy,
the underlying mechanisms of carcinogenesis may be bile acids, the main component of the duodenal juice,
different from site to site. have been implicated in gastric cancer due to
duodenogastric reflux [37]. In humans, duodenogas-
4.1. Upper aerodigestive tract tric reflux also appears to be implicated in gastric
stump carcinoma [38].
In the upper aerodigestive tract (pharynx and
larynx) bile acid exposure occurs during reflux arising 4.4. Gallbladder and bile duct
in the duodenum and passing through the stomach and
esophagus. This reflux contains not only bile acid from The gallbladder and bile duct are exposed to high
the duodenum, but also acid from the stomach. concentrations of bile acids, as discussed above in
Laryngopharyngeal reflux (LPR) is a common event in Section 2. The role of bile acids in gallbladder and bile
patients with head and neck cancer, suggesting that duct carcinogenesis is unclear. In a study carried out in
LPR has a role in the carcinogenic process of India, raised biliary deoxycholic acid and lithocholic
squamous cell carcinoma of the head and neck [27]. acid concentrations were present in patients with
The relative importance of acidity and bile acid carcinoma of the gallbladder, and it was suggested that
exposure has not yet been determined. Bile salts or this may be a factor in gallbladder carcinogenesis [39].
acidic conditions, or both, induce cyclooxygenase-2 On the other hand, in a study carried out in Mexico and
(COX-2) expression in normal pharyngeal mucosa Bolivia, patients with carcinoma of the gallbladder did
[28]. Since it is known that bile acid is associated with not have significantly raised conjugated deoxycholic
tumor formation in the esophagus and involves over- or lithocholic acids in the gallbladder [20]. Dietary
expression of COX-2, it was proposed that bile acids differences between India on the one hand, and Mexico
have a similar role in the tumorigenesis of the upper and Bolivia on the other, may have contributed to these
aerodigestive tract [28]. different observations.
Choledochal cysts are dilated lesions in the biliary
4.2. Esophagus tree, and biliary tract carcinoma occurs 5–35 times
more frequently in patients with these cysts than for
Barrett’s metaplasia of the esophagus is a major those without them [40]. The bile acid composition of
risk factor for the development of esophageal the cyst contents, in one case, showed 2% lithocholate,
adenocarcinoma. Individuals with Barrett’s metapla- 88% deoxycholate, 5% chenodeoxycholate, and 5%
sia have a higher exposure to bile acids in their cholate, with virtually all the bile acids in unconju-
H. Bernstein et al. / Mutation Research 589 (2005) 47–65 53

gated form. In contrast, the bile acids found in the environmental factor appears to be diet. Epidemiolo-
hepatic bile of this individual were fully conjugated, gical studies indicate a positive association between
and were composed of 0% lithocholate, 34% dietary fat consumption and colon cancer incidence
deoxycholate, 43% chenodeoxycholate and 23% (e.g. [49–54]). A high fat diet is associated with
cholate. The patient was found to have metaplasia increased bile acid secretion [55]. This association
of the epithelial lining of the cyst. These data presumably reflects the role of bile acids in
suggested that stasis of bile within choledochal cysts emulsifying dietary fat for absorption by the small
contributes to bacterial overgrowth and generation of intestine. Epidemiological studies have found that
unconjugated secondary bile acids [40]. However, fecal bile acid concentrations are elevated in popula-
another study of secondary bile acids in choledochal tions with a high incidence of colon cancer (e.g.
cysts in eleven patients did not find alterations in [17,54,56–63]). The most important bile acids in the
relative compsition or absolute concentrations of etiology of colon cancer in humans appear to be the
secondary bile acids [41]. secondary bile acids, deoxycholic acid and lithocholic
acid [17]. Deoxycholate is also higher in the serum of
4.5. Pancreas patients with colorectal adenomas than in individuals
without adenomas [64,65].
Epidemiological studies have demonstrated a
positive correlation between ingestion of a western-
style high fat diet and the incidence of pancreatic 5. Effects of vitamin D, the vitamin D receptor,
cancer [42–44]. Most adenocarcinomas of the and calcium on colon cancer and bile acid
pancreas occur in the head of the gland, which is in regulation
close proximity to bile. On the basis of the finding that
bile acids induce cyclooxygenase-2 expression in Epidemiologic evidence indicates that dietary
human pancreatic cancer cell lines, a possible role for intake of vitamin D and exposure to sunlight
bile acids in the pathogenesis of pancreatic cancer has contribute to the prevention of colon cancer [66,67].
been suggested [45]. The mechanisms by which vitamin D3 prevents colon
cancer may involve increased apoptosis [68],
4.6. Small intestine decreased proliferation [69,70], enhanced cellular
differentiation [71] and/or increased transcription that
About 57% of all adenocarcinomas of the small results from activation of the vitamin D3 receptor
intestine of humans occur in a 7 cm length which [72,73]. In addition, vitamin D3 may protect against
comprises less than 1% of the entire length of the oxidative DNA damage [69]. Some of the beneficial
small intestine [46]. The majority of these adeno- activities of vitamin D3 may ameliorate damaging
carcinomas were pinpointed to a small region around effects of bile acids. Of particular interest is the role of
the Ampulla of Vater, where bile and pancreatic the vitamin D3 receptor as a bile acid sensor [74].
secretions enter the small intestine. This finding led to Activation of the vitamin D3 receptor by the secondary
the suggestion that bile acids may be carcinogenic bile acid, lithocholic acid, or by vitamin D3 induces
[46]. Individuals with familial adenomatous polyposis expression of CYP3A, a cytochrome P450 enzyme
(FAP) have an increased risk for adenomas and that detoxifies lithocholic acid [74] through catabolic
cancers of the small and large intestines. In the small reactions (see review by Wolf [75]).
intestine these lesions also occur mainly around the Epidemiologic and animal studies indicate that
Ampulla of Vater, where their distribution parallels increased dietary intake of calcium lowers the
mucosal exposure to bile [47,48]. incidence of colon cancer [72,76–88]. The mechan-
isms by which calcium reduces colon cancer risk are
4.7. Colon probably multifactorial. Calcium may interact directly
with bile acids in the intestinal lumen [82] to form
The development of colon cancer in humans insoluble soaps [76], thereby reducing the deleterious
involves genetic and environmental factors. A major effects of bile acid interactions with epithelial cells.
54 H. Bernstein et al. / Mutation Research 589 (2005) 47–65

Calcium may also activate signal-transduction path- some of the generation of ROS results from a direct
ways that reduce cellular proliferation, enhance detergent effect of bile acids on membrane enzymes,
cellular differentiation and induce apoptosis [81–87]. such as phospholipase A2 (PLA2), which upon
activation release arachidonic acid. Arachidonic
acid is then acted on by enzymes such as cycloox-
6. Mechanisms of actions of bile acids: ygenase (COX) and lipoxygenase (LOX) to release
bile acids introduce oxidative/nitrosative ROS through the partial reduction of molecular
stress oxygen (O2), which accompanies formation of
prostaglandins and leukotrienes. Bile acid produc-
Deoxycholate and other hydrophobic bile acids tion of ROS has been shown to depend, in part, on
increase oxidative/nitrosative stress through the COX and LOX activity [89–92]. The ROS produced
generation of reactive oxygen/nitrogen species are likely, in turn, to cause oxidative DNA damage
(ROS/RNS). The source of the ROS/RNS is likely [93], in part, through the iron-catalyzed Fenton
multifactorial (see Fig. 3). As indicated in Fig. 3, reaction [92].

Fig. 3. Formation of ROS and RNS by hydrophobic bile acids and the pathways by which they induce DNA damage.
H. Bernstein et al. / Mutation Research 589 (2005) 47–65 55

