<<

REETA TALKA Effects of on Neuronal Nicotinic Receptors Recent Publications in this Series

10/2017 Veera Pohjolainen Health-Related Quality of Life and Cost-Utility in Bulimia Nervosa and Nervosa in Women 11/2017 Lotta von Ossowski Interaction of GluA1 AMPA with Synapse-Associated Protein 97 12/2017 Emma Andersson dissertationes scholae doctoralis ad sanitatem investigandam Characterization of Mature T-Cell Leukemias by Next-Generation Sequencing and universitatis helsinkiensis 30/2017 Sensitivity Testing 13/2017 Solja Nyberg Job Strain as a Risk Factor for Obesity, Physical Inactivity and Type 2 Diabetes – a Multi-cohort Study 14/2017 Eero Smeds REETA TALKA Cortical Processes Related to Motor Stability and Proprioception in Human Adults and Newborns Effects of Opioids on Neuronal Nicotinic 15/2017 Paavo Pietarinen Effects of Genotype and Phenotype in Personalized Drug Therapy Acetylcholine Receptors 16/2017 Irene Ylivinkka Netrins in Glioma Biology: Regulators of Tumor Cell Proliferation, Motility and Stemness 17/2017 Elisa Lázaro Ibáñez Extracellular Vesicles: Prospects in Prostate Cancer Biomarker Discovery and Drug Delivery 18/2017 Anu Kaskinen Measurement of Lung Liquid and Outcome after Congenital Cardiac Surgery 19/2017 Taru Hilander Molecular Consequences of Transfer-RNA Charging Defects 20/2017 Laura Teirilä Activation of the Inflammatory Response by Fungal Components 21/2017 Laura Sokka Burnout in the at Work 22/2018 Martti Rechardt Metabolic and Inflammatory Factors in Upper Extremity Soft-Tissue Disorders 23/2017 Jaana Hautala Improving the Palatability of Minitablets for Feline 24/2017 Satu Lehti Extracellular Lipid Particles in Atherosclerosis and Aortic Stenosis 25/2017 Asko Wegelius Influence of Birth Weight on the Risk and Clinical Presentation of Schizophrenia 26/2017 Siva P.R. Maddirala Venkata Public Health and Patient Care Aspects in Pharmacy Education and Pharmacists’ Role in National Health Care Programs in India 27/2017 Kristyna Spillerova The Role of the Angiosome Concept in the Treatment of below the knee Critical Limb Ischemia 28/2017 Anna-Riia Holmström Learning from Medication Errors in Healthcare — How to Make Medication Error Reporting Systems Work? 29/2017 Aaro Haapaniemi DIVISION OF AND PHARMACOTHERAPY Laryngeal Cancer Recurrence, Prognostic Factors and Management FACULTY OF PHARMACY DOCTORAL PROGRAMME IN DRUG RESEARCH UNIVERSITY OF HELSINKI 30/2017

Helsinki 2017 ISSN 2342-3161 ISBN 978-951-51-3126-3 Division of Pharmacology and Pharmacotherapy Faculty of Pharmacy University of Helsinki Finland

Effects of Opioids on Neuronal Nicotinic Acetylcholine Receptors

Reeta Talka

ACADEMIC DISSERTATION

To be presented, with the permission of the Faculty of Pharmacy, University of Helsinki, for public examination at Viikki Biocenter 2, Auditorium 1041, on May 19th 2017, at 12 noon.

