232
i AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE IMPACT OF PHARMACOLOGICAL INHIBITION AND GENETIC INFLUENCES ON NICOTINE PHARMACOLOGY By Eric Chun Kit Siu A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Graduate Department of Pharmacology University of Toronto ©Copyright by Eric Chun Kit Siu 2010

AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

i

AN INVESTIGATION OF NICOTINE METABOLISM IN MICE:

THE IMPACT OF PHARMACOLOGICAL INHIBITION AND GENETIC INFLUENCES

ON NICOTINE PHARMACOLOGY

By

Eric Chun Kit Siu

A thesis submitted in conformity with the requirements for the degree of

Doctor of Philosophy

Graduate Department of Pharmacology

University of Toronto

©Copyright by Eric Chun Kit Siu 2010

Page 2: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

ii

AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE IMPACT OF PHARMACOLOGICAL INHIBITION AND GENETIC INFLUENCES

ON NICOTINE PHARMACOLOGY Eric Chun Kit Siu

Doctor of Philosophy, 2010 Graduate Department of Pharmacology

University of Toronto

ABSTRACT

INTRODUCTION: Smoking is one of the single greatest causes of numerous

preventable diseases. We were interested in developing an animal model of nicotine

metabolism that can be used to examine the effects of potential CYP2A6 inhibitors on

nicotine metabolism and nicotine-mediated behaviours. Pharmacogenetic studies have

demonstrated that in humans, smoking behaviour is associated with rates of nicotine

metabolism by the CYP2A6 enzyme. Mouse CYP2A5 shares structural and functional

similarities to human CYP2A6 and has been implicated in nicotine self-administration

behaviours in mice, therefore the mouse represents a potential animal model for

studying nicotine metabolism. METHODS: We characterized nicotine and cotinine

metabolism in two commonly used mouse strains (DBA/2 and C57Bl/6). We also

examined the association between nicotine self-administration behaviours and nicotine

metabolism, and the impact of direct manipulation (i.e. inhibition) of nicotine metabolism

on nicotine pharmacodynamics (hot-plate and tail-flick tests) in mice. Finally, we studied

the effect of selegiline (a known cytochrome P450 mechanism-based inhibitor) on

nicotine metabolism in mice and in human CYP2A6. RESULTS: Nicotine metabolism in

mice in vitro was mediated by CYP2A5, and this enzyme was responsible for over 70%

and 90% of the metabolism of nicotine to cotinine and cotinine to 3-hydroxycotinine as

shown by immuno-inhibition studies, respectively. A polymorphism in CYP2A5 between

mouse strains, known to alter the probe substrate coumarin’s metabolism, did not affect

Page 3: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

iii

nicotine metabolism but dramatically altered cotinine metabolism. Nicotine self-

administration behaviour in mice was associated with level of hepatic CYP2A5 proteins

and rates of nicotine metabolism in male mice. In inhibition studies, the CYP2A5/6

inhibitor methoxsalen inhibited both in vitro and in vivo nicotine metabolism in mice and

substantially increased the anti-nociceptive effect of nicotine. Finally, selegiline was

found to be an inhibitor of CYP2A5 decreasing nicotine metabolism in vitro and in vivo in

mice. Moreover, we showed that selegiline is a mechanism-based inhibitor of CYP2A6

inhibiting nicotine metabolism irreversibly. CONCLUSION: The above data suggested

that the mouse model may be suitable for examining the impact of inhibition (and genetic

variation) on nicotine metabolism and nicotine-mediated behaviours and may potentially

be used to screen for novel inhibitors of nicotine metabolism.

Page 4: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

iv

ACKNOWLEDGEMENTS

As author of this thesis, I would like to thank those individuals who contributed their time and effort to the creation of this work. First, I would like to thank my supervisor, Professor Rachel Tyndale, for her guidance and support throughout my graduate training. I would also like to thank the past and present colleagues in the laboratory who shared their technical experience, particularly those who contributed to my various projects. These include Ewa Hoffmann, Sharon Miksys, Helma Nolte, Raj Sharma, Zhao Bin, and Linda Liu. I would also like to thank Howard Kaplan for his statistical assistance. I would like to acknowledge the support of my fellow graduate students in the laboratory, particularly Jill Mwenifumbo and Nael Al Koudsi. I am grateful to Dr. Denis Grant, Dr. Paul Fletcher, and Dr. Jack Uetrecht for sharing their knowledge and insight for the successful completion of my projects. I would also like to thank Dr. Dieter Wildenauer and Dr. Imad Damaj for their collaborations on my various projects. I am mostly graciously indebted to OGS, CIHR TUSP, CIHR STPTR, and CTCRI for the financial support that has insured my continuing success in this program. I am also very grateful to the support of Mr. and Mrs. Binh Chow through the years of my stay in Toronto. Finally, I am forever indebted to my loving parents Simon and Maggie Siu and my fiancé Teresa Tsui, which without their inspiration and support the completion of this thesis would not have been possible.

Page 5: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

v

TITLE PAGE Page i

ABSTRACT Page ii-iii

ACKNOWLEDGEMENTS Page iv

TABLE OF CONTENTS Page v-ix

LIST OF PUBLICATIONS Page x-xi

SUMMARY OF ABBREVIATIONS Page xii

LIST OF TABLES Page xiii

LIST OF FIGURES Page xiv-xv

STATEMENT OF RESEARCH PROBLEM Page 1

CHAPTER 1 INTRODUCTION Page 3

Section 1 Nicotine Metabolism Page 3

Section 1.1 Nicotine Metabolism in Mice Page 3

Section 1.2 Nicotine Pharmacokinetics in Mice Page 3

Section 1.3 Mouse Nicotine Metabolizing Enzymes Page 4

Section 1.3.1 Mouse Cytochrome P450 2A5 (CYP2A5; P450Coh; Page 4 Coumarin Hydroxylase)

Section 1.3.2 Mouse Flavin-containing Monoxidase 1 (FMO1) Page 6

Section 1.3.3 Mouse Aldehyde Oxidase 1 (AOX1) Page 7

Section 1.4 Nicotine Metabolism in Humans Page 8

Section 1.5 Nicotine Pharmacokinetics in Humans Page 8

Section 1.6 Human Nicotine Metabolizing Enzymes Page 11

Section 1.6.1 Human Cytochrome P450 2A6 (CYP2A6; Page 11 Coumarin 7-hydroxylase)

Section 1.6.2 Human Cytochrome P450 2A13 (CYP2A13) Page 14

Section 1.6.3 Human Cytochrome P450 2B6 (CYP2B6) Page 15

Section 1.6.4 Human Aldehyde Oxidase 1 (AOX1) Page 17

Section 1.6.5 Human Flavin-containing Monooxygenase 3 (FMO3) Page 18

Section 1.6.6 Human Uridine Diphosphate (UDP)-Glucuronosyl- Page 19 Transferase (UGT)

Section 2 Pharmacological Treatments for Smoking Cessation Page 21

Section 2.1 Primary Treatments Page 21

Page 6: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

vi

Section 2.1.1 Nicotine Replacement Therapies (NRTs) Page 21

Section 2.1.2 Bupropion Page 22

Section 2.1.3 Varenicline Page 23

Section 2.2 Secondary and Investigative (Non-approved) Page 25 Smoking Cessation Therapies

Section 2.2.1 Nortriptyline Page 25

Section 2.2.2 Clonidine Page 25

Section 2.2.3 Serotonergic Agents Page 26

Section 2.2.3.1 Buspirone Page 26

Section 2.2.3.2 Selective Serotonin Reuptake Inhibitors (SSRIs) Page 27

Section 2.2.4 Naltrexone Page 27

Section 2.2.5 Rimonabant (SR141716) Page 28

Section 2.2.6 Nicotine Vaccines Page 30

Section 2.2.7 Monoamine Oxidase Inhibitors (MAOIs) Page 30

Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32

Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32

Section 2.2.7.1.2 Selegiline Metabolism Page 33

Section 2.3 The Inhibition of Nicotine Metabolism and its Page 34 Application to Smoking Cessation

Section 2.3.1 Inhibitors of Human CYP2A6 Page 34

Section 2.3.2 Inhibitors of Mouse CYP2A5 Page 35

Section 2.3.3 Mechanism-Based Inhibition Page 36

Section 2.3.3.1 Methoxsalen (8-methoxypsoralen; 8-MOP): Page 37 A CYP2A5 and CYP2A6 MBI

Section 2.3.3.2 Selegiline as a MBI Page 37

Section 2.3.4 Improving Smoking Cessation through CYP2A6 Page 38 Inhibition

Section 2.3.5 Reduction of Tobacco-Specific Nitrosamine Page 40 Activation through Human CYP2A6 Inhibition

Section 3 Nicotine Pharmacological Behavioural Models Page 42

Section 3.1 Nicotine Self-Administration Page 42

Section 3.2 Analgesic Effect of Nicotine and Application in Page 43 Nicotine Studies

Section 4 Outline of Chapters Page 45

Page 7: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

vii

Section 4.1 Chapter 2: Characterization and Comparison of Page 45 Nicotine and Cotinine Metabolism In Vitro and In Vivo in DBA/2 and C57BL/6 Mice

Section 4.2 Chapter 3: Nicotine Self-Administration in Mice is Page 46 Associated with Rates of Nicotine Inactivation by CYP2A5

Section 4.3 Chapter 4: Inhibition of Nicotine Metabolism by Page 47 Methoxysalen: Pharmacokinetic and Pharmacological Studies in Mice

Section 4.4 Chapter 5: Selegiline (L-Deprenyl) is a Mechanism- Page 48 based Inactivator of CYP2A6 Inhibiting Nicotine Metabolism in Humans and Mice

CHAPTER 2 CHARACTERIZATION AND COMPARISON OF Page 50 NICOTINE AND COTININE METABOLISM IN VITRO AND IN VIVO IN DBA/2 AND C57BL/6 MICE

Abstract Page 51

Introduction Page 52

Materials and Methods Page 54

Results Page 59

Discussion Page 73

Significance of Chapter Page 79

CHAPTER 3 NICOTINE SELF-ADMINISTRATION IN MICE IS Page 80 ASSOCIATED WITH RATES OF NICOTINE INACTIVATION BY CYP2A5

Abstract Page 81

Introduction Page 82

Materials and Methods Page 84

Results Page 88

Discussion Page 99

Significance of Chapter Page 104

CHAPTER 4 INHIBITION OF NICOTINE METABOLISM BY Page 105 BY METHOXSALEN: PHARMACOKINETIC AND PHARMACOLOGICAL STUDIES IN MICE

Page 8: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

viii

Abstract Page 106

Introduction Page 107

Materials and Methods Page 109

Results Page 113

Discussion Page 128

Significance of Chapter Page 134

CHAPTER 5 SELEGILINE (L-DEPRENYL) IS A MECHANISM- Page 135 BASED INACTIVATOR OF CYP2A6 INHIBITING NICOTINE METABOLISM IN HUMANS AND MICE

Abstract Page 136

Introduction Page 137

Materials and Methods Page 141

Results Page 147

Discussion Page 157

Significance of Chapter Page 163

CHAPTER 6 GENERAL DISCUSSION Page 164

Section 5 Metabolism Topics Page 164

Section 5.1 In Vitro Metabolism Page 164

Section 5.1.1 In Vitro Nicotine Metabolism Page 164

Section 5.1.2 In Vitro Cotinine Metabolism Page 167

Section 5.1.3 Effect of a CYP2A5 Polymorphism on Substrate Page 167 Metabolism

Section 5.2 In Vivo Metabolism Page 168

Section 5.2.1 Biomarker of Nicotine Metabolism Page 170

Section 5.3 Mice (CYP2A5) as a Model to Study CYP2A6 Page 171 Inhibition

Section 5.3.1 Methoxsalen Page 172

Section 5.3.2 Selegiline Page 173

Section 6 Factors Affecting Nicotine Intake Page 176

Section 6.1 Impact of Genetic Variations Page 176

Page 9: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

ix

Section 6.1.1 Variations in Nicotine Metabolism Page 176

Section 6.1.2 Non-Nicotine Metabolism Variations Page 177

Section 6.2 Non-Genetic Factors Page 179

Section 7 Sex-Dependent Association between Nicotine Page 180 Metabolism and NSA

Section 8 Effect of Chronic Nicotine Treatment on Nicotine Page 182 Metabolism

Section 9 Future Directions Page 183

Section 9.1 Selegiline Page 184

Section 9.2 Mouse Model of Nicotine Metabolism Page 184

Section 10 Conclusion Page 187

REFERENCES Page 189

Page 10: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

x

LIST OF PUBLICATIONS Research Articles Grieder, T.E., Sellings, L.H., Vargas-Perez, H., Ting-A-Kee, R., Siu, E.C., Tyndale, R.F., and van der Kooy, D., 2009. Dopaminergic signaling mediates the motivational response underlying the opponent process to chronic but not acute nicotine. Neuropsychopharmacology In press. Shram, M.J., Siu, E., Li, Z., Tyndale, R.F., and Le, A.D., 2008. Adolescent and adult rats respond differently to the aversive effects of mecamylamine-precipitated nicotine withdrawal. Psychopharmacology (Berl) 198(2):181-90. Siu, E.C. and Tyndale R.F., 2008. Selegiline (L-deprenyl) is a mechanism-based inactivator of CYP2A6 inhibiting nicotine metabolism in humans and mice. J Pharmacol Exp Ther. 324(3):992-9. Siu, E.C. and Tyndale, R.F., 2007. Characterization and comparison of nicotine and cotinine metabolism in vitro and in vivo in DBA/2 and C57Bl/6 mice. Mol Pharm. 71(3):826-34. Damaj, M.I., Siu, E.C., Sellers, E.M., Tyndale, R.F. and Martin, B.R., 2007. Inhibition of nicotine metabolism by methoxysalen: pharmacokinetic and pharmacological studies in mice. J Pharmacol Exp Ther. 320(1):250-7. Siu, E.C., Wildenauer, D.B., and Tyndale, R.F., 2006. Nicotine self-administration in mice is associated with rates of nicotine inactivation by CYP2A5. Psychopharmacology (Berl). 184(3-4):401-408. Siu, E., Carreno, B.M., and Madrenas, J., 2003. TCR subunit specificity of CTLA-4-mediated signaling. J Leukoc Biol. 74(6):1102-1107. Darlington, P.J., Baroja, M.L., Chau, T.A., Siu, E., Ling, V., Carreno, B.M., and Madrenas, J., 2002. Surface cytotoxic T lymphocyte-associated antigen 4 partitions within lipid rafts and relocates to the immunological synapse under conditions of inhibition of T cell activation. J Exp Med. 195(10):1337-1347. Review Article Siu, E.C. and Tyndale, R.F.,2007. Non-nicotinic therapies for smoking cessation. Annu Rev Pharmacol Toxicol. 47:541-64. Conference Presentations Siu, E.C., Wildenauer, D.B., and Tyndale, R.F., Effect of genetic variation on the nicotine metabolizing enzyme (CYP2A5) on nicotine self-administration behaviours in mice. Society for Research on Nicotine and Tobacco's 12th Annual Meeting. Orlando, FL, 2006.

Page 11: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

xi

On-line Publications Siu, E.C., Al Koudsi, N., Ho, M.K., and Tyndale, R.F., 2006. Smoking, quitting and genetics. The Doctor Will See You Now (http://www.thedoctorwillseeyounow.com/articles/behavior/smoking_14/). Siu, E.C., Al Koudsi, N., Ho, M.K., and Tyndale, R.F., 2006. New frontiers in the treatment and management of smoking cessation. Cyberounds (http://www.cyberounds.com/conf/psychiatryneuroscience/2006-08-11/index.html). Abstracts 1. Siu, E.C. and Tyndale, R.F., Selegiline (l-deprenyl) inhibits nicotine metabolism in

mice and humans and is a mechanism-based inhibitor of CYP2A6. Society for Research on Nicotine and Tobacco's 14th Annual Meeting and the The 4th Annual Invitational Symposium for Research to Inform Tobacco Control. Portland, OR, 2008.

2. Siu, E.C. and Tyndale, R.F., Contrasting in vivo and in vitro metabolism of nicotine

and cotinine in DBA/2 and C57Bl/6 mice. Visions in Pharmacology Symposium. Toronto, ON, 2007.

3. Siu, E.C. and Tyndale, R.F., Contrasting in vivo and in vitro metabolism of nicotine and cotinine in DBA/2 and C57Bl/6 mice. Society for Research on Nicotine and Tobacco's 13th Annual Meeting. Austin, TX, 2007.

4. Siu, E.C. and Tyndale, R.F., In vivo and in vitro metabolism of nicotine and cotinine in DBA/2 and C57Bl/6 mice. The 3rd Annual Invitational Symposium for Research to Inform Tobacco Control. Toronto, ON, 2006.

5. Siu, E.C., Boutros, A., Wildenauer, D.B., and Tyndale, R.F., Genetic variation in Cyp2a5-mediated nicotine metabolism correlates with differential oral nicotine consumption in male mice. Visions in Pharmacology Symposium. Toronto, ON, 2006.

6. Siu, E.C., Wildenauer, D.B., and Tyndale, R.F., Effect of genetic variation on the nicotine metabolizing enzyme (CYP2A5) on nicotine self-administration behaviours in mice. The 2nd Annual Invitational Symposium for Research to Inform Tobacco Control. Toronto, ON, 2005.

7. Siu, E.C., Boutros, A., Miksys, S., Wildenauer, D.B., and Tyndale, R.F., Effect of genetic variation and chronic nicotine exposure on nicotine (CYP2A5) and ethanol (CYP2E1) metabolizing enzymes in mice. Society for Research on Nicotine and Tobacco's 11th Annual Meeting and the 7th Annual SRNT European Conference. Prague, Czech Republic, 2005.

8. Siu, E.C., Miksys, S., Klein, L.C., Blendy, J.A., and Tyndale, R.F., Effects of in vivo nicotine treatment on mouse hepatic CYP2A5 and CYP2E1 level. Visions in Pharmacology Research Symposium. Toronto, ON, 2004.

9. Siu, E.C., Boutros, A., Miksys, S., Wildenauer, D.B., and Tyndale, R.F., An Investigation into the effect of nicotine exposure and genetic variation on ethanol (CYP2E1) and nicotine (CYP2A5) inactivating enzymes in mice. The 1st Annual Invitational Meeting of Tobacco Control Investigators and Trainees. Toronto, ON, 2004.

10. Madrenas, J., Baroja, M.L., Carreno, B.M., Chau, L.A., Chau, T.A., Darlington, P.J., Ling, V., Siu, E.C., Regulation of CTLA-4-mediated signalling. Keystone Symposium on Lymphocyte Activation and Signaling. Steamboat Spring, CO, 2004.

Page 12: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

xii

SUMMARY OF ABBREVIATIONS

%MPE maximum possible effect

β-PEA phenylethylamine 3-HC 3’-hydroxycotinine 8-MOP 8-methoxypsoralen Ala alanine AOX aldehyde oxidase AUC area under the concentration CI confidence interval CL clearance Cmax maximum concentration COT cotinine COU coumarin CYP cytochrome P450 DES desmethylselegiline ED

50 effective dose 50%

F bioavailability FAD flavin adenine dinucleotide FMO flavin-containing monoxidase HLM human liver microsome HNC high nicotine consumption

HNF-4α hepatic nuclear factor-4 alpha IC50 inhibition concentration 50% L-AMP L-amphetamine LC liquid chromatography L-MAMP L-methamphetamine LNC low nicotine consumption MAOI monoamine oxidase inhibitor MBI mechanism-based inhibitor MLM mouse liver microsome NADPH nicotinamide adenine dinculeotide phosphate NCO nicotine C-oxidation NIC nicotine NNAL 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanol NNK 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone NRT nicotine replacement therapy NSA nicotine self-administration OR odds ratio QTL quantitative trait loci SEL selegiline SNP single nucleotide polymorphism SSRI selective serotonin reuptake inhibitor T1/2 half-life Tmax time at maximum concentration UGT uridine diphosphate-glucuronosyl-transferase UTR untranslated region Val valine

Page 13: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

xiii

LIST OF TABLES Table 1. In vitro nicotine metabolism (C-oxidation) parameters, page 61 Table 2. Pharmacokinetic parameters of plasma nicotine and cotinine in mice

treated with nicotine (1 mg/kg, s.c.), page 63 Table 3. Pharmacokinetic parameters of plasma cotinine and 3’-hydroxycotinine in

mice treated with cotinine (1 mg/kg, s.c.), page 68 Table 4. In vitro cotinine hydroxylation parameters, page 70

Table 5. Average nicotine consumption (µg/day; mean ± SD) in F2 male and female mice in a two-bottle choice paradigm, page 90

Table 6. Estimated in vitro NCO kinetic parameters of male LNC and HNC mice

(mean ± SD), page 95 Table 7. Concentrations (mean ± SD) of nicotine, cotinine, and 3-hydroxycotinine

(ng/mL) from the in vitro NCO activity assays (at 30 µM nicotine), page 96 Table 8. Kinetic parameters of in vitro nicotine C-oxidation activities of adult male

ICR mouse liver microsomes, page 115 Table 9. Pharmacokinetic parameters of nicotine in adult male ICR mouse after

pretreatment with methoxsalen, page 119 Table 10. Blockade of methoxsalen’s effects on nicotine acute pharmacological

responses after s.c. administration in mice, page 124 Table 11. Pharmacokinetic parameters of plasma nicotine in mice treated

intraperitoneally with nicotine (1 mg/kg) or nicotine plus selegiline (1 mg/kg), page 152

Page 14: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

xiv

LIST OF FIGURES

Figure 1A. Metabolism and in vivo disposition of nicotine and metabolites in humans, page 9

Figure 1B. Comparison of nicotine metabolism in humans and mice by known

enzymes, page 10 Figure 2. In vitro metabolism of nicotine to cotinine in DBA/2 and C57Bl/6 mice,

page 60 Figure 3. Plasma nicotine and cotinine concentrations following nicotine treatment,

page 62 Figure 4. Characterization of 3’-hydroxycotinine in mice by HPLC and LC/MS/MS,

page 65 Figure 5. Plasma cotinine and 3’-hydroxycotinine concentrations following cotinine

injections, page 67 Figure 6. In vitro metabolism of cotinine to 3’-hydroxycotinine in DBA/2 and C57Bl/6

mice, page 69 Figure 7. Antibody inhibits in vitro nicotine and cotinine metabolism in DBA/2 and

C57Bl/6 mice, page 72 Figure 8. Selectivity and linearity of CYP2A5 immunoblotting for mouse liver

microsomes, page 89 Figure 9. CYP2A5 levels are higher in male high nicotine consuming mice, page 91 Figure 10. Optimization of nicotine-C-oxidation assay, page 92 Figure 11. Nicotine-C-oxidation activity in male high nicotine consuming mice is

higher compared to the low nicotine consuming mice, page 94 Figure 12. CYP2A4/5 levels and in vitro NCO activities do not differ between female

LNC and HNC mice, page 97 Figure 13. In vitro nicotine C-oxidation activity of ICR mice, page 114 Figure 14. A representative Dixon plot of inhibition of cotinine formation by ICR

mouse microsomes in presence of varying concentrations of methoxsalen, page 116

Figure 15. Time course of nicotine and cotinine plasma concentration in mice

pretreated with methoxsalen, page 118

Page 15: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

xv

Figure 16. Time-course of methoxsalen effects on nicotine-induced (A) antinociception and (B) hypothermia after s.c. administration of 10 mg/kg in mice, page 120

Figure 17. Dose-response enhancement of methoxsalen after s.c. administration in

mice, page 121 Figure 18. Blockade of methoxsalen’s effects on nicotine-induced antinociception in

the tail-flick test after s.c. administration in mice, page 123 Figure 19. Effects of methoxsalen on the time-course of nicotine’s effects in (A) the

tail-flick test, (B) the hot-plate assay and (C) body temperature in mice, page 125

Figure 20. Relationship between nicotine-induced antinociception in the tail-flick test

and plasma levels of nicotine in the ICR mice after methoxsalen treatment, page 127

Figure 21. Selegiline and its metabolites, page 139 Figure 22. Selegiline and its metabolites inhibited nicotine metabolism, page 148 Figure 23. Pre-incubation of MLMs, but not CYP2A5, with selegiline increased

inhibition of nicotine metabolism, page 150 Figure 24. Selegiline decreased nicotine metabolism in vivo in DBA/2 mice, page

151 Figure 25. Pre-incubation of human hepatic microsomes and CYP2A6 with selegiline

increased inhibition of nicotine metabolism, page 154 Figure 26. Selegiline inhibited CYP2A6-mediated nicotine metabolism in a time-,

concentration-, and NADPH-dependent manner, page 155

Page 16: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

1

STATEMENT OF RESEARCH PROBLEM

According to the World Health Organization, over one billion men and 250 million women

smoke daily (WHO, 2007). Smoking is the leading preventable cause of morbidity and

mortality worldwide and accounts for almost five million deaths annually (Ezzati and

Lopez, 2004). The primary causes of mortality from smoking are cardiovascular disease,

lung cancer, and chronic obstructive pulmonary disease (Ezzati and Lopez, 2004).

Currently there are several approved therapies available for smoking cessation that

include nicotine replacement products (patch, gum, nasal spray, lozenges), bupropion,

and the recently marketed varenicline (Cahill et al., 2007; Siu and Tyndale, 2007b; Stead

et al., 2008). We have been interested in understanding the effects of compounds that

inhibit nicotine metabolism, and ultimately their therapeutic roles in smoking cessation.

The human cytochrome P450 2A6 (CYP2A6) is responsible for the primary metabolism

of nicotine and it is genetically and functionally variable (Mwenifumbo and Tyndale, 2007;

Nakajima et al., 1996b). Pharmacogenetic studies have demonstrated that among

smokers, slow metabolizers of nicotine (reduced CYP2A6 function) smoke fewer

cigarettes and have higher rates of quitting compared to normal nicotine metabolizers

(Malaiyandi et al., 2006; Patterson et al., 2008; Schoedel et al., 2004). More importantly,

preliminary studies have shown that inhibitors of CYP2A6 are effective in decreasing

nicotine metabolism as well as reducing smoking behaviours (Sellers et al., 2002;

Sellers et al., 2000; Tyndale et al., 1999). Considering the potential of CYP2A6 inhibitors

in smoking cessation, development and characterization of an animal model of nicotine

pharmacokinetics and pharmacodynamics will be important in studying these novel

inhibitors and their specificities, efficacies, toxicity profiles, and effects on nicotine

pharmacokinetics and pharmacodynamics. Furthermore, an animal model of nicotine

metabolism may be useful to examine the impact of genetic variation (e.g. the effects of

Page 17: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

2

increased or decreased nicotine metabolism) on nicotine pharmacology and or nicotine-

induced behaviours.

In this thesis we investigated the mouse as an experimental model on the basis

that the mouse CYP2A5 is highly identical in amino acid sequence to the human

CYP2A6 and a preliminary quantitative trait loci analysis has suggested the possible

involvement of CYP2A5 in oral nicotine self-administration behaviours in mice

(Wildenauer et al., 2005). Furthermore, the rat CYP2A enzymes do not appreciately

metabolize nicotine (Nakayama et al., 1993), therefore, the mouse may represent a

useful small animal model of nicotine metabolism and pharmacodynamics.

Page 18: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

3

CHAPTER 1 INTRODUCTION

Section 1 Nicotine Metabolism

Section 1.1 Nicotine Metabolism in Mice

Early in vitro mouse studies focused on the metabolism of nicotine and the distribution of

metabolites in tissue preparations using thin-layer chromatography but did not provide

the identification of various nicotine metabolites other than cotinine (Hansson et al., 1964;

Stalhandske, 1970; Stalhandske et al., 1969). Later, it was found using mouse (DBA/Jax)

hepatocytes, nicotine is metabolized primarily to cotinine, as well as nornicotine and

nicotine N-oxide (cotinine – 26.4%, nornicotine – 10.3%, and nicotine N-oxide 8.9% of

measurable metabolites formed) (Kyerematen et al., 1990), suggesting the involvement

of cytochromes P450 (C-oxidation) and flavin monooxygenases (N-oxidation). One study

found that neither nicotine nor cotinine were N-glucuronidated in vitro (Ghosheh and

Hawes, 2002), suggesting a potential lack of involvement of UGTs in nicotine

metabolism in mice.

Section 1.2 Nicotine Pharmacokinetics in Mice

The majority of in vivo nicotine disposition kinetic studies in mice were limited in details

(Hatchell and Collins, 1980; Mansner and Mattila, 1977). Nicotine and its metabolites are

mainly found in the liver and kidney, but they are also distributed in other tissues such as

bronchi, nasal mucosa, salivary gland, stomach, spleen, pancreas, intestine, bone,

gallbladder, and adrenal medulla (Waddell and Marlowe, 1976). A more comprehensive

study examining nicotine pharmacokinetics in three strains of mice (C3H/Ibg [C3H],

DBA/2Ibg [DBA/2], C57BL/6Ibg [C57BL/6]) showed that following a single intraperitoneal

injection, plasma nicotine levels peaked at around 5 minutes followed by rapid (~6-7 min)

Page 19: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

4

and monophasic elimination (Petersen et al., 1984). This elimination half-life is

considerably faster than rats (~0.92-1.1 hr) (Lesage et al., 2007; Miller et al., 1977) and

humans (~1.6-2.8 hr) (Benowitz and Jacob, 1993; Benowitz et al., 2006b). In mice,

nicotine is present in the brain and liver, and its elimination from these organs is similar

(~6-9 min) compared to plasma (Petersen et al., 1984). Cotinine is the major nicotine

metabolite in plasma. It is formed rapidly but has a significantly slower elimination half-

life compared to nicotine. Its rate of elimination differs between strains (e.g. 39.8±2.0 min

for DBA/2, 20.1±2.3 min for C57Bl/6 and 24.8±2.6 min for C3H) (Petersen et al., 1984).

Similar to nicotine, the rate of cotinine elimination in the brain and liver approximate that

of the plasma. Nicotine N-oxide is a minor metabolite compared to cotinine; however, its

plasma elimination half-life is longer than nicotine but shorter than cotinine (Petersen et

al., 1984).

Section 1.3 Mouse Nicotine Metabolizing Enzymes

Section 1.3.1 Mouse Cytochrome P450 2A5 (CYP2A5; P450Coh; Coumarin

Hydroxylase)

CYP2A5 was originally identified as coumarin hydroxylase due to its exclusive capability

to hydrolyze coumarin to 7-hydroxycoumarin in mice (Lang et al., 1989). Recently it was

reported that the initial step in nicotine to cotinine conversion in mice is the 5’-oxidation

of the pyrrolidine ring of nicotine to the ∆5’(1’)-iminium ion which is mediated by CYP2A5;

this is followed by its conversion to cotinine by cytosolic aldehyde oxidase (Gorrod and

Hibberd, 1982; Murphy et al., 2005b). The affinity for nicotine, as measured by the total

formation of ∆5’(1’)-iminium ion and cotinine, by cDNA-expressed CYP2A5 is ~8 µM

(Murphy et al., 2005b). In humans, cotinine 3’-hydroxylation is thought to be exclusively

Page 20: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

5

mediated by CYP2A6 (Nakajima et al., 1996a; Yamanaka et al., 2004); however, the

enzyme(s) responsible for this pathway in mice is unknown.

The mouse Cyp2a5 transcript is found in several tissues including liver, kidney,

lung, olfactory mucosa, brain, small intestines, but not in the heart or spleen (Su and

Ding, 2004). The tissue-specificity of the protein is not well characterized but coumarin

hydroxylase activity was found to be the highest in the liver followed by kidney and lung

(Kaipainen and Lang, 1985). In terms of sex differences, females have higher coumarin

7-hydroxyase activities (specific to CYP2A5) in some strains examined (e.g. DBA/2, 129,

CBA/Ca) but not in others (e.g. C3H/He, BALB/c, A/J) (van Iersel et al., 1994).

The 494-amino acid protein encoded by the Cyp2a5 gene is located on

chromosome 7 (6.5 cM; 9 exons) (Mouse Genome Informatics 4.01) and shares 86%

amino acid sequence identity with human CYP2A6. Mouse coumarin hydroxylase

activity is genetically variable between mouse strains (Wood and Taylor, 1979). Thus far

there are 38 identified single nucleotide polymorphisms (SNPs) that exist in Cyp2a5.

Two of the identified SNPs are in the coding region and result in amino acid changes

(non-synonymous): Valine117Alanine (V117A) and Threonine486Asparagine (T486N).

Twenty-nine of the SNPs are located in the introns; one is 1853-bp downstream of the

coding sequence, three are in the 3’-UTR of the mRNA, two are located in coding

sequence but are synonymous changes and one is located in the 5’-flanking region. The

full updated list of SNPs is found at Mouse Genome Informatics (MGI 4.01).

Polymorphisms in Cyp2a5 can have functional consequences; for example, the DBA/2J

strain expresses the high coumarin hydroxylase activity variant (valine at position 117)

whereas the C57BL/6J strain expresses the low activity variant (alanine at position 117).

The reported Km for coumarin 7-hydroxylase activities of the fractional purified CYP2A5

from the high (DBA/2) and low (C57Bl/6) activity variants are 0.35 µM and 4.0 µM,

respectively (Kaipainen et al., 1984).

Page 21: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

6

Section 1.3.2 Mouse Flavin-containing Monoxidase 1 (FMO1)

In general, FMOs are endoplasmic reticulum enzymes responsible for the oxidation of

nucleophilic oxygen, nitrogen, sulphur, phosphorus, or selenium atoms from a wide

range of substrates such as amines, amides, thiols, and sulfides (Eswaramoorthy et al.,

2006). In mice, the FMO1 enzyme appears to be responsible for the conversion of

nicotine to nicotine N-oxide (Itoh et al., 1997), although the contributions of

polymorphisms and other FMO isoforms to this process have not been determined. In

vitro, the metabolism of nicotine in mouse hepatocytes (DBA/Jax) showed that nicotine

N-oxide constitutes ~26% of all metabolites formed from nicotine (Kyerematen et al.,

1990). It is unclear if nicotine N-oxide mediates any pharmacological actions in mice.

In mice (129/Sv), the Fmo1 mRNA is found in similar levels in the liver, lung, and

kidney, while the brain has over 10-fold lower levels of the transcript (Janmohamed et al.,

2004). The relative levels of this protein in various tissues have not been examined. In

terms of sex, the mRNA is found at 10-20% higher in male liver, lung, and brain

compared to females. In contrast, the Fmo1 mRNA is two times higher in the kidneys of

female compared to male mice (Janmohamed et al., 2004).

The mouse Fmo1 gene is located on chromosome 1 (92.6 cM) and encodes a

532-amino acid protein (Mouse Genome Informatics 4.01). Bioinformatic searches

indicated there are 152 polymorphisms in this gene: eight upstream, nine downstream,

two synonymous, 25 in the 3’-UTR of mRNA, and the remaining SNPs are found in the

intron (MGI 4.01). Between DBA/2J and C57BL/6J mice, there are 50 SNPs: four

upstream, two non-synonymous, nine 3’-UTR, four downstream, and the remaining

SNPs are found in introns (MGI 4.01). Despite the large number of polymorphisms

present in Fmo1, there is little difference in the mRNA transcript levels and catalytic

Page 22: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

7

activities between mouse strains (Falls et al., 1995; Siddens et al., 2007), suggesting a

small impact of these SNPs on enzyme function.

Section 1.3.3 Mouse Aldehyde Oxidase 1 (AOX1)

Aldehyde oxidase is a member of the molybdo-flavoenzymes that require both flavin

adenine dinucleotide and molybdenum cofactor for its catalytic activity (Garattini et al.,

2003). The substrate specificity of mouse AOX1 is not well studied (Vila et al., 2004).

Aldehyde oxidase is responsible for the conversion of the nicotine-∆5’(1’)-iminium ion to

cotinine (Brandange and Lindblom, 1979). In mouse hepatic microsomes (plus cytosol)

nicotine-∆5’(1’)-iminium ion is metabolized to cotinine, although the enzyme responsible

for this reaction was not identified (Gorrod and Hibberd, 1982). However, it is likely that

mouse AOX1 is involved in this process due to its amino acid sequence similarity with

the human counterpart (83% identity, NCBI Blast). It is thought that, at least in humans,

the conversion of the nicotine-∆5’(1’)-iminium ion to cotinine is not rate limiting in this two-

step process of cotinine formation since the Km of aldehyde oxidase for the iminium ion

(Km = 0.85 µM, Vmax = 5300 pmol/min/mg; step 2) is much lower than the Km of nicotine

C-oxidation (step 1) (Fig. 1) (Murphy et al., 2005b; Obach, 2004). In addition, in our in

vitro studies we used sufficient amount of cytosol to ensure that aldehyde oxidase is not

rate limiting (Fig. 10d).

In mice, the expression of AOX1 is tissue- and sex-dependent. In CD-1 mice,

Aox1 mRNA is found in the liver, heart, lung, and esophagus; in contrast, the protein is

found in liver, lung, brain, and spinal cord, suggesting transcriptional or post-translational

control of enzyme expression (Kurosaki et al., 1999). With regards to sex, the AOX1

protein levels are higher in the liver but lower in the lung of males compared to females

(Kurosaki et al., 1999). Such sex difference may be determined by levels of sex

Page 23: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

8

hormones as hepatic AOX1 protein increases in both sexes upon treatment with

testosterone (Vila et al., 2004).

The mouse Aox1 gene is located on chromosome 1 (23.2 cM) and encodes a

300-kDa homo-dimer (Kurosaki et al., 1999). Based on mouse genome informatic

searches, there are 81 polymorphisms in the mouse Aox1 gene: 14 insertion-deletion

and/or SNPs in the upstream locus; three synonymous; one non-synonymous; 55 intron

polymorphisms; and eight in the 3’-UTR region (MGI 4.01). Between DBA/2J and

C57BL/6J mice, Aox1 has 56 SNPs: two intronic and one upstream deletion with the

remaining found in the introns. The functional consequences of these SNPs have not

been determined; however, DBA/2 mice have significantly less hepatic AOX1 protein

compared to C57Bl/6 (Vila et al., 2004), suggesting possible influence of these

polymorphisms on protein levels.

Section 1.4 Nicotine Metabolism in Humans

The in vivo metabolism and disposition of nicotine and its metabolites in humans have

been examined in detail. A summary of the nicotine metabolic pathway and major

nicotine metabolites are shown in figure 1A (a comparison of known enzymes

responsible for nicotine metabolism between human and mice is shown in figure 1B). In

addition to these metabolites, there are other minor metabolites that are present at

extremely low levels (Hecht et al., 1999).

Section 1.5 Nicotine Pharmacokinetics in Humans

Nicotine is distributed widely but is found mainly in blood, brain, lung, liver, kidney,

spleen, muscle, but are very low in adipose tissues (Urakawa et al., 1994). During

Page 24: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

9

Nic

oti

ne (

8-1

0%

)

Co

tin

ine

(10-1

5%

)

Trans-3

’-h

yd

rox

yc

oti

nin

e (

33-4

0%

)

Tra

ns-3

’-h

ydro

xyco

tin

ine O

’-glu

cu

ronid

e(7

-9%

)

Nic

otine

N’-o

xid

e (

4-7

%)

Nic

otine

N’-glu

cu

ron

ide

(3-5

%)

Cotinin

eN

’-o

xid

e (

2-5

%)

Co

tin

ine

N’-glu

curo

nid

e(1

2-1

7%

)

No

rco

tin

ine

(1-2

%)

No

rnic

otine

(0.4

-0.8

%)

Nic

otine iso

me

thon

ium

ion

(0

.4-1

.0%

)

5’-h

ydro

xyco

tin

ine

(1

.2-1

.6%

)5’-h

ydro

xyn

orc

otinin

e (

~0

.1%

)

[Nic

otine-∆

1’(

5’)-i

min

ium

ion]

4-o

xo

-4-(

3-p

yrid

yl)-

bu

tanoic

acid

(1-2

%)

4-h

ydro

xy-4

-(3

-pyrid

yl)-

bu

tano

ica

cid

(7

-9%

)

CY

P2A

6

CY

P2A

6

AO

X1

CY

P2A

6

CY

P2A

6FM

O3

UG

T2B

10

UG

T2B

10

UG

T2B

7

Fig

ure

1a. M

eta

bolis

m a

nd

in

viv

od

isp

ositio

n o

f n

icotine

and

meta

bolit

es in h

um

ans.

Appro

xim

ate

ly 8

0%

of

nic

otin

e is m

eta

boliz

ed (

5’

C-o

xid

ation)

to the n

ico

tine ∆

5’(1

’)-im

iniu

m ion b

y C

YP

2A

6, th

is is f

ollo

wed

by the c

onvers

ion

of th

eim

iniu

mio

n t

o c

otin

ine

by a

ldehyde

oxid

ase.

A s

ign

ific

ant

port

ion (

~60%

) of

cotin

ine

is f

urt

her

meta

bo

lized

to tra

ns-3

’-h

ydro

xyco

tin

ine b

y C

YP

2A

6. N

um

be

rs in the b

rackets

repre

se

nt th

e a

ppro

xim

ate

% r

ecovery

of

each c

om

po

und in

urine

fro

m a

syste

mic

dose o

f nic

otine.

Str

uctu

res o

f n

icotin

e, nic

otine

-

∆1

’(5’)-i

min

ium

ion

, cotinin

e, a

nd tra

ns-3

’-hydro

xycotinin

e a

re s

how

n in insets

. R

efe

ren

ce

s (

Be

now

itz

et a

l 1

994,

Beno

witz

and J

acob

2001

, C

hen e

t al. 2

007, H

ukkan

en

et

al 200

5, K

aiv

ors

aari

et a

l 2

007, K

yere

ma

ten

et al 19

90,

Murp

hy e

t al 19

99, Y

am

anaka e

t al

2005

a, Y

am

ana

ka e

t al 2005

b).

CY

P2A

6/

CY

P2B

6

CY

P2A

6

N

N

H

CH

3

N+

N

H

CH

3

N

N

H

CH

3

O

N

N

H

CH

3

OH

Nic

oti

ne (

8-1

0%

)

Co

tin

ine

(10-1

5%

)

Trans-3

’-h

yd

rox

yc

oti

nin

e (

33-4

0%

)

Tra

ns-3

’-h

ydro

xyco

tin

ine O

’-glu

cu

ronid

e(7

-9%

)

Nic

otine

N’-o

xid

e (

4-7

%)

Nic

otine

N’-glu

cu

ron

ide

(3-5

%)

Cotinin

eN

’-o

xid

e (

2-5

%)

Co

tin

ine

N’-glu

curo

nid

e(1

2-1

7%

)

No

rco

tin

ine

(1-2

%)

No

rnic

otine

(0.4

-0.8

%)

Nic

otine iso

me

thon

ium

ion

(0

.4-1

.0%

)

5’-h

ydro

xyco

tin

ine

(1

.2-1

.6%

)5’-h

ydro

xyn

orc

otinin

e (

~0

.1%

)

[Nic

otine-∆

1’(

5’)-i

min

ium

ion]

4-o

xo

-4-(

3-p

yrid

yl)-

bu

tanoic

acid

(1-2

%)

4-h

ydro

xy-4

-(3

-pyrid

yl)-

bu

tano

ica

cid

(7

-9%

)

CY

P2A

6

CY

P2A

6

AO

X1

CY

P2A

6

CY

P2A

6FM

O3

UG

T2B

10

UG

T2B

10

UG

T2B

7

Fig

ure

1a. M

eta

bolis

m a

nd

in

viv

od

isp

ositio

n o

f n

icotine

and

meta

bolit

es in h

um

ans.

Appro

xim

ate

ly 8

0%

of

nic

otin

e is m

eta

boliz

ed (

5’

C-o

xid

ation)

to the n

ico

tine ∆

5’(1

’)-im

iniu

m ion b

y C

YP

2A

6, th

is is f

ollo

wed

by the c

onvers

ion

of th

eim

iniu

mio

n t

o c

otin

ine

by a

ldehyde

oxid

ase.

A s

ign

ific

ant

port

ion (

~60%

) of

cotin

ine

is f

urt

her

meta

bo

lized

to tra

ns-3

’-h

ydro

xyco

tin

ine b

y C

YP

2A

6. N

um

be

rs in the b

rackets

repre

se

nt th

e a

ppro

xim

ate

% r

ecovery

of

each c

om

po

und in

urine

fro

m a

syste

mic

dose o

f nic

otine.

Str

uctu

res o

f n

icotin

e, nic

otine

-

∆1

’(5’)-i

min

ium

ion

, cotinin

e, a

nd tra

ns-3

’-hydro

xycotinin

e a

re s

how

n in insets

. R

efe

ren

ce

s (

Be

now

itz

et a

l 1

994,

Beno

witz

and J

acob

2001

, C

hen e

t al. 2

007, H

ukkan

en

et

al 200

5, K

aiv

ors

aari

et a

l 2

007, K

yere

ma

ten

et al 19

90,

Murp

hy e

t al 19

99, Y

am

anaka e

t al

2005

a, Y

am

ana

ka e

t al 2005

b).

CY

P2A

6/

CY

P2B

6

CY

P2A

6

N

N

H

CH

3

N+

N

H

CH

3

N

N

H

CH

3

O

N

N

H

CH

3

OH

Page 25: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

10

Nic

oti

ne

Co

tin

ine

Trans-3

’-h

yd

rox

yc

oti

nin

e

[Nic

otine

-∆1’(

5’)-i

min

ium

ion

]

CY

P2

A6

CY

P2

A6

AO

X1

FM

O3

Fig

ure

1b. C

om

pari

so

n o

f nic

otine m

eta

bolis

m in h

um

an

s a

nd m

ice b

y know

n e

nzym

es.

Re

fere

nces (

Benow

itz

et

al 1994,

Beno

witz

and J

acob 2

001, C

hen

et al. 2

007, G

orr

od

and H

ibbert

1982,

Hu

kkanen

et al 2005, Itoh

et

al 1

997, K

aiv

ors

aari

et al 2007, K

yere

mate

n

et al 1990, M

urp

hy e

t a

l 1

999, Y

am

anaka e

t al 2005a, Y

am

anaka e

t al 20

05b).

Nic

oti

ne

Co

tin

ine

Trans-3

’-h

yd

rox

yc

oti

nin

e

[Nic

otin

e-∆

1’(5

’)-i

min

ium

io

n]

CY

P2

A5

CY

P2

A5

AO

X1

Nic

otine

N’-o

xid

eF

MO

1

Hu

ma

ns

Mic

e

Nic

oti

ne

Co

tin

ine

Trans-3

’-h

yd

rox

yc

oti

nin

e

[Nic

otine

-∆1’(

5’)-i

min

ium

ion

]

CY

P2

A6

CY

P2

A6

AO

X1

FM

O3

Fig

ure

1b. C

om

pari

so

n o

f nic

otine m

eta

bolis

m in h

um

an

s a

nd m

ice b

y know

n e

nzym

es.

Re

fere

nces (

Benow

itz

et

al 1994,

Beno

witz

and J

acob 2

001, C

hen

et al. 2

007, G

orr

od

and H

ibbert

1982,

Hu

kkanen

et al 2005, Itoh

et

al 1

997, K

aiv

ors

aari

et al 2007, K

yere

mate

n

et al 1990, M

urp

hy e

t a

l 1

999, Y

am

anaka e

t al 2005a, Y

am

anaka e

t al 20

05b).

Nic

oti

ne

Co

tin

ine

Trans-3

’-h

yd

rox

yc

oti

nin

e

[Nic

otin

e-∆

1’(5

’)-i

min

ium

io

n]

CY

P2

A5

CY

P2

A5

AO

X1

Nic

otine

N’-o

xid

eF

MO

1

Hu

ma

ns

Mic

e

Page 26: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

11

smoking, nicotine is rapidly absorbed via the lung and reaches the brain within 20

seconds (Benowitz, 1990) and its plasma levels peak within five minutes (Armitage et al.,

1975). However, unlike mice, the elimination of nicotine follows a two compartmental

model with an initial rapid distribution (half-life 8.1±6.4 min) followed by elimination (half-

life 140±16 min) (Benowitz and Jacob, 1993; Benowitz et al., 1991b). In vivo, ~80% of

nicotine is metabolized to cotinine and 98% of a given nicotine dose can be recovered in

the urine as parent and metabolites within 24 hours (Benowitz et al., 1994).

Cotinine is the major metabolite of nicotine and it represents ~26% (free and

conjugated) of an intravenously infused dose of nicotine (Benowitz and Jacob, 1994).

Cotinine has a half-life of ~17.5 hr, although there is considerable variation between

subjects (range 8.1 to 29.3 hr) (Benowitz et al., 1983). The primary metabolite of cotinine

is trans 3’-hydroxycotinine (3-HC), and it represents ~40% (free and conjugated) of an

intravenously infused dose of nicotine (Benowitz and Jacob, 2001). The plasma half-life

of a single dose of 3-HC is ~6.6 hr (range 4.6 to 8.3 hr) (Benowitz and Jacob, 2001), but

the elimination of 3-HC is generation-limited (Dempsey et al., 2004). The primary

metabolite of 3-HC is trans 3’-hydroxycotinine glucuronide (Neurath and Pein, 1987).

Section 1.6 Human Nicotine Metabolizing Enzymes

Section 1.6.1 Human Cytochrome P450 2A6 (CYP2A6; Coumarin 7-

hydroxylase)

Similar to mouse CYP2A5, human CYP2A6 was originally named coumarin 7-

hydroxylase due to its exclusive ability to catalyze the formation of 7’-hydroxycoumarin

from coumarin (Pelkonen et al., 2000; Yamano et al., 1990). CYP2A6 is the primary

enzyme responsible for the C-oxidation of nicotine (Messina et al., 1997; Nakajima et al.,

1996b). It also contributes to the metabolism of a variety of compounds such as the

Page 27: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

12

anesthetic halothane (Spracklin et al., 1996), the platelet activator factor inhibitor SM-

12502 (Nunoya et al., 1996), the anti-convulsants losigamone (Torchin et al., 1996) and

valproic acid (Sadeque et al., 1997), the anti-neoplastic agent tegafur (Ikeda et al., 2000),

and it also activates the tobacco-specific nitrosamine 4-(methylnitrosamino)-1-(3-

pyridyl)-1-butanone (NNK), which is further described in section 2.3.5 (Tiano et al., 1993).

CYP2A6 also metabolizes several endogenous substrates such as trans-retinoic acid,

testosterone, estradiol, and progesterone althought its contribution to their metabolism is

minor (Di et al., 2009).

In humans, the two-step inactivation pathway of nicotine to cotinine is the same

as mice: nicotine is C-oxidized by CYP2A6 to form the ∆5’(1’)-iminium ion, which is

subsequently converted to cotinine by aldehyde oxidase (Gorrod and Hibberd, 1982). In

the cDNA-expressed system, CYP2A6 has an apparent Km of 11 to 58 µM for nicotine

(Ho et al., 2008; Messina et al., 1997; Nakajima et al., 1996b; Yamazaki et al., 1999). In

human liver microsomes, the Km of cotinine formation also varies greatly (range 13-131

µM) (Messina et al., 1997; Nakajima et al., 1996b). Similarly, in microsomes the Vmax of

cotinine formation from nicotine ranges from 4.2 to 120 nmol/mg protein/min (Messina et

al., 1997). These variations may be due to the genetic (e.g. CYP2A6 or transcription

factor variants) and/or environmental (e.g. dietary) influences. Chemical inhibitors and

antibodies were used to determine that the contribution of hepatic microsomal CYP2A6

to the metabolism of nicotine to cotinine to be ~90% (Messina et al., 1997; Nakajima et

al., 1996b). Consistent with this, individuals genetically lacking CYP2A6 produce little

cotinine (Kitagawa et al., 1999; Mwenifumbo and Tyndale, 2007; Yamanaka et al., 2004).

In addition to nicotine, CYP2A6 also metabolizes cotinine to 3-HC (Nakajima et

al., 1996a). In vitro, hepatic microsomes produce 3-HC from cotinine with a Km of ~234.5

µM (range 201.9 to 288 µM). This activity is highly correlated with coumarin 7-

Page 28: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

13

hydroxylase activity and is inhibited by both coumarin and a CYP2A6 inhibitory antibody

(Nakajima et al., 1996a). In vivo, individuals lacking CYP2A6 do not produce 3-HC

(Dempsey et al., 2004; Yamanaka et al., 2004). Finally, CYP2A6 can also metabolize

nicotine to nornicotine, and cotinine to 5’-hydroxycotinine and norcotinine (Fig. 1)

(Murphy et al., 1999; Yamanaka et al., 2005a).

CYP2A6 mRNA is found predominantly in the liver, but is also found in lung,

trachea, nasal mucosa, and skin (Ding and Kaminsky, 2003; Janmohamed et al., 2001).

