inhibition of brain cyp2d lowers codeine-induced analgesia ... · pdf fileinhibition of brain...

157
Inhibition of brain CYP2D lowers codeine-induced analgesia in rats by Kaidi Zhou A thesis submitted in conformity with the requirements for the degree of Master of Science Graduate Department of Pharmacology and Toxicology University of Toronto © Copyright by Kaidi Zhou 2012

Upload: nguyenphuc

Post on 16-Mar-2018

221 views

Category:

Documents


3 download

TRANSCRIPT

Inhibition of brain CYP2D lowers codeine-induced

analgesia in rats

by

Kaidi Zhou

A thesis submitted in conformity with the requirements

for the degree of Master of Science

Graduate Department of Pharmacology and Toxicology

University of Toronto

© Copyright by Kaidi Zhou 2012

ii

Inhibition of brain CYP2D lowers codeine-induced analgesia in

rats

Kaidi Zhou

Master of Science

Graduate Department of Pharmacology and Toxicology

University of Toronto

2012

Abstract

CYP2D6 metabolizes codeine to morphine, the active analgesic metabolite.

Variation in brain CYP2D6 activity may affect brain morphine levels after codeine

administration and thereby influence analgesia. We investigate the effect of

inhibiting brain CYP2D on codeine-induced analgesia. METHODS: Rats received

intracerebroventricular (i.c.v.) injections of CYP2D inhibitors or vehicle controls.

Rats were then given subcutaneous codeine injections and analgesia was measured

with the tail-flick test. Morphine and codeine concentrations in brain and plasma

were measured. CYP2D activity in brain and liver were assessed in vitro. RESULTS:

Compared to vehicle treatment, i.c.v. inhibitor treatments resulted in lower codeine-

induced analgesia, lower morphine levels in brain but not in plasma after codeine

injections, and lower CYP2D activity in brain membranes but not in liver

microsomes. CONCLUSIONS: Inhibiting brain CYP2D reduces codeine’s metabolism

to morphine, resulting in less analgesia. Variation in brain CYP2D6 activity may

influence response to codeine and other CYP2D6 substrates.

iii

Acknowledgements

First and foremost, I would like to thank my supervisor Dr. Rachel F. Tyndale for all

of her support and guidance throughout the course of my graduate study. I am

grateful to her for providing me with this amazing opportunity to learn and grow

intellectually and professionally. Her ambition, hard work and professionalism are

inspiring and have encouraged me to be a more effective student and scientist.

I would like to thank Dr. Jose N. Nobrega for being my M.Sc. advisor, as well

as Dr. Ali Salahpour, Dr. Daniel J. Mueller and Dr. John W. Semple for serving on my

defense committee. Their sharing of time, knowledge and constructive criticism are

greatly appreciated.

I am thankful to each and every member of Dr. Tyndale’s lab for generously

providing their advice, time, knowledge, and technical expertise during my M.Sc. I

am truly fortunate to have been part of a group with such intelligent, caring,

encouraging and supportive people. Special thanks to Dr. Sharon Miksys for her

advice and teaching expertise and for fostering a positive and encouraging lab

environment, to Jibran Khokhar for his enthusiasm, knowledge and help on multiple

aspects of this project, and to Dr. Bin Zhao and Steven Lo for their excellent technical

assistance and time contributions.

Last but not least, I would like to thank my family and friends for their ongoing

love, care and support. I am especially indebted to my parents for their hard work,

patience and dedication, which made all of my opportunities possible.

iv

Table of Contents

Abstract ...................................................................................................................... ii

Acknowledgements ................................................................................................... iii

Table of Contents ....................................................................................................... iv

List of Figures............................................................................................................ vii

Summary of Abbreviations ........................................................................................ ix

Section 1: Introduction ................................................................................................ 1

Statement of Research Problem .................................................................................. 1

Purpose of the Study and Objective ............................................................................ 2

Statement of Research Hypotheses and Rationale ....................................................... 3

Review of the Literature .............................................................................................. 5

1.1 Cytochrome P450 2D6 (CYP2D6) .......................................................................... 5

1.1.1 Cytochromes P450 ............................................................................................. 5

1.1.2 CYP2D6 substrates ............................................................................................. 6

1.1.3 CYP2D6 inhibitors .............................................................................................. 6

1.1.4 CYP2D6 regulation ............................................................................................. 7

1.1.5 CYP2D6 genetic variation ................................................................................... 8

1.1.5a Interethnic variability in CYP2D6 ...................................................................... 9

1.1.6 CYP2D expression in different species ............................................................ 10

1.2 Brain Cytochromes P450 ..................................................................................... 11

1.2.1 Brain CYP expression ....................................................................................... 12

1.2.2 Brain CYP activity ............................................................................................. 14

1.2.3 Brain CYP regulation ........................................................................................ 16

1.2.4 Brain CYP2D6 ................................................................................................... 17

1.2.4a Brain CYP2D expression ................................................................................. 18

1.2.4b Brain CYP2D function and activity .................................................................. 19

1.2.4c Brain CYP2D regulation .................................................................................. 24

1.3 Opioid Analgesics ............................................................................................... 26

1.3.1 Codeine ........................................................................................................... 26

1.3.1a Codeine metabolism in humans ..................................................................... 27

1.3.1b Codeine metabolism in rats ........................................................................... 31

1.3.2 Morphine .......................................................................................................... 33

1.3.2a Spinal mechanisms of morphine-induced analgesia ...................................... 33

1.3.2b Supraspinal mechanisms of morphine-induced analgesia ............................. 34

v

1.4 Rat Tail-Flick Test: Animal Model of Nociception ................................................ 35

1.4.1 Tail-flick reflex ................................................................................................. 35

1.4.2 Effect of opioid analgesics in the tail-flick test .................................................. 37

1.5 Study design ....................................................................................................... 38

Section 2: Materials and Methods ............................................................................. 39

Section 3: Results ...................................................................................................... 50

3.1 Inhibition of brain CYP2D reduced codeine-induced analgesia ......................... 50

3.2 Inhibition of brain CYP2D lowered codeine-induced area under the analgesia

time curve ................................................................................................................. 53

3.3 Inhibiting brain CYP2D did not affect baseline tail-flick latency ......................... 58

3.4 Inhibiting brain CYP2D did not affect morphine-induced analgesia ................... 60

3.5 Inhibiting brain CYP2D did not alter morphine-induced area under the analgesia

time curve ................................................................................................................. 62

3.6 There was no tolerance to the analgesic effects of codeine or morphine ............ 65

3.7 Codeine and morphine doses used resulted in similar levels of analgesia ......... 70

3.8 Inhibitor-treated rats had lower morphine levels in the brain but not plasma at 30

min after codeine injection ....................................................................................... 73

3.9 Analgesia correlated with brain, and not plasma, morphine levels .................... 81

3.10 Inhibitor-treated rats did not have lower morphine levels in the brain at 60 or 90

min after codeine injection ....................................................................................... 86

3.11 Inhibiting brain CYP2D in vivo lowered in vitro codeine metabolism in the brain

but not liver .............................................................................................................. 91

3.12 Inhibiting brain CYP2D in vivo lowered in vitro dextromethorphan metabolism

in the brain but not liver ........................................................................................... 93

Section 4: Discussion, Conclusions, Future Directions .............................................. 95

4.1 Summary and further implications ...................................................................... 95

4.1.1 Rat model of reduced brain CYP2D activity ..................................................... 96

4.1.2 Inhibition of brain CYP2D lowers codeine-induced analgesia ......................... 98

4.1.3 Analgesia correlates with morphine levels in the brain and not plasma ........ 100

4.1.4 Inhibiting brain CYP2D in vivo lowers in vitro enzyme activity in brain

membranes and not liver microsomes .................................................................... 102

4.1.5 Limitations ...................................................................................................... 102

4.2. Clinical relevance of brain CYP2D activity ...................................................... 105

4.2.1 Centrally-acting drugs ................................................................................... 106

4.2.2 Drugs of abuse ............................................................................................... 108

4.2.3 Endogenous substrates .................................................................................. 109

vi

4.2.4 Disease ........................................................................................................... 111

4.3 Other brain CYPs .............................................................................................. 114

4.4 Future directions ............................................................................................... 116

4.4.1 Other uses of rat models of differing levels of brain CYP2D activity .............. 116

4.4.1a Microdialysis ................................................................................................ 116

4.4.1b Different pain model .................................................................................... 116

4.4.1c Role of rat brain CYP2D in meditating drug inactivation .............................. 117

4.4.1d Effect of rat brain CYP2D induction on drug response ................................. 117

4.4.1e Role of rat brain CYP2D in neurotoxin inactivation ....................................... 118

4.4.2 Therapeutic uses of brain CYP2D induction ................................................... 118

4.5 Conclusions ....................................................................................................... 119

References .............................................................................................................. 120

List of Abstracts ....................................................................................................... 148

vii

List of Figures

Figure 1. Morphine levels in rat plasma and brain 30 min after peripheral injection of

codeine or morphine ................................................................................................ 22

Figure 2. Metabolic pathways of codeine in humans ................................................ 29

Figure 3. Inhibition of brain CYP2D reduced codeine-induced analgesia ................ 52

Figure 4. Inhibiting brain CYP2D with propranolol lowered codeine-induced area

under the analgesia time curve between 0-60 min after codeine injection ............... 54

Figure 5. Inhibiting brain CYP2D with propranolol did not lower codeine-induced

area under the analgesia time curve at 60-120 min or 0-120 min after codeine

injection .................................................................................................................... 55

Figure 6. Inhibiting brain CYP2D with propafenone lowered codeine-induced area

under the analgesia time curve between 0-60 min after codeine injection ............... 56

Figure 7. Inhibiting brain CYP2D with propafenone did not lower codeine-induced

area under the analgesia time curve at 60-120 min or 0-120 min after codeine

injection .................................................................................................................... 57

Figure 8. Inhibiting brain CYP2D did not affect baseline tail-flick latency ................ 59

Figure 9. Inhibiting brain CYP2D did not affect morphine-induced analgesia .......... 61

Figure 10. Inhibiting brain CYP2D with propranolol did not alter morphine-induced

area under the analgesia time curve ......................................................................... 63

Figure 11. Inhibiting brain CYP2D with propafenone did not alter morphine-induced

area under the analgesia time curve ......................................................................... 64

Figure 12. Rats treated with propranolol or vehicle did not develop tolerance to

codeine ..................................................................................................................... 66

Figure 13. Rats treated with propafenone or vehicle did not develop tolerance to

codeine ..................................................................................................................... 67

Figure 14. Rats treated with propranolol or vehicle did not develop tolerance to

morphine .................................................................................................................. 68

Figure 15. Rats treated with propafenone or vehicle did not develop tolerance to

morphine .................................................................................................................. 69

Figure 16. Codeine and morphine doses used resulted in similar levels of analgesia

after ACSF (i.c.v. vehicle) treatment ......................................................................... 71

Figure 17. Codeine and morphine doses used resulted in similar levels of analgesia

after cyclodextrin (i.c.v. vehicle) treatment .............................................................. 72

Figure 18. Inhibitor-treated rats had lower morphine levels in the brain but not in

plasma at 30 min after codeine injection ................................................................... 75

Figure 19. Inhibitor-treated rats had lower morphine to codeine ratios in the brain

but not in plasma at 30 min after codeine injection ................................................... 76

viii

Figure 20. Inhibitor-treated rats had lower morphine to total drug ratios in the brain

but not in plasma at 30 min after codeine injection ................................................... 77

Figure 21. Propranolol-treated rats did not have lower codeine levels or total drug

levels in the brain or in plasma at 30 min after codeine injection ............................. 78

Figure 22. Propafenone-treated rats did not have lower codeine levels or total drug

levels in the brain or in plasma at 30 min after codeine injection ............................. 79

Figure 23. Inhibitor-treated rats had similar morphine levels and morphine to

codeine ratios between the anterior and the posterior parts of the brain. ................ 80

Figure 24. Analgesia correlated with brain, and not plasma, morphine levels ......... 82

Figure 25. Analgesia correlated with brain, and not plasma, morphine to codeine

ratios ......................................................................................................................... 83

Figure 26. Analgesia correlated with brain, and not plasma, morphine to total drug

ratios ......................................................................................................................... 84

Figure 27. Analgesia did not correlate with codeine levels or total drug levels in

brain or plasma ......................................................................................................... 85

Figure 28. Inhibitor-treated rats did not have lower morphine levels in the brain at

60 or 90 min after codeine injection .......................................................................... 87

Figure 29. Inhibitor-treated rats did not have lower morphine to codeine ratios in the

brain or plasma at 60 or 90 min after codeine injection ............................................ 88

Figure 30. Inhibitor-treated rats did not have lower morphine to total drug ratios in

the brain or plasma at 60 or 90 min after codeine injection ...................................... 89

Figure 31. Inhibitor-treated rats did not have lower codeine levels or total drug

levels in the brain or plasma at 60 or 90 min after codeine injection ........................ 90

Figure 32. Inhibiting brain CYP2D in vivo lowered in vitro codeine metabolism to

morphine in brain membranes but not in liver microsomes ..................................... 92

Figure 33. Inhibiting brain CYP2D in vivo lowered in vitro dextromethorphan

metabolism to dextrorphan in brain membranes but not in liver microsomes ......... 94

ix

Summary of Abbreviations

ACSF Artificial cerebrospinal fluid

BBB Blood-brain barrier

CNS Central nervous system

COD Codeine

CYP Cytochrome P450

CYP2D Cytochrome P450 2D

CYP2D1 Cytochrome P450 2D1

CYP2D6 Cytochrome P450 2D6

EM Extensive metabolizer

h Hour

HPLC High-performance liquid chromatography

i.c.v. Intracerebroventricular

i.p. Intraperitoneal

MBI Mechanism-based inhibitor

min Minute

MOR Morphine

PM Poor metabolizer

s.c. Subcutaneous

sec Second

UM Ultra-rapid metabolizer

1

Section 1: Introduction

Statement of Research Problem

Cytochrome P450 2D6 (CYP2D6) is an oxidative enzyme that metabolizes many

centrally-acting drugs, including clinically prescribed drugs (e.g. risperidone,

fluoxetine, codeine) as well as drugs of abuse (e.g. amphetamine, MDMA) (Zanger,

Raimundo et al. 2004). CYP2D6 is primarily expressed in the liver but is also

expressed in the brain. Brain CYPs are active in vivo (Miksys and Tyndale 2009), and

in some cell types (e.g. frontal cortex pyramidal neurons) brain CYPs are expressed

at levels as high as those in the liver (Miksys, Hoffmann et al. 2000). Therefore,

CYP2D6 may metabolize centrally-acting drugs locally in the brain and have a

significant impact on drug effect. Examining the role of brain CYPs in drug

metabolism and response will help elucidate the function and importance of brain

CYP activity.

There is large interindividual variation in the response to centrally-acting

drugs, which does not always correlate with plasma drug levels (Michels and

Marzuk 1993a). This may be caused by variation in the degree of metabolism by

brain CYPs, which may affect local drug and metabolite levels in the brain, and in

turn influence drug response.

Brain CYP2D6 levels can vary independently from hepatic CYP2D6 levels.

Brain CYP2D6, unlike hepatic CYP2D6, is induced in animals by nicotine and ethanol

2

(Warner and Gustafsson 1994, Mann, Miksys et al. 2008, Yue, Miksys et al. 2008), and

its levels are higher in smokers and alcoholics (Miksys, Rao et al. 2002, Miksys and

Tyndale 2004, Mann, Miksys et al. 2008). Brain CYP2D6 levels also increase with age

while hepatic CYP2D6 remain the same or even decrease with age (Parkinson,

Mudra et al. 2004, Mann, Miksys et al. 2012). Therefore, environmental factors and

age may result in variation in brain CYP2D6 expression and activity, and thereby

alter the metabolism of, and response to, centrally-acting drugs.

While studies have suggested that brain CYP2D can metabolize drugs in vitro

and in vivo, no studies have examined whether brain CYP2D-mediated drug

metabolism can affect drug response. We thus seek to elucidate the effects of

altering brain CYP2D activity (without affecting liver CYP2D) on drug response.

Clarifying the role of brain CYP-mediated metabolism in drug response may help

explain, at least in part, the interindividual variation in the response to centrally-

acting drugs and the poor correlation between the plasma levels and effects of these

drugs.

Purpose of the Study and Objective

Brain CYPs have been shown to be active in vitro (Albores, Ortega-Mantilla et al.

2001) and in vivo (Miksys and Tyndale 2009), yet it is not clear what their precise

function is in drug metabolism and effect. Thus, the purpose of this study is to

elucidate the impact of brain CYPs on drug biotransformation and response.

CYP2D’s many centrally-acting substrates and in vitro activity in rat brain make it a

3

suitable enzyme for examining the role of brain CYP-mediated metabolism in drug

action. Our objective is to examine the influence of brain CYP2D on the metabolic

activation of a centrally-acting drug (codeine) and on the drug’s subsequent effect

(codeine-induced analgesia).

The role of brain CYP2D could be investigated by manipulating its activity

while leaving hepatic CYP2D activity unchanged. Rat brain CYP2D could be

selectively inhibited in vivo by intracerebroventricular (i.c.v.) injection of CYP2D

inhibitors, without affecting hepatic CYP2D. This provides an animal model of

reduced brain CYP2D activity that could be used to assess the contribution of brain,

as opposed to liver, codeine metabolism to codeine-induced analgesia. Clarifying

the impact of brain CYP2D-mediated metabolism on a drug effect will help us

understand the possible role brain CYPs have in interindividual variation in drug

response.

Statement of Research Hypotheses and Rationale

CYP2D metabolizes the opioid analgesic codeine (prodrug) to morphine (active

metabolite) (Adler, Fujimoto et al. 1955). Since morphine has a 3000-fold greater

affinity for the mu-opioid receptor than does codeine (Pert and Snyder 1973), and

CYP2D6 poor metabolizers produce no morphine from codeine and experience no

analgesia (Sindrup, Brosen et al. 1990, Chen, Somogyi et al. 1991), analgesia from

codeine is dependent on its metabolism to morphine. Codeine is metabolized to

morphine mainly by hepatic CYP2D; morphine then crosses into the brain where it

4

can interact with mu-opioid receptors to elicit analgesia. However, because

morphine is less permeable across the blood-brain barrier than is codeine, there is

a delay in morphine’s entry into the brain compared to codeine’s entry (Oldendorf,

Hyman et al. 1972). Thus, the initial morphine present in the brain after codeine

administration may be solely due to local metabolism of codeine in the brain. This is

supported by rat brain CYP2D’s ability to metabolize codeine in vitro (Chen, Irvine

et al. 1990), and the finding that morphine can be detected in rat brain at 30 min after

intraperitoneal codeine, but not morphine, injections. This implies that at 30 min

after codeine injection, morphine formed from hepatic metabolism has not yet

crossed into the brain, and that morphine found in the brain at this time is due to

brain CYP2D-mediated codeine metabolism.

Hypothesis: We hypothesize that inhibiting rat brain CYP2D will reduce

analgesia during the initial period after codeine administration by decreasing the

metabolism of codeine to morphine in the brain. More specifically, we hypothesize

that i.c.v. injection of CYP2D inhibitors will result in 1) lower brain morphine

concentrations and 2) shorter tail-flick latencies (a measure of analgesia) during the

first 30 min after peripheral codeine injection, compared to i.c.v. injection of vehicle.

5

Review of the Literature

1.1 Cytochrome P450 2D6 (CYP2D6)

1.1.1 Cytochromes P450

Cytochromes P450 (CYPs) are a superfamily of heme-containing enzymes that

oxidize a wide range of substrates (Estabrook 1999, Coon 2005), including drugs,

toxins, and endogenous substances (Rendic and Di Carlo 1997). Most drug-

metabolizing CYPs of the CYP2 family are found mainly in the liver, but also in other

organs such as the brain, intestines and lungs. The liver is thought to be responsible

for systemic drug metabolism, while the other organs may take part in localized, in

situ substrate metabolism (Ding and Kaminsky 2003). The expression and activity

level of a CYP may differ between individuals because of genetic variation and/or

exposure to environmental inducers or inhibitors (Lee, Miksys et al. 2006b, Ai, Li et

al. 2009).

Cytochrome P450 2D6 (CYP2D6), one of the 57 members of the CYP

superfamily in humans (Nelson 2006), makes up only 5% of total CYP content in the

liver (Guengerich 2003, Emoto, Murase et al. 2006) yet is involved in the metabolism

of ~30% of clinically used drugs (Zanger, Raimundo et al. 2004). Its extensive role in

drug metabolism makes it important to understand the function of this enzyme.

6

1.1.2 CYP2D6 substrates

CYP2D6 is capable of metabolically activating or inactivating a wide variety of both

exogenous and endogenous substances (Zanger, Raimundo et al. 2004). These

include clinically used drugs (analgesics, antidepressants, antipsychotics, -

adrenergic blockers, antiarrhythmics), recreational drugs (amphetamine, MDMA) as

well as neurotoxins, neurosteroids, and biogenic amines.

When a CYP is the main contributor (>80%) to the metabolic reaction of a

substrate, that substrate can be used as a probe drug (Frank, Jaehde et al. 2007).

That is, the ratio of parent compound to metabolite resulting from the enzymatic

reaction may be used as an indicator of that CYP’s activity. One such probe drug is

dextromethorphan, which is oxidized predominantly by CYP2D6 to dextrorphan

(Frank, Jaehde et al. 2007, Zhou 2009).

1.1.3 CYP2D6 inhibitors

The activity level of CYPs can be reduced by inhibitors. This can affect drug efficacy

or cause adverse drug reactions. For example, individuals pretreated with the

CYP2D6 inhibitor quinidine produce little morphine from codeine and experience

reduced analgesia (Sindrup, Arendt-Nielsen et al. 1992).

There are different types of CYP inhibitors, including competitive inhibitors

and mechanism-based inhibitors (MBIs; also known as suicide or irreversible

inhibitors). A competitive inhibitor binds at, and thereby blocks, the CYP’s

substrate-binding site (de Groot, Wakenhut et al. 2009). A MBI is a CYP substrate

7

that forms a reactive metabolite which binds covalently to the enzyme (Bertelsen,

Venkatakrishnan et al. 2003, Van, Heydari et al. 2006). MBIs cause the irreversible

loss of enzyme function, requiring the synthesis of new enzyme before activity is

restored.

The antiarrhythmic propafenone is a CYP2D6 competitive inhibitor with a Ki of

2.9 M (Kroemer, Fischer et al. 1991). It also blocks sodium channels and is used to

treat cardiac arrhythmias (Dukes and Vaughan Williams 1984).

The -adrenergic receptor blocker, propranolol, is a selective substrate with

a high affinity for CYP2D6 (Yamamoto, Suzuki et al. 2003). It is used to treat

hypertension, angina pectoris, and cardiac arrhythmias (Komura and Iwaki 2005). It

is also a potent MBI of CYP2D6 (Masubuchi, Narimatsu et al. 1994), with a Ki of 1 μM

(Rowland, Yeo et al. 1994). Propranolol undergoes 4-hydroxylation by CYP2D which

is associated with the formation of a reactive metabolite in the active site of the

enzyme. This reactive species covalently binds to the active site, thereby

inactivating the enzyme (Rowland, Yeo et al. 1994, Narimatsu, Arai et al. 2001).

CYP2D6 inactivates the neurotoxin 1-methyl-4-phenylpyridinium (MPP+); treating

cells with 0.1–30 M propranolol significantly increased the neurotoxicity and cell

death caused by MPP+ (Mann and Tyndale 2010).

1.1.4 CYP2D6 regulation

Hepatic CYP2D6 is constitutively expressed during adulthood (Transon, Lecoeur et

al. 1996, Stevens, Marsh et al. 2008) and is uninducible by common CYP inducers

8

such as phenobarbital (Rae, Johnson et al. 2001, Edwards, Price et al. 2003). Even so,

the promoter of CYP2D6 contains binding sites for positively and negatively acting

transcription factors such as Oct-1, YY-1, heterogeneous nuclear ribonucleoprotein

K and GABP (Yokomori, Kobayashi et al. 1995, Mizuno, Takahashi et al. 2003, Sakai,

Sakamoto et al. 2009). Hepatic CYP2D6 is transcriptionally regulated in large part by

the hepatocyte nuclear factor-4α (HNF-4α) transcription factor (Cairns, Smith et al.

1996, Jover, Bort et al. 1998, Corchero, Granvil et al. 2001). Expression of this

transcription factor is correlated with CYP2D6 expression (Cairns, Smith et al. 1996,

Corchero, Granvil et al. 2001). Expression of HNF-4α is highest in the liver and

kidneys, which is also where CYP2D6 expression is highest (Gonzalez 1990, Xie,

Liao et al. 2009).

1.1.5 CYP2D6 genetic variation

CYP2D6’s role in metabolizing an extensive range of commonly prescribed drugs

makes the wide interindividual variation in its functional levels important. There are

more than 80 known CYP2D6 allelic variants

(http://www.cypalleles.ki.se/cyp2d6.htm) which include gene deletions, frameshift

mutations, insertions, synonymous and non-synonymous substitutions, and copy

number variants (Gaedigk, Simon et al. 2008). These variants can be grouped into

null, reduced, normal, and increased function alleles. Individuals can be grouped

based on CYP2D6 genotype into four CYP2D6 phenotypic categories: poor

metabolizer (PM) (Rae, Johnson et al. 2001), intermediate metabolizer (IM),

9

extensive metabolizer (EM), and ultra-rapid metabolizer (UM) (Gaedigk, Simon et al.

2008).

Genetic variation in CYP2D6 is a major contributor to the interethnic and

interindividual differences in CYP2D6 activity. This variation in CYP2D6 activity can

then affect an individual’s response to the numerous drugs which are metabolically

activated or inactivated by CYP2D6. For example, PMs experience no analgesia

from codeine, which is activated by CYP2D6 (Sindrup, Brosen et al. 1990, Chen,

Somogyi et al. 1991). PMs experience increased side effects from the antipsychotics

haloperidol and risperidone, which are metabolized by CYP2D6 (de Leon, Susce et

al. 2005, Ingelman-Sundberg, Sim et al. 2007). UMs have lower drug efficacy from

the antidepressant imipramine, which is inactivated by CYP2D6 (Schenk, van

Fessem et al. 2008).

1.1.5a Interethnic variability in CYP2D6

The frequencies of CYP2D6 alleles that result in different levels of enzyme activity

vary substantially between ethnic groups, and this contributes to the interethnic

variability in CYP2D6 activity. Caucasians have the highest prevalence of PMs (5-

10%) (Zanger, Raimundo et al. 2004), which is largely (70-90%) due to the high

frequency (20-25%) of the CYP2D6*4 allele in this population (Zanger, Raimundo et

al. 2004, Neafsey, Ginsberg et al. 2009, Abraham, Maranian et al. 2010). Individuals

of African descent have the highest frequency (20-34%) of the CYP2D6*17 reduced-

function allele, which results in much lower CYP2D6 activity than wild-type

10

(Gaedigk, Bhathena et al. 2005). The most prevalent (37-70%) variant among Asians

is the reduced-function allele CYP2D6*10, which results in lower CYP2D6 activity

than wild-type (Garcia-Barcelo, Chow et al. 2000, Neafsey, Ginsberg et al. 2009).

CYP2D6 copy number variants, which result in a UM phenotype, are the most

frequent in North African (28-56%) and Middle Eastern (3-10%) populations.

Ethiopians, Saudi Arabians, and Spaniards have the highest occurrence CYP2D6

UMs, which comprise ~29%, ~20% and ~10% of the population, respectively

(Agundez, Ledesma et al. 1995, Dahl, Johansson et al. 1995, Aklillu, Persson et al.

1996, McLellan, Oscarson et al. 1997, Sachse, Brockmoller et al. 1997). This

interethnic variation in CYP2D6 may make some populations more susceptible to

adverse drug reactions or altered drug efficacy (De Gregori, Allegri et al. 2010).

1.1.6 CYP2D expression in different species

Members of the CYP2D subfamily have been identified among many mammalian

species including human, monkey, rat and mouse. While CYPs are mostly well

conserved across species (Lin 1995), small genetic differences can cause

interspecies differences in CYP expression, activity, substrate specificity and

regulation. Such is the case for CYP2D.

Whereas CYP2D6 is the only functional CYP2D isozyme in humans, mice have

nine different isozymes of Cyp2d. While Cyp2d22 is the most similar out of these

nine to CYP2D6 in terms of amino acid identity (Yu and Haining 2006, McLaughlin,

Dickmann et al. 2008), it differs from CYP2D6 in substrate specificity (Blume,

11

Leonard et al. 2000, Yu and Haining 2006). In contrast to human CYP2D6, Cyp2d22

and other mouse Cyp2ds have only a weak ability to 4-hydroxylate debrisoquine

and O-demethylate dextromethorphan (Lofgren, Hagbjork et al. 2004, Yu, Idle et al.

2004, Yu and Haining 2006, McLaughlin, Dickmann et al. 2008, Shen and Yu 2009).

Rats have six different CYP2D isozymes: CYP2D1, 2, 3, 4, 5 and 18. These vary

in substrate specificity, metabolism, and inhibition profiles (Strobl, von Kruedener et

al. 1993, Hiroi, Chow et al. 2002); they also have different tissue-specific expression

patterns, with CYP2D1 and CYP2D2 being the isozymes most abundant in the liver

(Wyss, Gustafsson et al. 1995, Haduch, Bromek et al. 2011). CYP2D1 has 71% amino

acid identity to human CYP2D6 (Funae, Kishimoto et al. 2003) and is believed to be

the rat homologue of human CYP2D6 (Miksys, Rao et al. 2000). CYP2D1 is capable of

performing many CYP2D6-mediated reactions, such as codeine and

dextromethorphan O-demethylation, and debrisoquine 4-hydroxylation (Matsunaga,

Zanger et al. 1989, Xu, Aasmundstad et al. 1997, Miksys, Rao et al. 2000). These

similarities between rat and human CYP2D enzymes make the rat a useful model of

human CYP2D6-mediated drug metabolism.

1.2 Brain Cytochromes P450

Most CYPs are expressed mainly in the liver, where they are responsible for the

majority of drug metabolism and drug clearance in the body. However, metabolism

by CYPs in extrahepatic tissues may significantly affect drug efficacy by changing

the local, target-site drug concentrations. Moreover, because of the tissue-specific

12

ways that CYPs are regulated, different tissues may respond differently to the same

drug. For example, many centrally-acting drugs are both substrates and inducers of

brain CYPs (e.g. nicotine, phenytoin, ethanol (Miksys, Hoffmann et al. 2000,

Schoedel, Sellers et al. 2001, Howard, Miksys et al. 2003, Meyer, Gehlhaus et al.

2007)), and thus may alter the brain’s sensitivity to these drugs and other brain CYP

substrates. Since the total CYP content in the brain is only a fraction of that in the

liver (Hedlund, Gustafsson et al. 2001, Gervasini, Carrillo et al. 2004), it is unlikely

that metabolism by brain CYPs affects plasma drug levels (Hedlund, Gustafsson et

al. 2001). However, their highly localized expression in different brain regions might

produce microenvironments in which brain CYP-mediated metabolism has a

significant impact on local drug levels and effect (Miksys and Tyndale 2002, Ghosh,

Gonzalez-Martinez et al. 2010). Metabolism by brain CYPs may be particularly

important for centrally-acting substrates which have active metabolites that are not

able to cross the blood-brain barrier. In such cases, the local production of

metabolites in brain may be crucial to the effect of drugs, toxins, and endogenous

neurochemicals.

1.2.1 Brain CYP expression

Of the 57 human CYP transcripts, 41 have been identified in the brain so far (Dauchy,

Dutheil et al. 2008, Dutheil, Dauchy et al. 2009). Only a fraction of these (i.e., CYP1A,

CYP1B, CYP2B, CYP2C, CYP2D, CYP2E, and CYP3A families) have been examined

in the brain at the transcript, protein, and/or activity level (Haining 2007, Dauchy,

13

Dutheil et al. 2008, Dutheil, Dauchy et al. 2009). In the brain, CYPs are expressed in

various cellular membranes including the plasma membrane, endoplasmic

reticulum, and mitochondrial membrane (Miksys, Rao et al. 2000, Howard, Miksys et

al. 2003, Miksys, Lerman et al. 2003, Haining 2007, Woodland, Huang et al. 2008,

Dutheil, Dauchy et al. 2009).

The expression of brain CYPs varies greatly depending on region and cell-

type (Dutheil, Dauchy et al. 2009). Within brain regions, CYPs are expressed at

different levels in pyramidal, Purkinje, granular, neuronal, astrocytic, and glial cells

(Miksys, Hoffmann et al. 2000, Howard, Miksys et al. 2003, Miksys, Lerman et al.

