thuoc từ curcumin

176
"rug Biotransformation En0yme Interactions Studies with 9urcumin, 9urcumin analogues and other Plant=derived 9omponents !egina (ppiah+,pong

Upload: r0ke1984

Post on 01-Dec-2014

95 views

Category:

Documents


3 download

TRANSCRIPT

Page 1: thuoc từ curcumin

"rug Biotransformation En0yme Interactions

Studies with 9urcumin, 9urcumin analogues and other Plant=derived 9omponents

!egina (ppiah+,pong

Page 2: thuoc từ curcumin
Page 3: thuoc từ curcumin

Drug Biotransformation Enzyme Interactions

Studies with Curcumin, Curcumin analogues and other Plant-derived Components

Regina Appiah-Opong

Page 4: thuoc từ curcumin

Drug Biotransformation Enzyme Interactions: Studies with Curcumin, Curcumin analogues and other Plant-derived Components Regina Appiah-Opong KNAW and GETFUND scholarship schemes are gratefully acknowledged for the financial support in the printing of this thesis. Printed by Printpartners Ipskamp © Regina Appiah-Opong, Legon, Ghana 2009. All rights reserved. No part of this thesis may be reproduced in any form or by any means without permission from the author.

Page 5: thuoc từ curcumin

VRIJE UNIVERSITEIT

Drug Biotransformation Enzyme Interactions Studies with Curcumin, Curcumin analogues and

other Plant-derived Components

ACADEMISCH PROEFSCHRIFT

ter verkrijging van de graad Doctor aan de Vrije Universiteit Amsterdam,

op gezag van de rector magnificus prof.dr. L.M. Bouter,

in het openbaar te verdedigen ten overstaan van de promotiecommissie

van de faculteit der Exacte Wetenschappen op woensdag 6 mei 2009 om 13.45 uur

in de aula van de universiteit, De Boelelaan 1105

door

Regina Appiah-Opong

geboren te Kumasi, Ghana

Page 6: thuoc từ curcumin

promotor : prof.dr. N.P.E. Vermeulen copromotor : dr. J.N.M. Commandeur

Page 7: thuoc từ curcumin

Reading Committee: prof.dr. Ivonne M.C.M. Rietjens prof.dr. Rob Leurs dr. Iwan J.P. de Esch dr. Chris Oostenbrink dr. Martijn Rooseboom

The investigations described in this thesis were carried out in the Leiden Amsterdam Center for Drug Research (LACDR)/Division of Molecular Toxicology, Department of Chemistry and Pharmaceutical Science, Faculty of Sciences, Vrije Universiteit, De Boelelaan 1083, 1081 HV Amsterdam, The Netherlands.

Page 8: thuoc từ curcumin
Page 9: thuoc từ curcumin

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

!

By wisdom the Lord laid the earths foundation, by understanding He set the Heavens in place; !

by His knowledge the deeps were divided and the clouds let drop the dew. Prov. 3: 19, 20

Page 10: thuoc từ curcumin

Contents Introduction ! Chapter !* +eneral introduction 3 Chapter 4* Curcumin* pharmacokinetics, metabolism, and potential for drug-

drug=food Interactions ?! Inhibition of C1P3GST activities by natural products and derivatives @3 Chapter 3* Inhibition of human recombinant cytochrome B?CDs by curcumin

and curcumin decomposition products @C Chapter ?* Structure-activity relationship of inhibition of recombinant human

cytochrome B?CD mediated metabolism by curcumin analogues G3 Chapter C* Inhibition of human glutathione S-transferases by curcumin and

analogues !DH Chapter @* Interactions between cytochromes B?CD, glutathione S-

transferases and +hanaian medicinal plants! !4J Summary, conclusions and perspectives Chapter J* Summary, conclusions and perspectives !?C Kppendices* Lutch summary !CH Mist of publications !@3 Epilogue !@C Mist of abbreviations !@@

Page 11: thuoc từ curcumin

1

"#$%&'()$*&#

Page 12: thuoc từ curcumin

2

!

Page 13: thuoc từ curcumin

!"

Chapter 1

General Introduction

Drug disposition/Biotransformation

#$%$&'" ()'*&$+,+" *)-" .(&+/*&/01" -23(+-4" /(" 2-&(5$(/$.+6" 70$,$&*/$(&" (8" /9-+-"

2-&(5$(/$.+:" $+" +$'&$8$.*&/01" -&9*&.-4" 51" /9-" 3)(.-++" ;&(<&" *+" 5$(/)*&+8(),*/$(&:"

<9-)-51" /9-" 0$3(39$0$."+=5+/*&.-" $+".(&%-)/-4" $&/("*",()-"914)(39$0$." 8()," /9*/" 8*%()+"

-2.)-/$(&" $&"=)$&-:"5$0-"()" 8*-.-+6">9-"5$(/)*&+8(),*/$(&"(8".-)/*$&"4)='+" )-+=0/+" $&" /9-"

8(),*/$(&"(8"*",-/*5(0$/-"/9*/"$+",()-"39*),*.(0('$.*001"*./$%-"*&4"0-++"/(2$."/9*&"/9-"

3*)-&/" 4)='" ?@A6" B$(/)*&+8(),*/$(&" ,*1" *0+(" )-+=0/" $&" /9-" 8(),*/$(&" (8" )-*./$%-"

,-/*5(0$/-+"*&4"+=5+-C=-&/01"*4%-)+-"-88-./+" ?D:!A6">9-" 0$%-)" $+" /9-",*E()"+$/-"(8"4)='"

5$(/)*&+8(),*/$(&:"*0/9(='9"5$(/)*&+8(),*/$(&"*0+("(..=)+" $&"(/9-)"()'*&+"*&4"/$++=-+6"

F-&(5$(/$." 5$(/)*&+8(),*/$(&" )-*./$(&+" *)-" '-&-)*001" 4$%$4-4" $&/(" /9)--" ')(=3+:" .*00-4"

39*+-" G:" 39*+-" GG" *&4" 39*+-" GGG6" H9*+-" G" )-*./$(&+" $&%(0%-" (2$4*/$(&:" )-4=./$(&" *&4"

914)(01+$+:" <9-)-" *" 8=&./$(&*0" ')(=3" $+" -23(+-4" ()" $&/)(4=.-4:" =+=*001" )-+=0/$&'" $&" *"

+,*00"$&.)-*+-"$&"2-&(5$(/$."914)(39$0$.$/16"I&"/9-"(/9-)"9*&4:"39*+-"GG")-*./$(&+"$&%(0%-"

.(&E='*/$(&" )-*./$(&+" $&.0=4$&':" .(&E='*/$(&" <$/9" '0=/*/9$(&-" *&4" (/9-)" *,$&(" *.$4+:"

'0=.=)(&$4*/$(&:"+=08*/$(&:"*.-/10*/$(&"*&4",-/910*/$(&6">9-+-")-*./$(&+"=+=*001")-+=0/"$&"

+/)(&'-)" $&.)-*+-+" $&" 914)(39$0$.$/1" )-C=$)-4" 8()" -0$,$&*/$(&" (8" +=5+/*&.-+6" H9*+-" GGG"

)-*./$(&+" $&%(0%-" -23()/$&'" (8" 3)(4=./+" (8" 39*+-" G" *&4" 39*+-" GG" )-*./$(&+" 51" %*)$(=+"

/)*&+3()/-)+" +=.9" *+" HJ'01.(3)(/-$&" KHJ'3L:" ()'*&$." *&$(&" /)*&+3()/$&'" 3(013-3/$4-+"

KIM>H+L"*&4",=0/$4)='")-+$+/*&.-"3)(/-$&+"KNOH+L"?PA6""

" N(4=0*/(&" (8" ,*E()" 5$(/)*&+8(),*/$(&" -&Q1,-+" +=.9" *+" .1/(.9)(,-" HPRS+"

KTUH+L"51"4)='+:"8((4".(,3(&-&/+"*&4"9-)5*0"3)(4=./+"$&"/9-"3)-+-&.-"(8"+(,-"4)='+"

.(=04" )-+=0/" $&"4)='J4)='V8((4V9-)5" $&/-)*./$(&+"<$/9"*4%-)+-"+$4-"-88-./+" ?RJWA6"I&" /9-"

(/9-)" 9*&4",(4=0*/$(&" (8" '0=/*/9$(&-" XJ/)*&+8-)*+-+" 51" /9-+-" 2-&(5$(/$.+" .(=04" 9*%-"

/(2$.(0('$.*0" .(&+-C=-&.-+:" 4=-" /(" /9-" ,*E()" )(0-" 30*1-4" 51" /9-+-" -&Q1,-+" $&"

4-/(2$8$.*/$(&6" >9=+:" /9-)-" $+" /9-" &--4" 8()" -%*0=*/$(&" (8" 8((4" .(,3(&-&/+" *&4" 9-)5*0"

3)(4=./+"8()"/9-"3(/-&/$*0"8()"4)='J8((4V9-)5"$&/-)*./$(&+:"/("4-.)-*+-"/9-")$+;"(8"9*),8=0"

+$4-"-88-./+6">*50-"@"+9(<+" /9-" /9)--"39*+-+"(8" /9-"3)(.-++"(8"5$(/)*&+8(),*/$(&:" /9-"

-&Q1,-+"*&4".9-,$.*0")-*./$(&+"$&%(0%-46"

"

Page 14: thuoc từ curcumin

4

Table 1. Biotransformation enzymes and reactions involved in the biotransformation

Phase Enzyme Type of reaction

I Cytochrome P450

Oxidation, epoxidation,

reduction, dealkylation,

dehalogenation

Monoamine oxidase Oxidation

Flavin containing mono-oxygenase Oxidation

NAD(P)H quinine oxidoreductase Reduction

Peroxidase Peroxidation

Alcohol dehydrogenase Oxidation

Aldehyde dehydrogenase Oxidation

II Glutathione S-transferase Conjugation

UDP-glucuronyl transferase Conjugation

Sulfotransferase Conjugation

Amino acid transferase Conjugation

N-acetyl transferase Acetylation

Methyl transferase Methylation

Epoxide hydrolase Hydration

III MRP/Pgp Transportation

OATP Transportation

N-acetyl transferase Acetylation, formation of

mercapturic acids

!-lyase !-elimination

Amino acid oxidases Deamination

Adapted from references [1] and [4].

Drug-drug/herb/food interactions

The effects of a particular drug on the efficacy and/or toxicity of another drug, which is

commonly referred to as drug-drug interactions have become an important healthcare

issue [8]. Drug-drug interaction is a phenomenon that can elude both regulatory

agencies and pharmaceutical companies. A drug that apparently is safe after extensive

evaluation in both preclinical and clinical trials can be extremely harmful when co-

administered with another drug. A typical example is the interaction between the non-

sedative antihistamine drug, terfenadine and ketoconazole, an antifungal drug [5]. The

Page 15: thuoc từ curcumin

5

co-administration of these drugs resulted in cardiotoxicity, and subsequently death of

some patients. Ketoconazole is a potent inhibitor of the CYP3A4 metabolism of

terfenadine, consequently co-administration of these two drugs leads to an increase in

plasma concentration of terfenadine to a cardio toxic level. Thus, drug-drug interactions,

remain important drug properties that should be well defined before humans are

exposed to any new drug.

Natural products such as foods and herbal products may also possess the

potential to cause harmful effects upon concomitant intake with drugs. Food-drug

interactions that have been investigated include that of grapefruit juice, orange juice,

broccoli, cabbage, Brussels sprouts, charcoal-grilled meats and garlic (Tables 2 and 3).

Grapefruit juice, is a potent inhibitor of CYP3A4-mediated drug metabolism [9].

CYP3A4-mediated metabolism of several drugs has been shown to be affected by

grapefruit juice. Clinically relevant interactions with grapefruit juice seem likely for most

dihydropyridines, terfenadine, saquinavir, cyclosporine, midazolam, triazolam and

verapamil and may also occur with lovastatin, cisapride and astemizole [9]. Grapefruit

and other fruit juices have also been shown to be potent in vitro inhibitors of a number

of organic anion-transporting polypeptides (OATPs). These juices were also found to

decrease the absorption of the non-metabolized OATP substrate, fexofenadine, hence

decrease drug bioavailability. Such findings enhance our understanding of the complex

nature of food-drug interactions, their possible influence on the clinical effects of

medications, and also underscore the need for similar investigations on other foods.

Herbal medicines are usually mixtures of more than one active ingredient; hence

the likelihood of herbal interactions is theoretically higher than drug-drug interactions.

Herb-drug interactions are a major concern, especially due to the potential to cause

adverse effects. These herbal products may modulate drug metabolizing enzymes or

transporters (e.g. CYPs or P-glycoprotein (Pgp)). Herbal medicines such as St. John’s

wort, ginseng and gingko, which are freely available over the counter, have been

reported to cause serious clinical interactions when co-administered with prescribed

medicines [8]. St. John’s wort, which is popularly used as an anti-depressant, has been

implicated in a potentially fatal interaction with cyclosporin. There is evidence that the

mechanism of interaction involves induction of CYP3A4 through the activation of

Page 16: thuoc từ curcumin

6

pregnane X-receptor, as well as induction of intestinal multidrug resistance transporter

protein MDR1/Pgp [10,11]. Thus, drug-drug/food/herb interactions may occur at the

level of receptors, transporters and/or biotransformation enzymes.

Drug-drug/food/herb interactions associated with transporters and receptors

Drug-drug interactions involving inhibition or induction of transport proteins have been

demonstrated [11,12]. Drug transporters play a significant role in absorption,

distribution, metabolism and excretion of drugs. Several drug transporter proteins have

been identified, however, the interactions mediated by P-gp are the most widely studied

and best understood. Approximately 50% of marketed drugs have been identified to be

Pgp substrates and/or inhibitors and it has broad substrate specificity. Drug transporters

and metabolizing enzymes may also act synergistically to cause drug-drug interaction.

For instance, in cardiovascular treatment, drug interactions result when therapeutic

doses of verapamil inhibit Pgp and CYP3A4, causing a marked increase in cyclosporine

absorption and availability [13]. The transporters are also vulnerable to inhibition,

induction and activation by foods and herbal products. Thus, drug-food/herb interactions

may result upon concomitant intake of some food and herbal products with drugs.

Curcumin, ginsenosides, piperine, some catechins from green tea, and silymarin from

milk thistle were found to be inhibitors of Pgp, while St. John's wort induced the

intestinal expression of Pgp in vitro and in vivo [14]. Some components (e.g.,

bergamottin and quercetin) of grapefruit juice were also reported to modulate Pgp

activity. Herbal constituents, in particular flavonoids, have been reported to modulate

Pgp by directly interacting with the vicinal ATP-binding site, the steroid-binding site, or

the substrate-binding site.

Drug receptors have also been shown to mediate drug-drug interactions. The

nuclear pregnane X receptor (PXR) and constitutive androstane receptor (CAR) have

now been discovered and their roles in rifampicin-mediated drug-drug interactions

demonstrated [15]. Rifampicin activates the nuclear PXR that in turn affects

cytochromes P450, glucuronosyltransferases and Pgp activities [16]. Some herbal

constituents (e.g., hyperforin and kava) were also shown to activate PXR [14]. Hence,

investigations on the effects of drugs, foods and herbal products on PXR are important

Page 17: thuoc từ curcumin

7

since co-administration of drugs with foods, herbal products or drugs that are PXR

modulators could result in harmful drug-drug/food/herb interactions.

Biotransformation enzymes and drug-drug interactions

Many known pharmacokinetic drug interactions are associated with phase I

biotransformation enzymes, particularly CYP enzymes. Pharmacokinetic CYP-mediated

drug interactions, one of the major causes of attritions in drug development, involves

induction and inhibition of the CYP enzymes with the latter being more common [17,18].

Drug-drug interactions involving CYPs have also been identified as important causes of

adverse drug reactions and therapeutic failure [19]. Intake of multiple drugs increases

the chances of drug-drug interactions and adverse reactions in patients.

Table 2. Human drug-drug/food interactions mediated by CYP inhibition

CYP inhibitor drug/substance involved

Inhibited CYP

Known or possible effect

Ref

Drug

Ketoconazole, erythromycin

Terfenadine, cyclosporine, tolbutamide

CYP3A4

Renal toxicity, cardiotoxicity

20,21,22

Quinidine, codeine amiadarone, haloperidol,

Fluoxetine CYP2D6 Anorexia, nervousness, tremor, seizures

23

Isoniazid Phenytoin, carbamazepine, diazepam

CYP3A4, CYP2C19

Adversities due to slow elimination of the drugs

24

Fluvoxamine Caffeine CYP1A2 Mental disorder 25,26

Fluconazole, ketoconazole

S-warfarin, phenytoin CYP2C9 Increased anticoagulation

27

Food

Grapefruit, Seville orange juice

Felodipine, terfenadine, saquinavir, cyclosporine, midazolam, triazolam, verapamil lovastatin, cisapride, astemizole

CYP3A4, CYP1A2

Renal toxicity and other toxicities

7,22,23,2428

Table 2 was partly adapted [8].

Page 18: thuoc từ curcumin

8

In contrast to the relatively narrow substrate specificity characteristics of other enzymes,

such as epoxidases, most drug-metabolizing CYP enzymes exhibit broad substrate

specificity [5]. Examples of CYP-mediated drug interactions and possible adverse

effects are shown in Tables 2 and 3.

Table 3. Drug interactions mediated by CYP induction

CYP inducer Drug/substance involved

Induced CYP

Known or possible effect

Ref

Drug

Rifampicin,Phenobarbital, phenytoin

Ethinylestradiol

CYP3A4

Ineffective contraceptive

29

Nicotine, omeprazole

Clozapine CYP1A2 Increased clearance of the drug leading to loss of antipsychotic properties

29,30

Food

Cruciferous vegetables (broccoli, cabbage, Brussels, sprouts) charcoal-grilled meats

Theophylline, warfarin, clozapine

CYP1A2 Increased plasma clearance leading to possible loss of control of asthma, reduced anticoagulation

23

Herb

St John’s wort Cyclosporine CYP3A4 Rejection of organ transplant

31

Ginkgo Depakote, Dilantin CYP2C9 Epileptic seizures 32

Table 3 was partly adapted [8].

Thus, inhibition of CYP enzymes could result in accumulation of drugs, and

subsequently lead to serious clinically important drug interactions [33]. Serious toxicity

may develop shortly if the drug has a narrow therapeutic window. A drug can be both a

substrate and an inhibitor for a particular isoenzyme. A non-substrate drug can also

inhibit the activity of an isoenzyme. Regardless of the mechanism, CYP inhibition may

result in decreased metabolism of a drug and alter its pharmacokinetic profile. In vitro

Page 19: thuoc từ curcumin

9

CYP-associated metabolic studies have been considered cost-effective for predicting

the potential for clinical drug-drug interactions [17]. Although in vitro findings on drug

interactions may not always correlate with in vivo situations, in vitro findings remain

useful and rapid indicators of potentially harmful drug interactions.

The observed induction and inhibition of CYP enzymes by natural products in the

presence of a prescribed drug has led to the general acceptance that natural products

can have adverse effects, contrary to the popular beliefs of their safety, especially in

countries where there is an active practice of ethnomedicine [8]. Hence, it is imperative

that foods and herbal products be assessed for their potential to cause harmful drug-

food/herb interactions.

CYP inhibition

Enzyme inhibition refers to the reduction in enzyme activity due to the presence of an

inhibitor. CYP inhibition can be reversible, comprising of competitive, non-competitive,

uncompetitive or irreversible inhibition. Irreversible inhibition usually results from

activation of a drug by CYPs into a reactive metabolite, which covalently binds to the

active site of the enzyme causing its inactivation, a process known as suicide, time-

dependent or mechanism-based inhibition (TDI or MBI) [34]. The level of accumulation

and the therapeutic window of a drug are important determinants of the clinical

relevance of a specific drug-drug interaction. CYP inhibition with clinical relevance

includes competitive inhibition and MBI [17]. Figure 1 shows the effect of grapefruit juice

on the CYP3A4-mediated metabolism of simvastatin, an inactive lactone prodrug [35].

Area analysis of figure 1, indicates that the bioavailability of simvastatin increased 13.5

fold (compared to the control, water) in the presence of grapefruit juice [35]. The

dramatic increase in simvastatin plasma concentration has been attributed to the

inhibition of CYP3A4 by grapefruit juice. The exposure to unmetabolized parent drug,

simvastatin, could result in toxicity as observed in the interaction between itraconazole,

a CYP3A4 inhibitor and simvastatin, that is an association with increased risk of skeletal

muscle toxicity [36].

Page 20: thuoc từ curcumin

10

Figure 1. Mean serum concentrations of simvastatin in 10 healthy volunteers after single oral doses of 40

mg simvastatin. Simvastatin was taken with 200 ml water (open circles), with 200 ml double strength

grapefruit juice after injestion of 200 ml grapefruit juice 3X daily for 2 days (solid triangles) or with 200 ml

water 24 h (solid diamonds), 3 days (open triangles) or 7 days (solid stars) after last dose of grapefruit

juice (adapted from ref. 35).

The mechanism-based inhibition (MBI) has recently received increasing focus,

with the notion that it could occur more frequently than anticipated, partly due to the

redox cycling-allied enzymatic action of CYPs [17]. Figure 2 shows a scheme of types of

enzyme inhibition [37].

Figure 2. Competitive inhibition (A); noncompetitive inhibition (B); mixed-type of inhibition (C); irreversible inhibition (D); E, enzyme; S, substrate; P, product; I, inhibitor; EI, enzyme-inhibitor complex; EI*, ‘dead-end’ enzyme inhibitor complex; ESI, enzyme-substrate-inhibitor complex; Ki, inhibitor constant; kinact, inactivation rate

A B

C D

E + S ES1

P1 + E

ESIEI

+ IKi'+ IK

i

E + S ES1

P1 + E

ESI

+ IKi

E + S ES1

P1 + E

EI

+ IKi

E + S ES1

P1 + E

ESIEI

EI*

+ IKi'+ IK

i

kinact

C D

Page 21: thuoc từ curcumin

11

CYP inhibition studies are extremely valuable, as they allow extrapolation of data to

other compounds, facilitate drug development and also indicate possible drug

interactions in organs other than the liver. Studies on prediction of in vivo drug-drug

interactions via inhibition CYP-mediated metabolism from in vitro data have been

performed [38]. Equation 1 can be used for quantitative prediction of in vivo area under

the concentration versus time curve (AUC) ratio, when the fraction of substrate

metabolized by the inhibited CYP pathway (fmCYP value) and the maximum hepatic

inhibitor concentration are known. Values for fmCYP may be estimated by assessing

exposure differences between extensive and poor metabolizers using probe substrates

[39]. Alternatively, fmCYP values for probe substrates could be obtained by calculating the

difference between the urinary recovery of metabolites in both the presence and

absence of the CYP selective inhibitor. The fmCYP values could also be estimated by the

combination of urinary recovery of metabolites, biliary excretion and the recovery of

unchanged drug [38].

AUCinhibitor = 1 (Equation 1) AUCcontrol fmCYP + (1-fmCYP) 1+[I]/Ki

Where, [I] is maximum hepatic inhibitor concentration and Ki is inhibition constant.

This prediction is possible when the other CYP pathways are not subject to inhibition

[40]. The greatest uncertainty in predicting the magnitude of in vivo drug interactions

resides in the values used for [I] [41]. The most appropriate value would be the

concentration available to the enzyme in the liver, but this value cannot be determined

in vivo. Possible representative in vivo concentrations that could be used include the

free or total systemic concentrations or the free or total hepatic inlet concentrations

estimated to occur during the absorption phase after oral administration [41].

Investigations on the use of the maximum hepatic inhibitor concentration at the inlet to

the liver ([I]in) have been performed. Calculation of this parameter (Equation 2) relies on

information on hepatic blood flow (QH), inhibitor dose (D), fraction absorbed from the

gastrointestinal tract (fa), the absorption rate constant (ka) and the systemic plasma

concentration ([I]av) [38].

Page 22: thuoc từ curcumin

12

[I]in = [I]av + ka.fa.D Equation 2 QH

In vivo clinical studies do not usually report ka values, but to avoid false-negative

prediction and obtain the largest [I]in, a maximum ka of 0.1 min-1 has been suggested to

be appropriate, assuming the gastric emptying is the rate limiting step for absorption.

The ka value for each inhibitor could be calculated, using the time to reach maximum

plasma concentration (Tmax) and the elimination rate constant (k) as shown in equation

3.

Tmax = ln(ka / k) Equation 3

(ka – k)

The use of fmCYP and ka values to refine estimates of [I]in has been shown to provide a

useful estimate of [I] and successful predictions [38]. However, it is possible that active

hepatic uptake processes could complicate the use of unbound plasma concentrations

for the inhibitor. This remains a limitation for application of these principles for the

prediction of drug-drug interactions.

The availability of specific probe substrates, human liver tissue and cDNA-

expressed CYP enzymes are valuable tools for the in vitro assessment of the potential

for varied xenobiotics to inhibit CYP isoenzymes [42]. It is appropriate to use tissue from

individual human donors for inhibition studies, if the enzyme activity is sufficiently

present. On the other hand recombinant CYP isoenzymes may be used for investigation

of specific isoenzymes. Human liver microsomes and recombinant CYP isoenzymes are

preferable test systems since they are more readily available than hepatocytes. The

effects of human liver microsomes and recombinant CYP on kinetic measurements are

not confounded by other metabolic processes or cellular uptake that would occur with

the use of hepatocytes [42]. However, the disadvantage of human liver microsomes and

recombinant isoenzymes is that they do not represent the true physiological milieu. In

addition, clinical relevance of the acquired data also has to be established.

Page 23: thuoc từ curcumin

13

Table 4. CYP substrate properties, substrates, reactions and interacting drug/food/herb [8,43,44]

CYP Substrates properties

Drug/Substrates Reaction Interacting drug /food/herb*,**

CYP1A2 Planar, polyaromatic compounds

Clozapine Phenacetin Olanzapine

Imipramine Propranolol Theophylline

Oxidation O-deethylation Oxidation

Demethylation 4-hydroxylation Hydroxylation

Cimetidine inh* Ciprofloxacin inh Fluvoxamine inh

Cruciferous vegetables ind** Cigarette smoke ind

CYP2B6 Cyclophosphamide Efavirenz Nevirapine Bupropion Artemisinin

Methadone Profofol

Hydroxylation Hydroxylation 3-ydroxylation Hydroxylation Unclear

N-demethylation Oxidation

Orphenadrine inh Ticlopidine inh Troleandomycin inh Ketoconazole inh Clopidogrel inh

Phenytoin ind Rifampin ind

CYP2C9 Polar, weakly acidic compounds

Diclofenac Flurbiprofen Ibuprofen Naproxen Phenytoin

Piroxicam Tolbutamide Warfarin

4-Hydroxylation 4-hydroxylation Oxidation O-demethylation 4-hydroxylation

5-hydroxylation Methyl hydroxylation 7-hydroxylation

Amiodarone inh Fluconazole inh Fluoxetine inh Isoniazid inh Ticlopidine inh

Rifampin ind Kava inh

CYP2C19 Polar acidic compounds

Amitriptyline Cyclophosphamide Diazepam Imipramine Omeprazole

Phenytoin

Demethylation Oxidation Demethylation N-demethylation Demethylation

4-hydroxylation

Ticlopidine inh Cimetidine inh Fluconazole inh Fluoxetine inh Carbamazepine ind

Norethindrone ind

CYP2D6 Lipophilic, basic, medium sized, positively charged at neutral pH

Amitriptyline Imipramine Propranolol Codeine Dextromethorphan Desipramine

Bufuralol

Benzyl hydroxylation 2-hydroxylation 4-hydroxylation O-demethylation O-demethylation Aromatic hydroxylation

Benzyl hydroxylation

Quinidine inh Indinavir inh Amiodarone inh Fluoxetine inh Haloperidol inh Codeine inh

Kava inh

CYP2E1 Small molecular weight compounds

Acetaminophen Chlorzoxazone

Dehydrogenation 6-hydroxylation

Isoniazid inh Disulfiram inh Garlic inh

CYP3A4 Lipophilic, large molecular weight compounds

Alprazolam Carbamazepine Testosterone

Cyclosporine Midazolam Simvastatin Triazolam Diazepam

Hydroxylation Epoxidation 6!-hydroxylation

Oxidation Oxidation 1-hydroxylation Hydroxylation N-demethylation

Cimetidine inh Erythromycin inh Intraconazle inh

Carbamazepine ind Rifampin ind Phenytoin ind Grapefruit juice ind SJW ind

*inh, inhibitor; **ind, inducer

Page 24: thuoc từ curcumin

14

CYP induction

CYP enzymes are susceptible to induction by structurally diverse xenobiotics, including

drugs, foods and herbal products. Induction occurs when a xenobiotic stimulates the

synthesis of CYP isoenzymes. Mechanisms of CYP induction include Ah-receptor

mediated transcriptional activation, as observed with CYP1A induction by polyaromatic

hydrocarbons. Additionally, nuclear hormone receptors have been identified to be

mediators of CYP2B induction by phenobarbitol (CAR, constitutively active receptor)

and CYP3A induction by rifampicin (PXR) [4]. These mechanisms involve binding of the

xenobiotic to the ligand binding domain of the receptor, which leads to conformational

change in the receptor. This facilitates transportation of the complex to the nucleus by

the respective nuclear translocater and binds to the response elements in the DNA

thereby promoting transcription of the gene.

CYP induction enhances its metabolizing capacity and can affect the efficacy of

medications through increased rate of drug metabolism and hepatic clearance of all

substrates through that specific pathway. This results in sub-therapeutic drug

concentrations. Induction of some CYPs is a risk factor in several cancers, since these

enzymes can convert pro-carcinogens to carcinogens. For example, CYP1A2 activates

heterocyclic amines (including 2-amino-1-methyl-6-phenylimidazol[4,5-b]pyridine

(PhIP)), and polycyclic aromatic hydrocarbons (such as benzo(a)pyrene (BaP)), present

in high temperature cooked foods, especially meat [45]. Thus, drug-food/herb

interactions could result upon administration of drugs, capable of inducing CYP1A2

together with foods or herbal products that can be activated by this enzyme.

Induction potential of drug candidates or compounds is difficult to access pre-

clinically and is often inferred from animal studies, which are not necessarily predictive

for humans [42]. However, in vitro and in vivo approaches for prediction have been

developed. In vitro test systems using cultured primary human hepatocytes,

cryopreserved hepatocytes and liver slices have been employed for studying CYP

induction [42,46,47]. CYP inductive response can be assessed in vitro by measuring the

changes in CYP activity. This is evident in changes of kinetic parameters such as a

decrease in area under the curve (AUC) on a dose response chart [42]. On the other

hand metabolism assays can be performed directly using whole cells. Western blot

Page 25: thuoc từ curcumin

15

analyses are also useful for determining gross changes in protein levels. In addition,

changes in mRNA levels can be used to measure changes in the expression of CYP

genes [46]. The successful use of CYP induction work of CYP1A, CYP2C, CYP2B and

CYP3A enzymes is well documented [42,46-48].

In vivo methods for evaluation of CYP induction by drugs, foods and herbal

products are also available. These include treatment of experimental animals with these

xenobiotic over a period and then measuring CYP activity in the liver and mRNA levels

[49].

Cytochrome P450

The cytochrome P450 enzymes (CYPs) are phase I enzymes and constitute a

superfamily of hemeproteins that are expressed in almost all organisms [50]. In

eukaryotes, they are usually bound to the endoplasmic reticulum or inner mitochondrial

membranes. The name cytochrome P450 is derived from the characteristic absorption

spectra of the reduced CO-bound complex formed at 450 nm [51]. They function

primarily as monooxygenases, which catalyze the incorporation of a single atom of

oxygen into a substrate (Figure 2) [52]. The active site of CYP is composed of an

iron(III) protoporphyrin(IX) moiety with a cysteine amino acid from the protein backbone

as an axial ligand to the iron.

In the resting state (1), a water molecule is the second axial ligand bound to the

iron. The cycle is initiated by a substrate binding to the iron(III) P450 active site (1),

followed by the displacement of the axial water molecule. Subsequently the iron(III)

heme (2) is reduced to the iron (II) state (3) by the addition of a single electron by the

cytochrome P450 NADPH-reductase, so that the ferrous iron can bind molecular

oxygen (4). The binding of molecular oxygen leads to the formation of an iron(II)-

superoxide spieces, after which a second electron is transferred from a redox partner,

yielding a negatively charged iron peroxo intermediate (5). This species is protonated to

the hydroperoxo-iron (6), which after the loss of water yields the active oxenoid iron

species (7), also referred to as compound I. The latter oxidizes the substrate, the

product is released and a water molecule binds again to the ferric iron [52]. In the

catalytic cycle, uncoupling may occur, whereby the cycle is started and molecular

Page 26: thuoc từ curcumin

16

oxygen reduced without product formation as indicated by arrows in figure 3 (a1, a2,

a3).

Figure 3. Catalytic cycle of CYPs, adapted from [52]. Seven stages are shown in the cycle. RH represents

the substrate, ROH the product and –Fe- the iron porphyrin IX, three uncoupling pathways indicated by

a1-a3 and the two peroxide shunt pathways by b1 and b2.

On the other hand, the catalytic cycle can be short-circuited in the presence of artificial

oxygen delivery agents such as peroxides. Thus the cycle turns immediately from stage

2 to stage 6 or 7, depending on the nature of the oxidant as indicated by arrows in figure

2 (b1 and b2) [53].

CYPs catalyze the biotransformation of several endogenous substrates (such as

bile acids, steroids and cholesterol) as well as xenobiotics (drugs, pollutants and dietary

components). Examples of CYP catalyzed reactions are shown in Figure 4. CYPs

mediate biotransformation such as oxidation reactions including hydroxylation of

aliphatic or aromatic carbon, epoxidation of a double bond, heteroatom oxygenation and

Page 27: thuoc từ curcumin

17

N-hydroxylation, heteroatom dealkylation, oxidative group transfer, cleavage of esters

and dehydrogenation [1]. Recently 57 active CYP genes and 58 pseudogenes have

been reported to be present in the human genome [54], belonging to 21 families and 20

subfamilies. CYP enzymes are divided into various subfamilies based on amino acid

homology sequence.

CYP reactions

Heteroatom oxygenation RH ROH

Cleavage of ester

X O

R1

R2

R3

R1

R2

+ HXR3

Heteroatom dealkylation

O

N CH3

OH

N CH3

Epoxidation

O

Reduction

H

O

HCOOHR R +

Hydroxylation

OH

NO2

OH

NO2

OH

Oxidative deamination

CH

2R

NH2

CR

O

H NH3+

Dehalogenation

CR

R

ClH

CR

R

O

Figure 4. Some reactions catalyzed by CYPs.

Page 28: thuoc từ curcumin

18

Drug metabolizing CYPs belonging to the subfamilies 1, 2 and 3 are responsible for

about 90% of all phase I dependent metabolism of clinically used drugs [55,56] and

participate in the metabolism of a huge number of xenobiotic chemicals. CYP

isoenzymes in the same family have at least 40% sequence similarity, while those in the

same subfamily have at least 60% sequence similarity. A majority of CYPs are

expressed in human liver endoplasmic recticulum, although they are also expressed in

extra hepatic tissues.

Drug metabolizing CYPs are extensively polymorphic and this property

influences the outcome of drug therapy causing lack of response or adverse drug

reactions [57]. For example, approximately 5 to 10% of Caucasians are poor

metabolizers of CYP2D6, whilst 20% Japanese and Chinese are poor metabolizers of

CYP2C19 [18,58]. Poor metabolism increases the bioavailability of drugs, and

consequently the possibility of toxicity [57]. Thus, increased drug accumulation and

toxicities are more likely to occur in poor metabolizers via drug-drug interactions upon

multiple administrations of drugs. Adverse drug reactions are much more common at

ordinary dosing among poor metabolizers for CYP2D6 [56]. On the other hand, rapid

metabolizers decrease the bioavailability of drugs resulting in therapeutic failure. Hence,

multiple drug co-administrations could result in further decrease of bioavailability if

induction of a drug metabolizing enzyme occurs.

Human CYP isoforms

Human CYP isoforms of particular importance for drug metabolism are CYP2C9,

CYP2C19, CYP2D6 and CYP3A4, whilst CYP1A1, CYP1A2, CYP1B1, CYP2E1, and

CYP3A4 are the most important isoforms responsible for metabolic activation of

procarcinogens [59]. Large inter-individual variability in the expression has been

observed with CYPs. The average relative abundance of hepatic CYPs is as follows,

CYP1A2 (13%), CYP2A6 (4%), CYP2B6 (<1%), CYP2C (20%), CYP2D6 (2%), CYP2E1

(7%) and CYP3A4 (30%) [60].

CYP1A2 is involved in the metabolism of various endogenous substrates, such

as melatonin and estrogens [2], and in the activation of procarcinogens, such as

heterocyclic amines, arylamines and aflatoxin B1 [61]. Substrates and inhibitors of

Page 29: thuoc từ curcumin

19

CYP1A2 are usually planar small-volume molecules that are neutral or weakly basic.

The binding pocket of CYP1A2 enzyme comprises mostly of hydrophobic and aromatic

amino acids with polar amino acids for hydrogen bonding located near the heme centre

[62]. Drug-drug/food/herb interactions at the level of CYP1A2 may occur as a result of

concomitant administration of drugs metabolized by this enzyme and another drug, food

or herbal product capable of modulating the activity of the enzyme (Tables 2, 3 and 4).

Clinically significant increases in levels of CYP1A2 drug substrate theophylline often

occur after the addition of one of the known inhibitors of this enzyme [Table 4]. Such

interaction may also occur upon intake of foods (such as grapefruit juice) and herbal

products (such as kava extract) that inhibit CYP1A2.

CYP2A6 is the major enzyme catalyzing the oxidative metabolism of nicotine and

cotinine, as well as the metabolism of pharmaceuticals (eg. fadrozole, tegafur, SM-

12502), nitrosamines and a number of coumarin-type alkaloids [63,64]. The CYP2A6

structure shows an enzyme that is well adapted for the oxidation of small planar

substrates, such as coumarin, that fit within the narrow, hydrophobic active site cavity of

the enzyme. CYP2A6 may be inducible by anti-epileptic drugs and it is decreased in

alcohol-induced severe cirrhosis [63]. CYP2A6 does not seem to have an extensive role

in human drug metabolism [65] and genetic variation affecting CYP2A6 activity is not

generally associated with adverse effects on drug clearance [66], suggesting that

CYP2A6 inhibition is unlikely to alter the metabolism of other drugs and be involved in

drug-drug/food/herb interactions.

Earlier studies indicated that CYP2B6 levels were only approximately 0.2% of the

total P450 content in human liver microsomes [55,67]. However, later studies have

demonstrated a greater frequency of detection and a higher percentage of CYP2B6

relative to the total P450 content [68]. This enzyme has proven to be increasingly

important for drug metabolism in the last few years 69]. The substrates of CYP2B6 are

usually non-planar molecules, neutral or weakly basic, fairly lipophilic with one or two

hydrogen-bond acceptors [70]. There are a number of important drugs metabolized by

CYP2B6, including bupropion, efavirenz, cyclophosphamide, ifosamide, pethidine,

artemisinin, propofol, ketamine and selegiline. Drug–drug interactions resulting from

inhibition or induction of CYP2B6 can have serious consequences in the case of

Page 30: thuoc từ curcumin

20

substrate drugs with a narrow therapeutic index, such as cyclophosphamide [71].

Known inhibitors of CYP2B6 include 2-phenyl-2-(1-piperidinyl)propane (PPP),

ticlopidine, and clopidogrel [ 72]. Potent inhibition of CYP2B6 activity was observed in

vitro with a dietary supplement and herbal cold remedy, as well as its individual

components lonicera, isatis root, and schizonepeta, respectively [73]. Such herbal

remedies and food components have the potential to cause drug-food/herb interactions

when taken with CYP2B6 substrate drugs, due to their inhibitory properties.

The CYP2C subfamily is one of the largest and most diverse in mammals, and

the human forms show one of the widest substrate ranges compared to other

subfamilies. This subfamily consists of four members in humans, namely CYP2C8,

CYP2C9, CYP2C18 and CYP2C19 [74]. It has been suggested that CYP2C8, CYP2C9

and CYP2C19 are expressed as functional enzymes in the human small intestine [75].

In addition, CYP2C genes are independently regulated in the human intestinal and liver.

Although, overall, the expression and activity of CYP2C enzymes is lower in the gut

than in the liver, the surface area of the proximal small intestine is large and intestinal

CYP2C9 and CYP2C19 may well contribute to the first-pass metabolism of their

substrate drugs [75]. The CYP2C enzymes are genetically polymorphic and share

significant sequence identity (approximately 70%), but have differences in their

localization and substrate profile [76]. Of them, CYP2C9 is the principal drug

metabolizing enzyme in the liver, and metabolizes many clinically important drugs

[72,76]. Substrates of CYP2C9 are generally weakly acidic with one or two hydrogen

bond acceptors. CYP2C9 is particularly interesting due to its implication in adverse drug

reactions as a result of polymorphism and drug-drug interactions due the narrow

therapeutic window of several substrates eg. (S)-warfarin, phenytoin and tolbutamide

[77,78]. Drugs, food components or herbal products that inhibit CYP2C9 such as

amiodarone, fluconazole, fluoxetine, curcumin, black tea extract and kava extract

[43,44,79,80] taken with CYP2C9 drug substrates, could result in drug-drug/food/herb

interactions of clinical relevance.

CYP2D6 is a highly polymorphic enzyme, therefore large inter-individual

differences exist in its activity [57]. Despite representing only 2% of the total human

hepatic CYPs, CYP2D6 plays an important role in the oxidation of xenobiotics, and is

Page 31: thuoc từ curcumin

21

involved in the metabolism of about 30% of the currently marketed drugs including

antidepressants, neuroleptics, beta-blockers, opioids and anti-arythmics [81,82].

Common characteristics of substrate and inhibitors of CYP2D6 include a flat

hydrophobic region, hydrogen-bonding properties and basic amines [83,84]. Drug-

drug/food/herb interaction could result upon modulation of the activity of CYP2D6 by a

drug, food or herbal product in the presence of other drugs metabolized by this enzyme.

One of the most potent known inhibitors of CYP2D6 is quinidine and it is used for drug-

drug interaction screening assays [84]. Other known inhibitors of CYP2D6 are

Haloperidol, paroxetine, indinavir, codeine, and extracts of Catharanthus roseus and

Artemisia vulgaris [43,44,85].

CYP2E1 is responsible for the biotransformation of a large number of low

molecular weight compounds with hydrogen bond-forming groups [69]. It is also

involved in the disposition of a many chemicals and xenobiotics, the most important of

these substrates being ethanol. Chronic alcohol consumption can induce CYP2E1. The

disadvantage of metabolism by this enzyme is that metabolites produced are usually

hepatotoxic. Substrates of CYP2E1 include acetaminophen, chlorzoxazone and several

inhalation anesthetics [86]. There are few known inhibitors of CYP2E1 and these

include disulfiram, isoniazid and garlic [8,43]. Concomittant intake of CYP2E1 drug

substrates with other drugs, food and herbal products that inhibit this enzyme could

result in harmful drug-drug/food/herb interactions.

Among the human CYP enzymes, CYP3A4 is considered the most versatile and

the most abundant in both liver and small intestine and hence is the most important

CYP [87]. It is responsible for the metabolism of half of the drugs currently on the

market [88]. Thus the inhibition of CYP3A4 by drugs often causes unfavourable and

long-lasting drug-drug interactions with the potential for morbidity and mortality.

CYP3A4 metabolizes structurally diverse substrates with a wide range of molecular

size, shape, enzyme affinity and turnover number [89]. Unusual kinetic interactions

involving CYP3A4 have been observed due to the binding of multiple substrates within

the active site of the enzyme [90]. CYP3A4 appears to be a key enzyme in food-drug

and herb-drug interactions. For example, interactions of grapefruit juice with felodipine

or cyclosporine, red wine with cyclosporine, and St John's wort with various medicines

Page 32: thuoc từ curcumin

22

including cyclosporine, have been reported [43,91,92]. In this thesis, curcumin, a dietary

component has been shown to be a potent inhibitor of CYP3A4 in vitro [80]. The

CYP3A4-related interaction with food components may be related to the high level of

expression of CYP3A4 in the small intestine, as well as its broad substrate specificity.

If potential drug interactions are to be predicted, it is essential that the ability of

food components and herbal products to interfere with drug-metabolizing enzyme

systems is fully established.

Glutathione S-transferases

Drugs, foods and herbal products could also interfere with the activities of Glutathione

S-transferases (GSTs). GSTs are principal phase II biotransformation enzymes

belonging to a superfamily of multifunctional proteins with fundamental roles in the

metabolism and detoxification of a wide range of exogenous and endogenous

compounds [93]. GSTs have also been implicated in cellular physiology,

pathophysiology and other processes such as drug resistance in cancer chemotherapy

[94]. Moreover, genetic polymorphisms in GSTs have been associated with cancer

susceptibility and increased susceptibility for drug interactions or worse outcomes in

diseases [95]. Their main reaction is to catalyze the conjugation of the tripeptide

glutathione (GSH: !-Glu-Cys-Gly), to electrophilic compounds (e.g. Figure 4), to form

more soluble and usually non-toxic peptide derivatives, to be excreted by phase III

enzymes [96]. However, some xenobiotics are converted to reactive intermediates by

the GST-catalyzed reactions. Figure 5 shows a scheme of an overview of enzymatic

detoxification by GST [97]. Almost all soluble GSTs are active as dimers, of either

identical (homodimers) of different (heterodimers) subunits (23-30 kDa), each of them

being encoded by independent genes. Each GST-subunit possesses two ligand-binding

sites: a very specific glutathione-binding site (G-site) and a hydrophobic substrate-

binding site (H-site). Sequence identity within a class is typically >40% while interclass

identities are significantly lower, usually <25% in mammals [93]. Human GSTs are

comprised of at least seven distinct classes of soluble enzymes including Alpha (A), Mu

(M), Pi (P), Theta (T), Omega (O), Sigma (S) and Zeta (Z) [93,98]. GSTs are

differentially expressed both quantitatively and qualitatively in different tissues.

Page 33: thuoc từ curcumin

23

N

O

CH3

O

N

OH

SH

CH3

O

+ GSH

Figure 4. Examples of GST-mediated conjugation reactions between electrophiles and GSH. Conjugation

of GSH with NAPQI (N-acetyl-p-benzoquinone imine), oxidized acetaminophen (APAP) to form 3’-GS-

APAP and conjugation of GSH with ETA (ethacrynic acid) to form ETASG (ethacrynic acid glutathione

conjugate).