The transcript is not found in the kidney, duodenum, peripheral lymphocytes, or uterine

endometrium (Koskela et al., 1999). CYP2A6 protein is expressed primarily in the liver

(Ding and Kaminsky, 2003; Yamano et al., 1990) and is also found in the lungs (Zhang

et al., 2007). Its protein levels in other tissues have not been described. CYP2A6 protein

is found at higher levels in females than males (Al Koudsi et al., 2006b). This finding is

consistent with pharmacokinetic studies showing females have significantly faster

nicotine to cotinine metabolism (~20%) (Benowitz et al., 2006a).

CYP2A6 is 494 amino acids in length and is found on chromosome 19 (19q13.2;

9 exons) (Ensembl Genome Browser). CYP2A6 is highly polymorphic – currently there

are 34 named alleles (*1 to *34). Variant alleles contain non-synonymous SNP including

deletion, duplication, frameshift, and gene conversion, and promoter region alleles. The

full updated list of CYP2A6 polymorphisms can be found at

http://www.cypalleles.ki.se/cyp2a6.htm. The functional aspects of the alleles vary

ranging from a significant drop in nicotine metabolic activity (*2, *4, *7, *10, *17, *19, *20,

*23, *26, and *27) (Benowitz et al., 2001; Fukami et al., 2005a; Fukami et al., 2005b;

Fukami et al., 2004; Ho et al., 2008; Mwenifumbo et al., 2008a; Xu et al., 2002), a

moderate drop in activity (*9 and *12) (Benowitz et al., 2006b; Oscarson et al., 2002;

Yoshida et al., 2003), no change in activity (*8, *14, *16, *18, *21, *24, *25, and *28) (Al

Koudsi et al., 2006a; Fukami et al., 2005b; Ho et al., 2008; Mwenifumbo et al., 2008a;

Page 29: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

14

Nakajima et al., 2006a; Xu et al., 2002), to increased activity (*1B, *1X2A, and *1X2B)

(Fukami et al., 2007; Mwenifumbo et al., 2008b; Rao et al., 2000; Wang et al., 2006a).

The functional impact of several other SNPs (*5, *6, *11, *13, *15, and *22) are predicted

to have reduced activities (Daigo et al., 2002; Haberl et al., 2005; Kitagawa et al., 2001;

Kiyotani et al., 2002; Oscarson et al., 1999) while others (*29, *30, *32, *33, and *34)

remain to be published. The impact of reduced nicotine metabolism on smoking

behaviour is briefly discussed in section 2.3.4.

Section 1.6.2 Human Cytochrome P450 2A13 (CYP2A13)

CYP2A13 is 93.5% identical in amino acid sequence with CYP2A6 (NCBI Pubmed).

Despite this homology, CYP2A13 has only approximately one-tenth the coumarin 7-

hydroxylase activity (Vmax/Km) of CYP2A6 (He et al., 2004b). CYP2A13 also metabolizes

tobacco alkaloids such as nicotine, cotinine (Bao et al., 2005), and NNK (Su et al., 2000).

CYP2A13 exhibits a similar affinity (Km) for cotinine formation from nicotine as

CYP2A6 (20 vs. 26 µM, respectively) in a heterologous expression system (Bao et al.,

2005). However, the enzyme is much more efficient in metabolizing cotinine to 3HC

compared to CYP2A6 (45 vs. 265 µM, respectively) (Bao et al., 2005). However, due to

the absence of CYP2A13 protein in the liver (Zhu et al., 2006), CYP2A6 remains the

main hepatic nicotine and cotinine metabolizing enzyme.

CYP2A13 mRNA is primarily found in the respiratory tract (Su et al., 2000). The

transcript is also found in liver, brain, mammary gland, prostate, testis, and uterus but

not in kidney, heart, small intestine, spleen, and stomach (Koskela et al., 1999; Su et al.,

2000). Using a CYP2A13-specific antibody the protein was found to be expressed in the

Page 30: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

15

lung but not the liver (Zhu et al., 2006). Thus far, sex differences in CYP2A13 expression

have not been examined.

CYP2A13 is a CYP2A6 homologue located in the same gene cluster on

chromosome 19 (19q13.2) with the same number (nine) of exons encoding the same

number (464) of amino acids. As with CYP2A6, CYP2A13 is polymorphic. Currently

there are nine named alleles (*1 to *9) which include one wild-type allele, seven alleles

that contain at least one non-synonymous SNPs, and one frame-shift allele. The full

updated list of CYP2A13 polymorphisms can be found at

http://www.cypalleles.ki.se/cyp2a13.htm. The functional consequences of the variation

range from decrease in function (*2, *5, *6, *8, and *9) (Schlicht et al., 2007; Zhang et al.,

2002) to loss-of-function (*4 and *7) (Cauffiez et al., 2004; Wang et al., 2006b). The

contribution of CYP2A13 to systemic nicotine and cotinine levels is likely to be minor

considering that the protein is not expressed in the liver (Zhu et al., 2006) and individuals

who are poor metabolizers of CYP2A6 do not produce 3’-hydroxycotinine (Yamanaka et

al., 2004). However, it may contribute to local metabolism of xenobiotics such as nicotine

and NNK, particularly in the lung where the protein is predominantly expressed (Bao et

al., 2005; Su et al., 2000). There is no identified structural/functional homologue of this

enzyme in mice.

Section 1.6.3 Human Cytochrome P450 2B6 (CYP2B6)

CYP2B6 was originally named as the phenobarbital-inducible cytochrome P450IIB;

CYP2B6 metabolizes a wide range of important compounds including nicotine

(Yamazaki et al., 1999), the smoking cessation drug/anti-depressant bupropion

(Faucette et al., 2000), the HIV treatment drug efavirenz (Ward et al., 2003), and various

Page 31: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

16

endogenous substrates such as testosterone and estrone (Imaoka et al., 1996; Shou et

al., 1997).

In an expression system, CYP2B6 has a low affinity for nicotine C-oxidation (Km

of 105±25 µM) (Yamazaki et al., 1999). Therefore, the contribution of CYP2B6 in the

metabolism of nicotine to cotinine is expected to be negligible at low concentrations, but

it may increase significantly when the nicotine concentration is high (Nakajima et al.,

1996b). The in vivo involvement of CYP2B6 in cotinine formation is debatable. For

instance, in a study of individuals deficient in CYP2A6 (CYP2A6*4), the metabolism of

nicotine was faster in those who had a CYP2B6 functional genetic variant (CYP2B6*6)

(Jinno et al., 2003), while this variant had no impact in the individuals with wild-type

CYP2A6 (Ring et al., 2007). In contrast, in another study, the same CYP2B6*6 variant

had no impact on in vivo nicotine plasma concentrations (obtained from nicotine patch)

regardless of the activity of CYP2A6 (Lee et al., 2007). Considering that a recent study

found CYP2B6*6 to be transcriptionally unstable resulting in a significant decrease in

protein production (Hofmann et al., 2008), additional studies will be required to reconcile

the observed differences. Nonetheless, a role for CYP2B6 may underlie the observation

that individuals lacking CYP2A6 still produce cotinine (~5% of normal level) following

nicotine administration, but the contribution of other enzymes has not been ruled out

(Yamanaka et al., 2004).

The contribution of CYP2B6 to nicotine metabolism may also be important under

situations of CYP2B6 enzyme induction. For example, in non-human primates

phenobarbital treatment led to greater than 6-fold increase in CYP2B6agm protein and

doubled its contribution to nicotine metabolism (Schoedel et al., 2003). CYP2B6 also

takes part in the N-demethylation of nicotine to form nornicotine although its involvement

is secondary to CYP2A6 (Yamanaka et al., 2005a).

Page 32: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

17

CYP2B6 mRNA and protein are found at low levels in the liver and are also found

in the brain, lung, and kidney, and intestine (Gervot et al., 1999; Miksys et al., 2003).

Sex difference exists for CYP2B6 as females tend to have higher levels of mRNA,

protein, and activity (S-mephenytoin N-demethylation) (Lamba et al., 2003; Stresser and

Kupfer, 1999).

CYP2B6 is located in the same gene cluster as CYP2A6 and CYP2A13 on

chromosome 19 (19q13.3; 9 exons) and the protein is 491 amino acids in length

(Ensembl Genome Browser). CYP2B6 is also polymorphic and currently there are 29

named alleles for CYP2B6 (*1 to *29). There are at least 25 alleles; the effects of many

of these variants on nicotine metabolism have not been studied. Mouse Cyp2b10 is the

most similar to human CYP2B6 in nucleic acid sequence similarity (~75%; NCBI Blast);

however, whether CYP2B10 metabolizes nicotine or its metabolites in mice has not been

studied.

Section 1.6.4 Human Aldehyde Oxidase 1 (AOX1)

AOX1 is responsible for the metabolism of a large number of nitrogen-containing

xenobiotics including the antiviral famcyclovir, the antineoplastics methotrexate and 6-

mercaptopurine, and the ethanol metabolite acetaldehyde (Garattini et al., 2003). And as

discussed in the previous section (1.3.3), AOX1 metabolizes the nicotine-∆5’(1’)-iminium

ion to cotinine (Brandange and Lindblom, 1979).

In humans, AOX1 mRNA is found in highest level in the liver followed by the

kidney and adrenal gland, and it is found in trace levels in various peripheral tissues

(Nishimura and Naito, 2006). There are no apparent sex differences in aldehyde oxidase

activity (Al-Salmy, 2001).

Page 33: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

18

AOX1 is located on chromosome 2 (2q33; 35 exons) and encodes a 1338 amino-

acid protein (Ensemble Genome Browser). The human AOX1 is highly polymorphic with

125 polymorphisms identified thus far, but only one non-synonymous SNP is predicted

to have any functional consequence (Steinberg et al., 2007). Despite the lack of the

impact of genetic variation, there is considerable variation in the activity of hepatic AOX1

activity (up to 50-fold depending on substrate tested) (Al-Salmy, 2001; Sugihara et al.,

1997). AOX1 variation is not expected to affect the formation of cotinine because up to

90% of variability in cotinine formation can be explained by CYP2A6 immunoreactivity

(Messina et al., 1997; Mwenifumbo and Tyndale, 2009)

Section 1.6.5 Human Flavin-containing Monooxygenase 3 (FMO3)

FMO3 metabolizes various compounds such as the pesticide fenthion (Furnes and

Schlenk, 2004), the anti-thyroid drug methimazole (Krueger and Williams, 2005),

triethylamine, as well as the conversion of nicotine to nicotine N-oxide (Cashman et al.,

1992). In humans, ~4-7% of administered nicotine is excreted in the urine as nicotine N-

oxide (Fig. 1) (Benowitz et al., 1994; Meger et al., 2002) and this percentage is three- to

four-fold higher in individuals lacking CYP2A6 (Yamanaka et al., 2004).

Like many other drug metabolizing enzymes the FMO3 transcript is mainly found

in the liver with minor transcript levels in both kidney and the lung (Zhang and Cashman,

2006), consistent with its protein expression (Krause et al., 2003; Overby et al., 1997).

There is marked inter-individual variation (over 9-fold) in the protein’s expression

(Krause et al., 2003; Overby et al., 1997). On the other hand, FMO3 expression and

nicotine N-oxide activity is uninfluenced by gender (Cashman et al., 1992; Krause et al.,

2003).

Page 34: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

19

The FMO3 gene is located on chromosome 1 (1q23; 8 coding exons and 1 non-

coding exon) and encodes a 532 amino-acid protein (Ensemble Genome Browser).

There are currently 40 identified FMO3 alleles with the majority being non-synonymous

SNPs. In addition, there are two deletion alleles and one truncation allele. The list of

alleles can be found at http://human-

fmo3.biochem.ucl.ac.uk/Human_FMO3/Tables/tableall.html. Thus far there have been

no studies examining the direct impact of FMO3 polymorphism on nicotine disposition

kinetics; however, in rare cases (<1%) individuals with loss-of-function FMO variants

there may be minor re-routing of nicotine metabolism.

Section 1.6.6 Human Uridine Diphosphate (UDP)-Glucuronosyltransferase

(UGT)

UGTs are phase II enzymes responsible for the transfer of glucuronic acid from UDP-

glucuronic acid to various endogenous and exogenous substrates which increases their

water-solubility and subsequent excretion (Kiang et al., 2005). In humans, nicotine N-

glucuronide represents approximately 3-5% of total urinary metabolite of a dose of

nicotine (Benowitz et al., 1994). Individuals deficient for CYP2A6 may excrete up to

~45% of nicotine as the glucuronidated product over 24 hours due to metabolic re-

routing (Yamanaka et al., 2004). Cotinine and trans 3’-hydroxycotinine are also found as

glucuronidated products in the urine (12-17% and 7-9%, respectively) (Benowitz et al.,

1994).

The enzymes responsible for the N-glucuronidation of nicotine was thought to be

UGT1A4 with some contribution by UGT1A9 (Kuehl and Murphy, 2003; Nakajima et al.,

2002). Similarly, cotinine was thought to be N-glucuronidated by UGT1A4 (Kuehl and

Murphy, 2003; Nakajima et al., 2002). However, with improved detection method, it was

Page 35: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

20

later identified that UGT2B10 to be the primary UGT enzyme that catalyzes the N-

glucuronidation of nicotine and cotinine (Kaivosaari et al., 2007). This was confirmed by

pharmacogenetic studies showing that hepatic microsomes from individuals with the

loss-of-function UGT2B10 enzyme (UGT2B10*2) showed a 5-fold reduction in nicotine

N-glucuronide production (Chen et al., 2007). This isoform also plays a predominant role

in the N-glucuronidation of cotinine (Kaivosaari et al., 2007) and UGT2B10*2 individuals

showed a 16-fold reduction in cotinine N-glucuronide production (Chen et al., 2007).

Trans 3’-hydroxycotinine is mainly O-glucuronidated and the primary enzyme involved in

this reaction is UGT2B7 (Yamanaka et al., 2005b).

Both UGT2B10 and UGT2B7 transcripts are found at high levels in the liver and

kidney (Turgeon et al., 2001). However, the variation in the expression of these two

genes between males and females has not been examined. Numerous UGT2B variants

have been identified in recent years; a current list of the variants can be found at

http://galien.pha.ulaval.ca. Both the UGT2B10 and UGT2B7 genes are located on

chromosome 4 (4q13; both are encoded by 6 exons) and are 528 and 529 amino acids

in length, respectively (Ensemble Genome Browser). Currently there are two identified

UGT2B10 and four UGT2B7 alleles with the variants coding for non-synonymous SNPs.

In mice neither nicotine and cotinine are glucuronidated (3-HC was not examined)

(Ghosheh and Hawes, 2002), suggesting a lack of involvement of UGT enzymes in

these metabolic pathways.

Page 36: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

21

Section 2 Pharmacological Treatments for Smoking Cessation

Section 2.1 Primary Treatments

Section 2.1.1 Nicotine Replacement Therapies (NRTs)

Nicotine replacement products are the first line treatments for nicotine dependence (Siu

and Tyndale, 2007b). Since nicotine is the primary addictive component of cigarettes

(Russel, 1987), NRTs promote smoking cessation by delivering nicotine to replace

nicotine from cigarettes. The potential use of nicotine to alleviate withdrawal symptoms

was recognized in the 1970s (Schneider et al., 1977). Currently there are several

nicotine delivery devices available that include gum (2 and 4 mg), transdermal patch (7,

14, and 21 mg), vapor inhaler, nasal spray, sublingual tablet (2 mg), and lozenge (2 and

4 mg).

The general feature of NRTs is that the nicotine is absorbed through the skin or

the nasal/buccal mucosal membrane in order to avoid the extensive hepatic first-pass

metabolism (~80%; unclear in mice) associated with oral ingestion (Benowitz et al.,

1991a; Benowitz et al., 1991b; Zins et al., 1997). However, in practice, a large portion of

the nicotine delivered via the gum (Benowitz et al., 1987), tablets (Molander and Lunell,

2001), and lozenges (Choi et al., 2003) is swallowed and metabolized before reaching

systemic circulation. After a single cigarette puff, nicotine reaches peak plasma

concentration in 5 to 8 min (Hukkanen et al., 2005b). On the other hand, NRTs in

general have slower nicotine delivery kinetics with slower peak times after administration

(gum – 30 min; patch – 3 to12 hr; inhaler – 30 min; nasal spray – 11 to 18 min; tablet

and lozenges – 1 hr) (Hukkanen et al., 2005b). The slower delivery and absorption

kinetics of nicotine from NRTs significantly reduces their abuse liabilities (Henningfield

and Keenan, 1993; West et al., 2000) but may also reduce their long term success as

smoking cessation treatments.

Page 37: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

22

A systemic review of 111 studies examining the efficacies of various NRT

products found that the relative risk (RR) of smoking abstinence after six months of

follow up with NRTs was 1.58 (95% CI 1.50 to 1.66) (Stead et al., 2008). Between NRTs,

nasal spray appeared to be the most effective [RR 2.02 (95% CI 1.49 to 3.73)] while the

gum appeared to be the least effective [RR 1.43 (95% CI 1.33 to 1.53)]. The RRs for

patch, inhaler, and lozenges are 1.66 (95% CI 1.53 to 1.81), 1.90 (95% CI 1.36 to 2.67),

and 2.00 (95% CI 1.63 to 2.45), respectively. It should be noted that not all studies in

each of the NRT forms showed definitive benefits over smoking cessation. There

appeared to be no overall effect of the type of NRTs on success of treatment outcome

(Stead et al., 2008). Behavioural support (e.g. advice, counseling) plus NRT

pharmacotherapy provided only minor advantage compared to pharmacotherapy alone

(Stead et al., 2008). There are conflicting evidence as to whether higher dose of nicotine

provided better outcome. For instance, 4 mg gum has been found to be provide better

cessation rates compared to 2 mg gum in some studies but not others (Garvey et al.,

2000; Herrera et al., 1995; Kornitzer et al., 1987). Nicotine patch studies provided less

evidence of dose-dependent efficacies (Killen et al., 1999; Tonnesen et al., 1999). The

primary adverse affects of nicotine replacement therapies depend on the type of

products used (e.g. skin irritation for patches), but common side effects include dizziness,

nausea, and headache (Stead et al., 2008).

Section 2.1.2 Bupropion

Smoking and depression are strongly associated perhaps because both are influenced

by dopamine levels (Paperwalla et al., 2004; Quattrocki et al., 2000). Bupropion is an

atypical antidepressant and was the first non-nicotine product approved as a first-line

therapy for smoking cessation. At therapeutic doses, bupropion binds to striatal

Page 38: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

23

dopamine transporters (~20% occupancy). This potentially prevents dopamine reuptake

(Argyelan et al., 2005; Learned-Coughlin et al., 2003) which contribute to the mechanism

for reducing withdrawal symptoms. The noradrenergic system is also implicated in the

pharmacological effects of bupropion, possibly via noradrenaline reuptake inhibition

(Foley et al., 2006). The above mechanisms may explain, in part, bupropion’s ability to

reduce cravings in abstinent smokers and alleviate certain withdrawal symptoms which

may explain its usefulness in smoking cessation (Durcan et al., 2002; Jorenby et al.,

1999; Shiffman et al., 2000; Tashkin et al., 2001). Finally, bupropion can also act as a

nicotinic receptor antagonist via a non-competitive inhibition mechanism, suggesting that

it may attenuate the rewarding effect of nicotine (Alkondon and Albuquerque, 2005;

Fryer and Lukas, 1999; Slemmer et al., 2000).

A Cochrane systemic review of 31 studies found that the odds ratio of smoking

abstinence after six months of follow up with bupropion alone was 1.94 (95% CI 1.72 to

2.19) (Hughes et al., 2007). As with NRT, not all studies showed significant benefit of

bupropion over placebo. The intensity of behavioural support did not alter the cessation

rate although it is not known if counseling plus bupropion had significant advantages

over bupropion alone (Hughes et al., 2007). The analysis also found no effect of dose on

cessation rate from two studies (150 mg vs 300 mg at 12 months after target quit date).

However, a study directly comparing different dose of bupropion (100, 150 and 300

mg/day) found that bupropion was effective in smoking cessation in a dose-dependent

manner, although statistical comparisons between dosing groups were not made (Hurt et

al., 1997). Bupropion was not advantageous over NRTs for relapse prevention (Hughes

et al., 2007). The major adverse reactions associated with bupropion are insomnia,

nausea/vomiting, and dizziness (Boshier et al., 2003).

Section 2.1.3 Varenicline

Page 39: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

24

Varenicline is a partial agonist of the α4β2 nicotinic cholinergic receptor (Coe et al.,

2005). In mouse studies, the β2 receptor subunit is required for nicotine reward, as

measured by conditioned place preference, and the α4 receptor subunit is important for

reward and development of tolerance to nicotine (McCallum et al., 2006; Tapper et al.,

2004). Varenicline is hypothesized to reduce activation of the dopaminergic system via

partial blockade of nicotinic receptors stimulated by high levels of nicotine during

smoking, thus lowering the rewarding effects of nicotine. In addition, as a weak agonist,

varenicline may also stimulate dopamine release by the mesolimbic neurons and reduce

craving as well as withdrawal during abstinence (Coe et al., 2005; Papke and

Heinemann, 1994).

Four studies that examined the 52-week continuous abstinence with varenicline

compared to placebo have all found varenicline to provide significantly better quit rate

compared to placebo (Gonzales et al., 2006; Jorenby et al., 2006; Nides et al., 2006;

Oncken et al., 2006). When analyzed together the odds of quitting were 3.22 (95% CI

2.43 to 4.27) favouring varenicline (Cahill et al., 2007). Two out of three studies that

compared the efficacy of varenicline to bupropion have found varenicline to be superior

(statistically significant) in maintaining 52-week continuous abstinence rate compared to

bupropion (Gonzales et al., 2006; Jorenby et al., 2006; Nides et al., 2006). Furthermore,

varenicline was more effective in reducing withdrawal symptoms compared to bupropion

(West et al., 2008). Varenicline was significantly more effective in preventing relapse

compared to placebo [198 (95% CI 159-260) days vs 87 (95% CI 58-143) days,

respectively; log-rank P<0.001] (Tonstad et al., 2006). The main adverse events

associated with varenicline are relatively mild and includes nausea, abnormal dreams,

and insomnia (Oncken et al., 2005). However, recent evidence indicated that varenicline

Page 40: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

25

may induce depression and suicidal ideation (Kuehn, 2008) and a public health advisory

had been placed on this drug by the FDA

(http://www.fda.gov/bbs/topics/NEWS/2008/NEW01788.html).

Section 2.2 Secondary and Investigative (Non-approved) Smoking

Cessation Therapies

Section 2.2.1 Nortriptyline

Nortriptyline is a second-line therapy for smoking cessation but has not yet been

approved for this indication by the US Federal Drug Administration. Nortriptyline is a

tricyclic antidepressant that inhibits the reuptake of serotonin and noradrenalin. A meta-

analysis comparing the efficacy of nortriptyline vs. placebo found the pooled odds ratio

to be 2.34 (95% CI 1.61 to 3.41) (Hughes et al., 2007). The smoking cessation rate

achieved with nortriptyline was comparable to those achieved with bupropion [pooled

ORs: 2.1 (1.5 – 3.1) vs. 2.0 (1.7 – 3.4) for nortriptyline and bupropion, respectively]

(Hughes et al., 2005). However, nortriptyline appears to cause greater adverse events

compared to bupropion (Hall et al., 2002; Wagena et al., 2005) and may increase

abnormal cardiac rhythms (QTc interval) and increased adverse cardiac events in

ischemic heart disease patients (Roose et al., 1998; van Noord et al., 2009), thus

preventing its wide use.

Section 2.2.2 Clonidine

The α2-adrenergic receptor agonist clonidine is primarily used for the treatment of

hypertension; however, it has also been used for treatment of opiate and alcohol

withdrawal (Gossop, 1988; Manhem et al., 1985). Clonidine can alleviate smoking

Page 41: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

26

withdrawal symptoms (Glassman et al., 1984; Ornish et al., 1988). A Cochrane analysis

of six studies found that clonidine treatment was associated with increased smoking

cessation [pooled OR: 1.89 (1.30 – 2.74)]; however, only one study showed significant

improvement in smoking cessation when the studies were analyzed separately (Gourlay

et al., 2004). Finally, adverse effects such as dry mouth, sedation, and postural

hypotension limit its general use.

Section 2.2.3 Serotonergic Agents

Section 2.2.3.1 Buspirone

The serotonergic system is responsible for the regulation of mood and appetite.

Buspirone is serotonin receptor (5-HT1a) partial agonist used for the treatment of

generalized anxiety disorder. The drug appears to function via the inhibition of the

release of serotonin and may potentially increase noradrenergic and dopaminergic

activity (Eison and Temple, 1986). Buspirone appears to benefit smokers with high level

of anxiety. For instance, in an eight-week treatment with maximum dose of 60 mg of

buspirone, the one-month abstinence in the high-anxiety group was 88% (treatment)

compared to 61% (placebo) but in the low-anxiety group the abstinence was 60%

(treatment) compared to 89% (placebo). At 12 months, there was no difference between

treatment and placebo in both anxiety group (Cinciripini et al., 1995). However, a

subsequent study revealed that buspione (six-week treatment; maximum dose of 60 mg)

was ineffective in smoking cessation regardless of degree of anxiety (Schneider et al.,

1996). In another study, buspirone (up to 40 mg for four weeks) was found to be equally

as effective as nicotine patch (fixed: 21/22 mg for six weeks; tapered: 21/22 mg for four

weeks, 14 mg for 4 weeks, and 7 mg for 4 weeks) (Hilleman et al., 1994). While the drug

Page 42: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

27

appears to be safe, with little adverse effects, the overall usefulness of buspirone in

smoking cessation is inconclusive.

Section 2.2.3.2 Selective Serotonin Reuptake Inhibitors (SSRIs)

SSRIs are used for the treatment of depression and anxiety (Nutt et al., 1999).

Fluoxetine has not been demonstrated to be effective for smoking cessation but may

decrease the negative and increase the positive affects after quitting (Blondal et al.,

1999; Cook et al., 2004; Niaura et al., 2002; Saules et al., 2004). Paroxetine was not

effective beyond four weeks (during a 9-week treatment protocol) (Killen et al., 2000)

and sertraline did not increase abstinence compared to placebo (Covey et al., 2002).

SSRIs do not appear to be effective smoking cessation aides.

Section 2.2.4 Naltrexone

Endogenous opioids (endorphins, enkephalins, dynorphins), together with their receptors

(mu, kappa, delta) are responsible for various biological processes which include

autonomic, neuroendocrine, hedonic, and emotional responses (Kieffer and Evans,

2009). There are several lines of evidence that suggested the involvement of the opioid

system in nicotine addiction. First, animals exhibit similar somatic withdrawal signs from

both nicotine and opiates (e.g. tremors, scratches, chewing, etc.) (Malin and Goyarzu,

2009). Second, in rats nicotine administration led to increased turnover of Tyr-Gly-Gly, a

marker of enkephalin release (Houdi et al., 1991), and smoking cigarettes containing

more nicotine led to greater plasma levels of β-endorphin compared to cigarettes with

low nicotine (Gilbert et al., 1992). Third, mice with the preproenkephalin or the mu-opioid

receptor genes knocked-out do not respond to the rewarding properties of nicotine as

Page 43: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

28

demonstrated by lack of nicotine-induced place preference (Berrendero et al., 2002;

Berrendero et al., 2005). Finally, in rats the opioid antagonist naloxone precipitated

nicotine withdrawal as well as prevented nicotine alleviation of withdrawal (Malin et al.,

1993; Malin et al., 1996) further suggesting the involvement of the opioid system in

nicotine dependence.

Naltrexone is a non-specific antagonist of the opioid receptors used for the

treatment of opioid-dependence (Modesto-Lowe and Van Kirk, 2002). Due to the

connection between nicotine dependence and opioid system, naltrexone has been

investigated as a potential therapy for smoking cessation. However, it should be noted

that several studies have found that naltrexone, when administered alone, led to

significant adverse effects due to precipitation of nicotine withdrawal (Covey et al., 1999;

Wong et al., 1999). On the contrary, studies of naltrexone with nicotine replacement

have found the drug combination to be more effective. For example, in a preliminary

study naltrexone (50 mg) plus nicotine patch (21 mg) resulted in higher 2-week

continuous abstinence rate (weeks 3-4 in a 4 week study) compared to placebo plus

patch (56.3% vs. 31.3%) (Krishnan-Sarin et al., 2003). In a larger study examining

naltrexone (0, 25, 50, and 100 mg/d) plus nicotine patch (21 mg) for smoking cessation,

it was found that naltrexone only at 100 mg/d (but not 25 and 50 mg/day) plus nicotine

patch led to significantly better 6-week continuous abstinence rate compared to placebo

plus patch. However, subsequent exploratory analzysis of 7-day point-prevalence

abstinence at 3-, 6-, and 12-months did not find any difference in cessation between

naltrexone vs placebo (O'Malley et al., 2006). As mentioned previously, naltrexone

causes withdrawal symptoms such as nausea, vomiting, diahrea, depression, and

insomnia which may limit its utility (O'Malley et al., 2006).

Section 2.2.5 Rimonabant (SR141716)

Page 44: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

29

The involvement of the endocannabinoid system with nicotine dependence was

established when it was observed that chronic nicotine administration increased the level

of the endogenous cannabinoid receptor agonists anandamide (AEA) and 2-

arachidonoly-glycerol (2-AG) (for CB1 and CB2 receptors, respectively). The increases in

these agonists are localized to the brainstem, a region involved in drug withdrawal

(Gonzalez et al., 2002; Shoemaker et al., 2005; Steffens et al., 2005). In rats,

administration of the CB1 receptor antagonist SR141716 (rimonabant) blocked nicotine-

induced dopamine release in the nucleus accumbens (Cohen et al., 2002). Rimonabant

also prevented IV nicotine self-administration, conditioned stimuli-induced nicotine-

seeking behaviors and nicotine-conditioned place preference in rats (Cohen et al., 2005;

Cohen et al., 2002; Le Foll and Goldberg, 2004). The precise mechanism of actions of

CB1 receptor antagonism on nicotine reward/reinforcement remains unclear.

A recent Cochrane review found that the pooled odds ratio (from STRATUS-US

and STRATUS-EUROPE) of quitting smoking at one year with 20 mg rimonabant to be

1.61 (95% CI 1.12 to 2.30) with the 5 mg showing no benefit over placebo (Cahill and

Ussher, 2007). When analyzed separately, only the STRATUA-US study showed a

significant benefit in cessation. Rimonabant showed potential benefit in relapse

prevention (STRATUS-Worldwide); both 5 mg and 20 mg treatment provided

approximately 1.5-fold increase in remaining abstinent (Cahill and Ussher, 2007). The

most common side effects were nausea and upper respiratory tract infection (Anthenelli,

2005). Due to the inconclusive evidence showing rimonabant to be effective in smoking

cessation, the drug is currently not approved for the indication in the United States. As of

2008, the European Medicines Agency has suspended the use of rimonabant in the

treatment of obesity due to increased risk of psychiatry disorders in overweight or obese

patients (Allchurch, 2008).

Page 45: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

30

Section 2.2.6 Nicotine Vaccines

The nicotine vaccine is another potential treatment for smoking. The mechanism behind

this therapeutic strategy is to prevent nicotine from entering the brain via binding to

nicotine-specific antibodies (Hieda et al., 2000; Pentel et al., 2000). An advantage of

nicotine vaccines is that only occasional booster shots, rather than daily administration,

should be needed for therapeutic effect. In a 38-week study it was found that during ad

libitum smoking, nicotine vaccine, at all doses tested (50, 100, and 200 µg; 1 primary

challenge shot and 3 booster shots at each dose), did not cause compensatory smoking

behaviors or precipitate withdrawal (Hatsukami et al., 2005). In addition, the 30-day

continuous abstinence rates were higher in patients receiving the highest vaccine dose

(~38%) compared with placebo (~10%) (Hatsukami et al., 2005). Careful interpretation of

these results is needed as the sample size was small and the study was primarily

designed to evaluate the safety and immunogenicity of the vaccine. In a subsequent

study with larger sample size (N=341) and a different vaccine, it was found that the four-

month continuous abstinent rate was significantly higher in individuals with the highest

antibody titer compared to placebo (56.6% [N=30 out of 53] vs. 31.3% [N=25 out of 80];

OR = 2.9; p<0.01) (Cornuz et al., 2008). The vaccine is associated with greater adverse

events such as flu-like symptoms and pyrexia (Cornuz et al., 2008). Currently nicotine

vaccine is still being investigated for efficacy in Phase III clinical trials (Clinicaltrial.gov).

Section 2.2.7 Monoamine Oxidase Inhibitors (MAOIs)

MAO enzymes are responsible for the metabolism of various endogenous substrates

(e.g. adrenaline, noradrenaline, dopamine, 5-hydroxytryptamine, 2-phenylethylamine,

Page 46: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

31

tryptamine, tyramine) and thus they can play an important role in smoking behaviours

(Youdim et al., 2006). The relationship between MAO activities and smoking was first

observed almost three decades ago (Oreland et al., 1981). In smokers, the brain MAO-A

and MAO-B activities are ~30 and ~40% lower compared to non-smokers, respectively

(Fowler et al., 1996a; Fowler et al., 1996b). MAO-B is the major isoform of the

monoamine oxidase in the brain (~75-80%) (Saura Marti et al., 1990; Squires, 1972).

The inhibition of MAO-B is reversed after long-term abstinence from smoking, likely due

to the synthesis of new enzymes (Gilbert et al., 2003). These findings suggest MAO-B

inhibitor(s) may be present in tobacco smoke (Khalil et al., 2000) that may inhibit MAO-

mediated dopamine metabolism, thereby enhancing the rewarding effects of nicotine

during smoking (Lewis et al., 2007). However, the specific tobacco smoke compounds

that lead to long term inhibition of MAO-B have not been identified.

Through inhibition of MAO-A and MAO-B, the metabolism of dopamine can

potentially be reduced and this may enhance the rewarding effects of nicotine during

smoking. In a preliminary study, the reversible MAO-A inhibitor moclobemide

significantly reduced self-reported smoking rates, although at one year there was no

difference in abstinence compared with placebo (Berlin et al., 1995). The reversible

inhibitor of MAO-B lazabemide has also been evaluated for smoking cessation, but the

study was terminated due to concerns of hepatotoxicity (Berlin et al., 2002).

The irreversible MAO-B inhibitor selegiline has also demonstrated some efficacy

in several small studies. In one study, oral selegiline (10 mg) decreased physical

symptoms (e.g. hand shakes, increased heart rate) and craving during abstinence as

well as reducing smoking satisfaction during smoking (Houtsmuller et al., 2002). In

another study, 5 mg b.i.d. oral selegiline increased the trial end point (8-week) 7-day

point prevalent abstinence compared with placebo [OR 4.64 (95% CI: 1.02-21.00

p<0.05)]; however at 6 months the 7-day point prevalent abstinence was not significant

Page 47: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

32

even though the OR was 4.75. A limitation of the study was the small sample size (N=40)

(George et al., 2003). In a third study that used a combination of oral selegiline and

nicotine patch, selegiline plus nicotine patch doubled the 52-week continuous abstinence

rate compared to nicotine patch alone (25% vs. 11%), although the difference was not

significant possibly due to small number of subjects (N=38) (Biberman et al., 2003).

Common adverse effects associated with selegiline monotherapy (5 mg b.i.d.)

are nausea and dizziness (Heinonen and Myllyla, 1998). Despite the dopamine

potentiating effect of selegiline, as well as the production of L-methamphetamine and L-

amphetamine metabolites (instead of the more potent D-isomers) from selegiline, there

is no indication that selegiline is addictive (Schneider et al., 1994).

Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®)

Section 2.2.7.1.1 Selegiline Pharmacodynamics

Selegiline is used clinically in conjunction with levodopa for the treatment of Parkinson’s

disease (Gerlach et al., 1996). The recommended dose of selegiline for Parkinson’s

disease is 10 mg (5 mg p.o. b.i.d.) (Gerlach et al., 1996). This is the same dose of

selegiline that was used in several smoking cessation trials. At these clinical doses,

selegiline is a selective and irreversible inhibitor of MAO-B (Gerlach et al., 1996). The

drug inhibits MAO-B by covalently binding to its flavine-adenine dinucleotide (FAD) co-

factor (Youdim, 1978), effectively decreasing the amount of enzyme available to

metabolize dopamine. At higher doses the selectivity of selegiline decreases and it also

inhibits MAO-A (Gerlach et al., 1992).

In addition to enhancing dopamine availability by inhibiting MAO-B, selegiline

may have another mechanism of action that modulates dopaminergic activities that may

be important to smoking cessation. Phenylethylamine (β-PEA), an endogenous

Page 48: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

33

substrate of MAO-B, exists in trace concentrations in the brain and was found to be

significantly higher (up to 10-fold) in selegiline-treated Parkinson’s patients compared to

non-treated patients (Reynolds et al., 1978). In monkeys, selegiline increased brain β-

PEA levels (Paterson et al., 1995). In rats, selegiline increased β-PEA levels in the

striatum and enhanced the neuronal response to dopamine stimulation (Berry et al.,

1994; Paterson et al., 1991). In addition, direct administration of β-PEA into the rat

striatum caused the release of dopamine (Bailey et al., 1987; Kuroki et al., 1990). β-PEA

has been proposed to be a neuromodulator of dopaminergic transmission although the

mechanism by which this occurs has not been determined (Paterson et al., 1990).

Therefore, the selegiline-mediated increase in β-PEA may enhance neuronal stimulation

by dopamine. The selegiline metabolite desmethylselegiline can also inhibit MAO-B

activity in vivo in humans although its contribution may be minor (Heinonen et al., 1997).

Section 2.2.7.1.2 Selegiline Metabolism

In humans, selegiline is extensively metabolized; less than 1% of an oral dose is

excreted unchanged in the urine (Shin, 1997). In correlation analyses, CYP2B6 accounts

for the largest variation in the metabolism of selegiline to its primary metabolites

desmethylselegiline (77%) and L-methamphetamine (79%) (Benetton et al., 2007). Both

CYP2A6 and CYP3A4 contribute to these metabolic pathways to some extent (14-15%

for both desmethylselegiline and L-methamphetamine by CYP2A6 and 33-35% for both

desmethylselegiline and L-methamphetamine by CYP3A4) (Benetton et al., 2007). The

plasma half-life of 10 mg of oral selegiline is 1.0 ± 0.9 hr; however, this increased to 2.7

± 1.8 hr following chronic (8 days) administration, indicating the metabolism slows with

chronic selegiline treatment. In addition, the observed accumulation ratio of the drug is

Page 49: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

34

higher at 2.7 compared to the estimated accumulation ratio of 1.0 based on elimination

half-life (although not statistically significant), indicating possibly that selegiline, or a

metabolite, down-regulates the enzyme(s) responsible for its metabolism or inhibits

selegiline’s clearance (Laine et al., 2000). Age and sex were not found to contribute to

differences in selegiline pharmacokinetics (Barrett et al., 1996a; Barrett et al., 1996b).

The plasma elimination half-lives of desmethylselegiline and L-

methamphetamine are 31 ± 8 and 161 ± 38 hr, respectively. L-amphetamine, the major

metabolite of desmethylselegiline, has a half-life of 40 ± 7 hr (Laine et al., 2000). Similar

to selegiline, the three metabolites showed accumulation over eight days to some extent.

L-methamphetamine and L-amphetamine are further metabolized to various p-

hydroxylation and β-hydroxylation products whose pharmacokinetic profiles have not

been examined in detail (Shin, 1997).

In mice, an inhibitor study showed that CYP2E1 is responsible for more than

50% of the metabolism of selegiline to all three of its metabolites (desmethylselegiline,

methamphetamine, and amphetamine) with a minor contribution from mouse CYP3A

enzyme(s) (Valoti et al., 2000). The contribution of CYP2A5 to the metabolism of

selegiline was not examined (Valoti et al., 2000). Similar to humans, the plasma half-life

of selegiline in mice is between 0.9 to 1.8 hr (Magyar et al., 2007).

The observation that CYP2A6 is involved in selegiline metabolism and chronic

administration of this drug decreases its own metabolism suggests that selegiline may

inhibit CYP2A6. These findings have implications for selegiline in the smoking cessation

and will be further described in sections 2.3.3.2 and 5.3.2.

Section 2.3 The Inhibition of Nicotine Metabolism and its Application to

Smoking Cessation

Section 2.3.1 Inhibitors of Human CYP2A6

Page 50: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

35

CYP2A6 has a compact, narrow, hydrophobic active site (Yano et al., 2005); therefore,

compounds that are good inhibitors of CYP2A6 are small and planar (e.g. structurally

related to nicotine) (Denton et al., 2005; Rahnasto et al., 2005; Yano et al., 2006). A

variety of compounds can inhibit CYP2A6 in vitro including pharmaceuticals [e.g.

methoxsalen (Maenpaa et al., 1993), pilocarpine (Kimonen et al., 1995), clotrimazole,

miconazole (Draper et al., 1997), tranylcypromine, tryptamine (Zhang et al., 2001),

amphetamine (Rahnasto et al., 2005)], naturally occurring substances [furanocoumarins

(Maenpaa et al., 1993), various nicotine-related alkaloids (in vitro) (Denton et al., 2004;

Denton et al., 2005), soy isoflavones (in vivo) (Nakajima et al., 2006b), grapefruit juice

(in vivo) (Hukkanen et al., 2006), star fruit juice (in vitro) (Yang, 2007)], and endogenous

substances (in vitro) [e.g. β-PEA (Rahnasto et al., 2003), dopamine (Higashi et al.,

2007b)]. Of the above, methoxsalen can also inhibit nicotine metabolism in vivo and may

be clinically significant in its application in smoking cessation (further discussed in

sections 2.3.3.1 and 5.3.1).

Section 2.3.2 Inhibitor of Mouse CYP2A5

Some CYP2A6 inhibitors also appear to be inhibitors of CYP2A5. Examples are

furanocoumarins-derivatives such methoxsalen (Maenpaa et al., 1993), pilocarpine

(Kimonen et al., 1995), and amphetamine (Rahnasto et al., 2005). Other sample

compounds that inhibit CYP2A5 include, various lactone (coumarin-like) compounds

(Juvonen et al., 2000), as well as etomidate, metomidate and metyrapone (Kojo et al.,

1989). These results suggest a similar inhibitor profile between the human CYP2A6 and

mouse CYP2A5 and the mouse may be useful in the evaluation of novel CYP2A6

inhibitors.

Page 51: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

36

Section 2.3.3 Mechanism-Based Inhibition

Mechanism-based (suicide) inhibitors (MBIs) are compounds that inhibit enzyme

activities by covalent modification. As a result, the enzyme is no longer functional, and

its activity can only be recovered via synthesis of new enzyme. There are several ways

in which mechanism-based inhibition can occur (Osawa and Pohl, 1989). First is the

covalent modification of the prosthetic heme in which the reactive metabolite binds to the

heme and together they dissociate from the enzyme. Another is the covalent binding of a

reactive metabolite to the holoenzyme. Third is the covalent binding of the heme to the

protein mediated by the reactive metabolite. In addition, the reactive metabolite can also

covalently bind to the enzyme co-factor [e.g. binding of selegiline reactive metabolite to

the flavin adenine dinucleotide (FAD) of MAO-B (Youdim, 1978)].

There are seven experimental criteria that can be used to establish whether a

compound is a MBI (Silverman, 1995). First, there is time-dependent loss of activity due

to reduction in active enzyme. Second, the inhibition is saturable – the rate of

inactivation of enzyme is proportional to the number of inactivator molecules added until

all enzyme molecules are saturated. Third, a substrate or a competitive (reversible)

inhibitor should slow down inactivation by the potential MBI (substrate protection). Fourth,

the inhibition is irreversibe such that dialysis or filtration (gel or membrane) does not

restore enzyme activity. Fifth, there should be inactivator stoichiometry in which only one

inactivator molecule is attached to each enzymatic site. Sixth, the involvement of a

catalytic step – the enzyme needs to convert the inhibitor to an active species that

actually carries out the inactivation step and therefore an enzymatic reaction is required.

Finally, the active species formed from enzymatic activity must inactivate the enzyme

prior to its release from the active site.

Page 52: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

37

Section 2.3.3.1 Methoxsalen (8-methoxypsoralen; 8-MOP): a CYP2A5 and

CYP2A6 MBI

Methoxsalen is a naturally occurring psoralen compound that is used for the treatment of

psoriasis and cutaneous T-cell lymphoma in conjunction with UV exposure (Bond et al.,

1981; Young, 1993). Methoxsalen is a competitive (Ki), non-competitive (Kiu), and a

mechanism-based (KI and kinact) inhibitor of CYP2A6 inhibiting both coumarin (Ki =1.5 µM

[human liver microsomes]; Kiu =0.5 µM; KI =0.3-0.8 µM [kinact =0.5 min-1]) and nicotine (Ki

=0.8 µM) metabolism (Draper et al., 1997; Koenigs et al., 1997; Maenpaa et al., 1994;

Zhang et al., 2001). In mice, methoxsalen is a competitive (Ki not determined), non-

competitive (Kiu =1.7 µM), and mechanism-based (kinact =0.17 min-1) inhibitor of

microsomal CYP2A5-mediated coumarin metabolism (Visoni et al., 2008). Methoxsalen

also inhibits human CYP1A2, CYP2B6, CYP2A13, and to some extent several other

P450 enzymes (von Weymarn et al., 2005; Zhang et al., 2001) in addition to human

CYP2A6 and mouse CYP2A5.

Section 2.3.3.2 Selegiline as a MBI

As mentioned in section 2.2.7.1, selegiline is an irreversible MBI of MAO-B. Selegiline

belongs to the acetylene group of chemicals that contain a carbon-carbon triple bond,

which are known to be potent MBIs (Correia and Ortiz de Montellano, 2005). Selegiline

is a MBI of CYP2B1, the rat nicotine metabolizing enzyme, although the precise

mechanism of inactivation is unclear (Sharma et al., 1996). It is unknown what other

additional rodent P450 enzymes selegiline may inhibit; however, in humans selegiline

Page 53: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

38

has little to no inhibitory effects on CYP1A2, CYP2C9, CYP2C19, CYP2D6, and CYP3A

in vitro (Polasek et al., 2006). Considering that CYP2A6 is involved in the metabolism of

selegiline in vitro (Benetton et al., 2007) and chronic selegiline treatment reduces its own

metabolism (Laine et al., 2000), it is possible that selegiline is a MBI of CYP2A6.

Section 2.3.4 Improving Smoking Cessation through CYP2A6 Inhibition

The impact of CYP2A6 genotype on nicotine metabolism and smoking behaviors has

been examined in detail (Ho and Tyndale, 2007; Malaiyandi et al., 2005; Mwenifumbo

and Tyndale, 2007). The most striking observation of the influence of CYP2A6 genetics

on smoking behaviours comes from slow metabolizers of nicotine (50% or less predicted

CYP2A6 activity). For instance, slow metabolizers are less likely to be current smokers

(OR 0.52 [95% CI: 0.29-0.95], p=0.03) (Schoedel et al., 2004), smoke significantly fewer

(20-25%) cigarettes (Malaiyandi et al., 2006; Schoedel et al., 2004), and have

significantly reduced (~30%) total inhalation volume (Strasser et al., 2007). In addition,

when smokers are ranked according to rates of nicotine metabolism (as measured by

the 3-hydroxycotinine to cotinine ratio), individuals in slowest quartile are more likely to

quit smoking after counseling (32%) compared to those in the fastest quartile (10%)

(Patterson et al., 2008). Furthermore, with nicotine patch treatment, individuals in the

upper quartiles (faster nicotine metabolism) have lower chance of quitting smoking

compared to those in the lowest quartile at the end of treatment and 6 months’ follow-up

(OR 0.72 [95% CI: 0.83-1.33], P=0.005) (Lerman et al., 2006). Similar observation is

also seen in light-smokers treated with nicotine gum (Ho et al., 2009).

Based on the above findings, it was hypothesized that one could

pharmacologically mimic impairments in the CYP2A6 enzyme (phenocopying) to

increase nicotine levels and decrease the number of cigarettes smoked (Sellers et al.,

Page 54: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

39

2000; Tyndale and Sellers, 2002). By inhibiting nicotine metabolism, the systemic

availability of nicotine is increased; therefore, the concomitant use of CYP2A6 inhibitors

with NRT products should enhance smoking cessation. For example, some of the

nicotine released from nicotine gum is absorbed by the buccal cavity, while a significant

portion is swallowed (Benowitz et al., 1987). Of the nicotine swallowed (from nicotine

gum, lozenges, or pills), approximately 80% is metabolized via first-pass metabolism

before entering systemic circulation (Benowitz et al., 1991b). Inhibition of CYP2A6 would

decrease first-pass metabolism and increase the bioavailability of nicotine as well as

other CYP2A6 substrates (see Section 1.6.1). In a preliminary study, after three days of

treatment with methoxsalen (10 mg t.i.d.) plus nicotine gum (4 mg), the mean plasma

nicotine levels were significantly higher compared to placebo plus nicotine gum (15.3 vs.

10.1 ng/ml; p<0.01) (Sellers et al., 2002). Similarly, the pill form of nicotine for smoking

cessation may also benefit from the co-administration of a CYP2A6 inhibitor. This is due

to the fact that higher doses of oral nicotine that are required to overcome first-pass

metabolism can cause gastrointestinal irritation (Benowitz et al., 1991b). Compared with

oral nicotine alone, the addition of methoxsalen (10 and 30 mg) to 4 mg of oral nicotine

significantly increased mean plasma nicotine levels (>70%) and decreased the desire to

smoke (p<0.05) (Tyndale et al., 1999). In another study, methoxsalen (30 mg) plus oral

nicotine (4 mg) also significantly altered smoking behaviours. Participants on this

treatment smoked fewer cigarettes (3.1 vs. 4.1; p<0.01) during a 90 minute free smoking

session, had lower levels of breath carbon monoxide (4.6 vs 8.7 ppm; p<0.01), lower

numbers of puffs (~28 vs. ~38; p<0.01) and inhalation intensity compared with placebo.

The subjects also had increased latency to lighting the second cigarette (~43 vs. ~32

mins; p<0.01) on methoxsalen plus nicotine compared to placebo plus nicotine (Sellers

et al., 2000). Finally, CYP2A6 inhibitor alone may be able to reduce smoking and

Page 55: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

40

increase quitting, similar to the pharmacogenetic findings in slow metabolizers

(Malaiyandi et al., 2006; Patterson et al., 2008; Rao et al., 2000; Schoedel et al., 2004);

however, this remains to be tested. An advantage of this strategy is that the long-term

behavioral effects of mimicking a defect in nicotine metabolism can be predicted based

on available pharmacogenetic and behavioral studies.

Section 2.3.5 Reduction of Tobacco-Specific Nitrosamine Activation

through Human CYP2A6 Inhibition

One goal of smoking cessation is to reduce smoking-associated morbidity and mortality.

The tobacco-specific nitrosamine NNK is a potent procarcinogen found in tobacco

smoke (Hecht and Hoffmann, 1988). NNK is primarily detoxified by conversion to 4-

(methylnitrosamino)-1-(3-pyridyl)-1-butanol (NNAL), followed by conjugation to NNAL-

glucuronide (Hecht et al., 1993; Maser et al., 1996). Alternatively, NNK can be

metabolically activated to a reactive metabolite capable of forming DNA adducts via α-

hydroxylation (Hecht et al., 1988; Peterson et al., 1990); one enzyme responsible for this

process is CYP2A6 (Nishikawa et al., 2004). The other enzyme that exhits the highest in

vitro activity towards NNK activity is CYP2A13 (with at least 100-fold greater activity than

CYP2A6) and it is primarily expressed in the lung and may be one of the enzyme

responsible for NNK activation in situ (Jalas et al., 2005). However, CYP2A13 level in

lung tissues is variable between individuals and thus CYP2A6, as well as other enzymes

with lower in vitro NNK activation activities, such as CYP1A1/2, CYP2B6, CYP2D6,

CYP2E1, and CYP3A4, may contribute to NNK activation in those lacking CYP2A13

(Jalas et al., 2005; Zhang et al., 2007). The involvement of CYP2A6 in lung

carcinogenesis is supported in a recent epidemiology study: individuals that were

CYP2A6-poor metabolizers had significantly less risk for developing lung cancer, even

Page 56: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

41

after adjusting for the number of cigarettes smoked (Fujieda et al., 2004). Furthermore,

the expression of CYP2A6 is higher in large lung adenocarcinoma tissues and tumors

that have metastasized compared to smaller tumor and non-metastasized tissues,

respectively, suggesting the involvement of CYP2A6 in lung carcinogenesis (Matsuda et

al., 2007).