2003, Dutheil, Beaune et al. 2008). While the level of CYPs in the brain has been

estimated to be 1-10% of that in the liver (Hedlund, Wyss et al. 1996, Gervasini,

Carrillo et al. 2004), brain tissue is not homogenous and CYPs are not uniformly

expressed across regions, so this percentage range is unlikely to reflect all brain

regions or all CYPs. For example, there is a ~2.5-fold difference in CYP2B

expression between brain regions of highest and lowest expression in both humans

and rats (Miksys, Hoffmann et al. 2000, Miksys, Lerman et al. 2003). In fact,

expression levels of CYPs in brain cells (e.g., CYP2B in frontal cortex pyramidal

neurons) can be equal to, or higher than, levels in hepatocytes (Miksys, Hoffman et

al. 2000). The highly localized expression of brain CYPs is thought to create

microenvironments in which local, in situ drug metabolism occurs in the brain (Britto

and Wedlund 1992, Miksys and Tyndale 2009). This may in turn alter the local

pharmacokinetics and effect of these drugs.

CYPs are present at the BBB (Miksys, Rao et al. 2000, Miksys, Lerman et al.

2003, Dauchy, Dutheil et al. 2008, Ghosh, Gonzalez-Martinez et al. 2010) as well as in

14

areas lacking BBB such as the choroid plexus and posterior pituitary (Volk,

Hettmannsperger et al. 1991, Ghersiegea, Perrin et al. 1993, Miksys, Rao et al. 2000).

This suggests that brain CYPs may play a role in assisting the BBB in preventing

drugs and toxins from entering the brain. Because some metabolites made in the

periphery are less permeable across the BBB, the formation of metabolites within the

brain can be crucial to the effect of centrally-acting drugs.

1.2.2 Brain CYP activity

Studies of in vitro brain CYP activity using brain homogenates have suggested that

brain CYPs are able to carry out the same reactions as hepatic CYPs. This has been

shown using different substrates including nicotine, chlorpyrifos and codeine

(Chambers and Chambers 1989, Chen, Irvine et al. 1990, Jacob, Ulgen et al. 1997).

Other in vitro studies have revealed brain CYPs to have similar substrate

specificities and affinities (Km) as their hepatic forms (Forsyth and Chambers 1989,

Lin, Kumagai et al. 1992, Bhamre, Anandatheerathavarada et al. 1993, Ghersiegea,

Perrin et al. 1993, Narimatsu, Yamamoto et al. 1999, Tyndale, Li et al. 1999, Bhagwat,

Boyd et al. 2000, Voirol, Jonzier-Perey et al. 2000). However, as cofactors were

added to these reactions, it is possible that there are insufficient endogenous levels

of these cofactors in the brain to carry out these reactions in vivo.

Studying in vivo brain CYP activity is challenging since peripheral metabolites

formed from hepatic metabolism can enter into the brain, thus making it hard to

distinguish metabolites formed from hepatic versus brain metabolism. Also, heme

15

levels in the brain are rate-limiting for at least some CYP functions, suggesting that

brain CYPs may not all be functional in vivo (Meyer, Lindberg et al. 2005). Another

rate-limiting factor may be a potential lack of the coenzyme NADPH-cytochrome

P450 oxidoreductase (POR), which is required for CYP function, near brain CYPs

(Miksys and Tyndale 2009). These factors contribute to the shortage of evidence for

in situ brain CYP activity.

The finding that the coenzyme POR is expressed in the same regions as brain

CYPs (Haglund, Kohler et al. 1984, Ghersiegea, Minn et al. 1989, Bergh and Strobel

1992, Bergh and Strobel 1996, Riedl, Watts et al. 1996, Conroy, Fang et al. 2010)

lends support to the feasibility of in situ brain CYP activity. It has recently been

shown that brain CYP2B protein is active in situ in living rats (Miksys and Tyndale

2009). Rats were pretreated on one side of the brain with a CYP2B MBI before

receiving bilateral intracerebral injections of a different radiolabeled CYP2B MBI.

This radiolabeled MBI becomes bound upon being metabolized by CYP2B. There

was significantly lower radiolabel binding on the inhibitor-treated side of the brain

compared to the untreated side (Miksys and Tyndale 2009). This demonstrated that

rat brain CYP2B is active in vivo, without the addition of cofactors.

The function of brain CYPs may be to protect against exogenous drugs and

toxins and/or to metabolize or catalyze the formation of endogenous compounds

such as neurosteroids and biogenic amines (Haining 2007). Certain CYPs can

metabolize or catalyze the formation of serotonin, dopamine, arachidonic acid,

pregnenolone, estradiol, androstenedione, testosterone, and melatonin (Rifkind, Lee

et al. 1995, Doostzadeh and Morfin 1997, Rosenbrock, Hagemeyer et al. 1999, Ohe,

Hirobe et al. 2000, Wang, Napoli et al. 2000, Fradette, Yamaguchi et al. 2004, Ma,

16

Idle et al. 2005, Bromek, Haduch et al. 2010). CYP2B metabolically inactivates the

anaesthetic propofol; inhibiting brain (but not hepatic) CYP2B resulted in longer

propofol-induced sleep times in rats (Khokhar and Tyndale 2011), suggesting that

brain CYP activity can have a meaningful impact on drug response.

1.2.3 Brain CYP regulation

Brain CYPs are induced in different ways depending on the CYP, brain region, cell

type, and inducer, and are also regulated differently from their hepatic forms

(Miksys and Tyndale 2002, Miksys and Tyndale 2004). Inducers of hepatic CYPs do

not always induce the corresponding CYPs in the brain, and vice versa. For

example, nicotine and ethanol can induce CYPs in vivo in an organ- and CYP-specific

way. Nicotine induces CYP2E1 in both the brain and liver in monkeys and rats, but it

induces CYP2B and CYP2D only in the brain (Miksys, Hoffmann et al. 2000, Joshi and

Tyndale 2006, Lee, Miksys et al. 2006a, Mann, Miksys et al. 2008, Yue, Miksys et al.

2008, Yue, Khokhar et al. 2009). Ethanol induces CYP2E1 in both the brain and liver

in rats, but it induces CYP2B and CYP2D in the liver only (Warner and Gustafsson

1994, Schoedel, Sellers et al. 2001, Howard, Miksys et al. 2003, Schoedel and

Tyndale 2003). In humans, cigarette smoking and alcohol use are both associated

with higher levels of CYP2B6, CYP2D6 and CYP2E1 in certain brain regions (Miksys,

Rao et al. 2002, Howard, Miksys et al. 2003, Miksys, Lerman et al. 2003, Miksys and

Tyndale 2004).

17

The induction of brain CYPs can alter drug effect and contribute to the

interindividual variation in response to centrally-acting drugs (Chimbira and

Sweeney 2000, Jabs, Bartsch et al. 2003, Funck-Brentano, Boelle et al. 2005,

Lysakowski, Dumont et al. 2006, George, Sacco et al. 2008). Drug plasma levels do

not always correlate well with drug response, and this is particularly the case for

certain centrally-acting drugs such as antipsychotics and antidepressants (Michels

and Marzuk 1993a, Nelson, Mazure et al. 1995, Lane, Chiu et al. 2000, Spina, Avenoso

et al. 2001, Riedel, Schwarz et al. 2005 ). Even at plasma levels of these drugs that

are expected to produce maximal therapeutic effects and minimal adverse effects,

there can be either no therapeutic effect or adverse side effects (Michels and Marzuk

1993a, Michels and Marzuk 1993b). This phenomenon may be partly accounted for

by drug metabolism occurring in the brain. Induction of brain CYPs as a result of

exposure to nicotine or alcohol may magnify this effect and contribute to

interindividual variation in drug response. In support of this, smokers require

higher doses of the anaesthetic propofol, which is inactivated by CYP2B6, in order to

achieve loss of consciousness, consistent with smokers having increased CYP2B6

levels and experiencing less adverse side effects from propofol (Chimbira and

Sweeney 2000, Lysakowski, Dumont et al. 2006).

1.2.4 Brain CYP2D6

CYP2D6 may be an important enzyme in the brain as it metabolizes many centrally-

acting substrates which include clinically prescribed drugs (Zanger, Raimundo et al.

18

2004) as well as drugs of abuse, neurotoxins and endogenous neurochemicals

(Miksys, Rao et al. 2000, Mann, Miksys et al. 2008).

1.2.4a Brain CYP2D expression

CYP2D expression has been detected in the brain of rat, mouse, dog, monkey, and

human (Fonne-Pfister, Bargetzi et al. 1987, Niznik, Tyndale et al. 1990, Tyndale,

Sunahara et al. 1991, Tyndale, Li et al. 1999, Siegle, Fritz et al. 2001, Miksys, Cheung

et al. 2005). In rats, CYP2D1, CYP2D4, CYP2D5 and CYP2D18 have been detected in

the brain (Komori 1993, Wyss, Gustafsson et al. 1995, Coleman, Spellman et al. 2000,

Miksys, Rao et al. 2000). CYP2D4 is mainly expressed in the brain and is the most

abundant CYP2D isozyme in the rat brain (Komori 1993, Wyss, Gustafsson et al.

1995), while CYP2D18 is thought to be expressed only in the brain (Coleman,

Spellman et al. 2000). Rat CYP2D2 and CYP2D3 have yet to be found in the brain

(Miksys, Rao et al. 2000).

In humans, CYP2D6 is expressed in most brain regions, including the

neocortex, caudate, putamen, globus pallidus, nucleus accumbens, hippocampus,

hypothalamus, thalamus, substantia nigra, cerebellum, and medulla oblongata

(Gilham, Cairns et al. 1997, McFayden, Melvin et al. 1998, Siegle, Fritz et al. 2001,

Miksys, Rao et al. 2002). CYP2D6 protein levels are highest in the caudate, putamen,

cortex, and cerebellum (Miksys, Rao et al. 2002).

In rats, CYP2D protein is also expressed in most brain regions, and moderate

to high levels are found in the cerebellum, hippocampus, medulla oblongata, pons,

19

cerebral cortex, striatum (caudate/putamen), thalamus, substantia nigra, choroid

plexus, and amygdaloid complex (Miksys, Rao et al. 2000).

Human CYP2D6 protein is expressed in a cell type-specific manner in the

brain. High levels are found in pigmented neurons of the substantia nigra, pyramidal

cells of the hippocampus and frontal cortex, Purkinje cells of the cerebellum, glial

cells, astrocytes, and endothelial cells at the BBB (Gilham, Cairns et al. 1997, Siegle,

Fritz et al. 2001, Miksys, Rao et al. 2002, Dauchy, Miller et al. 2009, Dutheil, Jacob et

al. 2010), similar to the cell types in which rat brain CYP2D is expressed (Michels

and Marzuk 1993a, Michels and Marzuk 1993b, Watts, Riedl et al. 1998, Riedl, Watts

et al. 1999).

1.2.4b Brain CYP2D function and activity

CYP2D6 activity in the brain may be important as many of CYP2D6’s substrates act

within the CNS. These include clinically prescribed drugs such as the

antidepressants fluoxetine and paroxetine, the analgesics codeine and oxycodone,

and the antipsychotics risperidone and haloperidol (Zanger, Raimundo et al. 2004).

Some CYP2D6 substrates are also commonly abused drugs such as codeine,

oxycodone, MDMA and dextromethorphan (Zanger, Raimundo et al. 2004). Brain

CYP2D6-mediated metabolism of these drugs could alter their disposition within the

brain and thereby alter their efficacy, side-effect profile and abuse liability. We are

thus interested in studying brain CYP2D activity and its role in drug metabolism and

response.

20

Brain CYP2D activity has been demonstrated in vitro using hydroxylation of

bufuralol and MDMA and demethylation of codeine and dextromethorphan in rats,

and sparteine demethylation in dogs (Chen, Irvine et al. 1990, Tyndale, Sunahara et

al. 1991, Lin, Kumagai et al. 1992, Jolivalt, Minn et al. 1995, Tyndale, Li et al. 1999,

Coleman, Spellman et al. 2000, Voirol, Jonzier-Perey et al. 2000). CYP2D substrate

affinities (Km) in the brain are comparable to those in the liver; however, because

CYP expression is lower in the brain, the maximal velocity (Vmax) and substrate

turnover (Vmax/Km) are lower (Tyndale, Sunahara et al. 1991, Coleman, Spellman et

al. 2000).

Cultured SH-SY5Y (a neuron-like cell line) cells can metabolize codeine to

morphine, a reaction catalyzed by CYP2D6 (Poeaknapo, Schmidt et al. 2004). These

cells can also metabolize the CYP2D probe drug 3-[2-(N,N-diethyl-N-

methylammonium)-ethyl]-7-methoxy-4-methylcoumarin, and this reaction was

inhibited by CYP2D inhibitors (Mann and Tyndale 2010). These findings suggest that

brain CYP2D is active and carries out these enzymatic reactions.

The formation of dextrorphan from dextromethorphan has been demonstrated

in rat brain membranes (Tyndale, Li et al. 1999). This reaction was inhibited by

classic CYP2D inhibitors and by antibodies raised against CYP2D1, but not by

inhibitors or antibodies against CYP2B, CYP2C or CYP3A. Rat CYP2D activity varies

across brain regions, and there was also a strong correlation of dextromethorphan

O-demethylation with brain CYP2D1 mRNA levels, as well as with brain CYP2D

protein levels. The rat cerebellum displayed the highest dextromethorphan

metabolism and CYP2D protein levels. These findings suggest that CYP2D1 is

responsible for this reaction in rat brain. Because dextromethorphan is a CYP2D6

21

probe drug and its O-demethylation is measure of human CYP2D6 activity, this

suggests that CYP2D1 is the rat homologue of human CYP2D6 (Miksys, Rao et al.

2000, Frank, Jaehde et al. 2007). CYP2D1 has been shown to perform other CYP2D6-

mediated reactions as well, such as codeine O-demethylation and debrisoquine 4-

hydroxylation (Matsunaga, Zanger et al. 1989, Xu, Aasmundstad et al. 1997).

The activity of brain CYP2D in vivo has been suggested by the local

metabolism of codeine to morphine in rat brain during the first 30 minutes after

peripheral codeine injection, which is believed to be the responsible for brain

morphine levels and analgesia at this time point (Chen, Irvine et al. 1990). Rats were

injected with either codeine (20 mg/kg, i.p.) or morphine (1 mg/kg, i.p.) at doses

that produced equal plasma morphine levels between the two drugs. At 30 min after

injection, while plasma morphine levels were the same between codeine- and

morphine-treated rats, morphine in the brain was found in the codeine-treated rats

but could not be detected in the morphine-treated rats (Figure 1). This strongly

suggests that at this time point, morphine formed from hepatic metabolism has not

yet crossed into the brain.

22

1.a) Plasma b) Brain

Figure 1. Morphine levels in rat plasma and brain 30 min after peripheral

injection of codeine or morphine. Morphine concentrations in (a) plasma and (b)

brain after i.p. injection of 20 mg/kg codeine phosphate (white bar) or 1 mg/kg

morphine sulphate (grey bar) (n=4/group). # no morphine detected. This figure is

adapted from Chen, Irvine et al. (1990).

0

400

800

1200[M

orp

hin

e] (n

g/m

l, m

ean

+ S

EM)

Codeine Morphine 0

12

24

36

[Mo

rph

ine]

(ng

/g, m

ean

+ S

EM)

Codeine Morphine

#

23

These findings are in concordance with the fact that morphine has one less

methyl group than codeine, which is expected to make morphine less lipid soluble

and therefore less able to cross the BBB. In one study (Oldendorf, Hyman et al.

1972), the brain uptake of codeine or morphine (i.e., the brain content of the drug as

a percentage of a highly diffusible reference substance injected simultaneously) was

measured in rats. At 15 sec after intra-arterial injection, the uptake of codeine was

24%, whereas the uptake of morphine was below quantification. At 30 sec after

intravenous injection of codeine or morphine, the uptake of codeine was nearly

complete whereas this was much lower with morphine. These findings indicate that

codeine enters the brain faster than morphine does. Morphine is also transported

out of the brain by efflux transporters at the BBB, and this could contribute to the

delay in antinociception after morphine administration in rats (Bouw, Gardmark et

al. 2000). These findings suggests that morphine found in the brain during the first 30

min after codeine administration may be due to local morphine formation in the

brain (Chen, Irvine et al. 1990). In addition, in rat brains infused with the neurotoxin

MPTP, there was local inactivation of MPTP to PTP in the striatum, a reaction

catalyzed by CYP2D (Vaglini, Pardini et al. 2004). Thus, while the study by Chen,

Irvine et al. (1990) did not explicitly determine whether morphine found in the brain

after codeine injection was formed by brain CYP2D-mediated metabolism, together

these studies suggest that CYP2D is capable of metabolizing substrates in situ in the

brain.

In addition to metabolizing many centrally-acting drugs and inactivating

neurotoxins, CYP2D6 can also metabolize endogenous compounds found in the

brain, such as biogenic amines (Yu, Idle et al. 2003a). For example, cDNA expressed

24

CYP2D6 and CYP2D6 transgenic mice can metabolize 5-methoxytryptamine to

serotonin (Yu, Idle et al. 2003a). Rat brain membranes, as well as CYP2D6 expressed

in yeast cells, can convert tyramine to dopamine (Hiroi, Imaoka et al. 1998, Bromek,

Haduch et al. 2010). These findings suggest that brain CYP2D6 activity may have an

impact on behaviour or personality (Hiroi, Imaoka et al. 1998, Yu, Idle et al. 2003a),

which is further supported by the association of CYP2D6 genotype with personality

traits (Gan, Ismail et al. 2004, Roberts, Luty et al. 2004, Kirchheiner, Lang et al. 2006,

Penas-Lledo, Dorado et al. 2009, Gonzalez, Penas-Lledo et al. 2008).

1.2.4c Brain CYP2D regulation

Brain CYP2D, unlike hepatic CYP2D, is inducible by various centrally-acting drugs

in a drug- and brain region-specific way (Rae, Johnson et al. 2001, Edwards, Price et

al. 2003). For example, the antipsychotic clozapine increased levels of rat CYP2D

protein in neurons of the substantia nigra, ventral tegmental area, olfactory bulb,

and cerebellum but it did not alter hepatic levels (Hedlund, Wyss et al. 1996).

CYP2D mRNA levels in the brain were unaltered, which suggests that brain CYP2D is

induced at the posttranscriptional level. The antidepressant fluoxetine produced an

increase in CYP2D protein and activity in rat cerebellum (Haduch, Bromek et al.

2011). Rats treated with thioridazine had an increase in CYP2D protein and activity in

the substantia nigra and cerebellum (Haduch, Bromek et al. 2011). Toluene caused

an increase in CYP2D4 mRNA, protein, and activity in rat brain (Mizuno, Hiroi et al.

2003).

25

Nicotine and ethanol have been shown to induce brain CYP2D in multiple

species. In rats, chronic (7 day) nicotine treatment increased CYP2D protein in the

cerebellum, hippocampus, and striatum (Yue, Miksys et al. 2008). The CYP2D mRNA

was unchanged, suggesting that induction is due to posttranscriptional modification.

In addition, hepatic CYP2D was unaltered. Acute ethanol treatment induced rat brain

CYP2D (Warner and Gustafsson 1994). In monkeys, nicotine treatment increased

CYP2D protein in the brain when compared to saline-treated animals; there was no

change in hepatic CYP2D (Mann, Miksys et al. 2008). Ethanol self-administration in

monkeys increased brain CYP2D without altering hepatic CYP2D (Miller, Miksys et

al. 2012). In humans, we have observed elevated brain CYP2D6 levels in alcoholics

compared to non-alcoholics and smokers compared to non-smokers (Miksys, Rao et

al. 2002, Miksys and Tyndale 2004).

In addition to regulation by environmental inducers, brain CYP2D is also

under developmental and hormonal regulation. Brain CYP2D6 levels increase with

age (Mann, Miksys et al. 2012). Estrogen and testosterone can change brain CYP2D

mRNA levels in ovariectomized rats (Bergh and Strobel 1996). Testosterone induced

brain CYP2D, whereas testosterone combined with estrogen treatment reduced

brain CYP2D induction, suggesting that estrogen may block the induction of brain

CYP2D levels.

The mechanism(s) of brain CYP2D induction is as of yet unknown. Because

nicotine does not increase CYP2D mRNA levels (Hedlund, Wyss et al. 1996, Mann,

Miksys et al. 2008, Yue, Miksys et al. 2008), this implies that CYP2D induction by

nicotine occurs via post-transcriptional events. These could potentially include

decreased splicing, increased translation, increased enzyme stabilization, and

26

decreased protein degradation. In rats, nicotine regulates certain ubiquitinating

proteins in the brain (Kane, Konu et al. 2004), which may alter CYP2D levels by

affecting degradation.

Increased levels of brain CYP2D6 may result in increased substrate

metabolism, which in turn may lead to altered efficacy of clinical drugs, as well as

altered susceptibility to adverse drug reactions. Evidence for this comes from the

observations that smokers have less extrapyramidal side effects from antipsychotics

(inactivated by CYP2D6) than nonsmokers (Jabs, Bartsch et al. 2003), and that

smokers and seniors experience less efficacy from antidepressants (inactivated by

CYP2D6) than nonsmokers and younger patients (Nelson, Mazure et al. 1995,

George, Sacco et al. 2008). The higher levels of brain CYP2D6 in smokers and older

individuals may increase the inactivation of these drugs in the brain, resulting in

reduced therapeutic and/or adverse effects.

1.3 Opioid Analgesics

1.3.1 Codeine

CYP2D6 metabolizes the opioid analgesic codeine (prodrug) to morphine (active

metabolite) (Adler, Fujimoto et al. 1955). Opioids confer their analgesic effects

through interacting with mu-opioid receptors, which are ‘Gi/Go-coupled’ receptors

(Law, Wong et al. 2000). Morphine has much greater (3000-fold) affinity for mu-

opioid receptors than does codeine (Pert and Snyder 1973), so even though

27

morphine is a minor metabolite, codeine-induced analgesia is dependent on its

metabolism to morphine. Codeine must be metabolized to morphine to produce

analgesia in both humans (Chen, Somogyi et al.1991) and rats (Mikus, Somogyi et al.

1991, Cleary, Mikus et al. 1994), and this reaction is performed solely by CYP2D

(Thorn, Klein et al. 2009). In humans, CYP2D6 PMs and individuals pretreated with

the CYP2D6 inhibitor quinidine produce little to no morphine from codeine and

experience no analgesia (Sindrup, Brosen et al. 1990, Chen, Somogyi et al. 1991,

Sindrup, Arendt-Nielsen et al. 1992). In rats pretreated with i.p. injections of the

CYP2D1 inhibitor quinine, there was a substantial reduction in codeine-induced

analgesia compared to untreated rats (Cleary, Mikus et al. 1994). Furthermore,

female Dark-Agouti rats, which lack CYP2D1 and are an animal model of CYP2D6

PMs, experienced no analgesia from codeine (Cleary, Mikus et al. 1994). Therefore,

analgesia from codeine requires its conversion to morphine by CYP2D. Variation in

brain CYP2D activity may affect morphine levels in the brain after codeine

administration, which in turn may affect the analgesic response to codeine.

1.3.1a Codeine metabolism in humans

In the human liver, 50-70% of codeine is glucuronidated by UGT2B7 (Coffman, Rios

et al. 1997) and UGT2B4 (Court, Krishnaswamy et al. 2003) to codeine-6-glucuronide,

10-15% is N-demethylated by CYP3A4 to norcodeine (Caraco, Tateishi et al. 1996,

Yue and Sawe 1997), 0-15% is O-demethylated by CYP2D6 to morphine (Thorn,

Klein et al. 2009), and 5-15% is excreted unchanged. About 60% of morphine is

28

glucuronidated to morphine-3-glucuronide, and 5-10% is glucuronidated to

morphine-6-glucuronide (Lotsch, Stockmann et al. 1996, Ohno, Kawana et al. 2008).

Both of these conjugations are performed mainly by UGT2B7, with a small

contribution by UGT1A1 (Holthe, Klepstad et al. 2002). Morphine is also N-

demethylated to normorphine, mainly by CYP3A4 with CYP2C8 playing a smaller

role (Projean, Morin et al. 2003). Normorphine can also be formed by the O-

demethylation of norcodeine by CYP2D6 (Yue, Hasselstrom et al. 1991). Norcodeine

can be glucuronidated to norcodeine-6-glucuronide (Yue, Hasselstrom et al. 1991).

Besides morphine, the other metabolites of codeine that have an analgesic effect are

normorphine and morphine-6-glucuronide (Lasagna and De Kornfeld 1958,

Osborne, Joel et al. 1988)). Both of these first require the O-demethylation of codeine

(or of norcodeine) by CYP2D6 in order to be formed (Yue, Hasselstrom et al. 1991).

Therefore, O-demethylation by CYP2D6 is necessary in the formation of all of

codeine’s analgesic metabolites (Figure 2). The difference in the O-demethylation

of codeine can be as large as 25-fold between EMs and PMs and 45-fold between

UMs and the PMs (Yue, Alm et al. 1997).

29

Figure 2. Metabolic pathways of codeine in humans. Analgesic metabolites are

in uppercase.

Codeine

Codeine-6-glucuronide MORPHINE Norcodeine Unchanged codeine

MORPHINE- 6-GLUCURONIDE

Morphine- 3-glucuronide

NORMORPHINE Norcodeine- 6-glucuronide

30

In the human brain, CYP2D6 expression and activity are much higher than

those of CYP3A, which are undetectable in some areas (Voirol, Jonzier-Perey et al.

2000, Dutheil, Dauchy et al. 2009). Therefore, CYP2D6 may play a larger role in

codeine metabolism than CYP3A in the brain.

In addition to altering codeine-induced analgesia, variation in CYP2D6

activity also affects susceptibility to codeine toxicity and abuse. For example, when

a CYP2D6 UM who had 3 or more functional alleles received a small dose of codeine,

this lead to serious toxicity as a result of the high levels and fast rates of morphine

and morphine-6-glucuronide formed (Gasche, Daali et al. 2004). CYP2D6 PMs are

underrepresented among individuals dependent on oral opioid drugs, which

suggests that the O-demethylated metabolites of codeine confer its reinforcing

effects, and that low CYP2D6 activity may reduce susceptibility to codeine abuse

(Tyndale, Droll et al. 1997). This is further supported by studies which have

examined the effect of CYP2D6 inhibitors on codeine abuse liability. Individuals

pretreated with quinidine had lower plasma levels of O-demethylated metabolites

and experienced fewer positive subjective effects from codeine

(Kathiramalainathan, Kaplan et al. 2000). Fourteen long-term users of oral opioid

drugs (primarily codeine) who were treated with fluoxetine, a CYP2D6 inhibitor, had

a decrease in CYP2D6 activity as well as a 30% to 100% decrease in opioid use

(Romach, Otton et al. 2000). Therefore, variation in brain CYP2D6 activity due to

genetics, environmental inducers or age may lead to differences in brain morphine

levels after codeine administration, which may have implications for the analgesia as

well as abuse liability of codeine.

31

1.3.1b Codeine metabolism in rats

At 24 h after rats were given 10 mg/kg s.c. codeine, 23.9% of the dose was excreted

in the urine as morphine-3-glucuronide, 4.3% as free morphine, 1.6% as unchanged

codeine, and 0.2% as codeine glucuronide; morphine-6-glucuronide was not

detected (Oguri, Hanioka et al. 1990). Rat UGT2B1 only forms the 3-glucuronide, and

whereas rat UGT2B7 can glucuronidate at both the 3- and 6-positions, rat UGT2B7 is

ten times more efficient at catalyzing the 3-glucuronidation (Ritter 2000).

CYP2D1 is responsible for the O-demethylations of codeine to morphine and

norcodeine to normorphine in rats. When antibodies for different rat CYPs were

tested in rat liver microsomes, only the anti-CYP2D1 antibody significantly inhibited

the O-demethylation of these substrates (Xu, Aasmundstad et al. 1997). Also, the

specific CYP2D1 inhibitor quinine inhibited the codeine and norcodeine O-

demethylations, whereas the CYP3A inhibitor did not (Xu, Aasmundstad et al. 1997).

Furthermore, these O-demethylations were impaired in the liver microsomes of

female Dark Agouti rats, which are known to have reduced CYP2D activity (Xu,

Aasmundstad et al. 1997). These findings provide evidence that the metabolism of

codeine to morphine and norcodeine to normorphine in rats is mediated by

CYP2D1.

Codeine can be O-demethylated in rat brain in vitro as demonstrated by the

formation of morphine from codeine by rat brain homogenates (Chen, Irvine et al.

1990). There is also in vivo evidence of codeine metabolism in rat brain. As

described previously in Section 2.4.2, in rats that received peripheral injection of

either codeine or morphine, morphine was found in codeine-treated rats but could

32

not be detected in morphine-treated at 30 min after drug injection, even though

plasma morphine levels were similar from the two drugs. This indicates that at this

time point, morphine formed from hepatic metabolism has not yet crossed into the

brain. This is consistent with the lower lipophilicity (and therefore lesser ease of

crossing the BBB) of morphine compared to codeine and the quicker uptake of

codeine than morphine into the brain (Oldendorf, Hyman et al. 1972). Thus, during

the first 30 min after codeine administration, morphine found in the brain may be

due to local conversion of codeine to morphine in the brain (Chen, Irvine et al.

1990).

In further support of the impact of local codeine metabolism in the brain, both

mu-opioid receptors and CYP2D are widely distributed throughout the brain, and

many brain regions where rat CYP2D levels are moderate to high (cerebral cortex,

striatum (caudate/putamen), hippocampus, brainstem) also are dense with mu-

opioid receptors (Arvidsson, Riedl et al. 1995; Miksys, Rao et al. 2000). Thus,

morphine formed in the brain can immediately bind to the proximate mu-opioid

receptors. Variation in brain CYP2D6 activity may therefore affect morphine levels

in the brain after codeine administration, which in turn may affect response to

codeine.

33

1.3.2 Morphine

1.3.2a Spinal mechanisms of morphine-induced

analgesia

Nociception is the process by which stimuli capable of causing tissue damage

(noxious thermal, mechanical, or chemical stimuli) are detected by primary sensory

neurons called nociceptors (Basbaum and Jessell 2000). Nociceptors convey this

noxious information by projecting to the dorsal horn of the spinal cord, where they

release glutamate and peptides to excite second order neurons (Fields 2004). A

subset of these second order neurons, in turn, project and transmit pain messages to

areas in the brain including the thalamus, brainstem, and ultimately the cerebral

cortex (Basbaum, Bautista et al. 2009).

Because the dorsal horn of the spinal cord is where the first synapse in pain

transmission is located, it is an effective target for the inhibition of pain transmission

by opioids (Heinricher, Tavares et al. 2009). Morphine’s interaction with mu opioid

receptors within the dorsal horn results in the suppression of the release of

neurotransmitters by nociceptors, as well as the hyperpolarization of second order

neurons, thereby reducing the transmission of pain information to higher centres

(McFadzean 1988, Simonds 1988, Lipp 1991).

34

1.3.2b Supraspinal mechanisms of morphine-induced analgesia

Supraspinal (or descending) control of pain transmission at the dorsal horn of the

spinal cord is mediated by several brain areas, all of which have mu-opioid

receptors (Mansour, Khachaturian et al. 1988, Arvidsson, Riedl et al. 1995, Mansour,

Fox et al. 1995, Akil, Owens et al. 1998). The most studied of these areas is the

periaqueductal gray (PAG) - rostral ventromedial medulla (RVM) system, which is

regarded as the principal site of action of opioids (Yaksh, Yeung et al. 1976,

Hohmann, Suplita et al. 2005, Leith, Wilson et al. 2007). The PAG receives input from

the hypothalamus and limbic forebrain structures including the amygdala, as well as

from the spine. The PAG synapses with the RVM, which in turn terminates in the

dorsal horn (Heinricher, Tavares et al. 2009). There is evidence that activation of the

PAG-RVM system results in the release of serotonin and norepinephrine at the spinal

level, and that this mediates its pain-modulatory effects (Proudfit and Hammond

1981, Jensen and Yaksh 1986, Pang and Vasko 1986).

The PAG-RVM system can exert both inhibitory and facilitatory effects on pain

transmission (Heinricher, Tavares et al. 2009). These two opposing effects result

from the activity of two cell classes found in the PAG and RVM called ON-cells and

OFF-cells (Heinricher, Cheng et al. 1987, Fields, Heinricher et al. 1991). In keeping

with their role in pain regulation, RVM ON- and OFF-cells project specifically to

dorsal horn laminae involved in nociceptive transmission (Fields, Malick et al. 1995).

