[30]Dietary flavonoids effects on xenobiotic and carcinogen metabolism

background image

Mini-review

Dietary flavonoids: Effects on xenobiotic and carcinogen metabolism

Young Jin Moon, Xiaodong Wang, Marilyn E. Morris

*

Department of Pharmaceutical Sciences, School of Pharmacy and Pharmaceutical Sciences, University at Buffalo, State University of New York,

517 Hochstetter Hall, Amherst, NY 14260-1200, United States

Received 1 May 2004; accepted 1 June 2005

Available online 11 November 2005

Abstract

Flavonoids are present in fruits, vegetables and beverages derived from plants (tea, red wine), and in many dietary supplements or

herbal remedies including Ginkgo Biloba, Soy Isoflavones, and Milk Thistle. Flavonoids have been described as health-promoting, dis-
ease-preventing dietary supplements, and have activity as cancer preventive agents. Additionally, they are extremely safe and associated
with low toxicity, making them excellent candidates for chemopreventive agents. The cancer protective effects of flavonoids have been
attributed to a wide variety of mechanisms, including modulating enzyme activities resulting in the decreased carcinogenicity of xeno-
biotics. This review focuses on the flavonoid effects on cytochrome P450 (CYP) enzymes involved in the activation of procarcinogens
and phase II enzymes, largely responsible for the detoxification of carcinogens.

A number of naturally occurring flavonoids have been shown to modulate the CYP450 system, including the induction of specific

CYP isozymes, and the activation or inhibition of these enzymes. Some flavonoids alter CYPs through binding to the aryl hydrocarbon
receptor (AhR), a ligand-activated transcription factor, acting as either AhR agonists or antagonists. Inhibition of CYP enzymes, includ-
ing CYP 1A1, 1A2, 2E1 and 3A4 by competitive or mechanism-based mechanisms also occurs. Flavones (chrysin, baicalein, and galan-
gin), flavanones (naringenin) and isoflavones (genistein, biocha\nin A) inhibit the activity of aromatase (CYP19), thus decreasing
estrogen biosynthesis and producing antiestrogenic effects, important in breast and prostate cancers. Activation of phase II detoxifying
enzymes, such as UDP-glucuronyl transferase, glutathione S-transferase, and quinone reductase by flavonoids results in the detoxifica-
tion of carcinogens and represents one mechanism of their anticarcinogenic effects. A number of flavonoids including fisetin, galangin,
quercetin, kaempferol, and genistein represent potent non-competitive inhibitors of sulfotransferase 1A1 (or P-PST); this may represent
an important mechanism for the chemoprevention of sulfation-induced carcinogenesis.

Importantly, the effects of flavonoids on enzymes are generally dependent on the concentrations of flavonoids present, and the dif-

ferent flavonoids ingested. Due to the low oral bioavailability of many flavonoids, the concentrations achieved in vivo following dietary
administration tend to be low, and may not reflect the concentrations tested under in vitro conditions; however, this may not be true
following the ingestion of herbal preparations when much higher plasma concentrations may be obtained. Effects will also vary with
the tissue distribution of enzymes, and with the species used in testing since differences between species in enzyme activities also can
be substantial. Additionally, in humans, marked interindividual variability in drug-metabolizing enzymes occurs as a result of genetic
and environmental factors. This variability in xenobiotic metabolizing enzymes and the effect of flavonoid ingestion on enzyme expres-
sion and activity can contribute to the varying susceptibility different individuals have to diseases such as cancer. As well, flavonoids may
also interact with chemotherapeutic drugs used in cancer treatment through the induction or inhibition of their metabolism.
Ó 2005 Elsevier Ltd. All rights reserved.

Keywords: Dietary flavonoids; Xenobiotic metabolism; CYP; Phase II enzymes; Carcinogen

0887-2333/$ - see front matter

Ó 2005 Elsevier Ltd. All rights reserved.

doi:10.1016/j.tiv.2005.06.048

*

Corresponding author. Tel.: +1 716 645 2842x230; fax: +1 716 645 3693.
E-mail address:

memorris@buffalo.edu

(M.E. Morris).

www.elsevier.com/locate/toxinvit

Toxicology in Vitro 20 (2006) 187–210

background image

Contents

1.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188

2.

Effect of bioflavonoids on cytochrome P450 (CYP450) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188

2.1.

CYP1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

192

2.2.

CYP2E. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

197

2.3.

CYP3A4. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

197

2.4.

CYP19 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

199

3.

Effect of bioflavonoids on phase II enzymes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200

3.1.

UDP-glucuronyltransferase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

200

3.2.

Glutathione-S-transferase (GST) and quinone reductase (QR). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

202

3.3.

Sulfotransferases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

203

4.

Difficulties in the prediction of in vivo metabolic effects in humans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204

5.

Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205

1. Introduction

Flavonoids are part of a family of naturally occurring

polyphenolic compounds and represent one of the most
prevalent classes of compounds in vegetables, nuts, fruits
and beverages such as coffee, tea, and red wine (

Hollman

and Katan, 1997

) as well as medical herbs (e.g., Silybum

marianum, Alpina officinarum, Hypericum perforatum).
The average total intake of flavonoids in the United States
was estimated to be 1 g/day (

Kuhnau, 1976; Scalbert and

Williamson, 2000

), but recent studies have indicated that

the intake varies widely (

Beecher, 2003

). More than 8000

compounds with a flavonoid structure have been identified.
The large number of compounds arises from the various
combinations of multiple hydroxyl and methoxyl group
substituents on the basic flavonoid skeleton (

Hodek

et al., 2002

). The classes of flavonoids include chalcones,

flavones, flavonols, flavanones, flavanols, anthocyanins
and isoflavones (

Table 1

). The flavonoid natural products

exert a wide range of biochemical and pharmacological
properties, with one of the most investigated effects being
their cancer preventive activities. The cancer protective
effects of flavonoids have been attributed to a wide variety
of mechanisms, including free radical scavenging, modify-
ing enzymes that activate or detoxify carcinogens, and
inhibiting the induction of the transcription factor activa-
tor protein-1 (AP-1) activity by tumor promoters (

Caniv-

enc-Lavier et al., 1996; Shih et al., 2000

). Flavonoids also

have inhibitory effect on the activities of many enzymes,
including b-glucuronidase (

Kim et al., 1994

), lipoxygenase

(

Laughton et al., 1991; Schewe et al., 2002

), cyclooxygen-

ase (

Laughton et al., 1991

), inducible nitric oxide synthase

(

Raso et al., 2001

), monooxygenase (

Siess et al., 1995

), thy-

roid peroxidase (

Doerge and Chang, 2002

), xanthine oxi-

dase (

Sheu et al., 1998

), mitochondrial succinoxidase and

NADH-oxidase (

Hodnick et al., 1994

), phosphodiesterase

(

Picq et al., 1989

), phospholipase A

2

(

Gil et al., 1994

)

,

and

protein kinase (

Cushman et al., 1991

).

The focus of this paper will be the effects of flavonoids

on cytochrome P450 (CYP) and phase II enzymes which
are key enzymes involved in the metabolism of xenobiotics.

Table 2

presents a summary of the flavonoid-induced alter-

ations in enzyme expression and/or activity described in
this review.

2. Effect of bioflavonoids on cytochrome P450 (CYP450)

Cytochrome P450 enzymes (phase I monooxygenase

enzymes) are widely known for their role in the metabolism
of drugs and other foreign compounds. Thus, modulation
of this enzyme system can influence the metabolism of
xenobiotics, producing effects of pharmacological and tox-
icological importance. A number of naturally occurring
flavonoids have been shown to modulate the CYP450 sys-
tem, including the induction of specific CYP isozymes, and
the activation or inhibition of these enzymes (

Wood et al.,

1986

).

Wattenberg et al. (1968)

began investigating the

effects of flavonoids on the CYP system over 36 years
ago, and a number of subsequent studies have shown differ-
ent modulatory effects of flavonoids on CYP activity
in vitro and in vivo. Many carcinogens are metabolized
by CYP enzymes to either biologically inactive metabolites
or to chemically reactive electrophilic metabolites that
covalently

bind

to

DNA

producing

carcinogenicity

(

Fig. 1

) (

Conney, 2003

). The reactive metabolites may

undergo additional metabolism by phase I or II enzymes
to inactive products. Therefore, induction of either phase
I or phase II enzymes can result in increased detoxification
of carcinogens (

Conney, 2003

). Since many chemical car-

cinogens are metabolized by CYP enzymes to both inac-
tive, as well as to carcinogenic metabolites, the effects of
inducers of these enzymes on the carcinogenicity of a chem-
ical will depend on the inducerÕs effects on the different met-
abolic pathways. In animal studies, it has been reported
that inducers of CYPs usually decrease the carcinogenicity
of chemical carcinogens in vivo; this suggests that induc-

188

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

background image

Table 1
The chemical structures and major food sources of the 6 major flavonoid subgroups

Structure

Example

Major food sources

Basic structure of flavonoids

Chalcone

Hop chalcones (xanthohumol and
dehydrocyclo-xanthohumol hydrate)

Hops, beer

Flavone

Acacetin

Parsley, thyme, celery, sweet red peppers,
honey, propolis

Apigenin
Baicalein
Chrysin
Diosmetin
Luteolin
Tangeretin

Flavonol

Galangin

Onions, kale, broccoli, apples, cherries,
berries, tea, red wine

Kaempferol
Morin
Myricetin
Quercetin

Flavanone

Eriodictyol

Citrus

Hesperetin
Homoeriodictyol
Naringenin

Flavanol

Catechin

Cocoa, green tea, chocolate, red wine,
hawthorn, bilberry, motherwort,
and other herbs

Epicatechin
Proanthocyanidins

Anthocyanin

Cyanidins

Cherries, grapes, berries, red cabbage

Pigmented
compounds

Isoflavone

Biochanin A

Red clover, alfalfa, peas, soy
and other legumes

Genistein
Diadzein

Equol
Formononetin

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

189

background image

Table 2
Summary of flavonoid-induced enzyme induction or inhibition (

"induction, #inhibition, * in vivo study)

Phase I

Flavonoids

Effects

References

CYP1A1

Quercetin

"mRNA expression and activity

Ciolino et al. (1999)

Kaempferol

#TCDD or B[a]P-induced transcription

Kang et al. (1999)

Galangin

#Activity

Ciolino and Yeh (1999)

"mRNA expression

Ciolino and Yeh (1999)

#TCDD or DMBA-induced transcription

Ciolino and Yeh (1999)

Green tea extracts

"Activity

*

Liu et al. (2003)

Theaflavins

#Activity in the intestine

*

Catterall et al. (2003)

Xanthohumol

#Activity

Henderson et al. (2000)

Baicalein

#DMBA-induced transcription

Chan et al. (2002)

#Activity

Chan et al. (2002)

Dimethoxyflavone

#Activity

Wen et al. (2005)

#B[a]P-induced mRNA and protein expression, and Activity

Wen et al. (2005)

Chrysin

"Activity

Wen et al. (2005)

Diosmin

"mRNA expression

Ciolino et al. (1998)

Diosmetin

"mRNA expression

Ciolino et al. (1998)

#DMBA-induced transcription

Ciolino et al. (1998)

#Activity

Doostdar et al. (2000)

Acacetin

#Activity

Doostdar et al. (2000)

Flavone

#Activity

Zhai et al. (1998)

Hesperetin

#Activity

Doostdar et al. (2000)

Biochanin A

#DMBA-induced transcription and activity

Chan and Leung (2003)

Genistein

#DMBA-induced transcription and activity

Chan and Leung (2003)

2

0

-Hydroxychalcone

#mRNA expression and activity

Wang et al. (2005)

CYP1A2

Quercetin

#Activity

Tsyrlov et al. (1994)

Biapigenin

#Activity

Obach (2000)

Galangin

#Activity

Zhai et al. (1998)

Green tea extracts

#TPA-induced transcription

Shih et al. (2000)

"Activity

*

Liu et al. (2003)

8-Prenylnaringenin

#Activity

Henderson et al. (2000)

Isoxanthohumol

#Activity

Henderson et al. (2000)

Diosmetin

#Activity

Doostdar et al. (2000)

Acacetin

#Activity

Doostdar et al. (2000)

Flavone

#Activity

Zhai et al. (1998)

Tangeretin

#Activity

Obermeier et al. (1995)

Naringin

#Activity

*

Fuhr et al. (1993)

Genistein

#Activity

*

Helsby et al. (1998)

#Activity

*

Chen et al. (2004)

Equol

#Activity

*

Helsby et al. (1998)

Diadzein

#Activity

*

Peng et al. (2003)

CYP1B1

Xanthohumol

#Activity

Henderson et al. (2000)

Baicalein

#DMBA-induced transcription

Chan et al. (2002)

Diosmetin

#Activity

Doostdar et al. (2000)

Acacetin

#Activity

Doostdar et al. (2000)

Hesperetin

#Activity

Doostdar et al. (2000)

Biochanin A

#DMBA-induced transcription and activity

Chan and Leung (2003)

Genistein

#DMBA-induced transcription and activity

Chan and Leung (2003)

Diadzein

#Activity

Roberts et al. (2004)

Formononetin

#Activity

Roberts et al. (2004)

2

0

-hydroxychalcone

#mRNA expression and activity

Wang et al. (2005)

CYP2E1

Genistein

#Activity

*

Helsby et al. (1998)

Equol

#Activity

*

Helsby et al. (1998)

Theaflavins

#Protein level

*

Catterall et al. (2003)

Silybin (=Silibinin, Silybinin)

#Activity

Zuber et al. (2002)

CYP3A4

Naringenin

#Activity

*

Fuhr (1998)

Hyperforin

#Activity

Obach (2000)

"mRNA expression

Zhou et al. (2003)

"Activity

*

Bray et al. (2002)

Biapigenin

#Activity

Obach (2000)

Quercetin

#Activity

Obach (2000)

Silymarin

#Activity

Venkataramanan et al. (2000)

Silybin

#Activity (at high conc.)

Beckmann-Knopp et al. (2000)

190

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

background image

Table 2 (continued)

Phase I

Flavonoids

Effects

References

"Activity (at low conc.)

Genistein

#Activity

Foster et al. (2003)

Diadzein

#Activity

Foster et al. (2003)

Glabridin

#Activity

Zhou et al. (2004)

Myricetin

#Activity

Ho et al. (2001)

CYP19

Biochanin A

#Activity

Kao et al. (1998)

#Activity

Le Bail et al. (2000)

#Activity

Almstrup et al. (2002)

Formononetin

#Activity

Almstrup et al. (2002)

Equol

#Activity

Pelissero et al. (1996)

Chrysin

#Activity

Kao et al. (1998)

#Activity

Le Bail et al. (2000)

#Activity

Almstrup et al. (2002)

#Activity

*

Gambelunghe et al. (2003)

Naringenin

#Activity

Kao et al. (1998)

#Activity

Le Bail et al. (2000)

#Activity

Almstrup et al. (2002)

Flavone

#Activity

Le Bail et al. (2000)

Apigenin

#Activity

Le Bail et al. (2000)

Bicalein

#Activity

Kao et al. (1998)

Galangin

#Activity

Kao et al. (1998)

Phase II
UGT1A1

Chrysin

"mRNA expression

Walle et al. (2000)

"activity

Galijatovic et al. (2001)

Galangin

"mRNA expression

Walle and Walle (2002)

Isorhamnetin

"mRNA expression

Walle and Walle (2002)

Tangeretin

#Activity

Williams et al. (2002)

Naringenin

#Activity

Williams et al. (2002)

Flavone

#Activity

Williams et al. (2002)

Quercetin

#Activity

Williams et al. (2002)

UGT

Flavone

"Activity

*

Brouard et al. (1988)

"Activity

*

Canivenc-Lavier et al. (1996)

"Activity

*

van der Logt et al. (2003)

Quercetin

"Activity

*

van der Logt et al. (2003)

"Activity

Sun et al. (1998)

Green tea

"Activity

*

Bu-Abbas et al. (1995)

"Activity

*

Sohn et al. (1994)

Biochanin A

"Activity

Sun et al. (1998)

Formononetin

"Activity

Sun et al. (1998)

Genistein

"Activity

Sun et al. (1998)

Diadzein

"Activity

Sun et al. (1998)

Naringenin

"Activity

Sun et al. (1998)

Galangin

"Activity

Sun et al. (1998)

Kaempferol

"Activity

Sun et al. (1998)

GST

Genistein

"or #mRNA expression

Ansell et al. (2004)

(cell type specific)
"Activity

*

Appelt and Reicks (1999)

Diadzein

"Activity

*

Appelt and Reicks (1999)

Genistein + Diadzein

Prevent TPA-downregulated Activity

*

Sharma and Sultana (2004)

Flavone

"Activity

*

Nijhoff et al. (1995)

Green tea extract

"Activity

*

Maliakal et al. (2001)

Morin

"Activity

*

Tanaka et al. (1999)

"Activity

*

Kawabata et al. (1999)

Silymarin

"Activity

*

Kohno et al. (2002)

GSTA2

Flavone

"mRNA expression

Kang et al. (2003)

2

0

-amino-3

0

methoxyflavone

"mRNA expression

Kang et al. (2003)

GSTP1-1

Quercetin

#Activity

van Zanden et al. (2003)

QR

Tephropurpurin

"Activity

Chang et al. (1997)

Xanthohumol

"Activity

Miranda et al. (2000)

(continued on next page)

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

191

background image

tion of Phase I and II detoxification pathways may occur to
a greater extent than induction of CYPs involved in the
formation of carcinogenic metabolites (

Fig. 1

) (

Conney,

2003

).

2.1. CYP1

The CYP1 family consists of 1A1, 1A2, and 1B1 mem-

bers that are capable of activating procarcinogens.
CYP1A1 and 1B1 are both involved in the biotransforma-
tion of polycyclic aromatic hydrocarbons (PAHs), a class
of ubiquitous environmental chemicals, to carcinogenic
metabolites (

Guengerich and Shimada, 1991

). This process

is believed to contribute to pulmonary carcinogenesis,
because increased lung CYP1A1 expression and activity
are associated with a high risk of lung cancer (

McLemore

et al., 1990

). High CYP1A1 activity is also associated with

colorectal cancer (

Sivaraman et al., 1994

). CYP1A2 mainly

metabolizes important drugs such as phenacetin, theophyl-
line, caffeine, imipramine, and propranolol (

Brosen, 1995

),

and also activates some procarcinogens to carcinogens.