Factors such as age, disease and exposure to inducing concentrations and

inhibiting xenobiotics also result in changes in enzyme levels and activities. Kinetic

properties, such as substrate specificities and inhibitor sensitivities can, to some extent

be used as a tool to distinguish between different GST isoenzymes [90]. Chlorinated

nitrobenzenes (1-chloro-2,4-dinitrobenzene, CDNB; 1,2-dichloro-4-nitrobenzene,

DCNB) have long served as standard substrates for nearly all GSTs [99]. The theta

class GSTs, however do not catalyze these reactions. Several GST-isoforms belong to

the Alpha and Mu classes, whilst the Pi class originally contained only one isoform [93].

Alpha, Mu and Pi classes

Alpha class GSTs (GSTA), are highly expressed in liver, constituting 80% of the total

GST protein expressed [100]. The enzyme is also expressed in the plasma, kidney and

testis [99]. Of the five human GSTA enzymes four have been fully characterized, and

these include GSTA1-1, GSTA2-2, GSTA3-3 and GSTA4-4. In hepatic cytosol, they

occur principally as GSTA1-1 and GSTA2-2 homodimers or as GSTA1-2 heterodimers

[101].

O

Cl

CH2

O

OH

ClO

+ GSH

ETA ETASG

O

Cl

O

OH

ClO

CH2

SGO

Cl

CH2

O

OH

ClO

+ GSH

ETA ETASG

O

Cl

O

OH

ClO

CH2

SG

NAPQI 3’-GS-APAP

Page 34: thuoc từ curcumin

24

Figure 5. Overview of enzymatic detoxification

A bioactivation-detoxification pathway in the case of benzo(a)pyrene is illustrated. The xenobiotic diffuses

freely across the plasma membrane into the cell, where it becomes a substrate for CYP1A1, resulting in

the formation of an epoxide. This in turn becomes a substrate for epoxide hydratase. The diol product is a

substrate again for CYP3A4 to form a carcinogenic and mutagenic diol-epoxide derivative. Both enzymes

are microsomal enzymes. GSTs are phase II enzymes which catalyze conjugation of the electrophilic

dihydrodiol epoxide to GSH. The GSH-conjugate is too hydrophilic to diffuse freely from the cell, and can

be pumped out actively by a transmembrane GS-X transporter. This results in the excretion of the GSH-

conjugate from the cell. Ultimately this conjugate is usually metabolized by Phase III enzymes and

excreted in urine as mercapturic acid (adapted from ref. 97).

Similarly five human Mu class GST (GSTM) enzymes have been identified, and these

include, GSTM1-1, GSTM2-2, GSTM3-3, GSTM4-4 and GSTM5-5 [98]. These enzymes

are expressed to varying extents in different tissues depending on the specific subfamily

[99]. For example, human GSTM1 is expressed at highest concentrations in the liver in

individuals carrying 1 or 2 functional alleles while GSTM2 is expressed in highest

concentration in the brain and hardly in the liver. A variety of GSTs are expressed in the

testis, but GSTM3 is expressed almost uniquely in this tissue [102]. GSTM1 deletion

Cell

OH

OH

HO

GS

O

OH

HO

OH

HO

O

OH

OH

HO

GS

Phase I II of

detoxification

Phase I I of

detoxification

Phase I of

detoxification

GSH

GST

O2CYP1A1

H2OEpoxide

hydratase

CYP3A4 O2

Transmembrane

transporter

Cell

OH

OH

HO

GS

O

OH

HO

OH

HO

O

OH

OH

HO

GS

Phase I II of

detoxification

Phase I I of

detoxification

Phase I of

detoxification

GSH

GST

O2CYP1A1

H2OEpoxide

hydratase

CYP3A4 O2

OH

OH

HO

GS

O

OH

HO

OH

HO

O

OH

OH

HO

GS

OH

OH

HO

GS

O

OH

HO

OH

HO

O

OH

OH

HO

GS

O

OH

HO

OH

HO

O

OH

OH

HO

GS

Phase I II of

detoxification

Phase I I of

detoxification

Phase I of

detoxification

GSH

GST

O2CYP1A1

H2OEpoxide

hydratase

CYP3A4 O2

Transmembrane

transporter

Page 35: thuoc từ curcumin

25

frequencies range from 42% to 60% in Caucasians [103]. Some initial studies suggest

that the GSTM1 null genotype confers an increased risk of lung cancer although this

has not been corroborated in subsequent reports [104,105].

The Pi class GSTs (GSTP] appears to be the most widely distributed GST

isoenzyme. It is overexpressed in cancer cells, hence it is regarded as a prognostic

factor in cancer treatment [99]. Regulation of GSTP is of particular interest, in cancer

chemotherapy. Insignificant amounts of GSTP are expressed in the liver and it is the

only class of GSTs expressed in erythrocytes [106]. Unlike GSTA and GSTM, GSTP1-1

is the only isoform recognized in the GSTP class [90].

Although inhibition of GSTs have been considered beneficial in cancer

chemotherapy, in normal cells it may result in harmful consequences such as oxidative

stress and inhibition of important synthesis and signaling pathways [93,107]. The

therapeutic agents, adriamycin, 1,3-bis(2-chloroethyl)-1-nitrosourea (BCNU), busulfan,

carmustine, chlorambucil, cisplatin, cyclophosphamide, melphalan, mitozantrone and

thiotepa are detoxified by GSTs [93]. One of the major heterocyclic amines found in

cooked food is 2-amino-1-methyl-6-phenylimidazol [4,5-b]pyridine (PhIP) (metabolized

by CYP1A2), and GSTs have been shown to detoxify the activated metabolite N-

acetoxy-PhIP. Inhibition of the detoxification of drugs and activated metabolites by

drugs, food components and herbal products, could result in toxicities. Drugs, food

components and herbal products shown to inhibit GSTs include ethacrynic acid,

disulfiram, curcumin, ellagic acid, Thonningia sanguinea, Phyllanthus amarus (this

thesis, chapter 6) [108,109,110]. Understanding of the inhibitory profile of drugs, foods

and herbal remedies towards GSTs is needful to avoid toxicities.

Plants and plant derivatives

Herbal remedies remain one of the forms of therapies used by a large part of the world’s

population [111]. Moreover, a wide range of conventional drugs have originally been

derived from plants. An analysis of the origin of the drugs developed between 1981 and

2002 showed that natural products or natural product-derived drugs comprised 28% of

all new chemical entities (NCEs) launched onto the market [112]. In addition, 24% of

these NCEs were synthetic or natural mimic compounds, based on pharmacophores

Page 36: thuoc từ curcumin

26

related to natural products [113]. This data suggests that natural products are important

sources for new drugs and drug candidates and are also useful lead compounds for

further optimization during drug development. For examples the drugs morphine,

codeine, noscapine, and papaverine isolated from P. somniferum (refer to Figure 6)

were developed as single chemical drugs and are still clinically used [113].

O N

OH

OH

O N

OH

O

N

O

O

O

O

O

O

O

O

O

Figure 6. Chemical structures of natural compounds developed as drugs from plant sources

While many of the plant products may be benign in nature, some of these have the

potential to cause harmful effects resulting from herb-drug interactions in humans. The

consequences of herb-drug interactions can be a) beneficial effects, such as cancer

prevention, b) undesirable effects, such as pharmacokinetic interactions with co-

administered drugs, c) harmful effects, such as organ toxicity or carcinogenesis [114].

Case reports demonstrate that patients taking a number of non-steroidal anti-

inflammatory drugs including aspirin and ibuprofen experienced severe bleeding after

self-prescribing ginkgo extracts at recommended doses [115,116]. Adverse effects were

particularly severe for aspirin (spontaneous hyphema) and for ibuprofen (comatose

state with an intra-cerebral mass bleeding of which the patient died). Interactions of

herbal preparations with immunosuppressant drugs have also been reported. The

interaction between cyclosporine and St. John’s wort is one of the most serious and

potentially fatal interactions between a herbal remedy and a conventional drug [31].

Morphine Codeine

Papaverine Artemisinin

Page 37: thuoc từ curcumin

27

The widely held belief of lay people is that ‘natural’ can be equated to ‘safe’ [111].

For this reason, the use, often self-prescribed is frequently not communicated to the

doctor, with potentially dangerous implications. Recently, interactions of herbal

medicines with synthetic have become of particular interest. Between 1999-2002, more

that 50 papers were published on interactions between St John’s wort and prescribed

drugs only (for a summary see [92,117,118]). The study of herb-drug interactions is

complex, since herbal extracts are multi-component mixtures and the active

constituents are often unknown. Furthermore, in most cases the general information on

the phytochemicals present and their amounts are unavailable. Thus, unspecific effects

such as protein-tannin interaction instead of an enzyme inhibition cannot be excluded.

Similarly, food-drug interactions are complex since foods are multi-component

mixtures. Food drug interactions are often overlooked although a particular food may

modulate the activity of a drug-metabolizing enzyme system, resulting in altered

pharmacokinetics of drugs metabolized by that enzyme system. For example,

consumption of foods such as cruciferous vegetables, grapefruit juice, broccoli,

watercress, red wine and garlic orange juice, have been reported to significantly alter

the pharmacokinetics of several drugs [44]. Investigations have shown that grapefruit

juice is a potent inhibitor of CYP3A4-mediated drug metabolism. It has been

demonstrated to interact with more than 25 drugs including dihydropyridine calcium

channel blockers, cyclosporine, terfenadine and lovastatin, through inhibition of

CYP3A4 [35,119].

The Ghanaian medicinal plants employed in this study, Phyllanthus amarus,

Cassia alata, Cassia siamea, Lactuca taraxicifolia, Momordica charantia, Morinda

lucida and Tridax procumbens are commonly used remedies for various ailments in

Ghana (this thesis, chapter 6). However, there is no information on the potential for

herb-drug interactions in humans and physicians are usually not informed about the use

of such remedies by patients. Hence, adverse effects resulting from herb-drug

interactions are not suspected, investigated and documented. In vitro investigations

have shown the potentials for herb-drug interaction of these medicinal plants at the level

of CYPs and GSTs (in this thesis, chapter 6). Similarly, studies performed on a plant

Page 38: thuoc từ curcumin

28

derivative and dietary component, curcumin have indicated it’s potential to cause drug

interactions. The clinical relevance of these findings remains to be established.

Curcumin and curcumin derivatives

Curcumin is a polyphenolic constituent of Curcuma longa L., (turmeric) obtained from

the powdered root of the plant. It is a component of the spice curry, and gives a unique

flavour and colour to food. Curcumin (Figure 7) can also be synthesized by heating

vanillin, acetylacetone and boric anhydride, over a free flame for 30 min, with a yield of

10% [120]. Turmeric has been and still is used in traditional medicine, especially in

South Eastern Asia, where it is used in the treatment of hepatic and biliary disorders,

rheumatism, cough, and diabetic wounds [121]. Interestingly, curcumin derived from

turmeric possesses several other biological activities, including cancer

chemopreventive, anti-inflammatory, antioxidant, anti-plasmodial and anti-HIV activities

[122-125].

OO

OH

O

OH

OCH3

CH3

Figure 7. Chemical structure of curcumin.

Despite the various therapeutic claims, curcumin certainly has some shortfalls, the most

important includes an extremely low bioavailability which is hampering systemic effects

upon oral administration [126]. Curcumin is also chemically unstable at neutral to basic

pH and exhibits potent inhibition of human GSTs A1-1, A2-2, M1-1, M2-2 and P1-1

[108,127]. However, this instability has been observed to be blocked by antioxidants

such as glutathione, ascorbic acid, N-acetyl-L-cysteine and protein from microsomal or

cytosolic fractions of rat liver [127]. In addition, curcumin is a potent inhibitor of human

CYPs 3A4 and 2C9 [80]. The potential of curcumin to modulate multiple targets may be

a disadvantage due to the possibility of multiple cellular pathways influencing each other

with the likely consequence of undesirable effects. The low bioavailability of curcumin is

a major problem with regards to its therapeutic application. However, various studies

are being performed to overcome this disadvantage. These include the co-

Page 39: thuoc từ curcumin

29

administration of curcumin with piperine, the formulation of curcumin with

phoshatidylcholine or polyethylene glycol, and use of curcumin nanoparticles, liposomal

curcumin and structural analogues of curcumin [128-133].

On the basis of the biological activity and ease of synthesis, curcumin appears

to be useful as a lead compound for the design of derivatives with better biological and

pharmacokinetic properties. For this reason investigators have synthesized various

curcumin analogues [134-136]. Sardjiman et al [134] omitted the active methylene

group and one carbonyl group of curcumin leading to the production of three series of

analogues, which are more stable compounds while retaining antioxidant and anti-

inflammatory properties. Thus, a series of 1,5-diphenyl-1,4-pentadiene-3-one (C),

together with cyclopentanone (B) and cyclohexanone (A) analogues were synthesized

(Figure 8). Some of these curcumin analogues showed potent inhibitory activities

towards human CYPs essential for drug metabolism, demonstrating potential for drug-

drug interactions [137]. Potent GST inhibition has also been observed with some of

these compounds (in this thesis, Chapter 5).

33

1

2

O

R

R

R R

R

R

2

1

1

33

R1

R

R

O

R

R

R

22

1

33

R1

R

R

O

R

R

R

22

Figure 8. Synthetic derivatives of curcumin, including 2,6-dibenzylidenecyclohexanones (A), 2,5-

dibenzylidenecyclopentanones (B) and 1,5-diphenyl-1,4-pentadiene-3-ones (C). These curcumin

analogues are derivatives of benzylidine having either electron-withdrawing, electron-donating or steric

groups.

Quantitative structure-activity relationship (QSAR)

QSAR approaches attempt to identify and quantify the physicochemical properties of a

drug, and to relate these properties to the biological activities of a drug. In drug

A B

C

Page 40: thuoc từ curcumin

30

discovery, QSARs are widely used to identify ligands with high affinity for a given

macromolecular target. More recently, QSAR approaches have also been extended to

rationalize and predict absorption, distribution, metabolism, elimination, toxicity

(ADMET) properties or the oral bioavailability of compounds [138,139]. The basis for

QSAR modeling comes from the concept of linear free-energy relationships, where

variations in the binding behaviour of small molecules to biological targets may be

quantitatively attributed to changes in their structure [140]. In the absence of large

structural changes, useful correlations may be generated between molecular properties

and biological activities and sometimes these may be used to predict the activities of

unknown compounds. For true QSAR relationships, equations can be constructed for

the rationalization of known and the prediction of biological activities of unknown

compounds. If a compound does not fit the equation, it implies that other molecular

features are likely also important, and this provides a lead for further development.

QSAR studies are based on two fundamental assumptions: (i) the affinity data refer to

the same target and (ii) all ligands bind in the same fashion to the receptor [141].

Many physical, structural and chemical properties have been studied by QSAR

approaches, but traditionally the most commonly studied properties emphasize

hydrophobic, electronic and steric effects. Various additional dimensions have more

recently also been employed in classification of QSAR (Table 5).

Table 5. Classification of more recent QSAR approaches based on their dimensionality

Dimension Method

1D Affinity is correlated with global molecular properties of ligands, that is one value per property and ligand (pKa, log P, etc.)

2D Affinity is correlated with structural patterns (connectivity, 2D pharmacophore, etc.) without

consideration of an explicit 3D representation of these properties

3D Affinity is correlated with the three-dimensional structure of the ligands

4D Ligands are represented as an ensemble of configurations

5D As 4D-QSAR + explicit representation of different induced-fit models

6D As 5D-QSAR + representation of different solvation scenarios

Table partly adapted [142].

Page 41: thuoc từ curcumin

31

The Molecular Operating Environment (MOE) software (Chemical Computing Group

Inc. Montreal) is a useful recent source to obtain structural descriptors. The QSAR

approach could play a major role in drug development in predicting compounds with

potentials for drug interactions.

Scope and Objectives

As outlined in this introduction, drug-drug/food/herb interactions mediated by the

inhibition of CYPs and GSTs, two of the most important biotransformation enzymes, are

a major cause of adverse drug reactions. Thus, evaluation of drugs and drug candidates

for CYP, GST inhibitory potential during drug discovery and development has been

considered cost-effective for predicting potentially relevant clinical drug-drug

interactions. Such drug-drug interactions are one of the major causes of attritions in

drug development [17]. Also, plant products such as food and herbal medicines may

have the potential to inhibit CYPs and GSTs and thus result in significant undesirable

effects upon co-ingestion with prescribed drugs or other substances requiring a

particular CYP for metabolism. Several in vitro test systems including well-established

isoenzyme selective assays for drug-drug interaction (i.e. enzyme inhibition or

induction) properties based on microsomal (CYP), cytosolic preparations and

hepatocytes are available [143-145]. The CYPs known to be important for the

metabolism of several drugs currently on the market include CYP3A4, CYP2D6,

CYP2C9, CYP1A2 and CYP2B6 [144].

At the commencement of these studies in 2004, the potential of curcumin, its

synthetic analogues and the seven Ghanaian medicinal plants, to cause drug-drug

interactions at the level of major human drug-metabolizing CYPs, had not yet been

evaluated. However, earlier studies had shown that curcumin is a potent inhibitor of rat

liver microsomal CYP1A1/1A2 and CYP2B1/2B2 [127]. In addition Phyllanthus amarus,

one of the medicinal plants employed, had previously been shown to be a potent

inhibitor of rat liver CYP1A1/1A2, but animal studies often lack predictive power for the

human model [146]. Earlier studies had also shown that some of the present synthetic

analogues had stronger anti-inflammatory and antioxidant properties than curcumin

[134]. The focus of this study was to evaluate the CYP- and GST-mediated drug-drug

Page 42: thuoc từ curcumin

32

interaction potential of plant derived components. In this regard curcumin was

considered a clinically relevant model compound for our purpose. Curcumin, a

frequently used food component isolated from Curcuma longa, possesses interesting

pharmacological properties such as the potential for anti-cancer, anti-oxidant and anti-

inflammatory activities. Thus, investigations are being carried out on its potential as a

chemopreventive agent. Synthetic curcumin analogues were also tested, and QSARs

performed to identify analogues with less potential to cause drug interactions compared

to curcumin and guide future design of curcumin analogues.

Major human recombinant CYP have been used in this work to determine the

CYP inhibitory potentials of curcumin, curcumin decomposition products, thirty-four

structural analogues of curcumin and seven Ghanaian medicinal plants. Secondly, due

to the major role of GSTs in detoxification, oxidative stress and synthesis of important

biomolecules and signal transduction, inhibition of these enzymes by drug candidates,

herbal products or foods could be harmful to normal cells. Thus, in these studies we

have employed the GST isoenzymes GSTA1-1, GSTM1-1 and GSTP1-1 as well as rat

and human liver cytosol, to investigate the inhibitory potencies of the curcumin

analogues and medicinal plants on these enzyme activities.

Aim of this thesis

The primary aim of the investigations described in this thesis is to determine the CYP-

and the GST mediated drug-drug/food/herb interactions via inhibitory potential of plant -

derived components and their derivatives, which includes curcumin, its analogues and

medicinal plants of Ghana. This work particularly focuses on three lines of research:

(i) Drug-drug interaction potential mediated by curcumin and its derivatives,

evaluated as inhibition of curcumin, its decomposition products and synthetic

analogues towards important human CYPs and the mechanisms of inhibition

by curcumin. Structure-activity relationships of analogues were also analyzed.

(ii) The potential of curcumin as well as thirty-four synthetic curcumin analogues

to inhibit human and rat GSTs, and related structure-activity relationships

were analyzed. Earlier studies have shown that curcumin is a potent inhibitor

of rat cytosolic GSTs [127].

Page 43: thuoc từ curcumin

33

(iii) The drug-herb interactions potential of seven important Ghanaian medicinal

herb extracts towards major CYPs were investigated. The inhibitory potentials

of these extracts towards human and rat GSTs were also assessed.

Outline of this thesis

Three lines of research (i-iii) are addressed in the following chapters. In Chapter 1, a

general introduction of the background, aim and the scope of the research described in

this thesis, is given. In Chapter 2 research findings on the pharmacokinetics,

metabolism and drug-drug/-food interactions potentials of curcumin are reviewed. As

reviewed, due to its low bioavailability, curcumin lacks the potential for adequate

therapy in organs and tissues distant from the intestines, however, several new

formulations of curcumin appear to have better potential in this regard. In chapter 3, the

inhibitory effects of curcumin and its decomposition products on human CYP1A2,

CYP2B6, CYP2C9, CYP2D6 and CYP3A4 were assessed, to evaluate potential drug-

drug interaction properties at the level of CYPs. Mechanisms of inhibition of the

enzymes by curcumin were also investigated to obtain insight into the mode of action,

and the possible clinical implications. In chapter 4, thirty-three curcumin analogues

belonging to three series of dibenzylidene derivatives were investigated for their

inhibitory potentials towards the major human drug metabolizing CYPs for reference

purposes compared to curcumin itself. Subsequently, structure-activity relationships

were also analyzed to guide future designing of curcumin analogues with less

susceptibility to CYP inhibition.

The GST inhibitory potential of the thirty-four curcumin analogues was

determined in Chapter 5, and structure-activity relationships were analyzed as well.

These results may be useful in designing and synthesizing curcumin analogues with

less susceptibility to GST inhibition. In Chapter 6, seven important medicinal herbs from

Ghana were assessed for their inhibitory potentials towards human CYPs and GSTs,

since most of these herbal extracts are often consumed together with prescribed drugs.

Finally, Chapter 7 summarizes the results and conclusions of all investigations in this

thesis.

Page 44: thuoc từ curcumin

34

References 1. Parkinson, A., 2001. Biotransformation of xenobiotic. In Casarett and Doull’s Toxicology. The basic science of poisons. Ed. Klaassen C.D., McGraw-Hill Medical Publishing Division, New York, USA, 133-224. 2. Ma, X., Idle, J.R., Krausz, K.W., Gonzalez, F.J., 2005. Metabolism of melatonin by human cytochromes p450. Drug Metab Dispos 33:489–494. 3. Hayes, J.D., Flanagan, J.U., Jowsey, I.R., 2005. Glutathione transferases. Ann Rev Pharmacol Toxicol 45:51-88. 4. Xu, C., Li, C.Y., Kong, A.N., 2005. Induction of phase I, II and III drug metabolism/transport by xenobiotics. Arch Pharm Res 28:249-268. 5. Honig, P.K., Wortham, D.C., Zamani, K., Conner, D., Mullin, J.C., Cantilena, L.R., 1993. Terfenadine ketoconazole interaction. J Am Med Assoc 269:1513-1518. 6. Brandin, H., Myrberg, O., Rundlof, T., Arvidsson, A-K., Brenning, G., 2007. Adverse effects of artificial grapefruit seed extract products in patients on wafarin therapy. Eur J Clin Pharmacol 63:565- 570. 7. Sorensen, J.M., 2002. Herb-drug, food-drug, nutrient-drug and drug-drug interactions: mechanisms involved and their medical applications. J Altern Complement Med 8:293-308. 8. Delgoda, R., Westlake, A.C.G., 2004. Herbal interactions involving cytochrome P450 enzymes. Toxicol Rev 23:239-249. 9. Arayne, M.S., Sultana, N., Bibi, Z., 2005. Grape fruit juice-drug interactions. Pak J Pharm Sci 18:45-57. 10. 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. John's wort induces hepatic drug metabolism through activation of the pregnane X receptor. Proc Natl Acad Sci 97:7500-7502. 11. Dürr, D., Stieger, B., Kullak-Ublick, G.A., Rentsch, K.M., Steinert, H.C., Meier, P.J., Fattinger, K., 2000. St John's Wort induces intestinal P-glycoprotein/MDR1 and intestinal and hepatic CYP3A4. Clin Pharmacol Ther 68:598-604. 12. Li, M., Anderson, G.D., Phillips, B.R., Kong, W., Shen, D.D., Wang, J., 2006. Interactions of

amoxicillin and cefaclor with human renal organic anion and peptide transporters. Drug Metab Dispos 34:547-555.

13. Abernethy, D.R., Flockhart, D.A., 2000. Molecular Basis of Cardiovascular Drug Metabolism: Implications for Predicting Clinically Important Drug Interactions. Circulation 101:1749.

14. Zhou, S., Lim, L.Y., Chowbay, B., 2004. Herbal modulation of P-glycoprotein. Drug Metab Rev 36:57-104.

15. Tirona, R.G., Lee, W., Leak, B.F., 2003. The orphan nuclear receptor HNF-4 alpha determines PXR- and CAR-mediated xenobiotic induction of CYP3A4. Nat med 9:220-224. 16. Chen, J., Raymond, K., 2006. Roles of rifampicin in drug-drug interactions: underlying molecular mechanisms involving the nuclear pregnane X receptor. Ann Clin Microbiol Antimicrob 5: 3. 17. Zhang, Z-Y., Wong, Y.N., 2005. Enzyme kinetics for clinically relevant CYP inhibition. Curr Drug Metab 6:241-257. 18. Zafar, A., Sharif, M.D., 2003. Pharmacokinetics, metabolism, and metabolism of atypical antipsy- chotics in special populations. Primary care companion J Clin Psychiatry 5:22-25. 19. Pea, F., Furlanut, M., 2001. Pharmacokinetic aspects of treating infections in the intensive care

unit: focus on drug interactions. Clin Pharmacokinet 40:833-868. 20. Guengerich, F.P., 1997. Role of cytochrome P450 in enzymes in drug-drug interactions. In: Li, A.P.,

editor. Drug-drug interactions. London: Scientific and regulatory perspectives. 7-35. 21. Kronbach, T., Fischer, V., Meyer, U.A., 1988. Cyclosporin metabolism in human liver: identification

of a cytochrome P450 III gene family as the major cyclosporinmetabolizing enzyme explains interactions of cyclosporine with othr drugs. Clin Pharmacol Ther 43:630-635.

22. Kivisto, K.Y., Neuvonen, P.J., Klotz, U., 1994. Inhibition of terfenadine metabolism: pharmacokin-etic and pharmacodynamic consequences. Clin Pharmacokinet 27:1-5.

23. Fontana, R.J., Lown, K.S., Paine, M.F., Fortlage, L., Santella, R.M., Felton, F.S., et al., 1999. Effects of a chargrilled meat diet on expression of CYP3A, CYP1A, and P-glycoprotein levels in healthy volunteers. Gastroenterology 117:89-98.

24. Desta, Z., Soukhova, N., Flockhart, D.A., 2001. Inhibition of cytochrome P450 isoforms by isonia-zid: potent inhibition of CYP2C19 and CYP3A. Antimicrobial Agents Chemother 45:382-392.

Page 45: thuoc từ curcumin

35

25. Jeppensen, U., Loft, S., Poulsen, H.E., Brsen, K., 1996. Fluvoxamine-caffeine interaction. Pharmacogenetics 6:3213-3222.

26. Carrillo, J.A., Benitez, J., 2000. Clinically significant pharmacokinetic interactions between dietary caffeine and medications. Clin Pharmacokinet 39:127-153.

27. Neal, J.M., Kunze, K.L., Levy, R.H., O’Reilly, R.A., Trager, W.F., 2003. Kiiv, an in vivo parameter for predicting the magnitude of drug interactions arising from competitive enzyme inhibition. Drug Metab Dispos 31:1043-1048.

28. Di Marco, M.P., Edwards, D.J., Wainer, I.W., Ducharme, M.P., 2002. The effect of grapejuice and serville orange juice on the pharmacokinetics of dextromethorphan: the role of gut CYP3A and P-glycoprotein. Life Sci 71:1149-1160.

29. Guengerich, F.P., 1988. Roles of cytochrome P450 enzymes in chemical carcinogenesis and cancer chemotherapy. Cancer Res 48:2946-2954.

30. Prior, T.I., Chue, P.S., Tibbo, P., Baker, G.B., 1999. Drug metabolism and atypical antipsychotics. Eur Neuropsychopharmacol 9:301-309.

31. Barone, G.W., Gurley, B.J., Ketel, B.L., Lightfoot, M.L., Abul-Ezz, S.R., 2000. Drug interaction between St. John’s Wort and cyclosporine. Ann Pharmacother 34:1013-1016.

32. Kupiec, T., Raj, V., 2005. Fatal seizures due to potential herb-drug interactions with Ginkgo biloba. J Anal Toxicol. 29:755-758.

33. Lahoz, A., Donato, M.T., Montero, S., Castell, J.V., Gomez-Lechon, M.J., 2008. A new in vitro approach for the simultaneous determination of phase I and phase II enzymatic activities of human hepatocyte preparations. Rapid Commun Mass Spectrom 22:240-244.

34. Fontana, E., Dansette, P.M., Poli, S.M., 2005. Cytochrome P450 enzymes mechanism-based inhibitors: common sub-structures and reactivity. Curr Drug Metab 6:413-454.

35. Lilja, J.J., Kivisto, K.T., Neuvonen, P.P., 2000. Duration of effect of grapefruit juice on pharmacokin-etics of the CYP3A4 substrate simvastatin. Clin Pharmacol Ther 68:384-390.

36. Neuvonen, J.P., Cantola, T., Kivisto, K.T., 1998. Simvastatin but not pravastatin is very susceptible to interaction with the CYP3A4 inhibitor itraconazole. Clin Pharmacol Ther 63:332-341.

37. Maurer, T., Fung, H.L., 2000. Comparison of methods for analyzing kinetic data from mechanism-based enzyme inactivation: Application to nitric oxide synthase. AAPS PharmSci 2:E8.

38. Brown, H.S., Ito, K., Aleksandra G., Houston, B., 2005. Prediction of in vivo drug-drug interactions from in vitro data: impact of incorporating parallel pathways of drug elimination and inhibitor absorp-tion rate constant. Br J Clin Pharmacol 60:508-518.

39. Scordo, M.G., Pengo, V., Spina, E., Dahl, M.L., Gusella, M., Padrini, R., 2002. Influence of CYP2C9 and CYP2C19 genetic polymorphisms on warfarin maintenance dose and metabolic clearance. Clin Pharmacol Ther 72:702-710.

40. Ito, K., Hallifax, D., Obach, R.S., Houston, J.B., 2005. Impact of parallel pathways of drug elimination and multiple CYP involvement on drug-drug interactions: CYP2D6 paradigm. Drug Metab Dispos 33:837-844.

41. Obach, R.S., Walsky, R.L., Venlatakrisnan, K., Houston, J.B., Tremaine, L.M., 2005. In vitro cytochrome P450 inhibition data and prediction of drug-drug interactions: Qualitative relationships, quantitative predictions and the rank order approach. Clin Pharmacol Ther 78:582-592.

42. Bjornsson, T.D., Callaghan, J.T., Einolf, H.J., Fischer, V., Gan, L., Grimm, S., et al., 2003. The conduct of in vitro and in vivo drug-drug interaction studies: A PhRMA perspective. J Clin Pharmacol 43:443-469.

43. Goshman, L., Fish, J., Roller, K., 1999. Clinically significant cytochrome P450 drug interactions. J Pharm Soc Winsconsin May/June.

44. Butterweck, V., Derendorf, H., Gaus, W., Nahrstedt, A., Schulz, V., Unger, M., 2004. Pharmacok-inetic herb-drug interactions: Are preventive screenings necessary and appropriate. Plant Med 70:784-791.

45. Cross, A.J., Peters, U., Kirsh, V.A., Andriole, G.L., Reding, D., Hayes, R.B., Sinha, R., 2005. A Prospective study of meat and meat mutagens and prostate Cancer Risk. Cancer Res 65:11779-11784.

46. Worboys, P.D., Carlile, D.J., 2001. Implications and consequences of enzyme induction on preclinical and clinical drug development. Xenobiotica 31:539-556.

Page 46: thuoc từ curcumin

36

47. Einolf, H.J., Bedi-Singh, S., Fischer, V., 2002. Precision-cut human liver slices as a model to examine changes in cytochrome P450 expression to predict or preclude drug-drug interactions. The Toxicologist 46:19.

48. Edwards, R.J., Price, R.J., Watts, P.S., Renwick, A.B., Tredger, J.M., Boobis, A.R., et al., 2003. Induction of cytochrome P450 enzymes in cultured precision-cut human liver slices. Drug Metab Dispos 31:282-288.

49. Martignoni, M., de Kanter, R., Grossi, P., Mahnke, A., Saturno, G., Monshouwer, M., 2004. An in vivo and in vitro comparison of CYP induction in rat liver and intestine using slices and quantitative RT-PCR. Chem-Biol Interact 151:1-11.

50. Ortiz de Montellano, P/R., 2005. Cytochromes P450: Structure, mechanism and biochemistry, third ed., Kluwer Academic, New York.

51. Omura, T., Sato, R., 1964. The carbon monoxide binding pigment of liver microsomes. I. solubilization, purification and properties. J Biol Chem 239:2379-2385.

52. Goenhof, A.R., Swart, M., Ehlers, A.W., Lammertsama, K., 2005. Electronic ground states of iron porphyrin and of the first species in the catalytic reaction cycle of cytochrome P450s. J Phys Chem 109:3411-3417.

53. Guengerich, F.P., 2001. Common and uncommon cytochrome P450 reactions related to metabolism and chemical toxicity. Chem Res Toxicol 14:611-650.

54. Nelson, D.R., Zelden, D.C., Hoffman, S.M., Maltais, L.J., Wain, W.H., Nerbert, D.W., 2004. Comparison of cytochrome P450 genes from the mouse and human genomes, including nomenclature recommendations for genes, pseudogenes and alternative-splice variants. Pharmacogenetics 14:1-18.

55. Shimada, T., Yamazaki, H., Mimura, M., Inui, Y., Guengerich, F.P., 1994. Interindividual variations in human liver cytochrome P-450 enzymes involved in the oxidation of drugs, carcinogens and toxic chemicals: studies with liver microsomes of 30 Japanese and 30 Caucasians. J Pharmacol Exp Ther 270:414–423.

56. Ingelman-Sundberg, M., 2004. Pharmacogentics of cytochrome P450 and its applications in drug therapy: the past, present and future. Trends Pharmacol Sci 25:193-200.

57. Ingelman-Sundberg, M., 2005. The human genome project and novel aspects of cytochrome P450 research. Toxicol Appl Pharmacol 207:S52-S56.

58. Nakamura, K., Goto, F., Ray, W. A., McAllister, C. B., Jacqz, E., Wilkinson, G. R., et al., 1985. Interethnic differences in genetic polymorphism of debrisoquin and mephenytoin hydroxylation between Japanese and Caucasian populations. Clin Pharmacol Ther 38:402-408.

59. Nebert, D.W., Russell, D.W., 2002. Clinical importance of cytochromes P450. Lancet 360:11155-1162.

60. Rendic, S., Carlos, F.J.D., 1997. Human cytochrome P450 enzymes: A status report summarizing their reactions, substrates, inducers and inhibitors. Drug Metab Rev 29:413-580.

61. Shimada, T., Guengerich, F.P., 2006. Inhibition of human cytochrome P450 1A1-, 1A2-, and 1B1-mediated activation of procarcinogens to genotoxic metabolites by polycyclic aromatic hydrocarbons. Chem Res Toxicol 19:288-294.

62. Korhonen, L.E., Rahnasto, M., Mahonen, N.J., Wittekindt, C., Poso, A., Juvonen, R.O., et al., 2005. Predictive three-dimensional quantitative structure-activity relationship of cytochrome P450 1A2 inhibitors. J Med Chem 48:3808-3815.

63. Pelkonen, O., Rautio, A., Raunio, H., Pasanen, M., 2000. CYP2A6: a human coumarin 7- hydroxylase. Toxicology 144:139-147.

64. Raunio, H., Rautio, A., Gullsten, H., Pelkonen, O., 2001. Polymorphisms of CYP2A6 and its practical consequences. Br J Clin Pharmacol 52:357-363.

65. Guengerich, F.P.,2005. In Cytochrome P450: Structure, Mechanism, and Biochemistry (ed. Ortiz de Montellano, P.R.) 377-530 (Kluwer Academic/Plenum Publishers, New York, 2005).

66. Yano, J.K., Hsu, M-H., Griffin, K.J., Stout, C.D., Johnson, E.F., 2005. Structures of human microsomal cytochrome P450 2A6 complexed with coumarin and methoxsalen. Nature Structural Mol Biol 12:822-823.

67. Mimura, M., Baba, T., Yamazaki, H., Ohmori, S., Inui, Y., Gonzale, F.J., et al., 1993. Characteriz-ation of cytochrome P-450 2B6 in human liver microsomes. Drug Metab Dispos 21:1048–1056.

68. Stresser, D.M., Kupfer, D., 1999. Monospecific antipeptide antibody to cytochrome P-450 2B6. Drug Metab Dispos 27:517–525.

Page 47: thuoc từ curcumin

37

69. Spatzengger, M., Liu, H., Wang, Q., Debarber, A., Koop, D.R., Halpert, J.R., 2003. Analysis of differential substrate slectivities of CYP2B6 and CYP2E1 by site-directed mutagenesis and molecular modeling. J Pharmacol Exp Ther 304:477-487.

70. Lewis, D.F., 2000. On the recognition of mammalian microsomal cytochrome P450 substrates and their characteristics: towards the prediction of human P450 substrate specificity and metabolism. Biochem Pharmacol 60:293–306.

71. Turpeinen, M., Raunio, H., Pelkonen, O., 2006. The functional role of CYP2B6 in human drug metabolism: substrates and inhibitors in vitro, in vivo and in silico. Curr Drug Metab 7:705–714.

72. Walsky, R.L., Obach, R.S., 2007. A comparison of 2-phenyl-2-(1-piperidinyl)propane (ppp), 1,1',1''-phosphinothioylidynetrisaziridine (thioTEPA), clopidogrel, and ticlopidine as selective inactivators of human cytochrome P450 2B6. Drug Metab Dispos 35:2053-2059.

73. Foti, R.S., Wahlstrom, J.L., Wienkers, L.C., 2007. The in Vitro Drug Interaction Potential of Dietary Supplements Containing Multiple Herbal Components. Drug Metab Dispos 35:185-188.

74. Goldstein, J.A., 2001. Clinical relevance of genetic polymorphisms in the human CYP2C subfamily. Br J Clin Pharmacol 52:349-355.

75. Lapple, F., von Richter, O., Fromm, M.F., Richter, T., Thon, K.P., Wisser, H., Griese, E.U., Eichelbaum, M., Kivisto, K.T., 2003. Differential expression and function of CYP2C isoforms in human intestine and liver. Pharmacogentics 13:565-575.

76. Polgar, T., Menyhard, D.K., Keseru, G.M., 2007. Effective virtual screening protocol for CYP2C9 ligands using a screening site constructed from flurbiprofen and S-warfarin pockets. J Comput Aided Mol Des 21:539-548.

77. Pirmohamed, M., Park ,B.K., 2003. Cytochrome P450 enzyme polymorphisms and adverse drug reactions. Toxicology 192:23–32.

78. Rettie, A.E., Jones, J.P., 2005. Clinical and toxicological relevance of CYP2C9: Drug-drug interactions and pharmacogenetics. Ann Rev Pharmacol Toxicol 45:477-494.

79. Gunes, A., Bilir, E., Zengil, H., Babaoglu, M.O., Bozkurt, A., Yasar, U., 2007. Inhibitory effect of valproic acid on cytochrome P450 2C9 activity in epilepsy patients. Basic Clin Pharmacol Toxicol 100:383-386.

80. Appiah-Opong, R., Commandeur, J.N.M., van Vugt-Lussenburg, B., Vermeulen, N.P.E., 2007. Inhibition of human recombinant cytochrome P450s by curcumin and curcumin decomposition products. Toxicology 235:83-91.

81. Zanger, U.M., Raimundo, S., Eichelbaum, M., 2004. Cytochrome P450 2D6: overview and update on pharmacology, genetics, biochemistry. Naunyn Schmiedebergs Arch Pharmacol 369:23-37.

82. Bertilsson, L., Dahl, M.L., Dalen, P., Al-Shurbaji, A., 2002. Molecular genetics of CYP2D6: clinical relevance with focus on psychotropic drugs. Br J Clin Pharmacol 53:111-122.

83. Koymans, L., Vermeulen, N.P., van Acker, S.A., te Koppele, J.M., Heykants, J.J., Lavrijsen, K., et al., 1992. A predictive model for substrates of cytochrome P450-debrisoquine (2D6). Chem Res Toxicol 5:211-219.

84. Hutzler, J.M., Walker, G.S., Wienkers, L.C., 2003. Inhibition of cytochrome P450 2D6: structure-activity studies using a series of quinidine and quinine analogues. Chem Res Toxicol 16:450-459.

85. Usia, T., Iwata, H., Hiratsuka, A., Watabe, T., Kadota, S., Tezuka, Y., 2006. CYP3A4 and CYP2D6 inhibitory activities of Indonesian medicinal plants. Phytomedicine 13:67-73.

86. Klotz, U., Ammon, E., 1998. Chemical and toxicological consequences of the inductive potential of ethanol. Eur J Clin Pharmacol 54:7-12.

87. Shu-Feng, Z., Changli, C., Xue-Qing, Y., Chunguang, L., Guangji, W., 2007. Clinically Important Drug Interactions Potentially Involving Mechanism-based Inhibition of Cytochrome P450 3A4 and the Role of Therapeutic Drug Monitoring. Ther Drug Monit 29:687-710.

88. Zhou, S.F., Xue, C.C., Yu, X.Q., Wang, G., 2007. Clinically important drug interactions potentially involving mechanism-based inhibition of cytochrome P450 3A4 and the role of therapeutic drug monitoring. Ther Drug Monit 29:687-710.

89. Lu, P., Lin, Y., Rodrigues, A.D., Rushmore, T.H., Baillie, T.A., Shou, M., 2001. Testosterone, 7- benzyloxyquinoline, and 7-benzyloxy-4-trifluoromethyl-coumarin bind to different domains within the active site of cytochrome P450 3A4. Drug Metab Dispos 29:1473-1479.

90. Shou, M., Lin, Y., Lu, P., Tang, C., Mei, Q., Cui, D., et al., 2001. Enzyme kinetics of cytochrome P450-mediated reactions. Curr Drug Metab 2:17-36.

Page 48: thuoc từ curcumin

38

91. Tsunoda, S.M., Harris, R.Z., Christians, U., Velez, R.L., Freeman, R.B., Benet, L.Z., Warshaw, A., 2001. Red wine decreases cyclosporine bioavailability. Clin Pharmacol Ther 70:462-467.

92. Hammerness, P., Basch, E., Ulbricht, C., barrette, E.P., Foppa, I., Bent, S., et al., 2003. St John’s wort: a systematic review of adverse effects and drug interactions for the consultation psychiatrist. Psychosomatics 44:271-282.

93. Frova, C., 2006. Glutathione transferases in the genomics era: New insights and perspectives. Biomol Eng 23:149-169.

94. Lo, H-W., Ali-Osman, F., 2007. Genetic polymorphism and function of glutathione S-transferases in tumor drug resistance. Curr Opinion Pharmacol 7:367-374.

95. Hayes, J.D., Strange, R.C., 2000. Glutathione S-transferase polymorphisms and their biological consequences. Pharmacology 61:154-166.

96. van Bladeren, P.J., 2000. Glutathione conjugation as a bioactivation reaction. Chem Biol Interact 129:61-76.

97. Sheehan, D., Meade, G., Foley, V.M., Dowd, C.A., 2001. Structure, function and evolution of glutathione transferases: implications for classification of non-mammalian members of an ancient enzyme superfamily. Biochem J 360:1-16.

98. Mannervik, B., Board, P.G., Hayes, J.D., Listowsky, I., Pearson, W.R., 2005. Nomenclature for mammalian soluble glutathione transferase. Methods Enzymol 401:1-8.

99. Eaton, D.L., Bammler, T.K., 1999. Concise review of the glutathione S-transferases and their significance to toxicology. Toxicol Sci 49:156-164.

100. van Ommen, B., Bogaards, J.J.P., Peters, W.H.M., Blaauboer, B., van Bladeren, P.J., 1990. Quantification of human hepatic glutathione S-transferase. Biochem J 269:609-613.

101. Dajani, L.K., Paus, E., Warren, D.J., 2001. Development of a rapid and sensitive immunofluorom- etric assay for glutathione S-transferase A. Clin Chem 47:867-873.

102. Listowsky, I., Rowe, J.D., Patskosky, Y.V., Tchaikovskaya, T., Shintani, N., Novikova,E., Nieves, E., 1998. Human testicular glutathione S-transferases: insights into tissue-specific expression of the diverse subunit classes. Chem Biol Interact 111/112:103-112.

103. Garte, S., Gaspari, L., Alexandrie, A.K., Ambrosone, C., Autrup, H., Autrup, J.L., et al., 2001. Metabolic gene polymorphism frequencies in control populations. Cancer Epidemiol Biomarkers Prev 10:1239-1248.

104. Taioli, E., Gaspari, L., Benhamou, S., Boffetta, P., Brockmoller, J., Butkiewicz, D., et al., 2003. Polymorphisms in CYP1A1, GSTM1, GSTT1 and lung cancer below the age of 45 years. Int J Epidemiol 32:60-63.

105. Demir, A., Demir, I., Alt!n, S., Koksal, V., Cetincelik, U., Dincer, S."., 2005. GSTM1 Gene Polymorphisms on Lung Cancer Development in the Turkish Population. Turkish Respir J 6:131- 134.

106. Awasthi, Y.C., Sharma, R., Singhal, S.S., 1994. Human glutathione S-transferases. Int J Biochem 26:295-308.

107. Morrow, C.S., Simitherman, P.K., Townsend, A.J., 2000. Role of multidrug-resistance protein 2 in glutathione S-transferase P1-1-mediated resistance to 4-nitroquinoline 1-oxide toxicities in HepG2 cells. Mol Carcinog 29:170-178.

108. Hayeshi, R., Mutingwende, I., Mavengere, W., Masiyanise, V., Mukanganyama, S., 2007. Inhibition of human glutathione S-transferases activity by plant polyphenolic compounds ellagic acid and curcumin. Food Chem Toxicol 45:286-295.

109. Gyamfi, M.A., Hokama, O., Oppong-Boachie, K., Aniya, Y., 2000. Inhibitory effects of the medicinal herb, Thonningia sanguinea, on liver drug metabolizing enzymes of rats. Hum Exp Toxicol 19:623- 631.