It is the OFF-cells that function as the pain-inhibiting output from the PAG-RVM

system and the ON-cells that are the pain-facilitating output (Heinricher and Ingram

35

2008). Pain threshold is lowest when ON-cells are active and OFF-cells are silent

(Heinricher, Barbaro et al. 1989, Heinricher, Haws et al. 1991).

Opioids analgesics produce their effects by modulating this system

(Heinricher, Morgan et al. 1994, Heinricher, McGaraughty et al. 1999). Opioids, by

acting on mu-opioid receptors, directly inhibit ON-cells. Opioids indirectly activate

OFF-cells, which do not express mu-opioid receptors (Fields, Heinricher et al. 1991,

Heinricher, Morgan et al. 1992, Heinricher, Morgan et al. 1994, Heinricher,

McGaraughty et al. 1999). OFF cells are inhibited by GABA (γ-aminobutyric acid)-

releasing cells (Heinricher, Morgan et al. 1992). These cells have mu-opioid

receptors, and opioids act on them to inhibit the release of GABA, and thereby

activate (disinhibit) OFF-cells (Vaughan and Christie 1997, Vaughan, Bagley et al.

2003).

1.4 Rat Tail-Flick Test: Animal Model of Nociception

1.4.1 Tail-flick reflex

The rat tail-flick test is one of the most widely used models of nociception (Hardy

1953, Hardy, Stoll et al. 1957, Le Bars, Gozariu et al. 2001). It consists of applying a

thermal stimulus in the form of an infrared heat beam to a rat’s tail, which triggers

the withdrawal of the tail by a quick, abrupt movement called the tail-flick reflex

(D'Amour and Smith 1941, Smith, D'Amour et al. 1943). The measured parameter is

36

the time from start of heat exposure to tail-flick reflex, and this is referred to as “tail-

flick latency” (TFL). A prolonging of TFL is an indication of analgesia.

The advantages of this method are as follows. It requires a simple apparatus

and is easy to perform on rats that have been habituated to manipulation. The tail-

flick reflex is easily observed and there is small interanimal variability in baseline

TFL (Le Bars, Gozariu et al. 2001). TFL stays the same with repeated testing if heat

intensity is kept constant and if tissue damage is avoided (Grossman, Basbaum et al.

1982). Tail flicks rarely occur spontaneously (Grossman, Basbaum et al. 1982).

Furthermore, this test is very sensitive to opioids (Le Bars, Gozariu et al. 2001).

Opioids are the only drugs that at nontoxic doses can inhibit tail-flick reflex during

extended (20-30 seconds) exposure to noxious heat (Grumbach 1975). This test is

able to predict the analgesic effects of opioids in humans (Archer and Harris 1965,

Grumbach 1966).

The nociceptors mediating the tail-flick reflex project to the superficial

laminae of the dorsal horn of the spinal cord (Grossman, Basbaum et al. 1982). There

is a high concentration of opioid receptors at this site (Pert, Kuhar et al. 1975, Atweh

and Kuhar 1977) and many of these are located on nociceptor terminals from the tail

(Lamotte, Pert et al. 1976, Fields, Emson et al. 1980). This is in line with the

observation that opioid drugs inhibit tail-flick reflex.

The tail-flick reflex is a spinal reflex, but it is under the influence of

supraspinal structures (Yaksh and Rudy 1978, Millan 2002). The tail-flick reflex can

be completely inhibited by electrical stimulation of brainstem regions (Grossman,

Basbaum et al. 1982). Microinjection of morphine or other opioids into the PAG or

RVM of the rat increases TFL (Jacquet and Lajtha 1973, Yaksh, Yeung et al. 1976,

37

Lewis and Gebhart 1977). When descending pathways are disrupted, such as after

spinal cord transection or cold-block, systemically administered morphine is less

effective in increasing TFL (Irwin, Houde et al. 1951, Basbaum, Clanton et al. 1976,

Basbaum, Marley et al. 1977, Sinclair, Main et al. 1988). These findings suggest that

the tail-flick test is a useful model for assessing supraspinally-mediated analgesia.

1.4.2 Effect of opioid analgesics in the tail-flick test

During instances of pain such as the tail-flick reflex, ON-cells become active and

OFF-cells are silenced (Barbaro, Heinricher et al. 1989, Heinricher, Barbaro et al.

1989). When opioids are microinjected into the PAG or RVM, or administered

systemically, OFF-cells fire continuously and more rapidly, and ON cells become

silent, with concurrent inhibition of tail-flick reflex (Fields 2004). When OFF-cell

activation is selectively blocked, morphine-induced analgesia is abolished

(Heinricher, McGaraughty et al. 1999). This indicates that OFF-cell activation is

necessary for morphine-induced analgesia to occur whether given systemically or

supraspinally (Heinricher, McGaraughty et al. 1997, Heinricher, McGaraughty et al.

2001).

Following electrical stimulation or microinjection of morphine in the PAG or

NRM, there is a release of serotonin in the spinal cord (Yaksh and Tyce 1979,

Hammond, Tyce et al. 1985). The analgesic effect on tail-flick reflex of morphine

microinjected into the PAG or NRM was diminished by intrathecal administration of

antagonists of serotonin and -adrenergic, but not opioid, receptors (Yaksh 1979,

38

Yaksh and Wilson 1979, Camarata and Yaksh 1985, Jensen and Yaksh 1986). These

findings suggest that supraspinal morphine activates downstream serotonergic and

-adrenergic systems which regulate spinal pain transmission (Jones and Gebhart

1988).

1.5 Study design

1. Rats received i.c.v. injections of either propranolol (mechanism-based CYP2D

inhibitor), propafenone (competitive CYP2D inhibitor) or the respective

vehicle.

2. Rats were used in one of five different types of experiments, to assess the

effect of i.c.v. CYP2D inhibitor treatment on:

a) Analgesia after codeine injection, as measured by tail-flick latency (TFL).

b) In vivo codeine metabolism, as measured by morphine and codeine levels

in brain and plasma after codeine injection.

c) In vitro CYP2D activity, as measured by morphine or dextrorphan

formation by brain membranes and liver microsomes incubated with

codeine or dextromethorphan.

d) Baseline nociception, as measured by TFL.

e) Analgesia after morphine injection, as measured by TFL.

39

Section 2: Materials and Methods

Animals

Male adult Wistar rats (250–300 g; Charles River, St-Constant, QC, Canada) were

kept in pairs or triplets under a 12 h artificial light/dark cycle (lights on at 6:00 AM).

Rats were handled, towel-restrained, and placed on the tail-flick meter daily to

acclimate them to testing procedures. All procedures were approved by the Animal

Care Committee at the University of Toronto.

Drug treatment

Propranolol hydrochloride (Sigma-Aldrich), a CYP2D mechanism-based inhibitor

(MBI) (Masubuchi, Narimatsu et al. 1994), was dissolved in artificial cerebrospinal

fluid (ACSF; 126 mM NaCl, 2.68 mM KCl, 1 mM Na2HPO4, 0.88 mM MgSO4, 22 mM

NaHCO3, 1.45 mM CaCl2, 11 mM D-glucose, pH 7.4), and 20-40 μg of the base was

injected intracerebroventricularly (i.c.v.) in a 1-4 μl volume. Propafenone

hydrochloride (Sigma-Aldrich), a CYP2D competitive inhibitor (Xu, Aasmundstad et

al. 1995), was dissolved in a 20% w/v solution of 2-hydroxypropyl-β-cyclodextrin

(cyclodextrin; Sigma-Aldrich) in water, and 40 μg of the base was given i.c.v. in a 4

µl volume. These two inhibitors were chosen for a variety of reasons as follows. They

inhibit CYP2D through different mechanisms, propranolol through mechanism-

based inhibition and propafenone through competitive inhibition (Kroemer, Fischer

et al. 1991; Masubuchi, Narimatsu et al. 1994). They selectively inhibit CYP2D

(Masubuchi, Fujita et al. 1991, Turpeinen, Korhonen et al. 2006, McGinnity, Waters et

40

al. 2008). They have different pharmacological actions and neither is expected to

have any central effects on its own that would influence analgesia or nociception

(Dukes and Vaughan Williams 1984, Komura and Iwaki 2005). The inhibitor doses

were chosen based on pilot studies showing that these doses did not inhibit hepatic

CYP2D (whereas higher doses did). The i.c.v. route of administration was used to

selectively target CYP2D in the brain (and not liver). Also, i.c.v. injection allows the

inhibitors to distribute across the entire brain and inhibit the total CYP2D content in

the brain.

Codeine phosphate (PCCA) was dissolved in sterile saline (0.9% NaCl; pH 7)

and injected subcutaneously (s.c.) at 30 mg base/kg body weight. This dose was

chosen based on a previous study that showed that this dose induced analgesia in

the tail-flick test in all animals tested over a 2 h time period (Cleary, Mikus et al.

1994). The s.c. route of administration was used because it avoids first-pass

metabolism by the liver, thus enhancing our ability to examine the potential role of

brain CYP2D-mediated codeine metabolism. Morphine sulfate (PCCA) was

dissolved in sterile distilled water and injected s.c. at 5 mg base/kg body weight.

This dose was chosen as it is the dose that produces equivalent analgesia in the tail-

flick test as the 30 mg/kg codeine dose used (Lewis, Sherman et al. 1981).

Inhibition of rat brain CYP2D via i.c.v. injection of CYP2D inhibitor

Rats were anesthetized with isoflurane and placed in a stereotaxic frame. In rats that

were to receive propranolol and be used for in vivo codeine metabolism

experiments or in vitro CYP2D activity experiments, rats received an i.c.v. injection

into the right lateral ventricle (Bregma coordinates: dorsal-ventral, -3.6; anterior-

41

posterior, -0.9; lateral, -1.4) (Paxinos and Watson 1986) of either 20 μg propranolol

(CYP2D MBI) dissolved in 1 μl ACSF, or 1 μl ACSF (vehicle control). All i.c.v.

injections were made over 120 sec, and the Hamilton syringe was left in place for

120 sec after injection.

In all rats that were to receive propafenone, as well as in rats that were to

receive propranolol and undergo tail-flick testing, i.c.v. cannulation was required to

allow for i.c.v. inhibitor injection immediately before injection of opioid drug (as in

the case of propafenone) or to allow for more than one i.c.v. injection per rat (as in

the case for tail-flick testing). Intracerebroventricular cannulas were inserted into

the right lateral ventricle (same coordinates as above). After a one week recovery

period, rats received an i.c.v. injection into the cannula of 20 or 40 μg propranolol

(CYP2D MBI) dissolved in 4 μl ACSF, 4 μl ACSF (vehicle control), 40 μg propafenone

(CYP2D competitive inhibitor) dissolved in 4 μl cyclodextrin, or 4 μl cyclodextrin

(vehicle control). All i.c.v. injections were made over 60 sec, and the injector was

left in place in the cannula for 60 sec after injection. To confirm the patency and

correct placement of the cannulas in the lateral ventricle, dipsogenic response to

i.c.v. administration of angiotensin II (Sigma-Aldrich) was tested before each i.c.v.

inhibitor injection (Vento and Daniels 2010).

Effect of i.c.v. CYP2D inhibitor injection on codeine-induced analgesia

In drug-naïve rats, TFL was measured three times in each rat, and the mean of the

three TFLs was used as that rat’s baseline TFL. All TFLs were taken by restraining the

rat with a towel and positioning it on the tail-flick meter so that the heat beam was

~2-3 cm from the end of its tail. Heat intensity was adjusted to produce baseline TFLs

42

of ~2-4 sec, and the same heat intensity was used for all rats for all TFL

measurements. Codeine was injected at 30 mg/kg s.c. 24 h after i.c.v. injection of a

CYP2D MBI (20 μg propranolol) or vehicle control (ACSF), or 5 min after i.c.v.

injection of a CYP2D competitive inhibitor (40 μg propafenone) or vehicle control

(cyclodextrin). The reason for the difference in timing is that a MBI results in a longer

lasting inhibition and can be given ahead of time before the opioid test drug, as

opposed to a competitive inhibitor which has to be given during the same time as

the test drug. This is because MBIs result in permanent loss of enzyme function that

can only be restored by synthesis of new enzyme (Bertelsen, Venkatakrishnan et al.

2003, Van, Heydari et al. 2006). Thus, an additional advantage of using propranolol is

that it is cleared from the body before codeine is given and therefore not expected

to have an effect on analgesia.

TFL was measured for 2 h after codeine injection. This length of time was

chosen because, while analgesia from codeine can last longer than this in rats

(Cleary, Mikus et al. 2004), we are interested in the earlier period after codeine

administration during which brain CYP2D is expected to play a larger role in

analgesia. Also, previous studies have shown this 2 h period to include the rise in the

analgesia time curve during the tail-flick test, as well as the plateau (Cleary, Mikus et

al. 2004, Lewis, Sherman et al. 1981). A cut-off of 10 sec was used to avoid damaging

the tail skin. Each rat’s baseline TFL was subtracted from its TFLs after opioid

injection, and these are the values used in the results. A within-animal design was

used in which, after a 2 week washout period, rats were crossed over (i.c.v. inhibitor

vs. vehicle) and retested with codeine. Thus, each rat acted as its own control, which

43

accounts for possible genetic variation between rats. In both phases, half the rats

received inhibitor and the other half received vehicle.

As a control, TFL was measured 24 h after 40 μg i.c.v. propranolol and 5 min

after 40 μg i.c.v. propafenone injection in the absence of codeine, and this was

compared with baseline TFL. This dose of propranolol was chosen to confirm that,

even at higher doses than used before codeine, propranolol would not have any

effect on nociception on its own. As another control, rats received morphine at 5

mg/kg, s.c. (Lewis, Sherman et al. 1981) 24 h after i.c.v. injection of 40 μg

propranolol or ACSF, or 5 min after i.c.v. injection of 40 μg propafenone or

cyclodextrin. This dose of propranolol was chosen to verify that, even at higher

doses than used before codeine, propranolol would not have any effect on

morphine’s analgesic actions. TFL was measured for 2 h after morphine injection. A

within-animal design was used in which after a 2 week washout period, rats were

crossed over (inhibitor vs. vehicle) and retested with morphine. In this way, each rat

could act as its own control. In both phases, half the rats received inhibitor and the

other half received vehicle.

Measurement of codeine and morphine levels in brain and plasma

Rats received codeine (30 mg/kg, s.c.) 24 h after propranolol (20 μg, i.c.v.) or ACSF

injection, or 5 min after propafenone (40 μg, i.c.v.) or cyclodextrin injection.

Propranolol-treated rats were decapitated at 30, 60 or 90 min after codeine injection,

and propafenone-treated rats were decapitated at 30 min after codeine injection.

Trunk blood and brains were collected, and brains were halved into hemispheres,

and the right hemisphere was further split into anterior and posterior portions. Blood

44

was centrifuged at 5000 g for 10 min, and supernatants were kept. Based on previous

methods (Kudo, Ishida et al. 2006), half brains were homogenized 1:3 (w/v) in 0.01 M

HCl, centrifuged at 5000 g for 10 min, and supernatants were kept. Morphine and

codeine concentrations in plasma and brain samples were measured using HPLC as

described below (Freiermuth and Plasse 1997, He, Shay et al. 1998).

In vitro codeine and dextromethorphan oxidation in brain and liver

Brain and hepatic CYP2D activity were measured using in vitro oxidation of codeine

as the drug of interest as well as dextromethorphan which is a typical CYP2D probe

drug (Von Moltke, Greenblatt et al. 1998, Frank, Jaehde et al. 2007). Rat brain

membranes were prepared as previously described (Tyndale, Li et al. 1999) from

rats that received an i.c.v. injection of MBI (20 μg propranolol) or vehicle control

(ACSF) 24 h prior to sacrifice. Briefly, immediately after sacrifice, each whole brain

was homogenized in 10 ml ACSF (approximately 1:5 (w:v)), centrifuged at 3000 g for

5 min, and the supernatant was kept. The pellet was resuspended in 10 ml ACSF and

centrifuged at 3000 g for 5 min, and the supernatant was combined with the initial

supernatant. The combined supernatant was centrifuged at 100,000 g for 60 min, and

the pellet was resuspended in ACSF. Protein concentration was measured using the

Bradford assay with a Bio-Rad Protein Assay kit (Bio-Rad Laboratories, Mississauga,

Canada). Because brain CYPs are more labile than their hepatic forms and their

activity is reduced by freezer storage (Tyndale, Li et al. 1999, Voirol, Jonzier-Perey

et al. 2000), membranes were prepared immediately after sacrifice and freshly

prepared (i.e., never frozen) membranes were used for all incubations.

45

For brain codeine oxidation, freshly prepared membranes (6 mg protein/ml)

were incubated with 500 μM (10 x Km (Mikus, Somogyi et al. 1991)) codeine, 5 mM

MgCl2 and 1 mM NADPH in ACSF (pH 7.4) for 120 min at 37 °C and 95% O2/5% CO2

in a final volume of 1 ml. The reaction was stopped using 200 μl bicarbonate buffer

solution (1 M, pH 9.7).

For brain dextromethorphan oxidation, as before (Tyndale, Li et al. 1999),

freshly prepared membranes (3 mg protein/ml) were incubated with 25 μM (10 x

Km) dextromethorphan and 1 mM NADPH in ACSF (pH 7.4) for 120 min at 37 °C and

95% O2/5% CO2 in a final volume of 1 ml. The reaction was stopped using 200 μl

bicarbonate buffer solution (1 M, pH 9.7). Linearity of dextrorphan formation at this

protein concentration and incubation time was demonstrated previously (Tyndale, Li

et al. 1999).

Rat liver microsomes were prepared as previously described (Siu,

Wildenauer et al. 2006) from rats that had received an i.c.v. injection of MBI (20 μg

propranolol) or vehicle control (ACSF) 24 h prior to sacrifice. Briefly, liver tissue was

homogenized 1:4 (w/v) in 1.15% KCl, centrifuged at 9000 g for 20 min, and the

supernatants were further centrifuged at 100,000 g for 60 min. The pellets were

resuspended in KCl. Protein concentrations were measured using the Bradford assay

using a Bio-Rad Protein Assay kit (Bio-Rad Laboratories, Mississauga, Canada).

For hepatic codeine oxidation, as before (Xu, Aasmundstad et al. 1997),

microsomes (0.25 mg protein/ml) were incubated with 500 μM codeine (10 x Km

(Mikus, Somogyi et al. 1991)), 5 mM MgCl2 and 1 mM NADPH in 100 mM

NaH2PO4*H2O buffer (pH 7.4) for 20 min at 37 °C in a final volume of 0.5 ml. The

reaction was stopped using 100 μl acetonitrile. Linearity of morphine formation at

46

this protein concentration and incubation time was demonstrated previously (Xu,

Aasmundstad et al. 1997).

For hepatic dextromethorphan oxidation, as before (Kerry, Somogyi et al.

1993, Vuppugalla and Mehvar 2005), microsomes (0.5 mg protein/ml) were

incubated with 25 μM dextromethorphan (10 x Km) and 1 mM NADPH in 100 mM

NaH2PO4*H2O buffer (pH 7.4) for 5 min at 37 °C in a final volume of 0.5 ml. The

reaction was stopped using 100 μl acetonitrile. Linearity of dextrorphan formation at

this protein concentration and incubation time was demonstrated previously (Kerry,

Somogyi et al. 1993, Vuppugalla and Mehvar 2005).

For all liver and brain incubations, controls were used in which there was one

incubation mixture without substrate and one incubation mixture without protein.

Substrate and metabolite levels after incubation were measured using HPLC.

HPLC

HPLC methods were based on modifications to previous techniques for measuring

codeine and morphine (Freiermuth and Plasse 1997, He, Shay et al. 1998), and

dextromethorphan and dextrorphan (Hendrickson, Gurley et al. 2003, Flores-Perez,

Flores-Perez et al. 2004).

Sample preparation

For all brain and plasma samples containing codeine, the internal standard 2-

benzoxazolinone (1 μg) was added. Extractions were performed using Bond Elut C18

(3 ml, 200 mg) solid phase extraction cartridges (Agilent, USA). Cartridges were first

conditioned with methanol (2 ml), followed by Milli-Q water (2 ml). The sample was

47

then applied to the cartridge and washed with 2 x 1 ml of Milli-Q water followed by 2

x 0.5 ml of a 40% acetonitrile solution in water. The drugs were eluted with 3 x 0.5 ml

of a 0.05 M HCl solution at 10% in acetonitrile. The eluate was evaporated to dryness

at 37ºC under a nitrogen stream and reconstituted with 110 μl of mobile phase and

100 μl of the solution was injected into the HPLC system.

For all brain samples containing dextromethorphan, the internal standard 2-

benzoxazolinone (50 ng) was added with 5 ml hexane-butanol (95:5 v/v). The

mixture was vortexed for 10 sec, mechanically shaken for 10 min and centrifuged at

1400 g for 10 min. The organic layer was then transferred to a 10 ml tube and

evaporated to dryness at 37ºC under a nitrogen stream. The residue was redissolved

in 110 μl of mobile phase and 100 μl of the solution was injected into the HPLC

system.

For all liver samples containing codeine or dextromethorphan, based on

previous methods (Mikus, Somogyi et al. 1991, Kerry, Somogyi et al. 1993), after

stopping the incubation reaction, samples were centrifuged at 13,250 g for 10 min

and 90 μl of the supernantant was directly injected into the HPLC system.

Chromatographic conditions

All samples containing codeine were analyzed by HPLC with ultraviolet detection

(HPLC-UV) (Agilent 1200 Separation Module). The limits of quantification were 25

ng/ml for morphine and 250 ng/ml for codeine from plasma as well as liver and

brain in vitro incubation mixtures, and 5 ng/g for morphine and 50 ng/g for codeine

assessed from brain homogenates. For plasma and the liver and brain in vitro

incubation mixtures, the assay was linear from 25 to 500 ng/ml for morphine and 250

48

to 2000 ng/ml for codeine, with an extraction efficiency of 76.9% for morphine and

83.5% for codeine. For brain homogenates, the assay was linear from 5 to 500 ng/g

for morphine and 50 to 1000 ng/g for codeine, with an extraction efficiency of 82.7%

for morphine and 90.1% for codeine. The HPLC-UV system was set for detection at

214 nm and morphine, codeine and internal standard were separated on an Agilent

ZORBAX SB-C18 Column (250 x 4.6 mm I.D.; particle size, 5 um). The mobile phase

used was methanol - 0.05M phosphate buffer, pH 5.8 (29.3/70.7, v/v) and the flow

rate was 1 ml/min. The retention times were 4.1 min for morphine, 9.4 min for

codeine, and 16.7 min for the internal standard.

All samples containing dextromethorphan were analyzed by HPLC with

fluorescence detection (HPLC-FLD) (Agilent 1200 Separation Module), with a limit of

quantification of 5 ng/ml for both dextromethorphan and dextrorphan. The assay

was linear from 5 to 150 ng/ml with an extraction efficiency of 80.3% for

dextromethorphan and 72.8% for dextrorphan. The HPLC-FLD system was set for

detection at 230 excitation wavelength and 330 nm emission wavelength, and

dextromethorphan, dextrorphan and internal standard were separated on an

Agilent ZORBAX Bonus-RP Column (150 x 4.6 mm I.D.; particle size, 5 um). The

mobile phase used was acetonitrile/potassium phosphate buffer (22.3:77.7 v/v, pH

5.07) containing 34 mM potassium phosphate monobasic, 34 mM citric acid, 3.3 mM

heptane sulfonic acid and 0.5% triethylamine, and the flow rate was 1.2 ml/min. The

retention times were 5.1 min for dextrorphan, 6.6 min for internal standard, and 12.5

min for dextromethorphan.

49

Statistical analyses

Paired t-tests (because of the within-animal design) were used for comparisons of

TFLs and AUCs following codeine or morphine injection after inhibitor treatment

versus after vehicle treatment, of TFLs and AUCs between the two phases of codeine

or morphine injection, of morphine concentrations and morphine to codeine ratios in

the anterior versus the posterior part of the brain, and of baseline TFLs versus TFLs

after inhibitor treatment. Unpaired t-tests (because of the between-animal design)

were used for comparisons of TFLs and AUCs after codeine injection versus after

morphine injection, of morphine concentrations, morphine to codeine ratios,

morphine to total drug ratios, codeine concentrations and total drug concentrations

in brain and plasma between inhibitor- and vehicle-treated rats, and of velocities

from brain membrane and liver microsome in vitro metabolism studies between

inhibitor- and vehicle-treated rats.

50

Section 3: Results

3.1 Inhibition of brain CYP2D reduced codeine-induced

analgesia

Codeine is metabolized by CYP2D to the active morphine metabolite which confers

analgesia. To test the behavioural effects of altering brain CYP2D activity, we

examined whether inhibiting brain CYP2D would reduce analgesia following

codeine administration. We used two different CYP2D inhibitors: propranolol – a

mechanism-based inhibitor (MBI) and propafenone – a competitive inhibitor.

Rats were injected with either 20 μg i.c.v. propranolol, 40 μg i.c.v.

propafenone, or their respective vehicles. Codeine was injected at 30 mg/kg s.c. 24

h after propranolol injection and 5 min after propafenone injection. Tail-flick latency

(TFL) was measured for 2 h after codeine injection. After a two week washout period,

rats were crossed over so that those that had received inhibitor in the first phase

received vehicle in the second phase, and vice versa (the same number of rats

received each treatment in each phase). They were then retested with codeine.

Because each rat received both inhibitor and vehicle, TFL after inhibitor treatment

could be compared with TFL after vehicle treatment within the same animal, with

each rat acting as its own control.

Compared to vehicle treatment, propranolol treatment resulted in

significantly shorter TFL at 15 (p<0.02; n=4 total), 20 (p<0.02; n=16 total), 30

(p<0.004; n=16 total), and 40 (p<0.005; n=16 total) min after codeine injection

51

(Figure 3.a). Compared to vehicle treatment, propafenone treatment resulted in

significantly shorter TFL at 20 (p<0.03; n=12 total), 25 (p<0.05; n=7 total), 30 (p<0.03;

n=12 total), and 40 (p<0.04; n=12 total) min after codeine injection (Figure 3.b). The

n-values are not the same at all each time point because two different sets of rats

were used, and TFL was not measured at the 5, 15 and 25 min time points in the first

set of rats. In the second set of experiments TFL was measured at these additional

time points.

52

3.a)

b)

Figure 3. Inhibition of brain CYP2D reduced codeine-induced analgesia. (a)

Compared to vehicle treatment, propranolol treatment resulted in significantly

shorter tail-flick latency (TFL) at 15 (p<0.02), 20 (p<0.02), 30 (p<0.004) and 40

(p<0.005) min after codeine injection (n=4-16 total/time point). (b) Compared to

vehicle treatment, propafenone treatment resulted in significantly shorter TFL at 20

(p<0.03), 25 (p<0.05), 30 (p<0.03), and 40 (p<0.04) min after codeine injection (n=7-

12 total/time point). *p<0.05, **p<0.01.

-1

1

3

5

7

0 30 60 90 120

TFL

(sec

, mea

n +

SEM

)

Time after codeine injection (min)

Vehicle

Propranolol

* *

** **

-1

1

3

5

7

0 30 60 90 120

TFL

(sec

, mea

n +

SEM

)

Time after codeine injection (min)

Vehicle

Propafenone

* *

* *

53

3.2 Inhibition of brain CYP2D lowered codeine-induced area

under the analgesia time curve

Another way to assess the behavioural effects of brain CYP2D inhibition is to

investigate the area under the analgesia time curve (AUC), which is a measure of

total analgesia over the time period examined. We chose to examine the period up

to 30 min after codeine injection as this was hypothesized based on Chen’s data

(1990) to be the time period during which inhibitor treatment would result in lower

analgesia. However, based on our tail-flick data, analgesia appeared to be lower

after inhibitor treatment up to 60 min after codeine injection, so we also analyzed

analgesia AUC for 30-60 min and 0-60 min after codeine injection. We also measured

AUC for 60-120 min after codeine injection during which no difference in analgesia

after inhibitor treatment was expected. Additionally, we measured AUC for 0-120

min after codeine injection to investigate the effect of inhibitor treatment on

analgesia over the total time period examined.

Compared to vehicle treatment, 20 μg i.c.v. propranolol treatment resulted in

significantly lower AUC for 0-30, 30-60 and 0-60 min after 30 mg/kg s.c. codeine

injection (p<0.01, p<0.05, p<0.02, respectively; Figure 4.a-c), but not for 60-120

min or 0-120 min after codeine injection (p>0.8, p>0.5, respectively; Figure 5.a,b)

(n=16 total). Compared to vehicle treatment, 40 μg i.c.v. propafenone treatment

resulted in significantly lower AUC for 0-30, 30-60 and 0-60 min after codeine

injection (p<0.01, p<0.05, p<0.02, respectively; Figure 6.a-c), but not for 60-120

min or 0-120 min after codeine injection (p>0.2, p>0.07, respectively; Figure 7.a,b)

(n=12 total).

54

4.a) 0-30 min b) 30-60 min

c) 0-60 min

Figure 4. Inhibiting brain CYP2D with propranolol lowered codeine-induced

area under the analgesia time curve between 0-60 min after codeine injection.

Compared to vehicle treatment, propranolol treatment resulted in significantly

lower area under the analgesia time curve (AUC) for (a) 0-30 min (p<0.01), (b) 30-60

min (p<0.05) and (c) 0-60 min (p<0.02) after codeine injection (n=16 total). *p<0.05,

**p<0.01.

0

110

220

330

AU

C 0

-60

m

in*s

ec, m

ean

+ S

EM

Vehicle Propranolol

*

0

40

80

120

AU

C 0

-30

m

in*s

ec, m

ean

+ S

EM

Vehicle Propranolol

**

0

70

140

210

AU

C 3

0-6

0

min

*se

c, m

ean

+ S

EM

Vehicle Propranolol

*

55

5.a) 60-120 min b) 0-120 min

Figure 5. Inhibiting brain CYP2D with propranolol did not lower codeine-

induced area under the analgesia time curve at 60-120 min or 0-120 min after

codeine injection. Compared to vehicle treatment, propranolol treatment did not

result in significantly different area under the analgesia time curve (AUC) for (a) 60-

120 min (p>0.8) or (b) 0-120 min (p>0.5) after codeine injection (n=16 total).

0

120

240

360A

UC

60

-12

0

min

*sec

, mea

n +

SEM

Vehicle Propranolol 0

230

460

690

AU

C 0

-12

0

min

*sec

, mea

n +

SEM

Vehicle Propranolol

56

6.a) 0-30 min b) 30-60 min

c) 0-60 min

Figure 6. Inhibiting brain CYP2D with propafenone lowered codeine-induced

area under the analgesia time curve between 0-60 min after codeine injection.

Compared to vehicle treatment, propafenone treatment resulted in significantly

lower area under the analgesia time curve (AUC) for (a) 0-30 min (p<0.01), (b) 30-60

min (p<0.04) and (c) 0-60 min (p<0.02) after codeine injection (n=12 total). *p<0.05,

**p<0.01.

0

110

220

330

AU

C 0

-60

m

in*s

ec, m

ean

+ S

EM

Vehicle Propafenone

*

0

40

80

120

AU

C 0

-30

m

in*s

ec, m

ean

+ S

EM

Vehicle Propafenone

**

0

70

140

210

AU

C 3

0-6

0

min

*sec

, mea

n +

SEM

Vehicle Propafenone

*

57

7.a) 60-120 min b) 0-120 min

Figure 7. Inhibiting brain CYP2D with propafenone did not lower codeine-

induced area under the analgesia time curve at 60-120 min or 0-120 min after

codeine injection. Compared to vehicle treatment, propafenone treatment did not

result in significantly different area under the analgesia time curve (AUC) for (a) 60-

120 min (p>0.2) or (b) 0-120 min (p>0.07) after codeine injection (n=12 total).

0

120

240

360A

UC

60

-12

0

min

*sec

, mea

n +

SEM

Vehicle Propafenone 0

230

460

690

AU

C 0

-12

0

min

*sec

, mea

n +

SEM

Vehicle Propafenone

58

3.3 Inhibiting brain CYP2D did not affect baseline tail-flick

latency

As a control we assessed whether either of the two CYP2D inhibitors on their own

would have an effect on nociception and cause tail-flick latency (TFL) to deviate from

baseline. In drug-naïve rats, TFL was measured three times in each rat within a 15

min period. The mean of the three TFLs in each rat was used as that rat’s own

baseline TFL. Baseline TFLs were similar in all the rats (mean+SEM=3.09+0.09 sec,

range=2.6-3.64 sec) and were also similar within each rat (within-animal SEM

ranged from 0.009-0.14). Immediately after measuring baseline TFL, rats were

injected with either 40 μg i.c.v. propranolol or 40 μg i.c.v. propafenone.