CYP1A2 plays a role in human tobacco-related cancers
(

Smith et al., 1996

). CYP1A1 is poorly expressed in human

liver although its synthesis can be markedly induced in
many extrahepatic tissues, notably the lungs (

Rendic and

Di Carlo, 1997

). In contrast, CYP1A2 is expressed princi-

pally in the liver (

Rendic and Di Carlo, 1997

). CYP1B1

is an extrahepatic estradiol 4-hydroxylase that activates
procarcinogens and elevated levels have been associated
with estrogen carcinogenesis (

Jefcoate et al., 2000

). Normal

human breast and human breast tumor tissues are known
to express CYP1B1, producing carcinogenic 4-hydroxy
estrogen. Inhibition of CYP1B1 affects the production of
mutagenic estrogen 3,4-catechols (

Roberts et al., 2004

).

Xenobiotic responsive elements (XRE) are cis-acting

enhancer elements located in the promoter regions of xeno-
biotic responsive genes, which include genes encoding for
CYP1A1 and 1B1. The expression of these xenobiotic
responsive genes can be regulated through pathway involv-
ing aryl hydrocarbon receptor (AhR), which is a cytosolic
protein that can be activated by PAH. The activated
AhR then translocates to the nucleus, dimerizes with

Table 2 (continued)

Phase II

Flavonoids

Effects

References

Morin

"Activity

*

Tanaka et al. (1999)

"Activity

*

Kawabata et al. (1999)

Pinostrobin

"Activity

Fahey and Stephenson (2002)

Kaempferol

"Activity

Uda et al. (1997)

Galangin

"Activity

Uda et al. (1997)

Quercetin

"Activity

Uda et al. (1997)

Silymarin

"Activity

*

Kohno et al. (2002)

Genistein

"Activity

Yannai et al. (1998)

"Activity and mRNA expression

Wang et al. (1998)

"Activity

*

Appelt and Reicks (1999)

Diadzein

"Activity

Yannai et al. (1998)

"Activity

*

Appelt and Reicks (1999)

Flavone

"Activity

*

Siess et al., 1996

SULT1A1

Fisetin

#Activity

Eaton et al. (1996)

(P-PST or TS-PST)

Galangin

#Activity

Eaton et al. (1996)

Quercetin

#Activity

Eaton et al. (1996)

#Activity

Ghazali and Waring (1999)

#Activity

Harris et al. (2004)

Catechin

#Activity

Ghazali and Waring (1999)

Equol

#Activity

Ghazali and Waring (1999)

Flavone

#Activity

Ghazali and Waring (1999)

Myricetin

#Activity

Eaton et al. (1996)

Kaempferol

#Activity

Eaton et al. (1996)

Chrysin

#Activity

Eaton et al. (1996)

Apigenin

#Activity

Eaton et al. (1996)

Genistein

#Activity

Eaton et al. (1996)

#Activity

Ghazali and Waring (1999)

#Activity

Harris et al. (2004)

Diadzein

#Activity

Harris et al. (2004)

#Activity

Ghazali and Waring (1999)

SULT1A3

Baicalein

#Activity

Harris et al. (2004)

SULT1E1

Equol

#Activity

Harris et al. (2004)

Quercetin

#Activity

Otake et al. (2000)

#Activity

Ohkimoto et al. (2004)

Genistein

#Activity

Ohkimoto et al. (2004)

Diadzein

#Activity

Ohkimoto et al. (2004)

#Activity

Wong and Keung (1997)

192

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

background image

AhR nuclear translocator (ARNT), and interacts with
XRE (

Kronenberg et al., 2000

). Induction of CYP1A by

flavonoids proceeds by various mechanisms, including the
direct stimulation of gene expression via specific receptor(s)
and/or CYP protein, or mRNA stabilization (

Lin and Lu,

1998; Shih et al., 2000

). Some flavonoids induce CYPs

through binding to AhR, a ligand-activated transcription
factor, by acting in a similar way as 2,3,7,8-tetrachlo-
rodibenzo-p-dioxin (TCDD) (

Kohn et al., 2001

). Gener-

ally, substrates for AhR are planar aromatic compounds
with few bulky substituent groups. That might partly
explain the activity of flavonoids, which have similar pla-
nar structures as AhR (

Hodek et al., 2002; Kohn et al.,

2001; MacDonald et al., 2001

). Other flavonoids have been

shown to directly inhibit CYP1A1 activty (

Doostdar et al.,

2000; Tsyrlov et al., 1994; Wen et al., 2005

), commonly

demonstrated to be a competitive-type of inhibition, and
to affect CYP1A1 transcription (

Ciolino et al., 1999; Kang

et al., 1999

).

The most abundant flavonoids, flavonols quercetin and

kaempferol, are both dietary ligands of the AhR, but they
exert different effects on CYP1A1 expression (

Ciolino

et al., 1999

). Treatment of MCF-7 cells with quercetin

resulted in a concentration- (

Fig. 2

A) and time (

Fig. 2

B)

dependent increase in the amount of CYP1A1 mRNA.
The activity of enzyme CYP1A1 was measured by a 7-eth-
oxyresorufin-O-dealkylation (EROD) assay. The increase
in EROD activity followed the increase in CYP1A1 mRNA.
Quercetin also causes a concentration- (

Fig. 2

C) and time-

(

Fig. 2

D) dependent increase in CYP1A1 enzyme activity

in MCF-7 cells (

Ciolino et al., 1999

). Kaempferol, by itself,

does not affect CYP1A1 expression (

Fig. 2

A), but it can

interact with the AhR, and act as an antagonist of TCDD-
induced CYP1A1 transcription (

Ciolino et al., 1999

).

Despite the structural similarity between quercetin and
kaempferol, their differential effects might be due to the
absence of an additional hydroxy group on the B-ring of
kaempferol (

Fig. 3

), preventing it from achieving an optimal

fit into the binding site on AhR to produce transcriptional
activation. The binding of kaempferol may block the bind-
ing of AhR ligands, and thus inhibit the activity of other
ligands such as TCDD (

Ciolino et al., 1999

). The importance

of the ortho-orientation of the hydroxyl group on the B ring
was tested by comparing the effects of quercetin with a 3

0

,4

0

-

substitution pattern and morin (2

0

,4

0

-substitution pattern)

(

Tsyrlov et al., 1994

). 7-Methoxyresorufin-O-dealkylation

(MROD) activities of mouse c-DNA-expressed CYP1A1
and 1A2, and human CYP1A2 were more effectively inhib-
ited with quercetin than with morin (

Tsyrlov et al., 1994

).

Orientation of the methyl group can also influence inhibition
(

Tsyrlov et al., 1994

). Quercetin, a component in St. JohnÕs

wort inhibits CYP1A2 (

Obach, 2000

). St. JohnÕs wort

(Hypericum perforatum) extracts are commercially available
preparations used in the treatment of depression. Quercetin
inhibits phenacetin O-deethylase activity (mediated by
CYP1A2) in a recombinant CYP1A2 enzyme preparation
with an IC

50

value of 7.5 lM. Another flavonoid in St.

JohnÕs wort, I3,II8-biapigenin was also shown to be a

Fig. 1. Flavonoids that block or suppress multistage carcinogenesis. Carcinogenesis is initiated with the transformation of the normal cell into a mutant
cell. These cells undergo tumor promotion into benign tumor cells, which progress to malignant cells. Flavonoids can interfere with different steps of this
process. Some flavonoids (for example, kaempferol, diosmetin, theaflavin, and biochanin A) can inhibit the metabolic activation of the procarcinogens to
their ultimate electrophilic species by phase I enzymes (predominantly CYPs), or their subsequent interaction with DNA. Therefore these agents block
tumor initiation (blocking agents). Alternatively, dietary flavonoids (for example, naringenin, quercetin, biochanin A, and prenylchalcones) can stimulate
the detoxification of carcinogens by inducing phase II enzymes, leading to their elimination from the body. Flavonoids such as genistein and EGCG
suppress the later steps (promotion and progression) of multistage carcinogenesis (suppressing agents) by affecting cell cycle, angiogenesis, invasion, and
apoptosis. Adapted from (

Chen and Kong, 2004

).

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

193

background image

potent, competitive inhibitor of CYP1A2 activity (IC

50

=

3.7 lM) (

Obach, 2000

).

Galangin, a flavonol found in honey, is a potent inhibi-

tor of CYP1A1 activity, as measured by inhibition of
EROD activity, in intact cells and in microsomes isolated
from dimethylbenz[a]anthracene (DMBA) treated cells
(

Ciolino and Yeh, 1999

). The effect is dose-dependent. Gal-

angin is a non-competitive inhibitor of CYP1A1 activity
(

Ciolino and Yeh, 1999

). Galangin increased the level of

CYP1A1 mRNA, indicating that it may be an agonist of
the AhR, but it inhibited the induction of CYP1A1 mRNA
by DMBA or by 2,3,5,7- TCDD (

Ciolino and Yeh, 1999

).

Galangin also inhibited the DMBA- or TCDD-induced
transcription of a reporter vector containing the CYP1A1
promoter. Thus, galangin is a potent inhibitor of DMBA
metabolism and an agonist/antagonist of the AhR (

Ciolino

and Yeh, 1999

). Similarly, quercetin causes an increase in

the level of CYP1A1 mRNA (

Ciolino et al., 1999

), whereas

it significantly inhibits benzo(a)pyrene (B[a]P)-induced
CYP1A1 mRNA and protein expression in human Hep
G2 cells (

Kang et al., 1999

). Galangin is a very potent

CYP1A2 inhibitor, too (

Zhai et al., 1998

). It showed the

mixed-type inhibition, indicating that this compound can
compete for substrate binding at the active site and also
may bind to a region that does not participate directly in
substrate binding (

Zhai et al., 1998

).

Activator protein-1 (AP-1) may be involved in the regu-

lation of human CYP1A2 which contains two AP-1 bind-
ing sites (

Shih et al., 2000

).

Shih et al. (2000)

developed a

cell line T2Luc, which is a HepG2-derived cell line stably
integrated with a region of the human CYP1A2 5

0

-flanking

gene containing two AP-1 binding sites linked to the thymi-

Fig. 2. Effect of quercetin or kaempferol on expression of mRNA CYP1A1 and CYP1A1 enzyme activity. (A) Concentration response of CYP1A1
mRNA to quercetin (hatched bars) and kaempferol (open bars). MCF-7 cells were treated with quercetin the indicated concentration of quercetin. The
amount of CYP1A1 was normalized to the GPDH level. nd, not determined. (B) Time course of CYP1A1 mRNA increase caused by quercetin. MCF-7
cells were treated with 0.5 lM quercetin for the times indicated. The amount of CYP1A1 mRNA was normalized to GPDH levels. (C) The activity of
CYP1A1 in intact MCF-7 cells was determined by EROD assay. Cells were treated with the indicated concentrations of quercetin (j) or kaempferol ()
for 48 h. (D) Cells were treated with 5 lM quercetin for the times indicated. Reproduced from (

Ciolino et al., 1999

) with permission from the Biochemical

Society.

194

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

background image

dine kinase promoter-driven firefly luciferase reporter gene.
Green tea extracts (GTEs) inhibited the phorbol 12-o-
tetradecanoate 13-acetate (TPA)-induced AP-1 transcrip-
tional activation of a human CYP1A2 enhancer element,
while quercetin enhanced this activity. Green tea (Camellia
sinensis) and its extracts are rich source of catechins, a class
of flavonoids (

Shih et al., 2000

). Catechins are the main

compounds in green tea; they consist of (

)-epicatechin,

(

)-epicatechin-3-gallate

(ECG),

(

)-epigallocatechin,

and

(

)-epigallocatechin-3-gallate

(EGCG)

(

Graham,

1992

). The mechanism of inhibition may be due to changes

in the composition of the AP-1 complex upon treatment of
cells with GTEs. An important and unexpected finding is
that GTEs themselves increased binding of nuclear pro-
teins to the AP-1 site to the same extent as TPA, yet did
not activate transcription of the luciferase reporter gene.
Unlike TPA, which induces AP-1 binding as early as 6 h
after its addition, GTE-induced AP-1 binding did not
occur until after 6 h, suggesting that other upstream signal-
ing pathways may also be targeted by this dietary agent
(data not shown) (

Shih et al., 2000

). In some in vitro exper-

iments, tea polyphenols had an inhibitory effect on micro-
somal CYP enzyme system (

Mukhtar et al., 1992; Wang

et al., 1988

). In contrast, long-term consumption of green

tea increases CYP1A1 and 1A2 activities in rats (

Liu

et al., 2003

). The contents of CYPs (measured by the

method of Omura and Sato (

Heffernan and Winston,

2000

)) in the livers of male rats in high dose (tea polyphe-

nols 833 mg/kg/day, intragastrically, for 6 months) groups
were significantly increased (

Liu et al., 2003

).

Theaflavins in black tea are catechins generated by oxi-

dation of flavanols during the fermentation of fresh tea leaf
(

Graham, 1992

). It has been shown to have antiprolifera-

tive and anticarcinogenic activities (

Yang et al., 1997

).

Male rats received intragastric administration of theaflav-
ins (20 mg/kg) for four weeks, and the liver and intestine
mucosa samples were obtained. Theaflavin treatment
markedly suppressed the CYP1A1 EROD activity in the
intestine (

Catterall et al., 2003

). However, there was no

effect on any CYP activity in the liver (

Catterall et al.,

2003

); these results contrast with those from another study

where theaflavins suppressed CYP1A1 activity in hepatic
cell cultures (

Feng et al., 2002

). This discrepancy is pro-

bably due to the poor bioavailability of theaflavins as a
result of poor absorption and/or extensive pre-systemic
metabolism.

Henderson et al. (2000)

reported the in vitro inhibition

of cDNA-expressed human CYP1A1, CYP1B1, and
CYP1A2 by flavonoids from hops (Humulus lupulus) (

Hen-

derson et al., 2000

). At 10 lM, the prenylated chalcone,

xanthohumol, almost completely inhibited the EROD
activity of CYP1A1. At the same concentration, other
hop flavonoids decreased the EROD activity by 27.0–
90.8%. At 10 lM, xanthohumol completely eliminated
CYP1B1 EROD activity, whereas the other hop flavonoids
showed varying degrees of inhibitory action ranging from
1.8% to 99.3% (

Henderson et al., 2000

). The most effective

inhibitors of CYP1A2 acetanilide 4-hydroxylase activity
were the two prenylated flavonoids, 8-prenylnaringenin
and isoxanthohumol, which produced >90% inhibition
when added at concentrations of 10 lM. CYP1A2 meta-
bolism of the carcinogen aflatoxin B1 was also inhibited
by isoxanthohumol and 8-prenylnaringenin as shown by
decreased appearance of dihydrodiols and aflatoxin M1
as analysed by HPLC (

Henderson et al., 2000

).

Baicalein, a flavone extracted from the root of the Scu-

tellaria species, is a strong competitive inhibitor of EROD
activity induced by DMBA in MCF-7 cells (

Chan et al.,

2002

). Baicalein can reduce the CYP1A1/1B1 mRNA

expression induced by DMBA, and the effect on mRNA
abundance of CYP1A1 was greater than that of CYP1B1.
An XRE-luciferase gene reporter assay also indicated that
AhR transactivation was suppressed (

Chan et al., 2002

).

5,7-Dimethoxyflavone (DMF) is a major constituent of

the leaves of a Malaysian Piper species. DMF reduced
CYP1A1 EROD activity of the HepG2 cells virtually down
to zero (

Wen et al., 2005

). It also inhibited benzo(a)pyrene

(B[a]P)-induced DNA binding (

Wen et al., 2005

). B[a]P, a

major PAH procarcinogen, induces CYP1A1 protein and
its catalytic activity as well as CYP1A1 mRNA; DMF
clearly inhibits this transcriptional activation by decreasing
CYP1A1 catalytic activity as well as protein and mRNA
expression. DMF can also directly inhibit CYP1A1 protein
catalytic activity, as determined with recombinant protein
(

Wen et al., 2005

). It is remarkable that two so seemingly

Fig. 3. Structures of quercetin, kaempferol and 2,3,7,8-tetrachlo-
rodibenzo-p-dioxin (TCDD).

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

195

background image

similar compounds as DMF and chrysin can have such dif-
ferent effects, i.e. the 5,7-dimethoxy compound is a potent
inhibitor of the CYP1A1 protein whereas the 5,7-dihy-
droxy compound is a potent inducer (

Wen et al., 2005

).

Thus, methylation of flavonoids seems to be an important
feature determining enzyme inhibitory properties.

The flavone diosmetin is ingested (as its glucoside, dios-

min) in the commercially available drug ÔDaflon 500 mgÕ
(90% diosmin and 10% hesperetin). Both diosmin and dios-
metin are agonists of the AhR, causing a dose-dependent
increase in expression of CYP1A1 mRNA in MCF-7
human breast epithelial cancer cells (

Ciolino et al., 1998

);

however, diosmetin, but not diosmin, inhibits CYP1A1
activity in a non-competitive manner in microsomes iso-
lated from 7,12-dimethylbenz-[a]anthracene (DMBA) trea-
ted cells, as assayed by EROD activity (

Ciolino et al.,

1998

).

Doostdar and co-workers (

Doostdar et al., 2000

) dem-

onstrated that six common flavonoids present in citrus
juices at 70–200 mg/L are potent inhibitors of CYP1A1,
1A2 and 1B1 EROD activity in vitro (

Doostdar et al.,

2000

). The flavones acacetin and diosmetin were more

potent inhibitors than the flavanones eriodictyol, hespere-
tin, homoeriodictyol and naringenin. This is probably
due to the presence of the reduced 2,3-bond in the C ring
of the structure in the flavones (

Table 1

) (

Dai et al.,

1998

). Potent inhibition of CYP1A1 and 1A2 enzyme activ-

ity by flavone has also been reported in other studies (

Lee

et al., 1994; Zhai et al., 1998

). Flavone inhibits CYP1A1

and CYP1A2, with a 2-fold greater potency toward the lat-
ter (

Zhai et al., 1998

).

Tangeretin, a polymethoxylated flavone present in large

amounts in citrus fruits, is a potent inhibitor of CYP1A2
EROD activity, with IC

50

of 0.8 lM in rat liver micro-

somes and 16 lM in human microsomes (

Obermeier

et al., 1995

). The inhibition was competitive with a inhibi-

tion constant (K

i

) value of 68 nM (

Obermeier et al., 1995

).