110. Ploemen, P.J., van Iersel, M.L., Wormhoudt, L.W., Commandeur, J.N., Vermeulen, N.P., van Bladeren, P.J., 1996. In vitro inhibition of rat and human glutathione S-transferase isoenzymes by disulfiram and diethyldithiocarbamate. Biochem Pharmacol 52:197-204.

111. Ernst, E., 2006. Herbal medicines – they are popular, but are they also safe? Eur J Clin Pharmacol 62:1-2.

112. Newman, D.J., Cragg, G.M., Snader, K.M., 2003. Natural products as sources of new drugs over the period 1981-2002. J Nat Prod 66:1022-1037.

113. Newman, D.J., Cragg, G.M., Snader, K.M., 2000. The influence of natural products upon drug discovery. Nat Prod Rep 17:215-234.

Page 49: thuoc từ curcumin

39

114. Mandlekar, S., Hong, J.L., Kong, A.N., 2006. Modulation of metabolic enzymes by dietary phytochemicals: a review of mechanisms underlying beneficial versus unfavorable effects. Curr Drug Metab 7:661-675.

115. Matthews, M.K., Jr. 1998. Association of Ginkgo biloba with intracerebral hemorrhage. Neurology 50:1933-1934.

116. Meisel, C., Johne, A., Roots, I., 2003. Fatal intracerebral mass bleeding associated with Ginkgo biloba and ibuprofen. Atherosclerosis 167:367.

117. Henderson, L., Yue, Q.Y., Bergquist, C., gerden, B., Arlett, P., 2002. St John’s wort (Hypericum perforatum): drug interactions and clinical outcomes. Br J Clin Pharmacol 54:349-356.

118. Schulz, V., 2006. Safety of St. John's Wort extract compared to synthetic antidepressants. Phytomedicine 13:199-204.

119. Bailey, D.G., Malcolm, J., Arnold, O., Spence, J.D., 1998. Grapefruit juice drug interactions. Br J Clin Pharmacol 46:101-110.

120. Pabon, H.J.J., 1964. A synthesis of curcumin and related compounds. Rec Trav Chim Pays-Bas 83 :379-386.

121. Ammon, H.P., Wahl, M.A., 1991. Pharmacology of Curcuma longa. Planta Med 57:1-7. 122. Leu, T-H., Maa, M-C., 2002. The molecular mechanisms for the antitumorigenic effect of curcumin.

Curr Med Chem-Anti-Cancer Agents 2:357-370. 123. Vajragupta, O., Boonchoong, P., Morris, G.M., Olson, A.J., 2005. Active site binding modes of

curcumin in HIV-1 protease and integrase. Bioorg Med Chem Lett 15:3364-3368. 124. Cole, G.M., Morihara, T., Lim, G.P., Yang, F., Begum, A., Frautschy, S.A., 2004. NSAID and antio-

xidant prevention of Alzheimer's disease: Lessons from in vitro and animal models. Annals NY Acad Sci 1035:68-84.

125. Reddy, R.C., Vatsaala, P.G., Keshamouni, V.G., Padmanaban, G.,Rangarajan, P.N., 2005. Curcumin for malaria therapy. Biochem Biophy Res Comm 326:472-474.

126. Sharma, R.A., McLelland, H.R., Hill, K.A., Ireson, C.R., Euden, S.A., Manson, M.M., et al., 2001. Pharmacodynamic and pharmacokinetic study of oral Curcuma extract in patients with colorectal cancer. Phytomedicine 13:1894-1900.

127. Oetari, S., Sudibyo, M., Commandeur, J.N.M., Samhoedi, R., Vermeulen, N.P., 1996. Effects of curcumin on cytochrome P450 and glutathione S-transferase activities in rat liver. Biochem Pharmacol 51:39-45

128. Shoba, G., Joy, D., Joseph, T. Majeed, M., 1998. Influence of piperine on the pharmacokinetics of curcumin in animals and human volunteers. Planta Med 64:353-356.

129. Marczylo, T.H., Verschoyle, R.D., Cooke, D.N., Morazzoni, P., Steward, W.P., Gescher, A.J., 2007. Comparison of systemic availability of curcumin with that of curcumin formulated with phosphatidylcholine. Cancer Chemother Pharmacol 60:171-177.

130. Ma, Z., Shayeganpour, A., Brocks, D.R., Lavasanifar, A., Samuel, J., 2007. High performance liquid chromatography analysis of curcumin in rat plasma: application to pharmacokinetics of polymeric micellar formulation of curcumin. Biomed Chromatogr 21:546-552.

131. Bisht, S., Feldmann, G., Soni, S., Ravi, R., Karikar, C., Maitra, A., Maitra, A., 2007. Polymeric nanoparticle-encapsulated curcumin (“nanocurcumin”): a novel strategy for human cancer therapy. J Nanobiotechnology 5:3.

132. Li, L., Ahmed, B., Mehta, K., Kurzrock, R., 2007. Liposomal curcumin with and without oxaliplatin: effects on cell growth, apoptosis, and angiogenesis in colorectal cancer. Mol Cancer Ther 6:1276- 1282.

133. Mosley, C.A., Liotta, D.C., Snyder, J.P., 2007. Highly active anticancer curcumin analogues. Adv Exp Med Biol 595:77-103.

134. Sardjiman, S., Reksohadiprodjo, M., Hakim, L., van der Goot, H., Timmerman, H., 1997. 1,5- Diphenyl-1,4-pentadiene-3-ones and cyclic analogues as antioxidative agents. Synthesis and structure-activity relationship. Eur J Med Chem 32:625-636.

135. Youssef, K.M., El-Sherbeny, M.A., El-Shafie, F.S., Farag, H.A., Al-Deeb, O.A., Awadalla, S.A.A., 2004. Synthesis of curcumin analogues as potential antioxidant, cancer preventive agents. Arch Pharm Pharm Med Chem 337:42-54.

136. Lin, L., Shi, Q., Nyarko, A.K., Bastow, K.F., Wu, C.C., Su, C.Y., et al., 2006. Antitumor agents. 250. Design and synthesis of new curcumin analogues as potential anti-prostate cancer agents. J Med Chem 49:3963-3972.

Page 50: thuoc từ curcumin

40

137. Appiah-Opong, R., de Esch, I., Commandeur, J.N., Andarini, M., Vermeulen, N.P., 2008. Structure- activity relationships for the inhibition of recombinant human cytochromes P450 by curcumin analogues. Eur J Med Chem 43:1621-1631.

138. Norinder, U., 2005. In silico modelling of ADMET-a mini review of work from 2000 to 2004. SAR QSAR Environ Res 16:1–11.

139. Martin, Y.C., 2005. A bioavailability score. J Med Chem 48:3164–3170. 140. Mazza, C.B., Whitehead, C.E., Brenerman, C.M., Cramer, S.M., 2002. Predictive quantitative

structure retention relationship models for ion-exchange chromatography. Chromatographia 56:147-152.

141. Nicolotti, O., Pellegrini-Calace, M., Altomare, C., Carrieri, A., Carotti, A., Sanz, F., 2002. Ligands of neuronal nicotinic acetylcholine receptor (nAChR): inteferences from the Hansch and 3-D quantitative structure-activity relationship (QSAR) models. Curr Med Chem 9:1-29.

142. Lill, M.A., 2007. Multidimensional QSAR in drug discovery. Drug Discovery Today 12:1013-1017. 143. Naritomi, Y., Teramura, Y., Terracita, S., Kagayama, A., 2004. Utility of microtiter plate assays for

human cytochrome P450 inhibition studies in drug discovery: application of simple method for detecting quasi-reversible and irreversible inhibitors. Drug Metab Pharmacokinet 19:55-61.

144. Crespi, C.L., Stresser, D.M., 2000. Fluorometric screening for metabolism based drug-drug intera- ctions. J Pharmacol Toxicol Methods 44:325-331.

145. Walsky, R.L., Obach, R.S., Validated assays for human cytochrome P450 activities. Drug Metab Dispos 32:647-660.

146. Eagling, V.A., Tjia, J.F., Back, D.J., 1998. Differential selectivity of cytochrome P450 inhibitors aga- inst probe substrates in human and rat liver microsomes. Br J Clin Pharmacol 45:107-114.

Page 51: thuoc từ curcumin

41

Chapter 2

Curcumin: Pharmacokinetics, Metabolism, and Potential for Drug-

Drug/Food Interactions

Appiah-Opong R., Commandeur J.N.M. and Vermeulen N.P.E.

Adapted from Proceedings of The International Symposium on recent progress in Curcumin Research,

11-12 September 2006, Yogyakarta - Indonesia

Curcumin, the yellow pigment in turmeric, and a component of a commonly used spice

(curry) derived from the rhizomes of Curcuma longa prevents carcinogenesis in a

variety of tissues in rodents, especially in the colon and also in humans. In addition, it

possesses several other pharmacological activities. This review highlights reported

findings on the pharmacokinetics, metabolism, and drug-drug/food interactions of

curcumin. Doses up to 8 g/day showed no overt toxicity. Peak serum concentrations of

curcumin were remarkably low, after oral intake and gradually decreased within 12 h. In

both rats and humans, the in vivo metabolites were, hexahydrocurcumin,

hexahydrocurcuminol (octahydrocurcumin), hexahydrocurcumin glucuronide and

curcumin glucuronide, whilst in vitro metabolites were hexahydrocurcumin,

hexahydrocurcuminol, tetrahydocurcumin, curcumin glucuronide, and curcumin sulfate.

In addition, O-demethylated curcumin, bis O-demethylated curcumin, O-demethylated

curcumin glucuronide and curcumin bis-glucuronide have been identified as

metabolites. Some of these metabolites may contribute to the observed

pharmacological properties. Approaches used to enhance bioavailability include co-

administration of curcumin with piperine or formulation of curcumin with

phoshatidylcholine or polyethylene glycol. Drug-drug/food interaction potential,

evaluated by inhibition of cytochrome P450 (CYP) has shown curcumin as a potent

inhibitor of CYP2C9 and CYP3A4.

Page 52: thuoc từ curcumin

42

1. Introduction

Curcumin (diferuloylmethane) – (1,7-bis (4-hydroxy-3-methoxyphenyl)-1,6-heptadiene-

3,5-dione) is a natural component of turmeric (Curcuma longa L.) that has emerged as a

promising anti-cancer and chemopreventive agent [1]. The wide spectrum of other

biological activities including chemoprotective, anti-oxidant, anti-inflammatory, anti-HIV

and anti-parasitic properties exhibited by curcumin, has engaged many researchers

over the years [2-6]. Figure 1 summarizes reported therapeutic potentials of curcumin

[7].

Antiangiogenic Diabetes

P latelet

Aggregation

Chemopreventive

skin, liver, colon,

stomach

Multiple

sclerosis

Cardiotoxicity

Wound healing

Curcumin

Cataract formation

Antiinflamatory

Anti -HIV

replication

Cholesterol, smooth

muscle proliferation

Antiangiogenic

nephrotoxicity

Antioxidant

Arth ritis

Lung fibrosis

Gall stones

formation

Figure 1. Curcumin therapeutic potentials (adapted from Aggarwal et al., 2003.

Curcumin has been reported to be non-toxic to humans at doses up to 8 g/day taken for

3 months [8]. Subsequently, a daily oral dose of 3.6 g has been advocated for phase II

evaluation in the prevention or treatment of cancers outside the gastro-intestinal tract

[9].

Investigations on the pharmacokinetics, metabolism and drug-drug/-food

interactions of curcumin have been reported [10-14]. Although several interesting

Page 53: thuoc từ curcumin

43

pharmacological properties have been reported on curcumin, it has a drawback of poor

systemic bioavailability [15]. For this reason studies have been performed on

enhancement of its pharmacokinetic properties. These include the use of piperine,

curcumin phospholipids complex, curcumin nanoparticles, liposomal curcumin and

structural analogues of curcumin [16]. Studies on metabolism of curcumin have also

shown that it undergoes reductive metabolism as well as conjugation reactions such as

conjugation with glutathione, glucuronidation and sulfation [17-19]. In addition, its

potential to cause drug-drug/-food interactions by inhibition or induction of CYP

enzymes has been studied [10,14,20]. This review will therefore primarily focus on

recent findings on the pharmacokinetics, metabolism and drug-drug/-food interactions of

curcumin.

2. Pharmacokinetics of curcumin

In a series of animal experiments, Min/+ mouse model of familial adenomatous

polyposis were administered 14C-labeled curcumin (100 mg/kg) via the intra-peritoneal

Figure 2. Elimination of radioactivity derived from [14

C]curcumin from the intestinal tract mucosa (A),

plasma (B), and liver (C) of C57B1/6J mice, which had received a single dose of [14

C]curcumin (100

mg/kg) via the i.p. route. Values are expressed as nmol curcumin equivalents per gram (g tissue; A, C) or

milliliters (ml) of plasma (B), and are the mean + SD of four mice (adapted from [21]).

Page 54: thuoc từ curcumin

44

route and monitored for the appearance and disappearance of radioactivity from the

intestinal tract mucosa, plasma and liver (Figure 2) [21].

Table 1. Disposition and toxicity data on curcumin

Dose (g/day)

Period (days)

Subjects Serum/plasma / tissue concn.

Toxicity Reference

2.0 0.17 Healthy volunteers

Serum 0.016 µM None [22]

0.02 1 Healthy volunteers

N None [23]

0.5-12 1 Healthy volunteers

Serum 0.0-0.155 µM None [24]

0.036-0.18

120 Colorectal cancer patients

Serum None Nausea, diarrhea

[12]

0.5-12 90 Malignant lessions patients

Serum 0.5-1.77 µM None [8]

0.45-3.60

7 Cancer patients

Plasma 0.0-3 nmol/L N [25]

0.4 Healthy rats 1.35 µg/ml N [22]

0.2 0.04-0.06

Healthy rats Serum 0.5-1.5 µg/ml N [26]

0.003* 0.02 Adenomatus Mice

Plasma 25 nmol/ml None [21]

0.04 Intestinal mucosa

200 nmol/g None

0.08-0.17

Brain 2.9 nmol/g None

0.01 Liver 73 nmol/g None

0.08-0.17

Kidney 78 nmol/g None

0.08-0.17

Lung 16 nmol/g None

0.08-0.17

Heart 9.1 nmol/g None

N: Not reported; *: i.p. administration

Page 55: thuoc từ curcumin

45

At different time points, samples of brain, heart, lung, liver, spleen, kidney, small

intestine and blood were collected, solubilized and analyzed by a liquid scintillation

counting. Peak levels of radioactivity in plasma and tissues, reached in less than four

hours, are shown in Table 1.

Beyond the peak levels, radioactivity declined rapidly to reach levels between 20

and 33% of peak values 4 h after dosing, or 8 h in the case of the small intestine. After

these time points minimal or no decrease in radioactivity levels were observed any more

up to 24 h.

Another group of the mice received diets containing 150-750 mg/kg/day curcumin

for 8 days [21]. Liver, small intestine, colon tissue and plasma were analyzed for

curcumin and its metabolites. Irrespective of the dose, curcumin was detected in plasma

at levels near the limit of detection (5 pmol/ml) and the parent compound was not

detectable in the urine.

In the liver tissue of mice fed a 0.2% curcumin diet, the concentration of curcumin

was 119 + 31 pmol/g of tissue i.e. approximately 0.001 of that observed in the intestinal

mucosa. Significantly higher concentrations of curcumin i.e. up to 3770 nmol/g tissue

were found in the mucosa of the small intestine, colon and faeces, and the latter

contained the largest concentrations (Table 2). Curcumin levels in the small intestines

related to the dose are shown in Table 2, with the exception of levels in colon and

faeces.

Table 2. Concentration of curcumin in small intestinal and colonic mucosa and faeces of mice that

received curcumin at 0.1, 0.2, or 0.5% in their diet for 1 week

Curcumin levelsa (nmol/g)

Curcumin content of diet

(%)

Small intestine

mucosa

Colon

mucosa

Faeces

0.1 39 + 9 15 + 9 3770 + 1246

0.2 111 + 40 508 + 149 3590 + 231

0.5 240 + 69 715 + 448 3186 + 2411

a Values are the means ± SD of four animals (adapted from [21]).

Page 56: thuoc từ curcumin

46

Conjugative or reductive metabolites of curcumin were not detected except in the

colonic mucosa and faeces, where HPLC analysis revealed traces of curcumin sulfate.

In mice first fed 300 mg/kg/day of curcumin for 1 week and then changed onto a

curcumin-free diet [21], levels of curcumin in plasma, gastro-intestinal and hepatic

tissues, rapidly declined to unquantifiable concentrations within 3 to 6 h after starting the

curcumin-free diet, however, faecal curcumin declined more slowly with a half-life of

about 23 h. Thus, the rapid removal of curcumin from the rodent tissues, including

target tissues must be taken into account, if sustained levels of curcumin are to be

achieved to elicit its biological effects. Secondly, some metabolites of curcumin may

also contribute to biological activities.

Studies aiming at increasing curcumin systemic bioavailability have recently been

performed. A complex of curcumin and phoshatidylcholine (Meriva), improved the

systemic bioavailability of curcumin in rat plasma and tissues [15]. Levels of curcumin in

the gut mucosa 2 h after administration of 340 mg/kg formulated curcumin by oral

gavages, were moderately lower and in plasma moderately higher than those subjected

to unformulated curcumin. Both formulated and unformulated curcumin were completely

cleared from plasma within 2 h, however, peak plasma levels of curcumin being

approximately 5-fold higher in the formulated form (Table 3).

Table 3. Estimated plasma peak levels for unformulated and formulated (Meriva) Curcumin in rats

Cmax (nM) Tmax (min) AUC (!g min/ml)a

Unformulated

Curcumin 6.5 + 4.5 30 4.8

Curcumin glucuronide 225.0 + 0.6 30 200.7

Curcumin sulfate 7.0 + 11.5 60 15.5

Meriva (formulated)

Curcumin 33.4 + 7.1 15 26.7

Curcumin glucuronide 4420.0 + 292 30 4764.7

Curcumin sulfate 21.2 + 3.9 60 24.8

Cmax, estimated plasma peak levels; Tmax, time of peak levels; aAUC was calculated using WinNonLin and

employing a non-compartmental model (adapted from [15])

Page 57: thuoc từ curcumin

47

Nonetheless, maximum systemic concentrations of the curcumin phospholipid complex

were still considerably below the levels eliciting pharmacological effects in cells and cell

free systems. Plasma levels of curcumin glucuronide, curcumin sulfate,

tetrahydrocurcumin and hexahydrocurcumin after administration of formulated curcumin

were 3 to 20 fold higher than after administration of unformulated curcumin (Table 3).

Obviously, a curcumin phospholipid complex significantly enhances systemic and

hepatic bioavailability of parent curcumin and metabolites as compared to unformulated

curcumin [15].

Other strategies to increase bioavailability involve the use of piperine, (1-[5-(1,3-

benzodioxol-5-yl)-1-oxo-2,4-pentadienyl]piperidine), extracted from black pepper to

enhance the bioavailability of curcumin in rats and healthy volunteers [22]. Co-

administration of curcumin (2 g/kg) and this piperine (20 mg/kg) in rats, resulted in

increased serum concentrations of curcumin for a period of 1-2 h after administration.

Elimination half-life of curcumin was significantly increased and clearance was

decreased, whilst bioavailability was increased by 154% with no adverse effects.

Whereas in humans undetectable serum levels of curcumin were obtained after a dose

of 2 g, subsequent co-administration with 20 mg piperine produced a 2000% increase in

bioavailability [22]. This significant increase in bioavailability needs to be verified. The

results suggest that for chemoprevention, targeting sites other than the gastrointestinal

tract, curcumin formulated with phosphatidylcholine or piperine may well be more

advantageous than unformulated curcumin.

Recently, curcumin solubilized with N,N-dimethylacetamide, polyethylene glycol

(PEG 400) and 40% of isotonic dextrose as well as micellar formulation of curcumin,

were administered to rats at doses of 10 and 5 mg/kg body weight [27]. The half-life of

solubilized curcumin was less than 1 h, whereas that of the curcumin encapsulated in

the polymeric micellar formulation was over 60 h. Moreover, a 3-fold decrease in

clearance was observed. In another similar attempt to overcome the low aqueous

solubility and bioavailability of curcumin, curcumin-polyethylene glycol conjugates of two

differently sized polyethylene glycol molecules (PEG 750 and PEG 3500) were used.

These conjugates were employed in treatment of some human cancer cell lines, and

Page 58: thuoc từ curcumin

48

were found to exhibit enhanced cytotoxicity as compared to that of the parent drug [28].

Although supporting data is still lacking, these water-soluble conjugates of curcumin

may be useful for injectable curcumin conjugates.

The use of liposomes, a drug delivery system with curcumin has also been

explored using in vitro and in vivo systems [29]. Studies on antitumour activity of

liposomal curcumin towards human pancreatic carcinoma cells have shown that it

inhibits pancreatic carcinoma growth and exhibits antiangiogenic effects. These

properties were compared with that of untreated and liposomal vehicle treated mice and

comparable or greater growth inhibition was observed. However, the bioavailability of

liposomal curcumin as compared to free curcumin is yet to be established.

Application of nanoparticles as in drug delivery systems to improve bioavailabilty

of curcumin has also been explored [30]. Nanocurcumin with less than 100 nm size, has

been synthesized, tested and found to have similar in vitro activity as that of free

curcumin in pancreatic cell lines. Comparison of bioavailability of nanocurcumin to free

curcumin remains to be evaluated.

Another approach to overcome the poor systemic bioavailability was the

synthesis of curcumin analogues [16,31]. A curcumin analogue EF-24 given to CD2F1

mice has been shown to be absorbed rapidly after both oral and i.p. administration [32].

The elimination half-life was 73.6 min and plasma clearance 0.482 L/min/kg. Plasma

peak concentrations detected 3 min after i.p. dosing were about 1000 nM. The EF-24

exhibited 60% and 35% bioavailability upon oral and i.p. administration respectively. EF-

24 was reported to be a lead compound possessing antitumour activity in vitro and in

vivo as compared to curcumin [32]. These analogues showed no in vivo toxicity.

Phase 1 clinical trials of curcumin showed that serum concentrations peaked at 1

to 2 h after oral intake of up to 12 g/day of curcumin, and gradually declined within 12 h

[8]. Remarkably, after administration of 4-8 g of curcumin, only low average peak serum

concentrations of <2.0 !M, were recorded (Table 1). Urine did not contain detectable

amounts of curcumin, and no treatment-related toxicity was observed at curcumin

concentrations < 8 g. In related pharmacokinetic studies of oral curcuma extract in

patients with colorectal cancer, patients were given, 36-180 mg of curcumin, after at

least a 2 h fast [12]. Curcumin and its potential metabolites including curcumin

Page 59: thuoc từ curcumin

49

glucuronide, curcumin sulfate, hexahydrocurcumin and hexahydrocurcuminol

(octahydrocurcumin) were not measurable in plasma or urine samples up to 29 days of

daily treatment. However, on days 8 and 29, after 144 and 180 mg consumption of

curcuma extract, faecal samples from patients showed the presence of curcumin

concentrations of < 519 nmol/g and < 1054 nmol/g, respectively. Curcumin sulfate was

the only metabolite identified in the faeces. On the other hand only unchanged curcumin

at a low concentration of 11.1 nmol/L, was detected in plasma samples taken 0.5 and 1

hour after administration to 3 patients consuming 3.6 g of curcumin daily [9]. The only

discernable toxicity upon administration of 0.5-3.6 g curcumin daily for up to 4 months

was a mild diarrhea. Table 1 shows a summary of the serum or plasma concentrations

and toxicity of curcumin in relation to administered doses in both healthy volunteers and

patients. These findings suggest that oral administration of doses up to 3.6 g of

curcumin daily for several months does not result in a significant systemic uptake of

curcumin, nor in tissue accumulation. Oral administration of curcumin will thus not be

efficacious for therapeutic effects in target organs distant from the gastro-intestinal tract,

due to its very low systemic bioavailability.

The low systemic bioavailability of curcumin might imply that the pharmacological

activity especially in tissues other than the colon, is mediated in part by curcumin

metabolites, since several biological activities have been observed at sites distant from

the locus of absorption in rodents, such as breast, prostate and liver [33-35].

Enhancement of curcumin bioavailability clearly might increase the therapeutic

application of this promising drug candidate. Further evaluation of drug-drug interactions

and other toxicities of solubilized or formulated curcumin at administered doses

however remain imperative.

3. Metabolism of curcumin

Metabolism is an integral component of the processes that govern pharmacokinetics,

since it makes drugs more soluble, thus facilitating transport to target organs and

elimination from the body and also usually renders them less toxic. Metabolism studies

on curcumin using slices and subcellular fractions from rat liver identified five reductive

but no oxidative metabolites of curcumin using HPLC and GC-MS analysis [19]. The

Page 60: thuoc từ curcumin

50

major reductive metabolites in rat liver slices originate from the reduction of the double

bonds of the heptadiene-3,5-dione chain, resulting in hexahydro-, tetrahydro- and

octahydro-curcumin and the minor products dihydrocurcumin and octahydrocurcumin

(Figure 3) [19]. These metabolites were predominantly present as glucuronides,

although a significant proportion of sulfate conjugates were also observed. No oxidative

metabolites of curcumin nor reductive metabolites were found using rat liver

microsomes and cytosol [19]. The biological activities of most of the reductive

metabolites of curcumin have not yet been established. However, tetrahydrocurcumin

was found to be a more potent antiinflammatory agent [36] and an equipotent

antioxidant as curcumin [37].

Figure 3. Scheme of curcumin metabolism in both phase I and phase II biotransformations.

Page 61: thuoc từ curcumin

51

The reduced metabolites appear to be conjugated in vitro and in vivo to a

monoglucuronide, monosulfate and a mixed sulfate/glucuronide [11]. Metabolites

identified in related studies include curcumin glucuronide, curcumin sulfate,

hexahydrocurcumin glucuronide and mixed glucuronide and sulfate conjugates

[11,17,18]. Figure 3 shows an overall scheme of curcumin metabolism in both phase I

and phase II reactions as yet reported in literature.

The metabolism of curcumin in subcellular cytosolic and microsomal fractions of

human liver and intestines has also been reported [18]. Spectrophotometric analysis

demonstrated the reduced metabolites, tetrahydro- and hexahydrocurcumin, and the

phase II metabolites, curcumin sulfate and curcumin glucuronide, containing the intact

yellow-pigmented 1,7-bis (4-hydroxy-3-methoxyphenyl)-1,6-hepta-diene-3,5-dione

structure (Figure 4).

Figure 4. HPLC chromatograms of extracts of incubations of curcumin (100 µM) with cytosol (A and C)

and microsomes (B) from human intestinal tissue and with cytosol (D and F) and microsomes (E) from rat

intestinal tissue. Incubation periods were 90 min for metabolic reduction (A and D) and 60 min for

conjugation (B, C, E and F) and chromatographic peaks were detected at 280nm and 420 nm

respectively. Hexahydrocurcumin (1), tetrahydrocurcumin (2), curcumin (3), curcumin glucuronide (4) and

curcumin sulfate (5) (adapted from [18]).

Page 62: thuoc từ curcumin

52

Human intestinal cytosol resulted primarily in curcumin sulfate and hexahydrocurcumin.

Intestinal human microsomes did not generate detectable levels of curcumin reduction

products [18]. Quantitative evaluation of curcumin metabolism in human tissue fractions

suggests that gut metabolism contributes considerably to the total metabolism of

curcumin in vivo [18,21]. This suggestion is in line with results of experiments in which

[3H] labeled curcumin was incubated with everted rat gut sacs [38].

Curcumin and its metabolites in humans were also measured in portal and

peripheral blood, bile and liver tissue, after daily oral intake of 0.45, 1.8 and 3.6 g of

curcumin for one week [39]. Samples of peripheral blood were taken 1 h immediately

after curcumin dosing and hepatic resection was performed 6-7 h after the last dose of

curcumin. Samples of portal blood and bile were taken intra-operatively. Curcumin was

poorly bioavailable following oral administration, with low nanomolar levels of curcumin

and its glucuronide and sulfate conjugates found in peripheral or portal circulation [39].

Although curcumin itself was not found in human liver tissue, trace levels of its

metabolic reduction products, hexahydrocurcumin and octahydrocurcumin were

detected. The low levels of curcumin in plasma are consistent with other clinical findings

indicating that oral doses up to 180 mg of curcumin do not result in detectable plasma

levels [12] while high doses of up to 8 g yield only approximately 0.5-2 µM of curcumin

in serum within 1h of oral administration [8].

Metabolic studies on curcumin have recently also been performed using mouse

and human liver microsomes in the presence of the cofactors NADPH for phase I

reactions and of NADPH and UDPGA for phase II reactions [40]. Analysis by LC-

MS/MS was done in full mass range (m/z: 90-800) and several metabolites including

oxidative metabolites and those previously identified, were found (Figure 3). These

included O-demethylated curcumin (m/z: 355), bis O-demethylated curcumin (m/z:

341), the respective di-hydrogenated derivatives (m/z: 357 and m/z: 343), O-

demethylated curcumin glucuronide (m/z: 517) and curcumin bisglucuronide (m/z:

721) (Figure 3) [40]. In a related study, activities of human hepatic and intestinal

microsomes and nine human recombinant UGT isoforms towards curcumin,

demethoxycurcumin and bis-demethoxycurcumin and their hexahydro- metabolites

Page 63: thuoc từ curcumin

53

were determined [41]. Two curcumin monoglucuronides were observed with human

liver microsomes, as previously reported by Pfeiffer et al [42], the major product

having a glucuronic acid at the phenolic position and the minor at the enolic hydroxyl

group, whereas in human intestinal microsomes only the latter conjugate was

observed [41]. Glucuronidation of curcumin was mainly catalyzed by UGT1A1, 1A8

and 1A10, and hexahydrocurcumin by UGT1A8, 1A9 and 2B7 (Table 4). All UGT

isoforms except UGT1A9 formed the phenolic glucuronide of curcumin. Due to its

preference for non-planar phenolic substrates, UGT1A9 preferred hexahydrocurcumin

containing saturated aliphatic chain, hence being less planar than curcumin [41,43].

Table 4. Glucuronidation of curcumin and hexahydrocurcumin by UGT isoforms

UGT activity (pmol/min/mg

protein)

Isoenzyme Curcumin Hexahydrocurcumin

UGT1A1 1875 375

UGT1A3 550 65

UGT1A6 40 nd

UGT1A7 180 210

UGT1A8 1585 950

UGT1A9 100 790

UGT1A10 1540 375

UGT2B7 300 535

Rates of glucuronidation were determined at 20 µM concentrations for each substrate and 0.1-0.2 mg

protein. Reaction mixtures were incubated at 37oC for up to 2 h with linear product formation; nd, not

determined (data adapted from [41]).

These results reaffirm the fact that apart from the liver, the intestinal tract is substantially

contributing to first pass glucuronidation of curcumin.

The studies reported provide clear evidence that after oral intake curcumin is

metabolized in humans predominantly in the intestinal tissue. The rat may not be a good

model for elucidation of the extra hepatic metabolic disposition of curcumin in humans

Page 64: thuoc từ curcumin

54

due to the greater ability of human intestinal and liver tissues to metabolize curcumin

than rats. Moreover, it has been observed in humans that cytosol or alcohol

dehydrogenase is required for the formation of tetrahydrocurcumin and

hexahydrocurcumin, while microsomes are needed for the reduction of

hexahydrocurcumin to octahydrocurcumin [18]. The specific enzymes responsible for

the phase I metabolism of curcumin and the pharmacological implications of curcumin

metabolites warrant further investigation. Identification of novel curcumin metabolites

using hybrid quadrupole linear ion trap mass spectrometer coupled with liquid

chromatography [40] have resolved the presence of several additional metabolites in

various tissues, plasma and faeces (Figure 3). Probable curcumin metabolites may

have to be synthesized and used for further identification, pharmacological screening

and safety assessment.

4. Drug-drug and drug-food interactions

Drug-drug interactions are major causes of attrition during drug development [44,45], as

well as of adverse drug reactions in humans [46]. Food-drug or dietary supplement-drug

interactions have also been associated with significant alterations in the

pharmacokinetic profile of various drugs that may have clinical implications [47]. For

example, interactions of red wine with cyclosporine and grapefruit juice with

cyclosporine and felodipine are potential causes of alterations in pharmacokinetics [48],

with possible therapeutic failure and adverse effects [49]. In a study of Jang et al [50],

75% of 116 food supplements were found to induce at least one rat liver microsomal

CYP expression, including CYP1A1, CYP2C11, CYP2D1, CYP2E1 and CYP3A1.

Compounds isolated from St John’s wort, an antidepressant of natural origin, have been

shown to possess potent inhibitory activity towards CYP1A2, CYP2C9, CYP2D6 and

CYP3A4 [51] and induction of CYP3A4 with chronic exposure [52]. Although drug-drug

or drug-food interactions caused by induction of CYP enzymes are known, interactions

due to CYP inhibitions are much more common [53]. Table 5 shows the percent

inhibition of major human CYP isoforms by 625 µg extract of selected natural

products/ml, including curcuma extracts [13].

Page 65: thuoc từ curcumin

55

In a recent study, 5 human recombinant CYP isoforms, known to be important

for metabolism of several drugs currently on the market namely, CYP3A4, CYP2D6,

CYP1A2, CYP2C9 and CYP2B6 were evaluated for CYP inhibition by curcumin. The

results showed curcumin as a moderate to potent inhibitor of CYP2C9 and CYP3A4 and

a less potent inhibitor of CYP1A2, CYP2D6 and CYP2B6 (Table 5) [14]. Competitive

type of inhibition was observed in the cases of CYP3A4, CYP1A2 and CYP2B6,

however, in the cases of CYP2C9 and CYP2D6 non-competitive inhibition was

observed. Further experiments showed that curcumin is not a mechanism-based

inhibitor of any of the 5 human CYPs mentioned above [14]. Inhibition of rat liver

microsomal CYPs by curcumin has also been investigated [10].

Table 5. Percent inhibition of major human CYP isoforms, by 625 µg natural products extracts/ml and IC50

values (µM) for curcumin

Extract CYP1A2 CYP2B6 CYP2C9 CYP2D6 CYP3A4

Artemisia vulgaris nd nd 97 100 97

Thyme nd nd 93 96 97

Cloves nd nd 99 98 94

Ginger nd nd 53 70 94

Black tea (5

varieties)

nd nd 92 - 98 76 – 93 77 – 84

Curcuma nd

(40.0*)

nd

(24.5*)

82 (4.3*) 49

(50.3*)

93

(16.3*)

The extracts or infusions were each tested at a single concentration indicated above (adapted from [13]).

*IC50 values for curcumin were determined within a concentration range of 0.4-181.8 µM [14]. nd, not

determined

Curcumin was found to be a potent competitive inhibitor of rat liver CYP1A1/1A2

measured as ethoxyresorufin deethylation (EROD) activity in !-naphthoflavone-induced

microsomes, a less potent competitive inhibitor of CYP2B1/2B2, measured as

pentoxyresorufin depentylation (PROD) activity in phenobarbital-induced microsomes

and a weak inhibitor of CYP2E1, measured as p-nitrophenol (PNP) hydroxylation

Page 66: thuoc từ curcumin

56

activity in pyrazole (Pyr)-induced microsomes [10]. Curcumin was found to be a very

weak inhibitor of CYP2E1.

Induction studies on the effects of dietary flavonoids including curcumin on

human CYP1A1 expression have been performed on the 101L cell line (derived from

human hepatoma cell line HepG2), transfected with a plasmid containing the human

CYP1A1 promoter [20]. The 101L cells were plated at a density of 7.5 x 104 cells/well in

96-well plates, and dose response curves for the various flavonoids were generated at

doses ranging from 1 to 20 !M with 18 h of exposure (Figure 5). A 3-fold increase in

activity of CYP1A1 was measured as elevation in luciferase activity. Compared to

2,3,7,8-tetrachlorodibenzo-"-dioxin (TCDD), omeprazole or benzanthracene where

increases in activity ranged from 12- to 35-fold, curcumin is a weak inducer of CYP1A1

[20].

Figure 5. Dose response curves for various flavonoids. Doses ranged from 1-20 µM and cells were

exposed to each agent for 18 h. Each time point represents the mean of data from three experiments +

SD. (adapted from [20]).

Apart from drug-drug interactions, CYP inhibition has been related to

chemopreventive activity against benzopyrene (B[a]P)-induced carcinogenesis, due to

Page 67: thuoc từ curcumin

57

inhibition of CYP1A1-mediated bioactivation of B[a]P. CYP inhibition may affect the

pharmacokinetics of drugs by decreasing the elimination-half life, hence increasing the

plasma concentration increasing the potential for toxic consequences. Animal data is

often poorly predictive of human situations due to species differences, including

differences in the properties of metabolic enzymes [54]. This is exemplified by high

inhibitory potency of curcumin towards rat liver CYP1A1/1A2 compared with the lower

potency obtained with human CYP1A2. The drug-drug interaction potential of curcumin

resulting from inhibition of CYPs needs to be further investigated for clinical relevance.

Curcumin has also been demonstrated to inhibit the expression of P-glycoprotein

(Pgp) in the multi-drug resistance (MDR) human cervical carcinoma KB-VI cells,

increase rhodamine-123 (Rh123) accumulation and inhibit Rh123 efflux in these cells

[55]. MDR is a challenge, limiting the success of chemotherapy, mainly due to the

overexpression of the Pgp, which causes a decrease in drug accumulation in cancer

cells [56]. Both Pgp and CYP3A4 have been suggested to act synergistically to limit the

bioavailability of orally administered agents [57]. Thus the inhibition of both enzymes by

curcumin may be related to the chemopreventive role of the latter.

Curcumin is also a potent inhibitor of glutathione S-transferase (GST) activity in

cytosol of rat liver treated with PB, !NF and Pyr, with 1-chloro-2,4-dinitrobenzene

(CDNB) as substrate. Similarly, curcumin has been found to be a potent inhibitor of

humans recombinant GSTs [58, in this thesis, Chapter 5]. Curcumin was shown to

inhibit GSTs A1-1, A2-2, M1-1, M2-2 and P1-1 with IC50 values ranging from 0.04 to 5

!M. GSTs are often overexpressed in drug-resistant cell lines including cancer cells.

Elevated GST activity is regarded as an indicator for the resistance to chemotherapy

[58]. In addition, multidrug resistance has been associated with a decrease in

intracellular drug accumulation in patient tumour cells due to enhanced drug efflux or

enhanced metabolism through GSTs. The inhibition of GST by curcumin therefore

renders it a promising anticancer agent [58]. On the other hand inhibition of GSTs

reduces their protective role in detoxification of electrophilic substances through

glutathione (GSH) conjugation [59]. Thus prolonged GST inhibition could also result in

toxicity of electrophilic chemicals or metabolites [60].

Page 68: thuoc từ curcumin

58

Investigation of the effect of three series of curcumin analogues designed and

synthesized by Sardjiman et al [61] on GSTA1-1, GSTM1-1, GSTP1-1 and human and

rat cytosolic GSTs revealed a variety of GST activities in the presences of varying

chemical structures [in this thesis Chapter 5]. Seven of the thirty-four compounds were

more potent inhibitors of GSTA1-1, whilst three and four compounds respectively, were

more potent inhibitors of GSTM1-1 and GSTP1-1 than curcumin itself. One of the three

series of curcumin analogues lacked inhibitory activity towards GSTP1-1. Although the

strong inhibitory activities of curcumin and some of its analogues towards GSTs could

have useful applications in cancer chemotherapy, these activities could also have

implications for toxicity in normal cells in the presence of potentially toxic compounds

with electrophilic properties, and/or electrophilic metabolite, since GSTs are major

detoxification enzymes in the body. Knowledge on the structure-activity relationship can

be useful in the designing of curcumin analogues with less or more GST-inhibitory

properties. The clinical relevance of these inhibitory activities, also in view of the organ

and tissue distribution (for example GSTP1-1 is primarily found in erythrocyte and not

the liver [62], as well as polymorphisms of GSTs, remains to be established.

5. Conclusion

This review has focused on the pharmacokinetics, metabolism and drug-drug interaction

potential of curcumin, all aspects of potential value in human applications of curcumin.

The extremely poor bioavailability and subsequent high exposure of the intestinal

mucosa to curcumin support the clinical evidence of its potential as therapeutic agent

for colorectal cancer. In line with this observation, orally administered curcumin is not

likely to become clinically useful in prevention of tumours distant from the locus of

absorption, nor other clinical applications for which systemic availability is needed.

However, local administration may be required where higher concentrations are

necessary in systemic, target organ or tissues. On the other hand, curcumin

formulations resulting in enhanced bioavailability are more promising alternatives. Apart

from significant phase II metabolism notably, glucuronidation, sulfation and GSH

conjugation, limited phase I metabolism notably, reductive and minor oxidative

metabolism have been demonstrated. The pharmacological significance of curcumin

Page 69: thuoc từ curcumin

59

metabolism and its metabolites need to be further investigated. The same holds true for

easily occurring decomposition products of curcumin. Knowledge of structure activity

relationships of curcumin analogues may be useful in redesigning analogues with less

potential for drug-drug interactions. Drug-food interactions have the potential to cause

harmful effects. Therefore, a rational approach is required to screen curcumin

formulations and foods for in vitro CYP inhibitory activities, since curcumin itself has

been shown to inhibit human important CYPs. The mechanisms underlying the

interactions between several enzymes and transporters with the properties of curcumin

also warrant further investigations. The strength of curcumin is also its weakness, thus

findings on the pharmacokinetics, metabolism and potential for drug-drug interactions

need to be considered for a more useful application of the compound.

References

1. Singh, S., Khar, A., 2006. Biological effects of curcumin and its role in cancerchemoprevention and therapy. Anti-Cancer Agents Med Chem 6:259-270.

2. Donatus, I.A,, Sardjoko, Vermeulen, N.P., 1990. Cytotoxic and cytoprotective activities curcumin. Effects on paracetamol-induced cytotoxicity, lipid peroxidation and glutathione depletion in rat hepatocytes. Biochem Pharmacol 39:1869-1875.

3. Leu, T-H., Maa, M-C., 2002. The molecular mechanisms for the antitumorigenic effect of curcumin. Curr Med Chem Anti-Cancer Agents 2:357-370.

4. Cole, G.M., Morihara, T., Lim, G.P., Yang, F., Begum, A., Frautschy, S.A., 2004. NSAID and antioxidant prevention of Alzheimer's disease: Lessons from in vitro and animal models. Annals NY Acad Sci 1035:68-84.

5. Vajragupta, O., Boonchoong, P., Morris, G.M., Olson, A.J., 2005. Active site binding modes of curcumin in HIV-1 protease and integrase. Bioorg Med Chem Letters 15:3364-3368.

6. Reddy, R.C., Vatsaala, P.G., Keshamouni, V.G., Padmanaban, G., Rangarajan, P.N., 2005. Curcumin for malaria therapy. Biochem Biophy Res Comm 326:472-474.

7. Aggarwal, B.B., Kumar, A., Bharti, A.C., 2003. Anticancer potential of curcumin: preclinical and clinical studies. Anticancer Res 23:363-398.

8. Cheng, A.L., Hsu, C.H., Lin, J.K., Hsu, M.M., Ho, Y.F., Shen, T.S., et al., 2001. Phase I clinical trial of curcumin, a chemopreventive agent, in patients with high-risk or pre-malignant lesions. Anticancer Res 21:2895-2900.

9. Sharma, R.A., Euden, S.A., Platton, S.L., Cooke, D.N., Shafayat, A., Hewitt, H.R., et al., 2004. Phase I clinical trial of oral curcumin: biomarkers of systemic activity and compliance. Clin Cancer Res 10:6847-6854.

10. Oetari, S., Sudibyo, M., Commandeur, J.N.M., Samhoedi, R., Vermeulen, N.P., 1996. Effects of curcumin on cytochrome P450 and glutathione S-transferase activities in rat liver. Biochem Pharmacol 51:39-45. 11. Ireson, C., Orr, S., Jones, D.J., Verschoyle, R., Lim, C.K., Luo, J.L., et al., 2001. Characterization of

metabolites of the chemopreventive agent curcumin in human and rat hepatocytes and in the rat in vivo, and evaluation of their ability to inhibit phorbol ester-induced prostaglandin E2 production. Cancer Res 61:1058-1064.

Page 70: thuoc từ curcumin

60

12. Sharma, R.A., McLelland, H.R., Hill, K.A., Ireson, C.R., Euden, S.A., Manson, M.M., et al., 2001. Pharmacodynamic and pharmacokinetic study of oral curcuma extract in patients with colorectal cancer. Clin Cancer Res 7:1894-1900.

13. 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.

14. Appiah-Opong, R., Commandeur, J.N.M., van Vugt-Lussenburg, B., Vermeulen, N.P.E., 2007. Inhibition of human recombinant cytochrome P450s by curcumin and curcumin decomposition products. Toxicology 235:83-91.

15. Marczylo, T.H., Verschoyle, R.D., Cooke, D.N., Morazzoni, P., Steward, W.P., Gescher, A.J., 2007. Comparison of systemic availability of curcumin with that of curcumin formulated with phosphatidylcholine. Cancer Chemother Pharmacol 60:171-177.

16. Anand, P., Kunnumakkara, A.B., Newman, R.A., Aggarwal, B.B., 2007. Bioavailability of curcumin: Problems and promises. Mol Pharmaceut DOI: 10.1021/mp700113r

17. Asai, A ., Miyazawa, T., 2000. Occurrence of orally administered curcuminoid as glucuronide and glucuronide/sulfate conjugates in rat plasma. Life Sci 67:2785-2793. 18. Ireson, C.R., Jones, D.J., Orr, S., Coughtrie, M.W., Boocock, D.J., Williams, M.L., Farmer, P.B.,

Steward, W.P., Gescher, A.J., 2002. Metabolism of the cancer chemopreventive agent curcumin in human and rat intestine. Cancer Epidemiol Biomarkers Prev 11:105-111.

19. Hoehle, S.I., Pfeiffer, E., Solyom, A.M., Metzler, M., 2006. Metabolism of curcuminoids in tissue slices and subcellular fractions from rat liver. J Agric Food Chem 54:756-764.

20. Allen, S.W., Mueller, L., Williams, S.N., Quattrochi, L.C., Raucy, J., 2001. The use of a high-volume screening procedure to assess the effects of dietary flavonoids on human CYP1A1 expression. Drug Metab Dispos 29:1074-1079.

