the effects of cannabidiol on the electrical and

105
United Arab Emirates University Scholarworks@UAEU eses Electronic eses and Dissertations 10-2014 THE EFFECTS OF CANNABIDIOL ON THE ELECTRICAL AND CONTCTILE PROPERTIES OF CARDIOMYOCYTES Ramez Ali Mansour Follow this and additional works at: hps://scholarworks.uaeu.ac.ae/all_theses Part of the Medicine and Health Sciences Commons is esis is brought to you for free and open access by the Electronic eses and Dissertations at Scholarworks@UAEU. It has been accepted for inclusion in eses by an authorized administrator of Scholarworks@UAEU. For more information, please contact [email protected]. Recommended Citation Mansour, Ramez Ali, "THE EFFECTS OF CANNABIDIOL ON THE ELECTRICAL AND CONTCTILE PROPERTIES OF CARDIOMYOCYTES" (2014). eses. 44. hps://scholarworks.uaeu.ac.ae/all_theses/44

Upload: others

Post on 26-Oct-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

United Arab Emirates UniversityScholarworks@UAEU

Theses Electronic Theses and Dissertations

10-2014

THE EFFECTS OF CANNABIDIOL ON THEELECTRICAL AND CONTRACTILEPROPERTIES OF CARDIOMYOCYTESRamez Ali Mansour

Follow this and additional works at: https://scholarworks.uaeu.ac.ae/all_theses

Part of the Medicine and Health Sciences Commons

This Thesis is brought to you for free and open access by the Electronic Theses and Dissertations at Scholarworks@UAEU. It has been accepted forinclusion in Theses by an authorized administrator of Scholarworks@UAEU. For more information, please contact [email protected].

Recommended CitationMansour, Ramez Ali, "THE EFFECTS OF CANNABIDIOL ON THE ELECTRICAL AND CONTRACTILE PROPERTIES OFCARDIOMYOCYTES" (2014). Theses. 44.https://scholarworks.uaeu.ac.ae/all_theses/44

United Arab Emirates University

College of Medicine and Health Sciences

Department of Pharmacology and Therapeutics

THE EFFECTS OF CANNABIDIOL ON THE ELECTRICAL AND

CONTRACTILE PROPERTIES OF CARDIOMYOCYTES

Ramez Ali Mansour

This thesis is submitted in partial fulfillment of the requirements for the degree of Master

of Medical Sciences

Under the Supervision of Dr. Murat Oz

October 2014

ii

Declaration of Original Work

I, Ramez Ali Mansour, the undersigned, a graduate student at the United Arab Emirates

University (UAEU) and the author of this thesis entitled “The effects of cannabidiol on

the electrical and contractile properties of cardiomyocytes”, hereby solemnly declare that

this thesis is an original research work that has been done and prepared by me under the

supervision of Dr. Murat Oz, in the College of Medicine and Health Sciences at the

UAEU. This work has not previously formed the basis for the award of any academic

degree, diploma or similar title at this or any other university. The materials borrowed

from other sources and included in my dissertation have been properly cited and

acknowledged.

Student’s Signature___________________ Date___________________

iii

Copyright © 2014 by Ramez Ali Mansour

All Rights Reserved

iv

Approval of the Master Thesis

This Master Thesis is approved by the following Examining Committee Members:

1) Advisor (Committee Chair):

Title:

Department of …

College of …

Signature________________________________ Date___________________________

2) Member:

Title:

Department of …

College of …

Signature________________________________ Date___________________________

3) Member:

Title:

Department of …

College of …

Signature________________________________ Date___________________________

4) Member (External Examiner):

Title:

Department of …

Institution:

Signature________________________________ Date___________________________

v

This Master thesis is accepted by:

Dean of the College of Medicine and Health Science: Professor Dennis Templeton

Signature___________________________________ Date________________

Dean of the College of Graduate Studies: Professor Nagi Wakim

Signature___________________________________ Date________________

Copy ____ of ____

vi

Abstract

In earlier studies, cannabidiol (CBD), a major nonpsychotropic cannabinoid found

in cannabis plant, has been shown to influence cardiovascular functions under various

physiological and pathological conditions. In the present study, the effects of CBD on

contractility and electrical properties of rat ventricular myocytes were investigated. Video

edge detection was used to measure myocyte shortening. Intracellular Ca2+

was measured

in cells loaded with the fluorescent indicator fura-2 AM. CBD (1 µM) caused a

significant decrease in the amplitudes of electrically-evoked myocyte shortening and Ca2+

transients. However, the amplitudes of caffeine-evoked Ca2+

transients and the rate of

recovery of electrically-evoked Ca2+

transients following caffeine application were not

altered. Whole-cell patch-clamp technique was employed to investigate the effect of CBD

on the characteristics of action potentials (APs) and L-type Ca2+

channels. CBD (1 µM)

significantly decreased the duration of APs. Further studies on L-type Ca2+

channels

indicated that CBD inhibits these channels with IC50 of 0.1 µM in a voltage-independent

manner.

Keywords: Cannabidiol, I type calcium channels, ventricular myocyte.

vii

Title and Abstract in Arabic

الكهربية و الانقباضية الخاصة بالخلايا العضلية على الخواص الكانابيديول آثار

القلبية

الملخص

في دراسات سابقة، الكانابيديول، و هو من المؤثرات العقلية الكانابينويدية الرئيسية الموجودة في نبات

ف الظروف الفيسيولوجية و المرضية. الحشيش، وجد أن له تأثيرات على وظائف القلب و الأوعية الدموية تحت مختل

الهدف في هذه الأطروحة هو دراسة اثار الكانابيديول على الانقباضات و الخصائص الكهربية لخلايا الفأر العضلية.

Caو تم قياس انقباض الخلايا العضلية بالاستعانة بطريقة الكشف بالفيديو إيدج. و تم قياس الموجودة داخل الخلايا +2

(.fura-2 AMمؤشر فلوريسنت ) باستخدام

تسبب في انخفاض ملحوظ في سعة انقباض الخلايا (μM 1)الكانابيديول أهم نتائج هذه الدراسة هو أن

Caالعضلية المثارة كهربيا و Ca)العابر +2

2+ transients) غير أن ال .Ca

العابر المثار بالكافيين و معدل +2

Caإنعاش ال الكانابيديول على التالية لعملية إضافة الكافيين لم تتغير. و تم التحقق من تأثير العابر المثار كهربيا +2

L- type Ca( و قنوات ال action potentialsخصائص جهد الفعل )بالاستعانة بتقتية الالتقاط الرقعي لكامل +2

أدى لإنخفاض (μM 1)يول الكانابيد(. بالإضافة إلى ذلك، دلت النتائج أن Whole-cell patch-clampالخلية )

L- type Caملحوظ في المدة الزمنية لجهد الفعل. مزيد من الدراسات على قنوات ال الكانابيديول يمنع بينت أن +2

(.voltage-independent mannerبطريقة الجهد المستقلة ) μM 0.1من IC50هذه القنوات ب

viii

Acknowledgments

I would like to express my deep gratitude to my master thesis advisor, Professor

Murat Oz who has supported me throughout my thesis with his patience and knowledge

whilst allowing me the room to work in my own way. I appreciate his encouragement and

efforts without which this thesis would not have been completed. One simply could not

wish for a better or friendlier supervisor. I am also grateful to Professor Chris Howarth,

Professor Oleg Krishtal and Dr. Rajesh Mohanraj for spending time in reading and

providing useful suggestions about the thesis.

My gratitude also goes to my colleagues to thank them for their assistance to me:

Mr. Anwar, Dr. Nour Alain, Dr. Lina and Khawla. Additionally, I want to acknowledge

some friends who inspirited my efforts to overcome difficulties, so a special thanks to,

Mohammed and Abrar.

I am grateful to my mother, my father and my uncle for all of the sacrifices that

they have made on my behalf. I would not have got to this point without you. Special

thanks go to my brother Ahmed and my sister Yousra for their continuous

encouragement.

ix

Dedication

To my parents, my uncle, my brother Ahmed and my

sister Yousra

x

Table of Contents

Title .................................................................................................................................................. i

Declaration of Original Work .......................................................................................................... ii

Copyright ........................................................................................................................................ iii

Approval of the Master Thesis ....................................................................................................... iv

Abstract .......................................................................................................................................... vi

Title and Abstract in Arabic .......................................................................................................... vii

Acknowledgments ........................................................................................................................ viii

Dedication ...................................................................................................................................... ix

Table of Contents ............................................................................................................................ x

List of Tables ................................................................................................................................. xii

List of Figures .............................................................................................................................. xiii

Glossary ........................................................................................................................................ xiv

Chapter 1: Introduction ................................................................................................................... 1

1.1 Aims and Objectives: ...................................................................................................... 1

1.2 History ............................................................................................................................. 1

1.3 Cannabinoids ................................................................................................................... 6

1.3.1 Phytocannabinoids ................................................................................................... 7

1.3.2 Endocannabinoids .................................................................................................... 8

1.3.3 Synthetic Cannabinoids ........................................................................................... 9

1.4 Cannabidiol ................................................................................................................... 13

1.4.1 Mechanisms of Action of Cannabidiol .................................................................. 13

1.4.2 Clinical Applications ............................................................................................. 20

1.5 Contraction of Cardiac Muscle ...................................................................................... 28

1.5.1 Excitation–Contraction Coupling .......................................................................... 31

1.5.2 Sarcoplasmic Reticulum and Ryanodine Receptors .............................................. 40

1.5.3 The Contractile Machinery .................................................................................... 41

1.5.4 Calcium Sensitivity of Myofilaments .................................................................... 43

Chapter 2: Methods ....................................................................................................................... 44

2.1 Ventricular Myocyte Isolation ....................................................................................... 44

2.2 Measurement of Ventricular Myocyte Shortening ........................................................ 46

2.3 Measurement of Intracellular Ca2+

Concentration ......................................................... 46

2.4 Measurement of Sarcoplasmic Reticulum Ca2+

Content ............................................... 47

2.5 Assessment of Myofilament Sensitivity to Ca2+

............................................................ 48

2.6 Whole-Cell Recordings ................................................................................................. 49

Chapter 3: Results ......................................................................................................................... 53

xi

3.1 Effects of Cannabidiol on Ventricular Myocyte Shortening ......................................... 53

3.2 Effects of Cannabidiol on Intracellular Ca2+

................................................................. 57

3.3 Effect of Cannabidiol on Sarcoplasmic Reticulum Ca2+

Transport ............................... 59

3.4 Effect of Cannabidiol on Myofilament Sensitivity to Ca2+

........................................... 61

3.5 Effects of Cannabidiol on the Action Potentials of Ventricular Myocytes ................... 63

3.6 Effect of Cannabidiol on L-type Ca2+

Currents ............................................................. 65

Chapter 4: Discussion .................................................................................................................... 70

Chapter 5: Limitation and Future Directions ................................................................................. 77

Bibliography .................................................................................................................................. 78

xii

List of Tables

Table 1. Class of Cannabinoids ..................................................................................................... 11

Table 2. Classification of Calcium Channels ................................................................................ 36

xiii

List of Figures

Figure 1. Cannabis Sativa leaves. .................................................................................................... 2

Figure 2. Major Phytocannabinoids ................................................................................................ 2

Figure 3. Major Phytocannabinoids Synthetic Pathways. ............................................................... 5

Figure 4. Heart's electrical system ................................................................................................. 28

Figure 5. Cardiac Myocytes. ......................................................................................................... 29

Figure 6. Ventricular Action Potential Membrane ........................................................................ 30

Figure 7. Cardiac Excitation Contraction Coupling. ..................................................................... 34

Figure 8. Calcium Channel Structure ............................................................................................ 38

Figure 9. Sarcomere Structure. ...................................................................................................... 42

Figure 10. Langendorff Heart ........................................................................................................ 45

Figure 11. Video Edge Detection and Ca2+ Imaging ................................................................... 48

Figure 12. Patch Clamp Setup ....................................................................................................... 50

Figure 13.Microelectrode Puller. ................................................................................................... 51

Figure 14. Glass Electrode Forming Giga Seal with Ventricular Myocyte ................................... 52

Figure 15. Ventricular Myocyte .................................................................................................... 52

Figure 16. Effects of Cannabidiol on Ventricular Myocyte Shortening... ..................................... 55

Figure 17. Concentration-Inhibition Relationship of Cannabidiol Effect.. ................................... 56

Figure 18. Effects of Cannabidiol on Amplitude and Time-Course of Intracellular Ca2+

............. 58

Figure 19. Effect of Cannabidiol on Sarcoplasmic Reticulum Ca2+

Transport ............................. 60

Figure 20. Effect of Cannabidiol on Myofilament Sensitivity to Ca2+

. ......................................... 62

Figure 21. Effect of Cannabidiol on the Excitability of Ventricular Myocytes. ........................... 64

Figure 22. Effect of Cannabidiol on Ca2+

Currents Mediated by L-type Ca2+

Channels ............... 66

Figure 23. The effect of Cannabidiol on Voltage-Current Characteristics ..................................... 66

Figure 24. Steady State Activation and Inactivation of L-type Ca2+

Currents .............................. 68

Figure 25. Effect of Cannabidiol on the Kinetics of L-type Ca2+

Currents ................................... 69

xiv

Glossary

THC Tetrahydrocannabinol

CBN Cannabinol

CBD Cannabidiol

THCA Tetrahydrocannabinolic Acid

CBGA Cannabigerolic Acid

CBG Cannabigerol

CBDA Cannabidiolic Acid

CBCA Cannabichromenic Acid

CBC Cannabichromene

AEA Anandamide

2-AG 2-Arachidonylglycerol

FAAH Fatty Acid Amidehydrolase

MAGL Monoacylglycerol

AJA Ajalemic Acid

TRPV1 Transient Vanniloid Aeceptors 1

PPARγ Peroxisome Proliferator Activated Receptor

VR1 Vannilloid Receptors

PLC Phospholipase C

PKC Phosphokinase C

NADA N-arachidonoyl-Dopamine

NCX Soduim- Calcium Exchanger

APD Action Potential Duration

RyRs Ryanodine Receptors

PKA Protein Kinase A

ADP Adenosine Diphosphate

ATP Adenosine Triphosphate

NT Normal Tyrode

RCL Resting Cell Length

SR Sarcoplasmic Reticulum

TTX Tetrodotoxin

AMP Amplitudes

RU Ratio Unit

SSA Steady-State Activation

SSL Steady-State Inactivation

AP Action Potential

APD Action Potential Duration

Chapter 1: Introduction

1.1 Aims and Objectives:

The aims of this research are to examine the therapeutic potential and mechanism

by which the phytocannabinoid Cannabidiol CBD affects cardiac contractility. This will

be achieved by investigating the effects of CBD on action potential duration, ventricular

shortening, myofilament sensitivity, and L-type calcium current.

1.2 History

Marijuana is a crude preparation of flowering tops, leaves, seeds and stems from

female Indian hemp cannabis sativa plants (Figure 1). It has been used for centuries for

both recreational and therapeutic purposes. Currently cannabis is being investigated for

the treatment of various pathological conditions such as pain, nausea, vomiting and

cancer. It is one of the most extensively studied plants. There were more than 12,000

PubMed articles published in January 2014 using the key word cannabis in relation to the

various biological and pharmacological effects of this drug.

2

Figure 1. Cannabis Sativa Leaves.ScienceDaily, n.d. Web. 30 Apr. 2014

Figure 2. Major Phytocannabinoids

3

The medicinal use of marijuana has a long history of therapeutic use, going back

several thousands of years. It was often used for the same medical conditions it is used to

treat today such as muscle spasms, pain, nausea, vomiting, epilepsy, and glaucoma.

However, use of marijuana is also associated with distorted perception (sights, sounds,

time and touch), panic attacks, trouble with thinking, difficulties in learning and problem

solving, and loss of motor coordination. Due to these adverse effects, the use of

marijuana was banned in the U.S.A and throughout most of the world almost eighty years

ago.

During the second half of the 20th

century, researchers discovered that the

biological actions of marijuana are mediated mainly by a group of chemicals collectively

called cannabinoids. Although cannabis plant extracts contain more than 460 compounds,

based on pharmacological properties and radioligand binding studies, currently about

60 of the compounds found in this plant, are collectively named phytocannabinoids.

Among these phytocannabinoids, Δ9-Tetrahydrocannabinol (THC) is a major chemical

that mediates the psychoactive actions of the cannabis plant (Figure 2). In addition to

THC, other cannabinoids such as Cannabigerol, Cannabichromene, Cannabinol (CBN)

and Cannabidiol (CBD), constitute major non-psychoactive components of the cannabis

plant. These compounds are highly lipophilic and bear significant similarities in their

chemical structure. For example, CBD which is the focus of this study, is a structural

isomer of THC i.e. it has the same chemical composition but the atoms are arranged

differently.

4

While the majority of the adverse psychological effects of marijuana are known to

be mediated by THC, the contribution of other cannabinoids such as CBD and CBN to

the overall influence of marijuana remains unknown. In recent years, various

combinations of these psychoactive and non-psychoactive cannabinoids have been used

in cannabis based treatments. Different extracts of cannabis have been used

therapeutically to relieve, asthma, pain, whooping cough, epilepsy and inflammation.

However, due to the psychoactive effects of cannabis such as anxiety, cognitive

impairment and hallucinations, medicinal use of these compounds was discontinued

(Robson, 2001).

Nevertheless, its therapeutic potential to solve a myriad of medical problems has

attracted considerable research interest. Thus, in recent years research has focused on

the therapeutic potential of the plant to develop new chemicals that retain the therapeutic

benefits without causing psychotropic effects.

5

Figure 3. Illustrating Major Phytocannabinoids Synthetic Pathways.