At 24 h after propranolol injection or 5 min after propafenone injection, TFLs

were measured three times in each rat within a 15 min period. The mean of the three

TFLs in each rat was used as that rat’s own TFL after propranolol or propafenone

treatment. There was no significant difference between baseline TFL and TFL after

propranolol treatment (p>0.3; n=10 total; Figure 8.a). There was also no significant

difference between baseline TFL and TFL after propafenone treatment (p>0.1; n=8

total; Figure 8.b).

59

8.a) b)

Figure 8. Inhibiting brain CYP2D did not affect baseline tail-flick latency.

(a) There was no significant difference between baseline tail-flick latency (TFL) and

TFL after propranolol treatment (p>0.3, n=10 total). (b) There was no significant

difference between baseline TFL and TFL after propafenone treatment (p>0.1, n=8

total).

0

1

2

3

4TF

L (s

ec, m

ean

+ S

EM)

Baseline Propranolol 0

1

2

3

4

TFL

(sec

, mea

n +

SEM

)

Baseline Propafenone

60

3.4 Inhibiting brain CYP2D did not affect morphine-induced

analgesia

As another control, we assessed whether either of the two CYP2D inhibitors would

affect analgesia following morphine injection. Since morphine is the active analgesic

compound, and it is not metabolized by CYP2D, its analgesic effect should not be

altered by changes in CYP2D activity.

Rats were injected with either 40 μg i.c.v. propranolol, 40 μg i.c.v.

propafenone, or their respective vehicles. Morphine was injected at 5 mg/kg s.c. 24

h after propranolol injection and 5 min after propafenone injection. Tail-flick latency

(TFL) was measured for 2 h after morphine injection. After a two week washout

period, rats were crossed over so that those that had received inhibitor in the first

phase received vehicle in the second phase, and vice versa (the same number of

rats received each treatment in each phase). They were then retested with

morphine. Because each rat received both inhibitor and vehicle, TFL after inhibitor-

treatment could be compared with TFL after vehicle treatment within the same

animal, with each rat acting as its own control.

Compared to vehicle treatment, propranolol treatment did not result in

significantly different TFL at any time point after morphine injection (p>0.08 at all

time points; n=12 total; Figure 9.a). Compared to vehicle treatment, propafenone

treatment did not result in significantly different TFL at any time point after morphine

injection (p>0.1 at all time points; n=6 total; Figure 9.b).

61

9.a)

b)

Figure 9. Inhibiting brain CYP2D did not affect morphine-induced analgesia.

(a) Compared to vehicle treatment, propranolol treatment did not result in

significantly different TFL after morphine injection (p>0.08 at all time points, n=12

total). (b) Compared to vehicle treatment, propafenone treatment did not result in

significantly different TFL after morphine injection (p>0.1 at all time points, n=6

total).

-1

1

3

5

7

0 30 60 90 120

TFL

(sec

, mea

n +

SEM

)

Time after morphine injection (min)

Vehicle

Propranolol

-1

1

3

5

7

0 30 60 90 120

TFL

(sec

, mea

n +

SEM

)

Time after morphine injection (min)

Vehicle

Propafenone

62

3.5 Inhibiting brain CYP2D did not alter morphine-induced

area under the analgesia time curve

As another way to assess whether either of the two CYP2D inhibitors would affect

analgesia following morphine injection, we investigated the area under the

analgesia time curve (AUC) for morphine. We chose to examine the AUCs for the

same periods of time following morphine injection as we did for codeine, so we

could compare the AUCs of these two opioids.

Compared to vehicle treatment, 40 μg i.c.v. propranolol treatment did not

result in significantly different AUC for 0-30, 30-60, 0-60, 60-120 or 0-120 min after 5

mg/kg s.c. morphine injection (p>0.4 for all time periods; n=12 total; Figure 10.a-

e). Compared to vehicle treatment, 40 μg i.c.v. propafenone treatment also did not

result in significantly different AUC for 0-30, 30-60, 0-60, 60-120 or 0-120 min after

morphine injection (p>0.2 for all time periods; n=6 total; Figure 11.a-e).

63

10.a) 0-30 min b) 30-60 min

c) 0-60 min d) 60-120 min

e) 0-120 min

Figure 10. Inhibiting brain CYP2D with propranolol did not alter morphine-

induced area under the analgesia time curve. Compared to vehicle treatment,

propranolol treatment did not result in significantly different AUC for (a) 0-30 min,

(b) 30-60, (c) 0-60 min, (d) 60-120 min or (e) 0-120 min after morphine injection

(p>0.4 for all time periods, n=12 total).

0

230

460

690

AU

C 0

-12

0

min

*sec

, mea

n +

SEM

Vehicle Propranolol

0

40

80

120

AU

C 0

-30

m

in*s

ec, m

ean

+ S

EM

Vehicle Propranolol 0

70

140

210

AU

C 3

0-6

0

min

*sec

, mea

n +

SEM

Vehicle Propranolol

0

110

220

330

AU

C 0

-60

m

in*s

ec, m

ean

+ S

EM

Vehicle Propranolol 0

120

240

360

AU

C 6

0-1

20

m

in*s

ec, m

ean

+ S

EM

Vehicle Propranolol

64

11.a) 0-30 min b) 30-60 min

c) 0-60 min d) 60-120 min

e) 0-120 min

Figure 11. Inhibiting brain CYP2D with propafenone did not alter morphine-

induced area under the analgesia time curve. Compared to vehicle treatment,

propafenone treatment did not result in significantly different AUC for (a) 0-30 min,

(b) 30-60, (c) 0-60 min, (d) 60-120 min or (e) 0-120 min after morphine injection

(p>0.2 for all time periods, n=6 total).

0

230

460

690

AU

C 0

-12

0

min

*sec

, mea

n +

SEM

Vehicle Propafenone

0

40

80

120

AU

C 0

-30

m

in*s

ec, m

ean

+ S

EM

Vehicle Propafenone 0

70

140

210

AU

C 3

0-6

0

min

*sec

, mea

n +

SEM

Vehicle Propafenone

0

110

220

330

AU

C 0

-60

m

in*s

ec, m

ean

+ S

EM

Vehicle Propafenone 0

120

240

360

AU

C 6

0-1

20

m

in*s

ec, m

ean

+ S

EM

Vehicle Propafenone

65

3.6 There was no tolerance to the analgesic effects of codeine

or morphine

We compared the tail-flick latencies (TFLs) and areas under the analgesia time

curves (AUCs) between the two phases of codeine or morphine treatment (i.e., the

first phase of opioid testing and the crossover phase two weeks later) to confirm that

the rats did not develop tolerance to either of the opioids within this testing

schedule. In rats treated with propranolol or vehicle and tested with codeine, there

was no significant difference between the two phases in TFL (p>0.1 at all time points;

Figure 12.a) or AUC (p>0.6 at all time periods; Figure 12.b-f) (n=4-16 total). In rats

treated with propafenone or vehicle and tested with codeine, there was no

significant difference between the two phases in TFL (p>0.1 at all time points; Figure

13.a) or AUC (p>0.2 at all time periods; Figure 13.b-f) (n=7-12 total). In rats treated

with propranolol or vehicle and tested with morphine, there was no significant

difference between the two phases in TFL (p>0.08 at all time points; Figure 14.a) or

AUC (p>0.4 at all time periods; Figure 14.b-f) (n=12 total). In rats treated with

propafenone or vehicle and tested with morphine, there was no significant

difference between the two phases in TFL (p>0.08 at all time points; Figure 15.a) or

AUC (p>0.1 at all time periods; Figure 15.b-f) (n=6 total).

66

12.a) b) 0-30 min

c) 30-60 min d) 0-60 min

e) 60-120 min f) 0-120 min

Figure 12. Rats treated with propranolol or vehicle did not develop tolerance to

codeine. In rats treated with propranolol or vehicle and tested with codeine, there

was no significant difference between the two phases in (a) tail-flick latency (TFL)

(p>0.1 at all time points) or (b-f) area under the analgesia time curve (AUC) (p>0.6

at all time periods) (n=4-16 total).

-1

1

3

5

7

0 30 60 90 120TFL

(sec

, mea

n +

SEM

)

Time after codeine injection (min)

Phase 1

Phase 2

0

40

80

120

AU

C 0

-30

m

in*s

ec, m

ean

+ S

EM

Phase 1 Phase 2

0

70

140

210

AU

C 3

0-6

0

min

*sec

, mea

n +

SEM

Phase 1 Phase 2 0

110

220

330

AU

C 0

-60

m

in*s

ec, m

ean

+ S

EM

Phase 1 Phase 2

0

120

240

360

AU

C 6

0-1

20

m

in*s

ec, m

ean

+ S

EM

Phase 1 Phase 2 0

230

460

690

AU

C 0

-12

0

min

*se

c, m

ean

+ S

EM

Phase 1 Phase 2

67

13.a) b) 0-30 min

c) 30-60 min d) 0-60 min

e) 60-120 min f) 0-120 min

Figure 13. Rats treated with propafenone or vehicle did not develop tolerance to

codeine. In rats treated with propafenone or vehicle and tested with codeine, there

was no significant difference between the two phases (a) in tail-flick latency (TFL)

(p>0.1 at all time points) or (b-f) area under the analgesia time curve (AUC) (p>0.2

at all time periods) (n=7-12 total).

-1

1

3

5

7

0 30 60 90 120TFL

(sec

, mea

n +

SEM

)

Time after codeine injection (min)

Phase 1

Phase 2

0

40

80

120

AU

C 0

-30

m

in*s

ec, m

ean

+ S

EM

Phase 1 Phase 2

0

70

140

210

AU

C 3

0-6

0

min

*sec

, mea

n +

SEM

Phase 1 Phase 2 0

110

220

330

AU

C 0

-60

m

in*s

ec, m

ean

+ S

EM

Phase 1 Phase 2

0

120

240

360

AU

C 6

0-1

20

m

in*s

ec, m

ean

+ S

EM

Phase 1 Phase 2 0

230

460

690

AU

C 0

-12

0

min

*sec

, mea

n +

SEM

Phase 1 Phase 2

68

14.a) b) 0-30 min

c) 30-60 min d) 0-60 min

e) 60-120 min f) 0-120 min

Figure 14. Rats treated with propranolol or vehicle did not develop tolerance to

morphine. In rats treated with propranolol or vehicle and tested with morphine,

there was no significant difference between the two phases in (a) tail-flick latency

(TFL) (p>0.08 at all time points) or (b-f) area under the analgesia time curve (AUC)

(p>0.4 at all time periods) (n=12 total).

-1

1

3

5

7

0 30 60 90 120TFL

(sec

, mea

n +

SEM

)

Time after morphine injection (min)

Phase 1

Phase 2

0

40

80

120

AU

C 0

-30

m

in*s

ec, m

ean

+ S

EM

Phase 1 Phase 2

0

70

140

210

AU

C 3

0-6

0

min

*sec

, mea

n +

SEM

Phase 1 Phase 2 0

110

220

330

AU

C 0

-60

m

in*s

ec, m

ean

+ S

EM

Phase 1 Phase 2

0

120

240

360

AU

C 6

0-1

20

m

in*s

ec, m

ean

+ S

EM

Phase 1 Phase 2 0

230

460

690

AU

C 0

-12

0

min

*sec

, mea

n +

SEM

Phase 1 Phase 2

69

15.a) b) 0-30 min

c) 30-60 min c) 0-60 min

e) 60-120 min f) 0-120 min

Figure 15. Rats treated with propafenone or vehicle did not develop tolerance to

morphine. In rats treated with propafenone or vehicle and tested with morphine,

there was no significant difference between the two phases in (a) tail-flick latency

(TFL) (p>0.08 at all time points) or (b-f) area under the analgesia time curve (AUC)

(p>0.1 at all time periods) (n=6 total).

-1

1

3

5

7

0 30 60 90 120TFL

(sec

, mea

n +

SEM

)

Time after morphine injection (min)

Phase 1

Phase 2

0

40

80

120

AU

C 0

-30

m

in*s

ec, m

ean

+ S

EM

Phase 1 Phase 2

0

70

140

210

AU

C 3

0-6

0

min

*sec

, mea

n +

SEM

Phase 1 Phase 2 0

110

220

330

AU

C 0

-60

m

in*s

ec, m

ean

+ S

EM

Phase 1 Phase 2

0

120

240

360

AU

C 6

0-1

20

m

in*s

ec, m

ean

+ S

EM

Phase 1 Phase 2 0

230

460

690

AU

C 0

-12

0

min

*sec

, mea

n +

SEM

Phase 1 Phase 2

70

3.7 Codeine and morphine doses used resulted in similar

levels of analgesia

To validate our choice of codeine and morphine doses, we compared the tail-flick

latencies (TFLs) and areas under the analgesia curves (AUCs) resulting from each

opioid after i.c.v. vehicle treatment to check that they produced equivalent

analgesia. Following ACSF (vehicle for propranolol) treatment, there was no

significant difference in TFL (p>0.3 for all time points; Figure 16.a) or AUC (p>0.1

for all time periods; Figure 16.b-f) after 30 mg/kg s.c. codeine (n=16) compared to

after 5 mg/kg s.c. morphine (n=12). Following cyclodextrin (vehicle for

propafenone) treatment, there was no significant difference in TFL (p>0.1 for all time

points; Figure 17.a) or AUC (p>0.5 for all time periods; Figure 17.b-f) after 30

mg/kg s.c. codeine (n=12) compared to after 5 mg/kg s.c. morphine (n=6).

71

16.a) b) 0-30 min

b) 30-60 min c) 0-60 min

d) 60-120 min e) 0-120 min

Figure 16. Codeine and morphine doses used resulted in similar levels of

analgesia after ACSF (i.c.v. vehicle) treatment. Following ACSF (vehicle for

propranolol) treatment, there was no significant difference in (a) tail-flick latency

(TFL) (p>0.3) or (b-e) area under the analgesia time curve (AUC) (p>0.1) after 30

mg/kg s.c. codeine (n=16) compared to after 5 mg/kg s.c. morphine (n=12).

-1

1

3

5

7

0 30 60 90 120 TFL

(se

c, m

ean

+ S

EM)

Time after opioid injection (min)

Codeine

Morphine

0

40

80

120

AU

C 0

-30

m

in*

sec,

mea

n +

SEM

Codeine Morphine

0

70

140

210

AU

C 3

0-6

0

min

*sec

, mea

n +

SEM

Codeine Morphine 0

110

220

330

AU

C 0

-60

m

in*s

ec, m

ean

+ S

EM

Codeine Morphine

0

120

240

360

AU

C 6

0-1

20

m

in*s

ec, m

ean

+ S

EM

Codeine Morphine 0

230

460

690

AU

C 0

-12

0

min

*sec

, mea

n +

SEM

Codeine Morphine

72

17.a) b) 0-30 min

b) 30-60 min c) 0-60 min

d) 60-120 min e) 0-120 min

Figure 17. Codeine and morphine doses used resulted in similar levels of

analgesia after cyclodextrin (i.c.v. vehicle) treatment. Following cyclodextrin

(vehicle for propafenone) treatment, there was no significant difference in (a) tail-

flick latency (TFL) (p>0.1) or (b-e) area under the analgesia time curve (AUC)

(p>0.5) after 30 mg/kg s.c. codeine (n=12) compared to after 5 mg/kg s.c. morphine

(n=6).

-1

1

3

5

7

0 30 60 90 120 TFL

(se

c, m

ean

+ S

EM)

Time after opioid injection (min)

Codeine

Morphine

0

40

80

120

AU

C 0

-30

m

in*s

ec, m

ean

+ S

EM

Codeine Morphine

0

70

140

210

AU

C 3

0-6

0

min

*sec

, mea

n +

SEM

Codeine Morphine 0

110

220

330

AU

C 0

-60

m

in*

sec,

mea

n +

SEM

Codeine Morphine

0

120

240

360

AU

C 6

0-1

20

m

in*s

ec, m

ean

+ S

EM

Codeine Morphine 0

230

460

690

AU

C 0

-12

0

min

*sec

, mea

n +

SEM

Codeine Morphine

73

3.8 Inhibitor-treated rats had lower morphine levels in the

brain but not plasma at 30 min after codeine injection

To assess the pharmacokinetic effects of the CYP2D inhibitors and to see if these are

consistent with differences in analgesia, we examined morphine and codeine levels

in brain and plasma at 30 min after codeine injection in inhibitor- and vehicle-

treated rats. This time point was chosen because it was when we hypothesized,

based on Chen’s data (1990), that brain morphine levels would be lower in inhibitor-

treated rats. This is also when the largest difference in behaviour (i.e., TFL) after

inhibitor treatment, compared to after vehicle treatment, occurred (Figure 3).

Morphine concentration was measured because morphine is the active metabolite

which confers analgesia. The ratio of morphine concentration to codeine

concentration (i.e., ratio of parent drug to metabolite) was calculated because this is

a measure of metabolism. Codeine concentration was measured to verify that it does

not correlate with analgesia. Total drug concentration (morphine plus codeine) was

calculated to ensure that it was the same after inhibitor treatment compared to after

vehicle treatment. The ratio of morphine concentration to total drug concentration

was calculated to confirm that morphine levels correlate with analgesia. These

parameters were examined in both brain and plasma to check that the inhibitors

were selectively affecting brain and not hepatic metabolism.

Rats were injected with either 20 μg i.c.v. propranolol, 40 μg i.c.v.

propafenone, or their respective vehicles. Codeine was injected at 30 mg/kg s.c. 24

h after propranolol injection and 5 min after propafenone injection. Rats were

sacrificed 30 min after codeine injection.

74

Compared to vehicle-treated rats, propranolol-treated rats had significantly

lower morphine concentrations, morphine to codeine ratios and morphine to total

drug ratios in the brain (p<0.02, p<0.05, p<0.05, respectively) but not in plasma

(p>0.6, p>0.7, p>0.7, respectively) (n=11/group; Figure 18-20). Codeine

concentrations and total drug concentrations were not significantly different

between propranolol- and vehicle-treated rats in brain (p>0.6, p>0.6, respectively)

or plasma (p>0.6, p>0.6, respectively) (n=11/group; Figure 21). Compared to

vehicle-treated rats, propafenone-treated rats also had significantly lower morphine

concentrations, morphine to codeine ratios, and morphine to total drug ratios in the

brain (p<0.006, p<0.03, p<0.03, respectively) but not in plasma (p>0.6, p>0.8,

p>0.8, respectively) (n=8/group; Figure 18-20). Codeine concentrations and total

drug concentrations were not significantly different between propafenone- and

vehicle-treated rats in brain (p>0.7, p>0.7, respectively) or plasma (p>0.9, p>0.9,

respectively) (n=8/group; Figure 22). As well, morphine levels and morphine to

codeine ratios were not significantly different between the anterior and the posterior

portion of the brains of propranolol-treated rats (p>0.6, p>0.4, respectively; n=8

total; Figure 23.a,c) or propafenone-treated rats (p>0.3, p>0.3, respectively; n=4

total; Figure 23.b,d), suggesting that the inhibitors distributed and inhibited CYP2D

throughout the brain.

75

18.a) Brain b) Plasma

c) Brain d) Plasma

Figure 18. Inhibitor-treated rats had lower morphine levels in the brain but not

in plasma at 30 min after codeine injection. Compared to vehicle-treated rats,

propranolol-treated rats had significantly lower morphine concentrations in (a) the

brain (p<0.02) but not in (b) plasma (p>0.6) at 30 min after codeine injection

(n=11/group). Compared to vehicle-treated rats, propafenone-treated rats had

significantly lower morphine concentrations in (c) the brain (p<0.006) but not in (d)

plasma (p>0.6) at 30 min after codeine injection (n=8/group). *p<0.05, **p<0.01.

MOR=morphine.

0

15

30

45

MO

R (

ng

/ml,

mea

n +

SEM

)

Vehicle Propranolol 0

15

30

45

MO

R (

ng

/g, m

ean

+ S

EM)

Vehicle Propranolol

*

0

15

30

45

MO

R (

ng

/ml,

mea

n +

SEM

)

Vehicle Propafenone 0

15

30

45

MO

R (

ng

/g, m

ean

+ S

EM)

Vehicle Propafenone

**

76

19.a) Brain b) Plasma

c) Brain d) Plasma

Figure 19. Inhibitor-treated rats had lower morphine to codeine ratios in the

brain but not in plasma at 30 min after codeine injection. Compared to vehicle-

treated rats, propranolol-treated rats had significantly lower morphine to codeine

ratios in (a) the brain (p<0.05) but not in (b) plasma (p>0.7) at 30 min after codeine

injection (n=11/group). Compared to vehicle-treated rats, propafenone-treated rats

had significantly lower morphine to codeine ratios in (c) the brain (p<0.03) but not in

(d) plasma (p>0.8) at 30 min after codeine injection (n=8/group). *p<0.05.

MOR=morphine, COD=codeine.

0

0.01

0.02

0.03

MO

R/C

OD

(m

ean

+ S

EM)

Vehicle Propranolol

*

0

0.01

0.02

0.03

MO

R/C

OD

(m

ean

+ S

EM)

Vehicle Propranolol

0

0.01

0.02

0.03

MO

R/C

OD

(m

ean

+ S

EM)

Vehicle Propafenone

*

0

0.01

0.02

0.03

MO

R/C

OD

(m

ean

+ S

EM)

Vehicle Propafenone

77

20.a) Brain b) Plasma

c) Brain d) Plasma

Figure 20. Inhibitor-treated rats had lower morphine to total drug ratios in the

brain but not in plasma at 30 min after codeine injection. Compared to vehicle-

treated rats, propranolol-treated rats had significantly lower morphine to total drug

ratios in (a) the brain (p<0.05) but not in (b) plasma (p>0.7) at 30 min after codeine

injection (n=11/group). Compared to vehicle-treated rats, propafenone-treated rats

had significantly lower morphine to total drug ratios in (c) the brain (p<0.03) but not

in (d) plasma (p>0.8) at 30 min after codeine injection (n=8/group). *p<0.05.

MOR=morphine, COD=codeine.

0

0.01

0.02

0.03M

OR

/(M

OR

+CO

D)

m

ean

+ S

EM

Vehicle Propranolol 0

0.01

0.02

0.03

MO

R/(

MO

R+C

OD

) m

ean

+ S

EM

Vehicle Propranolol

0

0.01

0.02

0.03

MO

R/(

MO

R+C

OD

) m

ean

+ S

EM

Vehicle Propafenone 0

0.01

0.02

0.03

MO

R/(

MO

R+C

OD

) m

ean

+ S

EM

Vehicle Propafenone

*

*

78

21.a) Brain b) Plasma

c) Brain d) Plasma

Figure 21. Propranolol-treated rats did not have lower codeine levels or total

drug levels in the brain or in plasma at 30 min after codeine injection. Codeine

concentrations (a,b) and total drug concentrations (c,d) were not significantly

different between propranolol- and vehicle-treated rats in brain or plasma (p>0.6 for

each measurement, n=11/group). COD=codeine, MOR=morphine.

0

1500

3000

4500

CO

D (

ng

/ml,

me

an +

SEM

)

Vehicle Propranolol 0

1500

3000

4500C

OD

(n

g/g

, mea

n +

SEM

)

Vehicle Propranolol

0

1500

3000

4500M

OR

+CO

D (

ng

/ml,

mea

n +

SEM

)

Vehicle Propranolol 0

1500

3000

4500

MO

R+C

OD

(n

g/g

, mea

n +

SEM

)

Vehicle Propranolol

79

22.a) Brain b) Plasma

c) Brain d) Plasma

Figure 22. Propafenone-treated rats did not have lower codeine levels or total

drug levels in the brain or in plasma at 30 min after codeine injection. Codeine

concentrations (a,b) and total drug concentrations (c,d) were not significantly

different between propafenone- and vehicle-treated rats in brain (p>0.7 for both

measurements) or plasma (p>0.9 for both measurements) (n=8/group).

COD=codeine, MOR=morphine.

0

1500

3000

4500

CO

D (

ng

/ml,

mea

n +

SEM

)

Vehicle Propafenone 0

1500

3000

4500

CO

D (

ng

/g, m

ean

+ S

EM)

Vehicle Propafenone

0

1500

3000

4500

MO

R+C

OD

(n

g/m

l, m

ean

+ S

EM)

Vehicle Propafenone 0

1500

3000

4500

MO

R+C

OD

(n

g/g

, mea

n +

SEM

)

Vehicle Propafenone

80

23.a) Propranolol b) Propafenone

c) Propranolol d) Propafenone

Figure 23. Inhibitor-treated rats had similar morphine levels and morphine to

codeine ratios between the anterior and the posterior parts of the brain.

Morphine levels and morphine to codeine ratios were not significantly different

between the anterior and the posterior portion of the brains of (a,c) propranolol-

treated rats (p>0.6, p>0.4, respectively; n=8 total) or (b,d) propafenone-treated rats

(p>0.3, p>0.3, respectively; n=4 total).

0

11

22

33

MO

R (

ng

/g, m

ean

+ S

EM)

Anterior Posterior 0

11

22

33

MO

R (

ng

/g, m

ean

+ S

EM)

Anterior Posterior

0

0.005

0.01

0.015

MO

R/C

OD

(m

ean

+ S

EM)

Anterior Posterior 0

0.005

0.01

0.015

MO

R/C

OD

(m

ean

+ S

EM)

Anterior Posterior

81

3.9 Analgesia correlated with brain, and not plasma, morphine

levels

For rats in which plasma and brain morphine and codeine levels were measured at

30 min after codeine injection, tail-flick latency (TFL) was measured right before

sacrifice. This allowed us to assess the correlations between TFL and the various

pharmacokinetic parameters measured, and to confirm which of these parameters is

responsible for analgesia.

TFL trended toward correlating with brain morphine concentration (p=0.054,

Figure 24.a), and correlated significantly with brain morphine to codeine ratio

(p<0.006, Figure 25.a) and brain morphine to total drug ratio (p<0.006, Figure

26.a) There was no correlation between TFL and brain or plasma codeine

concentration (p>0.5, Figure 27.a; p>0.8, Figure 27.b), brain or plasma total drug

concentration (p>0.5, Figure 27. c; p>0.8; Figure 27.d), plasma morphine

concentration (p>0.8, Figure 24.b), plasma morphine to codeine ratio (p>0.7,

Figure 25.b), or plasma morphine to total drug ratio (p>0.7, Figure 26.b) (n=20

total for all brain measurements, n=21 total for all plasma measurements).

82

24.a) Brain b) Plasma

Figure 24. Analgesia correlated with brain, and not plasma, morphine levels.

Tail-flick latency (TFL) trended toward correlating with morphine concentration in

(a) the brain (p=0.054, n=20 total) and not in (b) plasma (p>0.8, n=21 total). Dark

diamonds and dark squares represent propranolol- and propafenone-treated rats,

and pale diamonds and pale squares represent the respective vehicle-treated rats.

MOR=morphine.

-2

0

2

4

6

8

0 20 40 60

TFL

(sec

)

Brain MOR (ng/g)

-2

0

2

4

6

8

0 20 40 60

TFL

(sec

)

Plasma MOR (ng/ml)

R = 0.43 R = 0.04

83

25.a) Brain b) Plasma

Figure 25. Analgesia correlated with brain, and not plasma, morphine to

codeine ratios. Tail-flick latency (TFL) correlated significantly with morphine to

codeine ratios in (a) the brain (p<0.006, n=20 total) and not in (b) plasma (p>0.7,

n=21 total). Dark diamonds and dark squares represent propranolol- and

propafenone-treated rats, and pale diamonds and pale squares represent the

respective vehicle-treated rats. COD=codeine, MOR=morphine.

-2

0

2

4

6

8

0 0.01 0.02 0.03 0.04

TFL

(sec

)

Brain MOR/COD

-2

0

2

4

6

8

0 0.01 0.02 0.03 0.04

TFL

(sec

)

Plasma MOR/COD

R = 0.59 R = 0.08

84

26.a) Brain b) Plasma

Figure 26. Analgesia correlated with brain, and not plasma, morphine to total

drug ratios. Tail-flick latency (TFL) correlated significantly with morphine to total

drug ratios in (a) the brain (p<0.006, n=20 total) and not in (b) plasma (p>0.7, n=21

total). Dark diamonds and dark squares represent propranolol- and propafenone-

treated rats, and pale diamonds and pale squares represent the respective vehicle-

treated rats. COD=codeine, MOR=morphine.

R = 0.59

-2

0

2

4

6

8

0 0.01 0.02 0.03 0.04

TFL

(se

c)

Brain MOR/(MOR+COD)

R = 0.08

-2

0

2

4

6

8

0 0.01 0.02 0.03 0.04

TFL

(se

c)

Plasma MOR/(MOR+COD)

85

27.a) Brain b) Plasma

c) Brain d) Plasma

Figure 27. Analgesia did not correlate with codeine levels or total drug levels in

brain or plasma. There was no correlation between tail-flick latency (TFL) and

codeine concentration (a,b) or total drug concentration (c,d) in brain (p>0.5, p>0.5,

respectively; n=20 total) or plasma (p>0.8, p>0.8, respectively; n=21 total). Dark

diamonds and dark squares represent propranolol- and propafenone-treated rats,

and pale diamonds and pale squares represent the respective vehicle-treated rats.

COD=codeine, MOR=morphine.

-2

0

2

4

6

8

0 2000 4000 6000

TFL

(sec

)

Brain COD (ng/g)

-2

0

2

4

6

8

0 2000 4000 6000

TFL

(sec

)

Plasma COD (ng/ml)

-2

0

2

4

6

8

0 2000 4000 6000

TFL

(sec

)

Brain MOR+COD (ng/g)

-2

0

2

4

6

8

0 2000 4000 6000

TFL

(sec

)

Plasma MOR+COD (ng/ml)

R = 0.13 R = 0.03

R = 0.13 R = 0.03

86

3.10 Inhibitor-treated rats did not have lower morphine levels

in the brain at 60 or 90 min after codeine injection

To further investigate whether the pharmacokinetic effects of brain CYP2D inhibition

on codeine were consistent with its behavioural effects, we examined brain and

plasma morphine and codeine levels at 60 and 90 min after codeine injection in

inhibitor- and vehicle-treated rats. These time points were chosen because they

were when there was no longer a significant difference in behaviour (i.e., TFL) after

inhibitor compared to after vehicle treatment.

Rats were injected with either 20 μg i.c.v. propranolol or vehicle. Codeine was

injected at 30 mg/kg s.c. 24 h after propranolol injection. Rats were sacrificed 60 or

90 min after codeine injection.

At both 60 and 90 min after codeine injection, there was no significant

difference in morphine concentrations in brain (p>0.1, p>0.9, respectively; Figure

28.a) or plasma (p>0.5, p>0.2, respectively; Figure 28.b), morphine to codeine

ratios in brain (p>0.4, p>0.2, respectively; Figure 29.a) or plasma (p>0.2, p>0.8,

respectively; Figure 29.b), morphine to total drug ratios in brain (p>0.4, p>0.2,

respectively; Figure 30.a) or plasma (p>0.2, p>0.8, respectively; Figure 30.b),

codeine concentrations in brain (p>0.6, p>0.4, respectively; Figure 31.a) or plasma

(p>0.1, p>0.5, respectively; Figure 31.b), or total drug concentrations in brain

(p>0.6, p>0.4, respectively; Figure 31.c) or plasma (p>0.1, p>0.5, respectively;

Figure 31.d) between propranolol- and vehicle-treated rats (n=7/group for 60 min;

n=8/group for 90 min).

87

28.a) Brain b) Plasma

Figure 28. Inhibitor-treated rats did not have lower morphine levels in the brain

at 60 or 90 min after codeine injection. At both 60 min (n=7/group) and 90 min

(n=8/group) after codeine injection, there was no significant difference in morphine

concentrations in (a) brain (p>0.1, p>0.9, respectively) or (b) plasma (p>0.5, p>0.2,

respectively) between propranolol-treated rats and vehicle-treated rats.

COD=codeine, MOR=morphine.