ROS are also produced by mitochondria, as 7. Mechanisms of actions of bile acids: bile acids
indicated in Fig. 3. Mitochondria are known to be induce DNA damage, a likely direct cause of
damaged by bile acids [94–98], although the cancer risk
mechanism of this damage is not known. Mitochon-
drial damage due to bile acids can arise from several As described above, bile acids cause oxidative/
causes, as illustrated in Fig. 3; (1) the endogenous nitrosative stress and liberate ROS and RNS
generation of arachidonic acid [99], (2) truncated bid [91,93,95,96,98,102,113–115]. ROS have a plethora
(tbid) [100] from the ligand-independent activation of of biological effects, but perhaps of greatest sig-
the Fas receptor (FasR) [101], (3) Bak released from nificance, in terms of carcinogenesis, is their ability to
the stressed endoplasmic reticulum (ER), (4) ROS induce DNA damage. Oxidative DNA damage can
themselves, or (5) the decline of mitochondrial NAD+ cause mutation [116]. Bile acids also increase NOS2
levels as a result of DNA damage-induced poly(ADP- expression [117] leading to increased production of
ribose) polymerase (PARP) activity. reactive nitrogen species and increased DNA damage
Nitrosative stress, as illustrated in Fig. 3, can be [118].
generated through the activities of both NOS2 Table 3 lists 15 studies showing that the bile acids
(inducible nitric oxide synthase) and NOS3 (endothe- deoxycholate and lithocholate cause DNA damage in
lial NOS). Deoxycholate activates the redox-sensitive mammalian cells. These studies, plus additional
transcription factor, NF-kB [102], resulting in studies with bacteria and with isolated DNA, are
increased levels of NOS2 [103]. Bile acids also described next.
induce the activation of phospholipase C, which The bile acid lithocholic acid produces single-
releases inositol 3-phosphate (IP3) [104,105] resulting strand breaks in the DNA of mouse lymphoblastoma
in the release of Ca2+ from the ER [106]. The increase cells or of colon crypt cells [119–121]. Foreskin
in cytosolic Ca2+ can, in turn, activate the Ca2+/ fibroblasts treated with chenodeoxycholate and deox-
calmodulin-dependent NOS3 [107], known to be ycholate were found to experience significantly more
expressed in colon epithelial cells [108] in addition to DNA damage than untreated fibroblasts (as indicated
endothelial cells. NOS3, which produces RNS, can by increased DNA repair, measured by unscheduled
also be activated by hydrogen peroxide (H2O2) [109] DNA synthesis) [122]. DNA repair-deficient Chinese
resulting from the dismutation of O2. Deoxycholate hamster ovary (CHO) cells were found to be more
activates the peroxynitrite (ONOO) pathway, as sensitive than wild-type cells to killing by cheno-
evidenced by the formation of nitrotyrosine residues deoxycholate and deoxycholate, which again indicates
[110]. DNA damage may be caused by the direct induction of DNA damage by these agents [122]. Bile
action of NO and ONOO or the generation of the acids have also been shown to induce a DNA repair
hydroxyl radical (OH) through Fenton chemistry. NO response in bacteria [122,123]. Bile acids, or their
can also enhance oxidative DNA damage through the activated metabolites, have been reported to interact
S-nitrosylation and inhibition of DNA repair enzymes directly with naked DNA [124–126]. However, these
[111] that have a critical thiol group in their catalytic interactions may not be physiologically relevant to
site or interact with DNA through zinc-finger domains. bile acid exposure of mammalian cells, since bile
In addition, an important source of ROS generated acids may not gain direct access to nuclear DNA. In
by high levels of hydrophobic bile acids, is from the HeLa cells, deoxycholate and lithocholate activate the
action of microflora within the intestinal lumen. The gadd153 promoter, a promoter activated by DNA
bile tolerant Bacteroides strains, for example, synthe- damaging agents [127]. Deoxycholate also activates
size and utilize menaquinones, and vitamin K1 is a co- the gadd153 promoter in HepG2 cells [128]. Canine
factor in the dehydrogenation of bile acid substrates to bile containing a high level of deoxycholic acid,
form reduced K vitamins [112]. It has been suggested obtained in an experimental model for the anomalous
that these reduced K vitamins can be transported into arrangement of the pancreaticobiliary ducts, induced
mature colonocytes as mixed micelles and initiate DNA strand breaks in HeLa cells [129].
superoxide (O2) generating redox cycling reactions Extensive DNA damage was introduced into
[112]. colonic cells by treatment, at physiologic concentra-
56
Table 3
Bile acids cause DNA damage in mammalian cells
Bile acid(s)/salt(s) Evidence for DNA damage Source of DNA Year [ref.]
Lithocholate Single strand breaks (nucleoid Mouse lymphoblastoma cells 1980 [119]
sedimentation and alkaline elution)
Lithocholate Single strand breaks (nucleoid sedimentation) Mouse lymphoblastoma cells 1982 [120]
Lithocholate Single strand breaks (nucleoid sedimentation) Colon crypt cells 1985 [121]
Chenodeoxycholate Unscheduled DNA synthesis Human foreskin fibroblasts 1991 [122]

H. Bernstein et al. / Mutation Research 589 (2005) 47–65


deoxycholate
Chenodeoxycholate Increased cytotoxity when applied to two mutant Chinese Hamster Ovary cells 1991 [122]
deoxycholate cell lines with different DNA repair defects
Deoxycholate lithocholate Activation of gadd153 promoter, a promoter Human HeLa cells 1996 [127]
activated by DNA damage
Lithocholate deoxycholate DNA breakage seen by single cell electrophoresis Human colonic epithelial HT-29 cells 1997 [92]
‘‘Comet’’ assay; likely
oxidative DNA damage
Lithocholate deoxycholate Oxidative DNA damage seen by endonuclease III + Human colonic epithelial CACO-2 cells 1997 [93]
‘‘Comet’’ assay
Deoxycholate DNA strand breaks seen by nick translation method Canine bile used to test strand breaks in HeLa cells 1997 [129]
Lithocholate DNA breakage seen by single cell electrophoresis Human colon mucosal cells (from biopsies) 1997 [132]
‘‘Comet’’ assay
Lithocholate deoxycholate DNA breakage seen by single cell electrophoresis Human colonic epithelial HT-29 cells 1998 [133]
‘‘Comet’’ assay; likely
oxidative DNA damage
Deoxycholate Induction of DNA repair enzyme PARP (responds Human Jurkat cells, rat colon epithelial cells 1998 [102]
to breaks in DNA)
Deoxycholate Activation of gadd153 promoter, a promoter HepG2 cells 1999 [128]
activated by DNA damage
Deoxycholate DNA breakage seen by single cell electrophoresis Human colonic epithelial HCT-116 cells (p53+/+), HCT-15 cells (p53/) 2001 [130]
‘‘Comet’’ assay
Deoxycholate Single strand breaks seen by single cell electrophoresis Human colonic epithelial HT-29 cells, HCT-116 cells 2002 [131]
‘‘Comet’’ assay
Deoxycholate Induces DNA repair gene BRCA-1 Human colonic epithelial HCT-116 cells 2003 [138]
H. Bernstein et al. / Mutation Research 589 (2005) 47–65 57

tions, with deoxycholate [92,130,131] or lithocholate Lithocholate inhibits polymerase b, an enzyme


[92,132] as measured by the comet assay to detect and required for DNA base excision repair [137]. This
visualize DNA damage (strand breaks). Deoxycho- inhibition should have the effect of increasing the level
late-induced DNA damage could be reduced by of unrepaired DNA damages. Deoxycholate, at low
inhibiting release of a membrane-bound phospholipid, doses, increases expression of the DNA repair gene
arachidonic acid, and by inhibiting metabolism of BRCA-1 at the mRNA and protein level, suggesting
arachidonic acid through the lipoxygenase pathway that DNA damage caused by deoxycholate may up-
[133]. The lipoxygenase pathway, through a one regulate this repair protein [138].
electron reduction of molecular O2 produces the Deoxycholate has also been shown to cause the
superoxide free radical. Similarly, incubation of cells degradation of p53 [139]. The functions of p53 in
with CuDIPSH [copper(II) 3,5-diisopropyl salicylate response to DNA damage include, first, inducing cell
hydrate], a superoxide dismutase mimetic compound cycle arrest to allow time for DNA repair; second,
(which removes superoxide), also protects cells acting directly in DNA repair processes, and third,
against deoxycholate-induced DNA damage [133]. promoting apoptosis when DNA damage is excessive
In addition, the antioxidant, vitamin E, a potent [140]. Thus, a reduction in p53 by deoxycholate
inhibitor of lipid peroxidation, reduced deoxycholate- should increase genomic instability, because of
and lithocholate-induced DNA damage close to reduced DNA repair and reduced apoptotic removal
negative control values [92]. These results indicate of cells with DNA damage.
that ROS produced during lipid peroxidation may be a
major source of DNA damage produced by deox-
ycholate and lithocholate. Deoxycholate and litho- 8. Bile acids as likely mutagens in a pathway to
cholate, at levels often found in patients with colon gastrointestinal cancer
cancer associated pathology, produced DNA damage
as measured by the comet assay [93]. Endonuclease Spontaneous mutagenesis in mammalian cells
III, an enzyme which nicks DNA specifically at sites of appears to be caused mainly by oxidative events
oxidized pyrimidines, is used in the comet assay to [141]. We reviewed, above, the evidence that bile
determine whether the DNA damage is oxidative in acids cause DNA damage, particularly oxidative
nature. Endonuclease III treatment significantly DNA damage. A certain fraction of the oxidative
increased the genotoxicity of the bile acids [93], DNA damages induced by repeated exposure to high
again suggesting that the DNA damage is caused by levels of bile acids can be expected to remain
ROS. unrepaired until the next cycle of DNA replication.
Activation of poly(ADP-ribose) polymerase Such damages will cause replication errors as the
(PARP) is an immediate cellular reaction to DNA replication apparatus inaccurately copies the
strand breakage as induced by oxidizing agents as well damaged template [142]. Such replication errors lead
as other agents [134]. The resulting formation of to mutation, and some of the mutations will cause
protein-coupled poly(ADP-ribose) facilitates survival aberrant expression of oncogenes and tumor sup-
of proliferating cells under conditions of DNA pressor genes, leading to cancer (Fig. 4). In addition,
damage, probably via its contribution to DNA base- oxidative stress inactivates the human mismatch
excision repair [134]. Deoxycholate activates PARP, repair system [143] and this would also be expected
and specific inhibition of PARP sensitizes cells to the to enhance mutagenesis. In an azoxymethane-treated
induction of apoptosis by deoxycholate [102]. These rat model of colon tumorigenesis, deoxycholate
findings suggest that deoxycholate produces DNA increased the incidence of colon tumors as well as
strand breaks whose repair is mediated by PARP. the incidence of K-ras point mutations in the tumors
Deoxycholate and chenodeoxycholate, at physiologic [144]. This suggests that the promotion effect of bile
concentrations, stimulate hydroxylation of DNA bases acids on colon carcinogenesis in rodents may involve
mediated by the hydroxyl radical (a highly reactive induction of K-ras mutations. The bile acid litho-
oxygen species) to form 8-hydroxydeoxyadenine and cholic acid has also been reported to induce cell
8-hydroxydeoxyguanine [135,136]. transformation of CHO cells [145], which could
58 H. Bernstein et al. / Mutation Research 589 (2005) 47–65