Supervisors: Docent Outi Salminen, PhD Division of Pharmacology and Pharmacotherapy Faculty of Pharmacy University of Helsinki Finland

Professor Raimo K. Tuominen, MD, PhD Division of Pharmacology and Pharmacotherapy Faculty of Pharmacy University of Helsinki Finland

Reviewers: Docent Petri Hyytiä, PhD Department of Pharmacology Faculty of University of Helsinki Finland

Professor Jyrki Kukkonen, PhD Department of Veterinary Biosciences Faculty of Veterinary Medicine University of Helsinki Finland

Opponent: Professor Neil Millar, PhD Department of Neuroscience, Physiology and Pharmacology University College London United Kingdom

©Reeta Talka Cover photo by H. Zell Dissertationes Scholae Doctoralis Ad Sanitatem Investigandam Universitatis Helsinkiensis ISBN 978-951-51-3126-3 (Paperback), ISBN 978-951-51-3127-0 (PDF), ISSN 2342-3161 (Paperback), ISSN 2342-317X (PDF) Hansaprint Oy, Turenki, Finland 2017

ABSTRACT

Tobacco use is the leading cause of preventable death worldwide. is the primary addictive component of , and repeated nicotine exposure often leads to dependence in humans. Nicotine is one of the most commonly co-used substances among polysubstance abuse patients and combined use of nicotine and other of abuse, such as opioids, increases the use of one or both substances. The health consequences associated with polysubstance abuse exceed those of either drug alone. The current pharmacotherapeutic options are ineffective among -substituted patients and the levels of successful smoking cessation are low. At the cellular level, nicotine and opioids have their own molecular mechanisms of action, yet both drugs increase the activity of the reward pathway by increasing (DA) transmission.

The purpose of these studies was to investigate the possible effects of different opioid ligands on human neuronal nicotinic acetylcholine receptors (nAChRs) expressed in cell cultures. In the first part, the effects of on nAChRs was investigated with -binding and functional studies, and in the second part the effects of were studied with similar methods. Since these opioids showed effects on nAChRs, the next step was to study the effect of prolonged drug treatments on nAChR numbers and function because the nAChRs are known to be upregulated by chronic nicotine exposure. Additionally, the effects of other opioid ligands, , , , , and were also studied.

Our results showed that morphine has a partial effect at α4β2 nAChRs, a very weak antagonist effect at α3* nAChRs (where * denotes other nAChR subunits that may not have been identified) and a positive synergistic effect with nicotine on α7 nAChR function. We found that methadone acts as a non- competitive antagonist (NCA) at α4β2 and α3* nAChRs. We also confirmed that methadone is a human α7 nAChR agonist. In the prolonged studies with methadone and morphine, we found that human α3*, α4β2 and α7 nAChRs are differentially regulated by prolonged exposure to methadone and morphine. Methadone and morphine up-regulate α3* and α7 nAChRs, whereas α4β2 nAChRs are down-regulated. Methadone-induced up-regulation of α3* nAChRs has no effect on the function of cell surface receptors, while methadone and morphine-induced down-regulation of α4β2 nAChRs changes the function of receptors on the cell surface. Buprenorphine was shown to be a weak antagonist at α4β2, α3*, and α7 nAChRs, and codeine had a positive modulatory effect on α4β2 nAChRs and a weak NCA effect on α3* nAChRs. Oxycodone seemed to have a mixed competitive/non-competitive effect on α4β2 nAChRs and a weak NCA effect on α3* nAChRs. Tramadol was shown to be a NCA of α3* nAChRs and a weak NCA of α4β2 nAChRs. Naloxone and naltrexone were mixed competitive/non-competitive antagonists of α4β2 nAChRs, weak NCAs of α3* nAChRs and weak antagonists of α7 nAChRs.

Taken together, these studies showed that many opioid ligands have effects on nAChRs that are independent of their agonist or antagonist properties at opioid receptors. These findings suggest that some effects of the nicotine–opioid interaction seen in humans can be partially mediated through the receptor-level interplay of these substances. These results, together with earlier findings, highlight the

complexity of different nAChRs and the multiplicity of responses to opioid ligands. This variability should be taken into account when designing treatments for polysubstance dependence.