6

1.3 Cannabinoids

Despite its use for centuries, the active ingredients and chemical constituents

Of the cannabis plant remained largely unknown until the beginning of the 1960s when

the main psychoactive constituent, THC was extracted and identified (Mechoulam &

Gaoni, 1965). In the 1980s, other compounds with chemical structures similar to that of

THC were synthesized by various drug companies in an attempt to discover new drugs to

alleviate pain. Some of these newly synthesized compounds such as CP55,940,

CP55,244, HU-210, JWH-018 displaced radioactively labeled THC in radioligand

binding experiments and lead to the discovery of the first cannabinoid receptor

(CB1) in the brain (Herkenham et al., 1990). In the 1990’s CB1 and CB2 receptors

were cloned, expressed functionally and identified pharmacologically (Pertwee, 1997).

Further studies in this field, indicated that, similar to opioids, there are

endogenous ligands that can bind and activate cannabinoid receptors. These findings

eventually lead to the discovery of “endocannabinoids” i.e. endogenously produced

cannabinoids. Endocannabinoids mainly N-arachidonoylethanolamine (Anandamide;

AEA) and 2-arachidonylglycerol (2-AG) have considerable agonistic effects on the

cannabinoid receptors (Mechoulam et al., 1995), and mimic most of the pharmacological

actions of THC.

7

1.3.1 Phytocannabinoids

As mentioned earlier the most abundant constituents of the cannabis plant are

THC and CBD. Within the plant, phytocannabinoids are synthesized from fatty acid

precursors via a series of transferase and synthase enzymes (Figure 3). The two major

phytocannabinoids, THC and CBD, are derived from a common synthetic precursor,

cannabigerol. From a pharmaco-chemical perspective, whilst THC and CBD have pentyl

side chains, major homologues are Δ9-tetrahydrocannabivarinand cannabidivarin,

respectively, with propyl side chains, derived from cannabigerovarin. Importantly,

despite only small differences in chemical structure, these compounds appear to exhibit

markedly different pharmacological properties (Hill, Williams, Whalley, & Stephens,

2012).

THC has a wide spectrum of pharmacological effects ranging from anti-

inflammatory and analgesic actions to cancer treatment, mainly through the activation of

the cannabinoid receptors (Mechoulam, Parker, & Gallily, 2002). However, CBD has no

agonistic activity on cannabinoid receptors, but it is capable of producing various

pharmacological effects by acting on diverse groups of target proteins ranging from

enzymes and receptors to neurotransmitter transporters and ion channels (Hill et al.,

2012).

8

1.3.2 Endocannabinoids

Endogenous ligands for the cannabinoid receptor are synthesized from

polyunsaturated fatty acids and fatty acid derivatives from glycerol. They mimic the

actions of phytocannabinoids by binding to cannabinoid receptors (Di Marzo, 2008). In

addition to AEA and 2-AG, other endocannabinoids with similar fatty acid-based

chemical structures have also been identified in the last decade (Pertwee et al., 2010).

Some of these endocannabinoids include 2-arachidonylglycerol ether (Noladin ether);

(Hanus et al., 2001), NADA, and virodhamine (Pertwee et al., 2010; Porter et al., 2002).

Endocannabinoids, unlike neurotransmitters, are neither synthesized nor stored in

the neurons. Therefore they are not released by a synaptic process, but produced in

response to cellular activity (Di Marzo et al., 1994). It has been shown that biological

processes such as membrane depolarization and activation of metabotropic receptors

cause increases in intracellular Ca2+

levels and activate enzymes involved in the synthesis

of endocannabinoids (Di Marzo, 2008). Following their synthesis, AEA and 2-AG are

degraded by enzymatic hydrolysis by fatty acid amide hydrolase (FAAH) and

monoacylglycerol lipase (MAGL), respectively (Cravatt & Lichtman, 2002).

Endocannabinoids can also be transported out of the cell by diffusion through cell

membranes or by a mechanism that involves an unidentified transporting protein located

in the plasma membrane (Cravatt & Lichtman, 2002; Glaser, Kaczocha, and Deutsch,

2005).

9

1.3.3 Synthetic Cannabinoids

Following the discovery of cannabinoid receptors there have been attempts to

produce compounds with therapeutic effects such as anti-inflammatory/analgesic and

anti-emetic effects without inducing the psychotropic effects. As a result various

chemicals such as CP55,940, CP55,244,HU-210, JWH-018, and JWH-073 have been

synthesized (Huffman et al., 1998). These compounds are known collectively as synthetic

cannabinoids (Ashton, 2012). Currently, little is known about the biological activities of

these compounds with their diverse chemical structures. However, recent reports indicate

that various combinations of synthetic cannabinoids are employed for recreational

purposes throughout the world in order to avoid legal regulations related to

phytocannabinoids (Burstein and Zurier, 2009). However, the synthetic cannabinoid

Nabilone has recently been approved by the FDA (U.S Food and Drug Administration)

and recommended for the treatment of chemotherapy-induced nausea and vomiting that is

unresponsive to conventional antiemetics (Ware, Daeninck, and Maida, 2008).In addition

nabilone is used for the treatment of anorexia and weight loss in AIDS patients (Burstein

and Zurier, 2009; Pertwee et al., 2010). Nabilone has also been reported to be an effective

pain killer for patients with fibromyalgia (Skrabek, Galimova, Ethans, & Perry, 2008).

Currently the following cannabinoid based medications have been approved by

the FDA and are commercially available (Schicho & Storr, 2011): 1– Sativex® (GW

Pharmaceuticals), for pain and spasticity in patients with multiple sclerosis. Currently,

four different formulations of Sativex® are under investigation, including the high

THC extract (Tetranabinex®), THC:CBD (narrow ratio), THC:CBD (broad ratio)

and the high CBD extract (Nabidiolex®) (Scuderi et al. 2009). currently, three

10

Sativex® delivery systems exist: the oromucosal spray, sublingual tablets and inhaled

dosage forms. In 2005, the oromucosal spray administration of Sativex® was approved

for the treatment of multiple sclerosis symptoms (Perras 2005). To date, Sativex

preparations have been licensed in Canada, the UK and Spain. 2– Marinol® (Solvay

Pharmaceuticals, Belgium), oral capsules containing dronabinol (a synthetic THC), are

recommended as an appetite stimulant and antiemetic. 3– Cesamet® (Valeant

Pharmaceuticals International), is an oral capsule containing nabilone (a synthetic THC

analog), for patients with chemotherapy-induced nausea and vomiting. This drug is

currently licensed in Canada, the USA and the UK.

11

Table 1. Class of Cannabinoids

Major Class Name Action of CB

Receptors

Phytocannabinoids ∆9 Tetrahydrocannabinol (∆9 THC) Non selective CB1/CB2

agonist

Cannabigerol (CBG) CB1 selective

antagonist

Cannabidiol (CBD ) CB1 antagonist

CB2 inverse agonist

Cannabichromene (CBC) No action on CBRs

Cannabinol (CBN) Non selective CB1/CB2

agonist

Endocannabinoids 2-Arachidonylethanol Amide

(Anandamide)

Non selective CB1/CB2

agonist

2-Arachidonylglycerol (2-AG) Non selective CB1/CB2

agonist

O-Arachidonylethanolamine

(Virodamine)

Non selective CB1/CB2

agonist

12

Synthetic

Cannabinoids

CP55,940 Non selective CB1/CB2

agonist

WIN55,212 Non selective CB1/CB2

agonist

Sch.336 CB2 selective

Ajulemic acid (AJA) Non selective CB1/CB2

agonist

13

1.4 Cannabidiol

CBD is the main non-psychoactive component of cannabis and constitutes up to

40% of cannabis extracts (Grlic, 1962). It was first isolated in 1940 and its chemical

structure was identified in 1963 (Mechoulam and Hanus, 2002). In recent years, CBD has

been considered to be a promising therapeutic agent for clinical use, since, in addition to

its non-psychotropic properties, it exhibits as well as low toxicity in humans and has a

high therapeutic index and a low teratogenic potential (Rosenkrantz, Fleischman, and

Grant, 1981; Campos, Moreira, Gomes, Del Bel, and Guimaraes, 2012; Scuderi et al.,

2009). An orally-administered liquid containing CBD has received orphan drug status in

the US, for use as a treatment for Dravet Syndrome, a rare-form of epilepsy that is

resistant to treatment with current anti-epileptic drugs (Gloss, Nolan, and Staba, 2014).

Cannabidiol is insoluble in water but soluble in organic solvents, such as DMSO

and alcohol. It is rapidly distributed when administered intravenously, easily passes the

blood-brain barrier and has a prolonged elimination from the body (Samara, Bialer and

Harvey, 1990).

1.4.1 Mechanisms of Action of Cannabidiol

In general, CBD has been shown to act on a wide range of molecular and cellular

target sites. Some of these molecular targets are various types of membrane receptors

such as GPR55 (Ryberg et al., 2007), opioid (Kathmann, Flau, Redmer, Trankle and

Schlicker, 2006) and 5HT1A receptors as well as Transient Vanniloid receptors 1

(TRPV1) (Bisogno et al., 2001), and Peroxisome Proliferator Activated Receptor

(PPARγ) (Bishop-Bailey, 2000). Significantly, CBD, at low µM concentrations acts as an

antagonist for CB1 and CB2 receptors (Pertwee, Ross, Craib and Thomas, 2002). Other

14

proteins that CBD interacts with include voltage-gated and ligand-gated ion channels, and

various types of enzymes (for review see Pertwee, 2008; Izzo et al., 2009; Campos et al.,

2012;Garcia-Arencibia et al., 2007). These cellular and molecular targets are discussed in

detail in the following sections.

a. Actions of cannabidiol on the cannabinoid system:

In earlier clinical studies, CBD, with no detectable activity by itself, was reported

to attenuate some of the effects of THC such as anxiety and tachycardia (Karniol,

Shirakawa, Kasinski, Pfeferman and Carlini, 1974). Similarly, it was found that the

pharmacological actions of cannabinoid receptor agonists can be antagonized by low (nM

range) concentrations of CBD in a mouse’s vas deferens (Pertwee et al., 2002). These

effects of CBD are unlikely to involve direct interaction at cannabinoid receptor CB1 and

CB2 binding sites except at high concentrations of CBD, since CBD appears to have a

relatively low affinity for cannabinoid receptors. Ki values for CBD-induced

displacement of a radiolabelled ligand from cannabinoid CB1 and CB2 receptor binding

sites have been reported to be 4.35 and 2.86 µM, respectively in one study (Showalter,

Compton, Martin and Abood, 1996) and >10 µM in other experiments (Devane, Dysarz,

Johnson, Melvin and Howlett, 1988; Bisogno et al., 2001). Further studies indicated that

CBD behaves as a high-potency antagonist for cannabinoid receptor agonists in a

mouse’s brain tissue and in membranes from CHO cells transfected with cannabinoid

receptors (Thomas et al., 2007). It was also reported that the function of a novel receptor

GPR55 is inhibited by low concentrations of CBD (Ryberg et al., 2007).

15

The interaction of CBD with the cannabinoid system is not limited to cannabinoid

receptors. Tissue levels of endogenously produced cannabinoids are also affected by

CBD. For example, it has been shown that CBD inhibits the uptake and the enzymatic

hydrolysis of AEA (Bisogno et al., 2001). Furthermore, CBD has been shown to cause

activation of cannabinoid receptors indirectly by enhancing tissue concentrations of

endocannabinoids (Fowler, 2004). In functional experiments it has been suggested that

some of the effects of CBD sensitive to CB receptor antagonists are mediated by an

entourage effect of CBD on tissue levels of endocannabinoids (Fowler, 2004).

b. Effects of cannabidiol on other membrane receptors:

CBD has been shown to stimulate Vanilloid Receptors (VR1) with an efficacy

comparable to that of capsaicin, the natural agonist of this receptor (Bisogno et al.,

2001). The stimulation of the VR1 receptor by capsaicin and some of its analogues leads

to fast desensitization, with subsequent analgesic and anti-inflammatory effects. It has

been suggested that, similar to capsaicin, CBD also causes desensitization of VR1 and

enhances the effect of capsaicin on VR1, suggesting that CBD exerts its anti-

inflammatory actions, in part, by desensitization of sensory nociceptors (Bisogno et al.,

2001).CBD has been reported to act as a weak agonist (EC50 = 3.5 µM) on human

TRPV1 receptors inHEK293 cells expressing these receptors (De Petrocellis et al., 2011).

CBD has also been demonstrated to act as an agonist on TRPV2 (Qin et al., 2008) and

TRPA1 receptors, while it acts as an antagonist on the TRPM8 receptor (De Petrocellis et

al., 2008).

It was also shown that at a µM concentration range, CBD displaces the 8-OH-

DPAT, 5-HT1A receptor agonist, from cloned human 5-HT1A receptors expressed in

16

cultured cells obtained from a Chinese hamster ovary (E. B. Russo, Burnett, Hall and

Parker, 2005). In subsequent studies, results of several in vivo experiments have

supported the involvement of 5-HT1A receptors in the effect of CBD (Campos et al.,

2012). At high µM concentrations (10-100 µM), CBD has also been shown to interact

with other G-protein coupled receptors such as 5HT2 and opioid receptors in an allosteric

manner (E. B. Russo et al., 2005; Kathmann, et al., 2006).

c. Effect of cannabidiol on second messengers:

Calcium is considered one of the most important second messengers involved in

the regulation of various cell functions, such as muscle contraction and the release of

neurotransmitters. It has been shown that CBD treatment increases intracellular

concentrations of Ca2+

in cultured hippocampus neurons (Drysdale, Ryan, Pertwee and

Platt, 2006; Ryan, Drysdale, Lafourcade, Pertwee and Platt, 2009). In another study, the

anti-epileptic activity of CBD has been suggested to be due to its bidirectional regulatory

role on intracellular Ca2+

levels (Ryan et al., 2009). In another investigation, it was

proposed that CBD increases intracellular Ca2+

levels under normal physiological

conditions, but decreases Ca2+

levels under highly excitable conditions (Mato, Victoria

Sanchez-Gomez and Matute, 2010; Rao and Kaminski, 2006). This action of CBD was

believed to be due to the regulation of Ca2+

transport in mitochondria which acts as a sink

for intracellular Ca2+

(Ryan et al., 2009).

In addition to its influence on intracellular stores, CBD also acts on active Ca2+

transport proteins such as Ca2+

-ATPase in myocytes (Gilbert, Pertwee and Wyllie, 1977)

and neurons (Gilbert et al., 1977; Ryan et al., 2009). The highly lipophilic nature of CBD

grants it easy access to intracellular targets such as cellular organelles (endoplasmic

17

reticulum and mitochondria) and intracellularly located enzymes that contribute to CBD-

induced alterations in intracellular Ca2+

levels (Collins and Haavik, 1979).In this context,

it is important to note that, the functions of various Ca2+

-activated enzymes such as

phospholipase C (PLC) (Howlett, Scott and Wilken, 1989), phospholipaseA2,

lipoxygenase (Takeda, Usami, Yamamoto and Watanabe, 2009) and Phosphokinase C

(PKC) (Hillard and Auchampach, 1994; White and Tansik, 1980) are also modulated by

low µM concentrations of CBD.

In addition to Ca2+

, CBD acts on the uptake process of several neurotransmitters

(for a review see Pertwee 2008). CBD decreases the uptake of [3H] adenosine in both

murine microglia and macrophages, and binding studies show that CBD binds to the

nucleoside transporter (Carrier, Auchampach and Hillard, 2006). The enhancement of

adenosine signaling through inhibition of its uptake was suggested as providing non-

cannabinoid receptor mechanism by which CBD can decrease inflammation (Carrier et

al., 2006).

d. Cannabidiol effects on voltage-gated ion channels:

Earlier studies of motor neurons showed that CBD depresses the amplitude of

action potential suggesting that it inhibits the function of neuronal voltage-gated Na+

channels (Turkanis and Karler, 1986). However, there has been no further study

investigating the effect of CBD on Na+ channels. Another study indicated that CBD

increased the L-type Ca2+

current recorded in the hippocampal neurons (Drysdale et al.,

2006). Similarly to these Na+ channels, there have been no further studies investigating

the actions of CBD on L-type Ca2+

channels. On the other hand, the effect of CBD on T-

type voltage-gated Ca2+

channels has been investigated in recent years. It has been shown

18

that CBD inhibits different subtypes of CaV3 family channels (Ross, Napier and Connor,

2008).

e. Cannabidiol effects on ligand-gated ion channels:

In HEK-293 cells expressing glycine receptors, CBD showed a positive allosteric

modulatory effect in a low µM concentration range (Ahrens et al., 2009). The EC50

values for the potentiating effects of CBD were 12.3 µM for α1 and 18.1 µM for α1β1

subunit combinations. Direct activation of glycine receptors was also observed at CBD

concentrations above 100 µM. Similarly, glycine-mediated currents recorded in dorsal

horn neurons of rat spinal cord slices were potentiated by CBD applications (Xiong et al.,

2012). Further investigations indicated that mutations of the α1 subunit TM2 serine

residue to isoleucine abolished the co-activation and the direct activation, of the glycine

receptor by CBD (Foadi et al., 2010). In in vivo experiments, it has been shown that

systemic and intrathecal administration of CBD significantly suppressed chronic

inflammatory and neuropathic pain without causing apparent analgesic tolerance in

rodents (Xiong et al., 2012). The analgesic potency of CBD and 11, similarly structured

cannabinoids, is positively correlated with potentiation of the α3-subunitglycine receptor.

In contrast, analgesia induced by these cannabinoids is neither correlated with their

binding affinity for CB1 and CB2 receptors nor with their psychoactive side effects.

Furthermore, NMR analysis revealed a direct interaction between CBD and S296 in the

third transmembrane domain of purified α3 GlyR. More importantly, the CBD-induced

analgesic effect was absent in mice lacking the α3 subunit from the glycine receptor

suggesting that the α3 subunit mediates glycinergic CBD-induced suppression of chronic

pain (Xiong et al., 2012).