0

15

30

45

60 90

MO

R (

ng

/g, m

ean

+ S

EM)

Time after COD injection (min)

Vehicle

Propranolol

0

15

30

45

60 90

MO

R (

ng

/g, m

ean

+ S

EM)

Time after COD injection (min)

Vehicle

Propranolol

88

29. a) Brain b) Plasma

Figure 29. Inhibitor-treated rats did not have lower morphine to codeine ratios

in the brain or plasma at 60 or 90 min after codeine injection. At both 60 min

(n=7/group) and 90 min (n=8/group) after codeine injection, there was no

significant difference in morphine to codeine ratio in (a) brain (p>0.4, p>0.2,

respectively) or (b) plasma (p>0.2, p>0.8, respectively) between propranolol-

treated rats and vehicle-treated rats. COD=codeine, MOR=morphine.

0

0.01

0.02

0.03

60 90

MO

R/C

OD

(m

ean

+ S

EM)

Time after COD injection (min)

Vehicle

Propranolol

0

0.01

0.02

0.03

60 90

MO

R/C

OD

(m

ean

+ S

EM)

Time after COD injection (min)

Vehicle

Propranolol

89

30.a) Brain b) Plasma

Figure 30. Inhibitor-treated rats did not have lower morphine to total drug

ratios in the brain or plasma at 60 or 90 min after codeine injection. At both 60

min (n=7/group) and 90 min (n=8/group) after codeine injection, there was no

significant difference in morphine to total drug ratio in (a) brain (p>0.4, p>0.2,

respectively) or (b) plasma (p>0.2, p>0.8, respectively) between propranolol-

treated rats and vehicle-treated rats. COD=codeine, MOR=morphine.

0

0.01

0.02

0.03

60 90

MO

R/(

MO

R+C

OD

) (m

ean

+ S

EM)

Time after COD injection (min)

Vehicle

Propranolol

0

0.01

0.02

0.03

60 90

MO

R/(

MO

R+C

OD

) (m

ean

+ S

EM)

Time after COD injection (min)

Vehicle

Propranolol

90

31.a) Brain b) Plasma

c) Brain d) Plasma

Figure 31. Inhibitor-treated rats did not have lower codeine levels or total drug

levels in the brain or plasma at 60 or 90 min after codeine injection. At both 60

min (n=7/group) and 90 min (n=8/group) after codeine injection, there was no

significant difference in codeine concentration (a,b) or total drug concentration (c,d)

in (a,c) brain (p>0.6 at 60 min and p>0.4 at 90 min for both measurements) or (b,d)

plasma (p>0.1 at 60 min and p>0.5 at 90 min for both measurements) between

propranolol-treated rats and vehicle-treated rats. COD=codeine, MOR=morphine.

0

1500

3000

4500

60 90

CO

D (

ng

/g, m

ean

+ S

EM)

Time after COD injection (min)

Vehicle

Propranolol

0

1500

3000

4500

60 90

CO

D (

ng

/g, m

ean

+ S

EM)

Time after COD injection (min)

Vehicle

Propranolol

0

1500

3000

4500

60 90

MO

R+C

OD

(n

g/g

, mea

n +

SEM

)

Time after COD injection (min)

Vehicle

Propranolol

0

1500

3000

4500

60 90

MO

R+C

OD

(n

g/g

, mea

n +

SEM

)

Time after COD injection (min)

Vehicle

Propranolol

91

3.11 Inhibiting brain CYP2D in vivo lowered in vitro codeine

metabolism in the brain but not liver

To confirm that i.c.v. injections of CYP2D inhibitors could selectively decrease

morphine formation from codeine in the brain, we injected rats with i.c.v.

propranolol or vehicle, sacrificed the animals, and then measured the effect on

codeine metabolism ex vivo in brain membranes and liver microsomes. If

propranolol had irreversibly inhibited CYP2D in the brain and not in the liver, then

brain membranes should have more in vitro CYP2D activity in vehicle- versus

propranolol-treated rats, while in liver microsomes they should be similar. This

would provide further evidence to support that the lower morphine levels in the

brains of inhibitor-treated was indeed due to irreversible brain CYP2D inhibition

and reduced codeine metabolism in the brain.

Rats were injected with either 20 μg i.c.v. propranolol or vehicle and then

sacrificed 24 h later. In brain membranes incubated with codeine (500 μM=10 x Km

(Mikus, Somogyi et al. 1991)), velocity of morphine formation was significantly lower

in propranolol-treated rats than in vehicle-treated rats (p<0.04; n=8/group; Figure

32.a). In liver microsomes incubated with 500 μM codeine, there was no significant

difference in the velocity of morphine formation between propranolol- and vehicle-

treated rats (p>0.9; n=8/group; Figure 32.b).

92

32.a) Brain membranes b) Liver microsomes

Figure 32. Inhibiting brain CYP2D in vivo lowered in vitro codeine metabolism

to morphine in brain membranes but not in liver microsomes. (a) In brain

membranes incubated with 500 μM codeine, velocity of morphine formation was

significantly lower in propranolol-treated rats than in vehicle-treated rats (p<0.04,

n=8/group). (b) In liver microsomes incubated with 500 μM codeine, there was no

significant difference in the velocity of morphine formation between propranolol-

treated rats and vehicle-treated rats (p>0.9, n=8/group). *p<0.05. V=velocity of

morphine formation.

0

2

4

6V

(p

mo

l/m

g/h

, mea

n +

SEM

)

Vehicle Propranolol

*

0

4

8

12

V (

nm

ol/

mg

/h, m

ean

+ S

EM)

Vehicle Propranolol

93

3.12 Inhibiting brain CYP2D in vivo lowered in vitro

dextromethorphan metabolism in the brain but not liver

To further confirm that i.c.v. injections of CYP2D inhibitors could reduce brain

CYP2D activity, we measured the effect of i.c.v. propranolol treatment on the ex vivo

metabolism of a CYP2D probe drug, dextromethorphan, in brain membranes and

liver microsomes. If i.c.v. propranolol irreversibly inhibits CYP2D in the brain and

not in the liver, then brain membranes should have more in vitro CYP2D activity in

vehicle- versus propranolol-treated rats, while in liver microsomes they should be

similar. If i.c.v. propranolol can inhibit the metabolism of an additional CYP2D

substrate other than codeine in brain membranes but not in liver microsomes ex

vivo, this would provide further evidence that i.c.v. inhibitor treatment inhibits brain,

but not hepatic, CYP2D activity.

Rats were injected with either 20 μg i.c.v. propranolol or vehicle and then

sacrificed 24 h later. In brain membranes incubated with dextromethorphan (25

μM=10x Km (Kerry, Somogyi et al. 1993)), velocity of dextrorphan formation trended

toward being lower in propranolol-treated rats than in vehicle-treated rats (p=0.095;

n=12/group; Figure 33.a). In liver microsomes incubated with 25 μM

dextromethorphan, there was no significant difference in the velocity of dextrorphan

formation between propranolol- and vehicle-treated rats (p>0.8; n=12/group;

Figure 33.b).

94

33.a) Brain membranes b) Liver microsomes

Figure 33. Inhibiting brain CYP2D in vivo lowered in vitro dextromethorphan

metabolism to dextrorphan in brain membranes but not in liver microsomes.

(a) In brain membranes incubated with 25 μM dextromethorphan, velocity of

dextrorphan formation trended toward being lower in propranolol-treated rats than

in vehicle-treated rats (p=0.095, n=12/group). (b) In liver microsomes incubated

with 25 μM dextromethorphan, there was no significant difference in the velocity of

dextrorphan formation between propranolol-treated rats and vehicle-treated rats

(p>0.8, n=12/group). V=velocity of dextrorphan formation.

0

4

8

12V

(p

mo

l/m

g/h

, mea

n +

SEM

)

Vehicle Propranolol 0

20

40

60

V (

nm

ol/

mg

/h, m

ean

+ S

EM)

Vehicle Propranolol

95

Section 4: Discussion, Conclusions, Future Directions

4.1 Summary and further implications

While previous studies have demonstrated the activity (Tyndale, Li et al. 1999) and

inducibility (Mann, Miksys et al. 2008, Yue, Miksys et al. 2008) of brain CYP2D, they

did not indicate whether brain CYP2D expression or activity levels were high

enough to influence drug response. This is the first study to show that brain CYP2D-

mediated metabolism can alter the effect of a centrally-acting drug. We decreased

rat brain CYP2D activity through the use of CYP2D inhibitors, which decreased

codeine-induced analgesia. We also demonstrated that morphine concentrations in

the brain correlated with analgesia. Our results reveal that brain CYP2D activity in

the rat plays a significant role in the metabolism and effect of codeine, suggesting

that brain CYP2D-mediated metabolism may have an important impact on the

response to the numerous other centrally-acting CYP2D substrates as well. The

response to centrally-acting drugs can display large interindividual variation as well

as poor correlations with plasma drug levels (Michels and Marzuk 1993a).

Differences in the level of drug metabolism in the brain, which does not affect

plasma drug levels as shown here, may contribute to this variation in drug response.

96

4.1.1 Rat model of reduced brain CYP2D activity

A factor that has made it difficult in the past to ascertain the contribution of brain

CYP2D to drug metabolism is the challenge of differentiating between metabolites

that were formed in the brain versus those that were formed in the liver and crossed

into the brain. Establishing the role of brain CYP2D in local drug metabolism and

response requires a model in which CYP2D activity can be altered in the brain

without changing hepatic activity. Using i.c.v. injection of CYP2D inhibitors, we were

able to selectively inhibit rat brain CYP2D activity. The decrease in brain but not

plasma morphine concentrations, and brain but not hepatic enzyme activity, after

i.c.v. injections of CYP2D inhibitors offers convincing evidence that the decrease in

codeine-induced analgesia was due to a decrease in brain, and not liver, CYP2D-

mediated metabolism of codeine to morphine.

To assess the effect of inhibiting brain CYP2D on the analgesic response to

codeine, the use of propranolol and propafenone has several advantages over other

CYP2D inhibitors. Propranolol, a -adrenergic receptor blocker, and propafenone, a

sodium channel blocker, are not expected to affect the CNS in a way that would

affect behaviour in the tail-flick test, and there have been no reports of either of

these drugs having an effect on nociception or analgesia. This attribute makes them

more suitable for our purposes than other CYP2D inhibitors. For example, the

antidepressant paroxetine is a potent CYP2D6 MBI (Bertelsen, Venkatakrishnan et al.

2003), but it has antinociceptive effects, possibly through opioidergic mechanisms

(Duman, Kesim et al. 2004). Quinidine, a commonly used CYP2D competitive

inhibitor (Brosen, Gram et al. 1987), enhances morphine-induced analgesia by

97

inhibiting efflux transporters (Okura, Morita et al. 2009). Quinine, a potent CYP2D

competitive inhibitor in rats (Kobayashi, Murray et al. 1989, Kerry, Somogyi et al.

1993, Xu, Aasmundstad et al. 1995), has antinociceptive effects possibly through

dopaminergic mechanisms (Amabeoku, Ewesuedo et al. 1992). Fluoxetine, another

CYP2D competitive inhibitor, has its own antinociceptive effects in the tail-flick test,

possibly through central opioid pathways (Schreiber, Backer et al. 1996, Singh, Jain

et al. 2001), and it potentiates morphine-induced analgesia (Erjavec, Coda et al.

2000, Nayebi, Hassanpour et al. 2001). Furthermore, propranolol and propafenone

are relatively selective inhibitors of CYP2D (Masubuchi, Fujita et al. 1991,

Turpeinen, Korhonen et al. 2006, McGinnity, Waters et al. 2008), unlike paroxetine

which is also a MBI of CYP3A (Obach, Walsky et al. 2007) and fluoxetine which also

potently inhibits CYP2C19 (Turpeinen, Korhonen et al. 2006, McGinnity, Waters et

al. 2008). Thus, propranolol and propafenone were better choices than many other

CYP2D inhibitors.

An advantage of using a MBI, propranolol, was that it could be administered

well ahead of the test drug. Covalent modifications of enzymes by MBIs are

permanent and activity can only be recovered via synthesis of new enzyme, which

therefore results in a longer lasting inhibition that is not reversed by the presence of

a higher affinity substrate. Propranolol’s elimination half-life in rat blood and brain

are both approximately 1 h (Bianchetti, Elghozi et al. 1980). Therefore, there was

virtually no propranolol left in the body by the time codeine was given 24 h later,

which limits the chance of off-target effects. Showing that brain morphine levels and

analgesia from codeine were reduced by two CYP2D inhibitors with different

mechanisms of inhibition and different pharmacological actions supports the notion

98

that these results were indeed due to brain CYP2D inhibition as opposed to some

other property or action specific to either inhibitor.

4.1.2 Inhibition of brain CYP2D lowers codeine-induced

analgesia

We showed that inhibitor treatment resulted in lower tail-flick latency (TFL) and area

under the analgesia time curve (AUC) compared to vehicle treatment during the first

30 min after codeine injection, which is consistent with our hypothesis that inhibiting

brain CYP2D would reduce analgesia during the initial period after codeine

injection. However, TFL at 40 min after codeine injection and AUC from 30-60 min

and 0-60 min after codeine injection were also significantly lower after inhibitor

treatment, suggesting that the effects of brain CYP2D inhibition last longer than

estimated from Chen, Irvine et al. (1990). TFLs later than 40 min after codeine

injection and AUC for 60-120 min and 0-120 min after codeine injection were not

significantly different after inhibitor treatment compared to after vehicle treatment,

suggesting that inhibiting brain CYP2D only affects the earlier period after codeine

injection. We interpret these results as such: during the initial 40 min period after

codeine injection in rats, analgesia is mediated mainly by morphine formed by brain

CYP2D-mediated codeine metabolism, as relatively less morphine formed by

hepatic CYP2D has crossed into the brain at this time. Thus, during this period,

inhibition of brain CYP2D activity and the ensuing reduction in brain morphine

levels results in lower analgesia. Over time, morphine formed by hepatic CYP2D

99

accumulates in the periphery and crosses the BBB, and by 60 min after codeine

injection, sufficient amounts of morphine have entered the brain to offset the initial

differences in brain morphine levels that were due to differences in brain CYP2D

activity. Because more morphine is formed by the liver than by the brain, analgesia

at these later time points is mediated mainly by morphine formed by hepatic CYP2D

that crosses into the brain. Inhibition of brain CYP2D did not affect hepatic CYP2D

activity, and thus did not result in lower analgesia past 60 min after codeine

injection. Our interpretation is supported by the finding that morphine could be

detected in the brains of codeine- but not morphine-injected rats at 30 min after

peripheral injection (Chen, Irvine et al. 1990), implying that morphine in the brain at

this time is due to local codeine metabolism as morphine from the periphery had not

yet entered the brain. Also in line with our interpretation are findings that brain

uptake of codeine is faster than that of morphine (Oldendorf, Hyman et al. 1972), that

morphine is transported out of the brain by efflux transporters (Bouw, Gardmark et

al. 2000), and the fact that morphine has one less methyl group than codeine, which

is expected to make morphine less lipid soluble and therefore less able to cross the

BBB. Altogether, our findings indicate that inhibiting brain CYP2D reduces analgesia

during the first hour after codeine injection in rats, and that variation in brain CYP2D

activity may influence the onset of analgesia from codeine.

We did not observe a difference between baseline TFL and TFL after inhibitor

treatment, which indicates that the inhibitors by themselves do not affect

nociception. We also saw that inhibitor treatment did not result in different TFL or

AUC compared to vehicle treatment during any time after morphine injection, which

was expected as morphine, being the active drug, does not depend on CYP2D

100

activity in order to confer analgesia. This finding also indicates that the inhibitors did

not lower codeine-induced analgesia by way of interfering with morphine’s actions

(e.g., by affecting mu-opioid receptors or drug transporters). TFL and AUC did not

differ between the two phases of codeine or morphine testing, indicating that the

rats did not develop tolerance to either of the opioids. Following vehicle treatment,

there was no difference in TFL or AUC after codeine injection compared to after

morphine injection, which verifies that the codeine and morphine doses we used

produce equivalent analgesia. Altogether, these findings indicate that inhibiting

brain CYP2D reduces analgesia from codeine but not from an equianalgesic dose of

morphine.

4.1.3 Analgesia correlates with morphine levels in the brain

and not plasma

We showed that at 30 min after codeine injection, which is when the largest

difference in TFL after inhibitor treatment compared to after vehicle treatment

occurred, inhibitor-treated rats had lower brain morphine concentrations,

corresponding to the lower TFLs. We also showed that at this time, inhibitor-treated

rats had lower morphine to codeine ratios in the brain, consistent with the notion that

less codeine was metabolized to morphine in the brains of these animals. The ratio of

morphine to total (morphine plus codeine) drug concentration in the brain was also

lower in inhibitor-treated rats at this time, indicating that a smaller portion of the

total drug was found as morphine in the inhibited animals. There was no difference

101

between inhibitor- and vehicle-treated rats in codeine levels (as morphine is a minor

metabolite) or in total drug levels in brain or plasma at this time, indicating that

differences in analgesia were not due to differences in either of these two

parameters. There was also no difference between inhibitor- and vehicle-treated

rats in any plasma parameters at this time, indicating that hepatic CYP2D activity was

not affected by inhibitor treatment and that differences in analgesia were not due to

plasma drug levels. At this time, TFL correlated with brain morphine levels, brain

morphine to codeine ratios, and brain morphine to total drug ratios, but not with

brain codeine levels, brain total drug levels or any parameter in the plasma.

At 60 or 90 min after codeine injection, when there was no difference in TFL

after inhibitor treatment compared to after vehicle treatment, there was

correspondingly no difference in any of the pharmacokinetic parameters between

inhibitor- and vehicle-treated rats. This is in line with the belief that at these later

time points, analgesia is mediated by morphine formed in the liver that crosses into

the brain, and that initial differences in brain morphine levels between inhibitor-

and vehicle-treated rats seen at 30 min are now masked by the large amount of

morphine that has entered the brain from the periphery. Overall, these results

suggest that analgesia is mediated by morphine in the brain, and that the differences

in codeine-induced analgesia we observed after inhibitor treatment compared to

after vehicle treatment were due to differences in brain morphine levels.

102

4.1.4 Inhibiting brain CYP2D in vivo lowers in vitro enzyme

activity in brain membranes and not liver microsomes

CYP2D velocity, measured using codeine and dextromethorphan as substrates, was

lower in brain membranes of rats that had received i.c.v. inhibitor injections

compared to rats that had received i.c.v. vehicle injections, indicating that inhibitor

treatment did indeed result in the inhibition of brain CYP2D activity. In liver

microsomes, there was no difference in CYP2D velocity between inhibitor- and

vehicle-treated rats, indicating that hepatic CYP2D was not inhibited by i.c.v.

inhibitor injection. These findings provide evidence that the lower brain morphine

levels seen in inhibitor-treated rats were indeed due to reduced brain CYP2D-

mediated metabolism of codeine to morphine. Therefore, while hepatic CYP2D may

be responsible for the bulk of morphine formed from the systemic codeine injection,

localized, in situ codeine metabolism in the brain can meaningfully impact brain

morphine levels and, in turn, response to codeine.

4.1.5 Limitations

A limitation of this study was that propafenone, being a competitive inhibitor, had to

be administered during the same time period as codeine, which could potentially

put more stress on the animal and interfere with the response that is being

observed. It is also not known whether propafenone interacts with codeine in a way

that would reduce analgesia. However, given the very similar responses seen

following the two different inhibitors, this does not appear to have been an issue.

103

Another limitation was that propranolol, being lipophilic, is able to cross the

BBB (Rowland, Yeo et al. 1994, Komura and Iwaki 2005). Therefore, higher doses of

i.c.v. administered propranolol can cross into the periphery and inhibit hepatic

CYP2D, as was the case with the 40 g dose used in our pilot study (but not with the

lower 20 g dose used in our subsequent experiments). This put an upper limit to the

inhibitor dose we could use and thus we may not have completely inhibited CYP2D

in the brain; after inhibitor treatment there was still morphine in the brain at 30 min

following codeine injection and ex vivo CYP2D activity in the brain membranes.

Identifying a CYP2D inhibitor that does not cross the BBB would be advantageous as

it would allow higher doses to be used which may achieve more complete brain

CYP2D inhibition and result in an even larger effect on drug metabolism and

response, more analogous to those who are CYP2D6 PMs.

An additional limitation was that the heat exposure cut-off of 10 sec (which was

to prevent damage to tail skin) did not allow us to detect differences in analgesia in

the period when analgesia from codeine peaked, during which TFLs of 10 sec were

reached both after inhibitor treatment as well as after vehicle treatment. Therefore, it

is unclear whether the lack of difference seen in TFL between inhibitor versus

vehicle treatment after 40 min following codeine injection was due to the cut-off or to

an actual lack of difference in analgesia. It is therefore possible that inhibition of

brain CYP2D had a larger or longer effect on codeine-induced analgesia than what

we were able to detect with this model at the dose of codeine used. This may have

more likely been the case for the propafenone-treated rats, in which 8 out of 12 rats

reached TFLs of 10 sec after both vehicle and inhibitor treatment, compared to the

propranolol-treated rats in which only 6 out of 16 did.

104

We used a subcutaneous route of administration for codeine instead of the

oral route which is typically used in humans. This route was chosen because it

minimizes first pass metabolism by the liver, thereby allowing us to better detect the

effects of brain-mediated metabolism. However, a disadvantage of this method is

that it may not accurately represent the impact of brain CYP2D on the response to

codeine when it is administered by its typical route. Oral codeine is subject to first

pass effects, so even though we have shown that brain CYP2D-mediated metabolism

influences response to subcutaneously injected codeine, the same may not hold

when codeine is given orally. It would be useful to verify whether the impact of brain

CYP2D on codeine response will remain in the presence of more extensive hepatic

metabolism from the oral mode of administration. This is suggested to be the case

by the finding that, inhibiting brain, but not hepatic, CYP2B altered the sleep-times

induced by intaperitoneal injection of the anaesthetic propofol (metabolized by

CYP2B) (Khokhar and Tyndale 2011). Since intraperitoneal injection, like oral

administration, is subject to first pass metabolism, this finding suggests that brain

CYP activity can have a meaningful impact on drug response even in the presence of

hepatic metabolism.

In summary, by using i.c.v. injections of CYP2D inhibitors, we inhibited

CYP2D in rat brain and not liver, and established an animal model of reduced brain

CYP2D activity. We used this model to evaluate the role of brain CYP2D in the

metabolism and effect of codeine, a centrally-acting drug. We found that inhibition

of brain CYP2D could decrease the metabolic activation of codeine to morphine in

the brain and reduce codeine-induced analgesia. Differences in brain CYP activity

may contribute to interindividual variation in the response to centrally-acting drugs.

105

For example, the higher CYP2D6 levels in smokers and seniors might help explain

the altered efficacy or side-effect profiles of centrally-acting drugs seen in these

individuals (Nelson, Mazure et al. 1995, Jabs, Bartsch et al. 2003, Mann, Miksys et al.

2008, Mann, Miksys et al. 2012).

4.2. Clinical relevance of brain CYP2D activity

We have demonstrated that rat brain CYP2D is functional in vivo and can have a

significant contribution to drug effect. While our findings may not have definite

clinical implications in codeine-induced analgesia in humans, they do point to the

functionality of rat brain CYP2D and its ability to affect drug response through local

metabolism. Brain CYP2D-mediated metabolism may not only have a bearing on

drug efficacy, but the involvement of CYP2D in the formation or metabolism of

endogenous compounds such as serotonin (Yu, Idle et al. 2003a) and dopamine

(Bromek, Haduch et al. 2010), indicates that these CYPs may also play a role in

normal brain function. This notion is supported by associations of CYP2D6 genotype

with resting brain activity (Kirchheiner, Seeringer et al. 2011) and personality traits

(Bijl, Luijendijk et al. 2009). CYP2D also inactivates neurotoxins such as 1-methyl-4-

phenylpyridinium (MPP+) (Mann and Tyndale 2010). Inhibition of CYP2D increased

the neurotoxicity induced by MPP+ in a human neuronal cell line (Mann and Tyndale

2010). The induction of brain CYP2D by commonly used drugs such as alcohol and

nicotine (Warner and Gustafsson 1994, Miksys, Rao et al. 2002, Miksys and Tyndale

2004, Mann, Miksys et al. 2008, Yue, Miksys et al. 2008, Miller, Miksys et al. 2012)

106

may affect the likelihood of neurotoxicity following exposure to, and subsequent

metabolism of, neurotoxic substances. The multiple families of drug-metabolizing

CYPs in the brain, including CYP2B, CYP2D, CYP2E1, CYP3A and CYP4 (Strobel,

Thompson et al. 2001, Meyer, Gehlhaus et al. 2007) and their diverse range of

centrally-acting substrates, lends support to the brain being an important organ in

drug metabolism.

4.2.1 Centrally-acting drugs

Brain CYP2D expression and activity can vary among individuals based on genetics,

exposure to environmental inducers or differences in age. Differences in brain

CYP2D levels may affect the local pharmacokinetics of its many centrally-acting

substrates, which include clinical drugs, drugs of abuse, neurotoxins and

endogenous substances. The induction of brain CYP2D6 may result in increased

substrate metabolism, which in turn may lead to altered efficacy of clinical drugs, as

well as altered susceptibility to adverse drug reactions. Smokers respond differently

to centrally-acting CYP2D6 substrates such as antipsychotics and antidepressants

compared to nonsmokers (Jabs, Bartsch et al. 2003, George, Sacco et al. 2008).

Smokers also have elevated CYP2D6 levels in the brain, and not liver (Miksys and

Tyndale 2004, Mann, Miksys et al. 2008). We can therefore speculate that these

differences in drug response are due to differences in the level of brain CYP2D6-

mediated metabolism.

107

CYP2D6 metabolizes various antipsychotics including haloperidol,

perphenazine, thioridazine and risperidone (Ingelman-Sundberg 2005). Plasma

levels of risperidone and its active metabolite do not correlate with drug response

(Spina, Avenoso et al. 2001) or extrapyramidal side effects (EPS) (Lane, Chiu et al.

2000, Riedel, Schwarz et al. 2005) in schizophrenia patients, and this may be due to

variation in drug metabolism in the brain. Smokers have less EPS from

antipsychotics than non-smokers (Jabs, Bartsch et al. 2003). CYP2D in the human

brain is elevated in smokers, particularly in the basal ganglia which is involved in

EPS (Jabs, Bartsch et al. 2003, Mann, Miksys et al. 2008). Higher brain CYP2D6

activity, and therefore increased clearance of antipsychotics, may contribute to the

reduced EPS in smokers (Funck-Brentano, Boelle et al. 2005). Moreover,

antidepressants (inactivated by CYP2D6) were less effective in smokers compared

to non-smokers (George, Sacco et al. 2008). This may be attributable to the higher

levels of brain CYP2D6 in smokers which may increase inactivation of

antidepressants in the brain.

Because brain CYP2D6 levels increase with age, the impact of drug

metabolism in the brain by CYP2D6 may be augmented in seniors (Mann, Miksys et

al. 2012). Desipramine, an antidepressant inactivated by CYP2D6, was less effective

in older (>75 years) individuals than in younger patients when controlling for drug

dose and plasma levels (Nelson, Mazure et al. 1995). The increased levels of brain

CYP2D6 in individuals 60 to 80 years of age may increase rates of desipramine

inactivation and contribute to its reduced efficacy. In sum, variation in brain CYP2D6

levels, whether due to genetics, environmental induction or age may alter the

therapeutic and/or adverse effects of centrally-acting drugs.

108

4.2.2 Drugs of abuse

Codeine, along with being commonly prescribed for pain relief, is also widely

abused. Our observation that brain CYP2D activity influences the early analgesic

effects of codeine also has implications for codeine’s abuse liability. The abuse

liability of a drug depends, at least partly, on the quickness of onset of its reinforcing

effects (Griffiths and Wolf 1990, Farre and Cami 1991). As previously suggested

(Tyndale, Droll et al. 1997, Kathiramalainathan, Kaplan et al. 2000), the reinforcing

qualities of codeine come from its O-demethylated metabolites (morphine,

morphine-6-glucuronide), which are formed by CYP2D. Therefore, analogous to the

analgesic effects of codeine, the initial reinforcing effects of codeine should be

mediated by its metabolism to morphine in the brain, before morphine from the

periphery has crossed into the brain. Since the formation of morphine in the brain

should occur at a higher rate in those with elevated brain CYP2D activity (smokers,

alcoholics, seniors), these individuals are expected to experience a quicker onset of

reinforcing effects from codeine and may be more prone to codeine abuse. Higher

brain CYP2D6 activity can also be speculated to be a risk factor for the abuse of

recreational drugs that are inactivated by CYP2D6 (e.g., amphetamine), as the

abuse liability of a drug also depends partly on the quickness of offset of its

reinforcing effects (Griffiths and Wolf 1990, Farre and Cami 1991). Overall, altered

brain CYP2D levels, due to genetics, environmental induction or aging, could affect

the disposition of drugs of abuse metabolized by CYP2D (e.g. amphetamine, MDMA,

dextromethorphan), and thus lead to interindividual differences in the susceptibility

to abusing these drugs.

109

4.2.3 Endogenous substrates

CYP2D is involved in the synthesis and metabolism of a range of centrally-acting

endogenous substances (Funae, Kishimoto et al. 2003), which suggests a role for this

enzyme in basal brain function. Because CYP2D6 can catalyze the formation of

dopamine, serotonin, epinephrine and norepinephrine, differences in CYP2D6

levels are believed to have an impact on personality or mood (Funae, Kishimoto et

al. 2003, Yu, Idle et al. 2003a, Bromek, Haduch et al. 2010). Genetic variation in

CYP2D6 is associated with various personality and behavioural traits (Roberts, Luty

et al. 2004, Kirchheiner, Lang et al. 2006, Gonzalez, Penas-Lledo et al. 2008, Ahlner,

Zackrisson et al. 2010). For example, CYP2D6 PMs have been reported to have

higher anxiety (Llerena, Edman et al. 1998, Gonzalez, Penas-Lledo et al. 2008), less

success in being socialized (Llerena, Edman et al. 1998), higher impulsivity

(Gonzalez, Penas-Lledo et al. 2008, Penas-Lledo, Dorado et al. 2009), higher ease of

decision making (Bertilsson, Alm et al. 1989), lower harm avoidance (Roberts, Luty

et al. 2004), more novelty seeking behaviours (Roberts, Luty et al. 2004), and lower

competitiveness (Kirchheiner, Lang et al. 2006, Gonzalez, Penas-Lledo et al. 2008).

In addition to catalyzing the conversion of endogenous precursor compounds

to neurotransmitters, CYP2D6 also metabolizes neurosteroids such as progesterone

and allopregnanolone (Hiroi, Kishimoto et al. 2001, Kishimoto, Hiroi et al. 2004,

Niwa, Okada et al. 2008). Progesterone can regulate the synthesis and release of

neurotransmitters and neuropeptides (Pluchino, Luisi et al. 2006). Progesterone

increases serotonin turnover, catechol-O-methyltransferase activity, and MAO

activity in rats (Genazzani, Stomati et al. 2000). In women, progesterone treatment

110

changes mood (Hlatky, Boothroyd et al. 2002). Allopregnanolone, a metabolite of

progesterone, acts on γ-aminobutyric acid receptors and thus induces anxiolysis and

sedation (Pluchino, Luisi et al. 2006). Changes in allopregnanolone levels have been

linked with depression, anxiety, and irritability (Pluchino, Luisi et al. 2006). CYP2D6

also metabolizes anandamide (Snider, Sikora et al. 2008), an endocannabinoid that is

implicated in mood, anxiety, and emotional processing (Bambico and Gobbi, 2008).

In summary, variation in brain CYP2D6 may alter levels of these endogenous

neuromodulators and consequently influence personality, behaviour and mood.

A further function of brain CYP2D6 may be the production of endogenous

morphine. Various studies have shown that humans (Zhu, Cadet et al. 2005) and

animals (Kodaira and Spector 1988, Amann, Roos et al. 1995, Zhu, Mantione et al.

2005) are able to synthesize morphine, and that CYP2D is important in this process

as it metabolizes multiple morphine precursors (Zhu, Cadet et al. 2005, Zhu,

Mantione et al. 2005, Kream, Stefano et al. 2006). The formation of morphine from

tyramine, tyrosine and codeine was demonstrated both in vitro and in vivo in the

marine invertebrate Mytilus edulis; morphine formation was substantially reduced by

the CYP2D inhibitor quinidine (Zhu, Mantione et al. 2005, Zhu 2008). Human white

blood cells express CYP2D6 which can synthesize morphine from tyramine,

norlaudansoline and codeine; morphine production was reduced by the CYP2D6

inhibitors bufuralol, quinidine and paroxetine (Zhu, Cadet et al. 2005). Together,

these findings suggest that CYP2D is involved in endogenous morphine synthesis.