21. Perkins, S., Verschoyle, R.D., Hill, K., Parveen, Threadgill, M.D., Sharma, R.A., Williams, M.I., Steward, W.P., Gescher, A.J., 2002. Chemopreventive efficacy and pharmacokinetics of curcumin in the Min/+ mouse, a model of familial adenomatous polyposis. Cancer Epid Biomarkers & prevention 11:535-540.

22. Shoba, G., Joy, D., Joseph, T., Majeed, M., 1998. Influence of piperine on the pharmacokinetics of curcumin in animals and human volunteers. Planta Med 64:353-356.

23. Rasyid, A., Lelo, A., 1999. The effect of curcumin and placebo on human gall-bladder function: an ultrasound study. Alimentary Pharmacol Ther 13:245-249. 24. Lao, C.D., Ruffin, M.T., Normolle, D., Heath, D.D., Murray, S.I., Bailey, J.M., et al., 2006. Dose

escalation of a curcuminoid formulation. Complement Altern Med 6:10. 25. Garcea, G., Berry, D.P., Jones, D.J., Singh, R., Dennison, A.R., Farmer, P.B., Sharma, R.A.,

Steward, W.P., Gescher, A.J., 2005. Consumption of the putative chemopreventive agent curcumin by cancer patients: assessment of curcmin levels in the colorectum and their pharmacodynamic consequences. Cancer Epidermiol Biomarkers Prev 14:120-125.

26. Maiti, K., Mukherjee, K., Gantait, A., Saha, B.P., Mukherjee, P.K., 2007. Curcumin-phospholipid complex; preparation therapeutic evaluation and pharmacokinetics in rats. Int J Pharm 330:155-163.

27. Ma, Z., Shayeganpour, A., Brocks, D.R., Lavasanifar, A., Samuel, J., 2007. High performance liquid chromatography analysis of curcumin in rat plasma: application to pharmacokinetics of polymeric micellar formulation of curcumin. Biomed Chromatogr 21:546-552.

28. Safavy, A., Raisch, K.P., Mantena, S., Sanford, L.L., Sham, S.W., Krishna, N.R., Bonner, J.A., 2007. Design and development of water-soluble curcumin conjugates as potential anticancer agents. J Med Chem 50:6284-6288.

29. Li, L., Ahmed, B., Mehta, K., Kurzrock, R., 2007. Liposomal curcumin with and without oxaliplatin: effects on cell growth, apoptosis, and angiogenesis in colorectal cancer. Mol Cancer Ther 6:1276-1282. 30. Bisht, S., Feldmann, G., Soni, S., Ravi, R., Karikar, C., Maitra, A., Maitra, A., 2007. Polymeric nanoparticle-encapsulated curcumin (“nanocurcumin”): a novel strategy for human cancer therapy. J Nanobiotechnology 5:3. 31. Preetha, A., Banerjee, R., Huilgol, N., 2007. Tensiometric profiles and their modulation by cholesterol: implications in cervical cancer. Cancer Invest 25:172-181.

Page 71: thuoc từ curcumin

61

32. Mosley, C.A., Liotta, D.C., Snyder, J.P., 2007. Highly active anticancer curcumin analogues. Adv Exp Med Biol 595:77-103. 33. Inano, H., Onoda, M., 2002. Prevention of radiation-induced mammary tumours. Int J Radiat Oncol

52:212-223. 34. Frank, N., Knauft, J., Amelung, F., Nair, J., Wesch, H., Bartsch, H., 2003. No prevention of liver and

kidney tumours in Long-Evans Cinnamon rats by dietary curcumin, but inhibition at other sites and of metastases. Mutat Res 127-135:523-524.

35. Nanji, A.A., Jokelainen, K., Tipoe, G.L., Rahemtulla, A., Thomas, P., Dannenberg, A.J., 2003. Curcumin prevents alcohol-induced liver disease in rats by inhibiting the expression of NF-kappa B-dependent genes. Am J Physiol Gastrointest Liver Physiol 284:G321-G327.

36. Mukhopadhyay, A., Basu, N., Ghatak, N., Gujral, P.K., 1982. Anti-inflammatory and irritant activities of curcumin analogues in rats. Agents Actions 12:508-515.

37. Okada, K., Wangpoentrakul, C., Tanaka, T., Toyokuni, S., Uchida, K., Osawa, T., 2001. Curcumin and especially tetrahydrocurcumin ameliorate oxidative stress-induced renal injury in mice. J Nutri 131:2090-2095.

38. Ravindranath, V., Chandrasekhara, N., 1981. In vitro studies on the intestinal absorption of curcumin in rats. Toxicology 20:251-257.

39. Garcea, G., Jones, D.J., Singh, R., Dennison, A.R., Farmer, P.B., Sharma, R.A., Steward, W.P., Gescher, A.J., Berry, D.P., 2004. Detection of curcumin metabolites in hepatic tissues and portal blood of patients following oral administration. Br J Cancer 90:1011-1015.

40. Tamvakopoulos, C., Sofianos, Z.D., Garbis, S.D., Pantazis, P., 2007. Analysis of the in vitro metabolites of diferuloylmethane (curcumin) by liquid chromatography – tandem mass spectrometry on a hybrid quadrupole linear ion trap system: newly identified metabolites. Eur J Drug Metab Pharmacokinet 32:51-57.

41. Hoehle, S.I., Pfeiffer, E., Metzler, M., 2007. Glucuronidation of curcuminoids by human microsomal and recombinant UDP-glucuronosyltransferases. Mol Nutr Food Res 51:932-938.

42 Pfeiffer, E., Hoehle, S.I., Walch, S.G., Riess, A., Solyom, A.M., Metzler, M., 2007. Curcuminoids form reactive glucuronides in vitro. J Agric Food Chem 55:538-544.

43. Ebner, T., Burchell, B., 1993. Substrate specificies of two stably expressed human liver UDP-glucuronosyltransferases of the UGT1 gene family. Drug Metab Dispos 21:50-55.

44. Riley, R.J., Kenna, J.G., 2004. Cellular models for ADMET predictions and evaluation of drug-drug interactions. Curr Opin Drug Discov Devel 7:86-99.

45. Brown, H.S., Galetin, A., Hallifax, D., Houston, J.B., 2006. Prediction if in vivo drug-drug interactions from in vitro data: Factors affecting prototypic drug-drug interactions involving CYP2C9, CYP2D6 and CYP3A4. Clin Pharmacokinet 45:1035-1050.

46. Pea, F., Furlanut, M., 2001. Pharmacokinetic aspects of treating infections in the intensive care unit: focus on drug interactions. Clin Pharmacokinet 40:833-868.

47. Mandlekar, S., Hong, J.L., Kong, A.N., 2006. Modulation of metabolic enzymes by dietary phytochemicals: a review of mechanisms underlying beneficial versus unfavourable effects. Curr Drug Metab 7:661-675.

48. Harris, R.Z., Jang, G.R., Tsunda, S. 2003. Dietary effects on drug metabolism and transport. Clin Pharmacokinet 42:1071-1088.

49. Bailey, D.G., Dresser, G.K., Bend, J.R., 2003. Bergamottin, lime juice, and red wine as inhibitors of cytochrome P450 3A4 activity: Comparison with grapefruit juice. Clin Pharmacol Ther 73:529- 537. 50. Jang, E.H., Park, Y.C., Chung, W.G., 2004. Effects of dietary suppliments on induction and

inhibition of cytochrome P450s protein expression in rats. Food Chem Toxicol 42:1749-1756. 51. Obach, R.S., 2000. Inhibition of human cytochrome P450 enzymes by constituents of St. John's

wort, an herbal preparation used in the treatment of depression. J Pharmacol Exp Ther 294:88-95. 52. Komoroski, B.J., Zhang, S., Cai, H., Hutzler, J.M., Frye, R., Tracy, T.S., et al., 2004. Induction and

inhibition of cytochromes P450 by the St. John's wort constituent hyperforin in human hepatocyte cultures. Drug Metab Dispos 32:512-518.

53. Zafar, A., Sharif, M.D., 2003. Pharmacokinetics, metabolism, and metabolism of atypical antipsychotics in special populations. Primary care companion J Clin Psychiatry 5:22-25. 54. Eagling, V.A., Tjia, J.F., Back, D.J., 1998. Differential selectivity of cytochrome P450 inhibitors

against probe substrates in human and rat liver microsomes. Br J Clin Pharmacol 45:107-114.

Page 72: thuoc từ curcumin

62

55. Anuchapreeda, S., Leechanacha, P., Smith, M.M., Ambudkar, S.V., Limtrakul, P.N., 2002. Modulation of P-glycoprotein expression and function by curcumin in multi-drug resistant human KB cells. Biochem Pharmacol 64:573-582.

56. Limtrakul, P., Khantamat, O., Pintha, K., 2004. Inhibition of P-glycoprotein activity and reversal of cancer multi-drug resistance by Momordica charantia extracts. Cancer Chemother Pharmacol 54:525-530. 57. Achira, M., Suzuki, H., Ito, K., Sugiyama, Y., 2001. Comparative studies to determine the selective

inhibitors for P-glycoprotein and cytochrome P450 3A4. AAPS Pharm Sci 3:18 DOI: 10.1208/ps030218.

58. Hayeshi, R., Mutingwende, I., Mavengere, W., Masiyanise, V., Mukanganyama, S., 2007. Inhibition of human glutathione S-transferases activity by plant polyphenolic compounds ellagic acid and curcumin. Food Chem Toxicol 45:286-295.

59. Commandeur, J.N., Stijntjes, G.J., Vermeulen, N.P., 1995. Enzymes and transport systems involved in the formation and disposition of glutathione S-conjugates. Role in bioactivation and detoxication mechanisms of xenobiotics. Pharmacol Rev 47:271-330.

60. DeLeve, L.D., Wang, X., 2000. Role of oxidative stress glutathione in busulfan toxicity in cultured murine hepatocytes. Pharmacology 60:143-154.

61. Sardjiman, S., Reksohadiprodjo, M., Hakim, L., van der Goot, H., Timmerman, H., 1997. 1,5- Diphenyl-1,4-pentadiene-3-ones and cyclic analogues as antioxidative agents. Synthesis and structure-activity relationship. Eur J Med Chem 32:625-636. 62. Awashti, Y.C., Sharma, R., Singhal, S.S., 1994. Human glutathione S-transferases. Int J Biochem

26:295-308.

Page 73: thuoc từ curcumin

63

Inhibition of CYP/GST activities by natural products and derivatives

Page 74: thuoc từ curcumin

64

Page 75: thuoc từ curcumin

65

Chapter 3

Inhibition of human recombinant cytochrome P450s by curcumin and

curcumin decomposition products

Regina Appiah-Opong, Jan N. M. Commandeur, Barbara van Vugt-

Lussenburg, and Nico P. E. Vermeulen

Adapted from Toxicology 2007 235:83-91

Curcumin (diferuloylmethane) is a major yellow pigment and dietary component derived

from Curcuma longa. It has potent anti-inflammatory, anti-carcinogenic, anti-oxidant and

chemoprotective activities among others. We studied the interactions of curcumin, a

mixture of its decomposition products, and four of its individually identified

decomposition products (vanillin, vanillic acid, ferulic aldehyde and ferulic acid) on five

major human drug metabolizing cytochrome P450s (CYPs). Curcumin inhibited

CYP1A2 (IC50, 40.0 µM), CYP3A4 (IC50, 16.3 µM), CYP2D6 (IC50, 50.3 µM), CYP2C9

(IC50, 4.3 µM) and CYP2B6 (IC50, 24.5 µM). Curcumin showed a competitive type of

inhibition towards CYP1A2, CYP3A4 and CYP2B6, whereas a non-competitive type of

inhibition was observed with respect to CYP2D6 and CYP2C9. The inhibitory activity

towards CYP3A4, shown by curcumin may have implications for drug-drug interactions

in the intestines, in case of high exposure of the intestines to curcumin upon oral

administration. In spite of the significant inhibitory activities shown towards the major

CYPs in vitro, it remains to be established, whether curcumin will cause significant drug-

drug interactions in the liver, given the reported low systemic exposure of the liver to

curcumin. The decomposition products of curcumin showed no significant inhibitory

activities towards the CYPs investigated, and therefore, are not likely to cause drug-

drug interactions at the level of CYPs.

Page 76: thuoc từ curcumin

66

1. Introduction

Multiple drug therapy is a common therapeutic practice, especially in patients with

multiple complications [1,2]. If two or more drugs with affinity for the same CYP enzyme

are co-administered, their biotransformation may be compromised, leading to

undesirable accumulation of the drugs with toxic side-effects as possible consequence.

Drug-drug interactions involving CYPs have been identified as an important cause of

adverse drug reactions and therapeutic failure [3,4]. Drug-drug interactions may be due

to enzyme induction or inhibition, the latter being more common [2,5]. Next to drugs,

several natural compounds have also been shown to cause significant interactions at

the level of drug-metabolizing enzymes [6,7].

Curcumin, a polyphenolic component of turmeric (Curcuma longa), is a yellow

pigment widely used for coloring of foods. It has been shown previously to exhibit anti-

cancer, anti-oxidant, anti-inflammatory, anti-parasitic and anti-HIV properties [8-11].

Curcumin has also been shown to have chemoprotective, chemopreventive and

immuno-modulating properties [9,12,13]. Curcumin can be considered as a safe

compound, because oral doses as high as 8 g/day administered to humans did not

result in overt side effects [13]. Clinical trials for the use of curcumin as an anti-cancer

agent are currently ongoing [14].

Because relatively high doses of curcumin are evaluated in human studies, it

might be anticipated that curcumin might cause drug-drug interactions at the level of

intestinal and/or liver drug metabolism. Several in vivo and in vitro animal studies have

shown that curcumin can significantly modulate the activity of several drug metabolizing

enzymes by down-regulation, induction or by direct inhibition. Oetari et al [15] and

Thapliyal and Maru [16] reported potent inhibition of rat liver microsomal CYP1A1,

CYP1A2 and CYP2B1 enzymes by curcumin. In in vivo studies repetitive administration

of curcumin to rats resulted in down-regulation of intestinal CYP3A-enzymes, whereas

hepatic and renal CYP3A-levels were significantly induced [17]. Also, down-regulation

of esophagal CYP2B1 and CYP2E1 was reported after intragastric treatment of rats,

Page 77: thuoc từ curcumin

67

which might partially explain the chemopreventive activity of curcumin against

carcinogenic N-nitrosamines [18].

As yet, the effects of curcumin on the major human drug metabolizing CYPs

have not been studied. Due to the large species differences in the properties of

metabolic enzymes and metabolic profiles of drugs, the animal studies described above

are poorly predictive for the human situation [19,20]. The present in vitro investigation

therefore was designed to assess the potential of curcumin to cause drug-drug

interactions via inhibition of the five most important human drug metabolizing CYPs.

Major human CYP isoforms responsible for the metabolism and disposition of about

90% of the therapeutic drugs on the market include CYP1A2, CYP2D6, CYP2B6,

CYP2E1, CYP2C9 and CYP3A4 [21].

Figure 1. Chemical structures of curcumin and its decomposition products at pH 7.4

Because curcumin has been shown to be chemically unstable under neutral and

alkaline conditions [22], the inhibitory properties of these decomposition products were

also studied as mixture and individually, if available. Degradation products which have

been identified include trans-6(4’-hydroxy-3’-methoxyphenyl)-2,4-dioxo-5-hexenal, and

O O

CHO COOH

OH

CH CHCOOHCH CHCHO

OH OHOH

OHOH

HO

O

O CHOH

OCH3

OCH3OCH

3OCH3

OCH3

CH3

O

CH3

O

Curcumin

Vanillin Vanillic acid Ferulic acidFerulic aldehyde

Trans-6-(4'-hydroxy-3'-methoxyphenyl)-4-dioxo-5-hexenal

Page 78: thuoc từ curcumin

68

minor products being vanillin, vanillic acid, ferulic aldehyde and ferulic acid (Figure 1)

[22]. Apart from reversible inhibition, mechanism-based inhibition was also taken into

consideration.

2. Materials and methods

2.1. Materials

Methoxyresorufin and benzyloxyresorufin were synthesized by the method of Burke et

al [23] and the purity was determined by HPLC, mass spectrometry and 1H NMR. The

plasmid, pSP19T7LT_2D6 containing human CYP2D6 bicistronically co-expressed with

human cytochrome P450 NADPH reductase was kindly provided by Prof. M. Ingelman-

Sundberg (Stockholm, Sweden). The plasmids, BMX100/h1A2 and pCWh3A4 with

human cytochrome P450 NADPH reductase were kindly donated by Dr. M. Kranendonk

(Lisbon, Portugal). Expression plasmids, pCWh2B6hNPR and pCWh2C9hNPR with

human cytochrome P450 NADPH reductase were kindly provided by Prof. F.P.

Guengerich (Nashville, Texas, USA). All other chemicals were of analytical grade and

obtained from standard suppliers.

2.1. CYP expression and membrane isolation

The plasmids containing cDNA of five human CYPs were transformed into Escherichia

coli strain JM109. Expression of the CYPs was carried out in 3-litre flasks containing

300 ml terrific broth (TB) with 1mM !-aminolevulinic acid, 0.5 mM thiamine, 400 µl/L

trace elements, 100 µg/ml ampicillin, 1 mM isopropyl-"-D-thiogalactopyranoside (IPTG),

0.5 mM FeCl3 (for CYP2D6 and CYP3A4 only) and 30 µg/ml kanamycin (for CYP3A4

only). The culture media were inoculated with 3 ml overnight cultures of bacteria

containing plasmids for the various CYPs. The cell cultures were incubated for about 40

h at 28 oC and 125 rpm and CYP contents were determined using the carbon monoxide

(CO) difference spectra as described by Omura and Sato [24]. Cells were pelleted by

centrifugation (4000 g, 4 oC, 15 min) and resuspended in 30 ml Tris-Sucrose-EDTA

(TSE) buffer (50 mM Tris-acetate buffer pH 7.6, 250 mM sucrose, 0.25 mM EDTA).

Cells were treated with 0.5 mg/ml lysozyme prior to disruption by French press (1000

psi, 3 repeats). The membranes containing the human CYPs were isolated by

Page 79: thuoc từ curcumin

69

ultracentrifugation in a Beckmann 50.2Ti rotor (60 min, 40,000 rpm, 4 oC), resuspended

in TSE buffer and stored at –80 oC until use.

2.2. Decomposition of curcumin

Curcumin decomposition was performed according to the method of Wang et al [22] in

phosphate buffer of pH 7.4. Briefly, aliquots of 20 µl of 5 mM curcumin (dissolved in

methanol) were added to 980 µl of 0.1 M potassium phosphate buffer pH 7.4. After 1 h

incubation at 37 oC Samples were analyzed by HPLC (Jasco Separations FP 1575).

Vanillin, vanillic acid, ferulic aldehyde and ferulic acid were used as standards. Gradient

reversed-phase HPLC separations were performed using a C18 column (150 mm x 3.2

mm, 5 µm particle size, Phenomenex) and a mobile phase consisting of 0.1% acetic

acid (solvent A) and 98% methanol with 0.1% acetic acid (solvent B). HPLC-gradient

separation was carried out with a linear gradient eluting from 15% to 95% of solvent B in

60 min. The carrier flow rate was 0.4 ml/min and chromatographic peaks of

decomposition products were monitored by UV detection (! = 280 nm). Under these

chromatographic conditions curcumin and commercially available reference compounds

of possible degradation products eluted at: 28.5 min (curcumin), 14.3 min (vanillic acid),

15.2 min (vanillin), and 17.9 min (ferulic acid) 18.5 min (ferulic aldehyde).

2.3. CYP inhibition assays

2.3.1. 7-Methoxyresorufin, 7-benzyloxyresorufin, dibenzylfluorescein, 7-benzyloxy-4-

trifluo-romethylcoumarin and 7-benzyloxyquinoline O-dealkylation

Inhibition of the activities of the human CYP isoforms 1A2, 3A4 and 2B6, by curcumin

and its decomposition products was determined using microplate fluorimetric assays

[23,25]. Incubation conditions (eg. enzyme concentration, substrates, incubation time)

and wavelengths for detection for each assay are shown in Table 1. Under these

conditions kinetics of the reactions were linear over the periods indicated. The inhibitory

activity of curcumin towards CYP3A4 was tested using four different substrates of

CYP3A4, namely DBF, BFC and BQ, because inhibition of CYP3A4 activity is known to

be substrate-dependent [25]. In general, the incubations were carried out in a total

Page 80: thuoc từ curcumin

70

volume of 200 µl and in the presence of 100 µM NADPH (freshly prepared) in a black

coaster 96-well plate. Membranes were pre-incubated for 5 min at 37 oC with 0.1 M

potassium phosphate buffer (pH 7.4), substrates, and inhibitors (curcumin and its

decomposition products) with DMSO in concentration 0.5% (v/v) or less in all the CYP

assays.

Table 1. Experimental conditions for fluorescence CYP assays

CYP Enzyme Incubation Substrate Substrate Excitation Emission

amount time conc wavelength wavelength

nM min µM nm

1A2 13.2 10 MRes 5.0 530 586

3A4 14.3 30 BRes 5.0 530 586

2.8 10 DBF 0.5 485 535

14.3 20 BFC 80.0 410 538

14.3 30 BQ 40.0 410 538

2B6 15.3 30 BRes 20.0 530 586

MRes, methoxyresorufin; BRes, benzyloxyresorufin; DBF, dibenzylfluorescein; BQ, 7-benzyloxyquinoline;

BFC, 7-benzyloxy-4-trifluoromethylcoumarin

DMSO concentration was consistent in each assay. At these DMSO concentrations, the

studied human CYPs are not affected [26,27]. Stability of curcumin in phosphate buffer

pH 7.4, has been reported to be strongly improved by addition of rat liver microsomes or

cytosol, glutathione (GSH), N-acetyl L-cysteine (NAC) or ascorbic acid [15]. Therefore

enzymes were always added before the incorporation of curcumin in all inhibition

assays performed.

Initially, the inhibitory effect of curcumin and its decomposition products on the

CYP isoforms was studied at a concentration of 300 !M. For IC50 determinations,

concentration ranges for curcumin used were from 0.9 to 100 µM and for the

decomposition products, from 2.5 to 1000 µM. Incubations were commenced by the

addition of 100 µM NADPH, maintained at 37 oC for the periods defined. Reactions

Page 81: thuoc từ curcumin

71

were terminated with 75 µl of 80% acetonitrile and 20% 0.5 M Tris solution or 2 N NaOH

in the case of dibenzylfluorescein (DBF). Product formation was linear for all incubation

times. Concentrations of the probe substrates in all reaction mixtures were chosen near

the Michaelis-Menten’s (Km) value for each of the CYPs tested. The Km values obtained

using the alkoxyresorufins and all other substrates were within the range of reported

literature values [28]. All measurements were performed in triplicate. Metabolite

formation was measured spectrophotometrically on a Victor2 1420 multilabel counter.

2.3.2. Diclofenac hydroxylation

For the CYP2C9 inhibition assay, reaction mixtures in 500 µl total volume consisted of

49 nM enzyme, 100 µM NADPH, 0.1 M potassium phosphate buffer (pH 7.4), 6 µM

diclofenac and inhibitor. Screening for inhibitory effects of 300 µM concentrations of

curcumin, its decomposition mixture and the individual decomposition products on the

CYP2C9 was done. For IC50 determinations, curcumin concentrations used were of the

range 0.4 to 100 µM and 3.9 to 2000 µM for the individual decomposition products,

vanillin, vanillic acid, ferulic aldehyde and ferulic acid. After preincubation for 5 min at 37

oC, reactions were initiated by adding NADPH and terminated after 10 min with the

addition of 200 !l methanol. The reaction mixtures were centrifuged at 14,000 rpm for 3

min. Product formed was measured using an isocratic HPLC method [29], and was

linear in 10 min. A C18 column (150 mm x 3.2 mm, 5 µm particle size, Phenomenex)

was used and the carrier flow rate was 0.6 ml/min. The mobile phase consisted of 60%

(v/v) 20 mM potassium phosphate buffer (pH 7.4), 22.5% (v/v) methanol and 17.5%

(v/v) acetonitrile. Peaks were monitored at the wavelength of 280 nm. Retention times

for 4-hydroxydiclofenac and diclofenac were 5.0 and 24.1 min, respectively.

2.3.3. Dextromethorphan O-demethylation

Inhibition of CYP2D6 activity by curcumin and its decomposition products was

evaluated by the method described by Ko et al [30]. Inhibitory effects of 300 µM

concentrations of curcumin, its decomposition mixture and the individual decomposition

products on the CYP2D6 were first assessed. The reaction mixture had a total volume

of 500 µl and consisted of 18.2 nM enzyme, 4.5 µM dextromethorphan, 90.9 µM

Page 82: thuoc từ curcumin

72

NADPH, 0.1 M potassium phosphate buffer and inhibitor. For IC50 determination,

curcumin concentration range used was 0.4 to 181.8 µM and the decomposition

products 3.6 to 1818.2 µM. Reactions were initiated by the addition of NADPH and

allowed to proceed for 45 min before termination with the addition of 60 mM zinc

sulphate solution. Product formed was measured using an isocratic HPLC fluorescence

detection method and a C18 column (100 mm x 3 mm, 5 µm. particle size, Chromspher)

and was linear in 45 min. The mobile phase consisted of 24% (v/v) acetonitrile and

0.1% (v/v) triethylamine adjusted to pH 3 with perchloric acid. The carrier flow rate was

0.6 ml/min. Peaks were monitored at 280 nm (excitation) and 310 nm (emission). The

retention times of dextrorphan and dextromethorphan were 3.4 and 24.5 min,

respectively.

2.4. Type of inhibition

To determine the types of inhibition occurring in the reactions involving CYP1A2 and

CYP3A4, the substrate concentrations used were ranging from 0.65 to 10 µM for both

methoxyresorufin and benzyloxyresorufin. In the case of the other CYPs, substrate

concentration ranges were from 3.1 to 50 µM benzyloxyresorufin (CYP2B6), 0.3 to 20

µM diclofenac (CYP2C9) and 1.4 to 22.7 µM dextromethorphan (CYP2D6). Five

different substrate concentrations were used in each assay. The concentration of

curcumin used for the assays are indicated in Table 3. Reactions were carried out as

described above for all CYPs.

2.5. Mechanism-based inhibition

The potential of curcumin for mechanism-based inhibition of CYP1A2, CYP3A4 and

CYP2B6 was evaluated according to the method of Heydari et al [31] with slight

modifications. Briefly preincubation mixtures of total volumes 600 µl contained 13 to 16

nM CYP enzymes, 100 µM NADPH and curcumin solution (0, 10 and 50 µM). The

preincubations were performed for 20 min at 37 oC, and at 5 min intervals 100 µl

aliquots were taken to determine the remaining CYP activity. The aliquots of

preincubation mixtures were added to tubes containing 400 µl of the respective

substrates (5 µM methoxyresorufin for CYP1A2, 5 µM benzyloxyresorufin for CYP3A4

Page 83: thuoc từ curcumin

73

and 20 µM benzyloxyresorufin for CYP2B6) and NADPH (100 µM), and incubated as

described above. After stopping the reactions the remaining activities were determined

using a fluorescence spectrophotometer (Perkin Elmer Model 3000). Duplicate

experiments were performed.

For CYP2D6 and CYP2C9, preincubation tubes contained 40-49 nM CYP

enzyme, 100 µM NADPH, 50 µM and 20 µM curcumin solution. Preincubations were

performed for 20 min at 37 oC and at time intervals of 5 min the remaining activities of

the enzymes were re-assessed. Aliquots (100 µl) of the 600 µl preincubation mixtures

were transferred into tubes containing 400µl dextromethorphan (4.5 µM) or diclofenac

(6.0 µM) and NADPH (100 µM) and incubated at 37 oC for 30 min or 10 min to

determine remaining activity of CYP2D6 or CYP2C9, respectively.

2.6. Data analysis

Percent inhibition of CYP activity by curcumin, its decomposition products were

calculated from the ratio of the activity of treated to control samples. Statistical analysis

was performed using the Student t test. The enzyme kinetic parameters (Km and Vmax)

for metabolism of the various substrates and IC50 values were analyzed using

GraphPad Prism 4.0 version (GraphPad Prism software Inc. San Diego CA). The

inhibitor constant (Ki) values for competitive inhibition were calculated from according to

the following equation: for competitive inhibition, Ki = Km (inhibited) [I]/ Km (uninhibited) -

Km(inhibited) where substrate concentration is equal to the Km, and for non-competitive

inhibition, K i = Vmax (inhibited) [I]/ Vmax (uninhibited) - Vmax (inhibited) where K m, Vmax, S and [I] are

Michaelis constant, maximal enzyme activity, substrate concentration and inhibitor

concentration, respectively.

3. Results

3.1. Decomposition of curcumin

Degradation of curcumin under various pH conditions and the stability of curcumin in

physiological matrices have been previously reported [22]. To identify the inhibitory

potentials of the decomposition products of curcumin towards human CYPs,

decomposition experiments were also performed in the present study. Curcumin was

Page 84: thuoc từ curcumin

74

treated with 0.1 M phosphate buffer of pH 7.4 at 37 oC for 1 h, according to the

procedure described by Wang et al [22]. Figure 2 shows the HPLC-chromatogram of the

resulting mixture of decomposition products of curcumin.

Figure 2. HPLC chromatogram of decomposition products of curcumin at pH 7.4: V, Vac,

Fac, Fal and major product (vanillin, vanillic acid, ferulic acid, ferulic aldehyde and trans-6-(4’-hydroxy-3-

methoxyphenyl)-2,4-dioxo-5-hexenal with retention time 15.2, 14.3, 17.4, 18.5 and 16.9 min respectively).

Chromatographic peaks observed included a major peak and seven minor peaks, four

of which were identified as vanillin, vanillic acid, ferulic aldehyde and ferulic acid, by co-

eluting with commercially available reference compounds [data not shown]. This major

decomposition product with retention time of 16.9 min most likely represents trans-6-(4'-

hydroxy-3'-methoxyphenyl)-4-dioxo-5-hexenal, as the same procedure of decomposition

was used as that described by Wang et al [22].

3.2. CYP inhibition and types of inhibition

The inhibitory effects of 300 µM concentrations of curcumin, four of the individual

decomposition products and a complete mixture of decomposition products, on the

Page 85: thuoc từ curcumin

75

activities of CYP-1A2, -3A4, -2D6, -2C9, and -2B6 are shown in Figure 3. At this

concentration, curcumin appeared to inhibit almost completely the activities of CYP3A4,

CYP2C9, and CYP1A2, and 72% and 69.1% of the activities of CYP2D6 and CYP2B6,

respectively. The complete decomposition mixture and the four decomposition products,

vanillin, vanillic acid, ferulic aldehyde and ferulic acid, showed milder inhibitory effects

on the above CYPs. Percentages of inhibition of CYP activities in the range 1 - 50%

were observed in assays with the decomposition products (Figure 3). The

decomposition mixture caused 57.4% and 74.8% inhibition of CYP3A4 and CYP2C9

respectively. For the decomposition products showing more than 50% inhibition of

enzyme activity at 300 µM, the IC50 values were determined (Table 2).

Figure 3. Inhibition of CYP3A4 (A), CYP1A2 (B), CYP2B6 (C), CYP2C9 (D) and CYP2D6 (E) activities by

curcumin, four of its decomposition products (each at a concentration of 300 µM) and a mixture of all

curcumin decomposition products after incubation for 1 h at 37 oC (pH 7.4). General assay conditions are

described under materials and methods. The charts represent means based on n = 3 for samples

incubated with CYP1A2, CYP3A4, CYP2B6 and n = 2 for CYP2C9 and CYP2D6. The symbol ‘# ‘

represents statistically significant difference (P <0.05) from uninhibited reactions as determined by

Student’s t-test. Cur, curcumin; Van, vanillin; Vac, vanillic acid; Fal, ferulic aldehyde; Fac, ferulic acid.

Page 86: thuoc từ curcumin

76

The four decomposition products of curcumin that exhibited inhibition of the CYP

isoforms, appeared to be rather weak inhibitors as shown in table 2. For curcumin, the

inhibition of the CYPs in decreasing order of potency was CYP2C9 > CYP3A4 >

CYP2B6 > CYP1A2 > CYP2D6.

Results obtained on inhibition of CYP3A4 by curcumin, using the substrates 7-

benzyloxyquinoline (BQ) and 7-benzyloxy-4-trifluoromethylcoumarin (BFC) could not be

analyzed due to interference by curcumin at the wavelength for detection of the

metabolites.

Table 2. Concentrations of curcumin and four of its decomposition products required to reduce the

activities of five different human CYPs by 50% (IC50 value, µM)

Enzyme Curcumin Vanillic acid Ferulic aldehyde

CYP1A2 40.0 + 12.7 nd 227.5 + 23.4

CYP3A4 16.3 + 1.7a nd nd

13.9 + 3.4b nd nd

CYP2D6 50.3 + 2.0 nd 537.6 + 34.9

CYP2C9 4.3 + 0.8 250.6 + 17.5 259.3 + 22.8

CYP2B6 24.5 + 0.8 nd nd

All values are the means + standard deviation (S.D.) of at least two experiments as described in the

Methods section. nd, not determined due to low percent inhibition observed (<50%). a, b

Substrates used,

benzyloxyresorufin and DBF respectively.

Enzyme kinetic parameters for all CYPs and the corresponding types of inhibition

with curcumin are show in Table 3. In the MROD (methoxyresorufin O-deethylase, with

CYP1A2) and BROD (benzyloxyresorufin O-debenzylase, with CYP3A4, CYP2B6)

assays, the presence of curcumin resulted in an increase of the respective Km values,

whilst the Vmax values did not change significantly when compared to the control

experiment, thus indicating competitive inhibition according to Michaelis-Menten’s

kinetics. However, with respect to CYP2D6 and CYP2C9 curcumin caused significant

decreases in Vmax values with no significant changes in the Km values, thus indicating

non-competitive inhibition.

Page 87: thuoc từ curcumin

77

3.3. Mechanism-based inhibition

The effects of pre-incubation of curcumin with NADPH-fortified CYP isozymes on

inhibition potency was also evaluated with all five human CYPs. Pre-incubation of

curcumin for 20 min with NADPH-supplemented CYP did not increase inhibition of

activities by curcumin in any of the CYP isoforms (data not shown). Therefore,

mechanism-based inhibition does not seem to occur with any of the CYPs studied.

Table 3. Enzyme kinetic parameters of the effects of curcumin on human CYP activities

Enzyme Curcumin Vmax Km Ki Type of

(µM) (nM/min/nM) (µM) (µM) inhibition

CYP1A2 0.0 2.43 + 0.51 5.0 + 1.6 43.3 + 10.9

25.0 2.27 + 0.04 7.8 + 0.8 competitive

CYP3A4* 0.0 0.21 + 0.02 2.8 + 0.2 7.4 + 3.5

2.5 0.16 + 0.03 3.8 + 0.1 competitive

CYP2D6 0.0 1.68 + 0.01 3.2 + 0.1 51.0 + 3.9

45.5 0.89 + 0.04 3.3 + 0.1 Non-competitive

CYP2C9 0.0 7.83 + 0.05 6.1 + 1.2 11.5 + 0.8

6.0 5.14 + 0.07 6.4 + 0.2 Non-competitive

CYP2B6 0.0 0.13 + 0.04 34.0 +10.4 33.2 + 14.0

50.0 0.11 + 0.01 69.0 + 12.8 competitive

Values are means + S.D. of at least two experiments as described in the Methods section. Curcumin

concentrations used in the experiments are indicated in the Table. *Substrate use in CYP3A4 inhibition

assay was benzyloxyresorufin.

4. Discussion

The purpose of this study was to evaluate the inhibitory potential of curcumin and its

decomposition products on the five important human drug-metabolizing CYPs, namely

CYP1A2, CYP3A4, CYP2B6, CYP2C9 and CYP2D6. Earlier reports on the inhibition of

rat liver microsomal CYPs by curcumin showed that curcumin is a strong inhibitor of

CYP1A and CYP2B [15,16]. In the present study, curcumin and its decomposition

Page 88: thuoc từ curcumin

78

products were first screened at a high concentration of 300 µM, for their inhibitory

potential towards the five CYPs used. At the high concentration, compounds exhibiting

less than 50% inhibitory activities were excluded from further tests with the particular

CYPs. Curcumin was found to possess higher inhibitory potentials against human

CYP1A2, CYP2B6, CYP3A4, CYP2D6 and CYP2C9 than the decomposition products.

However, in contrast to CYP inhibition data on rat mentioned above, curcumin is a less

potent inhibitor of human CYP1A2 and CYP2B6. These results support findings that

animal data may be poorly predictive of the human situation [19]. A recent study on

inhibitory activities of Indonesian medicinal plants, showed plant extracts of three

curcuma species possessing 65 to 73% inhibitory activities towards CYP3A4 and 30 to

54% inhibition towards CYP2D6 [32]. Although these activities are due to the whole

extracts with components including curcumin, they lend support to our findings that

curcumin significantly inhibits CYP3A4 but less significantly inhibits CYP2D6. Curcumin

inhibited CYP1A2, CYP2B6 and CYP3A4 competitively, while non-competitive inhibition

was observed with CYP2C9 and CYP2D6. The structural difference in active sites of the

enzymes used may have contributed to the two different types of inhibition observed.

The insignificant inhibitory activities of the decomposition products towards the tested

CYPs, clearly indicate that the decomposition products of curcumin are not likely to

cause drug-drug interaction at the level of major drug-metabolizing CYPs. Moreover,

decomposition of curcumin is not likely to occur significantly at the low pH in the gut, in

addition to the presence of the stabilizing factors such as enzymes and GSH [15].

The inhibitory potential of curcumin towards CYP3A4 (IC50 = 16.3 µM and Ki =

7.4 µM) could have implications for drug-drug interactions in the intestines because of

the direct exposure of the intestines to curcumin upon oral administration and the high

levels of CYP3A4 in the intestinal epithelial cells. Inhibition of CYP3A4 in the intestines

upon co-administration of drugs could result in a significantly increased bioavailability of

drugs, and consequently increased plasma concentrations of drugs, with the potential

result of adverse drug reactions. Inhibition of CYP3A4 by co-administered drugs has

been shown to result in adverse clinical drug-drug interactions, including fatalities [3].

For example, concomitant intake of grapefruit juice with drugs has also been shown to

increase plasma concentrations of many drugs in humans [33]. This effect appears to

Page 89: thuoc từ curcumin

79

be mediated mainly by the inhibition of CYP3A4 in the intestinal wall. It is worth noting

that inhibitors of CYP3A4 are often also known to inhibit P-glycoprotein (P-gp) function

[34]; both phenomena have been suggested to synergistically influence bioavailability of

orally administered agents. Inhibition of P-gp by curcumin at an effective concentration

of 15 µM has also been documented [35]. It is anticipated that inhibition of CYP3A4 and

P-gp by curcumin may be advantageous in mitigating first pass elimination of orally

administered drugs [36].

It is still unknown whether inhibition of the activities of the presently tested human

CYPs in the liver may cause significant systemic drug-drug interactions. As yet, there

are no reports on interaction between curcumin and drugs at the level of hepatic CYPs

in humans. Currently available human pharmacokinetic data show an extremely low

exposure of the liver to curcumin even at very high doses. Plasma concentrations of

curcumin and its metabolites in humans were found to be in nanomolar ranges. High

concentrations of curcumin were found in the faeces [14]. This implies that the inhibitory

effect of curcumin on activities of the CYPs in the liver may be insignificant. The

inhibition parameters determined in the present study, including IC50 and Ki values,

which are important to estimate the CYP-inhibitory potential of curcumin [37] are

relatively high (in micromolar ranges) compared to the anticipated amounts of curcumin

in the liver. However, in recent rodent studies it was demonstrated that repeated oral

administration of curcumin resulted in a 2-fold upregulation of both CYP3A and P-gp in

the liver, whereas a downregulation of these proteins was observed in the intestines

[17]. The significant increase in the area-under-curve and decreases of oral clearances

of midazolam and celiprolol, suggest that the effects on the intestinal activities was

more significant than those on the hepatic activities.

In conclusion, curcumin appears to inhibit five of the important human CYPs, with

the increasing order of potency as CYP2D6 < CYP2B6 < CYP1A2 < CYP3A4 <

CYP2C9. Our results suggest that inhibition of CYP3A4, and to a lesser extent

CYP2C9, by curcumin has the potential to cause clinically significant and harmful drug-

drug interactions upon oral co-administration of curcumin and other drugs metabolized

by these CYPs. Further investigation is required to evaluate the in vivo relevance of the

inhibitory activities of curcumin observed in the present in vitro study.

Page 90: thuoc từ curcumin

80

Acknowledgement

We thank Ed Groot of the Molecular Toxicology Section and Ben Bruyneel of the

Analytical Chemistry and Spectroscopy Section of the Vrije Universiteit, for their

technical assistance.

References 1. Nadler, E.P., Reblock, K.K., Ford, H.R., Gaines, B.A., 2003. Monoterapy versus multiple therapy for the treatment of perforated appendicitis in children. Surg Infect (Larcmt) 4:327-333. 2. Hemaiswarya, S., Doble, M., 2006. Potential synergism of natural products in the treatment of

cancer. Phytother Res 20:239-249. 3. Honig, P.K., Wortham, D.C., Zamani, K., Conner, D., Mullin, J.C., Cantilena, L.R., 1993. Terfen adine ketoconazole interaction. J Am Med Assoc 269:1513-1518. 4. Pea, F., Furlanut, M., 2001. Pharmacokinetic aspects of treating infections in the intensive care

unit: focus on drug interactions. Clin Pharmacokinet 40:833-868. 5. Zafar, A., Sharif, M.D., 2003. Pharmacokinetics, metabolism, and metabolism of atypical antipsyc- hotics in special populations. Primary care companion J Clin Psychiatry 5:22-25. 6. Ioannides, C., 2002. Pharmacokinetic interactions between herbal remedies and medicinal drugs. Xenobiotica 32:451-478. 7. Obach, R.S., 2000. Inhibition of human cytochrome P450 enzymes by constituents of St. John's

wort, an herbal preparation used in the treatment of depression. J Pharmacol Exp Ther 294:88-95. 8. Leu, T-H., Maa, M-C., 2002. The molecular mechanisms for the antitumorigenic effect of curcumin. Curr Med Chem-Anti-Cancer Agents 2:357-370. 9. Cole, G.M., Morihara, T., Lim, G.P., Yang, F., Begum, A., Frautschy, S.A., 2004. NSAID and 5-23

antioxidant prevention of Alzheimer's disease: Lessons from in vitro and animal models. Annals NY Acad Sci 1035:68-84.

10. Vajragupta, O., Boonchoong, P., Morris, G.M., Olson, A.J., 2005. Active site binding modes of curcumin in HIV-1 protease and integrase. Bioorg Med Chem Lett 15:3364-3368.

11. Reddy, R.C., Vatsaala, P.G., Keshamouni, V.G., Padmanaban, G., Rangarajan, P.N., 2005. Curcumin for malaria therapy. Biochem Biophy Res Comm 326:472-474.

12. Donatus, I.A., Sardjoko, Vermeulen, N.P.E., 1990. Cytotoxic and cytoprotective activities of curcumin. Effect on paracetamol-induced cytotoxicity, lipid peroxidation and glutathione depletion in rat hepatocytes. Biochem Pharmacol 39:1869-1875. 13. Cheng, A.L., Hsu, C.H., Lin, J.K., Hsu, M.M., Ho, Y.F., Shen, T.S., et al., 2001. Phase I clinical trial of curcumin, a chemopreventive agent, in patients with high-risk or primalignant lesions. Anticancer Res 21:2895-2900. 14. Sharma, R.A., Euden, S.A., Platton, S.L., Cooke, D.N., Shafayat, A., Hewitt, H.R., et al., 2004. Phase I clinical trial of oral curcumin: biomarkers of systemic activity and compliance. Clin Cancer Res 10:6847-6854. 15. Oetari, S., Sudibyo, M., Commandeur, J.N.M., Samboedi, R., Vermeulen, N.P.E., 1996. Effects of curcumin on cytochrome P450 and glutathione S-transferase activities in rat liver. Biochem Pharmacol 51:39-45. 16. Thapliyal, R., Maru, G.B., 2001. Inhibition of cytochrome P450 isozymes by curcumins in vitro and in vivo. Food Chem Toxicol 39:541-547. 17. Zhang, W., Tan, T.M.C., Lim, L.-Y., 2007. Impact of curcumin-induced changes in P-gp and CYP3A4 expression on the pharmacokinetics of peroral celiprolol and midazolam in rats. Drug Metab Dispos 35:110-115. 18. Mori, Y., Tatematsu, K., Koide, A., Sugie, S., Tanaka, T., Mori, H., 2006. Modification by curcumin of mutagenic activation of carcinogenic N-nitrosamines by extrahepatic cytochromes P-450 2B1 and 2E1 in rats. Cancer Sci 97:896-904.

Page 91: thuoc từ curcumin

81

19. Eagling, V.A., Tjia, J.F., Back, D.J., 1998. Differential selectivity of cytochrome P450 inhibitors against probe substrates in human and rat liver microsomes. Br J Clin Pharmacol 45:107-114. 20. Guengerich, F.P., 1997. Comparisons of catalytic selectivity of cytochrome P450 subfamily enzymes from different species. Chem Biol Interact 106:161-182. 21. Shimada, T., Yamazaki, H., Mimura, M., Inui, Y., Guengerich, F.P., 1994. Interindividual variations

in human liver cytochrome P-450 enzymes involved in the oxidation of drugs, carcinogens and toxic chemicals: studies with liver microsomes of 30 Japanese and 30 Caucasians. J Pharmacol Exp Ther 270:414-423.

22. Wang, Y.J., Pan, M.H., Cheng, A.L., Lin, L.I., Ho, Y.S., Hsieh, C.Y., Lin, J.K., 1997. Stability of curcumin in buffer solutions and characterization of its degradation products. J Pharm Biomed Anal 15:1867-1876.