In mice, NNK exposure causes the formation of lung tumors (Hecht et al., 1978;

Hecht et al., 1994; Okubo et al., 2005), and administration of the mouse CYP2A5

inhibitor, methoxsalen, reduced lung malignancies (Miyazaki et al., 2005; Takeuchi et al.,

2003). The finding is corroborated with findings in humans in which methoxsalen

administration in smokers re-routed excretion of NNK to NNAL and NNAL-glucuronide,

suggesting the α-hydroxylation of NNK was inhibited (Sellers et al., 2003a).

CYP2A13 also contributes to the α-hydroxylation of NNK (He et al., 2004a). Due

to the high level of CYP2A13 in the respiratory tract (Su et al., 2000), it is thought that

CYP2A13 plays a significant role in the bioactivation of NNK in the lung. This is

supported by the finding that NNK activation is significantly correlated with CYP2A13

levels in lung microsomes (Zhang et al., 2007). In contrast, the association between

CYP2A13 levels/activities with NNK metabolism was only found in a subgroup of lung

cancer biopsy samples (Brown et al., 2007). It is possible that inhibition of CYP2A

enzymes (CYP2A6 and CYP2A13) may decrease the risk of development of NNK-

induced lung cancer.

Page 57: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

42

Section 3 Nicotine Pharmacological Behavioural Models

Section 3.1 Nicotine Self-Administration

Nicotine has reinforcing properties in mice as demonstrated through the use of self-

administration paradigms. Intravenous nicotine self-administration (NSA) is a widely

used method of testing nicotine’s reinforcing properties in rats (Corrigall and Coen, 1991;

Corrigall et al., 1994) but is more technically demanding in mice due to their small size

(Sparks and Pauly, 1999). Oral NSA, whilst not modeling smoking behaviours directly,

has been studied extensively in mice (Adriani et al., 2002; Aschhoff et al., 1999; Meliska

et al., 1995; Robinson et al., 1996; Stolerman et al., 1999) and has some similar features

to smoking, which may make it a suitable method of studying nicotine pharmacology. For

instance, oral NSA occurs intermittently primarily during the wake period. There exhibits

a wide spectrum of preference for oral nicotine between mouse strains (Aschhoff et al.,

1999; Robinson et al., 1996), suggesting a role of genetic influences in NSA, similar to

smoking behaviours in humans (Ho and Tyndale, 2007). Chronic oral NSA produces

pharmacologically relevant levels of nicotine in the plasma of mice. When nicotine is

provided in a free-access two-bottle choice paradigm, mice consume an amount of

nicotine that results in a plasma concentration (~35 ng/ml) similar to those seen in

humans during ad libitum smoking (10 to 50 ng/ml) (Hukkanen et al., 2005b; Rowell et

al., 1983). Chronic oral NSA can also lead to physical dependence as measured by

somatic signs (paw tremors, backing, and head shakes) although this occurs at high

nicotine levels (200 µg/ml) (Grabus et al., 2005). In addition, oral NSA up-regulates both

α4 and α7 nicotinic acetylcholine receptors in mouse brain (Sparks and Pauly, 1999);

this is consistent with findings that chronic nicotine and smoking increase nicotinic

receptors in baboons and smokers, respectively (Kassiou et al., 2001; Perry et al., 1999).

Page 58: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

43

Oral nicotine also alters the activities of enzymes (e.g. CREB, pERK) in pathways

thought to mediate nicotine reward (Brunzell et al., 2003). Behaviourally, in mice oral

nicotine is rewarding at an optimal range: too low a dose and nicotine is not rewarding

and at high doses it is aversive (Robinson et al., 1996). This finding is in line with human

and other animal studies (Rose and Corrigall, 1997). Together, these data suggest that

oral NSA in mice may be a suitable method of studying some aspects of the

pharmacology of nicotine.

Section 3.2 Analgesic Effect of Nicotine and Application in Nicotine

Studies

Nicotine’s analgesic properties have long been documented (Meyer et al., 2000). The

mechanism of action is thought to be mediated indirectly by the activation of the opioid

system. In a study examining the anti-nociceptive effect of nicotine on tail-pinch in mice,

nicotine induced the release of the endogenous opioid peptide, Met-enkephalin (agonist

at both µ- and δ-opioid receptors), but not dynorphins and endorphins, in the spinal cord.

The analgesic action was mediated in part by the nicotinic acetylcholine receptors as

mecamylamine ameliorated the effect (Kiguchi et al., 2008). In addition to analgesia,

nicotine also has both hypothermic and ataxic effects (Kita et al., 1988; Marks et al.,

1984).

The tail-flick and hot-plate tests are used to study nicotine’s analgesic effects

(Marubio et al., 1999; Rogers and Iwamoto, 1993). The tail-flick response is a spinal

reflex (Caggiula et al., 1995); therefore, the ataxic-properties of nicotine are not thought

to be confounding variables in this behavioural test (Rogers and Iwamoto, 1993). On the

other hand, the hot-plate response is thought to be mediated by supra-spinal (i.e. brain)

processes, thus it is conceivable that ataxia induced by nicotine may confound

Page 59: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

44

interpretation of results from this method (Caggiula et al., 1995). A recent study has

found that the tail-flick reflex requires α4β2 receptors whereas response to hot-plate

requires α4β2 and possibly α7 or other receptor subsets (Damaj et al., 2007a), both

indicating the direct involvement of nicotinic receptors in nicotine-induced analgesia.

Page 60: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

45

Section 4 Outline of Chapters

Section 4.1 Chapter 2: Characterization and Comparison of Nicotine and

Cotinine Metabolism In Vitro and In Vivo in DBA/2 and

C57BL/6 Mice

Rationale

In order to determine whether the mouse is a suitable model of nicotine metabolism, the

disposition kinetics of nicotine and its metabolites, as well as the enzyme responsible for

the primary metabolism of nicotine, need to be assessed. Second, to examine whether

interstrain differences in nicotine pharmacology could be contributed to by differences in

nicotine metabolism, nicotine pharmacokinetics between mouse strains also need to be

studied. We hypothesized that the mouse CYP2A5 is responsible for the primary

metabolism of nicotine to cotinine and that strain differences (in CYP2A5) could result in

differences in nicotine pharmacokinetics.

Outline

This chapter describes the studies undertaken to examine the pharmacokinetics of

nicotine and cotinine in mice. Immuno-inhibition studies indicated that, similar to human

CYP2A6, the mouse CYP2A5 is responsible for the primary metabolism of nicotine to

cotinine and cotinine to 3-hydroxycotinine in vitro. The metabolism of nicotine and

cotinine, both in vitro and in vivo, between two commonly studied mouse strains (DBA/2

and C57BL/6) were also described. Pharmacokinetic studies and in vitro assays

indicated that the metabolism of cotinine, but not nicotine, is different between these two

mouse strains.

Contribution to Thesis

Page 61: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

46

This chapter contributes to the overall thesis by first describing the pharmacokinetics of

nicotine and cotinine in mice, and second, by defining the contribution of a specific

enzyme (i.e. CYP2A5) responsible for the (in vitro) metabolism of these compounds. The

latter observation provided a line of evidence that the mouse could be a suitable model

for studying inhibitors of human nicotine metabolism. The chapter also suggests that

interstrain genetic variations may affect xenobiotic metabolism in a substrate-dependent

manner.

Section 4.2 Chapter 3: Nicotine Self-Administration in Mice is Associated

with Rates of Nicotine Inactivation by CYP2A5

Rationale

In humans, smoking (nicotine self-administration) is significantly affected by CYP2A6

variations (structure/function). We have previously demonstrated that CYP2A5 is

responsible for the primary metabolism of nicotine to cotinine in mice. However, the

association of nicotine self-administration behaviours with CYP2A5 variations has not

been examined. We hypothesized that oral nicotine self-administration in mice is

associated with CYP2A5 levels and rates of nicotine inactivation.

Outline

This chapter describes the study carried out to determine the amount of CYP2A5 protein

as well as the rates of nicotine metabolism in a group of F2 mice. The F2 mice were

previously generated by breeding a high nicotine consumption strain with a low nicotine

consumption strain. The F2 mice were also tested for nicotine self-administration in a 2-

bottle choice paradigm. In male mice, the amount of nicotine self-administered was

Page 62: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

47

significantly associated with the amount of CYP2A5 protein as well as rates of in vitro

nicotine metabolism.

Contribution to Thesis

The above findings further demonstrated that: 1) nicotine metabolism is associated with

CYP2A5; 2) genetic variation in CYP2A5 between DBA/2 and C57Bl/6 mice can alter

nicotine metabolism; and 3) CYP2A5 (and nicotine metabolism) is associated with, and

potentially influencing, nicotine pharmacology including, importantly, nicotine self-

administration. This chapter contributes to the overall thesis by demonstrating that

CYP2A5 influences nicotine metabolism through its expression/activity. In addition,

CYP2A5 can potentially alter the behavioural pharmacology of nicotine. These

observations strengthen the application of the mouse model of nicotine metabolism for

both inhibitor studies as well as genetic studies.

Section 4.3 Chapter 4: Inhibition of Nicotine Metabolism by Methoxysalen:

Pharmacokinetic and Pharmacological Studies in Mice

Rationale

We have previously demonstrated that CYP2A5 levels/activities are significantly

associated with nicotine behavioural pharmacology (i.e. oral nicotine self-administration

behaviour) in mice. This finding is analogous to human studies in which CYP2A6

function is significantly correlated with levels of smoking. We are interested in

determining whether direct manipulation (i.e. inhibition) of CYP2A5, can alter nicotine

metabolism as well as the resulting behavioural response to nicotine. We hypothesized

that inhibition of CYP2A5 would significantly reduce nicotine metabolism and enhance

the pharmacological effects of nicotine.

Page 63: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

48

Outline

This chapter describes the use of the CYP2A6 inhibitor, methoxsalen, to determine

whether it can inhibit nicotine metabolism in mice. Initially the examination of the

inhibition of nicotine metabolism in mouse hepatic microsomes was performed to

determine whether this inhibitor is effective in mice (specifically in inhibiting CYP2A5-

mediated nicotine metabolism). This was followed by the in vivo administration of

methoxsalen and the characterization on its impact on nicotine pharmacokinetics and

resulting behavioural pharmacology. The anti-nociception effect of nicotine on tail-flick

and hot-plate procedures was measured as the indicator of the pharmacological effects

of nicotine. Methoxsalen significantly inhibited nicotine metabolism both in vitro and in

vivo. More importantly, the reduction in clearance of nicotine was associated with a

dramatic prolongation in its anti-nociception effect as measured by tail-flick and hot-plate

procedures.

Contribution to Thesis

This chapter contributes to the thesis by demonstrating that: 1) Mice may be suitable for

studying CYP2A6 inhibitors; and 2) inhibition of nicotine metabolism can directly alter the

behavioural pharmacological effects of nicotine. These observations further support the

use of the mouse for the in vivo screenings of novel inhibitors of nicotine metabolism,

especially for the purpose of smoking cessation.

Section 4.4 Chapter 5: Selegiline (L-Deprenyl) is a Mechanism-based

Inactivator of CYP2A6 Inhibiting Nicotine Metabolism in

Humans and Mice

Page 64: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

49

Rationale

Selegiline is an anti-Parkinsonian drug currently being investigated as a potential

therapy for smoking cessation due to its ability to decrease dopamine metabolism

through inhibition of MAO-B. In humans chronic treatment with selegiline decreases its

own metabolism. Selegiline belongs to the acetylene group of compounds that are

known mechanism-based inhibitors. Since CYP2A6 has been demonstrated to

metabolize selegiline, we investigated whether selegiline could also inhibit mouse and

human nicotine metabolism. We hypothesized that selegiline can inhibit nicotine

metabolism by mouse CYP2A5 and human CYP2A6. This may be an additional

mechanism by which selegiline mediates smoking cessation.

Outline

This chapter described the various studies performed to characterize the inhibitory effect

of selegiline on nicotine metabolism in mouse and human hepatic microsomes, in cDNA-

expressed CYP2A5 and CYP2A6, and in mouse in vivo. Inhibition of nicotine metabolism

in vitro with major selegiline metabolites (desmethylselegline, L-methamphetamine, and

L-amphetamine) was also performed. Selegiline significantly inhibited nicotine

metabolism in all in vitro enzyme systems studied. Selegiline also inhibited nicotine

metabolism in mice in vivo. Most importantly, selegiline was found to be a potent

mechanism-based inhibitor of CYP2A6-mediated nicotine metabolism.

Contribution to Thesis

This chapter contributes to the thesis by demonstrating that in vitro, and especially in

vivo, the mouse model of nicotine metabolism is a potentially useful for the screening of

novel inhibitors of CYP2A6.

Page 65: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

50

CHAPTER 2: CHARACTERIZATION AND COMPARISON OF NICOTINE AND COTININE METABOLISM IN VITRO AND IN VIVO IN DBA/2 AND C57Bl/6 MICE

Eric C. K. Siu and Rachel F. Tyndale As published in: Molecular Pharmacology, 2007, Mar; 71(3):826-834.

All of the experiments, data analyses, and writing of the manuscript were performed by

ECK Siu except for the mass-spectrometry analysis which was performed at the

Proteomic and Mass Spectrometry Centre at the University of Toronto. The study was

designed by ECK Siu and RF Tyndale.

Reprinted from Molecular Pharmacology, Volume 71, Siu et al., “CHARACTERIZATION AND COMPARISON OF NICOTINE AND COTININE METABOLISM IN VITRO AND IN VIVO IN DBA/2 AND C57Bl/6 MICE”, pp 826-834, Copyright (2007) with permission from The American Society for Pharmacology and Experimental Therapeutics. © 2007 The American Society for Pharmacology and Experimental Therapeutics

Page 66: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

51

Abstract

DBA/2 and C57Bl/6 are two commonly used mouse strains that differ in response to

nicotine. Previous studies have shown that the nicotine metabolizing enzyme CYP2A5

differs in coumarin metabolism between these two strains suggesting differences in

nicotine metabolism. Nicotine was metabolized to cotinine in vitro by two enzymatic

sites. The high-affinity sites exhibited similar parameters (Km: 10.7 ± 4.8 vs. 11.4 ± 3.6

µM; Vmax: 0.58 ± 0.18 vs. 0.50 ± 0.07 nmol/min/mg, for DBA/2 and C57Bl/6,

respectively). In vivo, the elimination half-lives of nicotine (1 mg/kg, s.c.) were also

similar between DBA/2 and C57Bl/6 mice (8.6 ± 0.4 vs. 9.2 ± 1.6 min, respectively);

however, cotinine levels were much higher in DBA/2 mice. The production and identity of

the putative cotinine metabolite 3’-hydroxycotinine in mice was confirmed by LC/MS/MS.

The in vivo half-life of cotinine (1 mg/kg, s.c.) was significantly longer in the DBA/2 mice

compared to the C57Bl/6 mice (50.2 ± 4.7 vs. 37.5 ± 9.6 min, respectively, p<0.05). The

in vitro metabolism of cotinine to 3’-hydroxycotinine was also less efficient in DBA/2 than

C57Bl/6 mice (Km: 51.0 ± 15.6 vs. 9.5 ± 2.1 µM, p<0.05; Vmax: 0.10 ± 0.01 vs. 0.04 ± 0.01

nmol/min/mg, p<0.05, respectively). Inhibitory antibody studies demonstrated that the

metabolism of both nicotine and cotinine was mediated by CYP2A5. Genetic differences

in Cyp2a5 potentially contributed to similar nicotine but different cotinine metabolism,

which may confound interpretation of nicotine pharmacological studies and studies

utilizing cotinine as a biomarker.

Page 67: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

52

Introduction

Nicotine is the primary component of cigarettes that is responsible for the addictive

properties of smoking, which include feelings of pleasure and reward (Henningfield and

Keenan, 1993). Rodents, in particular mice, have been widely used for studying the

pharmacological effects of nicotine (Aschhoff et al., 1999; Stolerman et al., 1999). Two

of the most commonly used mouse strains for studying nicotine behavioral effects are

the DBA/2 and C57Bl/6. A large number of studies have examined various aspects of

nicotine-mediated behaviors such as discrimination, self-administration, tolerance, and

withdrawal, and the majority of these studies have found some differences in nicotine

effects between these strains (Aschhoff et al., 1999; Stolerman et al., 1999).

The amino acid sequence of CYP2A5 is 84% identical to the human CYP2A6,

the main enzyme responsible for the metabolic inactivation of nicotine (Messina et al.,

1997; Nakajima et al., 1996b). The mouse Cyp2a5 gene is genetically polymorphic

(Lindberg et al., 1992). Specifically, the DBA/2 mice express the amino acid Val117 in

hepatic CYP2A5 and metabolized coumarin, a selective probe substrate for mouse

CYP2A5 and human CYP2A6, much more efficiently than C57Bl/6 mice, which express

the amino acid Ala117 (Lindberg et al., 1992). Similarly, mutagenesis of CYP2A6,

substituting the valine with alanine at the same position, also significantly reduced its

catalytic efficiency for coumarin (He et al., 2004b).

Genetic variation in human CYP2A6 can alter nicotine metabolism resulting in

altered smoking behaviors (Malaiyandi et al., 2006; Schoedel et al., 2004). For instance,

individuals who are homozygous for the CYP2A6 deletion variant (CYP2A6*4) produce

minimal cotinine (Yamanaka et al., 2004). These individuals smoke fewer cigarettes and

are less likely to be dependent on tobacco (Malaiyandi et al., 2006; Schoedel et al.,

2004). Likewise, in male mice we have previously shown that lower nicotine self-

Page 68: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

53

administration behaviors were associated with lower CYP2A5 protein levels and rates of

nicotine metabolism (Siu et al., 2006). Furthermore, inhibition of CYP2A5-mediated

nicotine metabolism significantly enhanced the pharmacological (i.e. anti-nociceptive)

effects of nicotine in mice (Damaj et al., 2006). These data together suggest that, as in

humans, nicotine metabolism can significantly affect nicotine-mediated behaviors in mice.

Therefore, the main objective of the study was to characterize nicotine and cotinine

metabolism (both in vitro and in vivo) in both the DBA/2 and C57Bl/6 mouse strains.

Such differences may account for the variations observed in the pharmacological effects

of nicotine in these mice.

Page 69: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

54

Materials and Methods

Animals

Adult male C57Bl/6 and DBA/2 mice (22-24g) were obtained from Charles River

Laboratories Inc. (Saint-Constant, PQ). Animals were housed in groups of three to four

on a 12-hour light cycle and had free access to food and water. We restricted the study

to male mice as we have previously found large variation in CYP2A levels and nicotine

metabolism among female mice (Siu et al., 2006). This may be due to hormonal

influences (mouse estrous cycle is approximately 3-7 days) as estrogen-only oral

contraceptives increased nicotine metabolism in human females (Benowitz et al., 2006a).

In addition a second CYP2A enzyme, CYP2A4, is present in female mice that may

metabolize nicotine (Murphy et al., 2005b) complicating the interpretation of our CYP2A5

studies.

Reagents

(-)-Nicotine hydrogen tartrate and (-)-cotinine were purchased from Sigma-Aldrich (St.

Louis, MO). Both nicotine and cotinine were dissolved in physiological saline (0.9%

sodium chloride) for use in in vivo studies. Trans-3’-hydroxycotinine was custom-made

by Toronto Research Chemicals Inc. (Toronto, ON). The internal standard 5-

methylcotinine was a generous gift from Dr. Peyton Jacob III at UCSF. All doses are

expressed as the free base of the drug. Inhibitory antibodies against human CYP2A6,

CYP2B6, and CYP2D6 were purchased from BD Biosciences (Mississauga, ON).

Membrane Preparations

Microsomal membranes were prepared from mouse livers for in vitro nicotine

metabolism assays as previously described (Messina et al., 1997; Siu et al., 2006) and

Page 70: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

55

stored at -80°C in 1.15% KCl. The cytosolic fractions were acquired during membrane

preparation and were used as a source of aldehyde oxidase. All livers were collected

and frozen prior to 3 pm to avoid circadian effects on CYP2A5 expression.

Nicotine C-Oxidation Assay

Prior to determining the in vitro kinetic parameters (Km and Vmax) for nicotine metabolism

in C57Bl/6 and DBA mice, assay conditions were optimized as previously described (Siu

et al., 2006). Linear formation of cotinine from nicotine was obtained under assay

conditions of 0.5 mg/ml protein concentration with an incubation time of 15 min.

Incubation mixtures contained 1 mM NADPH and 1 mg/ml of mouse liver cytosol in 50

mM Tris-HCl buffer, pH 7.4 and were performed at 37oC in a final volume of 0.5 ml. The

reaction was stopped with a final concentration of 4% v/v Na2CO3. After incubation 5-

methylcotinine (70 µg) was added as the internal standard and the samples were

prepared and analyzed for nicotine and metabolites by HPLC system I as described

previously (Siu et al., 2006). The limits of quantification were 5 ng/ml for nicotine, 12.5

ng/ml for cotinine, and 10 ng/ml for 3’-hydroxycotinine.

Cotinine Hydroxylation Assay

Prior to determining the in vitro kinetic parameters (Km and Vmax) of cotinine metabolism

in C57Bl/6 and DBA mice, assay conditions were optimized. Linear formation of 3’-

hydroxycotinine from cotinine was obtained under assay conditions of 1 mg/ml protein

with an incubation time of 20 min. The incubation mixture was the same as above with

the exception that aldehyde oxidase was not added as cotinine metabolism to 3’-

hydroxycotinine does not require this cytosolic enzyme. Samples were then analyzed by

HPLC system I.

Page 71: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

56

In Vivo Nicotine and Cotinine Treatments and Plasma Nicotine, Cotinine, and 3’-

Hydroxycotinine Measurements

To determine the in vivo kinetic parameters of nicotine and cotinine in C57Bl/6 and

DBA/2 mice, animals were injected with nicotine (1 mg/kg, s.c.) or cotinine (1 mg/kg,

s.c.). Blood samples were drawn by cardiac puncture at baseline from untreated animals

and from treated animals at various times after the injections. Immediately after

collection plasmas were prepared by centrifugation at 3000 x g for 10 min and frozen at

–20˚C until analysis. Sample collection took place prior to 3 pm. Total nicotine, cotinine,

and 3’-hydroxycotinine levels (free and glucuronides) were measured following

deconjugation by β-glucuronidase at a final concentration of 5 mg/ml in 0.2 M acetate

buffer, pH 5.0, at 37oC overnight. Samples were then analyzed by HPLC system I.

LC/MS/MS Analysis of Cotinine Metabolite

An alternative HPLC system (system II) suitable for separation of eluate for mass-

spectrometry was used for the characterization of the cotinine metabolite. This system

was similar to that previously described with minor modifications (Murphy et al., 1999).

Briefly, using the same column as HPLC system I, cotinine and its metabolites were

eluted with a linear gradient from 100% A’ (10 mM ammonium acetate buffer, pH 6.5) to

70% A’ and 30% acetonitrile over the course of 30 minutes at a flow rate of 1 ml/min.

Mass-spectrometry analysis was performed at the Proteomic and Mass

Spectrometry Centre at the University of Toronto (Toronto, ON). Data were acquired with

the Q TRAP LC/MS/MS System (Applied Biosystems/MDS Sciex, Toronto, ON). The

sample was injected into the sample loop and delivered to the mass spectrometer by

65% acetonitrile and 0.1% formic acid in water at 20 µl/min. Liquid chromatography

conditions were as described above (system II) except a flow rate of 0.8 ml/min was

Page 72: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

57

used. The liquid was introduced to the mass spectrometer directly after 40:1 splitting.

Electrospray ionization was performed in enhanced mass scan (EMS) mode with

positive ionization. Nitrogen was used as curtain gas (25 psi), nebulizer gas (25 psi), and

heater gas (0 psi). The spray needle voltage was set at 5.5KV and collision-induced

dissociation gas was set at high. The decluster potential was 20V, collision energy was

30eV, and entrance potential was 10V. Enhanced product ion (EPI) was performed at a

collision energy of 30eV, all other parameters were same as described for EMS.

Antibody Inhibition of Nicotine and Cotinine Metabolism

We have previously demonstrated that the anti-CYP2A6 antibody was able to cross-

react with mouse CYP2A5 (Siu et al., 2006). Microsomes were preincubated with anti-

human selective P450 antibodies (anti-CYP2A6, anti-CYP2B6, and anti-CYP2D6), at

concentrations of 0, 2, 40, and 80 µl antibodies per mg microsomal protein, for 15

minutes on ice according to manufacturer's instruction. Substrate concentrations used

represented the high-affinity Km value concentrations for nicotine and cotinine

metabolism, specifically 11 µM for nicotine and 51 µM for cotinine for DBA/2 mice

microsomes and 11 µM for nicotine and 9.5 µM for cotinine for C57Bl/6 mice

microsomes.

In Vitro Kinetic and Pharmacokinetic Parameters Analyses

The Michaelis-Menten kinetic parameters Km and Vmax from in vitro metabolism studies

were calculated using Graphpad Prism (Graphpad Software Inc., San Diego, CA) and

were verified by the Eadee-Hofstee method. The equation used to determine Km and

Vmax for one and two enzymatic sites were v = Vmax [S] / (Km + [S]) and v = [Vmax1 [S] /

Page 73: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

58

(Km1 + [S])] + [Vmax2 [S] / (Km2 + [S])], respectively, where [S] denotes substrate

concentration.

The in vivo pharmacokinetic parameters were determined using non-

compartmental analysis: AUC0-480, peak plasma concentration (Cmax), maximum plasma

concentration (Tmax). AUC0-480 was calculated using the trapezoidal rule. Elimination half-

life (T1/2) was estimated by the terminal slope. Since the bioavailabilities (F) of nicotine

and cotinine were unknown following subcutaneous injection in mice, CL (clearance)

was determined as a hybrid parameter CL/F and was calculated as Dose/AUC0-480. The

average weights of the animals of the strains were similar (24.8 ± 1.7 vs. 25.5 ± 1.1 g for

DBA/2 and C57Bl/6, respectively, n=50 for each strain), therefore the dose of 25 µg (1

mg/kg) was used for the calculation of CL/F for nicotine.

Statistical Analyses

Statistical analyses of in vitro kinetic parameters were tested by Mann-Whitney U test.

Assessment of in vivo nicotine, cotinine and 3’-hydroxycotinine plasma levels for the

entire time course was not possible from individual animals due to limited blood volume,

therefore each time point represented data from multiple mice. Due to this experimental

design pharmacokinetic parameters (e.g. half life) were estimated by resampling

methods using the PKRandTest software (H. L. Kaplan, Toronto, ON) (Damaj et al.,

2006).

Page 74: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

59

Results

In Vitro Nicotine C-Oxidation in DBA/2 and C57Bl/6 Mice

We first assessed the in vitro kinetic parameters of nicotine C-oxidation in hepatic

microsomes prepared from DBA/2 and C57Bl/6 mice. Nicotine metabolism to cotinine,

demonstrated with Michaelis-Menten kinetics (Fig. 2A) and Eadee-Hofstee plotting (Fig.

2B), revealed two enzymatic sites in both strains. The high-affinity sites for hepatic

microsomes from both DBA/2 and C57Bl/6 showed similar Km, Vmax, and Vmax/Km values

for nicotine (Table 1). In contrast, the low-affinity enzymatic sites exhibited differing, and

much lower nicotine metabolic activities with a slight non-significantly higher Km for

C57Bl/6 and Vmax for DBA/2. The Vmax/Km value was significantly higher for the DBA/2

microsomes (Table 1).

Characterization of In Vivo Nicotine Metabolism in DBA/2 and C57Bl/6 Mice

Since the rate of drug metabolism in vitro does not necessarily reflect drug clearance in

vivo (e.g. presence of non-hepatic elimination processes), we determined whether the in

vivo clearance of nicotine was similar between the two mouse strains. Adult male mice

from both strains were treated with 1 mg/kg subcutaneous nicotine, a dose used

previously in nicotine behavioral studies (Zarrindast et al., 2003). In both mouse strains,

nicotine concentrations peaked at 10 minutes with DBA/2 mice having a significantly

greater maximum concentration compared to C57Bl/6 mice (Fig. 3A; Table 2). The

overall AUC0-480 of nicotine was also modestly higher for DBA/2 than for C57Bl/6 mice.

Both strains had similar elimination half-lives for nicotine but the clearance of nicotine

was slower in the DBA/2 mice compared to C57Bl/6 mice.

Page 75: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

60

Figure 2. In vitro metabolism of nicotine to cotinine in DBA/2 and C57Bl/6 mice. In vitro kinetic parameters of nicotine metabolism were investigated using hepatic microsomes. A) Rate of formation of cotinine from nicotine (NIC) (mean ± SD of 4 to 5 mice). B) Eadee-Hofstee plots of the kinetic data for DBA and C57Bl/6 mice. See Table 1 for values of kinetics parameters. V: velocity; S: substrate

Page 76: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

61

Table 1. In vitro nicotine metabolism (C-oxidation) parameters.

DBA/2 (n=5 mice) mean ± S.D.

C57Bl/6 (n=4 mice) mean ± S.D.

High Affinity Km (µM) 10.7 ± 4.8 11.4 ± 3.6

Vmax (nmol/min/mg) 0.58 ± 0.18 0.50 ± 0.07 Vmax/Km 0.05 ± 0.03 0.04 ± 0.02 Low Affinity Km (µM) 234 ± 77 306 ±126

Vmax (nmol/min/mg) 1.60 ± 0.63 1.18 ± 0.36 Vmax/Km 0.007 ± 0.001* 0.004 ± 0.001 *p < 0.05 compared to C57Bl/6 mice

Page 77: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

62

Figure 3. Plasma nicotine and cotinine concentrations following nicotine treatment. Adult male DBA/2 and C57Bl/6 mice were injected with subcutaneous nicotine (1 mg/kg). A) The AUC0-480 of nicotine (NIC) was 15% higher while the elimination half-life was similar in the DBA/2 mice compared to C57Bl/6 mice. B) The AUC0-480 of cotinine (COT) in the DBA/2 mice was twice that of the C57Bl/6 mice. Each time point represents mean (± SD) of 5 to 7 animals for each strain. Values below the limit of quantification are not shown.

Page 78: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

63

Table 2. Pharmacokinetic parameters of plasma nicotine and cotinine in mice treated with nicotine (1 mg/kg, s.c.). Results are derived using data from 5 to 7 animals at each time point.

Nicotine Cotinine DBA/2 C57Bl/6 DBA/2 C57Bl/6

AUC0-480 (ng•hr/ml) 92.9 ± 2.9* 80.8 ± 3.2 245 ± 10* 120 ± 7 T1/2 (min) 8.6 ± 0.4 9.2 ± 1.6 51.0 ± 4.1* 23.7 ± 2.0 CL/F (ml/min) 4.5 ± 0.1* 5.2 ± 0.2 ND ND Tmax (min) 10.0 ± 2.2 10.0 ± 4.7 45.0 ± 13.3* 20.0 ± 13.3 Cmax (ng/min) 201 ± 15* 160 ± 15 141 ± 7* 112 ± 12 *p < 0.05 compared to C57Bl/6 mice, ND – not determined

Page 79: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

64

When examining the disposition kinetics of cotinine formed from injected nicotine, we

observed that the appearance of the cotinine metabolite was rapid and similar between

the two mouse strains and achieving peak concentrations around 15 minutes (Fig 3B). In

contrast, compared to C57Bl/6 mice, DBA/2 mice showed a significantly larger AUC0-480

and longer elimination half-life (Table 2).

LC/MS/MS Characterization of the Putative Cotinine Metabolite 3’-Hydroxycotinine

In humans cotinine is metabolized exclusively to trans-3’-hydroxycotinine by CYP2A6

(Dempsey et al., 2004; Nakajima et al., 1996a; Yamanaka et al., 2004). To our

knowledge no prior studies have examined or confirmed the production of trans-3’-

hydroxycotinine from cotinine in mice, therefore our immediate goal was to determine

whether mice metabolize cotinine to 3’-hydroxycotinine. In a preliminary in vitro study we

identified a cotinine metabolite that displayed the same retention time as the trans-3’-

hydroxycotinine standard (Fig. 4A). To confirm the identity of the putative trans-3’-

hydroxcotinine compound, a second HPLC system compatible with MS/MS analysis was

used. Both the trans-3’-hydroxycotinine standard and the cotinine metabolite eluted with

the same retention time (10.4 min) using the new HPLC system (Fig. 4B).

In the LC/MS/MS analyses, EMS indicated that the trans 3’-hydroxycotinine

standard had a retention time of 11.8 minutes with a m/z of 193 (Fig. 4C) and when the

metabolite of in vitro cotinine metabolism was monitored at m/z 193, a major peak was

present at 11.1 minutes (Fig. 4D). Fragmentation of the trans-3’-hydroxycotinine

standard ion (m/z 193) gave two fragments of m/z 80 and m/z 134 (Fig. 4E).

Fragmentation of the cotinine metabolite at m/z 193 also gave two major ions of m/z 80

and m/z 134 (Fig. 4F). The peak area ratios of m/z 80/134 for the trans-3’-

hydroxycotinine and the cotinine metabolite were 3.44 and 3.41, respectively. The m/z

Page 80: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

65

Figure 4. Characterization of 3’-hydroxycotinine in mice by HPLC and LC/MS/MS. A) Cotinine and its metabolite (COT metabolite) from a pilot in vitro cotinine metabolism study using mouse liver microsomes (DBA/2) were separated with HPLC system I in the absence (left) or presence (right) of the trans-3’-hydroxycotinine standard (3-HC). B) Cotinine and trans-3’-hydroxycotinine standards (left) as well as products from a pilot in vitro cotinine metabolism study (right) were separated with HPLC system II. C) Enhanced mass scan at m/z 193 showed that the trans-3’-hydroxycotinine standard had a retention time of 11.8 minute. D) Liquid chromatography of products from in vitro cotinine metabolism monitored at m/z 193 identified a putative 3’-hydroxycotinine compound with a retention time of 11.1 minutes. E) Fragmentation of the m/z 193 ion (from C) at 30eV produced m/z 80 and m/z 134 ions. F) Fragmentation of the m/z 193 ion (from D) at 11.1 minutes at 30eV produced m/z 80 and m/z 134 ions.

Page 81: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

66

80 and the m/z 134 fragments corresponded to (C5H5N)H+ and pyridyl-C3H4O+,

respectively (Murphy et al., 1999).

Characterization of In Vivo Cotinine Metabolism in DBA/2 and C57Bl/6 Mice

Having confirmed that 3’-hydroxycotinine is produced from cotinine in mice, we

proceeded with in vivo injections of cotinine (1 mg/kg, s.c.). Plasma cotinine

concentrations were maximal between 5 to 15 minutes and were similar for both DBA/2

and C57Bl/6 mice (Fig. 5A, Table 3). Similar to cotinine derived from nicotine injection,

following cotinine injection the cotinine AUC0-180 was much higher in the DBA/2 mice

compared to the C57Bl/6 mice. The clearance of cotinine was slower in DBA/2 mice,

which resulted in longer elimination half-life of cotinine compared to C57Bl/6 mice.

The plasma levels of 3’-hydroxycotinine formed from cotinine injections were also

monitored. The plasma AUC0-180 of 3’-hydroxycotinine was higher in the DBA/2 mice

compared to the C57Bl/6 mice (Fig. 5B, Table 3).

In Vitro Cotinine Metabolism in DBA/2 and C57Bl/6 Mice

To determine whether cotinine was metabolized to 3’-hydroxycotinine differently

between the two mouse strains, accounting for the differences in cotinine plasma

concentrations seen in vivo, we performed in vitro cotinine metabolism studies. We

found that cotinine metabolism to 3’-hydroxycotinine was characterized by Michaelis-

Menten kinetics (Fig. 6A), mediated by a single enzymatic site in both strains (Fig. 6B).

The DBA/2 mice had a significantly higher Km compared to the C57Bl/6 mice (Table 4),

while the Vmax for cotinine was much greater for DBA/2 than for C57Bl/6 mice. This

resulted in an overall lower catalytic efficiency (Vmax/Km) for DBA/2 compared to C57Bl/6.

Page 82: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

67

Figure 5. Plasma cotinine and 3’-hydroxycotinine concentrations following cotinine injections. Adult male DBA/2 and C57Bl/6 mice were injected with subcutaneous cotinine (1 mg/kg). A) The AUC0-180 of cotinine (COT) was 37% higher and its elimination half-life was 34% longer in DBA/2 mice compared C57Bl/6 mice. B) The AUC0-180 of 3’-hydroxycotinine (3-HC) was 94% higher in the DBA/2 mice compared to C57Bl/6 mice. Each time point represents mean ± SD of 3 to 8 animals for each strain.

Page 83: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

68

Table 3. Pharmacokinetic parameters of plasma cotinine and 3’-hydroxycotinine in mice treated with cotinine (1 mg/kg, s.c.). Results are derived using data from 3 to 8 animals at each time point.

Cotinine 3’-hydroxycotinine DBA/2 C57Bl/6 DBA/2 C57Bl/6

AUC0-180 (ng•hr/ml) 1087 ± 33* 796 ± 33 518 ± 30* 268 ± 19 T1/2 (min) 50.2 ± 4.7* 37.5 ± 9.6 ND ND CL/F (ml/min) 0.38 ± 0.01* 0.52 ± 0.02 ND ND Tmax (min) 15.0 ± 7.5 15.0 ± 0.0 60.0 ± 0.0 60.0 ± 23.6 Cmax (ng/ml) 748 ± 35 759 ± 39 255 ± 23* 110 ± 10 *p < 0.05 compared to C57Bl/6, ND – not determined

Page 84: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

69

Figure 6. In vitro metabolism of cotinine to 3’-hydroxycotinine in DBA/2 and C57Bl/6 mice. In vitro kinetic parameters of cotinine metabolism were investigated using hepatic microsomes. A) Rate of formation of 3’-hydroxycotinine from cotinine (mean ± SD; n=4 samples per strain). DBA/2 and C57Bl/6 mice. Regression lines and intercepts for both DBA/2 and C57Bl/6 were calculated from Km and Vmax values derived from Michaelis-Menten fitting and no data points were censored. See Table 4 for values of kinetic parameters. V: velocity; S: substrate

Page 85: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

70

Table 4. In vitro cotinine hydroxylation parameters.

DBA/2 (n=4) mean ± S.D.

C57Bl/6 (n=4) mean ± S.D.

Km (µM) 51.0 ± 15.6* 9.5 ±2.1

Vmax (nmol/min/mg) 0.10 ± 0.01* 0.04 ± 0.01 Vmax/Km 0.002 ± 0.000* 0.005 ± 0.002 *p < 0.05 compared to C57Bl/6 mice

Page 86: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

71

Inhibition of In Vitro Nicotine and Cotinine Metabolism

Previously the mouse CYP2A5 was identified as the enzyme responsible for the high-

affinity metabolism of nicotine using cDNA-expressed CYP2A5 (Murphy et al., 2005b).

To extend these studies characterizing the enzyme involved, we tested the effect of

inhibitory antibodies on in vitro nicotine metabolism. Anti-CYP2A6 inhibitory antibodies

dose-dependently inhibited the formation of cotinine from nicotine in DBA/2 microsomes

with maximal inhibition of 70% at 40 µl antibody/mg microsomal protein (Fig. 7A, filled

symbols). Similar results were seen in hepatic microsomes from C57Bl/6 mice, tested at

80 µl antibody / mg protein (Fig. 7A, open symbols). Inhibitory antibodies against

CYP2B6 and CYP2D6, enzymes postulated to be involved in the remaining small

percentage of metabolism of nicotine in humans (Messina et al., 1997; Nakajima et al.,

1996b; Yamazaki et al., 1999), did not inhibit nicotine metabolism in either mouse strain

(Fig. 7A).

In humans, CYP2A6 is exclusively responsible for the metabolism of cotinine to

trans-3’-hydroxycotinine (Dempsey et al., 2004; Nakajima et al., 1996a). To determine

whether mouse CYP2A5 was also responsible for this metabolic pathway we performed

inhibition studies on cotinine metabolism. Anti-CYP2A6 inhibitory antibodies dose-

dependently inhibited the formation of 3’-hydroxycotinine from cotinine in DBA/2 mouse

liver microsomes (Fig. 7B, filled symbols), with a maximal inhibition of 90% at 40 µl

antibody / mg microsomal protein. Anti-CYP2A6 inhibitory antibodies also inhibited

cotinine metabolism (no detectable metabolite peak) in C57Bl/6 hepatic microsomes (Fig.

7B, open symbols). Inhibitory antibodies against CYP2B6 and CYP2D6 had no effect on

cotinine metabolism (Fig. 7B).

Page 87: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

72

Figure 7. Antibody inhibits in vitro nicotine and cotinine metabolism in DBA/2 and

C57Bl/6 mice. A) At Km (11 µM) nicotine, anti-CYP2A6 inhibitory antibodies inhibited cotinine formation maximally by 70% in hepatic microsomes in DBA/2 mice (filled

symbols). At Km (11 µM) nicotine, anti-CYP2A6 inhibitory antibodies (80 µl antibodies / mg protein) inhibited cotinine formation by 70% in C57Bl/6 hepatic microsomes (open

symbols). B) At Km (51 µM) cotinine, anti-CYP2A6 inhibitory antibodies inhibited 3’-hydroxycotinine formation by >90% in hepatic microsomes DBA/2 mice (filled symbols).

At Km (9.5 µM) cotinine, anti-CYP2A6 inhibitory antibodies inhibited 3’-hydroxycotinine formation in C57Bl/6 hepatic microsomes (open symbols). *Activity of cotinine metabolism was below limit of quantification. Anti-CYP2B6 and anti-CYP2D6 antibodies had no effect on nicotine or cotinine metabolism. Values are expressed as a percent of activity in the absence of antibodies and were determined from pooled sample of 2 animals.

Page 88: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

73

Discussion

In the present study we examined both in vitro metabolism and in vivo pharmacokinetics

of nicotine and cotinine in two commonly used inbred mouse strains, DBA/2 and C57Bl/6.

Mice have been used extensively for the study of nicotine behaviors; however, the

effects of nicotine vary widely between strains (Aschhoff et al., 1999; Stolerman et al.,

1999). One potential contributing factor may be differences in nicotine pharmacokinetics

which can significantly alter its pharmacology. CYP2A5 metabolizes nicotine (Murphy et

al., 2005b) and previous studies found that DBA/2 and C57Bl/6 mice differed in CYP2A5

structure and function (Lindberg et al., 1992). Therefore the primary goal of this study

was to determine whether nicotine and cotinine metabolism differed between these two

strains, which may be contributed by the Cyp2a5 polymorphism between these mice

(Lindberg et al., 1992).

The in vitro metabolism of nicotine to cotinine was mediated by a high- and a

low-affinity enzyme site in both mouse strains. The Km values for the high-affinity sites

reported here are consistent with those seen using cDNA-expressed CYP2A5 (7.7 ± 0.8

µM; 129/J mouse strain) (Murphy et al., 2005b) but are modestly higher affinity relative

to hepatic microsomes from ICR mice (18.6 ± 5.9 µM) (Damaj et al., 2006). The identity

of the high-affinity site was confirmed, as CYP2A5 inhibitory antibodies inhibited up to

70% of nicotine metabolism at Km for nicotine in both strains. The low-affinity sites in our

mice could potentially belong to the 2B family. In humans, cDNA-expressed CYP2B6

metabolizes nicotine but with much lower affinity and activity compared to CYP2A6

(Yamazaki et al., 1999). In monkeys, CYP2B6agm is a minor enzyme compared to

CYP2A6agm for the metabolism of nicotine to cotinine (Schoedel et al., 2003). In

contrast, rat CYP2B1/2 is the primary enzyme responsible for this process (Nakayama et

al., 1993). In our experiments, however, no indication of inhibition of nicotine metabolism

Page 89: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

74

was seen at the highest CYP2B antibody concentration tested. Considering that at the

highest plasma nicotine concentrations observed (~200 ng/ml ≅ 1.2 µM) following

nicotine injection the estimated contribution of the low-affinity enzymes to nicotine

metabolism was only ≤10%, the identity of this enzyme was not pursued.

To determine if nicotine was metabolized similarly between DBA/2 and C57Bl/6

mice in vivo we administered nicotine subcutaneously as this route kinetically mimics,

somewhat, the route of nicotine intake from smoking in that it bypasses first-pass

metabolism. The in vivo kinetics of nicotine in these two strains differed though not

dramatically. The higher nicotine Cmax in DBA/2 mice may be due to a smaller volume of

distribution of nicotine: male DBA/2 mice have, on average, 34-40% more body fat and

10% lower lean mass compared to C57Bl/6 mice (Mouse Phenome Database, Jackson

Labs) which could result in higher levels of nicotine in the plasma (and other highly

perfused organs such as liver, kidneys, and the lung) (Urakawa et al., 1994). Despite

similar nicotine clearance, however, we found that cotinine was removed more slowly in

DBA/2 mice as demonstrated by the two-fold longer elimination half-life and higher

cotinine AUC0-480.

In humans, the main metabolites of cotinine recovered in urine are trans-3’-

hydroxycotinine and its glucuronide which account for 40-60% of the total administered

dose of nicotine (Hukkanen et al., 2005b). Initially we demonstrated that mice produced

3’-hydroxycotinine from cotinine with LC/MS/MS. We then confirmed that the metabolism

of cotinine to 3’-hydroxycotinine was mediated by CYP2A5 – up to 90% of cotinine

metabolism to 3’-hydroxycotinine was inhibited by anti-CYP2A6 inhibitory antibodies.

This is consistent with the metabolism of cotinine to 3’-hydroxycotinine being mediated

exclusively by human CYP2A6 in vitro and in vivo (Dempsey et al., 2004; Nakajima et al.,

1996a). Following cotinine injections, DBA/2 mice showed slower clearance of cotinine

Page 90: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

75

compared to C57Bl/6 mice, which was consistent with the pharmacokinetics of cotinine

formed from nicotine injections. It is likely that DBA/2 mice have a slower hepatic

(intrinsic) clearance of cotinine to 3’-hydroxycotinine compared to C57Bl/6; this is

supported by our in vitro findings that cotinine was metabolized to 3’-hydroxycotinine

significantly slower in the DBA/2 compared to C57Bl/6 mice. Even at the maximum

plasma cotinine concentrations observed (~760 ng/ml ≅ 4.3 µM) our in vitro data

indicated that the DBA/2 mice metabolized cotinine slower than the C57Bl/6 mice (v =

0.008 vs. 0.0012 nmol/min/mg, respectively).

When examining the plasma concentrations of 3’-hydroxycotinine formed from

cotinine, we found that DBA/2 mice had a larger AUC0-180 compared to C57Bl/6 mice.

The higher level of 3’-hydroxycotinine in DBA/2 mice was likely to be due to reduced

elimination of 3’-hydroxycotinine. This could occur through slower rates of conjugation to

O-glucuronide and/or slower renal excretion of 3’-hydroxycotinine and its glucuronidated

metabolite. In mice, N-glucuronides of nicotine and its proximal metabolites have not

been detected nor identified (although O-glucuronides were not measured) (Ghosheh

and Hawes, 2002), while in humans, ~80% of trans-3’-hydroxycotinine is excreted

unchanged (Hukkanen et al., 2005b). Thus we believe the higher level of 3’-

hydroxycotinine was most likely due to slower renal excretion in the DBA/2 mice.

This study showed the in vitro and in vivo metabolism of nicotine was similar

between DBA/2 and C57Bl/6 mice. In contrast to nicotine, DBA/2 mice metabolized

cotinine to 3’-hydroxycotinine with lower efficiency compared to C57Bl/6 mice both in

vitro and in vivo. These data indicate that genetic differences in the structure of CYP2A5

between the two strains can potentially alter the rate of metabolism depending on the

specific substrate. The DBA/2 CYP2A5, which has valine at position 117, is more

efficient at metabolizing coumarin compared to C57Bl/6, which has an alanine in this

position (Lindberg et al., 1992; van Iersel et al., 1994). In contrast, this genetic variant

Page 91: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

76

does not appear to alter the metabolism of nicotine, but rather, it reduces the catalytic

activity for cotinine. Mouse CYP2A5 oxidizes nicotine to ∆5’(1’)-iminium ion followed by

conversion to cotinine by aldehyde oxidase (Murphy et al., 2005b); however, no

functional polymorphisms for the mouse aldehyde oxidase genes have been reported

(Mouse Genome Informatics), and our in vivo data was consistent with our in vitro data

suggesting the variations in aldehyde oxidases were not contributing substantively to our

observations. Substrate-selective metabolisms by genetic variants of human CYP2A6

have been observed. For example, the CYP2A6*7 variant has reduced nicotine

metabolic activity, but coumarin metabolism was minimally affected (Ariyoshi et al.,

2001). In addition, different levels of cotinine following similar nicotine intake have been

observed in smokers (Benowitz et al., 1999), and this may be partly related to genetic

variations in CYP2A6 that have minor impacts on the metabolism of nicotine relative to

the impact on cotinine metabolism. Finally, these observations warrant further studies on

the metabolic activation of CYP2A5/6 substrates such as NNK, a tobacco-specific

nitrosamine known to cause lung cancer (Miyazaki et al., 2005), and the consequence of

these genetic variants on NNK activation. Future studies will focus on the expression of

variants V117A (found in CYP2A5 [Lindberg et al., 1992] and CYP2A13 [NCBI]) as well

as F118L (found in CYP2A6 [NCBI]), in all three enzymes and their impact on multiple

substrates including NNK.

The finding that cotinine is differentially metabolized relative to nicotine, between

strains, has implications with respect to interpreting nicotine pharmacological studies.

Cotinine can pass through the blood brain barrier (Lockman et al., 2005). In rats, cotinine

can bind to epibatidine-sensitive nicotinic receptors in frontal cortex and hippocampus

tissues, although with lower affinity than nicotine (Vainio and Tuominen, 2001).

Furthermore, administration of cotinine in rat striatal tissues evoked dopamine overflow

in a dose-dependent manner (Dwoskin et al., 1999). Thus, differing levels of cotinine

Page 92: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

77

may result in altered pharmacological effects despite similar nicotine levels between the

two mouse strains, and the effects may be erroneously attributed to nicotine

pharmacology. On the other hand, the effects of cotinine in humans are less clear

(Buccafusco and Terry, 2003; Crooks and Dwoskin, 1997). Cotinine alone did not show

pharmacological effects, but it did interfere with the ability of nicotine patch to reduce

withdrawal symptoms (Hatsukami et al., 1998b). Cotinine also increased plasma nicotine

levels in smokers, possibly through increased smoking to compensate for the

interference of nicotine action by cotinine (Hatsukami et al., 1998a). In other studies

cotinine appeared to have some effects in reducing withdrawal symptoms (Benowitz et

al., 1983; Keenan et al., 1994).

Consideration should also be taken when using cotinine as a biomarker for

environmental tobacco smoke exposure (Benowitz, 1999) and cigarette intake (de Leon

et al., 2002) in humans, or nicotine consumption in mice (Sparks and Pauly, 1999), as

cotinine can be metabolized at different rates by human CYP2A6 and mouse CYP2A5.

Differing levels of cotinine could be erroneously interpreted as different levels of

exposure rather than differing rates of removal.

In conclusion, we have characterized nicotine and cotinine metabolism in two

mouse strains which differed in CYP2A5 enzyme structure (Lindberg et al., 1992). While

coumarin metabolism differed (Lindberg et al., 1992) we observed no substantial

difference in nicotine metabolism. In contrast, CYP2A5-mediated cotinine metabolism to

3’-hydroxycotinine was different between the mouse strains, which may confound

interpretation of pharmacological and biomarker studies.

Page 93: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

78

Acknowledgments

We would like to thank Linda Liu for assistance in HPLC and Dr. Sharon Miksys and Jill

Mwenifumbo for critical review of this manuscript. RFT is a shareholder and chief

scientific officer of Nicogen Inc., a company focused on the development of novel

smoking cessation therapies; no funds were received from Nicogen for these studies

and this manuscript was not reviewed by other people associated with Nicogen prior to

submission or revision.

Footnote

This study was supported by CAMH and CIHR grant # MOP 53248. CIHR-Special

Training Program in Tobacco Control and OGS to ECKS and Canada Research Chair in

Pharmacogenetics to RFT.

Page 94: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

79

SIGNIFICANCE OF CHAPTER

This study described the in vitro and in vivo metabolism of both the nicotine and cotinine

in mice. Similar to human CYP2A6, the mouse CYP2A5 is responsible for the majority of

metabolism of nicotine to cotinine, and it appears to be exclusively responsible for the

metabolism of cotinine to 3-HC. This provided a line of evidence that the mouse may be

a suitable model for studying inhibitors of nicotine metabolism. In addition, the study also

showed that polymorphism in CYP2A5 can affect cotinine metabolism differentially

compared with nicotine. This suggested that careful interpretations are required in

nicotine pharmacology studies as well as in studies where cotinine are used as

biomarkers.