conjugates were mutagenic in the standard Ames


Salmonella test system and in a human lymphoblast
cell line [149].
The standard mutation experiment, performed on
petri dishes, is effective for detecting an induced
mutation rate only if it is considerably more than 10
times the spontaneous mutation rate [150]. Thus, this
approach is relatively insensitive to moderate or weak
mutagens. To avoid this limitation, a test was
developed based on the classical Luria and Delbruck
fluctuation test, which is approximately 100-fold more
sensitive than the conventional tests [150]. Using this
more sensitive bacterial fluctuation test, deoxycholate,
chenodeoxycholate and cholic acid were found to be
significantly mutagenic in one study [151], and
deoxycholate and cholic acid were non-mutagenic
in another study [152].

9. Bile acids cause apoptosis, which may select for


apoptosis resistance and increased cancer risk

Deoxycholate is the bile acid present at highest


Fig. 4. Pathway to GI cancer through bile acid induced DNA
damage.
concentration in the human colon (Table 1 and [153]).
Deoxycholate causes production of ROS (see Section
6, above), and ROS are a common mediator of
indicate induction of mutations or stable epigenetic apoptosis [154]. Apoptosis is induced by deoxycholate
alterations. Bile acids were also reported to increase in mammalian cells at concentrations comparable to
the mutation frequency in transgenic rat fibroblasts at the bile acid concentration found in the colon after a
pH 5, but not at pH 7 [146]. high fat meal [127,130,131,155–166]. DNA damage
Attempts have been made to measure the muta- appears to be an initiating event in deoxycholate-
genicity of bile acids using the Ames Salmonella induced apoptosis [131].
mutagen test system. In the standard experiment, It might be expected that frequent exposure of the
which measures induced reversion to amino acid colonic epithelium to high concentrations of cytotoxic
independence in an auxotrophic strain of bacteria, the bile acids would allow selective growth of cells
cells are treated with mutagen and plated on selective resistant to bile acid-induced apoptosis, leading to an
agar containing a growth limiting supplement of the apoptosis resistant cell population. In a rat model,
particular amino acid. Of 30 bile acids tested in the repeated feeding of a diet containing 0.2% cholic acid
standard Ames Salmonella mutagen test system, none resulted in the development of increased resistance to
was mutagenic by itself, but lithocholic acid was co- apoptosis of the colon crypt cells both in aberrant crypt
mutagenic with 2-aminoanthracene activated by foci and normal crypts [167]. In addition, repeated
phenobarbital-stimulated rat liver homogenate long-term exposure of a human colonic epithelial cell
[147]. Deoxycholic acid and lithocholic acid were line to sublethal concentrations of deoxycholate
also co-mutagenic towards 1,2-dimethylhydrazine selected for cells that are resistant to deoxycholate
[148]. The N-nitroso bile acid conjugates N-nitroso- induced apoptosis [168]. Most importantly, when
taurocholic acid and N-nitrosoglycocholic acid are epithelial cells of the flat non-neoplastic colonic
formed naturally by a reaction of bile acids with N- mucosa of individuals with colon cancer are evaluated,
nitrosamides catalyzed by bacterial enzymes. These they are found to have reduced capacity to undergo
H. Bernstein et al. / Mutation Research 589 (2005) 47–65 59

deoxycholate-induced apoptosis [154,160,164,165]. tries or of vegetarians, where the incidence of colon


The reduction in apoptosis capability in cancer patients cancer is low [175,176]. Thus, the high colon cancer
would be expected if repeated long term exposure to incidence in Western countries appears to be largely a
high levels of bile acids selected for apoptosis resis- reflection of recent improvements associated with
tance. Loss of apoptosis capability is associated with an increased life span. Cancer of the digestive tract was
increased rate of mutation [169–171]. This may probably a negligible problem for our hunter-gatherer
predispose to development of malignancy. However, ancestors ([173], p. 10) and thus would have exerted
it is not known if the development of apoptosis insignificant selective pressure for an alternative
resistance precedes development of malignancy as manner of fat digestion.
expected by this view, or that the apoptosis resistance
results from selection by bile acids.
11. Conclusions

10. Is it reasonable to hypothesize that bile acids, The evidence reviewed here indicates that bile
natural detergents with an important function in acids play a key etiologic role in gastrointestinal
digestion, can also be carcinogens? cancers. Experiments in rodent model systems
indicate that bile acids act as promoters, but are
A possible objection to the hypothesis that bile not, by themselves, carcinogenic. However, the
acids are carcinogenic could be made on evolutionary rodent experiments do not preclude the possibility
grounds. It may seem contrary to our understanding of that bile acids act as carcinogens in humans,
how natural selection operates that a natural substance considering the great difference in time of tumor
produced by the body for a beneficial purpose (the development in rodent models compared to humans.
emulsification of dietary fats to aid in their digestion) Many different studies indicate that bile acids cause
could be carcinogenic. Would not natural selection DNA damage, strongly suggesting mutagenic and
have operated to allow an alternative benign solution carcinogenic potential. The DNA damaging mechan-
to the problem of fat digestion, if bile acids reduce ism seems to be indirect, involving production of
fitness by causing cancer? In answer to this objection ROS/RNS. Several mutagenesis experiments indi-
we make the following points. cate that bile acids are mutagenic. On the basis of
Colon cancer is mainly a disease of individuals over several lines of evidence we consider it likely that
age 50 [172]. However, the hunter-gatherer ancestors of bile acids are carcinogens in human gastrointestinal
modern humans probably had an average life span of cancers.
about 40 years ([173], p. 15), and those individuals who Some common types of non-gastrointestinal
survived past 50 years were well past their reproductive cancers (e.g. lung and bronchus, prostate, urinary
prime. Thus, in our ancestors, the influence of colon bladder) are strongly age-related, where a 70-year-old
cancer on natural selection would have been weak. human is roughly 50–100 times as likely to be
Life span in Western societies began to increase in diagnosed with a malignancy as a 27-year-old [172].
the 18th century, apparently due to such factors as Major gastrointestinal cancers also are strongly age-
improved sanitation and agricultural development related [e.g. the relative incidence for a 70-year-old
[173]. The substantial increase in life span in the 19th compared to a 27-year-old is approximately 27-fold
and 20th centuries resulted largely from a decline in (colon and rectum), 62-fold (esophagus), 71-fold
death rates due to infectious disease [173]. Important (pancreas) and 79-fold (larynx)] [172]. Thus, age-
factors influencing the incidence of gastrointestinal related gastrointestinal cancers are members of
cancers in current, longer lived populations may be a a larger group of strongly age-related cancers.
decline in beneficial calorie restriction and increased Perhaps age-related cancers generally reflect chronic
meat consumption [174]. The exposure of the (decades-long) exposure to specific agents that cause
gastrointesinal tract to bile acids in the ancestors of cellular stress, DNA damage and accumulation of
Western man may have been similar to that of mutations, as appears to be the case for gastrointestinal
contemporary populations in underdeveloped coun- cancers.
60 H. Bernstein et al. / Mutation Research 589 (2005) 47–65

Acknowledgments chromatography-electrospray tandem mass spectrometry, J.