The potential functional significance of this was demonstrated in CYP2D6 PMs and

EMs who were assessed for pain thresholds with the cold pressor test (Sindrup,

Poulsen et al. 1993). Peak pain ratings and area under the pain rating-time curve

111

were significantly higher in CYP2D6 PMs than in EMs (Sindrup, Poulsen et al. 1993).

PMs may therefore have lower pain tolerance than EMs, and this is speculated to be

due to a lack of endogenous morphine production by brain CYP2D6 (Sindrup,

Poulsen et al. 1993). Levels of endogenous morphine excreted in urine were not

found to differ between CYP2D6 PMs and EMs, suggesting that other CYPs may also

be involved in morphine biosynthesis (Mikus, Bochner et al. 1994). However, brain

CYP2D6 may synthesize endogenous morphine locally in the brain, which may

modulate nociception and suggests an additional role that brain CYP2D6 may have

in analgesia (Zhu 2008). In further support of this, there is high expression of

CYP2D6 in the thalamus, a brain region that is involved in nociception (Mann, Miksys

et al. 2008, Wilson, Uhelski et al. 2008). Interindividual variability in thalamus

activity correlates with pain threshold (Dostrovsky 2000, Lenz, Weiss et al. 2004,

Ochsner, Ludlow et al. 2006), and thalamus activity (as measured by resting brain

perfusion) also correlates with CYP2D6 genotype (Kirchheiner, Seeringer et al.

2011). Altogether, variation in brain CYP2D may alter analgesia by affecting the

formation of morphine from exogenous compounds such as codeine as well by

influencing the synthesis of endogenous morphine.

4.2.4 Disease

CYP2D may have a protective role against Parkinson’s disease (PD). PD is a

neurodegenerative disorder characterized by the loss of dopaminergic neurons in

the substantia nigra that project to the striatum (Fahn 2010). A risk factor for PD is

exposure to toxins such as pesticides (Di Monte 2003, Olanow 2007). CYP2D6 can

112

inactivate numerous neurotoxins including 1-methyl-4-phenylpyridinium (MPP+)

(Mann and Tyndale 2010), 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)

(Modi, Gilham et al. 1997), tetrahydroisoquinolines (Suzuki, Fujita et al. 1992),

harmaline, harmine (Yu, Idle et al. 2003b) and N-methyl-β-carbolines (Herraiz,

Guillen et al. 2006). CYP2D6 PM status is associated with an increased risk for PD

(McCann, Pond et al. 1997), which is further increased when these individuals are

exposed to pesticides (Deng, Newman et al. 2004, Elbaz, Levecque et al. 2004). This

suggests that a lack of CYP2D6 activity, and thus an inability to inactivate

environmental neurotoxins, makes individuals more vulnerable to their effects.

Among PD cases, CYP2D6 PMs were overrepresented, and EMs had ~40-50% lower

levels of brain CYP2D6 compared to control EMs (Mann, Miksys et al. 2012).

Therefore, low levels of brain CYP2D6 may increase the risk of PD, while high levels

may protect against it.

The observation that PD cases had lower brain CYP2D6 levels compared to

controls (Mann, Miksys et al. 2012) is in line with the findings that inhibiting CYP2D

increased toxicity of MPP+ (a Parkinsonian neurotoxin inactivated by CYP2D) in

human neuronal cells (Mann and Tyndale 2010) and that over-expressing CYP2D6

protected against MPP+ toxicity in PC12 cells (Matoh, Tanaka et al. 2003).

Furthermore, brain CYP2D6 is located in areas that are ideal for the inactivation of

Parkinsonian neurotoxins (e.g., basal ganglia, BBB, dopaminergic neurons) (Gilham,

Cairns et al. 1997, Miksys, Rao et al. 2002, Dutheil, Jacob et al. 2010). Thus, brain

CYP2D6 may protect against specific neurotoxins and thereby reduce the risk for

PD.

113

Smokers have high levels of brain CYP2D6, especially in the basal ganglia

including the substantia nigra (Mann, Miksys et al 2008), and smokers also have a

reduced risk for PD (Alves, Kurz et al. 2004). Nicotine has been shown to protect

against Parkinsonian neurotoxins both in vitro and in vivo (Khwaja, McCormack et al.

2007, Quik, O’Neill et al. 2007). For example, pretreating monkeys with nicotine

protected against nigrostriatal damage from MPTP (Quik, Parameswaran et al. 2006).

Moreover, nicotine significantly induced monkey CYP2D in the basal ganglia

including the substantia nigra (Mann, Miksys et al 2008). Therefore, the increased

levels of brain CYP2D resulting from induction by smoking and nicotine may protect

against PD.

Brain CYP2D6, through its ability to synthesize dopamine from tyramine, may

help offset the loss of dopamine in PD and thereby alleviate Parkinsonian symptoms.

Brain CYP2D6 is ideally located for this purpose in the basal ganglia and in

dopaminergic neurons (Gilham, Cairns et al. 1997, Riedl, Watts et al. 1999, Bromek,

Haduch et al. 2010, Gonzalez-Hernandez, Cruz-Muros et al. 2010). Nicotine treatment

has been demonstrated to improve dopaminergic tone and motor symptoms in PD

cases, and this may partly be due to increased dopamine formation by nicotine-

induced brain CYP2D6 (Fagerstrom, Pomerleau et al. 1994, Kelton, Kahn et al. 2000,

Quik, Cox et al. 2007). Altogether, variation in brain CYP2D6, due to genetics or

environmental inducers, may alter the risk for and symptom severity of PD. Genetic

variation in CYP2D6 may also be linked with predispositions to psychiatric illnesses

such as schizophrenia, eating disorders, and major depressive disorder (Dorado,

Penas-Lledo et al. 2007, Llerena, Dorado et al. 2007, Gonzalez, Penas-Lledo et al.

2008, Ahlner, Zackrisson et al. 2010, Penas-Lledo, Dorado et al. 2012).

114

4.3 Other brain CYPs

Our data suggests that other brain CYPs may also be functional in vivo and have an

important role in local substrate metabolism. This has been demonstrated to be the

case with brain CYP2B in rats. CYP2B metabolically inactivates the anaesthetic

propofol (Court, Duan et al. 2001). Inhibiting rat brain, and not liver, CYP2B resulted

in an increase in propofol-induced sleep times, while inducing brain CYP2B

decreased sleep times (Khokhar and Tyndale 2011). CYP2B is induced by nicotine

in the brain but not in the liver (Miksys, Hoffmann et al. 2000, Lee, Miksys et al.

2006a, Yue, Khokhar et al. 2009), and smokers have higher CYP2B levels in the

brain, and not liver compared to nonsmokers (Miksys, Lerman et al. 2003, Miksys

and Tyndale 2004). Accordingly, case reports have suggested that smokers require

higher propofol doses to achieve loss of consciousness, and also have fewer side

effects from propofol, compared to nonsmokers (Chimbira and Sweeney 2000,

Lysakowski, Dumont et al. 2006), possibly due to higher brain CYP2B-mediated

propofol inactivation (Tate and Cook 1996, Trapani, Altomare et al. 2000). Therefore,

variation in brain CYP2B activity may alter response to centrally-acting drugs.

CYP2B can also metabolize neurotoxins such as the insecticide, chlorpyrifos, which

is activated to a neurotoxic metabolite. Inhibition of rat brain CYP2B blocked

chlorpyrifos activation and reduced the neurotoxic effects of chlorpyrifos treatment

(Khokhar and Tyndale 2012). Chronic nicotine treatment in rats increased the

neurotoxic effects of chlorpyrifos (Abou-Donia, Abdel-Rahman et al. 2003), possibly

due to induction of brain CYP2B by nicotine and a resultant increase in chlorpyrifos

activation. Together, these findings indicate that differences in brain CYP2B activity

115

impact the metabolism and effects of propofol, chlorpyrifos, and probably the

numerous other centrally-acting CYP2B substrates as well (e.g. diazepam,

methadone, nicotine, cocaine, phencyclidine).

CYP2E1 is also expressed in the brain and can metabolize a variety of

centrally-acting substrates including ethanol, anaesthetics, and neurotoxins (Caro

and Cederbaum 2004). While ethanol metabolism in the brain is mainly mediated by

catalase, CYP2E1 also plays a significant role (Vasiliou, Ziegler et al. 2006). CYP2E1

knockout mice had longer ethanol-induced sleep-times than did wild-type mice,

suggesting decreased ethanol metabolism in the brains of knockout mice (Vasiliou,

Ziegler et al. 2006). However, these animals had lower CYP2E1 activity the liver as

well, thereby making it unclear how much of the effect on sleep-times can be

attributed to brain versus hepatic CYP2E1. Brain CYP2E1 is induced by ethanol, and

this induction is associated with higher oxidative damage in astrocytes (Montoliu,

Sancho-Tello et al. 1995). There are also higher brain CYP2E1 levels in smokers and

in animals treated with nicotine and ethanol (Howard, Miksys et al. 2003, Schoedel

and Tyndale 2003, Miksys and Tyndale 2004, Joshi and Tyndale 2006, Yue, Khokhar

et al. 2009). Thus, variable levels of brain CYP2E1 could affect the local metabolism

of CYP2E1 substrates as well the health of brain cells.

116

4.4 Future directions

4.4.1 Other uses of rat models of differing levels of brain

CYP2D activity

4.4.1a Microdialysis

Microdialysis could be used to measure codeine and morphine levels in the brain in

vivo after peripheral codeine injection in rats that had received i.c.v. CYP2D

inhibitor treatment versus i.c.v. vehicle treatment. These drug levels can then be

correlated with analgesia in the tail-flick test. Microdialysis allows drug levels to be

measured at multiple time points in the same rat (whereas with the method used in

our study, rats had to be sacrificed at the time point being analyzed). Therefore,

microdialysis may provide more comprehensive data on the pharmacokinetics of

codeine and morphine in the brain and plasma over time, and may allow a more

precise understanding of the relative importance of brain versus hepatic CYP2D-

mediated metabolism in analgesia at different times after codeine injection.

4.4.1b Different pain model

It would be useful to determine if our results can be replicated using a different pain

model, such as the warm plate test. This consists of placing a rat onto a metallic plate

that is heated to a constant temperature by a thermode or hot liquid (Woolfe and

MacDonald 1944, Eddy and Leimbach 1953, O'Callaghan and Holtzman 1975). The

117

reaction times of two behaviours, paw licking and jumping, are measured, with an

increase in reaction time indicating analgesia. Both behaviours are under

supraspinal control, and both are modulated by opioids (Le Bars, Gozariu et al.

2001).

4.4.1c Role of rat brain CYP2D in meditating drug inactivation

While we have demonstrated that inhibiting brain CYP2D decreases the metabolic

activation of a prodrug and thereby decreases drug efficacy, future studies should

determine whether inhibiting brain CYP2D also decreases the metabolic inactivation

of a drug and results in enhanced drug effects. Examples of drugs that are

inactivated by CYP2D include amphetamine (stimulant, drug of abuse), desipramine

and imipramine (antidepressants).

4.4.1d Effect of rat brain CYP2D induction on drug response

While we have demonstrated the effects of inhibiting brain CYP2D on drug

metabolism and response, future studies should also determine whether inducing

brain CYP2D results in an increase in brain CYP2D-mediated drug activation or

inactivation and the corresponding increase or decrease in drug effect. Induction of

brain CYP2D (without altering hepatic CYP2D) can be achieved through chronic

nicotine treatment in rats as has been shown previously (Yue, Miksys et al. 2008),

and drug response can be tested in the resultant rat model of increased CYP2D

118

activity. This model may reflect the elevated levels of brain CYP2D in human

smokers and be used to elucidate the altered response to centrally-acting CYP2D

substrates observed in these individuals.

4.4.1e Role of rat brain CYP2D in neurotoxin inactivation

While studies suggest that brain CYP2D protects against Parkinson’s disease (PD), it

is not known how important a role brain CYP2D plays in this function in vivo.

Manipulating brain CYP2D levels using inhibitors and inducers in an in vivo rat

model of PD (e.g., neurotoxicity caused by MPTP or MPP+ exposure) could be used

to determine the influence of brain CYP2D on the metabolism and effects of

Parkinsonian neurotoxins.

4.4.2 Therapeutic uses of brain CYP2D induction

Future studies should evaluate the efficacy of increasing brain CYP2D6 activity as a

new therapeutic technique for protecting against neurotoxicity (e.g. from MPTP or

MPP+). The induction of brain CYP2D6 would be valuable in circumstances where

low CYP2D6 levels may increase the risk of disease, such as the case with PD.

Nicotine may be used as a preventative measure to induce brain CYP2D6 and

increase local neurotoxin inactivation in brain regions affected by PD (e.g., striatum

and substantia nigra). This may lower the risk for, or delay the onset of, PD. Nicotine

may also be used after the onset of PD to help ameliorate symptoms (Kelton, Kahn et

119

al. 2000, Vieregge, Sieberer et al. 2001). Nicotine’s induction of brain CYP2D6 may

increase the CYP2D6-mediated synthesis of dopamine, which may help compensate

for the loss of dopamine in PD.

4.5 Conclusions

CYP2D6 is an important enzyme that metabolizes a wide variety of drugs, toxins and

endogenous substrates. CYP2D6 is genetically polymorphic, and the expression and

activity of CYP2D in the brain can be altered by environmental inducers and

increases in age, as demonstrated in animal and human studies. Using a rat model of

reduced brain CYP2D activity, we have shown that inhibiting brain CYP2D results in

lower brain morphine levels and less analgesia after codeine administration,

suggesting that brain CYP2D has an important role in the metabolism and effect of

codeine, a centrally-acting CYP2D substrate. Therefore, brain CYP2D6 may have a

significant impact on response to codeine (including analgesia and abuse liability)

and other drugs metabolized by CYP2D6 (e.g. antidepressants, antipsychotics).

Differences in brain CYP2D6 activity, whether through genetic variation, exposure to

alcohol or nicotine, or age, may contribute to the interindividual variation in

therapeutic efficacy and side effect profiles of centrally-acting drugs metabolized by

CYP2D6. The expression of multiple families of CYPs in the brain, and their diverse

array of centrally-acting substrates, supports the brain as being an important organ

in the metabolism of drugs, toxins and endogenous substrates.

120

References

Abou-Donia, M. B., A. Abdel-Rahman, L.B. Goldstein, A.M. Dechkovskaia, D.U. Shah,

S.L. Bullman and W.A. Khan (2003). "Sensorimotor deficits and increased brain

nicotinic acetylcholine receptors following exposure to chlorpyrifos and/or nicotine

in rats." Arch Toxicol 77(8): 452-458.

Abraham, J. E., M. J. Maranian, K. E. Driver, R. Platte, B. Kalmyrzaev, C. Baynes, C.

Luccarini, M. Shah, S. Ingle, D. Greenberg, H. M. Earl, A. M. Dunning, P. D. P.

Pharoah and C. Caldas (2010). "CYP2D6 gene variants: association with breast

cancer specific survival in a cohort of breast cancer patients from the United

Kingdom treated with adjuvant tamoxifen." Breast Cancer Res 12(4): R64.

Adler, T. K., J. M. Fujimoto, E. L. Way and E. M. Baker (1955). "The metabolic fate of

codeine in man." J Pharmacol Exp Ther 114(3): 251-262.

Agundez, J. A. G., M. C. Ledesma, J. M. Ladero and J. Benitez (1995). "Prevalence of

CYP2D6 gene duplication and its repercussion on the oxidative phenotype in a white

population." Clin Pharmacol Ther 57(3): 265-269.

Ahlner, J., A. L. Zackrisson, B. Lindblom and L. Bertilsson (2010). "CYP2D6, serotonin

and suicide." Pharmacogenomics 11(7): 903-905.

Ai, C. Z., Y. Li, Y. H. Wang, Y. D. Chen and L. Yang (2009). "Insight into the effects of

chiral isomers quinidine and quinine on CYP2D6 inhibition." Bioorg Med Chem Lett

19(3): 803-806.

Akil, H., C. Owens, H. Gutstein, L. Taylor, E. Curran and S. Watson (1998).

"Endogenous opioids: overview and current issues." Drug Alcohol Depend 51(1-2):

127-140.

Aklillu, E., I. Persson, L. Bertilsson, I. Johansson, F. Rodrigues and M.

IngelmanSundberg (1996). "Frequent distribution of ultrarapid metabolizers of

debrisoquine in an Ethiopian population carrying duplicated and multiduplicated

functional CYP2D6 alleles." J Pharmacol Exp Ther 278(1): 441-446.

Albores, A., G. Ortega-Mantilla, A. Sierra-Santoyo, M. E. Cebrian, J. L. Munoz-

Sanchez, J. V. Calderon-Salinas and M. Manno (2001). "Cytochrome P450 2B

(CYP2B)-mediated activation of methyl-parathion in rat brain extracts." Toxicol Lett

124(1-3): 1-10.

Alves, G., M. Kurz, S.A. Lie and J.P. Larsen (2004). "Cigarette smoking in Parkinson's

disease: influence on disease progression." Mov Disord 19(9): 1087-1092.

Amabeoku, G., R. Ewesuedo and F. Abuh (1992). "Influence of some dopaminergic

agents on antinociception produced by quinine in mice." Prog

Neuropsychopharmacol Biol Psychiatry 16(3): 351-360.

121

Amann, T., P.H. Roos, H. Huh and M.H. Zenk (1995). "Purifi cation and

characterization of a cytochrome P450 enzyme from pig liver, catalyzing the phenol

oxidative coupling of (R)-reticuline to salutaridine, the critical step in morphine

biosynthesis." Heterocycles 40: 425–440

Archer, S. and L. S. Harris (1965). "Narcotic antagonists." Fortschr Arzneimittelforsch

8: 261-320.

Arvidsson, U., M. Riedl, S. Chakrabarti, J. H. Lee, A. H. Nakano, R. J. Dado, H. H. Loh,

P. Y. Law, M. W. Wessendorf and R. Elde (1995). "Distribution and targeting of a mu-

opioid receptor (MOR1) in brain and spinal cord." J Neurosci 15(5 Pt 1): 3328-3341.

Atweh, S. F. and M. J. Kuhar (1977). "Autoradiographic localization of opiate

receptors in rat brain. I. Spinal cord and lower medulla." Brain Res 124(1): 53-67.

Bambico, F.R. and G. Gobbi (2008). "The cannabinoid CB1 receptor and the

endocannabinoid anandamide: possible antidepressant targets." Expert Opin Ther

Targets 12(11): 1347-1366.

Barbaro, N. M., M. M. Heinricher and H. L. Fields (1989). "Putative nociceptive

modulatory neurons in the rostral ventromedial medulla of the rat display highly

correlated firing patterns." Somatosens Mot Res 6(4): 413-425.

Basbaum, A. I., D. M. Bautista, G. Scherrer and D. Julius (2009). "Cellular and

Molecular Mechanisms of Pain." Cell 139(2): 267-284.

Basbaum, A. I., C. H. Clanton and H. L. Fields (1976). "Opiate and stimulus produced

analgesia: Functional anatomy of a medullospinal pathway." Proc Natl Acad Sci USA

73(12): 4685-4688.

Basbaum, A. I. and T. Jessell (2000). The perception of pain. Principles of

Neuroscience. E. R. Kandel, J. Schwartz and T. Jessell. New York, Appleton and

Lange 472–491.

Basbaum, A. I., N. J. E. Marley, J. Okeefe and C. H. Clanton (1977). "Reversal of

morphine and stimulus-produced analgesia by subtotal spinal cord lesions." Pain

3(1): 43-56.

Bergh, A. F. and H. W. Strobel (1992). "Reconstitution of the brain mixed function

oxidase system: purification of NADPH-cytochrome P450 reductase and partial

purification of cytochrome P450 from whole rat brain." J Neurochem 59(2): 575-581.

Bergh, A. F. and H. W. Strobel (1996). "Anatomical distribution of NADPH-

cytochrome P450 reductase and cytochrome P4502D forms in rat brain: Effects of

xenobiotics and sex steroids." Mol Cell Biochem 162(1): 31-41.

Bertelsen, K. M., K. Venkatakrishnan, L. L. Von Moltke, R. S. Obach and D. J.

Greenblatt (2003). "Apparent mechanism-based inhibition of human CYP2D6 in vitro

122

by paroxetine: Comparison with fluoxetine and quinidine." Drug Metab Dispos

31(3): 289-293.

Bertilsson, L., C. Alm, C. De Las Carreras, J. Widen, G. Edman and D. Schalling

(1989). "Debrisoquine hydroxylation polymorphism and personality." Lancet

1(8637): 555.

Bhagwat, S. V., M. R. Boyd and V. Ravindranath (2000). "Multiple forms of

cytochrome P450 and associated monooxygenase activities in human brain

mitochondria." Biochem Pharmacol 59(5): 573-582.

Bhamre, S., H. K. Anandatheerathavarada, S. K. Shankar, M. R. Boyd and V.

Ravindranath (1993). "Purification of multiple forms of cytochrome P450 from a

human brain and reconstitution of catalytic activities." Arch Biochem Biophys 301(2):

251-255.

Bianchetti, G., J. L. Elghozi, R. Gomeni, P. Meyer and P. L. Morselli (1980). "Kinetics

of distribution of dl-propranolol in various organs and discrete brain areas of the

rat." J Pharmacol Exp Ther 214(3): 682-687.

Bijl, M. J., H. J. Luijendijk, J.F. van den Berg, L.E. Visser, R.H. van Schaik, A. Hofman,

A.G. Vulto, T. van Gelder, H. Tiemeier and B.H. Stricker. (2009). "Association

between the CYP2D6*4 polymorphism and depression or anxiety in the elderly."

Pharmacogenomics 10(4): 541-547.

Blume, N., J. Leonard, Z. J. Xu, O. Watanabe, H. Remotti and J. Fishman (2000).

"Characterization of Cyp2d22, a novel cytochrome P450 expressed in mouse

mammary cells." Arch Biochem Biophys 381(2): 191-204.

Bouw, M. R., M. Gardmark and M. Hammarlund-Udenaes (2000). "Pharmacokinetic-

pharmacodynamic modelling of morphine transport across the blood-brain barrier

as a cause of the antinociceptive effect delay in rats - A microdialysis study." Pharm

Res 17(10): 1220-1227.

Britto, M. R. and P. J. Wedlund (1992). "Cytochrome P-450 in the brain. Potential

evolutionary and therapeutic relevance of localization of drug-metabolizing

enzymes." Drug Metab Dispos 20(3): 446-450.

Bromek, E., A. Haduch and W. A. Daniel (2010). "The ability of cytochrome P450 2D

isoforms to synthesize dopamine in the brain: An in vitro study." Eur J Pharmacol

626(2-3): 171-178.

Brosen, K., L. F. Gram, T. Haghfelt and L. Bertilsson (1987). "Extensive metabolizers

of debrisoquine become poor metabolizers during quinidine treatment." Pharmacol

Toxicol 60(4): 312-314.

Cairns, W., C. A. D. Smith, A. W. McLaren and C. R. Wolf (1996). "Characterization of

the human cytochrome P4502D6 promoter - A potential role for antagonistic

123

interactions between members of the nuclear receptor family." J Biol Chem 271(41):

25269-25276.

Camarata, P. J. and T. L. Yaksh (1985). "Characterization of the spinal adrenergic

receptors mediating the spinal effects produced by the microinjection of morphine

into the periaqueductal gray." Brain Res 336(1): 133-142.

Caraco, Y., T. Tateishi, F. P. Guengerich and A. J. Wood (1996). "Microsomal codeine

N-demethylation: cosegregation with cytochrome P4503A4 activity." Drug Metab

Dispos 24(7): 761-764.

Caro, A. A. and A. I. Cederbaum (2004). "Oxidative stress, toxicology, and

pharmacology of CYP2E1." Ann Rev Pharmacol Toxicol 44: 27-42.

Chambers, J. E. and H. W. Chambers (1989). "Oxidative desulfuration of

chlorpyrifos, chlorpyrifos-methyl, and leptophos by rat brain and liver." J Biochem

Toxicol 4(3): 201-203.

Chen, Z. R., R. J. Irvine, F. Bochner and A. A. Somogyi (1990). "Morphine formation

from codeine in rat brain: a possible mechanism of codeine-induced analgesia." Life

Sci 46(15): 1067-1074.

Chen, Z. R., A. A. Somogyi, G. Reynolds and F. Bochner (1991). "Disposition and

metabolism of codeine after single and chronic doses in one poor and seven

extensive metabolisers." Br J Clin Pharmacol 31(4): 381-390.

Chimbira, W. and B. P. Sweeney (2000). "The effect of smoking on postoperative

nausea and vomiting." Anaesthesia 55(6): 540-544.

Cleary, J., G. Mikus, A. Somogyi and F. Bochner (1994). "The influence of

pharmacogenetics on opioid analgesia: studies with codeine and oxycodone in the

Sprague-Dawley/Dark Agouti rat model." J Pharmacol Exp Ther 271(3): 1528-1534.

Coffman, B. L., G. R. Rios, C. D. King and T. R. Tephly (1997). "Human UGT2B7

catalyzes morphine glucuronidation." Drug Metab Dispos 25(1): 1-4.

Coleman, T., E. F. Spellman, A. Rostami-Hodjegan, M. S. Lennard and G. T. Tucker

(2000). "The 1 '-hydroxylation of rac-bufuralol by rat brain microsomes." Drug Metab

Dispos 28(9): 1094-1099.

Conroy, J. L., C. Fang, J. Gu, S. O. Zeitlin, W. Z. Yang, J. Yang, M. A. VanAlstine, J. W.

Nalwalk, P. J. Albrecht, J. E. Mazurkiewicz, A. Snyder-Keller, Z. X. Shan, S. Z. Zhang,

M. P. Wentland, M. Behr, B. I. Knapp, J. M. Bidlack, O. P. Zuiderveld, R. Leurs, X. X.

Ding and L. B. Hough (2010). "Opioids activate brain analgesic circuits through

cytochrome P450/epoxygenase signaling." Nat Neurosci 13(3): 284-286.

Coon, M. J. (2005). Cytochrome P450: Nature's most versatile biological catalyst.

Annu Rev Pharmacol Toxicol. 45: 1-25.

124

Corchero, J., C. P. Granvil, T. E. Akiyama, G. P. Hayhurst, S. Pimprale, L.

Feigenbaum, J. R. Idle and F. J. Gonzalez (2001). "The CYP2D6 humanized mouse:

effect of the human CYP2D6 transgene and HNF4alpha on the disposition of

debrisoquine in the mouse." Mol Pharmacol 60(6): 1260-1267.

Court, M. H., S. X. Duan, L.M. Hesse, K. Venkatakrishnan and D.J. Greenblatt (2001).

"Cytochrome P-450 2B6 is responsible for interindividual variability of propofol

hydroxylation by human liver microsomes." Anesthesiology 94(1): 110-119.

Court, M. H., S. Krishnaswamy, Q. Hao, S. X. Duan, C. J. Patten, L. L. Von Moltke and

D. J. Greenblatt (2003). "Evaluation of 3'-azido-3'-deoxythymidine, morphine, and

codeine as probe substrates for UDP-glucuronosyltransferase 2B7 (UGT2B7) in

human liver microsomes: specificity and influence of the UGT2B7*2 polymorphism."

Drug Metab Dispos 31(9): 1125-1133.

D'Amour, F. E. and D. L. Smith (1941). "A method for determining loss of pain

sensation." J Pharmacol Exp Ther 72(1): 74-79.

Dahl, M. L., I. Johansson, L. Bertilsson, M. Ingelmansundberg and F. Sjoqvist (1995).

"Ultrarapid hydroxylation of debrisoquine in a Swedish population. Analysis of the

molecular genetic basis." J Pharmacol Exp Ther 274(1): 516-520.

Dauchy, S., F. Dutheil, R. J. Weaver, F. Chassoux, C. Daumas-Duport, P. O. Couraud,

J. M. Scherrmann, I. De Waziers and X. Decleves (2008). "ABC transporters,

cytochromes P450 and their main transcription factors: expression at the human

blood-brain barrier." J Neurochem 107(6): 1518-1528.

Dauchy, S., F. Miller, P. O. Couraud, R. J. Weaver, B. Weksler, I. A. Romero, J. M.

Scherrmann, I. De Waziers and X. Decleves (2009). "Expression and transcriptional

regulation of ABC transporters and cytochromes P450 in hCMEC/D3 human cerebral

microvascular endothelial cells." Biochem Pharmacol 77(5): 897-909.

De Gregori, M., M. Allegri, S. De Gregori, G. Garbin, C. Tinelli, M. Regazzi, S.

Govoni and G. N. Ranzani (2010). "How and Why to Screen for CYP2D6

Interindividual Variability in Patients Under Pharmacological Treatments." Curr

Drug Metab 11(3): 276-282.

de Groot, M. J., F. Wakenhut, G. Whitlock and R. Hyland (2009). "Understanding

CYP2D6 interactions." Drug Discov Today 14(19-20): 964-972.

de Leon, J., M. T. Susce, R. M. Pan, M. Fairchild, W. H. Koch and P. J. Wedlund (2005).

"The CYP2D6 poor metabolizer phenotype may be associated with risperidone

adverse drug reactions and discontinuation." J Clin Psychiatry 66(1): 15-27.

Deng, Y., B. Newman, M.P. Dunne, P.A. Silburn and G.D. Mellick (2004). "Further

evidence that interactions between CYP2D6 and pesticide exposure increase risk for

Parkinson's disease." Ann Neurol 55(6): 897.

125

Di Monte, D.A. (2003). "The environment and Parkinson's disease: is the nigrostriatal

system preferentially targeted by neurotoxins?" Lancet Neurol 2(9): 531-538.

Ding, X. X. and L. S. Kaminsky (2003). "Human extrahepatic cytochromes P450:

Function in xenobiotic metabolism and tissue-selective chemical toxicity in the

respiratory and gastrointestinal tracts." Ann Rev Pharmacol Toxicol 43: 149-173.

Doostzadeh, J. and R. Morfin (1997). "Effects of cytochrome P450 inhibitors and of

steroid hormones on the formation of 7-hydroxylated metabolites of pregnenolone in

mouse brain microsomes." J Endocrinol 155(2): 343-350.

Dorado, P., E. M. Penas-Lledo and A. Llerena (2007). "CYP2D6 polymorphism:

implications for antipsychotic drug response, schizophrenia and personality traits."

Pharmacogenomics 8(11): 1597-1608.

Dostrovsky, J.O. (2000). "Role of thalamus in pain." Prog Brain Res 129: 245-257.

Dukes, I. D. and E. M. Vaughan Williams (1984). "The multiple modes of action of

propafenone." Eur Heart J 5(2): 115-125.

Duman, E. N., M. Kesim, M. Kadioglu, E. Yaris, N. I. Kalyoncu and N. Erciyes (2004).

"Possible involvement of opioidergic and serotonergic mechanisms in

antinociceptive effect of paroxetine in acute pain." J Pharmacol Sci 94(2): 161-165.

Dutheil, F., P. Beaune and M. A. Loriot (2008). "Xenobiotic metabolizing enzymes in

the central nervous system: Contribution of cytochrome P450 enzymes in normal and

pathological human brain." Biochimie 90(3): 426-436.

Dutheil, F., S. Dauchy, M. Diry, V. Sazdovitch, O. Cloarec, L. Mellottee, I. Bieche, M.

Ingelman-Sundberg, J. P. Flinois, I. de Waziers, P. Beaune, X. Decleves, C.

Duyckaerts and M. A. Loriot (2009). "Xenobiotic-Metabolizing Enzymes and

Transporters in the Normal Human Brain: Regional and Cellular Mapping as a Basis

for Putative Roles in Cerebral Function." Drug Metab Dispos 37(7): 1528-1538.

Dutheil, F., A. Jacob, S. Dauchy, P. Beaune, J. M. Scherrmann, X. Decleves and M. A.

Loriot (2010). "ABC transporters and cytochromes P450 in the human central nervous

system: influence on brain pharmacokinetics and contribution to neurodegenerative

disorders." Expert Opin Drug Metab Toxicol 6(10): 1161-1174.

Eddy, N. B. and D. Leimbach (1953). "Synthetic analgesics. II. Dithienylbutenyl- and

dithienylbutylamines." J Pharmacol Exp Ther 107(3): 385-393.

Edwards, R. J., R. J. Price, P. S. Watts, A. B. Renwick, J. M. Tredger, A. R. Boobis and

B. G. Lake (2003). "Induction of cytochrome P450 enzymes in cultured precision-cut

human liver slices." Drug Metab Dispos 31(3): 282-288.

Elbaz, A., C. Levecque, J. Clavel, J.S. Vidal, F. Richard, P. Amouyel, A. Alperovitch,

M.C. Chartier-Harlin and C. Tzourio (2004). "CYP2D6 polymorphism, pesticide

exposure, and Parkinson's disease." Ann Neurol 55(3): 430-434.