19

Another ligand-gated ion channel reported to be modulated by CBD is the α7-

nicotinic cholinergic receptor (Mahgoub et al., 2013). It was shown that the function of

human α7-nicotinic receptors expressed in xenopus oocytes was inhibited by CBD with

an IC50 value of 11.3 µM in a non-competitive manner. Significantly, the 5-HT3

receptor, with structural similarities to the α7-nicotinic receptor was also shown to be

inhibited in xenopus oocytes and HEK-293 cells in an allosteric manner (K. H. Yang et

al., 2010).

20

1.4.2 Clinical Applications

a. Anti-inflammatory actions

As the most abundant non-psychoactive cannabinoid in the plant, the anti-

inflammatory actions of CBD and its analogs have been studied extensively in recent

years (Burstein and Zurier, 2009; Booz, 2011). CBD reduces joint inflammation in

collagen-induced arthritis in mice (Sumariwalla et al., 2004) and carrageenan paw edema

in rats (Costa et al., 2004). In addition, oral administration of CBD (2.5–20 mg/kg)

reduces neuropathic and inflammatory pain in rats. This effect is reversed by vanilloid

but not by CB receptor antagonists (Costa, Trovato, Comelli, Giagnoni and Colleoni,

2007). CBD also reduces intestinal inflammation in mice (Capasso et al., 2008). In

addition to its ability to suppress production of the inflammatory cytokine TNFα, CBD

appears to exert anti-inflammatory pressure by suppressing the fatty acid amidohydrolase

(FAAH), thereby increasing concentrations of the anti-inflammatory endocannabinoid

anandamide. Further, insight into mechanisms whereby CBD exerts therapeutic effects

was provided by experiments which indicated that CBD attenuates inflammation induced

by high glucose levels in diabetic mice (Rajesh et al., 2007). Specifically, CBD treatment

reduces mitochondrial superoxide, inducible nitric oxide synthase, nuclear factor kappa β

activation, and transendothelial migration of monocytes (Burstein and Zurier, 2009;

Booz, 2011).Another potential therapeutic use of CBD may lie in its ability to counter

some undesirable effects of THC (sedation, psychotropic effects, tachycardia), thus

suggesting that if given together with THC, it may allow higher doses of THC (E. Russo

& Guy, 2006). THC and CBD have been administered as an oral mucosal spray to 58

patients with rheumatoid arthritis (Blake, Robson, Ho, Jubb and McCabe, 2006). Treated

21

patients had a significant reduction in pain and an improvement in sleep compared to

patients given a placebo.

b. Neuroprotection

In earlier studies, CBD has been shown to normalize glutamate homeostasis

(Hampson et al., 2000; El-Remessy et al., 2003), reduce oxidative stress (Hampson,

Grimaldi, Axelrod and Wink, 1998; Marsicano, Moosmann, Hermann, Lutz, and Behl,

2002), attenuate glial activation and the occurrence of local inflammatory events (Ruiz-

Valdepenas et al., 2011; Martin-Moreno et al., 2011). It appears that there may be two

key mechanisms underlying the neuroprotective effects of CBD. The first is the

capability of CBD to restore the normal balance between oxidative events and antioxidant

endogenous mechanisms (Fernandez-Ruiz, 2012) that are frequently disrupted in

neurodegenerative disorders, thereby enhancing neuronal survival. The second is CBD as

a neuroprotective compound, and its anti-inflammatory activity. The anti-inflammatory

effects of CBD are related to the control of microglial cell migration and the toxicity

exerted by these cells, i.e.the production of pro-inflammatory mediators (Fernandez-Ruiz,

2012).

The neuroprotective effects of CBD were also evaluated in animal models with

Parkinson’s disease, Huntington’s disease and neonatal ischemia. Pathological changes

induced by 6-hydroxydopamine, a toxic compound that targets catecholaminergic cells,

were significantly reduced by CBD treatment (Lastres-Becker, Molina-Holgado, Ramos,

Mechoulam and Fernandez-Ruiz, 2005; Mechoulam, et al., 2002). It appears that these

reductions after CBD treatment were irreversible, as they did not recover when CBD was

halted (Mechoulam, Peters, Murillo-Rodriguez and Hanus, 2007). CBD was also useful

22

in preventing beta-amyloid–induced neurodegeneration in in vivo and in vitro models of

Alzheimer disease (Fernandez-Ruiz, 2012). Increased production of reactive oxygen

species, lipid peroxidation, DNA fragmentation, and intracellular Ca2+

concentrations

induced by beta-amyloid peptide were significantly reduced after CBD treatment of PC12

cells (Iuvone et al., 2004).

CBD has also been shown to be protective against neuronal damage due to

ischemia (Hampson et al., 2000). In rats subjected to middle cerebral artery occlusion,

infarct size and neurological impairment were reduced by 50-60% by CBD. Similarly,

post-ischemic administration of CBD protected against hyperlocomotion and neuronal

injury following middle cerebral artery occlusion in gerbils (Braida et al., 2003).

Furthermore, it was recently shown that CBD treatment reduced the infarct volume in

brain ischemia, and that this effect was independent of cannabinoid receptor type 1and

transient receptor potential V1, but sensitive to the 5-HT1A receptor antagonist,

WAY100135 (Hayakawa et al., 2004; Mishima et al., 2005) suggesting that activation of

5-HT1A receptors mediate the neuroprotective effects of CBD.

c. Treatment of nausea and vomiting

Cannabis has long been used to treat nausea and vomiting induced by various

drugs and pathological conditions (Parker, Rock and Limebeer, 2011). Several

combinations of phytocannabinoids and synthetic cannabinoids have currently been

approved for use in the treatment of chemotherapy-induced nausea and vomiting (Parker

et al., 2011).The reduction of nausea and vomiting demonstrated by cannabis sativa use

in vivo and in vitro has been attributed to the presence of THC and CBD. Unlike THC,

which exerts its action via CB1 receptors, CBD mimics the anti-nausea and anti-vomiting

23

properties of THC through mechanisms unrelated to CB receptor activation (Mechoulam

et al., 2007). Interestingly, one of the most effective antiemetic drugs used in clinics is a

5HT3 receptor antagonist (Thompson, Chau, Chan and Lummis, 2006). 5HT3 receptors

belong to a superfamily of ligand-gated ion channels that mediate serotonin-induced

rapid depolarizations in intestinal neurons and regulate peristaltic activity of intestinal

smooth muscle (Izzo and Camilleri, 2009). CBD and THC have been shown to inhibit

5HT3receptors (K. H. Yang et al., 2010)and suppress emesis induced by various chemical

and physical stimuli (Parker et al., 2011).

24

d. Cardiovascular effects

i) Anti-arrhythmic role of CBD:

Studies examining the cardioprotective effects of CBD showed that CBD has

antiarrythmic effects after an ischemic reperfusion injury (Walsh, Hepburn, Kane and

Wainwright, 2010). In models of myocardial infarction following ligation of the left

anterior descending coronary artery, in vivo treatment with CBD resulted in a reduction

in infarct size (Rajesh et al., 2010). This finding has been attributed to an immuno-

modulatory effect in CBD, since it was accompanied by a reduction in leukocyte

infiltration and interleukin 6 concentrations. A single dose of CBD given 10 minutes

prior to ischemia or reperfusion resulted in a reduction in infarct size, and a significant

reduction in the incidence of ventricular tachycardia and total number of ventricular

ectopic beats (Stanley, Hind and O'Sullivan, 2013).

The CB1 receptor antagonist, AM251, also demonstrates antiarrhythmic

properties, and treatment with AM251 co-administered with CBD has a synergistic

effect, suggesting that activation of CB1 receptors is proarrhythmic (Walsh et al., 2010).

The synergism observed, which persists when CB1 receptors are blocked prior to CBD

administration, is suggestive of cross talk between CB1 and other CB receptors in the

ischemic heart (Walsh et al., 2010).Furthermore, CBD significantly reduces cardiac

dysfunction in mice with diabetes. CBD treatment has been shown to decrease

myocardial inflammation, reduce oxidative stress as demonstrated by a reduction in

nuclear factor-kB activation, suppress mitogen-activated protein kinase activation, and

reduce the expression of adhesion molecules and tumor necrosis factor (Walsh et al.,

2010).

25

ii) Vascular effects of CBDs:

CBDs exert vasodilative effects as demonstrated by in vivo and in vitro models

(see Stanley, Hind and O'Sullivan, 2013). However, the mechanism and potency of CBD

appear to differ between various experimental models. For example, in some studies,

CBD acts through cyclooxygenase, while other experiment use through CB1 receptor

antagonism and the inhibition of potassium (K+) channel hyperpolarization. The size of

the blood vessel is also a key factor, as maximal responses to anandamide (AEA) are

observed in small resistance vessels (O'Sullivan, Kendall and Randall, 2005).

In rat aorta, time-dependent CBD vasodilatation has been shown to be mediated

through Gi/o protein that is sensitive to pertussis toxin, but is not sensitive to CB1

antagonism or TRPV1 desensitization (O'Sullivan, Sun, Bennett, Randall and Kendall,

2009). Additional differences are observed in other species. The mechanism by which

vasodilatation occurs is also dependent on the type of cannabinoid. For instance, AEA

and N-arachidonoyl-dopamine (NADA) exert similar vasodilatation effects in rat aorta;

however, these effects occur through different mechanisms (Stanley, et al., 2013).

It has also been demonstrated that CBD induces vasodilatation of segments of

human mesenteric artery preconstricted with U46619 and endothelin-1. The authors of

this study determined that mesenteric artery relaxation is endothelium-dependent, and

involves CB1 receptor and TRPV channel activation, NO release and K+ channel

hyperpolarization (Stanley et al., 2013).

iii) Relationship between CBD and peroxisome proliferator-activated receptor

gamma:

26

Nuclear receptors are proteins found within cells that regulate diverse functions,

such as homeostasis, reproduction, development, and metabolism. Peroxisome

proliferator-activated receptor gamma (PPARγ) belongs to a family of nuclear receptors

and plays a crucial role in glucose and lipid homeostasis, in addition to cell proliferation,

cell differentiation, and inflammatory responses (Bishop-Bailey, 2000; Hsueh and

Bruemmer, 2004). There is evidence to suggest that cannabinoids bind to and activate

PPARγ, thereby causing PPAR-mediated responses. THC, which is the major component

of cannabis, causes time-dependent endothelium-dependent and PPARγ-mediated

vasodilatation of rat aorta. In contrast, CBD, which is a weak agonist of PPARγ

receptors, causes increased PPARγ transcriptional activity in PPARγ-overexpressing

HEK293 cells (Hsueh and Bruemmer, 2004).

Since CBD is a weak PPARγ receptor agonist, side effects normally associated with

PPARγ receptor activation, such as weight gain, edema, and increased plasma lipoprotein

can be avoided. Therefore, CBD may prove to have therapeutic potential as a low affinity

agonist (O'Sullivan et al., 2009; Walsh et al., 2010).

iv) Hemodynamic effects of CBD:

Studies on the hemodynamic effects of CBD proved inconclusive. In some studies,

a 16-mm Hg reduction in mean arterial blood pressure was noted, with no change in heart

rate (Walsh et al., 2010). The authors of this study concluded that the effects of CBD are

best seen in models with elevated blood pressure (Walsh et al., 2010). However, other

studies did not demonstrate significant hemodynamic changes (Bergamaschi, Queiroz,

Zuardi and Crippa, 2011).

27

CBD is reported to have anxiolytic properties, and is effective as a treatment for a

fear of public speaking. CBD reduced heart rate and blood pressure responses in Wistar

rats conditioned by fear. This response is believed to be mediated by 5HT1A receptors,

since the effects were inhibited by the 5HT1A receptor antagonist, WAY100635 (Gomes,

Resstel and Guimaraes, 2011; Zuardi, 2008).

v) Vascular protective effects of CBD:

High glucose levels are known to contribute to endothelial dysfunction in patients

with diabetes. High glucose levels promote the inhibition of endothelial NO, decrease the

vasodilatation effects and increase the vasoconstrictor effects of prostanoids, increase

superoxide production, and increase ROS production (Vanhoutte, Shimokawa, Tang and

Feletou, 2009). The aforementioned effects of high glucose were reduced when cells

were co-incubated with CBD (Rajesh et al., 2007). Key elements of atherosclerosis, such

as monocyte adhesion and transendothelial lmigration are reduced by CBD. Neither CB1

nor CB2 receptors appear to be responsible for mediating these effects of CBD (Rajesh et

al., 2007). Treatment with CBD may also protect against diabetic retinopathy. CBD

prevented vascular hyperpermeability at the blood-retinal barrier (BRB), and protected

the retina against oxidative damage, inflammation, and increased levels of cell adhesion

molecules in an in vivo model of diabetic retinopathy (El-Remessy et al., 2006).

28

1.5 Contraction of Cardiac Muscle

Contraction of cardiac muscle cells or cardiomyocytes shows certain characteristics

specific to this muscle. For example, contraction of cardiomyocytes is not initiated by

neurons as in skeletal muscle but by electrical excitation originating from the heart’s own

pacemaker, the sinoatrial node, which generates spontaneous and periodic action

potentials. When an action potential is initiated in one cell, current flows through the gap

junctions and depolarizes neighboring cells. If depolarization causes the membrane

potential to be more positive than the threshold, self-propagating action potentials occur

in the neighboring cells as well. Thus, the generation of an action potential is just as

critical for initiating contraction in cardiac muscle as it is in skeletal muscle, but it is

triggered by the sinoatrial node and the specialized conduction system of the heart

(Figure 4).

Figure 4. Anatomy and Function of the Heart's

Electrical System. Johns Hopkins Medicine,

Baltimore, Maryland.

29

Separate tubular structures, the transverse tubules (T tubules), cross the cell. In the

cardiac myocyte, the T tubule crossings occur at the Z-line, in contrast to the A-I junction

in skeletal muscle. The lumen of the T tubule is continuous with the extracellular fluid

surrounding the cell and, as in skeletal muscles, the action potential is propagated down

the T tubule (Ferrantini et al., 2013). Adjacent cardiac myocytes are joined end-to-end at

structures known as intercalated disks. These always occur at a Z-line (Figure 5). At these

points, the cell membranes form a number of parallel folds and are tightly held together

by desmosomes. This results in strong cell-to-cell cohesion, thus allowing the contraction

of one myocyte to be transmitted axially to the next. Gap junctions exist between cardiac

muscle cells, providing low resistance pathways for the spread of excitation from one cell

to another.

Figure 5. Adjacent cardiac myocytes are joined end-to-end at

structures known as intercalated disks.

30

Figure 6. Ventricular action potential membrane currents that generate a

normal action potential. Resting (4), upstroke (0), early repolarization (1), pla-

teau (2) and final repolarization are the 5 phases of the action potential. A decline of

potential at the end of phase 3 in pacemaker cells, such as the sinus node, is shown as

a broken line. The inward currents, INa, ICa and If, are shown in yellow boxes; the

sodium-calcium exchanger (NCX) is also shown in yellow. It is electrogenic and may

generate inward or outward current. IKAch, IK1, Ito, IKur, IKr, and IKs are shown in

gray boxes. The action potential duration (APD) is approximately 200 ms (Nattel &

Carlsson, 2006).

31

1.5.1 Excitation–Contraction Coupling

a. The ventricular action potential

The cardiac action potential represents changes in the membrane potential due to

the movement of ions across the cell membrane through voltage-gated ion channels,

pumps, and exchangers, and has distinct characteristics in different regions of the heart.

In ventricular cardiomyocytes (Amin, Tan and Wilde, 2010). According to kinetic

properties, the cardiac action potential is divided into 4 distinct phases (Figure 6). The

first phase of the action potential is characterized by the activation of fast Na+ channels,

which open briefly to produce an influx of positive Na+ ions causing rapid depolarization

of the cell membrane and an upstroke of the action potential. This initial phase is called

phase 0. Following phase 0, a brief repolarization is observed at the peak of the action

potential and is the consequence of both fast Na+ channel inactivation and the activation

of initial repolarizing currents. This brief repolarization phase is called phase 1.

Following phase 1, a long lasting plateau of depolarization (100-200 ms), named phase 2,

is observed. Depolarizing currents, mainly late Na+ and L-type Ca

2+ channels stay open

and continue to offset repolarizing currents during the plateau phase. Finally, L-type Ca2+

channels begin to inactivate and several cardiac K+ channels such as delayed-rectifier and

inward-rectifier K+ channels begin to open and cause the repolarization of the membrane

to resting membrane potential. This repolarization is called phase 3. The ventricular

action potential is regenerative, and continues to excite ventricular cells if the stimulus

exceeds the critical threshold for depolarization as determined by availability of Na+

current. Phase 4 is the resting membrane potential, and describes the membrane potential

when the cardiomyocyte is not being stimulated. In a great majority of contracting

32

cardiomyocytes phase 4 has a low slope (almost a horizontal line). However, in

pacemaking cells, phase 4 is unstable (phase 4 - is the pacemaker potential). In

pacemaker cells (such as the Sinoatrial node), the membrane slowly depolarizes during

phase 4, until it reaches a threshold potential (around -40mV) or until it is depolarized by

an electrical impulse coming from another cell. The reason for this pacemaker potential is

an increased inward current of sodium (Na+) through voltage-dependent channels, but

also an increased inward Ca2+

current and a decrease in the K+ outward current. These

Na+ channels, in cardiac pacemaker cells, contrary to what usually happens in other cells,

open when the voltage is more negative.

b. Calcium-induced calcium release

The action potential propagates along the sarcolemma, invades the transverse

tubules and causes release of Ca2+

from internal Ca2+

stores such as sarcoplasmic

reticulum and leads eventually to the initiation of muscle contraction. This sequence of

events, beginning with action potential generation and resulting in muscle contraction is

known as excitation-contraction coupling (Figure 7). Initially, depolarization caused by

the action potential activates voltage-gated L-type Ca2+

channels which are concentrated

in the Tubules (Shepherd and McDonough, 1998). Activation of these channels induces a

conformational change in the ion channel structure and causes a subsequent influx of

Ca2+

into the cell. This influx of Ca2+

is known as sparklets (Navedo and Santana, 2013).