Lipid Res. 42 (2001) 114–119.
[13] G.T. Krishnamurthy, P.H. Brown, Comparison of fatty meal
This work was supported in part by NIH and intravenous cholecystokinin infusion for gallbladder
Institutional Core Grant #CA23074, NIHPPG ejection fraction, J. Nucl. Med. 43 (2002) 1603–1610.
#CA72008, Arizona Disease Control Research Com- [14] R.N. Redinger, The coming of age of our understanding of the
mission Grants #10016 and #6002, VAH Merit Review enterohepatic circulation of bile salts, Am. J. Surg. 185
Grant 2HG, NCI SPORE Grant 1 P50CA95060-01 (2003) 168–172.
[15] B. Angelin, I. Bjorkhem, K. Einarsson, S. Ewerth, Hepatic
and Biomedical Diagnostics & Research Inc., Tucson, uptake of bile acids in man; fasting and postprandial con-
AZ. centrations of individual bile acids in portal venous and
systemic blood serum, J. Clin. Invest. 70 (1982) 724–731.
[16] F.M. Nagengast, M.J.A.L. Grubben, I.P. van Munster, Role of
References bile acids in colorectal carcinogenesis, Eur. J. Cancer 31
(1995) 1067–1070.
[1] J.W. Cook, E.L. Kennaway, N.M. Kennaway, Production of [17] M.J. Hill, Bile flow and colon cancer, Mutat. Res. 238 (1990)
tumours in mice by deoxycholic acid, Nature 145 (1940) 627. 313–320.
[2] N.N. Mahmoud, A.J. Dannenberg, R.T. Bilinski, J.R. Mestre, [18] T.M.C.M. De Kok, A. Van Faassen, B. Glinghammar, D.M.F.A.
A. Chadburn, M. Churchill, C. Martucci, M.M. Bertagnolli, Pachen, M. Eng, J.J. Rafter, C.G.M.I. Baeten, L.G.J.B. Engels,
Administration of an unconjugated bile acid increases duo- J.C.S. Kleinjans, Bile acid concentrations, cytotoxicity, and pH
denal tumors in a murine model of familial adenomatous of fecal water from patients with colorectal adenomas, Digest.
polyposis, Carcinogenesis 20 (1999) 299–303. Dis. Sci. 44 (1999) 2218–2225.
[3] T. Narisawa, N.E. Magadia, J.H. Weisburger, E.L. Wynder, [19] J. Stadler, H.S. Stern, K.S. Yeung, V. McGuire, R. Furrer, N.
Promoting effect of bile acids on colon carcinogenesis after Marcon, W.R. Bruce, Effect of high fat consumption on cell
intrarectal instillation of N-methyl-N0 -nitro-N-nitrosoguani- proliferation activity of colorectal mucosa and on soluble
dine in rats, J. Natl. Cancer Inst. 53 (1974) 1093–1097. faecal bile acids, Gut 29 (1988) 1326–1331.
[4] B.S. Reddy, T. Narasawa, J.H. Weisburger, E.L. Wynder, [20] B.L. Strom, R.D. Soloway, J.L. Rios-Dalenz, H.A. Rodri-
Promoting effect of sodium deoxycholate on colon adeno- guez-Martinez, S.L. West, J.L. Kinman, R.S. Crowther, D.
carcinomas in germ-free rats, J. Natl. Cancer Inst. 56 (1976) Taylor, M. Polansky, J.S. Berlin, Biochemical epidemiology
441–442. or gallbladder cancer, Hepatology 23 (1996) 1402–1411.
[5] B.S. Reddy, K. Watanabe, J.H. Weisburger, E.L. Wynder, [21] W.K.H. Kauer, J.H. Peters, T.R. DeMeester, H. Feussner, A.P.
Promoting effect of bile acids in colon carcinogenesis in Ireland, H.J. Stein, R.J. Siewert, Composition and concentra-
germ-free and conventional F344 rats, Cancer Res. 37 (1977) tion of bile acid reflux into the esophagus of patients with
3238–3242. gastroesophageal reflux disease, Surgery 122 (1997) 874–881.
[6] C.K. McSherry, B.I. Cohen, V.D. Bookenheuser, E.H. Mos- [22] E.L. Wynder, T. Shigematsu, Environmental factors of cancer
bach, J. Winter, N. Matoba, J. Scholes, Effects of calcium and of the colon and rectum, Cancer 20 (1967) 1520–1561.
bile acid feeding on colon tumors in the rat, Cancer Res. 49 [23] A.J. Cameron, C.T. Lomboy, Barrett’s esophagus: age pre-
(1989) 6039–6043. valence and extent of columnar epithelium, Gastroenterology
[7] B.A. Magnuson, I. Carr, R.P. Bird, Ability of aberrant crypt 103 (1992) 1241–1245.
foci characteristics to predict colonic tumor incidence in rats [24] P.D. Bowman, Aging and the cell cycle in vivo and in vitro, in:
fed cholic acid, Cancer Res. 53 (1993) 4499–4504. V.J. Cristofolo (Ed.), Handbook of Cell Biology of Aging,
[8] W.F. Weidema, E.E. Deschner, B.I. Cohen, J.J. DeCosse, CRC Press, Boca Raton, Florida, 1985, pp. 117–136.
Acute effects of dietary cholic acid and methylazoxymetha- [25] H.J. Eyssen, G. DePauw, J. Van Eldere, Formation of hyo-
nol acetate on colon epithelial cell proliferation; metabolism deoxycholic acid from muricholic acid and hyocholic acid by
of bile salts and neutral sterols in conventional and germfree an unidentified gram-positive rod termed HDCA-1 isolated
SD rats, J. Natl. Cancer Inst. 74 (1985) 665–670. from rat intestinal microflora, Appl. Environ. Microbiol. 65
[9] O. Kobori, T. Shimizu, M. Maeda, Y. Atomi, J. Watanabe, Y. (1999) 3158–3163.
Morioka, Enhancing effect of bile and bile acid on stomach [26] D.S. Alberts, J.G. Einspahr, D.L. Earnest, M.F. Krutzsch, P.
tumorigenesis induced by N-methyl-N0 -nitro-N-nitrosoguanide Pin, L.M. Hess, D.K. Heddens, D.J. Roe, M.E. Martinez, G.
in Wistar rats, J. Natl. Cancer Inst. 73 (1984) 853–861. Salen, A.K. Batta, Fecal bile acid concentrations in a sub-
[10] I. Bjorkhem, G. Eggertsen, Genes involved in initial steps of population of the wheat bran fiber colon polyp trial, Cancer
bile acid synthesis, Curr. Opin. Lipidol. 12 (2001) 97–103. Epidemiol., Biomark. Prevent. 12 (2003) 197–200.
[11] G.A. Kullak-Ublick, B. Stieger, P.J. Meier, Enterohepatic bile [27] J.S. Lewin, A.M. Gillenwater, J.D. Garrett, J.K. Bishop-
salt transporters in normal physiology and liver disease, Leone, D.D. Nguyen, D.L. Callender, G.D. Ayers, J.N.
Gastroenterology 126 (2004) 322–342. Myers, Characterization of laryngopharyngeal reflux in
[12] S. Perwaiz, B. Tuchweber, D. Mignault, T. Gilat, I.M. Yousef, patients with premalignant or early carcinomas of the larynx,
Determination of bile acids in biological fluids by liquid Cancer 97 (2003) 1010–1014.
H. Bernstein et al. / Mutation Research 589 (2005) 47–65 61