The flavanone hesperetin is a selective substrate of

human CYP1A1 and CYP1B1 in the lymphoblastoid cell
line AHH-1, and it is a competitive inhibitor of CYP1B1
(

Fig. 4

) (

Doostdar et al., 2000

). Hesperidin, the glycoside

of hesperetin, is the major flavonoid in orange juice. In vivo
studies have demonstrated that grapefruit juice consump-
tion increased the plasma half-lives of drugs such as caf-
feine. This effect was attributed to inhibition of CYP1A2
by naringin, the major flavonone in grapefruit juice (

Fuhr

et al., 1993

).

Biochanin A, the red clover (Trifolium pretense) isoflav-

one (

Chan et al., 2003

), and the soybean isoflavone geni-

stein (

Chan and Leung, 2003

) are effective inhibitors of

DMBA-induced DNA damage in MCF-7 cells by inhibit-
ing CYP1A1 and CYP1B1. Both isoflavones could reduce
xenobiotic-induced CYP1A1 and 1B1 mRNA expression
through interference with XRE-dependent transactivation.
Enzyme kinetic studies also indicated that biochanin A
inhibits CYP1A1 and 1B1 with K

i

values of 4.00 and

0.59 lM, respectively (

Chan et al., 2003

), and genistein is

an effective inhibitor of recombinant human CYP1A1
and CYP1B1 with K

i

values of 15.35 and 0.68 lM, respec-

tively (

Chan and Leung, 2003

). Extrahepatic CYP1B1

catalyzes the O-demethylation of biochanin A and formo-
nonetin to produce genistein and daidzein, respectively,
which inhibit CYP1B1 (

Roberts et al., 2004

). Inhibition

of CYP1B1 EROD activity by genistein was primarily
non-competitive (K

i

of 1.9 lM), and daidzein exhibited

mixed, but predominantly non-competitive inhibition of
CYP1B1 EROD activity (K

i

of 3.7 lM) (

Roberts et al.,

2004

). Therefore, biochanin A and/or formononetin may

exert anticarcinogenic effects directly by acting as compet-
itive substrates for CYP1B1 or indirectly through their
metabolites daidzein and genistein, which inhibit CYP1B1
(

Roberts et al., 2004

).

Genistein and equol inhibit CYP1A2 in liver micro-

somes from b-naphthoflavone-induced mice with IC

50

val-

ues of 5.6 mM and 1.7 mM, respectively (

Helsby et al.,

1998

). Using human CYP1A2 from a specific expression

system, non-competitive inhibition was seen with both iso-
flavones. This inhibition offers a possible explanation for
the chemopreventive effects of genistein in animals, but
inhibition of CYP1A2 is not likely to be achieved from
the concentration of genistein present in the human diet
(

Helsby et al., 1998

).

Dietary soy isoflavone containing 155 mg/g of genistein,

127 mg/g of daidzein, and other minor isoflavones had no
effect on the hepatic mRNA abundance of CYP1A1 and
1A2 in rats, determined by real-time quantitative RT-
PCR (

Kishida et al., 2004

). This indicates that dietary iso-

flavones do not induce CYPs in either the transcriptional
step or through post-transcriptional mRNA stabilization
(

Kishida et al., 2004

). Another in vivo study has reported

that genistein and equol only affected the protein content
or activity of CYP1A1 and 1A2 following the administra-

Fig. 4. Lineweaver-Burk plot of EROD activity of lymphoblastoid
expressed human CYP1B1 (3.7 pmol) without hesperetin (d) and in
reactions containing 0.01 lM (j) or 0.02 lM hesperetin (m). Reproduced
from (

Doostdar et al., 2000

), with permission from Elsevier Science Inc.

196

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

background image

tion of daily intraperitoneal injections to mice, when doses
of 40 mg/kg or higher of these isoflavones were adminis-
tered (

Helsby et al., 1997

).

Daidzein, a principal isoflavone in soybean, can inhibit

CYP1A2 activity and alter the pharmacokinetics of the-
ophylline in healthy volunteers (

Peng et al., 2003

). Theoph-

ylline is a bronchodilator with a narrow therapeutic index
(5–20 mg/L), and it is primarily eliminated by hepatic
metabolism mediated by CYP1A2. In the study by

Peng

et al. (2003)

, a single dose of 100 mg theophylline was taken

on day 3. Thereafter, one group received 200 mg daidzein
twice daily for 10 days, and the other group received pla-
cebo. On day 12, the test group received 200 mg daidzein
with 100 mg theophylline. Comparing the kinetic parame-
ters of theophylline determined on day 1 (without co-med-
ication) with those determined on day 12 (10-day daidzein),
the AUC, C

max

, and t

1/2

were significantly increased (

Peng

et al., 2003

).

Genistein as well as combination of three (genistein,

daidzein, and glycitein) to five (plus biochanin A and for-
mononetin) isoflavones inhibited tamoxifen a-hydroxyl-
ation in female rat liver microsomes in vitro via
inhibition of CYP1A2 (

Chen et al., 2004

). The inhibition

of a-hydroxylation by genistein was mixed-type with a K

i

value of 10.6 lM (

Chen et al., 2004

). a-Hydroxytamoxifen

and its sulfate conjugate are thought to be responsible for
DNA adduct formation (

Umemoto et al., 2001

). Thus, gen-

istein and its isoflavone analogs have the potential to
decrease the side effects of tamoxifen through metabolic
interactions that inhibit the formation of a-hydroxylation
(

Chen et al., 2004

).

Chalcones belong to the flavonoid family and, naturally

occurring chalcones are predominantly metabolized by
hydroxylation (

De Vincenzo et al., 2000

). Among them,

2

0

-hydroxyl substituted chalcones have attracted much

attention, because they have biologically active properties,
such as prevention of platelet aggregation (

Lin et al., 1997

)

and LPS-induced septic shock (

Batt et al., 1993

) in animal

models. A XRE-luciferase reporter assay indicated that 2

0

-

hydroxychalcone was most effective among five hydroxy-
chalcones in reducing CYP1A1 and 1B1 expression
through the disruption of XRE-transactivation (

Wang

et al., 2005

). The inhibition on CYP1A1 was competitive

and that of CYP1B1 was non-competitive. A decrease in
DMBA-DNA covalent binding was demonstrated in cul-
tures co-treated with 2

0

-hydroxychalcone and DMBA

(

Wang et al., 2005

). On the other hand, 2-hydroxychalcone

showed different effects; it inhibited CYP1A1 and 1B1
activities in recombinant microsomal preparation but
potentiated their expression in MCF-7 cells (

Wang et al.,

2005

). These findings suggested that the position and num-

ber of hydroxyl groups in hydroxychalcone might affect the
CYP1 enzyme inhibition and gene expression.

A series of flavonoids isolated from a resin of the tree

Dracaena cinnabari Balf. inhibited CYP1A activity in
hepatic microsomes isolated from TCDD treated mice
(

Machala et al., 2001

). The resin has been used in folk

medicine as an antiseptic and antiulcerative remedy. The
inhibitory potencies decreased in the following order:
galangin

> 7-hydroxyflavone > quercetin,

chrysin,

2-

hydroxychalcone > 4,6-dihydroxychalcone > homoisoflav-
anes (

Machala et al., 2001

).

2.2. CYP2E

CYP2E1, an ethanol-inducible enzyme, is important in

the field of toxicology and carcinogenesis, and it also has
a role in drug metabolism (

Guengerich et al., 1991

). For

example,

following

an

overdose

of

acetaminophen

CYP2E1 converts acetaminophen to toxic quinones, which
is responsible for the initiation of centrilobular liver toxic-
ity (

Lindros et al., 1990

). The activity of CYP2E1 is known

to be mainly regulated by post-transcrpitional protein sta-
bilization, but the contribution of the transcriptional step is
also significant (

Novak and Woodcroft, 2000

).

Genistein and equol, isoflavones in soy products, inhib-

ited p-nitrophenol (CYP2E1 substrate) metabolism in liver
microsomes from acetone-induced mice with IC

50

values of

approximately 10 mM and 560 lM, respecively (

Helsby

et al., 1998

). The 5-hydroxyl group and the 2,3-double

bond as well as hydroxyl groups in the flavonoid B ring
may be essential for inhibition of aryl hydroxylation by
CYP2E1 (

Helsby et al., 1998

).

Theaflavins, catechins in black tea, decreased the protein

level of CYP2E1 in rat intestinal microsomes after oral
intake for four weeks (

Catterall et al., 2003

). Theaflavins

have been shown to antagonize the carcinogenicity of
nitrosamines in mice (

Shukla and Taneja, 2002

) and to

have antimutagenic activity (

Apostolides et al., 1997

).

Silybin (also known as silibinin or silybinin) is the main

component of silymarin, an extract from milk thistle. Sily-
bin was able to inhibit p-nitrophenol hydroxylation via
CYP2E1 in human liver microsomes (

Zuber et al., 2002

).

It displayed dose-dependent inhibition of enzyme activity
with IC

50

values in the micromolar range. However,

plasma concentrations of the individual flavonolignans fol-
lowing dietary uptake do not exceed 0.5 lM (

Zuber et al.,

2002

), so inhibition of CYP2E1 by flavonolignans in the

diet may be unlikely.

2.3. CYP3A4

CYP3A4 is the largest subfamily of CYP enzymes

expressed in the human liver and gastrointestinal tract. It
is involved in the metabolism of 50% of therapeutic agents
as well as in the activation of toxic and carcinogenic sub-
stances. It has been reported that grapefruit juice alters
the pharmacokinetics of various drugs, including cyclo-
sporine (

Yee et al., 1995

), midazolam (

Kupferschmidt

et al., 1995

), dihydropyridine-type calcium channel block-

ers (

Bailey et al., 1994

), lovastatin (

Kantola et al., 1998

),

simvastatin (

Lilja et al., 1998

), ethinylestradiol (

Weber

et al., 1996

), and triazolam (

Hukkinen et al., 1995

). The

major mechanism for grapefruit juice-drug interactions is

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

197

background image

thought to be due to the inhibition of intestinal CYP3A4
by flavonoids in grapefruit juice (

Ameer and Weintraub,

1997; Evans, 2000; Zhang and Benet, 2001

), although the

inhibition of intestinal p-glycoprotein by flavonoids has
also been observed. Particular interest has focused on the
inhibitory effects of naringin (naringenin-7-rhamnogluco-
side), the major flavonoid in grapefruit juice (>200 mg/
L), on the activity of intestinal CYP3A4 (

Ameer and Wein-

traub, 1997; Bailey et al., 2000; Evans, 2000

). Naringin is

partially metabolized by gastrointestinal bacteria to form
the flavanone, naringenin (

Bailey et al., 2000

). Naringenin

exerts an inhibitory effect on intestinal CYP3A4 within
30 min and impairs the metabolism of the calcium channel
blockers, felodipine, nitrendipine, nisoldipine, and verapa-
mil, when co-administered with grapefruit juice (

Fuhr,

1998

).

Veronese et al. (2003)

demonstrated that consump-

tion of large amounts of grapefruit juice (double strength
three times daily for 3 days) inhibits both intestinal and
hepatic CYP3A4 activity, as quantified by the erythromy-
cin breath test and oral midazolam pharmacokinetics in
healthy male volunteers (

Veronese et al., 2003

). Consump-

tion of typical amounts (single strength) of grapefruit juice
caused less inhibition of intestinal CYP3A4 (

Veronese

et al., 2003

).

Another clinically important inhibitor of CYP3A4 is St.

JohnÕs wort (

Obach, 2000

). St. JohnÕs wort (Hypericum per-

foratum) is one of the most commonly used herbal medi-
cines in the United States. Its major constituents include
flavonols, flavonol glycosides and biflavones (

Barnes

et al., 2001; Jurgenliemk and Nahrstedt, 2002; Nahrstedt
and Butterweck, 1997

).

Obach (2000)

using cDNA-

expressed enzymes showed that hyperforin was a potent
competitive inhibitor of CYP3A4 activities. The flavonoid
compound I3, II8-biapigenin in St. JohnÕs wort was also
shown to be a potent, competitive inhibitor of CYP3A4
activity (

Obach, 2000

). Quercetin in St. JohnÕs wort was

found to inhibit CYP3A4 as well (

Obach, 2000; Zou

et al., 2002

). However, enzyme induction with St. JohnÕs

wort has also been reported. In vitro studies using primary
cultures of human hepatocytes have demonstrated that St.
JohnÕs wort extract is a potent inducer of CYP3A4, and the
responsible component is hyperforin (

Goodwin et al., 2001;

Moore et al., 2000; Wentworth et al., 2000; Zhou et al.,
2003

). Hyperforin is a potent ligand (K

i

of 27 nM) for the

pregnane X receptor (

Moore et al., 2000

), which is an

orphan nuclear receptor regulating expression of CYP3A4,
as well as other enzymes and transporters (

Durr et al.,

2000; Moore et al., 2000; Wentworth et al., 2000

).

Animal studies using probe drugs have provided addi-

tional evidence that St. JohnÕs wort is a potent modulator
of various CYP enzymes, and the induction of CYPs
depends on the dosing regimen. Mice studies have indi-
cated that short-term treatment (four consecutive days)
of St. JohnÕs wort extract (435 mg/kg/day) or hyperforin
(10 mg/kg/day) did not alter erythromycin N-demethylase
(CYP3A) activity (

Bray et al., 2002

). In contrast, adminis-

tration of St. JohnÕs wort extract (140 or 280 mg/kg/day) to

mice for three weeks resulted in a two-fold increase in the
CYP3A activity (

Bray et al., 2002

). The protein level of

CYP3A was also increased six-fold (

Bray et al., 2002

).

The short-term treatment may not activate the pregnane
X receptor. Similarly, human studies indicated that long-
term ingestion (14-days) of St. JohnÕs wort administration
significantly induced the activity of CYP3A4 as measured
by changes in alprazolam pharmacokinetics (

Markowitz

et al., 2003; Wang et al., 2001

), but short-term administra-

tion had no effect on CYP3A4 activity (

Markowitz et al.,

2000

).

Silymarin, milk thistle extract, is a naturally occurring

mixture of flavonolignans (silibinin A, silibinin B, silichri-
stin, silidianin, taxifolin) (

Simanek et al., 2001

). It has been

used to treat liver diseases for hundreds of years. It is
known to protect cardiomyocytes against doxorubicin-
induced oxidative stress, due mainly to its radical scaveng-
ing and iron chelating potency (

Chlopcikova et al., 2004

).

Silymarin significantly decreases CYP3A4 activity (

Fig. 5

)

in primary cultures of human hepatocytes (

Venkatarama-

nan et al., 2000

). The reason for the reduced activity of

CYP3A4 in silymarin-treated cells is not clear at this time.

Beckmann-Knopp et al. (2000)

used two model substrates,

denitronifedipine and erythromycin, and showed that sily-
bin, the major constituent of silymarin, inhibits oxidation
of denitronifedipine in a non-competitive manner, whereas
the effect on erythromycin demethylation involves activa-
tion at low and inhibition at high silybin concentrations.
CYP3A4 shows very complex behavior, in that it may have
multiple conformations with distinct substrate specificities
(

Koley et al., 1996

). Furthermore, cooperativity in oxida-

tions catalyzed by the enzyme have been described and
attributed to the existence of more than one substrate bind-
ing site on the CYP3A4 molecule (

Hosea et al., 2000; Ueng

Treatments

UN

RIF

Sily 0.1 mM Sily 0.25 mM

Form

a

ti

on of 6-beta tes

tos

terone

(nmol/min/mg protein)

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

*

*

Fig. 5. Effect of silymarin on CYP3A4 activity in primary cultures of
human hepatocytes. The formation rate of 6b (OH) testosterone was
measured in primary cultures of human hepatocytes that were not treated
or treated for 48 h with rifampin, or treated for 48 h with silymarin (0.1
and 0.25 mM). *, indicates significantly different from untreated cells.
Reproduced from (

Venkataramanan et al., 2000

), with permission from

the American Society for Pharmacology and Therapteutics.

198

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

background image

et al., 1997

). CYP3A4 activity, assessed by oxidation of

nifedipine, was also decreased by silybin in human liver
microsomes (

Zuber et al., 2002

). The inhibition was dose-

dependent with IC

50

values in the micromolar range.

CYP activities were unaffected by silybin at physiologically
relevant lower concentrations (

Zuber et al., 2002

). Addi-

tionally, milk thistle had no significant effect on CYP3A4
in vivo. Healthy female and male subjects have been
administered milk thistle (175 mg, twice daily, standardized
to 80% silymarins) for 28 days with no significant changes
on phenotype ratios (1-hydroxymidazolam/midazolam) for
CYP3A4 (

Gurley et al., 2004

). After 21 days of milk thistle

extract administration (153 mg silymarin, 3 times daily), no
clinically significant changes in the pharmacokinetics of
indinavir (a CYP3A4 substrate) were noted in humans
(

Piscitelli et al., 2002

). Another study of healthy volunteers

also found that 20 days of silymarin ingestion (160 mg 3
times daily) had no effect on the pharmacokinetics of indi-
navir (

DiCenzo et al., 2003

). Administration of silymarin

(70 mg 3 times daily) for 28 days to healthy volunteers
had no effect on the pharmacokinetics of aminopyrine or
phenylbutazone, two non-specific CYP probes (

Leber and

Knauff, 1976

).

Green tea (Camellia sinensis) extract does not alter

CYP3A4 activity in healthy volunteers (

Donovan et al.,

2004

). The probe drug alprazolam (2 mg, CYP3A4 activ-

ity)

was

administered

orally

to

healthy

volunteers

(n = 11) at baseline, and again after treatment with four
decaffeinated green tea (DGT) capsules/day for 14 days.
Each DGT capsule contained 211 ± 25 mg of green tea cat-
echins and <1 mg of caffeine. The plasma concentration of
the green tea flavonoid, (

)-epigallocatechin gallate

(EGCG), reached 1.3 ± 1.8 lM, 2 h after DGT treatment.
There were no significant differences in alprazolam phar-
macokinetics at baseline and after DGT treatment. In this
study, EGCG was bioavailable from this supplement
(

Donovan et al., 2004

). Therefore the lack of effect of

DGT on CYP3A4 activity cannot be due to a lack of bio-
availability of the extract used in this study (

Donovan

et al., 2004

).