The presence of sparklets is not sufficient to cause activation of the contractile

machinery. However, it is sufficient to initiate the release of furtherCa2+

from intracellular

stores by interacting with ryanodine receptors (RyRs) on the sarcoplasmic reticulum. The

local amplification of intracellular Ca2+

mediated by RyR activation is also known as a

33

Ca2+

spark (Fearnley, Roderick and Bootman, 2011), and is essential for causing a Ca2+

transient that activates the contractile myofilaments resulting in muscle contraction. In

general, cytosolic Ca2+

influx in ventricular cells occurs mainly through two distinct

sources. About 8 to 23% of Ca2+

enters extracellularly through the opening of L-type

Ca2+

channels and 77% to 92% of total Ca2+

employed in muscle contraction is released

from the sarcoplasmic reticulum through RyRs, although some inter-species differences

do exist (Bers, 2008). The contribution of NCX and mitochondria to cytosolic Ca2+

is

usually not more than 1% (Bers, 2000). Following its release, Ca2+

is pumped back into

the sarcoplasmic reticulum by a Ca2+

-ATPase pump and also removed from the cell via

the Na+/Ca

2+ exchanger located at the plasma membrane (Wang, Song, Lakatta and

Cheng, 2001).In cardiac muscle, the activity of a Ca2+

-ATPase in the sarcoplasmic

reticulum is inhibited by the regulatory protein phospholamban. When phospholamban is

phosphorylated by cAMP-dependent protein kinase, its ability to inhibit the Ca2+

-ATPase

is lost. Thus, activators of cAMP-dependent protein kinase, such as the neurotransmitter

epinephrine, may enhance the rate of cardiac myocyte relaxation.

34

Figure 7. Cardiac excitation contraction coupling. Inset shows the time course of an action

potential, Ca2+

transient and contraction measured in a rabbit ventricular myocyte at 37 °C. NCX,

Na+/Ca

2+ exchange; ATP, ATPase; PLB, phospholamban; SR, sarcoplasmic reticulum (Bers -

‎2002).

Calcium Channels

Due to the importance of Ca2+

in various vital cellular events such as muscle

contraction, neurotransmitter release, and exocytosis; the regulation of Ca2+

entrance

through voltage-gated Ca2+

channels has been studied in detail (Catterall, Perez-Reyes,

Snutch and Striessnig, 2005). According to pharmacological and biophysical

characteristics three different types of Ca2+

channels have been characterized. These

different types of Ca2+

channels are summarized in Table 2.

35

In the heart, both T-type and L-type Ca2+

channels have been shown to be

expressed (Zhou and January, 1998). However, T-type Ca2+

channels appear to be

localized mainly in cells involved in heart rhythm and automaticity such as pacemaker

and atrial cells, while L-type Ca2+

channels are localized mainly in ventricular myocytes.

As mentioned above, due to their slow activation kinetics, L-type Ca2+

channels require

long depolarizations lasting 100-200 ms. After the activation process, L-type Ca2+

channels are inactivated by various factors which include time, voltage, cytosolic Ca2+

concentration and calmodulin (Bers, 2000).

36

Table 2. Classification of calcium channels

Ca2+

current

type

α1

subunits

Specific

blocker

Principal physiological functions Inherited diseases

L Cav 1.1 DHPs Excitation-contraction coupling in

skeletal muscle, regulation of

transcription

Hypokalemic periodic

paralysis

Cav1.2 DHPs Excitation contraction coupling in

cardiac and smooth muscle, endocrine

secretion, neuronal Ca2+

transients in

cell bodies and dendrites, regulation of

enzyme activity, regulation of

transcription

Timothy syndrome: cardiac

arrhythmia with

developmental abnormalities

and autism spectrum

disorders

Cav1.3 DHPs Visual transduction Stationary night blindness

N Cav2.1 ω-CTx-GVIA Neurotransmitter release, Dendritic

Ca2+

transients

P/Q Cav2.2 ω-Agatoxin Neurotransmitter release, Dendritic

Ca2+

transients

Familial hemiplegic

migraine, cerebellar ataxia

R Cav2.3 SNX-482 Neurotransmitter release, Dendritic

Ca2+

transients

T Cav3.1 None Pacemaking and repetitive firing

Cav3.2 Pacemaking and repetitive firing Absence seizures

Cav3.3

Abbreviations: DHP, dihydropyridine; Ѡ-CTx-GVIA, Ѡ –conotoxin GVIA from the cone snail

Conusgeographus; SNX-482, a synthetic version of a peptide toxin from the tarantula

Hysterocratesgigas (Catterall, 2011).

As shown in Figure 8, the L-type Ca2+

channel consists of 4 subunits: the major

pore-forming 1 subunit and a variety of accessory subunits. Theα1 subunit: Formed of 4

homologous domains (I–IV), each consisting of 6 transmembrane (TM) helices joined by

intracellular links. The loops between TM5 and TM6form the channel pore, while TM4

contains arginine or lysine residues that form the voltage sensor, which is thought to

move or twist as the membrane potential changes.

Ion selectivity is governed by a glutamate residue preserved in the pore-lining

loop of all 4 domains (Dolphin, 2006; J. Yang, Ellinor, Sather, Zhang and Tsien, 1993).

Two Ca2+

ions enter the cell at a time, are regulated by 2 glutamate residues and flow

down the Ca2+

concentration gradient (Heinemann, Terlau, Stuhmer, Imoto and Numa,

1992; J. Yang et al., 1993). Lanthanum, cobalt and cadmium inhibit the α1 subunit

37

(Dolphin, 2009). Expression of the α1 subunit without the auxiliary subunits, results in

low L-type Ca2+

channel expression levels, and abnormal voltage and kinetic properties

in the channel. The β subunit is located on the intracellular side of the channel. It interacts

with the subunit through the -interaction domain (AID), a guanylate cyclase–like

region located between domains I and II (Catterall, 2011; Pragnell et al., 1994). The AID

has a well-conserved gene sequence and also occurs in N-type and P/Q-type

Ca2+

channels (Richards, Butcher and Dolphin, 2004).The β subunit alters channel

expression, voltage dependence, and gating kinetics. Four β subunits have been cloned,

and all except for β2a hyperpolarize activation and steady-state inactivation of Ca2+

(Dolphin, 2003). The α2δ subunit is an accessory subunit consisting of the

transmembrane domain, and 2, an extracellular domain. Four 2 isoforms have been

identified. Co-expression of α2δ subunits increases Ca2+

channel expression and function.

The subunit is a 30 to 40kD inhibitory protein consisting of 4 TM helices (Takahashi,

Seagar, Jones, Reber and Catterall, 1987).

38

Figure 8. Calcium channel subunits. Structure of calcium channels showing the five subunits α,

β, α2δ and (Figure modified from Catterall, 2011).

Secretion of catecholamines from the adrenal medulla and the autonomic nervous

system stimulates β-adrenergic receptors via adenylcyclase, cAMP, and protein kinase A

(PKA) to produce positive chronotropic, inotropic and lusitropic effects. Cardiac L-type

Ca2+

channels are modulated through the β-adrenergic/cAMP signaling pathway; α1 and β

subunits are phosphorylated by PKA (De Jongh et al., 1996; Haase, Bartel, Karczewski,

Morano& Krause, 1996; Hell et al., 1993; Puri, Gerhardstein, Zhao, Ladner and Hosey,

1997). β-adrenergic receptor activation causes a 2-fold increase in the L-type Ca2+

current

by increasing the number of channels and their activation probability. The latter is

mediated through PKA-mediated phosphorylation of the Cav1.2 (1C-subunit) channel

protein and/or associated proteins (Catterall, 2010). As opposed to the adrenergic system,

reciprocal control over Ca2+

entry is provided by ACh, acting through muscarinic Ach

39

receptors, raising intracellular cGMP concentrations and causing cGMP-dependent

phosphorylation of the L-type Ca2+

channels. In turn, the cGMP-dependent

phosphorylation of L-type Ca2+

channels, at sites distinct from those phosphorylated by

the cAMP-dependent kinase, causes a decrease in Ca2+

influx during the cardiac action

potential and thus a decrease in the force of contraction.

40

1.5.2 Sarcoplasmic Reticulum and Ryanodine Receptors

The sarcoplasmic reticulum has many roles in the regulation of muscle

contraction and intracellular Ca2+

homeostasis. The release of Ca2+

from this organelle

causes increased Ca2+

levels to activate various Ca2+

dependent enzymes such as

calmodulin and adenyl cyclase (Fearnley et al., 2011). It is also the site of phospholipid

generation and it communicates with other intracellular organelles, such as the

mitochondria. However, it’s most important function is the regulation of Ca2+

homeostasis. It is well known that cytosolic Ca2+

levels during resting conditions are in

the nM range, while extracellular Ca2+

concentrations are much higher (μM range). As

mentioned above, following an action potential, cytosolic Ca2+

concentration increases

and activates ryanodine receptors, which amplify the Ca2+

signal and stimulate Ca2+

-

induced Ca2+

release. Several isoforms of ryanodine receptors (RyR) have been identified:

RyR1 is expressed in skeletal muscle, RyR2 in cardiac muscle and the brain, and RyR3 in

the brain (Lanner, Georgiou, Joshi and Hamilton, 2010). Although the ryanodine receptor

is a protein of thr SR membrane, 80% of this protein is located in the cytoplasm. The

intracellular position of ryanodine differs from one cell type to another (Zucchi and

Ronca-Testoni, 1997). It can be directly attached (as in skeletal muscles) or detached (as

in cardiac muscles) to the L-type Ca2+

channel. In skeletal muscle, direct interaction of

the ryanodine receptor with the L-type Ca2+

channel causes fast excitation-contraction

coupling and rapid initiation of skeletal muscle contraction (within a few minutes). In

cardiac muscle on the other hand, excitation-contraction is a relatively slow process

(within hundreds of minutes). Function of the ryanodine receptor is also regulated by

several ligands, including Ca2+

, calmodulin and caffeine (Fearnley et al., 2011).

41

1.5.3 The Contractile Machinery

Myofilaments occupy about half of the cell volume in mammalian ventricular

myocytes. As with skeletal muscle, cardiac myocytes contain the contractile proteins

actin (thin filaments) and myosin (thick filaments) together with the regulatory proteins

troponin and tropomyosin. In their structural configuration, the thin filament comprises

actin with a spiraling backbone of tropomyosin, along with troponin complexes (C, I and

T) that occur at certain intervals (Figure 9). Each actin filament consists of a site where

myosin can bind, but these sites are blocked by tropomyosin when cytosolic Ca2+

concentration is low. When cytosolic Ca2+

concentration increases due to Ca2+

release,

Ca2+

binds to troponin C causing tropomyosin to shift, thereby exposing the myosin

binding sites. Adenosine diphosphate (ADP) and phosphate, present on the myosin head

since the previous movement cycle, are released when myosin attaches to myosin binding

sites on the actin filaments. The energy generated in the head is used to initiate upstroke

and sliding, until the actin–myosin complex is liberated by adenosine triphosphate (ATP)

binding the myosin head. ATP is hydrolyzed to ADP and phosphate, releasing energy

that is stored in the myosin head to be used in the subsequent stroke.

42

Figure 9. Sarcomere structure. The thin filament is composed of actin, tropomyosin, and the

troponin which consists of the troponins Tn-T, Tn-C, and Tn-I. The thick filament is composed of

myosin, a globular head portion (S1), a hinged stalk region (S2) and a rod section. S1 consists of

an ATP hydrolysis domain and the actin binding domain, it is associated with the light-chain 1

and 2 (LC-1; LC-2) (Hamdani et al., 2008).

43

1.5.4 Calcium Sensitivity of Myofilaments

The term myofilament Ca2+

sensitivity came into use after it was discovered that

the force of contraction is not only dependent on the free cytosolic Ca2+

concentration,

but also on the affinity of troponin C for Ca2+

. Biochemical and physiological

experiments have shown that several factors affect myofilamentCa2+

sensitivity,

including pH, temperature, ionic strength and troponin I phosphorylation (Ruegg, 2003).

Furthermore, these factors can be categorized by whether they affect thick or thin

filaments (Bers, 2001). Factors that affect thick filaments tend to affect the relationship

between intracellular Ca2+

concentration and the extent of myosin light chain

phosphorylation. Whereas factors that affect thin filaments, affect phosphorylation of

myosin light chain and force generation due to contractile proteins (Bers, 2001).

44

Chapter 2: Methods

2.1 Ventricular Myocyte Isolation

Ventricular myocytes were isolated from adult male Wistar rats (264 ± 19 g)

according to previously described techniques. This study was carried out in accordance

with the recommendations in the Guide for the Care and Use of Laboratory Animals from

the National Institutes of Health. The protocol was approved by the Committee on the

Ethics of Animal Experiments at the United Arab Emirates University. In brief, the

animals were euthanized using a guillotine and their hearts were removed rapidly and

mounted for retrograde perfusion according to the Langendorff method (Figure 10). The

hearts were perfused at a constant flow of 8 ml g heart-1

min-1

and at 36–37 °C with a

solution containing (mM): 130 NaCl, 5.4 KCl, 1.4 MgCl2, 0.75 CaCl2, 0.4 NaH2PO4, 5

HEPES, 10 glucose, 20 taurine and 10 creatine set to pH 7.3 with NaOH. When the heart

had stabilized perfusion was continued for 4 minutes with a Ca2+

-free isolation solution

containing 0.1 mM EGTA, and then for 6 minutes with cell isolation solution containing

0.05 mM Ca2+

, 0.75 mg/ml collagenase (type 1; Worthington Biochemical Corp., USA)

and 0.075 mg/ml protease (type X1V; Sigma, Germany). Ventricles were excised from

the heart, minced and gently shaken in collagenase-containing isolation solution

supplemented with 1% BSA. Cells were filtered from this solution at 4 minutes intervals

and resuspended in an isolation solution containing 0.75 mM Ca2+

.

45

Figure 10. Langendorff Heart

Solution

Chamber

Heating

Coil

Water jacketed

sugar chamber

Peristaltic

pump

46

2.2 Measurement of Ventricular Myocyte Shortening

Ventricular myocytes were allowed to settle on the glass bottom of a Perspex

chamber mounted on the stage of an inverted microscope (Axiovert 35, Zeiss, Germany).

Myocytes were superfused (3–5 ml/min) with normal Tyrode (NT) containing (mM): 130

NaCl, 5 KCl, 1 MgCl2, 10 glucose, 5 HEPES, 20 taurine, 10 creatine and 0.75 CaCl2 (pH

7.4). Shortening of myocytes was recorded using a video edge detection system (VED-

114, Crystal Biotech, USA) according to previously described techniques (Howarth,

Qureshi and White, 2002). Resting cell length (RCL) time to peak (TPK), time to half

(THALF) relaxation and amplitude of shortening (expressed as a percentage of resting

cell length) were measured in electrically stimulated (1 Hz) myocytes maintained at 35–

36 °C. Data was acquired and analyzed with signal averager software v 6.37 (Cambridge

Electronic Design, UK). Experimental solutions were prepared from stock immediately

prior to each experiment.

2.3 Measurement of Intracellular Ca2+

Concentration

Myocytes were loaded with the fluorescent indicator fura-2 AM (F-1221,

Molecular Probes, USA) as described (Howarth et al., 2002). In brief, 6.25 μl of a 1 mM

stock solution of fura-2 AM (dissolved in dimethyl sulphoxide) was added to 2.5 ml of

cells to give a final fura-2 concentration of 2.5 μM. Myocytes were gently shaken for

10 minutes at 24 °C (room temperature). After loading, myocytes were centrifuged,

washed with NT to remove extracellular fura-2 and then left for 30 min to ensure

complete hydrolysis of the intracellular ester. Intracellular Ca2+

was measured according

to previously described techniques (Howarth et al., 2011). Myocytes were alternately

illuminated by 340 and 380 nm light using a monochromator (Cairn Research, UK)

47

which changed the excitation light every 2 minutes. The resulting fluorescence emitted at

510 nm was recorded by a photomultiplier tube and the ratio of the fluorescence emitted

at the two excitation wavelengths (340/380 ratio) was calculated to provide an index of

intracellular Ca2+

concentration. Resting fura-2 ratio, time to peak (TPK), time to half

(THALF) decay of the Ca2+

transient, and the amplitude of the Ca2+

transient were

measured in electrically stimulated (1 Hz) myocytes maintained at 35-36 . Data were

acquired and analyzed with signal averager software v 6.37 Cambridge Electronic

Design, UK)

2.4 Measurement of Sarcoplasmic Reticulum Ca2+

Content

Sarcoplasmic reticulum (SR) Ca2+

release was assessed using previously described

techniques (Howarth et al., 2002). After establishing a steady state of Ca2+

transients in

electrically stimulated (1 Hz) myocytes maintained at 35–36 °C and loaded with fura-2,

stimulation was paused for a period of 5 seconds (Figure 11). Caffeine (20 mM) was then

applied for 10 seconds using a solution switching device customized for rapid solution

exchange. Electrical stimulation was resumed and the Ca2+

transients were allowed to

recover to a steady state. SR-releasable Ca2+

was assessed by measuring the area under

the curve of the caffeine-evoked Ca2+

transient. Fractional release of SR Ca2+

was

assessed by comparing the amplitude of the electrically evoked steady state

Ca2+

transients with that of the caffeine-evoked Ca2+

transient and refilling of the SR was

assessed by measuring the rate of recovery of electrically evoked Ca2+

transients

following the application of caffeine.