[28] M.W. Sung, J.L. Roh, B.J. Park, S.W. Park, T.K. Kwon, S.J. [44] E.L. Wynder, K. Mabuchi, N. Maruchi, J.G. Fortner, Epide-
Lee, Bile acid induces cyclo-oxygenase-2 expression in miology of cancer of the pancreas, J. Natl. Cancer Inst. 50
cultured human pharyngeal cells: A possible mechanism of (1973) 645–667.
carcinogenesis in the upper aerodigestive tract by laryngo- [45] O.N. Tucker, A.J. Dannenberg, E.K. Yang, T.J. Fahey, Bile acids
pharyngeal reflux, Laryngoscope 113 (2003) 1059–1063. induce cyclooxygenase-2 expression in human pancreatic can-
[29] D. Nehra, P. Howell, C.P. Williams, J.K. Pye, J. Beynon, cer cell lines, Carcinogenesis 25 (2004) 419–423.
Toxic bile acids in gastro-esophageal reflux disease: influence [46] R.K. Ross, N.M. Hartnett, L. Bernstein, B.E. Henderson,
of gastric acidity, Gut 44 (1999) 598–602. Epidemiology of adenocarcinomas of the small intestine: is bile
[30] M. Menges, M. Muller, M. Zeitz, Increased acid and bile a small bowel carcinogen, Br. J. Cancer 63 (1991) 143–145.
reflux in Barrett’s esophagus compared to reflux esophagitis, [47] D.G. Jagelman, J.J. DeCosse, H.J.R. Bussey, Upper gastro-
and effect of proton pump inhibitor therapy, Am. J. Gastro- intestinal cancer in familial adenomatous polyposis, Lancet 1
enterol. 96 (2001) 331–337. (1988) 1149–1151.
[31] H.J. Stein, W.K. Kauer, H. Feussner, J.R. Siewert, Bile reflux [48] A.D. Spigelman, C.B. Williams, I.C. Talbot, P. Domizio,
in benign and malignant Barrett’s esophagus: effect of med- R.K.S. Philips, Upper gastrointestinal cancer in patients
ical acid suppression and nissen fundoplication, J. Gastro- with familial adenomatous polyposis, Lancet 2 (1989)
intestinal. Surg. 2 (1998) 333–341. 783–785.
[32] P. Gillen, P. Keeling, P.J. Byrne, M. Healy, M.M. Omoore, [49] B.S. Drasar, D. Irving, Environmental factors and cancer of
T.P.J. Hennessey, Implication of duodenogastric reflux in the the colon and breast, Br. J. Cancer 27 (1973) 167–172.
pathogenesis of Barrett’s oesophagus, Br. J. Surg. 75 (1988) [50] E.G. Knox, Foods and diseases, Br. J. Prev. Soc. Med. 31
540–543. (1977) 71–80.
[33] X. Chen, C.S. Yang, Esophageal adenocarcinoma: a review [51] A.B. Miller, G.R. Howe, M. Jain, K.J.P. Craib, L. Harrison,
and perspectives on the mechanism of carcinogenesis and Food items and food groups as risk factors in a case-control
chemoprevention, Carcinogenesis 22 (2001) 1119–1129. study of diet and colo-rectal cancer, Int. J. Cancer 32 (1983)
[34] D.H. Stamp, Three hypotheses linking bile to carcinogenesis 155–161.
in the gastrointestinal tract: certain bile salts have properties [52] G.E.M. McKeown-Eyssen, E. Bright-See, Dietary factors in
that may be used to complement chemotherapy, Med. Hypo. colon cancer: international relationships, Nutr. Cancer 6
59 (2002) 398–405. (1984) 160–170.
[35] W.K. Kauer, H.J. Stein, Role of acid and bile in the genesis of [53] W. Willett, The search for the causes of breast and colon
Barrett’s esophagus, Chest Surg. Clin. North Am. 12 (2002) cancer, Nature 338 (1989) 389–394.
39–45. [54] P.Y. Cheah, Hypotheses for the etiology of colorectal can-
[36] M.F. Vaezi, S. Singh, J.E. Richter, Role of acid and duode- cer—an overview, Nutr. Cancer 14 (1990) 5–13.
nogastric reflux in esophageal mucosal injury: a review of [55] B.S. Reddy, Diet and excretion of bile acids, Cancer Res. 41
animal and human studies, Gastroenterology 108 (1995) (1981) 3766–3768.
1897–1907. [56] M.J. Hill, B.S. Drasar, V.C. Aries, J.S. Crowther, G.S. Hawks-
[37] A. Kuwahara, T. Saito, M. Kobayashi, Bile acids promote worth, R.E.O. Williams, Bacteria and etiology of cancer of
carcinogenesis in the remnant stomach of rats, J. Cancer Res. large bowel, Lancet 1 (1971) 95–100.
Clin. Oncol. 115 (1989) 423–428. [57] M.J. Hill, A.J. Taylor, M.H. Thompson, R. Wait, Faecal
[38] K. Kondo, Duodenogastric reflux and gastric stump carci- steroids and urinary volatile phenols in four Scandinavian
noma, Gastr. Cancer 5 (2002) 16–22. populations, Nutr. Cancer 4 (1982) 67–73.
[39] V.K. Shukla, S.C. Tiwari, S.K. Roy, Biliary bile acids in [58] J.S. Crowther, B.S. Drasar, M.J. Hill, R. Maclennan, D.
cholelithiasis and carcinoma of the gall bladder, Eur. J. Magnin, S. Peach, C.H. Teoh-Chan, Faecal steroids and
Cancer Prev. 2 (1993) 155–160. bacteria and large bowel cancer in Hong Kong by socio-
[40] R.M. Reveille, G. Van Stiegmann, G.T. Everson, Increased economic groups, Br. J. Cancer 34 (1976) 191–198.
secondary bile acids in a choledochal cyst. Possible role in [59] B.S. Reddy, E.L. Wynder, Large bowel carcinogenesis: faecal
biliary metaplasia and carcinoma, Gastroenterology 99 constituents of populations with diverse incidence ratio of
(1990) 525–527. coln cancer, J. Natl. Cancer Inst. 50 (1973) 1437–1442.
[41] K. Chijiiwa, E. Nagai, I. Makino, K. Shimada, are secondary [60] B.S. Reddy, E.L. Wynder, Metabolic epidemiology of colon
bile acids in choledochal cysts important as a risk factor in cancer: fecal bile acids and neutral sterols in colon cancer
biliary tract carcinoma? Aust. N.Z. J. Surg. 63 (1993) 109–112. patients and patients with adenomatous polyps, Cancer 39
[42] P. Ghadirian, A. Simard, J. Baillargeon, P. Maisonneuve, P. (1977) 2533–2539.
Boyle, Nutritional factors and pancreatic cancer in the fran- [61] B.S. Reddy, A.R. Hedges, K. Laakso, E.L. Wynder, Meta-
cophone community in Montreal, Canada, Int. J. Cancer 47 bolic epidemiology of large bowel cancer: fecal bulk and
(1991) 1–6. constituents of high-risk North American and low-risk Fin-
[43] M. Binstock, D. Krakow, J. Stamler, J. Reiff, V. Persky, K. nish population, Cancer 42 (1978) 2832–2838.
Liu, D. Moss, Coffee and pancreatic cancer: an analysis of [62] O.M. Jensen, R. MacLennan, J. Wahrendorf, Diet, bowel
international mortality data, Am. J. Epidemiol. 118 (1983) function, fecal characteristics, and large bowel cancer in
630–640. Denmark and Finland, Nutr. Cancer 4 (1982) 5–19.
62 H. Bernstein et al. / Mutation Research 589 (2005) 47–65