Page 95: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

80

CHAPTER 3: NICOTINE SELF-ADMINISTRATION IN MICE IS ASSOCIATED WITH RATES OF NICOTINE INACTIVATION BY CYP2A5

Eric C. K. Siu, Dieter B. Wildenauer, and Rachel F. Tyndale As published in: Psychopharmacology, 2006, Mar; 184(3-4):401-408.

The immunoblotting and in vitro enzyme activities assays, data analyses, and writing of

the manuscript were performed by ECKS. The nicotine self-administration study was

performed by DB Wildenauer. The study was designed by ECK Siu and RF Tyndale.

Reprinted from Pyschopharmacology, Volume 184, Siu et al., “NICOTINE SELF-ADMINISTRATION IN MICE IS ASSOCIATED WITH RATES OF NICOTINE INACTIVATION BY CYP2A5”, pp 401-408, Copyright (2006) with permission from Springer-Verlag. © 2006 Springer-Verlag

Page 96: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

81

Abstract

Rationale: Cyp2a5, the mouse homologue of human CYP2A6, encodes for the enzyme

responsible for the primary metabolism of nicotine. Variation in human CYP2A6 activity

can alter the amount smoked such as numbers of cigarettes smoked per day and

smoking intensity. Different mouse strains self-administer different amounts of oral

nicotine and quantitative trait loci analyses in mice suggested that Cyp2a5 may be

involved in differential nicotine consumption behaviours. Objectives: The goal of this

study was to examine whether in vivo nicotine consumption levels were associated with

CYP2A5 protein levels and in vitro nicotine metabolism in mice. Methods: F2 mice

propagated from high (C57Bl/6) and low (St/bJ) nicotine consuming mice were analyzed

for CYP2A5 hepatic protein levels and in vitro nicotine metabolizing activity. Results:

We found that F2 male high nicotine (n=8; 25.1±1.2 µg nicotine/day) consumers had

more CYP2A5 protein compared to low (n=11; 3.8±1.4 µg nicotine/day) consumers

(10.2±1.0 vs. 6.5±1.3 CYP2A5 units). High consumers also metabolized nicotine faster

than the low consumers (6 µM: 0.18±0.04 vs. 0.14±0.07; 30 µM: 0.36±0.06 vs.

0.26±0.13; 60 µM: 0.49±0.05 vs. 0.32±0.17 nmol/min/mg). In contrast, female high

(25.1±2.1 µg nicotine/day) and low (4.7±1.4 µg nicotine/day) nicotine consumers did not

show pronounced differences in nicotine metabolism or CYP2A4/5 protein levels; this is

consistent with other studies of sex differences in response to nicotine. Conclusions:

These data suggested that among male F2 mice, increased nicotine self-administration

is associated with increased rates of nicotine metabolism, most likely as a result of

greater CYP2A5 protein levels.

Page 97: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

82

Introduction

Nicotine is the main constituent of tobacco responsible for the addictive property

of cigarettes (Stolerman and Jarvis, 1995). Once dependent, smokers adjust their

smoking behaviours (e.g. number of cigarettes smoked, puff intensity) to maintain

constant nicotine plasma levels (McMorrow and Foxx, 1983; Russel, 1987). As a

consequence, factors affecting nicotine clearance alter an individual’s smoking

behaviours. In humans approximately 80% of nicotine is metabolically inactivated to

cotinine by the liver and approximately 90% of this reaction is carried out by CYP2A6

(Benowitz et al., 1994; Messina et al., 1997; Nakajima et al., 1996a). Individuals with

different reduced activity variants of CYP2A6 smoke lower numbers of cigarettes; for

instance, smokers who are slow metabolic inactivators of nicotine smoke fewer

cigarettes compared to those who are normal or faster metabolizers (Rao et al., 2000;

Schoedel et al., 2004). To further investigate the impact of nicotine metabolism on

nicotine consumption, a genetic and kinetic study in mice was initiated.

The mouse is a better suited model for the study of nicotine metabolism and

behaviours compared to rats for several reasons. First, the predominant enzymes

responsible for the metabolism of nicotine in rats belong to the CYP2B family

(Nakayama et al., 1993; Schoedel et al., 2001). Second, CYP2A5 is approximately 84%

identical to CYP2A6 in amino acid sequence, and cDNA-expressed CYP2A5

metabolizes nicotine with similar efficiency to human CYP2A6 (Murphy et al., 2005a).

Finally, the close structural similarity between CYP2A5 and CYP2A6 allows for the

development of inhibitors and in vivo screening of these compounds in mice. In addition,

the behavioural and physiological effects of nicotine in mice have been well documented

(Collins et al., 1988; Faraday et al., 2005; Grabus et al., 2005; Marks et al., 1985).

Page 98: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

83

With respect to nicotine intake, it has been shown that different mouse strains self-

administer different amounts of oral nicotine (Aschhoff et al., 1999; Robinson et al.,

1996), and a variety of mechanisms may explain the differential preferences. These

include differences in the rewarding properties of nicotine and taste preference for bitter

compounds (i.e. nicotine) (Lush and Holland, 1988; Merlo Pich et al., 1999).

Acetylcholine receptor density and receptor affinity for nicotine could also potentially

alter nicotine preference (Marks et al., 1989; Marks et al., 1986). However, there are no

studies examining the effect of differing rates of nicotine metabolism on oral nicotine

self-administration in mice. Recently, a quantitative trait loci study revealed that the

Cyp2a5 gene is situated close to a marker on Chromosome 7 associated with differential

nicotine self-administration behaviours, suggesting a possible involvement of Cyp2a5 in

nicotine consumption behaviours in mice (Aschhoff et al., 1999; Wildenauer et al., 2005).

The goal of this study was to determine whether CYP2A5 protein levels, and resulting

nicotine metabolism, differed between the high and low nicotine consuming mice

examined in the QTL study. The mice were selected from the F2 generation and would

allow us to address the relationship between gene function and behaviour.

Page 99: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

84

Materials and Methods

Animals

The F2 mice were established by crossing the F1 mice generated from a high

nicotine-consuming strain (C57Bl/6) and a low-nicotine consuming strain (St/bJ)

(Aschhoff et al., 1999). The mice were housed individually in polycarbonate cages with

wood-shaving bedding. Each cage contained two graduated bottles with one on the left

and one on the right side of each cage cover. The bottles were filled with tap water and

total consumption from each bottle was estimated from the decrease of the fluid cylinder

along the graduation of the tube every day during a 10 day acclimation period. On day

11 one bottle was filled with tap water, containing 5 µg/ml nicotine tartrate, while the

other one contained tap water only. Subsequently, nicotine concentrations were raised

to 10, 25, 50, 75 and 100 µg/ml at 5-day intervals. To minimize problems associated with

side preference, the position of the bottles at the right or left side of the cage was

changed every day after measuring consumption. During the experiment, animals were

fed ad libitum (24-hour free-access). There were a total of 485 F2 mice, of which the 46

lowest nicotine-consuming mice and the 47 highest nicotine-consuming mice were

selected for genotyping and QTL analyses (Wildenauer et al., 2005). From these we

analyzed the highest and lowest nicotine-consuming mice (males and females). High

nicotine consumers (HNC, n=22) and low nicotine consumers (LNC, n=23) were

selected based on the highest and lowest average amount of nicotine consumed over

the testing period, respectively. Animals were sacrificed after the last exposure at the

beginning of the light cycle at the end of the test. Following sacrifice the livers were

removed and frozen in liquid nitrogen and transferred to -80oC for storage.

Page 100: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

85

Membrane Preparations Microsomal membranes were prepared from mouse livers for the in vitro nicotine

metabolism assay and stored at 80°C in 1.15% KCl as previously described (Messina

et al., 1997). Briefly, the livers were homogenized in 1.15% KCl buffer and centrifuged at

9,000 g for 20 minutes. The supernatant fractions were collected and centrifuged again

at 100,000 g for 90 minutes. The resultant microsomal pellets were resuspended in

1.15% KCl buffer for in vitro NCO assay. The cytosolic fractions (supernatant) were also

collected from each of the membrane preparations and were pooled for used as the

common source of aldehyde oxidase. For immunoblotting, microsomal membranes were

prepared as above with the use of 100 mM Tris, pH 7.4, 0.1 mM EDTA, 0.1 mM

dithiothreitol, 1.15% (w/v) KCl, and 20% (v/v) glycerol. Protein concentrations were

determined with the Bradford reagent according to manufacturer's protocol (Bio-Rad

Labs Ltd., Mississauga, ON).

Immunoblotting Immunoblotting was performed essentially as previously described (Schoedel et al.,

2003). To determine the linear range of CYP2A5 detection for the immunoblotting assay

mouse liver microsomes were serially diluted and used to construct standard curves

(0.05 to 12.8 µg protein). Membrane proteins from livers (4.5 µg) were separated by

SDS-polyacrylamide gel electrophoresis (4% stacking and 8% separating) and

transferred overnight onto nitrocellulose membranes. Human lymphoblastoid cDNA-

expressed CYP2A6 (BD Biosciences, San Jose, CA) was used as a positive control. For

the detection of CYP2A proteins, nitrocellulose membranes were preincubated for 1 h in

a blocking solution containing 1% skim milk powder (w/v), and 0.1% bovine serum

albumin (w/v) in Tris-buffered saline-Triton X-100 [50 mM Tris, pH 7.4, 150 mM NaCl,

Page 101: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

86

0.1% (v/v) Triton X-100]. Membranes were probed with a monoclonal antibody to human

CYP2A6 (1:3000 dilution) (BD Biosciences), a horseradish peroxidase-conjugated anti-

mouse secondary antibody (1:20000 dilution) (Pierce Biotechnology, Rockfort, IL),

followed by detection using enhanced chemiluminescence (Pierce Biotechnology).

Controls included immunoblots that were incubated without primary antibody as well as

immunoblots loaded with several other cDNA-expressed CYPs for the assessment of

cross-reactivity. Nitrocellulose membranes were exposed to Progene autoradiographic

film (Ultident Scientific, St. Laurent, P.Q.) for 0.5 to 2 min. Immunoblots were analyzed

using the MCID imaging system (Imaging Research Inc., St. Catherines, ON, Canada).

Nicotine C-Oxidation (NCO) Assay To establish assay linearity, increasing concentrations of hepatic microsomal proteins (0,

0.25, 0.5, 1, 1.5, 2, 2.5, 3 mg/ml) were incubated with 65 and 650 µM of nicotine for 30

minutes; microsomal protein (0.5 mg/ml) were incubated for various durations (0, 5, 10,

15, 30, 45, 60 mins) with 65 and 650 µM of nicotine. These concentrations of nicotine

were originally examined based on human pharmacokinetic variables (Messina et al.,

1997), and then refined as data on the mice was generated. Varying concentrations of

cytosolic protein (0.3, 0.6, 0.75, 0.9, 1.1, 1.2 mg/ml) were incubated with 0.5 mg/ml of

microsomal protein and with 30 and 300 µM of nicotine for 15 minutes to ensure that

aldehyde oxidase was not rate limiting. This is important since the conversion of nicotine

to cotinine follows two steps: the first of which is the conversion of nicotine to nicotine-

∆1’(5’)-iminium ion by CYP2A5. The imimium ion is then converted to cotinine by aldehyde

oxidase which is present in the cytosolic fraction (Brandange and Lindblom, 1979). From

these studies a protein concentration of 0.5 mg/ml and an incubation time of 15 min was

selected for all subsequent assays. Incubation mixtures also contained 1 mM NADPH

and 1 mg/ml of mouse liver cytosol in 50 mM Tris-HCl buffer, pH 7.4 and were

Page 102: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

87

performed at 37oC in a final volume of 0.5 ml. The reaction was stopped with 4%

Na2CO3, and 5-methylcotinine (70 µg) was added as the internal standard. Nicotine

metabolizing activity was analyzed as previously described (Messina et al., 1997) with

the modification of the solid phase extraction procedure for assessment of 3-

hydroxycotinine. Briefly, samples were extracted using ISOLUTE HM-N columns

(Argonaut Technologies, Redwood City, CA) with 13 ml of dichloromethane and were

dried under nitrogen. Samples were reconstituted with 105 µl of 0.01 M HCl and 90 µl of

each sample was analyzed by HPLC with UV detection (260 nm). Separation of nicotine,

cotinine, and 3-hydroxycotinine was achieved using a ZORBAX Bonus-RP column (5 µm,

150 × 4.6 mm; Agilent Technologies Inc., Mississauga, ON) and a mobile phase

consisting of acetonitrile/potassium phosphate buffer (10:90 v/v, pH 5.07) containing

3.3 mM heptane sulfonic acid and 0.5% triethylamine. The separation was performed

with isocratic elution at a flow rate of 0.9 ml/min. Nicotine, cotinine, and 3-

hydroxycotinine sample concentrations were determined from standard curves. The

quantitation limits were 5 ng/ml for nicotine, 12.5 ng/ml for cotinine, and 10 ng/ml for 3-

hydroxycotinine.

Statistical Analyses Data were analyzed by Graphpad Prism (version 4.00, San Diego, CA). Statistical

significance was calculated using Mann-Whitney U test and the significance level was

set at p=0.05. Correlation analyses were performed using Pearson’s method.

Page 103: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

88

Results

CYP2A5 detection in mouse liver An immunoblotting assay was developed in order to detect CYP2A5 in male mouse liver

microsomes. The CYP2A5 signal was linear between 2.5 to 10 µg of microsomal protein

(Fig. 8a & b). Selectivity was confirmed by incubating the immunoblots without the

primary antibody (Fig. 8c). The primary antibody did not react with several alternative

cDNA-expressed CYPs (Fig. 8d).

Increased oral nicotine self-administration in male mice is associated with higher CYP2A5 protein levels Male mice in the high nicotine consumption (HNC) group self-administered significantly

higher amounts of oral nicotine compared to the low nicotine consumption (LNC) group

(Table 5; Fig. 9a). We first determined if the increased consumption of nicotine in the

male mice was associated with elevated levels of hepatic CYP2A5 protein. HNC mice

had ~60% more CYP2A5 protein compared to the LNC mice (10.2 ± 1.0 vs. 6.5 ± 1.3

units, respectively, p=0.03, Fig. 9a). Figure 9b shows a representative immunoblot of

CYP2A5 expression in both the high and the low consumption mouse groups.

In vitro NCO rate is faster in high nicotine consuming male mice In addition to affecting levels of protein expression, genetic variation in CYP2A5 may

also affect protein structure altering the interaction with nicotine (Km). We compared the

kinetic parameters of in vitro NCO activities between the HNC and LNC mice. We first

optimized assay conditions and linearity for mouse NCO reactions. We established good

separation of peaks (Fig. 10a) and identified appropriate concentration of cytosol (1

Page 104: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

89

Figure 8. Selectivity and linearity of CYP2A5 immunoblotting for mouse liver microsomes. a) A dilution curve of mouse liver microsomal proteins detected by immunoblotting (± SEM, 3 experiments). Representative immunoblots incubated with (b) and without (c) primary antibody. d) cDNA-expressed CYP proteins as well as mouse liver microsomal protein detected (lanes 1:CYP1A1, 2:CYP1A2, 3:CYP2A2, 4:CYP3A2, 5:CYP1B1, 6:CYP2C11, 7:CYP2D, 8:CYP2E1, 9:CYP2A6; 0.2 pmol each; lane 10:mouse liver microsomal protein, 2 µg).

1047kDa

1 2 3 4 5 6 7 8 9

b)

c)

d)

2.4 3.2 4.8 6.4 9.6 µg0 12.8

a)

0.0 2.5 5.0 7.5 10.00

1

2

3

4

5

6

7

8

Protein (µµµµg)C

YP

2A

5 D

en

sit

y U

nit

s (

±± ±±S

EM

)

1047kDa

1 2 3 4 5 6 7 8 9

b)

c)

d)

2.4 3.2 4.8 6.4 9.6 µg0 12.8

a)

0.0 2.5 5.0 7.5 10.00

1

2

3

4

5

6

7

8

Protein (µµµµg)C

YP

2A

5 D

en

sit

y U

nit

s (

±± ±±S

EM

)

Page 105: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

90

Table 5. Average nicotine consumption (µg/day; mean ± SD) in F2 male and female mice in a two-bottle choice paradigm.

LNC Group HNC Group Male 3.8 ± 1.4 (n=11) 25.1 ± 1.2* (n=8)

Female 4.7 ± 1.4 (n=12) 25.1 ± 2.1* (n=14) *p<0.05 compared to the LNC group

Page 106: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

91

Figure 9. CYP2A5 levels are higher in male high nicotine consuming mice. a) Microsomal proteins from both male LNC (n=11) and HNC (n=8) mice were analyzed for CYP2A5 protein levels. For each animal, CYP2A5 protein levels were plotted against nicotine (NIC) consumption. Protein data are from 4 independent immunoblots for each animal. b) A representative immunoblot of CYP2A5 in LNC and HNC mice. *p=0.03

47 kDa

LNC HNCb)

a)

0 10 20 300

5

10

15

20

25LNC

HNC

NIC Consumption (µµµµg/day)

CY

P2

A5

De

ns

ity

Un

its

*

47 kDa

LNC HNCb)

a)

0 10 20 300

5

10

15

20

25LNC

HNC

NIC Consumption (µµµµg/day)

CY

P2

A5

De

ns

ity

Un

its

*

0 10 20 300

5

10

15

20

25LNC

HNC

NIC Consumption (µµµµg/day)

CY

P2

A5

De

ns

ity

Un

its

*

Page 107: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

92

Figure 10. Optimization of nicotine-C-oxidation assay. a) HPLC separation of a blank sample containing the standards nicotine (NIC), cotinine (COT), 3-hydroxycotinine (3HC), and the internal standard 5-methylcotinine (5MC). Assay conditions were optimized for: b) hepatic microsomal protein concentration; c) incubation time and; d) I cytosolic protein concentration. Arrows at the X-axis indicate conditions used for subsequent experiments.

0 15 30 45 600

5

10

15

20

25

30

35

40

4565 µM

650 µM

Incubation time (mins)

Nic

oti

ne C

-oxid

ati

on

(nm

ol co

tin

ine f

orm

ed

/mg

pro

tein

)

0.0 0.3 0.6 0.9 1.2 1.50.0

0.1

0.2

0.3

0.4

0.530 µM

300 µM

Cytosolic protein (mg/ml)

Nic

oti

ne C

-oxid

ati

on

(nm

ol c

oti

nin

e f

orm

ed

/min

/mg

pro

tein

)

0 1 2 30.0

0.1

0.2

0.3

0.4

0.565 µM

650 µM

Microsomal protein (mg/ml)

Nic

oti

ne

C-o

xid

ati

on

(ng

co

tin

ine

fo

rmed

/min

)

c)

b)

d)

a)

COT

NIC

5 10 15

Retention time (min)

3HC

Ab

so

rban

ce (

mA

U)

5MC

0 15 30 45 600

5

10

15

20

25

30

35

40

4565 µM

650 µM

Incubation time (mins)

Nic

oti

ne C

-oxid

ati

on

(nm

ol co

tin

ine f

orm

ed

/mg

pro

tein

)

0.0 0.3 0.6 0.9 1.2 1.50.0

0.1

0.2

0.3

0.4

0.530 µM

300 µM

Cytosolic protein (mg/ml)

Nic

oti

ne C

-oxid

ati

on

(nm

ol c

oti

nin

e f

orm

ed

/min

/mg

pro

tein

)

0 1 2 30.0

0.1

0.2

0.3

0.4

0.565 µM

650 µM

Microsomal protein (mg/ml)

Nic

oti

ne

C-o

xid

ati

on

(ng

co

tin

ine

fo

rmed

/min

)

c)

b)

d)

a)

COT

NIC

5 10 15

Retention time (min)

3HC

Ab

so

rban

ce (

mA

U)

5MC

COT

NIC

5 10 15

Retention time (min)

3HC

Ab

so

rban

ce (

mA

U)

5MC

Page 108: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

93

mg/ml, Fig. 10b) and microsomes (0.5 mg/ml, Fig. 10c), as well as the incubation time

(15 mins, Fig. 10d) for the assay.

Next, based on the recent publication of the mouse nicotine kinetics using

expressed cDNAs, the Km value (7.7 µM) from the cDNA-expressed CYP2A5 (Murphy et

al., 2005a) we examined nicotine at 6, 30, and 60 µM - concentrations predicted to allow

for the separation of structural (Km) and expression (Vmax) differences between the HNC

and LNC groups. We found that the NCO activities were significantly higher in the male

HNC group compared to the LNC group at all three nicotine concentrations (6 µM:

0.18±0.04 vs. 0.14±0.07 nmol/min/mg, p=0.02; 30 µM: 0.36±0.06 vs. 0.26±0.13

nmol/min/mg, p=0.01; 60 µM: 0.49±0.05 vs. 0.32±0.17 nmol/min/mg, p=0.01; Figs. 11a &

b), consistent with the higher levels of CYP2A5 protein. There was also a relatively good

association (R2=0.28; p=0.02; Fig. 11c) between NCO activities (60 µM of nicotine) and

CYP2A5 protein levels. We found that while the predicted Vmax was significantly higher in

the HNC mice, the apparent Km was not different between the two groups (Table 6). The

concentrations of nicotine, cotinine, and 3-hydroxycotinine, as measured by HPLC at 30

µM nicotine, are listed in Table 7.

Nicotine self-administration in female mice was not associated with CYP2A5 protein levels or in vitro NCO activity Similar to males, the female mice in the HNC group consumed significantly more

nicotine compared to the LNC group (Table 5; Fig. 12a & b). When we measured

CYP2A protein, no differences in CYP2A levels between the female HNC and LNC mice

were observed (Fig. 12a). When we assessed NCO activities, we found that the female

HNC (n=14) mice tended to metabolize nicotine (30 µM) at higher rates compared to the

LNC (n=12) but this did not reach significant difference (0.43 ± 0.18 vs. 0.39 ± 0.30

nmol/min/mg; p=0.12; Fig. 12b). There were also no differences in activities at 60 µM

Page 109: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

94

Figure 11. Nicotine-C-oxidation activity in male high nicotine consuming mice is higher compared to the low nicotine consuming mice. a) Liver microsomes from male LNC and HNC mice were tested for velocity (V) of cotinine formation (mean ± SD). b) Velocities at 30 µM nicotine were plotted against averaged daily nicotine consumption. c) Correlation between CYP2A5 protein levels and velocity at 60 µM nicotine. *p=0.02 and **p=0.01 compared to the LNC group.

6 uM 30 uM 60 uM0.0

0.1

0.2

0.3

0.4

0.5

0.6 LNC

HNC

*

**

**

6 µµµµM 30 µµµµM

NIC

V (

nm

ol/

min

/mg

)

60 µµµµM

a) b)

c)

0 10 200.00

0.25

0.50

0.75

1.00LNC

HNC

R2=0.28p=0.02

CYP2A5 Density Units

V (

nm

ol/m

in/m

g)

0 10 20 300.00

0.25

0.50

0.75

1.00LNC

HNC

NIC Consumption (µµµµg/day)

V (

nm

ol/m

in/m

g) *

6 uM 30 uM 60 uM0.0

0.1

0.2

0.3

0.4

0.5

0.6 LNC

HNC

*

**

**

6 µµµµM 30 µµµµM

NIC

V (

nm

ol/

min

/mg

)

60 µµµµM

a) b)

c)

0 10 200.00

0.25

0.50

0.75

1.00LNC

HNC

R2=0.28p=0.02

CYP2A5 Density Units

V (

nm

ol/m

in/m

g)

0 10 20 300.00

0.25

0.50

0.75

1.00LNC

HNC

NIC Consumption (µµµµg/day)

V (

nm

ol/m

in/m

g) *

0 10 20 300.00

0.25

0.50

0.75

1.00LNC

HNC

NIC Consumption (µµµµg/day)

V (

nm

ol/m

in/m

g) **

Page 110: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

95

Table 6. Estimated in vitro NCO kinetic parameters of male LNC and HNC mice (mean ± SD)

Km (µM) Vmax (nmol/min/mg) LNC (n=10) 7.8 ± 2.0 0.30 ± 0.16 HNC (n=8) 10.7 ± 4.7 0.48 ± 0.05* *p=0.003 compared to the LNC group

Page 111: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

96

Table 7. Concentrations (mean ± SD) of nicotine, cotinine, and 3-hydroxycotinine (ng/mL) from the in vitro NCO activity assays (at 30 µM nicotine)

Nicotine (µg/mL)

Cotinine (µg/mL)

3-hydroxycotinine (µg/mL)

LNC (n=11) 4.63 ± 0.42 0.17 ± 0.09 ND Male

HNC (n=8) 4.58 ± 0.41 0.24 ± 0.04 ND LNC (n=12) 4.41 ± 0.36 0.26 ± 0.20 ND

Female HNC (n=14) 4.37 ± 0.34 0.28 ± 0.12 ND

ND: not detected

Page 112: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

97

Figure 12. CYP2A4/5 levels and in vitro NCO activities do not differ between female LNC and HNC mice. Microsomal proteins from female LNC and HNC mice were analyzed for a) CYP2A4/5 protein levels and b) NCO activities at 30 µM of nicotine. Protein data are from 4 independent immunoblots for each animal.

0 10 20 30

0

5

10

15

20

25LNC

HNC

NIC Consumption (µµµµg/day)

CY

P2

A4

/5

a) b)

0 10 20 30

0.00

0.25

0.50

0.75

1.00LNC

HNC

NIC Consumption (µµµµg/day)

V (

nm

ol/m

in/m

g)

Page 113: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

98

nicotine between LNC (n=7) and HNC (n=8) mice (0.61 ± 0.48 vs. 0.59 ± 0.17

nmol/min/mg; p=0.40).

Page 114: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

99

Discussion

One determinant of the amount of smoking is the rate of nicotine metabolic inactivation,

whereby the systemic availability of nicotine influences the number of cigarettes smoked

and the intensity of inhalation (McMorrow and Foxx, 1983; Russel, 1987; Schoedel et al.,

2004). Here we showed that nicotine self-administration in male mice is associated with

the amount of nicotine metabolizing enzyme CYP2A5 as well as the rate at which

nicotine is metabolized in vitro. CYP2A5 protein levels were correlated with NCO

activities. Results from the current study are consistent with human data in which cyp2a6

genotype and/or phenotype can have a significant impact on cigarette consumption

behaviours (Rao et al., 2000; Schoedel et al., 2004) where reducing or increasing

nicotine clearance can decrease smoking (Sellers et al., 2000; Sellers et al., 2003b) or

increase smoking (Benowitz and Jacob, 1985), respectively.

There are several interpretations for the presence of higher CYP2A5 levels and

NCO activities in the HNC mice. Quantitive trait loci analyses have implicated the

potential involvement of CYP2A5 in differential nicotine consumption levels (Wildenauer

et al., 2005); therefore, genetic variation in CYP2A5 may contribute to the observed

nicotine intake behaviours. For instance, CYP2A5 levels may be related to

polymorphisms in cis- and/or trans-regulatory factors that influence transcriptional

efficiency, translational efficiency, as well as message stability. This is plausible since

the human CYP2A6 contains multiple polymorphisms that affect its expression and

activity (Nakajima and Yokoi, 2005). For example, the CYP2A6*9 polymorphism

encodes a TATA box mutation which reduces its mRNA production by 40-70% (Kiyotani

et al., 2003; Pitarque et al., 2001; Yoshida et al., 2003). Smokers who are homozygous

for CYP2A6*9 (slow metabolizers) have reduced nicotine metabolism by approximately

50% (Schoedel et al., 2004; Yoshida et al., 2003); and slow metabolizers smoke ~20%

Page 115: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

100

fewer cigarettes compared to individuals who are normal metabolizers (Schoedel et al.,

2004). Polymorphism in CYP2A6 may also affect the structural stability of the enzyme.

The CYP2A6*2 variant codes for a protein with a single amino-acid substitution (L160H)

that results in an unstable and inactive enzyme (Yamano et al., 1990). In mice a single

nucleotide polymorphism causing a change in amino acid (V117A) in CYP2A5, and the

resultant coumarin-hydroxylase activity, has also been described (Lindberg et al., 1992).

Another potential explanation for the greater CYP2A5 levels in male HNC mice is

that mice that consume more nicotine may have induced CYP2A5, thus further

increasing nicotine consumption. However, studies have shown that chronic nicotine

administration decreases both the plasma clearance of nicotine in humans and CYP2A6

protein expression in monkeys (Benowitz and Jacob, 1993; Lee et al., 1987; Schoedel et

al., 2003). Therefore it is unlikely that the higher CYP2A5 levels in the male HLC mice

are the result of induction by nicotine consumption.

In addition to the nicotine metabolic pathway, other components (and their

respective genetic variations) involved in nicotine sensitivity and reward properties could

also affect nicotine intake, and many of these candidate genes have been investigated in

humans (Tyndale, 2003). In inbred mice there are marked differences in response to

nicotine across strains (Collins et al., 1988; Crawley et al., 1997; Marks et al., 1983). It is

therefore possible that nicotine self-administration in mice may also be related to factors

such as fluid intake and taste preference. Water consumption between the high (C57Bl/6)

and the low (St/bJ) nicotine consumption parental strains do not differ, indicating that the

difference in nicotine consumption is not related to fluid intake (Aschhoff et al., 1999).

Lush and Holland (1998) found that the C57Bl/6 mice have a greater preference for the

bitter compound copper glycinate compared to the St/bJ mice. However, this may be

difficult to interpret with regards to nicotine since the ranking of copper glycinate

Page 116: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

101

preference does not match that of nicotine preference in several other strains (Aschhoff

et al., 1999; Lush and Holland, 1988).

In females, the higher overall CYP2A protein levels compared to males and lack

of difference in CYP2A protein between consumption groups is likely due to the

presence of the female specific CYP2A4 protein (Burkhart et al., 1985; Harada and

Negishi, 1984). The CYP2A4 protein is approximately 98% identical to CYP2A5 in amino

acid sequence and they are indistinguishable from one another using available

antibodies (Su et al., 1998). CYP2A4 metabolizes nicotine only at very high

concentrations (more than 15-fold lower in affinity) and should not contribute to

metabolism at the nicotine concentrations consumed (Murphy et al., 2005a). We

therefore investigated the in vitro NCO activities between the two female groups. We

observed no significant difference between the female high and low consumers in in vitro

nicotine metabolism; however, the distribution of nicotine metabolism within each group

is such that the majority of NCO activities were higher in the high consumption group

compared to the low consumption group.

It appears that components in addition to nicotine metabolism have a greater

contribution to nicotine consumption in females compared to males. The apparent lack of

association between nicotine consumption levels and nicotine metabolism in females

may be related to effects of steroids (Dempsey et al., 2002). For example, in humans,

pregnancy increases plasma nicotine and cotinine clearances, but nicotine intake from

cigarettes are similar between pregnant and post-partum females (Dempsey et al., 2002).

On the other hand, in ICR mice, administration of exogenous progesterone or 17β-

estradiol to female mice dose-dependently decreased the antinociceptive effect of

nicotine, but administration of testosterone to male mice did not alter pain sensitivity

(Damaj, 2001). The differential effects of these hormones were attributed to the ability of

progesterone and 17β-estradiol to antagonize the binding of nicotine to the α4β2

Page 117: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

102

nicotinic-receptors (Damaj, 2001). Therefore, female sex hormones may modulate

nicotine-mediated behaviours (e.g. consumption) reducing the relative impact of nicotine

metabolism in these mice.

In conclusion, we found that varying levels of nicotine self-administration

behaviour in male mice were associated with different rates of nicotine metabolism, most

likely mediated by genetic controls over the differing expression levels of the CYP2A5

protein. This observation is similar to that observed in smokers in which the amount

smoked is related to the functional levels of CYP2A6 (Rao et al., 2000; Schoedel et al.,

2004). Using inducers and inhibitors, as well as various mouse strains, we can model

the influence of nicotine metabolism on various stages of smoking seen in humans such

as initiation, acquisition, maintenance, and cessation.

Page 118: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

103

Acknowledgements

We would like to thank Amir Boutrous for his assistance in protein analysis; Sharon

Miksys for guidance and critical review of the manuscript; Helma Nolte and Raj Sharma

for their assistance with HPLC assessment.

Support

This study was supported by CAMH, CIHR 14173 & 53248. CIHR-Special Training

Program in Tobacco Use in Special Populations and OGS to ECKS and Canada

Research Chair in Pharmacogenetics.

Page 119: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

104

SIGNIFICANCE OF CHAPTER

This is the first study to demonstrate the association between CYP2A5 and the

pharmacology of nicotine in mice. Specifically, we found that nicotine metabolism is

associated with the amount of CYP2A5 (expression/activity) as contributed by genetic

variations. More importantly, CYP2A5-mediated nicotine metabolism is associated with

NSA in which mice exhibiting faster in vitro nicotine metabolism self-administered more

nicotine. These findings are in line with humans in which CYP2A6 phenotype can have a

significant impact on cigarette consumption behaviours.

Page 120: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

105

CHAPTER 4: INHIBITION OF NICOTINE METABOLISM BY METHOXSALEN: PHARMACOKINETIC AND PHARMACOLOGICAL STUDIES IN MICE

M. Imad Damaj, Eric C. K. Siu, Edward M. Sellers, Rachel F. Tyndale, and Billy R. Martin

As published in: Journal of Pharmacology and Experimental Therapeutics, 2007, Jan; 320(1):250-257

The in vitro enzyme activities assays, analyses of the pharmacokinetic data, writing of

the enzyme kinetics and pharmacokinetics section of the manuscript were performed by

ECKS. The behavioural work were performed by MI Damaj. The study was designed by

MI Damaj and RF Tyndale.

Reprinted from Molecular Pharmacology, Volume 320, Damaj et al., “INHIBITION OF NICOTINE METABOLISM BY METHOXSALEN: PHARMACOKINETIC AND PHARMACOLOGICAL STUDIES IN MICE”, pp 250-257, Copyright (2007) with permission from The American Society for Pharmacology and Experimental Therapeutics. © 2007 The American Society for Pharmacology and Experimental Therapeutics

Page 121: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

106

Abstract

Studies were undertaken to examine whether methoxsalen, a specific and relatively

selective inhibitor of human CYP2A6, inhibited CYP2A5-mediated nicotine metabolism in

vitro. Furthermore, studies were performed in vivo to determine whether methoxsalen

would modulate acute nicotine pharmacokinetics and pharmacological effects

(antinociception and hypothermia) in ICR mouse. Our results demonstrated that

methoxsalen competitively inhibits in vitro nicotine metabolism in mice. The inhibition

was potent, as seen in human inhibition studies, with a Ki of 0.32 µM. In addition, we

found that administration of methoxsalen significantly increased the plasma half-life of

nicotine (approximately doubled) and increased its AUC compared to saline treatment.

There was a dose-dependent enhancement in the pharmacological effects of nicotine

(body temperature and analgesia) after methoxsalen treatment. Methoxsalen prolonged

the duration of nicotine-induced antinociception and hypothermia (2.5 mg/kg) for periods

up to 180 min post-nicotine administration. Furthermore, this prolongation in nicotine’s

effects after methoxsalen was associated with a parallel prolongation of nicotine plasma

levels in mice. These data strongly suggest that variation in the rates of nicotine

metabolic inactivation substantially alter nicotine’s pharmacological effects. In

conclusion, these results confirmed that methoxsalen did indeed inhibit the conversion of

nicotine to cotinine both in vitro and in vivo. They also suggest that mice may represent

a suitable model for studying variation in nicotine metabolism and its impact on

mechanisms of nicotine dependence, including the use of inhibitors to reduce nicotine

metabolism.

Page 122: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

107

INTRODUCTION

Nicotine is known to induce several physiological and pharmacological effects

and to produce subjective feelings of reward and pleasure in humans and animals.

Several well-characterized behavioral models and tests are currently used for

investigating the biological and pharmacological mechanisms underlying nicotine

dependence. These models reveal a high level of variability among nicotine behavioral

responses due to factors such as differing genetic background, species, sex, age, initial

basal and behavioral state, route and regimen of administration. In addition, the narrow

effective dose range for many pharmacological and behavioral effects of nicotine further

compounds the variability (see review by (Picciotto, 2003)). Species differences also

represent an important consideration regarding in vivo responses to either acute or

chronic nicotine exposure, particularly mice versus rats. In general, mice are less

sensitive to the acute pharmacological effects of nicotine than rats. For example,

nicotine’s potency in the tail-flick and hot-plate tests after systemic administration in mice

is 5-10 fold lower than that reported for rats (Tripathi et al., 1982). One of the most

striking differences concerns nicotine’s locomotor effects. Although nicotine induces

locomotor hyperactivity in rats, it generally fails to induce locomotor hyperactivity in mice

at any dose (Castane et al., 2002; Damaj and Martin, 1993; Freeman et al., 1987;

Gaddnas et al., 2002; Itzhak and Martin, 1999; Kita et al., 1992; Marks et al., 1983;

Smolen et al., 1994). In addition, even though i.v. nicotine self-administration can be

accomplished in mice, a model using drug-naïve mice not restricted in their movement

and a limited access schedule similar to that frequently used in rats is not reported. This

rat-mouse variability can be attributed to various factors, such as pharmacokinetics,

stress-induced by behavioral manipulations, physiological status, and/or psychological

context. Given the very rapid metabolism of nicotine in mice (plasma and brain t1/2 = 5.9-

6.9 min) (Petersen et al., 1984) compared to that of the rat (45 min) (Hwa Jung et al.,

Page 123: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

108

2001), distribution and kinetic factors appear to play an important role in the behavioral

response to nicotine in mice. A recent study reported that nicotine oral self-

administration in male mice is associated with the amount of nicotine metabolizing

enzyme CYP2A5 as well as the rate at which nicotine is metabolized in vitro (Siu et al.,

2006).

The main purpose of this study was to elucidate the impact of nicotine

metabolism on nicotine’s acute pharmacological effects in mice. The mouse is a better-

suited model for the study of nicotine metabolism and behaviors, compared to rats, since

the predominant enzymes responsible for the metabolism of nicotine in rats belong to

the CYP2B family (Nakayama et al., 1993; Schoedel et al., 2001; Siu et al., 2006). In

contrast to rats, human CYP2A6 and mouse CYP2A5 (mouse orthologous form of

human CYP2A6) are the main enzymes involved in nicotine metabolism (Messina et al.,

1997; Murphy et al., 2005b). There is close structural (84% amino acid sequence

similarity) and functional (CYP2A5 metabolizes nicotine with similar efficiency to human

CYP2A6) similarity between CYP2A5 and CYP2A6 (Murphy et al., 2005b; Siu et al.,

2006). In addition a similarly large portion of nicotine is metabolized to cotinine in

humans and mice, relative to the rat, making the mouse a good model for the

pharmacological impacts of manipulating nicotine metabolism in vivo. Accordingly,

studies were undertaken to examine whether methoxsalen [a specific and relatively

selective inhibitor of human CYP2A6, (Zhang et al., 2001)] inhibited CYP2A5-mediated

nicotine metabolism in vitro. Subsequently studies were performed in vivo to determine

whether methoxsalen would modulate acute nicotine pharmacokinetics and

pharmacological effects (antinociception and hypothermia) in the mouse.

Page 124: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

109

Materials and Methods

Animals

Male ICR mice (20-25g) obtained from Harlan Laboratories (Indianapolis, IN) were used

throughout the study. Animals were housed in groups of six and had free access to food

and water. Animals were housed in an AAALAC approved facility and the study was

approved by the Institutional Animal Care and Use Committee of Virginia

Commonwealth University.

Drugs

Mecamylamine hydrochloride was supplied as a gift from Merck, Sharp and Dohme &

Co. (West Point, PA). (-)-Nicotine hydrogen tartrate was obtained from Aldrich Chemical

Company, Inc. (Milwaukee, WI) and converted to the ditartrate salt as described by

(Aceto et al., 1979) (1979). Methoxsalen was purchased from Sigma-Aldrich (St. Louis,

MO). All drugs were dissolved in physiological saline (0.9% sodium chloride). All doses

are expressed as the free base of the drug.

Membrane Preparations

Microsomal membranes were prepared from mouse livers for the in vitro nicotine

metabolism assay as previously described (Messina et al., 1997; Siu et al., 2006) and

stored at 80°C in 1.15% KCl. The cytosolic fractions were also acquired during the

membrane preparation and were used as a source of aldehyde oxidase.

Nicotine C-Oxidation Assay

Assay conditions were optimized for ICR mouse microsomes as previously described

(Siu et al., 2006). Linear formation of cotinine from nicotine was obtained under assay

Page 125: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

110

conditions of 0.5 mg/ml protein concentration with an incubation time of 15 min.

Incubation mixtures contained 0.5 mg/ml microsomal protein, 1 mM NADPH and 1

mg/ml of mouse liver cytosol in 50 mM Tris-HCl buffer, pH 7.4 and were performed at

37oC in a final volume of 0.5 ml. The reaction was stopped with 4% Na2CO3, and 5-

methylcotinine (70 µg) was added as the internal standard. In vitro kinetic parameters

(Km and Vmax) of nicotine metabolism in ICR mice were characterized using substrate

concentrations of 0, 3.0 to 1000 µM nicotine. Nicotine metabolizing activity was analyzed

by HPLC as previously described (Messina et al., 1997) with the modification of the solid

phase extraction procedure for assessment of 3-hydroxycotinine (Siu et al., 2006).

Separation of nicotine and cotinine was achieved using a ZORBAX Bonus-RP column (5

mm, 150 x 4.6 mm; Agilent Technologies Inc., Mississauga, ON) and a mobile phase

consisting of acetonitrile/potassium phosphate buffer (10:90 v/v, pH 5.07) containing 3.3

mM heptane sulfonic acid and 0.5% triethylamine. The separation was performed with

isocratic elution at a flow rate of 0.9 ml/min. Nicotine, cotinine, and 3-hydroxycotinine

sample concentrations were determined from standard curves. The limits of

quantification were 5 ng/ml for nicotine and 12.5 ng/ml for cotinine.

Methoxsalen Inhibition of Cotinine Formation from Nicotine

To determine the apparent Ki value of inhibition of cotinine formation by methoxsalen in

the ICR mouse microsomes, 0, 0.04, 0.1, 0.2, and 0.4 µM methoxsalen were used. The

substrate (nicotine) concentrations were 10, 20 and 40 µM, which were approximately

equal to 1/2Km, Km, and 2Km.

Plasma nicotine and cotinine levels measurement

To determine plasma nicotine and cotinine levels, blood samples were drawn by cardiac

puncture at 5, 15, 30, 45, 60, 120, and 180 minutes after nicotine administration (2.5

Page 126: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

111

mg/kg, s.c.). Animals were pretreated with saline or methoxsalen for a variety of times

before nicotine administration. Immediately afterwards the plasma samples were

prepared by centrifugation at 3000 x g for 10 min and frozen at –20˚C until analysis. To

measure total nicotine and cotinine levels (free and glucuronides) the samples were

incubated with β-glucuronidase at a final concentration of 5 mg/ml in 0.2 M acetate

buffer, pH 5.0, at 37oC overnight. After incubation the samples were processed and

analyzed for nicotine and metabolite levels by HPLC as described above.

Behavioral tests

1. Tail-flick test. Antinociception was assessed by the tail-flick method of D'Amour and

Smith (d'Amour and Smith, 1941) as modified by Dewey et al. (Dewey et al., 1970). A

control response (2-4 sec) was determined for each mouse before treatment, and a test

latency was determined after drug administration. In order to minimize tissue damage, a

maximum latency of 10 sec was imposed. Antinociceptive response was calculated as

percent maximum possible effect (% MPE), where % MPE = [(test-control)/(10-control)] x

100.

2. Hot-plate Test. Mice were placed into a 10 cm wide glass cylinder on a hot plate

(Thermojust Apparatus) maintained at 55.0oC. Two control latencies at least ten min

apart were determined for each mouse. The normal latency (reaction time) was 8 to 12

seconds. Antinociceptive response was calculated as percent maximum possible effect

(% MPE), where % MPE = [(test-control)/(40-control) x 100]. The reaction time was

scored when the animal jumped or licked its paws. In order to minimize tissue damage,

a maximum latency of 40 sec was imposed.

Groups of eight to twelve animals were used for each dose and for each

treatment. Studies were carried out by pretreating the mice with either saline or

Page 127: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

112

methoxsalen 30 min before nicotine. The animals were tested 5 min after administration

of nicotine.

3. Body temperature. Rectal temperature was measured by a thermistor probe

(inserted 24 mm) and digital thermometer (Yellow Springs Instrument Co., Yellow

Springs, OH). Readings were taken just before and at 30 min after the s.c. injection of

either saline or nicotine. The difference in rectal temperature before and after treatment

was calculated for each mouse. The ambient temperature of the laboratory varied from

21-24oC from day to day.

Statistical analysis

Kinetic (i.e. apparent Km, Vmax, Ki) parameters were calculated using Graphpad Prism

(Graphpad Software Inc., San Diego, CA) and were verified by the Eadee-Hofstee plots.

The type of inhibition by methoxsalen was further assessed by the Dixon method.

Assessment of in vivo nicotine and cotinine plasma levels for the entire time course from

individual animals was not possible due to limited blood volume; therefore, each time

point represents data from individual mice. Due to this restriction to the experimental

design, statistical parameters (i.e. half life) were estimated by resampling methods using

the PKR and Test software (H. L. Kaplan, Toronto, ON). Statistical analysis of all

analgesic and behavioral studies was performed using either t-test or analysis of

variance (ANOVA) with Tukey’s test post hoc test when appropriate. All differences

were considered significant if at p < 0.05. ED50 values with 95% CL for behavioral data

were calculated by unweighted least-squares linear regression as described by Tallarida

and Murray (Tallarida and Murray, 1987).

Page 128: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

113

Results

In vitro nicotine C-oxidation activity of mice

We first established the kinetic parameters apparent Km and Vmax for in vitro nicotine

metabolism in ICR mice. Nicotine metabolism followed two-site Michaelis-Menten

kinetics as illustrated by the Eadee-Hofstee plot (Fig. 13 & inset). The estimated kinetic

parameters for the in vitro nicotine metabolism, derived from five animals, for both sites

can be found in Table 8. One site exhibited high affinity for nicotine with an apparent Km

of 18.6 ± 5.9 µM. The second site had a much lower affinity for nicotine with an

apparent Km of 215.2 ± 60.0 µM. The overall catalytic efficiency (Vmax/Km) was greater

for the high-affinity site compared to the low-affinity site (0.013 ± 0.007 vs. 0.003 ±

0.001).

Inhibition of in vitro nicotine C-oxidation in mouse microsomes by methoxsalen

Previous studies indicated that CYP2A5 is the primary enzyme responsible for nicotine

metabolism in mice (Murphy et al., 2005b). We tested whether the human CYP2A6

inhibitor methoxsalen (Zhang et al., 2001) was also an inhibitor of nicotine metabolism in

ICR mouse microsomes. Methoxsalen inhibited nicotine metabolism with a Ki of 0.32 ±

0.03 µM (Fig. 14 & Table 8). Methoxsalen was a competitive inhibitor with some

indication of a mixed-type of inhibition consistent with it also being a mechanism-based

inhibitor in vitro.

Effects of methoxsalen on nicotine and cotinie plasma levels after in vivo treatment

Having established that methoxsalen can inhibit in vitro nicotine metabolism in the ICR

mice, we tested the effect of the inhibitor on in vivo plasma nicotine and cotinine levels.

Page 129: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

114

Figure 13. In vitro nicotine C-oxidation activity of ICR mice. Microsomal proteins (0.5 mg/ml) from adult male ICR mice were incubated with increasing concentrations of nicotine for 15 minutes. The plot shows the average cotinine formation velocity ± SD values for 5 individual animals. Inset: a representative Eadee-Hofstee plot indicating the presence of two enzymatic sites responsible for the metabolism of nicotine.

0 250 500 750 10000.0

0.2

0.4

0.6

0.8

1.0

1.2

NIC (µµµµM)

Nic

oti

ne C

-oxid

ati

on

V (

nm

ol co

tin

ine f

orm

ed

/min

/mg

pro

tein

)

0.00 0.01 0.02 0.030.00

0.25

0.50

0.75

1.00

1.25

V/S

V (

nm

ol/

min

/mg

)

0 250 500 750 10000.0

0.2

0.4

0.6

0.8

1.0

1.2

NIC (µµµµM)

Nic

oti

ne C

-oxid

ati

on

V (

nm

ol co

tin

ine f

orm

ed

/min

/mg

pro

tein

)

0.00 0.01 0.02 0.030.00

0.25

0.50

0.75

1.00

1.25

V/S

V (

nm

ol/

min

/mg

)

Page 130: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

115

Table 8. Kinetic parameters of in vitro nicotine C-oxidation activities of adult male ICR mouse liver microsomes. Km, Vmax and Ki values were determined independently for each animal and the average ± SD values are shown.

ICR (n=5)

High Affinity Km (µM) 18.6 ± 5.9

Vmax (nmol/min/mg) 0.22 ± 0.05

Vmax/Km 0.013 ± 0.007

Low Affinity Km (µM) 215.2 ± 60.0

Vmax (nmol/min/mg) 0.65 ± 0.08

Vmax/Km 0.003 ± 0.001

Methoxsalen Inhibition Ki (µM) 0.32 ± 0.03

Page 131: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

116

Figure 14. A representative Dixon plot of inhibition of cotinine formation by ICR mouse microsomes in presence of varying concentrations of methoxsalen. In such plots, the abscissa of the point where the regression lines intersect is the negative of the estimated value of Ki. Three animals were used to establish Ki value. Velocity is expressed as nmol/min/mg.

-0.4 -0.2 0.0 0.2 0.4 0.6

20

40

60

8010 µM

20 µM

40 µM

8-MOP (µµµµM)

1/V

Ki = 0.3 µM

-0.4 -0.2 0.0 0.2 0.4 0.6

20

40

60

8010 µM

20 µM

40 µM

8-MOP (µµµµM)

1/V

Ki = 0.3 µM

Page 132: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

117

Administration of methoxsalen significantly increased the plasma half-life of nicotine and

increased its AUC compared to saline treatment (Fig. 15A & Table 9). Methoxsalen had

minimal effect on the maximum nicotine concentration (Table 9) consistent with the

inhibition of nicotine given by the subcutaneous route. The plasma levels of the primary

CYP2A5-mediated metabolite of nicotine, cotinine, were also reduced as a result of

inhibition of this nicotine metabolic pathway by methoxsalen (Fig 15B).

Effects of methoxsalen on nicotine acute pharmacological effects (antinociception and

hypothermia): Time-course, potency and selectivity studies

Methoxsalen given s.c. was evaluated for its ability to enhance a 2.5 mg/kg dose (an

ED84 dose) of nicotine-induced antinociception (tail-flick and hot-plate procedures) and

hypothermia after nicotine s.c. injection of the drug. We first carried out a time-course

study of methoxsalen (10 mg/kg) enhancement of nicotine’s effects to determine an

optimal pretreatment time. Mice were given methoxsalen at different times and then

received nicotine (2.5 mg/kg s.c.). The mice were tested 45 min after nicotine

administration. As illustrated in Fig. 16A, the potentiation of nicotine’s antinociceptive

effects by methoxsalen pretreatment in the tail-flick and hot-plate tests was time-

dependent with maximum enhancement occurring between 30 and 60 min. A similar

time-course was seen with nicotine-induced hypothermia (Fig. 16B) with maximum

enhancement occurring between 30 and 120 min. Based on the time-course results,

subsequent studies were carried out by pretreating the mice with methoxsalen 30 min

before nicotine. We then determined its potency of enhancing nicotine’s effects at this

pretreatment time. As showed in Fig. 17, methoxsalen dose-dependently enhanced

nicotine-induced antinociception in the tail-flick (Fig. 17A), hot-plate (Fig. 17B) assays

and hypothermia (Fig. 17C).

Page 133: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

118

Figure 15. Time course of nicotine and cotinine plasma concentration in mice pretreated with methoxsalen. Nicotine was administered s.c. (2.5 mg/kg) 30 min after pretreatment with ( ) saline or ( ) methoxsalen. Each time point represents the mean ± SD of 4 to 6 animals. For the saline pretreatment nicotine plasma levels, five of seven values at 120 min, and all values at 180 min, were below the limits of detection. They were included in the plot as 1 ng/ml for visualization of the data.

(A) (B)(A) (B)

Page 134: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

119

Table 9. Pharmacokinetic parameters of nicotine in adult male ICR mouse after pretreatment with methoxsalen. Values are shown as mean ± standard deviation.