48

2.5 Assessment of Myofilament Sensitivity to Ca2+

In some cells shortening and fura-2 ratio were recorded simultaneously.

Myofilament sensitivity to Ca2+

was assessed from phase-plane diagrams of fura-2 ratio

versus cell length by measuring the gradient of the fura-2-cell length trajectory during

late relaxation of the twitch contraction. The position of the trajectory reflects the relative

myofilament response to Ca2+

and hence, can be used as a measure of myofilament

sensitivity to Ca2+

(Spurgeon et al., 1992; Howarth et al., 2002).

Figure 11. The recording system used for video edge detection and Ca2+ imaging

experiments: The system includes an anti-vibration table (A), a Faraday cage (B), an

inverted microscope (C), a cell superfusion system (D) with a temperature control

system (E), a video edge motion detector (F), a microfluorescence photometry

system (G) and an electrical stimulator (H).

49

2.6 Whole-Cell Recordings

Whole-cell recordings were performed in isolated ventricular myocytes. The

myocytes were placed in the recording chamber (0.4 mL) mounted on the stage of an

inverted microscope (CK2, Olympus, Tokyo, Japan) (Figure 12). After settling to the

bottom of the chamber, cells were superfused with an external solution for 10 min at a

rate of 2–3 mL.min-1

at room temperature. The external solution contained (in mM) 120

NaCl, 1 EGTA, 10 MgCl2, 10 Glucose, 10 HEPES, 1 NaH2PO4 and 5 KCl.

Transmembrane currents were recorded with an axopatch amplifier 200 B (Axon

Instruments, Inc., Foster City, CA, USA). The whole-cell bath solution contained (in

mM): 95 NaCl, 50 TEACl, 2 MgCl2, 2 CaCl2, 10 HEPES and 10 Glucose (adjusted to

pH 7.35 with NaOH). Glass microelectrodes were made using a microelectrode puller

(Model P-97, Sutter Instrument Co.,Novato, CA, USA) (Figure 13) by two-stage pulling,

and had a resistance of 2.0–4.0 MΩ when filled with electrode internal solution

containing (in mM) 140 CsCl, 10 TEACl, 2 MgCl2, 2 HEPES 1 MgATP and 10 EGTA

(adjusted to pH 7.25 with CsOH). Only rod shaped cells with clear cross-striations were

used for the experiments. Figure 14 shows typical ventricular myocytes with patch

pipette in solution. After gigaseal formation (1-10 GΩ) by suction through a pipette, the

membrane was ruptured with a gentle suction to obtain the whole-cell voltage clamp

configuration. Membrane capacitance and series resistance were compensated after

membrane rupture to minimize the duration of capacitive current. The experiment was

conducted under room Temperature (22-24 ºC). To record L-type Ca2+

currents, the

external solution was changed to the Na+-free solution in which Na

+ was replaced by

equimolar TEA (TEA–Cl). The Na+

current was also inactivated at the holding potential -

50 mV. K+currents were suppressed by substituting K

+ by Cs

+. Computer-generated

50

voltage pulses were programmed using the pCLAMP 10.0 software (Axon Instruments).

Cells with stable calcium currents were then tested using different drugs

Figure 12. Patch clamp Setup

Power

supply for

micropump

ss

Light Source

Axopatch

Amplifier

200 B

Micropumps

Solution

Reservoirs

Manipulator

51

Figure 13. Microelectrode puller (model P-97, Sutter Instrument)

52

Figure 14. Glass electrode forming giga seal with ventricular

myocyte

Figure 15. Ventricular myocyte

53

Chapter 3: Results

3.1 Effects of Cannabidiol on Ventricular Myocyte Shortening

In Normal Tyrode (NT) solution, amplitudes of myocyte shortening in response to

electrical stimulation (1 Hz) gradually decreased to 85-80 % of controls during

experiments lasting up to 20 minutes. No further running down of the shortening

amplitude was observed in the NT solution containing 0.02 % DMSO. And, unless it

stated otherwise, DMSO (in the concentration used in CBD containing solution) was also

included routinely in the control NT solutions during the shortening and Ca2+

transient

experiments.

In initial control experiments and in line with earlier studies conducted under

similar conditions it has been shown that 0.01 % DMSO did not significantly affect

contractile parameters and Ca2+

current in rat ventricular myocytes(Sun et al., 2010). The

application of control solution (NT) containing dimethyl sulfoxide at the highest

concentration used in contractility studies (0.02 % v/v; used as vehicle for 1 μM CBD)

caused 11-13 % inhibition of the shortening (n=7-10; P<0.05 compared to a 0 time

point).

In the first set of experiments, we tested the effect of CBD on the contractility of

acutely isolated rat ventricular myocytes. Figure16A shows typical records of shortening

in myocytes superfused with either NT (in the absence of CBD, NT contained 0.02 %

DMSO in all experiments) or NT + 1 µM CBD and during washout with NT. The time

frame of the effect of CBD on amplitude of shortening is shown in Figure 16B. The

effect of CBD reached a steady level within 5 minutes of CBD application. Increasing the

incubation time to 10 minutes did not cause any further change in the amplitude of

shortening. Recovery from CBD inhibition was incomplete during the 5-10 minutes

54

washout time. In summary, the amplitude of shortening measured after 5 minutes bath

application of CBD was significantly reduced (paired t-test; n=17; P<0.05) by up to 46.4

± 3.1 % in the controls (Figure 16C). The effect of CBD increased in a concentration-

dependent manner. The concentration-response relationship normalized to maximal CBD

inhibition indicated that CBD suppresses shortening of amplitudes with an IC50 of 0.6

µM (Figure17). Among other contraction parameters measured, resting cell length

(RCL), time to peak and time to half relaxation were not significantly altered

(n=17; P>0.05) by 10 minutes superfusion with CBD.

55

Figure 16. Effects of cannabidiol on ventricular myocyte shortening. (A) Typical records of

shortening in an electrically stimulated (1 Hz) ventricular myocytes superfused with either NT

(containing the vehicle, 0.02% DMSO) or NT + 1 µM CBD and during washout with NT. (B)

Time course of the mean amplitudes (AMP) of shortening expressed as a percentage of control

values in vehicle (NT + 0.02% DMSO), and in presence of CBD (1 µM). Myocytes were

maintained at 35-36 °C and superfused with CBD for 10 minutes. (C) Data shown as means ±

S.E.M., n = 6-8 cells. * indicates statistically significant difference at the level of P< 0.05.

56

Figure 17. Concentration-inhibition relationship of cannabidiol effect on the shortening of

myocytes. Concentration-response curve for the inhibitory effect of CBD on myocyte shortening.

The data was normalized to maximal inhibitory effect of CBD and plotted as a function of CBD

concentrations. Data are mean ± S.E.M. from n = 6-8 cells for each concentration.

57

3.2 Effects of Cannabidiol on Intracellular Ca2+

In the second set of experiments we investigated the effects of 5 minutes bath

application of 1 µM CBD on the resting intracellular Ca2+

levels, and on the amplitudes

and kinetics of Ca2+

transients elicited by electrical-field stimulation. Typical records of

Ca2+

transients in a myocytes superfused with NT (containing 0.02 % DMSO), NT + 1

µM CBD and during washout with NT are shown in Figure18A. The effects of 1 µM

CBD on resting fura-2 ratio, TPK Ca2+

transient, THALF decay of the Ca2+

transient and

AMP of Ca2+

transients are shown in Figure 18B-E, respectively. Although, CBD has

been shown to alter intracellular Ca2+

levels in various types of cells (see Pertwee,

2008;Pertwee, et al., 2010), application of 1 µM CBD for 5 to 10 minutes did not cause a

significant alteration in resting fura-2 ratio, TPK Ca2+

transients, or THALF decay of the

Ca2+

transient (paired t-test; n=11-14 cells, P>0.05). However, the AMP of the Ca2+

transient was significantly reduced by 1 µM CBD (0.337 ± 0.043 fura-2 ratio units)

compared to 0.488 ± 0.056 fura-2 ratio units (paired t-test, n =11 cells) in the control

experiments.

58

Figure18. Effects of cannabidiol on amplitude and time-course of intracellular Ca2+

in rat

ventricular myocytes.(A) Typical records of Ca2+

transients in an electrically stimulated (1 Hz)

ventricular myocytes superfused with either NT or NT + 1 µM CBD and during washout with

NT; scale bar indicates 0.1 fura-2 ratio unit (RU). Also shown resting fura-2 ratio (340/380 nm)

(B), time to peak (TPK) Ca2+

transient (C), time to half (THALF) decay of the Ca2+

transient (D)

and amplitude of the Ca2+

transient (E). Myocytes were maintained at 35-36 °C and superfused

with CBD for 10 minutes. Data are shown as means ± SEM, n=11-14 cell. * indicates statistically

significant difference at the level of P< 0.05.

59

3.3 Effect of Cannabidiol on Sarcoplasmic Reticulum Ca2+

Transport

The effect of CBD on sarcoplasmic reticulum Ca2+

transport was tested.

Figure19A shows a typical record illustrating the protocol used in these experiments.

Initially, the myocyte was electrically stimulated at 1Hz. Electrical stimulation was then

turned off for 5 seconds. Caffeine was then applied for 10 seconds using a rapid solution-

exchanger device. Electrical stimulation was then restarted, and the recovery of

intracellular Ca2+

was recorded during a 60 seconds period. The SR Ca2+

content was

assessed by measuring caffeine-evoked Ca2+

release (under the caffeine-evoked Ca2+

transient) and fractional release of Ca2+

by comparing the amplitude of the electrically

evoked steady-state Ca2+

transients with that of the caffeine-evoked Ca2+

transient in the

presence of either NT alone or NT with 1 µM CBD. Maximal amplitudes and the rates of

Ca2+

release by 20 mM caffeine remained unaltered after a 10 minute bath application of

1 µM CBD (Figure 19B) and (Figure 19C; paired t-test; n = 7 cells, P>0.05). Fractional

release of SR Ca2+

was not significantly altered in 1 µM CBD compared to NT (0.78 ±

0.04 in CBD versus 0.81 ± 0.07 in controls; paired t-test; n =8 cells; Figure 19D). The

recovery of the Ca2+

transients during electrical stimulation following application of

caffeine (Figure19E) was also not significantly altered in myocytes exposed to 1 µM

CBD myocytes compared to the control cells (paired t-test; n = 8 cells, P>0.05).

NT AN

0

0.2

0.4

0.6

0.8

Am

pli

tud

e o

f c

aff

ein

e-e

vo

ked

Ca

2+ t

ran

sie

nt

(Fu

ra 2

ra

tio

un

its)

60

Figure19. Effect of cannabidol on sarcoplasmic reticulum Ca2+

transport. (A) Typical records

illustrating the experimental protocol. (B) Comparison of the effect of caffeine application on the

area under caffeine-evoked Ca2+

transients in NT and CBD-treated ventricular myocytes. (C)

Area under the curve. (D)Mean amplitude of SR fractional release of Ca2+

. (E) Recovery of

electrically evoked intracellular Ca2+

after application of caffeine. Data are mean ± S.E.M., n = 7-

8 cells.

61

3.4 Effect of Cannabidiol on Myofilament Sensitivity to Ca2+

The effects of CBD on myofilament sensitivity to Ca2+

were also investigated.

These experiments tested whether CBD decreases the mechanical response by altering

the affinity of the contractile machinery of the ventricular myocytes to intracellular Ca2+

.

Atypical record of myocyte shortening, fura-2 ratio and phase-plane diagrams of

fura-2 ratio versus cell length in myocytes exposed to NT is shown in Figure 20A. The

gradient of the trajectory reflects the relative myofilament response to Ca2+

and hence,

has been used as a measure of myofilament sensitivity to Ca2+

(Spurgeon, et al., 1992;

Howarth, et al., 2002). The gradients of the fura-2-cell length trajectory during late

relaxation of the twitch contraction measured during the periods 500-600 ms (Figure

20B), 500-700 ms (Figure20C), and 500–800 ms (Figure20D) were not significantly

altered in CBD compared to NT suggesting that myofilament sensitivity to Ca2+

is not

reduced by CBD (CBD-treatment was compared to NT containing 0.02 % DMSO, paired

t-test; n = 17 cells; P>0.05).

62

Figure 20. Effect of cannabidiol on myofilament sensitivity to Ca2+

.(A) Typical record of

myocyte shortening and fura-2 ratio and phase-plane diagrams of fura-2 ratio vs. cell length in a

myocyte exposed to NT. The arrow indicates the region where the gradient was measured. B-D

shows the effects of 1 µM CBD on the mean gradient of the fura-2 cell length trajectory of the

twitch contraction during the periods 500-600 (B), 500-700 (C) and 500-800 ms (D) of late

relaxation. Data are mean ± S.E.M., n = 17 cells.

63

3.5 Effects of Cannabidiol on the Action Potentials of Ventricular Myocytes

In this set of experiments, patch-clamped ventricular myocyte were exposed to

CBD while continuously monitoring their Vrest and APs in current clamp mode. The

generation of APs was stimulated by 0.9-1 nA depolarizing current pulses of 4 ms

durations. Since the intracellular pipette solution did not contain Ca2+

-chelating agents,

the generation of each AP was accompanied by a myocyte contraction. Therefore, current

pulses were applied at a frequency of 0.2 Hz. During a typical experiment the following

protocols were employed: first, whole-cell configuration was established and a 4 to 5

minute dialysis of the myocytes with a pipette solution was allowed to ensure an

equilibrium of conditions between the pipette solution and the intracellular milieu. After

achieving stable recordings of baseline electrical activity (Vrest and AP parameters),

myocytes were exposed to CBD for 5 to 10 minutes and was subsequently washed out.

The resting membrane potentials (mean ± SEM) were -77.4 ± 1.9 and -79.2 ± 1.7

mV in the control and after CBD treatment (n=9) myocytes, respectively (Figure 21A,

inset and Figure 21B). Maximal amplitudes of AP and the rising rate of AP (maximal

velocity; V/s, Figure 21C and Figure 21D) were not significantly altered in the presence

of CBD. CBD (1 µM) consistently shortened the duration of AP (measured at 60 % of

repolarization, APD60) (Figure 21A inset and Figure 21E). Changes in AP shortening in

response to CBD (1 µM) applications were noticeable within 20 to 30 seconds

(Figure21A). Recoveries were usually partial and required a longer time.

64

Figure 21. Effect of cannabidiol on the excitability of ventricular myocytes. (A)

Representative recordings show the APs in controls (dark grey area), in the presence of 1 μM

CBD (light grey area) and after washout (striped area) in a ventricular myocytes; the insets on

panel A show the time course of action potential duration (APD60) and resting potential (Vrest)

changes in response to CBD application (indicated by horizontal bars). (B-E) show summary of

CBD effects on amplitude and shape of the AP in ventricular myocytes; quantification of the

changes in Vrest(B), AP amplitude (C), AP maximal velocity(D) and AP duration (E),

characterized by APD60 in controls (dark grey bars) and in response to 1 μM CBD (light grey

bars). Data are mean ± S.E.M., n = 6 – 9 cells from 6 to 9 for each group.

65

3.6 Effect of Cannabidiol on L-type Ca2+

Currents

We also investigated the effect of CBD (1 µM) on L-type Ca2+

currents. Figure

22A shows a typical record of L-type Ca2+

currents elicited by applying a single 300 ms

voltage pulse to +10 mV from a holding potential of -50 mV in rat ventricular myocyte

before and after a 10 minute superfusion with 1µM CBD. The time frame of the effect of

CBD on the density of L-type Ca2+

currents is shown in Figure 22B. The application of a

vehicle (0.02% DMSO) for 10 minutes caused a 10-15% inhibition of the current density

of L-type Ca2+

currents in experiments lasting up to 15 to 20 min. The effects of CBD

were also investigated on the biophysical properties of L-type Ca2+

currents in rat

ventricular myocytes. L-type Ca2+

currents were recorded in the presence of intracellular

Cs+ and extracellular TEA

+ to suppress K

+ currents while retaining 95 mM Na

+ in the

extracellular solution. The elimination of contaminating Na+ current during the recording

of L-type Ca2+

currents was achieved by applying voltage step-pulses from relatively

depolarized Vh of -50 mV, which produced steady-state inactivation of sodium channels

(Voitychuk, et al., 2012). As evident from the original recordings and I-V relationships

(Figure23A), L-type Ca2+

currents started to appear at Vm = -30 mV, reached a

maximum at around Vm = +10 mV, and decreased at higher voltages approaching to zero

at about Vm = +60 mV (Figure23B). The maximal amplitudes of L-type Ca2+

currents

were suppressed in the presence of CBD (1 µM).

66

Figure 22. Effect of cannabidiol on Ca2+

currents mediated by L-type Ca2+

channels in rat

ventricular myocytes.(A) CBD inhibits L-type Ca2+

currents recorded using whole-cell voltage-

clamp mode of patch clamp technique. Current traces are recorded before (control) and after 10

min application of 1 μM CBD. L-type Ca2+

currents were recorded during 300 ms voltage pulses

to +10 mV from a holding potential of -50 mV. (B) Averages of the maximal currents of VGCCs

presented as a function of time in the presence of vehicle (0.02% DMSO; filled circles) and 1 μM

CBD (n=5 cells; open circles). Application time for the agents was presented in horizontal bars

Figure 23. The effect of cannabidiol on voltage-current characteristics of L-type Ca2+

channels in rat ventricular myocytes. (A) Representative recordings of L-type Ca2+

currentsin

response to the depicted pulse protocol under control conditions and after application of 1μM

CBD. (B) Normalized and averaged I-V relationships of control L-type Ca2+ currents(filled

circles) and L-type Ca2+ currents in the presence of 1 μM CBD (open circles) determined by

applying a series of step depolarizing pulses from -70 mV to +70 mV in 10 mV increments for a

duration of 300ms. Data points (mean ± S.E.M.) are from 5 to 7 cells.