23. Burke, M.D., Thompson, S., Elcombe, C.R., Halpert, J., Haaparanta, T., Mayer, R.T., 1985. Ethoxy-, pentoxy- and benzyloxyphenoxazones and homologues: a series of substrates to distinguish between different induced cytochromes P-450. Biochem Pharmacol 34:3337-3345. 24. Omura, T., Sato, R., 1964. The carbon monoxide binding pigment of liver microsomes. I. solubilization, purification and properties. J Biol Chem 239:2379-2385. 25. Stresser, D.M., Blanchard, A.P., Turner, S.D., Erve, J.C.L., Dandeneau, A.A., Miller, V.P., et al., 2000. Substrate-dependent modulation of CYP3A4 catalytic activity: Analysis of 27 test compounds with 4 fluorometric substrates. Drug Metab Dispos 28:1440-1448. 26. Hickman D., Wang J.P., Unadkat J.D., 1997. Evaluation of the selected in vitro probes and suitability of organic solvents for the measurement of human cytochrome P450 monooxygenase activities. Drug Metab Dispos 26:207-215. 27. Busby, W.F., Ackermann, J.M., Crespi C.L., 1998. Effects of methanol, ethanol, dimethyl sulfoxide, and acetonitrile on cDNA-expressed human cytochromes P-450. Drug Metab Dispos 27:246- 249. 28. Staskal, D.F., Diliberto, J.J., De Vito, M.J., Bimbaum, 2005. Inhibition of human and rat CYP1A2 by TCDD and dioxin-like chemicals. Toxicol Sci 84:221. 29. Walsky, R.L., Obach, R.S., 2004. Validation assays for human cytochrome P450 activities. Drug Metab Dispos 32:647-660. 30. Ko, J.W., Desta, Z., Soukhova, N.V., Tracy, T., Flockhart, D.A., 2000. In vitro inhibition of the cytochrome P450 (CYP450) system by the antiplatelet drug ticlopidine: potent effect on CYP2C19 and CYP2D6. Br J Clin Pharmacol 49:343-351. 31. Heydari, A., Yeo, K.R., Lennard, M.S., Ellis, S.W., Tucker, G.T., Rostami-Hodjegan, A., 2004. Mechanism-based inactivation of CYP2D6 by methylenedioxymethamphetamine. Drug Metab Dispos 32:1212-1217. 32. Usia, T., Iwata, H., Hiratsuka, A., Watabe, T., Kadota, S., Tezuka, Y., 2006. CYP3A4 and CYP2D6

inhibitory activities of Indonesian medicinal plants. Phytomedicine 13:67-73. 33. Fuhr, U., 1998. Drug interactions with grapefruit juice: Extent, probable mechanism and clinical

relevance. Drug safety 18:251-272. 34. Wacher, V.J., Silverman, J.A., Zhang, Y., Benet, L.Z., 1998. Role of P-glycoprotein and cytochrome P450 3A in limiting oral absorption of peptides and peptidomimetics. J Pharm Sci 87:1322-1330. 35. Chearwae, W., Anuchapreeda, S., Nandigama, K., Ambudkar, S.V., Limtrakul, P., 2004. Biochemical mechanism of modulation of human P-glycoprotein (ABCB1) by curcumin I, II and III purified from tumeric powder. Biochem Pharmacol 68:2043-2052. 36. Kuppens, I.E., Breedveld, P., Beijnen, J.H., Schellens, J.H., 2005. Modulation of oral drug bioavailability: from preclinical mechanism to therapeutic application. Cancer Invest 23:443- 464. 37. Bapiro, T.E., Egnell, A.C., Hasler, J.A., Masimirembwa, C.M., 2001. Application of higher

throughput screening (HTS) inhibition assays to evaluate the interaction of anti-parasitic drugs with cytochrome P450s. Drug Metab Dispos 29:30-35.

Page 92: thuoc từ curcumin

82

Page 93: thuoc từ curcumin

83

Chapter 4

Structure-activity relationship of inhibition of recombinant Human Cytochrome

P450 mediated metabolism by Curcumin Analogues

Regina Appiah-Opong, Iwan de Esch, Jan N. M. Commandeur, Mayagustina Andarini

and Nico P. E. Vermeulen

Adapted from European Journal of Medicinal Chemistry 2008 43:1621-1631

Inhibition of cytochrome P450 (CYP) is a major cause of drug-drug interactions. In this

work, inhibitory potentials of thirty-three curcumin analogues, i.e. 2,6-

dibenzylidenecyclohexanone (A series), 2,5-dibenzylidenecyclopentanone (B series)

and 1,4-pentadiene-3-one (C series) substituted analogues of curcumin towards

recombinant human CYP1A2, CYP3A4, CYP2B6, CYP2C9 and CYP2D6, all important

for drug metabolism, were studied in vitro. Fluorescence plate reader and high

performance liquid chromatography (HPLC) assays were used to evaluate CYP

inhibitory activities. MOE-based Quantitative structure–activity relationship (QSAR)

analysis suggested that electrostatic and hydrophobic interactions and lipophilicity are

important factors for CYP inhibition. Apart from insights in important molecular

properties for CYP inhibition, the present results may also guide further design of

curcumin analogues with less susceptibility to drug-drug interactions.

Page 94: thuoc từ curcumin

84

1. Introduction

Curcumin is a well-known food additive and constituent of traditional medicine in

Southeast Asia and the Indian subcontinent, the latter being an area with low incidence

of colorectal cancer [1]. This naturally occurring and synthetic compound is regarded as

a promising drug and has received considerable attention due to its antioxidant,

anticancer, anti-inflammatory, anti-HIV and anti-malarial properties [2-6]. Reports from a

phase 1 clinical trial on curcumin have shown that it is non-toxic even at doses as high

as 8 g/day [3]. Curcumin is unstable at a pH 7.4, however, this stability is strongly

improved by lowering the pH or by adding glutathione (GSH), N-acetyl L-cysteine

(NAC), ascorbic acid or rat liver microsomes or cytosol [7]. The instability of curcumin at

neutral to basic pH conditions has been attributed to the presence of an active

methylene group (figure 1) [8].

Figure 1. Chemical structure of curcumin. The active methylene group is indicated by the arrow.

Omitting this methylene group leads to the formation of more stable and potent

antioxidative compounds [8]. Removal of the active methylene group and one carbonyl

group led to 1,4-pentadiene-3-ones, which still possess antioxidant properties [9].

Modification of groups on the terminal aromatic rings of curcumin has revealed that the

electron-donating substituents increase the anti-inflammatory activity [10].

Previously a series of curcumin analogues were synthesized, in which the

methylene and one carbonyl group have been omitted [9]. These analogues are

derivatives of benzylidine, having either electron-withdrawing, electron-donating or

steric groups. The derivatives include nine compounds of 2,6-

dibenzylidenecyclohexanone (A), thirteen of 2,5-dibenzylidenecyclopentanone (B) and

eleven of 1,5-diphenyl-1,4-pentadiene-3-one (C) (Schemes 1-3).

CH3CH

3

OH

OH

O

O

OH

O CH3CH

3

OH

OH

O

O

OH

O

Page 95: thuoc từ curcumin

85

These analogues exhibit antioxidant, anti-inflammatory, and antibacterial activities that

render them potential drug candidates. Interestingly, some of the analogues have

shown much stronger antioxidant activities than curcumin [9].

Scheme 1

33

1

2

O

R

R

R R

R

R

2

1

Cyclohexanones (A)

Scheme 2

1

33

R1

R

R

O

R

R

R

22

Cyclopentanone (B)

R1 R2 R3

A8 H N(CH3)2 H

A10 Cl H H

A11 CH3 OH CH3

A14 t-C4H9 OH t-C4H9

R1 R2 R3

A0 H OH H

A2 H H H

A4 H OCH3 H

A5 H CH3 H

A7 H CF3 H

R1 R2 R3

B0 H OH H

B2 H H H

B1 OCH3 OH H

B3 H Cl H

B4 H OCH3 H

B7 H CF3 H

R1 R2 R3

B10 Cl H H

B11 CH3 OH CH3

B12 C2H5 OH C2H5

B13 i-C3H7 OH i-C3H7

B14 t-C4H9 OH t-C4H9

B15 OCH3 OH OCH3

Page 96: thuoc từ curcumin

86

Scheme 3

1

33

R1

R

R

O

R

R

R

22

1,4-pentadiene-3-ones (C)

Schemes 1-3 show the chemical structures of the synthetic curcumin analogues.

Drug-drug interactions due to inhibition or induction of enzyme activity are among

the major causes of attritions in drug development [11]. Inhibition of CYP enzymes is a

cause of clinically significant drug-drug interactions [12]. Irrespective of the mechanism,

CYP inhibition may result in accumulation of drugs resulting in adverse drug reactions

or a decrease in metabolism of drugs and activation of pro-drugs, and hence alter their

pharmacokinetic profile. Therefore, in vitro CYP-associated inhibition studies for

evaluation of drug candidates during the early stages of drug discovery and

development are considered cost-effective for predicting potential clinical drug-drug

interactions, since these interactions may result in adverse drug reactions and

therapeutic failure [13,14].

Recently the inhibitory effect of curcumin on five major human drug metabolizing

CYPs have been reported [15]. Curcumin showed strong inhibition of CYP2C9 and

CYP3A4, with IC50 values 4.3 and 16.3 µM respectively and moderate inhibition of

CYP1A2, CYP2B6 and CYP2D6 (40.0, 24.5, 50.3 µM respectively). The inhibitory

effects of curcumin analogues on human CYPs have not yet been reported. In this

study, we investigated the inhibitory potentials of thirty-three curcumin analogues to the

five major recombinant human drug-metabolizing CYPs [16,17] mentioned above.

R1 R2 R3

C0 H OH H

C1 OCH3 OH H

C2 H H H

C3 H Cl H

C5 H CH3 H

C6 H t-C4H9 H

R1 R2 R3

C7 H CF3 H

C9 Cl Cl H

C10 Cl H H

C11 CH3 OH CH3

C15 OCH3 OH OCH3

Page 97: thuoc từ curcumin

87

Inhibitory structure-activity relationships (SARs) were subsequently evaluated and

quantitative structure-activity relationships (QSARs) were investigated using the

program MOE (Molecular Operating Environment). In these studies we tried to identify

the molecular features that cause inhibition of the different CYP isoenzymes tested.

These studies are important because the resulting information will guide design the

synthesis of new curcumin analogues with less CYP inhibitory properties and open a

way for in silico prediction of xenobiotic inhibition of CYPs.

2. Materials and methods

2.1. Materials

Methoxyresorufin (MRes) and benzyloxyresorufin (BRes) were synthesized by the

method of Burke et al [18], and the purity was determined by HPLC, mass spectrometry

and 1H NMR. The plasmid, pSP19T7LT_2D6 containing human CYP2D6 bicistronically

co-expressed with human cytochrome P450 NADPH reductase was kindly provided by

Prof. M. Ingelman-Sundberg (Stockholm, Sweden). The plasmids, BMX100/h1A2 and

pCWh3A4 with human cytochrome P450 NADPH reductase were kindly donated by Dr.

M. Kranendonk (Lisbon, Portugal). Expression plasmids, pCWh2B6hNPR and

pCWh2C9hNPR with human cytochrome P450 NADPH reductase were kindly provided

by Prof. F.P. Guengerich (Nashville, Texas, USA). Curcumin analogues were kindly

donated by Dr. S. Sardjiman (Jakarta, Indonesia). All other chemicals were of analytical

grade and obtained from standard suppliers.

2.2. CYP expression and membrane isolation

The plasmids containing cDNA of five human CYPs were transformed into Escherichia

coli strain JM109. Expression of the CYPs was carried out in 3-litre flasks containing

300 ml terrific broth (TB) medium, with 1mM !-aminolevulinic acid, 0.5 mM thiamine,

400 µl/L trace elements, 100 µg/ml ampicillin, 1 mM isopropyl-"-D-

thiogalactopyranoside (IPTG), 0.5 mM FeCl3 (for CYP2D6 and CYP3A4 only), 1 mg/L

chloramphenicol (for CYP2B6 only) and 30 µg/ml kanamycin (for CYP3A4 only). The

culture media were inoculated with 3 ml overnight cultures of bacteria containing

plasmids for the various CYPs. The cell cultures were incubated for about 40 h at 28 oC

Page 98: thuoc từ curcumin

88

and 125 rpm, and CYP contents were determined using the carbon monoxide (CO)

difference spectra as described by Omura and Sato [19]. Cells were pelleted by

centrifugation (4000 g, 4 oC, 15 min) and resuspended in 30 ml Tris-Sucrose-EDTA

(TSE) buffer (50 mM Tris-acetate buffer pH 7.6, 250 mM sucrose, 0.25 mM EDTA).

Cells were treated with 0.5 mg/ml lysozyme prior to disruption by French press (1000

psi, 3 repeats). The membranes containing the human CYPs were isolated by

ultracentrifugation in a Beckmann 50.2Ti rotor (60 min, 40,000 rpm, 4 oC), resuspended

in TSE buffer and stored at –80 oC until use.

2.3. Stability of curcumin analogues in buffer

Decomposition of curcumin analogues was investigated as described [7], in 0.1 M

potassium phosphate buffer of pH 7.4. Solutions of 25µM curcumin analogues in buffer

(in 0.5% DMSO) were scanned every 5 min between 200-600 nm for 30 min using an

Ultrospec 2000 Pharmacia Biotech UV/visible spectrophotometer. The above

experiment was also performed in the presence of 1 mM GSH and 13.2 nM enzyme

(CYP) as described [7], to determine the effects of these factors on the stability of the

analogues in buffer.

2.4. CYP inhibition assays

2.4.1. 7-Methoxy-, 7-Benzyloxyresorufin and O-dealkylation

Inhibition of the activities of human CYP isoforms 1A2, 3A4 and 2B6, by curcumin

analogues was determined by microplate reader assays using fluorescent substrates.

Incubation conditions (eg. enzyme concentration, substrates, incubation time) and

wavelengths for detection for each of the inhibition assays are shown in Table 1. In

general, the microsomal incubations were carried out in a total volume of 200 µl, and in

the presence of 100 µM NADPH (freshly prepared) in a black coaster 96-well plate.

Membranes were pre-incubated for 5 min at 37 oC with 0.1 M potassium phosphate

buffer (pH 7.4), substrates, and inhibitors (curcumin analogues) with minimal use of

DMSO, i.e. always 1.0 % (v/v) or less.

Page 99: thuoc từ curcumin

89

Table 1. Experimental conditions for fluorescence CYP assays

CYP Enzyme Incubation Substrate Substrate Excitation Emission

amount time concn wavelength wavelength

nM min µM nm

1A2 13.2 10 MRes 5.0 530 586

3A4 14.3 30 BRes 5.0 530 586

2B6 15.3 30 BRes 20.0 530 586

MRes, methoxyresorufin; BRes, benzyloxyresorufin

Experiments were performed in the absence and presence of GSH, to determine the

influence of a second curcumin stabilizing factor on the experiments. Subsequently, the

analogues (100 µM each) were all screened for CYP inhibitory activity. The screening

procedure for all CYPs involved pooling compounds into groups of three or four and

testing for CYP inhibition of the mixture. Groups showing >20% inhibition were selected

for further screening, for identification of potential CYP inhibitors.

Determination of IC50 values for curcumin analogues showing over 20% inhibition

of CYP activities was performed. The concentration range of the curcumin analogues

used was from 0.195 to 100 µM. Incubations were started by the addition of 100 µM

NADPH, and maintained at 37 oC for the periods defined (Table 1). Reactions were

terminated with 75 µl of 80% acetonitrile and 20% 0.5 M Tris solution. Product formation

was linear for all incubation times. Concentrations of the probe substrates in all reaction

mixtures were chosen near the Michaelis-Menten’s constant (Km) value for each of the

CYPs tested. The Km values obtained using the alkoxyresorufins and all other

substrates were within the range of reported literature values [20]. All measurements

were performed in triplicate. Metabolite formation was measured spectrophotometrically

on a Victor2 1420 multilabel counter.

2.4.2. Diclofenac hydroxylation

For the CYP2C9 inhibition assay, reaction mixtures in 500 µl total volume consisted of

49 nM enzyme, 100 µM NADPH, 0.1 M potassium phosphate buffer (pH 7.4), 6 µM

diclofenac and inhibitor. Hundred micromolar of each of the analogues was screened for

Page 100: thuoc từ curcumin

90

CYP2C9 inhibitory activity. For IC50 determinations on curcumin analogues with >20%

inhibition, concentrations of analogues used were of the range 0.39 – 200 µM. After

preincubation for 5 min at 37 oC, reactions were initiated by adding NADPH and

terminated after 10 min upon the addition of 200 !l methanol. The reaction mixtures

were centrifuged at 14,000 rpm for 3 min. Product formed was measured using an

isocratic HPLC method [21]. A C18 column (150 mm x 3.2 mm, 5 µm particle size,

Phenomenex) was used and the carrier flow rate was 0.6 ml/min. The mobile phase

consisted of 60% (v/v) 20 mM potassium phosphate buffer (pH 7.4), 22.5% (v/v)

methanol and 17.5% (v/v) acetonitrile. Peaks were monitored at the wavelength of 280

nm. Retention times for 4-hydroxydiclofenac and diclofenac were 5.0 and 24.1 min,

respectively.

2.4.3. Dextromethorphan O-demethylation

Inhibition of CYP2D6 activity by curcumin and its decomposition products was

evaluated by a method described [22]. The reaction mixture had a total volume of 500 µl

and consisted of 18.2 nM enzyme, 4.5 µM dextromethorphan, 90.9 µM NADPH, 0.1 M

potassium phosphate buffer and curcumin analogues. Hundred micromolar of each of

the analogues was screened for CYP2D6 inhibitory activity, and IC50 was determined for

analogues showing >20% inhibition (concentration range 0.39-200 µM). Reactions were

initiated by the addition of NADPH and allowed to proceed for 45min before termination

with the addition of 60 mM zinc sulphate solution. Product formed was measured using

an isocratic HPLC fluorescence detection method and a C18 column (100 mm x 3 mm,

5 µm particle size, Chromspher). The mobile phase consisted of 24% (v/v) acetonitrile

and 0.1% (v/v) triethylamine adjusted to pH 3 with perchloric acid. The carrier flow rate

was 0.6 ml/min. Peaks were monitored at 280 nm (excitation) and 310 nm (emission).

The retention times of dextrorphan and dextromethorphan were 3.4 and 24.5 min,

respectively.

2.4.4. SAR and QSAR analysis

Percent inhibition of CYP activities by the curcumin analogues was calculated from the

ratios of the activities of inhibited to control samples. The IC50 values were calculated

Page 101: thuoc từ curcumin

91

using GraphPad Prism 4.0 version (GraphPad Prism software Inc. San Diego CA).

Subsequent Quantitative structure-activity relationship (QSAR) analysis of the

respective CYP-inhibitory activities of the curcumin analogues were performed using the

MOE software (Version 2005.06, Chemical Computing Group Inc, Montreal). Structural

descriptors were obtained from the QuaSAR-Descriptors in MOE, version 2005.06.

Table 2. List of all molecular descriptors used in this study, obtained from QuaSAR-Descriptor

MOE 2005.06 version.

No. Descriptor Description

1 a_acc number of hydrogen bond atoms

2 a_nO number of oxygen atoms

3 E_nb value of potential energy with all bonded terms disabled

4 PEOE_VSA_FHYD Fractional hydrophobic van der Waals surface area

5 PEOE_VSA_FPNEG Fractional negative polar van der Waals surface area

6 PEOE_VSA_FPPOS Fractional positive van der Waals surface area

7 SlogP_VSA0 VSA with sum of surface area vi such that contribution of log of

octanol/ water partition coefficient calculated from the given

structure Li is –0.4

8 SlogP_VSA4 (see 7) Sum of vi such that Li is in 0.1, 0.15

9 SMR_VSA5 Molecular refractivity/VSA, sum of vi such that molecular

refractivity from atom I, Ri is in 0.44, 0.485

10 Std_dim2 square root of second largest own value of covariance matrix of

the atomic coordinates

11 Std_dim3 square root of the third largest own value of covariance matrix of

the atomic coordinate

12 TPSA total polar surface area

13 VdistEq sum of log2m-pi/m where m and pi are sum and number of

distance matrix entries respectively

14 Weight molecular weight

Page 102: thuoc từ curcumin

92

One hundred and thirty seven descriptors, including both 2D and 3D molecular

descriptors were calculated using the MOE program. Structure-activity models were

generated by multiple stepwise regression analysis (MRA) of the biological and

structural variables using the statistical analysis software SPSS for Windows version 13.

Fischer coefficients (F values) were also calculated using the latter. Table 2 provides

information on MOE-selected descriptors used in this study.

3. Results

3.1. Stability of curcumin analogues in buffer

The stability of curcumin analogues (25 µM) in 0.1 M phosphate buffer was measured

spectrophotometrically at 200-600 nm. At pH 7.4 the curcumin analogues decomposed

to various extents after 30 min of incubation. Table 3 shows the percent decomposition

of the analogues.

Table 3. Percent decomposition of curcumin analogues in buffer (pH 7.4)

Cmpd Decomp (%) Cmpd Decomp (%) Cmpd Decomp(%)

A0 17.5 B3 12.0 C1 nd

A2 97.5 B4 30.5 C2 35.0

A4 54.5 B7 25.5 C3 46.0

A5 65.5 B8 17.0 C6 73.0

A7 74.0 B9 18.5 C7 82.5

A8 35.5 B10 40.5 C9 25.0

A10 63.0 B11 7.0 C10 46.0

A11 13.0 B12 25.5 C11 7.5

A14 57.5 B13 41.5 C15 6.5

B0 13.5 B14 3.0 Curcumin 74.0

B1 7.4 B15 nd

B2 13.0 C0 53.0

Cmpd, compound; Decomp, decomposition; nd, not determined. Concentration of each curcumin

analogues used was 25 !M. The experiment performed is described in the Methods section.

Page 103: thuoc từ curcumin

93

Generally, compounds of group B (i.e. 2,5-dibenzylidenecyclopentanones) were more

stable at pH 7.4 than those in groups A and C (2,6-dibenzylidenecyclohexanones and

1,5-diphenyl-1,4-pentadiene-3-ones). Group A, demonstrated the greatest instability,

with A2 having the highest (97.5%) and B14 the lowest (3.0%) percent decomposition.

The degradation of the compounds in buffer pH 7.4 was significantly blocked in the

presence of CYP enzyme and GSH resulting in 8.2% and 19% decomposition

respectively. These effects are similar to those reported on the stability of curcumin in

buffer, by Oetari et al [7].

3.2. CYP inhibition

All the thirty-three curcumin analogues (each at 100 µM concentration) were screened for

inhibitory potentials towards human CYP1A2, CYP3A4, CYP2B6, CYP2C9 and CYP2D6.

Preceding the inhibitor screening experiments, tests were conducted with and without GSH

[7] to determine whether another stabilization factor was necessary to mitigate

decomposition. The results indicated that inhibition by curcumin analogues in the presence

of GSH, did not influence the inhibitory effect of the compounds on CYP activity (data not

shown). Initial screening results indicated that the curcumin analogues demonstrated a

wide range of inhibitory activities towards CYP-mediated metabolism of probe substrates

(data not shown). Results on 29 subsequently selected compounds (with % inhibition

>20%) revealed 75.8% (22), 27.5% (8), 13.7% (4), 62.0% (18) and 41.3% (12) of the

compounds inhibiting CYP1A2, CYP3A4, CYP2B6, CYP2C9 and CYP2D6, respectively. In

general the compounds showed a comparatively stronger inhibitory potency towards

CYP1A2 and CYP2C9 than towards CYP3A4, CYP2B6 and CYP2D6 activities (Tables 5-

8).

The least inhibited enzyme was CYP2B6, with only four compounds, i.e. B13, B15,

C11 and C15 exhibiting >20% inhibition at 100 µM concentration of compounds, and IC50

values being 74.3, 70.0, 44.8 and 24.8 µM, respectively. Compounds of group A showed

very weak or no inhibitory activities towards CYP3A4, CYP2B6 and CYP2D6.

Page 104: thuoc từ curcumin

94

3.3. Similarity studies

Concerning the structures of curcumin analogues, Similarity studies were performed

using the procedure as described by Labute et al [23]. Accordingly, calculations were

carried out using the flexible alignment module of MOE, employing the MMFF94 force

field. The results and best scoring fit both in terms of similarity and objective function

are shown in table 4 and figure 2. The respective superimpositions indicated no

significant difference between the A, B and C series of compounds, since they appear

to be perfectly superposed. However, differences may also arise due to steric bulk.

Table 4. Results of flexible alignment module of MOE

F S dU

A0-B0 119.6379 177.1833 1.7428

A0-C0 119.5509 171.6317 0.0000

B0-C0 119.0929 168.9706 0.0000

F, similarity; S, objective function; dU, potential energy difference between the lowest minimal energy

conformation and the global minimal (kcal/mol).

In the conformation A0-B0, the potential energies of A0 and B0 were 5.7 and 3.9

kcal/mol, above the respective global minima. For the conformation A0-C0, the potential

energy of A0 was 3.7 kcal above, and that of C0 was exactly at the global minimum.

The conformation B0-C0 resulted in potential energies of B0 and C0, 0.6 and 0.4

kcal/mol respectively above the global minima.

Page 105: thuoc từ curcumin

95

A B

C

Figure 2. Flexible alignment of curcumin analogues A0 (black), B0 (grey) (A), A0 (black), C0 (grey) (B)

and B0 (black), C0 (grey) (C).

3.4. Quantitative structure-activity relationships (QSARs)

Summaries of the relevant datasets employed for generating the QSARs relating the

various molecular descriptors to the CYP inhibitory potencies of curcumin analogues

used in this work are shown in tables 5-8.

Table 5 shows the data for twenty curcumin analogues that exhibited inhibitory

activities towards CYP1A2, and five relatively important descriptors, i.e. standard

dimension 3, the standard deviation along a principal component axis (std_dim3), polar

surface area (TPSA), number of oxygen atoms (a_nO), number of hydrogen bond

atoms (a_acc) and potential energy with bonded terms disabled (E_nb). A weak

correlation (R2 = 0.682) was found between experimental and predicted IC50 data (Fig.

3A) based on these molecular descriptors.

Page 106: thuoc từ curcumin

96

R2

= 0.682

0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.5 1 1.5 2 2.5 3 3.5

Experimental log1/IC50 (1/uM)

Pre

dic

ted

lo

g1

/IC

50

(1

/uM

)

R2

= 0.804

0.0

0.5

1.0

1.5

2.0

2.5

0.5 1 1.5 2 2.5

Experimental log1/IC50 (1/uM)

Pre

dic

ted

lo

g1

/IC

50

(1

/uM

)

R2

= 0.73

0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.5 1 1.5 2 2.5 3 3.5

Experimental log1/IC50 (1/uM)

Pre

dic

ted

lo

g1

/IC

50

(1

/uM

)

R2

= 0.522

0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.5 1 1.5 2 2.5 3 3.5

Experimental log1/IC50 (1/uM)

Pre

dic

ted

lo

g1

/IC

50

(1

/uM

)

Figure 3. Plots of experimental and predicted CYP inhibitory activities (Log1/IC50). Equations for these

plots are shown under Tables 3-6 (CYP1A2, CYP3A4, CYP2C9 and CYP2D6 respectively) and the

compounds involved are listed in the tables.

However, exclusion of four outliers (B1, B13, C3 and C6) resulted in a good correlation

(R2 = 0.907), with the descriptors being an electrostatic parameter

(PEOE_VSA_FPNEG), a lipophilicity feature (SlogP_VSA0), molecular weight (weight),

molecular refractivity (SMR_VSA5) and potential energy with bonded terms disabled

(E_nb).

A CYP1A2

C CYP2C9

B CYP3A4

D CYP2D6

Page 107: thuoc từ curcumin

97

Table 5. Dataset for QSARs in CYP1A2 inhibitors

Cmpd IC50 (mM) Log1/IC50Ea Log1/ IC50P

b std_dim3 TPSA a_nO a_acc E_nb

A2 0.0009 3.046 2.334 0.432 17.07 1 1 310.29

A4 0.0110 1.959 1.842 0.822 35.53 3 1 706.87

A5 0.0210 1.678 1.839 0.669 17.07 1 1 746.37

A7 0.0104 1.983 2.273 0.538 17.07 1 1 344.92

A8 0.0026 2.585 2.554 0.491 23.55 1 1 413.26

A10 0.0034 2.469 2.492 0.460 17.07 1 1 303.64

B0 0.0456 1.341 1.744 0.334 57.53 3 3 263.25

B1 0.0389 1.410 1.173 0.961 75.99 3 5 677.95

B2 0.0366 1.437 1.520 0.713 17.07 1 1 1054.22

B12 0.0367 1.435 1.591 0.875 57.53 3 3 280.39

B13 0.0284 1.547 1.101 1.181 57.53 3 3 692.98

B14 0.0428 1.369 1.382 1.018 57.53 3 3 453.52

C1 0.0415 1.382 1.342 1.104 75.99 3 5 473.23

C2 0.0037 2.432 2.168 0.556 17.07 1 1 445.43

C3 0.0266 1.575 2.073 0.900 17.07 1 1 453.90

C6 0.0280 1.553 1.915 1.193 17.07 1 1 538.34

C7 0.0040 2.398 2.010 1.016 17.07 1 1 487.56

C10 0.0034 2.469 2.285 0.077 17.07 1 1 448.75

C11 0.0304 1.517 1.227 1.668 57.53 3 3 444.70

C15 0.0752 1.124 1.181 1.651 94.45 3 7 498.76

QSAR expressions: n = 20, s = 0.356, R2 = 0.682, F = 19.51.

Log1/IC50 = 3.249 – 0.252std_dim3 + 0.052TPSA - 0.479a_nO – 0.904a_acc – 0.001E_nb

a,b, Experimental and predicted

The relevant dataset on seven curcumin analogues inhibiting CYP3A4, employed for

generating QSARs, is found in table 6. A fairly good correlation (R2 = 0.804) was found

between experimentally derived and predicted activities based on the descriptor,

VdistEq, a distance matrix parameter (Fig. 3B).

Page 108: thuoc từ curcumin

98

Table 6. Dataset for QSARs in CYP3A4 inhibitors

Cmpd IC50 (mM) Log1/IC50Ea Log1/IC50P

b VDistEq

B0 0.0051 2.292 2.125 3.492

B1 0.0649 1.188 1.473 3.628

B12 0.0735 1.134 1.157 3.694

B13 0.0776 1.110 1.166 3.693

B15 0.0383 1.417 1.157 3.694

C11 0.0403 1.395 1.540 3.614

C15 0.0132 1.119 1.027 3.722

QSAR expressions: n = 7, s = 0.205, R2 = 0.804, F = 20.39.

Log1/IC50 = 18.867 – 4.793VDistEq.

a,b, Experimental and predicted

However, these results are biased due to one outlier (i.e. B0) that appears to influence

the outcome significantly. Exclusion of the outlier resulted in no correlation between

experimental and predicted inhibitory activities. The dataset on 12 compounds inhibiting

CYP2C9 activity is presented in Table 7. Three relatively important descriptors that

appear in the QSAR equation are, fractional polar negative van der Waals surface area

(PEOE_VSA_FPNEG), fractional hydrophobic van der Waals surface area

(PEOE_VSA_FHYD), and log P of accessible van der Waals surface area for each

atom (SlogP_VSA4). Experimental inhibitory activities of the analogues towards

CYP2C9 weakly correlated (R2 = 0.738) with predicted activities based on these

descriptors (Fig. 3C). Elimination of the outlier, B0 resulted in a fairly good correlation

(R2 = 0.836) with the same descriptors.

Page 109: thuoc từ curcumin

99

Table 7. Dataset for QSARs in CYP2C9 inhibitors

Cmpd IC50 (mM) Log1/IC50Ea Log1/ IC50P

b P_V_FH P_V_FP SlogP_VSA4

A0 0.0010 3.000 2.599 0.835 0.105 6.371

B0 0.0099 2.004 2.533 0.825 0.103 6.371

B1 0.0028 2.553 2.115 0.843 0.098 6.371

B11 0.0283 1.548 1.216 0.859 0.083 19.113

B12 0.0377 1.424 1.385 0.882 0.069 19.113

B14 0.0550 1.260 1.565 0.918 0.048 19.113

C0 0.0018 2.745 2.544 0.821 0.105 6.371

C1 0.0075 2.125 2.077 0.840 0.100 6.371

C2 0.0673 1.172 1.188 0.947 0.053 6.371

C10 0.0213 1.672 1.493 0.954 0.046 6.371

C11 0.0628 1.202 1.221 0.857 0.084 19.113

C15 0.0590 1.229 1.811 0.853 0.096 6.371

QSAR expressions: n = 12, s = 0.329, R2 = 0.730, F = 11.85.

P_V_FP, PEOE_VSA_FPNEG; P_V_FH, PEOE_VSA_ FHYD

Log1/IC50 = 53.151 – 92.429PEOE_VSA_FPNEG – 48.904PEOE_VSA_ FHYD –0.118SlogP_VSA4.

a,b, Experimental and predicted

Table 8 contains dataset for six curcumin analogues with inhibitory activity towards

CYP2D6. The descriptor present in the QSAR equation and as apparently related to the

Table 8. Dataset for QSARs in CYP2D6 inhibitors

Cmpd IC50 (mM) Log1/IC50Ea Log1/IC50P

b P_V_FP

B14 0.1187 0.926 0.911 0.048

C0 0.0020 2.699 2.497 0.105

C1 0.0006 3.222 2.358 0.100

C2 0.0685 1.164 1.050 0.053

C11 0.0132 1.879 1.913 0.084

C15 0.0827 1.082 2.247 0.096

QSAR expressions: n = 6, s = 0.352, R2 = 0.522, F = 27.88.

Log1/IC50 = -0.782 + 35.382PEOE_VSA_FPNEG.

a,b, Experimental and predicted

Page 110: thuoc từ curcumin

100

observed inhibitory activity is PEOE_VSA_FPNEG. A weak correlation (R2 = 0.522) was

found between experimental and predicted data based on the descriptor (Fig. 3D). A

good correlation (R2 = 0.903) was obtained upon removal of outlier C15.

4. Discussion

Inhibition of CYPs can lead to drug-drug interactions and therefore it is considered

important to evaluate potential drug candidates for CYP inhibitory activities. Inhibitory

potentials of curcumin towards recombinant human CYP1A2, CYP3A4, CYP2B6,

CYP2C9 and CYP2D6 have recently been evaluated in vitro [15]. The inhibitory

potencies (IC50 values) towards CYP3A4 and CY2C9 suggested a potential for relevant

drug-drug interaction in the intestine upon oral co-administration with other drugs

metabolized by CYP3A4, considering the required high dose for therapeutic effects.

Less potent activities were observed with CYP1A2, CYP2B6 and CYP2D6. Three

groups of curcumin analogues [9], i.e. nine of 2,6-dibenzylidenecyclohexanone (group

A), thirteen of 2,5-dibenzylidenecyclopentanoe (group B) and eleven of 1,5-diphenyl-

1,4-pentadiene-3-one (group C) were analogously tested experimentally for inhibition

towards five important human drug-metabolizing CYPs mentioned above [15]. QSAR

analysis employing molecular descriptors that have >18% correlations with activity were

used as inputs for artificial neural networks (ANNs) [Ma et al., 2006 http://www.natural-

selection.com/library/2006/NN_antihiv_ligand_gbf.pdf ] and were generated by MOE.

Most of the curcumin analogues exhibited low inhibitory activities towards the

CYPs tested. Six of them, A2, A8, A10, C2, C7 and C10 showed potent inhibitory

activities with IC50 values in the range 0.9 – 4 µM, towards CYP1A2. These compounds

showed about ten to forty fold greater potency towards inhibition of CYP1A2 than

curcumin itself. However, these six compounds have not been reported to have potent

antioxidant activities [9]. The compounds A2 and C2, both lack substituents, in all

investigated substitutions. Similarly, compounds A10 and C10 both have chloride at the

R1, with the R2 and R3 positions unsubstituted (Schemes 1 and 3). The only difference

between series A and C compounds is the absence of a central six-membered ring in

the latter. Therefore it appears that these substitutions together with the presence or

absence of a central six-membered ring favour increased CYP1A2 inhibitory potency of

Page 111: thuoc từ curcumin

101

compounds of A and C series. However, compound B2, having the same substituents

as A2 and C2 but a central five-membered ring, showed lower inhibitory potency.

Generally, compounds of the B series showed rather weak inhibitory activities towards

CYP1A2. Apparently the central five-membered ring renders these compounds less

active towards CYP1A2.

The present QSAR analysis suggest that five descriptors influence inhibition of

CYP1A2, being standard dimension 3 (std_dim3), the standard deviation along a

principal component axis, potential energy with bonded terms disabled (E_nb), number

of oxygen atoms (a_nO), number of hydrogen bond acceptor atoms (a_acc) and total

polar surface area (TPSA). A weak correlation (R2 = 0.682) was obtained between the

experimental and predicted inhibitory activities. However, exclusion of four outliers

resulted in a good correlation (R2 = 0.907), with electrostatic and lipophilicity

descriptors, as well as molecular refractivity, molecular weight and potential energy of

bonded terms disabled. The reason for the outliers is not clear, and subject to ongoing

research, that also include the molecular features of the target site.

Earlier QSAR studies have indicated that CYP activities are related to substrate

lipophilicity [24-26], and these results lend some support to that suggestion. Substrates

and inhibitors of CYP1A2 are usually planar small-volume molecules that are neutral or

weakly basic. However, a proposal has been made that the binding pocket of CYP1A2

enzyme is composed of mostly hydrophobic and aromatic amino acids with polar amino

acids for hydrogen bonding being present near the heme centre [27]. Thus it is possible

that hydrogen bonding and hydrophobic interactions contribute to the observed

inhibitory activities. The presence of an ortho- or para-hydroxyl group has been earlier

shown to be relevant for the antioxidant activity of these and other curcumin analogues

[9,28,29]. However, in contrast to these activities the most potent inhibitors of CYP1A2

lack the para-hydroxyl moiety. It is worth noting that CYP1A2 is known to be involved in

the activation of procarcinogens [30] and consequently, that inhibition of CYP1A2 could

be pharmacologically beneficial.

Seven curcumin analogues exhibited IC50 values towards CYP3A4 within the

concentration range used in the experiments. Weak inhibitory activities were obtained,

except in the cases of B0 and C12 where potencies of 5.1 + 4.0 and 13.2 + 3.0 µM were

Page 112: thuoc từ curcumin

102

found. These activities are comparable with that determined for curcumin [15]. Five of

these compounds belong to the cyclopentanone (B) group and the remaining two to the

1,4-pentadiene-3-one (C) group. No inhibition of CYP3A4 was observed for compounds

from series A, which suggests that the presence of the central six-membered ring

renders them less active towards CYP3A4. The QSAR equation found in the present

study, contained a distance matrix feature, VdisEq (Vertex distance equation), which

suggests that the observed inhibition of CYP3A4 by the analogues was related to

distance matrices of the compounds. The correlation found (R2 = 0.804) was biased due

to an outlier, resulting in an over-estimation of the outcome. Exclusion of the outlier

resulted in no correlation between the experimental and predicted inhibitory activities.

Hydrophobicity of compounds has been reported to play an important role in oxidation

of CYPs and binding to liver microsomes [31,32]. Thus, considering the hydrophobicity

of curcumin analogues they would be expected to bind significantly to the hydrophobic

pockets of the CYP3A4 active site. However, this was not observed in most cases

perhaps due to other more significant factors, as shown in these results.

The present curcumin analogues generally exhibited weak inhibitory activities

towards CYP2B6. IC50 values were obtained for only four compounds and were similar

or weaker than that of curcumin, which is a weak inhibitor of CYP2B6 [15]. Compounds

from the A series did not show any inhibitory activity towards CYP2B6. The QSAR

analysis was considered unreliable, due to the small number of significant inhibitors.

Evaluation of inhibitory potentials of curcumin analogues towards CYP2C9 resulted in

twelve compounds most of which had lower inhibitory potencies than that reported for

curcumin [15]. Among these compounds A0, B1 and C0 exhibited strong inhibitory

activities (range 1.0 – 2.8 µM). Since all three series of compounds (A, B and C) are

represented in this list, it appears that the absence or presence of the central five- or

six-membered ring in the structure of the compounds is not influencing their inhibitory

effect towards CYP2C9, but rather the aromatic ring substituents. B0 and C1, with

similar substituents (OCH3 and/or OH) as the strong inhibitors above (Scheme 1-3), are

moderately strong inhibitors of CYP2C9 having IC50 values of 9.9 + 0.79 and 7.5 + 0.30

µM, respectively. The present QSAR analysis suggested that the observed inhibitory

activities are related to the descriptors, PEOE_VSA_FHYD, PEOE_VSA_FPNEG and

Page 113: thuoc từ curcumin

103

SlogP_VSA4. A weak correlation (R2 = 0.73) was obtained between experimental and

predicted data based on these descriptors. However, exclusion of the outlier B0 resulted

in a better correlation with the same descriptors (R2 = 0.836). The reason for this

compound being an outlier is not clear. However, electrostatic and hydrophobic

interactions as well as lipophilicity are likely implicated in the observed inhibitory

activities towards CYP2C9. Compound lipophilicity has been suggested to play a role in

overall substrate binding affinity of CYPs and molecular modeling studies also indicated

possible electrostatic interactions at the CYP2C9 active site due to the presence of

basic amino acid residues [26,33]. Among others the relevance of hydrophobic

interactions in the inhibition of CYP2C9 was also observed. Hydrophobic interactions

are primary driving forces and contributors to binding affinity and specificity of a ligand

to an active site [Ma et al., 2006 http://www.natural-selection.com/library/2006/NN_anti-

hiv_ligand_gbf.pdf]. This factor is therefore a potential candidate for systematic

modulation when developing SARs leading to more potent and specific inhibitors of

CYP2C9. Furthermore, the presence of a hydroxyl substituent at the para- position

appears to be prevalent in the strong inhibitors of CYP2C9 and that has been

suggested to have similar implication for other biological activities in curcumin

analogues [9].

Six curcumin analogues inhibited CYP2D6 with a wide range of IC50 values (0.6 –

111.7 µM), with C0 and C1 being the most potent. No inhibition of CYP2D6 was

observed with compounds from the A series. Therefore, it appears that the presence of

central six-membered ring in the A series contributes to the observed effect. The QSAR

analysis of the CYP2D6 inhibitors, revealed that the electrostatic descriptor,

PEOE_VSA_FPNEG, relates best with CYP2D6 inhibition, resulting in a weak

correlation (R2 = 0.522) between the experimental and predicted activities. However

removal of the outlier, C15 resulted in a good correlation with the same descriptor. The

reappearance of the electrostatic parameter indicates that the inhibitory potencies of

curcumin analogues towards CYPs is possibly related to the fraction of polar negative

van der Waals surface area present in the compounds. A limitation to our QSAR

analysis however, is the small number of compounds especially in case of CYP2D6.

Previous SAR studies on analogue series of stereoisomers of quinidine and quinine

Page 114: thuoc từ curcumin

104

suggested that hydrogen bonding by the hydroxyl group and not the basic nitrogen

interaction with an active site residue, would control the inhibitory potency of quinidine

[34]. It is likely that the presence of the hydroxyl substituent at the para- position

contributes to the observed inhibitory activity towards CYP2D6, since it is present in five

of the six CYP2D6 inhibitors. It is interesting to note that the compound B15, which has

meta-methoxy and para-hydroxyl substituents quite similar to curcumin and has been

shown to also have over ten times greater antioxidant activity than the latter [9] did not

exhibit any significant inhibitory activities towards the five CYPs tested.

Similarity studies indicated no significant differences between the ring structures

of the three series of compounds (Fig. 2), although differences may still lie in steric

bulkiness of substituents. Compound C0 clearly showed more flexibility than the A0 and

B0, and possessed the lowest minimal potential energy in each conformation as

determined by a stochastic conformational search. This is to be attributed to the

absence of the central aliphatic ring occurring in the other compounds. Clearly the

results of flexible alignments are unable to explain the observed activities of the

compounds, but we consider the similarity studies a useful approach in comparing

curcumin analogues and other structurally related compounds.

Experiments on the stability of curcumin analogues in buffer (pH 7.4) resulted in

varying degrees of degradation of the compounds. Generally compounds of group A

were more susceptible to degradation than those of groups B and C. The instability of

group A compounds may be attributed to the central six-membered ring since that is the

basic difference between the three series of compounds. Compounds in series B

appeared to be more stable in the buffer. The analogues were more stable in buffer

than curcumin, except A2, A7, C6 and C7 which were equally or less stable than

curcumin with A2 being the most unstable and B14 the most stable. Instability of group

A compounds at neutral to basic pH, may be due to aromatization of the central six-

membered ring in the presence of slightly basic conditions and subsequent dissociation

of resulting single bonds. Previous studies however, clearly demonstrated that the

degradation of curcumin in buffer at neutral to basic pH could be blocked by the

incorporation of enzyme, GSH, NAC or ascorbic acid in the incubations [7]. The present

Page 115: thuoc từ curcumin

105

results obtained indicate that like curcumin, degradation of its analogues in buffer at pH

7.4 is also significantly blocked (90%) in the presence of enzymes.

5. Conclusion

Thirty-three curcumin analogues were investigated for inhibition of human recombinant

CYP1A2, CYP3A4, CYP2B6, CYP2C9 and CYP2D6. Most of the curcumin analogues

showed low or negligible activities towards the CYPs tested. Six of the compounds (A2,

A8, A10, C2, C7 and C10) showed strong inhibitory activities towards CYP1A2, while

one compound (B0) strongly inhibited CYP3A4. CYP2C9 and CYP2D6 were strongly

inhibited by three (A0, B1, C0) and two (C0 and C1) compounds respectively. The

MOE-based QSAR analyses suggest that electrostatic and/or hydrophobic descriptors

notably PEOE_VSA_FPNEG and PEOE_VSA_FHYD, are important factors of the

compounds relating to inhibition of CYP1A2, CYP2C9 and CYP2D6. Consideration of

these (Q)SAR results might be relevant in the optimization of curcumin analogues with

less potential to cause CYP mediated drug-drug interactions.

Acknowledgement

We thank Enade Istyastono and Eva Stjernschantz for technical assistance and helpful

discussions.