Genistein and diadzein, isoflavones present in soybeans,

were found to inhibit 3A4-mediated metabolism, but their
glycosides were inactive in human microsome preparations
(

Foster et al., 2003

). Genistein exhibited a similar inhibi-

tory activity against the human 3A isoform CYP3A7 (

Fos-

ter et al., 2003

). The results of this study were consistent

with the observations of Evans, indicating that daidzein
and genistein can inhibit oxidative metabolism (

Evans,

2000

).

Glabridin is a major flavonoid in licorice (Glycyrrhiza

glabra). Kent et al. reported that the isoflavan glabridin
inactivated CYP3A4 in a time-, concentration-, and
NADPH-dependent manner, indicative of mechanism-
based inactivation (

Kent et al., 2002; Zhou et al., 2004

).

Metabolism of glabridin by CYP3A4 resulted in the
destruction of the heme moiety, and the loss of the
CYP3A4-reduced carbon oxide spectrum and detectable

heme were correlated with glabridin concentrations (

Kent

et al., 2002

). Glabridin may generate reactive intermediates

that result in heme fragmentation. Two hydroxyl groups on
the 2

0

and 4

0

position of the flavonoid B ring of glabridin

seems to be required for CYP3A4 inactivation (

Kent

et al., 2002

).

The number of hydroxyl groups, as well as the position

of hydroxylation, also plays an important role in the inhib-
itory effects of flavones. Flavones having more hydroxyl
substitutions showed stronger inhibition of CYP3A4 activ-
ity than those with fewer hydroxyl groups (

Ho et al., 2001

).

In this study, the order of inhibitory potency was myrice-
tin > quercetin,

morin > kaempferol > apigenin > flavone

(with 6, 5, 4, 3, 0 hydroxyl groups respectively). This was
confirmed by the existence of a positive significant correla-
tion (r = 0.89, p < 0.0005) between number of hydroxyl
groups and extent of CYP3A4 inhibition of these six com-
pounds. This finding is in agreement with a previous study
showing that the ability of flavonoid compounds either to
inhibit or stimulate benzo(a)pyrene hydroxylation in
human liver microsomes was related to the presence or
absence of hydroxyl groups respectively (

Ho et al., 2001

).

2.4. CYP19

CYP19, also known as aromatase (11b-hydroxysteroid

dehydrogenase), is another member of the cytochrome
P450 enzyme superfamily. This enzyme represents a crucial
enzyme of estrogen biosynthesis; CYP19 converts andro-
stenedione and testosterone to estrone and estradiol,
respectively (

Zhu and Conney, 1998

). Increased expression

of aromatase has been observed in breast cancer tissue
(

Miller et al., 1990; Zhou et al., 1996

). Since estrogens

cause cell proliferation and certain estrogen metabolites
are carcinogens, local expression of CYP19 has been corre-
lated with tumor initiation, promotion and progression.
Local regulation of aromatase by both endogenous factors
as well as exogenous medicinal agents influences the levels
of estrogen available for breast cancer growth (

Brueggeme-

ier et al., 2001

). Flavonoids and isoflavones are structurally

similar to the endogenous steroid hormone, estradiol, and
possess either estrogenic or antiestrogenic activities (

Brueg-

gemeier et al., 2001

), and they are known to be competitive

inhibitors of aromatase with respect to androgen sub-
strates. Flavonoids with antiestrogenic effects have been
reported to be cancer preventive agents in breast cancer
(

Kumar et al., 2004

).

Flavones have previously been reported to be competi-

tive inhibitors of aromatase with respect to the androgen
substrate, with K

i

values at low micromolar concentrations

(

Adlercreutz et al., 1993; Campbell and Kurzer, 1993; Ibra-

him and Abul-Hajj, 1990; Kellis and Vickery, 1984; Pelis-
sero et al., 1996; Wang et al., 1994

). The binding

characteristics and the structural requirements necessary
for the inhibition of human aromatase by flavonoids were
obtained using computer modeling and confirmed by site-
directed mutagenesis (

Kao et al., 1998

). Flavonoids bind

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

199

background image

to the active site of aromatase in an orientation in which
rings A and C mimic rings D and C of the androgen sub-
strate, respectively. Generally, flavones and flavanones
have higher aromatase inhibitory activity than isoflavones
and isoflavanones, which exhibit significant binding affini-
ties for the estrogen receptor (

Kao et al., 1998

). Based on

the study of a series of B ring substituted flavanones with
a 7-methoxy group on the A ring, the structure-activity
relationships suggested that hydroxylation at position 3

0

and/or 4

0

are the optimal pattern of B ring substitution that

enhanced the anti-aromatase activity (

Pouget et al., 2002

).

Kao et al. showed that flavones (chrysin, baicalein, and

galangin), flavanone (naringenin) and isoflavones (biocha-
nin A) inhibit the activity of human aromatase in chinese
hamster ovary (CHO) cells, thus decreasing estrogen bio-
synthesis and circulating estrogen levels (

Kao et al.,

1998

). Similarly, flavone, chrysin, apigenin, naringenin,

and biochanin A inhibited human placental aromatase
(

Le Bail et al., 2000

). Isoflavone, equol was potent inhibitor

of the ovarian aromatase activity in rainbow trout and also
showed inhibitory effect on human placental aromatase
(

Pelissero et al., 1996

). Daily treatment for 21 days with

natural supplements such as propolis and honey, contain-
ing chrysin, could block the conversion of androgens into
estrogens by inhibiting aromatase, with a consequent
increase of testosterone, eventually measurable in urine
samples (

Gambelunghe et al., 2003

).

Almstrup et al. have investigated several flavonoids

using a pS2 mRNA essay for aromatase inhibition in
MCF-7 cells (

Almstrup et al., 2002

). Since aromatase con-

verts testosterone to 17a-estradiol, and pS2 mRNA is reg-
ulated by estrogen, aromatase activity can be measured by
differences in the expression level of the pS2 mRNA after
exposure to testosterone and the test compounds. Biocha-
nin A, formononetin, naringenin, and chrysin are aroma-
tase inhibitors at low concentrations (<1 lM) (

Almstrup

et al., 2002

).

3. Effect of bioflavonoids on phase II enzymes

Activation of phase II detoxifying enzymes, such as

UDP-glucuronyl transferase (UGT), glutathione S-trans-
ferase (GST), and NAD(P)H:quinone oxidoreductase
(QR) by flavonoids results in the detoxification of carcino-
gens and represents one mechanism of their anticarcino-
genic effects (

Fig. 1

). The importance of induction of

Phase II metabolism in cancer prevention has been demon-
strated in studies of nrf-2 knockout mice; nrf-2 is a tran-
scription factor necessary for Phase II enzyme induction
(

Ramos-Gomez et al., 2001

). Nrf2 is normally localized

in the cytosol, where it is associated through protein-pro-
tein interactions with the chaperone Keap1 (

Itoh et al.,

1999

). The presence of an inducer disrupts the Keap1-

Nrf2 interactions, allowing Nrf2 to translocate to the
nucleus and bind to the antioxidant/electrophile response
element (ARE), in conjunction with small Maf proteins,
after activation (

Hayes and McMahon, 2001

). Nrf2 is

important for both the constitutive and inducible expres-
sion of several Phase II proteins, including QR, GST, c-
glutamylcysteine synthetase and UGT. In contrast to
normal mice, nrf2 knock-out mice are more susceptible
to benzo[a]pyrene carcinogenesis and are not protected
by Phase II inducers (

Jeyapaul and Jaiswal, 2000;

McMahon et al., 2001; Ramos-Gomez et al., 2001

).

Several papers have suggested that many of phase II

inducers have, or acquire by metabolism, a electrophilic
Michael reaction acceptor functionality (

Dinkova-Kost-

ova, 2002; Talalay et al., 1988

). Moreover, it was found

that the potency of inducers correlates with reactivity in
the Michael reaction (

Talalay et al., 1988

). The inducer

potency of Michael reaction acceptors is profoundly
increased by the presence of ortho- (but not other) hydro-
xyl substituent(s) on the aromatic ring(s) (

Dinkova-Kost-

ova et al., 2001

). Many flavonoids contain Michael

reaction center(s) in their molecules, thus this characteristic
may be related to their effects on phase II enzymes. Identi-
fied dietary flavonoids acting as phase II inducers include:
kaempferol (

Uda et al., 1997

), a flavonoid present in high

amounts in kale; a flavonoid fraction found in blueber-
ries/cranberries (

Bomser et al., 1996

); the flavolignan silib-

inin (also known as silybin or silybinin) obtained from milk
thistle (

Zhao and Agarwal, 1999

).

3.1. UDP-glucuronyltransferase

Glucuronidation, catalyzed by the UDP-glucuronyl-

transferase family of enzymes, is a major metabolic path-
way of endogenous steroids, bile acids, drugs, and
carcinogens. UGTs have been divided into two families,
termed UGT1 and UGT2 (

Mackenzie et al., 1997

).

UGT1 enzymes mainly catalyse glucuronidation of exoge-
nous agents (drugs, pesticides, benzo[a]pyrene, etc.),
whereas UGT2 enzymes glucuronidate endogenous agents
(steroid hormones and bile acids). However exceptions
exist: for instance, human UGT1A1 can metabolize the
toxic heme breakdown product bilirubin, as well as cate-
chol estrogens, and flavonoids (

King et al., 2001

). In

humans, glucuronidation capacity is prominently present
in the liver, but UGT activity toward bile acids, phenols,
and bilirubin is present in human intestinal, kidney, and
colon tissue (

Tukey and Strassburg, 2000

).

UGT1A1, 1A3, 1A4, 1A6, and 1A9 mRNA expression

is present in human livers. Of importance, UGT1A7 and
1A10 were discovered and cloned from gastric tissue and
UGT1A10 in biliary tissue, indicating these RNAs are
exclusively extrahepatic UGT1A gene products (

Tukey

and Strassburg, 2000

). UGT1A10 appears to be expressed

in all tissues of the gastrointestinal tract except liver. This is
significant because UGT1A10 has one of the widest range
of substrate specificities of any of the UGTs, from small
phenolics to steroids, an indication that it may play a vital
role in most extrahepatic tissues for the glucuronidation of
endogenous and xenobiotic substrates (

Tukey and Strass-

burg, 2000

). The expression of UGT2 genes also follows

200

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

background image

a tissue-specific pattern. The human olfactory UGT2A1,
which has been shown to be one of the more versatile of
the UGTs by recognizing all of the major classes of sub-
strates, is restricted in expression to olfactory tissue. One
could predict that UGT2A1 has evolved for the need to
serve as a first line of metabolic defense for many sub-
stances that enter the body through the nasal mucosa. In
human liver, cDNAs have been identified for UGT2B4,
2B7, 2B11, and 2B15 (

Tukey and Strassburg, 2000

).

UGT2B7 transcripts are also found in intestine, esophagus,
brain, kidney, and pancreas. A significant observation is
that UGT2B transcripts are abundantly expressed in ste-
roid-sensitive target tissues such as prostate and mammary
gland; UGT2B10, 2B11, 2B15, and 2B17 gene transcripts
have been identified in human prostate, and UGT2B11 is
also expressed in mammary gland tissue. The presence of
UGT2B17 may have a significant impact on cancer of the
prostate gland by glucuronidating androgens and thus pro-
tecting this tissue from the carcinogenic actions of these
steroids (

Tukey and Strassburg, 2000

).

As reported by Walle et al., in human hepatoma HepG2

cells and the human intestinal cell line Caco-2, there was
a high level of induction of UGT1A1 by treatment with
the flavone chrysin (

Walle et al., 2000

). Similarly, the

UGT1A1-mediated glucuronidation of quercetin was
greatly increased (

Galijatovic et al., 2001

). This induction

response was quite specific because UGT1A6, UGT1A9,
and UGT2B7 were not affected by chrysin treatment.
Two of the flavonoids that induced CYP1A1, galangin
and isorhamnetin, had no effect on the UGT1A1 activity,
suggesting that the inducing effect of UGT1A1 is not
related to the AhR (

Fig. 6

) (

Walle and Walle, 2002

).

The polymethoxyflavone tangeretin was the most potent

inhibitor of UGT1A1 catalyzed estradiol-3-glucuronida-
tion in human liver microsomes (IC

50

= 1 lM at 5 lM

estradiol concentration) (

Williams et al., 2002

). Naringenin

inhibited estradiol 3-glucuronidation to a similar extent at
all naringenin concentrations tested (5–100 lM) and there-
fore did not act as a competitive-type inhibitor (

Williams

et al., 2002

). Flavone and quercetin were weak inhibitors

of estradiol 3-glucuronidation at the concentrations tested
(5–100 lM) (

Williams et al., 2002

). For chrysin, flavanone,

nobiletin, and silymarin, the greatest inhibitory effect on
estradiol 3-glucuronidation was at substrate concentrations
above 25 lM.

Flavone has been shown to induce hepatic 4-nitrophenol

(NP) UGT activity in rats at a concentration of 0.25%
(

Brouard et al., 1988

), 0.3% (

Canivenc-Lavier et al.,

1996

), and 1% (

van der Logt et al., 2003

) (w/w) in the diet

for 2 weeks. Van der Logt et al. quantified UGT enzyme
activity in hepatic and intestinal (proximal, mid and distal
small intestine and colon) tissue from male rats using NP
and 4-methlyumbelliferone as substrate (

van der Logt

et al., 2003

). In their study, 1% (w/w) quercetin in the basal

diet for 2 weeks caused significant increase of UGT enzyme
activity in liver and proximal and distal small intestine of
the rats. When determined using 2-aminophenol as the sub-

strate, long-term ingestion of green tea increases UGT
activity in rats (

Bu-Abbas et al., 1995; Sohn et al., 1994

),

and this induction is considered to contribute to the anti-
carcinogenic effect of green tea. In contrast, three other
studies reported that quercetin has no effect on hepatic
NP UGT activity (

Brouard et al., 1988; Canivenc-Lavier

et al., 1996; Siess et al., 1996

). This difference may be

caused by different doses, food deprivation in animals or
other differences in experimental design. Since starvation
is known to rapidly decrease phase II enzyme activity (

Sieg-

ers et al., 1989

), an initial increase in UGT enzyme activity

may have disappeared during this food deprivation period.

Sun et al. (1998)

found that UGT was induced by

selected flavonoids (biochanin A, diadzein, formonone-
tin, genistein, prunetin, apigenin, galangin, kaempferol,
naringenin, and quercetin) at a concentration of 5 lM in
the prostate cancer cell line LNCaP. LNCaP cells were
exposed to each of the flavonoids for 6 days. Those flavo-
noids stimulated the activity of testosterone-UGT, which
conjugates testosterone to testosterone glucuronide. Bio-
chanin A was the most potent inducer of UGTs (10-fold
increase at 5 lM), with increased activity over the

Fig. 6. Induction of UGT1A1 (A) and CYP1A (B) in Hep G2 cells by
flavonoids and classical AhR inducers. Api, apigenin; Chry, chrysin; Gal,
galanin; Iso, isorhamnetin; *, significantly higher than control (DMSO),
p < 0.001; **, significantly higher than control, p < 0.05. Reproduced from
(

Walle and Walle, 2002

), with permission from the American Society for

Pharmacology and Therapeutics.

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

201

background image

concentration range of 0.5–50 lM (

Sun et al., 1998

). The

production and release of prostate specific antigen (PSA),
the prostatic tumor marker (

Catalona et al., 1991

), is tes-

tosterone dependent (

Montgomery et al., 1992

), and bioch-

anin A can significantly decrease PSA concentrations,
likely due to increased glucuronidation of testosterone
(

Sun et al., 1998

). These studies suggest that the modula-

tion of hormone metabolism by flavonoids may be impor-
tant in the prevention and treatment of prostate cancer
(

Sun et al., 1998

).

3.2. Glutathione-S-transferase (GST) and quinone
reductase (QR)

Flavonoids also contribute to the modulation of other

phase II detoxifying enzymes, such as GST and QR.
Human cytosolic GSTs are a family of dimeric biotransfor-
mation enzymes comprised of the four main classes; a, l, p,
and h (

Hayes and Pulford, 1995

). They catalyze the binding

of a large variety of electrophiles to the sulfydryl group of
glutathione, are involved in the detoxification of (oxygen)
radicals, and have a main function in the binding and
transport of a wide variety of harmful compounds. GSTs
have a considerably important role in the detoxification
of carcinogens (

Hayes and Pulford, 1995

). GSTs are pres-

ent in many species and tissues and also in relatively large
amounts in the epithelial tissues of the human gastrointes-
tinal tract (

Peters et al., 1991

). A significant negative corre-

lation was demonstrated between GST enzyme activity and
tumor incidence in the mucosa along the human gastroin-
testinal tract, suggesting the importance of GSTs in cancer
prevention (

Peters et al., 1993

). QR prevents quinine redox

cycling and lowers levels of electrophilic quinines (

Kelly

et al., 2000

). Hence, the induction of GST and QR by

flavonoids is possibly associated with cancer chemopreven-
tive effects.

Genistein regulates estrogen receptor (ER)/ARE-depen-

dent gene expression in vitro (

Ansell et al., 2004

). ER reg-

ulation of a mouse GST Ya reporter gene was determined
in two cell lines in the presence of 1 lM of genistein. In
COS I cells expressing ERa and ERb, genistein repressed
GST Ya ARE-dependent gene expression (

Ansell et al.,

2004

); however, treatment of C4-12-5 cells (ER-negative

breast cancer cell line derived from the MCF-7 cell line)
with genistein resulted in modest GST gene induction fol-
lowing transfection with ERa and ERb. This suggests that
the effects of genistein on GST through ER/ARE signaling
are cell type specific (

Ansell et al., 2004

).

Flavone and 2

0

-amino-3

0

methoxyflavone induce the

nuclear translocation of a transcriptional factor—CCAAT/
enhancer-binding protein b (C/EBPb) and induce GSTA2
(a form) gene expression (

Fig. 7

) (

Kang et al., 2003

). Die-

tary administration of flavone increases GST activities (a
and l isoforms) and the levels of glutathione in many tis-
sues of male Wistar rats (

Nijhoff et al., 1995

). Long-term

ingestion of green tea extracts (GTEs) increases cytosolic
GST activity in female rats (

Maliakal et al., 2001

). On

the other hand, quercetin has been shown to effectively
inhibit human placental GST (GSTP1-1), a subclass of
the GST family, in a time- and concentration-dependent
manner in vitro. GSTP1-1 activity is completely inhibited
following a 1 h-incubation with 100 lM quercetin or 2 h-
incubation with 25 lM quercetin. The inactivation mecha-
nism may involve the covalent modification of cysteine
47 in GSTP1-1 by quercetin quinone or its quinone met-
hides (

van Zanden et al., 2003

).

