227
Actions of benzophenanthridine alkaloids and various synthetic compounds on the cannabinoid-1 (CB 1 ) receptor pathway of mouse brain with particular reference to the effects on [ 3 H]CP55940 and [ 3 H]SR141716A binding, interference with basal and CP55940-stimulated [ 35 S]GTPγS binding, and modification of WIN55212-2-dependent inhibition of L-glutamate release from synaptosomes by Amey S. Dhopeshwarkar MSc., University of Abertay Dundee, 2007 B.Pharm., University of Pune, 2004 THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY in the Department of Biological Sciences Faculty of Science Amey S. Dhopeshwarkar 2012 SIMON FRASER UNIVERSITY Summer 2012 All rights reserved. However, in accordance with the Copyright Act of Canada, this work may be reproduced, without authorization, under the conditions for “Fair Dealing.” Therefore, limited reproduction of this work for the purposes of private study, research, criticism, review and news reporting is likely to be in accordance with the law, particularly if cited appropriately.

by Amey S. Dhopeshwarkar - Simon Fraser University

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Actions of benzophenanthridine alkaloids and various synthetic compounds on the cannabinoid-1 (CB1) receptor pathway of mouse brain with particular reference to the

effects on [3H]CP55940 and [3H]SR141716A binding, interference with basal and

CP55940-stimulated [35S]GTPγS binding, and modification of WIN55212-2-dependent

inhibition of L-glutamate release from synaptosomes

by Amey S. Dhopeshwarkar

MSc., University of Abertay Dundee, 2007 B.Pharm., University of Pune, 2004

THESIS SUBMITTED IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

in the

Department of Biological Sciences

Faculty of Science

Amey S. Dhopeshwarkar 2012

SIMON FRASER UNIVERSITY Summer 2012

All rights reserved. However, in accordance with the Copyright Act of Canada, this work may

be reproduced, without authorization, under the conditions for “Fair Dealing.” Therefore, limited reproduction of this work for the

purposes of private study, research, criticism, review and news reporting is likely to be in accordance with the law, particularly if cited appropriately.

ii

Approval

Name: Amey S. Dhopeshwarkar Degree: Doctor of Philosophy (Biological Sciences) Title of Thesis: Actions of benzophenanthridine alkaloids and various

synthetic compounds on the cannabinoid-1 (CB1) receptor pathway of mouse brain with particular reference to the effects on [3H]CP55940 and [3H]SR141716A binding, interference with basal and CP55940-stimulated [35S]GTPγS binding, and modification of WIN55212-2-dependent inhibition of L-glutamate release from synaptosomes.

Examining Committee: Chair: Dr Julian Christians, Associate Professor

Dr Russell A. Nicholson Senior Supervisor Associate Professor

Dr Christopher Kennedy Supervisor Professor

Dr Francis C.P. Law Supervisor Professor

Dr Gordon Rintoul Internal Examiner Associate Professor Department of Biological Sciences, SFU

Dr Andrew Gifford External Examiner Scientist, Medical Department Brookhaven National Laboratory

Date Defended/Approved: August 15, 2012

iii

Partial Copyright Licence

Ethics Statement

The author, whose name appears on the title page of this work, has obtained, for the research described in this work, either:

a. human research ethics approval from the Simon Fraser University Office of Research Ethics,

or

b. advance approval of the animal care protocol from the University Animal Care Committee of Simon Fraser University;

or has conducted the research

c. as a co-investigator, collaborator or research assistant in a research project approved in advance,

or

d. as a member of a course approved in advance for minimal risk human research, by the Office of Research Ethics.

A copy of the approval letter has been filed at the Theses Office of the University Library at the time of submission of this thesis or project.

The original application for approval and letter of approval are filed with the relevant offices. Inquiries may be directed to those authorities.

Simon Fraser University Library Burnaby, British Columbia, Canada

update Spring 2010

iii

Abstract

Benzophenanthridine alkaloids (chelerythrine and sanguinarine) inhibited the binding of

[3H]CP55940 and [3H]SR141716A to mouse brain membranes (IC50s approx. 1-2 µM).

Piperonyl butoxide and (S)-methoprene were more potent inhibitors of [3H]CP55940

binding (IC50s: 8.2 µM and 16.4 µM respectively) than of [3H]SR141716A binding (IC50s:

21 µM and 63 µM respectively). Binding experiments demonstrated selectivity towards

the brain CB1 versus spleen CB2 receptor.

Benzophenanthridines reduced the Kd of [3H]CP55940 binding to brain membranes

whereas (S)-methoprene and piperonyl butoxide lowered Bmax. These study

compounds reduced the association of [3H]CP55940 and [3H]SR141716A, however

benzophenanthridines were consistently more effective.

In the presence of a saturating concentration of SR141716A, (S)-methoprene and

piperonyl butoxide increased dissociation of [3H]SR141716A above that observed with

SR141716A alone. All compounds activated [3H]SR141716A dissociation when assayed

alone, but (S)-methoprene was the least effective. In separate studies, phthalate

diesters reduced the Bmax of [3H]SR141716A without affecting Kd, and increased

[3H]SR141716A dissociation above a saturating concentration of AM251.

Benzophenanthridines antagonized CP55940-stimulated and basal binding of

[35S]GTPγS to the G-protein of mouse brain, whereas piperonyl butoxide and (S)-

methoprene inhibited CP55940-stimulated [35S]GTPγS binding only. Inhibition of

CP55940-stimulated binding of [35S]GTPγS was also demonstrated with phthalates.

4-Aminopyridine- (4-AP-) induced release of L-glutamate from mouse brain

synaptosomes was partially inhibited by WIN55212-2. The inhibitory effect of

WIN55212-2 was completely neutralized by AM251, (S)-methoprene, piperonyl butoxide

and phthalate diesters, whereas in the presence of WIN55212-2, the

benzophenanthridines enhanced 4-AP-induced L-glutamate release above that caused

by 4-AP alone.

iv

The [3H]CP55940 and [3H]SR141716A binding data suggest that the study compounds

modify radioligand binding allosterically. The [35S]GTPγS binding results suggest that

chelerythrine and sanguinarine are inverse agonists of G-protein-coupled CB1 receptors,

while piperonyl butoxide, (S)-methoprene and phthalate diesters are neutral lower

potency antagonists. Modulation 4-AP-evoked L-glutamate release from synaptosomes

by the study compounds with WIN-55212-2 present strongly supports this latter profiling.

Although these compounds exhibit lower potencies versus many conventional CB1

receptor inhibitors, further studies are warranted, given their potential to 1) modify CB1

receptor-dependent behavioral/physiological outcomes in the whole animal, and 2) serve

as starting structures for synthesis of novel/more potent G-protein-coupled CB1 receptor

blocking drugs.

Keywords: Benzophenanthridines; (S)-methoprene; piperonyl butoxide; [3H]CP55940; [35S]GTPγS; L-glutamate; synaptosomes;cannabinoid-1 (CB1) receptor

v

Dedication

To My Beloved Mom and Dad

vi

Acknowledgements

I wish to express my deepest gratitude and appreciation to my senior supervisor, Dr

Russell A. Nichoson for his guidance, patience and indefatigable support throughout my

graduate research career. I remember the days when Dr Nicholson spared his time

even on weekends and holidays to discuss my research and his invaluable suggestions

and encouragements have always made me feel confident about my research work.

Thorough discussion sessions with him about project and related scientific issues and

perspectives have enriched my knowledge in this field. Without Dr Nicholson’s support

and effort, I would not have completed my PhD research in time. I believe that I was

lucky to have such a knowledgeable senior supervisor and I am fortunate to be his last

graduate student.

I am very much thankful to Dr Chris Kennedy and Dr Francis C.P. Law for serving as my

committe members and their valuable time and inputs during my PhD. They have

always been supportive during my studies at SFU.

I am also thankful to Mr Saurabh Jain and Ms Kathleen M. Bisset for their help and

advice during my research.

Finally, I would like to thank my family for their love, support encouragement and always

believing in me.

vii

Table of Contents

Approval .......................................................................................................................... ii Abstract .......................................................................................................................... iii Dedication ....................................................................................................................... v Acknowledgements ........................................................................................................ vi Table of Contents .......................................................................................................... vii List of Tables ................................................................................................................. xii List of Figures................................................................................................................xiv Glossary ........................................................................................................................xxi

1. Introduction .......................................................................................................... 1 1.1. Historical significance of cannabis use and cannabinoids ....................................... 1

1.1.1. The early Chinese/Indian era ...................................................................... 1 1.1.2. The period encompassing the early Christian era through to the 18th

century ........................................................................................................ 2 1.1.3. The Western medicine era of the 19th and 20th centuries ............................. 2

1.2. Cannabinoids ......................................................................................................... 5 1.2.1. G protein-coupled receptors (GPCRs) and their activation cycle ................. 7 1.2.2. The [35S]GTPγS binding assay .................................................................... 8

1.3. Other cannabinoid receptors .................................................................................. 8 1.4. Cannabinoid-1 Receptors (CB1-Rs) ........................................................................ 9

1.4.1. The structure and activation of CB1-Rs ....................................................... 9 1.4.2. The distribution of CB1-Rs in mammalian brain ......................................... 17

1.5. CB1-R-mediated intracellular signaling pathways .................................................. 18 1.5.1. Inhibition of cyclic AMP (cAMP) ................................................................ 18 1.5.2. Stimulation of cAMP production ................................................................ 20 1.5.3. CB1-Rs and the modulation of Ca2+ fluxes and phospholipases C

and A ........................................................................................................ 21 1.5.4. CB1-R-dependent regulation of ion channels ............................................. 21 1.5.5. Involvement of CB1-Rs in the suppression of neurotransmitter

release ...................................................................................................... 22 1.6. Homodimerization and heterodimerization of CB1-Rs ........................................... 24 1.7. Constitutive activity of CB1-Rs .............................................................................. 25 1.8. The biochemistry of endocannabinoids ................................................................. 25

1.8.1. Anandamide biosynthesis ......................................................................... 28 1.8.2. 2-Arachidonoyl glycerol (2-AG) biosynthesis ............................................. 30

1.9. Degradation pathways for endocannabinoids ....................................................... 32 1.10. Transport of endocannabinoids ............................................................................ 33 1.11. Endocannabinoid-mediated short term depression (DSI and DSE) ....................... 35 1.12. Endocannabinoids as synaptic circuit breakers and retrograde messengers ........ 35 1.13. Mechanisms of endocannabinoid mediated short term depression (eCB-

STD) ..................................................................................................................... 38 1.13.1. CaER ........................................................................................................ 38 1.13.2. Basal RER ................................................................................................ 38 1.13.3. Ca2+-assisted RER .................................................................................... 39

1.14. Termination of eCB-STD ...................................................................................... 39 1.15. Endocannabinoid-mediated long term depression (eCB-LTD) .............................. 41

viii

1.16. Other important aspects of endocannabinoid signaling ......................................... 41 1.16.1. Regulation of excitability ........................................................................... 41 1.16.2. Basal activity of endocannabinoid signaling .............................................. 42 1.16.3. Plasticity of endocannabinoid signaling ..................................................... 42

1.17. Subcellular distribution of various signaling molecules involved in regulation of the endocannabinoid system ............................................................................ 42 1.17.1. Gq Protein α subunit .................................................................................. 42 1.17.2. Phospholipase Cβ (PLCβ) ......................................................................... 43 1.17.3. Diacylglycerol lipase (DAGL) ..................................................................... 43 1.17.4. N-acyl-phosphatidylethanolamine-hydrolyzing phospholipase D

(NAPE-PLD) ............................................................................................. 43 1.17.5. Monoacylglycerol lipase (MAGL) ............................................................... 44 1.17.6. Fatty acid amide hydrolase (FAAH) ........................................................... 44

1.18. Physiological roles of the endocannabinoid system .............................................. 44 1.18.1. Learning and Memory ............................................................................... 44 1.18.2. Anxiety ...................................................................................................... 45 1.18.3. Depression ................................................................................................ 45 1.18.4. Addiction ................................................................................................... 46 1.18.5. Appetite ..................................................................................................... 46 1.18.6. Pain .......................................................................................................... 46

1.19. Classification of ligands that bind to cannabinoid receptors .................................. 47 1.19.1. Cannabinoid receptor agonists .................................................................. 47

1.19.1.1. Classical cannabinoids ............................................................... 47 1.19.1.2. Non-classical cannabinoids ........................................................ 47 1.19.1.3. Aminoalkylindoles ....................................................................... 47 1.19.1.4. Eicosanoids/Endocannabinoids .................................................. 48

1.19.2. Cannabinoid receptor antagonists/ Inverse agonists ................................. 48 1.19.2.1. Diarylpyrazoles ........................................................................... 48 1.19.2.2. Other inverse agonists primarily active at CB1-Rs ....................... 48

1.20. Cannabinoid receptor 2 (CB2-R) ........................................................................... 53 1.20.1. CB2-R receptor signaling ........................................................................... 53

1.20.1.1. Adenylyl cyclase regulation......................................................... 53 1.20.1.2. Mitogen-activated protein kinase regulation ................................ 53

1.20.2. Therapeutic aspects of CB2-R modulators ................................................. 54 1.21. Brief overview of the test chemicals used in my research ..................................... 55

1.21.1. Benzophenanthridine alkaloids ................................................................. 55 1.21.2. Piperonyl butoxide (PBO) .......................................................................... 56 1.21.3. Methoprene ............................................................................................... 57 1.21.4. Phthalate esters ........................................................................................ 58 1.21.5. Tributyl tin (TBT) compounds .................................................................... 59

1.22. Rationale behind my research and the general approach ..................................... 62 1.22.1. Summary of objectives .............................................................................. 62

2. The actions of benzophenanthridine alkaloids, piperonyl butoxide and (S)-methoprene at the G-protein coupled cannabinoid CB1 receptor in vitro. .................................................................................................................... 64

2.1. Abstract ................................................................................................................ 64 2.2. Introduction .......................................................................................................... 65 2.3. Materials and Methods ......................................................................................... 67

ix

2.3.1. Radioligands, drugs and study compounds ............................................... 67 2.3.2. Animals ..................................................................................................... 67 2.3.3. Determination of the effects of study compounds on the binding of

[3H]CP55940 to CB1 receptors in mouse brain membranes....................... 67 2.3.4. Determination of the effects of study compounds on basal and

CP55940-stimulated [35S]GTPγS binding to mouse brain membranes ............................................................................................... 69

2.3.5. Data analysis ............................................................................................ 70 2.4. Results ................................................................................................................. 70 2.5. Discussion ............................................................................................................ 71 2.6. Figures and Tables ............................................................................................... 75

3. The G protein-coupled cannabinoid-1 (CB1) receptor of mammalian brain: Inhibition by phthalate esters in vitro. ................................................... 88

3.1. Abstract ................................................................................................................ 88 3.2. Introduction .......................................................................................................... 89 3.3. Materials and methods ......................................................................................... 91

3.3.1. Animals ..................................................................................................... 91 3.3.2. Investigation of the effects of phthalate esters on the binding of

[3H]CP55940 and [3H]SR141716A to CB1 receptors of mouse brain. ......... 92 3.3.3. Investigation of phthalate interference with CB1 receptor agonist-

stimulated [35S]GTPγS binding to the Gα-protein. ...................................... 93 3.3.4. Data analysis ............................................................................................ 95

3.4. Results ................................................................................................................. 95 3.4.1. Effects of phthalate esters on binding of [3H]CP55940 to CB1

receptors. .................................................................................................. 95 3.4.2. Effects of selected phthalate esters on binding of [3H]SR141716A to

CB1 receptors. ........................................................................................... 95 3.4.3. Influence of selected phthalates on the saturation binding of

[3H]SR141716A to CB1 receptors .............................................................. 96 3.4.4. Effects of selected phthalates on [3H]SR141716A kinetics ........................ 96 3.4.5. Effects of phthalates on CB1 receptor agonist-stimulated [35S]GTPγS

binding to the Gα-protein .......................................................................... 96 3.5. Discussion ............................................................................................................ 97 3.6. Note in added proof ............................................................................................ 100

3.6.1. Background ............................................................................................. 100 3.6.2. Experimental approach ........................................................................... 101 3.6.3. Results .................................................................................................... 101 3.6.4. Conclusion .............................................................................................. 101

3.7. Figures and Tables ............................................................................................. 102

4. Benzophenanthridine alkaloid, piperonyl butoxide and (S)-methoprene action at the cannabinoid-1 receptor (CB1-R) pathway of mouse brain: interference with [3H]CP55940 and [3H]SR141716A binding and modification of WIN55212-2-dependent inhibition of synaptosomal L-glutamate release. ............................................................................................ 115

4.1. Abstract .............................................................................................................. 115 4.2. Introduction ........................................................................................................ 116

x

4.3. Materials and Methods ....................................................................................... 118 4.3.1. Chemicals and supplies .......................................................................... 118 4.3.2. Animals ................................................................................................... 119 4.3.3. Isolation of membranes from mouse brain for binding studies ................. 119 4.3.4. Effects of benzophenanthridines, (S)-methoprene and piperonyl

butoxide on equilibrium binding of [3H]CP55940 and [3H]SR141716 to brain CB1 receptors ............................................................................. 120

4.3.5. Effect of benzophenanthridines, (S)-methoprene and piperonyl butoxide on the association and dissociation kinetics of [3H]CP55940 and [3H]SR141716A .......................................................... 121

4.3.6. Interaction of benzophenanthridines, methoprene and piperonyl butoxide with CB2 receptors in mouse spleen ......................................... 121

4.3.7. Preparation of synaptosomes from mouse whole brain ........................... 122 4.3.8. Release of L-Glutamate from synaptosomes........................................... 123 4.3.9. Analysis of radioligand binding data and glutamate release data ............ 124

4.4. Results ............................................................................................................... 124 4.4.1. Effects of benzophenanthridines, piperonyl butoxide and (S)-

methoprene on binding of [3H]SR141716A to CB1 receptors ................... 124 4.4.2. Influence of study compounds on the saturation binding of

[3H]SR141716A to CB1 receptors of mouse brain .................................... 125 4.4.3. Effects of sanguinarine, chelerythrine, piperonyl butoxide, and (S)-

methoprene on the kinetics of CB1 receptor-selective radioligand binding .................................................................................................... 125

4.4.4. Effects of study compounds on mouse spleen CB2 receptors as assessed by inhibition of [3H]CP55940 binding ....................................... 126

4.4.5. Effects of study compounds on WIN55212-2-dependent inhibition of 4-aminopyridine- (4-AP-) evoked release of L-glutamate from mouse brain synaptosomes ................................................................................ 126

4.5. Discussion .......................................................................................................... 127 4.6. Figures and Table .............................................................................................. 132

5. Effects of organotins on the CB1 receptor pathway of mouse brain in vitro. .................................................................................................................. 150

5.1. Introduction ........................................................................................................ 150 5.2. Materials and methods ....................................................................................... 151 5.3. Results ............................................................................................................... 152

5.3.1. Displacement of [3H]CP55940 binding to mammalian CB1 receptors by organotin compounds ......................................................................... 152

5.3.2. Basal and CP55940-stimulated [35S]GTPγS binding to the Gα subunit as influenced by tributyltin compounds ....................................... 152

5.3.3. Modulation by tributyltin acetate and phenylethynyl tributyltin of WIN55212-2-dependent inhibition of 4-aminopyridine-evoked release of L-glutamate from mouse brain synaptosomes ........................ 153

5.4. Discussion .......................................................................................................... 153 5.5. Figures and Table .............................................................................................. 156

xi

6. Conclusion and future prospects .................................................................... 162

References ................................................................................................................. 164

xii

List of Tables

Table 2.1 Inhibition of specific [3H]CP55940 binding to mouse brain membranes by isoquinoline type compounds and PMSF. Isoquinolines were present in the assay at 30 µM and PMSF was present at 0.5 mM. Data represent mean ± S.E.M. of 3 independent experiments. ...................................................................... 84

Table 2.2 Inhibition of 100 nM CP55940-stimulated and basal [35S]GTPγS binding to mouse brain membranes by AM251. Data represent mean ± S.E.M. of 3 independent experiments. ND = not determined. Results provided by Mr Saurabh Jain. ................................ 85

Table 2.3 Lack of effect of isoquinoline type compounds on CP55940-stimulated and basal [35S]GTPγS binding to mouse brain membranes. Study compounds were present in the assay at 40 µM. Data represent mean ± S.E.M. of 3 independent experiments. ........ 86

Table 2.4 Lack of effect of piperonyl butoxide and (S)-methoprene on the basal binding of [35S]GTPγS to mouse brain membranes. Values represent mean ± S.E.M. of 3 independent experiments. ....................... 87

Table 3.1 Inability of PMSF to influence the inhibitory effects of n-butylbenzylphthalate (nBBP) and di-n-butylphthalate (DnBP) on [3H]CP55940 binding to mouse brain membranes. Phthalate esters were present in the assay at 20 µM and PMSF was used at 50 µM. Each value represents the mean ± S.E.M. of 3-6 independent experiments. .................................................................... 113

Table 3.2 Inhibitory effects of n-butylbenzylphthalate (nBBP), di-n-butylphthalate (DnBP), diethylhexylphthalate (DEHP), mono-isohexylphthalate (MiHP) and mono-n-butyl phthalate (MnBP) on the specific binding of [3H]SR141716A to mouse brain membranes. Diesters were present at concentrations producing 50% inhibition of [3H]CP55940 binding. Each value represents the mean ± S.E.M. of 3 independent experiments. ..................................... 114

xiii

Table 4.1 Inhibitory effects of chelerythrine, sanguinarine, piperonyl butoxide and (S)-methoprene on spleen CB2 receptors as determined with [3H]CP55940. Each study compound was added at a concentration that achieved an IC50 for [3H]CP55940 binding to brain CB1 receptors (Dhopeshwarkar et al. 2011). All values represent mean percentage inhibition ± S.E.M. of at least 3 independent experiments. Parallel experiments with [3H]CP55940 corroborated our previously published IC50s at brain CB1 receptors (2.2 µM chelerythrine gave 49.03 ± 0.94 % inhibition, 1.2 µM sanguinarine gave 51.33 ± 0.49 % inhibition, 8.2 µM piperonyl butoxide gave 47.50 ± 1.17 % inhibition and 16.4 µM methoprene gave 50.22 ± 1.10 % inhibition). ...................................... 149

Table 5.1 Inhibitory effects of tributyl and triphenyltins on the binding of [3H]CP55940 to CB1 receptors in mouse brain. All values are as IC50s (with 95% confidence intervals in brackets) calculated from curves based on at least 3 independent experiments except for triphenyltin chloride where the IC50 was estimated from 2 independent experiments). ................................................................... 161

xiv

List of Figures

Figure 1.1 The spread of the use of cannabis across the globe (Adapted from Zuardi, 2006). ........................................................................................... 3

Figure 1-2 Structure of two important phytocannabinoids. Structures redrawn using ChemDraw Ultra 11.0 from structures reported in Pertwee et al. (2010). ................................................................................................. 6

Figure 1.3 Two dimensional representation of the CB1-R (Adapted from Shim et al., 2011). ........................................................................................... 12

Figure 1.4 Diagramatic representation of the C terminal domain of the CB1-R (Adapted from Stadel et al., 2011) .......................................................... 13

Figure 1.5 Structures of prominent endocannbinoids (All structures redrawn using ChemDraw Ultra 11.0 from Kano et al., 2009). ............................. 27

Figure 1.6 Transacylation-phosphodiesterase pathway for biosynthesis of anandamide (Adapted from Cadas et al., 1997). .................................... 29

Figure 1.7 Metabolic pathways for biosynthesis of 2-AG (Adapted from Kano et al., 2009). ........................................................................................... 31

Figure 1.8 Blockade of DSI by CB1-R antagonists. .................................................. 37

Figure 1.9 The pathway involved in the termination of endocannabinoid-mediated short term depression (eCB-STD) (Adapted from Kano et al., 2009). ........................................................................................... 40

Figure 1.10 Structures of ∆9-THC, ∆8-THC, HU210, DALN, CP47497, CP55244, CP55940, WIN55212-2, JWH015 and L-768242. All structures redrawn using ChemDraw 11.0 ultra from Howlett et al. (2002). ................................................................................................... 49

Figure 1.11 Structures of anandamide, 2-AG ether and 2-AG. All structures redrawn using ChemDraw Ultra 11.0 from Howlett et al. (2002). ............ 50

Figure 1.12 Structures of SR141716A, AM251, AM281, LY320135 and AM630. All structures redrawn using ChemDraw Ultra 11.0 from Howlett et al. (2002). .............................................................................. 51

Figure 1.13 Structures of (S)-methoprene, piperonyl butoxide, sanguinarine, chelerythrine, nBBP and DnBP. Structures redrawn using ChemDraw 11.0 from Dhopeshwarkar et al. (2011) and Bisset et al., (2011). .............................................................................................. 52

Figure 1.14 Structures of selected phthalate esters and tributyl tin compounds. All structures redrawn using ChemDraw Ultra 11.0. ............................... 61

xv

Figure 2.1 The structures of sanguinarine, berberine, papavarine and possible comparison of conformations of piperonyl butoxide and (S)-methoprene with anandamide and 2-arachidonoyl glycrol. Also possible comparison of sanguinarine and (S)-methoprene with ∆9-tetrahydrocannabinol and ∆9-tetrahydrocannabivarin. ............................ 76

Figure 2.2 Concentration-dependent inhibition of [3H]CP55940 binding to mouse brain CB1 receptors by sanguinarine and chelerythrine. Values represent mean ± S.E.M. of at least 3 independent experiments each performed in duplicate. Ki values were 0.38 µM (sanguinarine) and 0.57 µM (chelerythrine). ........................................... 77

Figure 2.3a Inhibition of A) CP55940-stimulated and B) basal binding of [35S]GTPγS to mouse brain membranes by chelerythrine. Values represent mean ± S.E.M. of 3 independent experiments each performed in triplicate. ............................................................................ 78

Figure 2.3b Inhibition of A) CP55940-stimulated and B) basal binding of [35S]GTPγS to mouse brain membranes by chelerythrine. Values represent mean ± S.E.M. of 3 independent experiments each performed in triplicate.Basal binding data provided by Mr Saurabh Jain. ....................................................................................................... 79

Figure 2.4a Inhibition of A) CP55940-stimulated and B) basal binding of [35S]GTPγS to mouse brain membranes by sanguinarine. Values represent mean ± S.E.M. of 3 independent experiments each performed in triplicate. ............................................................................ 80

Figure 2.4b Inhibition of A) CP55940-stimulated and B) basal binding of [35S]GTPγS to mouse brain membranes by sanguinarine. Values represent mean ± S.E.M. of 3 independent experiments each performed in triplicate. Basal binding data provided by Mr Saurabh Jain. ......................................................................................... 81

Figure 2.5a A) Concentration-dependent inhibition of [3H]CP55940 binding to mouse brain CB1 receptors by (S)-methoprene and piperonyl butoxide. Ki values were 2.13 µM (methoprene) and 4.25 µM (piperonyl butoxide). B) Inhibition of CP55940-stimulated binding of [35S]GTPγS to mouse brain membranes by (S)-methoprene and piperonyl butoxide. Values represent mean ± S.E.M. of 3 independent experiments each performed in triplicate. .......................... 82

Figure 2.5b A) Concentration-dependent inhibition of [3H]CP55940 binding to mouse brain CB1 receptors by (S)-methoprene and piperonyl butoxide. Ki values were 2.13 µM (methoprene) and 4.25 µM (piperonyl butoxide). B) Inhibition of CP55940-stimulated binding of [35S]GTPγS to mouse brain membranes by (S)-methoprene and piperonyl butoxide. Values represent mean ± S.E.M. of 3 independent experiments each performed in triplicate. .......................... 83

xvi

Figure 3.1 (a-f) The structures of phthalate diesters: n-butylbenzylphthalate (nBBP); di-n-hexylphthalate (DnHP); di-n-butylphthalate (DnBP); di-ethylhexylphthalate (DEHP); di-isooctylphthalate (DiOP) and di-n-octylphthalate (DnOP).(g-i) The structures of phthalate monoesters: mono-2-ethylhexyl-phthalate (M2EHP), mono-isohexyl-phthalate (MiHP) and mono-n-butyl-phthalate (MnBP). All structures have been redrawn from Bissett et al. (2011) using IsisDraw. .............................................................................................. 102

Figure 3.2 Inhibitory effects of phthalate esters (DnBP, nBBP, DnOP, MiHP and MnBP) on the binding of [3H]CP55940 to mouse brain CB1 receptors in vitro. Each point represents the mean ± SEM of 3 independent experiments. Results provided by Ms Kathleen M. Bisset. .................................................................................................. 103

Figure 3.3 Inhibitory effects of phthalate esters (DEHP, DnHP, DiOP and M2EHP) on the binding of [3H]CP55940 to mouse brain CB1 receptors in vitro. Each point represents the mean ± SEM of 3 independent experiments. Results provided by Ms Kathleen M. Bisset. .................................................................................................. 104

Figure 3.4 The effect of nBBP and DnBP (both at 35 µM) on the equilibrium binding of of [3H]SR141716A to CB1 receptors of mouse whole brain. Kd and Bmax values are displayed for each treatment and 95% confidence intervals were as follows: control (Kd 0.628 to 0.859. Bmax 0.303 to 0.343), nBBP (Kd 0.761 to 1.333. Bmax 0.176 to 0.229) and DnBP (Kd 0.624 to 0.846. Bmax 0.120 to 0.136). R2 values were 0.9877 (control), 0.9756 (nBBP) and 0.9887 (DnBP). Data points represent the means ± SEMs of 3 independent experiments (most SEM bars are obscured by data symbols). ............. 105

Figure 3.5a Influence of nBBP (35 µM) and DnBP (50 µM) on the time course of association of [3H]SR141716A with CB1 receptors of mouse brain. In a) membranes received the standard 15 min preincubation with phthalate esters prior to [3H]SR141716A addition. In b) the phthalate ester and [3H]SR141716A were applied simultaneously. Data points represent the means ± SEMs of 3 independent experiments (most SEM bars are obscured by data symbols). ...................................................................................... 106

Figure 3.5b Influence of nBBP (35 µM) and DnBP (50 µM) on the time course of association of [3H]SR141716A with CB1 receptors of mouse brain. In a) membranes received the standard 15 min preincubation with phthalate esters prior to [3H]SR141716A addition. In b) the phthalate ester and [3H]SR141716A were applied simultaneously. Data points represent the means ± SEMs of 3 independent experiments (most SEM bars are obscured by data symbols). ...................................................................................... 107

xvii

Figure 3.6 Dissociation of the [3H]SR141716A:CB1 receptor complex (initiated by challenge with 5 µM AM251) in the absence (control) or in the presence of 35 µM nBBP or 50 µM DnBP. Data represent mean ± SEM of at least 3 independent experiments, each performed in triplicate. .......................................................................... 108

Figure 3.7 Inhibition of CP55940-stimulated binding of [35S]GTPγS to mouse whole brain membranes by phthalate esters. Phthalate esters were assayed at 75 µM throughout. Each column represents the mean, and error bar the ± SEM of 7 independent experiments. ............ 109

Figure 3.8 Relationship between the ability of study compounds to inhibit the binding of [3H]CP55940 and CP55940-stimulated binding of [35S]GTPγS in mouse whole brain membrane fractions. All assays were performed 75 µM; r2 = 0.7844. ..................................................... 110

Figure 3.9 With WIN55212-2 present, BBP (at 30 µM but not 5 µM) enhances 4-AP-evoked L-glutamate release above the level produced by 4-AP alone. ............................................................................................. 111

Figure 3.10 With WIN55212-2 present, MnBP (both at 30 µM and 5 µM) does not enhance 4-AP-evoked L-glutamate release above the level produced by 4-AP alone. ...................................................................... 112

Figure 4.1 Concentration dependency of inhibition by chelerythrine (open circles), sanguinarine (solid circles), piperonyl butoxide (solid triangles) and (S)-methoprene (squares) on [3H]SR141716A binding to mouse brain CB1 receptors. IC50 and 95% confidence interval values are provided in Section 4.4.1. ....................................... 132

Figure 4.2 Effect of chelerythrine (1 µM; open circles), sanguinarine (1 µM; solid circles), piperonyl butoxide (30 µM; solid triangles) and (S)-methoprene (60 µM; squares) on equilibrium binding of [3H]SR141716A to mouse brain CB1 receptors. Control data points are identified by the diamond symbols. Kd values (as nM): control 0.51 ± 0.04; chelerythrine 0.47 ± 0.08; sanguinarine 0.46 ± 0.04; (S)-methoprene 1.5 ± 0.6 and piperonyl butoxide 2.5 ± 1.1. Bmax values (as pmol [3H]SR141716A/mg protein): control 0.79 ± 0.02; chelerythrine 0.32 ± 0.02; sanguinarine 0.50 ± 0.01; (S)-methoprene 0.44 ± 0.08 and piperonyl butoxide 0.56 ± 0.13. ............... 133

xviii

Figure 4.3 Effect of chelerythrine (2.5 µM; open circles), sanguinarine (1.5 µM; solid circles), piperonyl butoxide (10 µM; solid triangles) and (S)-methoprene (20 µM; squares) on equilibrium binding of [3H]CP55940 to mouse brain CB1 receptors. Control data points are identified by the diamond symbols. Kd values (as nM): control 0.36 ± 0.07; chelerythrine 2.32 ± 0.43; sanguinarine 2.28 ± 0.77; (S)-methoprene 1.37 ± 0.25 and piperonyl butoxide 0.34 ± 0.19. Bmax values (as pmol [3H]SR141716A/mg protein): control 0.6 ± 0.03; chelerythrine 0.65 ± 0.06; sanguinarine 0.63 ± 0.11; (S)-methoprene 0.25 ± 0.02 and piperonyl butoxide 0.35 ± 0.05. ............... 134

Figure 4.4a Influence of study compounds on the time course of association of [3H]SR141716A and [3H]CP55940 with CB1 receptors of mouse brain. In a) membranes received a standard 15 min preincubation with sanguinarine (2.5 µM), chelerythrine (2.5 µM), piperonyl butoxide (30 µM) and (S)-methoprene (30 µM) prior to [3H]SR141716A addition. .................................................................... 135

Figure 4.4b Influence of study compounds on the time course of association of [3H]SR141716A and [3H]CP55940 with CB1 receptors of mouse brain.The same study compound concentrations were applied simultaneously with [3H]SR141716A.. .................................................. 136

Figure 4.4c The effects of benzophenanthridines (5 µM), piperonyl butoxide (20 µM) and (S)-methoprene (20 µM) on the association of [3H]CP55940 under preincubation conditions are shown Symbols: diamonds = control; solid circles = sanguinarine; open circles = chelerythrine; triangles = piperonyl butoxide and squares = (S)-methoprene.Data points represent the means ± SEMs of 3 independent experiments (a number of SEM bars are obscured by data symbols) ....................................................................................... 137

Figure 4.5a The influence of study compounds on the dissociation of CB1 receptor-selective radioligands. Figure 4.5a shows the effects of piperonyl butoxide (30 µM) and (S)-methoprene (60 µM) on the dissociation of [3H]SR141716A when initiated by challenge with a saturating concentration (5 µM) of SR141716A. ................................... 138

Figure 4.5b The influence of study compounds on the dissociation of CB1 receptor-selective radioligands. Figure 4 5b, defines the effects of sanguinarine (5 µM), chelerythrine (5 µM), piperonyl butoxide (30 µM) and (S)-methoprene (60 µM) when added alone on the dissociation of [3H]SR141716A from the [3H]SR141716A:CB1 receptor complex .................................................................................. 139

xix

Figure 4.5c The influence of study compounds on the dissociation of CB1 receptor-selective radioligands. In Figure 4 5c, the effects of sanguinarine (5 µM), chelerythrine (5 µM), piperonyl butoxide (20 µM) and (S)-methoprene (60 µM) on the dissociation of [3H]CP55940 when initiated by application of a saturating concentration (5 µM) of CP55940 are given. Symbols: diamonds = control; solid circles = sanguinarine; open circles = chelerythrine; triangles = piperonyl butoxide and squares = (S)-methoprene. Data represent mean ± SEM of at least 3 independent experiments, each performed in triplicate......................... 140

Figure 4.6 Relationship between concentration of (S)-methoprene and inhibition at CB2 receptors of mouse spleen based on interference with [3H]CP55940 binding. .................................................................... 141

Figure 4.7a Inhibition of 50 µM veratridine-evoked release of L-glutamate from mouse brain synaptosomes by 5 µM tetrodotoxin (TTX) ...................... 142

Figure 4.7b Failure of 5 µM TTX to modify 3 mM 4-AP-evoked release of L-glutamate from synaptosomes. ............................................................ 143

Figure 4.8 Partial inhibition of 4-AP-evoked release of L-glutamate from synaptosomes by the CB1-R agonist WIN55212-2, and full relief of WIN55212-2-dependent inhibition by the CB1-R antagonist AM251. ................................................................................................ 144

Figure 4.9 With WIN55212-2 present, sanguinarine (at 2 µM but not 0.25 µM) enhances 4-AP-evoked L-glutamate release above the level produced by 4-AP alone. ...................................................................... 145

Figure 4.10 With WIN55212-2 present, chelerythrine (at 2 µM but not 0.25 µM) enhances 4-AP-evoked L-glutamate release above the level produced by 4-AP alone. ...................................................................... 146

Figure 4.11 With WIN55212-2 present, (S)-methoprene (at 25 µM but not 5 µM) enhances 4-AP-evoked L-glutamate release above the level produced by 4-AP alone. ...................................................................... 147

Figure 4.12 With WIN55212-2 present, piperonyl butoxide (at 25 µM but not 5 µM) enhances 4-AP-evoked L-glutamate release above the level produced by 4-AP alone. ...................................................................... 148

Figure 5.1 Structures of tributyl and triphenyltin compounds examined in the present investigation. Structures were constructed using Isis Draw. ................................................................................................... 156

xx

Figure 5.2 Concentration-dependent inhibition of specific [3H]CP55940 binding to mouse brain CB1 receptors by tributyltin benzoate, tributyltin acetate and phenylethynyl tributyltin. Each data point represents the mean ± S.E.M. inhibition of specific [3H]CP55940 binding for at least three independent assays, each performed in triplicate. Experiments conducted by Mr. Saurabh Jain. This figure was originally published in the M.Sc. thesis of Mr. Saurabh Jain (Simon Fraser University, 2011). .......................................................... 157

Figure 5.3 Concentration-dependent inhibition of CP55940 (100 nM)-stimulated [35S]GTPγS binding by tributyltin benzoate and phenylethynyl tributyltin. Each data point represents the mean ± S.E.M. percentage inhibition of CP55940 stimulated [35S]GTPγS binding determined by three independent assays each performed in triplicate. These experiments were conducted by Mr Saurabh Jain and this figure was originally published in the M.Sc. thesis of Mr. Saurabh Jain (Simon Fraser University, 2011). .............................. 158

Figure 5.4 Modulation of WIN55212-2-dependent inhibition of 4-aminopyridine (4-AP-)-evoked release of L-glutamate from mouse brain synaptosomes by tributyltin acetate (TBT acetate). Typical release profiles are displayed with mean % changes (± SEM) to 4-AP-evoked and control release in the adjacent table. ........................... 159

Figure 5.5 Modulation of WIN55212-2-dependent inhibition of 4-aminopyridine (4-AP-)-evoked release of L-glutamate from mouse brain synaptosomes by phenylethynyl tributyltin (TBPE tin). Typical release profiles are displayed with mean % changes (± SEM) to 4-AP-evoked and control release in the adjacent table. .......... 160

xxi

Glossary

2-AG 2-Arachidonyl glycerol

2-AGE 2-Arachidonyl glycerol ether

4-AP 4-Aminopyridine

AEA Anandamide

Bmax Maximum concentration of binding sites

BSA Bovine serum albumin

CBD Cannabidiol

CB1-R Cannabinoid receptor-1

CB2-R Cannabinoid receptor-2

CHEL Chelerythrine

DAGL Diacylglycerol lipase

DSE Depolarization-induced suppression of excitation

DSI Depolarization-induced suppression of inhibition

DMSO Dimethylsulfoxide

EDTA Ethylenediamine tetraacetic acid

EGTA Ethylene glycol-bis(2-aminoethyl)-N,N,N’,N’-tetraacetic acid

EPSCs Excitatory post synaptic currents

eCB-STD Endocannabinoid mediated short term depression

eCB-LTD Endocannabinoid mediated long term depression

FAAH Fatty acid amide hydrolase

GPCR G-protein coupled receptor

GTP Guanosine-5’-triphosphate

GDP Guanosine-5’-diphosphate

GABA γ-Aminobutyric acid

IC50 Concentration effective in producing 50% inhibition

IPSCs Inhibitory post synaptic currents

KCl Potassium chloride

Kd Dissociation constant

Lys Lysine

L-GLU L-glutamic acid

MAGL Monoacylglycerol lipase

xxii

MnBP Mono-n-butyl phthalate

METHO Methoprene

NADA N-Arachidonyl dopamine

NAPE N-Arachidonyl-phosphatidylethanolamine

nBBP n-Butylbenzylphthalate

PLD Phospholipase D

PI Phosphatidyl inositol

PLA1 Phospholipase A1

PMSF Phenylmethane sulfonyl fluoride

PBO Piperonyl butoxide

SANG Sanguinarine

TTX Tetrodotoxin

TBT Tributyltin

∆9 -THC ∆9- Tetrahydrocannabinol

∆8-THC ∆8- Tetrahydrocannabinol

VGSCs Voltage-gated sodium channels

VTD Veratridine

1

1. Introduction

1.1. Historical significance of cannabis use and cannabinoids

Cannabis sativa and its preparations have been used throughout the millennia for

recreational and various therapeutic purposes (Hollister, 2001). Cannabis sativa is one

of the oldest cultivated plants in the history of humankind dating back at least 10,000

years (Jiang, 2006).

The history of cannabis use can be broadly classified into three eras, the early

Chinese/Indian era, the early Christian era through to the 18th century, and the era of

Western medicine of the 19th and 20th centuries (Zuardi, 2006).

1.1.1. The early Chinese/Indian era

The earliest references to the use of different parts of the cannabis plant were

documented in the Han dynasty in China (Zuardi, 2006). Fibers obtained from the stem

were used for preparing ropes, strings and paper, while fruits were used as food by the

ancient Chinese (Li, 1973).

The world’s oldest pharmacopoeia, Pen-ts’ao ching documented the use of

cannabis as a medicine for the treatment of rheumatic pain, constipation and disorders

of the female reproductive system (Zuardi, 2006). Other evidence for the use of

cannabis in ancient China was reported by Jiang et al. (2006), where a clay bowl

containing cannabis was discovered in the 2500 old Yanghai tombs of Northwestern

China. It is believed that stores such as this were probably used for medicinal purposes

and psychomanupulation (Russo et al., 2008).

The ancient Indian culture (around 1000 years B.C.) regularly employed

cannabis for medicinal and religious reasons (Zuardi, 2006). In ancient Indian medicine,

2

the plant was used for various purposes including induction of analgesia and hypnotic

states and reducing the occurence of epileptic seizures. It was also used as an

antiparasitic, an antispasmodic, an antibiotic, an expectorant and an aphrodisiac (Zuardi,

2006).

1.1.2. The period encompassing the early Christian era through to the 18th century

During this period, the use of cannabis gained increasing acceptance throughout

the Middle East and Africa. Around 1000 A.D., Arabic medical compendiums described

the use of cannabis as a plant beneficial in the treatment of diuretic disorders and

gastrointestinal problems including flatulence (Zuardi, 2006). In the 16th century,

cannabis was introduced to South America through the arrival of African slaves, while

Arab traders introduced cannabis to the European sub-continent firstly in Spain and then

to various Mediterranean countries including Italy (Zuardi, 2006).

1.1.3. The Western medicine era of the 19th and 20th centuries

Pioneering scientific studies and several books published by the Irish physicist

William B. O’Shaughnessy and the French psychiatrist Jacques-Joseph Moreau

facilitated the rapid introduction of cannabis to Western medicine. In their books, they

documented a range of therapeutic uses as well as psychomimetic and experimental

manipulations based on the use of cannabis and hashish (cannabis resin) (Di Marzo,

2006).

Western medicine readily accepted many of their proposed uses of cannabis

since during this period there were very few realistic therapeutic options available for the

treatment of disorders such as rheumatism, muscular spasms, pain and convulsive

states (Zuardi, 2006). Figure 1.1 summarizes the global spread of cannabis use from

South East Asia through Africa and South America to Europe and the USA (as published

by Zuardi (2006)), while Table 1.1 details the main landmarks in cannabinoid research

up until the late 1990s (Novarro and Fonseca, 1998).

3

Figure 1.1 The spread of the use of cannabis across the globe (Adapted from Zuardi, 2006).

4

Table 1.1 Major advances in the use of cannabis and research on the cannabinoid system of mammalian brain (Adapted from Navarro and Fonseca, 1998).

Event Date

Medical, ceremonial and recreational uses of Cannabis 3000 B.C. onwards

Isolation of psychoactive cannabinoids 1964

Discovery of synthetic cannabinoids 1980 onwards

Discovery of the cannabinoid-1 receptor ( CB1-R) in mammalian brain

1988

Mapping of the CB1-R in mammalian brain 1990

Cloning of the CB1-R 1990

Neuropharmacology of the CB1-R 1988-1995

Discovery and isolation of a natural cannabinoid anandamide in brain

1992-1995

Synthesis of diarylpyrazole CB1 receptor antagonists, (e.g. SR141716A and AM251)

1994

Isolation and identification of 2-arachidonyl glycerol (2-AG) as another important endocannabinoid

1995-1997

Functional neuroanatomy of CB1 receptors 1996-1997

Delineation of anandamide biosynthesis and its mechanism of uptake

1997

5

1.2. Cannabinoids

Gaoni and Mechoulam (1964) identified ∆9-tetrahydrocannabinol (∆9-THC)

(Figure 1.2) as the main psychoactive component of Cannabis sativa, a discovery that

eventually led to the synthesis of various analogs of ∆9-THC (Howlett et al., 2002).

Compounds that mimic the actions of the cannabis derivative ∆9-THC are defined

as cannabinoids (Howlett et al., 2002). A critical advance in cannabinoid research

occured with the discovery of specific membrane receptors to which ∆9-THC actively

binds in brain tissue (Devane et al., 1988). Matsuda et al. (1990) cloned and

characterized the first cannabinoid-1 receptor (CB1-R) while a cannabinoid-2 receptor

(CB2-R) was identified by Munro et al. in 1993.

Before the discovery of these receptors, the psychoactive actions of ∆9-THC and

related cannabinoids were assumed to arise from their ability to 1) dissolve in lipids

(Seeman et al., 1972), 2) modify the fluidity of synaptic plasma membranes (Hillard et

al., 1985) and 3) intercalate with lipids and other components of neuronal plasma

membranes (Pertwee, 1988).

Both CB1-Rs and CB2-Rs belong to the rhodopsin-like subfamily of receptors

which are G protein-coupled receptors (GPCRs) with seven transmembrane spanning

domains (TMH1-7). CB1-Rs and CB2-Rs were found to be sensitive to inhibition by

Pertussis toxin treatment, indicating that the response to cannabinoid drugs was

mediated through the Gi/o family of G proteins (Howlett et al., 1986).

Moreover, both CB1-Rs and CB2-Rs are found to have varied tissue distributions

in vertebrates. CB1-Rs are densely located in many regions of the central nervous

system with much lower levels in kidney, testis, uterus, heart and vascular tissue. On

the other hand, CB2-Rs are abundantly expressed in tissues of the immune system,

including spleen, tonsils and haematopoietic cells, but are found at much lower levels in

central nervous system (CNS) (Kano et al., 2009; Brown, 2007).

6

∆9-Tetrahydrocannabinol (∆9-THC)

∆8-Tetrahydrocannabinol (∆8-THC)

Figure 1-2 Structure of two important phytocannabinoids. Structures redrawn using ChemDraw Ultra 11.0 from structures reported in Pertwee et al. (2010).

7

G protein-coupled CB1-Rs and G protein-coupled CB2-Rs are differentiated on

the basis of predicted amino acid sequence, signaling mechanisms, affinity towards

specific agonists and antagonists and tissue distribution. They each share 48% amino

acid sequence homology and both have their G proteins coupled to adenylyl cyclase and

mitogen-activated protein kinase (MAPK) (Howlett et al., 2002). The CB1-R is larger than

the CB2-R with 13 more amino acid residues on the C terminal, an extra 72 amino acid

residues on the N terminal and 15 additional residues on the third extracellular loop

(Childers, 2006).