Cmax (ng/ml)

AUC

(ng*hr/ml) T1/2

(min)

Saline/Nicotine 314 ± 170 132 ± 14 14.3 ± 8.1

Methoxsalen/Nicotine 354 ± 26 334 ± 24* 30.1 ± 13.5*

* p < 0.05 compared to Saline/Nicotine

Page 135: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

120

Figure 16. Time-course of methoxsalen effects on nicotine-induced (A) antinociception and (B) hypothermia after s.c. administration of 10 mg/kg in mice. Nicotine was given at the dose of (2.5 mg/kg, s.c.) and animals were tested 45 min after injection. Each point represents the mean ± SE of 8 to 12 mice. *P <0.05 compared to correspondent zero time point.

0

20

40

60

80

100

0 5 15 30 60 120 180

Tail-FlickHot-Plate

% M

PE

Time (min)

*

* **

**

*

*

-7

-6

-5

-4

-3

-2

-1

0

0 5 15 30 60 120 180

²ÞC

Time (min)

** *

*

(A) (B)

∆o

% m

axim

um

po

ssib

le e

ffec

t

Page 136: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

121

Figure 17. Dose-response enhancement of nicotine by methoxsalen after s.c. administration in mice. Methoxsalen at different doses was administered s.c. 15 min before nicotine (2.5 mg/kg, s.c.) and mice were tested 45 min later in (A) the tail-flick and (B) the hot-plate tests. Each point represents the mean ± SE of 8 to 12 mice. *P <0.05 compared to saline (zero dose) treatment.

0

25

50

75

100

% M

PE

0 1 5 10 15

8-MOP (mg/kg)

0

25

50

75

100

% M

PE

0 1 5 10 15

8-MOP (mg/kg)

(A) (B)

*

* *

*

* *

% m

ax

imu

m p

oss

ible

eff

ect

% m

ax

imu

m p

oss

ible

eff

ect

Page 137: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

122

By itself, methoxsalen did not significantly change tail-flick or hot-plate basal latencies,

or body temperature at the indicated doses and times. Furthermore, the antinociceptive

and hypothermic effects of methoxsalen/nicotine combination were totally blocked by a

pretreatment with mecamylamine, a non-competitive nicotinic antagonist, at 2 mg/kg

(Fig. 18 & Table 10).

To determine whether methoxsalen produced similar effects on other analgesic

substances, it was given at a dose of 10 mg/kg before the administration of an inactive

dose of morphine. Pretreatment with methoxsalem failed to significantly enhance the

effect of morphine in the tail-flick test [Saline/Saline = 7 ± 5 % MPE; Saline/Morphine

(0.5 mg/kg, s.c.) = 20 ± 7 % MPE; Methoxsalen (10 mg/kg)/Morphine (0.5 mg/kg, s.c.) =

25 ± 8 % MPE]. Thus, the effect of methoxsalen appears not to be generalized to other

analgesic substances since it did not enhance the effects of morphine after systemic

administration.

Effects of methoxsalen on the time-course of nicotine’s pharmacological effects

The onset of action for nicotine (2.5 mg/kg, s.c.) in the tail-flick test was rapid with

maximum antinociception occurring between 0 and 5 min. The duration of

antinociception was relatively brief in that the effect had disappeared completely within

45 min after nicotine administration in mice (Fig. 19A). A similar time-course pattern was

observed in the hot-plate test (Fig. 19B). The duration of nicotine-induced hypothermia

was however longer compared to that of antinociception. Indeed, nicotine’s effect

disappeared completely within 180 min after s.c. administration in mice (Fig. 19C).

When animals were pretreated with methoxsalen (10 mg/kg, s.c.), the effects of nicotine

in both analgesic tests were significantly prolonged. Nicotine-induced antinociception

did not disappear completely till 180 min after nicotine administration in mice (Fig. 19A &

Page 138: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

123

Figure 18. Blockade of methoxsalen’s effects on nicotine-induced antinociception in the tail-flick test after s.c. administration in mice. Animals were pretreated with either saline or mecamylamine (2 mg/kg, s.c.) followed by methoxsalen (10 mg/kg, s.c). Mice received 30 min later nicotine at a dose of 2.5 mg/kg and were tested 45 min after injection. Each point represents the mean ± SE of 8 to 12 mice. MOP (10) = methoxsalen at 10 mg/kg; Sal: Saline; Nic (2.5): Nicotine at 2.5 mg/kg; Meca (2): Mecamylamine at 2 mg/kg. *P <0.05 compared to control.

0

25

50

75

100

% M

PE

Control MOP (10) Sal/Nic (2.5) MOP (10)/Nic (2.5) Meca (2)/MOP (10)/Nicc (2.5)

*%

max

imu

m p

oss

ible

eff

ect

Page 139: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

124

Table 10. Blockade of methoxsalen’s effects on nicotine acute pharmacological responses after s.c. administration in mice. Animals were pretreated with either saline or mecamylamine (2 mg/kg, s.c.) followed by methoxsalen (10 mg/kg, s.c). Mice received 30 min later nicotine at a dose of 2.5 mg/kg and were tested 45 min after injection. Methoxsalen (10) = methoxsalen at 10 mg/kg; Meca (2): Mecamylamine at 2 mg/kg. Each point represents the mean ± SE of 8 to 12 mice.

Treatment

(mg/kg) Tail-flick Test

% MPE (Mean ± SEM)

Hot-plate Test % MPE

(Mean ± SEM)

Body temperature ∆

oC (Mean ± SEM)

Saline/Saline/Saline 7 ± 2 8 ± 2 -0.1 ± 0.2

Saline/Methoxsalen (10)/Saline 10 ± 3 12 ± 5 -1.5 ± 0.2

Meca (2)/Methoxsalen (10)/Saline 15 ± 7 10 ± 5 -1.8 ± 0.5

Meca (2)/Saline/Saline 8 ± 3 6 ± 4 -0.5 ± 0.3

Meca (2)/Saline/Nicotine (2.5) 5 ± 2 6 ± 2 -1.9 ± 0.6*

Saline/Saline/Nicotine (2.5) 10 ± 5 15 ± 5 -4.1 ± 0.6

Saline/Methoxsalen (10)/Nicotine (2.5) 80 ± 10* 83 ± 8* -6.8 ± 0.4*

Meca (2)/Methoxsalen (10)/Nicotine (2.5) 15 ± 8 23 ± 10 -2.2 ± 0.5*

*P <0.05 compared to control.

Page 140: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

125

Figure 19. Effects of methoxsalen on the time-course of nicotine’s effects in (A) the tail-flick test, (B) the hot-plate assay and (C) body temperature in mice. Animals were pretreated with either ( ) saline or ( ) methoxsalen (10 mg/kg, s.c) and 30 min later received nicotine at a dose of 2.5 mg/kg. Mice were tested at different times after injection. Each point represents the mean ± SE of 8 to 12 mice. *P <0.05 compared to control.

0

25

50

75

100

% M

PE

0 30 60 90 120 150 180 0

25

50

75

100

% M

PE

0 30 60 90 120 150 180

Time (min)

-8

-6

-4

-2

0

2

²ÞC

0 30 60 90 120 150 180

Time (min)

(A) (B)

(C)

*

*

*

* * *

**

*

*

** * **

**

* *

*

*

**

* *

*

*

% m

ax

imu

m p

oss

ible

eff

ect

% m

ax

imu

m p

oss

ible

eff

ect

Page 141: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

126

B). On the other hand, the effects of nicotine on body temperature were still significant

beyond the 180 min time point (Fig. 19C). Further times were not evaluated.

The prolongation of nicotine’s acute pharmacological effects correlated well with

the prolongation of nicotine plasma concentrations after methoxsalen administration in

mice. Indeed, as shown in Figure 20, there was a corresponding increase in nicotine’s

sensitivity in the tail-flick test to the increased plasma concentration of nicotine after

methoxsalen pretreatment as measured by the time-course curves. A comparable

profile can also be seen in the hot-plate test and body temperature changes (data not

shown).

Page 142: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

127

Figure 20. Relationship between nicotine-induced antinociception in the tail-flick test and plasma levels of nicotine in the ICR mice after methoxsalen treatment. MOP = methoxsalen; Sal: Saline; TF = Tail-flick effect; Conc: plasma concentration of nicotine.

0

50

100

150

200

250

300

350

400

0

20

40

60

80

100

5 15 30 45 60 120 180

Sal/Conc

MOP/Conc

Sal/TF

MOP/TF

Co

ncen

trati

on

(n

g/m

l)

%M

PE

Time (min)

% m

ax

imu

m p

oss

ible

effe

ct

Page 143: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

128

Discussion

A previous study examining nicotine metabolism in C57Bl/6 and DBA/2 mice found that

nicotine is metabolized by two enzymatic sites with the high affinity sites having Km’s of

~10 µM (Siu et al., unpublished) consistent with the nicotine metabolism mediated by

cDNA-expressed CYP2A5 (Km ~ 7.7 ± 0.8 µM; 129/J mouse strain) (Murphy et al.,

2005b). This current study indicates that ICR mice microsomes also contain two

enzymatic sites; the high affinity site had a modestly lower affinity for nicotine (~18 µM)

compared to that observed in C57Bl/6 and DBA/2 mice. Such small differences may be

attributed to unidentified polymorphisms in the ICR mouse strain altering the structure of

the enzyme to a minor extent (Lindberg et al., 1992). The low affinity enzyme site in mice

may be a member of the CYP2B family consistent with the low affinity site in human and

monkey hepatic nicotine metabolism studies (Messina et al., 1997; Schoedel et al.,

2003); the rat CYP2B can also metabolize nicotine (Schoedel et al., 2001).

The present study was the first to demonstrate that the human CYP2A6 inhibitor

methoxsalen (Zhang et al., 2001) can inhibit in vitro nicotine metabolism in mice. The

inhibition was competitive with some indication of mixed inhibition. The inhibition was

potent, as seen in human inhibition studies, with a Ki of 0.32 µM, indicating a 60-fold

higher affinity for the mouse enzyme compared to nicotine (Km = 19 µM, Table 8). These

pharmacokinetic values are similar to human where the Km is estimated to be 65 µM in

human liver microsomes (Messina et al., 1997) and the Ki was found to be 0.01 µM

(Zhang et al., 2001). As methoxsalen may be able to inhibit both CYP2A5 and CYP2B

enzymes in mice (Maenpaa et al., 1994), it is possible that methoxsalen reduces nicotine

metabolism by inhibiting both the high affinity CYP2A5 enzyme site (Murphy et al.,

2005b; Siu et al., 2006) as well as the second low affinity site. The in vitro mouse

pharmacokinetic data (Table 8) suggest that even at the very high concentrations initially

Page 144: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

129

achieved following 2.5 mg/kg s.c. nicotine injections (100-300 ng/ml) that approximately

80% of the metabolism is mediated by the high affinity site kinetic site (Table 8).

Several studies have suggested that methoxsalen acts as a competitive, non-

competitive and/or mechanism-based inhibitor of human CYP2A6 in vitro (Draper et al.,

1997; Koenigs et al., 1997; Maenpaa et al., 1994; Zhang et al., 2001). Methoxsalen

potently inhibited CYP2A6-mediated nicotine metabolism in vivo in humans however a

clear indication of mechanism-based inhibition was not seen using either coumarin

(Kharasch et al., 2000) or nicotine (Sellers et al., 2000) as a substrate.

This is the first study to examine the pharmacokinetic impact of methoxsalen in

vivo in mice. We found that the half-life of plasma nicotine in mice treated with

methoxsalen was significantly longer (approximately doubled) compared to the saline

inhibition group. In addition, as evident in Fig. 16, methoxsalen’s effects increased with

pretreatment times from zero to 60 minutes suggesting that this pretreatment time

enhanced methoxsalen’s ability to inhibit nicotine metabolism. Together with the in vitro

data this suggests that methoxsalen may be acting in vivo in mice as a competitive

inhibitor and through an additional mechanism, possibly mechanism-based inhibition. It

has been shown in vitro that mouse CYP2A5 can metabolize methoxsalen (Maenpaa et

al., 1994). Mechanism-based inhibitors, also known as suicide substrates, are

compounds which are metabolically activated by the enzyme and the metabolite binds

irreversibly to the enzyme inactivating it. Therefore, it is possible that metabolites of

methoxsalen, even if produced by other enzymes such as CYP2B, may inhibit the

CYP2A5 through a variety of mechanisms including mechanism-based inhibition,

competitive inhibition and/or binding to the heme moiety of the enzyme. Together this

suggests that preinhibition with methoxsalen, prior to administration of a substrate such

as nicotine, may lead to enhanced metabolic inhibition via one of these mechanisms.

The similar maximal plasma nicotine levels seen with and without pretreatment with

Page 145: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

130

methoxsalen are consistent with the subcutaneous route of nicotine’s administration, but

predicts a large increase in Cmax if nicotine is given by an oral route where the inhibition

would also block a first pass metabolism.

Our results confirmed that methoxsalen did indeed inhibit the conversion of

nicotine to cotinine both in vitro and in vivo. Pretreatment with methoxsalen increased

the AUC 2.5 fold and prolonged the nicotine plasma levels; it was also associated with a

2-fold increase in the half-life of the drug (Table 9). The nicotine plasma levels remained

above 10 ng/ml in the saline pretreatment for approximately 90 minutes but with

pretreatment with methoxsalen the nicotine plasma levels were still above 10 ng/ml at

180 min. Similar to the marked effects of CYP2A5 inhibition on nicotine metabolism,

there were dramatic enhancement in the pharmacological effects of nicotine (body

temperature and analgesia) after methoxsalen treatment. The effects of methoxsalen on

nicotine were dose-dependent and reached maximal enhancement at a dose of 15

mg/kg. Furthermore, the duration of methoxsalen-induced potentiation of nicotine’s

antinociceptive and hypothermic lasted close to 180 min after drug pretreatment.

Pharmacogenetic studies in humans have showed that smoking behaviours are

affected by polymorphisms in CYP2A6 that alter nicotine metabolism (Ariyoshi et al.,

2002; Malaiyandi et al., 2006; Minematsu et al., 2006; Rao et al., 2000; Schoedel et al.,

2004). For instance, individuals who are slow metabolizers of nicotine (e.g. those with

50% of less activity due to having inactive and/or decreased function alleles) smoked

significantly fewer cigarettes compared to individuals that are normal nicotine

metabolizers (those without variants) (Ariyoshi et al., 2002; Malaiyandi et al., 2006; Rao

et al., 2000; Schoedel et al., 2004). Furthermore, inhibition of orally delivered nicotine in

smokers using methoxsalen (to phenocopy slower metabolism found in those with

defects in CYP2A6) led to increases in plasma nicotine levels, reduction of cravings for

nicotine, and fewer numbers of cigarettes smoked relative to nicotine alone;

Page 146: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

131

methoxsalen alone, in smokers, also increased the latency between lighting the first and

second cigarette (Sellers et al., 2000). Consistent with human data, we have shown that

CYP2A5 levels and in vitro nicotine metabolism rates are associated with nicotine oral

self-administration in male mice (Siu et al., 2006). These findings suggest that mice may

represent a suitable model for studying variation in nicotine metabolism, including the

use of inhibitors to reduce nicotine metabolism.

The main purpose of the current study was to determine the impact of nicotine

metabolism (i.e. inhibition) on nicotine’s acute pharmacological effects in mice.

Methoxsalen prolonged the duration of nicotine-induced antinociception and

hypothermia (2.5 mg/kg) for periods up to 180 min post-nicotine administration.

Although this increase in nicotine’s time course was seen in all responses measured, it

was more dramatic in antinociceptive tests (an additional 150 min compared to the

control group). This differential effect is probably due to the fact that nicotine-induced

antinociception time-course is much shorter than the effects on body temperature.

Furthermore, this prolongation in nicotine’s effects after methoxsalen was associated

with a parallel prolongation of nicotine plasma levels in mice (Figure 20). Consistent

with these present results, we have shown that CYP2A5 levels and in vitro nicotine

metabolism rates are associated with nicotine oral self-administration in male mice (Siu

et al., 2006). These findings suggest that mice may represent a suitable model for

studying variation in nicotine metabolism and its impact on mechanisms of nicotine

dependence, including the use of inhibitors to reduce nicotine metabolism. Although the

study of nicotine’s acute effects is important to the understanding of nicotine

dependence, one must also consider responses after chronic exposure of the drug in

animals. Animal models studying the different aspects of dependence such as

tolerance, withdrawal and reward involve a chronic/repeated exposure of the animals to

nicotine. Assessing the impact of nicotine metabolism in these behavioral models will

Page 147: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

132

enhance our understanding of the molecular mechanisms involved in nicotine

dependence and will facilitate the development of new strategies for smoking cessation

therapies.

These data strongly suggest that variation in the rates of nicotine metabolic

inactivation substantially alter nicotine’s pharmacological effects. These data suggest

that in mice, with very rapid rates of metabolism, small differences in levels or function of

CYP2A5 between different mouse strains may result in substantially differing nicotine

pharmacological effects. Thus pharmacokinetic differences likely contribute to

differences in effects seen between studies employing the use of different mouse

strains.

In conclusion we have shown that the human CYP2A6 inhibitor methoxsalen is

also a potent inhibitor of mouse CYP2A5, both in vitro and in vivo. We have shown that

the pharmacological effects of inhibiting nicotine’s metabolism are profound, illustrating

the differences in resulting effects of nicotine that will occur between individuals and also

between animal strains that differ in levels of CYP2A-mediated nicotine metabolism.

Page 148: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

133

Acknowledgments

The authors greatly appreciate the technical assistance of Tie Han and W Zhang.

Support

This work was supported by National Institute on Drug Abuse grant # DA-05274,

the Centre for Addiction and Mental Health, and the Canadian Institutes for

Health Research grants # MOP53248 and #MOP14173. E Siu acknowledges

the support of the CIHR STPTR scholarship and RFT is grateful for the support

of the Canada Research Chairs program.

Disclosures

EMS and RFT are shareholders in Nicogen, a company focused on novel

treatment approaches involving inhibition of hepatic CYP2A6. No support from

Nicogen was used and no benefit to the company was obtained.

Page 149: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

134

SIGNIFICANCE OF CHAPTER

This chapter demonstrated the application of the inhibition of nicotine metabolism on

nicotine pharmacology in a mouse model. Specifically, the human CYP2A6 inhibitor

methoxsalen significantly inhibited the CYP2A5-mediated nicotine metabolism in mice

both in vitro and in vivo. The inhibition of nicotine metabolism in mice led to a substantial

increase in nicotine’s pharmacological effect (i.e. tail-flick and hot-plate tests).

Page 150: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

135

CHAPTER 5: SELEGILINE (L-DEPRENYL) IS A MECHANISM-BASED INACTIVATOR OF CYP2A6 INHIBITING NICOTINE METABOLISM IN HUMANS AND MICE

Eric C. K. Siu and Rachel F. Tyndale As published in: Journal of Pharmacology and Experimental Therapeutics, 2008, Mar;

324(3):992-999

All of the experiments, data analyses, and writing of the manuscript were performed by

ECK Siu. The study was designed by ECK Siu and RF Tyndale.

Reprinted from Journal of Pharmacology and Experimental Therapeutics, Volume 324, Siu et al., “SELEGILINE (L-DEPRENYL) IS A MECHANISM-BASED INACTIVATOR OF CYP2A6 INHIBITING NICOTINE METABOLISM IN HUMANS AND MICE”, pp 992-999, Copyright (2008) with permission from The American Society for Pharmacology and Experimental Therapeutics. © 2008 The American Society for Pharmacology and Experimental Therapeutics

Page 151: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

136

Abstract

Selegiline (L-deprenyl) is in clinical treatment trials as a potential smoking cessation drug.

We investigated the impact of selegiline and its metabolites on nicotine metabolism. In

mice, selegiline was a potent inhibitor of nicotine metabolism in hepatic microsomes and

cDNA-expressed CYP2A5; the selegiline metabolites desmethylselegiline, L-

methamphetamine, and L-amphetamine, also inhibited nicotine metabolism. Pre-

treatment with selegiline and desmethylselegiline increased inhibition (IC50) in

microsomes by 3.3- and 6.1-fold, respectively. In mice in vivo, selegiline increased AUC

(90.7±5.8 vs. 57.4±5.3 ng • hr/ml, p < 0.05), decreased clearance (4.6±0.4 vs. 7.3±0.3

ml/min, p < 0.05), and increased elimination half-life (12.5±6.3 vs. 6.6±1.4 min, p < 0.05)

of nicotine. In vitro, selegiline was a potent inhibitor of human nicotine metabolism in

hepatic microsomes and cDNA-expressed CYP2A6; desmethylselegiline and L-

amphetamine also inhibited nicotine metabolism. Selegiline pre-incubation increased

inhibition in microsomes (3.7-fold) and CYP2A6 (14.8-fold); the Ki for CYP2A6 was 4.2

µM. Selegiline dose- and time-dependently inhibited nicotine metabolism by CYP2A6 (KI

= 15.6±2.7 µM; kinact = 0.34±0.04 min-1) and the inhibition was irreversible in the

presence of NADPH, indicating that it is a mechanism-based inhibitor of CYP2A6. Thus,

inhibition of mouse nicotine metabolism by selegiline was competitive in vitro, and

significantly increased plasma nicotine in vivo. In humans where selegiline is both a

competitive and mechanism-based inhibitor, it is likely to have even greater effects on in

vivo nicotine metabolism. Our findings suggest that an additional potential mechanism of

selegiline in smoking cessation is through inhibition of nicotine metabolism.

Page 152: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

137

Introduction

Nicotine is the primary psychoactive component in tobacco responsible for the addictive

properties of cigarettes (Henningfield and Keenan, 1993). One action of nicotine is the

binding to nicotinic receptors stimulating dopamine release in the nucleus accumbens,

an area of the brain responsible for reward (Balfour, 2004). In the brain, dopamine is

metabolized by monoamine oxidases A and B (MAO-A and MAO-B) (Youdim et al., 2006)

although MAO-B appears to be the major form (Fowler et al., 1987). The amount of

MAO-B activity in the brain of smokers is ~40% lower compared to non-smokers (Fowler

et al., 1996a) and this inhibition of MAO-B is reversed during long-term smoking

abstinence (Gilbert et al., 2003), suggesting the presence of a MAO-B inhibitor in

tobacco smoke (Khalil et al., 2000). These findings suggest that the reduced metabolism

of dopamine, as mediated by the inhibition of MAO, may enhance the rewarding effects

of nicotine during smoking (Lewis et al., 2007).

Selegiline (L-deprenyl) is a selective and irreversible MAO-B inhibitor used in

conjunction with L-dopa to alleviate symptoms associated with Parkinson’s disease

(Gerlach et al., 1996). Due to its ability to reduce dopamine metabolism, selegiline has

been investigated as a potential therapy for smoking cessation. Several small-scale

studies have shown that selegiline is effective in reducing withdrawal symptoms and

increasing abstinence compared to placebo. For instance, in one study 10 mg of oral

selegiline decreased craving during abstinence and reduced smoking satisfaction during

smoking (Houtsmuller et al., 2002). In another study, 5 mg b.i.d. oral selegiline increased

the trial end point (8-week) 7-day point prevalent abstinence compared with placebo by

three-fold (George et al., 2003). In a third study that used a combination of oral selegiline

and nicotine patch, selegiline plus nicotine patch doubled the 52-week continuous

abstinence rate compared to nicotine patch alone although the difference was not

Page 153: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

138

significant due to small subject numbers (Biberman et al., 2003). Other MAO inhibitors

have also been investigated for their effects on smoking cessation, but due to poor

efficacy (moclobemide) or toxicity (lazabemide) they are no longer studied (McRobbie et

al., 2005).

Selegiline is metabolized to desmethylselegiline and L-methamphetamine, both

of which can be further metabolized to L-amphetamine as well as other minor

metabolites (Fig. 21) (Shin, 1997; Valoti et al., 2000); despite this, there is no evidence

indicating that selegiline is addictive (Schneider et al., 1994). In humans, chronic

treatment with selegiline reduces the metabolism of selegiline and its metabolites,

suggesting that selegiline or its metabolites may inhibit or down-regulate its own

metabolic enzymes (Laine et al., 2000). Selegiline belongs to the acetylene group of

compounds that contain a carbon-carbon triple bond, which are known to be potent

mechanism-based inhibitors (Correia and Ortiz de Montellano, 2005). Since the main

human nicotine metabolizing enzyme CYP2A6 appears to play a role in the metabolism

of selegiline in vitro (Benetton et al., 2007), we examined whether selegiline and its

metabolites (desmethylselegiline, L-methamphetamine, and L-amphetamine) (Shin,

1997) could inhibit nicotine metabolism in vitro in human and mouse hepatic microsomes,

as well by both cDNA-expressed major human nicotine metabolizing enzyme CYP2A6

and mouse CYP2A5 (Murphy et al., 2005b; Siu and Tyndale, 2007a). Genetically slow

CYP2A6 metabolizers have a greater likelihood (1.75-fold) of quitting smoking (Gu et al.,

2000) suggesting that if selegiline inhibits nicotine metabolism, this may be an additional

mechanism through which it reduces smoking. In addition, selegiline could potentially be

used to enhance the efficacies of current nicotine

Page 154: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

139

Figure 21. Selegiline and its metabolites. Selegiline is metabolized to desmethyselegiline and L-methamphetamine, both of which can be metabolized to L-amphetamine (Shin, 1997; Valoti et al., 2000).

Page 155: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

140

replacement therapies, or it could be combined with nicotine as an oral combination

therapy with nicotine for smoking cessation.

Page 156: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

141

Materials and Methods

Animals

Adult male DBA/2 mice, previously characterized in nicotine metabolism studies (Siu and

Tyndale, 2007a), were obtained from Charles River Laboratories Inc. (Saint-Constant,

PQ). Animals were housed in groups of three to four on a 12-hour light cycle and had

free access to food and water.

Reagents

(-)-Nicotine hydrogen tartrate, (-)-cotinine, selegiline (R-(-)-deprenyl hydrochloride), R-(-

)-desmethylselegiline, thiamine hydrochloride, and 5-aminolevulinic acid (ALA) were

purchased from Sigma-Aldrich (St. Louis, MO). L-methamphetamine and L-

amphetamine were purchased from Research Biochemicals Inc. (Natwick, MA). The

internal standard 5-methylcotinine was custom-made by Toronto Research Chemicals

(Toronto, ON). All dosed drugs are expressed as the free base of the drug. Ampicillin

and lysozyme were purchased from BioShop Canada (Burlington, ON). IPTG (Isopropyl-

beta-D-thiogalactopyranoside) was purchased from MBI Fermentas (Burlington, ON).

The monoclonal antibody to human CYP2A6 was purchased from BD Biosciences

(Mississauga, ON). Horseradish peroxidase-conjugated anti-mouse secondary antibody

and enhanced chemiluminescence were purchased from Pierce Biotechnology (Rockfort,

IL). Nitrocellulose membrane was purchased from Bio-Rad Labs (Mississauga, ON).

Autoradiographic film was purchased from Ultident Scientific (St. Laurent, PQ).

Membrane Preparations

Microsomal membranes were prepared from mouse and human livers for in vitro nicotine

metabolism assays as previously described (Siu et al., 2006) and stored at -80°C in

Page 157: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

142

1.15% KCl. The human liver is from the previously characterized K-series liver bank

(Messina et al., 1997) and was a generous gift from Dr. T. Inaba at the University of

Toronto. The cytosolic fractions were acquired during membrane preparation and were

used as a source of aldehyde oxidase. All mouse livers and plasma samples were

collected and frozen prior to 3 pm to minimize circadian effects on CYP2A5 expression.

Membrane protein concentrations were determined with Bradford reagent according to

manufacturer's protocol (Bio-Rad Labs Ltd., Mississauga, ON).

Expression of CYP2A5

Cyp2a5 cDNA vector in Escherichia coli was a generous gift from Dr. Xinxin Ding

(Wadsworth Center, New York State Department of Health, NY) and prepared as

previously described with modifications (Gu et al., 1998). Briefly, E. coli colonies from

ampicillin plates were inoculated in 2 ml of LB broth with 100 µg/ml of ampicillin and

incubated overnight (no more than 16 hours) at 37oC shaking at 200 rpm. The culture

was then diluted (1:100) in TB broth, with final concentrations of 100 µg/ml of ampicillin,

1 mM thiamine, 0.5 ml of ALA, and 1 mM IPTG, and incubated for 48 hours at 25oC

shaking at 150 rpm. Following incubation the culture was centrifuged at 2800xg at 4oC

for 20 min and the pellet was resuspended in 1/20th culture volume of ice-cold TSE

buffer (100 mM Tris, pH 7.4, 0.1 mM EDTA, 0.1 mM dithiothreitol, 1.15% KCl, 110 g/L

sucrose) and 1/20th culture volume of ice-cold water. Lysozyme was then added to a

final concentration of 0.25 mg/ml and this was shaken gently at 4oC for 1 hr followed by

centrifugation at 2800xg at 4oC for 20 min. After centrifugation the pellet was

resuspended in 1/25th culture volume of ice-cold TE buffer. The resuspension was

placed on ice and sonicated in 3 x 30 sec bursts with a Branson Digital Sonifier (Model

S-450D; Branson Ultrasonics, Markham, ON) set at 30% output. The suspension was

Page 158: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

143

spun at 12,000xg at 4oC for 12 min and the supernatant was re-spun at 180,000xg at

4oC for 60 min. The pellet containing the bacterial membrane fraction was resuspended

in 1.15% KCl and stored at -80oC.

Quantification by Immunoblotting of CYP2A5

Immunoblotting was performed for CYP2A5 and the reference lymphoblastoid cDNA-

expressed human CYP2A6 (BD Biosciences, San Jose, CA) essentially as previously

described (Siu et al., 2006). To determine the linear ranges of detection of CYP2A5 and

CYP2A6, in order to quantify the amount of CYP2A5 in the bacterial membrane, the

bacterial membrane fraction and CYP2A6 were serially diluted from 0.25 to 2 µg protein

and from 0.06 to 1.5 pmol, respectively.

In Vitro Nicotine C-Oxidation Assay

The linear conditions for nicotine C-oxidation to cotinine (nicotine metabolism) in DBA

mouse liver microsomes were previously described (Siu et al., 2006) (0.5 mg/ml protein;

15 min). In human liver microsomes, the linear conditions of nicotine metabolism were

obtained under assay conditions of 0.5 mg/ml protein with an incubation time of 20 min.

For CYP2A5, the linear conditions of nicotine metabolism were obtained under assay

conditions of 60 pmol P450/ml, 60 pmol of reductase and cytochrome b5 (Invitrogen,

Carlsbad, CA) with an incubation time of 15 min. For expressed CYP2A6 (containing

P450 reductase and cytochrome b5) (BD Gentest, Woburn, MA), the linear conditions of

nicotine metabolism were obtained under assay conditions of 20 pmol P450/ml with an

incubation time of 15 min. All incubation mixtures contained 1 mM NADPH and 1 mg/ml

of liver cytosol in 50 mM Tris-HCl buffer, pH 7.4 and were performed at 37oC in a final

volume of 0.5 ml. Unless specified, nicotine concentrations used were 10 µM (mouse

liver microsomes), 60 µM (CYP2A5), 20 µM (human liver microsomes), and 100 µM

Page 159: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

144

(CYP2A6). The reactions were stopped with a final concentration of 4% v/v Na2CO3.

After incubation, 5-methylcotinine (70 µg) was added as the internal standard and the

samples were prepared and analyzed for nicotine and metabolites by HPLC system as

described previously (Siu et al., 2006). The limits of quantification were 5 ng/ml for

nicotine and 12.5 ng/ml for cotinine.

In Vitro Inhibition of Nicotine Metabolism

In all in vitro inhibition experiments assays were conducted as above with the addition of

the inhibitors at the same time as nicotine. In experiments with the pre-incubation step,

reactions containing increasing concentrations of the inhibitor were initiated by pre-

warming the mixture for two minutes prior to the addition of NADPH and pre-incubation

of 15 minutes at 37oC prior to addition of nicotine and incubation.

For the NADPH-dependent inactivation experiment, 0 or 100 µM of selegiline

was added to the mixture in the presence or absence of NADPH. Following pre-

incubation, samples were loaded into the Microcon YM-30 centrifuge membrane filters

(Dyck and Davis, 2001) (Millipore, Etobicoke, ON) and centrifuged for at least 30 min at

4oC. Retentates (typically 40-50 µl) were added to fresh reaction mixtures containing

nicotine and the reactions were allowed to proceed.

In Vivo Inhibition of Nicotine Metabolism

Nicotine and selegiline were dissolved in physiological saline (0.9% sodium chloride)

and adjusted to pH 7.4 for use in in vivo studies. All animals were injected

intraperitoneally with nicotine or nicotine plus selegiline (both at 1 mg/kg). The nicotine

dose was chosen based on previous pharmacokinetic studies (Siu and Tyndale, 2007a)

and relevance in behavioural models for nicotine in mice (Marks et al., 1985). The

selegiline dose was based on its relevance in the mouse MPTP-model of Parkinson’s

Page 160: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

145

disease (Fredriksson and Archer, 1995). Blood samples were drawn by cardiac puncture

at various times after the injection. Immediately after collection plasma was prepared by

centrifugation at 3000xg for 10 min and kept at –20˚C until analysis. Total plasma

nicotine levels (free and glucuronides) were measured following deconjugation by β-

glucuronidase at a final concentration of 5 mg/ml in 0.2 M acetate buffer, pH 5.0, at 37oC

overnight. Samples were then analyzed by HPLC.

In Vitro Kinetic and Pharmacokinetic Parameters Analyses and Statistical Analyses

The Michaelis-Menten kinetic parameters Km and Vmax were calculated using Graphpad

Prism (Graphpad Software Inc., San Diego, CA) and were verified by the Eadie-Hofstee

method. The equation used to determine Km and Vmax was v = Vmax [S] / (Km + [S]) (Eqn.

1) where [S] denotes substrate concentration. These kinetic parameters were used to

determine nicotine concentrations used. Statistical analyses of in vitro kinetic parameters

were tested by student’s t-test.

The in vivo pharmacokinetic parameters were determined using non-

compartmental analysis: AUC0-40 was calculated using the trapezoidal rule. Elimination

half-life (t1/2) was estimated by the terminal slope. Since the bioavailability (F) of nicotine

was unknown following intraperitoneal injection in mice, CL (clearance) was determined

as a hybrid parameter CL/F and was calculated as Dose/AUC0-40. The average weights

of the animals were similar (24.4 ± 0.7 g, n=24), therefore an estimated dose of 25 µg

was used for the calculation of CL/F for nicotine. Assessment of in vivo nicotine levels

for the entire time-course was not possible from individual animals due to limited blood

volume; therefore each time point represented data from multiple mice. Due to this

experimental design, pharmacokinetic parameters (e.g. half-life) were estimated by

Page 161: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

146

resampling methods using the PKRandTest software (H. L. Kaplan, Toronto, ON) (Siu

and Tyndale, 2007a).

Page 162: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

147

Results

Inhibition of Nicotine Metabolism by Selegiline and Metabolites

Selegiline and desmethylselegiline had the highest inhibitory activities on nicotine

metabolism in mouse liver microsomes (MLMs) while L-methamphetamine and L-

amphetamine had smaller effects (Fig. 22a). Selegiline also inhibited CYP2A5 to a great

extent while desmethylselegiline, L-methamphetamine, and L-amphetamine had similar

inhibitory effects (Fig. 22b). In human liver microsomes (HLMs), selegiline and L-

amphetamine appeared to show the greatest inhibitory activities on nicotine metabolism

closely followed by desmethylselegiline (Fig. 22c), while L-methamphetamine did not

inhibit cotinine formation (Fig. 22c). As with the three previous enzyme sources,

selegiline caused the greatest inhibition on nicotine metabolism in CYP2A6 (Fig. 22d).

Desmethylselegiline and L-amphetamine showed almost similar inhibition whereas L-

methamphetamine did not alter nicotine metabolism (Fig. 22d), as was seen with HLMs

(Fig. 22c).

The above findings suggested that selegiline had the greatest inhibitory effect on

CYP2A5-mediated nicotine metabolism in vitro in mice followed by selegiline’s three

metabolites. Likewise, selegiline, desmethylselegiline, and L-amphetamine could inhibit

CYP2A6-mediated human nicotine metabolism in vitro whereas L-methamphetamine did

not.

Effects of Selegiline and Desmethylselegiline Pre-incubation on Nicotine Metabolism in

Mouse Liver Microsomes and CYP2A5

As selegiline and desmethylselegiline contain a carbon-carbon triple bond found in

mechanism-based inhibitors (MBI), the effects of pre-incubation of these compounds on

Page 163: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

148

Figure 22. Selegiline and its metabolites inhibited nicotine metabolism. a) MLMs, b) CYP2A5, c) HLMs, and d) CYP2A6 were incubated with 0 (control), 10, and 100 µM of selegiline (SEL), desmethylselegiline (DES), L-methamphetamine (L-MAMP), or L-amphetamine (L-AMP) in the presence of nicotine and analyzed for cotinine formation. Activity remaining for each compound was compared to the corresponding control treatment. Each data point represented the average of 2 to 5 independent experiments. * p < 0.05 compared to control.

Page 164: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

149

nicotine metabolism in MLMs and CYP2A5 were investigated. If pre-treatment increased

inhibition, this could indicate that the inhibitor is either acting as a MBI or is metabolized

to a more potent inhibitor during the pre-incubation. Selegiline dose-dependently

inhibited cotinine formation in both MLMs and CYP2A5; pre-incubation with selegiline

decreased the IC50 in MLMs (3.3-fold) but not in CYP2A5 (Figs. 23a & b).

Desmethylselegiline also dose-dependently inhibited nicotine metabolism in MLMs but

its effect was much weaker in expressed CYP2A5 (Figs. 23c & d). As with selegiline,

pre-incubation with desmethylselegiline enhanced inhibition in MLMs (6.1-fold change in

IC50) but not in CYP2A5 (Figs. 23c & d). This indicated that selegiline and

desmethylselegiline were competitive inhibitors, but not MBIs, of CYP2A5 in vitro; and

other enzymes in MLMs can metabolize selegiline and desmethylselegiline to inhibitors

of greater potency.

Effect of Selegiline on Nicotine Metabolism in vivo in DBA/2 Mice

To determine if selegiline could also inhibit nicotine metabolism in vivo, we treated

DBA/2 mice with selegiline and nicotine. Selegiline co-administration decreased the

clearance of nicotine by ~40%, resulting in 58% greater AUC and almost doubling the

elimination half-life (Fig. 24, Table 11). These results demonstrated that selegiline is an

effective inhibitor of nicotine metabolism in vivo in mice.

Effects of Selegiline and Desmethylselegiline Pre-incubation on Nicotine Metabolism in

Human Liver Microsomes and CYP2A6

To further investigate the effect of selegiline and desmethylselegiline on nicotine

metabolism in humans, HLMs and CYP2A6 were pre-incubated with the inhibitors.

Selegiline dose-dependently inhibited cotinine formation in both HLMs and expressed

Page 165: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

150

Figure 23. Pre-incubation of MLMs, but not CYP2A5, with selegiline increased inhibition of nicotine metabolism. MLMs (a & c) and CYP2A5 (b & d) were pre-incubated (+PRE) for 15 minutes with increasing concentrations (0, 0.3, 1, 3, 10, 30, and 100 µM) of selegiline (a & b) or desmethylselegiline (c & d) prior to addition of nicotine and an additional incubation of 15 minutes. Samples were also incubated with the inhibitors and nicotine without pre-treatment (-PRE). Activity remaining was calculated as % control (no inhibitor, same incubation conditions) within each treatment group. IC50 was determined as the concentration of inhibitor required to decrease nicotine metabolism by 50%. Each data point represented the average of 2 to 5 independent experiments. * p < 0.05 compared to no pre-incubation.

Page 166: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

151

Figure 24. Selegiline decreased nicotine metabolism in vivo in DBA/2 mice. DBA/2 mice were injected with nicotine (1 mg/kg, i.p.) (-SEL) or with nicotine plus selegiline (1 mg/kg, i.p.) (+SEL). Following nicotine injection plasma samples were collected at the indicated time points and analyzed for nicotine levels. Each time point represents mean (± S.D.) of 3 to 6 animals for each treatment; statistical comparisons are listed in Table 1.

Page 167: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

152

Table 11. Pharmacokinetic parameters of plasma nicotine in mice treated intraperitoneally with nicotine (1 mg/kg) or nicotine plus selegiline (1 mg/kg). Results were derived using data (Fig. 3) from three to six animals at each time point.

-SEL (mean ± SD)

+SEL (mean ± SD)

AUC0-40 (ng • hr/ml) 57.4 ± 5.3 90.7 ± 5.8* t1/2 (min) 6.6 ± 1.4 12.5 ± 6.3* CL/F (ml/min) 7.3 ± 0.3 4.6 ± 0.4*

*p < 0.05 compared to –SEL

Page 168: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

153

CYP2A6 (Figs. 25a & b). In contrast to the mouse enzymes, pre-incubation of both

HLMs and CYP2A6 with selegiline decreased the IC50 by 3.7- and 14.8-fold, respectively

(Figs. 25a & b). Desmethylselegiline also dose-dependently inhibited nicotine

metabolism in HLMs (Fig. 25c) but seemed to be a weaker inhibitor of CYP2A6 (Fig.

25d). Pre-incubation with desmethylselegiline did not enhance inhibition in HLMs but

may modestly increase inhibition in CYP2A6 (Figs. 25c & d). The above findings

suggested that selegiline is acting as a MBI of CYP2A6 or that selegiline could be

metabolized by HLMs and CYP2A6 to metabolites that are even more potent inhibitors

than selegiline. Following preincubation, a Ki of 4.2 µM for the competitive inhibition of

CYP2A6 was observed (Fig 26a).

Time-, Concentration-, and NADPH-Dependent Irreversible Inhibition of CYP2A6-

Mediated Nicotine Metabolism

In addition, to determine if the inhibition of CYP2A6-mediated nicotine metabolism by

selegiline is mechanism-based, we first carried out time- and concentration-dependent

inactivation assays. Figure 26b indicates that the cotinine formation decreases with

increasing pre-incubation time and the decreases were dose-dependent. Using a

double-reciprocal plot, the KI and kinact were estimated to be 15.6 ± 2.7 µM and 0.34 ±

0.04 min-1, respectively (Fig. 26c).

To determine if the decrease in nicotine metabolism was due to covalent

modification of CYP2A6 by the metabolically activated selegiline, we used a centrifuge

filtering system that would allow for the removal of the unbound inhibitors from the pre-

incubation mixture (Dyck and Davis, 2001). In the absence of NADPH in the pre-

incubation mixture with selegiline, filter removal of unbound selegiline (-N+I) did not lead

to a significant decrease in enzyme activity compared to control (-N-I) (Fig. 26d),

indicating a requirement for the cofactor NADPH. In contrast, in the presence of NADPH

Page 169: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

154

Figure 25. Pre-incubation of human hepatic microsomes and CYP2A6 with selegiline increased inhibition of nicotine metabolism. HLMs (a & c) and CYP2A6 (b & d) were pre-incubated (+PRE) for 15 minutes with increasing concentrations (0, 1, 3, 10, 30, and 100 µM) of selegiline (a & b) or desmethylselegiline (c & d) prior to addition of nicotine and an additional incubation of 15 or 20 (HLMs) minutes. Samples were also incubated with the inhibitors and nicotine without pre-treatment (-PRE). Activity remaining was calculated as % control (no inhibitor, same incubation conditions) within each treatment group. IC50 was determined as the concentration of inhibitor required to decrease nicotine metabolism by 50%. Each data point represented the average of 2 to 5 independent experiments. * p < 0.05 compared to no pre-incubation.

Page 170: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

155

Figure 26. Selegiline inhibited CYP2A6-mediated nicotine metabolism in a time-, concentration-, and NADPH-dependent manner. a) A Dixon plot of competitive inhibition: CYP2A6 were pre-incubated with increasing concentrations of selegiline (0, 0.5, 1, 2, and 4 µM) for 15 minutes. Following pre-incubation 50, 100, or 200 µM of nicotine was added to each of the samples and reactions were carried out for an additional 15 min. b) CYP2A6 was pre-incubated with increasing concentrations of selegiline (0, 1, 2.5, 10, 20, 40 µM) for 0, 1, 3, and 5 minutes. Following pre-incubation nicotine was added to each sample and the reactions were carried out for an additional 15 minutes. Data was plotted as the natural log (Ln) of % activity remaining against the pre-incubation time. Regression lines represent the initial slope of the inhibition curve (Ghanbari et al., 2006). c) Double reciprocal plot of the rate of inactivation (kobs) of cotinine formation as a function of the concentration of selegiline. Data represents duplicate experiments. d) CYP2A6 was pre-incubated without NADPH or inhibitors (-N-I), without NADPH but with 100 µM of selegiline or desmethylselegiline (-N+I), or with NADPH and 100 µM of selegiline or desmethylselegiline (+N+I) for 15 minutes in the absence of nicotine. This was followed by removal of the unbound inhibitors through centrifugal filtration. Retentates were added to fresh reaction mixture containing nicotine and incubated for an additional 15 minutes. Data were averaged from 2 to 3 experiments. * p < 0.05 compared to the -N+I treatment group.

Page 171: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

156

during pre-incubation (+N+I), filter removal of unbound selegiline did not lead to

significant recovery of cotinine formation activity, suggesting that CYP2A6 was

inactivated by metabolic activation of selegiline. Similar findings were also seen with

desmethylselegiline (Fig. 26d). The above findings demonstrated that the inhibition of

CYP2A6-mediated nicotine metabolism by selegiline is irreversible in the presence of

NADPH and that selegiline is a MBI of CYP2A6.

Page 172: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

157

Discussion

It is estimated that 20% of adults in the US are current smokers (National, 2007).

Previous studies demonstrated that selegiline was an effective aide in smoking

cessation (Biberman et al., 2003; George et al., 2003). At least four clinical trials are

currently underway to evaluate the effectiveness of selegiline as a potential therapy in

smoking cessation (www.clinicaltrial.gov). Selegiline can prevent dopamine metabolism

by irreversibly inhibiting MAO-B (Youdim, 1978). Selegiline also appeared to be able to

inhibit its own metabolism in vivo (Laine et al., 2000), possibly via inhibition of CYP2A6.

In the present study we showed that nicotine metabolism in both mice and human liver

microsomes and expressed enzymes could be inhibited by selegiline,

desmethylselegiline, L-methamphetamine (only in mice), and L-amphetamine. In

addition, we also demonstrated that selegiline inhibited nicotine metabolism in vivo in

mice. More importantly, we showed that selegiline is a mechanism-based inhibitor of

human CYP2A6.

In all in vitro systems tested in our study, selegiline (and desmethylselegiline in

MLMs) inhibited nicotine metabolism with the greatest potency compared to its

metabolites. In mice, a greater level of inhibition was seen in MLMs compared to the

expressed CYP2A5. However, the enhanced inhibition of nicotine metabolism following

selegiline pre-incubation in MLMs, not observed with CYP2A5, suggested that selegiline

is unlikely to be a MBI of CYP2A5. The greater inhibition in MLMs (compared to

expressed CYP2A5; and with pre-incubation in MLMs) may be due to the formation of

more potent inhibitory metabolite(s) of selegiline. None of the tested selegiline

metabolites inhibited CYP2A5 with greater potency compared to selegiline. However,

another metabolite that is produced in mice is selegiline-N-oxide (Levai et al., 2004). It is

possible that in addition to the compounds tested here, that selegiline-N-oxide, produced

by other enzymes in MLMs, can also inhibit CYP2A5-mediated nicotine metabolism.

Page 173: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

158

For both HLMs and CYP2A6, pre-incubation with selegiline led to an increase in

inhibition potency. This enhancement in inhibition was more dramatic in CYP2A6

compared to HLMs, which might reflect the metabolism of selegiline in HLMs by other

CYPs such as CYP2B6 and CYP2C8 (Benetton et al., 2007), thus reducing its

availability to inhibit CYP2A6. More importantly, the inhibition of CYP2A6, in addition to

being competitive, was likely mechanism-based. This was supported by the findings that

in the time- and concentration-dependent inhibition study, the decrease in nicotine

metabolism was enhanced with increasing inhibitor pre-incubation time and with

increasing inhibitor concentration. Furthermore, the inhibition of CYP2A6 by selegiline

was irreversible in the presence of NADPH, suggesting that selegiline was metabolically

activated by CYP2A6 and the reactive intermediate formed a covalent linkage with the

holoenzyme. The general mechanism of acetylene-mediated destruction of P450s

appears to be the direct N-alkylation of the porphyrin (Correia and Ortiz de Montellano,

2005). However the precise mechanism by which selegiline irreversibly inhibited

CYP2A6 activity was not determined. The minor inhibition seen in the presence of

selegiline without NADPH was likely due to residual inhibitor left in the retentate

transferred to the fresh nicotine-containing reaction mixture.

Since desmethylselegiline also contains the carbon-carbon triple bond and

irreversibly inhibited MAO-B and decreased its activity by up to 65% in vivo (Heinonen et

al., 1997), we also characterized its effects on nicotine metabolism in both species. In

mice in the absence of pre-incubation, desmethylselegiline inhibited nicotine metabolism

in MLMs to the same extent as selegiline. The greater inhibition by desmethylselegiline

in MLMs (compared to CYP2A5), and the potentiation of inhibition with pre-incubation, is

likely due to the production of more potent metabolite(s) of desmethylselegiline by other

enzymes in MLMs. The only known metabolites of desmethylselegiline are L-

amphetamine and its hydroxylated products (Shin, 1997); however, L-amphetamine had

Page 174: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

159

weaker inhibitory activities. Thus the specific desmethylselegiline metabolite (e.g.

potentially N-oxides) that may be contributing to the inhibition in MLMs, but not in

CYP2A5, is unknown.

Interestingly in HLMs pre-incubation with desmethylselegiline did not increase

the inhibition potency but it appeared to show minor effects on CYP2A6. One possibility

is that desmethylselegiline had weaker affinity for CYP2A6 compared to other enzymes

present in HLMs (i.e. inhibiting or metabolized by other enzymes) and thus was less

available to inhibit nicotine metabolism by CYP2A6. The inhibition of CYP2A6 by

desmethylselegiline could potentially be mechanism-based since in the presence of

NADPH the inhibition of nicotine metabolism was irreversible.

In our study L-methamphetamine inhibited nicotine metabolism in mice but not in

humans. L-amphetamine inhibited nicotine metabolism almost to the same extent as

selegiline in CYP2A5 and HLMs but had the weakest effect in CYP2A6. Again, it is

possible that L-amphetamine was metabolized to more potent inhibitors of nicotine

metabolism by other enzymes in HLMs. The effects of the D-isomers of

methamphetamine and amphetamine on CYP2A5 or CYP2A6 activities are unknown;

however, one study found that racemic (D,L-) amphetamine was a weak inhibitor of

CYP2A6 and a much weaker inhibitor of CYP2A5 (Rahnasto et al., 2003).

We and others have previously demonstrated that nicotine, at relevant

pharmacological concentrations, is metabolized essentially exclusively by CYP2A5 in

mice (Murphy et al., 2005b; Siu and Tyndale, 2007a). We have also shown that

subcutaneous administration of the CYP2A5/6 inhibitor methoxsalen significantly

inhibited nicotine metabolism when nicotine was given subcutaneously and this increase

in nicotine plasma levels subsequently increased the pharmacological actions of nicotine

(Damaj et al., 2007b), demonstrating the value of mice in studying nicotine metabolism

and pharmacology. Since selegiline inhibited CYP2A5-mediated nicotine metabolism in

Page 175: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

160

MLMs and CYP2A5, we determined its effect on nicotine clearance in mice in vivo. We

treated mice with nicotine and selegiline intraperitoneally as this route mimics oral

delivery of both drugs in animals, the route of choice for a novel smoking cessation

product. Administration of selegiline together with nicotine caused almost a 40%

decrease in clearance of nicotine and doubled its elimination half-life. These data

suggested that selegiline and its metabolites can act as competitive inhibitors of nicotine

in vivo even though selegiline is not a MBI of CYP2A5. It is likely that in humans,

selegiline can decrease the first-pass metabolism and elimination half-life of nicotine via

mechanism-based inactivation as well as competitive inhibition of CYP2A6. Genetically

slow CYP2A6 metabolizers have an increased likelihood of quitting smoking (Gu et al.,

2000) suggesting that inhibition of CYP2A6 by selegiline may contribute to selegiline’s

ability as a smoking cessation therapeutic agent.