NT

CBD

67

CBD did not change the steady-state activation curve of the L-type Ca2+

current

(V1/2 values in the control experiment and in the presence of CBD, they are -19.1 ± 0.4

mV and -17.6 ± 0.3 mV, respectively; n=6, Paired t-test, P>0.05) (Figure 24A). However,

there was a hyperpolarizing shift of L-type Ca2+

currents in steady-state inactivation by

9.6 mV (i.e., from control value V1/2= -10.2 ± 0.9 mV to V1/2= -19.8 ± 1.1 mV in the

presence of CBD) (Figure 24B). However, there was little effect on the slopes of the

curves (k = 7.7 ± 0.5 mV and k = -4.5 ± 0.4 mV for the control activation and

inactivation, respectively, versus k = 7.9 ± 0.3 mV and k = -4.2 ± 0.3 mV for the CBD-

modified activation and inactivation, correspondingly).

In line with earlier reports (Soldatov, et al., 1998), kinetic analysis of L-type

Ca2+

currents were titled to a double-exponential function with fast (τf) and slow (τs)

inactivation time constants. Comparison of L-type Ca2+

currents in the absence and

presence of CBD revealed noticeable acceleration of the current's inactivation kinetics by

CBD. Quantification of the time constants of L-type Ca2+

currents inactivation showed

that CBD (1 μM) significantly reduced τi in the range of Vm -30 mV and +10 mV (Figure

25).

68

Figure 24. Effect of cannabidiol on steady state activation and inactivation ofL-type Ca2+

currents in rat ventricular myocytes. Steady-state activation (A) steady-state inactivation (SSI)

(B) curves of L-type Ca2+ currents in the absence (filled circles) and presence of 1 μM CBD

(open circles). Data are mean ± S.E.M, n= 5 cells are from 5 cells. Fit of experimental data points

with Boltzmann equation.

69

Figure25. Effect of cannabidiol on the kinetics ofL-type Ca2+currents in rat ventricular

myocytes. Voltage-dependent fast (triangles) and slow (circles) inactivation time constants (τi) of

L-type Ca2+ currents under control conditions (filled circles, and triangles) and in the presence of

1 μM CBD (open circles and triangles). Data are mean ± S.E.M., n= 5-6 cells.

70

Chapter 4: Discussion

Our results show that CBD has a negative inotropic effect on the contractility of rat

ventricular myocytes. The negative inotropic effect of CBD appears to be mainly due to

impaired Ca2+

signaling in cardiomyocytes. Furthermore, the results indicate that the

impaired Ca2+

signaling emanates from the inhibitory effects of CBD on the function of L

type Ca2+

channel.

Actions of CBD on the cardiovascular system influences multiple organ systems

and involves complex cellular and molecular mechanisms (see Baranowska-Kuczko,

MacLean, Kozlowska and Malinowska, 2012; Stanley et al., 2013). CBD has been shown

to have various effects on muscular structures, neuronal and endothelial cells and alter the

activities of several receptors and ion channels directly (see Hejazi et al., 2006; Pertwee

2008; 2010). Furthermore, other factors such as the activation of autonomic reflexes, the

presence of endothelial cells and fatty-acid based metabolic products have been reported

to contribute to the complexity of CBD actions on the heart (Pertwee 2008; Stanley et al.,

2013). On the other hand, acutely dissociated ventricular myocytes have several

advantages over in-vivo and traditional in-vitro conditions, since it excludes some of the

factors mentioned above and also the influences of reflex pathways, autonomic nerve

endings, neurotransmitter uptake systems, gap-junction connections and coronary

perfusion status.

In our experiments, using video edge detection techniques in individual myocytes,

it was found that CBD causes a significant reduction in the maximal amplitude of

shortening (expressed as a percentage of resting cell length) without altering the time

frame (such as TPK shortening and THALF relaxation) of myocytes shortening. Our

71

findings from myocyte shortening and intracellular Ca2+

measurement experiments are in

agreement with earlier studies reporting that CBD exerts a negative-inotropic effect in

isolated-cardiac muscle preparations (Smiley, Karler and Turkanis, 1976; Nahas and

Trouve, 1985). Furthermore, results from our study suggest that a direct interaction of

CBD with ventricular myocytes, rather than the actions of CBD on nerve endings and

neurotransmitter uptake systems are mainly responsible for the negative inotropic effects

of CBD in the heart.

The negative-inotropic actions of CBD might be attributed to the impaired release

of Ca2+

from the SR. In fact, CBD and other cannabinoids have been reported to

modulate the ryanodine sensitive intracellular Ca2+

stores in neurons (Drsydale et al.,

2008) glia (Mato et al., 2010; Pertwee et al., 2010). However, the amplitude and kinetics

of caffeine-induced Ca2+

release from intracellular Ca2+

stores were not changed by CBD

suggesting that ryanodine-sensitive intracellular Ca2+

stores are not involved in the

negative-inotropic effects of CBD observed in this study.

Cannabidiol can alter the levels of second messengers such as cAMP, cGMP and

protein kinase C which are known to be involved in tuning the Ca2+

sensitivity of the

contractile proteins. In an earlier study of cardiac muscle membranes, cAMP levels have

been shown to be unaltered in the presence of CBD (Hillard, Pounds, Boyer and Bloom,

1990), Furthermore, the sensitivity of contractile proteins to intracellular Ca2+

remained

unchanged in the presence of CBD suggesting that phosphorylation and de-

phosphorylation of the contractile proteins did not play a significant role in the negative-

inotropic action of CBD. Collectively, these results suggest that the effects of CBD on

myocyte contractility are not related to changes in intracellular Ca2+

release machinery or

72

sensitivity of myofilaments to Ca2+

. In addition, in the presence of CBD, resting levels of

intracellular Ca2+

, and cell length of ventricular myocytes, remained unaltered suggesting

that CBD does not significantly affect Ca2+

homeostasis under resting conditions.

During excitation-contraction coupling, alterations in the amplitudes and kinetics

of the cardiac action potential (AP) are closely associated with corresponding changes in

the contractility of myocytes. In our study, cardiac APs were recorded using the whole

cell patch clamp technique. It was found that the amplitudes and dV/dtmax of APs were

not altered in the presence of CBD suggesting that voltage-gated sodium channels are not

affected by CBD. For this reason, the effect of CBD on the functional properties of

voltage-gated sodium channels was not investigated any further. However, the decrease

of AP duration, and suppression of the second phase of the AP was evident in whole cell

recordings from cardiomyocytes. Since the plateau (phase 2) of AP is mainly due the

influx of Ca2+

through L-type voltage-gated Ca2+

channels, these results suggest that

CBD can act on L-type Ca2+

channels. Indeed, whole cell recordings of Ca2+

currents in

cardiomyocytes, in agreement with the recordings of APs, indicated that CBD inhibits

potently the function of L-type voltage-gated Ca2+

channels with an IC50 value of 0.6 µM.

An analysis of current kinetics indicated that the inactivation of Ca2+

currents can double

the exponential decay and CBD can decrease the amplitudes of the Ca2+

currents by

accelerating both fast and slow components of the inactivation process. The effects of

CBD on the gating characteristics of Ca2+

channels were further investigated by studying

the activation and inactivation curves in the absence and presence CBD. Although half

activation voltage and the slope of the activation curve were not altered, there was a

73

hyperpolarizing-shift of the inactivation curve further suggesting that CBD changes the

voltage-dependent inactivation process.

Collectively, these results suggest that during excitation-contraction coupling,

shortening of action potential due to the inhibition of L-type Ca2+

channels decreases

Ca2+

-induced Ca2+

release from SR and causes the negative-inotropic effect of CBD

observed in our experiments and reported in earlier studies. In line with these findings,

although caffeine induced contractures and myofilament sensitivity to Ca2+

remained

unchanged, electrically-induced Ca2+

transients were significantly depressed by CBD;

further suggesting that Ca2+

-induced Ca2+

release was impaired in the presence of CBD.

In earlier studies CBD has been reported to inhibit the uptake of dopamine and other

neurotransmitters (Poddar and Dewey, 1980; Pandolfo, et al., 2011). However, an uptake

mechanism is not likely to be involved in our studies on isolated ventricular myocytes.

Furthermore, inhibition of uptake mechanisms for dopamine and/or catecholamines such

as epinephrine and norepinephrine would lead to well-established positive inotropic

actions, rather than a negative inotropic effect as observed in the present study.

Although CBD does not activate known cannabinoid receptors, some of its

pharmacological actions have been shown to be reversed by cannabinoid receptor

antagonists (Pertwee, 2008). Recent studies have shown that the antagonist sensitivity of

CBD effects is due to the inhibition of fatty acid amide hydrolase activity, an enzyme that

inactivates endogenous cannabinoids and related to subsequent increases in tissue

endocannabinoids levels (Pertwee, 2010). However, both amplitudes of myocyte

shortening and the L-type Ca2+

currents were not altered by URB597, a potent inhibitor

74

of fatty acid amide hydrolase (Al Kury et al., 2014), suggesting that the activity of

increased fatty acid amide hydrolase was not involved in the actions observed in CBD.

The interaction between L-type Ca2+

channels and CBD has not been studied in

detail. In an earlier study, 45

Ca2+

-uptake in brain synaptosomes had been shown to be

inhibited by CBD (Harris and Stokes, 1982). Similarly, CBD induced increases in

intracellular Ca2+

levels were found to be sensitive to blockers of L-type Ca2+

channels

(Drysdale et al., 2006). In agreement with these results, contractions of smooth muscles

elicited by high-K+-containing solutions were also significantly inhibited by CBD

(Cluny, Naylor, Whittle and Javid, 2011). The interaction between T-type Ca2+

channels

and CBD was investigated in a recent study, on human CaV3 channels stably expressed

in human embryonic kidney 293 cells and T-type channels in mouse sensory neurons

(Ross et al., 2008). At moderately hyperpolarized potentials, CBD inhibited peak CaV3.1

and CaV3.2 currents with IC50 values of 1 µM but was less potent on CaV3.3 channels.

CBD inhibited sensory neuron T-type channels by about 45% at 1 µM. However, in

recordings made from a holding potential of -70 mV, 100 nM CBD inhibited more than

50% of the peak CaV3.1 current. CBD produced a significant hyperpolarizing shift in the

steady state inactivation potentials for each of the CaV3 channels (Ross et al., 2008). In

this respect our results indicate for the first time that CBD inhibits the function of L-type

Ca2+

channels in ventricular myocytes.

In clinical studies, it has been shown that acute CBD intake does not cause a

significant change in blood pressure and heart rate (Stanley, 2013a). However, several

earlier studies indicate that increases in blood pressure and heart rate during stressful

conditions are markedly attenuated by CBD (Gomes, et al., 2013; Granjeiro, Gomes,

75

Guimaraes, Correa and Resstel, 2011; Resstel, Joca, Moreira, Correa and Guimaraes,

2006; Resstel, et al., 2009). In fact, in in vivo studies, it has been demonstrated that

increased heart rate and blood pressure, one of the most consistent effects of cannabis

intoxication, are significantly decreased by CBD (Nahas and Trouve, 1985). These

studies suggest that CBD can exert negative inotropic and chronotropic actions under

certain clinical conditions such as stress and cannabis intoxication.

CBD, in the concentration range used in this study (0.1-10 µM), has been shown

to modulate functions of various receptors and ion channels such as Ca2+

channels (Ross

et al., 2008), 5-HT3 (Yang et al., 2010), and nicotinic receptors (Mahgoub et al., 2013).

The peak plasma concentrations of CBD in healthy volunteers following administration

of SativexTM

(1:1 ratio of ∆9-THC and CBD) was reported to be between 0.01µM to

0.05µM (Massi, et al., 2013). In earlier clinical studies, mean CBD levels of 0.036 μM

were determined in blood analysis following a 6-week oral treatment with CBD at doses

of 10 mg/kg/day (Consroe et al., 1991). However, CBD is a highly lipophilic compound

and easily passes biological membranes and accumulates in tissues and it can reach to

several fold higher tissue concentrations than in blood. Collectively, these findings

suggest that CBD, at relatively high concentrations, can suppress the function of the

cardiovascular system. In fact, cardiac failure due to depressed heart contractility has

been suggested to be the main cause of mortality through cannabis intoxication

(Bergamaschi et al., 2011). Thus it is likely that some of the effects of the cannabis plant

are mediated by CBD during cannabis intoxication.

76

In addition to cardiac contractility, the electrical activity of heart has also been

suggested to be affected by CBD. In an earlier study, CBD has been shown to have

beneficial actions in ischemia-induced cardiac arrhythmias (Walsh et al., 2010). The

shortening of the AP duration by CBD observed in our study can be beneficial or

harmful, depending on the underlying pathology. Thus, during acute ischemia, in which

the duration of the cardiac APD is already shortened, a further decrease should be

proarrhythmic (Den Ruijter et al., 2007). However, shortening of AP duration should be

beneficial in preventing those arrhythmias caused by triggered activities observed in

conditions such as heart failure (Den Ruijter et al., 2007). In conclusion, the results

indicate for the first time, that CBD inhibits myocyte contractility by acting on L-type

Ca2+

channels.

77

Chapter 5: Limitation and Future Directions

The scope of this research is mainly concerned with the effects of CBD on healthy

ventricular cells, although endocannabinoids are stress induced neurotransmitters. In

addition to the effects of CBD on L- type calcium channels its effect on NCX should also

be established due to their influence on cytosolic calcium levels and the viability of

ventricular cells in the presence of CBD.

Future research may also include expressing calcium channels onto HEK-293 cells,

this allows us to exclude or include the indirect effects of CBD on intracellular signaling

that can affect calcium channels. Another method is to express the calcium channel each

subunit at a time and find where CBD exerts it effects most effectively and thereby

establish the mechanism by which CBD exerts there effects.

78

Bibliography

Ahrens, J., Demir, R., Leuwer, M., de la Roche, J., Krampfl, K., Foadi, N., . . . Haeseler, G.

(2009). The nonpsychotropic cannabinoid cannabidiol modulates and directly activates

alpha-1 and alpha-1-Beta glycine receptor function. Pharmacology, 83(4), 217-222. doi:

10.1159/000201556

Amin, A. S., Tan, H. L., & Wilde, A. A. (2010). Cardiac ion channels in health and disease.

Heart Rhythm, 7(1), 117-126. doi: 10.1016/j.hrthm.2009.08.005

Ashton, J. C. (2012). Synthetic cannabinoids as drugs of abuse. Curr Drug Abuse Rev, 5(2), 158-

168.

Baranowska-Kuczko, M., MacLean, M. R., Kozlowska, H., & Malinowska, B. (2012).

Endothelium-dependent mechanisms of the vasodilatory effect of the endocannabinoid,

anandamide, in the rat pulmonary artery. Pharmacol Res, 66(3), 251-259. doi:

10.1016/j.phrs.2012.05.004

Bergamaschi, M. M., Queiroz, R. H., Zuardi, A. W., & Crippa, J. A. (2011). Safety and side

effects of cannabidiol, a Cannabis sativa constituent. Curr Drug Saf, 6(4), 237-249.

Bers, D. M. (2000). Calcium fluxes involved in control of cardiac myocyte contraction. Circ Res,

87(4), 275-281.

Bers, D. M. (2008). Calcium cycling and signaling in cardiac myocytes. Annu Rev Physiol, 70,

23-49. doi: 10.1146/annurev.physiol.70.113006.100455

Bishop-Bailey, D. (2000). Peroxisome proliferator-activated receptors in the cardiovascular

system. Br J Pharmacol, 129(5), 823-834. doi: 10.1038/sj.bjp.0703149

Bisogno, T., Hanus, L., De Petrocellis, L., Tchilibon, S., Ponde, D. E., Brandi, I., . . . Di Marzo,

V. (2001). Molecular targets for cannabidiol and its synthetic analogues: effect on

vanilloid VR1 receptors and on the cellular uptake and enzymatic hydrolysis of

anandamide. Br J Pharmacol, 134(4), 845-852. doi: 10.1038/sj.bjp.0704327

Blake, D. R., Robson, P., Ho, M., Jubb, R. W., & McCabe, C. S. (2006). Preliminary assessment

of the efficacy, tolerability and safety of a cannabis-based medicine (Sativex) in the

treatment of pain caused by rheumatoid arthritis. Rheumatology (Oxford), 45(1), 50-52.

doi: 10.1093/rheumatology/kei183

Booz, G. W. (2011). Cannabidiol as an emergent therapeutic strategy for lessening the impact of

inflammation on oxidative stress. Free Radic Biol Med, 51(5), 1054-1061. doi:

10.1016/j.freeradbiomed.2011.01.007

79

Braida, D., Pegorini, S., Arcidiacono, M. V., Consalez, G. G., Croci, L., & Sala, M. (2003). Post-

ischemic treatment with cannabidiol prevents electroencephalographic flattening,

hyperlocomotion and neuronal injury in gerbils. Neurosci Lett, 346(1-2), 61-64.

Burstein, S. H., & Zurier, R. B. (2009). Cannabinoids, endocannabinoids, and related analogs in

inflammation. Aaps j, 11(1), 109-119. doi: 10.1208/s12248-009-9084-5

Campos, A. C., Moreira, F. A., Gomes, F. V., Del Bel, E. A., & Guimaraes, F. S. (2012).

Multiple mechanisms involved in the large-spectrum therapeutic potential of cannabidiol

in psychiatric disorders. Philos Trans R Soc Lond B Biol Sci, 367(1607), 3364-3378. doi:

10.1098/rstb.2011.0389

Capasso, R., Borrelli, F., Aviello, G., Romano, B., Scalisi, C., Capasso, F., & Izzo, A. A. (2008).