CONTENTS

1. INTRODUCTION ...... 1 2. REVIEW OF THE LITERATURE ...... 3 2.1 Neuronal nicotinic acetylcholine receptors ...... 3 2.1.1 Structure and classification ...... 3 2.1.2 Distribution and localization ...... 5 2.1.3 Functionality of nAChRs ...... 7 2.1.4 Regulation of nAChR numbers ...... 9 2.2 Nicotinic ligands ...... 13 2.2.1 Nicotinic and antagonists ...... 13 2.2.2 Allosteric modulation ...... 14 2.3 The connection between opioid ligands, nAChRs and nicotine ...... 14 2.3.1 Mechanism of ...... 14 2.3.2 Opioid receptors and their signaling ...... 16 2.3.3 Opioid dependence and pharmacological treatment options ...... 17 2.3.4 Methadone...... 17 2.3.5 Morphine ...... 1 8 2.3.6 Other opioids ...... 18 3. AIMS OF THE STUDY ...... 20 4. MATERIALS AND METHODS ...... 21 4.1 Drugs and reagents ...... 21 4.2 Cell culture ...... 21 4.3 [3H] binding ...... 22 4.4 Calcium fluorometry ...... 2 3 4.5 86Rb+ efflux ...... 24 4.6 Statistical analysis ...... 24 5. RESULTS ...... 26 5.1 Methadone ...... 26 5.1.1 Effects of acute methadone treatment (II) ...... 26 5.1.2 Effects of prolonged methadone treatment (IV) ...... 27 5.2 Morphine ...... 29 5.2.1 Effects of acute morphine treatment (I) ...... 29 5.2.2 Effects of prolonged morphine treatment (IV) ...... 29 5.3 Buprenorphine (unpublished) ...... 30 5.3.1 Effects of acute buprenorphine treatment ...... 30 5.4 Codeine (unpublished) ...... 33 5.4.1 Effects of acute codeine treatment ...... 33 5.5 Oxycodone (unpublished) ...... 35 5.5.1 Effects of acute oxycodone treatment ...... 35 5.6 Tramadol (unpublished) ...... 36

5.6.1 Effects of acute tramadol treatment ...... 36 5.7 Naloxone and naltrexone (unpublished) ...... 38 5.7.1 Effects of acute naloxone and naltrexone treatments ...... 38 6. DISCUSSION ...... 41 6.1 Effects of opioids on α4β2 nAChRs ...... 41 6.2 Effects of opioids on α3* nAChRs ...... 42 6.3 Effects of opioids on α7 nAChRs ...... 43 6.4 General discussion ...... 45 7. SUMMARY AND CONCLUSIONS ...... 47 8. ACKNOWLEDGEMENTS ...... 48 9. REFERENCES...... 50

LIST OF ORIGINAL PUBLICATIONS

This thesis is based on the following publications:

I Talka R, Salminen O, Whiteaker P, Lukas RJ, Tuominen RK. Nicotine–morphine interactions at α4β2, α7 and α3* nicotinic acetylcholine receptors. Eur J Pharmacol. 701:57-64, 2013

II Talka R, Salminen O, Tuominen RK. Methadone is a non-competitive antagonist at the α4β2 and α3* nicotinic acetylcholine receptors and an agonist at the α7 nicotinic acetylcholine receptor. Basic Clin Pharmacol Toxicol. 116:321-8, 2015

III Talka R, Tuominen RK, Salminen O. Methadone's effect on nAChRs -a link between methadone use and smoking? Biochem Pharmacol. 97:542-9, 2015

IV Talka R, Tuominen RK, Salminen O. Effects of prolonged methadone and morphine treatment on nAChR function and receptor number in cell lines expressing α7, α4β2 and α3* nAChRs. Manuscript

The publications are referred to in the text by their roman numerals. Reprints were made with permission from the copyright holders.