[63] L. Domellof, L. Darby, D. Hanson, L. Mathews, B. Simi, B.S. [78] H.L. Newmark, M. Lipkin, N. Maheshwari, Colonic hyper-
Reddy, Fecal sterols and bacterial b-glucuronidase activity: a proliferation induced in rats and mice by nutritional-stress
preliminary metabolic epidemiology study of healthy volun- diets containing four components of a human Western-style
teers from Umea Sweden and Metropolitan New York, Nutr. diet (series 2), Am. J. Clin. Nutr. 54 (1991) 209S–214S.
Cancer 4 (1982) 120–127. [79] H.L. Newmark, M. Lipkin, Calcium, vitamin D, and colon
[64] E. Bayerdorffer, G.A. Mannes, T. Ochsenkuhn, P. Dirschedl, cancer, Cancer Res. 52 (Suppl.) (1992) 2067–2070.
G. Paumgartner, Variation of serum bile acids in ptients with [80] M. Lipkin, H. Newmark, Calcium and the prevention of colon
colorectal adenomas during a one-year follow-up, Digestion cancer, J. Cell. Biochem. 22 (Suppl.) (1995) 65–73.
55 (1994) 121–129. [81] J.F. Whitfield, R.P. Bird, B.R. Chakravarthy, R.J. Isaacs, P.
[65] E. Bayerdorffer, T. Mannes, T. Ochsenkuhn, P. Dirschedl, B. Morley, Calcium-cell cycle regulator, differentiator, killer,
Wiebecke, G. Paumgartner, Unconjugated secondary bile chemopreventor, and maybe, tumor promoter, J. Cell. Bio-
acids in the serum of patients with colorectal adenomas, chem. 22 (Suppl.) (1995) 74–91.
Gut 36 (1995) 268–273. [82] J.R. Lupton, G. Steinbach, W.C. Chang, B.C. O’Brien, S.
[66] T.E. Robsahm, S. Tretli, A. Dahlback, J. Moan, Vitamin D3 Wiese, C.L. Stoltzfus, G.A. Glober, M.J. Wargovich, R.S.
from sunlight may improve the prognosis of breast-, colon- McPherson, R.J. Winn, Calcium supplementation modifies
and prostate cancer (Norway), Cancer Causes Control 15 the relative amounts of bile acids in bile and affects key aspects
(2004) 149–158. of human colon physiology, J. Nutr. 126 (1996) 1421–1428.
[67] C.F. Garland, F.C. Garland, E.D. Gorham, Calcium and [83] R. Van der Meer, J.A. Lapre, M.J.A.P. Govers, J.H. Kleibeu-
vitamin D. Their potential roles in colon and breast cancer ker, Mechanisms of the intestinal effects of dietary fats and
prevention, Annal. N.Y. Acad. Sci. 889 (1999) 107–119. milk products on colon carcinogenesis, Cancer Lett. 114
[68] T.F. McGuire, D.L. Trump, C.S. Johnson, Vitamin D3-induced (1997) 75–83.
apoptosis of murine squamous cell carcinoma cells. Selective [84] M.J. Wargovich, A. Jimenez, K. McKee, V.E. Steele, M.
induction of caspase-dependent MEK cleavage and up-regula- Velasco, J. Woods, R. Price, K. Gray, G.J. Kelloff, Efficacy of
tion of MEKK-1, J. Biol. Chem. 276 (2001) 26365–26373. potential chemopreventive agents on rat colon aberrant crypt
[69] E. Kallay, P. Bareis, E. Bajna, S. Kriwanek, E. Bonner, S. formation and progression, Carcinogenesis 21 (2000) 1149–
Toyokuni, H.S. Cross, Vitamin D receptor activity and pre- 1155.
vention of colonic hyperproliferation and oxidative stress, [85] K. Wu, W.C. Willett, C.S. Fuchs, G.A. Colditz, E.L. Gio-
Food Chem. Toxicol. 40 (2002) 1191–1196. vannucci, Calcium intake and risk of colon cancer in women
[70] P.R. Holt, N. Arber, B. Halmos, K. Forde, H. Kissileff, K.A. and men, J. Natl. Cancer Inst. 94 (2002) 437–446.
McGlynn, S.F. Moss, N. Kurihara, K. Fan, K. Yang, M. [86] M. Lipkin, Early develoment of cancer chemoprevention
Lipkin, Colonic epithelial cell proliferation decreases with clinical trials: studies of dietary calcium as a chemopreven-
increasing levels of serum 25-hydroxy vitamin D. Cancer tive agent for human subjects, Eur. J. Cancer Prev. 11 (Suppl.
Epidemiol, Biomark. Prev. 11 (2002) 113–119. 2) (2002) 65–70.
[71] H.G. Palmer, J.M. Gonzalez-Sancho, J. Espada, M.T. Ber- [87] S. Umar, A.P. Morris, F. Kourouma, J.H. Sellin, Dietary
ciano, I. Puig, J. Baulida, M. Quintanilla, A. Cano, A.G. de pectin and calcium inhibit colonic proliferation in vivo by
Herreros, M. Lafarga, A. Munoz, Vitamin D3 promotes the differing mechanisms, Cell Prolif. 36 (2003) 361–375.
differentiation of colon carcinoma cells by the induction of E- [88] S.A. Lamprecht, M. Lipkin, Cellular mechanisms of calcium
cadherin and the inhibition of b-catenin signaling, J. Cell and vitamin D in the inhibition of colorectal carcinogenesis,
Biol. 154 (2001) 369–387. Annal. N.Y. Acad. Sci. 952 (2001) 73–87.
[72] S.A. Lamprecht, M. Lipkin, Chemoprevention of colon can- [89] F.R. DeRubertis, P.A. Craven, R. Saito, Bile salt stimulation
cer by calcium, vitamin D and folate: molecular mechanisms, of colonic epithelial proliferation. Evidence for involvement
Nat. Rev. 3 (2003) 601–614. of lipoxygenase products, J. Clin. Invest. 74 (1984) 1614–
[73] R. Lin, J.H. White, The pleiotropic actions of vitamin D, 1624.
BioEssays 26 (2003) 21–28. [90] P.A. Craven, J. Pfanstiel, F.R. DeRubertis, Role of reactive
[74] M. Makishima, T.T. Lu, W. Xie, G.K. Whitfield, H. Domoto, oxygen in bile salt stimulation of colonic epithelial prolifera-
R.M. Evans, M.R. Haussler, D.J. Mangelsdorf, Vitamin D tion, J. Clin. Invest. 77 (1986) 850–859.
receptor as an intestinal bile acid sensor, Science 296 (2002) [91] P.A. Craven, J. Pfanstiel, F.R. DeRubertis, Role of activation
1313–1316. of protein kinase C in the stimulation of colonic epithelial
[75] G. Wolf, Intestinal bile acids can bind to and activate the proliferation and reactive oxygen formation by bile acids, J.
vitamin D receptor. Brief critical reviews, Nutr. Rev. 60 Clin. Invest. 79 (1987) 532–541.
(2002) 281–288. [92] L.A. Booth, I.T. Gilmore, R.F. Bilton, Secondary bile acid
[76] M.J. Wargovich, V.W.S. Eng, H.L. Newmark, Calcium inhi- induced DNA damage in HT29 cells: are free radicals
bits the damaging and compensatory proliferative effects of involved? Free Rad. Res. 26 (1997) 135–144.
fatty acids on mouse colon epithelium, Cancer Lett. 23 (1984) [93] M. Venturi, R.J. Hambly, B. Glinghammer, J.J. Rafter, I.R.
253–258. Rowland, Genotoxic activity in human faecal water and the
[77] A.W. Sorenson, M.L. Slattery, M.H. Ford, Calcium and colon role of bile acids: a study using the alkaline comet assay,
cancer: a review, Nutr. Cancer 11 (1988) 135–145. Carcinogenesis 18 (1997) 2353–2359.
H. Bernstein et al. / Mutation Research 589 (2005) 47–65 63

[94] S. Krahenbuhl, C. Talos, S. Fischer, J. Reichen, Toxicity of [106] D.C. Devor, M.C. Sekar, R.A. Frizzell, M.E. Duffey, Taur-
bile acids on the electron transport chain of isolated rat liver odeoxycholate activates potassium and chloride conduc-
mitochondria, Hepatology 19 (1994) 471–479. tances via an IP3-mediated release of calcium from
[95] R.J. Sokol, B.M. Winklhofer-Roob, M.W. Devereaux, J.M. intracellular stores in a colonic cell line (T84), J. Clin. Invest.
McKim Jr., Generation of hydroperoxides in isolated rat 92 (1993) 2173–2181.
hepatocytes and hepatic mitochondria exposed to hydropho- [107] H.H. Schmidt, J.S. Pollock, M. Nakane, U. Forstermann, F.
bic bile acids, Gastroentrology 109 (1995) 1249–1256. Murad, Ca2+/calmodulin-regulated nitric oxide synthases,
[96] D. Washo-Stultz, C.L. Crowley-Weber, K. Dvorakova, C. Cell Calcium 13 (1992) 427–434.
Bernstein, H. Bernstein, K. Kunke, C.N. Waltmire, H. Gar- [108] M. Olesen, R. Middelveld, J. Bohr, C. Tysk, J.O. Lundberg, S.
ewal, C.M. Payne, Role of mitochondrial complexes I and II, Eriksson, K. Alving, G. Jarnerot, Luminal nitric oxide and
reactive oxygen species and arachidonic acid metabolism in epithelial expression of inducible and endothelial nitric oxide
deoxycholate-induced apoptosis, Cancer Lett. 177 (2002) synthase in collagenous and lymphocytic colitis, Scand. J.
129–144. Gastroenterol. 38 (2003) 66–72.
[97] A.P. Rolo, P.J. Oliveira, A.J. Moreno, C.M. Palmeira, Che- [109] S.R. Thomas, K. Chen, J.F. Keaney Jr., Hydrogen peroxide
nodeoxycholate induction of mitochondrial permeability activates endothelial nitric-oxide synthase through coordi-
transition pore is associated with increased membrane fluidity nated phosphorylation and dephosphorylation via a phosphoi-
and cytochrome c release: protective role of carvedilol, nositide 3-kinase-dependent signaling pathway, J. Biol.
Mitochondrion 2 (2003) 305–311. Chem. 277 (2002) 6017–6024.
[98] C.L. Crowley-Weber, K. Dvorakova, C. Crowley, H. Bern- [110] D. Washo-Stultz, N. Hoglen, H. Bernstein, C. Bernstein, C.M.
stein, C. Bernstein, H. Garewal, C.M. Payne, Nicotine Payne, Role of nitric oxide and peroxynitrite in bile salt-
increases oxidative stress, activates NF-kB and GRP78, induced apoptosis: relevance to colon carcinogenesis, Nutr.
induces apoptosis and sensitizes cells to genotoxic/xenobiotic Cancer 35 (1999) 180–188.
stresses by a multiple stress inducer, deoxycholate: relevance [111] M. Jaiswal, N.F. LaRusso, L.J. Burgart, G.J. Gores, Inflam-
to colon carcinogenesis, Chem. Biol. Interact. 145 (2003) matory cytokines induce DNA damage and inhibit DNA repair
53–66. in cholangiocarcinoma cells by a nitric oxide-dependent
[99] T. Cocco, M. Di Paola, S. Papa, M. Lorusso, Arachidonic acid mechanism, Cancer Res. 60 (2000) 184–190.
interaction with the mitochondrial electron transport chain [112] M. Valko, H. Morris, M. Mazur, P. Rapta, R.F. Bilton, Oxygen
promotes reactive oxygen species generation, Free Rad. Biol. free radical generating mechanisms in the colon: do the semi-
Med. 27 (1999) 51–59. quinones of vitamin K play a role in the aetiology of colon
[100] H. Higuchi, H. Miyoshi, S.F. Bronk, H. Zhang, N. Dean, G.J. cancer? Biochim. Biophys. Acta. 1527 (2001) 161–166.
Gores, Bid antisense attenuates bile acid-induced apoptosis [113] R.J. Sokol, J.M. McKim, M.C. Goff, S.Z. Ruyle, M.W.
and cholestatic liver injury, J. Pharmacol. Exp. Therapeut. Devereaux, D. Han, L. Packer, G. Everson, Vitamin E
299 (2001) 866–873. Reduces oxidant injury to mitochondria and the hepatotoxi-
[101] W.A. Faubion, M.E. Guicciardi, H. Miyoshi, S.F. Bronk, P.J. city of taurochenodeoxycholic acid in the rat, Gastroenter-
Roberts, P.A. Svingen, S.H. Kaufmann, G.J. Gores, Toxic bile ology 114 (1998) 164–174.
salts induce rodent hepatocyte apoptosis via direct activation [114] H. Bernstein, C.M. Payne, C. Bernstein, J. Schneider, S.E.
of Fas, J. Clin. Invest. 103 (1999) 137–145. Beard, C.L. Crowley, Activation of the promoters of genes
[102] C.M. Payne, C. Crowley, D. Washo-Stultz, M. Briehl, H. associated with DNA damage, oxidative stress, ER stress and
Bernstein, C. Bernstein, S. Beard, H. Holubec, J. Warneke, protein malfolding by the bile salt, deoxycholate, Toxicol.
The stress-response proteins poly(ADP-ribose) polymerase Lett. 108 (1999) 37–46.
and NF-kB protect against bile salt-induced apoptosis, Cell [115] S. Lechner, U. Muller-Ladner, K. Schlottmann, B. Jung, M.
Death Differ. 5 (1998) 623–636. McClelland, J. Ruschoff, J. Welsh, J. Scholmerich, F. Kull-
[103] C.M. Payne, C.N. Waltmire, C. Crowley, C.L. Crowley- mann, Bile acids mimic oxidative stress induced upregulation
Weber, K. Dvorakova, H. Bernstein, C. Bernstein, H. Holu- of thioredoxin reductase in colon cancer cell lines, Carcino-
bec, H. Garewal, Caspase-6 mediated cleavage of guanylate genesis 23 (2002) 1281–1288.
cyclase alpha 1 during deoxycholate-induced apoptosis: pro- [116] A.L. Jackson, R. Chen, L.A. Loeb, Induction of microsatellite
tective role of the nitric oxide signaling module, Cell Biol. instability by oxidative DNA damage, Proc. Natl. Acad. Sci.
Toxicol. 19 (2003) 373–392. U.S.A. 95 (1998) 12468–12473.
[104] S.M. Chung, A.D. Proia, G.K. Klintworth, S.P. Watson, E.G. [117] Y. Hirose, C.V. Rao, B.S. Reddy, Modulation of inducible
Lapetina, Deoxycholate induces the preferential hydrolysis of nitric oxide synthase expression in rat intestinal cells by colon
phosphoinositides by human platelet and rat corneal phos- tumor promoters, Int. J. Oncol. 18 (2001) 141–146.
pholipase C, Biochem. Biophys. Res. Commun. 129 (1985) [118] C.M. Payne, C. Bernstein, H. Bernstein, E.W. Gerner, H.
411–416. Garewal, Reactive nitrogen species in colon carcinogenesis,
[105] H. Nakanishi, Y. Takeyama, H. Ohyanagi, Y. Saitoh, Y. Takai, Antioxi. Redox Signall. 1 (1999) 449–467.
Mode of stimulatory action of deoxycholate in signal trans- [119] M.S. Kulkarni, P.M. Heidepriem, K.L. Yielding, Production
duction system of isolated rat pancreatic acini, Biochem. by lithocholic acid of DNA strand breaks in L1210 cells,
Biophys. Res. Commun. 170 (1990) 111–118. Cancer Res. 40 (1980) 2666–2669.
64 H. Bernstein et al. / Mutation Research 589 (2005) 47–65