These G protein-coupled cannabinoid receptors are activated by certain

cannabis-derived compounds as well as endogenous lipid molecules termed

endocannabinoids. The endocannabinoids, their receptors and associated biochemical

machinery (including precursors, critical biosynthetic enzymes, degradative enzymes,

mediators and transporters) collectively constitute the endocannabinoid system (ECS).

The ECS represents a highly conserved system within all vertebrate phyla as well as

some invertebrates, with subtle structural differences in the structure of receptors,

enzymes and other components, thus underscoring the importance of the ECS for

survival of many life forms (De Petrocellis et al., 2004).

1.2.1. G protein-coupled receptors (GPCRs) and their activation cycle

GPCRs are seven transmembrane spanning receptors and are coupled to

specific heterotrimeric guanine nucleotide-binding proteins (G proteins) (Drake et al.,

2006). G proteins transduce an extracellular signal to an intracellular effector (Drake et

al., 2006). These receptors represent an attractive target for drug discovery and it has

been estimated that nearly half of the drugs marketed today target GPCRs (Kroeze et

al., 2003).

G proteins are made up of a monomer (Gα subunit) and dimer (Gβ and Gγ

subunit). In their inactive state, the Gα subunit is bound to guanosine diphosphate (GDP)

and exists as Gα(GDP)βγ (Harrison and Traynor, 2003). When activated, the Gα subunit

exchanges GDP for guanosine-5’-triphosphate (GTP) and this binary complex (Gα-GTP)

then detaches from the Gβγ subunit to act on different effectors (Griffin et al., 1998;

8

Harrison and Traynor, 2003). Inactivation of GPCRs occurs by the intrinsic GTPase

activity of the Gγ subunit which hydrolyses GTP to GDP. Finally, Gα-GDP and Gβγ

subunits recombine to form the inactive Gα(GDP)βγ.

1.2.2. The [35S]GTPγS binding assay

The [35S]GTPγS binding assay is a functional assay that can be employed to

measure the extent of G protein activation following the binding of ligand(s) to the GPCR

(Breivogel and Childers, 2000; Breivogel et al., 2001; Harrison and Traynor, 2003). It is

an excellent assay to measure the primary functional event that immediately follows the

activation of the GPCR by its ligand (Harrison and Traynor, 2003).

This assay is characterized by the replacement of endogenous GTP by

[35S]GTPγS which binds to Gα subunit to form the Gα[35S]GTPγS complex. The γ-

thiophosphate bond on [35S]GTPγS is highly resistant to GTPase-mediated hydrolysis

and therefore inactivation of the GPCR cycle is blocked and the extent of activation can

be conveniently quantified by measuring the [35S]-label bound (Griffin et al., 1998;

Harrison and Traynor, 2003).

1.3. Other cannabinoid receptors

Besides CB1-Rs and CB2-Rs, other two GPCRs, GPR55 and GPR119 (Lambert

and Muccioli, 2007) have been proposed as novel cannabinoid receptors based on their

affinity towards endocannabinoids. There is much ongoing debate in the scientific

community regarding classification of these receptors as genuinely cannabinoid

selective (Okunu and Yokomizo, 2011). Oka et al., (2007) reported the activation of

GPR55 by lysophosphatidylinositol derivatives but not cannabinoids. In marked

contrast, Lauckner et al., (2008) found that AEA and ∆9-THC increased intracellular

calcium in a cell line expressing GPR55. Nevertheless, this review will be focused

mainly on CB1-Rs and CB2-Rs. Phylogenetic analysis by Brown (2007) revealed that

CB1-Rs and CB2-Rs belong to family of lipid receptors (formerly endothelial

differentiation gene receptors (EDG)) which are activated by acylethanolamide

9

analogues typified by 2-arachidonylglycerol (2-AG) and anandamide

(arachidonylethanolamide, AEA).

Sharir et al., (2010) described the work of researchers at AstraZeneca who found

that nanomolar concentrations of cannabinoid agonists stimulate [35S]GTPγS binding

and this response was antagonized by cannabidiol, a natural product cannabinoid

receptor antagonist. The ion channel, transient receptor potential vanilloid 1 (TRPV1)

share several similarities with cannabinoid receptors in terms of intracellular signaling,

shared ligand and tissue distribution, their role in pathophysiological conditions and

binding of the endocannabinoid/endovanilloid anandamide (AEA) (Starowicz et al.,

2007).

1.4. Cannabinoid-1 Receptors (CB1-Rs)

1.4.1. The structure and activation of CB1-Rs

The first cloning and expression of a 473 amino acid CB1-R from rat brain was

achieved by Matsuda et al. (1990) (Figure 1.3). The human homolog of 472 amino acids

was reported by Gerard et al. (1990) and a 473 amino acid sequence from mouse brain

was identified by Chakrabarti et al. (1995). Significantly these three CB1-Rs exhibit 97-

99% amino acid sequence homology (Kano et al., 2009). Like any GPCR, the CB1-R is

an integral membrane protein consisting of seven hydrophobic transmembrane helices

(7TMH) linked by three extracellular (E1-3) and three intracellular loops (I1-3) which are

flanked by an N-terminal on the periplasmic domain and a C terminal on the cytoplasmic

domain (Montero et al., 2005). Between the cytoplasmic extension of TMH7 and the

proximal C terminus lies another helix designated helix 8, which runs parallel to the

cytosolic membrane surface (Patny et al., 2006). The cytoplasmic regions are

responsible for G protein binding, desensitization and cellular signal trafficking. Binding

of an agonist triggers activation of the heterotrimeric G proteins by exchanging GTP for

GDP on the α subunit. This leads to the dissociation of G proteins from receptors and

cleavage of α and β/γ subunits which in turn modulate downstream effectors (Stadel et

al., 2011). According to the two state model proposed for GPCRs, CB1-Rs exist in the

active (R) state and the inactive (R*) state (Samama et al., 1993; Gullapalli, 2010).

10

These states are in equilibrium in the absence of ligand and, following ligand binding, the

equilibrium can shift to either state. Thus, an agonist will actively bind to the active state

of this receptor while an inverse agonist will bind to the inactive state. A classical

antagonist will be overall neutral having an equal affinity towards both R and R* the state

of the receptors. These properties can be conveniently studied in CB1-Rs due to the

availability of selectively acting agonists (e.g. ∆9-THC, CP55940 and WIN55212-2),

antagonists (e.g. cannabidiol and AM251) and the inverse agonist (SR141716A)

(Gullapalli, 2010; Gatley et al., 1997; Herkenham et al., 1990; Pertwee, 2006; Rinaldi-

Carmona et al., 1994). Studies by various research groups have greatly improved our

understanding the role of CB1 transmembrane helices (TMH), extracellular loops (ECs)

in particular the E2 loop and the carboxyl terminus in cannabinoid binding and receptor

activation (Shim et al., 2011a; Shim et al., 2011b, McAllister, 2003; Ahn et al., 2009;

Stadel et al., 2011). However, despite much effort, little information is available on the

way in which ligands orient and dock at their respective active sites within the CB1-R

binding pocket. Chin et al. (1998) and Song and Bonner (1996) showed that the

hydrogen bonding interaction between residue K192 on the TMH and the phenolic

hydroxyl group of CP55940 and the carboxamide oxygen of the inverse agonist

SR141716A (Hurst et al., 2002, 2006) were critical for ligand binding. Moreover, residue

S383, which has been proposed to induce a bend in the TMH7, again appears essential

for agonist (CP55940) binding (Kapur et al., 2007). In addition, C386 (on TMH7) has

also been implicated as a critical amino acid for SR141716A binding (Fay et al., 2005).

McAllister et al., (2003) demonstrated that aromaticity at TMH5 (imparted by

residues such as F201, W280 and W357) was important for accomodating the agonist

CP55940 and inverse agonist SR141716A within the binding pocket of the CB1-R. By

employing molecular modeling, McAllister and associates also proposed that the binding

pocket for various ligands was primarily located in the hydrophobic transmembrane helix

bundle of the receptor. In related studies, Shi and Javitch (2004) found that the

extracellular loops also play a vital role regards ligand recognition, ligand sensitivity and

access of the ligand to the binding pocket of dopamine receptors (which are also a

GPCRs). This observation also proved relevant for CB1-Rs, since Ahn and coworkers

(2009) showed that the second extracellular loop (E2) was important for ligand binding

and receptor localization. They were also able to demonstrate that alanine (Ala)

11

mutations in the C terminal residues on E2 led to reduced agonist binding but had no

effect on the binding of inverse agonists. In addition, Bertalovitz et al. (2010) reported

that point mutations on the C terminal region of E2 can lead to loss of agonist and

antagonist binding capacity of CB1-Rs but have no effect on inverse agonist binding.

Shim et al. (2011) further elucidated the structure of the E2 loop by using a

combination of simulated annealing and molecular dynamics simulations. They studied

the molecular structure of E2 in two forms, disulphide (E2disulphide) and dithiol (E2dithiol). It

was found that that E2disulphide helical segment has a amphipathic alignment (at the

membrane:water interface) which stabilizes the receptor and imparts greater flexibility

compared to E2dithio. This further led to E2/TMH coupling and rearrangement of the TMH.

These coupling/interactions and TMH rearrangement are important for receptor

activation. However, the extent of this coupling was distinct in both the forms of E2.

Since in CB1 E2disulphide E2 offers more flexibility, the C terminal region of E2 inserts into

the extracellular H3/H5 region causing H5 to move away from H3 and H6 to move into

H3 at the extracellular region, leading to efficient coupling of E2 to TMH. However, in

CB1 E2dithiol, E2 has reduced flexibility leading to less insertion of E2 C terminal residues

and thus weak coupling of E2/TMH. This work confirmed the importance of the E2 C

terminal for receptor activation and also served as supporting evidence of the original

findings by Ahn et al. (2009).

12

Figure 1.3 Two dimensional representation of the CB1-R (Adapted from Shim et al., 2011).

13

Figure 1.4 Diagramatic representation of the C terminal domain of the CB1-R (Adapted from Stadel et al., 2011)

14

The C terminal is responsible for interaction with G proteins, CRIP 1a, protein

kinases, arrestins and with itself or other receptors to form a dimer (Ahn et al., 2009)

(Figure 1.4). Several characteristic features of the C terminal sequence such as

transmembrane interaction sites, palmitoylation sites, phosphorylation sites and the PDZ

binding domain make it a potential candidate for roles in biogenesis, receptor localization

and activity (Ahn et al., 2009).

The CB1-R carboxy terminus has 73 amino acid residues (R400-L472) and differs

from CB2-R by being 14 residues longer. Despite rough similiarities in length, the sharing

of a few ligands and participation in similar signal pathways, there is no significant

homology between the C termini of these receptors (Bramblett et al., 1995; Xie and

Chen, 2005; Choi et al., 2005). Hydropathy plot analyses conducted by Kyte and Dolittel

(1982) predicted the C terminus of the CB1-R to be less hydrophobic than the equivalent

region of CB2-R, but there remained difficulties in purifying the full length CB1 terminus

because of its high flexibility and relatively unstructured nature. Not until 2009 was this

problem overcome when successful purification of a peptide corresponding to the full

length CB1-R carboxy terminus was achieved. NMR analysis using doubly tagged (15N

and 13C) full length C terminus in dodecylphosphocholine revealed the presence of two

amphipathic α helical domains (Ahn et al., 2009). Importantly, Ahn and associates were

also successful in identifying the specific amino acids of these helical domains (S401-

F412 and A440-M461, respectively) thus suggesting an amphipathic role for each. They

also reported that the hydrophobic face of each helix was intimately associated with the

membrane surface, thus stabilizing the helical domains for binding with other proteins

involved in receptor function, while the polar face of the helices was able to project into

the cytosol (Ahn et al., 2009). Besides its role in receptor function, the C terminus has

been found to be a requirement for receptor exit from endoplasmic reticulum (Tai et al.,

1999; Bermak et al., 2001; Duvernay et al., 2004, Robert et al., 2005).

The cytoplasmic extension of TMH7 is characterized by the presence of a highly

conserved motif within the rhodopsin class A GPCRs, NPXXY and is termed Helix 8

(Patny et al., 2006; Tiburu et al., 20011). This amphipathic helical domain is an integral

part of the intracellular GPCR binding connection to G proteins (Tiburu et al., 2011;

Rosenbaum et al., 2009, Fritze et al., 2003) and is reported to play a crucial role both in

15

ligand recognition and signal transduction (Tiburu et al., 2011). The distal region on the

C-terminus carries another helical domain, Helix 9. However, Helix 9 has only recently

been identified within the CB1-R and very little work has so far been done in defining the

precise structure and role of this helix (Ahn et al., 2010).

TMH7/H8 region has always been an area of interest in probing various structural

determinants involved in activation of the GPCR cycle. Tiburu et al. (2011), using solid

state nuclear magnetic resonance (NMR) and site directed spin labeling-electron

paramagnetic resonance (SDSL/EPR) demonstrated short range electrostatic

interactions between TMH/H8 (with its conserved motif and proline kink) and the

phospolipid bilayer/membrane microenvironment. By employing local helix distortion

studies, they postulated that the conserved but flexible NPXXY motif likely plays a vital

role in ligand binding and signaling events involving TMH7/H8 (Tiburu et al., 2011;

Tiburu et al., 2009; Tyukhtenko, 2009; Hall et al., 2009). Furthermore, experiments

conducted by Tiburu et al. (2011) demonstrated dynamic functional interactions between

TMH7/H8 and the membrane phospholipid environment which modifies the membrane

bilayer structure at discrete loci. These findings hinted at a potential mechanochemical

role of the phospholipid bilayer in mediating CB1/GPCR signal transmission and hence

signal transduction.

Helix 8 is also reported to have a contributory role towards receptor signaling by

interaction with intracellular loops (ICs). This helix interacts with distal part of the C

terminus, offering it rigidity and potential to interact with the third intracellular loop (IC3).

This interaction is essential for proper receptor signaling (Ahn et al., 2010). Moreover,

Swift et al., (2006) identified various noncovalent interactions between helix 8, TMH 7

and IC1, which again were important for receptor signaling. Several studies in various

laboratories indicated a possible role of Helix 8 in receptor biosynthesis, folding and

trafficking (Oksche et al., 1998; Duvernay et al., 2004; Thielen et al., 2005).

Studies by Ahn et al., (2010) helped to understand the importance of this helical

domain (H8) for ligand binding and activation. As mentioned earlier, this helix is

amphipathic with leucines and/or phenylalanines and basic residues common to both

hydrophobic and hydrophilic faces. Ahn and coworkers employed point mutations in

several key residues on both faces. The first mutant involved substitution of the

16

hydrophobic groups, Leu404, Phe408 and Phe412 with alanine to reduce hydrophobicity

while the second involved replacement of the basic residues, Lys402, Arg405 and

Arg409 with glutamine to remove positive charge. The first mutant yielded low Bmax

values (based on saturation binding isotherms), minimal Emax (from [35S]GTPγS binding

studies) and defective localization when compared to the wild type CB1-R. Intriguingly,

the second mutant was virtually identical when compared with the wild type with respect

to the same parameters as the first mutant. Circular dichroism spectroscopy further

revealed that intact hydrophobic residues were indeed vital for maintenance of the helix

while positively charged residues could be easily replaced by less polar or neutral

residues without affecting the structure/function of the helix. These data supported the

importance of the formation of this helical domain for receptor localization and hence

ligand binding and activation. Moreover, various other groups have independently

reported that a defective H8 leads to impaired receptor localization and β arrestin

translocation to the plasma membrane (Suvorova et al., 2009; Yasuda et al., 2009; Ahn

et al., 2010).

Apart from this, the L404F point mutation on H8 domain displayed faster agonist

induced internalization when compared to the wild type, thus underlying its importance in

CB1-R trafficking (Anavi-Goffer et al., 2007).

Helix 9 (H9) consists of charged residues which have been implicated in the

formation of contacts with the cytoplasmic helical extension of TMH5, TMH6 and IC3.

Schertler (2008) suggested the probable role of hydrophobic residues on H9 as a point

of contact with the G protein (Gq).

Besides, extracellular loops, TMH/H8 and H9, distal C terminus ((∆418-472) of

the CB1-R have also been implicated in receptor localization, receptor stability, G protein

binding, desensitization, intracellular sorting during internalization and cellular trafficking

of the receptor (Stadel et al., 2011; Ahn et al., 2009).

Truncation of the CB1 distal C terminal domain led to changes in the magnitude

and kinetics of Ca2+ current inhibition via G protein coupling, suggesting the role of this

domain in cellular signal regulation (Nia and Lewis, 2001). Chillakuri et al. (2007)

demonstrated that after deletion of this domain of the CB1-R expressed in insect cells,

17

Sf9 resulted in a two-fold increase in receptor production and increased basal activity as

compared to wild type. Moreover, residues 418-439 on this domain were found to be

important for receptor desensitization but not internalization, while internalization was

affected when residues within the 460-473 were phosphorylated (Hsieh et al., 1999; Jin

et al., 1999).

Despite the accumulation of much data concerning the role of various domains of

theCB1GPCR, much more research is needed to completely understand the subtle

complexities of this fundamental signaling unit.

1.4.2. The distribution of CB1-Rs in mammalian brain

Understanding the distribution pattern of cannabinoid receptors in brain was

made possible by employing ligand binding studies with the highly specific synthetic

agonist [3H]CP55940 (Herkenham et al., 1991; Herkenham et al., 1990; Mailleux and

Vanderhaeghen, 1992).

The binding of [3H]CP55940 was found to be widely distributed with intensities

dependent on the area of brain concerned, with the general pattern of binding conserved

across mammalian species. The inner most layers of the olfactory bulb, the

hippocampus (in particular the dentate molecular layer and the CA3 region), the lateral

striatum, globus pallidus, entopenduncular nucleus, substantia nigra, pars reticularis and

the cerebellar molecular layer displayed the highest levels of binding of [3H]CP55940,

while moderate levels were found in the cerebral cortex, septum amygdala,

hypothalamus, lateral subnucleus of interpeduncular nucleus, parabrachial nucleus,

nucleus of the solitary tract and the spinal dorsal horn. Low ligand binding was noted

particularly in the thalamus, and various other nuclei in the brain stem and the spinal

ventral horn (Kano et al., 2009; Herkenham, 1990).

The telencephalic and cerebellar regions, where high intensity binding of

[3H]CP55940 occurs, control motor and cognitive function which can explain the

profound effect of cannabinoids on motor and cognitive responses, while lower brain

stem which controls cardiovascular and respiratory functions exhibited low density

binding. This latter distribution is compatible with the fact that high doses of

cannabinoids are not fatal (Herkenham et al., 1991; Mailleux and Vanderhaeghen,

18

1992). Similarly, moderate binding in spinal dorsal horn is consistent with the analgesic

action of intrathecally administered cannabinoid. Moreover, the anti-anorexic and anti-

emetic actions of cannabinoid agonists are likely achieved by moderate binding in the

ventromedial hypothalamic nucleus (located in the hypothalamus and amygdala) which

forms a major part of the satiety center (Kano et al., 2009).

The CB1-R expression pattern has been further investigated by histochemical

analysis. Depending on the region of brain, two distinct forms or labelings of CB1-R

mRNA expression have been reported, specifically, uniform labeling and non-uniform

labeling (Mailleux and Vanderhaeghen, 1992; Matsuda et al., 1993). Uniform labeling

reflecting a large number of cells expressing high level of CB1-R mRNA was detected in

the striatum, thalamus, hypothalamus, cerebellum and lower brain stem. In contrast,

non-uniform labeling resulting from a low number of cells expressing high levels of CB1-

R mRNA was found in the cerebral cortex, amygdala and hippocampus (Kano et al.,

2009). Furthermore, CB1-R expression is invariably greater at inhibitory synapses than

excitatory synapses; however the density of receptors at inhibitory synapse varies

considerably according to brain region. CB1-R density as assessed by immunogold

labeling was 30 times higher on inhibitory synapses than excitatory ones for

hippocampal CA1 pyramidal cells, 6 times for Purkinje cells and nearly 4 times higher for

striatal medium spiny neurons (Kawamura et al., 2006; Uchigashima et al., 2007).

1.5. CB1-R-mediated intracellular signaling pathways

1.5.1. Inhibition of cyclic AMP (cAMP) Activation of the CB1-R following agonist binding triggers a cascade of multiple

signal transduction pathways, a finding supported by [35S]GTPγS assays and studies

examining Pertussis toxin sensitivity of cannabinoid-dependent effects (Pertwee, 1997).

The first characterized pathway for agonist-stimulated CB1-R activation was the

inhibition of adenylyl cyclase and this effect was completely blocked by Pertussis toxin

indicating that it was mediated through Gi/o proteins (Howlett, 1985; Howlett and Fleming,

1984; Howlett et al., 1986). Further evidence towards the role of CB1-R and Gi/o proteins

was supported by work of Derkinderen and co-workers (1996 and 2001) who reported

19

that cannabinoid receptor stimulation led to Tyr-phosphorylation of the focal adhesion

kinase in hippocampal slices and this effect was readily blocked by both SR141716A

and Pertussis toxin.

In their activated states, Gi/o proteins were found to regulate adenylyl cyclase

isoforms 1,3,5,6 or 8 since co-expression of these isoforms resulted in CB1-R-mediated

inhibition of cAMP accumulation (Rhee et al., 1998).

Depending upon the precise signal transduction pathway, different subtypes of

Gi/o were found to be activated by the same ligand, agonist, antagonist or inverse agonist

indicating a CB1-R-biased signaling mechanism (Kenakin, 2007; Houstan and Howlett,

1998; Glass and Northup, 1999; Mukhopadhyay et al., 2002).

In N18TG2 cells, WIN55212-2 displayed agonist behavior for Gi1, Gi2, and Gi3

while desacetyllevonantradol (a CB1-R agonist) was an agonist at Gi1, Gi2 and an inverse

agonist at Gi3. Methanandamide (an agonist and stable analog of anandamide) behaved

as an inverse agonist at Gi1 and Gi2, but it displayed full agonist activity at Gi3.

Interestingly, SR141716A acted as an inverse agonist at all three G protein subtypes

(Mukhopadhyay and Howlett, 2005; Turu and Hunyady, 2010). Furthermore, Houston

and Howlett (1998) suggested the existence of different affinity states for the CB1-R

representing different conformations. This proposal was supported by Georgieva and co-

workers (2008) where, using plasmon-waveguide resonance spectroscopy they found

that the CB1-R assumed distinctly different conformations when bound by CP55940 or

WIN55212-2.

Deadwyler et al. (1995) and Hampson et al. (1995) reported that in hippocampal

cells, agonist-mediated stimulation of the CB1-R decreased intracellular cAMP, lowered

net dephosphorylation of ion channels, activated A-type potassium currents and in turn

led to hyperpolarization of the membrane. The cyclic AMP-dependent protein kinase

(PKA) pathway as regulated by CB1-Rs is intimately involved in synaptic plasticity and

neuronal remodeling (Howlett, 2005).

20

1.5.2. Stimulation of cAMP production

Intriguingly and in contrast to the above studies, CB1-R-mediated increases in

cAMP concentrations have also been observed in response to treatment with

cannabinoids (Howlett, 2005). Maneuf and Brotchie (1997) found an increase in basal

cAMP production in globus pallidus slice preparations treated with cannabinoid agonists

such as CP55940 and WIN55212-2.

Surprisingly, the order of potency of these agonists was similar in mediating this

effect through the CB1-R, and SR141716A was a competitive inhibitor for both inhibitory

and stimulatory components of this mechanism (Bonhaus et al., 1998).

Three mechanisms have been put forward by different research groups to

explain this observation. The first mechanism suggested the possibility of endogenous

synthesis of an adenylyl cyclase activator possibly CB-R-mediated synthesis of

prostaglandins (Burstein et al., 1986, 1994). Here prostaglandins might operate as

cellular stimulators for cannabinoid-mediated cAMP production (Hillard and Bloom,

1983).

The second possibility relates to potential differences in the type of adenylyl

cyclase isoforms expressed by the target cells and the capacity of expressed isoform to

respond to Gi/o-mediated regulation (Howlett, 2005). Agonist stimulation of recombinant

cannabinoid receptors co-expressing adenylyl cyclase isoforms of the 5/6 family or the

1/3/8 family displayed inhibition of adenylyl cyclase as a result of inhibition by Gi (α-

subunit), while those expressing isoforms of the 2/4/7 family led to stimulation of

adenylyl cyclase as a result of the increased Gs response by the Gi (βγ subunit) released

following cannabinoid receptor stimulation (Rhee et al., 1998).

The third mechanism could involve direct interaction between CB1-Rs and Gs. It

was reported that neurons and CHO cells expressing recombinant CB1-Rs pre-treated

with Pertussis toxin, supported cannabinoid agonist-mediated stimulation of cAMP

(Glass and Felder, 1997; Felder et al., 1998; Bonhaus et al., 1998)

21

1.5.3. CB1-Rs and the modulation of Ca2+ fluxes and phospholipases C and A

Sugiura and co-workers (1996, 1997, 1999) using a fura-2 fluorescence assay,

reported cannabinoid and endocannabinoid-mediated increases in intracellular free

calcium in N18TG2 neuroblastoma and NG108-15 neuroblastoma-glioma hybrid cells.

This response was due to a CB1-R and Gi/o interaction since it was blocked by both

SR141716A and Pertussis toxin. Other data also support the ability of cannabinoid

receptors to signal through an inositol 1,4,5-triphosphate (IP3)-Ca2+ mobilization pathway

(Howlett, 2005). Ca2+ mobilization in the N18TG2 neuroblastoma cell line could be

blocked by a phospholipase C (PLC) inhibitor, suggesting that PLCβ could be the

effector (Sugiura et al., 1996, 1997). Netzeband and co-workers (1999) found that

activation of CB1-Rs caused an increase in Ca2+ levels in response to depolarization via

glutamate receptor activation or high K+ which primarily resulted in mobilization of Ca2+

from caffeine-sensitive and IP3 receptor-sensitive stores. This signal was reduced by

SR141716A, Pertussis toxin and a PLC inhibitor, again suggesting that this Ca2+

mobilization originated from a CB1-R dependent PLC mechanism (Netzeband et al.,

1999).

Cannabinoid receptors have also been implicated in the regulation of

phospholipase A2 activity. Cannabinoid pre-treatment of several cultured cell types has

been shown to release arachidonic acid, and this effect is believed to be mediated by

phospholipase (Burstein 1991; Burstein et al., 1994; Shivachar et al., 1996).

1.5.4. CB1-R-dependent regulation of ion channels

Agonist-mediated stimulation of CB1-Rs results in activation of G protein-coupled

inwardly-rectifying potassium channels (GIRKs) (Henry and Chavkin 1995; Mackie et al.,

1995) and inhibition of L-type (Gebremedhin et al., 1999), N-type (Mackie and Hill, 1992)

and P/Q -type (Mackie et al., 1995) calcium channels. When treated with anandamide or

WIN55212-2, AtT-20 pituitary tumor cells expressing CB1-Rs, were found to activate

inwardly rectifying potassium currents (Kir) and this response was Pertussis toxin-

sensitive, indicating the involvement of Gi/o proteins (Mackie et al., 1995; Henry and

Chavkin, 1995; McAllister et al., 1999). McAllister and co-workers (1999) also

demonstrated the activation of GIRKs in the Xenopus oocyte system.

22

The agonist stimulation of CB1-Rs expressed in N18 neuroblastoma and NG108-

15 neuroblastoma-glioma hybrid cells resulted in inhibition of N-type voltage gated Ca2+

channels (Mackie and Hill, 1992; Mackie et al., 1993; Howlett, 2005). Fura-2

fluorescence studies also confirmed that anandamide and 2-AG inhibited the

depolarization-induced intracellular free Ca2+ increase in NG108-15 cells (Sugiura et al.,

1997). When treated with WIN55212-2 and anandamide, AtT-20 pituitary cells

expressing recombinant CB1-Rs, were found to inhibit Q-type Ca2+ channels (Mackie et

al., 1995). Pertussis toxin sensitivity suggested the mediation of this response through

Gi/o proteins. Hampson and coworkers (1998), using the fura 2 assay, showed that P/Q-

type Ca2+ fluxes are also inhibited by anandamide and this inhibition was suppressed by

SR141716A and Pertussis toxin, again suggesting the involvement of CB1-R and Gi/o

proteins. Similar results were reported for inhibition of L-type Ca2+ currents in cat brain

arterial smooth muscle cells expressing CB1-Rs by WIN55212-2 and anandamide and

these effects were inhibited by SR141716A and Pertussis toxin.

1.5.5. Involvement of CB1-Rs in the suppression of neurotransmitter release

Gill et al. (1970) first reported the CB1-R-mediated inhibition of transmitter

release in guinea pig ileum by investigating electrically-evoked twitch response

recordings. Since then, numerous studies using electrophysiological and biochemical

techniques have been carried which demonstrate CB1-R-mediated suppression of

transmitter release (Schlicker and Kathmann, 2001).

CB1-Rs control the release of several neurotransmitters in mammalian brain

including glutamate (Levenes et al., 1998), GABA (Szabo et al., 1998), glycine (Jennings

et al., 2001), acetylcholine (Gifford et al., 1996), norephinephrine (Ishac et al., 1996),

dopamine (Cadogan et al., 1997), serotonin (Nakazi et al., 2000) and cholecystokinin

(CCK) (Beinfeld and Connolly, 2001).

Shen and co-workers (1996) using cultured hippocampal neurons were the first

to gain evidence for the suppression of glutamate release by a cannabinoid agonist.

They found that WIN55212-2 decreased excitatory postsynaptic currents (EPSCs) thus

indicating a reduction in neurotransmitter release. Importantly, this suppression of

EPSCs was reversed by SR141716A demonstrating involvement of CB1-Rs. Schlicker

23

and Kathamann (2001) detected a similar kind of CB1-R-mediated decrease in

neurotransmitter release in the cerebellum, the cortex and the striatum. Wang (2003)

using rat brain synaptosomes demonstrated a CB1-R agonist-mediated decrease in 4-

AP-evoked L-glutamate release in response to low micromolar concentrations of

WIN55212-2. This decrease was concentration-dependent and was reversed by the

CB1-R antagonist AM281. Similar results were obtained by Godino et al. (2007), where

they showed a moderate (approx. 30%) decrease in KCl-evoked L-glutamate release by

low micromolar concentrations of WIN55212-2, this effect again being reversed by low

micromolar concentrations of AM281.

Interestingly, similar effects were noted on GABA release by CB1-R agonists in

the striatum, substantia nigra, pars reticularis, hippocampus, nucleus accumbens and

cerebellum (Kano et al., 2009). In these neurons, GABAergic inhibitory postsynaptic

currents (IPSCs) were effectively suppressed by WIN55212-2, but the postsynaptic

response to muscimol (a GABAA receptor agonist) was unaffected, clearly pointing to a

presynaptic locus of action. This effect was also blocked by SR141716A, strongly

supporting the role of CB1-Rs (Chan and Yung, 1998).

Several possible mechanisms have been proposed by various research groups

regarding the molecular targets involved in CB1-R-mediated suppression of transmitter

release (Schlicker and Kathmann, 2001).

The first mechanism advanced involved the possible involvement of voltage-

dependent Ca2+ channels. It was found that cadmium, a non-selective blocker of Ca2+

channnels, readily reduced or blocked the WIN55212-2-mediated inhibitory effects on

spontaneous EPSCs in rat striatal slices and spontaneous GABAergic IPSCs in rat

hippocampal slices (Huang et al., 2001; Hoffman and Lupica, 2000).

Besides regulation via voltage-gated Ca2+ channels, evidence also exists for

involvement of sites downstream of the Ca2+ channels. Takahashi (2000) conducted

electrophysiological experiments and reported that Ca2+-independent miniature EPSCs

and miniature IPSCs were reduced by a cannabinoid receptor agonist indicating a direct

influence on release machinery. Another possible mechanism is through involvement of

K+ channels, including GIRKs and KA channels (Pertwee, 1997; Ameri, 1999), since the

24

inhibitory effects of WIN55212-2 on evoked EPSCs in the mouse nucleus accumbens

were reduced by BaCl2 (a KA channel blocker) or 4-aminopyridine (4-AP) (a GIRK

channel blocker) and completely blocked when BaCl2 and 4AP were co-administered

(Robbe et al., 2001).

However, the complexity of the target systems impacted by cannabinoids

increases further when sodium channel involvement is considered. Nicholson et al.

(2003), Liao et al. (2004) and Duan et al. (2008) in our laboratory showed that

endocannabinoid anandamide and synthetic cannabinoid agonists WIN55212-2,

CP55940 and AM404 along with the cannabinoid antagonists AM251 inhibit sodium

channels at low micromolar concentrations. Hence more studies are needed to fully

elucidate and understand this complex phenomenon of suppression of neurotransmitter

release by cannabinoid and cannabimimetic drugs.

1.6. Homodimerization and heterodimerization of CB1-Rs

Another intriguing aspect of the CB1-R is its capacity to undergo dimerization in

the presence of other receptors (Turu and Hunyady, 2010). Wager-Miller et al. (2002)

were first to show homodimerization of the CB1-R where they showed a high molecular

weight band of a dimerized CB1-R using Western blotting. Resonance energy transfer

studies have also revealed the heterodimerization of CB1-Rs with D2 and A2A receptors

(Carriba et al., 2008) and also with opiate receptors (Rios et al., 2006) in transfected cell

lines. Kearn et al. (2005) reported an interesting interaction between heterodimerized

CB1 and D2 receptors. They noted that independent activation of both receptors resulted

in decreased forskolin-induced adenylyl cyclase activity. However, when co-expressed,

activation of the CB1-R reversed the inhibition caused by D2 stimulation and in turn led

to elevated cAMP levels (Kearn et al., 2005).

Furthermore, there are also reports of coupled trafficking of CB1-Rs to and from

the plasma membrane as a consequence of dimerization (Ellis et al., 2006). Orexin-1

receptor distribution in cells was found to be dramatically altered (from plasma

membrane to intracellular vesicles) following the coexpression of CB1-Rs in same

system (Ellis et al., 2006)

25

To summarize, dimerization is an emerging and interesting area for further

exploration of CB1-R activity in relation to regulation of endocannabinoid system.

1.7. Constitutive activity of CB1-Rs

Like several other GPCRs, CB1-Rs have been found to exhibit constitutive

activity wherein they show high basal activity in absence of agonist binding, both in

expression systems and in native tissues (Gifford and Ashby, 1996; Turu and Hunyady,

2010; Bouaboula et al., 1997; Rinaldi-Carmona, 1998; Pertwee, 2005). Moreover, the

CB1-R inverse agonist, SR141716A has been found to act on certain other receptors

(Savinainen et al., 2003; Lauckner et al., 2008). This further complicates the

interpretation of data. Hence the introduction of a neutral antagonist can help study the

constitutive property of CB1-Rs (Turu and Hunyady, 2010). Neutral antagonists can be

defined as agents or compounds that do not change the basal activity of CB1-Rs (Turu

and Hunyady, 2010). This definition remains valid if there is an assumption of the

existence of CB1 constitutive activity however, if this activity was caused due to

endocannabinoids innately present in tissues, then these neutral antagonists can be

referred as partial agonists. Hence, more studies are warranted in this area of

cannabinoid research to fully elucidate the constitutive versus basal activity of the CB1-

Rs (Turu and Hunyady, 2010).

1.8. The biochemistry of endocannabinoids

The first endocannabinoid was isolated from porcine brain and was named

anandamide (arachidonyl ethanolamide, from the Sanskrit word ‘anand’ means bliss or

happiness) (Figure 1-5) (Devane et al., 1992).

Sugiura and co-workers (2002) found that anandamide was a partial agonist at

cannabinoid receptors and a full agonist at transient receptor potential vanilloid receptors

1 (TRPV1) (Kano et al., 2009). Another breakthrough was the discovery of another major

endocannabinoid, 2-arachidonyl glycerol (2-AG) which was isolated from canine gut and

rat brain (Mechoulam, 1995; Sugiura et al., 1995). 2-AG is more abundant in the brain

26

and acts as a full agonist at cannabinoid receptors and hence can be termed as true

natural endocannabinoid (Sugiura et al., 2006).

Other endocannabinoids so far identified are dihomo-γ-linolenoyl ethanolamide

(Hanus et al., 2001), decosatetraenoyl ethanolamide (Hanus et al., 2001), 2-AG ether

(noladin ether) (Hanus et al., 2001), O-arachidonylethanolamine (virodhamine) (Porter et

al., 2002) and N-arachidonoyldopamine (Huang et al., 2002). See Figure 1-5 for

structures.

Dihomo-γ-linolenoyl ethanolamide and decosatetraenoyl ethanolamide belong to

a broad family of N-acylethanolamides like anandamide. They are present in mammalian

brain and bind to CB1-Rs with medium affinity (Felder et al., 1993; Hanus et al., 1993)

and with lower affinity to CB2-Rs (Felder et al., 1995) respectively.

Noladin ether was isolated from pig brain and was found to bind CB1-Rs with

medium affinity, although this compound has lower affinity for CB2-Rs (Hanus et al.,

2001). Virodhamine (from Sanskrit word, ‘Virodh’ meaning opposition or antagonism)

was isolated from rat brain (Porter et al. (2002) and acts as an antagonist or a partial

agonist at the CB1-R but as a full agonist at CB2-Rs. N-Arachidonoyldopamine, first

identified in bovine and rat nervous tissue, displays agonist actions on both the

endocannabinoid and endovanilloid system.

27

Anandamide 2-Arachidonoyl glycerol ether 2-Arachidonoyl glycerol

Virodhamine Dihomo-γ-linolenoyl ethanolamide

Docosatetraenoyl ethanolamide N-Arachidonoyl dopamine

Figure 1.5 Structures of prominent endocannbinoids (All structures redrawn using ChemDraw Ultra 11.0 from Kano et al., 2009).

28

Among all of these endocannabinoids, anandamide and 2-AG have been widely

accepted as the major endocannabinoids acting on CB1-Rs. Hence, although several

reviews are available (Basavarajappa, 2007; Bisogno et al., 2005; Okamoto et al., 2007;

Sugiura et al., 2006; Vandevoorde and Lambert, 2007; Kano et al., 2009) that explore

the metabolic pathways for formation and degradation of a whole range of potential

endocannabinoids, only anandamide and 2-AG will be discussed further.

1.8.1. Anandamide biosynthesis

Di Marzo and co-workers (1994) were the first to show the Ca2+ ionophore

ionomycin or high K+ depolarization causes anandamide production in rat striatal or

cortical neurons. The extracellular Ca2+ chelator EGTA effectively abolished anandamide

production indicating a role of Ca2+ entry in promoting this mechanism. Interestingly,

Stella and Piomelli (2001) noted that carbachol-evoked generation of anandamide was

not blocked by EGTA (an extracellular Ca2+ chelator) but was effectively blocked by

BAPTA-AM (a membrane permeable Ca2+ chelator). Besides these biochemical

experiments, electrical stimulation has also been shown to raise anandamide levels in

rat hypothalamic slices as measured by mass spectrometric analysis (Di et al., 2005).

The classical ‘transacylation-phosphodiesterase pathway’ has been suggested

for biosynthesis of anandamide (Figure 1.6).

29

Figure 1.6 Transacylation-phosphodiesterase pathway for biosynthesis of anandamide (Adapted from Cadas et al., 1997).

30

This is a two step enzyme-catalyzed reaction where the initial step involves

transfer of an arachidonate group from the sn-1 position of the phospholipids to the

primary group of phosphatidylethanolamine (PE) in the presence of the enzyme N-

acyltransferase (NAT), yielding N-arachidonyl PE or NAPE (Figure 1.6). The second

step involves the catalytic hydrolysis of NAPE to anandamide and phosphatidic acid in

the presence of N-acylphosphatidylethanolamine-hydrolysing phospholipase D (NAPE-

PLD) enzyme. The rate limiting step in anandamide synthesis is the NAT activity which

is regulated by Ca2+ (Kano et al., 2009). The anandamide formed is then presumed to

diffuse from the membrane into the surrounding medium (Kreitzer and Regehr, 2002).

1.8.2. 2-Arachidonoyl glycerol (2-AG) biosynthesis

Unlike anandamide, 2-AG displays full agonism at the CB1-R. Electrical

stimulation (Stella et al., 1997) in hippocampus as well as ionomycin treatment in

N18TG2 neuroblastoma cells (Bisogno et al., 1997) leads to increased levels of 2-AG.

Several biochemical pathways have been suggested for biosynthesis of 2-AG

(Figure 1-7), however the more important pathway involves phospholipase C (PLC) and

diacylglycerol lipase (DAGL) (Kano et al., 2009). The first step involves formation of

diacylglycerol (DAG) (which has an arachidonic acid moiety) by enzymatic hydrolysis of

arachidonic acid containing membrane phospholipids (like phosphatidylinositol) by PLC.

In the second step, DAGL acts on DAG to generate 2-AG (Kano et al., 2009). Other

biosynthetic pathways of possible relevance include those reactions leading to the

production of lysophospholipid from membrane phospholipid initially mediated by

phospholipase A1 (PLA1) and final formation of 2-AG by the action of Lyso-PLC on PLA1

(Sugiura et al., 1995; Tsutsumi et al., 1994; Ueda et al., 1993). Nakane et al. (2002)

suggested that 2-AG could also be formed by the action of a phosphatase on 2-

arachidonoyl lysophosphatidic acid (2-arachidonoyl LPA), while the generation of 2-

arachidonoyl phosphatidic acid from 1-acyl-2-arachidonoylglycerol has been suggested

by Bisogno et al. (1999) and Carrier et al. (2004) (Figure 1.7).

31

Figure 1.7 Metabolic pathways for biosynthesis of 2-AG (Adapted from Kano et al., 2009).

32

1.9. Degradation pathways for endocannabinoids

After endocannbinoids are released they can be degraded by two main

mechanisms i.e. hydrolysis and oxidation (Vandevoorde and Lambert, 2007). The

hydrolytic pathway involves the breakdown of anandamide by the action of fatty acid

amide hydrolase (FAAH) and 2-AG by monoacylglycerol lipase (MAGL) while the

oxidative pathway includes oxidation of the arachidonic acid moiety of endocannabinoids

by cyclooxygenase (COX) and lipoxygenase (LOX).

FAAH was first identified, purified and cloned from rat liver by Cravatt et al.

(1996). Formerly named as ‘anandamide amidohydrolase’, FAAH was identified in the

brain and many other organs. FAAH was reported to sensitive to the serine protease

inhibitor phenylmethylsulfonyl fluoride (PMSF), and was also found to act on other fatty

acid amides but with anandamide as the preferred substrate (Cravatt et al., 1996;

McFarland and Baker, 2004). In vitro studies have revealed that FAAH has the ability to

hydrolyse the ester linkage of 2-AG; however, this activity was minimal in vivo (Cravatt et

al, 1996). However, the importance of FAAH in anandamide breakdown was underlined

further when FAAH knockout mice were reported to be more responsive towards

exogenously administered anandamide (Cravett et al, 2001).

Tornqvist and Belfrage (1976) were the first to describe MAGL but this enzyme

was not cloned until 1997 by Karlsson and co-workers (1997) who cloned it from a

mouse adipocyte cDNA library. In vivo, MAGL is found to be the main enzyme that

catalyzes the hydrolysis of 2-AG (Dinh et al., 2002; Dinh et al., 2004; Vandevoorde and

Lambert, 2007). MAGL has 303 amino acids and is present in various organs including

brain. This enzyme accounts for about 85% of 2-AG degradation, while the remainder is

attributed to two less characterized enzymes namely, ABHD6 and ABHD12 (Blankman

et al., 2007).

Among the three known forms of COX enzymes in mammalian tissues, COX-2

accepts anandamide as a substrate generating prostaglandin-ethanolamides

(Vandevoorde and Lambert, 2007). Anandamide and 2-AG also apparently serve as a

33

substrates for LOX, however much less work has been done in this area (Chen et al.,

1994).

1.10. Transport of endocannabinoids

A two step process mediates the extracellular removal of endocannabinoids, the

first being transport into the cells followed by enzymatic hydrolysis (McFarland and

Baker, 2004; Kano et al., 2009).

Anandamide (AEA) transport has been widely studied and compared to 2 AG.

AEA uptake has been reported in cortical neurons (Fegley et al., 2004), striatal neurons,

astrocytes (Di Marzo et al., 1994; Beltramo et al., 1997) and cerebellar granule cells

(Hillard et al., 1997). Various aspects of AEA transport have been noted by research

groups. It has been reported that AEA transport is (i) temperature sensitive (ii) inhibited

by certain fatty acid amide derivatives (iii) relatively fast (t1/2 approx. 2.5 minutes) (iv)

controlled by other signal transduction pathways and (v) saturable at 37oC (Di Marzo et

al., 1994; Beltramo et al., 1997; Hillard et al., 1997; Maccarrone et al., 1998; Maccarrone

et al., 2000; Rakshan et al., 2000).

Several mechanisms have been advanced to explain the cellular uptake of AEA

including a protein carrier-mediated process, facilitated diffusion regulated by FAAH,

AEA sequestration by cellular machinery and endocytotic uptake of AEA (McFarland and

Baker, 2004).

Accumulation of a substrate on the cis side of the membrane leads to carrier

protein accumulation on the trans side of the membrane thus causing movement of an

extracellular substrate into intracellular compartment against a concentration gradient, a

phenomenon termed as ‘trans flux coupling’ (McFarland and Baker, 2004). ‘Trans flux

coupling for AEA was observed in cerebellar granule neurons (Hillard and Jarrahian,

2000). The intracellular presence (or expression) of FAAH positively modulated the

cellular uptake of AEA (McFarland and Baker, 2004). Neuroblastoma or glioma cells

expressing FAAH when treated with the FAAH inhibitor,

methylarachidonylflurophosphonate (MAFP) display close to a 50% reduction in AEA

accumulation (Deutsch et al., 2001). Furthermore, Day et al., (2001) showed a 2-fold

34

increase in the uptake of AEA following FAAH transfection of HeLa cells (which are

otherwise devoid of FAAH activity). Glaser et al. (2003) suggested FAAH-dependent

facilitated diffusion as a mechanism for cellular uptake of AEA. Kinetic analysis by

Glaser et al. (2003) demonstrated that AM404 (an AEA transport inhibitor and FAAH

inhibitor, Hillard and Jarrahian, 2000), when added at the 5 min time point, significantly

decreased AEA cellular accumulation, while no such effect was reported at earlier time

points (25 sec and 45 sec). This observation was consistent in both neuroblatoma cells

(with FAAH activity) and astrocytoma cells (without FAAH activity) (Glaser et al. 2003).

Hence, these researchers concluded that AEA accumulation was saturable at 5 min time

and this coincided with the inhibition of downstream components of AEA uptake

especially FAAH (Glaser et al., 2003).

Besides the above mentioned mechanisms, Hillard and Jarrahian (2003)

hypothesized that AEA could be taken up through sequestration by cellular components.

They found that radiolabeled AEA reached intracellular concentrations that were higher

than those in the extracellular media (Hillard and Jarrahian, 2003). They suggested the

existence of two distinct intracellular pools of AEA, one was free AEA and the other was

AEA sequestered by a cellular component. This sequestration of AEA is saturable and

may involve certain membranous compartments that serve as reservoir for this lipophilic

molecule (Hillard and Jarrahian, 2003). Since the AEA that is sequestered or bound is

not available for free diffusion across plasma membrane, it generates a positive inward

concentration gradient of AEA that is maintained by FAAH (Hillard and Jarrahian, 2003;

McFarland and Lambert, 2004).

The final model or mechanism proposed for AEA degradation is through the

caveolae-related endocytotic process (McFarland et al., 2003; McFarland et al., 2004).

McFarland (2004) showed that by inhibiting the caveolae-related endocytosis process

(by treating cells with N-ethylmaleimide and tyrosine kinase inhibitor, genistein), there

was marked reduction of cellular accumulation of AEA (McFarland et al., 2004)

As mentioned earlier, few studies have been directed towards understanding the

cellular uptake and degradation of 2-AG (Kano et al., 2009). However, some reviews

suggest the existence of a very similar mechanism for 2-AG transport as for AEA

(Beltramo and Piomelli, 2000; Bisogno et al., 2001; Piomelli et al., 1999).

35

1.11. Endocannabinoid-mediated short term depression (DSI and DSE)

In 2001, four research groups demonstrated independently that

endocannabinoids mediate synaptic plasticity through retrograde signaling in the CNS

and play a vital role towards short term and long term synaptic tonicity (Wilson et al.,

2001; Kreitzer et al., 2001; Ohno-Shosaku et al., 2001; Maejima et al., 2001).