Our findings suggest that oral selegiline may also be combined with oral nicotine

to create an orally bioavailable and clinically effective form of nicotine due to the high

hepatic extraction of selegiline (Heinonen et al., 1994) (suggesting that selegiline will be

able to rapidly inhibit CYP2A6 prior to reaching the systemic circulation) and the

relatively smaller Ki and KI of selegiline towards nicotine by CYP2A6 (4.2 and 15.6 µM

respectively) compared to the Km for nicotine (~65 µM) (Messina et al., 1997). Currently

there is no pill form of nicotine replacement therapy as the higher doses required to

overcome first-pass metabolism may cause nausea and diarrhea due to gastrointestinal

irritation (Benowitz et al., 1991b). Oral administration of even low doses (4 mg) of

nicotine to those genetically deficient in CYP2A6 increased nicotine AUC by over 3-fold

(Xu et al., 2002), suggesting that inhibition of nicotine metabolism can make nicotine

orally bioavailable. We have also shown that the combination of the CYP2A6 inhibitor

methoxsalen and nicotine significantly increased the mean plasma level of nicotine in

humans as well as decreased the number of cigarettes smoked, reduced craving, the

Page 176: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

161

number of puffs, the inhalation intensity, and breath carbon monoxide levels, a

biomarker of smoke inhalation (Sellers et al., 2000). Furthermore, inhibition of CYP2A6

by methoxsalen also re-routed the tobacco-specific pro-carcinogen NNK to NNAL and

glucuronidation pathways (Sellers et al., 2003a), indicating the additional potential

benefit for harm-reduction.

Here we have shown that selegiline is a mechanism-based inhibitor of human

CYP2A6, the major enzyme responsible for nicotine metabolism. In mice, where

selegiline and its metabolites are competitive inhibitors, selegiline substantially reduced

nicotine metabolism in vivo resulting in prolonged elevated nicotine plasma levels. These

findings suggest that in humans, where selegiline is a competitive and MBI of CYP2A6-

mediated nicotine metabolism, selegiline can potentially be an effective inhibitor of

nicotine metabolism.

Page 177: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

162

Acknowledgments

We would like to thank Dr. Sharon Miksys for the assistance in the development of the in

vivo protocol and Dr. Bin Zhao for assistance with HPLC. We would also like to thank

members of the Tyndale laboratory for critical review of this manuscript. RFT is a

shareholder and chief scientific officer of Nicogen Inc., a company focused on the

development of novel smoking cessation therapies; no funds were received from

Nicogen for these studies nor was this manuscript reviewed by other people associated

with Nicogen.

Page 178: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

163

SIGNIFICANCE OF CHAPTER

In this study we found selegiline to be an inhibitor of CYP2A5-mediated in vitro nicotine

metabolism. Selegiline also inhibited in vivo nicotine metabolism in mice. More

importantly, in humans the pre-incubation of hepatic microsomes and cDNA-expressed

CYP2A6 with selegiline enhanced the inhibition nicotine metabolism. This with additional

studies showing that selegiline irreversibly inhibited CYP2A6 under condition of

metabolic activation demonstrated that selegiline to be a MBI of CYP2A6-mediated

nicotine metabolism. The latter may represent an additional mechanism by which

selegiline mediates smoking cessation.

Page 179: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

164

CHAPTER 6 GENERAL DISCUSSION

Prior to the approval of a candidate drug for use in humans, the compound must first be

tested for its pharmacological and toxicological effects in vivo in animal models. In this

thesis I examined the mouse as a potential model of nicotine metabolism and its

application in studying inhibitors and the consequence of genetic variation in nicotine

metabolism and pharmacology. In order to determine the suitability of the mouse as an

experimental model, a fundamental requirement is that the metabolism of nicotine to

cotinine occurs similarly to that seen in humans (e.g. the reaction is primarily catalyzed

by a homologous enzyme with similar activities). We were interested in the mouse as an

experimental model to study nicotine metabolism based on several reasons. First, a

preliminary quantitative trait loci study in mice suggested the possible involvement of

Cyp2a5 in nicotine self-administration (Wildenauer et al., 2005). Second, mouse

CYP2A5 shares high amino acid sequence similarity to human CYP2A6. Third, the

pharmacodynamics of nicotine in several inbred mouse strains have been well

established (Collins et al., 1988; Faraday et al., 2005; Grabus et al., 2005; Marks et al.,

1985) and this may be combined with nicotine metabolism data to study the contribution

of genetic variation to nicotine pharmacology. Fourth, mice can be genetically modified

(e.g. introducing human cytochrome P450 genes). Finally, mice are small, easily

maintained, and are economical compared to other animal species.

Section 5 Metabolism Topicss

Section 5.1 In Vitro Metabolism

Section 5.1.1 In Vitro Nicotine Metabolism

Page 180: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

165

The in vitro nicotine metabolism data from our studies are in accordance with similar

studies published during this period. The apparent Kms of cotinine formation in the

hepatic microsomes from DBA/2, C57BL/6 (Siu and Tyndale, 2007a), and the F2 hybrid

(C57BL/6 x ST/bJ) mice (Siu et al., 2006) are similar to the Km of the cDNA-expressed

CYP2A5 (from the 129/J strain) (range 8 to 11 µM) (Murphy et al., 2005b). The Kms

observed here are lower compared to the outbred ICR mice (~19 µM) (Damaj et al.,

2007b). Another study found the Km for cotinine formation to be much higher in the

CD2F1 mice (~68 µM) (Raunio et al., 2008). It is likely that polymorphisms in CYP2A5

account for the variation in the observed Kms. For example, substitution of amino acid

207 from glycine (polar, uncharged) to proline (non-polar, hydrophobic) substantially

increased the Km (~100-fold) of CYP2A5 for coumarin 7-hydroxylation, possibly through

structural alteration increasing the size of the substrate binding pocket (Juvonen et al.,

1993). The Cyp2a5 nucleic acid sequences of ICR and CD2F1 mice and further kinetic

studies are required to confirm this hypothesis. Also, one cannot rule out the contribution

of differences in experimental procedure. Nevertheless, compared to humans, the Kms of

cotinine formation in the examined mouse strains (cDNA-expressed CYP2A5 and

hepatic microsomes; range 8 to 68 µM) largely overlap with those seen in humans

(cDNA-expressed CYP2A6 and hepatic microsomes; range 13 to 162 µM) (Messina et

al., 1997; Nakajima et al., 1996b; Yamazaki et al., 1999), suggesting similar enzymatic

activities, at least for cotinine formation.

The DBA/2 and C57BL/6 mice also have higher Vmaxs for nicotine metabolism

compared to the CD2F1 mice (Raunio et al., 2008) and this could be due to several

reasons. First, between-strain variation in the cis-acting regulatory elements and/or

trans-acting regulatory factors can affect both Cyp2a5 transcriptional and post-

transcriptional processes, thereby altering protein levels. For example, the human

Page 181: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

166

CYP2A6*9 contains a polymorphism in the TATA box resulting in decreased

transcription (Kiyotani et al., 2003; Pitarque et al., 2001). Hepatic nuclear factor-4α

(HNF-4α) is a major transcriptional factor for CYP2A6 and Cyp2a5 (Kamiyama et al.,

2007; Ulvila et al., 2004) and bioinformatic database showed that both human and

mouse HNF-4α are highly polymorphic (MGI 4.12, Ensembl). Other regulatory factors

that can influence CYP2A6 expression include PXR, PPAR-γ, ER, C/EBPα, and hnRNP-

A1 (Christian et al., 2004; Di et al., 2009; Higashi et al., 2007a; Pitarque et al., 2005).

Genetic variations in HNF-4α or other regulatory elements could potentially alter the

expression of Cyp2a5. Second, polymorphisms affecting various steps in the P450

catalytic cycle and the interaction of the P450 oxidase with heme or oxidoreductase may

also affect the rate of substrate metabolism. However, such a change would likely alter

the Km to some extent as well since Km is defined as 1

21

k

kk +− where k2 is the turnover

number (i.e. the catalytic step) that is dependent on the overall efficiency of the catalytic

cycle (after taking into consideration the total number of catalytic sites). For instance,

CYP2A6*6 contains an amino acid substitution at position 128 where arginine (polar,

basic) is replaced with glutamine (polar, uncharged). The substitution resulted in a 10-

fold decrease in Vmax and 5-fold increase in Km for coumarin metabolism, and this was

due to decreased heme binding and disruption of the holoprotein structure (Kitagawa et

al., 2001). Random mutagenesis in CYP2A6 found that substitution of residue 476 from

lysine (polar, basic) to glutamine acid (polar, acidic) caused a 20-fold decrease in kcat

(=k2, and by extension Vmax) but only a 25% decrease in Km for coumarin hydroxylation,

a result thought to be due to reduced electron transfer efficiency (Kim et al., 2005).

Our in vitro studies demonstrated that, as in humans, there exists a second

nicotine metabolizing enzyme in mice. However, the catalytic efficiency of this enzyme

for cotinine formation is at least 10-fold lower compared to CYP2A5, suggesting that this

Page 182: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

167

enzyme does not contribute significantly to the overall metabolism of nicotine under

normal pharmacological levels/exposures. At the moment it is difficult to speculate which

cytochrome P450 is contributing to the minor formation of cotinine as inhibitory

antibodies against human CYP2B6 and CYP2D6 did not alter nicotine metabolism in

mice hepatic microsomes (Fig. 7A).

Section 5.1.2 In Vitro Cotinine Metabolism

We have shown, from in vitro kinetic and inhibitory antibody studies that cotinine is

metabolized to 3-HC and this process appeared to be mediated almost entirely by

CYP2A5 (Figs. 6 & 7B). This finding parallels those in humans in which CYP2A6 is

solely responsible for the formation of 3-HC (Dempsey et al., 2004; Nakajima et al.,

1996a). The affinity for 3-hydroxylation of cotinine is much higher in mice (~10 to ~50 µM)

hepatic microsomes compared to humans (~235 µM) (Nakajima et al., 1996a;

Yamanaka et al., 2004). Genetic variation may contribute to the difference in the affinity

between mouse strains. Overall, our findings show that mice exhibit similar

characteristics in the metabolism of nicotine and its primary metabolite cotinine, as

observed in humans.

Section 5.1.3 Effect of a CYP2A5 Polymorphism on Substrate Metabolism

It is interesting to note that despite the reported amino acid and coumarin hydroxylase

activity differences in CYP2A5 between DBA/2 (Val117; high COH activity) and C57BL/6

(Ala117; low COH activity) (both amino acids are non-polar and hydrophobic) (Lindberg et

al., 1992), there is no difference in their cotinine formation capacity as measured by Km

Page 183: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

168

in our study. In contrast, the Km for 3-HC formation is five-fold higher in DBA/2 mice.

Previous studies have also demonstrated that amino acid 117 determines whether

CYP2A5 possesses 7β- or 2α-hydroxylase activity for dehydroepiandrosterone (Uno et

al., 1997). In CYP2A6, mutagenesis of residue 117 from valine to alanine in the cDNA-

expressed protein did not affect its Km but resulted in a substantial drop in Vmax for

coumarin hydroxylation (He et al., 2004b; Yano et al., 2005). Although crystallography

studies of the CYP2A6 crystal structure coupled with coumarin showed that amino acid

117 is in close proximity with coumarin (He et al., 2004b; Yano et al., 2005), the

aforementioned effect was hypothesized to be an indirect role of residue 117 in

coumarin hydroxylation through modulation of the function of surrounding amino acids

potentially involved in proton/electron transfer and reaction stabilization (He et al.,

2004b). It may be that these observations are restricted only to the effect of valine and

alanine on coumarin. As yet this particular mouse polymorphism has not been found in

the general human population and the impact of this alteration on nicotine and cotinine

metabolism in CYP2A6 has not been studied. The high sequence similarity of CYP2A5

with CYP2A6 suggests that residue 117 of CYP2A5 should also be located in the

substrate binding pocket and potentially be in contact with the substrate. It is possible

that the changes in amino acid 117 in mouse CYP2A5 may affect binding and catalytic

activity depending on the substrate.

Section 5.2 In Vivo Metabolism

Our data demonstrated that the in vivo clearance of nicotine in mice following

subcutaneous administration is rapid (t1/2 ~8 to 9 min) consistent with findings in these

mice following intraperitoneal injection (~6 to 7 min) (Petersen et al., 1984). The half-life

of plasma nicotine is longer at 14 minutes in the ICR strain. This is consistent with in

Page 184: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

169

vitro data showing lower nicotine metabolic efficiency (Vmax/Km) in the ICR (0.01) mice

compared to the DBA/2 (0.05) and C57BL/6 (0.04) mice. The elimination of plasma

nicotine in mice is monophasic compared to the biphasic elimination in humans; this

could be due to the much more rapid rate of metabolism compared to distribution of

nicotine following absorption in mice. The percentage recoveries of nicotine and its

metabolites in the urine of these mice were not measured. However, urinary analysis in

the CD2F1 strain indicated that nicotine represented <1% of the systemic dose 24 hours

following its oral administration (Raunio et al., 2008), whereas in humans (normal

CYP2A6 activity), the recovery is higher (~8 to 10%) (Benowitz and Jacob, 1994; Meger

et al., 2002). The difference could be due to the high activity of CYP2A5 as well as the

rapid hepatic blood flow in mice (four-times faster) compared to humans (Bischoff et al.,

1971; Boxenbaum, 1980), resulting in the rapid metabolic clearance of nicotine.

The results from our study, and from others, indicate that in mice cotinine is the

primary metabolite of nicotine in plasma and has a significantly longer half-life (e.g. 24 to

51 min) compared to nicotine (DBA/2; C57BL/6; C3H) (Figs. 3 & 5A) (Petersen et al.,

1984). In the urine of CD2F1 mice, cotinine represented ~3-4% of the systemic nicotine

dose (Raunio et al., 2008). The above findings are compared with humans in which the

half-life of cotinine is much longer (>10 hours) (Benowitz and Jacob, 1994) and its

urinary recovery is ~10-15% of an oral dose of nicotine (Meger et al., 2002).

In mice (and in humans) unconjugated 3-HC is the predominate metabolite of

nicotine that is accumulated in the urine (77% in CD2F1 mice; 33-40% in humans)

(Benowitz and Jacob, 1994; Meger et al., 2002; Raunio et al., 2008). We found that in

the DBA/2 and C57BL/6 mice the clearance of 3-HC is slower compared to cotinine

whereas in humans, 3-HC clearance is formation dependent (Dempsey et al., 2004).

This may be due to differences in the speed of formation (e.g. slower in mice) and renal

clearance of 3-HC considering metabolic clearance (i.e. glucuronidation) does not

Page 185: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

170

contribute to the removal 3-HC in mice (Ghosheh and Hawes, 2002) and only plays a

small role (up to ~25%) in the clearance of 3-HC in humans (Nakajima and Yokoi, 2005).

One possibility is that the renal clearance of 3-HC is influenced by genetic differences.

For instance, a human twin study found the genetic component accounted for up to 61%

of the variance in the renal clearance of nicotine and cotinine (clearance of 3-HC not

examined) (Benowitz et al., 2008). The genetic variation was suggested to be

contributed by the variation in the active secretory or reabsorptive components for both

nicotine and cotinine, although this remains to be determined for 3-HC. Another

possibility is that CYP2A5 and CYP2A6 produce different isomers of 3-HC which have

different affinities for renal excretion, but this remains to be verified.

In our studies we did not examine the kinetics of nicotine N-oxide and cotinine N-

oxide formation. Raunio et al (Raunio et al., 2008) found that cotinine N-oxide was the

second most abundant mouse urinary metabolite (~16%) followed by nicotine N-oxide,

which was found at lower levels (~3-5%). In humans, the urinary recovery of nicotine N-

oxide and cotinine N-oxide are 4-7% and 2-5%, respectively.

Section 5.2.1 Biomarker of Nicotine Metabolism

Many studies in humans have used ratios of nicotine metabolites as indicators of

CYP2A6 activity as majority of nicotine is metabolized by CYP2A6 (Nakajima et al.,

1996a), and it is the only enzyme that metabolizes cotinine to 3-HC (Nakajima et al.,

1996a). For example, salivary and plasma 3-HC/cotinine ratios have been found to be

highly correlated with oral nicotine clearance (first-pass metabolism and systemic

clearance) (Dempsey et al., 2004). The 3-HC/cotinine ratio following systemic

(intravenous) administration of nicotine is also indicative of CYP2A6 genotype as

grouped by activity (Benowitz et al., 2006b). In addition, the urinary and plasma 3-

Page 186: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

171

HC/cotinine ratios are highly correlated within subjects and also with the clearance of

nicotine by the cotinine pathway, suggesting that urinary 3-HC/cotinine ratio may be

used as a marker of CYP2A6 activity (Benowitz et al., 2006b). In addition, those without

CYP2A6 have no 3-HC production making the ratio both selective and sensitive.

Levels of nicotine metabolites (particularly in urine) in mice can also potentially

be used as indicators of CYP2A5 activity in inhibition studies; this may minimize

between-animal variations when studying multiple inhibitors and reduce the number of

animals needed. We showed that administration of methoxsalen to the ICR mice

significantly increased nicotine and decreased cotinine plasma levels (Fig. 15) although

metabolite ratios were not measured in either plasma or urine. In the CD2F1 mice

(Raunio et al., 2008), urinary metabolite ratios were examined as a measure of CYP2A5

activity. Following in vivo administration of methoxsalen, 24-hour total urinary 3-HC and

cotinine N-oxide levels (the two most abundant metabolites) decreased and nicotine

increased leading to significant decreases in 3-HC/nicotine and cotinine N-oxide/nicotine

ratios (Raunio et al., 2008). However, since the production of both 3-HC and cotinine N-

oxide requires another metabolic step after cotinine formation, the use of nicotine as the

denominator in the ratio may increase variability. Furthermore, due to the high variability

in the level of urinary metabolites (Raunio et al., 2008), the metabolite ratio may be less

sensitive to detecting the effect of weaker inhibitors.

Section 5.3 Mice (CYP2A5) as a Model to Study CYP2A6 Inhibition

Apart from nicotine metabolism, with respect to the structural/functional differences

between CYP2A5 and CYP2A6, inhibitor and modeling studies showed that CYP2A5

possess a larger active site compared to CYP2A6 (Juvonen et al., 2000; Poso et al.,

2001) and this has been suggested to be important for the relative potency of inhibitors

Page 187: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

172

(Juvonen et al., 2000). For instance, compounds with the bulkier lactone ring structure

were demonstrated to be better inhibitors of CYP2A5 than CYP2A6 (Juvonen et al.,

2000). However, subsequent experiments showed substantial differences in the

inhibition efficacies between compounds of similar sizes, suggesting that the size of the

active site may play a secondary role to the chemical properties of the inhibitors and

their interactions with the binding site (Rahnasto et al., 2003).

Section 5.3.1 Methoxsalen

Methoxsalen is one of the most thoroughly studied inhibitors of CYP2A5 and CYP2A6. In

mice, methoxsalen decreased NNK-induced lung tumor formation (Section 2.3.5)

(Miyazaki et al., 2005; Takeuchi et al., 2003). In humans, methoxsalen increased plasma

nicotine level and decreased smoking behaviours (Section 2.3.3.1) (Sellers et al., 2002;

Sellers et al., 2000; Tyndale et al., 1999) as well as decreased NNK activation (Section

2.3.5) (Sellers et al., 2003a).

Our data showed that in vitro methoxsalen acted as a competitive and potentially

a mixed-type inhibitor of nicotine metabolism (Fig. 14). This is consistent with the

compound acting as a mixed-type (competitive and non-competitive) inhibitor of

CYP2A5-mediated coumarin metabolism (Visoni et al., 2008). Methoxsalen is also a MBI

of CYP2A5 but the mechanism through which this occurs has not been studied (Visoni et

al., 2008). Results from another study suggested that the active metabolite(s) of

methoxsalen covalently modifies the apo-protein of CYP2A5 (Mays et al., 1990). During

in vivo behavioural tests, pre-treatment of methoxsalen 30 to 60 minutes prior to

subcutaneous nicotine injection in the ICR mice led to the greatest change in the anti-

nociceptive and hypothermic effects of nicotine (tested 45 minutes after nicotine injection;

Fig. 16). This also led to a substantial change in nicotine pharmacokinetics as

Page 188: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

173

demonstrated by a ~30% decrease in its half-life (Fig. 15). However, since we only

examined nicotine pharmacokinetics following 30 minutes (but not 60 minutes) pre-

treatment with methoxsalen, one cannot be definitive regarding the direct correlation

between behavioural changes and plasma nicotine levels across time. The

pharmacokinetics of methoxsalen in the ICR mice has not been studied and therefore it

is uncertain which inhibition mechanism (i.e. competitive and/or mechanism-based

inhibition) is responsible for the in vivo inhibition (Visoni et al., 2008). It is also unknown

if metabolites of methoxsalen in mice are capable of inhibiting CYP2A5.

In humans, methoxsalen acts as both a competitive and a MBI of CYP2A6 in

vitro (Section 2.3.3.1) (Draper et al., 1997; Koenigs et al., 1997; Maenpaa et al., 1994;

Zhang et al., 2001). The mechanism through which methoxsalen inactivates CYP2A6 is

unclear but it may be due to denaturation of the apo-protein, heme modification, or

covalent binding of methoxsalen to the active site of CYP2A6 (Koenigs et al., 1997). In

vivo, pre-administration of methoxsalen at clinical doses resulted in a statistically

significant decrease in the Cmax of the coumarin metabolite 7-hydroxycoumarin and its

AUC (Kharasch et al., 2000) suggesting inhibition of coumarin metabolism; however, the

pharmacokinetics of coumarin was not determined.

Section 5.3.2 Selegiline

We showed for the first time that selegiline, and to some extent, its metabolites, are

capable of inhibiting CYP2A6- and CYP2A5-mediated nicotine metabolism. The

inhibition seemed to be slightly more potent for CYP2A6 than CYP2A5. However, the

mechanisms through which selegiline inhibits these enzymes differ as selegiline is an in

vitro MBI of CYP2A6 but not of CYP2A5. In vivo in mice, co-administration of selegiline

with nicotine led to a significant increase in the plasma half-life of nicotine but had no

Page 189: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

174

effect on Cmax (Fig. 24). It is likely that both selegiline and its metabolites contribute to

the in vivo inhibition observed in mice. A recent study of the disposition kinetics of

selegiline and its metabolites in NMRI mice showed that following intraperitoneal

injection of selegiline, selegiline was the primary compound found in the plasma (Magyar

et al., 2007). In contrast, when the drug was administered orally, the bioavailability of

selegiline was substantially lowered (by more than three-fold) (Magyar et al., 2007). The

reduced selegiline bioavailability was not thought to be due to incomplete oral absorption,

suggesting instead that the majority of selegiline has reached hepatic circulation and the

liver (i.e. inhibiting CYP2A5) prior to reaching the systemic circulation, in contrast to

intraperitoneal administration. Thus, these data suggests that intraperitoneal

administration (in our study) may be underestimating the impact of selegiline on the in

vivo inhibition of CYP2A5-mediated nicotine metabolism in mice compared to oral

administration. Nevertheless, confirmation is needed as our study was done in the

DBA/2 mice and strain differences may alter the pharmacokinetics of selegiline. Ideally

in clinical setting both the nicotine and inhibitor would be given orally and these data

suggest that an even greater inhibition of nicotine metabolism may be seen with oral

selegiline administration alone.

In humans, the effect of selegiline on nicotine metabolism in vivo has not been

examined. The fold-reduction in intrinsic clearance by the MBI can be roughly estimated

by the equation: fold reduction in deg

deg

int

][

][

k

KI

kIk

CLI

inact

+

×+

= where kdeg is the degradation

rate constant of the enzyme (CYP2A6) and [I] is the concentration of the inhibitor at the

active site, the latter of which can be determined from total and free concentrations in

the plasma (Rostami-Hodjegan and Tucker, 2004). The KI value needs to be corrected

for non-specific binding of the inhibitor to microsomes. The effect of the inhibitor in vivo

Page 190: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

175

also depends on several factors. These include the properties of the substrate co-

administered, the relative fraction of the substrate metabolized by the enzyme being

inhibited (Ghanbari et al., 2006), and the potential inhibition caused by different

metabolites (e.g. desmethylselegiline, L-amphetamine, and L-methamphetamine). For

example, the plasma AUCs’ of metabolites following oral selegiline administration are

substantially greater than selegiline (demethylselegiline: 31-fold; L-methamphetamine:

161-fold; L-amphetamine: 50-fold) as well as having longer elimination half-lives (Laine

et al., 2000). Since we showed that these metabolites are capable of inhibiting CYP2A6

in vitro (Fig. 22), they may potentially contribute to the inhibition of in vivo nicotine

metabolism. By determining the effect of the inhibitor on nicotine clearance in vivo, the

relative contribution of the inhibitory effect of selegiline on CYP2A6 vs. MAO-B inhibition

in smoking cessation can potentially be estimated. For instance, by comparing the

nicotine pharmacokinetics resulting from selegiline-mediated CYP2A6 inhibition to

individuals with a particular CYP2A6 genotype that produces similar nicotine

pharmacokinetics, the extent of smoking cessation contributed by CYP2A6 and MAO-B

inhibition can be inferred from smoking behaviours of individuals with the CYP2A6

genotype being compared. Alternatively, if a compound exhibits similar MAO-B inhibition

characteristics but does not interact with CYP2A6 or other nicotine metabolizing

enzymes, its effect on smoking cessation may be compared with and estimate the

contribution of inhibition of MAO-B or CYP2A6 by selegiline. Practically speaking, the

concomitant inhibition of nicotine and dopamine metabolism provides selegiline an

advantage by targeting two distinct aspects important in nicotine dependence. As

mentioned in Section 2.3.4., individuals with normal CYP2A6 activity are less likely to

quit smoking with nicotine replacement therapy (transdermal and gum) or with

counseling compared to individuals with impaired CYP2A6 activity (Ho et al., 2009;

Lerman et al., 2006; Patterson et al., 2008); therefore, normal metabolizers may obtain

Page 191: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

176

the greatest benefit from inhibition of CYP2A6 by selegiline. Selegiline at 10 mg daily

dose plus oral nicotine may be particularly useful for smokers with Parkinson’s disease

although this remain to be tested.

Section 6 Factors Affecting Nicotine Intake

Section 6.1 Impact of Genetic Variations

Section 6.1.1 Variations in Nicotine Metabolism

We showed that oral nicotine consumption is associated with CYP2A5 levels and in vitro

nicotine metabolism in male mice. This is consistent with observations in humans where

CYP2A6 activity is associated with smoking behaviours (Section 2.3.4). In our study we

examined in vitro nicotine metabolism in mice from two extreme ends of the

consumption spectrum rather than from across the entire consumption range; therefore,

we were unable to assess the contribution of nicotine metabolism (and Cyp2a5 genotype)

to the variability in oral NSA. Since one of the parental strains (ST/bJ) self-administrated

significantly less amount of oral nicotine compared to the C57BL/6 mice, it would be

advantageous to have the Cyp2a5 sequence data and hepatic microsomes from the

ST/bJ mice to verify that they have a different Cyp2a5 sequence and slower nicotine

metabolism than the C57BL/6 mice. Similarly, the lower affinity of the ICR hepatic

microsomes for nicotine C-oxidation may be due to a polymorphism in Cyp2a5 which

may contribute to differences in oral NSA (and other nicotine behavioural responses),

but this remains to be verified. A recent genetic study in mice using the F2 offspring of

C57BL/6J (high nicotine consumption strain) and C3H/HeJ (low nicotine consumption

strain) has revealed that at least four QTLs in the mouse genome are associated with

oral NSA (LOD score of >2.0 [suggestive association]), and these QTLs contributed to

~62% of the variance in this behaviour (Li et al., 2007). One of the loci is located on

Page 192: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

177

chromosome 7 and the highest combined LOD score found on this loci is 5.9 (highly

significant association). Overall, the QTL on chromosome 7 was estimated to contribute

~13% of the variance in oral NSA. However, the study only reported data derived from

10 cM and onward, and Cyp2a5 is located at ~6.5 cM, thus it is unclear what the

contribution of this region is (i.e. Cyp2a5) to the overall variance in oral NSA (Li et al.,

2007). It is also possible that other genes influencing nicotine consumption co-segregate

with Cyp2a5 but this remains to be investigated. In order to directly test the contribution

of CYP2A5-mediated nicotine metabolism on NSA, inhibitors of CYP2A5 could be used.

Similarly in humans, genetics play a significant role in various aspects of smoking

behaviour which include initiation (Li et al., 2003; Vink et al., 2005), persistence to

regular smoking (Maes et al., 2004), number of cigarettes smoked (Koopmans et al.,

1999), degree of nicotine dependence (Vink et al., 2005), withdrawal severity (Xian et al.,

2003), and smoking cessation (Broms et al., 2006). The heritability estimates for the

amount of cigarettes smoked (once initiated) was shown to be up to 86% (Koopmans et

al., 1999), which is similar to the results from the above mouse study. Vast numbers of

pharmacogenetic studies have demonstrated that the CYP2A6 genotype to be highly

associated with many aspects of smoking (Section 2.3.4). The heritability estimates of

CYP2A6 to different dimensions of smoking behaviour have not been specifically defined

but a genome-wide linkage scan found this gene to be situated near a locus associated

with smoking frequency (Swan et al., 2006).

Section 6.1.2 Non-Nicotine Metabolism Variations

In addition to nicotine metabolism, other genetic factors may contribute to variations in

oral NSA among mouse strains. One particular striking feature is that a significant

association exists between the amount of nicotine self-administered and the threshold of

Page 193: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

178

nicotine-induced seizures observed between 19 strains of mice (r=0.89, P<0.01)

(Crawley et al., 1997); although this may be related to the sensitivity of mice to nicotine

which is in part controlled by the bioavailability of oral nicotine following hepatic first-pass

metabolism. An earlier study conducted by Hatchell and Collins showed a lack of clear-

cut association between brain nicotine levels or hepatic nicotine clearance and motor

depression between mouse strains, suggesting differential sensitivity to the

pharmacological effects of nicotine (Hatchell and Collins, 1980). Nicotine sensitivity may

be influenced by polymorphisms in genes of the neurotransmitter system involved in

nicotine reward, which include the dopaminergic (David et al., 2006), serotonergic

(Villegier et al., 2006), cholinergic (Levin et al., 2008), GABAnergic (Markou et al., 2004),

opiate (Berrendero et al., 2002), and cannabinoid (Merritt et al., 2008) systems. Although

currently there are no known studies demonstrating the association between

polymorphisms in these genes with nicotine sensitivity in mice, studies using inhibitors

and knock-out animals have revealed their direct role in nicotine pharmacology. For

instance, administration of the dopamine D1 receptor antagonist SCH 23390, or the

nicotinic receptor antagonist DHβE, disrupted intra-cranial self-administration of nicotine

to the ventral tegmental area, a brain region important for drug reward (David et al.,

2006). Mice with the β2 nicotinic acetylcholine receptors knocked out showed reduced

oral NSA in the first weeks of testing but nicotine consumption level returned to the wild-

type levels within the first month. In contrast, α7 nicotinic acetylcholine receptors

knockout mice showed normal oral NSA at the beginning of test but their nicotine

consumption level declined after several weeks (Levin et al., 2008). In humans, evidence

indicates that polymorphisms in the neurotransmitter systems of the nicotine reward

pathways potentially influence smoking behaviours. These include dopamine receptors

(e.g. DRD2) (Comings et al., 1996), transporter (DAT) (Ling et al., 2004), and

Page 194: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

179

metabolism (e.g. TH) (Olsson et al., 2004); serotonin transport (5-HTT) (Gerra et al.,

2005) and metabolism (TPH1) (Sullivan et al., 2001); nicotinic acetylcholine receptors

(e.g. α4) (Li et al., 2005); GABA receptor (GABAB2) (Beuten et al., 2005a); opiate

receptor (OPRM1) (Swan et al., 2006); and brain-derived neurotrophic factor (Beuten et

al., 2005b).

Section 6.2 Non-Genetic Factors

Although in general oral nicotine self-administration appears to mimic nicotine

consumption via smoking (Section 3.1), the two behaviours may be influenced by

different factors. For instance, in addition to nicotine, cigarette smoke contains numerous

chemicals and additives that may affect smoking behaviour through their interactions

with various cranial nerves, such as the olfactory nerve (smell), trigeminal nerve

(irritation), as well as facial, glossopharyngeal, vagal nerves (taste) (Megerdichian et al.,

2007). Studies have shown that application of anesthetic to the airway significantly

reduced satisfaction and craving from smoking (Rose et al., 1985; Rose et al., 1984) and

blockade of olfactory/taste stimuli from smoking reduced self-administration of cigarette

puffs (Perkins et al., 2001). In addition, studies have also shown that neither

denicotinized cigarettes nor pure nicotine alone completely abolish cravings or provide

the subjective satisfaction of smoking regular cigarettes (Rose et al., 2003; Rose et al.,

2000). This evidence suggests the importance of non-nicotinic factors in contributing to

smoking behaviours and they may be different to oral NSA.

There appears to be another important distinction between nicotine self-

administration in mice and smoking in humans. Stress is considered to be an important

factor for the amount of cigarettes smoked in humans (Baker et al., 2004; Todd, 2004).

Page 195: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

180

However, a genetic study using F2 mice comprised of low, average, and high stress

responders revealed that despite showing different response to stress, the mice did not

show different level of NSA. A caveat of this study was that the degree of NSA was not

measured against the degree of stress induced within the parental strain (or within the

stress-responder groups), thus the potential influence of genetic components (e.g.

nicotine metabolism) in NSA cannot be ruled out (Bilkei-Gorzo et al., 2008).

Section 7 Sex-Dependent Association between Nicotine Metabolism

and NSA

Even though our study chose male and female mice that were matched for their NSA

level, on average the female mice tend to metabolize nicotine slightly faster compared to

males. Similar observations are also seen in humans (Benowitz et al., 2006a). It has

been reported that on average, both men and women smoke a similar number of

cigarettes (Benowitz and Hatsukami, 1998; Benowitz et al., 2006a; Zeman et al., 2002).

However, women in general have a faster rate of nicotine metabolism (e.g. fractional

clearance of nicotine to cotinine [>15%] and 3-HC/cotinine ratios [>20%]) compared to

men (Benowitz et al., 2006a; Zeman et al., 2002). This is supported by our previous

findings that human females expressed a greater amount of CYP2A6 protein and have

faster in vitro nicotine metabolism compared to males (Al Koudsi et al., 2006b).

We also noted that female mice tend to have more variable rates of nicotine

metabolism compared to males. One potential explanation is the fluctuation in the level

of steroid hormones in females altering the level of CYP2A5 (as the mouse oestrous

cycle is less than one week). 17β-estradiol has been shown to induce a significant rise in

CYP2A6 mRNA in human hepatocytes (likely via the estrogen receptor-α by interacting

Page 196: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

181

with the estrogen response element [AGGTCAnnnCAACCT]), although protein levels

were not measured (Higashi et al., 2007a). Immunostaining also revealed that CYP2A6

protein levels tend to be higher during the follicular phase compared to the luteal phase

in endometrial tissues (Higashi et al., 2007a). The effect of sex steroids on nicotine

metabolism in mice has not been examined; however, there also appears to be an

estrogen response element in mice as well [AGGGCAnnnTGACCT] although it is

unclear if it is involved in the regulation of mouse Cyp2a5 (Orimo et al., 1995). It is

possible that sex hormones may induce other nicotine-metabolizing hepatic enzymes in

mice.

In contrast to the above in vitro study, pharmacokinetic studies in humans have

revealed some discrepancies on the role of female sex hormones on nicotine

metabolism. One study found that the pharmacokinetics of nicotine or cotinine are not

affected by the menstrual cycle, as measured during the mid-follicular and mid-luteal

phases (Hukkanen et al., 2005a). In contrast, other studies have shown that pregnancy

and the use of estrogen-containing oral contraceptives significantly increases the

clearance of nicotine via the cotinine pathway (~25% and ~40%, respectively) (Benowitz

et al., 2006a; Dempsey et al., 2002). This inconsistency may be due to the varying

amount and the duration of exposure to estrogen during these two distinctive

physiological phases. Other available evidence indicates that smoking behaviour is

influenced by menstrual cycle (increased during late luteal phase) but does not change

during pregnancy (DeBon et al., 1995; Dempsey et al., 2002; Snively et al., 2000),

suggesting the role of other influences on nicotine intake in females. 17β-estradiol and

progesterone have been shown to decrease nicotine’s antinociceptive effect in female

mice, through antagonism of α4β2 nicotinic receptor, whereas testosterone had no effect

in males (Damaj, 2001). These findings suggest a balance exists between nicotine

Page 197: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

182

metabolism and nicotine pharmacodynamics as influenced by female steroids (both

appeared to be, in part, affected by 17β-estradiol) in controlling nicotine intake,

especially in females. This may also underlie our observation that nicotine metabolism is

less correlated with NSA in female mice compared to the male mice. The complex

interactions between nicotine metabolism and sex may also contribute to the conflicting

evidence showing reduced efficacy of NRTs in females compared to males in some

studies but not others (Bjornson et al., 1995; Perkins and Scott, 2008; Schnoll et al.,

2009; Wetter et al., 1999).

Section 8 Effect of Chronic Nicotine Treatment on Nicotine Metabolism

In our studies we examined nicotine metabolism in mice in acute settings (i.e. following a

single nicotine administration); however, in clinical settings, nicotine metabolism and in

particular, nicotine-mediated behavioural effects, take place during chronic nicotine

exposure. Therefore, nicotine metabolism during chronic exposure is of particular

interest.

In monkeys, chronic nicotine treatment (0.3 mg/kg b.i.d.) for three weeks led to a

~60% decrease in CYP2A6agm protein levels and ~40% decrease in in vitro nicotine

metabolism (Schoedel et al., 2003). Cigarette smoke exposure in human led to only a

small reduction in the fractional clearance of nicotine to cotinine (Benowitz and Jacob,

2000). However, 7-day tobacco abstinence led to a 39% increase in non-renal nicotine

clearance (Lee et al., 1987) and a recent study showed that the clearance of oral

nicotine is reduced by up to 25% in smokers compared to non-smokers (Mwenifumbo et

al., 2007). Therefore, chronic nicotine exposure would likely decrease the metabolism of

CYP2A6 substrates, in particular nicotine. We have previously examined the effect of

nicotine metabolism in mice following chronic nicotine treatments. In a first study, DBA/2

Page 198: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

183

mice were treated with a single injection of nicotine (1 mg/kg s.c.) once daily for seven

days. On the last day plasma concentrations of nicotine were measured but no

differences were found in its elimination kinetics compared to those of acute treatment

(unpublished observations). This may be due to the rapid clearance of nicotine in the

mice. In a subsequent study osmotic mini-pumps providing 1 mg/kg/hr or 5 mg/kg/hr of

nicotine for seven days were implanted in DBA/2 mice. On day seven the mice were

sacrificed and livers were removed to examine hepatic CYP2A5 protein levels.

Compared to saline, 1 mg/kg/hr nicotine did not caused significant changes in CYP2A5

levels. In contrast, 5 mg/kg/hr nicotine caused a drop in CYP2A5 levels although nicotine

at this level were toxic to mice, therefore we felt that nicotine at this level were of little

pharmacological relevance (unpublished observations). A recent study found that

chronic nicotine exposure in DBA/2 and C57Bl/6 mice, while leading to changes in

various behavioural measures, not all the changes were congruent in terms of direction

of change. For example, during withdrawal both strains showed decreases in somatic

withdrawal signs between day 1 and day 3; in contrast, in the plus-maze test, the DBA/2

mice increased time spent whereas C57Bl/6 mice decreased time spent in the opening.

Therefore, the contribution of chronic nicotine exposure to behavioural changes by

potential differences in nicotine metabolism could be secondary to and/or confounded by

the impact of other physiological/neurological changes (Jackson et al., 2009). The above

data suggested while in both humans and monkeys chronic nicotine exposure could lead

to reduction in CYP2A6-mediated nicotine metabolism, which may play a role in altered

nicotine pharmacology in chronic settings; its impact on mouse nicotine metabolism has

not yet been demonstrated or is minimal, possibly due to rapid elimination of the drug in

this species.

Section 9 Future Directions

Page 199: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

184

Section 9.1 Selegiline

Previous study indicated that mouse deficient for MAO-B showed no changes in brain

dopamine levels, suggesting its lack of involvement in dopamine metabolism (Grimsby et

al., 1997). In contrast, dopamine metabolism in mice appeared to be mediated in part by

MAO-A (Shih et al., 1999). Thus, although selegiline is an irreversible inhibitor of MAO-B,

it would have no impact on dopamine metabolism in mice. Consequently, the effect of

selegiline on nicotine pharmacology can be tested and the changes in nicotine-mediated

behaviour can be attributed to the CYP2A5-inhibition (i.e. inhibition of nicotine

metabolism) aspect of selegiline. On the other hand, this also excludes the use of mice

for studying the impact of selegiline on dopamine metabolism and its subsequent effects

on nicotine pharmacodynamics.

Based on our observation that selegiline is an MBI of CYP2A6, it is reasonable to

assume that it will exert its greatest impact on the first-pass metabolism (i.e. inhibition) of

orally delivered nicotine. Considering existing clinical trials only examine selegiline either

alone (capsule or patch) or in combination (capsule plus nicotine patch) for smoking

cessation (Clinicaltrials.gov), it would also be important to examine the effect of oral

selegiline in combination with oral nicotine capsule/pill. It is likely that the smoking

cessation effect may be greater when oral selegiline is used in combination with oral

nicotine due to inhibition of dopamine metabolism and decreased nicotine metabolism.

Consider the involvement of various non-nicotine factors in nicotine-dependence, the

above therapeutic combination may be particularly useful as an initial treatment in

conjunction with behavioural therapies for smokers consider quitting.

Section 9.2 Mouse Model of Nicotine Metabolism

Page 200: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

185

In this thesis we demonstrated the usefulness of the mouse model in studying nicotine

metabolism in the context of pharmacological inhibition. However, extrapolation of

pharmacokinetic data from animals to humans has limitations due to species differences

in physiology and conservation of drug metabolism enzymes. The former can be

addressed by allometric scaling which proposed that certain pharmacokinetic information

(e.g. plasma-concentration-time profile, clearance, and distribution) can be estimated

between species as a function of body weight (Boxenbaum, 1984; Martignoni et al.,

2006). It has also been suggested that, in general, small animals have a higher hepatic

P450 content per unit body weight due to a larger liver to body weight ratio as well as

faster hepatic blood flow (Davies and Morris, 1993). As a consequence, smaller animals

have a faster metabolic clearance of xenobiotics compared to humans (Davies and

Morris, 1993). However, species variations in P450 represents a major factor in drug

metabolism difference between species, particularly for compounds that are extensively

metabolized (Martignoni et al., 2006). Therefore in these circumstances pharmacokinetic

data from animals may not be reasonably applied to humans. For instance, while

existing evidence, and results from our study, indicate that both human CYP2A6 and

mouse CYP2A5 exhibit similar nicotine metabolizing properties, studies with selegiline

showed that this drug to be a MBI of CYP2A6 but not CYP2A5. The endogenous

compound 2-phenylethyamine is a far better inhibitor of CYP2A6 than CYP2A5 in vitro

(IC50 ratio of <0.04) (Rahnasto et al., 2003) and other experimental compounds also

exhibit different inhibition potencies for these enzymes (Juvonen et al., 2000). Thus,

evaluating the impact of CYP2A6 inhibitors in mice in vivo may yield different results.

Furthermore, potential pharmacokinetic differences in the metabolism of the inhibitors

between species can also alter their inhibitory properties. While the mouse model may

be imperfect, it represents the best small animal model we have to date and may prove

useful for the first line in vivo animal studies considering that the main nicotine

Page 201: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

186

metabolizing enzymes in rats, the other well studied small animal species, belong to the

CYP2B family (Nakayama et al., 1993).

In the past few years humanized mice have been developed using several

strategies. One method is to introduce human P450 cDNA with tissue-specific promoter

or the entire gene (with regulatory elements) into the pronuclei of the mice (Cheung and

Gonzalez, 2008). With this method, however, the presence of endogenous counterparts

(homologous or orthologous genes) may interfere with, and confound the effects of, the

introduced genes. Therefore mice with the corresponding gene knocked out should be

used. Another method is to introduce cDNA into the site of the endogenous gene

thereby disrupting its function (Cheung and Gonzalez, 2008). An advantage of this

model is that once generated, the transgenic line can be maintained indefinitely.

Conceivably this model is ideal for studying the impact of various Cyp2a5 and CYP2A6

genetic polymorphisms on nicotine-mediated behaviours. In the case of inhibitor studies,

(human) enzymes involved in the metabolism of the inhibitor should ideally be

introduced to increase the predictability of the animal model; however, this may not be

feasible since different inhibitors will likely be metabolized by different enzymes. We are

aware that humanized CYP2A6 mice (Cyp2a5-null background) are currently being

developed and this model may be invaluable in studying CYP2A6 inhibitors in vivo.

Another method that has been developed is the transplantation of human

hepatocytes into mice (Katoh et al., 2008). This model employs a urokinase-type

plasminogen activator+/+/ severe combined immunodeficient mouse strain in which

human hepatocytes are directly injected into the spleen of the mice and the resultant

mice have between 80-90% of their hepatocytes replaced (Tateno et al., 2004). This

model may provide a much more accurate picture of in vivo metabolism (and inhibition)

as all human hepatic enzymes and their regulatory factors are present (Katoh et al.,

2008). This model may also indicate whether the target compound is inducing the

Page 202: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

187

expression of other enzymes through human (rather than mice) xenobiotic receptors,

which is particularly important for drug-drug interactions (Katoh et al., 2008). A drawback

of this particular model is that the chimera mice have a shorter life-span due to organ

failure caused by graft vs. host disease and administration of protease inhibitor is also

required to prolong survival (Tateno et al., 2004). The mice also need to be constantly

generated and thus costly and time-consuming. Their short life-span and inability to

generate offspring carrying human hepatocytes can limit their use in behavioural studies.

Clearly, each of the above mouse models has its advantages and drawbacks in

studying nicotine metabolism. Further development of these models, either through

creating mice with multiple human transgenes, or refinement of the hepatocyte

transplant model, are needed to increase their applicability to humans; however, a

disadvantage associated with the humanized mouse model is that the generation of a

panel of mice each with different CYP2A6 genotype is both time and resource

consuming. Continual functional characterization of CYP2A5 (and in vivo nicotine

metabolism) from different mouse strains may also be useful in correlating their

metabolism phenotype with that of CYP2A6 and its functional variants. Generation of

mutations in mouse Cyp2a5 that mimic various CYP2A6 genotypes may also be useful

in understanding the contribution of these genotypes in altering nicotine pharmacology.

Finally, even though nicotine is largely metabolized by the liver in mice, the contribution

of other pharmacokinetic factors such as absorption, distribution, and extrahepatic

metabolism need to be considered.

Section 10 Conclusion

Based on our findings that: 1) CYP2A5, as with CYP2A6, is the primary enzyme

responsible for the metabolism of nicotine to cotinine and cotinine to 3-HC; 2) genetic

Page 203: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

188

variations in CYP2A5 may alter nicotine metabolism; 3) inhibition of CYP2A5-mediated

nicotine metabolism led to change in nicotine pharmacokinetics and pharmacodynamics;

and 4) identification of an existing drug (selegiline) as an inhibitor of CYP2A5 and

CYP2A6 suggested the mouse may be a suitable model for studying novel CYP2A6

inhibitory compounds. Overall, the use of CYP2A5 inhibitors in the mouse model may be

suitable for studying the direct effect of nicotine metabolism on nicotine pharmacology.

On the other hand, genetic studies across strains may be useful for dissecting the

genetic architecture of nicotine-mediated behaviours, in this case oral NSA.

Page 204: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

189

REFERENCES

Aceto MD, Martin BR, Uwaydah IM, May EL, Harris LS, Izazola-Conde C, Dewey WL, Bradshaw TJ and Vincek WC (1979) Optically pure (+)-nicotine from (+/-)-nicotine and biological comparisons with (-)-nicotine. J Med Chem 22(2):174-177.

Adriani W, Macri S, Pacifici R and Laviola G (2002) Restricted daily access to water and voluntary nicotine oral consumption in mice: methodological issues and individual differences. Behav Brain Res 134(1-2):21-30.

Al-Salmy HS (2001) Individual variation in hepatic aldehyde oxidase activity. IUBMB Life 51(4):249-253.

Al Koudsi N, Mwenifumbo JC, Sellers EM, Benowitz NL, Swan GE and Tyndale RF (2006a) Characterization of the novel CYP2A6*21 allele using in vivo nicotine kinetics. Eur J Clin Pharmacol 62(6):481-484.

Al Koudsi N, Seller EM and Tyndale RF (2006b) Genetic variation, age, and sex: impacts on hepatic CYP2A6 levels and nicotine metabolizing activity, in Society for Research on Nicotine and Tobacco p S235, Orlando, FL.

Alkondon M and Albuquerque EX (2005) Nicotinic receptor subtypes in rat hippocampal slices are differentially sensitive to desensitization and early in vivo functional up-regulation by nicotine and to block by bupropion. J Pharmacol Exp Ther 313(2):740-750.

Allchurch MH (2008) Press release: The European Medicines Agency recommends suspension of the marketing authorisation of Acomplia, (Agency EM ed), European Medicines Agency.

Anthenelli RM (2005) Rimonabant in smoking abstinence: the STRATUS-program., in Proceedings of the 11 Annual Meeting of the Society for Research on Nicotine and Tobacco, Prague, Czech Republic.

Argyelan M, Szabo Z, Kanyo B, Tanacs A, Kovacs Z, Janka Z and Pavics L (2005) Dopamine transporter availability in medication free and in bupropion treated depression: a 99mTc-TRODAT-1 SPECT study. J Affect Disord 89(1-3):115-123.

Ariyoshi N, Sawamura Y and Kamataki T (2001) A novel single nucleotide polymorphism altering stability and activity of CYP2a6. Biochem Biophys Res Commun 281(3):810-814.

Ariyoshi N, Sekine H, Saito K and Kamataki T (2002) Characterization of a genotype previously designated as CYP2A6 D-type: CYP2A6*4B, another entire gene deletion allele of the CYP2A6 gene in Japanese. Pharmacogenetics 12(6):501-504.

Armitage AK, Dollery CT, George CF, Houseman TH, Lewis PJ and Turner DM (1975) Absorption and metabolism of nicotine from cigarettes. Br Med J 4(5992):313-316.

Aschhoff S, Schroff K-C, Wildenauer DB and Richter E (1999) Nicotine consumption of several mouse strains using a two bottle choice paradigm. J Exp Anim Sci 40(4):171-177.

Bailey BA, Philips SR and Boulton AA (1987) In vivo release of endogenous dopamine, 5-hydroxytryptamine and some of their metabolites from rat caudate nucleus by phenylethylamine. Neurochem Res 12(2):173-178.

Baker TB, Brandon TH and Chassin L (2004) Motivational influences on cigarette smoking. Annu Rev Psychol 55:463-491.

Page 205: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

190

Balfour DJ (2004) The neurobiology of tobacco dependence: a preclinical perspective on the role of the dopamine projections to the nucleus accumbens [corrected]. Nicotine Tob Res 6(6):899-912.

Bao Z, He XY, Ding X, Prabhu S and Hong JY (2005) Metabolism of nicotine and cotinine by human cytochrome P450 2A13. Drug Metab Dispos 33(2):258-261.

Barrett JS, Hochadel TJ, Morales RJ, Rohatagi S, DeWitt KE, Watson SK and DiSanto AR (1996a) Pharmacokinetics and Safety of a Selegiline Transdermal System Relative to Single-Dose Oral Administration in the Elderly. Am J Ther 3(10):688-698.

Barrett JS, Rohatagi S, DeWitt KE, Morales RJ and DiSanto AR (1996b) The Effect of Dosing Regimen and Food on the Bioavailability of the Extensively Metabolized, Highly Variable Drug Eldepryl((R)) (Selegiline Hydrochloride). Am J Ther 3(4):298-313.

Benetton SA, Fang C, Yang YO, Alok R, Year M, Lin CC and Yeh LT (2007) P450 phenotyping of the metabolism of selegiline to desmethylselegiline and methamphetamine. Drug Metab Pharmacokinet 22(2):78-87.

Benowitz NL (1990) Clinical pharmacology of inhaled drugs of abuse: implications in understanding nicotine dependence, in NIDA Research Monograph pp 12-29.

Benowitz NL (1999) Biomarkers of environmental tobacco smoke exposure. Environ Health Perspect 107 Suppl 2:349-355.

Benowitz NL, Chan K, Denaro CP and Jacob P, 3rd (1991a) Stable isotope method for studying transdermal drug absorption: the nicotine patch. Clin Pharmacol Ther 50(3):286-293.