[120] M.S. Kulkarni, B.A. Cox, K.L. Yielding, Requirements for [135] H. Allgayer, M. Kolb, M.S. Stuber, W. Kruis, Modulation of
induction of DNA strand breaks by lithocholic acid, Cancer base hydroxylation by bile acids and salicylates in a model of
Res. 42 (1982) 2792–2795. human colonic mucosal DNA, Digest. Dis. Sci. 44 (1999)
[121] M.S. Kulkarni, K.L. Yielding, DNA damage and repair in 761–767.
epithelial (mucous) cells and crypt cells from isolated colon, [136] H. Allgayer, M. Kolb, V. Stuber, W. Kruis, Effects
Chem. Biol. Interact. 52 (1985) 311–318. of bile acids on base hydroxylation in a model of human
[122] R.L. Kandell, C. Bernstein, Bile salt/acid induction of DNA colonic mucosal DNA, Cancer Detect. Prevent. 26 (2002) 85–
damage in bacterial and mammalian cells: implications for 89.
colon cancer, Nutr. Cancer 16 (1991) 227–238. [137] A. Ogawa, T. Murate, M. Suzuki, Y. Nimura, S. Yoshida,
[123] Z.-Y. Zheng, C. Bernstein, Bile salt/acid induction of DNA Lithocholic acid, a putative tumor promoter, inhibits mam-
damage in bacterial cells: effect of taurine conjugation, Nutr. malian DNA polymerase b, Jpn. J. Cancer Res. 89 (1998)
Cancer 18 (1992) 157–164. 1154–1159.
[124] P.K. Zachariah, T.J. Slaga, D.L. Berry, W.M. Bracken, S.G. [138] D.F. Romagnolo, R.B. Chirnomas, J. Ku, B.D. Jeffy, C.M.
Buty, C.M. Martinsen, M.R. Juchau, The ability of enteric Payne, H. Holubec, L. Ramsey, H. Bernstein, C. Bernstein, K.
bacteria to catalyze the covalent binding of bile acids and Kunke, A. Bhattacharyya, J. Warneke, H. Garewal, Deox-
cholesterol to DNA and their inability to metabolize ben- ycholate, an endogenous tumor promoter and DNA damaging
zo(a)pyrene to a binding product and to known metabolites, agent, modulates BRCA-1 expression in apoptosis-sensitive
Cancer Lett. 3 (1977) 99–105. epithelial cells: loss of BRCA-1 expression in colonic ade-
[125] P.Y. Cheah, H. Bernstein, Modification of DNA by bile acids: nocarcinomas, Nutr. Cancer 46 (2003) 82–92.
a possible factor in the etiology of colon cancer, Cancer Lett. [139] D. Qiao, S.V. Gaitonde, W. Qi, J.D. Martinez, Deoxycholic
49 (1990) 207–210. acid suppresses p53 by stimulating proteasome-mediated p53
[126] K. Hamada, A. Umemoto, A. Kajikawa, M.J. Seraj, Y. protein degradation, Carcinogenesis 22 (2001) 957–964.
Monden, In vitro formation of DNA adducts with bile acids, [140] C. Bernstein, H. Bernstein, C.M. Payne, H. Garewal, DNA
Carcinogenesis 15 (1994) 1911–1915. repair/pro-apoptotic dual-role proteins in five major DNA
[127] Z.-Y. Zheng, H. Bernstein, C. Bernstein, C.M. Payne, J.D. repair pathways: fail-safe protection against carcinogenesis,
Martinez, E.W. Gerner, Bile acid activation of the gadd153 Mutat. Res. 511 (2002) 145–178.
promoter and of p53-independent apoptosis: relevance to [141] T.G. Rossman, E.I. Goncharova, Spontaneous mutagenesis in
colon cancer, Cell Death Differ. 3 (1996) 407–414. mammalian cells is caused mainly by oxidative events and
[128] H. Bernstein, C.M. Payne, C. Bernstein, J. Schneider, S. can be blocked by antioxidants and metallothionein, Mutat.
Beard, C. Crowley, Activation of the promoters of genes Res. 402 (1998) 103–110.
associated with DNA damage, oxidative stress and protein [142] M.S. Cooke, M.D. Evans, M. Dizdaroglu, J. Lunec, Oxidative
malfolding by the bile salt, deoxycholate, Toxicol. Lett. 108 DNA damage: mechanisms, mutation, and disease, FASEB J.
(1999) 37–46. 17 (2003) 1195–1214.
[129] K. Masamune, K. Kunitomo, K. Sasaki, K. Yagi, N. Komi, S. [143] C.L. Chang, G. Marra, D.P. Chauhan, H.T. Ha, D.K. Chang,
Tashiro, Bile-induced DNA strand breaks and biochemical L. Ricciardiello, A. Randolph, J.M. Carethers, C.R. Boland,
analysis of bile acids in an experimental model of anomalous Oxidative stress inactivates the human mismatch repair
arrangement of the pancreaticobiliary ducts, J. Med. Invest. system, Am. J. Physiol. Cell Physiol. 283 (2002) C148–C154.
44 (1997) 47–51. [144] H. Narahara, M. Tatsuta, H. Iishi, M. Baba, N. Uedo, N.
[130] A. Powolny, J. Xu, G. Loo, Deoxycholate induces DNA Sakai, H. Yano, S. Ishiguro, K-ras point mutation is asso-
damage and apoptosis in human colon epithelial cells expres- ciated with enhancement by deoxycholic acid of colon car-
sing either mutant or wild-type p53, Int. J. Biochem. Cell cinogenesis induced by azoxymethane, but not with its
Biol. 33 (2001) 193–203. attenuation by all-trans-retinoic acid, Int. J. Cancer 88
[131] B. Glinghammar, H. Inoue, J.J. Rafter, Deoxycholic acid causes (2000) 157–161.
DNA damage in colonic cells with subsequent induction of [145] M.I. Kelsey, R.J. Pienta, Transformation of hamster embryo
caspases. COX-2 promoter activity and the transcription factors cells by cholesterol and epoxide and lithocholic acid, Cancer
NF-kB and AP-1, Carcinogenesis 23 (2002) 839–845. Lett. 6 (1979) 143–149.
[132] B.L. Pool-Zobel, U. Leucht, Induction of DNA damage by [146] J. Theisen, J.H. Peters, H.J. Stein, Experimental evidence for
risk factors of colon cancer in human colon cells derived from mutagenic potential of duodenogastric juice on Barrett’s
biopsies, Mutat. Res. 375 (1997) 105–115. esophagus, World J. Surg. 27 (2003) 1018–1020.
[133] L.A. Booth, R.F. Bilton, Genotoxic potential of the secondary [147] S.J. Silverman, A.W. Andrews, Bile acids: comutagenic
bile acids: a role for reactive oxygen species, in: O.I. Arouma, activity in the Salmonella-mammalian-microsome mutageni-
B. Halliwell (Eds.), DNA and Free Radicals: Techniques, city test: brief communication, J. Natl. Cancer Inst. 59 (1977)
Mechanisms & Applications, OICA International, London, 1557–1559.
1998, pp. 161–177. [148] M. Wilpart, P. Mainguet, A. Maskens, M. Roberfroid, Struc-
[134] A. Burkle, Poly(ADP-ribosyl)ation, a DNA damage-driven ture-activity relationship amongst biliary acids showing co-
protein modification and regulator of genomic instability, mutagenic activity towards 1,2-dimethylhydrazine, Carcino-
Cancer Lett. 163 (2001) 1–5. genesis 4 (1983) 1239–1241.
H. Bernstein et al. / Mutation Research 589 (2005) 47–65 65