References

1. Greenlee, R.T., Murray, T., Bolden, S., Wingo, P.A., 2000. Cancer statistics. CA Cancer J Clin 50:7-33.

2. Cole, G.M., Morihara, T., Lim, G.P., Yang, F., Begum, A., Frautschy, S.A., 2004. NSAID and antioxidant prevention of Alzheimer's disease: Lessons from in vitro and animal models. Annals of the New York Academy of Sciences 1035:68-84. 3. Cheng, A.L., Hsu, C.H., Lin, J.K., Hsu, M.M., Ho, Y.F., Shen, et al., 2001. Phase I clinical trial of curcumin, a chemopreventive agent in patients with high-risk or pre-malignant lesions. Anticancer Res 21:2895-2900. 4. Gescher, A., 2004. Polyphenolic phytochemicals versus non-steroidal antiinflammatory drugs: which are better cancer chemopreventive agents? J Chemother 4:3-6. 5. Vajragupta, O., Boonchoong, P., Morris, G.M., Olson, A.J., 2005. Active site binding modes of cur- cumin in HIV-1 protease and integrase. Bioorg Med Chem Letters 15:3364-3368. 6. Reddy, R.C., Vatsaala, P.G., Keshamouni, V.G., Padmanaban, G., Rangarajan, P.N., 2005. Cur- cumin for malaria therapy. Biochem Biophy Res Comm 326:472-474. 7. Oetari, S., Sudibyo, M., Commandeur, J.N.M., Samboedi, R., Vermeulen, N.P.E., 1996. Effects of

curcumin on cytochrome P450 and glutathione S-transferase activities in rat liver. Biochem Pharmacol 51:39-45.

Page 116: thuoc từ curcumin

106

8. Youssef, K.M., El-Sherbeny, M.A., El-Shafie, F.S., Farag, H.A., Al-Deeb, O.A., Awadalla, S.A.A., 2004. Synthesis of curcumin analogues as potential antioxidant, cancer preventive agents. Arch Pharm Pharm Med Chem 337:42-54.

9. Sardjiman, S., Reksohadiprodjo, M., Hakim, L., van der Goot, H., Timmerman, H., 1997. 1,5-Diphe- nyl-1,4-pentadiene-3-ones and cyclic analogues as antioxidative agents. Synthesis and structure- activity relationship. Eur J Med Chem 32:625-636. 10. Nurfina, A., Reksohadiprodjo, M., Timmerman, H., Jenie, U., Sugiyanto, D., van dr Groot, H., 1997. Synthesis of some symmetrical curcumin derivatives and their anti-inflammatory activity. Eur J Med Chem 32:321-328. 11. Zhang, Z.Y., Wong, Y.N., 2005. Enzyme kinetics for clinically relevant CYP inhibition. Curr drug Metab 6:241-257. 12. Desta, Z., Soukhova, N.V., Flockhart, D.A., 2001. Inhibition of cytochrome P450 (CYP450) isofo- rms by isoniazid: potent inhibition of CYP2C19 and CYP3A. Antimicrob Agents Chemother 45:382- 392.

13. Honig, P.K., Wortham, D.C., Zamani, K., Conner, D., Mullin, J.C., Cantilena, L.R., 1993.Terfenadine ketoconazole interaction. J Am Med Assoc 269:1513-1518.

14. Pea, F., Furlanut, M., 2001. Pharmacokinetic aspects of treating infections in the intensive care unit: focus on drug interactions. Clin Pharmacokinet 40:833-868. 15. Appiah-Opong, R., Commandeur, J.N.M., van Vugt-Lussenburg, B., Vermeulen, N.P.E., 2007.

Inhibition of human recombinant cytochrome P450s by curcumin and curcumin decomposition products. Toxicology 235:83-91.

16. Nebert, D.W., Russell, D.W., 2002. Clinical importance of the cytochromes P450. Lancet 360:1155 –1162. 17. Lamba, V., Lamba, J., Yasuda, K., Strom, S., Davila, J., Hancock, M.L., et al., 2003. Hepatic CYP2B6 expression: gender and ethnic differences and relationship to CYP2B6 genotype and CAR (constitutive androstane receptor) expression. J Pharmacol Exp Ther 307:906-922. 18. Burke, M.D., Thompson, S., Elcombe, C.R., Halpert, J., Haaparanta, T., Mayer, R.T., 1985. Ethoxy-, pentoxy- and benzyloxyphenoxazones and homologues: a series of substrates to distinguish between different induced cytochromes P-450. Biochem Pharmacol 34:3337-3345. 19. Omura, T., Sato, R., 1964. The carbon monoxide binding pigment of liver microsomes. I. Solubilization, purification and properties. J Biol Chem 239:2379-2385. 20. Staskal, D.F., Diliberto, J.J., De Vito, M.J., Bimbaum. 2005. Inhibition of human and rat CYP1A2 by TCDD and dioxin-like chemicals. Toxicol Sci 84:225-31. 21. Walsky, R.L., Obach, R.S., 2004. Validation assays for human cytochrome P450 activities. Drug Metab Dispos 32:647-660. 22. Ko, J.W., Desta, Z., Soukhova, N.V., Tracy, T., Flockhart, D.A., 2000. In vitro inhibition of the cytochrome P450 (CYP450) system by the antiplatelet drug ticlopidine: potent effect on CYP2C19 and CYP2D6. Br J Clin Pharmacol 49:343-351. 23. Labute, P., Williams, C., Feher, M., Sourial, E., Schmidt, J.M., 2001. Flexible alignment of small

molecules. J Med Chem 44:1483-1490. 24. Al-Gailany, K.A.S., Houston, J.B., Bridges, J.W., 1978. The role of substrate lipophilicity in determ-

ining type 1 microsomal P450 binding characteristics. Biochem Pharmacol 27:783-788. 25. Lewis, D.F.V., Dickins, M., 2003. Baseline lipophilicty relationships in human cytochrome P450 associated with drug metabolism. Drug Metab Rev 35:1-18. 26. Lewis, D.F.V., Brian, G.K., Yuko, I., Pavel, A., 2006. Quantitative structure-activity relationships (QSARs) within cytochromes P450 2B (CYP2B) subfamily enzymes: The importance of lipophilicity for binding and metabolism. Drug Metab Drug Interact 21:213-231. 27. Korhonen, L.E., Rahnasto, M., Mahonen, N.J., Wittekindt, C., Poso, A., Juvonen, R.O., et al., 2005. Predictive three-dimensional quantitative structure activity relationship of cytochrome P450 1A2 inhibitors. J Med Chem 48:3808-3815. 28. Adam, B.K., Ferstl, E.M., Davis, M.C., Herold, M., Kurtkaya, S., Camalier, R.F., et al., 2004. Synthesis and biological evaluation of novel curcumin analogues as anti-cancer and anti- angiogenesis agents. Bioorg Med Chem 12:3871-3883. 29. Anto, R.J., George, J., Babu, K.V., Rajasekharan, K.N., Kuttan, R., 1996. Antimutagentic and anticarcinogentic activity of natural synthetic curcuminoids. Mutat Res 370:127-131.

Page 117: thuoc từ curcumin

107

30. Shimada, T., Guengerich, F.P., 2006. Inhibition of human cytochrome P450 1A1-,1A2, and 1B1- mediated activation of procarcinogens to genotoxic metabolites by polycyclic aromatic hydrocarbons. Chem Res Toxicol 19:288-294. 31. Delaforge, M., Pruvost, A., Perrin, L., Andre, F., 2005. Cytochrome P450-mediated oxidation of

glucuronide derivatives: example of estradiol-17beta-glucuronide oxidation to 2-hydroxy-estradiol-17beta-glucuronide by CYP2C8. Drug Metab Dispos 33:466-473.

32. Austin, R.P., Barton, P., Cockroft, S.L., Wenlock, M.C., 2002. The influence of non-specific micros- omal binding on apparent intrinsic clearance, and its prediction from physicochemical properties. Drug Metab Dispos 30:1497-1503. 33. Lewis, D.F.V., 2004. Quantitative structure activity relationships (QSARs) for substrates of human cytochrome P450 CYP2 family enzymes. Toxicol in vitro 18:89-97. 34. Hutzler, J.M., Walker, G.S., Wienkers, L.C., 2003. Inhibition of cytochrome P450 2D6: structure activity studies using a series of quinidine and quinine analogues. Chem Res Toxicol 16:450-459.

Page 118: thuoc từ curcumin

108

Page 119: thuoc từ curcumin

109

Chapter 5

Inhibition of human glutathione S-transferases by curcumin and analogues

Regina Appiah-Opong, Jan N. M. Commandeur, Enade Istyastono, Jan J. Bogaards,

Nico P. E. Vermeulen

Xenobiotica, 2009, in press

Glutathione S-transferases are important phase II drug metabolizing enzymes playing a

major role in protecting cells from the toxic insults of electrophilic compounds. Curcumin,

a promising chemotherapeutic agent inhibits human GSTA1-1, GSTM1-1 and GSTP1-1

isoenzymes. The effect of three series of curcumin analogues, 2,6-

dibenzylidenecyclohexanone (A series), 2,5-dibenzylidenecyclopent-anone (B series) and

1,4-pentadiene-3-one (C series) substituted analogues on these human isoenzymes,

human and rat liver cytosolic GSTs was investigated, using 1-chloro-2,4-dinitrobenzene

as substrate. Most of the curcumin analogues showed less potent inhibitory activities

towards GSTA1-1, GSTM1-1 and GSTP1-1 than curcumin. Compounds B14 and C10

were the most potent inhibitors of GSTA1-1 and human liver cytosolic GST, with IC50

values of 0.2-0.6 µM. The most potent inhibitors GSTM1-1 were C1, C3 and C10, with

IC50 values of 0.2-0.7 µM. Similarly, GSTP1-1 was predominantly strongly inhibited by

compounds of the series C, C0, C1, C2, C10 and A0, with IC50 values of 0.4-4.6 µM.

Compounds in the B series showed no significant inhibition of GSTP1-1. QSAR analyses

have suggested the relevance of van der Waal’s surface area and compound lipophilicity

factors, for the inhibition of GSTA1-1 and GSTM1-1. These results may be useful in

design and synthesis of curcumin analogues with less potency for GST inhibition.

Page 120: thuoc từ curcumin

110

1. Introduction

Curcumin, a common dietary component in curry, derived from the plant Curcuma longa

has several important biological activities, which include anti-cancer, anti-oxidant, anti-

inflammatory and anti-HIV [1-4]. Due to the important pharmacological properties of

curcumin, clinical trials are ongoing [5,6]. In spite of the vast biological properties of

curcumin there are drawbacks to the development of this potential chemotherapeutic

agent, which include low bioavailability and instability at neutral to basic conditions [6,7].

In addition, curcumin has been shown to be a potent inhibitor of drug metabolizing human

glutathione S-transferase (GST) A1-1, GSTM1-1 and GSTP1-1 and cytochrome P450

(CYP) 3A4 and CYP2C9 [8-10]. Thus, design and synthesis of curcumin analogues with

enhanced bioavailability and better pharmacological properties has been carried out [11-

13].

Glutathione S-transferases (GSTs) are a super-family of multifunctional proteins

with fundamental roles in the cellular detoxification of a wide range of xenobiotics [14,15].

Conjugation of electrophiles to the nucleophilic sulfur atom of the major intracellular thiol,

the tripeptide glutathione, constitutes a common detoxification pathway. A number of

compounds can however be activated to more reactive or toxic products through

conjugation [6]. Alternatively, GSTs play other roles including mediation of multi-drug

resistance in cancer chemotherapy, protection of tissues against oxidative damage,

targeting of endogenous substrates and xenobiotics for transmembrane transport, which

is essential in processes such as biosynthesis of leukotrienes and ligandins [15,17]. Most

GSTs exist as soluble enzymes and are active as dimeric proteins, with each subunit

having an active site composed of two distinct functional regions, comprising of a G-site

specific for binding of the co-substrate glutathione (GSH), and a hydrophobic H-site

binding structurally diverse electrophilic substrates [15]. The seven main classes of

mammalian GSTs are alpha (A), mu (M), pi (P), theta (T), sigma (S), Omega (O) and zeta

(Z) based on amino acid/nucleotide sequence identity and physical structure of the genes

[15]. The cytosolic GSTs are differentially expressed in various organs. GSTA and GSTM

are predominantly expressed in the liver, whereas insignificant levels of GSTP1 are

Page 121: thuoc từ curcumin

111

expressed in the liver [18]. Higher levels of GSTP1 are however found in human

erythrocytes, lungs, esophagus and placenta [19-22].

The crucial roles of GSTs in drug metabolism, enhancing elimination of

electrophilic toxic drugs and metabolites through GSH-conjugation and physiological roles

mentioned above, would be reduced or compromised as a result of inhibition of GSTs and

this could have profound toxicological or clinical implications. Inhibition of human GSTs by

many compounds, drugs and natural products such as RRR-!-tocopherol, quinine,

quinidine, tetracycline and artemisinin, have previously been studied in vitro [23-25]. For

example, with the anti-malarials quinine, quinidine, tetracycline and artemisinin, IC50

values towards GSTP1-1 were below peak plasma concentrations obtained during

therapy, therefore it is likely that this isoenzyme may be inhibited in vivo at doses

normally used in clinical practice. On the other hand, the role of GST inhibition in cancer

therapy has been well studied [26,27].

In the present study we investigated the inhibitory potentials of thirty-four curcumin

analogues towards human recombinant GSTA1-1, GSTM1-1, GSTP1-1, human and rat

liver cytosolic GSTs. The compounds were designed and synthesized in three series by

Sardjiman et al [11] (Schemes 1-3). Studies on anti-oxidant activities of these compounds

have shown that some of them possess similar or even more potent activities than

curcumin [11]. Most of these compounds have also shown weak inhibitory activities

towards recombinant human drug metabolizing CYPs [28]. In this study, quantitative

structure-activity relationships (QSARs) of GST-inhibition by these curcumin analogues

were also investigated using the MOE (Molecular Operating Environment) program, in

order to identify molecular features responsible for the inhibition of the three different GST

isoenzymes tested. These results may give leads for improved design of curcumin

analogues with less inhibitory potential towards GST inhibition.

Page 122: thuoc từ curcumin

112

Scheme 1

33

1

2

O

R

R

R R

R

R

2

1

Cyclohexanones (A)

Scheme 2

1

33

R1

R

R

O

R

R

R

22

Cyclopentanone (B)

R1 R2 R3

A8 H N(CH3)2 H

A10 Cl H H

A11 CH3 OH CH3

A14 t-C4H9 OH t-C4H9

R1 R2 R3

A0 H OH H

A2 H H H

A4 H OCH3 H

A5 H CH3 H

A7 H CF3 H

R1 R2 R3

B0 H OH H

B1 OCH3 OH H

B2 H H H

B3 H Cl H

B4 H OCH3 H

B7 H CF3 H

B8 H N(CH3)2 H

R1 R2 R3

B9 Cl Cl H

B10 Cl H H

B11 CH3 OH CH3

B12 C2H5 OH C2H5

B13 i-C3H7 OH i-C3H7

B14 t-C4H9 OH t-C4H9

B15 OCH3 OH OCH3

Page 123: thuoc từ curcumin

113

Scheme 3

1

33

R1

R

R

O

R

R

R

22

1,4-pentadiene-3-ones (C)

Schemes 1-3 show the chemical structures of the synthetic curcumin analogues.

2. Materials and methods

2.1. Reagents

Reduced glutathione was obtained from Sigma-Aldrich (Steinheim, Germany). 1-chloro-

2,4-dinitrobenzene (CDNB) was obtained from Aldrich-Europe (Beerse, Belgium). Pooled

human liver cytosolic fraction was purchased from CellzDirect Inc. (North Carolina, USA).

Curcumin analogues were kindly donated by Dr. S. Sardjiman (Jakarta, Indonesia).

Purified human recombinant GSTA1-1 and GSTM1-1 were donated by Dr. J.J. Bogaards

(TNO Zeist, The Netherlands) and human GSTP1-1 was a gift from Prof. M. Lo Bello

(University of Rome, Italy). The specific activities of the GSTs using CDNB as substrate

were, 123 U/mg, 262 U/mg and 24.7 U/mg protein, respectively [25]. All other chemicals

were of analytical grade and obtained from standard suppliers.

R1 R2 R3

C0 H OH H

C1 OCH3 OH H

C2 H H H

C3 H Cl H

C5 H CH3 H

C6 H t-C4H9 H

R1 R2 R3

C7 H CF3 H

C9 Cl Cl H

C10 Cl H H

C11 CH3 OH CH3

C15 OCH3 OH OCH3

Page 124: thuoc từ curcumin

114

2.2. Rat liver cytosol

Rat liver cytosol was prepared from untreated rats as described [29]. Briefly, isolated rat

liver samples were homogenized in 2 volumes of 50 mM potassium phosphate buffer (pH

7.4) containing 0.9% sodium chloride, using a Potter-Elvehjem homogenizer at 4oC.

Cytosolic fractions were obtained by centrifuging the homogenate fraction for 20 min at

12000 g, and subsequent centrifugation of the supernatant for 60 min at 100000 g.

Supernatant representing cytosolic proteins, were freezed at -20oC until use. Protein

concentration was determined by the method of Bradford [30].

2.3. GST inhibition assays

Inhibition of GST activity was assessed as described [25] with slight modification.

GST-mediated conjugation of CDNB to GSH was measured using an Ultrospec

2000 Pharmacia Biotech UV/visible spectrophotometer at the wavelength of 340 nm

for 2 min. Incubation mixtures (1 ml) contained 0.1 M potassium phosphate buffer

pH 6.5, 400 µM CDNB, 1 mM GSH, and enzyme solutions (2.1 µg/ml protein

(GSTA1-1), 1.5 µg/ml protein) GSTM1-1, 1.0 µg/ml protein (GSTP1-1), or 32.0 and

1.7 µg/ml protein, human and rat liver cytosolic fractions respectively). After the

addition of GSH and enzymes to the reaction mixture, curcumin analogues were

added and the reaction was started with the addition of the substrate, and monitored

for 2 min at room temperature (24oC). The thirty-four compounds were firstly tested

at a concentration of 100 µM. Subsequently, compounds showing > 70% inhibition

were selected for IC50 determination in the concentration range 0.043-100 µM. All

assays were linear functions of protein concentration and of time for at least 2 min.

The formation of GSH-conjugates of the curcumin analogues in the reaction mixture

was determined using gradient HPLC analysis. A C18 column (150 mm x 3.2 mm, 5

!M particle size, Phenomenex), a mobile phase consisting of 0.1% trifluoroacetic

acid (solvent A) and 99% acetonitrile (solvent B) were used, with a linear gradient

eluting from 0-10 min 0-29% B, 10-40 min 29% B, 40-50 min 29-95% B, and a flow

rate of 0.5 ml/min. Conjugate formation was monitored using a UV/visible detector at

the wavelengths 360 nm and 427 nm.

Page 125: thuoc từ curcumin

115

2.4. Data and QSAR analysis

Percent inhibition of GST activity by curcumin analogues was calculated from the

ratio of inhibited versus control samples. The inhibitor concentrations giving 50%

inhibition of enzyme activity (IC50) values were analyzed using GraphPad Prism

4.0 version (GraphPad Prism software Inc. San Diego CA). Descriptors for QSAR

analysis were calculated using Molecular Operating Environment (MOE) software

version 2006.08 developed by Chemical Computing Group Inc, Canada. A list of

the descriptors finally selected and applied in this study can be found in Table 1.

Table 1. Relevant descriptors in this QSAR studies, obtained from QuaSAR-Descriptors, MOE

2006.08 version.

No. Descriptors Description

1 CASA- Negative charge weighted surface area

2 dipole Dipole moment calculated from the partial charges of the

molecule

3 PEOE_VSA-0 Sum of the van der Waals surface area of atoms where partial

charge of atom (calculated using the Partial Equalization of

Orbital Electronegativities (PEOE) method) is in the range -

0.05 to 0.00.

4 PEOE_RPC+ Relative positive partial charge

5 SlogP_VSA4 Sum of the van der Waals surface area of atoms such that

contribution to log of octanol/water partition coefficient

calculated from the given structure is in the range 0.1 to 0.15.

6 SlogP_VSA8 Sum of the van der Waals surface area of atoms such that

contribution to log of octanol/water partition coefficient

calculated from the given structure is in the range 0.3 to 0.4.

7 SMR_VSA7 Sum of the van der Waals surface area of atoms such that

contribution to molar refractivity calculated from the given

structure is > 0.56.

Page 126: thuoc từ curcumin

116

Conformational analysis using stochastic conformation search was performed using the

conformational import module provided by MOE with no filters and no constraints. The

conformational analysis and energy minimization were performed using stochastic

conformational search with root mean square (RMS) gradient of 0.001 Å and iteration limit

of 10,000 using the MMFF94 force field. All non-quantum descriptors were calculated as

described previously [31]. The relationships between the log 1/IC50 (IC50 in Molar) and the

descriptors were identified by stepwise regression analysis using SPSS for Windows

version 14.0 developed by SPSS Inc, USA. The following statistical measures were used:

N=number of samples, F=coefficient of variance (test for quality of fit), r=coefficient of

correlation, R2=coefficient of correlation squared and s = standard error of estimation. The

descriptors selected for the “best model”, using stepwise regression analysis were

independent (i.e. the cross correlation between descriptors < 0.7; Pearson correlation

method using SPSS was performed to analyze the cross correlation). The leave-one-out

cross-validation (LOO-CV/q2) method was employed to determine cross-validated

coefficient (q2) as internal validation of the models.

3. Results

3.1. Inhibition of purified GST isoenzymes

Fifteen out of the thirty-four curcumin analogues tested, inhibited GSTA1-1 activity by less

than 70% at concentrations of 33.3 (due to solubility problems) or 100 µM (data not

shown). The IC50 values of compounds showing more than 70% inhibition of GSTA1-1

are shown in Tables 2-4. Only seven compounds (A0, B0, B14, C1, C9 and C10) showed

stronger inhibitory activities towards GSTA1-1, compared to curcumin (Tables 2-4).

Compounds B14 possessing a hydroxyl group at the para position (R2) in addition to the

bulky butyl substituents at the R1 and R3 positions and C10 having only a chloride group

at position R1 as the only substituent, were the most potent inhibitors of GSTA1-1 (IC50

values, 0.5 and 0.6 µM, respectively).

Page 127: thuoc từ curcumin

117

Table 2. Percent inhibition or IC50 values (µM) for inhibition of human recombinant GSTs by group A

curcumin analogues and curcumin

Cmpd R1 R2 R3 GSTA1-1 GSTM1-1 GSTP1-1

A0* H OH H 16.7 + 0.49 6.4 + 0.55 1.6 + 0.01

A2 H H H 73.3 + 0.02 12.2% 54.8 + 0.16

A4 OH OCH3 H 65.3% 27.3% 71.8 + 4.44

A7 H CF3 H 56.9 + 3.25 77.2 + 4.71 5.6%

A11 CH3 OH CH3 53.9% 17.1 + 4.14 49.6%

Cur OCH3 OH OCH3 18.8 + 0.77 0.3 + 0.10 15.1 + 1.12

Cmpd, compound code; *highest concentration tested, 33.3 µM; IC50 values are means + standard

deviation (S.D.) of two experiments as described in the Methods section.

Similarly sixteen of the compounds inhibited GSTM1-1 activity by less than 70%.

Tables 2-4 show the IC50 values of compounds exhibiting more than 70% inhibition of

GSTM1-1. Thirteen compounds exhibited potent inhibitory activities towards GSTM1-1

having IC50 values in the range 0.2-9.9 µM.

Table 3. Percent inhibition or IC50 values (µM) for inhibition of human recombinant GSTs by group B

curcumin analogues

Cmpd R1 R2 R3 GSTA1-1 GSTM1-1 GSTP1-1

B0 H OH H 10.7 + 2.52 4.1 + 3.04 57.0%

B1 OCH3 OH H 61.0% 1.0 + 0.18 52.4%

B2 H H H 48.5 + 0.40 59.9% 20.0%

B9 Cl Cl H 65.8% 31.9 + 0.78 31.1%

B11 CH3 OH CH3 28.7 + 4.77 9.9 + 2.40 40.4%

B12 C2H5 OH C2H5 24.5 + 10.1 30.0 + 2.93 26.3%

B14* t-C4H9 OH t-C4H9 0.5 + 0.13 58.0% 43.2%

Cmpd, compound code; *highest concentration tested, 33.3 µM; IC50 values are means + standard

deviation (S.D.) of two experiments as described in the Methods section.

Page 128: thuoc từ curcumin

118

Compound C1 (R1 = OCH3 R2 = OH) and compound C9 (R1 and R2 = C) inhibited

GSTM1-1 with equipotent activities compared to curcumin (0.3 !M) whilst compound C10

(R1= Cl) showed slightly more potency.

Table 4. Percent inhibition or IC50 values (µM) for inhibition of human recombinant GSTs by group C

curcumin analogues

Cmpd R1 R2 R3 GSTA1-1 GSTM1-1 GSTP1-1

C0 H OH H 43.5 + 3.06 1.5 + 0.49 0.6 + 0.14

C1 OCH3 OH H 6.9 + 1.04 0.3 + 0.10 4.6 + 0.44

C2 H H H 21.3 + 0.86 9.9 + 2.13 4.5 + 2.13

C3* H Cl H 77.7 + 3.15 0.7 + 0.12 6.1 +1.06

C5* H CH3 H 25.3 + 6.22 59.4% 55.2%

C7 H CF3 H 29.4% 2.0 + 1.11 51.5%

C9* Cl Cl H 1.6 + 0.44 0.3 + 0.06 22.0%

C10* Cl H H 0.6 + 0.02 0.2 + 0.03 0.4 + 0.11

C11* CH3 OH CH3 18.0 + 0.61 2.4 + 0.37 10.9 + 1.29

C15 OCH3 OH OCH3 21.0% 4.2 + 0.52 25.6%

Cmpd, compound code; *highest concentration tested, 33.3 µM; IC50 values are means + standard

deviation (S.D.) of two experiments as described in the Methods section.

The GSTP1-1 activity was inhibited by twenty-five of the curcumin analogues by less than

70%. Seven compounds showed more potent inhibitory activities towards this enzyme as

compared to curcumin. Six of these compounds belonged to the C series of curcumin

analogues, whereas only one was of the A series. None of the compounds in series B

inhibited GSTP1-1 activity by >70%. The two most potent inhibitors of GSTP1-1 among

the compounds were compounds C0 that has a hydroxyl group at the R2 position and C10

(IC50 values, 0.6 and 0.4 µM, respectively).

Page 129: thuoc từ curcumin

119

3.2. Human and rat liver cytosol

The thirty-four curcumin analogues shown in schemes 1-3 were also screened for

inhibitory activities towards pooled human and rat liver cytosolic GSTs. This resulted in

most of the compounds exhibiting less than 70% inhibitory activities towards the liver

cytosolic enzymes of both species. Seven compounds showed over 70% inhibition of

human liver GSTs, these are A0, B0, B14, C1, C3, C9 and C10, all having IC50 values in

the range 0.2-25.3 µM. The compounds B14 and C10 which were the most potent

inhibitors of GSTA1-1, were also found to be the most potent inhibitors of human liver

GSTs (Table 5).

Table 5. Percent inhibition or Percent inhibition or IC50 values (µM) for inhibition of human and rat

liver cytosolic GSTs by curcumin analogues and curcumin

Cmpd R1 R2 R3 Human liver Rat liver

A0* H OH H 25.3 + 0.86 53.0%

B0 H OH H 9.3 + 0.08 14.4 + 1.33

B14* t-C4H9 OH t-C4H9 0.2 + 0.01 12.8 + 1.51

C1 OCH3 OH H 18.2 + 1.53 0.8 + 0.17

C3* H Cl H 7.3 + 1.30 65.2%

C9* Cl Cl H 7.6 + 1.24 45.1%

C10* Cl H H 0.3 + 0.03 0.4 + 0.03

C11* CH3 OH CH3 61.8% 2.2 + 0.05

Cur OCH3 OH H 50.5% 4.2 + 0.23

Cmpd, compound code; *highest concentration tested, 33.3 µM; IC50 values are means + standard

deviation (S.D.) of two experiments as described in the Methods section.

On the other hand, six compounds inhibited rat liver cytosolic GST activities. The

compounds C1 and C10 were the most potent inhibitors of the rat liver GSTs, having IC50

values of 0.8 and 0.4 µM, respectively.

Page 130: thuoc từ curcumin

120

3.3. Quantitative structure-activity relationships (QSARs)

The QSARs relating the various molecular descriptors generated by the MOE program to

the GST inhibitory potencies of curcumin analogues were analyzed in this work. Equation

1 (Eq. 1) generated from the stepwise regression analysis is considered as the ‘best’

model for curcumin analogues as inhibitor of GSTA1-1.

Log (1/IC50) = 5.283 (± 0.307) + 0.007 (± 0.001) [SMR_VSA7]

- 0.070 (± 0.027) [SlogP_VSA4]

- 0.326 (± 0.239) [dipole] (Eq. 1)

N = 16, r = 0.813, R2 = 0.660, S = 0.452, F3, 12 = 7.771, F5%, 3, 12 = 3.490, q2 = 0.464.

The relatively important descriptors for this model are SMR_VSA7, SlogP_VSA4 and

dipole. The q2 value was 0.464, which indicates a low predictive power (i.e. <0.5), and the

difference between R2 and q2 was less than 0.2, but only a weak correlation was found

between the experimental and predicted inhibitory potencies. However, assigning C3 as

an outlier with the same descriptors as used in Eq. 1, resulted in a model with a good

predictive power where q2 is 0.617 and a stronger R2 value of 0.78.

In the QSAR analysis of GSTM1-1 inhibitors, equation 2 (Eq. 2) is considered as

the ‘best’ model of curcumin analogues as inhibitors of GSTM1-1.

Log (1/IC50 GSTM1-1) = 6.918 (± 0.510) - 0.009 (± 0.004) [PEOE_VSA-0]

- 0.014 (± 0.007) [SlogP_VSA8]

- 1.000 (± 0.999) [PEOE_RPC+] (Eq. 2)

N = 17, r = 0.807, R2 = 0.651, S = 0.502, F3, 13 = 8.087, F5%, 3, 13 = 3.411, q2 = 0.388.

The relevant descriptors for this model are PEOE_VSA-0, SlogP_VSA8 and

PEOE_RPC+. The difference between R2 and q2 is more than 0.2, suggesting the

presence of outlier(s) [32]. The q2 is also less than 0.5 (q2 = 0.338) and R2 value 0.651,

suggesting that the model possesses a weak predictive power. By assigning C15 as an

outlier and using the same descriptors, a model with better predictive power was

generated, with a q2 of 0.674 and R2 of 0.792.

Page 131: thuoc từ curcumin

121

The relatively important descriptor for this QSAR analysis of curcumin analogues

as GSTP1-1 inhibitors, CASA-, resulted in a good predictive model(Eq. 3)

Log (1/IC50 GSTP1-1) = 1.375 (± 1.086) - 0.003 (± 0.001) [CASA-] (Eq. 3)

N = 8, r = 0.832, R2 = 0.692, S = 0.498, F1, 6 = 13.486, F5%, 1, 6 = 5.990, q2 = 0.692.

GSTA1-1

r2 = 0.780

3.5

4.0

4.5

5.0

5.5

6.0

6.5

3.5 4.0 4.5 5.0 5.5 6.0 6.5

Experimental log 1/IC50 (1/M)

Pre

dic

ted

lo

g 1

/IC

50

(1

/M)

GSTM1-1

r2 = 0.792

4.0

4.5

5.0

5.5

6.0

6.5

4.0 4.5 5.0 5.5 6.0 6.5

Experimental log 1/IC50 (1/M)

Pre

dic

ted

lo

g 1

/IC

50

(1

/M)

GSTP1-1

r2 = 0.692

4.0

4.5

5.0

5.5

6.0

6.5

4.0 4.5 5.0 5.5 6.0 6.5

Experimental log 1/IC50 (1/M)

Pre

dic

ted

lo

g 1

/IC

50

(1

/M)

Figure 1. Plots of observed and calculated inhibitory activities (log 1/IC50). GSTA1-1; The log

(1/IC50)predicted values were calculated from Eq. 1 (A); GSTM1-1; The log (1/IC50)predicted values were

calculated from Eq. 2 (B); and GSTM1-1; The log (1/IC50)predicted values were calculated from Eq. 3 (C).

The q2 is more than 0.5 (q2 = 0.692), suggesting that the model possesses a good

predictive power, however a weak correlation was observed (R2 = 0.692). Plots of

C

A B

Page 132: thuoc từ curcumin

122

experimental and calculated inhibitory activities for all three human GST used are

presented in Figure 1.

Discussion

The effect of curcumin analogues on the activity of human recombinant GST activity was

assessed by measuring the inhibition of GST-mediated conjugation of GSH to CDNB.

Earlier studies have shown that curcumin is a potent inhibitor of GSTs A1-1, M1-1 and

P1-1, using CDNB as substrate, with IC50 values of 2, 0.04 and 5 µM [9]. The alpha (A)

and mu (M) GST isoforms are predominantly expressed in the human and rat livers,

whereas levels of the pi (P) form are insignificant [18]. Although ubiquitously expressed in

humans, the pi isoform is the most abundant GST in erythrocytes and significant levels

are also found in the lungs, esophagus and placenta [19,33].

The present study shows that most of the curcumin analogues are weaker

inhibitors of the human recombinant GSTs tested when compared to curcumin, and

patterns of selectivity have also been revealed. GSTA1-1 and GSTM1-1 were most

susceptible to inhibition by the curcumin analogues. Only seven compounds inhibited

GSTA1-1 more strongly than curcumin, three of these compounds belong to the C series,

and the most potent inhibitors were B14 and C10. Steric properties may be partly

responsible for the strong inhibitory activity of B14, because compounds with similarly

bulky substituents, as well as the central five-membered ring, such as B11 and B12

showed over 70% activities. On the other hand, the presence of the chloride substituent

at the meta position in C10 appears to increase the inhibitory activity towards GSTA1-1,

as observed also with C9. The compound C10 also showed strong inhibitory activity

towards CYP1A2, whilst B14 was a weak inhibitor [28].

Thirteen out of thirty-four compounds showed rather strong (<10 µM) inhibitory

activities towards GSTM1-1. However, comparing these activities with that of curcumin,

nine of the compounds showed weaker activities, with only four compounds of the C

series i.e. C1, C3, C9 and C10, showing similar inhibitory potencies towards this enzyme.

Since GSTM1 is expressed in only 60% of human individuals, inhibition of this enzyme

may not have significant clinical implications as observed in individuals lacking this

Page 133: thuoc từ curcumin

123

enzyme. However, it has also been reported that people lacking GSTM1-1, have higher

risk of developing lung cancer [23].

Similarly, GSTP1-1 was weakly inhibited by most of the curcumin analogues, with

twenty-five compounds out of thirty-four showing less than 70% inhibitory activities. None

of the compounds of series B showed any significant inhibition of GSTP1-1, suggesting

that the weak inhibitory activities of these compounds may be due to the presence of the

central five-membered ring. Strong inhibition of GSTP1-1, by compounds such as C0 and

C10, could have implications for oxidative stress in human tissues and erythrocytes

where the enzyme is highly expressed. This class as opposed to the other GST classes is

highly susceptible to oxidation due to a reactive cysteine residue [Cys47 in human, rat

and mouse GSTP1-1 and Cys45 in that of the pig] located near the glutathione-binding

site [34].

The compounds B14 and C10 were the most potent inhibitors of human liver

cytosolic GSTs and GSTA1-1 as expected, due to the high levels of GSTA in the liver.

Steric hindrance, together with the central five-membered ring could play a role in the

potent inhibitory activity of B14. The IC50 values with the recombinant human GSTs were

however, generally much lower compared to those with human liver cytosol. Especially in

the cases where moderate inhibition of GSTA1-1 and strong inhibition of GSTM1-1 were

recorded, a moderate to strong inhibition may be expected with the human liver cytosol, in

which GSTA1-1 and GSTM1-1 are known to be significantly expressed [18,25]. These

differences in inhibition could possibly be due to binding to proteins such as hemoglobin

and albumin, with higher affinity than to GSTs, as suggested by van Haaften et al [25].

The presence of the para-hydroxy and chloride substituents in C1 and C10, respectively,

possibly contributes to their strong GST inhibitory activities observed in rat liver cytosol.

Like curcumin, all the present analogues possess !, !-unsaturated carbonyl

groups, and hence have the potential to conjugate to GSH, which could influence the

outcome of the current inhibition experiments. Previous studies have shown that GSH

conjugates are good inhibitors of GSTs [35], however, after incubation for 2 min GSH-

conjugates of none of the analogues could be detected to any significant amount by

HPLC analysis, which is supporting the concept that the compounds are responsible for

the observed inhibitory activities. Generally, the strongest inhibitors of GSTs were of the

Page 134: thuoc từ curcumin

124

series C. Inhibition of GST may have toxicological consequences similar to that of

deficiency in GST expression, which has not been reported to result in abnormalities. On

the other hand overexpression of GSTs has been implicated in drug resistance in cancer

chemotherapy [36].

The QSAR model of GSTA1-1 inhibitors shows a positive correlation with

SMR_VSA7, and negative correlations with SlogP_VSA4 and dipole. This suggests that

in designing new curcumin analogues with less potent inhibition, the compounds should

possess low values of SMR_VSA7, but high values of SlogP_VSA4 and dipole. On the

other hand, the QSAR model of GSTM1-1 inhibitors has negative correlations with

PEOE_VSA-0, SlogP_VSA8, and PEOE_RPC+. This implies that to avoid inhibition of

GSTM1-1, new curcumin analogues should be designed to have high values of

PEOE_VSA-0, SlogP_VSA8, and PEOE_RPC+. With respect to the GSTP1-1 inhibitors,

a negative correlation with CASA- is shown by the QSAR model, thus to design new

curcumin analogues with weak inhibitory activities towards GSTP1-1, the compounds

should have high CASA-.

In conclusion this study has shown the inhibitory potencies of thirty-four

compounds, representing three series of curcumin analogues towards three important

human GST isoenzymes, human and rat cytosolic GSTs. The present study has shown

that 27, 31 and 27 curcumin analogues out of thirty four are less potent inhibitors of

GSTA1-1, GSTM1-1 and GSTP1-1, respectively than curcumin. Since GSTs are a major

group of phase II detoxification enzymes, potent inhibition by compounds such as C10

and B14 could have implications for toxicity in humans. The strong inhibitory activities

exhibited by some of the curcumin analogues could however have useful application in

chemotherapy. The MOE-based QSAR analyses have also suggested the relevance of

van der Waal’s surface area and compound lipophilicity factors for the inhibition of

GSTA1-1 and GSTM1-1. These results may be useful in designing of curcumin analogues

with less inhibitory activity towards GSH or in the consideration of these compounds as

chemotherapeutic agents.

Page 135: thuoc từ curcumin

125

References

1. Chen, A., Xu, J., Johnson, A.C., 2006. Curcumin inhibits human colon cancer cell growth by suppressing gene expression of epidermal growth factor receptor through reducing the activity of the transcription factor Egr-1. Oncogene 25:278-287.

2. Cole, G.M., Morihara, T., Lim, G.P., Yang, F., Begum, A., Frautschy, S.A., 2004. NSAID and antioxidant prevention of Alzheimer's disease: Lessons from in vitro and animal models. Annals NY Acad Sci 1035:68-84.

3. Nonn, L., Doung, D., Peehl, D.M., 2007. Chemopreventive anti- inflammatory activities of curcumin and other phytochemicals mediated by MAP kinase phosphatase-5 in prostate cells. Carcinogenesis 28:1188-96.

4. Vajragupta, O., Boonchoong, P., Morris, G.M., Olson, A.J., 2005. Active site binding modes of curcumin in HIV-1 protease and integrase. Bioorg Med Chem Letters 15:3364-3368. 5. Cheng, A.L., Hsu, C.H., Lin, J.K., Hsu, M.M., Ho, Y.F., Shen, T.S., et al., 2001. Phase I clinical trial of

curcumin, a chemopreventive agent, in patients with high-risk or premalignant lesions. Anticancer Res 21:2895-2900.

6. Sharma, R.A., Euden, S.A., Platton, S.L., Cooke, D.N., Shafayat, A., Hewitt, H.R., et al., 2004. Phase I clinical trial of oral curcumin: biomarkers of systemic activity and compliance. Clin Cancer Res 10:6847-6854.

7. Oetari, S., Sudibyo, M., Commandeur, J.N.M., Samboedi, R., Vermeulen, N.P.E., 1996. Effects of curcumin on cytochrome P450 and glutathione S-transferase activities in rat liver. Biochem Pharmacol 51:39-45.

8. van Iersel, M.L., Ploemen, J.P., Lo Bello, M., Federici, G., van Bladeren, P.J., 1997. Interactions of alpha, beta-unsaturated aldehydes and ketones with human glutathione S- transferase P1-1. Chem Biol Interact 108:67-78.

9 Hayeshi, R., Mutingwende, I., Mavengere, W., Masiyanise, V., Mukanganyama, S., 2007. Inhibition of human glutathione S-transferases activity by plant polyphenolic compounds ellagic acid and curcumin. Food Chem Toxicol 45:286-295.

10. Appiah-Opong, R., Commandeur, J.N.M., van Vugt-Lussenburg, B., Vermeulen, N.P.E., 2007. Inhibition of human recombinant cytochrome P450s by curcumin and curcumin decomposition products. Toxicology 235:83-91.

11. Sardjiman, S., Reksohadiprodjo, M., Hakim, L., van der Goot, H., Timmerman, H., 1997. 1,5- Diphenyl-1,4-pentadiene-3-ones and cyclic analogues as antioxidative agents. Synthesis and structure-activity relationship. Eur J Med Chem 32:625-636.

12. Youssef, K.M., El-Sherbeny, M.A., El-Shafie, F.S., Farag, H.A., Al-Deeb, O.A., Awadalla, S.A.A., 2004. Synthesis of curcumin analogues as potential antioxidant, cancer preventive agents. Arch Pharm Pharm Med Chem 337:42-54. 13. Lin, L., Shi, Q., Nyarko, A.K., Bastow, K.F., Wu, C.C., Su, C.Y., et al., 2006. Antitumor agents 250. Design and synthesis of new curcumin analogues as potential anti-prostate cancer agents. J Med Chem 49:3963-3972. 14. Commandeur, J.N., Stijntjes, G.J., Vermeulen, N.P., 1995. Enzymes and transport systems involved in the formation and disposition of glutathione S-conjugates. Role in bioactivation and detoxication mechanisms of xenobiotics. Pharmacol Rev 47:271-330. 15. Frova, C., 2006. Glutathione transferases in the genomics era: New insights and perspectives.

Biomol Eng 23:149-169. 16. van Bladeren, P.J., 2000. Glutathione conjugation as a bioactivation reaction. Chem Biol Interact

129:61-79. 17. Morrow, C.S., Simitherman, P.K., Townsend, A.J., 2000. Role of multidrug-resistance protein 2 in glutathione S-transferase P1-1-mediated resistance to 4-nitroquinoline 1-oxide toxicities in HepG2 cells. Mol Carcinog 29:170-178. 18. Eaton, D.L., Bammler, T.K., 1999. Concise review of glutathione S- transferases and the significa-

nce to toxicology. Toxicol Sci 49:156-164. 19. Awasthi, Y.C., Sharma, R., Singhal, S.S., 1994. Human glutathione S-transferases. Int J Biochem 26:295-308.

Page 136: thuoc từ curcumin

126

20. Prade, L., Huber, R., Manoharan, T.H., Fahl, W.E., Reuter, W., 1997. Structures of class pi glutathione S-transferase from human placenta in complex with substrate, transition-state analo- gue and inhibitor. Structure 5:1287-1295.

21. Chandra, R.K., Bentz, B.G., Haines, G.K., Robinson, A.M., Radosevich, J.A., 2002. Expression of glutathione S-transferase Pi in benign mucosa, Barrett's metaplasia, and adenocarcinoma of the esophagus. Head Neck 24:575–81. 22. Piipari, R., Nurminen, T., Savela, K., Hirvonen, A., Mantyla, T., Anttila, S., 2003. Glutathione S- transferases and aromatic DNA adducts in smokers bronchioaveolar macrophages. Lung Cancer 39:265-272.

23. van Bladeren, P.J., van Ommen, B., 1991. The inhibition of glutathione S-transferases: Mechan- isms, toxic consequences and therapeutic benefits. Pharmacol Ther 51:35-46. 24. Mukanganyama, S., Widersten, M., Naik, Y.S., Mannervik, B., Hasler, J.A., 2002. Inhibition of

glutathione S-transferases by antimalarial drugs possible implications for circumventing anticancer drug resistance. Int J Cancer 97:700-705.

25. van Haaften, R.I.M., Haenen, G.R.M.M., van Bladeren, P.J., Bogaards, J.J.P., Evelo, C.T.A., Bast, A., 2003. Inhbition of glutathione S-transferase isoenzymes by RRR-!-tocopherol. Toxicol in Vitro 17:245-251.

26. Townsend, D.M., Tew, K.D., 2003. The role of glutathione-S-transferase in anti-cancer drug resista-nce. Oncology 22:7369-7375.

27. Depeille, P., Cuq, P., Passagne, I., Evrard, A., Vian, L., 2005. Combined effects of GSTP1 and MRP1 in melanoma drug resistance. B J Cancer 93:216-223.

28. Appiah-Opong, R., de Esch, I., Commandeur, J.N.M., Andarini, M., Vermeulen, N.P.E., 2007. Structure-activity relationship for the inhibition of recombinant human cytochrome P450 mediated metabolism by curcumin analogues. Eur J Med Chem 43:1621-1631.

29. Rooseboom, M., Vermeulen, N.P.E., Groot, E.J., Commandeur, J.N.M., 2002. Tissue distribution of cytosolic beta-elimination reactions of selenocysteine Se-conjugates in rat and human. Chem Biol Interact 140:243-264.

30. Bradford, M.M., 1976. A rapid and sensitive method for the quantification of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem 72:248-254.

31. Shahapurkar, S., Pandya, T., Kawathekar, N., Chaturvedi, S.C., 2004. Quantitative structure activity relationship studies of diaryl furanones as selective COX-2 inhibitors. Eur J Med Chem 39:899-904.

32. Ericksson, L., Jaworska, J., Worth, A.P., Cronin, M.T., McDowell, R.M., Gramatica, P., 2003. Methods for reliability and uncertainty assessment and for applicability evaluations of classification- and regression-based QSARs. Environ Health Perspect 111:1361-1375.

33. Moscow, J.A., Fairchild, C,R., Madden, M.J., Ransom, D.T., Wieand, H.S., O’Brien, E.E., et al., 1989. Expression of anionic glutathione S-transferase and P-glycoprotein genes in human tissues and tumors. Cancer Res 49:1422-1428.

34. Sluis-Cremer, N., Naidoo, N., Dirr, H., 1996. Class-pi glutathione is unable to regain its native conformation after oxidative inactivation by hydrogen peroxide. Eur J Biochem 242:301-307.