126

Emoto, C., S. Murase and K. Iwasaki (2006). "Approach to the prediction of the

contribution of major cytochrome P450 enzymes to drug metabolism in the early

drug-discovery stage." Xenobiotica 36(8): 671-683.

Erjavec, M. K., B. A. Coda, Q. Nguyen, G. Donaldson, L. Risler and D. D. Shen (2000).

"Morphine-fluoxetine interactions in healthy volunteers: Analgesia and side effects."

J Clin Pharmacol 40(11): 1286-1295.

Estabrook, R. (1999). "An introduction to the cytochrome P450s." Mol Aspects Med

20(1-2): 5-12, 13-137.

Fagerstrom, K.O., O. Pomerleau, B. Giordani and F. Stelson (1994). "Nicotine may

relieve symptoms of Parkinson's disease." Psychopharmacology (Berl) 116(1): 117-

119.

Fahn, S. (2010). "Parkinson's disease: 10 years of progress, 1997-2007." Mov Disord

25 Suppl 1: S2-S14.

Farre, M. and J. Cami (1991). "Pharmacokinetic considerations in abuse liability

evaluation." Br J Addict 86(12): 1601-1606.

Fields, H. (2004). "State-dependent opioid control of pain." Nat Rev Neurosci 5(7):

565-575.

Fields, H. L., P. C. Emson, B. K. Leigh, R. F. T. Gilbert and L. L. Iversen (1980).

"Multiple opiate receptor sites on primary afferent fibers." Nature 284(5754): 351-

353.

Fields, H. L., M. M. Heinricher and P. Mason (1991). "Neurotransmitters in

nociceptive modulatory circuits." Annu Rev Neurosci 14: 219-245.

Fields, H. L., A. Malick and R. Burstein (1995). "Dorsal horn projection targets of on

and off cells in the rostral ventromedial medulla." J Neurophysiol 74(4): 1742-1759.

Flores-Perez, J., C. Flores-Perez, H. Juarez-Olguin, I. Lares-Asseff and M. Sosa-

Macias (2004). "Determination of dextromethorphan and dextrorphan in human

urine by high performance liquid chromatography for pharmacogenetic

investigations." Chromatographia 59(7-8): 481-485.

Fonne-Pfister, R., M. J. Bargetzi and U. A. Meyer (1987). "MPTP, the neurotoxin

inducing Parkinson's disease, is a potent competitive inhibitor of human and rat

cytochrome P450 isozymes (P450bufI, P450db1) catalyzing debrisoquine 4-

hydroxylation." Biochem Biophys Res Commun 148(3): 1144-1150.

Forsyth, C. S. and J. E. Chambers (1989). "Activation and degradation of the

phosphorothionate insecticides parathion and EPN by rat brain." Biochem Pharmacol

38(10): 1597-1603.

127

Fradette, C., N. Yamaguchi and P. du Souich (2004). "5-hydroxytryptamine is

biotransformed by CYP2C9, 2C19 and 2B6 to hydroxylamine, which is converted

into nitric oxide." Br J Pharmacol 141(3): 407-414.

Frank, D., U. Jaehde and U. Fuhr (2007). "Evaluation of probe drugs and

pharmacokinetic metrics for CYP2D6 phenotyping." Eur J Clin Pharmacol 63(4): 321-

333.

Freiermuth, M. and J. C. Plasse (1997). "Determination of morphine and codeine in

plasma by HPLC following solid phase extraction." J Pharm Biomed Anal 15(6): 759-

764.

Funae, Y., W. Kishimoto, T. Cho, T. Niwa and T. Hiroi (2003). "CYP2D in the brain."

Drug Metab Pharmacokinet 18(6): 337-349.

Funck-Brentano, C., P. Y. Boelle, C. Verstuyft, C. Bornert, L. Becquemont and J. M.

Poirier (2005). "Measurement of CYP2D6 and CYP3A4 activity in vivo with

dextromethorphan: sources of variability and predictors of adverse effects in 419

healthy subjects." Eur J Clin Pharmacol 61(11): 821-829.

Gaedigk, A., A. Bhathena, L. Ndjountche, R. E. Pearce, S. M. Abdel-Rahman, S. W.

Alander, L. D. Bradford and J. S. Leeder (2005). "Identification and characterization of

novel sequence variations in the cytochrome P4502D6 (CYP2D6) gene in African

Americans." Pharmacogenomics J 5(3): 173-182.

Gaedigk, A., S. D. Simon, R. E. Pearce, L. D. Bradford, M. J. Kennedy and J. S. Leeder

(2008). "The CYP2D6 activity score: Translating genotype information into a

qualitative measure of phenotype." Clin Pharmacol Ther 83(2): 234-242.

Gan, S.H., R. Ismail, W.A. Wan Adnan, W. Zulmi, N. Kumaraswamy and E.T. Larmie

(2004) "Relationship between Type A and B personality and debrisoquine

hydroxylation capacity." Br J Clin Pharmacol 57(6): 785-789.

Garcia-Barcelo, M., L. Y. Chow, H. F. K. Chiu, Y. K. Wing, D. T. S. Lee, K. L. Lam and

M. M. Y. Waye (2000). "Genetic analysis of the CYP2D6 locus in a Hong Kong

Chinese population." Clin Chem 46(1): 18-23.

Gasche, Y., Y. Daali, M. Fathi, A. Chiappe, S. Cottini, P. Dayer and J. Desmeules

(2004). "Codeine intoxication associated with ultrarapid CYP2D6 metabolism." N

Engl J Med 351(27): 2827-2831.

Genazzani, A.R., M. Stomati, A. Morittu, F. Bernardi, P. Monteleone, E. Casarosa, R.

Gallo, C. Salvestroni and M. Luisi (2000). "Progesterone, progestagens and the

central nervous system." Hum Reprod 15 Suppl 1: 14-27.

George, T. P., K. A. Sacco, J. C. Vessicchio, A. H. Weinberger and R. D. Shytle

(2008). "Nicotinic antagonist augmentation of selective serotonin reuptake inhibitor-

128

refractory major depressive disorder - A preliminary study." J Clin

Psychopharmacol 28(3): 340-344.

Gervasini, G., J. A. Carrillo and J. Benitez (2004). "Potential role of cerebral

cytochrome P450 in clinical pharmacokinetics - Modulation by endogenous

compounds." Clin Pharmacokinet 43(11): 693-706.

Ghersiegea, J. F., A. Minn, J. L. Daval, Z. Jayyosi, V. Arnould, H. Souhailielamri and

G. Siest (1989). "NADPH:cytochrome P-450(c) reductase: biochemical

characterization in rat brain and cultured neurons and evolution of activity during

development." Neurochem Res 14(9): 883-888.

Ghersiegea, J. F., R. Perrin, B. Leiningermuller, M. C. Grassiot, C. Jeandel, J. Floquet,

G. Cuny, G. Siest and A. Minn (1993). "Subcellular localization of cytochrome P450,

and activities of several enzymes responsible for drug metabolism in the human

brain." Biochem Pharmacol 45(3): 647-658.

Ghosh, C., J. Gonzalez-Martinez, M. Hossain, L. Cucullo, V. Fazio, D. Janigro and N.

Marchi (2010). "Pattern of P450 expression at the human blood-brain barrier: roles of

epileptic condition and laminar flow." Epilepsia 51(8): 1408-1417.

Gilham, D. E., W. Cairns, M. J. I. Paine, S. Modi, R. Poulsom, G. C. K. Roberts and C.

R. Wolf (1997). "Metabolism of MPTP by cytochrome P4502D6 and the demonstration

of 2D6 mRNA in human foetal and adult brain by in situ hybridization." Xenobiotica

27(1): 111-125.

Gonzalez, F. J. (1990). "Molecular genetics of the P-450 superfamily." Pharmacology

& Therapeutics 45(1): 1-38.

Gonzalez, I., E. M. Penas-Lledo, B. Perez, P. Dorado, M. Alvarez and L. L. A (2008).

"Relation between CYP2D6 phenotype and genotype and personality in healthy

volunteers." Pharmacogenomics 9(7): 833-840.

Gonzalez-Hernandez, T., I. Cruz-Muros, D. Afonso-Oramas, J. Salas-Hernandez and J.

Castro-Hernandez (2010). "Vulnerability of mesostriatal dopaminergic neurons in

Parkinson's disease." Front Neuroanat 4: 140.

Griffiths, R. R. and B. Wolf (1990). "Relative abuse liability of different

benzodiazepines in drug abusers." J Clin Psychopharmacol 10(4): 237-243.

Grossman, M. L., A. I. Basbaum and H. L. Fields (1982). "Afferent and efferent

connections of the rat tail-flick reflex (a model used to analyze pain control

mechanisms)." J Comp Neurol 206(1): 9-16.

Grumbach, L. (1966). The prediction of analgesic activity in man by animal testing. .

Pain, 15th International Symposium, Detroit, 1964. R. S. Knighton and P. R. Dumke.

Boston, Little Brown: 163–182.

129

Grumbach, L. (1975). Pain: Research and Treatment Further Observations From City

of Hope National Medical Centre. B. L. Crue. New York Academic Press: 163-182.

Guengerich, F. P. (2003). "Cytochromes P450, drugs, and diseases." Mol Interv 3(4):

194-204.

Haduch, A., E. Bromek and W. A. Daniel (2011). "The effect of psychotropic drugs on

cytochrome P450 2D (CYP2D) in rat brain." Eur J Pharmacol 651(1-3): 51-58.

Haglund, L., C. Kohler, T. Haaparanta, M. Goldstein and J. A. Gustafsson (1984).

"Presence of NADPH-cytochrome P450 reductase in central catecholaminergic

neurones." Nature 307(5948): 259-262.

Haining, R. L. (2007). Cytochrome P450 Reactions in the Human Brain. Handbook of

Neurochemistry and Molecular Neurobiology. A. Lajtha. New York, NY, Springer

Science 43-91.

Hammond, D. L., G. M. Tyce and T. L. Yaksh (1985). "Efflux of 5-hydroxytryptamine

and noradrenaline into spinal cord superfusates during stimulation of the rat

medulla." J Physiol 359: 151-162.

Hardy, J. D. (1953). "Thresholds of Pain and Reflex Contraction as Related to Noxious

Stimulation." J Appl Physiol 5(12): 725-739.

Hardy, J. D., A. M. Stoll, D. Cunningham, W. M. Benson and L. Greene (1957).

"Responses of the rat to thermal radiation." Am J Physiol 189(1): 1-5.

He, H., S. D. Shay, Y. Caraco, M. Wood and A. J. Wood (1998). "Simultaneous

determination of codeine and its seven metabolites in plasma and urine by high-

performance liquid chromatography with ultraviolet and electrochemical detection."

J Chromatogr B Biomed Sci Appl 708(1-2): 185-193.

Hedlund, E., J. A. Gustafsson and M. Warner (2001). "Cytochrome P450 in the brain;

a review." Curr Drug Metab 2(3): 245-263.

Hedlund, E., A. Wyss, T. Kainu, M. Backlund, C. Kohler, M. PeltoHuikko, J. A.

Gustafsson and M. Warner (1996). "Cytochrome P4502D4 in the brain: Specific

neuronal regulation by clozapine and toluene." Mol Pharmacol 50(2): 342-350.

Heinricher, M. M., N. M. Barbaro and H. L. Fields (1989). "Putative nociceptive

modulating neurons in the rostral ventromedial medulla of the rat: firing of on- and

off-cells is related to nociceptive responsiveness." Somatosens Mot Res 6(4): 427-

439.

Heinricher, M. M., Z. F. Cheng and H. L. Fields (1987). "Evidence for two classes of

nociceptive modulating neurons in the periaqueductal gray." J Neurosci 7(1): 271-

278.

130

Heinricher, M. M., C. M. Haws and H. L. Fields (1991). "Evidence for GABA-mediated

control of putative nociceptive modulating neurons in the rostral ventromedial

medulla: iontophoresis of bicuculline eliminates the off-cell pause." Somatosens Mot

Res 8(3): 215-225.

Heinricher, M. M. and S. L. Ingram (2008). The brainstem and nociceptive

modulation. The Senses, a Comprehensive Reference. Pain, Vol. 5. M. C. Bushnell

and A. I. Basbaum. San Diego, Academic Press: 593–626.

Heinricher, M. M., S. McGaraughty and D. A. Farr (1999). "The role of excitatory

amino acid transmission within the rostral ventromedial medulla in the

antinociceptive actions of systemically administered morphine." Pain 81(1-2): 57-65.

Heinricher, M. M., S. McGaraughty and D. K. Grandy (1997). "Circuitry underlying

antiopioid actions of orphanin FQ in the rostral ventromedial medulla." J

Neurophysiol 78(6): 3351-3358.

Heinricher, M. M., S. McGaraughty and V. Tortorici (2001). "Circuitry underlying

antiopioid actions of cholecystokinin within the rostral ventromedial medulla." J

Neurophysiol 85(1): 280-286.

Heinricher, M. M., M. M. Morgan and H. L. Fields (1992). "Direct and indirect actions

of morphine on medullary neurons that modulate nociception." Neuroscience 48(3):

533-543.

Heinricher, M. M., M. M. Morgan, V. Tortorici and H. L. Fields (1994). "Disinhibition

of off-cells and antinociception produced by an opioid action within the rostral

ventromedial medulla." Neuroscience 63(1): 279-288.

Heinricher, M. M., I. Tavares, J. L. Leith and B. M. Lumb (2009). "Descending control

of nociception: Specificity, recruitment and plasticity." Brain Res Rev 60(1): 214-225.

Hendrickson, H. P., B. J. Gurley and W. D. Wessinger (2003). "Determination of

dextromethorphan and its metabolites in rat serum by liquid-liquid extraction and

liquid chromatography with fluorescence detection." J Chromatogr B Analyt Technol

Biomed Life Sci 788(2): 261-268.

Herraiz, T., H. Guillen, V.J. Aran, J.R. Idle and F.J. Gonzalez (2006). "Comparative

aromatic hydroxylation and N-demethylation of MPTP neurotoxin and its analogs, N-

methylated betacarboline and isoquinoline alkaloids, by human cytochrome P450

2D6." Toxicol Appl Pharmacol 216(3): 387-398.

Hiroi, T., T. Chow, S. Imaoka and Y. Funae (2002). "Catalytic specificity of CYP2D

isoforms in rat and human." Drug Metab Dispos 30(9): 970-976.

Hiroi, T., S. Imaoka and Y. Funae (1998). "Dopamine formation from tyramine by

CYP2D6." Biochem Biophys Res Commun 249(3): 838-843.

131

Hiroi, T., W. Kishimoto, T. Chow, S. Imaoka, T. Igarashi, and Y. Funae (2001).

"Progesterone oxidation by cytochrome P450 2D isoforms in the brain."

Endocrinology, 142(2): 3901-3908.

Hlatky, M.A., D. Boothroyd, E. Vittinghoff, P. Sharp and M.A. Whooley (2002).

"Quality-of life and depressive symptoms in postmenopausal women after receiving

hormone therapy: results from the Heart and Estrogen/Progestin Replacement Study

(HERS) trial." JAMA, 287(5) 591-597.

Hohmann, A. G., R. L. Suplita, N. M. Bolton, M. H. Neely, D. Fegley, R. Mangieri, J. F.

Krey, J. M. Walker, P. V. Holmes, J. D. Crystal, A. Duranti, A. Tontini, M. Mor, G.

Tarzia and D. Piomelli (2005). "An endocannabinoid mechanism for stress-induced

analgesia." Nature 435(7045): 1108-1112.

Holthe, M., P. Klepstad, K. Zahlsen, P. C. Borchgrevink, L. Hagen, O. Dale, S. Kaasa,

H. E. Krokan and F. Skorpen (2002). "Morphine glucuronide-to-morphine plasma

ratios are unaffected by the UGT2B7 H268Y and UGT1A1*28 polymorphisms in

cancer patients on chronic morphine therapy." Eur J Clin Pharmacol 58(5): 353-356.

Howard, L. A., S. Miksys, E. Hoffmann, D. Mash and R. F. Tyndale (2003). "Brain

CYP2E1 is induced by nicotine and ethanol in rat and is higher in smokers and

alcoholics." Br J Pharmacol 138(7): 1376-1386.

Ingelman-Sundberg, M. (2005). "Genetic polymorphisms of cytochrome P450 2D6

(CYP2D6): clinical consequences, evolutionary aspects and functional diversity."

Pharmacogenomics J 5(1): 6-13.

Ingelman-Sundberg, M., S. C. Sim, A. Gomez and C. Rodriguez-Antona (2007).

"Influence of cytochrome P450 polymorphisms on drug therapies: Pharmacogenetic,

pharmacoepigenetic and clinical aspects." Pharmacol Ther 116(3): 496-526.

Irwin, S., R. W. Houde, D. R. Bennett, L. C. Hendershot and M. H. Seevers (1951).

"The effects of morphine, methadone and meperidine on some reflex responses of

spinal animals to nociceptive stimulation." J Pharmacol Exp Ther 101(2): 132-143.

Jabs, B. E., A. J. Bartsch and B. Pfuhlmann (2003). "Susceptibility to neuroleptic-

induced parkinsonism - age and increased substantia nigra echogenicity as putative

risk factors." Eur Psychiatry 18(4): 177-181.

Jacob, P., 3rd, M. Ulgen and J. W. Gorrod (1997). "Metabolism of (-)-(S)-nicotine by

guinea pig and rat brain: identification of cotinine." Eur J Drug Metab Pharmacokinet

22(4): 391-394.

Jacquet, Y. F. and A. Lajtha (1973). "Morphine action at central nervous system sites

in rat: analgesia or hyperalgesia depending on site and dose." Science 182(4111):

490-492.

132

Jensen, T. S. and T. L. Yaksh (1986). "Examination of spinal monoamine receptors

through which brainstem opiate-sensitive systems act in the rat." Brain Res 363(1):

114-127.

Jolivalt, C., A. Minn, M. Vincentviry, M. M. Galteau and G. Siest (1995).

"Dextromethorphan O-demethylase activity in rat brain microsomes." Neurosci Lett

187(1): 65-68.

Jones, S. L. and G. F. Gebhart (1988). "Inhibition of spinal nociceptive transmission

from the midbrain, pons and medulla in the rat: activation of descending inhibition

by morphine, glutamate and electrical stimulation." Brain Res 460(2): 281-296.

Joshi, M. and R. F. Tyndale (2006). "Induction and recovery time course of rat brain

CYP2E1 after nicotine treatment." Drug Metab Dispos 34(4): 647-652.

Jover, R., R. Bort, M. J. Gomez-Lechon and J. V. Castell (1998). "Re-expression of

C/EBP alpha induces CYP2B6, CYP2C9 and CYP2D6 genes in HepG2 cells." FEBS

Lett 431(2): 227-230.

Kane, J. K., O. Konu, J. Z. Ma and M. D. Li (2004). "Nicotine coregulates multiple

pathways involved in protein modification/degradation in rat brain." Mol Brain Res

132(2): 181-191.

Kathiramalainathan, K., H. L. Kaplan, M. K. Romach, U. E. Busto, N. Y. Li, J. Sawe, R. F.

Tyndale and E. M. Sellers (2000). "Inhibition of cytochrome P450 2D6 modifies

codeine abuse liability." J Clin Psychopharmacol 20(4): 435-444.

Kelton, M.C., H.J. Kahn, C.L. Conrath and P.A. Newhouse (2000). "The effects of

nicotine on Parkinson's disease." Brain Cogn 43(1-3): 274-282.

Kerry, N. L., A. A. Somogyi, G. Mikus and F. Bochner (1993). "Primary and secondary

oxidative metabolism of dextromethorphan. In vitro studies with female Sprague-

Dawley and Dark Agouti rat liver microsomes." Biochem Pharmacol 45(4): 833-839.

Khokhar, J. Y. and R. F. Tyndale (2011). "Drug metabolism within the brain changes

drug response: selective manipulation of brain CYP2B alters propofol effects."

Neuropsychopharmacology 36(3): 692-700.

Khokhar, J. Y. and R. F. Tyndale (2012). "Rat brain CYP2B-enzymatic activation of

chlorpyrifos to the oxon mediates cholinergic neurotoxicity." Toxicol Sci 126(2): 325-

335.

Khwaja, M., A. McCormack, J.M. McIntosh, D.A. Di Monte and M. Quik (2007).

"Nicotine partially protects against paraquat-induced nigrostriatal damage in mice;

link to alpha6beta2*nAChRs." J Neurochem 100(1): 180-190.

Kirchheiner, J., U. Lang, T. Stamm, T. Sander and J. Gallinat (2006). "Association of

CYP2D6 genotypes and personality traits in healthy individuals." J Clin

Psychopharmacol 26(4): 440-442.

133

Kirchheiner, J., A. Seeringer, A. L. Godoy, B. Ohmle, C. Maier, P. Beschoner, E. J. Sim

and R. Viviani (2011). "CYP2D6 in the brain: genotype effects on resting brain

perfusion." Mol Psychiatry 16(3): 333-341.

Kishimoto, W., T. Hiroi, M. Shiraishi, M. Osada, S. Imaoka, S. Kominami, T. Igarashi

and Y. Funae (2004). "Cytochrome P450 2D catalyze steroid 21-hydroxylation in the

brain." Endocrinology 145(2): 699-705.

Kobayashi, S., S. Murray, D. Watson, D. Sesardic, D. S. Davies and A. R. Boobis

(1989). "The specificity of inhibition of debrisoquine 4-hydroxylase activity by

quinidine and quinine in the rat is the inverse of that in man." Biochem Pharmacol

38(17): 2795-2799.

Kodaira, H. and S. Spector (1988). "Transformation of thebaine to oripavine, codeine,

and morphine by rat liver, kidney, and brain microsomes." Proc Natl Acad Sci USA

85(4): 1267-1271.

Komori, M. (1993). "A novel P450 expressed at high level in rat brain." Biochem

Biophys Res Commun 196(2): 721-728.

Komura, H. and M. Iwaki (2005). "Pharmacokinetics and metabolism of metoprolol

and propranolol in the female DA and female Wistar rat: the female DA rat is not

always an animal model for poor metabolizers of CYP2D6." J Pharm Sci 94(2): 397-

408.

Kream, R. M. and G. B. Stefano (2006). "De novo biosynthesis of morphine in animal

cells: an evidence-based model." Med Sci Monit 12(10): RA207-219.

Kroemer, H. K., C. Fischer, C. O. Meese and M. Eichelbaum (1991).

"Enantiomer/enantiomer interaction of (S)- and (R)-propafenone for cytochrome

P450IID6-catalyzed 5-hydroxylation: in vitro evaluation of the mechanism." Mol

Pharmacol 40(1): 135-142.

Kudo, K., T. Ishida, N. Nishida, N. Yoshioka, H. Inoue, A. Tsuji and N. Ikeda (2006).

"Simple and sensitive determination of free and total morphine in human liver and

kidney using gas chromatography-mass spectrometry." J Chromatogr B Analyt

Technol Biomed Life Sci 830(2): 359-363.

Lamotte, C., C. B. Pert and S. H. Snyder (1976). "Opiate receptor binding in primate

spinal cord: Distribution and changes after dorsal root section." Brain Res 112(2):

407-412.

Lane, H. Y., W. C. Chiu, J. C. Y. Chou, S. T. Wu, M. H. Su and W. H. Chang (2000).

"Risperidone in acutely exacerbated schizophrenia: Dosing strategies and plasma

levels." J Clin Psychiatry 61(3): 209-214.

Lasagna, L. and T. J. De Kornfeld (1958). "Analgesic potency of normorphine in

patients with postoperative pain." J Pharmacol Exp Ther 124(3): 260-263.

134

Law, P. Y., Y. H. Wong and H. H. Loh (2000). "Molecular mechanisms and regulation

of opioid receptor signaling." Annu Rev Pharmacol Toxicol 40: 389-430.

Le Bars, D., M. Gozariu and S. W. Cadden (2001). "Animal models of nociception."

Pharmacol Rev 53(4): 597-652.

Lee, A. M., S. Miksys, R. Palmour and R. F. Tyndale (2006a). "CYP2B6 is expressed in

African Green monkey brain and is induced by chronic nicotine treatment."

Neuropharmacology 50(4): 441-450.

Lee, A. M., S. Miksys and R. F. Tyndale (2006b). "Phenobarbital increases monkey in

vivo nicotine disposition and induces liver and brain CYP2B6 protein." Br J

Pharmacol 148(6): 786-794.

Leith, J. L., A. W. Wilson, L. F. Donaldson and B. M. Lumb (2007). "Cyclooxygenase-

1-derived prostaglandins in the periaqueductal gray differentially control C-versus

A-fiber-evoked spinal nociception." J Neurosci 27(42): 11296-11305.

Lenz, F.A., N. Weiss, S. Ohara, C. Lawson and J.D. Greenspan (2004). "The role of the

thalamus in pain." Suppl Clin Neurophysiol 57: 50-61.

Lewis, J. W., J. E. Sherman and J. C. Liebeskind (1981). "Opioid and non-opioid stress

analgesia: assessment of tolerance and cross-tolerance with morphine." J Neurosci

1(4): 358-363.

Lewis, V. A. and G. F. Gebhart (1977). "Evaluation of the periaqueductal central gray

(PAG) as a morphine-specific locus of action and examination of morphine-induced

and stimulation-produced analgesia at coincident PAG loci." Brain Res 124(2): 283-

303.

Lin, J. H. (1995). "Species similarities and differences in pharmacokinetics." Drug

Metab Dispos 23(10): 1008-1021.

Lin, L. Y., Y. Kumagai and A. K. Cho (1992). "Enzymatic and chemical

demethylenation of (methylenedioxy)amphetamine and

(methylenedioxy)methamphetamine by rat brain microsomes." Chem Res Toxicol

5(3): 401-406.

Lipp, J. (1991). "Possible mechanisms of morphine analgesia." Clin Neuropharmacol

14(2): 131-147.

Llerena, A., P. Dorado, E. M. Penas-Lledo, M. C. Caceres and A. De la Rubia (2007).

"Low frequency of CYP2D6 poor metabolizers among schizophrenia patients."

Pharmacogenomics J 7(6): 408-410.

Llerena, A., G. Edman, J. Cobaleda, J. Benitez, D. Schalling and L. Bertilsson (1993).

"Relationship between personality and debrisoquine hydroxylation capacity -

suggestion of an endogenous neuroactive substrate or product of the cytochrome-

P4502D6." Acta Psychiatr Scand 87(1): 23-28.

135

Lofgren, S., A. L. Hagbjork, S. Ekman, R. Fransson-Steen and Y. Terelius (2004).

"Metabolism of human cytochrome P450 marker substrates in mouse: a strain and

gender comparison." Xenobiotica 34(9): 811-834.

Lotsch, J., A. Stockmann, G. Kobal, K. Brune, R. Waibel, N. Schmidt and G.

Geisslinger (1996). "Pharmacokinetics of morphine and its glucuronides after

intravenous infusion of morphine and morphine-6-glucuronide in healthy

volunteers." Clin Pharmacol Ther 60(3): 316-325.

Lysakowski, C., L. Dumont, C. Czarnetzki, D. Bertrand, E. Tassonyi and M.R. Tramer

(2006). "The effect of cigarette smoking on the hypnotic efficacy of propofol."

Anaesthesia 61(9): 826-831.

Ma, X. C., J. R. Idle, K. W. Krausz and F. J. Gonzalez (2005). "Metabolism of melatonin

by human cytochromes P450." Drug Metab Dispos 33(4): 489-494.

Mann, A., S. Miksys, A. Lee, D. C. Mash and R. F. Tyndale (2008). "Induction of the

drug metabolizing enzyme CYP2D in monkey brain by chronic nicotine treatment."

Neuropharmacology 55(7): 1147-1155.

Mann, A., S. L. Miksys, A. Gaedigk, S. J. Kish, D. C. Mash and R. F. Tyndale (2012).

"The neuroprotective enzyme CYP2D6 increases in the brain with age and is lower in

Parkinson's disease patients." Neurobiol Aging 33(9): 2160-2171.

Mann, A. and R. F. Tyndale (2010). "Cytochrome P450 2D6 enzyme neuroprotects

against 1-methyl-4-phenylpyridinium toxicity in SH-SY5Y neuronal cells." Eur J

Neurosci 31(7): 1185-1193.

Mansour, A., C. A. Fox, H. Akil and S. J. Watson (1995). "Opioid-receptor mRNA

expression in the rat CNS: anatomical and functional implications." Trends Neurosci

18(1): 22-29.

Mansour, A., H. Khachaturian, M. E. Lewis, H. Akil and S. J. Watson (1988). "Anatomy

of CNS opioid receptors." Trends Neurosci 11(7): 308-314.

Masubuchi, Y., S. Fujita, M. Chiba, N. Kagimoto, S. Umeda and T. Suzuki (1991).

"Impairment of debrisoquine 4-hydroxylase and related monooxygenase activities

in the rat following treatment with propranolol." Biochem Pharmacol 41(6-7): 861-

865.

Masubuchi, Y., S. Narimatsu, S. Hosokawa and T. Suzuki (1994). "Role of the CYP2D

subfamily in metabolism-dependent covalent binding of propranolol to liver

microsomal protein in rats." Biochem Pharmacol 48(10): 1891-1898.

Matoh, N., S. Tanaka, M. Takehashi, M. Banasik, T. Stedeford, E. Masliah, S. Suzuki, Y.

Nishimura and K. Ueda (2003). "Overexpression of CYP2D6 attenuates the toxicity of

MPP+ in actively dividing and differentiated PC12 cells." Gene Expr 11(3-4): 117-

124.

136

Matsunaga, E., U. M. Zanger, J. P. Hardwick, H. V. Gelboin, U. A. Meyer and F. J.

Gonzalez (1989). "The CYP2D gene subfamily: analysis of the molecular basis of the

debrisoquine 4-hydroxylase deficiency in DA rats." Biochemistry 28(18): 7349-7355.

McCann, S.J., S.M. Pond, K.M. James and D.G. Le Couteur (1997). "The association

between polymorphisms in the cytochrome P-450 2D6 gene and Parkinson's disease:

a case-control study and meta-analysis." J Neurol Sci 153(1): 50-53.

McFadzean, I. (1988). "The ionic mechanisms underlying opioid actions."

Neuropeptides 11(4): 173-180.

McFayden, M. C., W. T. Melvin and G. I. Murray (1998). "Regional distribution of

individual forms of cytochrome P450 mRNA in normal adult human brain." Biochem

Pharmacol 55(6): 825-830.

McGinnity, D. F., N. J. Waters, J. Tucker and R. J. Riley (2008). "Integrated in vitro

analysis for the in vivo prediction of cytochrome P450-mediated drug-drug

interactions." Drug Metab Dispos 36(6): 1126-1134.

McLaughlin, L. A., L. J. Dickmann, C. R. Wolf and C. J. Henderson (2008). "Functional

expression and comparative characterization of nine murine cytochromes P450 by

fluorescent inhibition screening." Drug Metab Dispos 36(7): 1322-1331.

McLellan, R. A., M. Oscarson, J. Seidegard, D. A. Evans and M. Ingelman-Sundberg

(1997). "Frequent occurrence of CYP2D6 gene duplication in Saudi Arabians."

Pharmacogenetics 7(3): 187-191.

Meyer, R. P., M. Gehlhaus, R. Knoth and B. Volk (2007). "Expression and function of

cytochrome p450 in brain drug metabolism." Curr Drug Metab 8(4): 297-306.

Meyer, R. P., R. L. Lindberg, F. Hoffmann and U. A. Meyer (2005). "Cytosolic

persistence of mouse brain CYP1A1 in chronic heme deficiency." Biol Chem 386(11):

1157-1164.

Michels, R. and P. M. Marzuk (1993a). "Progress in psychiatry (1)." N Engl J Med

329(8): 552-560.

Michels, R. and P. M. Marzuk (1993b). "Progress in psychiatry (2)." N Engl J Med

329(9): 628-638.

Miksys, S., E. Hoffmann and R. F. Tyndale (2000). "Regional and cellular induction of

nicotine-metabolizing CYP2B1 in rat brain by chronic nicotine treatment." Biochem

Pharmacol 59(12): 1501-1511.

Miksys, S., C. Lerman, P. G. Shields, D. C. Mash and R. F. Tyndale (2003). "Smoking,

alcoholism and genetic polymorphisms alter CYP2B6 levels in human brain."

Neuropharmacology 45(1): 122-132.