Benowitz NL, Griffin C and Tyndale R (2001) Deficient C-oxidation of nicotine continued. Clin Pharmacol Ther 70(6):567.

Benowitz NL and Hatsukami D (1998) Gender differences in the pharmacology of nicotine addictioin. Addict Biol 3(4):383-404.

Benowitz NL and Jacob P, 3rd (1985) Nicotine renal excretion rate influences nicotine intake during cigarette smoking. J Pharmacol Exp Ther 234(1):153-155.

Benowitz NL and Jacob P, 3rd (1993) Nicotine and cotinine elimination pharmacokinetics in smokers and nonsmokers. Clin Pharmacol Ther 53(3):316-323.

Benowitz NL and Jacob P, 3rd (1994) Metabolism of nicotine to cotinine studied by a dual stable isotope method. Clin Pharmacol Ther 56(5):483-493.

Benowitz NL and Jacob P, 3rd (2000) Effects of cigarette smoking and carbon monoxide on nicotine and cotinine metabolism. Clin Pharmacol Ther 67(6):653-659.

Benowitz NL and Jacob P, 3rd (2001) Trans-3'-hydroxycotinine: disposition kinetics, effects and plasma levels during cigarette smoking. Br J Clin Pharmacol 51(1):53-59.

Benowitz NL, Jacob P, 3rd, Denaro C and Jenkins R (1991b) Stable isotope studies of nicotine kinetics and bioavailability. Clin Pharmacol Ther 49(3):270-277.

Benowitz NL, Jacob P, 3rd, Fong I and Gupta S (1994) Nicotine metabolic profile in man: comparison of cigarette smoking and transdermal nicotine. J Pharmacol Exp Ther 268(1):296-303.

Benowitz NL, Jacob P, 3rd and Savanapridi C (1987) Determinants of nicotine intake while chewing nicotine polacrilex gum. Clin Pharmacol Ther 41(4):467-473.

Benowitz NL, Kuyt F, Jacob P, 3rd, Jones RT and Osman AL (1983) Cotinine disposition and effects. Clin Pharmacol Ther 34(5):604-611.

Benowitz NL, Lessov-Schlaggar CN and Swan GE (2008) Genetic influences in the variation in renal clearance of nicotine and cotinine. Clin Pharmacol Ther 84(2):243-247.

Page 206: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

191

Benowitz NL, Lessov-Schlaggar CN, Swan GE and Jacob P, 3rd (2006a) Female sex and oral contraceptive use accelerate nicotine metabolism. Clin Pharmacol Ther 79(5):480-488.

Benowitz NL, Perez-Stable EJ, Fong I, Modin G, Herrera B and Jacob P, 3rd (1999) Ethnic differences in N-glucuronidation of nicotine and cotinine. J Pharmacol Exp Ther 291(3):1196-1203.

Benowitz NL, Swan GE, Jacob P, 3rd, Lessov-Schlaggar CN and Tyndale RF (2006b) CYP2A6 genotype and the metabolism and disposition kinetics of nicotine. Clin Pharmacol Ther 80(5):457-467.

Berlin I, Aubin HJ, Pedarriosse AM, Rames A, Lancrenon S and Lagrue G (2002) Lazabemide, a selective, reversible monoamine oxidase B inhibitor, as an aid to smoking cessation. Addiction 97(10):1347-1354.

Berlin I, Said S, Spreux-Varoquaux O, Launay JM, Olivares R, Millet V, Lecrubier Y and Puech AJ (1995) A reversible monoamine oxidase A inhibitor (moclobemide) facilitates smoking cessation and abstinence in heavy, dependent smokers. Clin Pharmacol Ther 58(4):444-452.

Berrendero F, Kieffer BL and Maldonado R (2002) Attenuation of nicotine-induced antinociception, rewarding effects, and dependence in mu-opioid receptor knock-out mice. J Neurosci 22(24):10935-10940.

Berrendero F, Mendizabal V, Robledo P, Galeote L, Bilkei-Gorzo A, Zimmer A and Maldonado R (2005) Nicotine-induced antinociception, rewarding effects, and physical dependence are decreased in mice lacking the preproenkephalin gene. J Neurosci 25(5):1103-1112.

Berry MD, Scarr E, Zhu MY, Paterson IA and Juorio AV (1994) The effects of administration of monoamine oxidase-B inhibitors on rat striatal neurone responses to dopamine. Br J Pharmacol 113(4):1159-1166.

Beuten J, Ma JZ, Payne TJ, Dupont RT, Crews KM, Somes G, Williams NJ, Elston RC and Li MD (2005a) Single- and multilocus allelic variants within the GABA(B) receptor subunit 2 (GABAB2) gene are significantly associated with nicotine dependence. Am J Hum Genet 76(5):859-864.

Beuten J, Ma JZ, Payne TJ, Dupont RT, Quezada P, Huang W, Crews KM and Li MD (2005b) Significant association of BDNF haplotypes in European-American male smokers but not in European-American female or African-American smokers. Am J Med Genet B Neuropsychiatr Genet 139B(1):73-80.

Biberman R, Neumann R, Katzir I and Gerber Y (2003) A randomized controlled trial of oral selegiline plus nicotine skin patch compared with placebo plus nicotine skin patch for smoking cessation. Addiction 98(10):1403-1407.

Bilkei-Gorzo A, Racz I, Michel K, Darvas M, Maldonado R and Zimmer A (2008) A common genetic predisposition to stress sensitivity and stress-induced nicotine craving. Biol Psychiatry 63(2):164-171.

Bischoff KB, Dedrick RL, Zaharko DS and Longstreth JA (1971) Methotrexate pharmacokinetics. J Pharm Sci 60(8):1128-1133.

Bjornson W, Rand C, Connett JE, Lindgren P, Nides M, Pope F, Buist AS, Hoppe-Ryan C and O'Hara P (1995) Gender differences in smoking cessation after 3 years in the Lung Health Study. Am J Public Health 85(2):223-230.

Blondal T, Gudmundsson LJ, Tomasson K, Jonsdottir D, Hilmarsdottir H, Kristjansson F, Nilsson F and Bjornsdottir US (1999) The effects of fluoxetine combined with nicotine inhalers in smoking cessation--a randomized trial. Addiction 94(7):1007-1015.

Bond CA, Grant K and Boh L (1981) Photochemotherapy of psoriasis with methoxsalen and longwave ultraviolet light (PUVA). Am J Hosp Pharm 38(7):990-995.

Page 207: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

192

Boshier A, Wilton LV and Shakir SA (2003) Evaluation of the safety of bupropion (Zyban) for smoking cessation from experience gained in general practice use in England in 2000. Eur J Clin Pharmacol 59(10):767-773.

Boxenbaum H (1980) Interspecies variation in liver weight, hepatic blood flow, and antipyrine intrinsic clearance: extrapolation of data to benzodiazepines and phenytoin. J Pharmacokinet Biopharm 8(2):165-176.

Boxenbaum H (1984) Interspecies pharmacokinetic scaling and the evolutionary-comparative paradigm. Drug Metab Rev 15(5-6):1071-1121.

Brandange S and Lindblom L (1979) The enzyme "aldehyde oxidase" is an iminium oxidase. Reaction with nicotine delta 1'(5') iminium ion. Biochem Biophys Res Commun 91(3):991-996.

Broms U, Silventoinen K, Madden PA, Heath AC and Kaprio J (2006) Genetic architecture of smoking behavior: a study of Finnish adult twins. Twin Res Hum Genet 9(1):64-72.

Brown PJ, Bedard LL, Reid KR, Petsikas D and Massey TE (2007) Analysis of CYP2A contributions to metabolism of 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone in human peripheral lung microsomes. Drug Metab Dispos 35(11):2086-2094.

Brunzell DH, Russell DS and Picciotto MR (2003) In vivo nicotine treatment regulates mesocorticolimbic CREB and ERK signaling in C57Bl/6J mice. J Neurochem 84(6):1431-1441.

Buccafusco JJ and Terry AV, Jr. (2003) The potential role of cotinine in the cognitive and neuroprotective actions of nicotine. Life Sci 72(26):2931-2942.

Burkhart BA, Harada N and Negishi M (1985) Sexual dimorphism of testosterone 15 alpha-hydroxylase mRNA levels in mouse liver. cDNA cloning and regulation. J Biol Chem 260(28):15357-15361.

Caggiula AR, Epstein LH, Perkins KA and Saylor S (1995) Different methods of assessing nicotine-induced antinociception may engage different neural mechanisms. Psychopharmacology (Berl) 122(3):301-306.

Cahill K, Stead LF and Lancaster T (2007) Nicotine receptor partial agonists for smoking cessation. Cochrane Database Syst Rev(1):CD006103.

Cahill K and Ussher M (2007) Cannabinoid type 1 receptor antagonists (rimonabant) for smoking cessation. Cochrane Database Syst Rev(4):CD005353.

Cashman JR, Park SB, Yang ZC, Wrighton SA, Jacob P, 3rd and Benowitz NL (1992) Metabolism of nicotine by human liver microsomes: stereoselective formation of trans-nicotine N'-oxide. Chem Res Toxicol 5(5):639-646.

Castane A, Valjent E, Ledent C, Parmentier M, Maldonado R and Valverde O (2002) Lack of CB1 cannabinoid receptors modifies nicotine behavioural responses, but not nicotine abstinence. Neuropharmacology 43(5):857-867.

Cauffiez C, Lo-Guidice JM, Quaranta S, Allorge D, Chevalier D, Cenee S, Hamdan R, Lhermitte M, Lafitte JJ, Libersa C, Colombel JF, Stucker I and Broly F (2004) Genetic polymorphism of the human cytochrome CYP2A13 in a French population: implication in lung cancer susceptibility. Biochem Biophys Res Commun 317(2):662-669.

Chen G, Blevins-Primeau AS, Dellinger RW, Muscat JE and Lazarus P (2007) Glucuronidation of nicotine and cotinine by UGT2B10: loss of function by the UGT2B10 Codon 67 (Asp>Tyr) polymorphism. Cancer Res 67(19):9024-9029.

Cheung C and Gonzalez FJ (2008) Humanized mouse lines and their application for prediction of human drug metabolism and toxicological risk assessment. J Pharmacol Exp Ther 327(2):288-299.

Choi JH, Dresler CM, Norton MR and Strahs KR (2003) Pharmacokinetics of a nicotine polacrilex lozenge. Nicotine Tob Res 5(5):635-644.

Page 208: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

193

Christian K, Lang M, Maurel P and Raffalli-Mathieu F (2004) Interaction of heterogeneous nuclear ribonucleoprotein A1 with cytochrome P450 2A6 mRNA: implications for post-transcriptional regulation of the CYP2A6 gene. Mol Pharmacol 65(6):1405-1414.

Cinciripini PM, Lapitsky L, Seay S, Wallfisch A, Meyer WJ, 3rd and van Vunakis H (1995) A placebo-controlled evaluation of the effects of buspirone on smoking cessation: differences between high- and low-anxiety smokers. J Clin Psychopharmacol 15(3):182-191.

Coe JW, Brooks PR, Vetelino MG, Wirtz MC, Arnold EP, Huang J, Sands SB, Davis TI, Lebel LA, Fox CB, Shrikhande A, Heym JH, Schaeffer E, Rollema H, Lu Y, Mansbach RS, Chambers LK, Rovetti CC, Schulz DW, Tingley FD, 3rd and O'Neill BT (2005) Varenicline: an alpha4beta2 nicotinic receptor partial agonist for smoking cessation. J Med Chem 48(10):3474-3477.

Cohen C, Perrault G, Griebel G and Soubrie P (2005) Nicotine-associated cues maintain nicotine-seeking behavior in rats several weeks after nicotine withdrawal: reversal by the cannabinoid (CB1) receptor antagonist, rimonabant (SR141716). Neuropsychopharmacology 30(1):145-155.

Cohen C, Perrault G, Voltz C, Steinberg R and Soubrie P (2002) SR141716, a central cannabinoid (CB(1)) receptor antagonist, blocks the motivational and dopamine-releasing effects of nicotine in rats. Behav Pharmacol 13(5-6):451-463.

Collins AC, Miner LL and Marks MJ (1988) Genetic influences on acute responses to nicotine and nicotine tolerance in the mouse. Pharmacol Biochem Behav 30(1):269-278.

Comings DE, Ferry L, Bradshaw-Robinson S, Burchette R, Chiu C and Muhleman D (1996) The dopamine D2 receptor (DRD2) gene: a genetic risk factor in smoking. Pharmacogenetics 6(1):73-79.

Cook JW, Spring B, McChargue DE, Borrelli B, Hitsman B, Niaura R, Keuthen NJ and Kristeller J (2004) Influence of fluoxetine on positive and negative affect in a clinic-based smoking cessation trial. Psychopharmacology (Berl) 173(1-2):153-159.

Cornuz J, Zwahlen S, Jungi WF, Osterwalder J, Klingler K, van Melle G, Bangala Y, Guessous I, Muller P, Willers J, Maurer P, Bachmann MF and Cerny T (2008) A vaccine against nicotine for smoking cessation: a randomized controlled trial. PLoS ONE 3(6):e2547.

Correia MA and Ortiz de Montellano PR (2005) Inhibition of Cytochrome P450 Enzymes. Kluwer Academic / Plenum, New York.

Corrigall WA and Coen KM (1991) Opiate antagonists reduce cocaine but not nicotine self-administration. Psychopharmacology (Berl) 104(2):167-170.

Corrigall WA, Coen KM and Adamson KL (1994) Self-administered nicotine activates the mesolimbic dopamine system through the ventral tegmental area. Brain Res 653(1-2):278-284.

Covey LS, Glassman AH and Stetner F (1999) Naltrexone effects on short-term and long-term smoking cessation. J Addict Dis 18(1):31-40.

Covey LS, Glassman AH, Stetner F, Rivelli S and Stage K (2002) A randomized trial of sertraline as a cessation aid for smokers with a history of major depression. Am J Psychiatry 159(10):1731-1737.

Crawley JN, Belknap JK, Collins A, Crabbe JC, Frankel W, Henderson N, Hitzemann RJ, Maxson SC, Miner LL, Silva AJ, Wehner JM, Wynshaw-Boris A and Paylor R (1997) Behavioral phenotypes of inbred mouse strains: implications and recommendations for molecular studies. Psychopharmacology (Berl) 132(2):107-124.

Page 209: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

194

Crooks PA and Dwoskin LP (1997) Contribution of CNS nicotine metabolites to the neuropharmacological effects of nicotine and tobacco smoking. Biochem Pharmacol 54(7):743-753.

d'Amour FE and Smith DL (1941) A method for determining loss of pain sensation. J Pharmacol Exp Ther 72(74-79).

Daigo S, Takahashi Y, Fujieda M, Ariyoshi N, Yamazaki H, Koizumi W, Tanabe S, Saigenji K, Nagayama S, Ikeda K, Nishioka Y and Kamataki T (2002) A novel mutant allele of the CYP2A6 gene (CYP2A6*11) found in a cancer patient who showed poor metabolic phenotype towards tegafur. Pharmacogenetics 12(4):299-306.

Damaj MI (2001) Influence of gender and sex hormones on nicotine acute pharmacological effects in mice. J Pharmacol Exp Ther 296(1):132-140.

Damaj MI, Fonck C, Marks MJ, Deshpande P, Labarca C, Lester HA, Collins AC and Martin BR (2007a) Genetic approaches identify differential roles for alpha4beta2* nicotinic receptors in acute models of antinociception in mice. J Pharmacol Exp Ther 321(3):1161-1169.

Damaj MI and Martin BR (1993) Is the dopaminergic system involved in the central effects of nicotine in mice? Psychopharmacology (Berl) 111(1):106-108.

Damaj MI, Siu EC, Seller EM, Tyndale RF and Martin BR (2006) Inhibition of nicotine metabolism by methoxysalen: pharmacokinetic and pharmacological studies in mice. J Pharmacol Exp Ther October 4:[Epub ahead of print].

Damaj MI, Siu EC, Sellers EM, Tyndale RF and Martin BR (2007b) Inhibition of nicotine metabolism by methoxysalen: Pharmacokinetic and pharmacological studies in mice. J Pharmacol Exp Ther 320(1):250-257.

David V, Besson M, Changeux JP, Granon S and Cazala P (2006) Reinforcing effects of nicotine microinjections into the ventral tegmental area of mice: dependence on cholinergic nicotinic and dopaminergic D1 receptors. Neuropharmacology 50(8):1030-1040.

Davies B and Morris T (1993) Physiological parameters in laboratory animals and humans. Pharm Res 10(7):1093-1095.

de Leon J, Diaz FJ, Rogers T, Browne D, Dinsmore L, Ghosheh OH, Dwoskin LP and Crooks PA (2002) Total cotinine in plasma: a stable biomarker for exposure to tobacco smoke. J Clin Psychopharmacol 22(5):496-501.

DeBon M, Klesges RC and Klesges LM (1995) Symptomatology across the menstrual cycle in smoking and nonsmoking women. Addict Behav 20(3):335-343.

Dempsey D, Jacob P, 3rd and Benowitz NL (2002) Accelerated metabolism of nicotine and cotinine in pregnant smokers. J Pharmacol Exp Ther 301(2):594-598.

Dempsey D, Tutka P, Jacob P, 3rd, Allen F, Schoedel K, Tyndale RF and Benowitz NL (2004) Nicotine metabolite ratio as an index of cytochrome P450 2A6 metabolic activity. Clin Pharmacol Ther 76(1):64-72.

Denton TT, Zhang X and Cashman JR (2004) Nicotine-related alkaloids and metabolites as inhibitors of human cytochrome P-450 2A6. Biochem Pharmacol 67(4):751-756.

Denton TT, Zhang X and Cashman JR (2005) 5-substituted, 6-substituted, and unsubstituted 3-heteroaromatic pyridine analogues of nicotine as selective inhibitors of cytochrome P-450 2A6. J Med Chem 48(1):224-239.

Dewey WL, Harris LS, Howes JF and Nuite JA (1970) The effect of various neurohumoral modulators on the activity of morphine and the narcotic antagonists in the tail-flick and phenylquinone tests. J Pharmacol Exp Ther 175(2):435-442.

Page 210: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

195

Di YM, Chow VD, Yang LP and Zhou SF (2009) Structure, Function, Regulation and Polymorphism of Human Cytochrome P450 2A6. Curr Drug Metab.

Ding X and Kaminsky LS (2003) Human extrahepatic cytochromes P450: function in xenobiotic metabolism and tissue-selective chemical toxicity in the respiratory and gastrointestinal tracts. Annu Rev Pharmacol Toxicol 43:149-173.

Draper AJ, Madan A and Parkinson A (1997) Inhibition of coumarin 7-hydroxylase activity in human liver microsomes. Arch Biochem Biophys 341(1):47-61.

Durcan MJ, Deener G, White J, Johnston JA, Gonzales D, Niaura R, Rigotti N and Sachs DP (2002) The effect of bupropion sustained-release on cigarette craving after smoking cessation. Clin Ther 24(4):540-551.

Dwoskin LP, Teng L, Buxton ST and Crooks PA (1999) (S)-(-)-Cotinine, the major brain metabolite of nicotine, stimulates nicotinic receptors to evoke [3H]dopamine release from rat striatal slices in a calcium-dependent manner. J Pharmacol Exp Ther 288(3):905-911.

Dyck LE and Davis BA (2001) Inhibition of rat liver microsomal CYP1A2 and CYP2B1 activity by N-(2-heptyl)-N-methyl-propargylamine and by N-(2-heptyl)-propargylamine. Drug Metab Dispos 29(8):1156-1161.

Eison AS and Temple DL, Jr. (1986) Buspirone: review of its pharmacology and current perspectives on its mechanism of action. Am J Med 80(3B):1-9.

Eswaramoorthy S, Bonanno JB, Burley SK and Swaminathan S (2006) Mechanism of action of a flavin-containing monooxygenase. Proc Natl Acad Sci U S A 103(26):9832-9837.

Ezzati M and Lopez AD (2004) Regional, disease specific patterns of smoking-attributable mortality in 2000. Tob Control 13(4):388-395.

Falls JG, Blake BL, Cao Y, Levi PE and Hodgson E (1995) Gender differences in hepatic expression of flavin-containing monooxygenase isoforms (FMO1, FMO3, and FMO5) in mice. J Biochem Toxicol 10(3):171-177.

Faraday MM, Blakeman KH and Grunberg NE (2005) Strain and sex alter effects of stress and nicotine on feeding, body weight, and HPA axis hormones. Pharmacol Biochem Behav 80(4):577-589.

Faucette SR, Hawke RL, Lecluyse EL, Shord SS, Yan B, Laethem RM and Lindley CM (2000) Validation of bupropion hydroxylation as a selective marker of human cytochrome P450 2B6 catalytic activity. Drug Metab Dispos 28(10):1222-1230.

Foley KF, DeSanty KP and Kast RE (2006) Bupropion: pharmacology and therapeutic applications. Expert Rev Neurother 6(9):1249-1265.

Fowler JS, MacGregor RR, Wolf AP, Arnett CD, Dewey SL, Schlyer D, Christman D, Logan J, Smith M, Sachs H and et al. (1987) Mapping human brain monoamine oxidase A and B with 11C-labeled suicide inactivators and PET. Science 235(4787):481-485.

Fowler JS, Volkow ND, Wang GJ, Pappas N, Logan J, MacGregor R, Alexoff D, Shea C, Schlyer D, Wolf AP, Warner D, Zezulkova I and Cilento R (1996a) Inhibition of monoamine oxidase B in the brains of smokers. Nature 379(6567):733-736.

Fowler JS, Volkow ND, Wang GJ, Pappas N, Logan J, Shea C, Alexoff D, MacGregor RR, Schlyer DJ, Zezulkova I and Wolf AP (1996b) Brain monoamine oxidase A inhibition in cigarette smokers. Proc Natl Acad Sci U S A 93(24):14065-14069.

Fredriksson A and Archer T (1995) Synergistic interactions between COMT-/MAO-inhibitors and L-Dopa in MPTP-treated mice. J Neural Transm Gen Sect 102(1):19-34.

Freeman GB, Sherman KA and Gibson GE (1987) Locomotor activity as a predictor of times and dosages for studies of nicotine's neurochemical actions. Pharmacol Biochem Behav 26(2):305-312.

Page 211: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

196

Fryer JD and Lukas RJ (1999) Noncompetitive functional inhibition at diverse, human nicotinic acetylcholine receptor subtypes by bupropion, phencyclidine, and ibogaine. J Pharmacol Exp Ther 288(1):88-92.

Fujieda M, Yamazaki H, Saito T, Kiyotani K, Gyamfi MA, Sakurai M, Dosaka-Akita H, Sawamura Y, Yokota J, Kunitoh H and Kamataki T (2004) Evaluation of CYP2A6 genetic polymorphisms as determinants of smoking behavior and tobacco-related lung cancer risk in male Japanese smokers. Carcinogenesis 25(12):2451-2458.

Fukami T, Nakajima M, Higashi E, Yamanaka H, McLeod HL and Yokoi T (2005a) A novel CYP2A6*20 allele found in African-American population produces a truncated protein lacking enzymatic activity. Biochem Pharmacol 70(5):801-808.

Fukami T, Nakajima M, Higashi E, Yamanaka H, Sakai H, McLeod HL and Yokoi T (2005b) Characterization of novel CYP2A6 polymorphic alleles (CYP2A6*18 and CYP2A6*19) that affect enzymatic activity. Drug Metab Dispos 33(8):1202-1210.

Fukami T, Nakajima M, Yamanaka H, Fukushima Y, McLeod HL and Yokoi T (2007) A novel duplication type of CYP2A6 gene in African-American population. Drug Metab Dispos 35(4):515-520.

Fukami T, Nakajima M, Yoshida R, Tsuchiya Y, Fujiki Y, Katoh M, McLeod HL and Yokoi T (2004) A novel polymorphism of human CYP2A6 gene CYP2A6*17 has an amino acid substitution (V365M) that decreases enzymatic activity in vitro and in vivo. Clin Pharmacol Ther 76(6):519-527.

Furnes B and Schlenk D (2004) Evaluation of xenobiotic N- and S-oxidation by variant flavin-containing monooxygenase 1 (FMO1) enzymes. Toxicol Sci 78(2):196-203.

Gaddnas H, Pietila K, Alila-Johansson A and Ahtee L (2002) Pineal melatonin and brain transmitter monoamines in CBA mice during chronic oral nicotine administration. Brain Res 957(1):76-83.

Garattini E, Mendel R, Romao MJ, Wright R and Terao M (2003) Mammalian molybdo-flavoenzymes, an expanding family of proteins: structure, genetics, regulation, function and pathophysiology. Biochem J 372(Pt 1):15-32.

Garvey AJ, Kinnunen T, Nordstrom BL, Utman CH, Doherty K, Rosner B and Vokonas PS (2000) Effects of nicotine gum dose by level of nicotine dependence. Nicotine Tob Res 2(1):53-63.

George TP, Vessicchio JC, Termine A, Jatlow PI, Kosten TR and O'Malley SS (2003) A preliminary placebo-controlled trial of selegiline hydrochloride for smoking cessation. Biol Psychiatry 53(2):136-143.

Gerlach M, Riederer P and Youdim MB (1992) The molecular pharmacology of L-deprenyl. Eur J Pharmacol 226(2):97-108.

Gerlach M, Youdim MB and Riederer P (1996) Pharmacology of selegiline. Neurology 47(6 Suppl 3):S137-145.

Gerra G, Garofano L, Zaimovic A, Moi G, Branchi B, Bussandri M, Brambilla F and Donnini C (2005) Association of the serotonin transporter promoter polymorphism with smoking behavior among adolescents. Am J Med Genet B Neuropsychiatr Genet 135B(1):73-78.

Gervot L, Rochat B, Gautier JC, Bohnenstengel F, Kroemer H, de Berardinis V, Martin H, Beaune P and de Waziers I (1999) Human CYP2B6: expression, inducibility and catalytic activities. Pharmacogenetics 9(3):295-306.

Ghanbari F, Rowland-Yeo K, Bloomer JC, Clarke SE, Lennard MS, Tucker GT and Rostami-Hodjegan A (2006) A critical evaluation of the experimental design of studies of mechanism based enzyme inhibition, with implications for in vitro-in vivo extrapolation. Curr Drug Metab 7(3):315-334.

Page 212: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

197

Ghosheh O and Hawes EM (2002) Microsomal N-glucuronidation of nicotine and cotinine: human hepatic interindividual, human intertissue, and interspecies hepatic variation. Drug Metab Dispos 30(12):1478-1483.

Gilbert DG, Meliska CJ, Williams CL and Jensen RA (1992) Subjective correlates of cigarette-smoking-induced elevations of peripheral beta-endorphin and cortisol. Psychopharmacology (Berl) 106(2):275-281.

Gilbert DG, Zuo Y, Browning RA, Shaw TM, Rabinovich NE, Gilbert-Johnson AM and Plath L (2003) Platelet monoamine oxidase B activity changes across 31 days of smoking abstinence. Nicotine Tob Res 5(6):813-819.

Glassman AH, Jackson WK, Walsh BT, Roose SP and Rosenfeld B (1984) Cigarette craving, smoking withdrawal, and clonidine. Science 226(4676):864-866.

Gonzales D, Rennard SI, Nides M, Oncken C, Azoulay S, Billing CB, Watsky EJ, Gong J, Williams KE and Reeves KR (2006) Varenicline, an alpha4beta2 nicotinic acetylcholine receptor partial agonist, vs sustained-release bupropion and placebo for smoking cessation: a randomized controlled trial. Jama 296(1):47-55.

Gonzalez S, Cascio MG, Fernandez-Ruiz J, Fezza F, Di Marzo V and Ramos JA (2002) Changes in endocannabinoid contents in the brain of rats chronically exposed to nicotine, ethanol or cocaine. Brain Res 954(1):73-81.

Gorrod JW and Hibberd AR (1982) The metabolism of nicotine-delta 1'(5')-iminium ion, in vivo and in vitro. Eur J Drug Metab Pharmacokinet 7(4):293-298.

Gossop M (1988) Clonidine and the treatment of the opiate withdrawal syndrome. Drug Alcohol Depend 21(3):253-259.

Gourlay SG, Stead LF and Benowitz NL (2004) Clonidine for smoking cessation. Cochrane Database Syst Rev(3):CD000058.

Grabus SD, Martin BR, Batman AM, Tyndale RF, Sellers E and Damaj MI (2005) Nicotine physical dependence and tolerance in the mouse following chronic oral administration. Psychopharmacology (Berl) 178(2-3):183-192.

Grimsby J, Toth M, Chen K, Kumazawa T, Klaidman L, Adams JD, Karoum F, Gal J and Shih JC (1997) Increased stress response and beta-phenylethylamine in MAOB-deficient mice. Nat Genet 17(2):206-210.

Gu DF, Hinks LJ, Morton NE and Day IN (2000) The use of long PCR to confirm three common alleles at the CYP2A6 locus and the relationship between genotype and smoking habit. Ann Hum Genet 64(Pt 5):383-390.

Gu J, Zhang QY, Genter MB, Lipinskas TW, Negishi M, Nebert DW and Ding X (1998) Purification and characterization of heterologously expressed mouse CYP2A5 and CYP2G1: role in metabolic activation of acetaminophen and 2,6-dichlorobenzonitrile in mouse olfactory mucosal microsomes. J Pharmacol Exp Ther 285(3):1287-1295.

Haberl M, Anwald B, Klein K, Weil R, Fuss C, Gepdiremen A, Zanger UM, Meyer UA and Wojnowski L (2005) Three haplotypes associated with CYP2A6 phenotypes in Caucasians. Pharmacogenet Genomics 15(9):609-624.

Hall SM, Humfleet GL, Reus VI, Munoz RF, Hartz DT and Maude-Griffin R (2002) Psychological intervention and antidepressant treatment in smoking cessation. Arch Gen Psychiatry 59(10):930-936.

Hansson E, Hoffmann PC and Schmiterloew CG (1964) Metabolism Of Nicotine In Mouse Tissue Slices. Acta Physiol Scand 61:380-392.

Harada N and Negishi M (1984) Mouse liver testosterone 15 alpha-hydroxylase (cytochrome P-450(15) alpha). Purification, regioselectivity, stereospecificity, and sex-dependent expression. J Biol Chem 259(2):1265-1271.

Hatchell PC and Collins AC (1980) The influence of genotype and sex on behavioral sensitivity to nicotine in mice. Psychopharmacology (Berl) 71(1):45-49.

Page 213: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

198

Hatsukami D, Lexau B, Nelson D, Pentel PR, Sofuoglu M and Goldman A (1998a) Effects of cotinine on cigarette self-administration. Psychopharmacology (Berl) 138(2):184-189.

Hatsukami D, Pentel PR, Jensen J, Nelson D, Allen SS, Goldman A and Rafael D (1998b) Cotinine: effects with and without nicotine. Psychopharmacology (Berl) 135(2):141-150.

Hatsukami DK, Rennard S, Jorenby D, Fiore M, Koopmeiners J, de Vos A, Horwith G and Pentel PR (2005) Safety and immunogenicity of a nicotine conjugate vaccine in current smokers. Clin Pharmacol Ther 78(5):456-467.

He XY, Shen J, Ding X, Lu AY and Hong JY (2004a) Identification of critical amino acid residues of human CYP2A13 for the metabolic activation of 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone, a tobacco-specific carcinogen. Drug Metab Dispos 32(12):1516-1521.

He XY, Shen J, Hu WY, Ding X, Lu AY and Hong JY (2004b) Identification of Val117 and Arg372 as critical amino acid residues for the activity difference between human CYP2A6 and CYP2A13 in coumarin 7-hydroxylation. Arch Biochem Biophys 427(2):143-153.

Hecht SS, Chen CB, Hirota N, Ornaf RM, Tso TC and Hoffmann D (1978) Tobacco-specific nitrosamines: formation from nicotine in vitro and during tobacco curing and carcinogenicity in strain A mice. J Natl Cancer Inst 60(4):819-824.

Hecht SS, Hatsukami DK, Bonilla LE and Hochalter JB (1999) Quantitation of 4-oxo-4-(3-pyridyl)butanoic acid and enantiomers of 4-hydroxy-4-(3-pyridyl)butanoic acid in human urine: A substantial pathway of nicotine metabolism. Chem Res Toxicol 12(2):172-179.

Hecht SS and Hoffmann D (1988) Tobacco-specific nitrosamines, an important group of carcinogens in tobacco and tobacco smoke. Carcinogenesis 9(6):875-884.

Hecht SS, Isaacs S and Trushin N (1994) Lung tumor induction in A/J mice by the tobacco smoke carcinogens 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone and benzo[a]pyrene: a potentially useful model for evaluation of chemopreventive agents. Carcinogenesis 15(12):2721-2725.

Hecht SS, Spratt TE and Trushin N (1988) Evidence for 4-(3-pyridyl)-4-oxobutylation of DNA in F344 rats treated with the tobacco-specific nitrosamines 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone and N'-nitrosonornicotine. Carcinogenesis 9(1):161-165.

Hecht SS, Trushin N, Reid-Quinn CA, Burak ES, Jones AB, Southers JL, Gombar CT, Carmella SG, Anderson LM and Rice JM (1993) Metabolism of the tobacco-specific nitrosamine 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone in the patas monkey: pharmacokinetics and characterization of glucuronide metabolites. Carcinogenesis 14(2):229-236.

Heinonen EH, Anttila MI, Karnani HL, Nyman LM, Vuorinen JA, Pyykko KA and Lammintausta RA (1997) Desmethylselegiline, a metabolite of selegiline, is an irreversible inhibitor of monoamine oxidase type B in humans. J Clin Pharmacol 37(7):602-609.

Heinonen EH, Anttila MI and Lammintausta RA (1994) Pharmacokinetic aspects of l-deprenyl (selegiline) and its metabolites. Clin Pharmacol Ther 56(6 Pt 2):742-749.

Heinonen EH and Myllyla V (1998) Safety of selegiline (deprenyl) in the treatment of Parkinson's disease. Drug Saf 19(1):11-22.

Henningfield JE and Keenan RM (1993) Nicotine delivery kinetics and abuse liability. J Consult Clin Psychol 61(5):743-750.

Herrera N, Franco R, Herrera L, Partidas A, Rolando R and Fagerstrom KO (1995) Nicotine gum, 2 and 4 mg, for nicotine dependence. A double-blind placebo-

Page 214: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

199

controlled trial within a behavior modification support program. Chest 108(2):447-451.

Hieda Y, Keyler DE, Ennifar S, Fattom A and Pentel PR (2000) Vaccination against nicotine during continued nicotine administration in rats: immunogenicity of the vaccine and effects on nicotine distribution to brain. Int J Immunopharmacol 22(10):809-819.

Higashi E, Fukami T, Itoh M, Kyo S, Inoue M, Yokoi T and Nakajima M (2007a) Human CYP2A6 is induced by estrogen via estrogen receptor. Drug Metab Dispos 35(10):1935-1941.

Higashi E, Nakajima M, Katoh M, Tokudome S and Yokoi T (2007b) Inhibitory effects of neurotransmitters and steroids on human CYP2A6. Drug Metab Dispos 35(4):508-514.

Hilleman DE, Mohiuddin SM and Delcore MG (1994) Comparison of fixed-dose transdermal nicotine, tapered-dose transdermal nicotine, and buspirone in smoking cessation. J Clin Pharmacol 34(3):222-224.

Ho M, Mwenifumbo J, Al Koudsi N, Okuyemi K, Ahluwalia J, Benowitz N and Tyndale R (2009) Association of Nicotine Metabolite Ratio and CYP2A6 Genotype With Smoking Cessation Treatment in African-American Light Smokers. Clin Pharmacol Ther.

Ho MK, Mwenifumbo JC, Zhao B, Gillam EM and Tyndale RF (2008) A novel CYP2A6 allele, CYP2A6*23, impairs enzyme function in vitro and in vivo and decreases smoking in a population of Black-African descent. Pharmacogenet Genomics 18(1):67-75.

Ho MK and Tyndale RF (2007) Overview of the pharmacogenomics of cigarette smoking. Pharmacogenomics J 7(2):81-98.

Hofmann MH, Blievernicht JK, Klein K, Saussele T, Schaeffeler E, Schwab M and Zanger UM (2008) Aberrant splicing caused by single nucleotide polymorphism c.516G>T [Q172H], a marker of CYP2B6*6, is responsible for decreased expression and activity of CYP2B6 in liver. J Pharmacol Exp Ther 325(1):284-292.

Houdi AA, Pierzchala K, Marson L, Palkovits M and Van Loon GR (1991) Nicotine-induced alteration in Tyr-Gly-Gly and Met-enkephalin in discrete brain nuclei reflects altered enkephalin neuron activity. Peptides 12(1):161-166.

Houtsmuller EJ, Thornton JA and Stitzer ML (2002) Effects of selegiline (L-deprenyl) during smoking and short-term abstinence. Psychopharmacology (Berl) 163(2):213-220.

Hughes JR, Stead LF and Lancaster T (2005) Nortriptyline for smoking cessation: a review. Nicotine Tob Res 7(4):491-499.

Hughes JR, Stead LF and Lancaster T (2007) Antidepressants for smoking cessation. Cochrane Database Syst Rev(1):CD000031.

Hukkanen J, Gourlay SG, Kenkare S and Benowitz NL (2005a) Influence of menstrual cycle on cytochrome P450 2A6 activity and cardiovascular effects of nicotine. Clin Pharmacol Ther 77(3):159-169.

Hukkanen J, Jacob P, 3rd and Benowitz NL (2005b) Metabolism and disposition kinetics of nicotine. Pharmacol Rev 57(1):79-115.

Hukkanen J, Jacob P, 3rd and Benowitz NL (2006) Effect of grapefruit juice on cytochrome P450 2A6 and nicotine renal clearance. Clin Pharmacol Ther 80(5):522-530.

Hurt RD, Sachs DP, Glover ED, Offord KP, Johnston JA, Dale LC, Khayrallah MA, Schroeder DR, Glover PN, Sullivan CR, Croghan IT and Sullivan PM (1997) A

Page 215: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

200

comparison of sustained-release bupropion and placebo for smoking cessation. N Engl J Med 337(17):1195-1202.

Hwa Jung B, Chul Chung B, Chung SJ and Shim CK (2001) Different pharmacokinetics of nicotine following intravenous administration of nicotine base and nicotine hydrogen tartarate in rats. J Control Release 77(3):183-190.

Ikeda K, Yoshisue K, Matsushima E, Nagayama S, Kobayashi K, Tyson CA, Chiba K and Kawaguchi Y (2000) Bioactivation of tegafur to 5-fluorouracil is catalyzed by cytochrome P-450 2A6 in human liver microsomes in vitro. Clin Cancer Res 6(11):4409-4415.

Imaoka S, Yamada T, Hiroi T, Hayashi K, Sakaki T, Yabusaki Y and Funae Y (1996) Multiple forms of human P450 expressed in Saccharomyces cerevisiae. Systematic characterization and comparison with those of the rat. Biochem Pharmacol 51(8):1041-1050.

Itoh K, Nakamura K, Kimura T, Itoh S and Kamataki T (1997) Molecular cloning of mouse liver flavin containing monooxygenase (FMO1) cDNA and characterization of the expression product: metabolism of the neurotoxin, 1,2,3,4-tetrahydroisoquinoline (TIQ). J Toxicol Sci 22(1):45-56.

Itzhak Y and Martin JL (1999) Effects of cocaine, nicotine, dizocipline and alcohol on mice locomotor activity: cocaine-alcohol cross-sensitization involves upregulation of striatal dopamine transporter binding sites. Brain Res 818(2):204-211.

Jackson KJ, Walters CL, Miles MF, Martin BR and Damaj MI (2009) Characterization of pharmacological and behavioral differences to nicotine in C57Bl/6 and DBA/2 mice. Neuropharmacology 57(4):347-355.

Jalas JR, Hecht SS and Murphy SE (2005) Cytochrome P450 enzymes as catalysts of metabolism of 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone, a tobacco specific carcinogen. Chem Res Toxicol 18(2):95-110.

Janmohamed A, Dolphin CT, Phillips IR and Shephard EA (2001) Quantification and cellular localization of expression in human skin of genes encoding flavin-containing monooxygenases and cytochromes P450. Biochem Pharmacol 62(6):777-786.

Janmohamed A, Hernandez D, Phillips IR and Shephard EA (2004) Cell-, tissue-, sex- and developmental stage-specific expression of mouse flavin-containing monooxygenases (Fmos). Biochem Pharmacol 68(1):73-83.

Jinno H, Tanaka-Kagawa T, Ohno A, Makino Y, Matsushima E, Hanioka N and Ando M (2003) Functional characterization of cytochrome P450 2B6 allelic variants. Drug Metab Dispos 31(4):398-403.

Jorenby DE, Hays JT, Rigotti NA, Azoulay S, Watsky EJ, Williams KE, Billing CB, Gong J and Reeves KR (2006) Efficacy of varenicline, an alpha4beta2 nicotinic acetylcholine receptor partial agonist, vs placebo or sustained-release bupropion for smoking cessation: a randomized controlled trial. Jama 296(1):56-63.

Jorenby DE, Leischow SJ, Nides MA, Rennard SI, Johnston JA, Hughes AR, Smith SS, Muramoto ML, Daughton DM, Doan K, Fiore MC and Baker TB (1999) A controlled trial of sustained-release bupropion, a nicotine patch, or both for smoking cessation. N Engl J Med 340(9):685-691.

Juvonen RO, Gynther J, Pasanen M, Alhava E and Poso A (2000) Pronounced differences in inhibition potency of lactone and non-lactone compounds for mouse and human coumarin 7-hydroxylases (CYP2A5 and CYP2A6). Xenobiotica 30(1):81-92.

Juvonen RO, Iwasaki M, Sueyoshi T and Negishi M (1993) Structural alteration of mouse P450coh by mutation of glycine-207 to proline: spin equilibrium, enzyme kinetics, and heat sensitivity. Biochem J 294 (Pt 1):31-34.

Page 216: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

201

Kaipainen P and Lang M (1985) Comparison of kidney, lung and liver coumarin 7-hydroxylases in phenobarbital pretreated DBA/2J and C57BL/6J mice. Biochem Pharmacol 34(11):1987-1991.

Kaipainen P, Nebert DW and Lang MA (1984) Purification and characterization of a microsomal cytochrome P-450 with high activity of coumarin 7-hydroxylase from mouse liver. Eur J Biochem 144(3):425-431.

Kaivosaari S, Toivonen P, Hesse LM, Koskinen M, Court MH and Finel M (2007) Nicotine glucuronidation and the human UDP-glucuronosyltransferase UGT2B10. Mol Pharmacol 72(3):761-768.

Kamiyama Y, Matsubara T, Yoshinari K, Nagata K, Kamimura H and Yamazoe Y (2007) Role of human hepatocyte nuclear factor 4alpha in the expression of drug-metabolizing enzymes and transporters in human hepatocytes assessed by use of small interfering RNA. Drug Metab Pharmacokinet 22(4):287-298.

Kassiou M, Eberl S, Meikle SR, Birrell A, Constable C, Fulham MJ, Wong DF and Musachio JL (2001) In vivo imaging of nicotinic receptor upregulation following chronic (-)-nicotine treatment in baboon using SPECT. Nucl Med Biol 28(2):165-175.

Katoh M, Tateno C, Yoshizato K and Yokoi T (2008) Chimeric mice with humanized liver. Toxicology 246(1):9-17.

Keenan RM, Hatsukami DK, Pentel PR, Thompson TN and Grillo MA (1994) Pharmacodynamic effects of cotinine in abstinent cigarette smokers. Clin Pharmacol Ther 55(5):581-590.

Khalil AA, Steyn S and Castagnoli N, Jr. (2000) Isolation and characterization of a monoamine oxidase inhibitor from tobacco leaves. Chem Res Toxicol 13(1):31-35.

Kharasch ED, Hankins DC and Taraday JK (2000) Single-dose methoxsalen effects on human cytochrome P-450 2A6 activity. Drug Metab Dispos 28(1):28-33.

Kiang TK, Ensom MH and Chang TK (2005) UDP-glucuronosyltransferases and clinical drug-drug interactions. Pharmacol Ther 106(1):97-132.

Kieffer BL and Evans CJ (2009) Opioid receptors: from binding sites to visible molecules in vivo. Neuropharmacology 56 Suppl 1:205-212.

Kiguchi N, Maeda T, Tsuruga M, Yamamoto A, Yamamoto C, Ozaki M and Kishioka S (2008) Involvement of spinal Met-enkephalin in nicotine-induced antinociception in mice. Brain Res 1189:70-77.

Killen JD, Fortmann SP, Davis L, Strausberg L and Varady A (1999) Do heavy smokers benefit from higher dose nicotine patch therapy? Exp Clin Psychopharmacol 7(3):226-233.

Killen JD, Fortmann SP, Schatzberg AF, Hayward C, Sussman L, Rothman M, Strausberg L and Varady A (2000) Nicotine patch and paroxetine for smoking cessation. J Consult Clin Psychol 68(5):883-889.

Kim D, Wu ZL and Guengerich FP (2005) Analysis of coumarin 7-hydroxylation activity of cytochrome P450 2A6 using random mutagenesis. J Biol Chem 280(48):40319-40327.

Kimonen T, Juvonen RO, Alhava E and Pasanen M (1995) The inhibition of CYP enzymes in mouse and human liver by pilocarpine. Br J Pharmacol 114(4):832-836.

Kita T, Nakashima T, Shirase M, Asahina M and Kurogochi Y (1988) Effects of nicotine on ambulatory activity in mice. Jpn J Pharmacol 46(2):141-146.

Kita T, Okamoto M and Nakashima T (1992) Nicotine-induced sensitization to ambulatory stimulant effect produced by daily administration into the ventral tegmental area and the nucleus accumbens in rats. Life Sci 50(8):583-590.

Page 217: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

202

Kitagawa K, Kunugita N, Katoh T, Yang M and Kawamoto T (1999) The significance of the homozygous CYP2A6 deletion on nicotine metabolism: a new genotyping method of CYP2A6 using a single PCR-RFLP. Biochem Biophys Res Commun 262(1):146-151.

Kitagawa K, Kunugita N, Kitagawa M and Kawamoto T (2001) CYP2A6*6, a novel polymorphism in cytochrome p450 2A6, has a single amino acid substitution (R128Q) that inactivates enzymatic activity. J Biol Chem 276(21):17830-17835.

Kiyotani K, Fujieda M, Yamazaki H, Shimada T, Guengerich FP, Parkinson A, Nakagawa K, Ishizaki T and Kamataki T (2002) Twenty one novel single nucleotide polymorphisms (SNPs) of the CYP2A6 gene in Japanese and Caucasians. Drug Metab Pharmacokinet 17(5):482-487.

Kiyotani K, Yamazaki H, Fujieda M, Iwano S, Matsumura K, Satarug S, Ujjin P, Shimada T, Guengerich FP, Parkinson A, Honda G, Nakagawa K, Ishizaki T and Kamataki T (2003) Decreased coumarin 7-hydroxylase activities and CYP2A6 expression levels in humans caused by genetic polymorphism in CYP2A6 promoter region (CYP2A6*9). Pharmacogenetics 13(11):689-695.

Koenigs LL, Peter RM, Thompson SJ, Rettie AE and Trager WF (1997) Mechanism-based inactivation of human liver cytochrome P450 2A6 by 8-methoxypsoralen. Drug Metab Dispos 25(12):1407-1415.

Kojo A, Honkakoski P, Jarvinen P, Pelkonen O and Lang M (1989) Preferential inhibition of mouse hepatic coumarin 7-hydroxylase by inhibitors of steroid metabolizing monooxygenases. Pharmacol Toxicol 65(2):104-109.

Koopmans JR, Slutske WS, Heath AC, Neale MC and Boomsma DI (1999) The genetics of smoking initiation and quantity smoked in Dutch adolescent and young adult twins. Behav Genet 29(6):383-393.

Kornitzer M, Kittel F, Dramaix M and Bourdoux P (1987) A double blind study of 2 mg versus 4 mg nicotine-gum in an industrial setting. J Psychosom Res 31(2):171-176.

Koskela S, Hakkola J, Hukkanen J, Pelkonen O, Sorri M, Saranen A, Anttila S, Fernandez-Salguero P, Gonzalez F and Raunio H (1999) Expression of CYP2A genes in human liver and extrahepatic tissues. Biochem Pharmacol 57(12):1407-1413.

Krause RJ, Lash LH and Elfarra AA (2003) Human kidney flavin-containing monooxygenases and their potential roles in cysteine s-conjugate metabolism and nephrotoxicity. J Pharmacol Exp Ther 304(1):185-191.

Krishnan-Sarin S, Meandzija B and O'Malley S (2003) Naltrexone and nicotine patch smoking cessation: a preliminary study. Nicotine Tob Res 5(6):851-857.

Krueger SK and Williams DE (2005) Mammalian flavin-containing monooxygenases: structure/function, genetic polymorphisms and role in drug metabolism. Pharmacol Ther 106(3):357-387.

Kuehl GE and Murphy SE (2003) N-glucuronidation of nicotine and cotinine by human liver microsomes and heterologously expressed UDP-glucuronosyltransferases. Drug Metab Dispos 31(11):1361-1368.

Kuehn BM (2008) Updated US smoking cessation guideline advises counseling, combining therapies. Jama 299(23):2736.

Kuroki T, Tsutsumi T, Hirano M, Matsumoto T, Tatebayashi Y, Nishiyama K, Uchimura H, Shiraishi A, Nakahara T and Nakamura K (1990) Behavioral sensitization to beta-phenylethylamine (PEA): enduring modifications of specific dopaminergic neuron systems in the rat. Psychopharmacology (Berl) 102(1):5-10.

Page 218: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

203

Kurosaki M, Demontis S, Barzago MM, Garattini E and Terao M (1999) Molecular cloning of the cDNA coding for mouse aldehyde oxidase: tissue distribution and regulation in vivo by testosterone. Biochem J 341 (Pt 1):71-80.

Kyerematen GA, Morgan M, Warner G, Martin LF and Vesell ES (1990) Metabolism of nicotine by hepatocytes. Biochem Pharmacol 40(8):1747-1756.

Laine K, Anttila M, Huupponen R, Maki-Ikola O and Heinonen E (2000) Multiple-dose pharmacokinetics of selegiline and desmethylselegiline suggest saturable tissue binding. Clin Neuropharmacol 23(1):22-27.

Lamba V, Lamba J, Yasuda K, Strom S, Davila J, Hancock ML, Fackenthal JD, Rogan PK, Ring B, Wrighton SA and Schuetz EG (2003) Hepatic CYP2B6 expression: gender and ethnic differences and relationship to CYP2B6 genotype and CAR (constitutive androstane receptor) expression. J Pharmacol Exp Ther 307(3):906-922.

Lang MA, Juvonen R, Jarvinen P, Honkakoski P and Raunio H (1989) Mouse liver P450Coh: genetic regulation of the pyrazole-inducible enzyme and comparison with other P450 isoenzymes. Arch Biochem Biophys 271(1):139-148.

Le Foll B and Goldberg SR (2004) Rimonabant, a CB1 antagonist, blocks nicotine-conditioned place preferences. Neuroreport 15(13):2139-2143.

Learned-Coughlin SM, Bergstrom M, Savitcheva I, Ascher J, Schmith VD and Langstrom B (2003) In vivo activity of bupropion at the human dopamine transporter as measured by positron emission tomography. Biol Psychiatry 54(8):800-805.

Lee AM, Jepson C, Shields PG, Benowitz N, Lerman C and Tyndale RF (2007) CYP2B6 genotype does not alter nicotine metabolism, plasma levels, or abstinence with nicotine replacement therapy. Cancer Epidemiol Biomarkers Prev 16(6):1312-1314.

Lee BL, Benowitz NL and Jacob P, 3rd (1987) Influence of tobacco abstinence on the disposition kinetics and effects of nicotine. Clin Pharmacol Ther 41(4):474-479.

Lerman C, Tyndale R, Patterson F, Wileyto EP, Shields PG, Pinto A and Benowitz N (2006) Nicotine metabolite ratio predicts efficacy of transdermal nicotine for smoking cessation. Clin Pharmacol Ther 79(6):600-608.

Lesage MG, Keyler DE, Burroughs D and Pentel PR (2007) Effects of pregnancy on nicotine self-administration and nicotine pharmacokinetics in rats. Psychopharmacology (Berl) 194(3):413-421.

Levai F, Fejer E, Szeleczky G, Szabo A, Eros-Takacsy T, Hajdu F, Szebeni G, Szatmari I and Hermecz I (2004) In vitro formation of selegiline-N-oxide as a metabolite of selegiline in human, hamster, mouse, rat, guinea-pig, rabbit and dog. Eur J Drug Metab Pharmacokinet 29(3):169-178.