35. Burg, D., Filippov, D.V., Hermanns, R., van der Marel, G.A., van Boom, J.H., Mulder, G.J., 2002. Peptidomimetic glutathione analogues as novel gamma GT stable inhibitors. Bioorg Med Chem 10:195-205. 36. Burg, D., Mulder, G.J., 2002. Glutathione conjugates and their synthetic derivatives as inhibitors of

glutathione-dependent enzymes involved in cancer and drug resistance. Drug Metab Rev 34:821- 863.

Page 137: thuoc từ curcumin

127

Chapter 6

Interactions between Cytochromes P450, Glutathione S-transferases and

Ghanaian Medicinal plants

Regina Appiah-Opong, Jan N. M. Commandeur, Civianny Axson and Nico P. E.

Vermeulen

Adapted from Food and Chemical Toxicology 2008 46:3598-3603

Inhibition of cytochrome P450s (CYPs) is a major cause of adverse drug-drug

interactions. Alternatively, inhibition of glutathione S-transferases (GSTs) may increase

harmful effects from electrophilic compounds. In the present study, aqueous extracts of

seven Ghanaian medicinal plants were investigated for inhibitory potentials towards

recombinant human CYP1A2, CYP2C9, CYP2D6 and CYP3A4 heterologously

expressed in Escherichia coli. Effects of these extracts on recombinant human GSTA1-

1, GSTM1-1, GSTP1-1, human and rat cytosolic GSTs were also investigated. Seven

extracts, including Phyllanthus amarus whole plant, leaf, stem and root, Cassia siamea

and Momordica charantia, inhibited CYP1A2 and CYP2C9 with IC50 values ranging from

28.3-134.3 !g/ml and 63.4-425.9 !g/ml, respectively. Similarly, both CYP2D6 and

CYP3A4 were each inhibited by five extracts including Phyllanthus amarus whole plant,

leaf, stem and root and Cassia alata, with IC50 values ranging from 45.8-182.0 !g/ml

and 79.2-158.8 !g/ml respectively. Human and rat liver cytosolic GSTs were inhibited

by the extracts with IC50 values ranging from 25.2-95.5 !g/ml and 8.5-139.4 !g/ml,

respectively. GSTM1-1 was most susceptible to the inhibition by the extracts, with IC50

values in the range 3.6-50.0 !g/ml, whilst IC50 values of 8.9-159.0 !g/ml and 68.6-157.0

!g/ml were obtained for GSTA1-1 and GSTP1-1, respectively. These findings show the

potential for CYP- and GST-mediated herb-drug interactions of the Ghanaian medicinal

plants investigated.

Page 138: thuoc từ curcumin

128

1. Introduction

Herbal medicines, increasingly used over the past decades [1,2], are usually assumed

to be harmless. Thus they are not subjected to the scrutiny of the approval process

applied to new drug applications. However, some of the herbal medicines may result in

CYP-mediated herb-drug interactions upon co-administration with prescribed drugs [3].

Similarly, interactions of plant extracts with glutathione S-transferases (GSTs) have

been reported [4]. Interactions of herbal medicines with human CYPs have been

associated with alterations in the pharmacokinetics of drugs such as midazolam [3]. The

interactions may involve both induction and inhibition of CYPs, the latter being more

common [5] and sometimes causing harmful side-effects [6]. Investigations in human

volunteers taking Hydrastis canadensis (goldenseal), 900 mg three times a day have

shown that this herbal supplement, taken to prevent common cold and upper respiratory

tract infections, significantly inhibits CYP2D6 and CYP3A4/5 by about 40% [3]. This

effect had been observed earlier in vitro [7]. Using in vitro approaches, many other

herbs and natural compounds have been identified as inhibitors of various CYPs.

Phyllanthus amarus, a medicinal plant, has been shown to possess anti-cancer

properties [8] (Table 1). Studies on inhibitory activities of a water/ethanol extract of

Phyllanthus amarus on HIV replication in vitro and ex vivo have also shown interference

with binding of HIV-1 gp120 to CD4 and inhibition of reverse transcriptase, integrase

and protease with IC50 values 2.65, 8.17, 0.48 and 21.8 µg/ml respectively [9].

Additionally, in human volunteers receiving a single dose of 1200 mg of dried

Phyllanthus amarus extract, virus replication was reduced by more than 30%, compared

to the known drug lamivudine (17%) [9]. Phyllanthus amarus extract has also been

reported to be a potent inhibitor of rat liver microsomal 7-ethoxyresorufin-O-deethylase

(CYP1A1), 7-methoxyresorufin-O-demethylase (CYP1A2) and also 7-pentoxyresorufin-

O-depentylase (CYP2B1/2) [10]. Furthermore, curcumin, a natural compound derived

from Curcuma longa, and possessing many important therapeutic activities such as

anti-cancer and anti-oxidant activities [11,12], has recently been shown to be a potent

inhibitor of recombinant human CYP3A4 and CYP2C9 [13]. Studies with Kava (Piper

methysticum) extract, a commercially available herbal anoxiolytic, product normalized to

100 µM total kavalactones, showed significant inhibition of human CYP1A2 (56%),

Page 139: thuoc từ curcumin

129

CYP2C9 (92%), CYP2C19 (86%), CYP2D6 (73%) and CYP3A4 (78%) in vitro [14]. A

case report described a coma of a woman after concomitant ingestion of kava and

alprazolam, a known CYP3A4 substrate [15]. Although direct evidence for an in vitro-in

vivo correlation is lacking, caution is imperative when kava is used in combination with

CYP3A4-substrate drugs [16]. These and many other adversities [17,18] have spurred

in vitro investigations on herbal medicines.

Herbal medicines may also modulate other important drug metabolizing enzymes

such as glutathione S-transferases (GSTs) [19]. GSTs are a super family of

multifunctional detoxification enzymes, which catalyze conjugation of glutathione (GSH)

to a wide variety of electrophilic compounds [20,21]. Inhibition of GSTs may also have

consequences on a large number of endogeneous processes, including the beneficial

inhibitory role of GSTs in drug resistance of tumors [22]. On the other hand, inhibition of

GST-mediated scavenging of electrophilic xenobiotic compounds may result in harmful

consequences [20,23].

In Ghana, herbal medicines are popular and often used by patients in

combination with prescribed drugs [24], in spite of possible adverse herb-drug

interactions. The effects of some herbal medicines such as Phyllanthus amarus extract

on rat liver CYPs has been studied [10]. However, extrapolation from animal to human

data remains unreliable [25,26]. The potential of Ghanaian herbal medicines for such

interactions with human CYPs has not yet been investigated. Since herb-drug

interactions is becoming an increasing concern [17], we selected seven Ghanaian

medicinal plants commonly employed for various ailments (Table 1) and evaluated the

potential of the aqueous extracts to cause CYP-mediated herb-drug interactions in vitro.

We investigated the effects of the plant extracts on four major human CYPs and three

major human GSTs and on human and rat liver cytosolic GSTs. Although the active

principles and phytochemical profile of the plants is not fully known, studies on

Phyllanthus amarus have shown the presence of lignans, phyllanthin and

hypophyllanthin, flavonoids such as quercetin and astragalin, ellagitannins and

hydrolysable tannins [8]. The presence of flavonoids, anthraquinones, alkaloids, tannins

and saponins in leaves of Cassia alata has also been reported [27].

Page 140: thuoc từ curcumin

1

30

Ta

ble

1.

Gh

an

aia

n m

ed

icin

al p

lan

ts,

the

ir f

am

ilie

s,

pa

rts u

se

d,

loca

l n

am

es,

the

rap

eu

tic u

se

s a

nd

vo

uch

er

sp

ecim

en

nu

mb

ers

Pla

nt

na

me

Fa

mily

P

art

use

d

Lo

ca

l n

am

e

Th

era

pe

utic u

se

Vo

uch

er

No

.

Ca

ssia

ala

ta L

. C

ae

sa

lpin

ace

ae

L

ea

ve

s

Ose

mp

a

An

ti-i

nfla

mm

ato

ry, a

nti-m

icro

bia

l G

C 4

59

33

a

nti-p

late

let

ag

gre

ga

tio

n, a

nti-d

iab

e-

tic,

an

ti-h

yp

ert

en

siv

e, a

nti-m

ala

ria

l

Ca

ssia

sia

me

a L

am

. C

ae

sa

lpin

ace

ae

L

ea

ve

s

Gye

du

a

An

ti-o

xid

an

t, a

nti-t

um

or,

G

C 4

59

32

a

nti-m

ala

ria

l

La

ctu

ca

ta

raxic

ifo

lia

Co

mp

osita

ce

ae

Le

ave

s

Nn

en

oa

A

nti-o

xid

an

t, a

nti-i

nfla

mm

ato

ry

GC

45

92

9

(Will

d.)

Sch

um

Mo

mo

rdic

a c

ha

ran

tia

L.

Cu

cu

rbita

ce

ae

L

ea

ve

s

Nyin

ya

A

nti-v

ira

l, a

nti-m

uta

ge

nic

, G

C 4

59

31

a

nti-d

iab

etic,

an

ti-i

nfla

mm

ato

ry

a

nti-c

an

ce

r, a

na

lge

sic

Mo

rin

da

lu

cid

a B

en

th.

Ru

bia

ce

ae

L

ea

ve

s

Oko

nkro

ma

A

nti-m

ala

ria

l, a

nti-d

iab

etic,

GC

45

93

0

a

nti-d

rep

an

ocyta

ry

Ph

ylla

nth

us a

ma

rus

Eu

ph

orb

iace

ae

W

ho

le p

lan

t A

wo

mm

ag

ua

kyi

An

ti-c

an

ce

r, a

nti-h

ep

atitis,

GC

47

81

3

(S

ch

um

) T

ho

nn

L

ea

ve

s,

Ste

ms

an

ti-H

IV,

an

ti-m

ala

ria

l,

R

oo

ts

an

ti-t

um

or,

Tri

da

x p

rocu

mb

en

s L

. C

om

po

sita

ce

ae

W

ho

le p

lan

t F

om

izig

be

A

nti-b

acte

ria

l, a

nti-p

roto

zo

al,

G

C 4

78

14

w

ou

nd

he

alin

g

Page 141: thuoc từ curcumin

131

Previous studies have also shown that Momordica charantia contains glycosides such

as momordin, vitamin C, flavonoids and other polyphenols [18].

2. Materials and methods

2.1. Reagents and chemicals

Methoxyresorufin was synthesized by methylation of resorufin by iodomethane in

acetone, in the presence of potassium carbonate, and confirmed by (1)H NMR [29]. The

plasmid, pSP19T7LT_2D6 containing human CYP2D6 bi-cistronically co-expressed with

human cytochrome P450 NADPH reductase was kindly provided by Prof. Ingelman-

Sundberg (Stockholm, Sweden). The plasmids BMX100/h1A2 and pCWh3A4 with CYP

NADPH reductase were donated by Dr. Michel Kranendonk (Lisbon, Portugal). The

plasmid pCWh2C9hNPR was kindly donated by Prof. F.P. Guengerich (Nashville,

Tennessee USA). Human liver cytosolic fraction representing a pool from 15 individual

donors, was obtained from CellzDirect Inc., North Carolina, (USA). Purified recombinant

human GST P1-1 was a gift from Prof. M. Lo Bello (University of Rome, Italy) and GST

A1-1 and GST M1-1 were kindly donated by Dr. Jan J.P. Bogaards (TNO Zeist, The

Netherlands). The specific activities with CDNB (1 mM) of these three isoenzymes are

24.7 U/mg, 123 U/mg and 262 U/mg protein, respectively [30]. The Ghanaian medicinal

plants were obtained from Baba Issah Yemoh of Haatso Accra, Ghana, and identified

by John Y. Amponsah, Herbarium technician at the University of Ghana herbarium,

Accra, where voucher specimens of all the plants used were deposited (Table 1). All

other chemicals were of analytical grade and obtained from standard suppliers.

2.2. Plant extracts

Aqueous extracts of the plant materials were prepared according to the method

described by Ayisi et al [31]. Briefly, 2 g of air dried samples of each of the plant

materials used, Cassia alata, Cassia siamea, Lactuca taraxicifolia, Momordica charantia

and Morinda lucida leaves, Phyllanthus amarus whole plant, leaves, stems, roots and

Tridax procumbens whole plant (Table 1) was boiled in 100 ml distilled water for 10 min

and filtered. The aqueous extracts were lyophilized and stored at –20oC until use.

Page 142: thuoc từ curcumin

132

2.3. CYP expression and membrane isolation

The plasmids BMX100/h1A2, pCWh2C9hNPR, pSP19T7LT_2D6 and pCWh3A4 were

transformed into Escherichia coli strain JM109. Expression of the CYPs was carried out

in 3-litre flasks containing 300 ml terrific broth (TB) with 1mM !-aminolevulinic acid, 0.5

mM thiamine, 400 µl/L trace elements, 100 µg/ml ampicillin, 1 mM isopropyl-"-D-

thiogalactopyranoside (IPTG), 0.5 mM FeCl3 (for CYP2D6 and CYP3A4 only) and 30

µg/ml kanamycin (for CYP3A4 only). The culture media were inoculated with 3 ml

overnight cultures of bacteria containing plasmids for the various CYPs. The cell

cultures were incubated for about 40 h at 28 oC and 125 rpm and CYP contents were

determined using the carbon monoxide (CO) difference spectra as described by Omura

and Sato [32]. Cells were pelleted by centrifugation (4000 g, 4 oC, 15 min) and

resuspended in 30 ml Tris-Sucrose-EDTA (TSE) buffer (50 mM Tris-acetate buffer pH

7.6, 250 mM sucrose, 0.25 mM EDTA). Cells were treated with 0.5 mg/ml lysozyme

prior to disruption by French press (1000 psi, 3 repeats). The membranes containing the

human CYPs were isolated by ultracentrifugation in a Beckmann 50.2Ti rotor (60 min,

100,000 g, 4 oC), resuspended in TSE buffer and stored at –80 oC until use.

2.4. Rat liver cytosol

Rat liver cytosol was prepared from untreated rats as described [33]. Briefly, isolated rat

liver samples were homogenized in 2 volumes of 50 mM potassium phosphate buffer

(pH 7.4) containing 0.9% sodium chloride, using a Potter-Elvehjem homogenizer at 4oC.

Cytosolic fractions were obtained by centrifuging the homogenate fraction for 20 min at

12000 g. The supernatant was further centrifuged for 60 min at 100000 g. Protein

concentration was determined by the method of Bradford [34]. The cytosolic fractions

were stored at –20 oC.

2.5. CYP inhibition assays

2.5.1. 7-Methoxyresorufin O-demethylation and dibenzylfluorescein O-debenzylation

Inhibition of the activities of human CYP1A2 and CYP3A4, by extracts of Cassia alata,

Cassia siamea, Lactuca taraxicifolia, Momordica charantia and Morinda lucida leaves,

Phyllanthus amarus whole plant, leaves, stems, roots and Tridax procumbens whole

Page 143: thuoc từ curcumin

133

plant, was determined by measuring the formation of resorufin from methoxyresorufin

[29] and fluorescein from dibenzylfluorescein (DBF) [35]. The extracts were first

screened at a high concentration (1000 !g/ml) for their inhibitory activities towards the

CYPs. IC50 values were subsequently only determined for extracts showing > 70%

inhibition. The assay mixture contained 0.1 M sodium phosphate buffer (pH 7.4), 5 µM

methoxyresorufin (CYP1A2) or 0.5 µM DBF (CYP3A4), 16 nM CYP1A2 or 17 nM

CYP3A4, plant extract and 100 µM NADPH in a black coaster 96-well plate in a final

volume of 200 µl. For IC50 determinations, the concentration range of each extract used

was 1.4-1000 µg/ml. The plant extracts were dissolved in water before use. Reaction

mixtures were pre-incubated for 5 min, and reactions were initiated by the addition of

NADPH and terminated after 10 min with 75 µl of 80% acetonitrile and 20% 0.5 M tris

solution (CYP1A2) or 2 N NaOH (CYP3A4). All measurements were performed in

duplicate or triplicate and at 37oC. Resorufin and fluorescein formation was measured

spectrophotometrically on a Victor2 1420 multilabel counter (!ex = 530 nm, !em = 586 nm

and !ex = 485 nm, !em = 535 nm respectively). Metabolite formation was linear for at

least 10 min (data not shown).

2.5.2. Diclofenac hydroxylation

Inhibition of the activities of human CYP2C9, by the plants extracts was determined by

measuring the inhibition of formation of 4-hydroxydiclofenac from diclofenac as

described [36] with slight modifications. Reaction mixtures (500 !l) consisted of

expressed enzyme (98 nM CYP2C9), 100 !M NADPH, 100 mM potassium phosphate

buffer (pH 7.4), substrate (6 !M diclofenac) and plant extract. The plant extracts were

initially screened for inhibitory activities towards CYP2C9 at a high concentration (1000

!g/ml). Subsequently, IC50 determinations were performed using only extracts that

showed >70% inhibition at high concentration (range 1.4-1000 !g/ml). Reactions were

initiated by adding NADPH after a pre-incubation period of 5 min. Incubations were

allowed to proceed at 37oC for 10 min and terminated by the addition of 200 !l

methanol. The incubation mixtures were centrifuged at 11250 g for 3 min, after which

the supernatant was analyzed by isocratic HPLC method. The mobile phase consisted

of 60% (v/v) 20 mM potassium phosphate buffer (pH 7.4), 22.5% (v/v) methanol and

Page 144: thuoc từ curcumin

134

17.5% (v/v) acetonitrile. Peaks were monitored by UV detection at wavelength 280 nm.

Under these conditions, retention times for 4-hydroxydiclofenac and diclofenac were 5.0

and 24.1 min respectively.

2.5.3. Dextromethorphan O-demethylation

Inhibition of CYP2D6 activity by the plant extracts were evaluated by the method of Ko

et al. [37] with slight modifications. The extracts were first screened for CYP2D6

inhibitory activity at 1000 µg/ml and then IC50 determinations were performed on

extracts showing > 70% inhibition. Briefly, the reaction mixture consisted of 4.5 µM

dextromethorphan, 18.2 nM CYP2D6 and 90.9 µM NADPH. Plant extracts were tested

in the concentration range 1.4-1000 µg/ml. Reactions were carried out at 37oC and

terminated after 40 min with 60 mM zinc sulphate solution. Product formed

(dextrorphan) was measured using an isocratic HPLC with fluorescence detection

method (!ex=280 nm, !em=310 nm), and a C18 column (100 mm x 3 mm, 5 µm particle

size, Chromspher). Product formation was linear for at least 45 min. The mobile phase

consisted of 24% (v/v) acetonitrile and 0.1% (v/v) triethylamine adjusted to pH 3 with

perchloric acid. The flow rate was 0.6 ml/min. The retention time of dextrorphan was 3.4

min and of dextromethorphan, 24.5 min.

2.6. GST inhibition assays

Inhibition of the activities of cytosolic GSTs by the plant extracts was assessed as

described previously [30] with slight modifications. GST-mediated conjugation of 1-

chloro-2,4-dinitrobenzene (CDNB) to glutathione (GSH) was measured at room

temperature (25oC) using an Ultrospec 2000 Pharmacia Biotech UV/visible

spectrophotometer at the wavelength of 340 nm for 2 min. Incubation mixtures (1 ml)

contained 0.1 M potassium phosphate buffer pH 6.5, 400 µM CDNB, 1 mM GSH, and

GST enzymes (1.04 µg/ml GST A1-1, 1.48 µg/ml GST M1-1, 1.25 µg/ml GST P1-1, or

32.0 µg/ml and 1.7 µg/ml human and rat liver cytosol, respectively). The plant extracts

were each firstly tested at a concentration of 500 µg/ml. Subsequently, extracts showing

> 70% inhibitions were selected for IC50 determination in the concentration range 0.69-

Page 145: thuoc từ curcumin

135

500 µg/ml. All assays were linear functions of protein concentration and of time for at

least 2 min.

3. Results

3.1. Inhibition of CYPs

The method used, in preparation of aqueous plant extracts in the present study, is

similar to that used in preparation of these aqueous extracts for consumption in Ghana,

where the plant material is boiled for at least 10 minutes. The lyophilization and

resuspension of plant material is not likely to cause significant changes in the properties

of the extracts. This procedure is also been used by several investigators [8,31,41,42].

The plant extracts that showed >70% inhibition of CYP1A2 (methoxyresorufin O-

methylase) activity at 1000 µg/ml include, Cassia alata and Lactuca taraxicifolia leaves,

Phyllanthus amarus whole plant, leaves, stems, roots, and Tridax procumbens whole

plant (data not shown). IC50 values subsequently determined (Table 2) showed

strongest inhibition of CYP1A2 by Cassia alata leaves (IC50 = 28.5 µg/ml), Lactuca

taraxicifolia leaves (IC50 = 39.9 µg/ml) and Phyllanthus amarus stems and whole plant

(IC50 = 38.1 and 47.5 µg/ml, respectively). At a concentration of 1000 µg/ml, all plant

extracts except Morinda lucida leaves, and Cassia alata leaves, showed >70%

inhibitory activity towards CYP2C9 (data not shown). Only Phyllanthus amarus, stems,

roots and whole plant exhibited comparatively strong inhibitory activities towards

CYP2C9, with IC50 values of 63.4, 74.1 and 86.0 µg/ml, respectively (Table 2). Two out

of the seven plants tested in this study, showed > 70% inhibition of CYP2D6 at a

concentration of 1000 µg/ml (data not shown). Only extracts of Cassia alata leaves and

Phyllanthus amarus demonstrated > 70% inhibition of CYP2D6 activity. IC50 values

subsequently obtained were ranging from 45.8-182.0 µg/ml. The roots of Phyllanthus

amarus showed the strongest inhibition of CYP2D6, and the leaves the weakest, with

IC50 values of 45.8 and 182.0 µg/ml, respectively (Table 2).

Page 146: thuoc từ curcumin

136

Table 2. IC50 values (µg/ml) for human recombinant CYPs with aqueous plant extracts

Plant extract CYP1A2 CYP3A4 CYP2C9 CYP2D6

Cassia alata (Leaves) 28.3 + 2.42 158.8 + 9.62 nd 165.5 + 7.50

Cassia siamea (Leaves)

nd nd 346.5 + 20.93 nd

Lactuca taraxicifolia (Leaves)

39.9 + 11.93 nd 396.3 + 2.26 nd

Momordica charantia (Leaves)

nd nd 425.9 + 2.33 nd

Morinda lucida (Leaves)

nd nd nd nd

Phyllanthus amarus (Leaves)

92.2 + 12.85 79.2 + 0.41 127.5 + 3.96 182.0 + 4.81

Phyllanthus amarus (Roots)

134.3 + 17.25 80.4 + 5.29 74.1 + 0.10 45.8 + 2.40

Phyllanthus amarus (Stems)

38.1 + 3.14 146.8 + 17.54 63.4 + 4.53 164.0 + 5.37

Phyllanthus amarus (Whole plant)

47.5 + 1.27 97.0 + 18.74 86.0 + 5.26 48.5 + 0.95

Tridax procumbens (Whole plant)

127.5 + 6.08 nd nd nd

Values are means + standard deviation of at least two experiments as described in the Methods section.

CYP3A4 was also inhibited by extracts of Cassia alata leaves and Phyllanthus amarus,

with >70% inhibition at 1000 µg/ml (data not shown) and IC50 values of these plant

materials were in the range 79.2-158.8 µg/ml (Table 2).

3.2. Inhibition of GSTs

Inhibition of GST mediated conjugation of GSH to CDNB by the plant extracts was

investigated with GSTs in human and rat liver cytosol and recombinant human GSTA1-

1, GSTM1-1 and GSTP1-1 isoenzymes. When tested at 500 µg/ml, all plant extracts

showed >70% inhibition of rat liver cytosolic GSTs except Cassia siamea and Morinda

lucida leaf extracts. The human liver cytosol i.e. GSTs activities, were inhibited at 500

Page 147: thuoc từ curcumin

137

µg/ml by >70% by six of the plant extracts, whereas Cassia siamea, Lactuca

taraxicifolia and, Morinda lucida leaves and Tridax procumbens whole plant showed

<70% inhibition (data not shown). The corresponding IC50 values obtained with rat liver

cytosol were in the range of 8.5-139.4 µg/ml, whilst those with human liver cytosol were

25.2-95.5 µg/ml (Table 3). Determination of the effects of the plant extracts (500 µg/ml)

on human recombinant GSTs, showed that all extracts, except Lactuca taraxicifolia

leaves, inhibited GST M1-1 by >70%, whilst GST A1-1 and GST P1-1 were each

inhibited by only two plants, including Phyllanthus amarus and Momordica charantia or

Cassia alata respectively.

Table 3. IC50 values (!g/ml) for human and rat liver cytosolic GSTs with aqueous plant extracts

Plant extract Human liver Rat liver

Cassia alata (Leaves) 83.2 + 6.94 41.9 + 3.09

Cassia siamea (Leaves) nd nd

Lactuca taraxicifolia (Leaves) nd 112.8 + 4.60

Momordica charantia (Leaves) 95.5 + 0.83 29.0 + 1.33

Morinda lucida (Leaves) nd nd

Phyllanthus amarus (Leaves) 58.8 + 7.79 50.0 + 4.74

Phyllanthus amarus (Roots) 47.0 + 4.54 8.5 + 0.98

Phyllanthus amarus (Stems) 25.2 + 4.67 28.9 + 0.24

Phyllanthus amarus (Whole plant) 46.6 + 1.14 21.6 + 1.70

Tridax procumbens (Whole plant) nd 139.4 + 3.46

Values are means + standard deviation of two experiments as described in the Methods section.

nd: not determined because percent inhibition at 500 !g/ml was < 70%

IC50 values for the extracts showing >70% inhibitions are shown in Table 4. The IC50

values for extracts inhibiting GST A1-1 ranged from 8.9-159.0 µg/ml, with Phyllanthus

amarus roots being the strongest inhibitor and Momordica charantia leaves the

Page 148: thuoc từ curcumin

138

weakest. With respect to GST M1-1, the IC50 values obtained were in the range 3.6-50.0

µg/ml, whereas that of GST P1-1 was 68.6-157.0 µg/ml (Table 4).

Table 4. Inhibition of human recombinant GSTs by aqueous plant extracts (IC50, !g/ml)

Plant extract GSTA1-1 GSTM1-1 GSTP1-1

Cassia alata (Leaves) nd 5.6 + 0.02 68.6 + 15.64

Cassia siamea (Leaves) nd 50.0 + 1.59 nd

Lactuca taraxicifolia (Leaves) nd nd nd

Momordica charantia (Leaves) 159.0 + 10.82 16.5 + 1.19 nd

Morinda lucida (Leaves) nd 42.4 + 2.69 nd

Phyllanthus amarus (Leaves) 36.6 + 7.92 15.6 + 1.73 157.0 + 21.10

Phyllanthus amarus (Roots) 8.9 + 0.50 3.6 + 0.21 98.2 + 6.02

Phyllanthus amarus (Stems) 42.0 + 6.65 8.8 + 0.42 133.8 + 22.00

Phyllanthus amarus (Whole plant) 47.5 + 0.99 9.7 + 0.99 96.2 + 16.33

Tridax procumbens (Whole plant) nd 40.6 + 12.51 nd

Values are means + standard deviation of two experiments as described in the Methods section.

nd: not determined because percent inhibition at 500 !g/ml was < 70%

In these cases Phyllanthus amarus roots and Cassia alata leaves were the strongest

inhibitors whilst Cassia siamea and Phyllanthus amarus leaves were the weakest

inhibitors respectively.

4. Discussion

In the present study, inhibitory potencies of seven Ghanaian medicinal plants towards

recombinant human CYP1A2, CYP2C9, CYP2D6 and CYP3A4, recombinant human

GSTA1-1, GSTM1-1 and GSTP1-1 and human and rat liver cytosolic GSTs were

investigated. Cassia alata extract was the most potent inhibitor of CYP1A2 activity.

Approximately, a 3 or 4 times more potent activity was observed with Cassia alata as

compared to the other extracts. A recent in vitro study reported a potent inhibition (IC50

Page 149: thuoc từ curcumin

139

= 7.8 µg/ml) of CYP1A1/2 (methoxyresorufin O-demethylase) activity in rat liver

microsomes, by methanol extracts of Phyllanthus amarus [10]. Interestingly, aqueous

extracts of Phyllanthus amarus have been found to also possess anti-carcinogenic

activity towards 20-methylcholanthrene induced sarcoma in mice [8]. A dose of 750

mg/kg body weight resulted in 75% survival in 180 days as compared to an untreated

control group with no survivors [8]. The observed anti-carcinogenic effect and the low

IC50 value for inhibition of CYP1A1/2 [10], suggest possible of interference at the level of

CYP1A1/2-mediated bioactivation.

CYP2C9 was inhibited by extracts of Phyllanthus amarus, Cassia siamea,

Lactuca taraxicifolia and Momordica charantia by more than 70%. However, comparing

IC50 values, extracts of Phyllanthus amarus inhibited CYP2C9 about 3-6 fold more

strongly than the other extracts. Extracts of Phyllanthus amarus, as well as Cassia alata

were the only extracts with more than 70% inhibition of CYP2D6 at the highest

concentration used. The roots and whole plant extracts of Phyllanthus amarus showed

about 3-fold more potent activities than the leaves, stems and Cassia alata extracts.

Inhibition of CYP3A4 activity by >70% was also observed with the extracts of

Phyllanthus amarus and Cassia alata. CYP3A4 is responsible for the metabolism of

about 50% of the drugs currently on the market [38]. Therefore herbs inhibiting this

enzyme do have a relatively high potential to cause significant herb-drug interactions.

Although the inhibitory constituents of the plant extracts used are not known,

flavonoids which are nearly ubiquitously present in plants [39], have been found to

exhibit varying potencies of inhibitory activities towards human CYP1A2. For examples,

quercetin, chrysin, galangin, morin and apigenin inhibited CYP1A2 with IC50 values of

169.0, 0.2, 3.1, 9.5 and 1.4 µM, respectively [40]. Especially in the case of Phyllanthus

amarus, flavonoids including quercetin have been identified [8]. Earlier studies on

extracts of Ginkgo biloba reported significant inhibition of CYP2C9 with Ki = 14.0 µg/ml

by the whole extract, whilst the flavonoidic and terpenoidic fractions showed inhibitory

activities with Ki values of 4.9 and 15 µg/ml, respectively [41]. Methanol extract of

Ginkgo biloba was also shown to inhibit recombinant human CYP3A4 with Ki value

155.0 µg/ml, whereas the flavonoidic fraction shown to be responsible for the

neuroprotective effect of this plant, showed Ki value of 43.0 µg/ml [41]. In addition,

Page 150: thuoc từ curcumin

140

recent studies on Indonesian medicinal plants showed the water/methanol extract of

Catharanthus roseus to be a potent inhibitor of CYP2D6 (IC50 = 11.0 µg/ml) [42]. Two

classes of compounds found in Catharanthus roseus, active as wound healing and anti-

diabetic agent, are alkaloids and tannins [43]. A methanol soluble dried aqueous extract

of this plant with the alkaloids, ajmalicine, serpentine and vindoline showed 94%

inhibition of CYP2D6 at 25 µg/ml [44].

GST activity has also been shown to be modulated by natural plant products [45].

The present study has revealed the GST inhibitory potential of seven Ghanaian

medicinal plants. Amongst the extracts tested those of Phyllanthus amarus as well as

Cassia alata and Momordica charantia were the strongest inhibitors of human liver

cytosolic GSTs. The human liver generally expresses high levels of the GST alpha (A)

class and insignificant levels of the pi (P) class [46]. The mu class of GSTs, is also

expressed in human liver, predominantly as GSTM1 and GSTM4 [46]. Rat liver GSTs

were inhibited by all the plant extracts except the extracts of Cassia siamea and

Morinda lucida. The strongest inhibition of rat liver GSTs was observed with the root

extracts of Phyllanthus amarus and not the stem, unlike in the human samples. No

correlation was found between the cytosolic human and rat data. This is likely attributed

to differences in enzyme structures and catalytic activities of GSTs between species

[26].

Investigating the effect of the plant extracts on human recombinant GSTA1-1,

GSTM1-1 and GSTP1-1 revealed significant differences in their inhibitory potential

towards these enzymes. Our study has shown a strong inhibitory potential of

Phyllanthus amarus extracts towards GSTA1-1 compared to the other extracts. Most of

the plant extracts inhibited GSTM1-1, while GSTP1-1 was also inhibited by extracts of

Phyllanthus amarus as well as Cassia alata. Flavonoids have been shown to inhibit

human GSTs in blood platelets [47]. Quercetin, kaempferol and genistein have been

shown to significantly inhibit GSTM1-1 and M2-2 [48] and quercetin GSTP1-1 [49].

Inhibition of GSTs by the Phyllanthus amarus as well as Cassia alata could possibly

result in reduced protection from toxic effects of electrophilic substances.

In conclusion, this study has shown the inhibitory potential of aqueous extracts of

seven Ghanaian medicinal plants, towards four of the most important human CYPs.

Page 151: thuoc từ curcumin

141

Most of the plant extracts appear to lack a strong potential to inhibit the CYPs tested.

Phyllanthus amarus generally appeared to be the most potent inhibitor of the CYPs.

Phyllanthus amarus was also the strongest inhibitor of the GSTs tested. Since usually

very high concentrations of plant extracts are consumed, regular human consumption of

these extracts could well inhibit GSTA1-1 and GSTM1-1 activities in the liver and

GSTP1-1 in erythrocytes. Inhibition of GSTs may be beneficial for cancer therapy, but in

normal cells GST inhibition can also result in increased toxicity due to a reduced

protection against electrophilic chemicals or metabolites. Therefore there is a need to

further identify plant products with GST inhibitory potentials. Further investigations are

required to determine the clinical implications of the present CYP and GST inhibition

results and to identify the compounds in the plant extracts responsible for the observed

inhibitory activities.

References

1. Eisenberg, D.M., Davis, R., Ettner, S, Appel S., Wilkey, S., Van Rompay M., Kessler, R.C., 1998. Trends in alternative medicine use in the Unites States, 1990-1997: results of a follow-up national survey. JAMA 280:1569-1575.

2. Menniti-Ippolito, F., Gargiulo, L., Bologna, E., Forcella, E., Raschetti, R., 2002. Use of unconventio- nal medicine in Italy: a nation-wide survey. Eur J Clin Pharmacol 58:61-64.

3. Gurley, B.J., Gardner, S.F., Hubbard, M.A., Williams, D.K., Gentry, W.B., Khan, I.K., Shah, A., 2005. In vivo effects of goldenseal, kava kava, black cohosh, and valerian on human cytochrome P450 1A2, 2D6, 2E1, and 3A4/5 phenotypes. Clin Pharmacol Ther 77:415-426.

4. Coruh, N., Celep, A.G.S., Ozgokce, F., 2007. Antioxidant properties of Prangos ferulacea (L.) Lindl., Chaerophllum macropodum Boiss. and Heracleum persicum Desf. from Apiaceae family used as food in Eastern Anatolia and their inhibitory effects on glutathione S-transferase. Food Chem 100:1237-1242.

5. Zafar, A., Sharif, M.D., 2003. Pharmacokinetics, metabolism, and metabolism of atypical antipsychotics in special populations. Primary care companion J Clin Psychiatry 5:22-25.

6. Ioannides, C., 2002. Pharmacokinetic interactions between herbal remedies and medicinal drugs. Xenobiotica 32:451-478.

7. Budzinski, J.W., Foster, B.C., Vandenhoek, S., Arnason, J.T., 2000. An in vitro evaluation of human cytochrome P450 3A4 inhibition by selected commercial herbal extracts and tinctures. Phytomedi- cine 7:273-282.

8. Rajeshkumar, N.V., Joy, K.L., Kuttan, G., Ramsewak, R.S., Nair, M.G., Kuttan, R., 2002. Antitum- our and anticarcinogenic activity of Phyllanthus amarus extract. J Ethnopharmcol 81:17-22.

9. Notka, F., Meier, G., Waner, R., 2004. Concerted inhibitory activities of Phyllanthus amarus on HIV replication in vitro and ex vivo. Antiviral Res 64:93-102.

10. Hari Kumar, K.B., Kuttan, R., 2006. Inhibition of drug metabolizing enzymes (cytochrome P450) in vitro as well as in vivo by Phyllanthus amarus Schum & Thonn. Biol Pharm Bull 29:1310-1313.

11. Leu, T-H., Maa, M-C., 2002. The molecular mechanisms for the antitumorigenic effect of curcumin. Curr Med Chem-Anti-Cancer Agents 2:357-370.

12. Cole, G.M., Morihara, T., Lim, G.P., Yang, F., Begum, A., Frautschy, S.A., 2004. NSAID and antioxidant prevention of Alzheimer's disease: Lessons from in vitro and animal models. Annals NY Acad Sci 1035:68-84.

Page 152: thuoc từ curcumin

142

13. Appiah-Opong, R., Commandeur, J.N.M., van Vugt-Lussenburg, B., Vermeulen, N.P.E., 2007. Inhibition of human recombinant cytochrome P450s by curcumin and curcumin decomposition products. Toxicology 235:83-91.

14. Mathews, J.M., Etheridge, A.S., Black, S.R., 2002. Inhibition of human cytochrome P450 activities by kava extract and kavalactones. Drug Metab Disops 30:1153-1157.

15. Almeida, J.C., Grimsley, E.W., 1996. Coma from the health food store: Interaction between kava and alprazolam. Annals Intern Med 125:940-941.

16. Sparreboom, A., Cox, M.C., Acharya, M.R., Figg, W.D., 2004. Herbal remedies in the United States: Potential adverse interactions with anticancer agents. J Clin Oncol 22:2489-2503.

17. Rupika, D., Andrew, W.C.G., 2004. Herbal interactions involving cytochrome P450 enzymes: A mini review. Tox Rev 23:239-249.

18. Izzo, A.A., 2004. Herb-drug interactions: An overview of the clinical evidence. Fundamental Clin Pharmacol 19:1-16.

19. Hayeshi, R., Mukanganyama, S., Hazra, B., Abegaz, B., Hasler, J., 2004. The interaction of selected natural products with human recombinant glutathione transferases. Phytother Res 18:877- 883. 20. Commandeur, J.N., Stijntjes, G.J., Vermeulen, N.P., 1995. Enzymes and transport systems involved in the formation and disposition of glutathione S-conjugates. Role in bioactivation and detoxication mechanisms of xenobiotics. Pharmacol Rev 47:271-330. 21. Frova, C., 2006. Glutathione transferases in the genomics era: New insights and perspectives.

Biomol Eng 23:149-169. 22. Mukanganyama, S., Widersten, M., Naik, Y.S., Mannervik, B., Hasler, J.A., 2002. Inhibition of glutathione S-transferases by antimalarial drugs possible implications for circumventing anticancer drug resistance. Int J Cancer 97:700-705. 23. Capela, J.P., Macedo, C., Branco, P.S., Ferreira, L.M., Lobo, A.M., Fernandes, E., Remiao,

F., Bastos, M.L., Dirnagl, U., Meisel, A., Carvalho, F., 2007. Neurotoxicity mechanisms of thioether ecstasy metabolites. Neuroscience 146:1743-1757.

24. Buabeng, K.O., Duwiejua, M., Dodoo, A.N., Matowe, L.K., Enlund, H., 2007. Self-reported use of anti-malarial drugs and health facility management of malaria in Ghana. Malar J 6:85.

25. Eagling, V.A., Tjia, J.F., Back, D.J., 1998. Differential selectivity of cytochrome P450 inhibitors against probe substrates in human and rat liver microsomes. Br J Clin Pharmacol 45:107-114.

26. Aurbek, N., Thiermann, H., Szinicz, L., Eyer, P., Worek, F., 2006. Application of kinetic-based computer modelling to evaluate the efficacy of HI 6 in percutaneous VX poisoning. Toxicology 224:74-80.

27. Idu, M., Oronsaye, F.E., Igeleke, C.L., Omonigho, S.E., Omogbeme, O.E., Ayinde, B.A., 2006. Preliminary Investigation on the Phytochemistry and Antimicrobial Activity of Senna alata L. Leaves. J Appl Sci 6:2481-2485.

28. Raj, S.K., Khan, M.S, Singh, R., Kumari, N., Prakash, D., 2005. Occurrence of yellow mosaic Gem- iniviral disease on bitter gourd (Momordica charantia) and its impact on phytochemical contents. Int. J. Food Sci Nutr 56:185-192.

29. Burke, M.D., Thompson, S., Elcombe, C.R., Halpert, J., Haaparanta, T., Mayer, R.T., 1985. Ethoxy-, pentoxy- and benzyloxyphenoxazones and homologues: a series of substrates to distinguish between different induced cytochromes P-450. Biochem Pharmacol 34:3337-3345.

30. van Haaften, R.I.M., Haenen, G.R.M.M., van Bladeren, P.J., Bogaards, J.J.P., Evelo, C.T.A., Bast,

A., 2003. Inhibition of various glutathione S-transferase isoenzymes by RRR-!-tocopherol. Toxicol

in vitro 17:245-351. 31. Ayisi, N.K., Nyadedzor,C., 2003. Comparative in vitro effects of AZT and extracts of Ocimum

gratissimum, Ficus polita, Clausena anisata, Alchornea cordifolia, and Elaeophorbia drupifera against HIV-1 and HIV-2 infections. Antiviral Res 58:25-33.

32. Omura, T., Sato, R., 1964. The carbon monoxide binding pigment of liver microsomes. I. Solubiliz- ation, purification and properties. J Biol Chem 239:2379-2385.

33. Rooseboom, M., Vermeulen, N.P.E., Groot, E.J., Commandeur J.N.M., 2002. Tissue distribution of cytosolic beta-elimination reactions of selenocysteine Se-conjugates in rat and human. Chem Biol Interact 140:243-264.

34. Bradford, M. M., 1976. A rapid and sensitive method for the quantification of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem 72:248-254.

Page 153: thuoc từ curcumin

143

35. Stresser, D.M., Blanchard, A.P., Turner, S.D., Erve, J.C.L., Dandeneau, A.A., Miller, V.P., Crespi C.L., 2000. Substrate-dependent modulation of CYP3A4 catalytic activity: Analysis of 27 test comp- ounds with 4 fluorometric substrates. Drug Metab Dispos 28:1440-1448.

36. Walsky, R.L., Obach, R.S., 2004. Validation assays for human cytochrome P450 activities. Drug Metab Dispos 32:647-660.

37. Ko, J.W., Desta, Z., Soukhova, N.V., Tracy, T., Flockhart, D.A., 2000. In vitro inhibition of the cytochrome P450 (CYP450) system by the antiplatelet drug ticlopidine: potent effect on CYP2C19 and CYP2D6. Br J Clin Pharmacol 49:343-351. 38. Sarlis, N.J., Gourgiotis, L., 2005. Hormonal effects on drug metabolism through the CYP system:

Perspectives on their potential significance in the era of pharmacogenomics. Curr Drug Targets- Immune Endocrine & Metabolic disorders 5:439-448.

39. Middleton, E. Jr., Kandaswami, C., Theoharides, T.C., 2000. The effect of plant flavonoids on mammalian cells: Implications for inflammation, heart disease and cancer. Pharmacol Rev 52:673- 751.

40. Lee, H., Yeom, H., Kim, Y.G., Yoon, C.N., Jin, C., Choi, J.S., Kim, B-R., Kim, D-H., 1998. Structure- related inhibition of human hepatic caffeine N3-demethylation by naturally occurring flavonoids. Biochem Pharmacol 55:1369-1375.

41. Gaudineau, C., Beckerman, R., Welbourn, S., Auclair, K., 2004. Inhibition of human P450 enzymes by multiple constituent of Ginkgo biloba extract. Biochem Biophys Res Commun 318:1072-1078.

42. Usia, T., Iwata, H., Hiratsuka, A., Watabe, T., Kadota, S., Tezuka, Y., 2006. CYP3A4 and CYP2D6 inhibitory activities of Indonesian medicinal plants. Phytomedicine 13:67-73.

43. Nayak, B.S., Pereira, L.M.P., 2006. Catharanthus roseus flower extract has wound-healing activity in Sprague Dawley rats. BMC Complement Altern Med 6:41.

44. Usia, T., Watabe, T., Kadota, S., Tezuka, Y., 2005. Cytochrome P450 2D6 (CYP2D6) inhibitory constituents of Catharanthus roseus. Biol Pharm Bull 28:1021-1024.

45. Zhang, K., Wong, K.P., Chow, P., 2003. Conjugation of chlorambucil with GSH by GST purified from human colon adenocarcinoma cells and its inhibition by plant polyphenols. Life Sci 72:2629- 2640.

46. Eaton, D.L., Bammler, T.K., 1999. Concise review of glutathione S- transferases and their significa- nce to toxicology. Toxicol Sci 49:156-164.

47. Ghazali, R., Waring, R.H., 1999. Effects of flavonoids on glutathione-S-transferase in human blood platelets, rat liver, rat kidney, HT-29 colon adenocarcinoma cell-lines: potential in drug metabolism and chemoprevention. Med Sci Res 27:449-451.

48. Burg, D., Mulder, G.J., 2002. Glutathione conjugates and their synthetic derivatives as inhibitors of glutathione-dependent encimes involved in cancer and drug resistance. Drug Metab Rev 34:821- 863.

49. van Zanden, J.J., Geraets, L., Wortelboer, H.M., van Bladeren, P.J., Rietjens, I.M.C.M., Cnubben, N.H.P., 2004. Structural requirements for the flavonoid-mediated modulation of glutathione S- transferase P1-1 and GS-X pump activity in MCF7 breast cancer cells. Biochem Pharmacol 67:1607-1617.

Page 154: thuoc từ curcumin

144

Page 155: thuoc từ curcumin

145

Summary, conclusions and perspectives

Page 156: thuoc từ curcumin

146

Chapter 7

Summary and conclusions

Cytochrome P450 (CYP)-mediated drug–drug interactions are major causes of

attrition during drug development, black-box warnings of marketed drugs or

withdrawals of drugs from the market, due to adverse effects [1,2]. Inhibition of CYP

activity is a major cause of drug-drug interactions, since it results in accumulation of

drugs which could otherwise be metabolized into less toxic products. Alternatively,

CYP induction may also contribute to serious drug-drug interactions. Thus, early

evaluation of new chemical entities (NCEs) for both inhibitory and inducing drug-

drug interaction potentials is considered useful and cost effective in drug

development. Inhibition and induction of glutathione S-transferases (GSTs) by drugs

and other xenobiotics may also results in toxicities and adverse drug reactions, since

these enzymes are primary detoxification, and in some cases toxification enzymes in

the body [3].