Cannabidiol, extracted from Cannabis sativa, selectively inhibits inflammatory

hypermotility in mice. Br J Pharmacol, 154(5), 1001-1008. doi: 10.1038/bjp.2008.177

Carrier, E. J., Auchampach, J. A., & Hillard, C. J. (2006). Inhibition of an equilibrative

nucleoside transporter by cannabidiol: a mechanism of cannabinoid immunosuppression.

Proc Natl Acad Sci U S A, 103(20), 7895-7900. doi: 10.1073/pnas.0511232103

Catterall, W. A. (2010). Signaling complexes of voltage-gated sodium and calcium channels.

Neurosci Lett, 486(2), 107-116. doi: 10.1016/j.neulet.2010.08.085

Catterall, W. A. (2011). Voltage-gated calcium channels. Cold Spring Harb Perspect Biol, 3(8),

a003947. doi: 10.1101/cshperspect.a003947

Catterall, W. A., Perez-Reyes, E., Snutch, T. P., & Striessnig, J. (2005). International Union of

Pharmacology. XLVIII. Nomenclature and structure-function relationships of voltage-

gated calcium channels. Pharmacol Rev, 57(4), 411-425. doi: 10.1124/pr.57.4.5

Cluny, N. L., Naylor, R. J., Whittle, B. A., & Javid, F. A. (2011). The effects of cannabidiolic

acid and cannabidiol on contractility of the gastrointestinal tract of Suncus murinus. Arch

Pharm Res, 34(9), 1509-1517. doi: 10.1007/s12272-011-0913-6

Collins, F. G., & Haavik, C. O. (1979). Effects of cannabinoids on cardiac microsomal CaATPase

activity and calcium uptake. Biochem Pharmacol, 28(15), 2303-2306.

Costa, B., Colleoni, M., Conti, S., Parolaro, D., Franke, C., Trovato, A. E., & Giagnoni, G.

(2004). Oral anti-inflammatory activity of cannabidiol, a non-psychoactive constituent of

cannabis, in acute carrageenan-induced inflammation in the rat paw. Naunyn

Schmiedebergs Arch Pharmacol, 369(3), 294-299. doi: 10.1007/s00210-004-0871-3

Costa, B., Trovato, A. E., Comelli, F., Giagnoni, G., & Colleoni, M. (2007). The non-

psychoactive cannabis constituent cannabidiol is an orally effective therapeutic agent in

rat chronic inflammatory and neuropathic pain. Eur J Pharmacol, 556(1-3), 75-83. doi:

10.1016/j.ejphar.2006.11.006

80

Cravatt, B. F., & Lichtman, A. H. (2002). The enzymatic inactivation of the fatty acid amide class

of signaling lipids. Chem Phys Lipids, 121(1-2), 135-148.

De Jongh, K. S., Murphy, B. J., Colvin, A. A., Hell, J. W., Takahashi, M., & Catterall, W. A.

(1996). Specific phosphorylation of a site in the full-length form of the alpha 1 subunit of

the cardiac L-type calcium channel by adenosine 3',5'-cyclic monophosphate-dependent

protein kinase. Biochemistry, 35(32), 10392-10402. doi: 10.1021/bi953023c

De Petrocellis, L., Ligresti, A., Moriello, A. S., Allara, M., Bisogno, T., Petrosino, S., . . . Di

Marzo, V. (2011). Effects of cannabinoids and cannabinoid-enriched Cannabis extracts

on TRP channels and endocannabinoid metabolic enzymes. Br J Pharmacol, 163(7),

1479-1494. doi: 10.1111/j.1476-5381.2010.01166.x

De Petrocellis, L., Vellani, V., Schiano-Moriello, A., Marini, P., Magherini, P. C., Orlando, P., &

Di Marzo, V. (2008). Plant-derived cannabinoids modulate the activity of transient

receptor potential channels of ankyrin type-1 and melastatin type-8. J Pharmacol Exp

Ther, 325(3), 1007-1015. doi: 10.1124/jpet.107.134809

Den Ruijter, H. M., Berecki, G., Opthof, T., Verkerk, A. O., Zock, P. L., & Coronel, R. (2007).

Pro- and antiarrhythmic properties of a diet rich in fish oil. Cardiovasc Res, 73(2), 316-

325. doi: 10.1016/j.cardiores.2006.06.014

Devane, W. A., Dysarz, F. A., 3rd, Johnson, M. R., Melvin, L. S., & Howlett, A. C. (1988).

Determination and characterization of a cannabinoid receptor in rat brain. Mol

Pharmacol, 34(5), 605-613.

Di Marzo, V. (2008). Endocannabinoids: synthesis and degradation. Rev Physiol Biochem

Pharmacol, 160, 1-24. doi: 10.1007/112_0505

Di Marzo, V., Fontana, A., Cadas, H., Schinelli, S., Cimino, G., Schwartz, J. C., & Piomelli, D.

(1994). Formation and inactivation of endogenous cannabinoid anandamide in central

neurons. Nature, 372(6507), 686-691. doi: 10.1038/372686a0

Dolphin, A. C. (2003). G protein modulation of voltage-gated calcium channels. Pharmacol Rev,

55(4), 607-627. doi: 10.1124/pr.55.4.3

Dolphin, A. C. (2006). A short history of voltage-gated calcium channels. Br J Pharmacol, 147

Suppl 1, S56-62. doi: 10.1038/sj.bjp.0706442

Dolphin, A. C. (2009). Calcium channel diversity: multiple roles of calcium channel subunits.

Curr Opin Neurobiol, 19(3), 237-244. doi: 10.1016/j.conb.2009.06.006

Drysdale, A. J., Ryan, D., Pertwee, R. G., & Platt, B. (2006). Cannabidiol-induced intracellular

Ca2+ elevations in hippocampal cells. Neuropharmacology, 50(5), 621-631. doi:

10.1016/j.neuropharm.2005.11.008

81

El-Remessy, A. B., Al-Shabrawey, M., Khalifa, Y., Tsai, N. T., Caldwell, R. B., & Liou, G. I.

(2006). Neuroprotective and blood-retinal barrier-preserving effects of cannabidiol in

experimental diabetes. Am J Pathol, 168(1), 235-244. doi: 10.2353/ajpath.2006.050500

El-Remessy, A. B., Khalil, I. E., Matragoon, S., Abou-Mohamed, G., Tsai, N. J., Roon, P., . . .

Liou, G. I. (2003). Neuroprotective effect of (-)Delta9-tetrahydrocannabinol and

cannabidiol in N-methyl-D-aspartate-induced retinal neurotoxicity: involvement of

peroxynitrite. Am J Pathol, 163(5), 1997-2008.

Fearnley, C. J., Roderick, H. L., & Bootman, M. D. (2011). Calcium signaling in cardiac

myocytes. Cold Spring Harb Perspect Biol, 3(11), a004242. doi:

10.1101/cshperspect.a004242

Fernandez-Ruiz, J. (2012). [Cannabinoid drugs for neurological diseases: what is behind?]. Rev

Neurol, 54(10), 613-628.

Ferrantini, C., Crocini, C., Coppini, R., Vanzi, F., Tesi, C., Cerbai, E., . . . Sacconi, L. (2013). The

transverse-axial tubular system of cardiomyocytes. Cell Mol Life Sci, 70(24), 4695-4710.

doi: 10.1007/s00018-013-1410-5

Foadi, N., Leuwer, M., Demir, R., Dengler, R., Buchholz, V., de la Roche, J., . . . Ahrens, J.

(2010). Lack of positive allosteric modulation of mutated alpha(1)S267I glycine

receptors by cannabinoids. Naunyn Schmiedebergs Arch Pharmacol, 381(5), 477-482.

doi: 10.1007/s00210-010-0506-9

Fowler, C. J. (2004). Oleamide: a member of the endocannabinoid family? Br J Pharmacol,

141(2), 195-196. doi: 10.1038/sj.bjp.0705608

Garcia-Arencibia, M., Gonzalez, S., de Lago, E., Ramos, J. A., Mechoulam, R., & Fernandez-

Ruiz, J. (2007). Evaluation of the neuroprotective effect of cannabinoids in a rat model of

Parkinson's disease: importance of antioxidant and cannabinoid receptor-independent

properties. Brain Res, 1134(1), 162-170. doi: 10.1016/j.brainres.2006.11.063

Gilbert, J. C., Pertwee, R. G., & Wyllie, M. G. (1977). Effects of delta9-tetrahydrocannabinol and

cannabidiol on a Mg2+-ATPase of synaptic vesicles prepared from rat cerebral cortex. Br

J Pharmacol, 59(4), 599-601.

Glaser, S. T., Kaczocha, M., & Deutsch, D. G. (2005). Anandamide transport: a critical review.

Life Sci, 77(14), 1584-1604. doi: 10.1016/j.lfs.2005.05.007

Gloss, D., Nolan, S. J., & Staba, R. (2014). The role of high-frequency oscillations in epilepsy

surgery planning. Cochrane Database Syst Rev, 1, Cd010235. doi:

10.1002/14651858.CD010235.pub2

Gomes, F. V., Alves, F. H., Guimaraes, F. S., Correa, F. M., Resstel, L. B., & Crestani, C. C.

(2013). Cannabidiol administration into the bed nucleus of the stria terminalis alters

82

cardiovascular responses induced by acute restraint stress through 5-HT(1)A receptor.

Eur Neuropsychopharmacol, 23(9), 1096-1104. doi: 10.1016/j.euroneuro.2012.09.007

Gomes, F. V., Resstel, L. B., & Guimaraes, F. S. (2011). The anxiolytic-like effects of

cannabidiol injected into the bed nucleus of the stria terminalis are mediated by 5-HT1A

receptors. Psychopharmacology (Berl), 213(2-3), 465-473. doi: 10.1007/s00213-010-

2036-z

Granjeiro, E. M., Gomes, F. V., Guimaraes, F. S., Correa, F. M., & Resstel, L. B. (2011). Effects

of intracisternal administration of cannabidiol on the cardiovascular and behavioral

responses to acute restraint stress. Pharmacol Biochem Behav, 99(4), 743-748. doi:

10.1016/j.pbb.2011.06.027

Grlic, Ljubiša. "United Nations Office on Drugs and Crime." UNODC. N.p., 1 Jan. 1962. Web.

01 May 2014.

Haase, H., Bartel, S., Karczewski, P., Morano, I., & Krause, E. G. (1996). In-vivo

phosphorylation of the cardiac L-type calcium channel beta-subunit in response to

catecholamines. Mol Cell Biochem, 163-164, 99-106.

Hamdani, N., Kooij, V., van Dijk, S., Merkus, D., Paulus, W. J., Remedios, C. D., . . . van der

Velden, J. (2008). Sarcomeric dysfunction in heart failure. Cardiovasc Res, 77(4), 649-

658. doi: 10.1093/cvr/cvm079

Hampson, A. J., Grimaldi, M., Axelrod, J., & Wink, D. (1998). Cannabidiol and (-)Delta9-

tetrahydrocannabinol are neuroprotective antioxidants. Proc Natl Acad Sci U S A, 95(14),

8268-8273.

Hampson, A. J., Grimaldi, M., Lolic, M., Wink, D., Rosenthal, R., & Axelrod, J. (2000).

Neuroprotective antioxidants from marijuana. Ann N Y Acad Sci, 899, 274-282.

Hanus, L., Abu-Lafi, S., Fride, E., Breuer, A., Vogel, Z., Shalev, D. E., . . . Mechoulam, R.

(2001). 2-arachidonyl glyceryl ether, an endogenous agonist of the cannabinoid CB1

receptor. Proc Natl Acad Sci U S A, 98(7), 3662-3665. doi: 10.1073/pnas.061029898

Hayakawa, K., Mishima, K., Abe, K., Hasebe, N., Takamatsu, F., Yasuda, H., . . . Fujiwara, M.

(2004). Cannabidiol prevents infarction via the non-CB1 cannabinoid receptor

mechanism. Neuroreport, 15(15), 2381-2385.

Heinemann, S. H., Terlau, H., Stuhmer, W., Imoto, K., & Numa, S. (1992). Calcium channel

characteristics conferred on the sodium channel by single mutations. Nature, 356(6368),

441-443. doi: 10.1038/356441a0

Hejazi, N., Zhou, C., Oz, M., Sun, H., Ye, J. H., & Zhang, L. (2006). Delta9-

tetrahydrocannabinol and endogenous cannabinoid anandamide directly potentiate the

83

function of glycine receptors. Mol Pharmacol, 69(3), 991-997. doi:

10.1124/mol.105.019174

Hell, J. W., Yokoyama, C. T., Wong, S. T., Warner, C., Snutch, T. P., & Catterall, W. A. (1993).

Differential phosphorylation of two size forms of the neuronal class C L-type calcium

channel alpha 1 subunit. J Biol Chem, 268(26), 19451-19457.

Herkenham, M., Lynn, A. B., Little, M. D., Johnson, M. R., Melvin, L. S., de Costa, B. R., &

Rice, K. C. (1990). Cannabinoid receptor localization in brain. Proc Natl Acad Sci U S A,

87(5), 1932-1936.

Hill, A. J., Williams, C. M., Whalley, B. J., & Stephens, G. J. (2012). Phytocannabinoids as novel

therapeutic agents in CNS disorders. Pharmacol Ther, 133(1), 79-97. doi:

10.1016/j.pharmthera.2011.09.002

Hillard, C. J., & Auchampach, J. A. (1994). In vitro activation of brain protein kinase C by the

cannabinoids. Biochim Biophys Acta, 1220(2), 163-170.

Hillard, C. J., Pounds, J. J., Boyer, D. R., & Bloom, A. S. (1990). Studies of the role of membrane

lipid order in the effects of delta 9-tetrahydrocannabinol on adenylate cyclase activation

in heart. J Pharmacol Exp Ther, 252(3), 1075-1082.

Howarth, F. C., Qureshi, M. A., Hassan, Z., Al Kury, L. T., Isaev, D., Parekh, K., . . . Adeghate,

E. (2011). Changing pattern of gene expression is associated with ventricular myocyte

dysfunction and altered mechanisms of Ca2+ signalling in young type 2 Zucker diabetic

fatty rat heart. Exp Physiol, 96(3), 325-337. doi: 10.1113/expphysiol.2010.055574

Howarth, F. C., Qureshi, M. A., & White, E. (2002). Effects of hyperosmotic shrinking on

ventricular myocyte shortening and intracellular Ca(2+) in streptozotocin-induced

diabetic rats. Pflugers Arch, 444(3), 446-451. doi: 10.1007/s00424-002-0830-0

Howlett, A. C., Scott, D. K., & Wilken, G. H. (1989). Regulation of adenylate cyclase by

cannabinoid drugs. Insights based on thermodynamic studies. Biochem Pharmacol,

38(19), 3297-3304.

Hsueh, W. A., & Bruemmer, D. (2004). Peroxisome proliferator-activated receptor gamma:

implications for cardiovascular disease. Hypertension, 43(2), 297-305. doi:

10.1161/01.HYP.0000113626.76571.5b

Huffman, J. W., Liddle, J., Duncan, S. G., Jr., Yu, S., Martin, B. R., & Wiley, J. L. (1998).

Synthesis and pharmacology of the isomeric methylheptyl-delta8-tetrahydrocannabinols.

Bioorg Med Chem, 6(12), 2383-2396.

Iuvone, T., Esposito, G., Esposito, R., Santamaria, R., Di Rosa, M., & Izzo, A. A. (2004).

Neuroprotective effect of cannabidiol, a non-psychoactive component from Cannabis

84

sativa, on beta-amyloid-induced toxicity in PC12 cells. J Neurochem, 89(1), 134-141.

doi: 10.1111/j.1471-4159.2003.02327.x

Izzo, A. A., & Camilleri, M. (2009). Cannabinoids in intestinal inflammation and cancer.

Pharmacol Res, 60(2), 117-125. doi: 10.1016/j.phrs.2009.03.008

Karniol, I. G., Shirakawa, I., Kasinski, N., Pfeferman, A., & Carlini, E. A. (1974). Cannabidiol

interferes with the effects of delta 9 - tetrahydrocannabinol in man. Eur J Pharmacol,

28(1), 172-177.

Kathmann, M., Flau, K., Redmer, A., Trankle, C., & Schlicker, E. (2006). Cannabidiol is an

allosteric modulator at mu- and delta-opioid receptors. Naunyn Schmiedebergs Arch

Pharmacol, 372(5), 354-361. doi: 10.1007/s00210-006-0033-x

Lanner, J. T., Georgiou, D. K., Joshi, A. D., & Hamilton, S. L. (2010). Ryanodine receptors:

structure, expression, molecular details, and function in calcium release. Cold Spring

Harb Perspect Biol, 2(11), a003996. doi: 10.1101/cshperspect.a003996

Lastres-Becker, I., Molina-Holgado, F., Ramos, J. A., Mechoulam, R., & Fernandez-Ruiz, J.

(2005). Cannabinoids provide neuroprotection against 6-hydroxydopamine toxicity in

vivo and in vitro: relevance to Parkinson's disease. Neurobiol Dis, 19(1-2), 96-107. doi:

10.1016/j.nbd.2004.11.009

Ma, Y., Zhang, L., Zhao, Y. F., Zang, W. J., & Chen, C. (2010). Effects of GH secretagogues on

contractility and Ca2+ homeostasis of isolated adult rat ventricular myocytes.

Endocrinology, 151(9), 4446-4454. doi: 10.1210/en.2009-1432

Mahgoub, M., Keun-Hang, S. Y., Sydorenko, V., Ashoor, A., Kabbani, N., Al Kury, L., . . . Oz,

M. (2013). Effects of cannabidiol on the function of alpha7-nicotinic acetylcholine

receptors. Eur J Pharmacol, 720(1-3), 310-319. doi: 10.1016/j.ejphar.2013.10.011

Marsicano, G., Moosmann, B., Hermann, H., Lutz, B., & Behl, C. (2002). Neuroprotective

properties of cannabinoids against oxidative stress: role of the cannabinoid receptor CB1.