Endocannabinoids are biosynthesized de novo and then released into the synaptic cleft

either tonically under basal conditions or in an activity-dependent manner (Kano et al.,

2009, Katona and Freund, 2008). The endocannbinoids that are released then travel

retrogradely (unlike usual anterograde transmission) and activate presynaptically-located

CB1-Rs (Katona and Freund, 2008), thus causing suppression of transmitter release

either transiently (endocannbinoid mediated short term depression, eCB-STD) or over

extended duration (eCB-long term depression, eCB-LTD) (Kano et al., 2009).

1.12. Endocannabinoids as synaptic circuit breakers and retrograde messengers

Pitler and Alger (1994) reported that following a train of action potentials from the

post synaptic cell, the spontaneous inhibitory postsynaptic potentials (IPSPs) from CA1

pyramidal cells of hippocampal slices were suppressed, thus indicating a reduction of

GABA release from presynaptic nerve endings: a novel phenomenon termed

‘depolarization-induced suppression of inhibition’ (DSI).

It was found that depolarization of the post synaptic neuron triggers Ca2+ entry

through voltage-dependent Ca2+ channels which then elevates Ca2+ levels in post

synaptic neurons. Support for this mechanism was clear since cerebellar DSI is

inhibited by chelation of extracellular Ca2+ or by adding Cd2+ to the bath solution (Llano

et al., 1991). Moreover, DSI was enhanced by the L-type Ca2+ channel activator, BAY K

8644 while blocked by calcium chelators like BAPTA and EGTA (Pitler and Alger, 1992;

Vincent and Marty, 1993). An ultimate presynaptic locus for DSI expression was

36

confirmed by Llano et al. (1991) when they found that DSI correlated directly with the

decrease in frequency of IPSCs (IPSCs).

Almost the same time, Ohno-Shosaku et al. (2001) using rat hippocampal

cultures and Wilson and Nicoll (2001) using rat hippocampal slices, reported that DSI

was completely abolished by the CB1 receptor antagonists SR141716A, AM251 and

AM281, while a metabotropic glutamate receptor antagonist failed to do this, indicating

that the retrograde signaling mechanism was mediated through CB1-Rs.

Meanwhile, in cerebellar Purkinje cells, Kreitzer and Regehr (2001) reported a

phenomenon similar to DSI where they observed that following postsynaptic

depolarization, the excitatory transmission was transiently suppressed, an effect termed

as ‘depolarization-induced suppression of excitation’ or DSE. The presynaptic Ca2+

currents generated in response to stimulation of the excitatory climbing fibers were found

to be inhibited or suppressed during DSE, thus providing good evidence for a

presynaptic locus of this effect (Kreitzer and Regehr, 2001). Moreover, DSE was

abolished by treatment with the calcium ion chelator BAPTA and occluded by the CB1-R

antagonist, AM251. This effect was not blocked by mGlu, adenosine and GABA

receptor antagonists suggesting that like DSI, DSE is also mediated through the

endocannabinoid system (Kreitzer and Regehr, 2001).

37

Figure 1.8 Blockade of depolarization-induced suppression of inhibition (DSI) CB1-R antagonists by AM251 and SR141716A in rat hippocampal neurons (Kano et al., 2009).

A: Examples of inhibitory postsynaptic currents (IPSCs) (right panel) and

control results showing that DSI can be evoked repetitively without change in its

magnitude (left panel)

B: Example of IPSCs, control and after treatment with AM281 (left panel).

Average time courses for DSI before and after treatment with AM281 (right panel)

C: Example of IPSCs, control and after treatment with SR141716a (left

panel). Average time courses for DSI before and after treatment with SR141716A (right

panel)

38

1.13. Mechanisms of endocannabinoid mediated short term depression (eCB-STD)

Unlike classical neurotransmitters, endocannabinoids are de novo synthesized

on demand and released and not stored in vesicles. Precisely what initiates the

endocannabinoid production and hence eCB-STD in neurons has been the subject of

much scientific debate (Kano et al., 2009). Two different pathways have been put forth

i.e. a PLCβ-independent pathway triggered by large rise in intracellular Ca2+

concentration alone (CaER), while the other is a PLCβ-dependent pathway that is

activated by stimulation of basal (receptor-driven) endocannabinoid release, (basal

RER) or elevated calcium-driven endocannabinoid release (Ca2+-assisted RER).

1.13.1. CaER

In the hippocampus, influx of Ca2+ ions into the postsynaptic cell triggers DSI

thus indicating vital role of Ca2+ elevation in initiating DSI and hence eCB-STD (Wilson

and Nicoll, 2001). The main sources for postsynaptic Ca2+ elevation are through voltage

gated Ca2+ channels (Ohno-shosaku et al., 2007; Pitler and Alger, 1992) or through

release from an intracellular Ca2+ reservoir (Isokawa and Alger, 2006).

According to this model, micromolar concentrations of Ca2+ trigger the activation

of voltage-dependent calcium channels which then produce 2-AG, likely through a

DAGL-mediated pathway and a PLCβ-independent pathway (Kano et al., 2009). 2-AG

released then initiates DSI/DSE by activating CB1-Rs.

1.13.2. Basal RER

The G-protein coupled receptors (Gq/11) mGlu1/5, M1/M3 muscarinic receptors,

glucocorticoid receptors, oxytocin receptors and orexin receptors were found to induce

eCB-STD when strongly activated (Hashimotodani et al., 2007; Hashimotodani et al.,

2007; Maejima et al., 2005). The most likely pathway for induction of eCB-STD by this

pathway is thought to involve stimulation of postsynaptic G-protein-coupled receptors

39

and DAG generation (through a PLCβ-dependent pathway), which then generates 2-AG

by the action of DAGL. The 2-AG then mediates induction of DSI or DSE (Kano et al.,

2009).

1.13.3. Ca2+-assisted RER

This model proposed to understand eCB-STD is a combination of the above two

models except here a weak stimulation of postsynaptic Gq/11 receptors cause a small rise

in Ca2+ to submicromolar concentrations which in turn activates PLCβ and thus 2-AG

and DSI/DSE (Kano et al., 2009; Hashimotodani et al., 2007).

1.14. Termination of eCB-STD

The 2-AG biosynthesized by the above mechanisms may then be partially

degraded by the COX-2 enzyme which is located in the postsynaptic cell and the

remaining 2-AG then diffuses rapidly into extracellular space by lateral diffusion

ultimately binding and activating presynaptic CB1-Rs. Presynaptically-located MAGL can

then degrade the remaining 2-AG thus terminating eCB-STD signalling (Kano et al.,

2009).

40

Figure 1.9 The pathway involved in the termination of endocannabinoid-mediated short term depression (eCB-STD) (Adapted from Kano et al., 2009).

41

1.15. Endocannabinoid-mediated long term depression (eCB-LTD)

As the name indicates, eCB-LTD induces a prolonged suppression of

neurotransmitter release in the brain and this effect is mediated through presynaptically

located CB1-Rs. However, the precise location of the initiation of eCB-LTD in specific

regions of brain also plays an important role as a deciding factor for involvement of CB1-

Rs and/or presynaptic components (Kano et al., 2009). It was found that application of

the CB1-R agonist (CP55940) at both excitatory synapses in the nucleus accumbens

(Robbe et al., 2002) and inhibitory synapses in the hippocampus (Chevaleyre et al.,

2007) induces LTD, indicating that CB1-Rs alone are sufficient to initiate the LTD

response in these regions of brain. However, at excitatory synapses in the dorsal

striatum, LTD cannot be evoked by CB1-R activation alone (Ronesi et al., 2004) since it

requires simultaneous activation of CB1-Rs and low frequency presynaptic activity

(Singla et al., 2007).

Studies have also been directed towards seeking an understanding of the way in

which a relatively short activation of CB1-Rs triggers LTD. It was reported that

presynaptic inhibition of the cAMP/PKA cascade and P/Q type voltage-gated calcium

channels was required for induction of LTD in the nucleus accumbens (Mato et al.,

2008), while presynaptic cAMP/PKA signalling and RIM1α were needed for expression

of LTD in the hippocampus and amygdala (Chavaleyre et al., 2007).

1.16. Other important aspects of endocannabinoid signaling

1.16.1. Regulation of excitability

Endocannabinoids have been found to control neuronal firing in several brain

regions (Kano et al., 2009). Kreitzer et al. (2002) showed in that in rat cellebellar cells,

interneuronal excitability was reduced in a CB1-R-dependent manner when Purkinje cells

were depolarized. They suggested that depolarization triggers generation and release of

endocannabinoids that then bind to CB1-Rs causing opening of K+ channels and hence

hyperpolarization (Kreitzer et al., 2002; Kano et al., 2009).

42

1.16.2. Basal activity of endocannabinoid signaling

CB1-Rs expressed exogenously using recombinant expression systems have

been found to support constitutive activity, but the case for such activity in native

membranes is weak (Howlett, 2004). SR141716A has shown to behave as an inverse

agonist in many native brain membrane preparations by inhibiting basal G protein

binding. This effect of SR141716A however, was at micromolar concentrations, in

contrast to the competitive antagonist behaviour in electrophysiological systems which

can occur at nanomolar concentrations (Sim-Selley et al., 2001). This again opens the

door to the debate as to whether CB1-Rs are constitutively active or this inverse agonist

activity of SR141716A (in the micromolar range) can be interpreted as its action on the

constitutively active adenosine receptor (Savinainen et al., 2003).

Moreover, CB1-Rs have been found to display basal activity and this basal effect

can be attributed to the tonically-released endocannabinoids (Kano et al., 2009;

Hoffman, 2003).

1.16.3. Plasticity of endocannabinoid signaling

Endocannabinoid system is tonically regulated and can be up or down regulated

in different situations (Kano et al., 2009). Chen et al., (2007) showed that DSI can be

potentiated by tetanic stimulation in hippocampal slice preparations and this was

mediated through CB1-Rs. Similarly, chronic exposure of nucleus accumbens to ∆9-THC

or WIN55212-2 reduced the sensitivity of CB1-R and abolished eCB-LTD (Hoffman et al.,

2003).

1.17. Subcellular distribution of various signaling molecules involved in regulation of the endocannabinoid system

1.17.1. Gq Protein α subunit

So far four isoforms of Gq protein α-subunits have been identified by

immunochemical studies. These include Gαq, Gα11, Gα14 and Gα15/16. Among these,

Gαq and Gα11 are the main isoforms found mostly in brain (Tanaka et al., 2000) and they

43

also represnt the major ones that are attached to the extrasynaptic membrane

containing mGluR1 in Purkinje cells and mGluR5 in hippocampal pyramidal cells

(Tanaka et al., 2000).

1.17.2. Phospholipase Cβ (PLCβ)

All four isoforms (1-4) of PLCβ are abundantly expressed in the brain largely in a

nonoverlapping expression manner (Kano et al., 2009). PLCβ1 is mainly expressed in

the telencephalon, PLCβ2 in white matter, PLCβ3 in the caudal cerebellum, PLCβ4 in

the raustral cerebellum, thalamus and brain stem (Ross et al., 1989; Roustan, 1995;

Tanaka and Kondo, 1994; Watanabe et al., 1998).

1.17.3. Diacylglycerol lipase (DAGL)

DAGLα is widely distributed in various regions of brain with its highest density

found in cerebellar Purkinje cells, pyramidal cells in hippocampus and medium spiny

neurons in the striatum (Kano et al., 2009). However the distribution of DAGLα varies

from region to region. For example, DAGLα is at the highest amount in the spine neck

(as compared to spine head) in Purkinje cells (Yoshida et al., 2006), while in

somatodendritic membranes, DAGLα levels were highest in spine followed by the

dendritic shaft with lower amounts in the soma (Katona et al., 2006; Yoshida et al.,

2006).

1.17.4. N-acyl-phosphatidylethanolamine-hydrolyzing phospholipase D (NAPE-PLD)

NAPE-PLD is reported to be widely expressed in presynatic regions with the

highest levels in granule cells of the dentate gyrus and medium to low levels in CA3

pyramidal cells of the hippocampus, olfactory bulb, piriform cortex and thalamic nuclei

(Kano et al., 2009). Interestingly, the presence of NAPE-PLD in the axonal terminal may

indicate a presynaptic locus for anandamide synthesis which in turn may be functioning

as an anterograde messenger (Kano et al., 2009).

44

1.17.5. Monoacylglycerol lipase (MAGL)

Dinh et al. (2002) described the MAGL mRNA expression patterns in various

brain regions. They found the presence of this enzyme in the synapse-rich neuropil

region of the hippocampus, the amygdala and the cerebral cortex, where this enzymes is

mostly distributed as a seat for axon terminals (Dinh et al., 2002). Interestingly, MAGL

activity in brain was also found to be responsible for determining basal endocannabinoid

tonicity and retrograde signaling since presynaptically-located MAGL was reported to

breakdown the 2-AG released from the postsynaptic cell, thus restricting the

accumulation of 2-AG in and around the synaptic cleft (Hashimotodani et al., 2007).

1.17.6. Fatty acid amide hydrolase (FAAH)

FAAH presence is complementary to the expression of MAGL and CB1-Rs, for

example, FAAH is absent in the globus pallidus where CB1-Rs are abundantly expressed

while it is present at high density in somatodendritic elements of principal neurons but

mostly absent in interneurons (where CB1 and MAGL are widely expressed) (Kano et al.,

2009).

1.18. Physiological roles of the endocannabinoid system

Behavioural studies have helped enormously in exploring the physiological role

of the endocannabinoid system in a variety of brain functions like learning and memory,

depression, addiction, appetite and feeding behaviour, pain, as well as neuroprotection.

1.18.1. Learning and Memory

It has been now well established that phytocannabinoids and synthetic

cannabinoids cause memory and learning impairment in humans and laboratory animals

(Davies et al., 2002; Kano et al., 2002). In laboratory animals, Morris water maze tests

have been employed to investigate the effects of cannabinoid agonists. Mice (and rats)

failed to perform well in this test following systemic administration of CB1-R agonists

(Verval et al., 2001). Moreover, CB1 knockout mice and wild type mice treated with

SR141716A (CB1-R antagonist) exhibited similar performance in the fixed hidden

45

platform water maze task. Interestingly, when the same task was repeated with change

in location of the hidden platform there was a marked difference in the behaviour of both

animals with wild-type returning to the new location and CB1 knock out returning to the

old location of the hidden platform, thus indicating impairment of the extinction process

(Varvel and Lichtman, 2002).

1.18.2. Anxiety

There is increasing evidence suggesting a relationship between anxiety and

endocannabinoid tone. The effects of cannabinoid agonists on anxiety in laboratory

animals can be studied by employing different tests like elevated plus-maze, the light-

dark crossing test, the vocalization test and the social interaction test. The results are

quite complex to interpret although generally, cannabinoid agonists at low doses were

anxiolytic but at high doses were anxiogenic (Viveros et al., 2005). A cross talk

mechanism was suggested by Berrendero and Maldonaldo (2002) between the

cannabinoid and opioid systems since in the light-dark crossing test, the anxiolytic

effects of ∆9-THC were completely blocked by the opoid antagonist, naltrindole.

Moreover, in the plus-maze test, the anxiogenic effect of CP55940 was almost

completely blocked by an opioid antagonist, nor-binaltorphimine (Marin et al., 2003).

1.18.3. Depression

The endocannabinoid system has also been implicated in depressive episodes

(Kano et al., 2009). Hill and Gorzalka (2005) reported an antidepressant effect in

rodents following activation of the CB1-R. They found that in the rat forced swim test

(FST), HU210 (a CB1-R agonist) (5-25µg/kg i.p.) and AM404 (an anandamide transport

inhibitor) (5 mg/kg, i.p.) had antidepressant effects similar to the well-know

antidepressant drug, desipramine and this antidepressant effect was reversed by AM251

(a CB1-R antagonist) (Hill and Gorzalka, 2005).

Moreover recently, the CB1-R antagonist SR141716A (Rimonabant) was

withdrawn from European markets after reports of suicidal tendencies in some patients

employing this drug to treat obesity (Christensen et al., 2007).

46

1.18.4. Addiction

The endocannabinoid system has been known for its drug seeking and drug

reward effects and its likely involvement in addiction (De Vries and Schoffelmeer, 2005;

Fattore et al, 2007; Maldonado et al, 2006). Castane et al. (2002) showed that there

was a significant rewarding effect in wild type mice administered with nicotine (0.5 mg/kg

sc) but not in CB1 knockout mice. Furthermore, this effect was reduced by

administration of SR141716A (Le Foll and Goldberg, 2004).

However, the exact mechanism by which the endocannabinoid system

contributes to drug seeking and addiction behaviour is still rather obscure (Kano et al.,

2009).

1.18.5. Appetite

Food intake, appetite and feeding behaviour are precisely regulated through

involvement of the endocannabinoid system, and the mechanisms involved are subject

to ongoing investigation (Kano et al., 2009). Pagotto et al. (2006) showed that CB1-R

agonists increase food intake in a dose- dependent manner in laboratory animals while

CB1-R antagonists lead to decreased food intake in wild type mice but not in CB1

knockout mice (Di Marzo et al., 2001).

Rodent models and clinical studies strongly support the notion that CB1-Rs can

be targeted for the treatment of appetite disorders and obesity (Kano et al., 2009). ∆9-

THC has been employed effectively as an appetite stimulant in patients with HIV-

induced wasting syndrome and cancer while, as mentioned earlier, rimonabant

(SR141716A) was briefly used in European markets for the treatment of obesity

(Christensen et al., 2007; Kano et al., 2009).

1.18.6. Pain

The endocannabinoid system is intimately involved in pain modulation, and the

antinociceptive effects of cannbinoids have been reported to be at par with the opiates

(Hohmann and Suplita, 2006). Evidence for this came from the studies by Calignano et

al. (1998) who showed that SR141716A induced hyperalgesia in the Formalin test and

47

the hot plate test (Richardson et al., 1998). Also, analgesia induced by electrical

stimulation of periaqueductal gray matter was blocked by SR141716A (Walker et al.,

1999), thus indicating the role of the endocannabinoid system in pain modulation.

1.19. Classification of ligands that bind to cannabinoid receptors

1.19.1. Cannabinoid receptor agonists

1.19.1.1. Classical cannabinoids

This group encompasses derivatives of ABC-tricyclic benzopyran compounds

obtained from the Cannabis plant or synthetic analogs. The prototypical example of this

class is ∆9-THC (Figure 1.10). Others includes ∆8-THC, (6aR,10aR)- 9-(hydroxymethyl)-

6,6-dimethyl-3-(2-methyloctan-2-yl)-6a,7,10,10a-tetrahydrobenzo[c]chromen-1-ol (HU-

210) and desacetyl-L-nantradol (DALN). Here, ∆9-THC and ∆8-THC are

phytocannabinoids while HU-210 and DALN are synthetic (Howlett et al., 2002) (Figure

1.10).

1.19.1.2. Non-classical cannabinoids

Much of the success in developing the non-classical cannabinoids can be

attributed to Pfizer researchers who synthesized new compounds lacking the

dihydropyran ring of ∆9-THC. CP47497, CP55244 and CP55940 are the examples of

this group (Melvin et al., 1993; Devane et al., 1988) (Figure 1.10).

1.19.1.3. Aminoalkylindoles

This group of compounds were developed by Sterling Winthrop researchers and

are structurally related analogs of pravadoline, structurally unrelated to ∆9-THC (unlike

other cannabimimetics) and typified by R-(+)-WIN55212 (Bell et al., 1991; Pacheco et

al., 1991; Howlett et al., 2002) (Figure 1.10).

Other examples of this group include JWH-015 and L-768242 (Figure 1.10).

48

1.19.1.4. Eicosanoids/Endocannabinoids

These compounds which belong to 20:4, n-6 series of fatty acid amides, are

mostly endogenous cannabinoid agonists that are de novo synthesized and released by

mammalian brain (Howlett et al., 2002). Examples of eicosanoids/endocannabinoids

include anandamide (AEA), 2-arachidonyl glycerol (2-AG),

docosatetraenoylethanolamide and 2-AG ether (noladin ether) (Figure 1.11).

1.19.2. Cannabinoid receptor antagonists/ Inverse agonists

1.19.2.1. Diarylpyrazoles

The first cannabinoid receptor antagonist namely, SR141716A (CB1-R-selective)

and SR144528 (CB2-R-selective) were synthesized at Sanofi (Rinaldi-Carmona et al.,

1994 and 1998). Initially identified as cannabinoid antagonists, they were later found in

some preparations to exert effects opposite to cannabinoid agonists and hence were

classified as inverse agonists (Pertwee, 1999). Two other analogs in this diarylpyrazole

series include AM251 and AM281 (Figure 1.12).

1.19.2.2. Other inverse agonists primarily active at CB1-Rs

The Eli Lilly compound, LY320135 (a substituted benzofuran) has a higher

affinity for CB1-Rs then CB2-Rs and displays an inverse agonist profile similar to

SR141716A (Howlett et al., 2002). Similarly, another aminoalkylindole, 6-

iodopravadoline (AM630) was reported to be an inverse agonist at CB2-Rs (Howlett et

al., 2002) (Figure 1-12).

Dhopeshwarkar et al. (2011) found that (S)-methoprene and piperonyl butoxide

were antagonistic at CB1-R while sanguinarine and chelerythrine displayed inverse

agonist-like profiles. Bisset et al. (2011) reported that certain phthalate diesters (nBBP,

DnBP) were antagonist at CB1-Rs. However, all these compounds require low to

moderate micromolar concentrations for activity at CB1-Rs (for structures of these

compounds see Figure 1.13).

49

∆9-THC ∆8-THC HU210

DALN CP47497 CP55244

CP55940 WIN55212-2

JWH015 L-768242

Figure 1.10 Structures of ∆9-THC, ∆8-THC, HU210, DALN, CP47497, CP55244, CP55940, WIN55212-2, JWH015 and L-768242. All structures redrawn using ChemDraw 11.0 ultra from Howlett et al. (2002).

50

Anandamide 2-Arachidonyl glycerol ether (noladin ether)

2-Arachidonyl glycerol (2-AG)

Figure 1.11 Structures of anandamide, 2-AG ether and 2-AG. All structures redrawn using ChemDraw Ultra 11.0 from Howlett et al. (2002).

51

SR141716A AM251 AM281

LY320135 AM630

Figure 1.12 Structures of SR141716A, AM251, AM281, LY320135 and AM630. All structures redrawn using ChemDraw Ultra 11.0 from Howlett et al. (2002).

52

Methoprene Piperonyl butoxide

Sanguinarine Chelerythrine

nBBP DnBP

Figure 1.13 Structures of (S)-methoprene, piperonyl butoxide, sanguinarine, chelerythrine, nBBP and DnBP. Structures redrawn using ChemDraw 11.0 from Dhopeshwarkar et al., (2011) and Bisset et al., (2011).

53

1.20. Cannabinoid receptor 2 (CB2-R)

1.20.1. CB2-R receptor signaling

Like CB1-Rs, CB2-Rs couple to Gi/o proteins and modulate adenylyl cyclase and

MAPK activity. However, unlike CB1-Rs, CB2-Rs do not couple to the Gs subunit of G

protein, and hence cannot modulate ion channel activity (Demuth and Molleman, 2006;

Felder et al., 1995; Kobayashi et al., 2001).

1.20.1.1. Adenylyl cyclase regulation

CB2-Rs are negatively coupled to adenylate cyclase thus they decrease cAMP

production. Slipetz et al. (1995) reported that in cell lines transfected with CB2-Rs,

forskolin-stimulated cAMP production was inhibited by CB2-R agonists and this effect

was concentration-dependent. Similar results were obtained by Felder et al. (1995) with

∆9THC and anandamide. Another line of evidence for coupling of CB2-R to Gi/o proteins

came from experiments where following pretreatment of CB2 transfected CHO cells with

Pertussis toxin was found to attenuate the inhibition of cAMP production (Pertwee,

1997).

1.20.1.2. Mitogen-activated protein kinase regulation

In common with CB1-Rs, CB2-Rs have been implicated in the regulation of MAP

kinase (Pertwee, 1997). CP55940 and WIN55212-2 were found to activate MAP kinase

in CHO cells transfected with CB2-Rs in a concentration-dependent manner (Bouaboula

et al., 1996). This effect of CP55940 was blocked by pretreatment of cells with Pertussis

toxin, indicating the involvement of Gi/o G proteins (Bouaboula et al., 1996). Bouaboula

and co-workers (1996) also found that MAP kinase regulation was mediated through

protein kinase C, since inhibitors of this enzyme decreased CB2-R-mediated activation of

MAP kinase. Interestingly, CHO cells stably transfected with the CB1-R showed

activation of MAP kinase but this activation of unaffected by treatment with a protein

kinase C inhibitor, indicating a distinct difference in the signal transduction mechanism of

each receptor (Bouaboula et al., 1996).

54

1.20.2. Therapeutic aspects of CB2-R modulators

Studies have indicated a potential role of CB2-R agonists in the treatment of

chronic pain (Mackie, 2006). Like CB1 receptor agonists, CB2-Rs agonists (like AM1241,

HU308, JWH133) can be effectively employed for neuropathic and inflammatory pain

treatment with the advantage that they are devoid of the major psychoactive side effects

that often accompany therapy with CB1-R agonists (Mackie, 2006). Moreover, analgesia

produced by CB2-R agonists was found to be sensitive to naloxone (an opioid

antagonist), implicating the opioid pathway and offering a novel indirect mechanism for

how these agonists might relieve neuropathic pain (Ibrahim et al., 2003; Elmes et al.,

2004).

Interestingly, CB2-Rs have also been implicated in bone development and bone

mass density increases in mouse models. CB2-R knockout mice displayed decreased

bone mass compared to normal mice thus hinting towards the possible role of CB2-R

agonists in treatment of bone disorders and osteoporosis in humans (Bab, 2005). Bab

(2005) also found that HU308, a CB2 receptor agonist, reduced bone mass loss in

ovariectomized mice.

Another interesting role of CB2 receptor agonists is their potential therapeutic

utility for treatment of atherosclerotic lesions. Steffens et al. (2005) showed that

treatment of mice on an atherogenic diet with a low dose of ∆9-THC, decreased the

progression of atherosclerotic lesions and this effect was abolished by SR144258, a

CB2-R antagonist. Benito et al. (2003) also predicted the possible role of impaired CB2-

R signaling in plaque formation in Alzheimer’s disease, as well as in the chronic

inflammatory response seen during retroviral encephalitis (Benito et al., 2003; Benito et

al., 2005).

CB2-R agonists may represent promising candidates for treatment of the various

diseases/disorders as mentioned above. However, more studies are clearly warranted

as the exact physiological role of CB2-Rs, their signaling transducers and their agonists

remain to be defined (Mackie, 2006). Preclinical studies and animal models have shown

promise in the development of the treatment of a variety of CB2-R mediated

diseases/disorders. However, detailed clinical trials are needed to understand the exact

role and the therapeutic potential of the CB2-R agonists in human subjects.

55

1.21. Brief overview of the test chemicals used in my research

1.21.1. Benzophenanthridine alkaloids

Alkaloids are a group of nitrogen-containing biological compounds which are

mostly generated as metabolic by-products by higher plants (Maiti and Kumar, 2007).

Their spectrum of activity ranges from anti-inflammatory, antitumor and anti-microbial to

analgesic and spasmolytic (Jursky and Baliova, 2011). Benzophenanthridine alkaloids

(as exemplified by sanguinarine and chelerythrine; Figure 1-13) are class of substances

that can be isolated from plants of the Papaveraceae, Fumariaceae, Rutaceae,

Meliaceae and Caprifoliaceae families. These plants are also known for antifungal,

nematocidal and various other beneficial properties (Simanek et al., 2003).

Benzophenanthridine also show an array of biological and physiological activities when

introduced into mammalian cell preparations; they therefore represent an area of

immense interest (Simanek et al., 2003).

In Russia, Sanguinaria® and Sanguinatrine® containing extract from S.

canadensis and M. cordata respectively is used in toothpastes and mouthwashes

primarily as antiplaque agents and also in feed additives (Simanek et al., 2003).

However, toxic effects attributed to benzophenanthridines are known (Das and Khanna,

1997). Das and Khanna (1997) found that the consumption of edible oil containing

Argemone mexicana seed oil (which contains about 0.5% benzophenanthridines) was

associated with a dropsy-like syndrome. Moreover, long term use of dental products

containing benzophenanthridines have been associated with increased frequency of

leukoplakia of the maxillary vestibule (Eversole et al., 2000).

At the molecular level, benzophenanthridine react with nucleophilic and anionic

moieties in various amino acids that form the backbones of peptides and proteins

(Schmeller et al., 1997). Benzophenanthridines also inhibit protein kinase C (Wang et

al, 1997) and are capable of intercalating with DNA and complexing with genetic material

(Maiti et al., 1982).

Sanguinarine and chelerythrine can exist in the iminium (charged) or

alkanolamine (neutral, pseudobase or hydroxide adduct) forms at pH 1.0-6.0 and 8.5-

56

11.0 respectively, with equilibrium between these two forms at pKa 7.4 (Maiti and

Kumar, 2007).

The biological activity of benzophenanthridines can in part be attributed to their

capacity to maintain equilibrium between both forms at physiological pH (Dvorak and

Simenak, 2007). Benzophenanthridines penetrate the cell membrane in the

alkanolamine form since this form imparts the lipophilicity necessary for efficient

penetration into membranes and the cell (Dvorak and Simenak, 2007; Simenak et al.,

2003). Once inside the cell benzophenanthridines convert to the iminium form (Simenak

et al., 2003). Both form can exert biological effects, and this would depend on how

readily the iminium or pseudobase can access and bind to their respective molecular

targets. Another important consideration of benzophenanthridine activity is the iminium

bond. Since this bond is prone to nucleophilic attack benzophenanthridines readily bind

to SH-groups on proteins (Simenak et al., 2003). Stiborova et al. (2002) also reported

the formation of DNA adducts following activation of sanguinarine and chelerythrine by

cytochrome P450 of human liver microsomes. Moreover, Dvorak and Simenak (2007)

reported the metabolism of sanguinarine by cytosolic and microsomal reductases in rats

and guinea pig intestine. The product (dihydrosanguinarine) can then be metabolized to

different conjugates by sulfation, glucuronidation, O-demethylation and N-demethylation

(Dvorak and Simenak, 2007).

1.21.2. Piperonyl butoxide (PBO)

Piperonyl butoxide (PBO) is a methylenedioxyphenyl compound which is

synthesized from natural saffrole derivatives found in many plant tissues (Wang et al.,

2012, Ugolini et al., 2005). PBO is commercially available as a potent synergist of

pyrethrin-, pyrethroid- and carbamate-based insecticides (Ugolini et al., 2005) (Figure 1-

13). Here synergists can be defined as agents that increase the insecticidal activity of

the insecticidal compound but virtually lack any insecticidal activity on their own

(Franklin, 1976). The popularity of PBO as a synergist for insecticidal compounds is due

to its broad spectrum of synergistic action and very low acute mammalian toxicity (acute

oral LD50 = 10 g/kg in the rat) (Franklin, 1976). PBO was also found to be safe on

chronic exposure at low doses (Franklin, 1976). However, chronic exposures at very

high doses have been reported to be hepatotoxic and to cause kidney damage (Franklin,

57

1976). Franklin (1976) demonstrated that the methylenedioxy moiety was very important

for synergistic activity and metabolic enzymes were capable of cleaving the methylene

carbon.

The synergistic mechanism of PBO action was first studied by Sun and Johnson

(1960) and this was later confirmed by Casida in 1970. PBO was found to inhibit the

cytochrome P450 enzymes responsible for oxidation of insecticides and xenobiotics.

Thus inhibition of cytochrome P450 spared insecticides from metabolic breakdown

therefore allowing them to exert their toxic actions on insects. Binding conformation

studies performed by Keseru et al. (1999) showed that PBO reduced the conformational

mobility of cytochrome P450 and also created a steric hindrance in the channel

responsible for substrate access ito the enzyme.

Schleier III and coworkers (2007) found that PBO undergoes exponential decay

in water over a 36 hour time period indicating a relatively short t1/2 of PBO in the

environment.

1.21.3. Methoprene

Methoprene is a non-cyclic, amber colored synthetic terpenoid commonly used

as a biochemical pesticide belonging to a potent class of insect growth regulators which

mimics the insect juvenile growth hormone (Monteiro et al., 2005; Wilson, 2004; Siddall,

1976)(Figure 1-13). Juvenile growth hormone is important for insect development,

reproduction, behaviour, pheromone production and adult diapauses (Wilson, 2004).

Methoprene, when applied to insects, mimics the action of this hormone (juvenile

hormone III) thus causing disruption of development which leads to death (Wilson,

2004). Another advantage of methoprene as an insect control chemical is its remarkable

insect specificity. Thus methoprene is extremely active against dipteran insects but

practically inactive against lepidopterans (Staal, 1975; Wilson, 2004).

Methoprene is practically non-toxic to mammals with an oral LD50 of > 34,600

mg/kg in rats (Hawkins et al., 1977). However, some toxic effects have been reported in

non-target aquatic insects and some fish at high concentrations (Miura and Takahashi,

1973, Breaud et al., 1977, Quistad et al., 1976). Using the model microorganism

Bacillus stearothermophilus, Monteiro et al. (2005) suggested that the toxic actions of

58

methoprene on non-target organisms may be due to membrane interactions and

disturbance of cell bioenergetics.

Methoprene is rapidly metabolised by mammals (Hawkins et al., 1977). Hawkins

et al. (1977) showed that following oral administration of methoprene to rats, some

methoprene was metabolised by gut flora. Methoprene was also extensively distributed

throughout rat tissues with the highest concentration in the adrenal cortex (Hawkins et

al., 1977). In the environment, methoprene is rapidly degraded by sunlight and it is also

rapidly broken down in soil (t1/2 approx. 10 days in soil; Siddall, 1976; Hawkins et al.,

1977).

1.21.4. Phthalate esters

Phthalate esters are an important class of industrial chemicals that are widely

used as additives in plastic products (Xu et al., 2010, Staples et al., 2011) (Figure 1-13).

Phthalates having higher molecular weights like di-2-ethylhexyl phthalate (DEHP) are

employed as plasticizers to soften polyvinylchloride products, while phthalates with lower

molecular weights like di-n-butyl phthalate (DBP) and butyl benzyl phthalate (BBP) are

primarily used to hold color and scents in various consumer products (Cao, 2010).

Phthalates have been widely reported environmental contaminants (Rudel et al., 2003;

Bornehag et al., 2004; Bornehag et al., 2005) including some foods (Cao, 2010). This

ubiquitous contamination by phthalates is a cause of increasing concern since

toxicological studies have indicated a number of adverse effects displayed by these

compounds. DEHP was found to be a liver carcinogen in rodents (Cao, 2010) while

DBP, BBP and other phthalates were reported to have teratogenic effects in mice and

rats (Blount et al., 2000). Bornehag and coworkers (2004) also showed that there was a

relationship between phthalates in house dust and asthma and allergy in children.

Another important source of human exposure to phthalate is through food (Cao, 2010).

Indeed, several reports have been published to date showing the migration of phthalates

from plastic containers into cooking oil, mineral water (Xu et al., 2010), soft drinks

(Bosnir et al., 2007) and infant foods packed in recycled paperboard (Gratner et al.,

2009).

59

In humans, phthalates tend not to accumulate in the body (Schmid and Schlatter,

1985) because they are quickly metabolized to their respective monoesters which then

undergo glucuronidation and excretion in urine and faeces (Blount et al., 2000).

Phthalates are rapidly photodegraded in air (half life ~1 day), but in water the t1/2 was

found to be slightly longer (Staples et al., 1997). Staples and coworkers (1997) also

showed that phthalates were hydrolysed to form their respective alcohols and phthalic

acid and these intermediates are then further degraded aerobically or anaerobically.

Factors affecting the rate of phthalate degradation were availability of oxygen,

temperature and nutrient content (Staples et al., 1997).

1.21.5. Tributyl tin (TBT) compounds

TBT compounds are organic derivatives of tin (Sn4+) which are widely used as

biocides or as stabilizers in industrial or agricultural sectors (Kannan et al., 1999, Okoro

et al., 2011) (Figure 1-14). Since the 1970s, TBTs have been employed as paint

additives to reduce bio-fouling on ship hulls, marine docks and fishing nets (Dubey and

Roy, 2002). However, a decade later, France became the first country to ban TBT

application as anti-fouling agent in boats less than 25 m long since researchers in

France and UK reported adverse effects of TBTs on nontarget aquatic organisms

(Dubey and Roy, 2002). TBTs have been found to weaken oyster and mussel shells

and also affect the growth and development of aquatic snails (Dubey and Roy, 2002).

The lipophilic nature of these compounds leads to easy entry into cells. TBTs

then often interfere with energy transduction and this can represent a major mechanism

leading to toxicity and death (Cooney and Wuertz, 1989; Cooney, 1995; Dubey and Roy,

2002). However, TBT resistant (or tolerant) bacteria have also been reported (Dubey

and Roy, 2002). These include Alteromonas sp M-1, E.coli, Pseudomonas fluorescens,

Pseudomonas aeruginosa, B. Subtilis and S.aureus. The TBT resistance (tolerance)

can be explained by the inherent capacity of these microorganisms to a) perform

dealkylation of TBTs generating less toxic compounds b) effectively pump TBTs out of

the cell (via Pgp efflux proteins) c) carry out metabolic utilization of TBTs as source of

carbon or d) complex TBTs to metallothionein-like proteins (Fukagawa et al., 1994; Blair

et al., 1982).

60

In humans, the greatest chance of exposure to TBTs is either in drinking TBT-

contaminated water and beverages or eating contaminated aquatic organisms (Antizar-

Ladislao, 2008). TBTs are potentially dangerous for pregnant humans since Nakanishi

(2007) showed that these compounds stimulate human placental oestrogen biosynthesis

and human chronic gonadotropin production in vitro. TBTs have also been showed to

cause cell necrosis or apoptosis in healthy mammalian cells (Nakanishi, 2007; Saitoh et

al., 2001).

The degradation of TBTs in soil, fresh water and marine and estuarine

environments is mostly mediated through biotic processes (Barug, 1981; Dowson et al.,

1996). The t1/2 of most TBTs in soil is >89 days (Wuertz et al., 1991), but in fresh water

it is much less (9 days; Seligman et al., 1986), while in sediments it can be up to 2.1

years (Sarradin et al., 1995). Abiotic degradation processes like breaking of the Sn-C

bond as a result of UV and gamma irradiation and also chemical and thermal cleavage

are also thought to occur (Sheldon, 1975).

61

R1 = R2 = -(CH2)3-CH3, DnBP MnBP

R1 = R2 = -(CH2)5-CH3, DnHP

R1 = -CH2-C6H6, R2= -(CH2)3-CH3, nBP

R= -H, TBT hydride Triphenyl tin chloride

R= -Br, TBT bromide

R= -OCH3, TBT methoxide

R= -OCOOCH3, TBT acetate

R= -C6H6, Tributyl phenyl tin

Figure 1.14 Structures of selected phthalate esters and tributyl tin compounds. All structures redrawn using ChemDraw Ultra 11.0.

62

1.22. Rationale behind my research and the general approach

The pioneering work by Quistad et al., (2002, 2006) at the University of

California, Berkeley has established that various organophosphorus pesticides inhibit the

binding of the radioligand [3H]CP55940 to brain CB1-Rs. However, we know very little of

the extent to which environmental chemicals might interfere with CB1-Rs in the brain.

Both synthetic and natural product chemicals represent a significant component of the

human environment and show an extraordinary diversity in structure. Given the power

of the endocannabinoid system as a fundamental regulator of synaptic strength in many

neuronal networks, even a relatively low level of interference by a moderately potent

xenobiotic would be anticipated to cause some physiological, behavioral or

psychological alterations. Such subtle outcomes, especially if they are psychological or

behavioral, could possibly be missed in the standard toxicological evaluation of

environmental chemicals. Another potential consequence of examining a diverse group

of environmental chemicals at CB1-Rs is that one might obtain a new perspective on a

chemical structure or molecular feature capable of offering a useful way forward in the

design of novel agonists, antagonists or inverse agonists. Such compounds may have

therapeutic value.

Initial exploratory experiments in our laboratory identified a number of chemicals

capable of interfering with the binding of [3H]CP55940 to receptors of mouse brain.

Included in this group were two benzophenanthridine alkaloids (sanguinarine and

chelerythrine), (S)-methoprene and piperonyl butoxide.

The objectives of each phase of my research are detailed below and elaboration

of each objective is provided in the Introductory sections of Chapters 2, 3, 4 and 5.

1.22.1. Summary of objectives

1) To confirm that benzophenanthridine alkaloids (sanguinarine and

chelerythrine), (S)-methoprene and piperonyl butoxide interfere with the binding of

63

[3H]CP55940 to CB1 receptors in mouse brain and establish concentration:inhibition

relationships to determine inhibitory potencies based on IC50 estimates.

2) To classify study compounds as agonists, antagonists or inverse agonists at

CB1-Rs of mouse brain. This phase of the pharmacological profiling utilized the

[35S]GTPγS binding assay, since it allows any modification by study compounds to the

primary functional response of the G-protein to be investigated. These experiments

required effects on basal [35S]GTPγS binding and CB1-R agonist-stimuated [35S]GTPγS

binding to be defined.

3) To explore in greater depth the mechanisms by which study compounds

inhibit the binding of [3H]CP55940 and [3H]SR141716A (radioligands specific for the

brain CB1-R binding pocket) using saturation binding and kinetic approaches. This

phase of the study was extended to phthalates.

4) To investigate the ability of benzophenanthridine alkaloids, (S)-methoprene

and piperonyl butoxide to interfere with the binding of [3H]CP55940 to CB2 receptors in

mouse spleen and explore concentration:inhibition relationships to a level sufficient to

draw conclusions on CB1-R vs CB2-R selectivities.

5) To pharmacologically classify study compounds based on the ultimate

functional presynaptic outcome of drugs interfering with CB1-Rs, that is the release of

neurotransmitter from the nerve ending. For this I employed a continuous fluorometric L-

glutamate release assay. Prior to profiling the study compounds, it was necessary to

verify the inhibitory effect of a standard cannabinoid agonist (WIN55212-2) and

demonstrate the neutralization of WIN55212-2's inhibitory effect by a diarylpyrazole

(AM251), along the lines reported by (Wang et al., 2003). This study compound testing

phase of the investigation was later extended to selected phthalate esters and

tributyltins.

Note: Other researchers in my laboratory were involved in pursuing some of the

above objectives. Appropriate acknowledgements to their work are given, both in the

Results sections and in Figure legends.

64

2. The actions of benzophenanthridine alkaloids, piperonyl butoxide and (S)-methoprene at the G-protein coupled cannabinoid CB1 receptor in vitro.

2.1. Abstract

This investigation focused primarily on the interaction of two

benzophenanthridine alkaloids (chelerythrine and sanguinarine), piperonyl butoxide and

(S)-methoprene with G-protein-coupled cannabinoid CB1 receptors of mouse brain in

vitro.

Chelerythrine and sanguinarine inhibited the binding of the CB1 receptor agonist

[3H]CP55940 to mouse whole brain membranes at low micromolar concentrations (IC50s:

chelerythrine 2.20 µM; sanguinarine 1.10 µM). The structurally related isoquinoline

alkaloids (berberine and papaverine) and the phthalide isoquinoline ((-)-β-hydrastine)

were either inactive or considerably below IC50 at 30 µM. Chelerythrine and sanguinarine

antagonized CP-55940-stimulated binding of [35S]GTPγS to the G-protein (IC50s:

chelerythrine 2.09 µM; sanguinarine 1.22 µM). In contrast to AM251, both compounds

strongly inhibited basal binding of [35S]GTPγS (IC50s: chelerythrine 10.06 µM;

sanguinarine 5.19 µM).

Piperonyl butoxide and S-methoprene inhibited the binding of [3H]CP55940

(IC50s: piperonyl butoxide 8.2 µM; methoprene 16.4 µM), and also inhibited agonist-

stimulated (but not basal) binding of [35S]GTPγS to brain membranes (IC50s: piperonyl

butoxide 22.5 µM; (S)-methoprene 19.31 µM). PMSF did not modify the inhibitory effect

of (S)-methoprene on [3H]CP55940 binding.

Our data suggest that chelerythrine and sanguinarine are effacacious

antagonists of G-protein-coupled CB1 receptors. They exhibit lower potencies compared

65

to many conventional CB1 receptor blockers but act differently to AM251. Reverse

modulation of CB1 receptor agonist binding resulting from benzophenanthridines

engaging with the G-protein component may explain this difference. Piperonyl butoxide

and (S)-methoprene are effacacious, low potency, neutral antagonists of CB1 receptors.

Certain of the study compounds may represent useful starting structures for

development of novel/more potent G-protein-coupled CB1 receptor blocking drugs.

Acknowledgement: This chapter adheres closely to our paper published by the

European Journal of Pharmacology as: Dhopeshwarkar A.S., Jain S., Liao C., Ghose

S.K., Bisset K.M., & Nicholson R.A. (2011). The actions of benzophenanthridine

alkaloids, piperonyl butoxide and (S)-methoprene at the G-protein coupled cannabinoid

CB1 receptor in vitro. European Journal of Pharmacology, 654(1), 26-32.

2.2. Introduction

Cannabinoid CB1 receptors are widely distributed in mammalian brain and occur

at high density in the cerebral cortex, hippocampus, cerebellum and basal ganglia

(Herkenham et al., 1991; Tsou et al., 1998). CB1 receptors are predominantly

presynaptic and interface directly with G-proteins in the neuronal membrane, forming the

initial presynaptic element of a negative feedback mechanism regulating transmitter

exocytosis (Howlett et al., 1986; Katona et al., 1999; Kawamura et al., 2006). During

heightened synaptic activity, postsynaptic neurons generate endocannabinoids which

translocate retrogradely to activate presynaptic CB1 receptors. Activation of the coupled

G-protein leads to inhibition of voltage-sensitive Ca++ channels (Mackie and Hille, 1992;

Twichell et al., 1997, Kushmerick et al., 2004), negative modulation of adenylate cyclase

(Howlett and Fleming, 1984; Howlett, 1985) and activation of K+ currents (Deadwyler et

al., 1993; Mackie et al., 1995). The net effect is a downward adjustment of transmitter

release from the nerve ending (Chevaleyre et al., 2006; Kreitzer and Regehr, 2001;

Wilson and Nicoll, 2001; Howlett, et al., 2002).

66

In addition to endocannabinoids, various other natural and synthetic compounds

including ∆9-tetrahydrocannabinol, CP55940 and WIN55212-2 exert agonist effects at

brain cannabinoid receptors (Devane et al., 1988; Compton et al., 1992). Selective CB1

receptor antagonists such as the diarylpyrazoles AM251 and SR141716A, and the

phytocannabinoid, ∆9-tetrahydrocannabivarin have also been discovered (Rinaldi-

Carmona et al., 1994; Lan et al., 1999; Thomas et al., 2005). These CB1 receptor

modulators exert potent effects in vitro, acting in the nanomolar range. Despite

unfavorable psychiatric side effects associated with the first group of CB1

antagonists/inverse agonists developed to treat obesity, compounds with this

pharmacological profile remain of substantial interest (Szabo et al., 2009; Wu et al.,

2009; Riedel et al., 2009).

Chelerythrine and sanguinarine are quaternary benzophenanthridine alkaloids of

plant origin. We considered that the pseudobase forms of chelerythrine and

sanguinarine might engage CB1 receptors in a similar way to ∆9-tetrahydrocannabinol or

∆9-tetrahydrocannabivarin (see Fig. 2.1) based on preliminary findings that both natural

products displace the binding of [3H]CP55940 to mouse brain membranes. In other

exploratory experiments, binding inhibition was noted for two synthetic chemicals used in

insect pest management: piperonyl butoxide, which we hypothesized may adopt an

endocannabinoid-like conformation; and (S)-methoprene, which may represent a highly

flexible analog of ∆9-tetrahydrocannabinol or ∆9-tetrahydrocannabivarin or perhaps

mimic 2-AG (see Figure 2.1).

The aim of the present work was to investigate the in vitro effects of these study

compounds on the G-protein coupled CB1 receptor in mouse brain in more depth.

Interactions with this signaling complex were evaluated on the basis of ability to 1)

displace the binding of [3H]CP55940, a radioligand that binds to a region of the CB1

receptor shared with the recognition sites for endocannabinoids, classical cannabinoids,

aminoalkylindoles and diarylpyrazoles (Devane et al., 1988; Song and Bonner, 1996;

McAllister et al., 2003) and 2) modify the binding of [35S]GTPγS to brain G-proteins in the

presence and absence of agonist, an assay which determines functional coupling of the

CB1 receptor to its G-protein (Selley et al., 1996; Petitet et al., 1997).