The aim of the investigations described in this thesis is to evaluate the drug-drug

-food and -herb interaction potentials at the level of CYP and GST inhibition of plant

and plant derived components. The CYP and GST inhibitory potentials of curcumin,

which was used as a model compound, were compared with that of three series of

synthetic curcumin analogues including, 2,6-dibenzylidenecyclohexanone (A series),

2,5-dibenzylidenecyclopentanone (B series) and 1,4-pentadiene-3-one (C series) [4].

Structure-activity relationships were analyzed to guide future synthesis of

compounds with less (or more) inhibitory potencies towards CYPs and GSTs. The

inhibitory potentials of seven Ghanaian medicinal plants on CYP and GST activities

were also assessed.

In Chapter 1 a general introduction of the background of CYP-mediated drug-drug

interactions, due to CYP-inhibition and -induction, and the detoxification and

toxification roles of GSTs are discussed. Co-administration of two or more drugs has

been shown to be a potential cause of drug-drug interactions with possible serious

adverse side effects [5]. CYPs contribute significantly to the elimination of drugs

Page 157: thuoc từ curcumin

147

from the body through CYP-mediated metabolism. Of the human CYP isoenzymes a

majority is important for the metabolism of drugs currently on the market [6]. These

enzymes include CYP3A4, CYP2D6, CYP2C9, CYP2C19, CYP2B6, CYP2A6 and

CYP1A2. Knowledge of CYP inhibitory properties of NCEs and analysis of structure-

activity relationships is generally seen as necessary during the early stages of drug

discovery and development to guide synthesis of safer drug candidates and to

minimize losses due to attrition. Since natural products, including foods and herbal

medicines, have also been shown to inhibit CYPs and to cause potentially significant

adverse effects, these products should be investigated for possible drug-herb/food

interactions as well [7]. The same holds true for the GST-inhibitory properties of

NCEs, drug candidates, foods and herbal or other plant products, therefore they

should be assessed as well in order to avoid potential toxicity. In the present studies,

human GSTA1-1, GSTM1-1 and GSTP1-1 were selected as targets.

Curcumin, a derivative of Curcuma longa and known to possess many important

pharmacological properties was used as a model compound in the present

investigations. The unique availability of a series of 40 synthetic structural analogues

of curcumin, 34 of which with appropriate solubility properties, (Figure 1) allowed us

to investigate structure-activity relationships for both the CYP- and GST-

isoenzymes. Using the experimental methods developed, we also investigated the

inhibitory potential of the Ghanaian medicinal plant extracts of, Cassia alata, Cassia

siamea, Lactuca taraxicifolia, Momordica charantia and Morinda lucida, Phyllanthus

amarus and Tridax procumbens toward the human CYP- and GST-isoenzymes.

OO

OH

O

OH

OCH3

CH3

Curcumin

33

1

2

O

R

R

R R

R

R

2

1

1

33

R1

R

R

O

R

R

R

22

1

33

R1

R

R

O

R

R

R

22

2,6-dibenzylidenecyclohexanone (A) 2,5-dibenzylidenecyclopentanone (B)

1,4-pentadiene-3-one (C)

Figure 1. Chemical structures of curcumin and curcumin analogues.

Page 158: thuoc từ curcumin

148

Chapter 2 presents a brief review of studies that have been performed to investigate

pharmacokinetic, metabolism and drug-drug interaction properties of curcumin.

Animal experiments performed, using radioactivity to monitor curcumin

concentrations in plasma and tissues after intraperitoneal administration, showed

low concentrations of about 5 pmol/ml in rat plasma, and rapid removal of the parent

compound from the tissues [8]. Similarly, phase I clinical trials on curcumin have

revealed low plasma concentrations of 11 nmol/L upon oral administration of 3.6 g of

Curcumin [9]. In order to enhance the bioavailability of curcumin, it has been

administered with piperine, solubilized with N,N-dimethylacetamide, polyethylene

glycol (PEG 400) and 40% isotonic dextrose or formulated with micelles or

phoshatidylcholine [10-13]. In addition, curcumin nanoparticles, liposomal curcumin

and structural analogues of curcumin have also been employed to enhance

bioavailability. The highest bioavailability reached (154% in rats) was with a

curcumin/piperine combination [11]. Curcumin has been shown to be metabolized in

sub cellular liver fractions of mouse, rat and humans in a similar way. Phase I

metabolites include reductive and oxidative metabolites of curcumin (including

hexahydrocurcuminol, hexahydro- and tetrahydrocurcumin), whilst phase II

metabolites were predominantly glucuronides and sulfates [14]. Early in vitro studies

with rat liver microsomes and cytosol showed that curcumin is a potent inhibitor of

CYP1A1/2, a less potent inhibitor of CYP2B1 and a potent inhibitor of GSTs [15].

Curcumin has been shown to inhibit strongly cytosolic GSTs isolated from human

melanoma cells [16]. In this thesis comprehensive inhibition has been done with

individual human CYPs and GSTs as well as with cytosolic fractions.

Chapter 3 focuses on the inhibitory activities of curcumin and curcumin

decomposition products towards the 5 major human recombinant CYPs, responsible

for the metabolism of majority of currently marketed drugs. Curcumin inhibited the

enzymes tested in decreasing order of potency (IC50) as follows: CYP2C9 >

CYP3A4 > CYP2B6 > CYP1A2 > CYP2D6 [17]. Competitive inhibition was observed

with CYP1A2, CYP2B6 and CYP3A4, whereas CYP2C9 and CYP2D6 were inhibited

Page 159: thuoc từ curcumin

149

non-competitively. Four decomposition products of curcumin did not show any

significant inhibition towards the human CYPs tested. Mechanism-based or time-

dependent inhibition by curcumin was neither observed with any of the 5 human

enzymes. Curcumin is a potent inhibitor of CYP2C9 and CYP3A4 and a moderate

inhibitor of CYP2B6, CYP1A2 and CYP2D6, with IC50 values ranging from 4.3 to

50.3 !M. It was concluded that, the inhibitory activity of curcumin towards CYP3A4

may well have implications for drug-drug interactions in the intestine, rather than in

the liver when the intestines are exposed to high concentrations upon oral ingestion

together with drugs metabolized by this enzyme. Further studies are needed to

establish the clinical implications of these results.

In Chapter 4 the inhibitory activities of thirty-three (selected out of 40) compounds

belonging to three series of curcumin analogues, 2,6-dibenzylidenecyclohexanone

(A series), 2,5-dibenzylidenecyclopentanone (B series) and 1,4-pentadiene-3-one (C

series) substituted analogues [4] towards the 5 major human recombinant CYPs

were experimentally determined. Subsequently, structure-activity relationships were

analyzed using the MOE software. Most of the curcumin analogues exhibited low

inhibitory activities towards the CYPs tested, when compared to curcumin itself [17].

Six compounds were potent inhibitors of CYP1A2, three potent towards CYP2C9

and two towards CYP2D6. None of the 2,6-dibenzylidenecyclohexanone derivatives

(A series) inhibited CYP3A4, CYP2B6 and CYP2D6, whilst one compound strongly

inhibited CYP2C9. The MOE-QSAR analysis lead to the conclusion that electrostatic

and hydrophobic descriptors, notably PEOE_VSA_FPNEG and PEOE_VSA_FHYD

are important factors of the compounds especially relating to inhibition of CYP2C9

and CYP2D6. The results of these studies are not only important because of the

new insights in structural properties for CYP-inhibition, but also because they

identified the structural analogues of curcumin without CYP-inhibitory properties. A

larger number of curcumin analogues will be required to enhance the QSAR

prediction of drug-drug interaction potentials of these compounds against all

individual CYPs.

Page 160: thuoc từ curcumin

150

GSTs are important phase II enzymes, involved in detoxification and toxification of

xenobiotics and in multidrug resistance in chemotherapy. In Chapter 5, the results

of investigations on the GST inhibitory potentials of curcumin itself and three series

of 34 (selected out of 40) structural analogues of curcumin, as well as structure-

activity relationships are presented and discussed. Curcumin was shown to be a

potent inhibitor of human recombinant GSTA1-1, GSTM1-1 and GSTP1-1 and

interestingly, the results were recently confirmed by Hayeshi et al. [18]. As observed

with CYP inhibition (Chapter 4), most of the curcumin-analogues lacked or

demonstrated weak inhibitory activities towards the human GST-isoenzymes,

GSTA1-1, GSTM1-1 and GSTP1-1, when compared to curcumin. The 2,5-

dibenzylidenecyclopentanone (B series) derivatives of curcumin showed no

significant inhibition of GSTP1-1. The most potent inhibitors, notably of GSTM1-1

and GSTP1-1, with IC50 values ranging from 0.2 to 6.1 µM were predominantly the

1,4-pentadiene-3-one derivatives of curcumin (C series). The inhibitory activities of

B14 and C10 towards GSTA1-1, with IC50 values of 0.5 and 0.6 µM respectively, and

towards human liver cytosolic GSTs could have implications for GST-mediated

protection against drug toxicities in the liver due to the high hepatic levels of this

GST-isoform [19]. This is in contrast to GSTP1-1, which is primarily expressed in

erythrocytes [20]. MOE-based QSAR analyses amongst others have delineated the

relevance of van der Waal’s surface area and lipophilicity factors (SMR_VSA7,

SlogP_VSA4 and dipole, PEOE_VSA-0, SlogP_VSA8 and PEOE_RPC,

respectively) for the inhibition of GSTA1-1 and GSTM1-1. The results of this chapter

may be useful in the design and synthesis of curcumin analogues with either more or

with less susceptibility to GST inhibition. This is important because of the role of

GSTs in detoxification, toxification of xenobiotics and multidrug resistance in

chemotherapy.

Herbal products and traditional medicines have also been shown to be able to inhibit

important drug metabolizing enzymes such as CYPs and GSTs, thus resulting in

clinically relevant adverse drug reactions or toxicities [21]. Therefore, Chapter 6

focuses on the interactions between the major human drug metabolizing CYPs and

Page 161: thuoc từ curcumin

151

GSTs and a selection of the most important Ghanaian medicinal plants. A standard

procedure for preparation of aqueous extracts of the seven Ghanaian plants,

involved freeze drying of the decoction of plants [22]. Most of the aqueous plant

extracts tested appeared to lack the potential to strongly inhibit the four human

CYPs and three GSTs tested. Generally, Phyllanthus amarus extracts were the most

potent inhibitors of CYP1A2, CYP2C9, CYP2D6, CYP3A4, with IC50 values ranging

from 38.1 to 97.0 µM and GSTA1-1, GSTM1-1 and GSTP1-1 with IC50 values

ranging from 3.6 to 98.2 µM. Although the inhibitory constituents of the Ghanaian

plant extracts are not known yet, flavonoids have been found in Phyllanthus amarus,

Cassia alata and Mormodica charantia. Recent studies on extracts of Ginkgo biloba

have shown that flavonoidic and terpenoidic fractions as well as whole plant extracts

strongly inhibit CYP2C9 [23]. Prangos ferulacea, also demonstrated strong inhibitory

activity towards sheep liver cytosolic GSTs [24]. Clinical implications of the present

CYP- and GST-inhibition results on the Ghanaian medicinal plant extracts are yet to

be established. However, obviously it is necessary that all herbal preparations are

screened for their potentials to cause herb-drug interactions.

Conclusion and perspectives

The major objective of the research described in this thesis was to evaluate the

interactions between plant-derived products and CYPs and GSTs, two important

human drug metabolizing enzyme systems. Such interactions might implicate

clinically relevant drug-drug interactions at the level of microsomal CYPs and

cytosolic GSTs. Firstly, the inhibition of five major human recombinant drug-

metabolizing CYPs by curcumin, derived from Curcuma longa and chosen as a

model compound, and four curcumin decomposition products have been

investigated. Subsequently, thirty-four synthetic curcumin analogues (based on

solubility criteria selected out of 40) were measured and analyzed by MOE-based

methods for quantitative structure-activity relationships (QSARs). Finally, the

inhibitory effects of seven important Ghanaian medicinal plants were also assessed

on the human drug-metabolizing CYPs. In an analogous way, the inhibition of

recombinant human GSTs and human and rat liver cytosolic GSTs by curcumin

Page 162: thuoc từ curcumin

152

analogues and seven Ghanaian medicinal plants were studied. The general

conclusions based on the findings in the experimental sections of this thesis, and

future perspectives are discussed below.

Inhibition of CYPs by curcumin, and curcumin decomposition products

The potential for drug-drug interactions of drugs and drug candidates and of new

chemical entities (NCEs) are usually evaluated during the early stages of drug

development since it is a major cause of attrition of drugs from the market [1].

Curcumin, a common food additive and a naturally occurring and synthetic

compound has been considered a promising drug candidate due to its several

pharmacological activities [25]. In this thesis, the potential for drug-drug interactions

of curcumin and four of its decomposition products was firstly assessed by their

inhibitory activities towards five important human CYPs, namely CYP1A2, CYP2B6,

CYP2C9, CYP2D6 and CYP3A4. These CYPs are responsible for the hepatic

metabolism of about 80% of drugs currently on the market and notably CYP3A4 is

also abundant in the intestine [6]. Strong competitive inhibition and mechanism-

based inhibition of enzymes are considered clinically important [1]. In the present

study, curcumin was not found to be a mechanism-based inhibitor of any of the

CYPs tested, however, it is a relatively strong competitive inhibitor of CYP3A4 and a

non-competitive inhibitor of CYP2C9. The potent inhibition of curcumin towards

CYP3A4 shown could have implications for drug-drug interactions in the intestines,

in case of high exposure of the intestines to curcumin upon oral administration.

Four decomposition products of curcumin showed no significant inhibitory

activity towards the CYPs tested, and are therefore not likely to contribute to drug-

drug interactions at the level of CYPs. The potentials for drug interactions involving

many herbs (e.g. St. John’s wort, garlic and kava) and natural compounds

(flavonoids, coumarins, caffeine and anthroquinones) have previously been

investigated [26], and some have been identified as inhibitors or inducers of various

CYP enzymes. Likewise, the inhibitory potencies of curcumin towards CYPs, have

been demonstrated in this thesis. Extrapolation of the present human in vitro data to

in vivo occurring drug-drug interactions could be achieved in principle, if the inhibitor

Page 163: thuoc từ curcumin

153

concentration in plasma and/or in the liver and the fraction of drug metabolized by a

particular CYP were known [27]. Further studies will be required, however, to predict

the clinical relevance of the strong inhibitory potential of curcumin towards CYP3A4

and CYP2C9 and the weaker inhibitory potentials towards the other human CYPs.

Inhibition of CYPs and GSTs by curcumin analogues

Due to several drawbacks of curcumin for human therapies, including an extremely

low bioavailability and significant instability at neutral to basic pH conditions [9,15],

many structural analogues of curcumin have been synthesized [4,28]. In this thesis,

we determined experimentally CYP and GST inhibitory potentials of three series of

thirty-four synthetic curcumin analogues [4] as compared to curcumin, using the five

major human CYPs mentioned above and three major human GSTs, i.e. GSTA1-1,

GSTM1-1 and GSTP1-1. Most of the analogues were less potent inhibitors of the

human CYPs and GSTs tested as compared to curcumin. The 1,4-pentadiene-3-one

derivatives (C series) contained some more potent inhibitors of CYP1A2, CYP2C9

and CYP2D6. The most potent inhibitors of CYP1A2 lacked a para-hydroxyl moiety.

Inhibition and induction of GST may have implications for detoxification,

toxification of endogenous and exogenous compounds and chemotherapy against

cancer cells [3]. With respect to GST inhibition, this work has shown that most of the

curcumin analogues tested are less potent inhibitors of the human GSTs employed

compared to curcumin. Interestingly, a recent study confirmed the fact that curcumin

is a potent inhibitor of GSTA1-1, GSTM1-1 and GSTP1-1 [18]. Our studies have

shown that GSTA1-1 and GSTM1-1 were most susceptible to inhibition by the

curcumin analogues. As observed with the human CYPs, the 1,4-pentadiene-3-one

derivatives of curcumin (C series) were generally the most potent inhibitors of the

GSTs tested, suggesting that the absence of the central ring (which is present only

in the A and B series) yields structures with increased GST-inhibitory potencies. The

GSTA1-1 and GSTM1-1 isoforms are predominantly expressed in the human liver,

whereas the hepatic levels of GSTP1-1 are insignificant [19]. However, GSTP1-1 is

the most abundant isoform in erythrocytes [20]. Thus, inhibition of GSTA1-1 and

GSTM1-1 in the liver could have implications for hepatotoxicity or hepatoprotection,

Page 164: thuoc từ curcumin

154

whilst inhibition of GSTP1-1 could lead to a decrease protection against

electrophiles and oxidative stress in erythrocytes.

Most of the curcumin analogues exhibited less inhibitory activities towards the

CYPs and GSTs tested compared to curcumin, from this point of view these could

be better alternative drug candidates. To our knowledge, no studies on drug-drug

interactions potential of curcumin analogues at the level of CYPs and GSTs have

been performed. Thus, the present findings could be an important basis for further

studies on other structurally related curcumin analogues and also for consideration

of these compounds as chemotherapeutic agents, anti-oxidant, anti-inflammatory or

other pharmacological activities [25]. Further investigations, however, are required to

establish the in vivo relevance of the present in vitro results, as well as the

bioavailability and toxicity of these compounds.

Structure-activity relationships for inhibition of CYPs and GSTs by curcumin

analogues

Except for a difference in the steric bulkiness of their aromatic substituents, chemical

similarity studies indicated no significant differences between the three series of

curcumin analogues. MOE-based QSAR analysis suggested that electrostatic and

hydrophobic interactions were important factors for CYP2C9 and CYP2D6 inhibition.

In addition, our results suggest that van der Waal’s surface area and compound

lipophilicity factors are significant for the inhibition of GSTA1-1 and GSTM1-1. As

yet, no other QSAR-studies on curcumin analogues have been reported. Thus the

present results provide new insights in structure-activity relationships of curcumin

analogues at the level of CYPs and GSTs, enabling a more rational design and

synthesis of new analogues with better properties in this regard. The statistical

MOE-based QSAR approach in the present studies, provide the means to predict to

some extent the interaction between other curcumin analogues and the CYPs and

GSTs used. Additional studies with a larger number of compounds will undoubtedly

refine the models and may account for some of the compounds being more poorly

predicted. A limiting feature of any 2D-QSAR approach is its insensitivity to the

stereochemistry of the members of the data sets used and the lack of easily

Page 165: thuoc từ curcumin

155

interpretable information useful for the design of new drugs. Thus, a 3D-QSAR

approach could be better used together with the 2D approach to provide more

enhanced models for drug-drug interaction predictions with the CYPs and GSTs.

Effects of Ghanaian medicinal plants on CYPs and GSTs

Natural products, including medicinal plants and foods are generally being

considered as harmless, and therefore they are not subjected to the scrutiny of the

approval process such as with new drug candidates. However, interactions of herbal

medicines with human CYPs have been associated with strong alterations in the

pharmacokinetics of drugs such as midazolam and alprazolam and strong adverse

effects such as with St. John’s wort [7,26]. In this thesis, the potential of seven

commonly used Ghanaian medicinal plants to inhibit human CYPs and GSTs, and

thus cause herb-drug interactions have been investigated for the first time.

Phyllanthus amarus, appeared to be the most potent inhibitor of the human CYPs

tested, while in the case of CYP1A2, Cassia alata and Lactuca taraxicifolia also

showed strong inhibition. In line with the present findings, recent in vitro and in vivo

animal studies on CYP inhibition by extracts of Phyllanthus amarus showed that it is

a potent inhibitor of CYP1A1, CYP1A2 and CYP2B1 [29].

GST activity has also been shown to be modulated by natural plant products,

such as Prangos ferulacea [24]. Since GSTs are major detoxification enzymes in

humans, inhibition of these enzymes by the medicinal plants could have important

clinical and toxicological consequences. Among the medicinal plants tested,

Phyllanthus amarus strongly inhibited GSTA1-1, GSTM1-1 and GSTP1-1.

Interestingly, GSTM1-1 was susceptible to strong inhibition by all the plants except

Lactuca taraxicifolia. The compounds responsible for the observed activities and the

clinical relevance of the inhibitory activities remain to be established. Generally,

Phyllanthus amarus also demonstrated the strongest inhibitory activities towards the

GSTs tested as observed with CYPs. Since most of the other plants tested lacked

strong inhibitory potencies towards the major human CYPs and GSTs studied, most

of these plants are not likely to cause clinically important herb-drug interactions.

However, the clinical relevance of results obtained remains to be established. Due to

Page 166: thuoc từ curcumin

156

the potential of herbs and foods to cause drug-food/-herb interactions with adverse

effects as observed with Ginkgo biloba and grapefruit juice, it is imperative that other

medicinal plants, foods and natural products that are consumed are subjected to

tests, to critically assess their potential for CYP and GST inhibition. As enzymes are

proteins, and tannins present in plants are known to precipitate and inactivate

proteins, it would be useful to determine the amount of tannin in plant extracts or

foods and to test its influence on enzyme assays.

In summary, in vitro studies on the inhibitory effects of drug candidates, herbal

products and food components are cost effective pre-clinical approaches to avoid

potential adverse effects resulting from drug-drug/-food/-herb interactions. In this

thesis, we have used recombinant human CYPs and GSTs to study the CYP and

GST inhibitory potential of plant derived components, including curcumin and

Ghanaian medicinal herbs as well as synthetic curcumin analogues. Subsequently,

structure-activity relationships were evaluated to understand underlying mechanisms

and to facilitate rational design and synthesis of curcumin analogues with less

susceptibility to drug-drug interactions at the level of CYPs and GSTs. We have

shown that curcumin, a very widely used therapeutic plant product and drug

candidate with several interesting pharmacological properties, is a potent inhibitor of

important human biotransformation enzymes, notably CYP3A4 and CYP2C9.

Curcumin inhibition of CYP3A4 may have implications for drug-drug interactions in

the intestines, also due to the high doses of curcumin usually administered. The

curcumin analogues tested were generally less potent inhibitors of CYPs and GSTs

employed, and thus they appear to be better drug candidates than curcumin from

this point of view.

The statistical 2D-QSAR approaches used to analyze the CYP and GST

inhibitory activities of the curcumin analogues revealed insights of the relationships

between structure properties of the analogues and the inhibitory activities obtained.

Increasing the number of curcumin analogues and combining 2D- and 3D- QSAR

approaches may enhance the predictive power for potential drug-drug interactions.

The inhibitory activities of the Ghanaian medicinal herbs, in particular Phyllanthus

Page 167: thuoc từ curcumin

157

amarus towards the CYPs and GSTs tested require further investigations. The

potential for clinically relevant drug-food/-herb interactions involving human CYP-

and GST-biotransformation enzymes for all herbal products and food components

must be further evaluated as well. Extrapolation of human in vitro data to the human

in vivo situation is important for prediction of drug-drug interactions in vivo. This

requires the estimation of the hepatic inhibitor concentrations as well as the fraction

of drug substrate metabolized by the CYP or GST inhibited pathway. Altogether, this

thesis has uncovered important new insights in the inhibitory potential towards

important drug metabolizing enzymes of plant-derived curcumin, synthetic curcumin

analogues and Ghanaian medicinal plant extracts.

References

1. Zhang, Z.Y., Wong, Y.N., 2005. Enzyme kinetics for clinically relevant CYP inhibition. Curr drug Metab 6:241-257.

2. Pea, F., Furlanut, M., 2001. Pharmacokinetic aspects of treating infections in the intensive care unit: focus on drug interactions. Clin Pharmacokinet 40:833-868.

3. Commandeur, J.N., Stijntjes, G.J., Vermeulen, N.P., 1995. Enzymes and transport systems

involved in the formation and disposition of glutathione S-conjugates. Role in bioactivation and detoxication mechanisms of xenobiotics. Pharmacol Rev 47:271-330.

4. Sardjiman, S., Reksohadiprodjo, M., Hakim, L., van der Goot, H., Timmerman, H., 1997. 1,5-Diphenyl-1,4-pentadiene-3-ones and cyclic analogues as antioxidative agents. Synthesis and structure-activity relationship. Eur J Med Chem 32:625-636.

5. Aparasu, R., Baer, R., Aparasu, A., 2007. Clinically important potential drug-drug Interactions in outpatient settings. Res Social Adm Pharm 3:426-437.

6. Lamb, D.C., Waterman, M.R., Kelly, S.L., Guengerich, F.P., 2007. Cytochrome P450 and drug discovery. Curr Opin Biotechnol 18:1-9.

7 Gurley, B.J., Gardner, S.F., Hubbard, M.A., Williams, D.K., Gentry, W.B., Khan, I.K., Shah, A., 2005. In vivo effects of goldenseal, kava kava, black cohosh, and valerian on human cytochrome P450 1A2, 2D6, 2E1, and 3A4/5 phenotypes. Clin Pharmacol Ther 77:415-426.

8. Perkins, S., Verschoyle, R.D., Hill, K., Parveen, I., Threadgill, M.D., Sharma, R.A., Williams, M.I., Steward, W.P., Gescher, A.J., 2002. Chemopreventive efficacy and pharmacokinetics of curcumin in the Min/+ mouse, a model of familial adenomatous polyposis. Cancer Epid Biomarkers & prevention 11:535-540.

9. Sharma, R.A., Euden, S.A., Platton, S.L., Cooke, D.N., Shafayat, A., Hewitt, H.R., Marczylo, T.H., Morgan, B., Hemingway, D., Plummer, S.M., Pirmohamed, M., Gescher, A.J., Steward, W.P., 2004.Phase I clinical trial of oral curcumin: biomarkers of systemic activity and compliance. Clin Cancer Res 10:6847-6854.

10. Marczylo, T.H., Verschoyle, R.D., Cooke, D.N., Morazzoni, P., Steward, W.P., Gescher, A.J., 2007. Comparison of systemic availability of curcumin with that of curcumin formulated with phosphatidylcholine. Cancer Chemother Pharmacol 60:171-177.

11. Shoba, G., Joy, D., Joseph, T., Majeed, M., 1998. Influence of piperine on the pharmacokinetics of curcumin in animals and human volunteers. Planta Med 64:353-356.

12. Ma, Z., Shayeganpour, A., Brocks, D.R., Lavasanifar, A., Samuel, J., 2007. High performance liquid chromatography analysis of curcumin in rat plasma: application to pharmacokinetics of polymeric micellar formulation of curcumin. Biomed Chromatogr 21:546-552.

Page 168: thuoc từ curcumin

158

13. Safavy, A., Raisch, K.P., Mantena, S., Sanford, L.L., Sham, S.W., Krishna, N.R., Bonner, J.A., 2007. Design and development of water-soluble curcumin conjugates as potential anticancer agents. J Med Chem Nov 1 (Epub ahead of print).

14. Tamvakopoulos, C., Sofianos, Z.D., Garbis, S.D., Pantazis, P., 2007. Analysis of the in vitro metabolites of diferuloylmethane (curcumin) by liquid chromatography – tandem mass spectrometry on a hybrid quadrupole linear ion trap system: newly identified metabolites. Eur J Drug Metab Pharmacokinet 32:51-57.

15. Oetari, S., Sudibyo, M., Commandeur, J.N.M., Samhoedi, R., Vermeulen, N.P., 1996. Effects of curcumin on cytochrome P450 and glutathione S-transferase activities in rat liver. Biochem Pharmacol 51:39-45.

16. van Iersel, M.L., Ploemen, J.P., Struik, I., van Amersfoort, C., Keyzer, A.E., Schefferlie, J.G., van Bladeren, P.J., 1996. Inhibition of glutathione S-transferase activity in human melanoma cells by alpha, beta-unsaturated carbonyl derivatives. Effects of acrolein, cinnamaldehyde, citral, crotonaldehyde, curcumin, ethacrynic acid and trans-2-hexenal. Chem Biol Interact 102:117-132.

17. Appiah-Opong, R., Commandeur, J.N.M., van Vugt-Lussenburg, B., Vermeulen, N.P.E., 2007. Inhibition of human recombinant cytochrome P450s by curcumin and curcumin decomposition products. Toxicology 235:83-91.

18. Hayeshi, R., Mutingwende, I., Mavengere, W., Masiyanise, V., Mukanganyama, S., 2007. Inhibition of human glutathione S-transferases activity by plant polyphenolic compounds ellagic acid and curcumin. Food Chem Toxicol 45:286-295.

19. Eaton, D.L., Bammler, T.K., 1999. Concise review of the glutathione S-transferases and their significance to toxicology. Toxicol Sci 49:156-164.

20. Awasthi, Y.C., Sharma, R., Singhal, S.S., 1994. Human glutathione S-transferases. Int J Biochem 26:295-308.

21. Ioannides, C., 2002. Pharmacokinetic interactions between herbal remedies and medicinal drugs. Xenobiotica 32:451-478.

22. Ayisi, N.K., Nyadedzor,C., 2003. Comparative in vitro effects of AZT and extracts of Ocimum gratissimum, Ficus polita, Clausena anisata, Alchornea cordifolia, and Elaeophorbia

drupiferagainst HIV-1 and HIV-2 infections. Antiviral Res 58:25-33. 23. Gaudineau, C., Beckerman, R., Welbourn, S., Auclair, K., 2004. Inhibition of human P450

enzymes by multiple constituent of Ginkgo biloba extract. Biochem Biophys Res Commun 318:1072-1078.

24. Coruh, N., Celep, A.G.S., Ozgokce, F., 2007. Antioxidant properties of Prangos ferulacea (L.) Lindl., Chaerophllum macropodum Boiss. and Heracleum persicum Desf. from Apiaceae family used as food in Eastern Anatolia and their inhibitory effects on glutathione S-transferase. Food Chem 100:1237-1242.

25. Singh, S., Khar, A., 2006. Biological effects of curcumin and its role in cancer chemoprevention and therapy. Anti-Cancer Agents Med Chem 6:259-270.

26. Izzo, A.A., 2004. Herb-drug interactions: An overview of the clinical evidence. Fundamental Clin Pharmacol 19:1-16.

27. Brown, H.S., Ito, K., Aleksandra G., Houston, B., 2005. Prediction of in vivo drug-drug interactions from in vitro data: impact of incorporating parallel pathways of drug elimination and inhibitor absorp- tion rate constant. Br J Clin Pharmacol 60:508-518.

28. Youssef, K.M., El-Sherbeny, M.A., El-Shafie, F.S., Farag, H.A., Al-Deeb, O.A., Awadalla, S.A.A., 2004. Synthesis of curcumin analogues as potential antioxidant, cancer preventive agents. Arch Pharm Pharm Med Chem 337:42-54.

29. Hari Kumar, K.B., Kuttan, R., 2006. Inhibition of drug metabolizing enzymes (cytochrome P450) in vitro as well as in vivo by Phyllanthus amarus Schum & Thonn. Biol Pharm Bull 29:1310-1313.

Page 169: thuoc từ curcumin

159

Samenvatting en conclusies

Geneesmiddel-geneesmiddel of geneesmiddel-voedsel interacties waarbij

Cytochrome P450 enzymen (CYPs) betrokken zijn, zijn een belangrijke oorzaak

voor het terugtrekken van de markt van geneesmiddelen als gevolg van ernstige

bijwerkingen, van het verlies van geneesmiddelkandidaat-moleculen tijdens de

ontwikkeling van nieuwe geneesmiddelen, en van waarschuwingen in bijsluiters

van geneesmiddelen. Inhibitie van CYPs door dergelijke interacties kan leiden tot

een verminderde biologische beschikbaarheid van geneesmiddelen en tot een

vertraagde metabole afbraak ervan, met langere halfwaarde-tijden als gevolg.

Ook kunnen geneesmiddel-geneesmiddel en geneesmiddel-voedsel interacties

leiden tot omgekeerde effecten als gevolg van inductie van de CYP-niveaus in

het maag-darm kanaal, in de lever of in andere weefsels en organen. Analoog

kan er ook inhibitie of inductie van Glutathion-S-transferases (GSTs) door

geneesmiddelen en andere xenobiotica optreden. Dit kan leiden tot een

verminderde bescherming tegen bepaalde vormen van toxiciteit van xenobiotica

en bijwerkingen van geneesmiddelen.

Het doel van de studies, die in dit proefschrift beschreven zijn, was om

interacties op het niveau van CYPs en GSTs te bestuderen tussen

geneesmiddelen onderling en tussen geneesmiddelen en plantaardige

producten. Curcumine, een gele farmacologisch actieve stof die voor het eerst

geïsoleerd werd uit Curcuma longa, is hierbij gebruikt als een modelstof. Voor

Curcumine zijn, veelal (maar niet uitsluitend) in vitro, anti-inflammatoire, anti-

oxidant, anti-tumor, chemoprotectieve, chemopreventieve en diverse andere

farmacologische activiteiten aangetoond. Bovendien zijn in dit proefschrift de

interacties op het niveau van CYPs en GSTs bestudeerd van 40 synthetische

Curcumine-analoga en van extracten van een 7-tal Ghanese medicinale planten.

Om inzicht te krijgen in relaties tussen de chemische structuur en de CYP- en

GST-interacties zijn structuur-werkingsrelaties afgeleid met statische methodes

gebaseerd op MOE.

Page 170: thuoc từ curcumin

160

In hoofstuk 1 is een algemene inleiding gegeven van geneesmiddel-

geneesmiddel en geneesmiddel-voedsel/planten interacties op het niveau van

CYPs en GSTs, zowel voor wat betreft de mogelijke consequenties op gewenste

biologische activiteiten (de farmacologie) en ongewenste biologische activiteiten

(de toxicologie). In hoofstuk 2 wordt een overzicht gegeven van de effecten van

Curcumine op de farmacokinetiek en het metabolisme van geneesmiddelen en

andere biologisch actieve stoffen. In hoofdstuk 3 ligt de focus op de bepaling

van de inhiberende activiteit van Curcumine op 5 van de belangrijkste humane

CYPs (CYP1A2, CYP2B6, CYP2C9, CYP3A4 en CYP2D6). Er wordt

geconcludeerd dat Curcumine een dusdanig inhiberend potentieel ten opzichte

van CYP3A4 heeft, dat er na orale toediening van geneesmiddelen klinisch

relevante geneesmiddel-geneesmiddel interacties kunnen optreden in het maag-

darm kanaal. In hoofdstuk 4 worden de resultaten beschreven van een studie

naar het inhiberend vermogen bepaald ten opzichte de 5 genoemde

recombinante humane CYPs van 33 analoga van Curcumine (op basis van

oplosbaarheidscriteria geselecteerd uit 40 analoga). Bovendien werd een op

MOE-gebaseerde QSAR analyse uitgevoerd op de gevonden IC50 waarden. De

resultaten leiden tot de conclusie dat electrostatische en hydrofobe moleculaire

descriptoren bruikbaar zijn om de CYP-inhiberende activiteiten van de

Curcumine analoga aan te relateren. In hoofstuk 5 worden de resultaten

besproken van een analoge studie naar het inhiberend vermogen van dezelfde

33 Curcumine analoga ten opzichte van een 3-tal recombinante humane

Glutathione S-transferases (GSTA1-1, GSTM1-1 en GSTP1-1) en humaan en

rattelever-cytosol. Een op MOE-gebaseerde QSAR analyse maakte duidelijk dat

Van der Waals oppervlakte en lipofiliciteits factoren een rol spelen in de GST-

inhiberende werking. Tot slot worden in hoofdstuk 6 de inhiberende

eigenschappen bestudeerd van 7 extracten van geselecteerd (veel gebruikte)

Ghanese medicinale planten op de 5 recombinante humane CYPs, op de 3

recombinante humane GSTs en humaan en rattelever-cytosol. Phyllantus

amarus extracten bleken het sterkste inhiberend vermogen te hebben ten

Page 171: thuoc từ curcumin

161

opzichten van het merendeel van deze recombinante enzymen en leverenzym-

fracties.

Algemene conclusies en perspectieven

Inhibitie van CYPs en GSTs, twee van de belangrijkste humane biotransformatie-

enzymsystemen voor geneesmiddelen en andere chemische stoffen, kan klinisch

relevante effecten hebben op de farmacokinetiek en op de farmacologische en

toxicologische werking van dergelijke lichaamsvreemde stoffen. In dit proefschrift

worden de resultaten beschreven van in vitro studies naar het inhiberend

vermogen van Curcumine, een bekende en zeer veel gebruikte biologisch

actieve component uit Curcuma longa, van 4 ontledingsproducten van

Curcumine, van 34 synthetische analoga van Curcumine (op basis van

oplosbaarheids-criteria geselecteerd uit 40 analoga) en van wortel- en stam-

extracten van 7 veelvuldig gebruikte Ghanese medicinale planten. Met behulp

van op MOE gebaseerde methoden werden de gemeten IC50-waarden

geanalyseerd op mogelijk significante structuur-werkingsrelaties (QSARs).

Behalve met 5 humane CYPs en 3 humane GSTs, verkregen met recombinante

gen-expressie in Ecoli-bacterien, werden ook inhibitiestudies gedaan met

humane lever fracties. Curcumine zelf bleek vooral sterk competitief inhiberend

op CYP3A4 en wel zodanig dat deze eigenschap na orale toediening van

geneesmiddelen, klinisch relevante geneesmiddel-interacties in de dunne darm

zou kunnen veroorzaken. CYP2C9 onderging in mindere mate inhibitie als

gevolg van Curcumine. Curcumine bleek geen ‘mechanism-based’ inhibitie

eigenschappen te vertonen.

Met behulp van de 34 synthetische Curcumine-analoga, konden zowel voor de

bestudeerde CYPs als voor de GSTs enkele goede structuur-werkingsrelaties

(QSARs) worden afgeleid door gebruik te maken van op MOE-gebaseerde

statistische QSAR analyse methodes. Hoewel de significantie van de QSAR-

relaties nog vergroot zouden kunnen worden, door meer en sterker inhiberende

Curcumine-analoga te bestuderen of door 3D-QSAR analyses (met inbegrip van

structurele informatie van de CYP- en GST-isoenzymen), zijn de in proefschrift

Page 172: thuoc từ curcumin

162

beschreven QSAR-resultaten zeer nuttig bij het ontwerpen en ontwikkelen van

analoga van Curcumine met meer of minder sterk CYP- of GST-inhiberende

eigenschappen. Inhiberende eigenschappen van deze enzymsystemen kunnen

als gunstig (b.v. een betere cytostatische activiteit van alkylerende cytostatica als

gevolg van GST-inhibitie) en als ongunstig (b.v. kans op geneesmiddel-

interacties als gevolg van CYP-inhibitie in het maagdarm kanaal of in de lever)

beschouwd worden.

Voor wat betreft de Ghanese medicinale planten bleken extracten van de

Phyllanthus amarus en in mindere mate Cassia alata and Lactuca taraxicifolia,

relatief sterke inhiberende activiteiten te vertonen op enkele CYPs en op enkele

GSTs. De klinische relevantie van deze waarnemingen moet echter nog

onderzocht worden. In de literatuur zijn veel klinisch relevante geneesmiddel-

voeding en geneesmiddel-plantenextract interacties beschreven, zoals

bijvoorbeeld voor Ginkgo biloba en grapefruit sap.

Concluderend kan gesteld worden dat het in dit proefschrift beschreven in vitro

onderzoek belangrijke nieuwe inzichten heeft opgeleverd in het inhiberend

vermogen van Curcumine, Curcumine ontledingsproducten, een reeks

Curcumine-analoga en van wortel- en stam-extracten van veel gebruikte

Ghanese medicinale planten ten opzichte van 5 van de belangrijkste humane

CYPs en 3 van de belangrijkste GSTs. De resultaten duiden op een mogelijke

klinische relevantie, hoewel zekerheid daarover alleen verkregen zal kunnen

worden in echte klinische studies. Bovendien hebben de resultaten nieuwe

inzichten verschaft in verbanden tussen moleculaire parameters van dergelijke

stoffen en hun inhiberend vermogen ten opzichte van humane CYPs en GSTs.

Deze relaties zijn waardevol bij het rationaliseren van deze biologische

activiteiten en bij het ontwerpen van nieuwe moleculen met een betere balans

tussen gewenste en ongewenste activiteiten.

Page 173: thuoc từ curcumin

163

Appendices

List of publications related to this thesis Appiah-Opong, R., Commandeur, J.N.M., van Vugt-Lussenburg, B., Vermeulen, N.P.E., 2007. Inhibition of human recombinant cytochrome P450s by curcumin and curcumin decomposition products. Toxicology 235:83-91. Appiah-Opong, R., Commandeur, J.N.M., Vermeulen, N.P.E. Curcumin: Pharmacokin-etics, metabolism, and potential for drug-drug/food interactions. Proceedings of Int Symp Recent Progress in Curcumin Res Yogyakarta Indonesia, 2007. Appiah-Opong, R., de Esch, I., Commandeur, J.N.M., Andarini, M., Vermeulen, N.P.E., 2008. Inhibition of recombinant human cytochrome P450 mediated metabolism by curcumin analogues and related structure-activity relationships. Eur J Med Chem 43:1621-1631. Appiah-Opong, R., Commandeur, J.N.M., Axson C., Vermeulen, N.P.E. 2008. Interactions between cytochromes P450 and glutathione S-transferases and Ghanaian medicinal plants. Food Chem Toxicol 46:3598-3603. Appiah-Opong, R., Commandeur, J.N.M., Istyastono, E., Bogaards, J. J., Vermeulen, N.P.E. Inhibition of glutathione S-transferases activity by curcumin analogues. Xenobi-otica submitted.

List of publications not related to the work in this thesis

Kinomoto, M.*, Appiah-Opong, R.*, Brandful, J.A.M., Yokoyama, M., Nii-Trebi, N., Ugly-Kwame, E., Sato, H., Ofori-Adjei, D., Kurata, T., Barre-Sinoussi, F., Sata, T., Tokunaga, K., 2005. HIV-1 proteases from drug-naïve West African patients are differentially less susceptible to protease inhibitors. Clin Infect Dis 41:243-251. *contributed equally Ankrah, N-A., Quaye, I. K. E., Appiah-Opong, R., Dzokoto, C., Ekuban, F. A., Teye, K., 2003. Association between low blood glutathione levels and haptoglobulin phenotypes in pregnant women. Ghana Med J 37:35-38. Ankrah, N-A., Appiah-Opong, R., Dzokoto, C., 2000. Human breast milk storage and the glutathione content. J Trop Ped 46:111-113. Ankrah, N-A., Appiah-Opong, R., 1999. Toxicity of low levels of methylglyoxal: depletion of blood glutathione and adverse effect on glucose tolerance in mice. Toxicol Lett 109:61-67.

Page 174: thuoc từ curcumin

164

Ankrah, N-A., Nyarko, A. K., Ofosuhene, M., Appiah-Opong, R., Akyeampon Y. A., 1998. Lead exposure in urban and rural school children in Ghana. Afr J Health Sci 5:85-88. Ankrah, N-A., Dunyo, S. K., Nyarko, A. K., Appiah-Opong, R., Ofosuhene, M., 1998. Biliary excretion in persons with low blood glutathione levels. East Afr Med J 75:204-207. Ankrah, N-A., Sittie, A., Appiah-Opong, R., Ackom, I., 1996. Glyoxalase–I activity levels in peripheral blood of Ghanaian Africans with or without Plasmodium falciparum. Afr J Health Sci 3:41-43. Ankrah, N-A., Kamiya, Y., Appiah-Opong, R., Akyeampon, Y. A., Addae, M. M., 1996. Lead levels and related findings occuring in Ghanaian subjects occupationally exposed to lead. East Afr Med J 73:375-379. Armah, G. E., Mingle, J. A. A., Dodoo, A. K., Anyanful, A., Antwi, R., Commey, J., Nkrumah, F. K., 1994. Seasonality of rotavirus infection in Ghana. Annals Trop Pediatr 14:223-230.

Page 175: thuoc từ curcumin

165

Epilogue My sincere gratitude is extended to all who contributed in one way or the other in

the outcome of this research work. I thank God for the strength, grace and

endurance He afforded me. Nico, as my promoter you have been a great source

of encouragement to me, and I really appreciate your help, direction and support

even in the use of the computer. Jan, thank you for being my copromoter. Your

criticism and instructions were helpful and very much appreciated. Thank you

Chris O., for your kind advice and support. I am grateful to Prof. Ron de Kloet for

kindly introducing me to my promoter. Laura, your encouragement and help were

very much appreciated. Claudia and Laura, I am very grateful to you for the

efforts you made towards the printing of this thesis. I thank all past and present

staff and colleagues of Moltox section namely, Jeroen Kool, Ed, Micaela, Chris

de Graaf, Peter, Sebastian, Jelle, Jolanda, Eva, Aldo, Anton, Barbara, Jozef,

Chris Vos, Jeroen, Bernardo, Nathan, Eduardo and Vanina. I am also grateful to

all other staff and colleagues of the Department of Pharmaceutical Sciences and

Chemistry for their support. My gratitude goes to students I have worked with, on

this project including Maya, Chimed, Civianny, Enade and Robbert. Appreciation

is also extended to the Ghana government and Getfund Scholarship Schemes of

the Republic of Ghana for funding this project. Prof. D. Ofori Adjei, former

Director of Noguchi Memorial Institute for Medical Research (NMIMR), thanks for

your effort in obtaining this scholarship for study. For your support and

encouragement, Prof. A.K. Nyarko, Director of NMIMR, I thank you. Ken, Dorcas

and Lois, I appreciate you, for your support and the sacrifice of allowing me to

stay so far away from home for four long years. Mummy, I am very grateful to

you for taking care of my children while I was away studying. My father, siblings,

Legon Interdenominational Church, Ghana, Pentecost International Worship

Centre, Amsterdam, friends and loved ones, thank you all for your support. May

God bless all of you.

Page 176: thuoc từ curcumin

166

List of abbreviations

BFC 7-benzyloxy-4-trifluoromethyl-coumarin

BRes benzyloxyresorufin

BROD benzyloxyresorufin O-dealkylation

BQ benzyloxyquinoline

CDNB 1-chloro-2,4-dinitrobenzene

CYP cytochrome P450

DBF dibenzylfluorescein

EROD ethoxyresorufin O-dealkylation

GSH reduced glutathione

GST glutathione S-transferase

HPLC high performance liquid chromatography

MOE molecular operating environment

MRes methoxyresorufin

MROD methoxyresorufin O-dealkylation

NAC N-acetyl L-cysteine

PROD pentoxyresorufin O-dealkylation

(Q)SAR (quantitative) structure-activity relationship

UDPGA uridine diphosphate glucuronic acid

UGT UDP-glucuronosyltransferase