Chang et al. (1997)

reported the isolation of a potent

QR inducer from the pantropical coastal shrub Tephrosia
purpurea, the chalcone, (+)-tephropurpurin. Hop flavo-
noids (prenylchalcones and prenylflavanones) can induce
GST enzymes and QR in the mouse hepatoma Hepa1c1c7
cell line (

Miranda et al., 2000

). In contrast, the flavanone,

naringenin, with no prenyl group, was ineffective in induc-
ing QR (

Miranda et al., 2000

). The hop chalcones, xan-

thohumol and dehydrocycloxanthohumol hydrate, also
induce QR activity in the AhR-defective mutant cell line,
Hepa1c1c7bp(r)c1 (

Miranda et al., 2000

). Dietary admin-

istration of morin to F344 rats led to significant increases
in the activities of QR and GST in the liver, large bowel
and tongue and was protective against azoxymethane-
induced adenocarcinoma of the large intestine (

Tanaka

et al., 1999

), as well as against 4-nitroquinone 1-oxide-

induced tongue carcinogenesis (

Kawabata et al., 1999

).

The flavonoid pinostrobin, present in honey and Thai gin-
ger (Boesenbergia pandurata) represents a potent inducer
of mammalian QR activity (

Fahey and Stephenson,

2002

).

Uda et al. (1997)

reported that a 2,3-double bond

Fig. 7. Immunoblot analysis of rGSTA2 in H4IIE cells treated with
flavone. A representative immunoblot shows the level of rGSTA2 protein
in cells treated with PD98059 (3–30 lM) or flavone (1–10 lM) for 24 h.
Each lane was loaded with 10 lg of cytosolic proteins. The relative
rGSTA2 level was assessed by scanning densitometry. Data represent the
mean ± SD with three separate experiments (significant as compared with
control,

*

P < 0.05;

**

P < 0.01; control level = 1). Reproduced from (

Kang

et al., 2003

), with permission of Oxford University Press.

202

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

background image

in the C ring of flavonoids is critical for the induction of
QR. Among the flavonoids, the flavonols are the most
effective inducers of QR activity in Hepa1c1c7 cells
(kaempferol, galangin > quercetin > myricetin, apigenin
(a flavone)) and flavanols and flavans are ineffective. The
2,3-double bond in the C ring may play a role as a
Michael reaction acceptor (

Dinkova-Kostova, 2002

). In

contrast to flavone, flavanone, which has a saturated
2,3-double bond in the C ring, cannot induce QR activity
(

Dinkova-Kostova, 2002

).

Silymarin can significantly elevate GST and QR activi-

ties in the liver and colon of the rat. These effects may be
related to the in vivo suppressive effects of silymarin on
the occurrence of aberrant crypt foci, a putative precursor
lesions for colonic adenocarcinoma, and azoxymethane-
induced carcinoma (

Kohno et al., 2002

).

Genistein and daidzein were found to increase QR activ-

ities in Hepa1c1c7 cells (

Yannai et al., 1998

). The inhibition

of benzo[a]pyrene metabolite-DNA binding by genistein
may result from QR induction (

Lee et al., 1999

). Induction

of QR activity by genistein has been also observed in the
human colon cancer cell line Colo205 (

Wang et al.,

1998

). QR induction was further confirmed by using

reverse transcription-polymerase chain reaction (RT-
PCR) techniques to measure mRNA expression. A signifi-
cant correlation between the expression of QR mRNA and
the corresponding QR activity was observed (r = 0.76,
P < 0.001) (

Wang et al., 1998

). The single topical applica-

tion of 12-O-tetradecanoyl phorbol-13-acetate (TPA), a
well-known tumor promoter, down-regulated the level of
GST activity (47%) in mouse skin model (

Sharma and Sul-

tana, 2004

). Topical applications of soy isoflavones (mix-

ture of 33 mg of genistein and 67 mg of daidzein), 30 min
prior to the application of TPA prevented the decrease in
GST activity, in a dose dependent manner. Dietary geni-
stein and diadzein are also able to elevate the activities of
GST in the kidney and QR in the colon of female rats
in vivo (

Appelt and Reicks, 1999

).

Siess et al. (1996)

showed that dietary administration of

flavone to male Wistar rats elevated the activities of phase
2 enzymes (GST and UGT) as well as phase I enzymes
(CYP 1A1/2 and 2B1/2) (

Siess et al., 1996

). Importantly,

the activities of CYP1A1/2 increased as early as 6 h after
the first dose, reaching maximal induction after 4 days,
while the earliest elevation of the activities of CYP2B1/2
and phase II enzymes was observed 24 h after flavone feed-
ing. This difference in the time course of induction may
suggest that the ultimate inducers of the CYP2B1/2 and
phase II enzymes are probably flavone metabolites and
not the parent compounds themselves, or that the induc-
tion occurs through different mechanisms for these different
enzymes.

3.3. Sulfotransferases

The cytosolic sulfotransferases (SULTs) catalyze the sul-

fate conjugation of many hormones, neurotransmitters,

and xenobiotic compounds (

Coughtrie et al., 1998

). They

are involved in the Phase II detoxification of xenobiotics,
as well as in the inactivation of endogenous compounds
such as steroid and thyroid hormones, catecholamines
and bile acids (

Coughtrie et al., 1998

). In contrast to this

protective function, sulfation is also a key step in the bio-
activation of a host of pro-mutagens and pro-carcinogens
(

Glatt, 2000; Lewis et al., 2000; Yamazoe et al., 1999

). Sul-

fation activates carcinogens such as hydroxymethyl polycy-
clic aromatic hydrocarbons, allylic alcohols, benzylic
alcohols, and N-hydroxyarylamines, since their sulfate
esters are electrophiles that covalently bind to nucleic acids
and other macromolecules (

Glatt, 2000; Meerman et al.,

1994

).

Flavonoids have been suggested as potential chemopre-

ventive agents in sulfation-induced carcinogenesis (

Cought-

rie and Johnston, 2001; Eaton et al., 1996; Ghazali and
Waring, 1999; Tamura and Matsui, 2000; Walle et al.,
1995

). A number of flavonoids exert inhibitory effects on

sulfotransferase activity. A previous study has demon-
strated that flavonoids (fisetin, galangin, quercetin, myrice-
tin, kaempferol, chrysin, apigenin and genistein) represent
potent inhibitors of the P-form phenolsulfotransferase (P-
PST or SULT1A1)-mediated sulfation of acetaminophen
and minoxidol by human liver cytosol (

Eaton et al.,

1996

). Quercetin, fisetin, and galangin demonstrated simi-

lar potencies for the inhibition of the P-form PST while
myricetin, chrysin, apigenin, and kaempferol were 3–10
times less potent.

The flavonoids daidzein, genistein, quercetin, (+)-cate-

chin, equol and flavone are non-competitive inhibitors of
human platelet P-PST with low K

m

and K

i

values (

Ghazali

and Waring, 1999

). The non-competitive nature of these

flavonoid inhibitors of P-form PST is consistent with the
observation that they are poor substrates for this enzyme
(

Table 3

) (

Ghazali and Waring, 1999

). Consistent with pre-

vious findings, Harris et al. reported that quercetin was the
most potent inhibitor (IC

50

of 60 nM) (

Harris et al., 2004

).

Genistein and daidzein inhibited SULT1A1 with IC

50

val-

ues of 500 and 600 nM, respectively (

Harris et al., 2004

).

Table 3
K

m

, V

max

and K

i

values for inhibition of phenol sulfotransferase (PST) by

flavonoids

Flavonoid

K

m

, lM

V

max

K

i

, lM

nmol/min/mg protein

Quercetin

0.051 ± 0.004

0.051 ± 0.007

0.10 ± 0.03

Genistein

0.047 ± 0.002

0.040 ± 0.001

0.21 ± 0.03

Daidzein

0.050 ± 0.007

0.049 ± 0.004

0.34 ± 0.08

Equol

0.058 ± 0.008

0.030 ± 0.007

0.49 ± 0.11

(+)-Catechin

0.053 ± 0.001

0.058 ± 0.009

0.76 ± 0.03

Flavone

0.061 ± 0.004

0.031 ± 0.006

0.94 ± 0.17

Control P-form-PST activity (without inhibitor) was 0.86 ± 0.09 nmol/
min/mg protein which also had a K

m

value of 0.050 ± 0.007 lM and V

max

of 0.080 ± 0.03 nmol/min/mg protein. Data are expressed as means ± SD.
Reproduced from (

Ghazali and Waring, 1999

), with permission from

Elsevier Science Inc.

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

203

background image

The most potent inhibitor of SULT1A3 in human platelets
was baicalein (IC

50

of 500 nM) (

Harris et al., 2004

).

Quercetin inhibits the sulfation of resveratrol in human

liver and duodenum in vitro (

De Santi et al., 2000

). Resve-

ratrol is a polyphenolic compound present in wine at con-
centration between 1 and 10 lM (

Soleas et al., 1997

). It has

beneficial effects against cancer and protective effects on the
cardiovascular system. The IC

50

for quercetin was 12 pM

(liver) and 15 pM (duodenum) for the inhibition of resvera-
trol sulfation (

De Santi et al., 2000

). The potent inhibition

of resveratrol sulfation by quercetin suggests that com-
pounds present in the diet may inhibit the sulphation of
resveratrol, thus improving its bioavailability.

Marchetti et al. (2001)

also showed that quercetin is a

potent inhibitor of human liver and duodenum sulfotrans-
ferase in vitro. Drugs are often administered orally, and the
intestine and liver are therefore important sites of drug
metabolism. In this study, cytosolic fractions of human
liver samples and biopsies of the duodenum were used.
The IC

50

values of quercetin for the sulfation of four clin-

ically used drugs (dopamine, (

)-salbutamol, minoxidil

and paracetamol) are greater in duodenum than those in
liver. Such a difference may reflect the different composi-
tion of sulfotransferase forms in the liver and duodenum,
since SULT1A1 is preferentially expressed in the liver
and SULT1A3 in the intestine (

Marchetti et al., 2001

).

In the case of estrogen sulfotransferase (SULT1E1),

equol was the most potent inhibitor. Equol has hydroxyl
groups that can potentially superimpose with the 3-hydro-
xyl group of 17b-estradiol (E

2

) (

Harris et al., 2004

). The

4

0

-hydroxyl group appears to play an important role in

the inhibitory effect because formononetin, which lacks
this group, was the least potent of all the compounds
tested (

Harris et al., 2004

). Quercetin competitively inhi-

bits the sulfation of E

2

in normal human mammary epi-

thelial cells (

Otake et al., 2000

). Another study reported

that quercetin, genistein and daidzein inhibit the sulfation
of 17b-estradiol by zebrafish SULT (

Ohkimoto et al.,

2004

). Kinetic analyses showed that the mechanism of

action by these flavonoids was competitive inhibition
(

Ohkimoto et al., 2004

).

The effect of the flavonoids may be mediated not just by

their inhibition of the bioactivating activity of the SULTs
on carcinogens, but by the effects of the sulfoconjugates
of the flavonoids produced by the activity of these enzymes
(

Pai et al., 2001

). It has been shown that daidzein sulfocon-

jugates are potent inhibitors of sterol sulfatase and
decrease the production of the biologically active estro-
genic steroids (shown to stimulate many breast tumors)
in mammary tissue from their inactive sulfoconjugates
(

Wong and Keung, 1997

).

4. Difficulties in the prediction of in vivo metabolic
effects in humans

It is clear that effects of flavonoids in vivo may not

always be predicted on the basis of in vitro results alone.

For example, in the study of

Ueng et al. (1999)

, naringenin

but not naringin is an inhibitor of benzo(a)pyrene hydrox-
ylase (AHH) activity in vitro, whereas naringin inhibited
AHH activity and the expression of CYP1A2 in vivo.
Moreover, flavonoids generally have low oral bioavailabil-
ity and can be degraded by gut bacteria. Therefore, concen-
trations in vivo may not reflect the concentrations tested
under in vitro conditions (

Kuhnau, 1976

). Isoflavone gly-

cosides are not absorbed intact across the enterocyte of
healthy adults, and their bioavailability requires initial
hydrolysis of the sugar moiety by intestinal b-glucosidases
for uptake to the peripheral circulation (

Setchell et al.,

2002

). Apigenin, naringenin, and chrysin, which are strong

inhibitors of aromatase in vitro, did not inhibit androstene-
dione-induced uterine growth, indicating a lack of aroma-
tase-inhibiting effect in vivo (

Saarinen et al., 2001

), a

difference that may be due to their relatively poor absorp-
tion and/or bioavailability. Tangeretin did not alter the
CYP3A4 activity in human volunteers, although it is a
potent stimulator of CYP3A4 activity in human liver
microsomes and microsomes containing cDNA-expressed
CYP3A4 (

Backman et al., 2000

). Quercetin has been shown

to inhibit the catalytic activity of P-PST using cell-free
enzyme preparations in vitro with an IC

50

value as low as

0.1 lM. In the intact human hepatoma cell line HepG2,
the potency of quercetin as an inhibitor of P-PST decreased
about 25-fold, yielding an IC

50

value of 2.5 lM (

Eaton

et al., 1996

). This difference is possibly due to the high

serum protein binding and poor plasma membrane perme-
ability of quercetin, and its metabolism to inactive metab-
olites. Silymarin also has inhibitory effect on CYP3A4
in vitro, but not in vivo. This lack of in vitro–in vivo cor-
relation may be due to poor bioavailability, large inter-
individual variations in silibinin (also known as silybin or
silybinin) absorption, lower CYP binding affinities of silib-
inin conjugates, product variability in silymarin content, or
poor dissolution characteristics of milk thistle dosage
forms (

DiCenzo et al., 2003; Gurley et al., 2004; Moore

et al., 2000

).

Flavonoid metabolites may have a higher or lower bio-

logical activity than the parent drug, and result in a change
of the overall cancer protective response.

Nielsen et al.

(2000)

has identified and quantified 10 different metabo-

lites, each bearing an intact flavan skeleton, in rat urine
and feces, after repeated administration of tangeretin
(abundant in citrus peal), as primarily demethylation
and/or hydroxylation products. The differences in metabo-
lite formation and disposition in vitro and in vivo may
account for some of the differences observed in in vivo
and in vitro studies.

Variable dietary exposure to a range of flavonoid com-

pounds may contribute to some of the interindividual var-
iation in the pharmacokinetics and pharmacological
responses observed for drugs such as phenacetin, caffeine,
and theophylline, which are substrates for CYP1A2 (

Ren-

dic and Di Carlo, 1997

) as well as that observed for drugs

that are substrates for other P450 isozymes (

Zhai et al.,

204

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

background image

1998

). The considerable inter-individual variability in the

inhibitory effects of grapefruit juice on drug metabolism
is due in part to different bacterial strains in the gut, to
CYP polymorphism, and to the differing amounts of flavo-
noids present in a brand of grapefruit juice (

Ameer and

Weintraub, 1997; Fuhr et al., 1993

). Individuals with differ-

ent CYP profiles may derive different benefits from dietary
flavonoids with regard to protection against cancer (

Brein-

holt et al., 2002

). Because of differences in potency and bio-

logical fate of a parent compound and its metabolites,
biotransformation, and the activity of specific CYPs, might
be of considerable importance in mediating the overall can-
cer protective response (

Breinholt et al., 2002

). It should be

also emphasized that metabolism of flavonoids by CYP
may affect the metabolism of other CYP substrates thus
adding to the complexity of this issue.

Gender differences in drug metabolism in rats have been

known for more than 60 years since it was reported that the
much shorter duration of drug action in the male was due to
the effects of testicular androgens (

Liu et al., 2003

). The

activities of hepatic drug-metabolizing enzymes, especially
CYPs and SULTs can be regulated through the sex-related
secretion pattern of growth hormone (

Kato, 1995

). Some

studies have reported an influence of gender on the flavo-
noid-mediated alterations of drug-metabolizing enzymes
(

Finnen and Hassall, 1984; Kobayashi et al., 2000

).

Liu

et al. (2003)

have shown a marked sex difference in the

effects of long-term treatment with tea polyphenols on
hepatic drug-metabolizing enzymes in rats (

Liu et al., 2003

).

Differences between species in enzymes, in drug metabo-

lism rates, and in toxicity or resistance to various chemicals
also can be substantial. Flavone inhibits EROD activity in
rat and human liver, with IC

50

values determined in human

liver microsomes that are 100 times lower than that for rat
liver microsomes (

Siess et al., 1995

).

Tsyrlov et al. (1994)

explored the effects of flavonoids on activities catalyzed
by mouse CYP1A1 and CYP1A2 and human CYP1A2.
The study showed that some flavonoids had different effects
on mouse and human CYP1A2. Therefore, care is needed
when extrapolating data from in vitro studies and in vivo
animal investigations to humans.

5. Conclusion

Although flavonoids have been studied for about 50

years, the cellular mechanisms involved in their biological
activity are still largely unknown (

Depeint et al., 2002

).

Within the last decade, reports on flavonoid activities have
been predominantly associated with enzyme inhibition or
induction and anti-proliferative activity. The modulation
of drug-metabolizing enzymes by flavonoids is important
in terms of human health since these enzymes can inacti-
vate carcinogens, which contributes to the cancer preven-
tive

properties

of

these

compounds.

Additionally,

flavonoids may also interact with chemotherapeutic drugs
used in cancer treatment through the induction or inhibi-
tion of their metabolism.

Acknowledgement

We acknowledge research support through the Susan G.

Komen Breast Cancer Foundation, U.S. Army Breast Can-
cer Research Program Contract DAMD17-00-1-0376 and
Pfizer Global Research Inc.

References

Adlercreutz, H., Bannwart, C., Wahala, K., Makela, T., Brunow, G.,

Hase, T., Arosemena, P.J., Kellis Jr., J.T., Vickery, L.E., 1993.
Inhibition of human aromatase by mammalian lignans and isoflavo-
noid phytoestrogens. J. Steroid Biochem. Mol. Biol. 44, 147–153.

Almstrup, K., Fernandez, M.F., Petersen, J.H., Olea, N., Skakkebaek,

N.E., Leffers, H., 2002. Dual effects of phytoestrogens result in u-
shaped dose–response curves. Environ. Health Perspect. 110, 743–748.

Ameer, B., Weintraub, R.A., 1997. Drug interactions with grapefruit juice.

Clin. Pharmacokinet. 33, 103–121.