67

2.3. Materials and Methods

2.3.1. Radioligands, drugs and study compounds

[3H]CP-55940[(1R,3R,4R)-3-[2-hydroxy-4-(1,1-dimethylheptyl)phenyl]-4-(3-

hydroxy-propyl)cyclohexan-1-ol; side chain-2,3,4-[3H]; sp. act. 139.6 and 174.6 Ci/

mmol) and guanosine 5'-O-(γ35S]thio)-triphosphate ([35S]GTPγS; sp. act. 1250 Ci/ mmol)

were obtained from Perkin Elmer Life and Analytical Sciences, Canada. Chelerythrine,

berberine, sanguinarine, (as chloride or hydrochloride salts), papaverine, (-)-b-

hydrastine, CP55940, N-piperidin-1-yl)-5-(4-iodophenyl)-1-(2,4-dichlorophenyl)-4-methyl-

1H-pyrazole-3-carboxamide (AM251), 2,3-dihydro-5-methyl-3-[(4-

morpholinyl)methyl]pyrrolo-[1,2,3-de]-1,4-benzoxazin-6-yl](1- naphthyl)methanone

(WIN55,212-2), phenylmethanesulfonylfluoride (PMSF) and piperonyl butoxide were

purchased from Sigma-Aldrich, Canada. (S)-Methoprene (98.5% purity) was kindly

supplied by Doug Vangundy, Director of Speciality Product Development, Wellmark

International (Dallas, Texas).

2.3.2. Animals

Male CD1 mice (20-25g) obtained from Charles River Laboratories, (Saint-

Constant, Quebec, Canada) were used for all experiments. Animals were maintained on

a 12 h light:dark cycle with food and water provided ad libitum. All procedures using

mice adhered to the Canadian Council on Animal Care standards regarding the use of

animals in research and had approval of the Simon Fraser University Animal Care

Committee.

2.3.3. Determination of the effects of study compounds on the binding of [3H]CP55940 to CB1 receptors in mouse brain membranes.

We evaluated several published procedures for the measurement of specific

binding of [3H]CP55,940 to CB1 receptors. The method described by Quistad et al.,

(2002) was adopted with minor modifications for the present investigation. Mice were

euthanized by rapid cervical dislocation and all isolation procedures were carried out at

0-4 oC. Mouse whole brains were homogenized (10 up/down strokes) in ice-cold buffer

68

(Trisma base (100 mM), EDTA (1 mM) adjusted to pH 9 with HCl; 1 brain/10 ml buffer)

using a motor driven homogenizer (pestle rotation approx. 1500 rpm). Homogenates

were centrifuged in a Beckman J2HS centrifuge at 900 x g for 10 min in a JA20 rotor.

The supernatant containing the neuronal membranes was centrifuged at 11,500 x g for

20 min. Pellets were thoroughly resuspended to a protein concentration of close to 6.5

mg/ml in storage buffer (Trisma base (50 mM), EDTA (1 mM) and MgCl2.6H2O (3 mM),

adjusted to pH 7.4 with HCl) and stored in aliquots at -80 oC. When required for

experiments, membranes were thawed on ice, taken up in a 5 ml syringe and thoroughly

resuspended by moving the suspension out and in (6 times) through an 18g needle (with

its square cut tip held close to the base of the tube) and then vortexed. For assay,

compounds (in DMSO; 5 µl) were added to borosilicate glass culture tubes (13 x 100

mm; Kimble-Chase; no siliconization), followed by binding buffer (500 µl; Trisma base

(50 mM), EDTA (1 mM), MgCl2.6H2O (3 mM), BSA (fatty acid free; 3 mg/ml) adjusted to

pH 7.4 with HCl). Membranes (154.3 + 3.5 µg protein) were then added to each tube

and the mixture vortexed and incubated for 15 minutes at room temperature. Following

addition of [3H]CP-55940 (added in 10 µl DMSO; final radioligand concentration 1.0 nM),

the tube contents were thoroughly mixed and incubations run for 90 min at 30 oC with

gentle shaking. Binding reactions were stopped by adding ice-cold wash buffer (0.9%

NaCl containing 2 mg/ml BSA; 1 ml) and membranes were collected by rapid vacuum

filtration on pre-soaked Whatman GF/C filters. Membranes trapped on the filter were

immediately washed (3 x 4 ml) with ice-cold wash buffer. Filters were thoroughly air

dried before adding scintillant (4 ml; BCS, Amersham Bioscience UK) and radioactivity

was quantitated using liquid scintillation counting. Non-specific binding, measured in the

presence of unlabeled CP55,940 or WIN55,212-2 (both at 10 µM), was subtracted from

total binding to yield the specific binding signal which averaged 80.9 + 4.7 % and 80.7 =

3.1% respectively. In each experiment, binding in the absence and presence of

unlabeled CP55,940 or WIN55212-2 was performed in triplicate and test compounds

were assayed in duplicate. A minimum of three experiments were conducted for every

treatment. All protein measurements were carried out as described by Peterson (1977).

69

2.3.4. Determination of the effects of study compounds on basal and CP55940-stimulated [35S]GTPγS binding to mouse brain membranes

The procedure for isolating brain membranes and measuring the effects of study

compounds on basal and agonist-stimulated [35S]GTPγS binding was adapted from that

of Breivogel et al., (2000). The isolation of brain membranes was carried out at 0-4 oC.

Immediately following the cervical dislocation procedure, whole brains were removed

from two mice and homogenized (Polytron Kinematica GmBH; speed setting 6 for 15

seconds) in isolation buffer (Trisma base (50 mM), MgCl2.6H20 (3 mM), EGTA (0.2 mM),

NaCl (100 mM) with pH adjusted to 7.4 with HCl). The homogenate was centrifuged in a

Beckman J2HS centrifuge (JA20 rotor) at 24,000 x g for 25 min, and the resulting pellet

was then resuspended in isolation buffer and re-centrifuged The final membrane pellet

was thoroughly homogenized in isolation buffer, the protein concentration adjusted to 7

mg/ml and aliquots transfered to the -80 oC freezer. After removal from storage at -80 oC, brain membranes were thawed on ice and thoroughly dispersed as described in

Section 2.3. [35S]GTPγS binding experiments were performed as follows. The test

compound (in DMSO; 5 µl) or DMSO control, as appropriate, was placed in the tube first

followed by assay buffer (500 µl; isolation buffer (pH 7.4) containing, bovine serum

albumin (fatty-acid free; 1 mg/ml), guanosine diphosphate (GDP; 100 µM), dithiothreitol

(20 µM), [35S]GTPγS (0.14 nM final concentration) and adenosine deaminase (0.004

units/ml). The brain membranes (70.1 + 4.2 µg protein) were then added and after

thorough vortexing, a 15 min preincubation at room temperature was carried out.

Following this, CP55940 (100 nM final concentration; in 5 µl DMSO) or DMSO control

was added (total assay volume = 535 µl), the samples were mixed thoroughly and the

incubation continued at 30 0C for 90 min with gentle shaking. Where effects on basal

binding were investigated, no agonist addition was made after the preincubation. The

incubation was terminated by the addition of ice-cold wash buffer (2 ml; Trisma base:

HCl; pH 7.4) and rapid filtration under vacuum through pre-soaked Whatman GF/B

filters. This was quickly followed by three 4 ml washes of the membranes on the filter.

Membrane-bound 35S was quantitated as described in Section 2.3. Assay tubes

(borosilicate glass culture tubes; 13 x 100 mm) were siliconized with Sigmacote (Sigma-

Aldrich Canada) 24h prior to assay. All assays were conducted in triplicate. Basal

binding, defined as the binding occuring in the absence of agonist (CP55940) minus

70

non-specific binding (measured with 100 µM unlabelled GTPγS present) averaged 95.4

+ 1.4%. 100 nM CP-55940 increased the basal binding of [35S]GTPγS by 65.6 + 2.8%.

2.3.5. Data analysis

Results are given as the mean ± S.E.M. Curve fitting by non-linear regression

analysis and estimation of IC50 (concentration of study compound producing 50%

inhibition) was carried out using Prism 4 (GraphPad Software Inc., San Diego, CA). Ki

values for study compounds were determined using the Cheng-Prusoff equation with

0.35 nM as the Kd for [3H]CP55940.

2.4. Results

Fig. 2.2 shows the concentration-dependent inhibition of [3H]CP55940 binding to

mouse brain CB1 receptors by benzophenanthridines under equilibrium conditions.

Sanguinarine and chelerythrine exhibited inhibitory potencies as estimated from IC50s of

1.10 µM (95% CI = 0.62-1.93 µM) and 2.20 µM (95% CI = 1.55-3.13 µM) respectively.

At higher concentrations both approached full inhibition of [3H] radioligand binding, with

sanguinarine slightly more effacacious than chelerythrine.

The isoquinoline alkaloids berberine and papaverine, as well as the phthalide

isoquinoline (-)-β-hydrastine individually achieved no greater than 17.6 % inhibition of

[3H] radioligand binding at 30 µM (Table 1). PMSF (0.5 mM) also had no effect on

[3H]CP55940 binding (Table 2.1).

Chelerythrine and sanguinarine inhibited both agonist- (CP55940-) stimulated

and basal binding of [35S]GTPγS binding to mouse brain membranes (Figs. 2.3 and 2.4).

In the agonist-stimulated [35S]GTPγS binding assays (Figs. 2.3a and 2.4a), chelerythrine

produced an additional 54.92 + 2.54% and 67.52 + 3.40% encroachment into the basal

signal at 20 µM and 50 µM respectively and sanguinarine caused an additional 50.12 +

2.84%, 68.28 + 0.56% and 74.76 + 1.05% encroachment into the basal signal at 4 µM,

10 µM and 50 µM respectively. The IC50s for agonist stimulation were: chelerythrine 2.09

µM (95% CI = 1.73-2.44 µM) and sanguinarine 1.22 µM (95% CI = 1.05-1.50 µM). The

IC50s for basal binding were: chelerythrine 10.06 µM (95% CI = 7.18-15.54 µM) and

71

sanguinarine 5.19 µM (95% CI = 4.59-5.89 µM). Under identical conditions, AM251

inhibited 100 nM CP55940-stimulated [35S]GTPγS binding by 59.59% and 98.93% at

0.01 µM and 1 µM respectively (Table 2.2), confirming that the G-protein under

investigation is coupled to CB1 receptors. However, at 20 µM AM251 produced minimal

(circa 10%) inhibition of basal [35S]GTPγS binding (Table 2.2). No significant inhibitory

effects of berberine, papaverine, (-)-β-hydrastine (all at 40 µM) on CP55940-stimulated

or basal [35S]GTPγS binding were detected (Table 2.3).

The effects of piperonyl butoxide and (S)-methoprene on [3H]CP55940 binding

and agonist-stimulated [35S]GTPγS binding are shown in Fig. 5a and 5b. Piperonyl

butoxide and (S)-methoprene inhibited [3H]CP55940 binding to CB1 receptors [IC50s:

piperonyl butoxide 8.2 µM (95% CI = 7.08-9.32 µM) and (S)-methoprene 16.4 µM (95%

CI = 13.7-19.06 µM)]. In parallel experiments, 4 µM (S)-methoprene inhibited [3H]CP-

55940 binding by 27.32 + 4.11% and 28.16 + 4.17% in the absence and presence of 50

µM PMSF respectively, showing that esterases were not limiting the inhibitory potency of

this compound. Both compounds also blocked CP55940-stimulated binding of

[35S]GTPγS to the G-protein [IC50s: piperonyl butoxide 22.5 µM (95% CI = 18.98-36.02

µM) and (S)-methoprene 19.31 µM (95% CI = 17.01-21.61 µM)]. No effects of piperonyl

butoxide or (S)-methoprene on basal binding of [35S]GTPγS were observed (Table 2.4).

2.5. Discussion

The results of this investigation demonstrate that benzophenanthridine alkaloids,

piperonyl butoxide and (S)-methoprene inhibit the G-protein-coupled CB1 receptor of

mammalian brain, and suggest a clear functional difference between the actions of the

natural product alkaloids and the synthetic compounds at this complex.

The IC50s of sanguinarine and chelerythrine in the [3H]CP55940 binding assay lie

in the 1-2 µM range, which places them very similar in potency to cannabidiol,

virodhamine, various ∆8-tetrahydrocannabinol derivatives and certain bicyclic resorcinols

(Devane et al., 1988; Compton et al., 1993; Steffens et al., 2005; Wiley et al., 2002),

however, these benzophenanthridines are considerably less potent than ∆9-

tetrahydrocannabinol and ∆9-tetrahydrocannabivarin, which inhibit the binding of

72

[3H]CP55940 at low nanomolar concentrations (Devane et al., 1988; Thomas et al.,

2005).

Antagonist-like actions for sanguinarine and chelerythrine at the G-protein-

coupled CB1 were strongly indicated since both alkaloids produce inhibition of CP55940-

stimulated [35S]GTPγS binding at IC50s identical to those for displacement of [3H]CP-

55940 binding to mouse brain membranes. The effect of these benzophenanthridines in

the agonist-stimulated [35S]GTPγS assay is qualitatively similar to that of the CB1

receptor-selective antagonist AM251. However, at greater than maximum effect

concentrations, and in marked contrast to AM251, sanguinarine and chelerythrine

showed considerable encroachment into the basal component of [35S]GTPγS binding

with CP55940 present. Again unlike AM251, chelerythrine and sanguinarine strongly

inhibited basal [35S]GTPγS binding. On the basis of these G-protein modulatory profiles,

chelerythrine and sanguinarine may be more reasonably classified as inverse agonists

at the G-protein-coupled CB1 receptor. The slightly higher inhibitory potency of

sanguinarine compared to chelerythrine in the [3H]CP55940 and agonist stimulated

[35S]GTPγS binding assays was likely due to the methylenedioxy moiety in place of the

two methoxy groups on ring A. We cannot say whether the cationic (quaternary

ammonium) form or the hydroxide adduct (pseudobase; which is formed at physiological

pH; Slalinova et al., 2001) binds to the G-protein-coupled CB1 receptor. However, the

existence of an equilibrium state raises the possibility that that lower concentrations of

an active species may cause the effects described here. The failure of berberine (which

can also form the pseudobase), papaverine and (-)-β-hydrastine to interact with the G-

protein-coupled CB1 receptor in these assays indicates that aromatic ring stacking and

position of the nitrogen atom are important for inhibitory activity.

Our original hypothesis that the benzophenanthridines may bind in a similar way

to phytocannabinoids at the CB1 receptor seems improbable. However, since it is known

that 1) GTP and non-hydrolysable GTP analogs including GTPγS and guanyl-5’-yl

imidodiphosphate allosterically dissociate [3H]CP55940 from the CB1 receptor (Devane

et al., 1988; Houston and Howlett, 1993) and that 2) low micromolar concentrations of

sanguinarine and chelerythrine inhibit the binding of a fluorescent GTP probe to the GTP

binding protein Rac1b in a competitive fashion (Beausoleil et al., 2009), we think it most

73

likely that chelerythrine and sanguinarine exert retrograde allosteric inhibition of agonist

binding to the CB1 receptor by targeting the guanine nucleotide recognition site on the

associated G-protein. In theory, a drug that acts selectively in this fashion could offer an

alternative mechanism to the diarylpyrazoles for downregulating endocannabinoid-

mediated signaling in the CNS, and therefore may have potential in weight reduction and

the treatment of various metabolic disorders in humans. It is also possible that the level

of inhibition of endocannabinoid activation of CB1 receptors in vivo may be more readily

managed with moderate potency drugs that are selective for the G-protein component of

this complex, potentially reducing psychiatric side effects. Some evidence exists for

selectivity of benzophenanthridines, since it is known that chelerythrine and

sanguinarine bind more avidly to certain GTP binding proteins than members of the

berberine series (Beausoleil et al., 2009). In addition, our data on basal and agonist-

stimulated [35S]GTPγS binding, which reflect respectively the sum of binding to all GTP

binding sites in the membrane fraction versus only those specifically activated by

CP55940, indicate that the latter response is 4-5 fold more sensitive to blockade by

chelerythrine and sanguinarine. Benzophenanthridines may therefore offer a novel area

of chemistry for development of drugs that negatively regulate the endocannabinoid

system, but achieving selectivity at CB1 over CB2 receptors through this mechanism may

be difficult. It is important to note that chelerythrine was reported originally as a potent

and specific protein kinase C inhibitor (Herbert et al., 1990), although considerable doubt

now exists (Lee et al. 1998), and studies by Garcia et al. (1998) found that stimulation of

protein kinase C with phorbol 12-myristate 13-acetate phosphorylates the CB1 receptor

which in turn suppresses both cannabinoid-induced activation of an inwardly rectifying

K+ current and depression of P/Q type Ca++ channel activity. If, in our experiments, the

benzophenanthridines were inhibiting protein kinase C through this particular

mechanism, we would predict they would not block (or possibly facilitate) CP-55940-

mediated activation of the G-protein. Consistent with the results presented here,

chelerythrine has been reported to inhibit desacetyllevonantradol-dependent activation

of the G-protein-coupled CB1 receptor in N18TG2 neuroblastoma cells, however, this

observation was interpreted as supporting other direct evidence for modulation of a

downstream protein kinase C by the CB1 receptor (Rubovitch et al., 2004).

74

We also considered the possibility that our study compounds may act on CB1

receptors indirectly by increasing the levels of anandamide and/or 2-AG through

inhibition of endocannabinoid degrading enzymes (eg FAAH or MAGL). Such a

mechanism is involved in the action of organophosphorus compounds (Nomura et al.,

2008). However, since these endocannabinoids are CB1 receptor agonists, our study

compounds would be expected to significantly increase the basal [35S]GTPγS binding

signal (as we show with CP55940). This is clearly not the case since the

benzophenanthridines strongly inhibit, and methoprene and piperonyl butoxide fail to

influence basal [35S]GTPγS binding. Furthermore, 0.5 mM PMSF (a concentration which

would be expected to fully inhibit FAAH and provide about 50% inhibition of MAGL) has

no influence on the binding of [3H]CP55940 to CB1 receptors in our in vitro system,

offering another line of evidence that our study compounds could not act by this

mechanism.

Piperonyl butoxide is widely used as a synergist for insecticides (Jones, 1998),

while methoprene exerts growth regulatory effects in insects rather than direct toxicity

(Henrick, 2007). Both exhibit low mammalian toxicity [acute oral LD50s in the rat: 7,500-

10,000 mg/kg (piperonyl butoxide) and >34,600 mg/kg (methoprene); Siddall, 1976;

Hawkins et al., 1977]. Our results indicate that in contrast to the benzophenanthridines,

piperonyl butoxide and (S)-methoprene act as neutral antagonists of the CB1 receptor,

giving some support to our hypothesis that they act as structural mimics of

phytocannabinoids or endocannabinoids. However, based on displacement of

[3H]CP55940 binding, their potencies are approximately 4-10 fold lower than the CB1

receptor inhibitor virodhamine (Steffens et al., 2005) and 2-hydroxyphenyl

arachidonamide (Edgemont et al., 1995), 12-24 fold lower than LH-21 (Chen et al.,

2008) and 130-260-fold lower than ∆9-tetrahydrocannabivarin (Thomas et al., 2005).

While a variety of hydroxylated analogs of anandamide have been evaluated at CB1

receptors (Van der Stelt et al., 2002), we are not aware of attempts to optimize

alkylalkoxymethylenedioxyphenyl systems, or methoprene. Such compounds may offer

alternative starting templates for optimization of novel CB1 receptor modulators.

75

2.6. Figures and Tables

N

O

O

O

O

C

H

3

C

H

3

N

C

H

3

O

O

O

O

C

H

3

C

H

3

(a) Sanguinarine (b) Berberine (c) Chelerythrine

N

O

O

C

H

3

O

O

C

H

3

C

H

3

C

H

3

N

O

O

O

C

H

3

O

O

C

H

3

O

H

H

C

H

3

O

H

N

H

O

(d) Papavarine (e) (-)(1R,9S)-β-Hydrastine (f) Anandamide

O

O

O

O

O

O

O

O

H

O

H

(g) Piperonyl butoxide (h) 2-Arachidonoyl glycerol

O

O

O

C

H

3

H

(i) (S)- Methoprene

Figure 2.1 The structures of sanguinarine, berberine, papavarine and possible comparison of conformations of piperonyl butoxide and (S)-methoprene with anandamide and 2-arachidonoyl glycrol. Also possible comparison of sanguinarine and (S)-methoprene with ∆9-tetrahydrocannabinol and ∆9-tetrahydrocannabivarin (continued on page 76).

76

N

C

H

3

O

H

O

O

O

O

O

O

H

(j) Sanguinarine(pseudobase form) (k) ∆9-Tetrahydrocannabinol

O

O

O

C

H

3

H

O

O

H

(l) (S)-methoprene(m) ∆9- Tetrahydrocannabivarin

Figure 2.1 (continued) The structures of sanguinarine, berberine, papavarine and possible comparison of conformations of piperonyl butoxide and (S)-methoprene with anandamide and 2-arachidonoyl glycrol. Also possible comparison of sanguinarine and (S)-methoprene with ∆9-tetrahydrocannabinol and ∆9-tetrahydrocannabivarin.

(a) - (e): The structures of sanguinarine, chelerythrine, berberine, papaverine and (-) (1R,9S)-β-hydrastine. (f) - (i): Comparisons of possible conformations of piperonyl butoxide and (S)-methoprene with anandamide and 2-arachidonoyl glycerol respectively. (j - m): Comparison of the pseudobase form of sanguinarine and a possible conformation of (S)-methoprene with ∆9-tetrahydrocannabinol and ∆9-tetrahydrocannabivarin.

All structures redrawn using IsisDraw from Dhopeshwarkar et al. (2011).

77

Figure 2.2 Concentration-dependent inhibition of [3H]CP55940 binding to mouse brain CB1 receptors by sanguinarine and chelerythrine. Values represent mean + S.E.M. of at least 3 independent experiments each performed in duplicate. Ki values were 0.38 µM (sanguinarine) and 0.57 µM (chelerythrine).

78

Figure 2.3a Inhibition of A) CP55940-stimulated and B) basal binding of [35S]GTPγS to mouse brain membranes by chelerythrine. Values represent mean + S.E.M. of 3 independent experiments each performed in triplicate.

79

Figure 2.3b Inhibition of A) CP55940-stimulated and B) basal binding of [35S]GTPγS to mouse brain membranes by chelerythrine. Values represent mean + S.E.M. of 3 independent experiments each performed in triplicate.Basal binding data provided by Mr Saurabh Jain.

80

Figure 2.4a Inhibition of A) CP55940-stimulated and B) basal binding of [35S]GTPγS to mouse brain membranes by sanguinarine. Values represent mean + S.E.M. of 3 independent experiments each performed in triplicate.

81

Figure 2.4b Inhibition of A) CP55940-stimulated and B) basal binding of [35S]GTPγS to mouse brain membranes by sanguinarine. Values represent mean + S.E.M. of 3 independent experiments each performed in triplicate. Basal binding data provided by Mr Saurabh Jain.

82

Figure 2.5a A) Concentration-dependent inhibition of [3H]CP55940 binding to mouse brain CB1 receptors by (S)-methoprene and piperonyl butoxide. Ki values were 2.13 µM (methoprene) and 4.25 µM (piperonyl butoxide). B) Inhibition of CP55940-stimulated binding of [35S]GTPγS to mouse brain membranes by (S)-methoprene and piperonyl butoxide. Values represent mean + S.E.M. of 3 independent experiments each performed in triplicate.

83

Figure 2.5b A) Concentration-dependent inhibition of [3H]CP55940 binding to mouse brain CB1 receptors by (S)-methoprene and piperonyl butoxide. Ki values were 2.13 µM (methoprene) and 4.25 µM (piperonyl butoxide). B) Inhibition of CP55940-stimulated binding of [35S]GTPγS to mouse brain membranes by (S)-methoprene and piperonyl butoxide. Values represent mean + S.E.M. of 3 independent experiments each performed in triplicate.

84

Table 2.1 Inhibition of specific [3H]CP55940 binding to mouse brain membranes by isoquinoline type compounds and PMSF. Isoquinolines were present in the assay at 30 µM and PMSF was present at 0.5 mM. Data represent mean + S.E.M. of 3 independent experiments.

Compound Inhibition (%)

Berberine 12.05 + 2.2

1R, 9S-(-)-β-Hydrastine 4.09 + 1.64

Papaverine 17.56 + 0.5

PMSF 1.27 + 3.12

85

Table 2.2 Inhibition of 100 nM CP55940-stimulated and basal [35S]GTPγS binding to mouse brain membranes by AM251. Data represent mean + S.E.M. of 3 independent experiments. ND = not determined. Results provided by Mr Saurabh Jain.

AM251

(µM)

Inhibition of CP55940-stimulated

[35S]GTPγS binding (%)

Inhibition of

basal [35S]GTPγS

binding (%)

0.010 59.59 + 2.66 ND

1.0 98.93 + 1.62 ND

10.0 100 7.61 + 6.32

20.0 ND 10.01 + 1.37

9.21 + 0.76% encroachment of AM251 on the basal component of [35S]GTPγS

binding was observed in experiments involving 10 µM CP55940 agonist, as observed by

others with the closely related analog SR141716A (Selley et al., 1996; Petitet et al.,

1997).

86

Table 2.3 Lack of effect of isoquinoline type compounds on CP55940-stimulated and basal [35S]GTPγS binding to mouse brain membranes. Study compounds were present in the assay at 40 µM. Data represent mean + S.E.M. of 3 independent experiments.

Compound

Inhibition of CP55940-

stimulated [35S]GTPγS

binding (%)

Effect on basal

[35S]GTPγS binding

(+ = % increase; - = % decrease)

Berberine 2.82 + 0.89 a 1.98 + 0.35

1R, 9S-(-)β-

Hydrastine

2.23 + 2.23 - 0.50 + 5.77

Papaverine 5.79 + 1.77 3.90 + 2.32

a represents a % increase in CP-55940-induced stimulation

87

Table 2.4 Lack of effect of piperonyl butoxide and (S)-methoprene on the basal binding of [35S]GTPγS to mouse brain membranes. Values represent mean + S.E.M. of 3 independent experiments.

Compound

Concentration

(% increase)

Piperonyl butoxide 20 µM 8.93 + 6.20

30 µM 7.21 + 1.61

40 µM 1.47 + 0.47

(S)-Methoprene 25 µM 4.33 + 6.32

40 µM 2.80 + 0.80

50 µM 0.03 + 5.59

88

3. The G protein-coupled cannabinoid-1 (CB1) receptor of mammalian brain: Inhibition by phthalate esters in vitro.

3.1. Abstract

This research examines the in vitro interaction of phthalate diesters and

monoesters with the G protein-coupled cannabinoid 1 (CB1) receptor, a presynaptic

complex involved in the regulation of synaptic activity in mammalian brain. The diesters,

n-butylbenzylphthalate (nBBP), di-n-hexylphthalate (DnHP), di-n-butylphthalate (DnBP),

di-2-ethylhexylphthalate (DEHP), di-isooctylphthalate (DiOP) and di-n-octylphthalate

(DnOP) inhibited the specific binding of the CB1 receptor agonist [3H]CP55940 to mouse

whole brain membranes at micromolar concentrations (IC50s: nBBP 27.4 µM; DnHP 33.9

µM; DnBP 45.9 µM; DEHP 47.4 µM; DiOP 55.4 µM; DnOP 75.2 µM). DnHP, DnBP and

nBBP achieved full (or close to full) blockade of [3H]CP55940 binding, whereas DEHP,

DiOP and DnOP produced partial (55-70%) inhibition. Binding experiments with

phenylmethane-sulfonylfluoride (PMSF) indicated that the ester linkages of nBBP and

DnBP remain intact during assay. The monoesters mono-2-ethylhexylphthalate

(M2EHP) and mono-isohexylphthalate (MiHP) failed to reach IC50 at 150 µM and mono-

n-butylphthalate (MnBP) was inactive. Inhibitory potencies in the [3H]CP55940 binding

assay were positively correlated with inhibition of CB1 receptor agonist-stimulated

binding of [35S]GTPγS to the G protein, demonstrating that phthalates cause functional

impairment of this complex. DnBP, nBBP and DEHP also inhibited binding of

[3H]SR141716A, whereas inhibition with MiHP was comparatively weak and MnBP had

no effect. Equilibrium binding experiments with [3H]SR141716A showed that phthalates

reduce the Bmax of radioligand without changing its Kd. DnBP and nBBP also rapidly

enhanced the dissociation of [3H]SR141716A. Our data are consistent with an allosteric

mechanism for inhibition, with phthalates acting as relatively low affinity antagonists of

CB1 receptors and cannabinoid agonist-dependent activation of the G-protein. Further

89

studies are warranted, since some phthalate esters may have potential to modify CB1

receptor-dependent behavioral and physiological outcomes in the whole animal.

Acknowledgement: This chapter adheres closely to our paper published by

Neurochemistry International as: Bisset, K.M., Dhopeshwarkar, A.S., Liao C., Nicholson

R.A. (2011). The G protein-coupled cannabinoid-1 (CB1) receptor of mammalian brain:

Inhibition by phthalate esters in vitro. Neurochemistry International, 59(1), 706-713.

3.2. Introduction

CB1 receptors together with their associated G-proteins form an integral part of

the endocannabinoid system which regulates the activity at many synapses in the brain

via a negative feedback mechanism. When synaptic activity intensifies, postsynaptic

neurons generate endocannabinoids which translocate retrogradely and bind to

presynaptic CB1 receptors. Activation of the G-protein then causes inhibition of voltage-

gated Ca++ channels (Mackie and Hille, 1992; Twichell et al., 1997, Kushmerick et al.,

2004) and activation of inwardly rectifying K+ channels (Deadwyler et al., 1993; Mackie

et al., 1995). In addition, the lowering of cAMP levels through negative modulation of

adenylate cyclase by the G-protein (Howlett and Fleming, 1984; Howlett, 1985; Bidault-

Russell et al., 1990) can reduce protein kinase A activity (Kim et al., 2006) which favors

activation of A-type K+ channels (Hoffman and Johnson 1998). Stimulation of

presynaptic CB1 receptors therefore downregulates Ca++-dependent release of

neurotransmitters (Chevaleyre et al., 2006; Kreitzer and Regehr, 2001; Wilson and

Nicoll, 2001) and changes in K+ conductances reduce presynaptic excitability (Mackie et

al., 1995; Hoffman et al., 1997; Zona et al., 2002).

In addition to endocannabinoids, several classes of synthetic CB1 receptor

activators have been identified (Devane et al., 1988; Compton et al., 1993; Felder et al.,

1995) and correlations between their binding affinities at CB1 receptors and in vivo

outcomes such as antinociception, hypothermia, catelepsy and psychoactivity have been

90

demonstrated (Compton et al., 1993). Moreover, various behavioral responses of

classical CB1 receptor agonists are typically blocked by antagonists of CB1 receptors

such as SR141716A (Compton et al., 1996; Rinaldi-Carmona et al., 1995).

Investigations involving phytocannabinoids of Cannabis sativa, various synthetic ligands

and endogenous cannabinoids of the CNS have radically improved our understanding of

the neurobiology and pharmacology of the endocannabinoid system to the extent that

novel therapeutic applications can now be pursued (Pertwee, 1997; Di Marzo, 2009;

Muccioli, 2007; Thakur et al., 2009).

By contrast, considerably less emphasis has been placed on synthetic

environmental chemicals and their potential to modify CB1 receptor function in

mammalian brain. However, it is known that binding of [3H]CP55940 to CB1 receptors is

strongly inhibited by certain organophosphorus and organosulfur compounds

incorporating longer chain alkyl moieties (Segall et al., 2003), and also by various

organophophorus pesticides (Quistad et al., 2002). We have demonstrated that at

micromolar concentrations, methoprene and piperonyl butoxide (chemicals of low acute

toxicity used in pest management) can inhibit CP55940 action at CB1 receptors in vitro,

possibly by adopting endocannabinoid-like conformations or by methoprene acting

alternatively as a flexible analog of ∆9-tetrahydrocannabinol (Dhopeshwarkar et al.,

2011).

During exploratory investigations we found that certain phthalate esters reduce

the binding of [3H]CP55940. Phthalate esters are found in a wide range of consumer and

industrial products (Schettler, 2006). They are utilized as formulation components in

paints, adhesives and medicines as well as in personal care and pest management

products. As phthalate esters remain relatively mobile within polymeric matrices, they

are frequently added to plastics to enhance the flexibility, elasticity and self-lubricating

properties of the end products. Significant quantities of phthalate esters are incorporated

into plastics used to manufacture toys, food wrappings and medical devices (Heudorf et

al., 2007). Human exposures to phthalate esters can occur following dermal, oral and

vapor phase contact and exposures resulting from medical procedures and PVC

manufacturing can be significant for some phthalates (Kavlock et al., 2002a and 2002b)

91

The acute toxicities of phthalate esters in rats and mice are generally quite low

(Calley et al., 1966; Heudorf et al., 2007). However, following chronic administration to

rodents, several phthalate esters have been identified as reproductive and

developmental toxicants (Huendorf et al., 2007) and evidence of hepatic and renal

impairment has also been demonstrated for example with DEHP in some rodent models

(Kavlock, et al., 2002a). Kim et al. (2009) reported a positive association between

phthalate ester exposure (measured as urinary phthalate metabolites) and symptoms of

attention-deficit/hyperactivity disorder in school-age children. A number of phthalate

esters influence brain function following intraperitoneal injection as evidenced by their

ability to modify hexobarbital sleep time in rodents (Calley et al., 1966). Additionally, the

cholinergic system of vertebrate brain is sensitive to phthalate esters. Various diesters

(and to a lesser extent monoesters) inhibit nicotinic acetylcholine receptor-mediated

calcium signaling in human neuroblastoma cells (Liu et al., 2009; Lu et al., 2004) and

fish exposed to dietary dibutyl- and diethylhexyl- phthalates show reduced brain

acetylcholinesterase activity (Jee et al., 2009).

In the present investigation we examined a group of phthalate diesters and

monoesters for their ability to influence the binding of [3H]CP55940 and [3H]SR141716A.

These radioligands and also the endocannabinoid anandamide bind to closely

associated loci within the CB1 receptor binding pocket (McAllister et al., 2003). We then

carried out experiments with [35S]GTPγS to evaluate potential interference of phthalates

with the functional coupling of the CB1 receptor to its G-protein (Selley et al., 1996;

Petitet et al., 1997).

3.3. Materials and methods

3.3.1. Animals

This research was carried out on male mice (weight approx. 25 g; CD1 strain)

that were purchased from Charles River Laboratories (Saint-Constant, Quebec,

Canada). Animals had continuous access to water and food (rodent Lab Diet 5001).

Mice were subjected to a 12 h light/12 h dark cycle and were maintained at 21 + 2 oC

and 50 + 10 % relative humidity. All experimental procedures using mice followed

92

guidelines developed by the Canadian Council on Animal Care and had prior approval of

the Animal Care Committee at Simon Fraser University.

3.3.2. Investigation of the effects of phthalate esters on the binding of [3H]CP55940 and [3H]SR141716A to CB1 receptors of mouse brain.

We used the method of Quistad et al., (2002) to investigate the effects of

phthalates on [3H]CP55940 binding. Membrane isolation procedures were carried out at

0-4 oC. Mouse whole brains were homogenized (10 up/down strokes) in ice-cold buffer

(Trisma base (100 mM), EDTA (1 mM) adjusted to pH 9 with HCl; 1 brain/10 ml buffer)

using a motor driven homogenizer (pestle rotation rate approx. 1500 rpm). The

homogenate was centrifuged at 900 x g (Beckman J2HS; JA20 rotor) for 10 min. The

supernatant was centrifuged at 11,500 x g for 25 min. After thorough resuspension of the

membrane pellet in a small volume of ice-cold buffer [Trisma base (50 mM), EDTA (1

mM) and MgCl2.6H2O (3 mM); adjusted to pH 7.4 with HCl], the protein concentration

was diluted to approx. 6.5 mg/ml and the binding preparation then stored in aliquots at -

80 oC. Just prior to experiments, the membrane suspension was thawed on ice, taken up

in a 5 ml syringe and slowly forced into a tube through an 18g needle (with its square cut

tip held close to the base of the tube). The membranes were moved in and out of the

syringe 5-6 times then thoroughly vortexed, a procedure that helped reduce variability

between replicates. Phthalate esters (see Figure 3.1 for structures), dissolved in 5 µl

DMSO (or 5 µl DMSO as the solvent control) were added to borosilicate glass culture

tubes (13 x 100 mm; without siliconization) and this was followed by addition of binding

buffer (500 µl; Trisma base (50 mM), EDTA (1 mM), MgCl2.6H2O (3 mM), BSA (fatty acid

free; 3 mg/ml) adjusted to pH 7.4 with HCl). Brain membranes (170.67 + 0.84) µg

protein) were then placed in each tube and the suspension was then thoroughly

vortexed and pre-incubated at room temperature for 15 minutes. Additions of [3H]CP-

55940 (side chain-2,3,4-[3H]; sp. act. 174.6 Ci/ mmol; Perkin Elmer Life and Analytical

Sciences, Canada) to each tube were made in 10 µl DMSO (final radioligand

concentration 1.0 nM; final DMSO concentration 2.8%), and after careful mixing,

incubations were run for 1.5 h at 30 oC with gentle shaking. Incubations were stopped by

the addition of 1 ml of ice-cold wash buffer (0.9% NaCl containing 2 mg/ml BSA) and

membranes were quickly harvested by vacuum filtration on pre-soaked Whatman GF/C

93

filters. Membranes trapped on filters were immediately washed (3 x 4 ml) with ice-cold

wash buffer. Filters were completely dried (in a fume hood). Scintillation cocktail (BCS,

Amersham Bioscience UK) was then added and radioactivity was measured using liquid

scintillation counting. Non-specific binding, measured in the presence of unlabeled

WIN55,212-2 (10 µM), was subtracted from total binding to calculate the specific binding

signal (76.8 + 1.1% of total binding). This is very similar to the 80% obtained by Quistad

et al. (2002) using 1 µM WIN55212-2. Under the present assay conditions the IC50 for

WIN55212-2 was 6 nM and maximum displacement of [3H]CP55940 by WIN55212-2

was achieved at both 1 and 10 µM.

Selected phthalate esters were also evaluated in binding assays using the CB1

receptor antagonist [3H]SR141716A (sp. act. 56 Ci/ mmol; Perkin Elmer Life and

Analytical Sciences, Canada). For competitive displacement assays an identical

experimental procedure to that described above was used. [3H]SR141716A was

present at 1.2 nM and AM251 (at 2 µM) was introduced to estimate the specific binding

signal, which averaged (71.0 + 0.7%). For association experiments, membranes were

either preincubated with the phthalate ester for 15 min. before [3H]SR141716A addition

or received simultaneous application of phthalate and radioligand. Dissociations were

initiated on CB1 receptors equilibrated with [3H]SR141716A using either a saturating

concentration of AM251 or this concentration of AM251 plus the phthalate ester.

In each experiment with [3H]CP55940 or [3H]SR141716A, binding in the

absence and presence of unlabeled WIN55212-2 or AM251 was performed in triplicate

and test compounds were assayed in duplicate. A minimum of three independent

experiments were performed for every treatment. Protein measurements were

conducted according to Peterson (1977).

3.3.3. Investigation of phthalate interference with CB1 receptor agonist-stimulated [35S]GTPγS binding to the Gα-protein.

The method we used to isolate the mouse whole brain membrane fraction and

determine the effects of phthalates on agonist-stimulated [35S]GTPγS binding generally

followed the procedure published by Breivogel et al. (2000). Whole brains were quickly

removed from two mice and homogenized for 15 sec in 10 ml of ice-cold isolation buffer

94

(Trisma base (50 mM), MgCl2.6H20 (3 mM), EGTA (0.2 mM), NaCl (100 mM) with pH

adjusted to 7.4) using a tissue fragmenter (Polytron Kinematica GmBH; setting 6). The

suspension was centrifuged in a Beckman J2HS centrifuge (24,000 x g for 25 min at 2 oC) and the pellet was then resuspended in fresh ice-cold isolation buffer and re-

centrifuged. The washed membrane pellet was thoroughly dispersed in isolation buffer,

then protein concentration was adjusted to approx. 7 mg/ml before aliquots were

transferred to a -80 oC freezer. Prior to experimentation, the membrane fractions were

thawed on ice and completely dispersed as described in the previous section. This

procedure helped improve the reproducibility between replicates without obvious loss in

agonist-stimulated [35S]radioligand binding. Binding experiments were performed using

guanosine 5'-O-(g-γ35S]thio)-triphosphate ([35S]GTPγS) of sp. act. 1250 Ci/mmol

purchased from Perkin Elmer Life and Analytical Sciences, Canada.

The phthalate esters (dissolved in DMSO; 5 µl) or DMSO control (as required)

were placed in borosilicate glass tubes (13 x 100 mm; siliconized 24 h prior to assay with

Sigmacote [Sigma-Aldrich Canada]) and then 500 µl of isolation buffer (pH 7.4) was

added which contained fatty-acid free bovine serum albumin (1 mg/ml), guanosine

diphosphate (GDP; 100 µM), dithiothreitol (20 µM), [35S]GTPγS (0.14 nM final

concentration) and adenosine deaminase (0.004 units/ml). The membrane fraction (70.1

+ 4.2 µg protein/assay) was then added and after thorough vortexing, initial 15 min

incubation was carried out at room temperature. The CB1 receptor agonist CP55940

(100 nM final concentration; in 5 µl DMSO) or DMSO solvent control were then added

and, after thorough mixing, incubations were continued for 1.5 h at 30 0C with gentle

shaking. The final concentration of DMSO was 1.9%. Incubations were terminated by

introduction of 2 ml of ice-cold wash buffer (50 mM Trisma base:HCl; pH 7.4) which was

immediately followed by rapid vacuum filtration through pre-soaked Whatman GF/B

filters. Membranes trapped on filters were subjected to three 4 ml washes with the same

buffer. Membrane-bound 35S was measured using liquid scintillation counting. All assays

were conducted in triplicate. Specific binding of [35S]GTPγS (90.7 + 1.4%) was

calculated by subtracting [35S]GTPγS bound in the presence of 100 µM unlabelled

GTPγS from total binding. 100 nM CP55940 stimulated the basal specific [35S]GTPγS

binding signal by 57.7 + 0.6% and the effect of phthalates on this signal was

investigated.

95

3.3.4. Data analysis

Values are given as mean ± S.E.M. All values of IC50 (concentration of phthalate

ester producing 50% inhibition) were estimated from the concentration:response

relationships defined by non-linear regression analysis using Prism 4 (GraphPad

Software Inc., San Diego, CA, USA). Other non-linear and linear regression analyses

were likewise carried out with Prism 4.

3.4. Results

3.4.1. Effects of phthalate esters on binding of [3H]CP55940 to CB1 receptors.

The effects of the di- and mono-esters on the binding of [3H]CP55940 to CB1

receptors in mouse whole brain membranes are shown in Figures 3.2 and 3.3. Apart

from MnBP, all compounds produced concentration-dependent inhibition of [3H]CP55940

binding. Within the diester series, nBBP and DnHP were the most potent as indicated by

IC50s of 27.4 µM (95% CI = 20.7-36.5 µM) and 33.9 µM (95% CI = 26.5-38.5 µM)

respectively. DnBP, DEHP and DiOP were of intermediate potency (IC50s of 45.9 µM

(95% CI = 35.9-58.6 µM), 47.4 µM (95% CI = 41.5-54.1 µM) and 55.4 µM (95% CI =

45.8-67.0 µM) respectively), while DnOP was the weakest (IC50: 75.2 µM (95% CI =

65.9-87.2 µM). Based on the level of inhibition at 150 µM (the maximum concentration

employed), BBP, DnHP and DnBP were the most effacacious (85-100% inhibition),

followed by DEHP and DiOP (60-70% inhibition), while DnOP displayed lower efficacy

(50-60% inhibition). At 150 µM MiHP and M2EHP achieved less than 50% inhibition of

[3H]CP55940 binding and MnBP was inactive. In a separate series of experiments,

PMSF failed to modify the inhibitory effects of nBBP and DnBP on [3H]CP55940 binding

(Table 3.1).

3.4.2. Effects of selected phthalate esters on binding of [3H]SR141716A to CB1 receptors.

Table 3.2 shows the inhibitory effects of selected di- and mono-esters on the

specific binding of [3H]SR141716A to CB1 receptors of mouse brain. In the case of the

diesters nBBP, DBP and DEHP, the extent of inhibition of [3H]SR141716A binding at

96

concentrations that achieve 50% inhibition of [3H]CP55940 binding resulted in inhibitory

effects that were approximately 35-45% higher (nBBP, DnBP) and 25% lower (DEHP).

The monoester MnBP (100 µM) had no effect on specific binding of [3H]SR141716A

while MiHP produced circa. 33% inhibition.

3.4.3. Influence of selected phthalates on the saturation binding of [3H]SR141716A to CB1 receptors

The control saturation binding curve was constructed by measuring the specific

binding of [3H]SR141716A to CB1 receptors at equilibrium over a range of radioligand

concentrations (0.032 to 2.8 nM). Experiments were concurrently performed in the

presence of nBBP or DnBP (Figure 3.4). Analyses revealed that phthalates have

negligible effect on the Kd of radioligand binding but reduce the Bmax by 37% (nBBP) and

60% (DnBP).

3.4.4. Effects of selected phthalates on [3H]SR141716A kinetics

The association time course of [3H]SR141716A between 0 and 3 min in the

absence of test compounds (Figures 3.5a and 3.5b) aligns with data published by

Rinaldi-Carmona et al., 1996 using synaptosomes. nBBP (35 µM) and DnBP (50 µM)

reduce the ability of [3H]SR141716A to equilibrate with CB1 receptors both when applied

in advance of the [3H]SR141716A (Figure 3.5a) and to a lesser extent when introduced

simultaneously with radioligand (Figure 3.5b). When combined with a saturating

concentration of AM 251, nBBP (35 µM) and DnBP (50 µM) increased the dissociation of

the [3H]SR141716A: CB1 receptor complex to levels much greater than that produced by

a saturating concentration of AM251 alone (Figure 3.6).

3.4.5. Effects of phthalates on CB1 receptor agonist-stimulated [35S]GTPγS binding to the Gα-protein

Both diesters and monoesters also had the capacity to inhibit CB1 receptor

agonist-activated binding of [35S]GTPγS to the Gα-protein and in agreement with

[3H]CP55940 binding data, the diesters were consistently more active (Figure 3.7).

Inhibitory effects of the study compounds on [3H]CP55940 binding and CP55940-

97

stimulated binding of [35S]GTPγS to the Ga-protein were closely associated (r2 = 0.7844;

Figure 3.8).

3.5. Discussion

The present investigation demonstrates that certain phthalate esters interfere

with the binding of [3H]CP55940 and [3H]SR141716A to CB1 receptors of mouse brain at

micromolar concentrations in vitro. Since we found that CB1 receptor agonist-stimulated

[35S]GTPγS binding is also decreased by phthalate esters, these compounds appear to

inhibit activation of the associated G-protein receptors by operating as low affinity CB1

receptor antagonists.

In the [3H]CP55940 binding assay, the IC50 values of the phthalate esters nBBP,

DnHP, DnBP and DEHP lie in the 27-47 µM range, which put them at similar potency to

cis-9,10-octadecenyl-a-methylethanolamide an analog of the sleep-inducing lipid cis-

oleamide (Boring et al., 1996), of higher potency than thujone, cis-oleamide and cis-

9,10-octadecenylethanolamide (Boring et al, 1996; Meschler and Howlett, 1999), but of

lower potency compared to the antagonists sanguinarine, chelerythrine, piperonyl

butoxide and (S)-methoprene (Dhopeshwarkar et al., 2011). In distinct contrast to

thujone and cis-oleamide and its analogs (Boring et al., 1996; Meschler and Howlett,

1999), the phthalate esters were able to inhibit CB1 receptor agonist activation of the G-

protein. It must be emphasized that nBBP, DnHP, DnBP and DEHP are obviously much

weaker as inhibitors of [3H]CP55940 binding, [3H]SR141716A binding and CB1 receptor

agonist-stimulated [35S]GTPγS binding when compared to the antagonists SR141716A

and AM251 which exert their effects in the low nanomolar range (Rinaldi-Carmona et al.,

1995; Rinaldi-Carmona et al., 1996; Gatley et al., 1997).