137

Miksys, S., Y. Rao, E. Hoffmann, D. C. Mash and R. F. Tyndale (2002). "Regional and

cellular expression of CYP2D6 in human brain: higher levels in alcoholics." J

Neurochem 82(6): 1376-1387.

Miksys, S., Y. Rao, E. M. Sellers, M. Kwan, D. Mendis and R. F. Tyndale (2000).

"Regional and cellular distribution of CYP2D subfamily members in rat brain."

Xenobiotica 30(6): 547-564.

Miksys, S. and R. F. Tyndale (2004). "The unique regulation of brain cytochrome

P450 2 (CYP2) family enzymes by drugs and genetics." Drug Metab Rev 36(2): 313-

333.

Miksys, S. and R. F. Tyndale (2009). "Brain drug-metabolizing cytochrome P450

enzymes are active in vivo, demonstrated by mechanism-based enzyme inhibition."

Neuropsychopharmacology 34(3): 634-640.

Miksys, S. L., C. Cheung, F. J. Gonzalez and R. F. Tyndale (2005). "Human CYP2D6

and mouse CYP2Ds: organ distribution in a humanized mouse model." Drug Metab

Dispos 33(10): 1495-1502.

Mikus, G., F. Bochner, M. Eichelbaum, P. Horak, A. A. Somogyi and S. Spector

(1994). "Endogenous codeine and morphine in poor and extensive metabolisers of

the CYP2D6 (debrisoquine/sparteine) polymorphism." J Pharmacol Exp Ther 268(2):

546-551.

Mikus, G., A. A. Somogyi, F. Bochner and M. Eichelbaum (1991). "Codeine O-

demethylation: rat strain differences and the effects of inhibitors." Biochem

Pharmacol 41(5): 757-762.

Miksys, S. L. and R. F. Tyndale (2002). "Drug-metabolizing cytochrome P450s in the

brain." J Psychiatry Neurosci 27(6): 406-415.

Millan, M. J. (2002). "Descending control of pain." Prog Neurobiol 66(6): 355-474.

Miller, R.T., S. Miksys and R.F. Tyndale (2012). "An investigation of the expression of

brain CYP2D6 in a monkey model of nicotine and ethanol exposure." Society for

Research on Nicotine and Tobacco, Houston, TX.

Mizuno, D., T. Hiroi, P. Ng, W. Kishimoto, S. Imaoka and Y. Funae (2003). "Regulation

of CYP2D expression in rat brain by toluene." Osaka City Med J 49(1): 49-56.

Mizuno, D., Y. Takahashi, T. Hiroi, S. Imaoka, T. Kamataki and Y. Funae (2003). "A

novel transcriptional element which regulates expression of the CYP2D4 gene by

Oct-1 and YY-1 binding." Biochim Biophys Acta 1627(2-3): 121-128.

Modi, S., D.E. Gilham, M.J. Sutcliffe, L.Y. Lian, W.U. Primrose, C.R. Wolf and G.C.

Roberts (1997). "1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine as a substrate of

cytochrome P450 2D6: allosteric effects of NADPH-cytochrome P450 reductase."

Biochemistry 36(15): 4461-4470.

138

Montoliu, C., M. Sancho-Tello, I. Azorin, M. Burgal, S. Valles, J. Renau-Pigueras and

C. Guerri (1995). "Ethanol increases cytochrome P4502E1 and induces oxidative

stress in astrocytes." J Neurochem 65(6): 2561-2570.

Narimatsu, S., T. Arai, Y. Masubuchi, T. Horie, M. Hosokawa, K. Ueno, H. Kataoka, S.

Yamamoto, T. Ishikawa and A. K. Cho (2001). "Inactivation of rat cytochrome P450 2D

enzyme by a further metabolite of 4-hydroxypropranolol, the major and active

metabolite of propranolol." Biol Pharm Bull 24(9): 988-994.

Narimatsu, S., S. Yamamoto, T. Koitabashi, R. Kato, Y. Masubuchi, T. Suzuki and T.

Horie (1999). "Biphasic kinetics of imipramine N-oxidation in rat brain microsomes."

Biol Pharm Bull 22(3): 253-256.

Nayebi, A. R. M., M. Hassanpour and H. Rezazadeh (2001). "Effect of chronic and

acute administration of fluoxetine and its additive effect with morphine on the

behavioural response in the formalin test in rats." J Pharm Pharmacol 53(2): 219-225.

Neafsey, P., G. Ginsberg, D. Hattis and B. Sonawane (2009). "Genetic polymorphism

in cytochrome P450 2D6 (CYP2D6): Population distribution of CYP2D6 activity." J

Toxicol Environ Health B Crit Rev 12(5-6): 334-361.

Nelson, D. R. (2006). "Cytochrome P450 nomenclature, 2004." Methods Mol Biol 320:

1-10.

Nelson, J. C., C. M. Mazure and P. I. Jatlow (1995). "Desipramine treatment of major

depression in patients over 75 years of age." J Clin Psychopharmacol 15(2): 99-105.

Niwa, T., K. Okada, T. Hiroi, S. Imaoka, S. Narimatsu and Y. Funae (2008). "Effect of

psychotropic drugs on the 21-hydroxylation of neurosteroids, progesterone and

allopregnanolone, catalyzed by rat CYP2D4 and human CYP2D6 in the brain." Biol

Pharm Bull 31(3): 348-351.

Niznik, H. B., R. F. Tyndale, F. R. Sallee, F. J. Gonzalez, J. P. Hardwick, T. Inaba and

W. Kalow (1990). "The dopamine transporter and cytochrome P45OIID1

(debrisoquine 4-hydroxylase) in brain: resolution and identification of two distinct

[3H]GBR-12935 binding proteins." Arch Biochem Biophys 276(2): 424-432.

Obach, R. S., R. L. Walsky and K. Venkatakrishnan (2007). "Mechanism-based

inactivation of human cytochrome p450 enzymes and the prediction of drug-drug

interactions." Drug Metab Dispos 35(2): 246-255.

O'Callaghan, J. P. and S. G. Holtzman (1975). "Quantification of the analgesic activity

of narcotic antagonists by a modified hot-plate procedure." J Pharmacol Exp Ther

192(3): 497-505.

Ochsner, K.N., D.H. Ludlow, K. Knierim, J. Hanelin, T. Ramachandran, G.C. Glover

and S.C. Mackey (2006). "Neural correlates of individual differences in pain-related

fear and anxiety." Pain 120(1-2): 69-77.

139

Oguri, K., N. Hanioka and H. Yoshimura (1990). "Species differences in metabolism

of codeine: urinary excretion of codeine glucuronide, morphine-3-glucuronide and

morphine-6-glucuronide in mice, rats, guinea pigs and rabbits." Xenobiotica 20(7):

683-688.

Ohe, T., M. Hirobe and T. Mashino (2000). "Novel metabolic pathway of estrone and

17beta-estradiol catalyzed by cytochrome P-450." Drug Metab Dispos 28(2): 110-

112.

Ohno, S., K. Kawana and S. Nakajin (2008). "Contribution of UDP-

glucuronosyltransferase 1A1 and 1A8 to morphine-6-glucuronidation and its kinetic

properties." Drug Metab Dispos 36(4): 688-694.

Okura, T., Y. Morita, Y. Ito, Y. Kagawa and S. Yamada (2009). "Effects of quinidine on

antinociception and pharmacokinetics of morphine in rats." J Pharm Pharmacol

61(5): 593-597.

Olanow, C.W. (2007). "The pathogenesis of cell death in Parkinson's disease--2007."

Mov Disord 22 Suppl 17: S335-342.

Oldendorf, W. H., S. Hyman, L. Braun and S. Z. Oldendorf (1972). "Blood-brain

barrier: penetration of morphine, codeine, heroin, and methadone after carotid

injection." Science 178(4064): 984-986.

Osborne, R., S. Joel, D. Trew and M. Slevin (1988). "Analgesic activity of morphine-6-

glucuronide." Lancet 1(8589): 828.

Pang, I. H. and M. R. Vasko (1986). "Effect of depletion of spinal cord norepinephrine

on morphine-induced antinociception." Brain Res 371(1): 171-176.

Parkinson, A., D. R. Mudra, C. Johnson, A. Dwyer and K. M. Carroll (2004). "The

effects of gender, age, ethnicity, and liver cirrhosis on cytochrome P450 enzyme

activity in human liver microsomes and inducibility in cultured human hepatocytes."

Toxicol Appl Pharmacol 199(3): 193-209.

Paxinos, G. and C. Watson (1986). The Rat Brain in Stereotaxic Coordinates, 2nd edn.

San Diego, Academic Press.

Penas-Lledo, E. M., P. Dorado, Z. Aguera, M. Gratacos, X. Estivill, F. Fernandez-

Aranda and A. Llerena (2012). "CYP2D6 polymorphism in patients with eating

disorders." Pharmacogenomics J 12(2): 173-175.

Penas-LLedo, E. M., P. Dorado, R. Pacheco, I. Gonzalez and L. L. A (2009). "Relation

between CYP2D6 genotype, personality, neurocognition and overall

psychopathology in healthy volunteers." Pharmacogenomics 10(7): 1111-1120.

Pert, C. B., M. J. Kuhar and S. H. Snyder (1975). "Autoradiographic localization of the

opiate receptor in rat brain." Life Sciences 16(12): 1849-1853.

140

Pert, C. B. and S. H. Snyder (1973). "Properties of opiate-receptor binding in rat

brain." Proc Natl Acad Sci U S A 70(8): 2243-2247.

Pluchino, N., M. Luisi, E. Lenzi, M. Centofanti, S. Begliuomini, L. Freschi, F. Ninni and

A.R. Genazzani (2006). "Progesterone and progestins: effects on brain,

allopregnanolone and beta-endorphin." J Steroid Biochem Mol Biol 102(1-5): 205-

213.

Poeaknapo, C., J. Schmidt, M. Brandsch, B. Drager and M. H. Zenk (2004).

"Endogenous formation of morphine in human cells." Proc Natl Acad Sci U S A

101(39): 14091-14096.

Projean, D., P. E. Morin, T. M. Tu and J. Ducharme (2003). "Identification of CYP3A4

and CYP2C8 as the major cytochrome P450 s responsible for morphine N-

demethylation in human liver microsomes." Xenobiotica 33(8): 841-854.

Proudfit, H. K. and D. L. Hammond (1981). "Alterations in nociceptive threshold and

morphine-induced analgesia produced by intrathecally administered amine

antagonists." Brain Res 218(1-2): 393-399.

Quik, M., H. Cox, N. Parameswaran, K. O'Leary, J.W. Langston and D. Di Monte

(2007). "Nicotine reduces levodopa-induced dyskinesias in lesioned monkeys." Ann

Neurol 62(6): 588-596.

Quik, M., M. O'Neill and X.A. Perez (2007). "Nicotine neuroprotection against

nigrostriatal damage: importance of the animal model." Trends Pharmacol Sci 28(5):

229-235.

Quik, M., N. Parameswaran, S.E. McCallum, T. Bordia, S. Bao, A. McCormack, A. Kim,

R.F. Tyndale, J.W. Langston and D.A. Di Monte (2006). "Chronic oral nicotine

treatment protects against striatal degeneration in MPTP-treated primates." J

Neurochem 98(6): 1866-1875.

Rae, J. M., M. D. Johnson, M. E. Lippman and D. A. Flockhart (2001). "Rifampin is a

selective, pleiotropic inducer of drug metabolism genes in human hepatocytes:

studies with cDNA and oligonucleotide expression arrays." J Pharmacol Exp Ther

299(3): 849-857.

Rendic, S. and F. J. Di Carlo (1997). "Human cytochrome P450 enzymes: a status

report summarizing their reactions, substrates, inducers, and inhibitors." Drug

Metab Rev 29(1-2): 413-580.

Riedel, M., M. J. Schwarz, M. Strassnig, I. Spellmann, A. Muller-Arends, K. Weber, J.

Zach, N. Muller and H. J. Moller (2005). "Risperidone plasma levels, clinical response

and side-effects." Eur Arch Psychiatry Clin Neurosci 255(4): 261-268.

141

Riedl, A. G., P. M. Watts, R. J. Edwards, A. R. Boobis, P. Jenner and C. D. Marsden

(1996). "Selective localisation of P450 enzymes and NADPH-P450 oxidoreductase in

rat basal ganglia using anti-peptide antisera." Brain Res 743(1-2): 324-328.

Riedl, A. G., P. M. Watts, R. J. Edwards, T. Schulz-Utermoehl, A. R. Boobis, P. Jenner

and C. D. Marsden (1999). "Expression and localisation of CYP2D enzymes in rat

basal ganglia." Brain Res 822(1-2): 175-191.

Rifkind, A. B., C. Lee, T. K. Chang and D. J. Waxman (1995). "Arachidonic acid

metabolism by human cytochrome P450s 2C8, 2C9, 2E1, and 1A2: regioselective

oxygenation and evidence for a role for CYP2C enzymes in arachidonic acid

epoxygenation in human liver microsomes." Arch Biochem Biophys 320(2): 380-389.

Ritter, J. K. (2000). "Roles of glucuronidation and UDP-glucuronosyltransferases in

xenobiotic bioactivation reactions." Chem Biol Interact 129(1-2): 171-193.

Roberts, R. L., S. E. Luty, R. T. Mulder, P. R. Joyce and M. A. Kennedy (2004).

"Association between cytochrome P450 2D6 genotype and harm avoidance." Am J

Med Genet B Neuropsychiatr Genet 127B(1): 90-93.

Romach, M. K., S. V. Otton, G. Somer, R. F. Tyndale and E. M. Sellers (2000).

"Cytochrome P450 2D6 and treatment of codeine dependence." J Clin

Psychopharmacol 20(1): 43-45.

Rosenbrock, H., C. E. Hagemeyer, I. Singec, R. Knoth and B. Volk (1999).

"Testosterone metabolism in rat brain is differentially enhanced by phenytoin-

inducible cytochrome P450 isoforms." J Neuroendocrinol 11(8): 597-604.

Rowland, K., W. W. Yeo, S. W. Ellis, I. G. Chadwick, I. Haq, M. S. Lennard, P. R.

Jackson, L. E. Ramsay and G. T. Tucker (1994). "Inhibition of CYP2D6 activity by

treatment with propranolol and the role of 4-hydroxy propranolol." Br J Clin

Pharmacol 38(1): 9-14.

Sachse, C., J. Brockmoller, S. Bauer and I. Roots (1997). "Cytochrome P450 2D6

variants in a Caucasian population: allele frequencies and phenotypic

consequences." Am J Hum Genet 60(2): 284-295.

Sakai, N., K. Q. Sakamoto, S. Fujita and M. Ishizuka (2009). "The importance of

heterogeneous nuclear ribonucleoprotein K on cytochrome P450 2D2 gene

regulation: its binding is reduced in Dark Agouti rats." Drug Metab Dispos 37(8):

1703-1710.

Schenk, P. W., M. A. van Fessem, S. Verploegh-Van Rij, R. A. Mathot, T. van Gelder,

A. G. Vulto, M. van Vliet, J. Lindemans, J. A. Bruijn and R. H. van Schaik (2008).

"Association of graded allele-specific changes in CYP2D6 function with imipramine

dose requirement in a large group of depressed patients." Mol Psychiatry 13(6): 597-

605.

142

Schoedel, K. A., E. M. Sellers and R. F. Tyndale (2001). "Induction of CYP2B1/2 and

nicotine metabolism by ethanol in rat liver but not rat brain." Biochem Pharmacol

62(8): 1025-1036.

Schoedel, K. A. and R. F. Tyndale (2003). "Induction of nicotine-metabolizing CYP2B1

by ethanol and ethanol-metabolizing CYP2E1 by nicotine: summary and

implications." Biochim Biophys Acta 1619(3): 283-290.

Schreiber, S., M. M. Backer, R. Weizman and C. G. Pick (1996). "The antinociceptive

effects of fluoxetine." Pain Clinic 9(3): 349-356.

Shen, H. W. and A. M. Yu (2009). "Difference in desipramine metabolic profile

between wild-type and CYP2D6-humanized mice." Drug Metab Lett 3(4): 234-241.

Siegle, I., P. Fritz, K. Eckhardt, U. M. Zanger and M. Eichelbaum (2001). "Cellular

localization and regional distribution of CYP2D6 mRNA and protein expression in

human brain." Pharmacogenetics 11(3): 237-245.

Simonds, W. F. (1988). "The molecular basis of opioid receptor function." Endocr Rev

9(2): 200-212.

Sinclair, J. G., C. D. Main and G. F. Lo (1988). "Spinal vs. supraspinal actions of

morphine on the rat tail-flick reflex." Pain 33(3): 357-362.

Sindrup, S. H., L. Arendt-Nielsen, K. Brosen, P. Bjerring, H. R. Angelo, B. Eriksen and

L. F. Gram (1992). "The effect of quinidine on the analgesic effect of codeine." Eur J

Clin Pharmacol 42(6): 587-591.

Sindrup, S. H., K. Brosen, P. Bjerring, L. Arendt-Nielsen, U. Larsen, H. R. Angelo and

L. F. Gram (1990). "Codeine increases pain thresholds to copper vapor laser stimuli

in extensive but not poor metabolizers of sparteine." Clin Pharmacol Ther 48(6): 686-

693.

Sindrup, S. H., L. Poulsen, K. Brosen, L. Arendt-Nielsen and L. F. Gram (1993). "Are

poor metabolisers of sparteine/debrisoquine less pain tolerant than extensive

metabolisers?" Pain 53(3): 335-339.

Siu, E. C., D. B. Wildenauer and R. F. Tyndale (2006). "Nicotine self-administration in

mice is associated with rates of nicotine inactivation by CYP2A5."

Psychopharmacology (Berl) 184(3-4): 401-408.

Smith, D. L., M. C. D'Amour and F. E. D'Amour (1943). "The analgesic properties of

certain drugs and drug combinations." Journal of Pharmacology and Experimental

Therapeutics 77(2): 184-193.

Snider, N.T., M.J. Sikora, C. Sridar, T.J. Feuerstein, J.M. Rae and P.F. Hollenberg

(2008). "The endocannabinoid anandamide is a substrate for the human polymorphic

cytochrome P450 2D6." J Pharmacol Exp Ther, 327(2): 538-545.

143

Spina, E., A. Avenoso, G. Facciola, M. Salemi, M. G. Scordo, M. Ancione, A. G. Madia

and E. Perucca (2001). "Relationship between plasma risperidone and 9-

hydroxyrisperidone concentrations and clinical response in patients with

schizophrenia." Psychopharmacology (Berl) 153(2): 238-243.

Stevens, J. C., S. A. Marsh, M. J. Zaya, K. J. Regina, K. Divakaran, M. Le and R. N.

Hines (2008). "Developmental changes in human liver CYP2D6 expression." Drug

Metab Dispos 36(8): 1587-1593.

Strobel, H. W., C. M. Thompson and L. Antonovic. (2001). "Cytochromes P450 in

brain: function and significance." Curr Drug Metab 2(2): 199-214.

Strobl, G. R., S. von Kruedener, J. Stockigt, F. P. Guengerich and T. Wolff (1993).

"Development of a pharmacophore for inhibition of human liver cytochrome P-450

2D6: molecular modeling and inhibition studies." J Med Chem 36(9): 1136-1145.

Suzuki, T., S. Fujita, S. Narimatsu, Y. Masubuchi, M. Tachibana, S. Ohta and M. Hirobe

(1992). "Cytochrome P450 isozymes catalyzing 4-hydroxylation of parkinsonism-

related compound 1,2,3,4-tetrahydroisoquinoline in rat liver microsomes." Faseb J

6(2): 771-776.

Tate, S. and H. Cook (1996). "Postoperative nausea and vomiting. 1: Physiology and

aetiology." Br J Nurs 5(16): 962, 964, 966 passim.

Thorn, C. F., T. E. Klein and R. B. Altman (2009). "Codeine and morphine pathway."

Pharmacogenet Genomics 19(7): 556-558.

Trapani, G., C. Altomare, G. Liso, E. Sanna and G. Biggio (2000). "Propofol in

anaesthesia. Mechanism of action, structure-activity relationships, and drug

delivery." Curr Med Chem 7(2): 249-271.

Transon, C., S. Lecoeur, T. Leemann, P. Beaune and P. Dayer (1996). "Interindividual

variability in catalytic activity and immunoreactivity of three major human liver

cytochrome P450 isozymes." Eur J Clin Pharmacol 51(1): 79-85.

Turpeinen, M., L. E. Korhonen, A. Tolonen, J. Uusitalo, R. Juvonen, H. Raunio and O.

Pelkonen (2006). "Cytochrome P450 (CYP) inhibition screening: comparison of three

tests." Eur J Pharm Sci 29(2): 130-138.

Tyndale, R. F., K. P. Droll and E. M. Sellers (1997). "Genetically deficient CYP2D6

metabolism provides protection against oral opiate dependence."

Pharmacogenetics 7(5): 375-379.

Tyndale, R. F., Y. Li, N. Y. Li, E. Messina, S. Miksys and E. M. Sellers (1999).

"Characterization of cytochrome P-450 2D1 activity in rat brain: high-affinity kinetics

for dextromethorphan." Drug Metab Dispos 27(8): 924-930.

Tyndale, R. F., R. Sunahara, T. Inaba, W. Kalow, F. J. Gonzalez and H. B. Niznik

(1991). "Neuronal cytochrome P450IID1 (debrisoquine/sparteine-type): potent

144

inhibition of activity by (-)-cocaine and nucleotide sequence identity to human

hepatic P450 gene CYP2D6." Mol Pharmacol 40(1): 63-68.

Vaglini, F., C. Pardini, C. Viaggi, C. Bartoli, D. Dinucci and G. U. Corsini (2004).

"Involvement of cytochrome P450 2E1 in the 1-methyl-4-phenyl-1,2,3,6-

tetrahydropyridine-induced mouse model of Parkinson's disease." J Neurochem

91(2): 285-298.

Van, L. M., A. Heydari, J. Yang, J. Hargreaves, K. Rowland-Yeo, M. S. Lennard, G. T.

Tucker and A. Rostami-Hodjegan (2006). "The impact of experimental design on

assessing mechanism-based inactivation of CYP2D6 by MDMA (Ecstasy)." J

Psychopharmacol 20(6): 834-841.

Vasiliou, V., T. L. Ziegler, P. Bludeau, D.R. Petersen, F.J. Gonzalez and R.A. Deitrich

(2006). "CYP2E1 and catalase influence ethanol sensitivity in the central nervous

system." Pharmacogenet Genomics 16(1): 51-58.

Vaughan, C.W., E. E. Bagley, G. M. Drew, A. Schuller, J. E. Pintar, S. P. Hack and M. J.

Christie (2003). "Cellular actions of opioids on periaqueductal grey neurons from

C57B16/J mice and mutant mice lacking MOR-1." Br J Pharmacol 139(2): 362-367.

Vaughan, C.W. and M. J. Christie (1997). "Presynaptic inhibitory action of opioids on

synaptic transmission in the rat periaqueductal grey in vitro." J Physiol 498 ( Pt 2):

463-472.

Vento, P.J. and D. Daniels (2010). "Repeated administration of angiotensin II reduces

its dipsogenic effect without altering saline intake." Exp Physiol 95(6): 736-745.

Vieregge, A., M. Sieberer, H. Jacobs, J.M. Hagenah and P. Vieregge (2001).

"Transdermal nicotine in PD: a randomized, double-blind, placebo-controlled

study." Neurology 57(6): 1032-1035.

Voirol, P., M. Jonzier-Perey, F. Porchet, M. J. Reymond, R. C. Janzer, C. Bouras, H. W.

Strobel, M. Kosel, C. B. Eap and P. Baumann (2000). "Cytochrome P-450 activities in

human and rat brain microsomes." Brain Res 855(2): 235-243.

Volk, B., U. Hettmannsperger, T. Papp, Z. Amelizad, F. Oesch and R. Knoth (1991).

"Mapping of phenytoin-inducible cytochrome P450 immunoreactivity in the mouse

central nervous system." Neuroscience 42(1): 215-235.

Von Moltke, L.L., D.J. Greenblatt, J.M. Grassi, B.W. Granda, K. Venkatakrishnan, J.

Schmider, J.S. Harmatz and R.I. Shader (1998). "Multiple human cytochromes

contribute to biotransformation of dextromethorphan in-vitro: role of CYP2C9,

CYP2C19, CYP2D6, and CYP3A." J Pharm Pharmacol 50(9): 997-1004.

Vuppugalla, R. and R. Mehvar (2005). "Enzyme-selective effects of nitric oxide on

affinity and maximum velocity of various rat cytochromes P450." Drug Metab Dispos

33(6): 829-836.

145

Wang, H., K. L. Napoli and H. W. Strobel (2000). "Cytochrome P450 3A9 catalyzes the

metabolism of progesterone and other steroid hormones." Mol Cell Biochem 213(1-

2): 127-135.

Warner, M. and J. A. Gustafsson (1994). "Effect of ethanol on cytochrome P450 in the

rat brain." Proc Natl Acad Sci U S A 91(3): 1019-1023.

Watts, P. M., A. G. Riedl, D. C. Douek, R. J. Edwards, A. R. Boobis, P. Jenner and C. D.

Marsden (1998). "Co-localization of P450 enzymes in the rat substantia nigra with

tyrosine hydroxylase." Neuroscience 86(2): 511-519.

Wilson, H.D., M.L. Uhelski, and P.N. Fuchs (2008). "Examining the role of the medial

thalamus in modulating the affective dimension of pain." Brain Res 1229: 90-99.

Woodland, C., T. T. Huang, E. Gryz, R. Bendayan and J. P. Fawcett (2008).

"Expression, activity and regulation of CYP3A in human and rodent brain." Drug

Metab Rev 40(1): 149-168.

Woolfe, G. and A. D. MacDonald (1944). "The evaluation of the analgesic action of

pethidine hydrochloride (Demerol)." J Pharmacol Exp Ther 80(3): 300-307.

Wyss, A., J. A. Gustafsson and M. Warner (1995). "Cytochromes P450 of the 2D

subfamily in rat brain." Mol Pharmacol 47(6): 1148-1155.

Xie, X., H. Liao, H. Dang, W. Pang, Y. Guan, X. Wang, J. Y. Shyy, Y. Zhu and F. M.

Sladek (2009). "Down-regulation of hepatic HNF4alpha gene expression during

hyperinsulinemia via SREBPs." Mol Endocrinol 23(4): 434-443.

Xu, B. Q., T. A. Aasmundstad, A. Bjorneboe, A. S. Christophersen and J. Morland

(1995). "Ethylmorphine O-deethylation in isolated rat hepatocytes. Involvement of

codeine O-demethylation enzyme systems." Biochem Pharmacol 49(4): 453-460.

Xu, B. Q., T. A. Aasmundstad, A. S. Christophersen, J. Morland and A. Bjorneboe

(1997). "Evidence for CYP2D1-mediated primary and secondary O-dealkylation of

ethylmorphine and codeine in rat liver microsomes." Biochem Pharmacol 53(4): 603-

609.

Yaksh, T. L. (1979). "Direct evidence that spinal serotonin and noradrenaline

terminals mediate the spinal antinociceptive effects of morphine in the

periaqueductal gray." Brain Res 160(1): 180-185.

Yaksh, T. L. and T. A. Rudy (1978). "Narcotic analgestics: CNS sites and mechanisms

of action as revealed by intracerebral injection techniques." Pain 4(4): 299-359.

Yaksh, T. L. and G. M. Tyce (1979). "Microinjection of morphine into the

periaqueductal gray evokes the release of serotonin from spinal cord." Brain Res

171(1): 176-181.

146

Yaksh, T. L. and P. R. Wilson (1979). "Spinal serotonin terminal system mediates

antinociception." J Pharmacol Exp Ther 208(3): 446-453.

Yaksh, T. L., J. C. Yeung and T. A. Rudy (1976). "Systematic examination in the rat of

brain sites sensitive to the direct application of morphine: observation of differential

effects within the periaqueductal gray." Brain Res 114(1): 83-103.

Yamamoto, T., A. Suzuki and Y. Kohno (2003). "High-throughput screening to

estimate single or multiple enzymes involved in drug metabolism: microtitre plate

assay using a combination of recombinant CYP2D6 and human liver microsomes."

Xenobiotica 33(8): 823-839.

Yokomori, N., R. Kobayashi, R. Moore, T. Sueyoshi and M. Negishi (1995). "A DNA

methylation site in the male-specific P450 (Cyp 2d-9) promoter and binding of the

heteromeric transcription factor GABP." Mol Cell Biol 15(10): 5355-5362.

Yu, A. M. and R. L. Haining (2006). "Expression, purification, and characterization of

mouse CYP2d22." Drug Metab Dispos 34(7): 1167-1174.

Yu, A. M., J. R. Idle and F. J. Gonzalez (2004). "Polymorphic cytochrome P450 2D6:

humanized mouse model and endogenous substrates." Drug Metab Rev 36(2): 243-

277.

Yu A.M., J.R. Idle, L.G. Byrd, K.W. Krausz, A. Kupfer and F.J. Gonzalez (2003a).

"Regeneration of serotonin from 5-methoxytryptamine by polymorphic human

CYP2D6." Pharmacogenetics 13(3):173–181.

Yu, A.M., J.R. Idle, K.W. Krausz, A. Kupfer and F.J. Gonzalez (2003b). "Contribution of

individual cytochrome P450 isozymes to the O-demethylation of the psychotropic

betacarboline alkaloids harmaline and harmine." J Pharmacol Exp Ther 305(1): 315-

322.

Yue, Q. Y., J. Hasselstrom, J. O. Svensson and J. Sawe (1991). "Pharmacokinetics of

codeine and its metabolites in Caucasian healthy volunteers: comparisons between

extensive and poor hydroxylators of debrisoquine." Br J Clin Pharmacol 31(6): 635-

642.

Yue, J., J. Khokhar, S. Miksys and R. F. Tyndale (2009). "Differential induction of

ethanol-metabolizing CYP2E1 and nicotine-metabolizing CYP2B1/2 in rat liver by

chronic nicotine treatment and voluntary ethanol intake." Eur J Pharmacol 609(1-3):

88-95.

Yue, J., S. Miksys, E. Hoffmann and R. F. Tyndale (2008). "Chronic nicotine treatment

induces rat CYP2D in the brain but not in the liver: an investigation of induction and

time course." J Psychiatry Neurosci 33(1): 54-63.

Yue, Q. Y., C. Alm, J. O. Svensson and J. Sawe (1997). "Quantification of the O- and N-

demethylated and the glucuronidated metabolites of codeine relative to the

147

debrisoquine metabolic ratio in urine in ultrarapid, rapid, and poor debrisoquine

hydroxylators." Ther Drug Monit 19(5): 539-542.

Yue, Q. Y. and J. Sawe (1997). "Different effects of inhibitors on the O- and N-

demethylation of codeine in human liver microsomes." Eur J Clin Pharmacol 52(1):

41-47.

Zanger, U. M., S. Raimundo and M. Eichelbaum (2004). "Cytochrome P450 2D6:

overview and update on pharmacology, genetics, biochemistry." Naunyn

Schmiedebergs Arch Pharmacol 369(1): 23-37.

Zhou, S. F. (2009). "Polymorphism of human cytochrome P450 2D6 and its clinical

significance: Part I." Clin Pharmacokinet 48(11): 689-723.

Zhu, W. (2008). "CYP2D6: a key enzyme in morphine synthesis in animals." Med Sci

Monit 14(11): SC15-18.

Zhu, W., P. Cadet, G. Baggerman, K.J. Mantione and G.B. Stefano (2005). "Human

white blood cells synthesize morphine: CYP2D6 modulation." J Immunol 175(11):

7357-7362.

Zhu, W., K. J. Mantione, L. Shen, P. Cadet, T. Esch, Y. Goumon, E. Bianchi, D. Sonetti

and G. B. Stefano (2005). "Tyrosine and tyramine increase endogenous ganglionic

morphine and dopamine levels in vitro and in vivo: cyp2d6 and tyrosine

hydroxylase modulation demonstrates a dopamine coupling." Med Sci Monit 11(11):

BR397-404.

148

List of Abstracts

Zhou, K., J.Y. Khokhar and R.F. Tyndale (2012). “The role of brain CYP2D metabolism

in codeine analgesia.” Visions in Pharmacology, Toronto.

Zhou, K., J.Y. Khokhar and R.F. Tyndale (2012). “The role of brain CYP2D metabolism

in codeine analgesia.” Canadian College of Neuropsychopharmacology, Vancouver.

Zhou, K., J.Y. Khokhar and R.F. Tyndale (2012). “The role of brain CYP2D metabolism

in codeine analgesia.” Southern Ontario Neuroscience Association, Toronto.