Ansell, P.J., Espinosa-Nicholas, C., Curran, E.M., Judy, B.M., Philips,

B.J., Hannink, M., Lubahn, D.B., 2004. In vitro and in vivo regulation
of antioxidant response element-dependent gene expression by estro-
gens. Endocrinology 145, 311–317.

Apostolides, Z., Balentine, D.A., Harbowy, M.E., Hara, Y., Weisburger,

J.H., 1997. Inhibition of PhIP mutagenicity by catechins, and by
theaflavins and gallate esters. Mutat. Res. 389, 167–172.

Appelt, L.C., Reicks, M.M., 1999. Soy induces phase II enzymes but does

not

inhibit

dimethylbenz[a]anthracene-induced

carcinogenesis

in

female rats. J. Nutr. 129, 1820–1826.

Backman, J.T., Maenpaa, J., Belle, D.J., Wrighton, S.A., Kivisto, K.T.,

Neuvonen, P.J., 2000. Lack of correlation between in vitro and in vivo
studies on the effects of tangeretin and tangerine juice on midazolam
hydroxylation. Clin. Pharmacol. Ther. 67, 382–390.

Bailey, D.G., Arnold, J.M., Spence, J.D., 1994. Grapefruit juice

and drugs. How significant is the interaction? Clin. Pharmacokinet.
26, 91–98.

Bailey, D.G., Dresser, G.K., Kreeft, J.H., Munoz, C., Freeman, D.J.,

Bend, J.R., 2000. Grapefruit-felodipine interaction: effect of unpro-
cessed fruit and probable active ingredients. Clin. Pharmacol. Ther. 68,
468–477.

Barnes, J., Anderson, L.A., Phillipson, J.D., 2001. St JohnÕs wort

(Hypericum perforatum L.): a review of its chemistry, pharmacology
and clinical properties. J. Pharm. Pharmacol. 53, 583–600.

Batt, D.G., Goodman, R., Jones, D.G., Kerr, J.S., Mantegna, L.R.,

McAllister, C., Newton, R.C., Nurnberg, S., Welch, P.K., Covington,
M.B., 1993. 2Õ-Substituted chalcone derivatives as inhibitors of
interleukin-1 biosynthesis. J. Med. Chem. 36, 1434–1442.

Beckmann-Knopp, S., Rietbrock, S., Weyhenmeyer, R., Bocker, R.H.,

Beckurts, K.T., Lang, W., Hunz, M., Fuhr, U., 2000. Inhibitory effects
of silibinin on cytochrome P-450 enzymes in human liver microsomes.
Pharmacol. Toxicol. 86, 250–256.

Beecher, G.R., 2003. Overview of dietary flavonoids: nomenclature,

occurrence and intake. J. Nutr. 133, 3248S–3254S.

Bomser, J., Madhavi, D.L., Singletary, K., Smith, M.A., 1996. In vitro

anticancer activity of fruit extracts from vaccinium species. Planta
Med. 62, 212–216.

Bray, B.J., Perry, N.B., Menkes, D.B., Rosengren, R.J., 2002. Toxicol. Sci.

66, 27–33.

Breinholt, V.M., Offord, E.A., Brouwer, C., Nielsen, S.E., Brosen, K.,

Friedberg, T., 2002. In vitro investigation of cytochrome P450-
mediated metabolism of dietary flavonoids. Food Chem. Toxicol. 40,
609–616.

Brosen, K., 1995. Drug interactions and the cytochrome P450 system. The

role of cytochrome P450 1A2. Clin. Pharmacokinet. 29 (Suppl 1), 20–
25.

Brouard, C., Siess, M.H., Vernevaut, M.F., Suschetet, M., 1988. Com-

parison of the effects of feeding quercetin or flavone on hepatic and

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

205

background image

intestinal drug-metabolizing enzymes of the rat. Food Chem. Toxicol.
26, 99–103.

Brueggemeier, R.W., Richards, J.A., Joomprabutra, S., Bhat, A.S.,

Whetstone, J.L., 2001. Molecular pharmacology of aromatase and
its regulation by endogenous and exogenous agents. J. Steroid
Biochem. Mol. Biol. 79, 75–84.

Bu-Abbas, A., Clifford, M.N., Ioannides, C., Walker, R., 1995. Stimula-

tion of rat hepatic UDP-glucuronosyl transferase activity following
treatment with green tea. Food Chem. Toxicol. 33, 27–30.

Campbell, D.R., Kurzer, M.S., 1993. Flavonoid inhibition of aromatase

enzyme activity in human preadipocytes. J. Steroid Biochem. Mol.
Biol. 46, 381–388.

Canivenc-Lavier, M.C., Vernevaut, M.F., Totis, M., Siess, M.H., Mag-

dalou, J., Suschetet, M., 1996. Comparative effects of flavonoids and
model inducers on drug-metabolizing enzymes in rat liver. Toxicology
114, 19–27.

Catalona, W.J., Smith, D.S., Ratliff, T.L., Dodds, K.M., Coplen, D.E.,

Yuan, J.J., Petros, J.A., Andriole, G.L., 1991. Measurement of
prostate-specific antigen in serum as a screening test for prostate
cancer. N. Engl. J. Med. 324, 1156–1161.

Catterall, F., McArdle, N.J., Mitchell, L., Papayanni, A., Clifford, M.N.,

Ioannides, C., 2003. Hepatic and intestinal cytochrome P450 and
conjugase activities in rats treated with black tea theafulvins and
theaflavins. Food Chem. Toxicol. 41, 1141–1147.

Chan, H.Y., Chen, Z.Y., Tsang, D.S., Leung, L.K., 2002. Baicalein

inhibits DMBA-DNA adduct formation by modulating CYP1A1 and
CYP1B1 activities. Biomed. Pharmacother. 56, 269–275.

Chan, H.Y., Leung, L.K., 2003. A potential protective mechanism of soya

isoflavones against 7,12-dimethylbenz[a]anthracene tumour initiation.
Br. J. Nutr. 90, 457–465.

Chan, H.Y., Wang, H., Leung, L.K., 2003. The red clover (Trifo-

lium pratense) isoflavone biochanin A modulates the biotransfor-
mation pathways of 7,12-dimethylbenz[a]anthracene. Br. J. Nutr. 90,
87–92.

Chang, L.C., Gerhauser, C., Song, L., Farnsworth, N.R., Pezzuto, J.M.,

Kinghorn, A.D., 1997. Activity-guided isolation of constituents of
Tephrosia purpurea with the potential to induce the phase II enzyme,
quinone reductase. J. Nat. Prod. 60, 869–873.

Chen, C., Kong, A.N., 2004. Dietary chemopreventive compounds and

ARE/EpRE signaling. Free Radic. Biol. Med. 36, 1505–1516.

Chen, J., Halls, S.C., Alfaro, J.F., Zhou, Z., Hu, M., 2004. Potential

beneficial metabolic interactions between tamoxifen and isoflavones
via cytochrome P450-mediated pathways in female rat liver micro-
somes. Pharm. Res. 21, 2095–2104.

Chlopcikova, S., Psotova, J., Miketova, P., Simanek, V., 2004. Chemo-

protective effect of plant phenolics against anthracycline-induced
toxicity on rat cardiomyocytes. Part I. Silymarin and its flavonolign-
ans. Phytother. Res. 18, 107–110.

Ciolino, H.P., Daschner, P.J., Yeh, G.C., 1999. Dietary flavonols

quercetin and kaempferol are ligands of the aryl hydrocarbon receptor
that affect CYP1A1 transcription differentially. Biochem. J. 340 (Pt 3),
715–722.

Ciolino, H.P., Wang, T.T., Yeh, G.C., 1998. Diosmin and diosmetin are

agonists of the aryl hydrocarbon receptor that differentially affect
cytochrome P450 1A1 activity. Cancer Res. 58, 2754–2760.

Ciolino, H.P., Yeh, G.C., 1999. The flavonoid galangin is an inhibitor of

CYP1A1 activity and an agonist/antagonist of the aryl hydrocarbon
receptor. Br. J. Cancer 79, 1340–1346.

Conney, A.H., 2003. Enzyme induction and dietary chemicals as

approaches to cancer chemoprevention: the Seventh DeWitt S.
Goodman Lecture. Cancer Res. 63, 7005–7031.

Coughtrie, M.W., Johnston, L.E., 2001. Interactions between dietary

chemicals and human sulfotransferases-molecular mechanisms and
clinical significance. Drug Metab. Dispos. 29, 522–528.

Coughtrie, M.W., Sharp, S., Maxwell, K., Innes, N.P., 1998. Biology

and function of the reversible sulfation pathway catalysed by
human sulfotransferases and sulfatases. Chem. Biol. Interact. 109,
3–27.

Cushman, M., Nagarathnam, D., Burg, D.L., Geahlen, R.L., 1991.

Synthesis and protein-tyrosine kinase inhibitory activities of flavonoid
analogues. J. Med. Chem. 34, 798–806.

Dai, R., Zhai, S., Wei, X., Pincus, M.R., Vestal, R.E., Friedman, F.K.,

1998. Inhibition of human cytochrome P450 1A2 by flavones: a
molecular modeling study. J. Protein Chem. 17, 643–650.

De Santi, C., Pietrabissa, A., Spisni, R., Mosca, F., Pacifici, G.M., 2000.

Sulphation of resveratrol, a natural product present in grapes and
wine, in the human liver and duodenum. Xenobiotica 30, 609–617.

De Vincenzo, R., Ferlini, C., Distefano, M., Gaggini, C., Riva, A.,

Bombardelli, E., Morazzoni, P., Valenti, P., Belluti, F., Ranelletti,
F.O., Mancuso, S., Scambia, G., 2000. In vitro evaluation of newly
developed chalcone analogues in human cancer cells. Cancer Chemo-
ther. Pharmacol. 46, 305–312.

Depeint, F., Gee, J.M., Williamson, G., Johnson, I.T., 2002. Evidence for

consistent patterns between flavonoid structures and cellular activities.
Proc. Nutr. Soc. 61, 97–103.

DiCenzo, R., Shelton, M., Jordan, K., Koval, C., Forrest, A., Reichman,

R., Morse, G., 2003. Coadministration of milk thistle and indinavir in
healthy subjects. Pharmacotherapy 23, 866–870.

Dinkova-Kostova, A.T., 2002. Protection against cancer by plant

phenylpropenoids: induction of mammalian anticarcinogenic enzymes.
Mini. Rev. Med. Chem. 2, 595–610.

Dinkova-Kostova, A.T., Massiah, M.A., Bozak, R.E., Hicks, R.J.,

Talalay, P., 2001. Potency of Michael reaction acceptors as inducers
of enzymes that protect against carcinogenesis depends on their
reactivity with sulfhydryl groups. Proc. Natl. Acad. Sci. USA 98,
3404–3409.

Doerge, D.R., Chang, H.C., 2002. Inactivation of thyroid peroxidase by

soy isoflavones, in vitro and in vivo. J. Chromatogr. B Analyt.
Technol. Biomed. Life Sci. 777, 269–279.

Donovan, J.L., Chavin, K.D., Devane, C.L., Taylor, R.M., Wang, J.S.,

Ruan, Y., Markowitz, J.S., 2004. Green tea (Camellia sinensis) extract
does not alter cytochrome p450 3A4 or 2D6 activity in healthy
volunteers. Drug Metab. Dispos. 32, 906–908.

Doostdar, H., Burke, M.D., Mayer, R.T., 2000. Bioflavonoids: selective

substrates and inhibitors for cytochrome P450 CYP1A and CYP1B1.
Toxicology 144, 31–38.

Durr, D., Stieger, B., Kullak-Ublick, G.A., Rentsch, K.M., Steinert, H.C.,

Meier, P.J., Fattinger, K., 2000. St JohnÕs Wort induces intestinal P-
glycoprotein/MDR1 and intestinal and hepatic CYP3A4. Clin. Phar-
macol. Ther. 68, 598–604.

Eaton, E.A., Walle, U.K., Lewis, A.J., Hudson, T., Wilson, A.A., Walle,

T., 1996. Flavonoids, potent inhibitors of the human P-form pheno-
lsulfotransferase. Potential role in drug metabolism and chemopre-
vention. Drug Metab. Dispos. 24, 232–237.

Evans, A.M, 2000. Influence of dietary components on the gastrointestinal

metabolism and transport of drugs. Ther. Drug Monit. 22, 131–136.

Fahey, J.W., Stephenson, K.K., 2002. Pinostrobin from honey and Thai

ginger (Boesenbergia pandurata): a potent flavonoid inducer of
mammalian phase 2 chemoprotective and antioxidant enzymes. J.
Agric. Food Chem. 50, 7472–7476.

Feng, Q., Torii, Y., Uchida, K., Nakamura, Y., Hara, Y., Osawa, T.,

2002. Black tea polyphenols, theaflavins, prevent cellular DNA
damage by inhibiting oxidative stress and suppressing cytochrome
P450 1A1 in cell cultures. J. Agric. Food Chem. 50, 213–220.

Finnen, M.J., Hassall, K.A., 1984. Effects of hypophysectomy on sex

differences in the induction and depression of hepatic drug-metaboliz-
ing enzymes in the rat. J. Pharmacol. Exp. Ther. 229, 250–254.

Foster, B.C., Vandenhoek, S., Hana, J., Krantis, A., Akhtar, M.H.,

Bryan, M., Budzinski, J.W., Ramputh, A., Arnason, J.T., 2003. In
vitro inhibition of human cytochrome P450-mediated metabolism of
marker substrates by natural products. Phytomedicine 10, 334–342.

Fuhr, U., 1998. Drug interactions with grapefruit juice. Extent probable

mechanism and clinical relevance. Drug Saf. 18, 251–272.

Fuhr, U., Klittich, K., Staib, A.H., 1993. Inhibitory effect of grapefruit

juice and its bitter principal, naringenin, on CYP1A2 dependent
metabolism of caffeine in man. Br. J. Clin. Pharmacol. 35, 431–436.

206

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

background image

Galijatovic, A., Otake, Y., Walle, U.K., Walle, T., 2001. Induction of

UDP-glucuronosyltransferase UGT1A1 by the flavonoid chrysin in
Caco-2 cells—potential role in carcinogen bioinactivation. Pharm.
Res. 18, 374–379.

Gambelunghe, C., Rossi, R., Sommavilla, M., Ferranti, C., Ciculi, C.,

Gizzi, S., Micheletti, A., Rufini, S., 2003. Effects of chrysin on urinary
testosterone levels in human males. J. Med. Food 6, 387–390.

Ghazali, R.A., Waring, R.H., 1999. The effects of flavonoids on human

phenolsulphotransferases: potential in drug metabolism and chemo-
prevention. Life Sci. 65, 1625–1632.

Gil, B., Sanz, M.J., Terencio, M.C., Ferrandiz, M.L., Bustos, G., Paya,

M., Gunasegaran, R., Alcaraz, M.J., 1994. Effects of flavonoids on
Naja naja and human recombinant synovial phospholipases A2 and
inflammatory responses in mice. Life Sci. 54, PL333–PL338.

Glatt, H., 2000. Sulfotransferases in the bioactivation of xenobiotics.

Chem. Biol. Interact. 129, 141–170.

Goodwin, B., Moore, L.B., Stoltz, C.M., McKee, D.D., Kliewer, S.A.,

2001. Regulation of the human CYP2B6 gene by the nuclear pregnane
X receptor. Mol. Pharmacol. 60, 427–431.

Graham, H.N., 1992. Green tea composition, consumption, and polyphe-

nol chemistry. Prev. Med. 21, 334–350.

Guengerich, F.P., Kim, D.H., Iwasaki, M., 1991. Role of human

cytochrome P-450 IIE1 in the oxidation of many low molecular
weight cancer suspects. Chem. Res. Toxicol. 4, 168–179.

Guengerich, F.P., Shimada, T., 1991. Oxidation of toxic and carcinogenic

chemicals by human cytochrome P-450 enzymes. Chem. Res. Toxicol.
4, 391–407.

Gurley, B.J., Gardner, S.F., Hubbard, M.A., Williams, D.K., Gentry,

W.B., Carrier, J., Khan, I.A., Edwards, D.J., Shah, A., 2004. In vivo
assessment of botanical supplementation on human cytochrome P450
phenotypes: Citrus aurantium, Echinacea purpurea, milk thistle, and
saw palmetto. Clin. Pharmacol. Ther. 76, 428–440.

Harris, R.M., Wood, D.M., Bottomley, L., Blagg, S., Owen, K., Hughes,

P.J., Waring, R.H., Kirk, C.J., 2004. Phytoestrogens are potent
inhibitors of estrogen sulfation: implications for breast cancer risk and
treatment. J. Clin. Endocrinol. Metab. 89, 1779–1787.

Hayes, J.D., McMahon, M., 2001. Molecular basis for the contribution of

the antioxidant responsive element to cancer chemoprevention. Cancer
Lett. 174, 103–113.

Hayes, J.D., Pulford, D.J., 1995. The glutathione S-transferase supergene

family: regulation of GST and the contribution of the isoenzymes to
cancer chemoprotection and drug resistance. Crit. Rev. Biochem. Mol.
Biol. 30, 445–600.

Heffernan, L.M., Winston, G.W., 2000. Distribution of microsomal CO-

binding chromophores and EROD activity in sea anemone tissues.
Mar. Environ. Res. 50, 23–27.

Helsby, N.A., Chipman, J.K., Gescher, A., Kerr, D., 1998. Inhibition of

mouse and human CYP 1A- and 2E1-dependent substrate metabolism
by the isoflavonoids genistein and equol. Food Chem. Toxicol. 36,
375–382.

Helsby, N.A., Williams, J., Kerr, D., Gescher, A., Chipman, J.K., 1997.

The isoflavones equol and genistein do not induce xenobiotic-metab-
olizing enzymes in mouse and in human cells. Xenobiotica 27, 587–
596.

Henderson, M.C., Miranda, C.L., Stevens, J.F., Deinzer, M.L., Buhler,

D.R., 2000. In vitro inhibition of human P450 enzymes by preny-
lated flavonoids from hops, Humulus lupulus. Xenobiotica 30,
235–251.

Ho, P.C., Saville, D.J., Wanwimolruk, S., 2001. Inhibition of human

CYP3A4 activity by grapefruit flavonoids, furanocoumarins and
related compounds. J. Pharm. Pharm. Sci. 4, 217–227.