Within the group of phthalate esters examined in the present investigation, a

broad range of CB1 receptor inhibitory effects were demonstrated. Our data indicate that

the diesters are more potent inhibitors of [3H]CP55940 binding than the monoesters. A

similar differential was also noted by Liu et al., (2009) in studies on the inhibitory effects

of phthalates on Ca++ transients triggered by nAChR activation in human neuroblastoma

cells. In contrast, the potency of the monoester MEHP as an inhibitor of follicle -

98

stimulating hormone binding to Sertoli cells was reported to be at least three orders of

magnitude higher than for the diester DEHP; the latter phthalate showing no activity at

100 µM (Grasso et al., 1993).

Our results suggest that the overall relationship between phthalate diester

structure and inhibitory effects on [3H]CP55940 binding to CB1 receptors is complex.

nBBP and DnHP showed the highest inhibitory potencies (IC50s = 27.4 and 33.9 µM

respectively) combined with robust efficacy (85% or greater inhibition at maximum

concentration). Reducing the length of each n-hexyl group of DnHP by 2 carbons (i.e.

giving DnBP), reduces inhibitory potency but efficacy is retained at circa. 85%. By

contrast, DEHP, which can be considered as a bis 2-ethyl analog of DnHP or a bis 2-

propyl analog of DnBP, demonstrates reduced potency and efficacy against DnHP, and

similar potency with reduced efficacy compared to DnBP. The phthalate diesters with the

longer alkyl substituents (DiOP and DnOP) exhibited lower inhibitory potencies (IC50s

55.4 and 75.2) and efficacies were also comparatively low (55-65%). The critical nature

of the diester configuration for inhibition of [3H]CP55940 binding is emphasized by

comparison of nBBP and DnBP (which are amongst the most effective compounds

studied) with MnBP (a phthalate devoid of activity).

For the experiments with PMSF, we reasoned that using phthalates of

intermediate (DnBP) and higher (nBBP) potency at < IC50 would offer a sensitive basis

for assessment. Moreover, DnBP and nBBP (study compounds with alkyl and aryl

substituents respectively) might be expected to show different susceptibilities to

breakdown by serine hydrolases. Nonetheless, based on our experiments, there was no

evidence that serine hydrolases limit the inhibitory effect of either of these analogs in the

[3H]CP-55940 binding assay.

The equilibrium binding and dissociation data using [3H]SR141716A provide a

useful insight into the mechanism by which nBBP and DnBP inhibit radioligand binding.

The saturation isotherms demonstrate that these phthalates act by eliminating binding

sites for radioligand (i.e. Bmax is reduced), without affecting the affinity of radioligand for

the remaining sites (i.e. Kd is unchanged). Moreover, the dissociation experiments

strongly suggest that these compounds act allosterically with respect to the

[3H]SR141716A binding site, since under our assay conditions any access by phthalate

99

diesters to the radioligand binding site is completely prevented by the saturating levels of

AM251. The dissociation data also argue against an irreversible or tight binding of

phthalate esters to the [3H]SR141716A recognition site, another potential explanation of

the reduced Bmax and unchanged Kd. The time courses for dissociation of

[3H]SR141716A in the presence of nBBP and DnBP indicate that the binding of

phthalates to this allosteric binding site and subsequent negative modulation of

radioligand binding occurs very rapidly. Rapid engagement of phthalates with a site

coupled allosterically to the [3H]SR141716A binding site is also consistent with the

reduced levels of nBBP and DnBP binding in the association experiments. However, the

association profiles in the presence of nBBP and DnBP are likely markedly influenced by

the effect of these compounds on availability of receptors (Bmax) that can bind

[3H]SR141716A. Overall, our results indicate that a critical mechanism underlying

inhibition of [3H]SR141716A to CB1 receptors involves phthalates engaging with a site

that is distinct from but negatively coupled to the radioligand recognition site. The

proposed binding region for phthalate esters on the CB1 receptor may represent a novel

target that could be exploited therapeutically by phthalate ester analogs or other drugs to

produce downregulation of endocannabinoid action in the brain.

Unlike phthalate diesters, MEHP and other monoesters inhibit the binding of

follicle stimulating hormone (FSH) to G-protein coupled FSH receptors, an action that

may involve direct engagement of MEHP with the G-protein (Grasso et al., 1993). The

allosteric inhibition of [3H]SR141716A binding to the CB1 receptor by phthalate diesters

could also arise from a direct interaction with its G-protein as we have postulated for

chelerythrine and sanguinarine (Dhopeshwarkar et al. 2011). However, in contrast to the

findings of Grasso et al., (1993), we found that monoesters are, at best, exceptionally

weak inhibitors of CB1 receptor radioligand binding. Therefore, negative allosteric

coupling between a phthalate diester recognition site on the CB1 receptor and the

radioligand binding site is likely a more productive area for future exploration.

Phthalate diesters have potential to access the brain, since a number of these

compounds interfere with barbiturate-induced sleep duration following systemic

administration (Calley et al., 1966) and phthalate ester exposure in school children has

been associated with a behavioral (attention-deficit/hyperactivity) disorder (Kim et al.

2009). The presynaptic CB1 receptor plays a fundamental role at many synapses in

100

mammalian brain and activation of this complex by endocannabinoids promotes a

variety of physiological and behavioral responses. Moreover, downregulation of CB1

receptors and other components of the endocannabinoid system in human epilepsy is

associated with increased excitability in neuronal networks and has been linked to

reduced seizure thresholds (Ludanyi et al., 2008). A critical question is whether brain

CB1 receptors are exposed to phthalate diesters in vivo at concentrations that are

sufficient to interfere with the activation of this signaling pathway by endocannabinoids.

Phthalate esters undergo extensive ester cleavage in the gastrointestinal tract and

hydrolysis would be expected to limit the ability of diesters to reach the brain particularly

after acute oral exposure. However, individuals receiving higher exposure to phthalate

esters on a continuous basis (perhaps as a result of occupational exposure) or hospital

patients exposed to phthalates released from medical devices may be more likely to

accumulate these chemicals in the brain. In the present investigation, threshold inhibitory

effects of DEHP, DnOP, DiOP and nBBP on [3H]CP55940 binding are evident between 1

and 10 µM. Even concentrations within this range in brain may be sufficient to

antagonize endocannabinoid-mediated signaling at CB1 receptors to an extent that

causes low level synaptic perturbations and subtle pathophysiological and affective

responses.

Finally, it must be stressed that further studies aimed at determining 1) phthalate

ester levels in brain following short term systemic and chronic exposures and 2) the

ability of these compounds to modify critical effects of cannabinoid agonists in intact

animals are essential to improve our understanding of the potential phthalate diesters

might have in modulating the endocannabinoid system in vivo.

3.6. Note in added proof

3.6.1. Background

Subsequent to our Neurochemistry International publication appearing, I

established the L-glutamate release assay originally described by Nicholls et al. (1987)

in our laboratory. This assay uses purified synaptosomes and has been utilized by

Wang (2003) to study inhibition of 4-aminopyridine- (4-AP-) evoked release of L-

101

glutamate by the CB1-R agonist WIN55212-2. Since inhibition by WIN55212-2 is

blocked by a diarylpyrazole CB1-R antagonist (Wang, 2003), this system provides a

rigorous functional test for compounds that might antagonize CB1-Rs in the brain.

3.6.2. Experimental approach

We selected two phthalates for investigation, BBP and MnBP. The former was

one of the more potent diesters, both for inhibition of [3H]CP55940 binding and inhibition

of CP55940-stimulated [35S]GTPγS binding. The latter, a monoester, gave no inhibition

of [3H]CP55940 binding and at best produced marginal (<20%) inhibition of [35S]GTPγS

binding. The methods for the isolation of synaptosomes and the fluorimetric

measurement of L-glutamate release are described in detail in Chapter 4.

3.6.3. Results

Consistent with the findings of Wang (2003), WIN55212-2 inhibited of 4-AP-

evoked release of L-glutamate and this effect was fully blocked by AM251. In agreement

with our [3H]CP55940 binding results and as predicted by the [35S]GTPγS binding data,

BBP at 30 µM (but not 5 µM) fully reversed the inhibition by WIN55212-2 (Figure 3.9).

Again in agreement with the [3H]CP55940 binding experiments and the [35S]GTPγS

assays, MnBP failed to modify the inhibitory effect of WIN55212-2 (Figure 3.10).

3.6.4. Conclusion

The L-glutamate assay results strongly supports the idea that 1) phthalate

diesters (as exemplified by BBP) exert antagonist actions at presynaptic CB1-Rs at low

to moderate micromolar concentrations in vitro, and 2) monoesters (as exemplified by

MnBP) are inactive.

102

3.7. Figures and Tables

O

O

O

O

O

O

O

O

O

O

O

O

(a) nBBP (b) DnHP (c) DnBP

O

O

O

O

O

O

O

O

O

O

O

O

(d) DEHP (e) DiOP (f) DnOP

O

O

O

O

H

O

O

O

O

H

O

O

O

O

H

(g) M2EHP (h) MiHP (i) MnBP

Figure 3.1 (a-f) The structures of phthalate diesters: n-butylbenzylphthalate (nBBP); di-n-hexylphthalate (DnHP); di-n-butylphthalate (DnBP); di-ethylhexylphthalate (DEHP); di-isooctylphthalate (DiOP) and di-n-octylphthalate (DnOP).(g-i) The structures of phthalate monoesters: mono-2-ethylhexyl-phthalate (M2EHP), mono-isohexyl-phthalate (MiHP) and mono-n-butyl-phthalate (MnBP). All structures have been redrawn from Bissett et al. (2011) using IsisDraw.

103

Figure 3.2 Inhibitory effects of phthalate esters (DnBP, nBBP, DnOP, MiHP and MnBP) on the binding of [3H]CP55940 to mouse brain CB1 receptors in vitro. Each point represents the mean + SEM of 3 independent experiments. Results provided by Ms Kathleen M. Bisset.

104

Figure 3.3 Inhibitory effects of phthalate esters (DEHP, DnHP, DiOP and M2EHP) on the binding of [3H]CP55940 to mouse brain CB1 receptors in vitro. Each point represents the mean + SEM of 3 independent experiments. Results provided by Ms Kathleen M. Bisset.

105

Figure 3.4 The effect of nBBP and DnBP (both at 35 µM) on the equilibrium binding of of [3H]SR141716A to CB1 receptors of mouse whole brain. Kd and Bmax values are displayed for each treatment and 95% confidence intervals were as follows: control (Kd 0.628 to 0.859. Bmax 0.303 to 0.343), nBBP (Kd 0.761 to 1.333. Bmax 0.176 to 0.229) and DnBP (Kd 0.624 to 0.846. Bmax 0.120 to 0.136). R2 values were 0.9877 (control), 0.9756 (nBBP) and 0.9887 (DnBP). Data points represent the means + SEMs of 3 independent experiments (most SEM bars are obscured by data symbols).

106

Figure 3.5a Influence of nBBP (35 µM) and DnBP (50 µM) on the time course of association of [3H]SR141716A with CB1 receptors of mouse brain. (a) membranes received the standard 15 min preincubation with phthalate esters prior to [3H]SR141716A addition.

107

Figure 3.5b Influence of nBBP (35 µM) and DnBP (50 µM) on the time course of association of [3H]SR141716A with CB1 receptors of mouse brain. (b) The phthalate ester and [3H]SR141716A were applied simultaneously. Data points represent the means + SEMs of 3 independent experiments (most SEM bars are obscured by data symbols).

108

Figure 3.6 Dissociation of the [3H]SR141716A:CB1 receptor complex (initiated by challenge with 5 µM AM251) in the absence (control) or in the presence of 35 µM nBBP or 50 µM DnBP. Data represent mean + SEM of at least 3 independent experiments, each performed in triplicate.

109

Figure 3.7 Inhibition of CP55940-stimulated binding of [35S]GTPγS to mouse whole brain membranes by phthalate esters. Phthalate esters were assayed at 75 µM throughout. Each column represents the mean, and error bar the SEM of 7 independent experiments.

110

Figure 3.8 Relationship between the ability of study compounds to inhibit the binding of [3H]CP55940 and CP55940-stimulated binding of [35S]GTPγS in mouse whole brain membrane fractions. All assays were performed 75 µM; r2 = 0.7844.

111

Figure 3.9 With WIN55212-2 present, BBP (at 30 µM but not 5 µM) enhances 4-AP-evoked L-glutamate release above the level produced by 4-AP alone.

112

Figure 3.10 With WIN55212-2 present, MnBP (both at 30 µM and 5 µM) does not enhance 4-AP-evoked L-glutamate release above the level produced by 4-AP alone.

113

Table 3.1 Inability of PMSF to influence the inhibitory effects of n-butylbenzylphthalate (nBBP) and di-n-butylphthalate (DnBP) on [3H]CP55940 binding to mouse brain membranes. Phthalate esters were present in the assay at 20 µM and PMSF was used at 50 µM. Each value represents the mean + S.E.M. of 3-6 independent experiments.

Treatment Inhibition (%)

PMSF -2.82 + 3.18

nBBP 28.25 + 2.11

nBBP + PMSF 27.01 + 4.55

DnBP 20.50 + 1.40

DnBP + PMSF 22.61 + 3.31

114

Table 3.2 Inhibitory effects of n-butylbenzylphthalate (nBBP), di-n-butylphthalate (DnBP), diethylhexylphthalate (DEHP), mono-isohexylphthalate (MiHP) and mono-n-butyl phthalate (MnBP) on the specific binding of [3H]SR141716A to mouse brain membranes. Diesters were present at concentrations producing 50% inhibition of [3H]CP55940 binding. Each value represents the mean + S.E.M. of 3 independent experiments.

Treatment Inhibition of specific binding (%)

nBBP (27 µM) 67.82 + 1.71

DnBP (46 µM) 72.30 + 3.23

DEHP (47 µM) 37.42 + 3.48

MiHP (100 µM) 33.23 + 4.15

MnBP (100 µM) 0

115

4. Benzophenanthridine alkaloid, piperonyl butoxide and (S)-methoprene action at the cannabinoid-1 receptor (CB1-R) pathway of mouse brain: interference with [3H]CP55940 and [3H]SR141716A binding and modification of WIN55212-2-dependent inhibition of synaptosomal L-glutamate release.

4.1. Abstract

Benzophenanthridine alkaloids (chelerythrine and sanguinarine) inhibited binding

of [3H]SR141716A to mouse brain membranes (IC50s: <1 µM). Piperonyl butoxide and

(S)-methoprene were less potent (IC50s: 21 and 63 µM respectively).

Benzophenanthridines and piperonyl butoxide were more selective towards brain CB1-

Rs versus spleen CB2-Rs.

All compounds reduced Bmax of [3H]SR141716A binding to CB1-Rs, but only

methoprene and piperonyl butoxide increased Kd (3-5-fold). Benzophenanthridines

increased the Kd of [3H]CP55940 binding (6-fold), but did not alter Bmax. (S)-methoprene

increased the Kd of [3H]CP55940 binding (by almost 4-fold) and reduced Bmax by 60%.

Piperonyl butoxide lowered the Bmax of [3H]CP55940 binding by 50%, but did not

influence Kd.

All compounds reduced [3H]SR141716A and [3H]CP55940 association with CB1-

Rs. Combined with a saturating concentration of SR141716A, only piperonyl butoxide

and (S)-methoprene increased dissociation of [3H]SR141716A above that of SR141716A

alone. Only piperonyl butoxide increased dissociation of [3H]CP55940 to a level greater

than CP55940 alone. Binding results indicate predominantly allosteric components to

the study compounds action.

116

4-Aminopyridine- (4-AP-) evoked release of L-glutamate from synaptosomes was

partially inhibited by WIN55212-2, an effect completely neutralized by AM251, (S)-

methoprene and piperonyl butoxide. With WIN55212-2 present, benzophenanthridines

enhanced 4-AP-evoked L-glutamate release above 4-AP alone. Modulatory patterns of

L-glutamate release (with WIN-55212-2 present) align with previous antagonist/inverse

agonist profiling based on [35S]GTPγS binding. Although these compounds exhibit lower

potencies compared to many classical CB1 receptor inhibitors, they may modify CB1-R-

dependent behavioral/physiological outcomes in the whole animal and could offer

templates for synthesis of novel and more potent CB1-R blocking drugs.

Note: The research described in this chapter will be submitted shortly for

publication in an appropriate neurochemical/neuropharmacological journal. The

submission will adhere closely to the format laid out here.

4.2. Introduction

Cannabinoid-1 receptors (CB1-Rs) are present in numerous regions of

mammalian brain and are particularly abundant within the cerebral cortex, hippocampus,

cerebellum and basal ganglia (Herkenham et al., 1991; Tsou et al., 1998). CB1-Rs

couple to G-proteins in the plasma membrane of nerve terminals and together they

constitute the primary presynaptic element of the endocannabinoid signaling pathway

that regulates transmitter release through negative feedback (Howlett et al., 1986;

Katona et al., 1999; Kawamura et al., 2006). Endocannabinoids, generated in

postsynaptic neuronal cell bodies when synaptic activity intensifies, migrate retrogradely

and bind to presynaptic CB1-Rs. G-protein activation leads to inhibition of voltage-

sensitive Ca++ channels (Mackie and Hille, 1992; Twichell et al., 1997, Kushmerick et al.,

2004; Guo and Ikeda, 2004), negative modulation of adenylate cyclase (Howlett and

Fleming, 1984; Howlett, 1985) and activation of K+ currents (Deadwyler et al., 1993;

Mackie et al., 1995, Childers and Deadwyler, 1996; Guo and Ikeda, 2004). Since these

various signaling mechanisms reduce the ability of action potentials impinging on the

nerve ending to depolarize and activate calcium entry, transmitter release is adjusted

downwards, thus completing the negative feedback loop (Chevaleyre et al., 2006;

117

Kreitzer and Regehr, 2001; Wilson and Nicoll, 2001; Howlett, et al., 2002; Freund et al.,

2003).

Certain plant natural products and synthetic drugs mimic endocannabinoid

activation of this signaling pathway by exerting potent (nanomolar) agonist actions at

CB1-Rs. Prominent xenocannabinoid agonists include ∆9-tetrahydrocannabinol, the main

psychoactive principle of Cannabis sativa (Razdan, 1986), CP55940 (Johnson and

Melvin, 1986) and the aminoalkylindole WIN55212-2 (Compton et al., 1992). Selective

high potency CB1-R antagonists, notably the phytocannabinoid ∆9-

tetrahydrocannabivarin and the diarylpyrazole antagonist/inverse agonists AM251 and

SR141716A, have also been reported (Rinaldi-Carmona et al., 1994; Lan et al., 1999;

Thomas et al., 2005).

There is considerable interest in possible therapeutic applications of CB1-R

modulators. Agonists and allosteric activators of agonist action have been considered in

the relief of pain, muscle spasms, anxiety states and depressive illness, and they can

also block emesis, improve sleep and stimulate appetite (Van Sickle et al., 2001;

Iversen, 2003, Ligresti et al., 2009; Bradshaw and Walker, 2005; Di Marzo, 2009). On

the other hand, CB1-R antagonists/inverse agonists such as the diarylpyrazole

rimonabant (SR141716A) have shown effectiveness in reducing body weight through

suppression of appetite (Colombo et al., 1998), but rimonabant use in human medicine

was curtailed due to adverse psychiatric side effects. Nevertheless, discovery of a CB1-

R inhibitor divorced of such unfavourable symptoms clearly remains of considerable

interest (Szabo et al., 2009; Wu et al., 2009; Riedel et al., 2009).

Research in our laboratory has focused on other natural products and synthetic

environmental chemicals capable of interacting with the endocannabinoid system.

Specifically we have demonstrated that at very low to moderate micromolar

concentrations, the benzophenanthridine alkaloids (sanguinarine and chelerythrine), the

pesticides (piperonyl butoxide and (S)-methoprene), certain phthalate dialkyl ester

plasticizers, as well as the more acutely toxic tributyltin derivatives (tributyltin acetate

and tributylethynyl tin) inhibit both the binding of [3H]CP55940 to CB1-Rs, as well as CB1-

R agonist-dependent activation of the G-protein (Dhopeshwarkar et al., 2011; Bisset et

118

al., 2011; Jain S., M.Sc. Thesis, Simon Fraser University, 2011; Dhopeshwarkar A.S.,

Ph.D. Thesis, Simon Fraser University, 2012, Chapter 5).

Endocannabinoids, the high potency synthetic agonists (e.g. CP55940 and

WIN55212-2) and dihydropyrazole antagonists/inverse agonists (e.g. AM251 and

SR141716A) engage with a discrete binding pocket on the CB1-R and available

evidence suggests individual binding domains are distinct or may have tendency to

partially overlap (Shim, 2010; Kapur et al., 2007). The purpose of this phase of our

research program was to examine the effects of sanguinarine, chelerythrine, piperonyl

butoxide and (S)-methoprene in greater detail with regards their ability to interfere with

the equilibrium binding and kinetic properties of radioligands that engage with the

binding pocket of CB1-Rs (specifically [3H]55940 and [3H]SR141716A). Moreover, since

we predicted that, based on their abilities to modulate basal and CB1-R-agonist-

(CP55940-) stimulated binding of [35S]GTPγS to the G protein, the benzophenanthridines

were likely inverse agonists and piperonyl butoxide and (S)-methoprene were neutral

antagonists of CB1-Rs (Dhopeshwarkar et al., 2011), it was critical to investigate whether

modification of neurotransmitter release (i.e. the ultimate consequence of presynaptic

CB1-R engagement by these compounds) was consistent with this functional profiling.

4.3. Materials and Methods

4.3.1. Chemicals and supplies

2-[(1R,2R,5R)-5-hydroxy-2-(3-hydroxypropyl)-cyclohexyl]-5-(2-methyloctan-2-yl)-

phenol (CP-55940),5-(4-iodophenyl)-1-(2,4-dichlorophenyl)-4-methyl-N-(piperidin-1-yl)-

1H-pyrazole-3-carboxamide (AM251), (R)-(+)-[2,3-dihydro-5-methyl-3-(4-morpholinyl-

methyl)pyrrolo[1,2,3-de]-1,4-benzoxazin-6-yl]-1-napthalenylmethanone mesylate

(WIN55212-2), dimethylsulfoxide (DMSO), sanguinarine, chelerythrine, piperonyl

butoxide, Ethylene diamine tetraacetic acid (EDTA), Ethylene glycol-bis(2-

aminoethylether)-N,N,N′,N′-tetraacetic acid (EGTA), 4-aminopyridine (4-AP), veratridine

(VTD), tetrodotoxin (TTX), acetone, Percoll®, glutamate dehydrogenase (Type II from

bovine liver), β-nicotinamide adenine dinucleotide phosphate (sodium salt hydrate;

NADP+), bovine serum albumin (BSA; fatty acid free) and all other chemicals required for

119

assay buffers and salines, were purchased from Sigma Aldrich, Canada. 5-(4-

chlorophenyl)-1-(2,4-dichlorophenyl)-4-methyl-N-(piperidin-1-yl)-1H-pyrazole-3-

carboxamide (SR141716A) was obtained from Cayman Chemical and (S)-methoprene

(98.5% purity) was supplied by Doug Vangundy, Director of Speciality Product

Development, Wellmark International (Dallas, Texas). Radioligands, [3H]CP55940 (side

chain-2,3,4-[3H]; specific activities 139.6 and 174.6 Ci/mmol) and [3H]SR141716A

(piperidine ring 3,4-[3H]; specific activity 56 Ci/mmol) were purchased from Perkin Elmer

Life and Analytical Sciences, Canada.

4.3.2. Animals

All experiments described in this report were performed with male CD1 mice (20-

25 g) purchased from Charles River Laboratories (Saint-Constant, Quebec). Animal

orders were placed by Animal Care Services of Simon Fraser University, Burnaby,

Canada. Upon receipt, mice were housed under standardized environmental conditions

(21 0C; 55 % relative humidity; 12 hour light/dark cycle) and allowed unlimited access to

food and water. Our animal housing, handling and experimental procedures conformed

to Canadian Council on Animal Care guidelines and were formally approved by the

Simon Fraser University Animal Care Committee prior to embarking on this investigation.

4.3.3. Isolation of membranes from mouse brain for binding studies

The isolation of brain membranes were carried out at 0-4 oC according to a

method published previously (Dhopeshwarkar et al., 2011). Mouse whole brains were

homogenized (10 pestle excursions; pestle rotation approx. 1500 rpm) in ice-cold buffer

[Trisma base (100 mM), EDTA (1 mM), adjusted to pH 9 with HCl; 1 brain per 10 ml

buffer]. The homogenate was centrifuged for 10 min at 900 x g in a JA20 rotor of a

Beckman J2HS centrifuge. Centrifugation of the supernatant at 11,500 x g for 25 min

produced a membrane pellet which was resuspended in ice-cold buffer [Trisma base (50

mM), EDTA (1 mM) and MgCl2.6H2O (3 mM); adjusted to pH 7.4 with HCl] at a protein

concentration of approx. 6.5 mg/ml. The membrane preparation was then frozen in

aliquots at -80 oC. Just prior to assay, the preparation was thawed on ice and carefully

120

dispersed by slowly moving the membrane suspension in and out of a syringe fitted with

a 18 gauge needle (6 times) followed by vortexing.

4.3.4. Effects of benzophenanthridines, (S)-methoprene and piperonyl butoxide on equilibrium binding of [3H]CP55940 and [3H]SR141716 to brain CB1 receptors

The saturation binding method of Steffens et al. (2004) was adopted (with

modifications as detailed below), and effects on radioligand saturation binding

characteristics were investigated by assaying the study compound (at ≥ IC50) with

different concentrations of [3H]CP55940 or [3H]SR141716A (0.032-3.5 nM). To achieve

these concentrations of [3H]CP55940 and [3H]SR141716A, radioligand specific activity

was reduced by addition of the required quantity of unlabelled CP55940 and

SR141716A. For assay, compounds formulated in DMSO (5 µl) or DMSO control (5 µl),

as appropriate, were rapidly injected into borosilicate glass tubes (13 x 100 mm; Kimble-

Chase) containing binding buffer [500 µl; Trisma base (50 mM), EDTA (1 mM),

MgCl2.6H2O (3 mM), BSA (fatty acid free; 3 mg/ml) adjusted to pH 7.4 with HCl]. After

addition of brain membranes (227.09 + 0.66 µg protein), the mixture was gently vortexed

and the incubation continued for 15 minutes at room temperature. [3H]CP55940 or

[3H]SR141716A was then added at the required concentrations and incubations were

run for 90 min to ensure equilibration of radioligand. Incubations were concluded by

addition of 4 ml ice-cold wash buffer (0.9% NaCl containing 2 mg/ml BSA (fatty acid

free)) and membranes were harvested on Whatman GF/C filters (presoaked with wash

buffer) using a (Hoefer FH 225V) vacuum filtration system attached to a vacuum pump

(Hoefer FH 225V). Membranes collected on filters were washed (3 x 4 ml) with ice-cold

wash buffer. Each filter was removed from its filtration well and placed in a scintillation

vial and allowed to dry completely. After drying, 4 ml of scintillation cocktail (BCS,

Amersham Biosciences, UK) was added and radioactivity measured using liquid

scintillation counting. Specific binding was calculated by subtracting non-specific binding

(binding in presence of 10 µM CP55940 or 10 µM SR141716A) from total binding and

this was determined for each concentration of [3H]CP55940/CP55940 or

[3H]SR141716A/SR141716A in the absence and presence of study compound. Values

were used to construct equilibrium binding isotherms which allowed calculation of the Kd

(equilibrium dissociation constant) and Bmax (number of receptors available for

121

radioligand binding). At least three experiments were conducted for each treatment and

protein measurement was done as described by Peterson (1977).

4.3.5. Effect of benzophenanthridines, (S)-methoprene and piperonyl butoxide on the association and dissociation kinetics of [3H]CP55940 and [3H]SR141716A

Association studies were initiated by addition of study compound in DMSO (5 µl)

or DMSO control (5 µl), as appropriate, to 500 µl assay buffer in borosilicate glass tubes.

Membranes (approx. 230 µg protein) were then added and the system allowed to

incubate for 15 min at room temperature. Meanwhile, the Hoefer FH 225V filtration

apparatus was prepared by inserting pre-soaked Whatman GF/C filters and allowing the

vacuum to draw down a few minutes before the end of the preincubation. [3H]CP55940

or [3H]SR141716A (1 nM final concentration) was added at t = 15 min, and the brain

membrane suspensions filtered at various time points between 0 and 180 secs. After

three 4 ml washes, filters were dried and radioactivity associated with the membranes

quantified. The time course of radioligand association was also tracked after each study

compound (or DMSO control) was added a few seconds before the radioligand (defined

as co-treatment situation).

Dissociation studies were conducted by equilibrating brain membranes (approx.

230 µg protein) with [3H]CP55940 or [3H]SR141716A (1 nM final concentration) for 90

minutes at 37 oC with gentle shaking. At equilibrium, either a saturating (5 µM)

concentration of (CP55940) or AM251 (added in 10 µl DMSO), or study compound at ≥

IC50 (added in 5 µl DMSO) plus 5 µM (CP55940) or AM251 (added in 5 µl DMSO) was

added, Dissociation of radioligand was monitored over 300 seconds.

4.3.6. Interaction of benzophenanthridines, methoprene and piperonyl butoxide with CB2 receptors in mouse spleen

After evaluation of several methods, the method of Hillard et al. (1999) was

adopted for this investigation. Two mice were euthanized by rapid cervical dislocation

and the spleens were rapidly removed and homogenized in 10 ml TME buffer (Tris-HCl

(50 mM), EDTA (1 mM) and MgCl2.6H2O (3 mM), titrated to pH 7.4 with HCl) using a

pre-chilled motor driven homogenizer (10 strokes up and down, pestle rotation 1500

122

rpm). The homogenate was then centrifuged at 500 x g for 10 minutes in a Beckman

J2HS centrifuge using JA-20 rotor. The pellet was discarded and the supernatant

centrifuged at 17,500 x g for 30 minutes. The fresh spleen membranes were

immediately used for experimentation. For assay, 500 µl of binding buffer (Tris-HCl (50

mM), EDTA (1 mM) and MgCl2. 6H2O (3 mM) and BSA (fatty acid free, 3 mg/ml), pH 7.4

with HCl) was transferred to borosilicate glass tubes (13 x 100 mm; Kimble-Chase).

Study compounds (in 5 µl DMSO) or DMSO controls, as appropriate, were then

introduced followed by spleen membranes (200.20 ± 2.07 µg of protein per tube). The

mixture was gently vortexed and incubated for 15 min at room temperature, whereupon

[3H]CP55940 (1 nM final concentration; added in 10 µl in DMSO) was injected, the tube

contents thoroughly mixed, and a 90 minute incubation at 30 oC with gentle shaking

carried out. The binding reactions were terminated by addition of ice-cold wash buffer

(0.9% NaCl containing 2 mg/ml BSA; 1 ml) and membranes were collected on pre-

soaked Whatman GF/C filter papers using vacuum filtration (Hoefer FH 225V).

Harvested membranes were then washed with three washes of 4 ml ice-cold wash

buffer. Scintillant (4 ml, BCS, Amersham Bioscience UK) was added after thorough

drying of the filter and radioactivity measured using liquid scintillation counting. Non-

specific binding (measured in the presence of 10 µM CP55940) was subtracted from

total binding to yield specific binding to CB2 receptors and this averaged 75.01 ± 1.4%.

All assays were performed in triplicate and a minimum of three experiments were

conducted for every treatment.

4.3.7. Preparation of synaptosomes from mouse whole brain

Synaptosomes (pinched-off nerve endings) were prepared and isolated from the

whole brains of male CD1 mice using the discontinuous isoosmotic gradient technique of

Dunkley et al. (2008). On the day of experiment, Percoll gradients were prepared in two

12 ml polycarbonate centrifuge tubes by carefully layering 2 ml of each Percoll gradient

solution, starting with 23% at the bottom followed by 15%, 10% and finally 3% at the top.

A flow rate of less than 2 ml/min was maintained while layering so that Schlieren lines

remained clearly visible at the interface between Percoll layers. Gradients were then

transferred onto ice and stored in the cold room until needed. All centrifuge tubes,

glassware and buffers were pre-chilled on ice before start of experiment and all

processing steps were performed at 0-40C. Two mice were sacrificed by cervical

123

dislocation and whole brains removed within 15 seconds. The brains were then

thoroughly rinsed with 5 ml of ice-cold homogenizing buffer [sucrose (0.32 M), EDTA (1

mM) and Trisma base (5 mM); with pH adjusted to 7.4 using HCl]. Next, brain tissue

was transferred to a specially constructed glass/plexiglass tissue homogenizer

containing 15 ml of homogenization buffer and homogenized (7 strokes up and down;

pestle rotation 700 rpm). The homogenate was immediately centrifuged in a Beckman

J2HS centrifuge using JA-20 rotor for 10 minutes (1000 x g). The supernatant obtained

was centrifuged at 15000 x g for 30 minutes and the pellet was then gently homogenized

(2 ml glass homogenizer; Uniform, England). The homogenate was then diluted to 4 ml

with homogenizing buffer and then 2 ml carefully layered on top of the 3% Percoll layer

in each polycarbonate tube. Tube contents were then centrifuged at 31,000 x g (at 4 0C)

for 5 minutes (excluding acceleration and deceleration time). Tubes were carefully then

transferred to ice and synaptosome-rich fractions at the 10-15% and 15-23% interfaces

were pooled in a chilled 100 ml beaker. Synaptosomal fractions were then slowly diluted

with 80 ml assay buffer [NaCl (130 mM), KCl (4.5 mM), NaHCO3 (5 mM), MgCl2.6H2O (1

mM), Na2HPO4 (1.2 mM), HEPES (10 mM), glucose (10 mM), BSA (1 mg/ml) with pH

adjusted to 7.4 with NaOH]. This dilution was performed over a 30 minute period

whereupon the suspension was centrifuged at 20,000 x g (at 4 0C) for 30 minutes. The

pellets containing purified synaptosomes were gently resuspended in 1 ml assay buffer

and 100 µl aliquots of pure synaptosomes [0.767 ± 0.11 µg of protein per aliquot as

determined by Peterson (1977)] were apportioned to 9 snap top vials which were held on

ice in the cold room until required.

4.3.8. Release of L-Glutamate from synaptosomes

The enzyme-linked fluorescence technique originally described by Nicholls et al.

(1987) was adopted with minor modifications to measure the release of endogenous

glutamate from mouse brain synaptosomes. Briefly, 100 µl synaptosomes were added

to 2 ml of ice-cold assay buffer and the suspension incubated in shaking water bath at

37 0C for 15 mins. The suspension was then transferred to stirred quartz cuvette

thermostated at 37 0C in a Perkin Elmer LS 50 spectrophotometer. NADP+ (1 mM),

glutamic dehydrogenase (100 units) and CaCl2 (as appropriate; 1.3 mM) were added

followed by the cannabinoid agonist (WIN55212-2) and/or antagonist (AM251) along

with study compounds, inhibitors, EGTA or solvent controls, as appropriate, and the

124

system allowed to incubate for another 5 minutes. Fluorescence recording was started

and 4-AP (3 mM) was added at 100 secs to activate glutamate release from

synaptosomes and the experiment was terminated at 500 secs. The excitation

wavelength was set at 360 nm (slit width 5) and the emission was sampled at 460 nm

(slit width 5). Fluorescence output was measured at 1 sec intervals. The amount of

glutamate released was monitored as an increase in fluorescence due to NADPH

forming from NADP+ as a result of the oxidative deamination of released glutamate by

glutamate dehydrogenase. Standard glutamate was added to allow quantitation of the

released glutamate as nmol glutamate/mg synaptosomal protein.

All compound additions were done with microsyringes (Hamilton, USA).

4.3.9. Analysis of radioligand binding data and glutamate release data

Curve fitting and calculation of binding parameters was performed using Prism,

GraphPad Software Inc., San Diego, CA, USA). In the glutamate release assays,

fluorescence changes were measured with the vertical analysis tool of the Perkin Elmer

LS-50 software and the percentage change to the 4-AP-induced fluorescence signal

above the control (assay buffer added at 100 secs instead of 4-AP) was calculated.

Time resolved fluorescence data were also transferred to GraphPad Prism (5.0) and full

fluorescence traces constructed. Percentage changes induced by standard

pharmacological agents and study compounds were calculated as mean ± S.E.M. based

on 3-4 independent experiments.

4.4. Results

4.4.1. Effects of benzophenanthridines, piperonyl butoxide and (S)-methoprene on binding of [3H]SR141716A to CB1 receptors

Figure 4.1 shows the inhibitory effects of benzophenanthridines, piperonyl

butoxide and (S)-methoprene on specific binding of [3H]SR141716A to CB1 receptors in

the mouse brain membrane preparation. The benzophenanthridines, sanguinarine and

chelerythrine were the most potent inhibitors based on IC50 values of 732 nM (95%

confidence interval (CI) = 364-1470 nM) and 911 nM (95% CI = 713-1164 nM)

125

respectively. Piperonyl butoxide and methoprene had weaker effects on [3H]SR141716A

binding, achieving IC50s of 21.1 µM (95% CI 17.4-25.7) and 62.8 µM (95% CI 48.4-81.5)

respectively. At maximum effect concentrations, the benzophenanthridines produced

>90% inhibition, in contrast to piperonyl butoxide (70-80%) and methoprene (approx.

50%).

4.4.2. Influence of study compounds on the saturation binding of [3H]SR141716A to CB1 receptors of mouse brain

The control saturation binding curve was constructed by measuring the specific

binding of [3H]SR141716A to CB1 receptors at equilibrium over a range of radioligand

concentrations (0.032 to 2.8 nM). Control experiments were always performed

concurrently with benzophenanthridines, piperonyl butoxide or (S)-methoprene assayed

at ≥ IC50 (Figure 4.2). All study compounds reduced the apparent Bmax, but only

methoprene and piperonyl butoxide increased the Kd of [3H]SR141716A binding (by

approx. 3- and 5-fold respectively). Analogous equilibrium binding experiments were

conducted with [3H]CP55940 (Figure 4.3) and differences between benzopenanthridines

and synthetic compounds were again indicated. The benzophenanthridines increased

the Kd of [3H]CP55940 binding by approximately 6-fold, but did not alter Bmax. In

contrast, methoprene increased the Kd of [3H]CP55940 binding by almost 4-fold and

reduced Bmax by approx. 60%. Piperonyl butoxide reduced the Bmax of [3H]CP55940

binding by close to 50%, but had no influence on the Kd.

4.4.3. Effects of sanguinarine, chelerythrine, piperonyl butoxide, and (S)-methoprene on the kinetics of CB1 receptor-selective radioligand binding

The effects of sanguinarine, chelerythrine, piperonyl butoxide, and (S)-

methoprene on the association of [3H]SR141716A with CB1 receptors over the initial

phase of the binding reaction are displayed in Figure 4.4. All compounds reduced the

ability of [3H]SR141716A to progressively bind to CB1 receptors, when applied both

before radioligand (Figure 4.4a) and together with radioligand (Figure 4.4b). In similar

experiments, all study compounds reduced the association of [3H]CP55940 (Figure

4.4c). When added together with a saturating concentration of SR141716A, piperonyl

butoxide (30 µM) and (S)-methoprene (60 µM) increased the dissociation of the

126

[3H]SR141716A:CB1 receptor complex to a rate greater that induced by the saturating (5

µM) concentration of SR141716A alone (Figure 4.5a). Sanguinarine and chelerythrine

(at 5 µM) failed to increase SR141716A-induced dissociation under these conditions

(data not shown). However, when applied alone, all study compounds initiated

dissociation of [3H]SR141716A, (S)-methoprene being the least effective (Figure 4.5b).

In the presence of a saturating (5 µM) concentration of CP55940, only piperonyl

butoxide accelerated the dissociation of [3H]CP55940 to a level greater than that induced

by CP-55940 alone (Figure 4.5c).

4.4.4. Effects of study compounds on mouse spleen CB2 receptors as assessed by inhibition of [3H]CP55940 binding

After verifying the IC50s of sanguinarine, chelerythrine, piperonyl butoxide and

(S)-methoprene that we reported previously for [3H]CP55940 binding to mouse brain CB1

receptors (Dhopeshwarkar et al. 2011), we employed identical concentrations to

investigate the inhibitory actions of study compounds on [3H]CP55940 binding to CB2

receptors of mouse spleen. In contrast to (S)-methoprene, which showed 51% inhibition

of [3H]CP55940 binding to CB2 receptors, the three other compounds were very much

weaker, achieving 4% inhibition (chelerythrine), 14% inhibition (sanguinarine) and 21%

inhibition (piperonyl butoxide), (Table 4.1). Since (S)-methoprene was of very similar

inhibitory potency at brain CB1 as spleen CB2 receptors, the relationship between

concentration and inhibition of [3H]CP55940 binding to spleen CB2 receptors was

explored in more detail. The results demonstrated (S)-methoprene to be a partial

inhibitor, unable to produce greater than 50% inhibition of [3H]CP55940 binding (Figure

4.6).

4.4.5. Effects of study compounds on WIN55212-2-dependent inhibition of 4-aminopyridine- (4-AP-) evoked release of L-glutamate from mouse brain synaptosomes

In marked contrast to 50 µM veratridine-evoked release of L-glutamate from

synaptosomes, the release of L-glutamate induced by 3 mM 4-AP was completely

insensitive to inhibition by 5 µM tetrodotoxin (TTX; Figure 4.7). In these preliminary

experiments, we also verified that 4-AP-evoked release of L-glutamate is partially (28.4 +

1.59 %) inhibited by the CB1-R agonist WIN55212-2 (8 µM), that this inhibition by

127

WIN55212-2 is completely relieved by the CB1-R antagonist AM251 (8 µM), and that

preincubation with AM251 alone has no influence on the amount of neurotransmitter

released by 4-AP (Figure 4.8). All study compounds were then tested at a low and high

concentration for their ability to modify WIN55212-2-dependent inhibition of L-glutamate

release from mouse brain synaptosomes. Sanguinarine and chelerythrine (both at 0.25

µM) and (S)-methoprene at 5 µM failed to affect inhibition by WIN55212-2 of 4-AP-

evoked release, however a weak (approx. 5%) reduction in inhibition by WIN55212-2

was observed with piperonyl butoxide at 5 µM (Figures 4.9-4.12). In the presence of

WIN55212-2, sanguinarine and chelerythrine (at 2 µM) enhanced the release of L-

glutamate over and above that of 4-AP alone (Figures 4.9 and 4.10), whereas (S)-

methoprene (25 µM) and piperonyl butoxide (25 µM) fully suppressed WIN55212-2-

dependent inhibition of L-glutamate release without any tendency for a

benzophenanthridine-like overshoot (Figures 4.11 and 4.12). Like AM251, chelerythrine

(2 µM), sanguinarine (2 µM), (S)-methoprene (25 µM) and piperonyl butoxide (2 µM), did

not enhance the baseline release of L-glutamate from mouse brain synaptosomes

(Figures 4.9-4.12).

4.5. Discussion

In an earlier investigation the benzophenanthridine alkaloids (chelerythrine and

sanguinarine) and the pesticide formulation components ((S)-methoprene and piperonyl

butoxide) were found to inhibit the binding of [3H]CP55940 to CB1-Rs of mouse brain and

inhibit CB1-R agonist-dependent activation of [35S]GTPγS binding to the G protein Gα

subunit (Dhopeshwarkar et al., 2011). The present research provides insight into how

these compounds inhibit the binding of [3H]CP55940 and [3H]SR141716A to CB1-Rs of

mouse brain and explores putative antagonist-like actions further by investigating the

ability of study compounds to modify evoked (CB1-R-sensitive) release of L-glutamate

from mouse brain synaptosomes.

Present evidence indicates that [3H]CP55940 and [3H]SR141716A bind to

specific regions within the CB1 receptor binding pocket (Shim, 2010). Our initial

experiments of this investigation established IC50s for study compounds in the

[3H]SR141716A binding assay. When compared to [3H]CP55940 values

128

(Dhopeshwarkar et al., 2011), the inhibitory potencies for [3H]SR141716A binding were 1

to 2-fold higher for benzophenanthridines and 2.5-fold and 3.8-fold lower for piperonyl

butoxide and (S)-methoprene respectively, indicating different capacities to impact on

radioligand-specific binding loci within the CB1-R binding pocket. Our saturation binding

and kinetic data for these radioligands support this and other mechanistic differences

between study compounds in their actions.

The saturation binding constants obtained for [3H]SR141716A agree closely with

those published by Rinaldi-Carmona et al. (1996) using a fraction from rat brain. We

found that chelerythrine and sanguinarine decrease the total number of binding sites

(Bmax) available to [3H]SR141716A without affecting the affinity (Kd). This may be

explained by the benzophenanthridines binding to an allosteric site and triggering a

profound conformational modification to the [3H]SR141716A recognition site such that

radioligand cannot bind. However our results do not exclude benzophenanthridines

binding irreversibly (or in a very slowly reversible manner) to the orthosteric site. At

similar concentrations to those used for the benzophenanthridines in this study,

Beausoleil et al., (2009) found that sanguinarine and chelerythrine competitively inhibit

the binding of a GTP fluoroprobe to Rac1b (a GTP binding protein), and this led us to

propose that benzophenanthridines may inhibit [3H]CP55940 binding to the CB1-R by

targeting the guanine nucleotide recognition site on its associated G-protein

(Dhopeshwarkar et al., 2011). However, while it is well recognized that GTP and its

GTP analogs allosterically dissociate [3H]CP55940 from the CB1 receptor (Devane et al.,

1988; Houston and Howlett, 1993), the binding of [3H]SR141716A is known to be

unaffected by 300 µM GTPγS (Rinaldi-Carmona et al., 1996). Nevertheless, our

saturation binding assays with [3H]CP55940 indicated that the benzophenanthridines

reduce radioligand binding affinity without changing Bmax, lending support to an apparent

competitive mechanism for inhibition of [3H]CP55940 binding by benzophenanthridine

alkaloids, with potential for allosteric involvement. The kinetic results support the idea

that benzophenanthridine-dependent changes to the equilibrium binding constants of

[3H]CP55940 and [3H]SR141716A can arise both through a slowing of the rate of

association and by increasing the rate of dissociation of these radioligands.

It is improbable that (S)-methoprene and piperonyl butoxide engage with the

[3H]SR141716A recognition site itself because the pattern of inhibition by these

129

compounds is mixed, with both compounds affecting Bmax and Kd of [3H]SR141716A

binding. The possibility of an allosteric mechanism with respect to [3H]SR141716A

binding is supported because (S)-methoprene and piperonyl butoxide increase the

dissociation of equilibrated radioligand from its recognition site, irrespective of whether a

saturating concentration of unlabeled SR1417126A is present or not.

Our findings regarding the mechanism of interference with the binding of

[3H]CP55940 by (S)-methoprene are not so clear. Whereas a reduction in the initial rate

of formation of the radioligand:recognition site complex by (S)-methoprene, combined

with its failure to influence the dissociation of [3H]CP55940 lends strong support to a

simple competition, this compound clearly increases Kd and reduces Bmax in saturation

binding assays, an outcome in apparent conflict with such a mechanism. We have

hypothesized that (S)-methoprene may represent a flexible analog of ∆9-

tetrahydrocannabinol (Dhopeshwarkar et al., 2011), a phytocannabinoid that binds to the

same site region of the CB1-R binding pocket as CP55940 (Gatley et al., 1997; Thomas

et al., 2005). A model that involves binding of (S)-methoprene to the [3H]CP55940

binding region in two conformations, one that allows radioligand to bind but with reduced

affinity, and one that blocks access to [3H]CP55940, could reconcile these observations.

The manner in which (S)-methoprene and piperonyl butoxide interfere with

[3H]CP55940 binding are evidently different since the ability of piperonyl butoxide when

combined with a maximum effect concentration of unlabeled CP55940 to accelerate

dissociation of radioligand over and above the rate observed with a maximum effect

concentration of CP55940 alone suggests a prominent allosteric component. We

conclude that the reduction in the Bmax of [3H]CP55940 binding observed with piperonyl

butoxide in saturation binding experiments is most likely allosterically-mediated and

therefore not inconsistent with our premise (Dhopeshwarkar et al., 2011) that piperonyl

butoxide could bind to the endocannabinoid receptor within the CB1-R binding pocket.

The L-glutamate release experiments reported in this investigation provide

strong support to the pharmacological profiling we originally proposed for

benzophenanthridines, (S)-methoprene and piperonyl butoxide that was based on

inhibition of CB1-R agonist-activated [35S]GTPγS binding (Dhopeshwarkar et al., 2011).