[149] S. Puju, D. Shuker, W.W. Bishop, K.R. Falchuk, S.R. Tan- [163] A.I. Haza, B. Glinghammar, A. Grandien, J. Rafter, Effect of
nenbaum, W.G. Thilly, Mutagenicity of N-nitroso bile acid colonic luminal components on induction of apoptosis in
conjugates in Salmonella typhimurium and diploid human human colonic cell lines, Nutr. Cancer 36 (2000) 79–89.
lymphoblasts, Cancer Res. 42 (1982) 2601–2604. [164] C. Bernstein, H. Bernstein, H. Garewal, P. Dinning, R. Jabi,
[150] M.H.L. Green, W.J. Muriel, B.A. Bridges, Use of a simplified R.E. Sampliner, M.K. McCuskey, M. Panda, D.J. Roe, L.
fluctuation test to detect low levels of mutagens, Mutat. Res. L’Heureux, C.M. Payne, A bile acid-induced apoptosis assay
38 (1976) 33–42. for colon cancer risk and associated quality control studies,
[151] J. Watabe, H. Bernstein, The mutagenicity of bile acids using Cancer Res. 59 (1999) 2353–2357.
a fluctuation test, Mutat. Res. 158 (1985) 45–51. [165] H. Bernstein, H. Holubec, J.A. Warneke, H. Garewal, D.L.
[152] Y. Mori, H. Nii, A. Masuda, K. Mizumoto, Y. Konishi, M. Earnest, C.M. Payne, D.J. Roe, H. Cui, E.L. Jacobson, C.
Une, T. Hoshita, Absence of mutagenic action of 5b-cholan- Bernstein, Patchy field defects of apoptosis resistance and
24-oic acid derivatives in bacterial fluctuation and standard dedifferentiation in flat mucosa of colon resections from
Ames tests, Mutat. Res. 262 (1991) 267–274. colon cancer patients, Ann. Surg. Oncol. 9 (2002) 505–517.
[153] U.G. Allinger, G.K. Johansson, J.A. Gustafsson, J.J. Rafter, [166] M.J. Redlak, M.S. Dennis, T.A. Miller, Apoptosis is a major
Shift from a mixed to a lactovegetarian diet: influence on mechanism of deoxycholate-induced gastric mucosal cell
acidic lipids in fecal water—a potential risk factor for colon death, Am. J. Physiol. Gastrointest. Liver Physiol. 285
cancer, Am. J. Clin. Nutr. 50 (1989) 992–996. (2003) G870–G879.
[154] C.M. Payne, C. Bernstein, H. Bernstein, Apoptosis overview [167] B.A. Magnuson, N. Shirtliff, R.P. Bird, Resistance of aberrant
emphasizing the role of oxidative stress. DNA damage and crypt foci to apoptosis induced by azoxymethane in rats
signal-transduction pathways, Leuk. Lymph. 19 (1995) 43–93. chronically fed cholic acid, Carcinogenesis 15 (1994)
[155] T. Patel, S.F. Bronk, G.J. Gores, Increases of intracellular 1459–1462.
magnesium promote glycodeoxycholate-induced apoptosis in [168] C.L. Crowley-Weber, C.M. Payne, M. Gleason-Guzman, G.S.
rat hepatocytes, J. Clin. Invest. 94 (1994) 2183–2192. Watts, B. Futscher, C.N. Waltmire, C. Crowley, K. Dvora-
[156] C.M. Payne, H. Bernstein, C. Bernstein, H. Garewal, kova, C. Bernstein, M. Craven, H. Garewal, H. Bernstein,
Role of apoptosis in biology and pathology: resistance to Development and molecular characterization of HCT-116
apoptosis in colon carcinogenesis, Ultrastruct. Pathol. 19 cell lines resistant to the tumor promoter and multiple
(1995) 224–248. stress-inducer, deoxycholate, Carcinogenesis 23 (2002)
[157] H. Samaha, E. Asher, C.M. Payne, C. Bernstein, H. Bernstein, 2063–2080.
Evaluation of cell death in EBV-transformed lymphocytes [169] C. Cherbonnel-Lasserre, M.K. Dosanjh, Suppression of apop-
using agarose gel electrophoresis, light microscopy and tosis by overexpression of Bcl-2 or Bcl-xL promotes survival
electron microscopy, Leuk. Lymph. 19 (1995) 95–105. and mutagenesis after oxidative damage, Biochimie 79
[158] H. Samaha, C. Bernstein, C.M. Payne, H.S. Garewal, R.E. (1997) 613–617.
Sampliner, H. Bernstein, Bile salt induction of apoptosis in [170] C. Cherbonnel-Lasserre, S. Guany, A. Kronenberg, Suppres-
goblet cells of the normal human colonic mucosa: relevance sion of apoptosis by Bcl-2 or Bcl-xL promotes susceptibility
to colon cancer, Acta Microsc. 4 (1995) 43–58. to mutagenesis, Oncogene 13 (1996) 1489–1497.
[159] A. Hague, D.J.E. Elder, D.J. Hicks, C. Paraskeva, [171] Y. Saintigny, A. Dumay, S. Lambert, B.S. Lopez, A novel role
Apoptosis in colorectal tumour cells: Induction by the for the Bcl-2 protein family: specific suppression of the
short chain fatty acids butyrate, propionate and acetate and RAD51 recombination pathway, EMBO J. 20 (2001)
by the bile salt deoxycholate, Int. J. Cancer 60 (1995) 400–406. 2596–2607.
[160] H. Garewal, H. Bernstein, C. Bernstein, R. Sampliner, C. [172] L.A.G. Ries, M.P. Eisner, C.L. Kosary, B.F. Hankey, B.A.
Payne, Reduced bile acid-induced apoptosis in ‘‘normal’’ Miller, L. Clegg, A. Mariotto, E.J. Feuer, B.K. Edwards
colorectal mucosa: a potential biological marker for cancer (Eds.). SEER Cancer Statistics Review, 1975-2001, National
risk, Cancer Res. 56 (1996) 1480–1483. Cancer Institute. Bethesda, MD, http://seer.cancer.gov/csr/
[161] M.C. Marchetti, G. Migliorati, R. Moraca, C. Riccardi, F.R. 1975_2001, 2004.
Nicoletti, V. Mastrandrea, G. Morozzi, Possible mechanisms [173] J. Cairns, Matters of Life and Death, Princeton University
involved in apoptosis of colon tumor cell lines induced by Press, Princeton, NJ, 1997.
deoxycholic acid, short-chain fatty acids, and their mixtures, [174] W.C. Willett, Diet and cancer: one view at the start of the
Nutr. Cancer 28 (1997) 74–80. millennium, Cancer Epidemiol. Biomarkers Prev. 10 (2001)
[162] J.D. Martinez, E.D. Stratagoules, J.M. LaRue, A.A. Powell, 3–8.
P.R. Gause, M.T. Craven, C.M. Payne, M.B. Powell, E.W. [175] S.H. Landis, T. Murray, S. Bolden, P.A. Wingo, Cancer
Gerner, D.L. Earnest, Different bile acids exhibit distinct statistics, CA Cancer J. Clin. 48 (1998) 6–29.
biological effects: The tumor promoter deoxycholic acid [176] G.E. Fraser, Associations between diet and cancer, ischemic
induces apoptosis and the chemopreventive agent ursodeoxy- heart disease, and all-cause mortality in non-Hispanic white
cholic acid inhibits cell proliferation, Nutr. Cancer. 31 (1998) California seventh-day adventists, Am. J. Clin. Nutr. 70
111–118. (1999) 532S–538S.

You might also like