Hodek, P., Trefil, P., Stiborova, M., 2002. Flavonoids-potent and versatile

biologically active compounds interacting with cytochromes P450.
Chem. Biol. Interact. 139, 1–21.

Hodnick, W.F., Duval, D.L., Pardini, R.S., 1994. Inhibition of mito-

chondrial respiration and cyanide-stimulated generation of reactive
oxygen species by selected flavonoids. Biochem. Pharmacol. 47, 573–
580.

Hollman, P.C., Katan, M.B., 1997. Absorption, metabolism and health

effects of dietary flavonoids in man. Biomed. Pharmacother. 51, 305–
310.

Hosea, N.A., Miller, G.P., Guengerich, F.P., 2000. Elucidation of distinct

ligand binding sites for cytochrome P450 3A4. Biochemistry 39, 5929–
5939.

Hukkinen, S.K., Varhe, A., Olkkola, K.T., Neuvonen, P.J., 1995. Plasma

concentrations of triazolam are increased by concomitant ingestion of
grapefruit juice. Clin. Pharmacol. Ther. 58, 127–131.

Ibrahim, A.R., Abul-Hajj, Y.J., 1990. Aromatase inhibition by flavonoids.

J. Steroid Biochem. Mol. Biol. 37, 257–260.

Itoh, K., Wakabayashi, N., Katoh, Y., Ishii, T., Igarashi, K., Engel, J.D.,

Yamamoto, M., 1999. Keap1 represses nuclear activation of antiox-
idant responsive elements by Nrf2 through binding to the amino-
terminal Neh2 domain. Genes Dev. 13, 76–86.

Jefcoate, C.R., Liehr, J.G., Santen, R.J., Sutter, T.R., Yager, J.D., Yue,

W., Santner, S.J., Tekmal, R., Demers, L., Pauley, R., Naftolin, F.,
Mor, G., Berstein, L., 2000. Tissue-specific synthesis and oxidative
metabolism of estrogens. J. Natl. Cancer Inst. Monogr., 95–112.

Jeyapaul, J., Jaiswal, A.K., 2000. Nrf2 and c-Jun regulation of antioxidant

response element (ARE)-mediated expression and induction of
gamma-glutamylcysteine synthetase heavy subunit gene. Biochem.
Pharmacol. 59, 1433–1439.

Jurgenliemk, G., Nahrstedt, A., 2002. Phenolic compounds from Hyper-

icum perforatum. Planta Med. 68, 88–91.

Kang, K.W., Park, E.Y., Kim, S.G., 2003. Activation of CCAAT/

enhancer-binding

protein

beta

by

2

0

-amino-3

0

-methoxyflavone

(PD98059) leads to the induction of glutathione S- transferase A2.
Carcinogenesis. 24, 475–482.

Kang, Z.C., Tsai, S.J., Lee, H., 1999. Quercetin inhibits benzo[a]pyrene-

induced DNA adducts in human Hep G2 cells by altering cytochrome
P-450 1A1 gene expression. Nutr. Cancer 35, 175–179.

Kantola, T., Kivisto, K.T., Neuvonen, P.J., 1998. Grapefruit juice greatly

increases serum concentrations of lovastatin and lovastatin acid. Clin.
Pharmacol. Ther. 63, 397–402.

Kao, Y.C., Zhou, C., Sherman, M., Laughton, C.A., Chen, S., 1998.

Molecular basis of the inhibition of human aromatase (estrogen
synthetase) by flavone and isoflavone phytoestrogens: a site-directed
mutagenesis study. Environ. Health Perspect. 106, 85–92.

Kato, R., 1995. Molecular pharmacological and toxicological studies of

drug-metabolizing enzymes. Yakugaku Zasshi 115, 661–680.

Kawabata, K., Tanaka, T., Honjo, S., Kakumoto, M., Hara, A., Makita,

H., Tatematsu, N., Ushida, J., Tsuda, H., Mori, H., 1999. Chemo-
preventive effect of dietary flavonoid morin on chemically induced rat
tongue carcinogenesis. Int. J. Cancer 83, 381–386.

Kellis Jr., J.T., Vickery, L.E., 1984. Inhibition of human estrogen

synthetase (aromatase) by flavones. Science 225, 1032–1034.

Kelly, V.P., Ellis, E.M., Manson, M.M., Chanas, S.A., Moffat, G.J.,

McLeod, R., Judah, D.J., Neal, G.E., Hayes, J.D., 2000. Chemopre-
vention of aflatoxin B1 hepatocarcinogenesis by coumarin, a natural
benzopyrone that is a potent inducer of aflatoxin B1-aldehyde
reductase, the glutathione S-transferase A5 and P1 subunits, and
NAD(P)H: quinone oxidoreductase in rat liver. Cancer Res. 60, 957–
969.

Kent, U.M., Aviram, M., Rosenblat, M., Hollenberg, P.F., 2002. The

licorice root derived isoflavan glabridin inhibits the activities of human
cytochrome P450S 3A4, 2B6, and 2C9. Drug Metab. Dispos. 30, 709–
715.

Kim, D.H., Jin, Y.H., Park, J.B., Kobashi, K., 1994. Silymarin and its

components are inhibitors of beta-glucuronidase. Biol. Pharm. Bull.
17, 443–445.

King, C., Tang, W., Ngui, J., Tephly, T., Braun, M., 2001. Character-

ization of rat and human UDP-glucuronosyltransferases respon-
sible for the in vitro glucuronidation of diclofenac. Toxicol. Sci. 61,
49–53.

Kishida, T., Nagamoto, M., Ohtsu, Y., Watakabe, M., Ohshima, D.,

Nashiki, K., Mizushige, T., Izumi, T., Obata, A., Ebihara, K., 2004.
Lack of an inducible effect of dietary soy isoflavones on the mRNA

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

207

background image

abundance of hepatic cytochrome P-450 isozymes in rats. Biosci.
Biotechnol. Biochem. 68, 508–515.

Kobayashi, Y., Ohshiro, N., Suzuki, M., Sasaki, T., Tokuyama, S.,

Yoshida, T., Yamamoto, T., 2000. Sex-related effect of hemin on
cytochrome P450 and drug-metabolizing enzymes in rat liver. J.
Toxicol. Sci. 25, 213–222.

Kohn, M.C., Walker, N.J., Kim, A.H., Portier, C.J., 2001. Physiological

modeling of a proposed mechanism of enzyme induction by TCDD.
Toxicology 162, 193–208.

Kohno, H., Tanaka, T., Kawabata, K., Hirose, Y., Sugie, S., Tsuda, H.,

Mori, H., 2002. Silymarin, a naturally occurring polyphenolic antiox-
idant flavonoid, inhibits azoxymethane-induced colon carcinogenesis
in male F344 rats. Int. J. Cancer 101, 461–468.

Koley, A.P., Robinson, R.C., Friedman, F.K., 1996. Cytochrome P450

conformation and substrate interactions as probed by CO binding
kinetics. Biochimie 78, 706–713.

Kronenberg, S., Esser, C., Carlberg, C., 2000. An aryl hydrocarbon

receptor conformation acts as the functional core of nuclear dioxin
signaling. Nucleic Acids Res. 28, 2286–2291.

Kuhnau, J., 1976. The flavonoids. A class of semi-essential food

components: their role in human nutrition. World Rev. Nutr. Diet
24, 117–191.

Kumar, N., Allen, K., Riccardi, D., Kazi, A., Heine, J., 2004. Isoflavones

in breast cancer chemoprevention: where do we go from here? Front
Biosci. 9, 2927–2934.

Kupferschmidt, H.H., Ha, H.R., Ziegler, W.H., Meier, P.J., Krahenbuhl,

S., 1995. Interaction between grapefruit juice and midazolam in
humans. Clin. Pharmacol. Ther. 58, 20–28.

Laughton, M.J., Evans, P.J., Moroney, M.A., Hoult, J.R., Halliwell, B.,

1991. Inhibition of mammalian 5-lipoxygenase and cyclo-oxygenase by
flavonoids and phenolic dietary additives. Relationship to antioxidant
activity and to iron ion-reducing ability. Biochem. Pharmacol. 42,
1673–1681.

Le Bail, J.C., Champavier, Y., Chulia, A.J., Habrioux, G., 2000. Effects of

phytoestrogens on aromatase, 3beta and 17beta-hydroxysteroid
dehydrogenase activities and human breast cancer cells. Life Sci. 66,
1281–1291.

Leber, H.W., Knauff, S., 1976. Influence of silymarin on drug metabo-

lizing enzymes in rat and man. Arzneimittelforschung 26, 1603–1605.

Lee, H., Wang, H.W., Su, H.Y., Hao, N.J., 1994. The structure–activity

relationships of flavonoids as inhibitors of cytochrome P-450 enzymes
in rat liver microsomes and the mutagenicity of 2-amino-3-methyl-
imidazo[4,5-f]quinoline. Mutagenesis 9, 101–106.

Lee, S.K., Song, L., Mata-Greenwood, E., Kelloff, G.J., Steele, V.E.,

Pezzuto, J.M., 1999. Modulation of in vitro biomarkers of the
carcinogenic process by chemopreventive agents. Anticancer Res. 19,
35–44.

Lewis, A.J., Otake, Y., Walle, U.K., Walle, T., 2000. Sulphonation of N-

hydroxy-2-acetylaminofluorene by human dehydroepiandrosterone
sulphotransferase. Xenobiotica 30, 253–261.

Lilja, J.J., Kivisto, K.T., Neuvonen, P.J., 1998. Grapefruit juice-simva-

statin interaction: effect on serum concentrations of simvastatin,
simvastatin acid, and HMG-CoA reductase inhibitors. Clin. Pharma-
col. Ther. 64, 477–483.

Lin, C.N., Lee, T.H., Hsu, M.F., Wang, J.P., Ko, F.N., Teng, C.M., 1997.

2

0

,5

0

-Dihydroxychalcone as a potent chemical mediator and cycloox-

ygenase inhibitor. J. Pharm. Pharmacol. 49, 530–536.

Lin, J.H., Lu, A.Y., 1998. Inhibition and induction of cytochrome P450

and the clinical implications. Clin. Pharmacokinet. 35, 361–390.

Lindros, K.O., Cai, Y.A., Penttila, K.E., 1990. Role of ethanol-inducible

cytochrome P-450 IIE1 in carbon tetrachloride-induced damage to
centrilobular hepatocytes from ethanol-treated rats. Hepatology 12,
1092–1097.

Liu, T.T., Liang, N.S., Li, Y., Yang, F., Lu, Y., Meng, Z.Q., Zhang, L.S.,

2003. Effects of long-term tea polyphenols consumption on hepatic
microsomal drug-metabolizing enzymes and liver function in Wistar
rats. World J. Gastroenterol. 9, 2742–2744.

MacDonald, C.J., Ciolino, H.P., Yeh, G.C., 2001. Dibenzoylme-

thane modulates aryl hydrocarbon receptor function and expres-
sion of cytochromes P50 1A1, 1A2, and 1B1. Cancer Res. 61, 3919–
3924.

Machala, M., Kubinova, R., Horavova, P., Suchy, V., 2001. Chemopro-

tective potentials of homoisoflavonoids and chalcones of Dracaena
cinnabari: modulations of drug-metabolizing enzymes and antioxidant
activity. Phytother. Res. 15, 114–118.

Mackenzie, P.I., Owens, I.S., Burchell, B., Bock, K.W., Bairoch, A.,

Belanger, A., Fournel-Gigleux, S., Green, M., Hum, D.W., Iyanagi,
T., Lancet, D., Louisot, P., Magdalou, J., Chowdhury, J.R., Ritter,
J.K., Schachter, H., Tephly, T.R., Tipton, K.F., Nebert, D.W., 1997.
The UDP glycosyltransferase gene superfamily: recommended nomen-
clature update based on evolutionary divergence. Pharmacogenetics 7,
255–269.

Maliakal, P.P., Coville, P.F., Wanwimolruk, S., 2001. Tea consumption

modulates hepatic drug metabolizing enzymes in Wistar rats. J. Pharm.
Pharmacol. 53, 569–577.

Marchetti, F., De Santi, C., Vietri, M., Pietrabissa, A., Spisni, R., Mosca,

F., Pacifici, G.M., 2001. Differential inhibition of human liver and
duodenum sulphotransferase activities by quercetin, a flavonoid
present in vegetables, fruit and wine. Xenobiotica 31, 841–847.

Markowitz, J.S., DeVane, C.L., Boulton, D.W., Carson, S.W., Nahas, Z.,

Risch, S.C., 2000. Effect of St. JohnÕs wort (Hypericum perforatum) on
cytochrome P-450 2D6 and 3A4 activity in healthy volunteers. Life Sci.
66, PL133–PL139.

Markowitz, J.S., Donovan, J.L., DeVane, C.L., Taylor, R.M., Ruan, Y.,

Wang, J.S., Chavin, K.D., 2003. Effect of St JohnÕs wort on drug
metabolism by induction of cytochrome P450 3A4 enzyme. Jama 290,
1500–1504.

McLemore, T.L., Adelberg, S., Liu, M.C., McMahon, N.A., Yu, S.J.,

Hubbard, W.C., Czerwinski, M., Wood, T.G., Storeng, R., Lubet,
R.A., et al., 1990. Expression of CYP1A1 gene in patients with lung
cancer: evidence for cigarette smoke-induced gene expression in
normal lung tissue and for altered gene regulation in primary
pulmonary carcinomas. J. Natl. Cancer Inst. 82, 1333–1339.

McMahon, M., Itoh, K., Yamamoto, M., Chanas, S.A., Henderson, C.J.,

McLellan, L.I., Wolf, C.R., Cavin, C., Hayes, J.D., 2001. The
CapÕnÕCollar basic leucine zipper transcription factor Nrf2 (NF-E2
p45-related factor 2) controls both constitutive and inducible expres-
sion of intestinal detoxification and glutathione biosynthetic enzymes.
Cancer Res. 61, 3299–3307.

Meerman, J.H., Ringer, D.P., Coughtrie, M.W., Bamforth, K.J., Gilissen,

R.A., 1994. Sulfation of carcinogenic aromatic hydroxylamines and
hydroxamic acids by rat and human sulfotransferases: substrate
specificity, developmental aspects and sex differences. Chem. Biol.
Interact. 92, 321–328.

Miller, W.R., Anderson, T.J., Jack, W.J., 1990. Relationship between

tumour aromatase activity, tumour characteristics and response to
therapy. J. Steroid Biochem. Mol. Biol. 37, 1055–1059.

Miranda, C.L., Aponso, G.L., Stevens, J.F., Deinzer, M.L., Buhler, D.R.,

2000. Prenylated chalcones and flavanones as inducers of quinone
reductase in mouse Hepa 1c1c7 cells. Cancer Lett. 149, 21–29.

Montgomery, B.T., Young, C.Y., Bilhartz, D.L., Andrews, P.E., Prescott,

J.L., Thompson, N.F., Tindall, D.J., 1992. Hormonal regulation of
prostate-specific antigen glycoprotein in the human prostatic adeno-
carcinoma cell line, LNCaP. Prostate 21, 63–73.

Moore, L.B., Goodwin, B., Jones, S.A., Wisely, G.B., Serabjit-Singh, C.J.,

Willson, T.M., Collins, J.L., Kliewer, S.A., 2000. St. JohnÕs wort
induces hepatic drug metabolism through activation of the pregnane X
receptor. Proc. Natl. Acad. Sci. USA 97, 7500–7502.

Mukhtar, H., Wang, Z.Y., Katiyar, S.K., Agarwal, R., 1992. Tea

components: antimutagenic and anticarcinogenic effects. Prev. Med.
21, 351–360.

Nahrstedt, A., Butterweck, V., 1997. Biologically active and other

chemical constituents of the herb of Hypericum perforatum L.
Pharmacopsychiatry 30 (Suppl 2), 129–134.

208

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

background image

Nielsen, S.E., Breinholt, V., Cornett, C., Dragsted, L.O., 2000. Biotrans-

formation of the citrus flavone tangeretin in rats. Identification of
metabolites with intact flavane nucleus. Food Chem. Toxicol. 38, 739–
746.

Nijhoff, W.A., Bosboom, M.A., Smidt, M.H., Peters, W.H., 1995.

Enhancement of rat hepatic and gastrointestinal glutathione and
glutathione S-transferases by alpha-angelicalactone and flavone.
Carcinogenesis 16, 607–612.

Novak, R.F., Woodcroft, K.J., 2000. The alcohol-inducible form of

cytochrome P450 (CYP2E1): role in toxicology and regulation of
expression. Arch. Pharm. Res. 23, 267–282.

Obach, R.S, 2000. Inhibition of human cytochrome P450 enzymes by

constituents of St. JohnÕs Wort, an herbal preparation used in the
treatment of depression. J. Pharmacol. Exp. Ther. 294, 88–95.

Obermeier, M.T., White, R.E., Yang, C.S., 1995. Effects of bioflavonoids

on hepatic P450 activities. Xenobiotica 25, 575–584.

Ohkimoto, K., Liu, M.Y., Suiko, M., Sakakibara, Y., Liu, M.C., 2004.

Characterization of a zebrafish estrogen-sulfating cytosolic sulfotrans-
ferase: inhibitory effects and mechanism of action of phytoestrogens.
Chem. Biol. Interact. 147, 1–7.

Otake, Y., Nolan, A.L., Walle, U.K., Walle, T., 2000. Quercetin and

resveratrol potently reduce estrogen sulfotransferase activity in normal
human mammary epithelial cells. J. Steroid Biochem. Mol. Biol. 73,
265–270.

Pai, T.G., Suiko, M., Sakakibara, Y., Liu, M.C., 2001. Sulfation of

flavonoids and other phenolic dietary compounds by the human
cytosolic sulfotransferases. Biochem. Biophys. Res. Commun. 285,
1175–1179.

Pelissero, C., Lenczowski, M.J., Chinzi, D., Davail-Cuisset, B., Sumpter,

J.P., Fostier, A., 1996. Effects of flavonoids on aromatase activity, an
in vitro study. J. Steroid Biochem. Mol. Biol. 57, 215–223.

Peng, W.X., Li, H.D., Zhou, H.H., 2003. Effect of daidzein on CYP1A2

activity and pharmacokinetics of theophylline in healthy volunteers.
Eur. J. Clin. Pharmacol. 59, 237–241.

Peters, W.H., Kock, L., Nagengast, F.M., Kremers, P.G., 1991. Biotrans-

formation enzymes in human intestine: critical low levels in the colon?
Gut 32, 408–412.