In preliminary experiments, we confirmed, as found originally by Wang (2003), that 4-

130

AP-dependent release of L-glutamate is inhibited by WIN55212-2, and that the inhibition

is relieved by a diarylpyrazole (AM251 in present studies). The lower level of inhibition

(approx. 28%) by WIN55212-2 observed in the present investigation aligns more closely

with the level of inhibition by WIN55212-2 of KCl-evoked L-glutamate release from

synaptosomes (Godino et al., 2007). We also found that this WIN55212-2-sensitive

component of release required extrasynaptosomal Ca++ (data not shown) and that 3 mM

4-AP-induced release of L-glutamate from synaptosomes was unaffected by

tetrodotoxin. This latter result confirmed a lack of participation of voltage-sensitive

sodium channels in the 4-AP response, a critical prerequisite because CB1-R drugs,

including WIN55212-2 and diarylpyrazoles inhibit voltage-sensitive sodium channels at

low micromolar concentrations (Nicholson et al. 2003; Kim et al., 2005; Liao et al., 2004),

and very similar concentrations of these drugs are necessary to reveal CB1-R-dependent

effects on L-glutamate release from synaptosomes.

A salient finding of the present study is that concentrations of sanguinarine,

chelerythrine, (S)-methoprene and piperonyl butoxide that have no effect on basal

release of L-glutamate are able to neutralize (reverse) the inhibitory effect of WIN55212-

2 on 4-AP-evoked release of this neurotransmitter. In parallel experiments, the classical

CB1-R antagonist AM251 displayed an identical profile in agreement with other studies

using the diarylpyrazole AM281 (Wang 2003; Godino et al., 2007). The effects of the

study compounds on WIN55212-2-dependent inhibition of 4-AP-evoked release are

concentration-dependent. Moreover, the concentrations of sanguinarine, chelerythrine,

(S)-methoprene and piperonyl butoxide that we show are capable of neutralizing the

inhibitory effects of WIN55212-2 on L-glutamate release are very similar to those needed

to produce significant inhibition of radioligand ([3H]CP55940 and [3H]SR141716A)

binding and inhibition of CB1-R agonist-stimulated [35S]GTPγS binding to the G protein.

In marked contrast to the neutral antagonist actions of (S)-methoprene and

piperonyl butoxide in the L-glutamate release assay, an inverse agonist-like action was

indicated for sanguinarine and chelerythrine since at 2 µM (and in the presence of 4-AP),

these alkaloids promote release of neurotransmitter that is greater than that achieved by

4-AP alone. Our results show that this latter phenomenon is exclusively dependent on

WIN 55212-2 being present, so for benzophenanthridines to act in this way in vivo, a

significant level of endocannabinoid tone would be necessary.

131

In conclusion, we demonstrate that binding sites for [3H]CP55940 and

[3H]SR141716A within the CB1-R binding pocket are differentially influenced in the very

low micromolar range by benzophenanthridine alkaloids (chelerythrine and

sanguinarine), and in the low to moderate micromolar range by piperonyl butoxide and

(S)-methoprene. Certain structural features of these study compounds or their binding

sites may be useful points of consideration for discovery of more potent G-protein-

coupled CB1 receptor blocking drugs that might be capable of downregulating the central

effects of endocannabinoid agonists.

132

4.6. Figures and Table

Figure 4.1 Concentration dependency of inhibition by chelerythrine (open circles), sanguinarine (solid circles), piperonyl butoxide (solid triangles) and (S)-methoprene (squares) on [3H]SR141716A binding to mouse brain CB1 receptors. IC50 and 95% confidence interval values are provided in Section 4.4.1.

133

Figure 4.2 Effect of chelerythrine (1 µM; open circles), sanguinarine (1 µM; solid circles), piperonyl butoxide (30 µM; solid triangles) and (S)-methoprene (60 µM; squares) on equilibrium binding of [3H]SR141716A to mouse brain CB1 receptors. Control data points are identified by the diamond symbols. Kd values (as nM): control 0.51 + 0.04; chelerythrine 0.47 + 0.08; sanguinarine 0.46 + 0.04; (S)-methoprene 1.5 + 0.6 and piperonyl butoxide 2.5 + 1.1. Bmax values (as pmol [3H]SR141716A/mg protein): control 0.79 + 0.02; chelerythrine 0.32 + 0.02; sanguinarine 0.50 + 0.01; (S)-methoprene 0.44 + 0.08 and piperonyl butoxide 0.56 + 0.13.

134

Figure 4.3 Effect of chelerythrine (2.5 µM; open circles), sanguinarine (1.5 µM; solid circles), piperonyl butoxide (10 µM; solid triangles) and (S)-methoprene (20 µM; squares) on equilibrium binding of [3H]CP55940 to mouse brain CB1 receptors. Control data points are identified by the diamond symbols. Kd values (as nM): control 0.36 + 0.07; chelerythrine 2.32 + 0.43; sanguinarine 2.28 + 0.77; (S)-methoprene 1.37 + 0.25 and piperonyl butoxide 0.34 + 0.19. Bmax values (as pmol [3H]SR141716A/mg protein): control 0.6 + 0.03; chelerythrine 0.65 + 0.06; sanguinarine 0.63 + 0.11; (S)-methoprene 0.25 + 0.02 and piperonyl butoxide 0.35 + 0.05.

135

Figure 4.4a Influence of study compounds on the time course of association of [3H]SR141716A and [3H]CP55940 with CB1 receptors of mouse brain. In a) membranes received a standard 15 min preincubation with sanguinarine (2.5 µM), chelerythrine (2.5 µM), piperonyl butoxide (30 µM) and (S)-methoprene (30 µM) prior to [3H]SR141716A addition. Symbols: diamonds = control; solid circles = sanguinarine; open circles = chelerythrine; triangles = piperonyl butoxide and squares = (S)-methoprene. Data points represent the means + SEMs of 3 independent experiments (a number of SEM bars are obscured by data symbols).

136

Figure 4.4b Influence of study compounds on the time course of association of [3H]SR141716A and [3H]CP55940 with CB1 receptors of mouse brain. b) The same study compound concentrations were applied simultaneously with [3H]SR141716A. Symbols: diamonds = control; solid circles = sanguinarine; open circles = chelerythrine; triangles = piperonyl butoxide and squares = (S)-methoprene.Data points represent the means + SEMs of 3 independent experiments (a number of SEM bars are obscured by data symbols).

137

Figure 4.4c Influence of study compounds on the time course of association of [3H]SR141716A and [3H]CP55940 with CB1 receptors of mouse brain. c) The effects of benzophenanthridines (5 µM), piperonyl butoxide (20 µM) and (S)-methoprene (20 µM) on the association of [3H]CP55940 under preincubation conditions are shown Symbols: diamonds = control; solid circles = sanguinarine; open circles = chelerythrine; triangles = piperonyl butoxide and squares = (S)-methoprene. Data points represent the means + SEMs of 3 independent experiments (a number of SEM bars are obscured by data symbols).

138

Figure 4.5a The influence of study compounds on the dissociation of CB1 receptor-selective radioligands. Figure 4.5a shows the effects of piperonyl butoxide (30 µM) and (S)-methoprene (60 µM) on the dissociation of [3H]SR141716A when initiated by challenge with a saturating concentration (5 µM) of SR141716A. Symbols: diamonds = control; triangles = piperonyl butoxide and squares = (S)-methoprene. Data represent mean + SEM of at least 3 independent experiments, each performed in triplicate.

139

Figure 4.5b The influence of study compounds on the dissociation of CB1 receptor-selective radioligands. Figure 4 5b, defines the effects of sanguinarine (5 µM), chelerythrine (5 µM), piperonyl butoxide (30 µM) and (S)-methoprene (60 µM) when added alone on the dissociation of [3H]SR141716A from the [3H]SR141716A:CB1 receptor complex. Symbols: diamonds = control; solid circles = sanguinarine; open circles = chelerythrine; triangles = piperonyl butoxide and squares = (S)-methoprene. Data represent mean + SEM of at least 3 independent experiments, each performed in triplicate.

140

Figure 4.5c The influence of study compounds on the dissociation of CB1 receptor-selective radioligands. Figure 4 5c, the effects of sanguinarine (5 µM), chelerythrine (5 µM), piperonyl butoxide (20 µM) and (S)-methoprene (60 µM) on the dissociation of [3H]CP55940 when initiated by application of a saturating concentration (5 µM) of CP55940 are given. Symbols: diamonds = control; solid circles = sanguinarine; open circles = chelerythrine; triangles = piperonyl butoxide and squares = (S)-methoprene. Data represent mean + SEM of at least 3 independent experiments, each performed in triplicate.

141

Figure 4.6 Relationship between concentration of (S)-methoprene and inhibition at CB2 receptors of mouse spleen based on interference with [3H]CP55940 binding.

142

Figure 4.7a Inhibition of 50 µM veratridine-evoked release of L-glutamate from mouse brain synaptosomes by 5 µM tetrodotoxin (TTX)

143

Figure 4.7b Failure of 5 µM TTX to modify 3 mM 4-AP-evoked release of L-glutamate from synaptosomes.

144

Figure 4.8 Partial inhibition of 4-AP-evoked release of L-glutamate from synaptosomes by the CB1-R agonist WIN55212-2, and full relief of WIN55212-2-dependent inhibition by the CB1-R antagonist AM251.

145

Figure 4.9 With WIN55212-2 present, sanguinarine (at 2 µM but not 0.25 µM) enhances 4-AP-evoked L-glutamate release above the level produced by 4-AP alone.

146

Figure 4.10 With WIN55212-2 present, chelerythrine (at 2 µM but not 0.25 µM) enhances 4-AP-evoked L-glutamate release above the level produced by 4-AP alone.

147

Figure 4.11 With WIN55212-2 present, (S)-methoprene (at 25 µM but not 5 µM) enhances 4-AP-evoked L-glutamate release above the level produced by 4-AP alone.

148

Figure 4.12 With WIN55212-2 present, piperonyl butoxide (at 25 µM but not 5 µM) enhances 4-AP-evoked L-glutamate release above the level produced by 4-AP alone.

149

Table 4.1 Inhibitory effects of chelerythrine, sanguinarine, piperonyl butoxide and (S)-methoprene on spleen CB2 receptors as determined with [3H]CP55940. Each study compound was added at a concentration that achieved an IC50 for [3H]CP55940 binding to brain CB1 receptors (Dhopeshwarkar et al. 2011). All values represent mean percentage inhibition + S.E.M. of at least 3 independent experiments. Parallel experiments with [3H]CP55940 corroborated our previously published IC50s at brain CB1 receptors (2.2 µM chelerythrine gave 49.03 + 0.94 % inhibition, 1.2 µM sanguinarine gave 51.33 + 0.49 % inhibition, 8.2 µM piperonyl butoxide gave 47.50 + 1.17 % inhibition and 16.4 µM methoprene gave 50.22 + 1.10 % inhibition).

Compound Inhibition of [3H]CP55940 binding to CB2 receptors of mouse spleen

Chelerythrine (2.2 µM) 4.14 + 0.14

Sanguinarine (1.2 µM 14.34 + 0.23

Piperonyl butoxide (8.2 µM) 20.86 + 0.23

(S)-Methoprene (16.4 µM) 50.98 + 0.21

150

5. Effects of organotins on the CB1 receptor pathway of mouse brain in vitro.

Note: The research described in this chapter will be submitted shortly for

publication in an appropriate neurochemical/neuropharmacological journal. The

submission will adhere closely to the format laid out here.

5.1. Introduction

This Chapter constitutes a report on the modulatory effects of eight tributyltins

and two triphenyltin compounds at G protein-coupled cannabinoid-1 receptors (CB1-Rs)

of mouse brain in vitro. Organotin compounds have been exploited extensively for their

ability to preserve lumber and prevent biofouling on marine structures. Concerns over

toxicity to many marine and terrestrial species has led to severe restrictions on organotin

use, however they continue to be used on large ocean-going cargo ships because of the

huge reductions in fuel costs (and CO2 output) that can be achieved by minimizing

biofouling and the associated frictional drag on hulls. Tributyltin derivatives have also

seen significant uses as chemical intermediates and also as catalysts for various

chemical reactions.

The nervous system is known to be a highly sensitive target of tributyltins and a

number of mechanisms appear to be involved in their toxic actions. For example, in vivo

administration of tributyltin chloride to pregnant mice elevates dopamine concentrations

in the striatum and 5-hydroxytryptamine (5-HT) concentrations in the medulla oblongata

of F1 offspring, while dams exhibit widespread decreases in brain 5-HT levels (Tsunoda

et al., 2006). Tributyltins have been reported to lower the binding of [3H]MK801 to NMDA

receptors in the cerebral cortex of mice both in vitro and in vivo (Konno et al., 2001).

151

Tributlytins modulate glutathione levels in the striatum, hippocampus and cortex (Fortier

et al., 2010). Measurements of extracellular glutamate concentrations indicate that

tributyltins cause significant release of L-glutamate from neurons, leading to

excitotoxicity (Nakatsu et al., 2006).

The potential of tributyltins to interfere with the binding of [3H]CP55940 was

demonstrated originally as a result of experiments conducted by Dr. Chengyong Liao in

our laboratory. Mr. Sudip Ghose and Mr. Saurabh Jain of our laboratory went on to test

other organotin compounds in the [3H]CP55940 binding assay. Selected tributyltin

compounds were subsequently assayed for their capacity to modulate basal and

CP55940-stimulated binding of [35S]GTPγS to the G protein by Mr. Saurabh Jain (M.Sc.

Thesis Simon Fraser University, 2011) These assays suggested that some tributyltins

could possibly act as inverse agonists of CB1-Rs at low micromolar concentrations. My

task was to examine selected tributyltins for their ability to modify WIN55212-2-

dependent inhibition of evoked transmitter (L-glutamate) release from synaptosomes,

since we hypothesized that the inhibitory effect of WIN55212-2 should be neutralized

and transmitter release may possibly undergo further enhancement.

5.2. Materials and methods

All organotin compounds were purchased from Sigma-Aldrich. Methodological

details of the binding assays for [3H]CP55940 and [35S]GTPγS have been adequately

described in Chapter 2, Section 2.3.3 and 2.3.4 and Chapter 3, Section 3.3.2 and 3.3.3

of this thesis. The isolation of synaptosomes from mouse brain and the assay of L-

glutamate release was described in sufficient detail in the previous Chapter 4 (Section

4.3.7 and 4.3.8).

152

5.3. Results

5.3.1. Displacement of [3H]CP55940 binding to mammalian CB1 receptors by organotin compounds

Eight tributyltins and 2 triphenyltins were tested for their ability to interfere with

[3H]CP55940 binding to mammalian CB1 receptors. The two triphenyltins and all

tributyltins except two (tributyltin hydride and tributylphenyltin; both inactive) achieved

IC50s at low micromolar concentrations (Table 5.1). All except the two inactive

compounds gave well-defined sigmoidal inhibition curves. Apart from phenylethynyl

tributyltin (which had the highest IC50 value of the group of active compounds and

produced about 80% inhibition at its maximum effect concentration), all other active

compounds achieved full (93-100%) inhibition. Concentration:inhibition curves are

displayed for tributyltin benzoate, tributyltin acetate and phenylethynyl tributyltin (Figure

5.1).

5.3.2. Basal and CP55940-stimulated [35S]GTPγS binding to the Gα subunit as influenced by tributyltin compounds

The inhibitory effects of tributyltin benzoate and phenylethynyl tributyltin on

[35S]GTPγS binding to the Ga subunit are shown in Figure 5.3. Tributyltin benzoate

reached IC50 at 1.43 µM (95% CI = 1.35 - 1.53 µM) and produced complete inhibition of

agonist-stimulated [35S]GTPγS binding around 2.5 µM. In addition tributyltin benzoate

encroached 51.03 ± 3.24% into basal binding at 5 µM with agonist present.

Phenylethynyl tributyltin achieved IC50 at 1.87 µM (95% CI = 1.71 - 2.02 µM) and

reached complete inhibition of agonist-stimulated [35S]GTPγS binding around 4 µM with

48.98 ± 14.64% and 68.53 ± 8.81% encroachment into the basal binding signal at 10 µM

and 20 µM respectively. Tributyltin benzoate (on its own) reduced basal [35S]GTPγS

binding by 22.12 ± 3.60% (at 2.5 µM) and 50.87 ± 1.87% (at 5 µM). Phenylethynyl

tributyltin (on its own) also reduced basal binding signal by 63.06 ± 3.62% at 20 µM.

153

5.3.3. Modulation by tributyltin acetate and phenylethynyl tributyltin of WIN55212-2-dependent inhibition of 4-aminopyridine-evoked release of L-glutamate from mouse brain synaptosomes

The capacity of tributyltin acetate and phenylethynyl tributyltin to modify CB1-R

agonist- (WIN55212-2-) dependent inhibition of 4-AP-evoked release of L-glutamate

from synaptosomes was explored to test the hypothesis that tributyltins negatively

modulate CB1-Rs by acting as inverse agonists. The fluorescence traces of L-glutamate

release from synaptosomes under these circumstances are displayed in Figures 5.4 and

5.5. Tributyltin acetate and phenylethynyl tributyltin (at 3 µM) fully relieved the inhibition

of 4-AP-evoked release caused by 8 µM WIN55212-2. This effect was identical to that of

the classical diarylpyrazole inverse agonist/antagonist AM251 at 8 µM. No relief of

WIN55212-2-dependent inhibition was observed at 0.5 µM with either compound. In

contrast to phenylethynyl tributyltin, tributyltin acetate (at 3 µM) actually enhanced the

release of L-glutamate over and above that observed with 4-AP alone. Little effect (2-8%

change) of the organotin compounds alone on background release of L-glutamate from

mouse brain synaptosomes was detected (Figures 5.4 and 5.5).

5.4. Discussion

This investigation demonstrates that a number of tin-containing compounds have

the capacity to inhibit the binding of [3H]CP55940 to CB1-Rs in mouse brain at very low

micromolar concentrations. However, the relationships between structural features and

inhibitory potencies are complex. Compounds containing a wide variety of substituents

attached to the central tin atom of the tributyltin system (including benzoate, acetate,

methoxide, hydroxide, bromide and trifluoromethane sulfonate), as well as the hydroxide

derivative of triphenyltin, all achieved IC50 between 2.0 and 3.3 µM. The chloride

derivative of triphenyltin and the phenylethynyl derivative of tributyltin were less potent

(IC50s approx. 5 and 15 µM respectively). Intriguingly, the highly bulky three phenyl

substituent system (in triphenyltin hydroxide and triphenyltin chloride) fails to reduce

inhibitory potency greatly. Furthermore, in the tributyltin series, if a single phenyl ring is

attached directly to the central tin atom (e.g. tributylphenyl tin), inhibitory activity is

eliminated. However, if an ethynyl or carboxylate spacer is inserted between the tin atom

154

and the phenyl ring (as in phenylethynyl tributyltin or tributyltin benzoate), the ability to

inhibit the binding of [3H]CP55940 to CB1-Rs dramatically recovers. In addition, within

the subtituted tributyltin series, the hydride is inactive; nevertheless, replacement of the

hydrogen with bromine produces the most potent organotin inhibitor we have found to

date.

The results of the [35S]GTPγS binding experiments (conducted by Mr. Saurabh

Jain) provided the first indication that tributyltins are able to interfere functionally with

CB1-R agonist-dependent activation of the CB1-R:G protein complex. For these

experiments, we focused specifically on tributyltin benzoate and phenylethynyl tributyltin.

Since the results demonstrated that both compounds inhibit basal binding of [35S]GTPγS,

CP55940-stimulated binding of [35S]GTPγS to the Gα subunit and they also inhibit the

basal [35S]GTPγS signal with CP55940 present, we infer that they likely act as inverse

agonists of CB1-Rs.

The possibility that tributyltins negatively modulate CB1-Rs by acting as inverse

agonists, was pursued further by investigating the ability of two analogs (tributyltin

acetate and phenylethynyl tributyltin) to modulate CB1-R agonist- (WIN55212-2-)

dependent inhibition of the release L-glutamate from synaptosomes following challenge

with 4-aminopyridine. This assay is well suited for such studies because it has been

shown by Wang (2003) and Godino et al. (2007) that inhibition of evoked L-glutamate

release from synaptosomes by WIN55212-2 is blocked by AM281, a classical

diarylpyrazole inverse agonist. Tributyltin acetate and phenylethynyl tributyltin fully

suppressed the inhibition of 4-AP-evoked release caused by WIN55212-2. However, in

contrast to phenylethynyl tributyltin, tributyltin acetate actually enhanced release over

and above that observed with 4-AP alone. Tributyltin acetate and phenylethynyl

tributyltin are certainly functional inhibitors of CB1-Rs as assessed by their modulatory

effects on presynaptic release of L-glutamate. Moreover, they closely mimic the standard

antagonist/inverse agonist AM281 in this assay. The profile of tributyltin acetate is

particularly consistent with an inverse agonist action at brain CB1-Rs. Given the findings

of this study, the ability of tributyltins to cause significant release of L-glutamate from

neurons and excitotoxicity (Nakatsu et al., 2006) may be caused in part by blockade of

endocannabinoid-dependent inhibition of L-glutamate release.

155

Although organotin compounds can be quite toxic to mammals, there may be

potential to develop from these structures a new class of drug that acts at CB1 receptors.

A practical approach may be to develop hybrid molecules that link the tin-containing

center with various pharmacophoric moieties present in the many classical CB1-R active

drugs. These hybrids may or may not be problematic from the toxicological standpoint.

Additionally, replacement of the central tetravalent tin atom with silicon or carbon could

generate useful structure activity data that could assist in drug design especially if tin-

containing prototypes turned out to be too toxic.

156

5.5. Figures and Table

Sn

O

O

Sn

O

Sn

O

S

O

C

F

3

O

TBT acetate TBT methoxide Tributylstannyl trifluoromethane sulphonate

Sn

Br

Sn

Sn

TBT bromide Tributylphenyl tin Tributyl(phenyl ethynyl)tin

Sn

O

O

Sn

O

H

Sn

Cl

TBT benzoate Triphenyl tin hydroxide Triphenyl tin chloride

Sn

H

TBT hydride

Figure 5.1 Structures of tributyl and triphenyltin compounds examined in the present investigation. Structures were constructed using Isis Draw.

157

Figure 5.2 Concentration-dependent inhibition of specific [3H]CP55940 binding to mouse brain CB1 receptors by tributyltin benzoate, tributyltin acetate and phenylethynyl tributyltin. Each data point represents the mean ± S.E.M. inhibition of specific [3H]CP55940 binding for at least three independent assays, each performed in triplicate. Experiments conducted by Mr. Saurabh Jain. This figure was originally published in the M.Sc. thesis of Mr. Saurabh Jain (Simon Fraser University, 2011).

158

Figure 5.3 Concentration-dependent inhibition of CP55940 (100 nM)-stimulated [35S]GTPγS binding by tributyltin benzoate and phenylethynyl tributyltin. Each data point represents the mean ± S.E.M. percentage inhibition of CP55940 stimulated [35S]GTPγS binding determined by three independent assays each performed in triplicate. These experiments were conducted by Mr Saurabh Jain and this figure was originally published in the M.Sc. thesis of Mr. Saurabh Jain (Simon Fraser University, 2011).

159

Figure 5.4 Modulation of WIN55212-2-dependent inhibition of 4-aminopyridine (4-AP-)-evoked release of L-glutamate from mouse brain synaptosomes by tributyltin acetate (TBT acetate). Typical release profiles are displayed with mean % changes (+ SEM) to 4-AP-evoked and control release in the adjacent table.

160

Figure 5.5 Modulation of WIN55212-2-dependent inhibition of 4-aminopyridine (4-AP-)-evoked release of L-glutamate from mouse brain synaptosomes by phenylethynyl tributyltin (TBPE tin). Typical release profiles are displayed with mean % changes (+ SEM) to 4-AP-evoked and control release in the adjacent table.

161

Table 5.1 Inhibitory effects of tributyl and triphenyltins on the binding of [3H]CP55940 to CB1 receptors in mouse brain. All values are as IC50s (with 95% confidence intervals in brackets) calculated from curves based on at least 3 independent experiments except for triphenyltin chloride where the IC50 was estimated from 2 independent experiments).

Organotin IC50 with 95% confidence interval (µM)

TBT benzoate 2.6 (1.7 - 3.9)

TBT acetate 2.7 (2.3 - 3.3)

TBT methoxide 3.3 (2.7-3.8)

TBT bromide 2.0 (1.6 - 2.5)

Tributylstannyl-TMS 2.9(2.3- 3.6)

Triphenyltin hydroxide 2.6 (1.5-4.5)

Triphenyltin chloride 5.1

Tributylphenylethynyl tin 14.8 (9.8 - 22.2)

TBT hydride >100

Tributylphenyltin >100

The above IC50 values were calculated from concentration:inhibition experiments carried out by Dr. Chengyong Liao, Mr. Sudip Ghose and Mr. Saurabh Jain. TBT = tributyltin, TMS = trifluoromethane sulfonate

162

6. Conclusion and future prospects

The conclusions of my Ph.D. research have already been discussed in detail

elsewhere (see Chapter 2, Section 2.5; Chapter 3, Section 3.5; Chapter 4, Section 4.5;

Chapter 5, Section 5.4).

In summary, I conclude that sanguinarine, chelerythrine, (S)-methoprene and

piperonyl butoxide exert inhibitory actions at CB1-Rs in mouse brain in vitro. These

compounds act at very low to moderate micromolar concentrations with an inhibitory

potency ranking (estimated from [3H]CP55940, [3H]SR141716A and [35S]GTPγS binding

data) of sanguinarine ~ chelerythrine > piperonyl butoxide > methoprene.

Based on my saturation binding and kinetic experiments I infer that these

compounds inhibit via predominently allosteric mechanisms with respect to [3H]CP55940

and [3H]SR141716A binding.

My experiments with mouse brain synaptosomes demonstrate that WIN-55212-2-

dependent inhibition of 4-AP-evoked L-glutamate release is blocked by sanguinarine,

chelerythrine, (S)-methoprene and piperonyl butoxide. The actions of (S)-methoprene

and piperonyl butoxide are indistinguishable from AM251 (a classical diarylpyrazole CB1-

R antagonist), demonstrating an antagonist action at CB1-Rs. In addition to blocking the

inhibitory effect of WIN55212-2 on evoked release of L-glutamate, sanguinarine and

chelerythrine (with WIN55212-2 present) increase the release of neurotransmitter to a

level greater than that caused by WIN55212-2 alone. This suggests an inverse agonist-

like action of the benzophenanthridines. The L-glutamate release results therefore align

with our previous profiling of these compounds in the [35S]GTPγS binding assay.

Building on the findings of Dr. Chengyong Liao, Ms. Kathleen M. Bisset and Mr.

Saurabh Jain in my laboratory, I further explored the pharmacological actions of

phthalate esters and tributyltins at brain CB1-Rs. Based on [35S]GTPγS binding and L-

163

glutamate release results I conclude that these common environmental pollutants are

antagonists of CB1-R function.

The environmental chemicals highlighted in this thesis represent a broad range

of structural classes. It is interesting that when their potential to modify functional

outcomes of the CB1-R signaling pathway was investigated (i.e. using [35S]GTPγS

binding and L-glutamate release) only inhibitory (antagonist or inverse agonist-like)

actions were revealed. From these observations I infer that environmental chemicals

possessing the structural features of a CB1-R agonist might be more rarely encountered.

It would be logically predicted that if the study compounds were able to enter the

brain and engage with CB1-Rs, they should reduce the effectiveness of

endocannabinoids (e.g. anandamide and 2-AG) in activating CB1-Rs. I recommend that

future studies examine the ability of these compounds to inhibit both [3H]anandamide

binding to CB1-Rs and anandamide-induced suppression of L-glutamate release from

synaptosomes. Another very important line of future research would be to examine how

in-vivo administration of study compounds might modify the classical behavioral

manifestations of CB1-R agonists. Should any compound show CB1-R antagonist or

inverse agonist-like effects in vivo, careful consideration should be given to its potential

as a prototype for rational design of more effective analogs. As mentioned in the

introduction, CB1-R antagonists are effective in body weight reduction. Despite of the

issue with Rimonabant (SR141716A) and the USFDA's recent approval of Lorcaserin (a

weight reducing 5-HT2C agonist), there is a huge demand for drugs with this property.

Certain study compounds (e.g. (S)-methoprene and piperonyl butoxide) would represent

very low acute toxicity starting points and I would also suggest that improvements in

potency to the level of SR141716A may not be needed as some CNS drugs are effective

and well tolerated at low micromolar concentrations.

164

References

Abadji, V., Lin, S., Taha, G., Griffin, G., Stevenson, L., Pertwee, R., & Makriyannis, A. (1994). (R)-Methanandamide-a chiral novel anandamide possessing higher potency and metabolic stability. Journal of Medicinal Chemistry, 37(12), 1889-1893.

Adams, R., Harfenist, M., & Loewe, S. (1949). New analogs of tetrahydrocannabinol Journal of the American Chemical Society, 71(5), 1624-1628.

Ahn, H., Bertalovitz, A., Mierke, D., & Kendall, A. (2009). Dual role of the second extracellular loop of the cannabinoid receptor 1: Ligand binding and receptor localization. Molecular Pharmacology, 76(4), 833-842.

Ahn, H., Nishiyama, A., Mierke, F., & Kendall, A. (2010). Hydrophobic residues in helix 8 of cannabinoid receptor 1 are critical for structural and functional properties. Biochemistry, 49(3), 502-511.

Ahn, H., Pellegrini, M., Tsomaia, N., Yatawara, K., Kendall, A., & Mierke, F. (2009). Structural analysis of the human cannabinoid receptor one carboxyl-terminus identifies two amphipathic helices. Biopolymers, 91(7), 565-573.

Alger, B. & Pitler, T. (1995). Retrograde signaling at GABAA-receptor synapses in the mammalian CNS. Trends in Neurosciences, 18(8), 333-340.

Ameri, A. (1999). The effects of cannabinoids on the brain. Progress in Neurobiology, 58(4), 315-348.

Anavi-Goffer, S., Fleischer, D., Hurst, P., Lynch, L., Barnett-Norris, J., Shi, S., Lewis, S., Mukhopadhyay. S., Howlett, C., Reggio, H., & Abood, E. (2007). Helix 8 leu in the CB1 cannabinoid receptor contributes to selective signal transduction mechanisms. Journal of Biological Chemistry, 282(34), 25100-25113.

Antizar-Ladislao, B. (2008). Environmental levels, toxicity and human exposure to tributyltin (TBT)-contaminated marine environment. A review. Environment International, 34(2), 292-308.

Bab I. (2005). The skeleton: stone bones or stoned heads? Cannabinoids as therapeutics. ed. R. Mechoulam, pp201-206. Basel, Switzerland: Birkhauser Verlag AG.

165

Barnett-Norris, J., Guarnieri, F., Hurst, D., & Reggio, P. (1998). The exploration of biologically relevant conformations of anandamide, 2-arachidonylglycerol and their analogues using conformational memories. Journal of Medicinal Chemistry, 41(24), 4861-4872.

Barnett-Norris, J., Hurst, D., Lynch, D., Guarnieri, F., Makriyannis, A., & Reggio, P. (2002). Conformational memories and the endocannabinoid binding site at the cannabinoid CB1 receptor. Journal of Medicinal Chemistry, 45(17), 3649-3659.

Barug, D. (1981). Microbial-degradation of bis(tributyltin) oxide. Chemosphere, 10(10), 1145-1154.

Basavarajappa, B. (2007). Critical enzymes involved in endocannabinoid metabolism. Protein and Peptide Letters, 14(3), 237-246.

Beausoleil, E., Chauvignac, C., Taverne, T., Lacombe, S., Pognante, L., Leblond, B., Pallares, D., De Oliveira, C., Bachelot, F., Carton, R., Peillon, H., Coutadeur, S., Picard, V., Lambeng, N., Desire, L., & Schweighoffer, F. (2009). Structure–activity relationship of isoform selective inhibitors of Rac1/1b GTPase nucleotide binding. Biorganic and Medicinal Chemistry Letters, 19 (1), 5594–5598.

Beinfeld, M., & Connolly, K. (2001). Activation of CB1 cannabinoid receptors in rat hippocampal slices inhibits potassium-evoked cholecystokinin release, a possible mechanism contributing to the spatial memory defects produced by cannabinoids. Neuroscience Letters, 301(1), 69-71.

Bell, M., Dambra, T., Kumar, V., Eissenstat, M., Herrmann, J., Wetzel, J., Rosi, D., Philion, R., Daum, S., Hlasta, D., Kullnig, R., Ackerman, J., Haubrich, D., Luttinger, D., Baizman, R., Miller, M., & Ward, S. (1991). Antinociceptive (Aminoalkyl)indoles. Journal of Medicinal Chemistry, 34(3), 1099-1110.

Beltramo, M., & Piomelli, D. (2000). Carrier-mediated transport and enzymatic hydrolysis of the endogenous cannabinoid 2-arachidonylglycerol. Neuroreport, 11(6), 1231-1235.

Beltramo, M., Stella, N., Calignano, A., Lin, S., Makriyannis, A., & Piomelli, D. (1997). Functional role of high-affinity anandamide transport, as revealed by selective inhibition. Science, 277(5329), 1094-1097.

Benito, C., Kim, W., Chavarria, I., Hillard, C., Mackie, K., Tolon, R., Williams, K., & Romero, J. (2005). A glial endogenous cannabinoid system is upregulated in the brains of macaques with simian immunodeficiency virus-induced encephalitis. Journal of Neuroscience, 25(10), 2530-2536.

Benito, C., Nunez, E., Tolon, R., Carrier, E., Rabano, A., Hillard, C., & Romero, J. (2003). Cannabinoid CB2 receptors and fatty acid amide hydrolase are selectively overexpressed in neuritic plaque-associated glia in Alzheimer's disease brains. Journal of Neuroscience, 23(35), 11136-11141.

166

Bermak, J., Li, M., Bullock, C., & Zhou, Q. (2001). Regulation of transport of the dopamine D1 receptor by a new membrane-associated ER protein. Nature Cell Biology, 3(5), 492-498.

Berrendero, F., & Maldonado, R. (2002). Involvement of the opoid system in the anxiolytic effects induced by ∆9-tetrahydrocannabinol. Psychopharmacology, 163(1), 111-117.

Bertalovitz, A., Ahn, K., & Kendall, D. (2010). Ligand binding sensitivity of the extracellular loop two of the cannabinoid receptor 1. Drug Development Research, 71(7), 404-411.

Bidaut-Russell, M., Devane, A., & Howlett, A. (1990). Cannabinoid receptors and modulation of cyclic AMP accumulation in the rat brain. Journal of Neurochemistry, 55(1), 21-26.

Bisset, K. M., Dhopeshwarkar, A. S., Liao, C., & Nicholson, R. A. (2011). The G protein-coupled cannabinoid-1 (CB1) receptor of mammalian brain: Inhibition by phthalate esters in vitro. Neurochemistry International, 59(5), 706-713.

Bisogno, T., Ligresti, A., & Di Marzo, V. (2005). The endocannabinoid signalling system: Biochemical aspects. Pharmacology Biochemistry and Behavior, 81(2), 224-238.

Bisogno, T., Maccarrone, M., De Petrocellis, L., Jarrahian, A., Finazzi-Agro, A., Hillard, C., & Di Marzo, V. (2001). The uptake by cells of 2-arachidonoylglycerol, an endogenous agonist of cannabinoid receptors. European Journal of Biochemistry, 268(7), 1982-1989.

Bisogno, T., Melck, D., De Petrocellis, L., & Di Marzo, V. (1999). Phosphatidic acid as the biosynthetic precursor of the endocannabinoid 2-arachidonoylglycerol in intact mouse neuroblastoma cells stimulated with ionomycin. Journal of Neurochemistry, 72(5), 2113-2119.

Bisogno, T., Sepe, N., Melck, D., Maurelli, S., DePetrocellis, L., & DiMarzo, V. (1997). Biosynthesis, release and degradation of the novel endogenous cannabimimetic metabolite 2-arachidonoylglycerol in mouse neuroblastoma cells. Biochemical Journal, 322(1), 671-677.

Blair, W., Olson, G., Brinckman, F., & Iverson, W. (1982). Accumulation and fate of tri-N-butyltin cation in estuarine bacteria. Microbial Ecology, 8(3), 241-251.

Blankman, J., Simon, G., & Cravatt, F. (2007). A comprehensive profile of brain enzymes that hydrolyze the endocannabinoid 2-arachidonoylglycerol. Chemistry & Biology, 14(12), 1347-1356.

Blount, B., Silva, M., Caudill, S., Needham, L., Pirkle, J., Sampson, E., Lucier, G., Jackson, R., & Brock, J. (2000). Levels of seven urinary phthalate metabolites in a human reference population. Environmental Health Perspectives, 108(10), 979-982.

167

Bonhaus, D., Chang, L., Kwan, J., & Martin, G. (1998). Dual activation and inhibition of adenylyl cyclase by cannabinoid receptor agonists: Evidence for agonist-specific trafficking of intracellular responses. Journal of Pharmacology and Experimental Therapeutics, 287(3), 884-888.

Boring, D., Berglund, B., & Howlett, A., (1996). Cerebrodiene, arachidonylethanolamide, and hybrid structures: potential for interaction with brain cannabinoid receptors. Prostaglandins, Leukotrienes & Essential Fatty Acids, 55(1), 207-210.

Bornehag, C., Sundell, J., Weschler, C., Sigsgaard, T., Lundgren, B., Hasselgren, M., & Hagerhed-Engman, L. (2004). The association between asthma and allergic symptoms in children and phthalates in house dust: A nested case-control study. Environmental Health Perspectives, 112(14), 1393-1397.

Bornehag, C., Lundgren, B., Weschler, C., Sigsgaard, T., Hagerhed-Engman, L., & Sundell, J. (2005). Phthalates in indoor dust and their association with building characteristics. Environmental Health Perspectives, 113(10), 1399-1404.

Bosnir, J., Puntaric, D., Galic, A., Skes, I., Dijanic, T., Klaric, M., Grgic, M., Curkovic, M., & Smit, Z. (2007). Migration of phthalates from plastic containers into soft drinks and mineral water. Food Technology and Biotechnology, 45(1), 91-95.

Bouaboula, M., Perrachon, S., Milligan, L., Canat, X., RinaldiCarmona, M., Portier, M., Barth, F., Calandra, B., Pecceu, F., Lupker, J., Maffrand, J., LeFur, G., & Casellas, P. (1997). A selective inverse agonist for central cannabinoid receptor inhibits mitogen-activated protein kinase activation stimulated by insulin or insulin-like growth factor 1: evidence for a new model of receptor/ligand interactions. Journal of Biological Chemistry, 272(35), 22330-22339.

Bouaboula, M., PoinotChazel, C., Marchand, J., Canat, X., Bourrie, B., RinaldiCarmona, M., Calandra, B., LeFur, G., & Casellas, P. (1996). Signaling pathway associated with stimulation of CB2 peripheral cannabinoid receptor-involvement of both mitogen-activated protein kinase and induction of krox-24 expression. European Journal of Biochemistry, 237(3), 704-711.

Bradshaw, H., & Walker, J., (2005). The expanding field of cannabimimetic and related lipid mediators. British Journal of Pharmacology, 144(1), 459-465.

Bramblett, R., Panu, A., Ballesteros, J., & Reggio, P. (1995). Construction of a 3D model of the cannabinoid CB1 receptor: determination of helix ends and helix orientation. Life Sciences, 56(23-24), 1971-1982

Breaud, T., Farlow, J., Steelman, C., & Schilling, P. (1977). Effects of insect growth-regulator methoprene on natural-populations of aquatic organisms in Louisiana intermediate marsh habitats. Mosquito News, 37(4), 704-712.

Breivogel, C., & Childers, S. (2000). Cannabinoid agonist signal transduction in rat brain: Comparison of cannabinoid agonists in receptor binding, G-protein activation, and adenylyl cyclase inhibition. Journal of Pharmacology and Experimental Therapeutics, 295(1), 328-336.

168

Breivogel, C., Griffin, G., Di Marzo, V., & Martin, B. (2001). Evidence for a new G protein-coupled cannabinoid receptor in mouse brain. Molecular Pharmacology, 60(1), 155-163.

Brown, A. J. (2007). Novel cannabinoid receptors. British Journal of Pharmacology, 152(5), 567-575

Burstein, S., Budrow, J., Debatis, M., Hunter, S., & Subramanian, A. (1994). Phospholipase participation in cannabinoid-induced release of free arachidonic-acid. Biochemical Pharmacology, 48(6), 1253-1264.

Burstein, S., Hunter, S., Latham, V., Mechoulam, R., Melcgior, D., Renzulli, L., & Tefft, R. (1986). Prostaglandins and cannabis: comparison of enantiomeric cannabinoids in stimulating prostaglandin synthesis in fibroblasts. Life Sciences, 39(19), 1813-1823.

Cadas, H., diTomaso, E., & Piomelli, D. (1997). Occurrence and biosynthesis of endogenous cannabinoid precursor, N-arachidonoyl phosphatidylethanolamine, in rat brain. Journal of Neuroscience, 17(4), 1226-1242.

Cadogan, A., Alexander, S., Boyd, E., & Kendall, D. (1997). Influence of cannabinoids on electrically evoked dopamine release and cyclic AMP generation in the rat striatum. Journal of Neurochemistry, 69(3), 1131-1137.

Calignano, A., La Rana, G., Giuffrida, A., & Piomelli, D. (1998). Control of pain initiation by endogenous cannabinoids. Nature, 394(6690), 277-281.

Callet, D., Autian, J., & Guess, W. (1966). Toxicology of a series of phthalate esters. Journal of Pharmaceutical Sciences, 55(2), 158-159.

Cao, X. (2010). Phthalate esters in foods: Sources, occurrence, and analytical methods. Comprehensive Reviews in Food Science and Food Safety, 9(1), 21-43.

Carriba, P., Navarro, G., Ciruela, F., Ferre, S., Casado, V., Agnati, L., Cortis, A., Mallol, J., Fuxe, K., Canela, E., Carmen, E., & Franco, R. (2008). Detection of heteromerization of more than two proteins by sequential BRET-FRET. Nature Methods, 5(8), 727-733

Carrier, E., Kearn, C., Barkmeier, A., Breese, N., Yang, W., Nithipatikom, K., Pfister, S., Campbell, W., & Hillard, C. (2004). Cultured rat microglial cells synthesize the endocannabinoid 2-arachidonylglycerol, which increases proliferation via a CB2 receptor-dependent mechanism. Molecular Pharmacology, 65(4), 999-1007.

Casida, J. (1970). Mixed-function oxidase involvement in biochemistry of insecticide synergists. Journal of Agricultural and Food Chemistry, 18(5), 753-760.

Castane, A., Valjent, E., Ledent, C., Parmentier, M., Maldonado, R., & Valverde, O. (2002). Lack of CB1 cannabinoid receptors modifies nicotine behavioural responses, but not nicotine abstinence. Neuropharmacology, 43(5), 857-867.

169

Chakrabarti, A., Onaivi, E., & Chaudhuri, G. (1995). Cloning, sequencing and characterization of mouse brain-type cannabinoid receptor gene. FASEB Journal, 9(3), 404-408.

Chan, P., & Yung, W. (1998). Occlusion of the presynaptic action of cannabinoids in rat substantia nigra pars reticulata by cadmium. Neuroscience Letters, 249(1), 57-60.

Chen, K., Neu, A., Howard, A. L., Foldy, C., Echegoyen, J., Hilgenberg, L., Smith, M., Mackie, K., & Soltesz, I. (2007). Prevention of plasticity of endocannabinoid signaling inhibits persistent limbic hyperexcitability caused by developmental seizures. Journal of Neuroscience, 27(1), 46-58.

Chen, R., Frassetto, A., Lao, J, Huang, R., Xiao, J., Clements, M., Walsh, T., Hale, J., Wang, J., Tong, X., & Fong, T. (2008). Pharmacological evaluation of LH-21, a newly discovered molecule that binds to cannabinoid CB1 receptor. European Journal of Pharmacology, 584(1), 338-342.

Chen, X., Sheller, J., Johnson, E., & Funk, C. (1994). Role of leukotrienes revealed by targeted disruption of the 5-lipoxygenase. Nature, 372(6502), 179-182.

Chevaleyre, V., Heifets, B. D., Kaeser, P. S., Sudhof, T. C., Purpura, D. P., & Castillo, P. E. (2007). Endocannabinoid-mediated long-term plasticity requires cAMP/PKA signaling and RIM1 alpha. Neuron, 54(5), 801-812.

Chevaleyre, V., Takahashi, K., & Castillo, P. (2006). Endocannabinoid-mediated synaptic plasticity in the CNS. Annual Reviews of Neuroscience, 29(1), 37-76.

Childers, S. (2006). Activation of G-proteins in brain by endogenous and exogenous cannabinoids. AAPS Journal, 8(1), 112-117.

Childers S., & Deadwyler, S. (1996). Role of cyclic AMP in the actions of cannabinoid receptors. Biochemical Pharmacology, 52(1), 819-827,

Chillakuri, C., Reinhart, C., & Michel, H. (2007). C-terminal truncated cannabinoid receptor 1 coexpressed with G protein trimer in Sf9 cells exists in a precoupled state and shows constitutive activity. FEBS Journal, 274(23), 6106-6115.

Chin, C., Lucas-Lenard, J., Abadji, V., & Kendall, D. (1998). Ligand binding and modulation of cyclic AMP levels depend on the chemical nature of residue 192 of the human cannabinoid receptor 1. Journal of Neurochemistry, 70(1), 366-373.

Choi, G., Guo, J., & Makriyannis, A. (2005). The conformation of the cytoplasmic helix 8 of the CB1 cannabinoid receptor using NMR and circular dichroism. Biochimica Et Biophysica Acta-Biomembranes, 1668(1), 1-9.

Christensen, R., Kristensen, P., Bartels, E., Blidda, H., & Astrup, A. (2007). Efficacy and safety of the weight-loss drug rimonabant: A meta-analysis of randomised trials. Lancet, 370(9600), 1706-1713.

170

Colombo, G., Agabio, R., Diaz, G., Lobina, C., Reali, R., & Gessa, G. (1998). Appetite suppression and weight loss after the cannabinoid antagonists SR141716. Life Sciences, 63(8), 113-117.

Compton, D., Gold, L., Ward, S., Balster, R., & Martin, B. (1992). Aminoalkylindole analogs: cannabimimetic activity of a class of compounds structurally distinct from Δ9-tetrahydrocannabinol. Journal of Experimental Therapeutics, 263(1), 1118-1126.

Compton, D., Rice, K., De Costa, B., Razdan, R., Melvin, L., Johnson, M., & Martin, B. (1993). Cannabinoid structure-activity relationships: correlation of receptor binding and in vivo activities. Journal of Experimental Therapeutics, 265(1), 218-226.

Compton, D., Aceto, M., Lowe, J., & Martin, B. (1996). In vivo characterization of a specific cannabinoid receptor antagonist (SR141716A): inhibition of ∆9-tetrahydrocannabinol-induced responses and apparent agonist activity. Journal of Pharmacology and Experimental Therapeutics, 277(1), 586-594.

Cooney, J. (1995). Organotin compounds and aquatic bacteria - a review. Helgolander Meeresuntersuchungen, 49(1-4), 663-677.

Cooney, J., & Wuertz, S. (1989). Toxic effects of tin-compounds on microorganisms. Journal of Industrial Microbiology, 4(5), 375-402.

Cravatt, B., Giang, D., Mayfield, S., Boger, D., Lerner, R., & Gilula, N. (1996). Molecular characterization of an enzyme that degrades neuromodulatory fatty-acid amides. Nature, 384(6604), 83-87.

Das, M., & Khanna, S. (1997). Clinicoepidemiological, toxicological, and safety evaluation studies on argemone oil. Critical Reviews in Toxicology, 27(3), 273-297.

Davies, S., Pertwee, R., & Riedel, G. (2002). Functions of cannabinoid receptors in the hippocampus. Neuropharmacology, 42(8), 993-1007.

Day, T., Rakhshan, F., Deutsch, D., & Barker, E. (2001). Role of fatty acid amide hydrolase in the transport of the endogenous cannabinoid anandamide. Molecular Pharmacology, 59(6), 1369-1375.

De Petrocellis, L., Cascio, M., & Di Marzo, V. (2004). The endocannabinoid system: A general view and latest additions. British Journal of Pharmacology, 141(5), 765-774.

De Vries, T., & Schoffelmeer, A. (2005). Cannabinoid CB1 receptors control conditioned drug seeking. Trends in Pharmacological Sciences, 26(8), 420-426.