Levin ED, Petro A, Pollard N, Christopher NC, Strauss M, Avery J, Nicholson J and Rezvani AH (2008) Nicotinic alpha7- or beta2-containing receptor knockout: Effects on radial-arm maze learning and long-term nicotine consumption in mice. Behav Brain Res.

Lewis A, Miller JH and Lea RA (2007) Monoamine oxidase and tobacco dependence. Neurotoxicology 28(1):182-195.

Li MD, Beuten J, Ma JZ, Payne TJ, Lou XY, Garcia V, Duenes AS, Crews KM and Elston RC (2005) Ethnic- and gender-specific association of the nicotinic acetylcholine receptor alpha4 subunit gene (CHRNA4) with nicotine dependence. Hum Mol Genet 14(9):1211-1219.

Li MD, Cheng R, Ma JZ and Swan GE (2003) A meta-analysis of estimated genetic and environmental effects on smoking behavior in male and female adult twins. Addiction 98(1):23-31.

Page 219: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

204

Li XC, Karadsheh MS, Jenkins PM, Brooks JC, Drapeau JA, Shah MS, Lautner MA and Stitzel JA (2007) Chromosomal loci that influence oral nicotine consumption in C57BL/6J x C3H/HeJ F2 intercross mice. Genes Brain Behav 6(5):401-410.

Lindberg RL, Juvonen R and Negishi M (1992) Molecular characterization of the murine Coh locus: an amino acid difference at position 117 confers high and low coumarin 7-hydroxylase activity in P450coh. Pharmacogenetics 2(1):32-37.

Ling D, Niu T, Feng Y, Xing H and Xu X (2004) Association between polymorphism of the dopamine transporter gene and early smoking onset: an interaction risk on nicotine dependence. J Hum Genet 49(1):35-39.

Lockman PR, McAfee G, Geldenhuys WJ, Van der Schyf CJ, Abbruscato TJ and Allen DD (2005) Brain uptake kinetics of nicotine and cotinine after chronic nicotine exposure. J Pharmacol Exp Ther 314(2):636-642.

Lush IE and Holland G (1988) The genetics of tasting in mice. V. Glycine and cycloheximide. Genet Res 52(3):207-212.

Maenpaa J, Juvonen R, Raunio H, Rautio A and Pelkonen O (1994) Metabolic interactions of methoxsalen and coumarin in humans and mice. Biochem Pharmacol 48(7):1363-1369.

Maenpaa J, Sigusch H, Raunio H, Syngelma T, Vuorela P, Vuorela H and Pelkonen O (1993) Differential inhibition of coumarin 7-hydroxylase activity in mouse and human liver microsomes. Biochem Pharmacol 45(5):1035-1042.

Maes HH, Sullivan PF, Bulik CM, Neale MC, Prescott CA, Eaves LJ and Kendler KS (2004) A twin study of genetic and environmental influences n tobacco initiation, regular tobacco use and nicotine dependence. Psychol Med 34(7):1251-1261.

Magyar K, Szatmary I, Szebeni G and Lengyel J (2007) Pharmacokinetic studies of (-)-deprenyl and some of its metabolites in mouse. J Neural Transm Suppl(72):165-173.

Malaiyandi V, Lerman C, Benowitz NL, Jepson C, Patterson F and Tyndale RF (2006) Impact of CYP2A6 genotype on pretreatment smoking behaviour and nicotine levels from and usage of nicotine replacement therapy. Mol Psychiatry 11(4):400-409.

Malaiyandi V, Sellers EM and Tyndale RF (2005) Implications of CYP2A6 genetic variation for smoking behaviors and nicotine dependence. Clin Pharmacol Ther 77(3):145-158.

Malin DH and Goyarzu P (2009) Rodent models of nicotine withdrawal syndrome. Handb Exp Pharmacol(192):401-434.

Malin DH, Lake JR, Carter VA, Cunningham JS and Wilson OB (1993) Naloxone precipitates nicotine abstinence syndrome in the rat. Psychopharmacology (Berl) 112(2-3):339-342.

Malin DH, Lake JR, Payne MC, Short PE, Carter VA, Cunningham JS and Wilson OB (1996) Nicotine alleviation of nicotine abstinence syndrome is naloxone-reversible. Pharmacol Biochem Behav 53(1):81-85.

Manhem P, Nilsson LH, Moberg AL, Wadstein J and Hokfelt B (1985) Alcohol withdrawal: effects of clonidine treatment on sympathetic activity, the renin-aldosterone system, and clinical symptoms. Alcohol Clin Exp Res 9(3):238-243.

Mansner R and Mattila MJ (1977) Pharmacokinetics of nicotine in adult and infant mice. Med Biol 55(6):317-324.

Markou A, Paterson NE and Semenova S (2004) Role of gamma-aminobutyric acid (GABA) and metabotropic glutamate receptors in nicotine reinforcement: potential pharmacotherapies for smoking cessation. Ann N Y Acad Sci 1025:491-503.

Page 220: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

205

Marks MJ, Burch JB and Collins AC (1983) Genetics of nicotine response in four inbred strains of mice. J Pharmacol Exp Ther 226(1):291-302.

Marks MJ, Miner L, Burch JB, Fulker DW and Collins AC (1984) A diallel analysis of nicotine-induced hypothermia. Pharmacol Biochem Behav 21(6):953-959.

Marks MJ, Romm E, Bealer SM and Collins AC (1985) A test battery for measuring nicotine effects in mice. Pharmacol Biochem Behav 23(2):325-330.

Marks MJ, Romm E, Campbell SM and Collins AC (1989) Variation of nicotinic binding sites among inbred strains. Pharmacol Biochem Behav 33(3):679-689.

Marks MJ, Romm E, Gaffney DK and Collins AC (1986) Nicotine-induced tolerance and receptor changes in four mouse strains. J Pharmacol Exp Ther 237(3):809-819.

Martignoni M, Groothuis GM and de Kanter R (2006) Species differences between mouse, rat, dog, monkey and human CYP-mediated drug metabolism, inhibition and induction. Expert Opin Drug Metab Toxicol 2(6):875-894.

Marubio LM, del Mar Arroyo-Jimenez M, Cordero-Erausquin M, Lena C, Le Novere N, de Kerchove d'Exaerde A, Huchet M, Damaj MI and Changeux JP (1999) Reduced antinociception in mice lacking neuronal nicotinic receptor subunits. Nature 398(6730):805-810.

Maser E, Richter E and Friebertshauser J (1996) The identification of 11 beta-hydroxysteroid dehydrogenase as carbonyl reductase of the tobacco-specific nitrosamine 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone. Eur J Biochem 238(2):484-489.

Matsuda Y, Yamakawa K, Saoo K, Hosokawa K, Yokohira M, Kuno T, Iwai J, Shirai T, Obika K, Kamataki T and Imaida K (2007) CYP2A6 overexpression in human lung cancers correlates with a high malignant status. Oncol Rep 18(1):53-57.

Mays DC, Hilliard JB, Wong DD, Chambers MA, Park SS, Gelboin HV and Gerber N (1990) Bioactivation of 8-methoxypsoralen and irreversible inactivation of cytochrome P-450 in mouse liver microsomes: modification by monoclonal antibodies, inhibition of drug metabolism and distribution of covalent adducts. J Pharmacol Exp Ther 254(2):720-731.

McCallum SE, Collins AC, Paylor R and Marks MJ (2006) Deletion of the beta 2 nicotinic acetylcholine receptor subunit alters development of tolerance to nicotine and eliminates receptor upregulation. Psychopharmacology (Berl) 184(3-4):314-327.

McMorrow MJ and Foxx RM (1983) Nicotine's role in smoking: an analysis of nicotine regulation. Psychol Bull 93(2):302-327.

McRobbie H, Lee M and Juniper Z (2005) Non-nicotine pharmacotherapies for smoking cessation. Respir Med 99(10):1203-1212.

Meger M, Meger-Kossien I, Schuler-Metz A, Janket D and Scherer G (2002) Simultaneous determination of nicotine and eight nicotine metabolites in urine of smokers using liquid chromatography-tandem mass spectrometry. J Chromatogr B Analyt Technol Biomed Life Sci 778(1-2):251-261.

Megerdichian CL, Rees VW, Wayne GF and Connolly GN (2007) Internal tobacco industry research on olfactory and trigeminal nerve response to nicotine and other smoke components. Nicotine Tob Res 9(11):1119-1129.

Meliska CJ, Bartke A, McGlacken G and Jensen RA (1995) Ethanol, nicotine, amphetamine, and aspartame consumption and preferences in C57BL/6 and DBA/2 mice. Pharmacol Biochem Behav 50(4):619-626.

Merlo Pich E, Chiamulera C and Carboni L (1999) Molecular mechanisms of the positive reinforcing effect of nicotine. Behav Pharmacol 10(6-7):587-596.

Merritt LL, Martin BR, Walters C, Lichtman AH and Damaj MI (2008) The endogenous cannabinoid system modulates nicotine reward and dependence. J Pharmacol Exp Ther 326(2):483-492.

Page 221: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

206

Messina ES, Tyndale RF and Sellers EM (1997) A major role for CYP2A6 in nicotine C-oxidation by human liver microsomes. J Pharmacol Exp Ther 282(3):1608-1614.

Meyer MD, Decker MW, Rueter LE, Anderson DJ, Dart MJ, Kim KH, Sullivan JP and Williams M (2000) The identification of novel structural compound classes exhibiting high affinity for neuronal nicotinic acetylcholine receptors and analgesic efficacy in preclinical models of pain. Eur J Pharmacol 393(1-3):171-177.

Miksys S, Lerman C, Shields PG, Mash DC and Tyndale RF (2003) Smoking, alcoholism and genetic polymorphisms alter CYP2B6 levels in human brain. Neuropharmacology 45(1):122-132.

Miller RP, Rotenberg KS and Adir J (1977) Effect of dose on the pharmacokinetics of intravenous nicotine in the rat. Drug Metab Dispos 5(5):436-443.

Minematsu N, Nakamura H, Furuuchi M, Nakajima T, Takahashi S, Tateno H and Ishizaka A (2006) Limitation of cigarette consumption by CYP2A6*4, *7 and *9 polymorphisms. Eur Respir J 27(2):289-292.

Miyazaki M, Yamazaki H, Takeuchi H, Saoo K, Yokohira M, Masumura K, Nohmi T, Funae Y, Imaida K and Kamataki T (2005) Mechanisms of chemopreventive effects of 8-methoxypsoralen against 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone-induced mouse lung adenomas. Carcinogenesis 26(11):1947-1955.

Modesto-Lowe V and Van Kirk J (2002) Clinical uses of naltrexone: a review of the evidence. Exp Clin Psychopharmacol 10(3):213-227.

Molander L and Lunell E (2001) Pharmacokinetic investigation of a nicotine sublingual tablet. Eur J Clin Pharmacol 56(11):813-819.

Murphy S, Raulinaitis V and Brown KM (2005a) Nicotine 5'-oxidation and methyl oxidation by P450 2A enzymes. Drug Metab Dispos 33(8):1166-1173.

Murphy SE, Johnson LM and Pullo DA (1999) Characterization of multiple products of cytochrome P450 2A6-catalyzed cotinine metabolism. Chem Res Toxicol 12(7):639-645.

Murphy SE, Raulinaitis V and Brown KM (2005b) Nicotine 5'-oxidation and methyl oxidation by P450 2A enzymes. Drug Metab Dispos 33(8):1166-1173.

Mwenifumbo JC, Al Koudsi N, Ho MK, Zhou Q, Hoffmann EB, Sellers EM and Tyndale RF (2008a) Novel and established CYP2A6 alleles impair in vivo nicotine metabolism in a population of Black African descent. Hum Mutat 29(5):679-688.

Mwenifumbo JC, Lessov-Schlaggar CN, Zhou Q, Krasnow RE, Swan GE, Benowitz NL and Tyndale RF (2008b) Identification of novel CYP2A6*1B variants: the CYP2A6*1B allele is associated with faster in vivo nicotine metabolism. Clin Pharmacol Ther 83(1):115-121.

Mwenifumbo JC, Sellers EM and Tyndale RF (2007) Nicotine metabolism and CYP2A6 activity in a population of black African descent: impact of gender and light smoking. Drug Alcohol Depend 89(1):24-33.

Mwenifumbo JC and Tyndale RF (2007) Genetic variability in CYP2A6 and the pharmacokinetics of nicotine. Pharmacogenomics 8(10):1385-1402.

Mwenifumbo JC and Tyndale RF (2009) Molecular genetics of nicotine metabolism. Handb Exp Pharmacol(192):235-259.

Nakajima M, Fukami T, Yamanaka H, Higashi E, Sakai H, Yoshida R, Kwon JT, McLeod HL and Yokoi T (2006a) Comprehensive evaluation of variability in nicotine metabolism and CYP2A6 polymorphic alleles in four ethnic populations. Clin Pharmacol Ther 80(3):282-297.

Nakajima M, Itoh M, Yamanaka H, Fukami T, Tokudome S, Yamamoto Y, Yamamoto H and Yokoi T (2006b) Isoflavones inhibit nicotine C-oxidation catalyzed by human CYP2A6. J Clin Pharmacol 46(3):337-344.

Page 222: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

207

Nakajima M, Tanaka E, Kwon JT and Yokoi T (2002) Characterization of nicotine and cotinine N-glucuronidations in human liver microsomes. Drug Metab Dispos 30(12):1484-1490.

Nakajima M, Yamamoto T, Nunoya K, Yokoi T, Nagashima K, Inoue K, Funae Y, Shimada N, Kamataki T and Kuroiwa Y (1996a) Characterization of CYP2A6 involved in 3'-hydroxylation of cotinine in human liver microsomes. J Pharmacol Exp Ther 277(2):1010-1015.

Nakajima M, Yamamoto T, Nunoya K, Yokoi T, Nagashima K, Inoue K, Funae Y, Shimada N, Kamataki T and Kuroiwa Y (1996b) Role of human cytochrome P4502A6 in C-oxidation of nicotine. Drug Metab Dispos 24(11):1212-1217.

Nakajima M and Yokoi T (2005) Interindividual variability in nicotine metabolism: C-oxidation and glucuronidation. Drug Metab Pharmacokinet 20(4):227-235.

Nakayama H, Okuda H, Nakashima T, Imaoka S and Funae Y (1993) Nicotine metabolism by rat hepatic cytochrome P450s. Biochem Pharmacol 45(12):2554-2556.

National CfHS (2007) Fast Stats: Smoking, (Services UDoHaH ed). Neurath GB and Pein FG (1987) Gas chromatographic determination of trans-3'-

hydroxycotinine, a major metabolite of nicotine in smokers. J Chromatogr 415(2):400-406.

Niaura R, Spring B, Borrelli B, Hedeker D, Goldstein MG, Keuthen N, DePue J, Kristeller J, Ockene J, Prochazka A, Chiles JA and Abrams DB (2002) Multicenter trial of fluoxetine as an adjunct to behavioral smoking cessation treatment. J Consult Clin Psychol 70(4):887-896.

Nides M, Oncken C, Gonzales D, Rennard S, Watsky EJ, Anziano R and Reeves KR (2006) Smoking cessation with varenicline, a selective alpha4beta2 nicotinic receptor partial agonist: results from a 7-week, randomized, placebo- and bupropion-controlled trial with 1-year follow-up. Arch Intern Med 166(15):1561-1568.

Nishikawa A, Mori Y, Lee IS, Tanaka T and Hirose M (2004) Cigarette smoking, metabolic activation and carcinogenesis. Curr Drug Metab 5(5):363-373.

Nishimura M and Naito S (2006) Tissue-specific mRNA expression profiles of human phase I metabolizing enzymes except for cytochrome P450 and phase II metabolizing enzymes. Drug Metab Pharmacokinet 21(5):357-374.

Nunoya K, Yokoi Y, Kimura K, Kodama T, Funayama M, Inoue K, Nagashima K, Funae Y, Shimada N, Green C and Kamataki T (1996) (+)-cis-3,5-dimethyl-2-(3-pyridyl) thiazolidin-4-one hydrochloride (SM-12502) as a novel substrate for cytochrome P450 2A6 in human liver microsomes. J Pharmacol Exp Ther 277(2):768-774.

Nutt DJ, Forshall S, Bell C, Rich A, Sandford J, Nash J and Argyropoulos S (1999) Mechanisms of action of selective serotonin reuptake inhibitors in the treatment of psychiatric disorders. Eur Neuropsychopharmacol 9 Suppl 3:S81-86.

O'Malley SS, Cooney JL, Krishnan-Sarin S, Dubin JA, McKee SA, Cooney NL, Blakeslee A, Meandzija B, Romano-Dahlgard D, Wu R, Makuch R and Jatlow P (2006) A controlled trial of naltrexone augmentation of nicotine replacement therapy for smoking cessation. Arch Intern Med 166(6):667-674.

Obach RS (2004) Potent inhibition of human liver aldehyde oxidase by raloxifene. Drug Metab Dispos 32(1):89-97.

Okubo C, Morishita Y, Minami Y, Ishiyama T, Kano J, Iijima T and Noguchi M (2005) Phenotypic characteristics of mouse lung adenoma induced by 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone. Mol Carcinog 42(2):121-126.

Olsson C, Anney R, Forrest S, Patton G, Coffey C, Cameron T, Hassett A and Williamson R (2004) Association between dependent smoking and a

Page 223: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

208

polymorphism in the tyrosine hydroxylase gene in a prospective population-based study of adolescent health. Behav Genet 34(1):85-91.

Oncken C, Gonzales D, Nides M, Rennard S, Watsky E, Billing CB, Anziano R and Reeves K (2006) Efficacy and safety of the novel selective nicotinic acetylcholine receptor partial agonist, varenicline, for smoking cessation. Arch Intern Med 166(15):1571-1577.

Oncken C, Watsky E, Reeves K and Anziano R (2005) Varenicline is efficacious and well tolerated in promoting smoking cessation: results from a 7-week, randomized, placebo- and bupropion-controlled trial, in Proceedings of the 11 Annual Meeting of the Society for Research on Nicotine and Tobacco, Prague, Czech Republic.

Oreland L, Fowler CJ and Schalling D (1981) Low platelet monoamine oxidase activity in cigarette smokers. Life Sci 29(24):2511-2518.

Orimo A, Inoue S, Ikeda K, Noji S and Muramatsu M (1995) Molecular cloning, structure, and expression of mouse estrogen-responsive finger protein Efp. Co-localization with estrogen receptor mRNA in target organs. J Biol Chem 270(41):24406-24413.

Ornish SA, Zisook S and McAdams LA (1988) Effects of transdermal clonidine treatment on withdrawal symptoms associated with smoking cessation. A randomized, controlled trial. Arch Intern Med 148(9):2027-2031.

Osawa Y and Pohl LR (1989) Covalent bonding of the prosthetic heme to protein: a potential mechanism for the suicide inactivation or activation of hemoproteins. Chem Res Toxicol 2(3):131-141.

Oscarson M, McLellan RA, Asp V, Ledesma M, Bernal Ruiz ML, Sinues B, Rautio A and Ingelman-Sundberg M (2002) Characterization of a novel CYP2A7/CYP2A6 hybrid allele (CYP2A6*12) that causes reduced CYP2A6 activity. Hum Mutat 20(4):275-283.

Oscarson M, McLellan RA, Gullsten H, Agundez JA, Benitez J, Rautio A, Raunio H, Pelkonen O and Ingelman-Sundberg M (1999) Identification and characterisation of novel polymorphisms in the CYP2A locus: implications for nicotine metabolism. FEBS Lett 460(2):321-327.

Overby LH, Carver GC and Philpot RM (1997) Quantitation and kinetic properties of hepatic microsomal and recombinant flavin-containing monooxygenases 3 and 5 from humans. Chem Biol Interact 106(1):29-45.

Paperwalla KN, Levin TT, Weiner J and Saravay SM (2004) Smoking and depression. Med Clin North Am 88(6):1483-1494, x-xi.

Papke RL and Heinemann SF (1994) Partial agonist properties of cytisine on neuronal nicotinic receptors containing the beta 2 subunit. Mol Pharmacol 45(1):142-149.

Paterson IA, Davis BA, Durden DA, Juorio AV, Yu PH, Ivy G, Milgram W, Mendonca A, Wu P and Boulton AA (1995) Inhibition of MAO-B by (-)-deprenyl alters dopamine metabolism in the macaque (Macaca facicularis) brain. Neurochem Res 20(12):1503-1510.

Paterson IA, Juorio AV, Berry MD and Zhu MY (1991) Inhibition of monoamine oxidase-B by (-)-deprenyl potentiates neuronal responses to dopamine agonists but does not inhibit dopamine catabolism in the rat striatum. J Pharmacol Exp Ther 258(3):1019-1026.

Paterson IA, Juorio AV and Boulton AA (1990) 2-Phenylethylamine: a modulator of catecholamine transmission in the mammalian central nervous system? J Neurochem 55(6):1827-1837.

Patterson F, Schnoll RA, Wileyto EP, Pinto A, Epstein LH, Shields PG, Hawk LW, Tyndale RF, Benowitz N and Lerman C (2008) Toward personalized therapy for

Page 224: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

209

smoking cessation: a randomized placebo-controlled trial of bupropion. Clin Pharmacol Ther 84(3):320-325.

Pelkonen O, Rautio A, Raunio H and Pasanen M (2000) CYP2A6: a human coumarin 7-hydroxylase. Toxicology 144(1-3):139-147.

Pentel PR, Malin DH, Ennifar S, Hieda Y, Keyler DE, Lake JR, Milstein JR, Basham LE, Coy RT, Moon JW, Naso R and Fattom A (2000) A nicotine conjugate vaccine reduces nicotine distribution to brain and attenuates its behavioral and cardiovascular effects in rats. Pharmacol Biochem Behav 65(1):191-198.

Perkins KA, Gerlach D, Vender J, Grobe J, Meeker J and Hutchison S (2001) Sex differences in the subjective and reinforcing effects of visual and olfactory cigarette smoke stimuli. Nicotine Tob Res 3(2):141-150.

Perkins KA and Scott J (2008) Sex differences in long-term smoking cessation rates due to nicotine patch. Nicotine Tob Res 10(7):1245-1250.

Perry DC, Davila-Garcia MI, Stockmeier CA and Kellar KJ (1999) Increased nicotinic receptors in brains from smokers: membrane binding and autoradiography studies. J Pharmacol Exp Ther 289(3):1545-1552.

Petersen DR, Norris KJ and Thompson JA (1984) A comparative study of the disposition of nicotine and its metabolites in three inbred strains of mice. Drug Metab Dispos 12(6):725-731.

Peterson LA, Carmella SG and Hecht SS (1990) Investigations of metabolic precursors to hemoglobin and DNA adducts of 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone. Carcinogenesis 11(8):1329-1333.

Picciotto MR (2003) Nicotine as a modulator of behavior: beyond the inverted U. Trends Pharmacol Sci 24(9):493-499.

Pitarque M, Rodriguez-Antona C, Oscarson M and Ingelman-Sundberg M (2005) Transcriptional regulation of the human CYP2A6 gene. J Pharmacol Exp Ther 313(2):814-822.

Pitarque M, von Richter O, Oke B, Berkkan H, Oscarson M and Ingelman-Sundberg M (2001) Identification of a single nucleotide polymorphism in the TATA box of the CYP2A6 gene: impairment of its promoter activity. Biochem Biophys Res Commun 284(2):455-460.

Polasek TM, Elliot DJ, Somogyi AA, Gillam EM, Lewis BC and Miners JO (2006) An evaluation of potential mechanism-based inactivation of human drug metabolizing cytochromes P450 by monoamine oxidase inhibitors, including isoniazid. Br J Clin Pharmacol 61(5):570-584.

Poso A, Gynther J and Juvonen R (2001) A comparative molecular field analysis of cytochrome P450 2A5 and 2A6 inhibitors. J Comput Aided Mol Des 15(3):195-202.

Quattrocki E, Baird A and Yurgelun-Todd D (2000) Biological aspects of the link between smoking and depression. Harv Rev Psychiatry 8(3):99-110.

Rahnasto M, Raunio H, Poso A and Juvonen RO (2003) More potent inhibition of human CYP2A6 than mouse CYP2A5 enzyme activities by derivatives of phenylethylamine and benzaldehyde. Xenobiotica 33(5):529-539.

Rahnasto M, Raunio H, Poso A, Wittekindt C and Juvonen RO (2005) Quantitative structure-activity relationship analysis of inhibitors of the nicotine metabolizing CYP2A6 enzyme. J Med Chem 48(2):440-449.

Rao Y, Hoffmann E, Zia M, Bodin L, Zeman M, Sellers EM and Tyndale RF (2000) Duplications and defects in the CYP2A6 gene: identification, genotyping, and in vivo effects on smoking. Mol Pharmacol 58(4):747-755.

Page 225: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

210

Raunio H, Pokela N, Puhakainen K, Rahnasto M, Mauriala T, Auriola S and Juvonen RO (2008) Nicotine metabolism and urinary elimination in mouse: in vitro and in vivo. Xenobiotica 38(1):34-47.

Reynolds GP, Riederer P, Sandler M, Jellinger K and Seemann D (1978) Amphetamine and 2-phenylethylamine in post-mortem Parkinsonian brain after (-)deprenyl administration. J Neural Transm 43(3-4):271-277.

Ring HZ, Valdes AM, Nishita DM, Prasad S, Jacob P, 3rd, Tyndale RF, Swan GE and Benowitz NL (2007) Gene-gene interactions between CYP2B6 and CYP2A6 in nicotine metabolism. Pharmacogenet Genomics 17(12):1007-1015.

Robinson SF, Marks MJ and Collins AC (1996) Inbred mouse strains vary in oral self-selection of nicotine. Psychopharmacology (Berl) 124(4):332-339.

Rogers DT and Iwamoto ET (1993) Multiple spinal mediators in parenteral nicotine-induced antinociception. J Pharmacol Exp Ther 267(1):341-349.

Roose SP, Laghrissi-Thode F, Kennedy JS, Nelson JC, Bigger JT, Jr., Pollock BG, Gaffney A, Narayan M, Finkel MS, McCafferty J and Gergel I (1998) Comparison of paroxetine and nortriptyline in depressed patients with ischemic heart disease. Jama 279(4):287-291.

Rose JE, Behm FM, Westman EC, Bates JE and Salley A (2003) Pharmacologic and sensorimotor components of satiation in cigarette smoking. Pharmacol Biochem Behav 76(2):243-250.

Rose JE, Behm FM, Westman EC and Johnson M (2000) Dissociating nicotine and nonnicotine components of cigarette smoking. Pharmacol Biochem Behav 67(1):71-81.

Rose JE and Corrigall WA (1997) Nicotine self-administration in animals and humans: similarities and differences. Psychopharmacology (Berl) 130(1):28-40.

Rose JE, Tashkin DP, Ertle A, Zinser MC and Lafer R (1985) Sensory blockade of smoking satisfaction. Pharmacol Biochem Behav 23(2):289-293.

Rose JE, Zinser MC, Tashkin DP, Newcomb R and Ertle A (1984) Subjective response to cigarette smoking following airway anesthetization. Addict Behav 9(2):211-215.

Rostami-Hodjegan A and Tucker GT (2004) 'In silico' simulations to assess the 'in vivo' consequences of 'in vitro' metabolic drug-drug interactions. Drug Discov Today: Technol 1(4):441-448.

Rowell PP, Hurst HE, Marlowe C and Bennett BD (1983) Oral administration of nicotine: its uptake and distribution after chronic administration to mice. J Pharmacol Methods 9(4):249-261.

Russel MSH (1987) Nicotine intake and its regulation by smokers, in Advances in Behavioral Biology. Plenum Press, New York.

Sadeque AJ, Fisher MB, Korzekwa KR, Gonzalez FJ and Rettie AE (1997) Human CYP2C9 and CYP2A6 mediate formation of the hepatotoxin 4-ene-valproic acid. J Pharmacol Exp Ther 283(2):698-703.

Saules KK, Schuh LM, Arfken CL, Reed K, Kilbey MM and Schuster CR (2004) Double-blind placebo-controlled trial of fluoxetine in smoking cessation treatment including nicotine patch and cognitive-behavioral group therapy. Am J Addict 13(5):438-446.

Saura Marti J, Kettler R, Da Prada M and Richards JG (1990) Molecular neuroanatomy of MAO-A and MAO-B. J Neural Transm Suppl 32:49-53.

Schlicht KE, Michno N, Smith BD, Scott EE and Murphy SE (2007) Functional characterization of CYP2A13 polymorphisms. Xenobiotica 37(12):1439-1449.

Schneider LS, Tariot PN and Goldstein B (1994) Therapy with l-deprenyl (selegiline) and relation to abuse liability. Clin Pharmacol Ther 56(6 Pt 2):750-756.

Page 226: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

211

Schneider NG, Olmstead RE, Steinberg C, Sloan K, Daims RM and Brown HV (1996) Efficacy of buspirone in smoking cessation: a placebo-controlled trial. Clin Pharmacol Ther 60(5):568-575.

Schneider NG, Popek P, Jarvik ME and Gritz ER (1977) The use of nicotine gum during cessation of smoking. Am J Psychiatry 134(4):439-440.

Schnoll RA, Patterson F, Wileyto EP, Tyndale RF, Benowitz N and Lerman C (2009) Nicotine metabolic rate predicts successful smoking cessation with transdermal nicotine: a validation study. Pharmacol Biochem Behav 92(1):6-11.

Schoedel KA, Hoffmann EB, Rao Y, Sellers EM and Tyndale RF (2004) Ethnic variation in CYP2A6 and association of genetically slow nicotine metabolism and smoking in adult Caucasians. Pharmacogenetics 14(9):615-626.

Schoedel KA, Sellers EM, Palmour R and Tyndale RF (2003) Down-regulation of hepatic nicotine metabolism and a CYP2A6-like enzyme in African green monkeys after long-term nicotine administration. Mol Pharmacol 63(1):96-104.

Schoedel KA, Sellers EM and Tyndale RF (2001) Induction of CYP2B1/2 and nicotine metabolism by ethanol in rat liver but not rat brain. Biochem Pharmacol 62(8):1025-1036.

Sellers EM, Dortok D and Tyndale RF (2002) CYP2A6 inhibition increases plasma nicotine concentrations during nicotine patch and gum. Clin Pharmacol Ther 71(2):17.

Sellers EM, Kaplan HL and Tyndale RF (2000) Inhibition of cytochrome P450 2A6 increases nicotine's oral bioavailability and decreases smoking. Clin Pharmacol Ther 68(1):35-43.

Sellers EM, Ramamoorthy Y, Zeman MV, Djordjevic MV and Tyndale RF (2003a) The effect of methoxsalen on nicotine and 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone (NNK) metabolism in vivo. Nicotine Tob Res 5(6):891-899.

Sellers EM, Tyndale RF and Fernandes LC (2003b) Decreasing smoking behaviour and risk through CYP2A6 inhibition. Drug Discov Today 8(11):487-493.

Sharma U, Roberts ES and Hollenberg PF (1996) Inactivation of cytochrome P4502B1 by the monoamine oxidase inhibitors R-(-)-deprenyl and clorgyline. Drug Metab Dispos 24(6):669-675.

Shiffman S, Johnston JA, Khayrallah M, Elash CA, Gwaltney CJ, Paty JA, Gnys M, Evoniuk G and DeVeaugh-Geiss J (2000) The effect of bupropion on nicotine craving and withdrawal. Psychopharmacology (Berl) 148(1):33-40.

Shih JC, Chen K and Ridd MJ (1999) Monoamine oxidase: from genes to behavior. Annu Rev Neurosci 22:197-217.

Shin HS (1997) Metabolism of selegiline in humans. Identification, excretion, and stereochemistry of urine metabolites. Drug Metab Dispos 25(6):657-662.

Shoemaker JL, Ruckle MB, Mayeux PR and Prather PL (2005) Agonist-directed trafficking of response by endocannabinoids acting at CB2 receptors. J Pharmacol Exp Ther 315(2):828-838.

Shou M, Korzekwa KR, Brooks EN, Krausz KW, Gonzalez FJ and Gelboin HV (1997) Role of human hepatic cytochrome P450 1A2 and 3A4 in the metabolic activation of estrone. Carcinogenesis 18(1):207-214.

Siddens LK, Henderson MC, Vandyke JE, Williams DE and Krueger SK (2007) Characterization of mouse flavin-containing monooxygenase transcript levels in lung and liver, and activity of expressed isoforms. Biochem Pharmacol.

Silverman RB (1995) Mechanism-Based Enzyme Inactivation: Chemistry and Enzymology, in Methods in Enzymology (Purich DL ed) pp 240-283.

Page 227: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

212

Siu EC and Tyndale RF (2007a) Characterization and comparison of nicotine and cotinine metabolism in vitro and in vivo in DBA/2 and C57BL/6 mice. Mol Pharmacol 71(3):826-834.

Siu EC and Tyndale RF (2007b) Non-nicotinic therapies for smoking cessation. Annu Rev Pharmacol Toxicol 47:541-564.

Siu EC, Wildenauer DB and Tyndale RF (2006) Nicotine self-administration in mice is associated with rates of nicotine inactivation by CYP2A5. Psychopharmacology (Berl) 184(3-4):401-408.

Slemmer JE, Martin BR and Damaj MI (2000) Bupropion is a nicotinic antagonist. J Pharmacol Exp Ther 295(1):321-327.

Smolen A, Marks MJ, DeFries JC and Henderson ND (1994) Individual differences in sensitivity to nicotine in mice: response to six generations of selective breeding. Pharmacol Biochem Behav 49(3):531-540.

Snively TA, Ahijevych KL, Bernhard LA and Wewers ME (2000) Smoking behavior, dysphoric states and the menstrual cycle: results from single smoking sessions and the natural environment. Psychoneuroendocrinology 25(7):677-691.

Sparks JA and Pauly JR (1999) Effects of continuous oral nicotine administration on brain nicotinic receptors and responsiveness to nicotine in C57Bl/6 mice. Psychopharmacology (Berl) 141(2):145-153.

Spracklin DK, Thummel KE and Kharasch ED (1996) Human reductive halothane metabolism in vitro is catalyzed by cytochrome P450 2A6 and 3A4. Drug Metab Dispos 24(9):976-983.

Squires RF (1972) Multiple forms of monoamine oxidase in intact mitochondria as characterized by selective inhibitors and thermal stability: a comparison of eight mammalian species. Adv Biochem Psychopharmacol 5:355-370.

Stalhandske T (1970) The metabolism of nicotine and cotinine by a mouse liver preparation. Acta Physiol Scand 78(2):236-248.

Stalhandske T, Slanina P, Tjalve H, Hansson E and Schmiterlow CG (1969) Metabolism in vitro of 14C-nicotine in livers of foetal, newborn and young mice. Acta Pharmacol Toxicol (Copenh) 27(5):363-380.

Stead LF, Perera R, Bullen C, Mant D and Lancaster T (2008) Nicotine replacement therapy for smoking cessation. Cochrane Database Syst Rev(1):CD000146.

Steffens M, Zentner J, Honegger J and Feuerstein TJ (2005) Binding affinity and agonist activity of putative endogenous cannabinoids at the human neocortical CB1 receptor. Biochem Pharmacol 69(1):169-178.

Steinberg KK, Relling MV, Gallagher ML, Greene CN, Rubin CS, French D, Holmes AK, Carroll WL, Koontz DA, Sampson EJ and Satten GA (2007) Genetic studies of a cluster of acute lymphoblastic leukemia cases in Churchill County, Nevada. Environ Health Perspect 115(1):158-164.

Stolerman IP and Jarvis MJ (1995) The scientific case that nicotine is addictive. Psychopharmacology (Berl) 117(1):2-10; discussion 14-20.

Stolerman IP, Naylor C, Elmer GI and Goldberg SR (1999) Discrimination and self-administration of nicotine by inbred strains of mice. Psychopharmacology (Berl) 141(3):297-306.

Strasser AA, Malaiyandi V, Hoffmann E, Tyndale RF and Lerman C (2007) An association of CYP2A6 genotype and smoking topography. Nicotine Tob Res 9(4):511-518.

Stresser DM and Kupfer D (1999) Monospecific antipeptide antibody to cytochrome P-450 2B6. Drug Metab Dispos 27(4):517-525.

Su T, Bao Z, Zhang QY, Smith TJ, Hong JY and Ding X (2000) Human cytochrome P450 CYP2A13: predominant expression in the respiratory tract and its high

Page 228: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

213

efficiency metabolic activation of a tobacco-specific carcinogen, 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone. Cancer Res 60(18):5074-5079.

Su T and Ding X (2004) Regulation of the cytochrome P450 2A genes. Toxicol Appl Pharmacol 199(3):285-294.

Su T, He W, Gu J, Lipinskas TW and Ding X (1998) Differential xenobiotic induction of CYP2A5 in mouse liver, kidney, lung, and olfactory mucosa. Drug Metab Dispos 26(8):822-824.

Sugihara K, Kitamura S, Tatsumi K, Asahara T and Dohi K (1997) Differences in aldehyde oxidase activity in cytosolic preparations of human and monkey liver. Biochem Mol Biol Int 41(6):1153-1160.

Sullivan PF, Jiang Y, Neale MC, Kendler KS and Straub RE (2001) Association of the tryptophan hydroxylase gene with smoking initiation but not progression to nicotine dependence. Am J Med Genet 105(5):479-484.

Swan GE, Hops H, Wilhelmsen KC, Lessov-Schlaggar CN, Cheng LS, Hudmon KS, Amos CI, Feiler HS, Ring HZ, Andrews JA, Tildesley E and Benowitz N (2006) A genome-wide screen for nicotine dependence susceptibility loci. Am J Med Genet B Neuropsychiatr Genet 141B(4):354-360.

Takeuchi H, Saoo K, Yokohira M, Ikeda M, Maeta H, Miyazaki M, Yamazaki H, Kamataki T and Imaida K (2003) Pretreatment with 8-methoxypsoralen, a potent human CYP2A6 inhibitor, strongly inhibits lung tumorigenesis induced by 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone in female A/J mice. Cancer Res 63(22):7581-7583.

Tallarida RJ and Murray RB (1987) Manual of Pharmacological Calculations with Computer Programs. Springer-Verlag, New York.

Tapper AR, McKinney SL, Nashmi R, Schwarz J, Deshpande P, Labarca C, Whiteaker P, Marks MJ, Collins AC and Lester HA (2004) Nicotine activation of alpha4* receptors: sufficient for reward, tolerance, and sensitization. Science 306(5698):1029-1032.

Tashkin D, Kanner R, Bailey W, Buist S, Anderson P, Nides M, Gonzales D, Dozier G, Patel MK and Jamerson B (2001) Smoking cessation in patients with chronic obstructive pulmonary disease: a double-blind, placebo-controlled, randomised trial. Lancet 357(9268):1571-1575.

Tateno C, Yoshizane Y, Saito N, Kataoka M, Utoh R, Yamasaki C, Tachibana A, Soeno Y, Asahina K, Hino H, Asahara T, Yokoi T, Furukawa T and Yoshizato K (2004) Near completely humanized liver in mice shows human-type metabolic responses to drugs. Am J Pathol 165(3):901-912.

Tiano HF, Hosokawa M, Chulada PC, Smith PB, Wang RL, Gonzalez FJ, Crespi CL and Langenbach R (1993) Retroviral mediated expression of human cytochrome P450 2A6 in C3H/10T1/2 cells confers transformability by 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone (NNK). Carcinogenesis 14(7):1421-1427.

Todd M (2004) Daily processes in stress and smoking: effects of negative events, nicotine dependence, and gender. Psychol Addict Behav 18(1):31-39.

Tonnesen P, Paoletti P, Gustavsson G, Russell MA, Saracci R, Gulsvik A, Rijcken B and Sawe U (1999) Higher dosage nicotine patches increase one-year smoking cessation rates: results from the European CEASE trial. Collaborative European Anti-Smoking Evaluation. European Respiratory Society. Eur Respir J 13(2):238-246.

Tonstad S, Tonnesen P, Hajek P, Williams KE, Billing CB and Reeves KR (2006) Effect of maintenance therapy with varenicline on smoking cessation: a randomized controlled trial. Jama 296(1):64-71.

Page 229: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

214

Torchin CD, McNeilly PJ, Kapetanovic IM, Strong JM and Kupferberg HJ (1996) Stereoselective metabolism of a new anticonvulsant drug candidate, losigamone, by human liver microsomes. Drug Metab Dispos 24(9):1002-1008.

Tripathi HL, Martin BR and Aceto MD (1982) Nicotine-induced antinociception in rats and mice: correlation with nicotine brain levels. J Pharmacol Exp Ther 221(1):91-96.

Turgeon D, Carrier JS, Levesque E, Hum DW and Belanger A (2001) Relative enzymatic activity, protein stability, and tissue distribution of human steroid-metabolizing UGT2B subfamily members. Endocrinology 142(2):778-787.

Tyndale RF (2003) Genetics of alcohol and tobacco use in humans. Ann Med 35(2):94-121.

Tyndale RF, Li Y, Kaplan HL and Seller EM (1999) Inhibition of nicotine's metabolism: a potential new treatment for tobacco dependence [abstract]. Clin Pharmacol Ther 65(2):145.

Tyndale RF and Sellers EM (2002) Genetic variation in CYP2A6-mediated nicotine metabolism alters smoking behavior. Ther Drug Monit 24(1):163-171.

Ulvila J, Arpiainen S, Pelkonen O, Aida K, Sueyoshi T, Negishi M and Hakkola J (2004) Regulation of Cyp2a5 transcription in mouse primary hepatocytes: roles of hepatocyte nuclear factor 4 and nuclear factor I. Biochem J 381(Pt 3):887-894.

Uno T, Mitchell E, Aida K, Lambert MH, Darden TA, Pedersen LG and Negishi M (1997) Reciprocal size-effect relationship of the key residues in determining regio- and stereospecificities of DHEA hydroxylase activity in P450 2a5. Biochemistry 36(11):3193-3198.

Urakawa N, Nagata T, Kudo K, Kimura K and Imamura T (1994) Simultaneous determination of nicotine and cotinine in various human tissues using capillary gas chromatography/mass spectrometry. Int J Legal Med 106(5):232-236.

Vainio PJ and Tuominen RK (2001) Cotinine binding to nicotinic acetylcholine receptors in bovine chromaffin cell and rat brain membranes. Nicotine Tob Res 3(2):177-182.

Valoti M, Fusi F, Frosini M, Pessina F, Tipton KF and Sgaragli GP (2000) Cytochrome P450-dependent N-dealkylation of L-deprenyl in C57BL mouse liver microsomes: effects of in vivo pretreatment with ethanol, phenobarbital, beta-naphthoflavone and L-deprenyl. Eur J Pharmacol 391(3):199-206.

van Iersel M, Walters DG, Price RJ, Lovell DP and Lake BG (1994) Sex and strain differences in mouse hepatic microsomal coumarin 7-hydroxylase activity. Food Chem Toxicol 32(4):387-390.

van Noord C, Straus SM, Sturkenboom MC, Hofman A, Aarnoudse AJ, Bagnardi V, Kors JA, Newton-Cheh C, Witteman JC and Stricker BH (2009) Psychotropic drugs associated with corrected QT interval prolongation. J Clin Psychopharmacol 29(1):9-15.

Vila R, Kurosaki M, Barzago MM, Kolek M, Bastone A, Colombo L, Salmona M, Terao M and Garattini E (2004) Regulation and biochemistry of mouse molybdo-flavoenzymes. The DBA/2 mouse is selectively deficient in the expression of aldehyde oxidase homologues 1 and 2 and represents a unique source for the purification and characterization of aldehyde oxidase. J Biol Chem 279(10):8668-8683.

Villegier AS, Salomon L, Blanc G, Godeheu G, Glowinski J and Tassin JP (2006) Irreversible blockade of monoamine oxidases reveals the critical role of 5-HT transmission in locomotor response induced by nicotine in mice. Eur J Neurosci 24(5):1359-1365.

Page 230: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

215

Vink JM, Willemsen G and Boomsma DI (2005) Heritability of smoking initiation and nicotine dependence. Behav Genet 35(4):397-406.

Visoni S, Meireles N, Monteiro L, Rossini A and Pinto LF (2008) Different modes of inhibition of mouse Cyp2a5 and rat CYP2A3 by the food-derived 8-methoxypsoralen. Food Chem Toxicol 46(3):1190-1195.

von Weymarn LB, Zhang QY, Ding X and Hollenberg PF (2005) Effects of 8-methoxypsoralen on cytochrome P450 2A13. Carcinogenesis 26(3):621-629.

Waddell WJ and Marlowe C (1976) Localization of nicotine-14C, cotinine-14C, and nicotine-1'-N-oxide-14C in tissues of the mouse. Drug Metab Dispos 4(6):530-539.

Wagena EJ, Knipschild PG, Huibers MJ, Wouters EF and van Schayck CP (2005) Efficacy of bupropion and nortriptyline for smoking cessation among people at risk for or with chronic obstructive pulmonary disease. Arch Intern Med 165(19):2286-2292.

Wang J, Pitarque M and Ingelman-Sundberg M (2006a) 3'-UTR polymorphism in the human CYP2A6 gene affects mRNA stability and enzyme expression. Biochem Biophys Res Commun 340(2):491-497.

Wang SL, He XY, Shen J, Wang JS and Hong JY (2006b) The missense genetic polymorphisms of human CYP2A13: functional significance in carcinogen activation and identification of a null allelic variant. Toxicol Sci 94(1):38-45.

Ward BA, Gorski JC, Jones DR, Hall SD, Flockhart DA and Desta Z (2003) The cytochrome P450 2B6 (CYP2B6) is the main catalyst of efavirenz primary and secondary metabolism: implication for HIV/AIDS therapy and utility of efavirenz as a substrate marker of CYP2B6 catalytic activity. J Pharmacol Exp Ther 306(1):287-300.

West R, Baker CL, Cappelleri JC and Bushmakin AG (2008) Effect of varenicline and bupropion SR on craving, nicotine withdrawal symptoms, and rewarding effects of smoking during a quit attempt. Psychopharmacology (Berl) 197(3):371-377.

West R, Hajek P, Foulds J, Nilsson F, May S and Meadows A (2000) A comparison of the abuse liability and dependence potential of nicotine patch, gum, spray and inhaler. Psychopharmacology (Berl) 149(3):198-202.

Wetter DW, Kenford SL, Smith SS, Fiore MC, Jorenby DE and Baker TB (1999) Gender differences in smoking cessation. J Consult Clin Psychol 67(4):555-562.

Wildenauer DB, Aschoff S, Hallmayer J, Schroff K-C, Schwab SG and Richter E (2005) Mapping Quantitative Trait Loci (QTL) for Nicotine Preference in Inbred Mice, in Society for Research on Nicotine and Tobacco, Prague, Czech Republic.

Wong GY, Wolter TD, Croghan GA, Croghan IT, Offord KP and Hurt RD (1999) A randomized trial of naltrexone for smoking cessation. Addiction 94(8):1227-1237.

Wood AW and Taylor BA (1979) Genetic regulation of coumarin hydroxylase activity in mice. Evidence for single locus control on chromosome. J Biol Chem 254(13):5647-5651.

Xian H, Scherrer JF, Madden PA, Lyons MJ, Tsuang M, True WR and Eisen SA (2003) The heritability of failed smoking cessation and nicotine withdrawal in twins who smoked and attempted to quit. Nicotine Tob Res 5(2):245-254.

Xu C, Rao YS, Xu B, Hoffmann E, Jones J, Sellers EM and Tyndale RF (2002) An in vivo pilot study characterizing the new CYP2A6*7, *8, and *10 alleles. Biochem Biophys Res Commun 290(1):318-324.

Yamanaka H, Nakajima M, Fukami T, Sakai H, Nakamura A, Katoh M, Takamiya M, Aoki Y and Yokoi T (2005a) CYP2A6 AND CYP2B6 are involved in nornicotine formation from nicotine in humans: interindividual differences in these contributions. Drug Metab Dispos 33(12):1811-1818.

Page 231: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

216

Yamanaka H, Nakajima M, Katoh M, Kanoh A, Tamura O, Ishibashi H and Yokoi T (2005b) Trans-3'-hydroxycotinine O- and N-glucuronidations in human liver microsomes. Drug Metab Dispos 33(1):23-30.

Yamanaka H, Nakajima M, Nishimura K, Yoshida R, Fukami T, Katoh M and Yokoi T (2004) Metabolic profile of nicotine in subjects whose CYP2A6 gene is deleted. Eur J Pharm Sci 22(5):419-425.

Yamano S, Tatsuno J and Gonzalez FJ (1990) The CYP2A3 gene product catalyzes coumarin 7-hydroxylation in human liver microsomes. Biochemistry 29(5):1322-1329.

Yamazaki H, Inoue K, Hashimoto M and Shimada T (1999) Roles of CYP2A6 and CYP2B6 in nicotine C-oxidation by human liver microsomes. Arch Toxicol 73(2):65-70.

Yang L (2007) Inhibition of human liver cytochrome p450 by star fruit juice. J Pharm Pharm Sci 10(4):496-503.

Yano JK, Denton TT, Cerny MA, Zhang X, Johnson EF and Cashman JR (2006) Synthetic inhibitors of cytochrome P-450 2A6: inhibitory activity, difference spectra, mechanism of inhibition, and protein cocrystallization. J Med Chem 49(24):6987-7001.

Yano JK, Hsu MH, Griffin KJ, Stout CD and Johnson EF (2005) Structures of human microsomal cytochrome P450 2A6 complexed with coumarin and methoxsalen. Nat Struct Mol Biol 12(9):822-823.

Yoshida R, Nakajima M, Nishimura K, Tokudome S, Kwon JT and Yokoi T (2003) Effects of polymorphism in promoter region of human CYP2A6 gene (CYP2A6*9) on expression level of messenger ribonucleic acid and enzymatic activity in vivo and in vitro. Clin Pharmacol Ther 74(1):69-76.

Youdim MB (1978) The active centers of monoamine oxidase types "A" and "B": binding with (14C)-clorgyline and (14C)-deprenyl. J Neural Transm 43(3-4):199-208.

Youdim MB, Edmondson D and Tipton KF (2006) The therapeutic potential of monoamine oxidase inhibitors. Nat Rev Neurosci 7(4):295-309.

Young JW (1993) Treatment of cutaneous T-cell lymphoma: an update. J Am Osteopath Assoc 93(5):591-598, 603.

Zarrindast MR, Faraji N, Rostami P, Sahraei H and Ghoshouni H (2003) Cross-tolerance between morphine- and nicotine-induced conditioned place preference in mice. Pharmacol Biochem Behav 74(2):363-369.

Zeman MV, Hiraki L and Sellers EM (2002) Gender differences in tobacco smoking: higher relative exposure to smoke than nicotine in women. J Womens Health Gend Based Med 11(2):147-153.

Zhang J and Cashman JR (2006) Quantitative analysis of FMO gene mRNA levels in human tissues. Drug Metab Dispos 34(1):19-26.

Zhang W, Kilicarslan T, Tyndale RF and Sellers EM (2001) Evaluation of methoxsalen, tranylcypromine, and tryptamine as specific and selective CYP2A6 inhibitors in vitro. Drug Metab Dispos 29(6):897-902.

Zhang X, D'Agostino J, Wu H, Zhang QY, von Weymarn L, Murphy SE and Ding X (2007) CYP2A13: variable expression and role in human lung microsomal metabolic activation of the tobacco-specific carcinogen 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone. J Pharmacol Exp Ther 323(2):570-578.

Zhang X, Su T, Zhang QY, Gu J, Caggana M, Li H and Ding X (2002) Genetic polymorphisms of the human CYP2A13 gene: identification of single-nucleotide polymorphisms and functional characterization of an Arg257Cys variant. J Pharmacol Exp Ther 302(2):416-423.

Page 232: AN INVESTIGATION OF NICOTINE METABOLISM IN MICE: THE ... · Section 2.2.7.1 Selegiline (l-deprenyl; Eldepryl®) Page 32 Section 2.2.7.1.1 Selegiline Pharmacodynamics Page 32 Section

217

Zhu LR, Thomas PE, Lu G, Reuhl KR, Yang GY, Wang LD, Wang SL, Yang CS, He XY and Hong JY (2006) CYP2A13 in human respiratory tissues and lung cancers: an immunohistochemical study with a new peptide-specific antibody. Drug Metab Dispos 34(10):1672-1676.

Zins BJ, Sandborn WJ, Mays DC, Lawson GM, McKinney JA, Tremaine WJ, Mahoney DW, Zinsmeister AR, Hurt RD, Offord KP and Lipsky JJ (1997) Pharmacokinetics of nicotine tartrate after single-dose liquid enema, oral, and intravenous administration. J Clin Pharmacol 37(5):426-436.