J Neurochem, 80(3), 448-456.

Martin-Moreno, A. M., Reigada, D., Ramirez, B. G., Mechoulam, R., Innamorato, N., Cuadrado,

A., & de Ceballos, M. L. (2011). Cannabidiol and other cannabinoids reduce microglial

activation in vitro and in vivo: relevance to Alzheimer's disease. Mol Pharmacol, 79(6),

964-973. doi: 10.1124/mol.111.071290

Mato, S., Victoria Sanchez-Gomez, M., & Matute, C. (2010). Cannabidiol induces intracellular

calcium elevation and cytotoxicity in oligodendrocytes. Glia, 58(14), 1739-1747. doi:

10.1002/glia.21044

85

Mechoulam, R., Ben-Shabat, S., Hanus, L., Ligumsky, M., Kaminski, N. E., Schatz, A. R., . . . et

al. (1995). Identification of an endogenous 2-monoglyceride, present in canine gut, that

binds to cannabinoid receptors. Biochem Pharmacol, 50(1), 83-90.

Mechoulam, R., & Gaoni, Y. (1965). A TOTAL SYNTHESIS OF DL-DELTA-1-

TETRAHYDROCANNABINOL, THE ACTIVE CONSTITUENT OF HASHISH. J Am

Chem Soc, 87, 3273-3275.

Mechoulam, R., & Hanus, L. (2002). Cannabidiol: an overview of some chemical and

pharmacological aspects. Part I: chemical aspects. Chem Phys Lipids, 121(1-2), 35-43.

Mechoulam, R., Parker, L. A., & Gallily, R. (2002). Cannabidiol: an overview of some

pharmacological aspects. J Clin Pharmacol, 42(11 Suppl), 11s-19s.

Mechoulam, R., Peters, M., Murillo-Rodriguez, E., & Hanus, L. O. (2007). Cannabidiol--recent

advances. Chem Biodivers, 4(8), 1678-1692. doi: 10.1002/cbdv.200790147

Mishima, K., Hayakawa, K., Abe, K., Ikeda, T., Egashira, N., Iwasaki, K., & Fujiwara, M.

(2005). Cannabidiol prevents cerebral infarction via a serotonergic 5-

hydroxytryptamine1A receptor-dependent mechanism. Stroke, 36(5), 1077-1082. doi:

10.1161/01.str.0000163083.59201.34

Nahas, G., & Trouve, R. (1985). Effects and interactions of natural cannabinoids on the isolated

heart. Proc Soc Exp Biol Med, 180(2), 312-316.

Nattel, S., & Carlsson, L. (2006). Innovative approaches to anti-arrhythmic drug therapy. Nat Rev

Drug Discov, 5(12), 1034-1049. doi: 10.1038/nrd2112

Navedo, M. F., & Santana, L. F. (2013). CaV1.2 sparklets in heart and vascular smooth muscle. J

Mol Cell Cardiol, 58, 67-76. doi: 10.1016/j.yjmcc.2012.11.018

O'Sullivan, S. E., Kendall, D. A., & Randall, M. D. (2005). Vascular effects of delta 9-

tetrahydrocannabinol (THC), anandamide and N-arachidonoyldopamine (NADA) in the

rat isolated aorta. Eur J Pharmacol, 507(1-3), 211-221. doi: 10.1016/j.ejphar.2004.11.056

O'Sullivan, S. E., Sun, Y., Bennett, A. J., Randall, M. D., & Kendall, D. A. (2009). Time-

dependent vascular actions of cannabidiol in the rat aorta. Eur J Pharmacol, 612(1-3),

61-68. doi: 10.1016/j.ejphar.2009.03.010

Pandolfo, P., Silveirinha, V., dos Santos-Rodrigues, A., Venance, L., Ledent, C., Takahashi, R.

N., . . . Kofalvi, A. (2011). Cannabinoids inhibit the synaptic uptake of adenosine and

dopamine in the rat and mouse striatum. Eur J Pharmacol, 655(1-3), 38-45. doi:

10.1016/j.ejphar.2011.01.013

86

Parker, L. A., Rock, E. M., & Limebeer, C. L. (2011). Regulation of nausea and vomiting by

cannabinoids. Br J Pharmacol, 163(7), 1411-1422. doi: 10.1111/j.1476-

5381.2010.01176.x

Pertwee, R. G. (1997). Pharmacology of cannabinoid CB1 and CB2 receptors. Pharmacol Ther,

74(2), 129-180.

Pertwee, R. G. (2008). The diverse CB1 and CB2 receptor pharmacology of three plant

cannabinoids: delta9-tetrahydrocannabinol, cannabidiol and delta9-

tetrahydrocannabivarin. Br J Pharmacol, 153(2), 199-215. doi: 10.1038/sj.bjp.0707442

Pertwee, R. G., Howlett, A. C., Abood, M. E., Alexander, S. P., Di Marzo, V., Elphick, M. R., . . .

Ross, R. A. (2010). International Union of Basic and Clinical Pharmacology. LXXIX.

Cannabinoid receptors and their ligands: beyond CB(1) and CB(2). Pharmacol Rev,

62(4), 588-631. doi: 10.1124/pr.110.003004

Pertwee, R. G., Ross, R. A., Craib, S. J., & Thomas, A. (2002). (-)-Cannabidiol antagonizes

cannabinoid receptor agonists and noradrenaline in the mouse vas deferens. Eur J

Pharmacol, 456(1-3), 99-106.

Poddar, M. K., & Dewey, W. L. (1980). Effects of cannabinoids on catecholamine uptake and

release in hypothalamic and striatal synaptosomes. J Pharmacol Exp Ther, 214(1), 63-67.

Porter, A. C., Sauer, J. M., Knierman, M. D., Becker, G. W., Berna, M. J., Bao, J., . . . Felder, C.

C. (2002). Characterization of a novel endocannabinoid, virodhamine, with antagonist

activity at the CB1 receptor. J Pharmacol Exp Ther, 301(3), 1020-1024.

Pragnell, M., De Waard, M., Mori, Y., Tanabe, T., Snutch, T. P., & Campbell, K. P. (1994).

Calcium channel beta-subunit binds to a conserved motif in the I-II cytoplasmic linker of

the alpha 1-subunit. Nature, 368(6466), 67-70. doi: 10.1038/368067a0

Puri, T. S., Gerhardstein, B. L., Zhao, X. L., Ladner, M. B., & Hosey, M. M. (1997). Differential

effects of subunit interactions on protein kinase A- and C-mediated phosphorylation of L-

type calcium channels. Biochemistry, 36(31), 9605-9615. doi: 10.1021/bi970500d

Qin, N., Neeper, M. P., Liu, Y., Hutchinson, T. L., Lubin, M. L., & Flores, C. M. (2008). TRPV2

is activated by cannabidiol and mediates CGRP release in cultured rat dorsal root

ganglion neurons. J Neurosci, 28(24), 6231-6238. doi: 10.1523/jneurosci.0504-08.2008

Rajesh, M., Mukhopadhyay, P., Batkai, S., Hasko, G., Liaudet, L., Drel, V. R., . . . Pacher, P.

(2007). Cannabidiol attenuates high glucose-induced endothelial cell inflammatory

response and barrier disruption. Am J Physiol Heart Circ Physiol, 293(1), H610-619. doi:

10.1152/ajpheart.00236.2007

87

Rao, G. K., & Kaminski, N. E. (2006). Cannabinoid-mediated elevation of intracellular calcium:

a structure-activity relationship. J Pharmacol Exp Ther, 317(2), 820-829. doi:

10.1124/jpet.105.100503

Resstel, L. B., Joca, S. R., Moreira, F. A., Correa, F. M., & Guimaraes, F. S. (2006). Effects of

cannabidiol and diazepam on behavioral and cardiovascular responses induced by

contextual conditioned fear in rats. Behav Brain Res, 172(2), 294-298. doi:

10.1016/j.bbr.2006.05.016

Resstel, L. B., Tavares, R. F., Lisboa, S. F., Joca, S. R., Correa, F. M., & Guimaraes, F. S. (2009).

5-HT1A receptors are involved in the cannabidiol-induced attenuation of behavioural and

cardiovascular responses to acute restraint stress in rats. Br J Pharmacol, 156(1), 181-

188. doi: 10.1111/j.1476-5381.2008.00046.x

Richards, M. W., Butcher, A. J., & Dolphin, A. C. (2004). Ca2+ channel beta-subunits: structural

insights AID our understanding. Trends Pharmacol Sci, 25(12), 626-632. doi:

10.1016/j.tips.2004.10.008

Robson, P. (2001). Therapeutic aspects of cannabis and cannabinoids. Br J Psychiatry, 178, 107-

115.

Rosenkrantz, H., Fleischman, R. W., & Grant, R. J. (1981). Toxicity of short-term administration

of cannabinoids to rhesus monkeys. Toxicol Appl Pharmacol, 58(1), 118-131.

Ross, H. R., Napier, I., & Connor, M. (2008). Inhibition of recombinant human T-type calcium

channels by Delta9-tetrahydrocannabinol and cannabidiol. J Biol Chem, 283(23), 16124-

16134. doi: 10.1074/jbc.M707104200

Ruiz-Valdepenas, L., Martinez-Orgado, J. A., Benito, C., Millan, A., Tolon, R. M., & Romero, J.

(2011). Cannabidiol reduces lipopolysaccharide-induced vascular changes and

inflammation in the mouse brain: an intravital microscopy study. J Neuroinflammation,

8(1), 5. doi: 10.1186/1742-2094-8-5

Russo, E., & Guy, G. W. (2006). A tale of two cannabinoids: the therapeutic rationale for

combining tetrahydrocannabinol and cannabidiol. Med Hypotheses, 66(2), 234-246. doi:

10.1016/j.mehy.2005.08.026

Russo, E. B., Burnett, A., Hall, B., & Parker, K. K. (2005). Agonistic properties of cannabidiol at

5-HT1a receptors. Neurochem Res, 30(8), 1037-1043. doi: 10.1007/s11064-005-6978-1

Ryan, D., Drysdale, A. J., Lafourcade, C., Pertwee, R. G., & Platt, B. (2009). Cannabidiol targets

mitochondria to regulate intracellular Ca2+ levels. J Neurosci, 29(7), 2053-2063. doi:

10.1523/jneurosci.4212-08.2009

88

Ryberg, E., Larsson, N., Sjogren, S., Hjorth, S., Hermansson, N. O., Leonova, J., . . . Greasley, P.

J. (2007). The orphan receptor GPR55 is a novel cannabinoid receptor. Br J Pharmacol,

152(7), 1092-1101. doi: 10.1038/sj.bjp.0707460

Samara, E., Bialer, M., & Harvey, D. J. (1990). Identification of urinary metabolites of

cannabidiol in the dog. Drug Metab Dispos, 18(5), 571-579.

Schicho, R., & Storr, M. (2011). Alternative targets within the endocannabinoid system for future

treatment of gastrointestinal diseases. Can J Gastroenterol, 25(7), 377-383.

Scuderi, C., Filippis, D. D., Iuvone, T., Blasio, A., Steardo, A., & Esposito, G. (2009).

Cannabidiol in medicine: a review of its therapeutic potential in CNS disorders.

Phytother Res, 23(5), 597-602. doi: 10.1002/ptr.2625

Shepherd, N., & McDonough, H. B. (1998). Ionic diffusion in transverse tubules of cardiac

ventricular myocytes. Am J Physiol, 275(3 Pt 2), H852-860.

Showalter, V. M., Compton, D. R., Martin, B. R., & Abood, M. E. (1996). Evaluation of binding

in a transfected cell line expressing a peripheral cannabinoid receptor (CB2):

identification of cannabinoid receptor subtype selective ligands. J Pharmacol Exp Ther,

278(3), 989-999.

Skrabek, R. Q., Galimova, L., Ethans, K., & Perry, D. (2008). Nabilone for the treatment of pain

in fibromyalgia. J Pain, 9(2), 164-173. doi: 10.1016/j.jpain.2007.09.002

Smiley, K. A., Karler, R., & Turkanis, S. A. (1976). Effects of cannabinoids on the perfused rat

heart. Res Commun Chem Pathol Pharmacol, 14(4), 659-675.

Spurgeon, H. A., duBell, W. H., Stern, M. D., Sollott, S. J., Ziman, B. D., Silverman, H. S., . . .

Lakatta, E. G. (1992). Cytosolic calcium and myofilaments in single rat cardiac myocytes

achieve a dynamic equilibrium during twitch relaxation. J Physiol, 447, 83-102.

Stanley, C. P., Hind, W. H., & O'Sullivan, S. E. (2013). Is the cardiovascular system a therapeutic

target for cannabidiol? Br J Clin Pharmacol, 75(2), 313-322. doi: 10.1111/j.1365-

2125.2012.04351.x

Sumariwalla, P. F., Gallily, R., Tchilibon, S., Fride, E., Mechoulam, R., & Feldmann, M. (2004).

A novel synthetic, nonpsychoactive cannabinoid acid (HU-320) with antiinflammatory

properties in murine collagen-induced arthritis. Arthritis Rheum, 50(3), 985-998. doi:

10.1002/art.20050

Takahashi, M., Seagar, M. J., Jones, J. F., Reber, B. F., & Catterall, W. A. (1987). Subunit

structure of dihydropyridine-sensitive calcium channels from skeletal muscle. Proc Natl

Acad Sci U S A, 84(15), 5478-5482.

89

Takeda, S., Usami, N., Yamamoto, I., & Watanabe, K. (2009). Cannabidiol-2',6'-dimethyl ether, a

cannabidiol derivative, is a highly potent and selective 15-lipoxygenase inhibitor. Drug

Metab Dispos, 37(8), 1733-1737. doi: 10.1124/dmd.109.026930

Thomas, A., Baillie, G. L., Phillips, A. M., Razdan, R. K., Ross, R. A., & Pertwee, R. G. (2007).

Cannabidiol displays unexpectedly high potency as an antagonist of CB1 and CB2

receptor agonists in vitro. Br J Pharmacol, 150(5), 613-623. doi: 10.1038/sj.bjp.0707133

Thompson, A. J., Chau, P. L., Chan, S. L., & Lummis, S. C. (2006). Unbinding pathways of an

agonist and an antagonist from the 5-HT3 receptor. Biophys J, 90(6), 1979-1991. doi:

10.1529/biophysj.105.069385

Turkanis, S. A., & Karler, R. (1986). Cannabidiol-caused depression of spinal motoneuron

responses in cats. Pharmacol Biochem Behav, 25(1), 89-94.

Vanhoutte, P. M., Shimokawa, H., Tang, E. H., & Feletou, M. (2009). Endothelial dysfunction

and vascular disease. Acta Physiol (Oxf), 196(2), 193-222. doi: 10.1111/j.1748-

1716.2009.01964.x

Walsh, S. K., Hepburn, C. Y., Kane, K. A., & Wainwright, C. L. (2010). Acute administration of

cannabidiol in vivo suppresses ischaemia-induced cardiac arrhythmias and reduces

infarct size when given at reperfusion. Br J Pharmacol, 160(5), 1234-1242. doi:

10.1111/j.1476-5381.2010.00755.x

Wang, S. Q., Song, L. S., Lakatta, E. G., & Cheng, H. (2001). Ca2+ signalling between single L-

type Ca2+ channels and ryanodine receptors in heart cells. Nature, 410(6828), 592-596.

doi: 10.1038/35069083

Ware, M. A., Daeninck, P., & Maida, V. (2008). A review of nabilone in the treatment of

chemotherapy-induced nausea and vomiting. Ther Clin Risk Manag, 4(1), 99-107.

White, H. L., & Tansik, R. L. (1980). Effects of delta 9-tetrahydrocannabinol and cannabidiol on

phospholipase and other enzymes regulating arachidonate metabolism. Prostaglandins

Med, 4(6), 409-417.

Xiong, W., Cui, T., Cheng, K., Yang, F., Chen, S. R., Willenbring, D., . . . Zhang, L. (2012).

Cannabinoids suppress inflammatory and neuropathic pain by targeting alpha3 glycine

receptors. J Exp Med, 209(6), 1121-1134. doi: 10.1084/jem.20120242

Yang, J., Ellinor, P. T., Sather, W. A., Zhang, J. F., & Tsien, R. W. (1993). Molecular

determinants of Ca2+ selectivity and ion permeation in L-type Ca2+ channels. Nature,

366(6451), 158-161. doi: 10.1038/366158a0

Yang, K. H., Galadari, S., Isaev, D., Petroianu, G., Shippenberg, T. S., & Oz, M. (2010). The

nonpsychoactive cannabinoid cannabidiol inhibits 5-hydroxytryptamine3A receptor-

90

mediated currents in Xenopus laevis oocytes. J Pharmacol Exp Ther, 333(2), 547-554.

doi: 10.1124/jpet.109.162594

Zhou, Z., & January, C. T. (1998). Both T- and L-type Ca2+ channels can contribute to

excitation-contraction coupling in cardiac Purkinje cells. Biophys J, 74(4), 1830-1839.

doi: 10.1016/s0006-3495(98)77893-2

Zuardi, A. W. (2008). Cannabidiol: from an inactive cannabinoid to a drug with wide spectrum of

action. Rev Bras Psiquiatr, 30(3), 271-280.

Zucchi, R., & Ronca-Testoni, S. (1997). The sarcoplasmic reticulum Ca2+ channel/ryanodine

receptor: modulation by endogenous effectors, drugs and disease states. Pharmacol Rev,

49(1), 1-51.