Deadwyler, S., Hampson, R., Bennett, B., Edwards, T., Mu, J., Pacheco, M., Ward, S., & Childers, S., (1993). Cannabinoids modulate potassium current in cultured hippocampal neurons. Receptors and Channels, 1(1), 121-134.

171

Deadwyler, S., Hampson, R., Mu, J., Whyte, A., & Childers, S. (1995). Cannabinoids modulate voltage-sensitive potassium A-current in hippocampal-neurons via a camp-dependent process. Journal of Pharmacology and Experimental Therapeutics, 273(2), 734-743.

Demuth, D., & Molleman, A. (2006). Cannabinoid signaling. Life Sciences, 78(6), 549-563.

Derkinderen, P., Toutant, M., Burgaya, F., LeBert, M., Siciliano, J., deFranciscis, V., Gelman, M., & Girault, J. (1996). Regulation of a neuronal form of focal adhesion kinase by anandamide. Science, 273(5282), 1719-1722.

Derkinderen, P., Toutant, M., Kadare, G., Ledent, C., Parmentier, M., & Girault, J. (2001). Dual role of fyn in the regulation of FAK(+) 6,7 by cannabinoids in hippocampus. Journal of Biological Chemistry, 276(41), 38289-38296.

Deutsch, D., Glaser, S., Howell, J., Kunz, J., Puffenbarger, R., Hillard, C., & Anbumrad, N. (2001). The cellular uptake of anandamide is coupled to its breakdown by fatty-acid amide hydrolase. Journal of Biological Chemistry, 276(10), 6967-6973.

Devane, W., Dysarz, F., Johnson, M., Melvin, L. & Howlett, A. (1988). Determination and characterization of a cannabinoid receptor in rat-brain. Molecular Pharmacology, 34(5), 605-613.

Devane, W., Hanus, L., Breuer, A., Pertwee, R., Stevenson, L., Griffin, G., Gibson, D., Mandelbaum, A., Etinger, A, & Mechoulam, R. (1992). Isolation and structure of a brain constituent that binds to the cannabinoid receptor. Science, 258(5090), 1946-1949.

Dhopeshwarkar, A., Jain, S., Liao, C., Ghose, S., Bisset, K., & Nicholson, R. (2011). The actions of benzophenanthridine alkaloids, piperonyl butoxide and (S)-methoprene at the G-protein coupled cannabinoid CB1 receptor in vitro. European Journal of Pharmacology, 654(1), 26-32.

Di Marzo, V., Goparaju, S., Wang, L., Liu, J., Batkai, S., Jarai, Z., Fezza, F., Miura, G., Palmiter, R., Sugiura, T., & Kunos, G. (2001). Leptin-regulated endocannabinoids are involved in maintaining food intake Nature, 410(6830), 822-825.

DiMarzo, V., Fontana, A., Cadas. H., Schinelli, S., Cimino, G., Schwartz, J., & Piomelli, D. (1994). Formation and inactivation of endogenous cannabinoid anandamide in central neurons. Nature, 372(6507), 686-691.

Di Marzo, V., (2009). The endocannabinoid system: Its general strategy of action, tools for its pharmacological manipulation and potential therapeutic exploitation. Pharmacological Research, 60 (1), 77–84.

172

Di, S., Boudaba, C., Popescu, I., Weng, F., Harris, C., Marcheselli, V., Bazan, N., & Tasker, J. (2005). Activity-dependent release and actions of endocannabinoids in the rat hypothalamic supraoptic nucleus. Journal of Physiology-London, 569(3), 751-760.

Dinh, T., Carpenter, D., Leslie, F., Freund, T., Katona, I., Sensi, S., Kathuria, S., & Piomelli, D. (2002). Brain monoglyceride lipase participating in endocannabinoid inactivation. Proceedings of the National Academy of Sciences of the United States of America, 99(16), 10819-10824

Dinh, T., Kathuria, S., & Piomelli, D. (2004). RNA interference suggests a primary role for monoacylglycerol lipase in the degradation of the endocannabinoid 2- arachidonoylglycerol. Molecular Pharmacology, 66(5), 1260-1264.

Dowson, P., Bubb, J., & Lester, J. (1996). Persistence and degradation pathways of tributyltin in freshwater and estuarine sediments. Estuarine Coastal and Shelf Science, 42(5), 551-562.

Drake, D., Jensen, R., Busch-Petersen, J., Kawakami, J., Fernandez-Garcia, M., Fan, P., Makriyannis, T., & Tius, M. (1998). Classical/nonclassical hybrid cannabinoids: Southern aliphatic chain-functionalized C-6 beta methyl, ethyl, and propyl analogues. Journal of Medicinal Chemistry, 41(19), 3596-3608.

Drake, M., Shenoy, S., & Lefkowitz, R. (2006). Trafficking of G protein-coupled receptors. Circulation Research, 99(6), 570-582.

Duan, Y., Liao, C., Jain, S., & Nicholson, R. A. (2008). The cannabinoid receptor agonist CP-55,940 and ethyl arachidonate interfere with [3H]batrachotoxinin A 20 alpha-benzoate binding to sodium channels and inhibit sodium channel function. Comparative Biochemistry and Physiology C-Toxicology & Pharmacology, 148(3), 244-249.

Dubey, S., & Roy, U. (2003). Biodegradation of tributyltins (organotins) by marine bacteria. Applied Organometallic Chemistry, 17(1), 3-8.

Dunkley, P. R., Jarvie, P. E., & Robinson, P. J. (2008). A rapid percoll gradient procedure for preparation of synaptosomes. Nature Protocols, 3(11), 1718-1728.

Duvernay, M., Zhou, F., & Wu, G. (2004). A conserved motif for the transport of G protein-coupled receptors from the endoplasmic reticulum to the cell surface. Journal of Biological Chemistry, 279(29), 30741-30750.

Dvorak, Z., & Simanek, V. (2007). Metabolism of sanguinarine: The facts and the myths. Current Drug Metabolism, 8(2), 173-176.

Edgemond, W., Campbell, W., & Hillard, C. (1995). The binding of novel phenolic derivatives of anandamide to brain cannabinoid receptors. Prostaglandins, Leukotrienes and Essential Fatty Acids, 52(1), 83-86.

173

Eissenstat, M., Bell, M., Dambra, T., Alexander, E., Daum, S., Ackerman, J., Gruett, M., Kumar, V., Estep, K., Olefirowicz, E., Wetzel, J., Alexander, M., Weaver, J., Haycock, D., Luttinger, D., Casiano, F., Chippari, S., Kuster, J., Stevenson, J., & Ward, S. (1995). Aminoalkylindoles structure-activity-relationships of novel cannabinoid mimetics. Journal of Medicinal Chemistry, 38(16), 3094-3105.

Elmes, S., Jhaveri, M., Smart, D., Kendall, D., & Chapman, V. (2004). Cannabinoid CB2 receptor activation inhibits mechanically evoked responses of wide dynamic range dorsal horn neurons in naive rats and in rat models of inflammatory and neuropathic pain. European Journal of Neuroscience, 20(9), 2311-2320.

Eversole, L., Eversole, G., & Kopcik, J. (2000). Sanguinaria-associated oral leukoplakia - comparison with other benign and dysplastic leukoplakic lesions. Oral Surgery Oral Medicine Oral Pathology Oral Radiology and Endodontics, 89(4), 455-464.

Fattore, L., Spano, M. S., Deiana, S., Melis, V., Cossu, G., Fadda, P., & Fratta, W. (2007). An endocannabinoid mechanism in relapse to drug seeking: A review of animal studies and clinical perspectives. Brain Research Reviews, 53(1), 1-16.

Fay, J., Dunham, T., & Farrens, D. (2005). Cysteine residues in the human cannabinoid receptor: Only C257 and C264 are required for a functional receptor, and steric bulk at C386 impairs antagonist SR141716A binding. Biochemistry, 44(24), 8757-8769.

Fegley, D., Kathuria, S., Mercier, R., Li, C., Goutopoulos, A., Makriyannis, A., & Piomelli, D. (2004). Anandamide transport is independent of fatty-acid amide hydrolase activity and is blocked by the hydrolysis-resistant inhibitor AM1172. Proceedings of the National Academy of Sciences of the United States of America, 101(23), 8756-8761.

Felder, C., Briley, E., Axelrod, J., Simpson, J., Mackie, K., & Devane, W. (1993). Anandamide, an endogenous cannabimimetic eicosanoid, binds to the cloned human cannabinoid receptor and stimulates receptor-mediated signal-transduction. Proceedings of the National Academy of Sciences of the United States of America, 90(16), 7656-766.

Felder, C., Joyce, K., Briley, E., Glass, M., Mackie, K., Fahey, K., Cullinan, G., Hunder, D., Johnson, D., Chaney, M., Koppel, G., & Brownstein, M. (1998). LY320135, a novel cannabinoid CB1 receptor antagonist, unmasks coupling of the CB1 receptor to stimulation of cAMP accumulation. Journal of Pharmacology and Experimental Therapeutics, 284(1), 291-297.

Felder, C., Joyce, K., Briley, E., Mansouri, J., Mackie, K., Blond, O., Lai, Y., Ma, A., & Mitchell, R. (1995). Comparison of the pharmacology and signal-transduction of the human cannabinoid CB1 and CB2 receptors. Molecular Pharmacology, 48(3), 443-450

Feng, W., & Song, Z. (2003). Effects of D3.49A, R3.50A, and A6.34E mutations on ligand binding and activation of the cannabinoid-2 (CB2) receptor. Biochemical Pharmacology, 65(7), 1077-1085.

174

Fortier, M., Frouin, H., Cloutier, A., Arseneault, M., Ramassamy, C., Badiwa-Bizowe, B., St-Louis, R., Pelletier, E., and Fournier, M. (2010). Toxicological effects of mouse diet contaminated with tributyltin on the immune and neurological systems of C57BL/6 mice. Toxicological and Environmental Chemistry, 92(1), 927-945.

Franklin, M. (1976). Methylenedioxyphenyl insecticide synergists as potential human health-hazards. Environmental Health Perspectives, 14(1), 29-37.

Freund TF, Katona I, Piomelli D (2003). Role of endogenous cannabinoids in synaptic signaling. Physiological Reviews, 83 (3), 1017–66.

Fritze, O., Filipek, S., Kuksa, V., Palczewski, K., Hofmann, K., & Ernst, O. (2003). Role of the conserved NPxxY(x)(5,6)F motif in the rhodopsin ground state and during activation. Proceedings of the National Academy of Sciences of the United States of America, 100(5), 2290-2295.

Fukagawa, T., & Suzuki, S. (1993). Cloning of gene responsible for tributyltin chloride (tbtcl) resistance in tbtcl-resistant marine bacterium, alteromonas sp M-1. Biochemical and Biophysical Research Communications, 194(2), 733-740.

Gaertner, S., Balski, M., Koch, M., & Nehls, I. (2009). Analysis and migration of phthalates in infant food packed in recycled paperboard. Journal of Agricultural and Food Chemistry, 57(22), 10675-10681.

Gamaleddin, I., Wertheim, C., Zhu, X., Coen, K., Vemuri, K., Makryannis, A., Goldberg, S., & Le Foll, B. (2012). Cannabinoid receptor stimulation increases motivation for nicotine and nicotine seeking. Addiction Biology, 17(1), 47-61.

Gaoni, Y., & Mechoulam, R. (1964). Isolation structure and partial synthesis of active constituent of hashish. Journal of the American Chemical Society, 86(8), 1646-1649.

Garcia, D.E., Brown, S., Hille, B., & Mackie, K. (1998). Protein kinase C disrupts cannabinoid actions by phosphorylation of the CB1 cannabinoid receptor. Journal of Neuroscience, 18(1), 2834-2841.

Gareau, Y., Dufresne, C., Gallant, M., Rochette, C., Sawyer, N., Slipetz, D., Tremblay, N., Weech, P., Metters, K., & Labelle, M. (1996). Structure activity relationships of tetrahydrocannabinol analogues on human can nabinoid receptors. Bioorganic & Medicinal Chemistry Letters, 6(2), 189-194.

Gatley, S., Lan., R., Pyatt, B., Gifford, A., Volkow, N., & Makriyannis, A. (1997). Binding of the non-classical cannabinoid CP55940 and the diarylpyrazole AM251 to rodent brain cannabinoid receptors. Life Sciences, 61(14), 191-197.

Gebremedhin, D., Lange, A., Campbell, W., Hillard, C., & Harder, D. (1999). Cannabinoid CB1 receptor of cat cerebral arterial muscle functions to inhibit L-type Ca2+ channel current. American Journal of Physiology-Heart and Circulatory Physiology, 276(6), 2085-2093.

175

Georgieva, T., Devanathan, S., Stropova, D., Park, C. K., Salamon, Z., Tollin, G., Hruby, V., Roeske, W., Yamamura, H., & Varga, E. (2008). Unique agonist-bound cannabinoid CB1 receptor conformations indicate agonist specificity in signaling. European Journal of Pharmacology, 581(1-2), 19-29.

Gerard, C., Mollereau, C., Vassart, G., & Parmentier, M. (1990). Nucleotide-sequence of a human cannabinoid receptor cDNA. Nucleic Acids Research, 18(23), 7142-7142.

Gifford, A., & Ashby, C. (1996). Electrically evoked acetylcholine release from hippocampal slices is inhibited by the cannabinoid receptor agonist, WIN 55212-2, and is potentiated by the cannabinoid antagonist, SR141716A. Journal of Pharmacology and Experimental Therapeutics, 277(3), 1431-1436.

Gill, E., Paton, W., & Pertwee, R. (1970). Preliminary experiments on chemistry and pharmacology of cannabis. Nature, 228(5267), 134-136.

Glaser, S., Abumrad, N., Fatade, F., Kaczocha, M., Studholme, K., & Deutsch, D. (2003). Evidence against the presence of an anandamide transporter. Proceedings of the National Academy of Sciences of the United States of America, 100(7), 4269-4274.

Glass, M., Dragunow, M., & Faull, R. (1997). Cannabinoid receptors in the human brain: A detailed anatomical and quantitative autoradiographic study in the fetal, neonatal and adult human brain. Neuroscience, 77(2), 299-318.

Glass, M., & Felder, C. (1997). Concurrent stimulation of cannabinoid CB1 and dopamine D2 receptors augments cAMP accumulation in striatal neurons: Evidence for a G(s) linkage to the CB1 receptor. Journal of Neuroscience, 17(14), 5327-5333.

Glass, M., & Northup, J. (1999). Agonist selective regulation of G proteins by cannabinoid CB1 and CB2 receptors. Molecular Pharmacology, 56(6), 1362-1369.

Godino del Carmen, M., Torres, M., & Sanchez-Prieto, J. (2007). CB1 receptors diminish both Ca2+ influx and glutamate release through two different mechanisms active in distinct populations of cerebrocortical nerve terminals. Journal of Neurochemistry, 101(6), 1471-1482.

Gouldson, P., Calandra, B., Legoux, P., Kerneis, A., Rinaldi-Carmona, M., Barth, F., Le Fur, G., Ferrara, P., & Shire, D. (2000). Mutational analysis and molecular modelling of the antagonist SR144528 binding site on the human cannabinoid CB2 receptor. European Journal of Pharmacology, 401(1), 17-25.

Goutopoulos, A., Fan, P., Khanolkar, A., Xie, X., Lin, S., & Makriyannis, A. (2001). Stereochemical selectivity of methanandamides for the CB1 and CB2 cannabinoid receptors and their metabolic stability. Bioorganic & Medicinal Chemistry, 9(7), 1673-1684.

176

Grasso, P., Heindel, J., Powell, C., & Reichert, L. Jr., (1993). Effects of mono (2-ethylhexylphthalate, a testicular toxicant, on follicle-stimulating hormone binding to membranes from cultured rat Sertoli cells. Biology of Reproduction, 48(1), 454-459.

Griffin, G., Atkinson, P., Showalter, V., Martin, B., & Abood, M. (1998). Evaluation of cannabinoid receptor agonists and antagonists using the guanosine-5 '-O-(3-[35S] thio)-triphosphate binding assay in rat cerebellar membranes. Journal of Pharmacology and Experimental Therapeutics, 285(2), 553-560.

Gullapalli, S., Amrutkar, D., Gupta, S., Kandadi, M. R., Kumar, H., Gandhi, M., Karande, V., & Narayanan, S. (2010). Characterization of active and inactive states of CB1 receptor and the differential binding state modulation by cannabinoid agonists, antagonists and inverse agonists. Neuropharmacology, 58(8), 1215-1219.

Guo, J., & Ikeda, S. (2004). Endocannabinoids modulate N-type calcium channels and G-protein-coupled inwardly rectifying potassium channels via CB1 cannabinoid receptors heterologously expressed in mammalian neurons. Molecular Pharmacology, 65 (3): 665–74.

Hall, S., Roberts, K., & Vaidehi, N. (2009). Position of helical kinks in membrane protein crystal structures and the accuracy of computational prediction. Journal of Molecular Graphics & Modelling, 27(8), 944-950.

Hampson, A., Bornheim, L., Scanziani, M., Yost, C., Gray, A., Hansen, B., Leonoudakis, D., & Bickler, P. (1998). Dual effects of anandamide on NMDA receptor-mediated responses and neurotransmission. Journal of Neurochemistry, 70(2), 671-676.

Hampson, R., Evans, G., Mu, J., Zhuang, S., King, V., Childers, S., & Deadwyler, S. (1995). Role of cyclic-AMP-dependent protein-kinase in cannabinoid receptor modulation of potassium A-current in cultured rat hippocampal-neurons. Life Sciences, 56(23-24), 2081-2088.

Hanus, L., Abu-Lafi, S., Fride, E., Breuer, A., Vogel, Z., Shalev, D., Kustanovich, I., & Mechoulam, R. (2001). 2-arachidonyl glyceryl ether, an endogenous agonist of the cannabinoid CB1 receptor. Proceedings of the National Academy of Sciences of the United States of America, 98(7), 3662-3665.

Hanus, L., Gopher, A., Almog, S., & Mechoulam, R. (1993). 2 new unsaturated fatty-acid ethanolamides in brain that bind to the cannabinoid receptor. Journal of Medicinal Chemistry, 36(20), 3032-3034.

Harrison, C., & Traynor, J. (2003). The [35S]GTPγS binding assay: Approaches and applications in pharmacology. Life Sciences, 74(4), 489-508.

Hashimotodani, Y., Ohno-Shosaku, T., & Kano, M. (2007). Ca2+-assisted receptor-driven endocannabinoid release: Mechanisms that associate presynaptic and postsynaptic activities. Current Opinion in Neurobiology, 17(3), 360-365.

177

Hashimotodani, Y., Ohno-Shosaku, T., & Kano, M. (2007). Presynaptic monoacylglycerol lipase activity determines basal endocannabinoid tone and terminates retrograde endocannabinoid signaling in the hippocampus. Journal of Neuroscience, 27(5), 1211-1219.

Hashimotodani, Y., Ohno-Shosaku, T., Watanabe, M., & Kano, M. (2007). Roles of phospholipase C beta and NMDA receptor in activity-dependent endocannabinoid release. Journal of Physiology-London, 584(2), 373-380.

Hawkins, D., Westin, K., Chasseaud, L., & Franklin, E. (1977). Fate of methoprene (isopropyl (2e, 4e)-11-methoxy-3,7,11-trimethyl-2,4-dodecadienoate) in rats. Journal of Agricultural and Food Chemistry, 25(2), 398-403.

Henrick, C. (2007). Methoprene. Journal of American Mosquito Control Association, 23(1), 225-239.

Henry, D., & Chavkin, C. (1995). Activation of inwardly rectifying potassium channels (GIRK1) by coexpressed rat-brain cannabinoid receptors in Xenopus oocytes. Neuroscience Letters, 186(2-3), 91-94.

Herbert, J., Augereau, J., Gleye, J., & Maffrand, J. (1990). Chelerythrine is a potent and specific inhibitor of protein kinase C. Biochemical and Biophysical Research communication, 172(1), 993-999.

Herkenham, M., Lynn, A., Johnson, M., Melvin, L., Decosta, B., & Rice, K. (1991). Characterization and localization of cannabinoid receptors in rat-brain -a quantitative in vitro autoradiographic study. Journal of Neuroscience, 11(2), 563-583.

Herkenham, M., Lynn, A., Little, M., Johnson, M., Melvin, L., Decosta, B., & Rice, K. (1990). Cannabinoid receptor localization in brain. Proceedings of the National Academy of Sciences of the United States of America, 87(5), 1932-1936.

Heudorf, U., Mersch-Sundermann, V., & Angerer J. (2007). Phthalates: Toxicology and exposure. International Journal of Hygiene and Environmental Helath. 210(1), 623–634.

Hill, M., & Gorzalka, B. (2005). Pharmacological enhancement of cannabinoid CB1 receptor activity elicits an antidepressant-like response in the rat forced swim test. European Neuropsychopharmacology, 15(6), 593-599.

Hillard, C., & Bloom, A. (1983). Possible role of prostaglandins in the effects of the cannabinoids on adenylate-cyclase activity. European Journal of Pharmacology, 91(1), 21-27.

Hillard, C., Edgemond, W., Jarrahian, A., & Campbell, W. (1997). Accumulation of N-arachidonoylethanolamine (anandamide) into cerebellar granule cells occurs via facilitated diffusion. Journal of Neurochemistry, 69(2), 631-638.

178

Hillard, C., Harris, R., & Bloom, A. (1985). Effects of the cannabinoids on physical-properties of brain membranes and phospholipid-vesicles fluorescence studies. Journal of Pharmacology and Experimental Therapeutics, 232(3), 579-588.

Hillard, C., & Jarrahian, A. (2000). The movement of N-arachidonoylethanolamine (anandamide) across cellular membranes. Chemistry and Physics of Lipids, 108(1-2), 123-134

Hillard, C., & Jarrahian, A. (2003). Cellular accumulation of anandamide: Consensus and controversy. British Journal of Pharmacology, 140(5), 802-808.

Hillard, C., Manna, S., Greenberg, M., Dicamelli, R., Ross, R., Stevenson, L., Murphy, V., Pertwee, R., & Campbell, W. (1999). Synthesis and characterization of potent and selective agonists of the neuronal cannabinoid receptor (CB1). Journal of Pharmacology and Experimental Therapeutics, 289(3), 1427-1433.

Hoffman, D., & Johnston, D. (1998). Downregulation of transient K+ channels in dendrites of hippocampal CA1 pyramidal neurons by activation of PKA and PKC. Journal of Neuroscience, 18(1), 3521–3528.

Hoffman, A., & Lupica, C. (2000) Mechanisms of cannabinoid inhibition of GABA(A) synaptic transmission in the hippocampus. Journal of Neuroscience, 20(7), 2470-2479.

Hoffman, D., Magee, J., Colbert, C. & Johnston, D. (1997). K+ channel regulation of signal propagation in dendrites of hippocampal pyramidal neurons. Nature, 387(1), 869–875.

Hoffman, A., Oz, M., Caulder, T., & Lupica, C. (2003). Functional tolerance and blockade of long-term depression at synapses in the nucleus accumbens after chronic cannabinoid exposure. Journal of Neuroscience, 23(12), 4815-4820.

Hohmann, A., & Suplita, R. (2006). Endocannabinoid mechanisms of pain modulation. AAPS Journal, 8(4), 693-708.

Hollister L. (2001). Marijuana (Cannabis) as medicine. Journal of cannabis therapeutics. 1(1), 5-27.

Hosohata, K., Quock, R., Hosohata, Y., Burkey, T., Makriyannis, A., Consroe, P., Roeske, W., & Yamamura, H. (1997a). AM630 is a competitive cannabinoid receptor antagonist in the guinea pig brain. Life Sciences, 61(9), 115-118

Hosohata, Y., Quock, R., Hosohata, K., Makriyannis, A., Consroe, P., Roeske, W., & Yamamura, H. (1997b). AM630 antagonism of cannabinoid-stimulated [35S]GTPγS binding in the mouse brain. European Journal of Pharmacology, 321(1), 1-3.

Houston, D., & Howlett, A. (1998). Differential receptor-G-protein coupling evoked by dissimilar cannabinoid receptor agonists. Cellular Signalling, 10(9), 667-674.

179

Houston, D. & Howlett, A., (1993). Solubilization of the cannabinoid receptor from rat brain and its functional interaction with guanine nucleotide-binding proteins. Molecular Pharmacology, 43(1), 17-22.

Howlett, A. (1985). Cannabinoid inhibition of adenylate-cyclase biochemistry of the response in neuroblastoma cell membranes. Molecular Pharmacology, 27(4), 429-436.

Howlett, A. (2004). Efficacy in CB1 receptor-mediated signal transduction. British Journal of Pharmacology, 142(8), 1209-1218.

Howlett, A.C. (2005). Cannabinoid receptor signaling. Handbook of Experimental Pharmacology, 168(1), 53-79.

Howlett, A., Barth, F., Bonner, T., Cabral, G., Casellas, P., Devane, W., Felder, C., Herkenham, M., Mackie, K., Martin, B., Mechoulam, R., & Pertwee, R. (2002). International Union of Pharmacology. XXVII. Classification of cannabinoid receptors. Pharmacological Reviews, 54(2), 161-202.

Howlett, A., & Fleming, R. (1984). Cannabinoid inhibition of adenylate-cyclase: pharmacology of the response in neuroblastoma cell membranes. Molecular Pharmacology, 26(3), 532-538.

Howlett, A., Qualy, J., & Khachatrian, L. (1986). Involvement of Gi in the inhibition of adenylate-cyclase by cannabimimetic drugs. Molecular Pharmacology, 29(3), 307-313.

Howlett, A., Wilken, G., Pigg, J., Houston, D., Lan, R., Liu, Q., & Makriyannis, A. (2000). Azido- and isothiocyanato-substituted aryl pyrazoles bind covalently to the CB1 cannabinoid receptor and impair signal transduction. Journal of Neurochemistry, 74(5), 2174-2181.

Hsieh, C., Brown, S., Derleth, C., & Mackie, K. (1999). Internalization and recycling of the CB1 cannabinoid receptor. Journal of Neurochemistry, 73(2), 493-501.

Huang, C., Lo, S., & Hsu, K. (2001). Presynaptic mechanisms underlying cannabinoid inhibition of excitatory synaptic transmission in rat striatal neurons. Journal of Physiology-London, 532(3), 731-748.

Huang, S., Bisogno, T., Trevisani, M., Al-Hayani, A., De Petrocellis, L., Fezza, F., Tognetto, M., Petros, T., Krey, J., Chu, C., Miller, J., Davies, S., Geppetti, P., Walker, J. & Di Marzo, V. (2002). An endogenous capsaicin-like substance with high potency at recombinant and native vanilloid VR1 receptors. Proceedings of the National Academy of Sciences of the United States of America, 99(12), 8400-8405.

Huffman, J., Bushell, S., Miller, J., Wiley, J., & Martin, B. (2002). 1-methoxy-, 1-deoxy-11-hydroxy- and 11-hydroxy-1-methoxy-∆8-tetrahydrocannabinols: New selective ligands for the CB2 receptor. Bioorganic & Medicinal Chemistry, 10(12), 4119-4129.

180

Huffman, J., Dai, D., Martin, B., & Compton, D. (1994). Design, synthesis and pharmacology of cannabimimetic indoles. Bioorganic & Medicinal Chemistry Letters, 4(4), 563-566

Huffman, J., Mabon, R., Wu, M., Lu, J., Hart, R., Hurst, D., Reggio, P., Wiley, J., & Martin, B. (2003). 3-indolyl-1-naphthylmethanes: New cannabimimetic indoles provide evidence for aromatic stacking interactions with the CB1 cannabinoid receptor. Bioorganic & Medicinal Chemistry, 11(4), 539-549.

Hurst, D., Lynch, D., Barnett-Norris, J., Hyatt, S., Seltzman, H., Zhong, M., Song, Z., Nie, J, Lewis, D., & Reggio, P. (2002). N-(piperidin-1-yl)-5-(4-chlorophenyl)-1-(2,4-dichlorophenyl)-4-methyl-1H-pyrazole-3-carboxamide (SR141716A) interaction with LYS3.28(192) is crucial for its inverse agonism at the cannabinoid CB1 receptor. Molecular Pharmacology, 62(6), 1274-1287.

Hurst, D., Umejiego, U., Lynch, D., Seltzman, H., Hyatt, S., Roche, M., McAllister, S., Fleischer, D., Kapur, A., Abood, M., Shi, S., Jones, J., Lewis, D., & Reggio, P. (2006). Biarylpyrazole inverse agonists at the cannabinoid CB1 receptor: Importance of the C-3 carboxamide oxygen/lysine3.28 (192) interaction. Journal of Medicinal Chemistry, 49(20), 5969-5987.

Hynes, J., Leftheris, K., Wu, H., Pandit, C., Chen, P., Norris, D., Chen, P., Norris, D., Chen, B., Zhao, R.,, Kiener, P., Chen, X., Turk, L., Patil-Koota, V., Gillooly, K, Shuster, D., & McIntyre, K. (2002). C-3 Amido-indole cannabinoid receptor modulators. Bioorganic & Medicinal Chemistry Letters, 12(17), 2399-2402.

Ibrahim, M., Deng, H., Zvonok, A., Cockayne, D., Kwan, J., Mata, H., Vanderah, T., Lai, J., Porreca, F., Makriyannis, A., & Malan, T. (2003). Activation of CB2 cannabinoid receptors by AM1241 inhibits experimental neuropathic pain: Pain inhibition by receptors not present in the CNS. Proceedings of the National Academy of Sciences of the United States of America, 100(18), 10529-10533.

Ishac, E., Jiang, L., Lake, K., Varga, K., Abood, M., & Kunos, G. (1996). Inhibition of exocytotic noradrenaline release by presynaptic cannabinoid CB1 receptors on peripheral sympathetic nerves. British Journal of Pharmacology, 118(8), 2023-2028.

Isokawa, M., & Alger, B. (2006). Ryanodine receptor regulates endogenous cannabinoid mobilization in the hippocampus. Journal of Neurophysiology, 95(5), 3001-3011.

Iverson, L. (2003). Cannabis and the brain. Brain, 126(1), 1252-1270.

Jee, J., Koo, J., Keum, Y., Park, K., Choi S., & Kang, J., (2009). Effects of dibutyl phthalate and di-ethylhexyl phthalate on acetylcholinesterase activity in bagrid catfish, Pseudobagrus fulvidraco (Richardson). Journal of Applied Ichthyology, 25(1), 771–775.

Jennings, E., Vaughan, C., & Christie, M. (2001). Cannabinoid actions on rat superficial medullary dorsal horn neurons in vitro. Journal of Physiology-London, 534(3), 805-812.

181

Jiang, H., Li, X., Zhao, Y., Ferguson, D. K., Hueber, F., Bera, S., Wang, Y, Zhao, L., Liu, C., & Li, C. (2006). A new insight into Cannabis sativa (cannabaceae) utilization from 2500-year-old Yanghai tombs, Xinjiang, China. Journal of Ethnopharmacology, 108(3), 414-422.

Jin, W., Brown, S., Roche, J., Hsieh, C., Celver, J., Kovoor, A., Chavkin, C., & Mackie, K. (1999). Distinct domains of the CB1 cannabinoid receptor mediate desensitization and internalization. Journal of Neuroscience, 19(10), 3773-3780.

Johnson M., & Melvin L. (1986). The discovery of nonclassical cannabinoid analgetics in Cannabinoids as Therapeutic Agents, ed Mechoulam R., (CRC, Boca Raton, FL), pp 121–145.

Jones, D. (1998). Piperonyl Butoxide: The Insecticide Synergist. Academic Press.

Jursky, F., & Baliova, M. (2011). Differential effect of the benzophenanthridine alkaloids sanguinarine and chelerythrine on glycine transporters. Neurochemistry International, 58(6), 641-647.

Kannan, K., Senthilkumar, K., & Giesy, J. (1999). Occurrence of butyltin compounds in human blood. Environmental Science & Technology, 33(10), 1776-1779.

Kano, M., Ohno-Shosaku, T., Hashimotodani, Y., Uchigashima, M., & Watanabe, M. (2009). Endocannabinoid-mediated control of synaptic transmission. Physiological Reviews, 89(1), 309-380.

Kapur, A., Hurst, D. P., Fleischer, D., Whitnell, R., Thakur, G. A., Makriyannis, A., Reggio, P., & Abood, M. (2007). Mutation studies of Ser7.39 and Ser2.60 in the human CB1 cannabinoid receptor: Evidence for a serine-induced bend in CB1 transmembrane helix 7. Molecular Pharmacology, 71(6), 1512-1524.

Karlsson, M., Contreras, J., Hellman, U., Tornqvist, H., & Holm, C. (1997). cDNA cloning, tissue distribution, and identification of the catalytic triad of monoglyceride lipase evolutionary relationship to esterases, lysophospholipases, haloperoxidases. Journal of Biological Chemistry, 272(43), 27218-27223.

Katoch-Rouse, R., Pavlova, O., Caulder, T., Hoffman, A., Mukhin, A., & Horti, A. (2003). Synthesis, structure-activity relationship, and evaluation of SR141716 analogues: Development of central cannabinoid receptor ligands with lower lipophilicity. Journal of Medicinal Chemistry, 46(4), 642-645.

Katona, I., Sperlagh, B., Sik, A., Kafalvi, A., Vizi, E., Mackie, K., & Freund F. (1999). Pre-synaptic located CB1 cannabinoid receptors regulate GABA release form axon terminals of specific hippocampal interneurons. Journal of Neuroscience, 19(1), 4555-4558.

Katona, I., Urban, G., Wallace, M., Ledent, C., Jung, K., Piomelli, D., Mackie, K., & Freund, T. (2006). Molecular composition of the endocannabinoid system at glutamatergic synapses. Journal of Neuroscience, 26(21), 5628-5637.

182

Katona, I., & Freund, T. (2008). Endocannabinoid signaling as a synaptic circuit breaker in neurological disease. Nature Medicine, 14(9), 923-930.

Kavlock, R., Boekelheide, K., Chapin, R., Cunningham, M., Faustman, E., Foster, P., Golub, M., Henderson, R., Hinberg, I., Little, R., Seed, J., Shea, K., Tabacova, S., Tyl, R., Williams, P., & Zacharewski, T. (2002a). NTP Center for the Evaluation of Risks to Human Reproduction: phthalates expert panel report on the reproductive and developmental toxicity of di (2-ethylhexyl) phthalate. Reproductive Toxicology, 16(1), 529-653.

Kawamura, Y., Fukaya, M., Maejima, T., Yoshida, T., Miura, E., Watanabe, M., Ohno-Shosaku, T., & Kano, M. (2006) The CB1 cannabinoid receptor is the major cannabinoid receptor at excitatory presynaptic sites in the hippocampus and cerebellum. Journal of Neuroscience, 26(11), 2991-3001.

Kearn, C., Blake-Palmer, K., Daniel, E., Mackie, K., & Glass, M. (2005). Concurrent stimulation of cannabinoid CB1 and dopamine D2 receptors enhances heterodimer formation: A mechanism for receptor cross-talk? Molecular Pharmacology, 67(5), 1697-1704.

Kelly, D. L., Gorelick, D. A., Conley, R. R., Boggs, D. L., Linthicum, J., Liu, F., Feldman, S., Ball, P., Wehring, H., McMahon, R., Huestis, M., Heischman, S., Warren, K., & Buchanan, R. W. (2011). Effects of the cannabinoid-1 receptor antagonist rimonabant on psychiatric symptoms in overweight people with schizophrenia A randomized, double-blind, pilot study. Journal of Clinical Psychopharmacology, 31(1), 86-91.

Kenakin, T. (2007). Functional selectivity through protean and biased agonism: Who steers the ship? Molecular Pharmacology, 72(6), 1393-1401.

Keseru, G., Kolossvary, I., & Bertok, B. (1999). Piperonyl butoxide-mediated inhibition of cytochrome P450-catalysed insecticide metabolism: A rational approach. Pesticide Science, 55(10), 1004-1006.

Khanolkar, A., Palmer, S., & Makriyannis, A. (2000). Molecular probes for the cannabinoid receptors. Chemistry and Physics of Lipids, 108(1-2), 37-52.

Kim, B., Cho, S., Kim, Y., Shin, M., Yoo, H., Kim, J., Yang, Y., Kim, H., Bhang, S., & Hong, Y. (2009). Phthalates exposure and attention-Deficit/Hyperactivity disorder in school-age children. Biological Psychiatry, 66(10), 958-963.

Kim, S., Won, S., Mao, X.O., Jin, K., & Greenberg, D. (2006). Molecular mechanisms of cannabinoid protection from neuronal excitotoxicity. Molecular Pharmacology, 69(1), 691-696.

Kim, H.I., Kim, T.H., Shin, Y.K., Lee, C.S., Park, M., & Song, J. (2005). Anandamide suppression of Na+ currents in rat dorsal root ganglion neurons. Brain Research, 1062(1), 39-47.

183

Kobayashi, Y., Arai, S., Waku, K., & Sugiura, T. (2001). Activation by 2-arachidonoylglycerol, an endogenous cannabinoid receptor ligand, of p42/44 mitogen-activated protein kinase in HL-60 cells. Journal of Biochemistry, 129(5), 665-669.

Konno, N., Tsunoda, M., Nakano, K., Liu, Y. (2001). Effect of tributyltin on the N-methyl-D-aspartate (NMDA) receptors in the mouse brain. Archives of Toxicology, 75(1), 549-554.

Kreitzer, A., Carter, A., & Regehr, W. (2002). Inhibition of interneuron firing extends the spread of endocannabinoid signaling in the cerebellum. Neuron, 34(5), 787-796.

Kreitzer, A., & Regehr, W. (2001). Retrograde inhibition of presynaptic calcium influx by endogenous cannabinoids at excitatory synapses onto Purkinje cells. Neuron, 29(3), 717-727.

Kreitzer, A., & Regehr, W. (2002). Retrograde signaling by endocannabinoids. Current Opinion in Neurobiology, 12(3), 324-330.

Krishnamurthy, M., Ferreira, A., & Moore, B. (2003). Synthesis and testing of novel phenyl substituted side-chain analogues of classical cannabinoids. Bioorganic & Medicinal Chemistry Letters, 13(20), 3487-3490.

Kriwacki, R., & Makriyannis, A. (1989). The conformational-analysis of ∆9-tetrahydrocannabinols and ∆9,11-tetrahydrocannabinols in solution using high-resolution nuclear magnetic-resonance spectroscopy. Molecular Pharmacology, 35(4), 495-503.

Kroeze, W., Sheffler, D., & Roth, B. (2003). G-protein-coupled receptors at a glance. Journal of Cell Science, 116(24), 4867-4869.

Kushmerick, C., Price, G.D., Taschenberger, H., Puente, N., Renden, R., Wadiche, J., Duvoisin, R., Grandes, P. Von Gersdorff, H., (2004). Retroinhibition of pre-synaptic Ca2+ currents by endocannabinoids released via postsynaptic mGluR activation at a Calyx synapse. Journal of Neuroscience, 24(1), 5955-5965.

Lambert, D., & Muccioli, G. (2007). Endocannabinoids and related N-acylethanolamines in the control of appetite and energy metabolism: Emergence of new molecular players. Current Opinion in Clinical Nutrition and Metabolic Care, 10(6), 735-744.

Lan, R., Liu, Q., Fan, P., Lin, S., Fernando, S., McCallion, D., Pertwee, R., & Makriyannis, A. (1999). Structure-activity relationships of pyrazole derivatives as cannabinoid receptor antagonists. Journal of Medicinal Chemistry, 42(4), 769-776.

184

Lange, J., Coolen, H., van Stuivenberg, H., Dijksman, J., Herremans, A., Ronken, E., Keizer, H., Tipker, K., McCreary, A., Veerman, W., Wals, H., Stork, B., Verveer, P., den Hartog, A., de Jong, N., Adolfs, T., Hoogendoorn, J., & Kruse, C. (2004). Synthesis, biological properties, and molecular modeling investigations of novel 3, 4-diarylpyrazolines as potent and selective CB1 cannabinoid receptor antagonists. Journal of Medicinal Chemistry, 47(3), 627-643.

Lauckner, J., Jensen, J., Chen, H., Lu, H., Hille, B., & Mackie, K. (2008). GPR55 is a cannabinoid receptor that increases intracellular calcium and inhibits M current. Proceedings of the National Academy of Sciences of the United States of America, 105(7), 2699-2704.

Le Foll, B., & Goldberg, S. (2004). Rimonabant, a CB1 antagonist, blocks nicotine-conditioned place preferences. Neuroreport, 15(13), 2139-2143.

Lee, C., Zaugg, H., Michaels, R., Dren, A., Plotnikoff, N., & Young, P. (1983). New azacannabinoids highly-active in the central nervous-system. Journal of Medicinal Chemistry, 26(2), 278-280.

Lee, S., Qing, W., Mar, W., Luyengi, L., Mehta, R., Kawanishi, K., Fong, H.., Beecher, C., Kinghorn, A., Pezzuto, J., (1998). Angoline and chelerythrine, benzophenanthridine alkaloids that do not inhibit protein kinase C. Journal of Biological Chemistry, 273(1), 19829-19833.

Levenes, C., Daniel, H., Soubrie, P., & Crepel, F. (1998). Cannabinoids decrease excitatory synaptic transmission and impair long-term depression in rat cerebellar Purkinje cells. Journal of Physiology-London, 510(3), 867-879.

Li, H. (1973). An archaeological and historical account of cannabis in china. Springer New York.

Liao, C., Zheng, J., David, L., & Nicholson, R. (2004). Inhibition of voltage-sensitive sodium channels by the cannabinoid-1 receptor antagonist AM251 in mammalian brain. Basic & Clinical Pharmacology & Toxicology, 94(2), 73-78.

Ligresti, A., Petrosino, S., & Di Marzo, V., (2009). From endocannabinoid profiling to ‘endocannabinoid therapeutics’. Current Opinion in Chemical Biology, 13(1), 321-331.

Lin, S., Khanolkar, A., Fan, P., Goutopoulos, A., Qin, C., Papahadjis, D., & Makriyannis, A. (1998). Novel analogues of arachidonylethanolamide (anandamide): Affinities for the CB1 and CB2 cannabinoid receptors and metabolic stability. Journal of Medicinal Chemistry, 41(27), 5353-5361.

Liu, P., Tseng, F., & Liu, J. (2009). Comparative suppression of phthalate monoesters and phthalate diesters on calcium signalling coupled to nicotinic acetylcholine receptors. Journal of Toxicological Sciences, 34(1), 255-263.

185

Little, P., Compton, D., Johnson, M., Melvin, L., & Martin, B. (1988). Pharmacology and stereoselectivity of structurally novel cannabinoids in mice. Journal of Pharmacology and Experimental Therapeutics, 247(3), 1046-1051.

Llano, I., Leresche, N., & Marty, A. (1991). Calcium entry increases the sensitivity of cerebellar Purkinje-cells to applied GABA and decreases inhibitory synaptic currents. Neuron, 6(4), 565-574.

Lu, K., Tseng, F., Wu, C., & Liu, P. (2004). Suppression by phthalates of the calcium signaling of human nicotinic acetylcholine receptors in human neuroblastoma SH-SY5Y cells. Toxicology, 200(1), 113-121.

Ludanyi, A., Eross, L., Czirjak, S., Vajda, J., Halasz, P., Watanabe, M., Palkovits, M., Magloczky, Z., Freund, T.F., & Katona, I. (2008). Downregulation of the CB1 cannabinoid receptor and related molecular elements of the endocannabinoid system in epileptic human hippocampus. Journal of Neuroscience, 28(1), 2976-2990.

Maccarrone, M., Bari, M., Lorenzon, T., Bisogno, T., Di Marzo, V., & Finazzi-Agro, A. (2000). Anandamide uptake by human endothelial cells and its regulation by nitric oxide. Journal of Biological Chemistry, 275(18), 13484-13492.

Maccarrone, M., van der Stelt, M., Rossi, A., Veldink, G., Vliegenthart, J., & Agrio, A. (1998). Anandamide hydrolysis by human cells in culture and brain. Journal of Biological Chemistry, 273(48), 32332-32339.

Mackie, K. (2006). Cannabinoid receptors as therapeutic targets. Annual Review of Pharmacology and Toxicology, 46, 101-122.

Mackie, K., Devane, W., & Hille, B. (1993). Anandamide, an endogenous cannabinoid, inhibits calcium currents as a partial agonist in N18 neuroblastoma-cells. Molecular Pharmacology, 44(3), 498-503.

Mackie, K., & Hille, B. (1992). Cannabinoids inhibit N-type calcium channels in neuroblastoma glioma-cells. Proceedings of the National Academy of Sciences of the United States of America, 89(9), 3825-3829.

Mackie, K., Lai, Y., Westenbroek, R., & Mitchell, R. (1995). Cannabinoids activate an inwardly rectifying potassium conductance and inhibit Q-type calcium currents in ATT20 cells transfected with rat-brain cannabinoid receptor. Journal of Neuroscience, 15(10), 6552-6561.

Maejima, T., Hashimoto, K., Yoshida, T., Aiba, A., & Kano, M. (2001). Presynaptic inhibition caused by retrograde signal from metabotropic glutamate to cannabinoid receptors. Neuron, 31(3), 463-475.

186

Maejima, T., Oka, S., Hashimotodani, Y., Ohno-Shosaku, T., Aiba, A., Wu, D., Waku, K., Suguira, T., & Kano, M. (2005). Synaptically driven endocannabinoid release requires Ca2+-assisted metabotropic glutamate receptor subtype 1 to phospholipase C beta 4 signaling cascade in the cerebellum. Journal of Neuroscience, 25(29), 6826-6835.

Magloczky, Z., Freund, T., & Katona, I. (2008). Downregulation of the CB1 cannabinoid receptor and related molecular elements of the endocannabinoid system in epileptic human hippocampus. Journal of Neuroscience, 28(1), 2976-2990.

Mailleux, P., & Vanderhaeghen, J. (1992). Distribution of neuronal cannabinoid receptor in the adult-rat brain - a comparative receptor-binding autoradiography and in-situ hybridization histochemistry. Neuroscience, 48(3), 655-668.

Maiti, M., Nandi, R., & Chaudhari, K. (1982). Sanguinarine - a monofunctional intercalating alkaloid. FEBS Letters, 142(2), 280-284.

Maiti, M., & Kumar, G. S. (2007). Molecular aspects on the interaction of protoberberine, benzophenanthridine, and aristolochia group of alkaloids with nucleic acid structures and biological perspectives. Medicinal Research Reviews, 27(5), 649-695.

Malan, T., Ibrahim, M., Deng, H., Liu, Q., Mata, H., Vanderah, T., Porreca, F., & Makriyannis, A. (2001). CB2 cannabinoid receptor-mediated peripheral antinociception. Pain, 93(3), 239-245.

Maldonado, R., Valverde, O., & Berrendero, F. (2006). Involvement of the endocannabinoid system in drug addiction. Trends in Neurosciences, 29(4), 225-232.

Maneuf, Y., & Brotchie, J. (1997). Paradoxical action of the cannabinoid WIN 55,212-2 in stimulated and basal cyclic AMP accumulation in rat globus pallidus slices. British Journal of Pharmacology, 120(8), 1397-1398.

Marin, S., Marco, E., Biscaia, M., Fernandez, B., Rubio, M., Guaza, C., Schmidhammer, H., & Viveros, M. (2003). Involvement of the kappa-opioid receptor in the anxiogenic-like effect of CP55940 in male rats. Pharmacology Biochemistry and Behavior, 74(3), 649-656.

Martin, B., Jefferson, R., Winckler, R., Wiley, J., Thomas, B., Crocker, P., Williams, W., & Razdan, R. (2002). Assessment of structural commonality between tetrahydrocannabinol and anandamide. European Journal of Pharmacology, 435(1), 35-42.

Mato, S., Lafourcade, M., Robbe, D., Bakiri, Y., & Manzoni, O. (2008). Role of the cyclic-AMP/PKA cascade and of P/Q-type Ca++ channels in endocannabinoid-mediated long-term depression in the nucleus accumbens. Neuropharmacology, 54(1), 87-94.

187

Matsuda, L., Bonner, T, & Lolait, S. (1993). Localization of cannabinoid receptor messenger-RNA in rat-brain. Journal of Comparative Neurology, 327(4), 535-550.

Matsuda, L., Lolait, S., Brownstein, M., Young, A., & Bonner, T. (1990). Structure of a cannabinoid receptor and functional expression of the cloned cDNA. Nature, 346(6284), 561-564.