Peters, W.H., Roelofs, H.M., Hectors, M.P., Nagengast, F.M., Jansen,

J.B., 1993. Glutathione and glutathione S-transferases in BarrettÕs
epithelium. Br. J. Cancer 67, 1413–1417.

Picq, M., Dubois, M., Prigent, A.F., Nemoz, G., Pacheco, H., 1989.

Inhibition of the different cyclic nucleotide phosphodiesterase isoforms
separated from rat brain by flavonoid compounds. Biochem. Int. 18,
47–57.

Piscitelli, S.C., Formentini, E., Burstein, A.H., Alfaro, R., Jagannatha, S.,

Falloon, J., 2002. Effect of milk thistle on the pharmacokinetics of
indinavir in healthy volunteers. Pharmacotherapy 22, 551–556.

Pouget, C., Fagnere, C., Basly, J.P., Besson, A.E., Champavier, Y.,

Habrioux, G., Chulia, A.J., 2002. Synthesis and aromatase inhibitory
activity of flavanones. Pharm. Res. 19, 286–291.

Ramos-Gomez, M., Kwak, M.K., Dolan, P.M., Itoh, K., Yamamoto, M.,

Talalay, P., Kensler, T.W., 2001. Sensitivity to carcinogenesis is
increased and chemoprotective efficacy of enzyme inducers is lost in
NRF2 transcription factor-deficient mice. Proc. Natl. Acad. Sci. USA
98, 3410–3415.

Raso, G.M., Meli, R., Di Carlo, G., Pacilio, M., Di Carlo, R., 2001.

Inhibition of inducible nitric oxide synthase and cyclooxygenase-2
expression by flavonoids in macrophage J774A.1. Life Sci. 68, 921–
931.

Rendic, S., Di Carlo, F.J., 1997. Human cytochrome P450 enzymes: a

status report summarizing their reactions, substrates, inducers, and
inhibitors. Drug Metab. Rev. 29, 413–580.

Roberts, D.W., Doerge, D.R., Churchwell, M.I., Gamboa da Costa, G.,

Marques, M.M., Tolleson, W.H., 2004. Inhibition of extrahepatic
human cytochromes P450 1A1 and 1B1 by metabolism of isoflavones
found in Trifolium pratense (red clover). J. Agric. Food. Chem. 52,
6623–6632.

Saarinen, N., Joshi, S.C., Ahotupa, M., Li, X., Ammala, J., Makela, S.,

Santti, R., 2001. No evidence for the in vivo activity of aromatase-
inhibiting flavonoids. J. Steroid Biochem. Mol. Biol. 78, 231–239.

Scalbert, A., Williamson, G., 2000. Dietary intake and bioavailability of

polyphenols. J. Nutr. 130, 2073S–2085S.

Schewe, T., Kuhn, H., Sies, H., 2002. Flavonoids of cocoa inhibit

recombinant human 5-lipoxygenase. J. Nutr. 132, 1825–1829.

Setchell, K.D., Brown, N.M., Zimmer-Nechemias, L., Brashear, W.T.,

Wolfe, B.E., Kirschner, A.S., Heubi, J.E., 2002. Evidence for lack of
absorption of soy isoflavone glycosides in humans, supporting the
crucial role of intestinal metabolism for bioavailability. Am. J. Clin.
Nutr. 76, 447–453.

Sharma, S., Sultana, S., 2004. Modulatory effect of soy isoflavones on

biochemical alterations mediated by TPA in mouse skin model. Food
Chem. Toxicol. 42, 1669–1675.

Sheu, S.Y., Lai, C.H., Chiang, H.C., 1998. Inhibition of xanthine oxidase

by purpurogallin and silymarin group. Anticancer Res. 18, 263–267.

Shih, H., Pickwell, G.V., Quattrochi, L.C., 2000. Differential effects of

flavonoid compounds on tumor promoter-induced activation of
the human CYP1A2 enhancer. Arch. Biochem. Biophys. 373, 287–
294.

Shukla, Y., Taneja, P., 2002. Anticarcinogenic effect of black tea on

pulmonary tumors in Swiss albino mice. Cancer Lett. 176, 137–
141.

Siegers, C.P., Bartels, L., Riemann, D., 1989. Effects of fasting and

glutathione depletors on the GSH-dependent enzyme system in the
gastrointestinal mucosa of the rat. Pharmacology. 38, 121–128.

Siess, M.H., Leclerc, J., Canivenc-Lavier, M.C., Rat, P., Suschetet, M.,

1995. Heterogenous effects of natural flavonoids on monooxygenase
activities in human and rat liver microsomes. Toxicol. Appl. Pharma-
col. 130, 73–78.

Siess, M.H., Mas, J.P., Canivenc-Lavier, M.C., Suschetet, M., 1996. Time

course of induction of rat hepatic drug-metabolizing enzyme activities
following dietary administration of flavonoids. J. Toxicol. Environ.
Health 49, 481–496.

Simanek, V., Walterova, D., Vicar, J., Urbanikova, J., Kren, V.,

Modriansky, M., Skottova, N., Ulrichova, J., 2001. Silymarin—an
extract from the milk thistle (Silybum marianum)—is it a drug or
nutritional supplement? Ceska Slov. Farm. 50, 66–69.

Sivaraman, L., Leatham, M.P., Yee, J., Wilkens, L.R., Lau, A.F., Le

Marchand, L., 1994. CYP1A1 genetic polymorphisms and in situ
colorectal cancer. Cancer Res. 54, 3692–3695.

Smith, T.J., Guo, Z., Guengerich, F.P., Yang, C.S., 1996. Metabolism of

4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone (NNK) by human
cytochrome P450 1A2 and its inhibition by phenethyl isothiocyanate.
Carcinogenesis 17, 809–813.

Sohn, O.S., Surace, A., Fiala, E.S., Richie Jr., J.P., Colosimo, S., Zang, E.,

Weisburger, J.H., 1994. Effects of green and black tea on hepatic
xenobiotic metabolizing systems in the male F344 rat. Xenobiotica 24,
119–127.

Soleas, G.J., Diamandis, E.P., Goldberg, D.M., 1997. Resveratrol: a

molecule whose time has come? and gone? Clin. Biochem. 30,
91–113.

Sun, X.Y., Plouzek, C.A., Henry, J.P., Wang, T.T., Phang, J.M., 1998.

Increased UDP-glucuronosyltransferase activity and decreased pros-
tate specific antigen production by biochanin A in prostate cancer
cells. Cancer Res. 58, 2379–2384.

Talalay, P., De Long, M.J., Prochaska, H.J., 1988. Identification of a

common chemical signal regulating the induction of enzymes that
protect against chemical carcinogenesis. Proc. Natl. Acad. Sci. USA.
85, 8261–8265.

Tamura, H., Matsui, M., 2000. Inhibitory effects of green tea and grape

juice on the phenol sulfotransferase activity of mouse intestines and
human colon carcinoma cell line, Caco-2. Biol. Pharm. Bull. 23, 695–
699.

Tanaka, T., Kawabata, K., Kakumoto, M., Makita, H., Ushida, J.,

Honjo, S., Hara, A., Tsuda, H., Mori, H., 1999. Modifying effects of a

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210

209

background image

flavonoid morin on azoxymethane-induced large bowel tumorigenesis
in rats. Carcinogenesis 20, 1477–1484.

Tsyrlov, I.B., Mikhailenko, V.M., Gelboin, H.V., 1994. Isozyme- and

species-specific susceptibility of cDNA-expressed CYP1A P-450s to
different flavonoids. Biochim. Biophys. Acta 1205, 325–335.

Tukey, R.H., Strassburg, C.P., 2000. Human UDP-glucuronosyltransfe-

rases: metabolism, expression, and disease. Ann. Rev. Pharmacol.
Toxicol. 40, 581–616.

Uda, Y., Price, K.R., Williamson, G., Rhodes, M.J., 1997. Induction

of the anticarcinogenic marker enzyme, quinone reductase, in
murine hepatoma cells in vitro by flavonoids. Cancer Lett. 120,
213–216.

Ueng, Y.F., Chang, Y.L., Oda, Y., Park, S.S., Liao, J.F., Lin, M.F., Chen,

C.F., 1999. In vitro and in vivo effects of naringin on cytochrome
P450-dependent monooxygenase in mouse liver. Life Sci. 65, 2591–
2602.

Ueng, Y.F., Kuwabara, T., Chun, Y.J., Guengerich, F.P., 1997. Coop-

erativity in oxidations catalyzed by cytochrome P450 3A4. Biochem-
istry 36, 370–381.

Umemoto, A., Komaki, K., Monden, Y., Suwa, M., Kanno, Y.,

Kitagawa, M., Suzuki, M., Lin, C.X., Ueyama, Y., Momen, M.A.,
Ravindernath, A., Shibutani, S., 2001. Identification and quantifica-
tion of tamoxifen-DNA adducts in the liver of rats and mice. Chem.
Res. Toxicol. 14, 1006–1013.

van der Logt, E.M., Roelofs, H.M., Nagengast, F.M., Peters, W.H., 2003.

Induction of rat hepatic and intestinal UDP-glucuronosyltransferases
by naturally occurring dietary anticarcinogens. Carcinogenesis 24,
1651–1656.

van Zanden, J.J., Ben Hamman, O., van Iersel, M.L., Boeren, S.,

Cnubben, N.H., Lo Bello, M., Vervoort, J., van Bladeren, P.J.,
Rietjens, I.M., 2003. Inhibition of human glutathione S-transferase P1-
1 by the flavonoid quercetin. Chem. Biol. Interact. 145, 139–148.

Venkataramanan, R., Ramachandran, V., Komoroski, B.J., Zhang, S.,

Schiff, P.L., Strom, S.C., 2000. Milk thistle, a herbal supplement,
decreases the activity of CYP3A4 and uridine diphosphoglucuronosyl
transferase in human hepatocyte cultures. Drug Metab. Dispos. 28,
1270–1273.

Veronese, M.L., Gillen, L.P., Burke, J.P., Dorval, E.P., Hauck, W.W.,

Pequignot, E., Waldman, S.A., Greenberg, H.E., 2003. Exposure-
dependent inhibition of intestinal and hepatic CYP3A4 in vivo by
grapefruit juice. J. Clin. Pharmacol. 43, 831–839.

Walle, T., Eaton, E.A., Walle, U.K., 1995. Quercetin, a potent and specific

inhibitor of the human P-form phenosulfotransferase. Biochem.
Pharmacol. 50, 731–734.

Walle, T., Otake, Y., Galijatovic, A., Ritter, J.K., Walle, U.K., 2000.

Induction of UDP-glucuronosyltransferase UGT1A1 by the flavonoid
chrysin in the human hepatoma cell line hep G2. Drug Metab. Dispos.
28, 1077–1082.

Walle, U.K., Walle, T., 2002. Induction of human UDP-glucuronosyl-

transferase UGT1A1 by flavonoids-structural requirements. Drug
Metab. Dispos. 30, 564–569.

Wang, C., Makela, T., Hase, T., Adlercreutz, H., Kurzer, M.S., 1994.

Lignans and flavonoids inhibit aromatase enzyme in human preadi-
pocytes. J. Steroid Biochem. Mol. Biol. 50, 205–212.

Wang, H., Wang, Y., Chen, Z.Y., Chan, F.L., Leung, L.K., 2005.

Hydroxychalcones exhibit differential effects on XRE transactivation.
Toxicology 207, 303–313.

Wang, W., Liu, L.Q., Higuchi, C.M., Chen, H., 1998. Induction of

NADPH: quinone reductase by dietary phytoestrogens in colonic
Colo205 cells. Biochem. Pharmacol. 56, 189–195.

Wang, Z., Gorski, J.C., Hamman, M.A., Huang, S.M., Lesko, L.J., Hall,

S.D., 2001. The effects of St JohnÕs wort (Hypericum perforatum)
on human cytochrome P450 activity. Clin. Pharmacol. Ther. 70, 317–
326.

Wang, Z.Y., Das, M., Bickers, D.R., Mukhtar, H., 1988. Interaction of

epicatechins derived from green tea with rat hepatic cytochrome P-450.
Drug Metab. Dispos. 16, 98–103.

Wattenberg, L.W., Page, M.A., Leong, J.L., 1968. Induction of increased

benzpyrene hydroxylase activity by flavones and related compounds.
Cancer Res. 28, 934–937.

Weber, A., Jager, R., Borner, A., Klinger, G., Vollanth, R., Matthey, K.,

Balogh, A., 1996. Can grapefruit juice influence ethinylestradiol
bioavailability? Contraception 53, 41–47.

Wen, X., Walle, U.K., Walle, T., 2005. 5,7-Dimethoxyflavone downreg-

ulates CYP1A1 expression and benzo[a]pyrene-induced DNA binding
in Hep G2 cells. Carcinogenesis 26, 803–809.

Wentworth, J.M., Agostini, M., Love, J., Schwabe, J.W., Chatterjee,

V.K., 2000. St JohnÕs wort, a herbal antidepressant, activates the
steroid X receptor. J. Endocrinol. 166, R11–R16.

Williams, J.A., Ring, B.J., Cantrell, V.E., Campanale, K., Jones, D.R.,

Hall, S.D., Wrighton, S.A., 2002. Differential modulation of UDP-
glucuronosyltransferase 1A1 (UGT1A1)-catalyzed estradiol-3-glucu-
ronidation by the addition of UGT1A1 substrates and other
compounds to human liver microsomes. Drug Metab. Dispos. 30,
1266–1273.

Wong, C.K., Keung, W.M., 1997. Daidzein sulfoconjugates are potent

inhibitors of sterol sulfatase (EC 3.1.6.2). Biochem. Biophys. Res.
Commun. 233, 579–583.

Wood, A.W., Smith, D.S., Chang, R.L., Huang, M.T., Conney, A.H.,

1986. Effects of flavonoids on the metabolism of xenobiotics. Prog.
Clin. Biol. Res. 213, 195–210.

Yamazoe, Y., Nagata, K., Yoshinari, K., Fujita, K., Shiraga, T., Iwasaki,

K., 1999. Sulfotransferase catalyzing sulfation of heterocyclic amines.
Cancer Lett. 143, 103–107.

Yang, G.Y., Liu, Z., Seril, D.N., Liao, J., Ding, W., Kim, S., Bondoc, F.,

Yang, C.S., 1997. Black tea constituents, theaflavins, inhibit 4-
(methylnitrosamino)-1-(3-pyridyl)-1-butanone (NNK)-induced lung
tumorigenesis in A/J mice. Carcinogenesis 18, 2361–2365.

Yannai, S., Day, A.J., Williamson, G., Rhodes, M.J., 1998. Character-

ization of flavonoids as monofunctional or bifunctional inducers of
quinone reductase in murine hepatoma cell lines. Food Chem. Toxicol.
36, 623–630.

Yee, G.C., Stanley, D.L., Pessa, L.J., Dalla Costa, T., Beltz, S.E., Ruiz, J.,

Lowenthal, D.T., 1995. Effect of grapefruit juice on blood cyclosporin
concentration. Lancet 345, 955–956.

Zhai, S., Dai, R., Friedman, F.K., Vestal, R.E., 1998. Comparative

inhibition of human cytochromes P450 1A1 and 1A2 by flavonoids.
Drug Metab. Dispos. 26, 989–992.

Zhang, Y., Benet, L.Z., 2001. The gut as a barrier to drug absorption:

combined role of cytochrome P450 3A and P-glycoprotein. Clin.
Pharmacokinet. 40, 159–168.

Zhao, J., Agarwal, R., 1999. Tissue distribution of silibinin, the major

active constituent of silymarin, in mice and its association with
enhancement of phase II enzymes: implications in cancer chemopre-
vention. Carcinogenesis 20, 2101–2108.

Zhou, C., Zhou, D., Esteban, J., Murai, J., Siiteri, P.K., Wilczynski, S.,

Chen, S., 1996. Aromatase gene expression and its exon I usage in
human breast tumors. Detection of aromatase messenger RNA by
reverse transcription-polymerase chain reaction. J. Steroid Biochem.
Mol. Biol. 59, 163–171.

Zhou, S., Gao, Y., Jiang, W., Huang, M., Xu, A., Paxton, J.W., 2003.

Interactions of herbs with cytochrome P450. Drug Metab. Rev. 35,
35–98.

Zhou, S., Koh, H.L., Gao, Y., Gong, Z.Y., Lee, E.J., 2004. Herbal

bioactivation: the good, the bad and the ugly. Life Sci. 74, 935–968.

Zhu, B.T., Conney, A.H., 1998. Functional role of estrogen metabolism in

target cells: review and perspectives. Carcinogenesis 19, 1–27.

Zou, L., Harkey, M.R., Henderson, G.L., 2002. Effects of herbal

components on cDNA-expressed cytochrome P450 enzyme catalytic
activity. Life Sci. 71, 1579–1589.

Zuber, R., Modriansky, M., Dvorak, Z., Rohovsky, P., Ulrichova, J.,

Simanek, V., Anzenbacher, P., 2002. Effect of silybin and its congeners
on human liver microsomal cytochrome P450 activities. Phytother.
Res. 16, 632–638.

210

Y.J. Moon et al. / Toxicology in Vitro 20 (2006) 187–210


Document Outline


Wyszukiwarka

Podobne podstrony:
Chung, Zhao Humor effect on memory and attitude moderating role of product involvement
Caffeine effect on mortality and oviposition in successive of Aedes aegypti
53 755 765 Effect of Microstructural Homogenity on Mechanical and Thermal Fatique
Changes in passive ankle stiffness and its effects on gait function in
31 411 423 Effect of EAF and ESR Technologies on the Yield of Alloying Elements
Effects of the Great?pression on the U S and the World
Heavy metal toxicity,effect on plant growth and metal uptake
Possible Effects of Strategy Instruction on L1 and L2 Reading
33 437 452 Primary Carbides in Spincast HSS for Hot Rolls and Effect on Oxidation
Effect of Drugs and Alcohol on Teenagers
Effect of heat treatment on microstructure and mechanical properties of cold rolled C Mn Si TRIP
48 671 684 Cryogenic Treatment and it's Effect on Tool Steel
Genetic and environmental effects on polyphenols
53 755 765 Effect of Microstructural Homogenity on Mechanical and Thermal Fatique
Changes in passive ankle stiffness and its effects on gait function in
EFFECTS OF EATING AND NOT EATING ON ENERGY STORES AND BODY WEIGHT

więcej podobnych podstron