Matsumato, K., Stark, P., & Meister, R. (1977). Cannabinoids,1.1-amino-7,8,9,10 and 1-mercapto-7,8,9,10-tetrahydro-6(h)-dibenzo[b,d]pyrans. Journal of Medicinal Chemistry, 20(1), 17-24.

McAllister, S., Griffin, G., Satin, L., & Abood, M. (1999). Cannabinoid receptors can activate and inhibit G protein-coupled inwardly rectifying potassium channels in a xenopus oocyte expression system. Journal of Pharmacology and Experimental Therapeutics, 291(2), 618-626.

McAllister, S., Rizvi, G., Anavi-Goffer, S., Hurst, D., Barnett-Norris, J., Lynch, D., Reggio, P., & Abood, M. (2003). An aromatic microdomain at the cannabinoid CB1 receptor constitutes an agonist/inverse agonist binding region. Journal of Medicinal Chemistry, 46(24), 5139-5152.

McFarland, M., & Barker, E. (2004) Anandamide transport. Pharmacology & Therapeutics, 104(2), 117-135.

McFarland, M., Porter, A., & Barker, E. (2003). Uptake of the endogenous cannabinoid anandamide by endocytic processes and accumulation in lipid rafts. FASEB Journal, 17(4), 641-642.

McFarland, M., Rakhshan, F., Wilson, J., & Barker, E. (2004). Endocytic and cellular trafficking processes involved with uptake and metabolism of anandamide. FASEB Journal, 18(4), 581-581.

Mechoulam, R., Benshabat, S., Hanus, L., Ligumsky, M., Kaminski, N., Schatz, A., Gopher, A., Almog, S., Martin, B., Compton, D., Pertwee, R., Griffin, G., Bayewitch, M., Barg, J., & Vogel, Z. (1995). Identification of an endogenous 2-monoglyceride, present in canine gut, that binds to cannabinoid receptors. Biochemical Pharmacology, 50(1), 83-90.

Mechalouam, R., Feigenbaum, J., Lander, N., Segal, M., Jarbe, T., Hiltunen, A., & Consroe, P. (1988). Enantiomeric cannabinoids stereospecificity of psychotropic activity. Experientia, 44(9), 762-764.

Melvin, L., Milne, G., Johnson, M., Subramaniam, B., Wilken, G. & Howlett, A. (1993). Structure-activity-relationships for cannabinoid receptor-binding and analgesic activity: studies of bicyclic cannabinoid analogs. Molecular Pharmacology, 44(5), 1008-1015.

188

Meschler, J. & Howlett, A. (1999). Thujone exhibits low affinity for cannabinoid receptors but fails to evoke cannabimimetic responses. Pharmacology Biochemistry and Behaviour, 62(1), 473–480.

Miura, T. & Takahash, R. (1973). Insect developmental inhibitors: effects on nontarget aquatic organisms. Journal of Economic Entomology, 66(4), 917-922.

Monteiro, J., Jurado, A., Moreno, A., & Madeira, V. (2005). Toxicity of methoprene as assessed by the use of a model microorganism. Toxicology in Vitro, 19(7), 951-956.

Montero, C., Campillo, N., Goya, P., & Paez, J. (2005). Homology models of the cannabinoid CB1 and CB2 receptors. A docking analysis study. European Journal of Medicinal Chemistry, 40(1), 75-83.

Muccioli, G. (2007). Blocking the cannabinoid receptors: Drug candidates and therapeutic promises. Chemistry and Biodiversity, 4(1), 1805-1827.

Mukhopadhyay, S., & Howlett, A. (2005). Chemically distinct ligands promote differential CB1 cannabinoid receptor-Gi protein interactions. Molecular Pharmacology, 67(6), 2016-2024.

Mukhopadhyay, S., Shim, J., Assi, A., Norford, D., & Howlett, A. (2002). CB1 cannabinoid receptor-G protein association: A possible mechanism for differential signaling. Chemistry and Physics of Lipids, 121(1-2), 91-109.

Munro, S., Thomas, K., & Abushaar, M. (1993). Molecular characterization of a peripheral receptor for cannabinoids. Nature, 365(6441), 61-65.

Nakanishi, T. (2007). Potential toxicity of organotin compounds via nuclear receptor signaling in mammals. Journal of Health Science, 53(1), 1-9.

Nakatsu, Y., Kotake, Y., Komasaka, K., Hakozaki, H., Taguchi, R., Kume, T., Akaike, A., Ohta, S. (2006). Glutamate excitotoxicity is involved in cell death caused by tributyltin in cultured rat cortical neurons. Toxicological Sciences, 89(1), 235-242.

Nakazi, M., Bauer, U., Nickel, T., Kathmann, M., & Schlicker, E. (2000). Inhibition of serotonin release in the mouse brain via presynaptic cannabinoid CB1 receptors. Naunyn-Schmiedebergs Archives of Pharmacology, 361(1), 19-24.

Navarro, M., & Rodriguez de Fonseca, F. (1998). Introduction to the neurobiology of cannabinoid transmission: From anandamide signaling to higher cerebral functions and disease. Neurobiology of Disease, 5(6), 379-385.

Netzeband, J., Conroy, S., Parsons, K., & Gruol, D. (1999). Cannabinoids enhance NMDA-elicited Ca2+ signals in cerebellar granule neurons in culture. Journal of Neuroscience, 19(20), 8765-8777.

189

Ng, E., Aung, M., Abood, M., Martin, B., & Razdan, R. (1999). Unique analogues of anandamide: Arachidonyl ethers and carbamates and norarachidonyl carbamates and ureas. Journal of Medicinal Chemistry, 42(11), 1975-1981.

Nicholson, R., Liao, C., Zheng, J., David, L., Coyne, L., Errington, A., Singh, G., & Lees, G. (2003) Sodium channel inhibition by anandamide and synthetic cannabimimetics in brain. Brain Research, 978(1-2), 194-204.

Nicholls, D., Sihra, T., & Sanchez-prieto, J. (1987). Calcium-dependent and calcium-independent release of glutamate from synaptosomes monitored by continuous fluorometry. Journal of Neurochemistry, 49(1), 50-57.

Nie, J., & Lewis, D. (2001). The proximal and distal C-terminal tail domains of the CB1 cannabinoid receptor mediate G protein coupling. Neuroscience, 107(1), 161-167.

Nomura, D., Blankman, J., Simon, G., Fujioka, K., Issa, R., Ward, A., Cravatt, B., & Casida, J. (2008). Activation of the endocannabinoid system by organophosphorus nerve agents. Nature Chemical Biology, 4(1), 373-378.

Ohno-Shosaku, T., Maejima, T., & Kano, M. (2001). Endogenous cannabinoids mediate retrograde signals from depolarized postsynaptic neurons to presynaptic terminals. Neuron, 29(3), 729-738.

Ohno-Shosaku, T., Hashimotodani, Y., Ano, M., Takeda, S., Tsubokawa, H., & Kano, M. (2007). Enclocannabinoid signalling triggered by NMDA receptor-mediated calcium entry into rat hippocampal neurons. Journal of Physiology-London, 584(2), 407-418.

Oka, S., Nakajima, K., Yamashita, A., Kishimoto, S., & Sugiura, T. (2007). Identification of GPR55 as a lysophosphatidylinositol receptor. Biochemical and Biophysical Research Communications, 362(4), 928-934.

Okamoto, Y., Wang, J., Morishita, J., & Ueda, N. (2007). Biosynthetic pathways of the endocannabinoid anandamide. Chemistry & Biodiversity, 4(8), 1842-1857.

Oksche, A., Dehe, M., Schulein, R., Wiesner, B., & Rosenthal, W. (1998). Folding and cell surface expression of the vasopressin V2 receptor: Requirement of the intracellular C-terminus. FEBS Letters, 424(1-2), 57-62.

Okuno, T., & Yokomizo, T. (2011). What is the natural ligand of GPR55? Journal of Biochemistry, 149(5), 495-497.

Okoro, H., Fatoki, O., Adekola, F., Ximba, B., Snyman, R., & Opeolu, B. (2011). Human exposure, biomarkers, and fate of organotins in the environment. Reviews in Environmental Contamination and Toxicology, 213(1), 27-54.

190

Osgood, P., Howes, J., Razdan, R., & Pars, H. (1978). Drugs derived from cannabinoids: tachycardia and analgesia structure-activity-relationships in ∆9-tetrahydrocannabinol and some synthetic analogs. Journal of Medicinal Chemistry, 21(8), 809-811.

Pacheco, M., Childers, S., Arnold, R., Casiano, F., & Ward, S. (1991).Aminoalkylindoles-actions on specific G-protein-linked receptors. Journal of Pharmacology and Experimental Therapeutics, 257(1), 170-183.

Pagotto, U., Marsicano, G., Cota, D., Lutz, B., & Pasquali, R. (2006). The emerging role of the endocannabinoid system in endocrine regulation and energy balance. Endocrine Reviews, 27(1), 73-100.

Palmer, S., Thakur, G., & Makriyannis, A. (2002). Cannabinergic ligands. Chemistry and Physics of Lipids, 121(1-2), 3-19.

Papahatjis, D., Kourouli, T., Abadji, V., Goutopoulos, A., & Makriyannis, A. (1998). Pharmacophoric requirements for cannabinoid side chains: Multiple bond and C1'-substituted ∆8-tetrahydrocannabinols. Journal of Medicinal Chemistry, 41(7), 1195-1200.

Papahatjis, D., Nikas, S., Tsotinis, A., Vlachou, M., & Makriyannis, A. (2001). A new ring-forming methodology for the synthesis of conformationally constrained bioactive molecules. Chemistry Letters, (3), 192-193.

Papahatjis, D., Nikas, S., Andreou, T., & Makriyannis, A. (2002). Novel 1', 1'-chain substituted ∆8-tetrahydrocannabinols. Bioorganic & Medicinal Chemistry Letters, 12(24), 3583-3586

Papahatjis, D., Nikas, S., Kourouli, T., Chari, R., Xu, W., Pertwee, R., & Makriyannis, A. (2003). Pharmacophoric requirements for the cannabinoid side chain: probing the cannabinoid receptor subsite at C1'. Journal of Medicinal Chemistry, 46(15), 3221-3229.

Pars, H., Granchelli, F., Razdan, R., Keller, J., Teiger, D., Rosenberg, F., & Harris, L. (1976). Drugs derived from cannabinoids: nitrogen analogs, benzopyranopyridines and benzopyranopyrroles. Journal of Medicinal Chemistry, 19(4), 445-454.

Patny, A., Desai, P., & Avery, M. (2006). Homology modeling of G-protein-coupled receptors and implications in drug design. Current Medicinal Chemistry, 13(14), 1667-1691.

Pertwee, R., Griffin, G., Fernando, S., Li, X., Hill, A., & Makriyannis, A. (1995). AM630, a competitive cannabinoid receptor antagonist. Life Sciences, 56(23-24), 1949-1955.

Pertwee, R. (1988). The central neuropharmacology of psychotropic cannabinoids. Pharmacology & Therapeutics, 36(2-3), 189-261.

191

Pertwee, R. (1997). Pharmacology of cannabinoid CB1 and CB2 receptors. Pharmacology & Therapeutics, 74(2), 129-180.

Pertwee, R. (2005). Inverse agonism and neutral antagonism at cannabinoid CB1 receptors. Life Sciences, 76(12), 1307-1324.

Pertwee, R. (2006). The pharmacology of cannabinoid receptors and their ligands: An overview. International Journal of Obesity, 30(1), 13-18.

Peterson, G. (1977). Simplification of protein assay method of Lowry et al, which is more generally applicable. Analytical Biochemistry, 83(2), 346-356.

Petitet, F., Jeantaud, B., Capet, M., & Doble, A., (1997). Interaction of brain cannabinoid receptors with guanine nucleotide binding protein; a radioligand binding study. Biochemical Pharmacology, 54(1), 1267-1270.

Pinto, J., Potie, F., Rice, K., Boring, D., Johnson, M., Evans, D., Wilken, G., Cantrell, C., & Howlett, A. (1994) Cannabinoid receptor binding and agonist activity of amides and esters of arachidonic-acid. Molecular Pharmacology, 46(3), 516-522.

Piomelli, D., Beltramo, M., Glasnapp, S., Lin, S., Goutopoulos, A., Xie, X., & Makriyannis, A. (1999). Structural determinants for recognition and translocation by the anandamide transporter. Proceedings of the National Academy of Sciences of the United States of America, 96(10), 5802-5807.

Pitler, T., & Alger, B. (1992). Postsynaptic spike firing reduces synaptic GABA(A) responses in hippocampal pyramidal cells. Journal of Neuroscience, 12(10), 4122-4132.

Pitler, T., & Alger, B. (1994). Depolarization-induced suppression of GABAergic inhibition in rat hippocampal pyramidal cells: G-protein involvement in a presynaptic mechanism. Neuron, 13(6), 1447-1455.

Porter, A., Sauer, J., Knierman, M., Becker, G., Berna, M., Bao, J., Nomikos, G, Carter, P., Bymaster, F., Leese, A., & Felder, C. (2002). Characterization of a novel endocannabinoid, virodhamine, with antagonist activity at the CB1 receptor. Journal of Pharmacology and Experimental Therapeutics, 301(3), 1020-1024.

Quistad, G., Schooley, D., Sraiger, L., Bergot, B., Sleight, B., & Macek, K. (1976). Environmental degradation of insect growth-regulator methoprene and metabolism by bluegill fish. Pesticide Biochemistry and Physiology, 6(6), 523-529.

Quistad, G., Nomura, D., Sparks, S., Segall, Y., & Casida, J. (2002). Cannabinoid CB1 receptor as a target for chlorpyrifos oxon and other organophosphorus pesticides. Toxicology Letters, 135(1-2), 89-93.

192

Quistad, G., Klintenberg, R., Caboni, P., Liang, S., & Casida, J. (2006). Monoacylglycerol lipase inhibition by organophosphorus compounds leads to elevation of brain 2-arachidonoylglycerol and the associated hypomotility in mice. Toxicology and Applied Pharmacology, 211(1), 78-83.

Rakhshan, F., Day, T., Blakely, R., & Barker, E. (2000). Carrier-mediated uptake of the endogenous cannabinoid anandamide in RBL-2H3 cells. Journal of Pharmacology and Experimental Therapeutics, 292(3), 960-967.

Razdan, R. (1986). Structure-activity-relationships in cannabinoids. Pharmacological Reviews, 38(2), 75-149.

Reggio, P., Basu-Dutt, S., Barnett-Norris, J., Castro, M., Hurst, D., Seltzman, H., Roche, M., Gilliam, A., Thomas, B., Stevenson, L., Pertwee, R., & Abood, M. (1998). The bioactive conformation of aminoalkylindoles at the cannabinoid CB1 and CB2 receptors: Insights gained from (E)- and (Z)-naphthylidene indenes. Journal of Medicinal Chemistry, 41(26), 5177-5187.

Reggio, P., Greer, K., & Cox, S. (1989). The importance of the orientation of the C9 substituent to cannabinoid activity. Journal of Medicinal Chemistry, 32(7), 1630-1635.

Rhee, M., Bayewitch, M., Avidor-Reiss, T., Levy, R., & Vogel, Z. (1998). Cannabinoid receptor activation differentially regulates the various adenylyl cyclase isozymes. Journal of Neurochemistry, 71(4), 1525-1534.

Rhee, M., Nevo, I., Levy, R., & Vogel, Z. (2000). Role of the highly conserved asp-arg-tyr motif in signal transduction of the CB2 cannabinoid receptor. FEBS Letters, 466(2-3), 300-304.

Rich, M. (1993). Conformational-analysis of arachidonic and related fatty-acids using molecular-dynamics simulations. Biochimica Et Biophysica Acta, 1178(1), 87-96.

Richardson, J., Aanonsen, L., & Hargreaves, K. (1998). Hypoactivity of the spinal cannabinoid system results in NMDA-dependent hyperalgesia. Journal of Neuroscience, 18(1), 451-457.

Riedel, G., Fadda, P., McKillop-Smith, S., Pertwee, R.G., Platt, B., & Robinson, L., (2009). Synthetic and plant-derived cannabinoid receptor antagonists show hypophagic properties in fasted and non-fasted mice. British Journal of Pharmacology, 156(1), 1154-1166.

Rinaldi-Carmona, M., Barth, F., Heaulme, M., Shire, D., Calandra, B., Congy, C., Martinez, S., Maruani, J., Neliat, G., Caput, D., Ferrara, P., Soubrie, P., Breliere, C., & Lefur, G. (1994) SR141716A, a potent and selective antagonist of the brain cannabinoid receptor. FEBS Letters, 350(2-3), 240-244.

193

Rinaldi-Carmona, M., Barth, F., Heaulme, M., Alonso, R., Shire, D., Congy, C., Soubrie, P., Breliere, J.C., & Le Fur, G. (1995). Biochemical and pharmacological characterization of SR141716A, the first potent and selective brain cannabinoid receptor antagonist. Life Sciences, 56(1), 1941–1947.

Rinaldi-Carmona, M., Pialot, F., Congy, C., Redon, E., Barth, F., Bachy, A., Breliere, J-C., Soubrie, P., & Le Fur, G., (1996). Characterization and distribution of binding sites for [3H]SR141716A, a selective brain (CB1) cannabinoid receptor antagonist, in rodent brain. Life Sciences, 58(1), 1239-1247.

Rinaldi-Carmona, M., Le Duigou, A., Oustric, D., Barth, F., Bouaboula, M., Carayon, P., Casellas, P. & Le Fur, G. (1998). Modulation of CB1 cannabinoid receptor functions after a long-term exposure to agonist or inverse agonist in the chinese hamster ovary cell expression system. Journal of Pharmacology and Experimental Therapeutics, 287(3), 1038-1047.

Rinaldi-Carmona, M., Barth, F., Congy, C., Martinez, S., Oustric, D., Perio, A., Poncelet, M., Maruani, J., Arnone, M., Finance, O., Soubrie., & Le Fur, G. (2004). SR147778 [5-(4-bromophenyl)-1-(2,4-dichlorophenyl)-4-ethyl-N-(1-piperidinyl)-1H-pyrazole-3-carboxamide], a new potent and selective antagonist of the CB1 cannabinoid receptor: Biochemical and pharmacological characterization. Journal of Pharmacology and Experimental Therapeutics, 310(3), 905-914.

Rios, C., Gomes, I., & Devi, L. (2006). Mu opioid and CB1 cannabinoid receptor interactions: Reciprocal inhibition of receptor signaling and neuritogenesis. British Journal of Pharmacology, 148(4), 387-395.

Robbe, D., Alonso, G., Duchamp, F., Bockaert, J., & Manzoni, O. (2001) Localization and mechanisms of action of cannabinoid receptors at the glutamatergic synapses of the mouse nucleus accumbens. Journal of Neuroscience, 21(1), 109-116.

Robbe, D., Kopf, M., Remaury, A., Bockaert, J., & Manzoni, O. (2002). Endogenous cannabinoids mediate long-term synaptic depression in the nucleus accumbens. Proceedings of the National Academy of Sciences of the United States of America, 99(12), 8384-8388.

Robert, J., Clauser, E., Petit, P., & Ventura, M. (2005). A novel C-terminal motif is necessary for the export of the vasopressin V1b/V3 receptor to the plasma membrane. Journal of Biological Chemistry, 280(3), 2300-2308.

Ronesi, J., Gerdeman, G., & Lovinger, D. (2004). Disruption of endocannabinoid release and striatal long-term depression by postsynaptic blockade of endocannabinoid membrane transport. Journal of Neuroscience, 24(7), 1673-1679.

Rosenbaum, D., Rasmussen, S., & Kobilka, B. (2009). The structure and function of G-protein-coupled receptors. Nature, 459(7245), 356-363.

194

Ross, C., Maccumber, M., Glatt, C., & Snyder, S. (1989). Brain phospholipase-C isozymes: differential messenger-RNA localizations by in-situ hybridization. Proceedings of the National Academy of Sciences of the United States of America, 86(8), 2923-2927.

Ross, R., Brockie, H., Stevenson, L., Murphy, V., Templeton, F., Makriyannis, A., & Pertwee, R. (1999). Agonist inverse agonist characterization at CB1 and CB2 cannabinoid receptors of L759633, L759656 and AM630. British Journal of Pharmacology, 126(3), 665-672.

Roustan, P., Abitbol, M., Menini, C., Ribeaudeau, F., Gerard, M., Vekemans, M., Mallet, J., & Dufier, J. (1995) The rat phospholipase C-beta-4 gene is expressed at high abundance in cerebellar Purkinje-cells. Neuroreport, 6(14), 1837-1841.

Rubovitch, V., Gafni, M., & Sarne, Y., (2004). The involvement of VEGF receptors and MAPK in the cannabinoid potentiation of Ca2+ flux into N18TG2 neuroblastoma cells. Molecular Brain Research, 120(1), 138-144.

Rueda, D., Navarro, B., Martinez-Serrano, A., Guzman, M., & Galve-Roperh, I. (2002). The endocannabinoid anandamide inhibits neuronal progenitor cell differentiation through attenuation of the Rap1/B-Raf/ERK pathway. Journal of Biological Chemistry, 277(48), 46645-46650.

Rudel, R., Camann, D., Spengler, J., Korn, L., & Brody, J. (2003). Phthalates, alkylphenols, pesticides, polybrominated diphenyl ethers, and other endocrine-disrupting compounds in indoor air and dust. Environmental Science & Technology, 37(20), 4543-4553.

Russo, E. B., Jiang, H., Li, X., Sutton, A., Carboni, A., del Bianco, F., Mandolino, G., Potter, D., Zhao, Y., Bera, S., Zhang, Y., Lue, E., Ferguson, D., Hueber, F., Zhao., L., Liu, C., Wang, Y., & Li, C. (2008). Phytochemical and genetic analyses of ancient cannabis from central Asia. Journal of Experimental Botany, 59(15), 4171-4182.

Ryan, W., Banner, W., Wiley, J., Martin, B., & Razdan, R. (1997). Potent anandamide analogs: The effect of changing the length and branching of the end pentyl chain. Journal of Medicinal Chemistry, 40(22), 3617-3625.

Saitoh, M., Yanase, T., Morinaga, H., Tanabe, M., Mu, Y., Nishi, Y., Nomura, M., Okabe, T., Goto, K., Takayanagi, R., & Nawata, H. (2001). Tributyltin or triphenyltin inhibits aromatase activity in the human granulosa-like tumor cell line KGN. Biochemical and Biophysical Research Communications, 289(1), 198-204.

Samama, P., Cotecchia, S., Costa, T., & Lefkowitz, R. (1993). A mutation-induced activated state of the β (2)-adrenergic receptor: extending the ternary complex model. Journal of Biological Chemistry, 268(7), 4625-4636.

Sarradin, P., Lapaquellerie, Y., Astruc, A., Latouche, C., & Astruc, M. (1995). Long-term behavior and degradation kinetics of tributyltin in marine sediment. Science of the Total Environment, 170(1-2), 59-70.

195

Savinainen, J., Saario, S., Niemi, R., Jarvinen, T., & Laitinen, J. (2003). An optimized approach to study endocannabinoid signaling: Evidence against constitutive activity of rat brain adenosine A (1) and cannabinoid CB1 receptors. British Journal of Pharmacology, 140(8), 1451-1459.

Schertler, G. (2008). Signal transduction: the rhodopsin story continued. Nature, 453(7193), 292-293.

Schettler, T. (2006). Human exposure to phthalates via consumer products. International Journal of Andrology, 29(1), 134–139.

Schlicker, E., & Kathmann, M. (2001). Modulation of transmitter release via presynaptic cannabinoid receptors. Trends in Pharmacological Sciences, 22(11), 565-572.

Schleier, J, Peterson, R., Macedo, P., & Brown, D. (2008). Environmental concentrations, fate, and risk assessment of pyrethrins and piperonyl butoxide after aerial ultralow-volume applications for adult mosquito management. Environmental Toxicology and Chemistry, 27(5), 1063-1068.

Schmeller, T., LatzBruning, B., & Wink, M. (1997). Biochemical activities of berberine, palmatine and sanguinarine mediating chemical defence against microorganisms and herbivores. Phytochemistry, 44(2), 257-266.

Seeman, P., Chauwong, M., & Moyyen, S. (1972). Membrane binding of morphine, diphenylhydantoin, and tetrahydrocannabinol. Canadian Journal of Physiology and Pharmacology, 50(12), 1193-1200.

Segall, Y., Quistad, G., Sparks, S., Nomura, D., & Casida, J. (2003). Toxicological and structural features of organophosphorus and organosulfur cannabinoid CB1 receptor ligands. Toxicological Sciences, 76(1), 131-137.

Seligman, P., Valkirs, A., & Lee, R. (1986). Degradation of tributyltin in San-Diego bay, California, waters. Environmental Science & Technology, 20(12), 1229-1235.

Selley, D.E., Stark, S., Sim, L.J., & Childers, S. (1996). Cannabinoid receptor stimulation of guanosine-5’-O-(3-[35S]thio)triphosphate binding in rat brain membranes. Life Sciences, 59(1), 659-668.

Seltzman, H., Fleming, D., Thomas, B., Gilliam, A., McCallion, D., Pertwee, R., Compton, D., & Martin, B. (1997). Synthesis and pharmacological comparison of dimethylheptyl and pentyl analogs of anandamide. Journal of Medicinal Chemistry, 40(22), 3626-3634.

Sharir, H., & Abood, M. (2010). Pharmacological characterization of GPR55, a putative cannabinoid receptor. Pharmacology & Therapeutics, 126(3), 301-313

Sheldon, A. (1975). Effects of organotin anti-fouling coatings on man and his environment. Journal of Paint Technology, 47(600), 54-58.

196

Shen, M., Piser, T., Seybold, V., & Thayer, S. (1996). Cannabinoid receptor agonists inhibit glutamatergic synaptic transmission in rat hippocampal cultures. Journal of Neuroscience, 16(14), 4322-4334.

Sheskin, T., Hanus, L., Slager, J., Vogel, Z., & Mechoulam, R. (1997). Structural requirements for binding of anandamide-type compounds to the brain cannabinoid receptor. Journal of Medicinal Chemistry, 40(5), 659-667.

Shim, J., Bertalovitz, A., & Kendall, D. (2011a) Identification of essential cannabinoid-binding domains: Structural insights into early dynamic events in receptor activation. Journal of Biological Chemistry, 286(38), 33422-33435.

Shim J., Rudd, J., & Ding, T. (2011b). Distinct second extracellular loop structures of the brain cannabinoid CB1 receptor: Implication in ligand binding and receptor function. Proteins, 79(1), 581-597.

Shim, Y-J. (2010). Understanding functional residues of the cannabinoid CB1 receptor for drug discovery. Current Topics in Medicinal Chemistry, 10(1), 779-798.

Shim, J., Welsh, W., Cartier, E., Edwards, J., & Howlett, A. (2002). Molecular interaction of the antagonist N-(piperidin-1-yl)-5-(4-chlorophenyl)-1(2,4-dichlorophenyl)-4-methyl-1H-pyrazole-3-carboxamide with the CB1 cannabinoid receptor. Journal of Medicinal Chemistry, 45(7), 1447-1459.

Shivachar, A., Martin, B., & Ellis, E. (1996). Anandamide- and ∆9-tetrahydro-cannabinol-evoked arachidonic acid mobilization and blockade by SR141716A [N-(piperidin-1-yl)-5-(4-chlorophenyl)-1-(2,4-dichlorophenyl)-4-methyl-1H-pyrazole-3-carboximide hydrochloride]. Biochemical Pharmacology, 51(5), 669-676.

Siddall, J. (1976). Insect growth-regulators and insect control: critical appraisal. Environmental Health Perspectives, 14(1), 119-126.

Simanek, V., Vespalec, R., Sedo, A., Ulrichova, J., & Vicar, J. (2003). Quaternary benzo[c]phenanthridine alkaloids: biological activities, intracellular targets of their action and application in human and veterinary medicine. Chemical Probes in Biology: Science at the Interface of Chemistry, Biology and Medicine, 129(1), 245-254.

Sim-Selley, L., Brunk, L., & Selley, D. (2001). Inhibitory effects of SR141716A on G-protein activation in rat brain. European Journal of Pharmacology, 414(2-3), 135-143.

Singla, S., Kreitzer, A., & Malenka, R. (2007). Mechanisms for synapse specificity during striatal long-term depression. Journal of Neuroscience, 27(19), 5260-5264.

Slalinova, I., Taborska, E., Bochorakova, H., & Slanina, J. (2001). Interaction of benzo[c]phenanthridine and protoberberine alkaloids with animal and yeast cells. Cell Biology and Toxicology, 17(1), 51-63.

197

Slipetz, D., Oneill, G., Favreau, L., Dufresne, C., Gallant, M., Gareau, Y., Guay, D., Labelle, M., & Metters, K. (1995). Activation of the human peripheral cannabinoid receptor results in inhibition of adenylyl cyclase. Molecular Pharmacology, 48(2), 352-361.

Song, Z., & Bonner, T. (1996). A lysine residue of the cannabinoid receptor is critical for receptor recognition by several agonists but not WIN55212-2. Molecular Pharmacology, 49(5), 891-896.

Staal, G. (1975). Insect growth-regulators with juvenile-hormone activity. Annual Review of Entomology, 20(1), 417-460.

Stadel, R., Ahn, K., & Kendall, D. (2011). The cannabinoid type-1 receptor carboxyl-terminus, more than just a tail. Journal of Neurochemistry, 117(1), 1-18.

Staples, C., Peterson, D., Parkerton, T., & Adams, W. (1997). The environmental fate of phthalate esters: A literature review. Chemosphere, 35(4), 667-749.

Staples, C., Guinn, R., Kramarz, K., & Lampi, M. (2011). Assessing the chronic aquatic toxicity of phthalate ester plasticizers. Human and Ecological Risk Assessment, 17(5), 1057-1076.

Starowicz, K., Nigam, S., & Di Marzo, V. (2007). Biochemistry and pharmacology of endovanilloids. Pharmacology & Therapeutics, 114(1), 13-33.

Steffens, M., Engler, C., Zentner, J., & Feuerstein, T. (2004). Cannabinoid CB1 receptor-mediated modulation of evoked dopamine release and of adenylyl cyclase activity in the human neocortex. British Journal of Pharmacology, 141(7), 1193-1203.

Steffens, S., Veillard, N., Arnaud, C., Pelli, G., Burger, F., Staub, C., Zimmer, A., Frossard, J., & Mach, F. (2005). Low dose oral cannabinoid therapy reduces progression of atherosclerosis in mice. Nature, 434(7034), 782-786.

Stella, N., & Piomelli, D. (2001). Receptor-dependent formation of endogenous cannabinoids in cortical neurons. European Journal of Pharmacology, 425(3), 189-196.

Stella, N., Schweitzer, P., & Piomelli, D. (1997). A second endogenous cannabinoid that modulates long-term potentiation. Nature, 388(6644), 773-778.

Stiborova, M., Simanek, V., Frei, E., Hobza, P., & Ulrichova, J. (2002). DNA adduct formation from quaternary benzo[c]phenanthridine alkaloids sanguinarine and chelerythrine as revealed by the 32P-postlabeling technique. Chemico-Biological Interactions, 140(3), 231-242.

Sugiura, T., Kobayashi, Y., Oka, S., & Waku, K. (2002). Biosynthesis and degradation of anandamide and 2-arachidonoylglycerol and their possible physiological significance. Prostaglandins Leukotrienes and Essential Fatty Acids, 66(2-3), 173-192.

198

Sugiura, T., Kodaka, T., Kondo, S., Nakane, S., Kondo, H., Waku, K., Ishima, Y., Watanabe, K., & Yamamoto, I. (1997). Is the cannabinoid CB1 receptor a 2-arachidonoylglycerol receptor? Structural requirements for triggering a Ca2+ transient in NG108-15 cells. Journal of Biochemistry, 122(4), 890-895.

Sugiura, T., Kodaka, T., Kondo, S., Tonegawa, T., Nakane, S., Kishimoto, S., Yamashita, A., & Waku, K. (1996). 2-arachidonoylglycerol, a putative endogenous cannabinoid receptor ligand, induces rapid, transient elevation of intracellular free Ca2+ in neuroblastoma x glioma hybrid NG108-15 cells. Biochemical and Biophysical Research Communications, 229(1), 58-64.

Sugiura, T., Kodaka, T., Nakane, S., Miyashita, T., Kondo, S., Suhara, Y., Takayama, H., Waku, K., Seki, C., Baba, N., & Ishima, Y. (1999). Evidence that the cannabinoid CB1 receptor is a 2-arachidonoylglycerol receptor: structure-activity relationship of 2-arachidonoylglycerol ether-linked analogues, and related compounds. Journal of Biological Chemistry, 274(5), 2794-2801.

Sugiura, T., Kondo, S., Sukagawa, A., Nakane, S., Shinoda, A., Ithoh, K., Yamashita, A., & Waku, K. (1995). 2-arachidonoylgylcerol: a possible endogenous cannabinoid receptor-ligand in brain. Biochemical and Biophysical Research Communications, 215(1), 89-97.

Sugiura, T., Kishimoto, S., Oka, S., & Gokoh, M. (2006). Biochemistry, pharmacology and physiology of 2-arachidonoylglycerol, an endogenous cannabinoid receptor ligand. Progress in Lipid Research, 45(5), 405-446.

Sun, Y., & Johnson, E. (1960). Synergistic and antagonistic actions of insecticide-synergist combinations and their mode of action. Journal of Agricultural and Food Chemistry, 8(4), 261-266.

Suvorova, E., Gripentrog, J., Jesaitis, A., & Miettinen, H (2009). Agonist-dependent phosphorylation of the formyl peptide receptor is regulated by the membrane proximal region of the cytoplasmic tail. Biochimica Et Biophysica Acta-Molecular Cell Research, 1793(2), 406-417.

Swift, S., Leger, A., Talavera, J., Lei, Z., Bohm, A., & Kuliopulos, A. (2006). Role of the PAR1 receptor 8th helix in signaling - the 7-8-1 receptor activation mechanism. Journal of Biological Chemistry, 281(7), 4109-4116.

Szabo, B., Dorner, L., Pfreundtner, C., Norenberg, W., & Starke, K. (1998). Inhibition of GABAergic inhibitory postsynaptic currents by cannabinoids in rat corpus striatum. Neuroscience, 85(2), 395-403.

Szabo, G., Varga, B., Payer-Lengyel, D., Szemzo, A., Erdelyi, P., Vukics, K., Szikra, J., Hegyi, E., Vastag, M., Kiss, B., Laszy, J., Gyertyn I., & Fischer, J. (2009). Chemical and biological investigation of cyclopropyl containing diaryl-pyrazole-3-carboxamides as novel and potent cannabinoid type 1 receptor antagonists. Journal of Medicinal Chemistry, 52(1), 4329-4337.

199

Tai, A., Chuang, J., Bode, C., Wolfrum, U., & Sung, C. (1999). Rhodopsin's carboxy-terminal cytoplasmic tail acts as a membrane receptor for cytoplasmic dynein by binding to the dynein light chain Tctex-1. Cell, 97(7), 877-887.

Takahashi, K., & Linden, D. (2000). Cannabinoid receptor modulation of synapses received by cerebellar Purkinje cells. Journal of Neurophysiology, 83(3), 1167-1180.

Tanaka, J., Nakagawa, S., Kushiya, E., Yamasaki, M., Fukaya, M., Iwanaga, T., Simon, M., Sakimura, K., Kano, M., & Watanabe, M. (2000). Gq protein alpha subunits Gαq and Gα11 are localized at postsynaptic extra-junctional membrane of cerebellar Purkinje cells and hippocampal pyramidal cells. European Journal of Neuroscience, 12(3), 781-792.

Tanaka, O., & Kondo, H. (1994). Localization of messenger-RNAs for 3 novel members (β-3, β-4 and γ-2) of phospholipase-C family in mature rat-brain. Neuroscience Letters, 182(1), 17-20.

Tao, Q., & Abood, M. (1998). Mutation of a highly conserved aspartate residue in the second transmembrane domain of the cannabinoid receptors, CB1 and CB2, disrupts G-protein coupling. Journal of Pharmacology and Experimental Therapeutics, 285(2), 651-658.

Tarzia, G., Duranti, A., Tontini, A., Spadoni, G., Mor, M., Rivara, S., Plazzi, P., & Piomelli, D. (2003). Synthesis and structure-activity relationships of a series of pyrrole cannabinoid receptor agonists. Bioorganic & Medicinal Chemistry, 11(18), 3965-3973.

Thakur, G. A., Tichkule, R., Bajaj, S., & Makriyannis, A. (2009). Latest advances in cannabinoid receptor agonists. Expert Opinion on Therapeutic Patents, 19(12), 1647-1673.

Thielen, A., Oueslati, M., Hermosilla, R., Krause, G., Oksche, A., Rosenthal, W., & Schulein, R. (2005). The hydrophobic amino acid residues in the membrane-proximal C tail of the G protein-coupled vasopressin V2 receptor are necessary for transport-competent receptor folding. FEBS Letters, 579(23), 5227-5235.

Thomas, A., Stevenson, L., Wease, K., Price, M., Baillie, G., Ross, R., Pertwee, R., (2005). Evidence that the plant cannabinoid ∆9-tetrahydrocannabivarin is a cannabinoid CB1 and CB2 receptor antagonist. British Journal of Pharmacology, 146(1), 917-926.

Tiburu, E. K., Gulla, S. V., Tiburu, M., Janero, D. R., Budil, D. E., & Makriyannis, A. (2009). Dynamic conformational responses of a human cannabinoid receptor-1 helix domain to its membrane environment. Biochemistry, 48(22), 4895-4904.

Tiburu, E. K., Tyukhtenko, S., Zhou, H., Janero, D., Struppe, J., & Makriyannis, A. (2011). Human cannabinoid-1 GPCR C-terminal domain interacts with bilayer phospholipids to modulate the structure of its membrane environment. AAPS Journal, 13(1), 92-98.

200

Tius, M., Makriyannis, A., Zou, X., & Abadji, V. (1994). Conformationally restricted hybrids of CP55940 and HHC: stereoselective synthesis and activity. Tetrahedron, 50(9), 2671-2680.

Tornqvist, H., & Belfrage, P. (1976). Purification and some properties of a monoacylglycerol-hydrolyzing enzyme of rat adipose-tissue. Journal of Biological Chemistry, 251(3), 813-819.

Tsou, K., Brown, S., Sanudo-Pena, M., Mackie, K., & Walker, J., (1998). Immunohistochemical distribution of cannabinoid CB1 receptors in the rat central nervous system. Neuroscience, 83(1), 393-411.

Tsunoda, M., Aizawa, Y., Konno, N., Kimura, K., & Sugita-Konishi, Y. (2006). Subacute administration of tributyltin chloride modulates neurotransmitters and their metabolites in discrete brain regions of maternal mice and their F1 offspring. Toxicology and Industrial Health, 22(1), 15-25.

Tsutsumi, T., Kobayashi, T., Ueda, H., Yamauchi, E., Watanabe, S., & Okuyama, H. (1994). Lysophosphoinositide-specific phospholipase-C in rat-brain synaptic plasma-membranes. Neurochemical Research, 19(4), 399-406.

Turu, G., & Hunyady, L. (2010). Signal transduction of the CB1 cannabinoid receptor. Journal of Molecular Endocrinology, 44(2), 75-85.

Twitchell, W., Brown, S., & Mackie, K. (1997). Cannabinoids inhibit N- and P/Q-type calcium channels in cultured rat hippocampal neurons. Journal of Neurophysiology, 78(1), 43-50.

Tyukhtenko, S., Tiburu, E., Deshmukh, L., Vinogradova, O., Janero, D., & Makriyannis, A. (2009). NMR solution structure of human cannabinoid receptor-1 helix 7/8 peptide: Candidate electrostatic interactions and microdomain formation. Biochemical and Biophysical Research Communications, 390(3), 441-446.

Uchigashima, M., Narushima, M., Fukaya, M., Katona, I., Kano, M., & Watanabe, M. (2007). Subcellular arrangement of molecules for 2-arachidonoyl-glycerol-mediated retrograde signaling and its physiological contribution to synaptic modulation in the striatum. Journal of Neuroscience, 27(14), 3663-3676.

Ueda, H., Kobayashi, T., Kishimoto, M., Tsutsumi, T., & Okuyama, H. (1993). A possible pathway of phosphoinositide metabolism through edta-insensitive phospholipase-A(1) followed by lysophosphoinositide-specific phospholipase-C in rat-brain. Journal of Neurochemistry, 61(5), 1874-1881.

Ugolini, L., Della Noce, I., Trincia, P., Borzatta, V., & Palmieri, S. (2005). Benzodioxole derivatives as negative effectors of plant proteases. Journal of Agricultural and Food Chemistry, 53(19), 7494-7501.

Vandevoorde, S., & Lambert, D. M. (2007). The multiple pathways of endocannabinoid metabolism: A zoom out. Chemistry & Biodiversity, 4(8), 1858-1881.

201

Van Sickle, M., Oland, L., Ho, W, Hillard, C., Mackie, K, Davison, J., & Sharkey, K. (2001) Cannabinoids inhibit emesis through CB1 receptors in the brainstem of the ferret. Gastroenterology, 121(4):767-74.

Van der Stelt, M., Van Kuik, J., Bari, M., Van Zadelhoff, G., Leeflang, B., Veldink, G., Finazzi-Agro, A., Vliegenthart, J., & Maccarrone, M. (2002). Oxygenated metabolites of anandamide and 2-arachidonoylglycerol: conformational analysis and interaction with cannabinoid receptors, membrane transporter, and fatty acid amide hydrolase. Journal of Medicinal Chemistry, 45(1), 3709-3720.

Varvel, S., Hamm, R., Martin, B., & Lichtman, A. (2001). Differential effects of ∆9-THC on spatial reference and working memory in mice. Psychopharmacology, 157(2), 142-150.

Varvel, S. & Lichtman, A. (2002). Evaluation of CB1 receptor knockout mice in the morris water maze. Journal of Pharmacology and Experimental Therapeutics, 301(3), 915-924.

Vincent, P. & Marty, A. (1993). Neighboring cerebellar Purkinje-cells communicate via retrograde inhibition of common presynaptic interneurons. Neuron, 11(5), 885-893.

Di Marzo, V. (2006). A brief history of cannabinoid and endocannabinoid pharmacology as inspired by the work of British scientists. Trends in Pharmacological Sciences, 27(3), 134-140.

Wager-Miller, J., Westenbroek, R., & Mackie, K. (2002). Dimerization of G protein-coupled receptors: CB1 cannabinoid receptors as an example. Chemistry and Physics of Lipids, 121(1-2), 83-89.

Walker, J., Huang, S., Strangman, N., Tsou, K., & Sanudo-Pena, M. (1999). Pain modulation by release of the endogenous cannabinoid anandamide. Proceedings of the National Academy of Sciences of the United States of America, 96(21), 12198-12203.

Wang, B., Lu, Z., & Polya, G. (1997). Inhibition of eukaryote protein kinases by isoquinoline and oxazine alkaloids. Planta Medica, 63(6), 494-498.

Wang, S. (2003) Cannabinoid CB1 receptor-mediated inhibition of glutamate release from rat hippocampal synaptosomes. European Journal of Pharmacology, 469(1-3), 47-55.

Wang, S., Liu, J., Qian, C., & Chen, X. (2012). Synthetic and mechanistic investigation of piperonyl butoxide from dihydrosafrole. Research on Chemical Intermediates, 38(1), 147-160.

Watanabe, M., Nakamura, M., Sato, K., Kano, M., Simon, M., & Inoue, Y. (1998) Patterns of expression for the mRNA corresponding to the four isoforms of phospholipase C beta in mouse brain. European Journal of Neuroscience, 10(6), 2016-2025.

202

Wiley, J., Jefferson, R., Grier, M., Mahadevan, A., Razdan, R., & Martin, B. (2001). Novel pyrazole cannabinoids: Insights into CB1 receptor recognition and activation. Journal of Pharmacology and Experimental Therapeutics, 296(3), 1013-1022.

Wiley, J., Beletskaya, I., Ng, E., Dai, Z., Crocker, P., Mahadevan, A., Razdan, R., & Martin, B. (2002). Bicyclic resorcinol derivatives: a novel template for the development of cannabinoid CB1/CB2 and CB2-selective agonists. Journal of Pharmacology and Experimental Therapeutics, 301(1), 679-689.

Wilson, T. (2004). The molecular site of action of juvenile hormone and juvenile hormone insecticides during metamorphosis: How these compounds kill insects. Journal of Insect Physiology, 50(2-3), 111-121.

Wilson, R., Kunos, G., & Nicoll, R. (2001). Presynaptic specificity of endocannabinoid signaling in the hippocampus. Neuron, 31(3), 453-462.

Wilson, R., & Nicoll, R. (2001). Endogenous cannabinoids mediate retrograde signalling at hippocampal synapses. Nature, 410(6828), 588-592.

Wrobleski, S., Chen, P., Hynes, J., Lin, S., Norris, D., Pandit, C., Spergel, S., Wu, H., Tokarski, J., Chen, X., Gillooly, K., Kiener, P., Mcintyre, K., Patil-Koota, V.,Shuster, D., Turk, L., Yang,G., & Leftheris, K. (2003). Rational design and synthesis of an orally active indolopyridone as a novel conformationally constrained cannabinoid ligand possessing antiinflammatory properties. Journal of Medicinal Chemistry, 46(11), 2110-2116.

Wu, C., Hung, M., Song, J., Yeh, T., Chou, M., Chu, C., Jan, J., Hsieh, M., Tseng, S., Chang, C., Hsieh, W., Lin, Y., Yeh, Y., Chung, W., Kuo, C., Lin, C., Shy, H., Chao, Y., & Shia, K. (2009). Discovery of 2-[5-(4-chlorophenyl)-1-(2,4-dichlorophenyl)-4-ethyl-1H-pyrazol-3-yl]-1,5,5-trimethyl-1,5-dihydroimidazol-4-thione (BPR-890) via an active metabolite. A novel, potent and selective cannabinoid-1 receptor inverse agonist with high antiobesity efficacy in DIO mice. Journal of Medicinal Chemistry, 52(1), 4496-4510.

Wuertz, S., Miller, C., Pfister, R., & Cooney, J. (1991). Tributyltin-resistant bacteria from estuarine and fresh-water sediments. Applied and Environmental Microbiology, 57(10), 2783-2789.

Xie, X. & Chen, J. (2005). NMR structural comparison of the cytoplasmic juxtamembrane domains of G-protein-coupled CB1 and CB2 receptors in membrane mimetic dodecylphosphocholine micelles. Journal of Biological Chemistry, 280(12), 12064-12064.

Xie, X., Chen, J., & Billings, E. (2003) 3D structural model of the G-protein-coupled cannabinoid CB2 receptor. Proteins-Structure Function and Genetics, 53(2), 307-319

203

Xu, Q., Yin, X., Wang, M., Wang, H., Zhang, N., Shen, Y., Xu, S., Zhang, L., & Gu, Z. (2010). Analysis of phthalate migration from plastic containers to packaged cooking oil and mineral water. Journal of Agricultural and Food Chemistry, 58(21), 11311-11317.

Yasuda, D., Okuno, T., Yokomizo, T., Hori, T., Hirota, N., Hashidate, T., Miyano, M., Shimizu, T. & Nakamura, M. (2009). Helix 8 of leukotriene B (4) type-2 receptor is required for the folding to pass the quality control in the endoplasmic reticulum. FASEB Journal, 23(5), 1470-1481.

Yoshida, T., Fukaya, M., Uchigashima, M., Miura, E., Kamiya, H., Kano, M., & Watanabe, M. (2006). Localization of diacylglycerol lipase-alpha around postsynaptic spine suggests close proximity between production site of an endocannabinoid, 2-arachidonoyl-glycerol, and presynaptic cannabinoid CB1 receptor. Journal of Neuroscience, 26(18), 4740-4751.

Zona, C., Tancredi, V., Longone, P., D’Arcangelo, G., D’Antuono, M., Manfredi, M., & Avoli, M. (2002). Neocortical potassium currents are enhanced by the antiepileptic drug lamotrigine. Epilepsia 43(1), 685–690.

Zuardi, Antonio Waldo (2006). History of cannabis as a medicine: a review. Revista Brasileira de Psiquiatria, 28(2), 153-157.

Zvonok, N., Xi, W., Williams, J., Janero, D. R., Krishnan, S. C., & Makriyannis, A. (2010). Mass spectrometry-based GPCR proteomics: Comprehensive characterization of the human cannabinoid 1 receptor. Journal of Proteome Research, 9(4), 1746-1753.