ABBREVIATIONS

ACh acetylcholine α-Bgtx α- ANOVA analysis of variance

Bmax maximum specific binding 2+ [Ca ]i concentration of free intracellular calcium CNS central nervous system CPP conditioned place preference DA dopamine

EC50 concentration activating 50% of maximum [3H]EPI [3H]epibatidine ER endoplasmic reticulum GABA γ-aminobutyric acid GPCR G protein–coupled receptor

IC50 concentration inhibiting 50% of maximum

Kd equilibrium binding constant

Ki equilibrium dissociation constant MAO Meca MHb medial habenula MLA mRNA messenger ribonucleic acid NAc nucleus accumbens nAChR neuronal nicotinic acetylcholine receptor NAM negative NCA non-competitive antagonist PAM positive allosteric modulator 86Rb+ rubidium-86 isotope SAM silent allosteric modulator SD standard deviation SEM standard error of mean TSS Tyrode's salt solution VOCCs voltage-operated Ca2+ channels VTA ventral tegmental area

Introduction

1. INTRODUCTION

Tobacco use is a worldwide public health problem and is the leading cause of preventable death. The vast majority of lung cancer deaths are caused by (Islami et al. 2015). In Finland, smoking caused approximately 4 300–4 500 deaths in 2012 (Jääskeläinen and Virtanen 2015). Nicotine is the primary addictive component in tobacco, and repeated nicotine exposure often leads to dependence in humans. In addition to the problem of nicotine dependence in the general population, tobacco smoking among individuals with substance-use disorders presents another public health concern. Nicotine is one of the most commonly co-used substances among polysubstance abuse patients (McClure et al. 2014; Goodwin et al. 2014). Clinical and epidemiological studies suggest that nicotine, when combined with other drugs of abuse, increases consumption of one or both substances (Mello et al. 1980; Henningfield and Griffiths 1981; Chait and Griffiths 1984; Higgins et al. 1994). The health consequences associated with polysubstance abuse exceed those of either drug alone and the mortality rate of polysubstance users is significantly increased due to tobacco-related causes (Hurt et al. 1996).

Almost all methadone-maintained patients are tobacco smokers (Nahvi et al. 2006; Richter et al. 2007; Elkader et al. 2009; Pajusco et al. 2012; Chisolm et al. 2013). In addition to methadone, the use of other opioid agonists, such as buprenorphine and , also increase smoking prevalence (Mello et al. 1980; Mello et al. 1985; Mutschler et al. 2002; Pajusco et al. 2012). Furthermore, smokers are more likely to be abusers of prescription opioids (Zale et al. 2015). Nicotine use increases methadone self-administration and reinforcing properties, and smoking combined with methadone use also increases subjective ratings of smoking satisfaction (Chait and Griffiths 1984; Spiga et al. 1998). Although smoking cessation is more difficult for maintenance patients than for the general population, quitting smoking also improves opioid abstinence, thus, effective treatment for tobacco dependence is particularly beneficial among these patients (Lemon et al. 2003).

At the cellular level, nicotine and opioids both have their own molecular mechanisms of action. Nicotine interacts directly with neuronal nicotinic acetylcholine receptors (nAChRs), which are pentameric ligand- gated ion channels. Opioid agonists act through G protein-coupled opioid receptors. A commonality among all addictive drugs is their ability to increase extracellular DA levels in the nucleus accumbens (NAc) (Nestler 2005; Ross and Peselow 2009). Both nicotine and opioids increase DA by indirect mechanisms that affect DA cell firing. The activation of μ-opioid receptors increases DA transmission by inhibiting γ-aminobutyric acid-ergic (GABAergic) interneurons that normally provide tonic inhibition to neurons in the ventral tegmental area (VTA) (Kosten and George 2002). Nicotine binds to the nAChRs located on the dopaminergic cell bodies and on GABAergic and neurons in the VTA (Pidoplichko et al. 1997; Mansvelder and McGehee 2002), which causes a shift from tonic firing of dopaminergic neurons to burst firing, resulting in an increase in DA levels in the NAc and the prefrontal cortex (Rao et al. 2003; Rice and Cragg 2004). Furthermore, the endogenous opioid system influences nicotine reward and antinociception. The levels of endogenous opioid peptides, such as and β-endorphin, are increased following nicotine administration (Dhatt et al. 1995). These endogenous peptides bind to μ-opioid receptors located on the GABA interneurons in the VTA, which further increases DA release in the NAc by disinhibiting the GABAergic interneurons (Davenport et al. 1990; Bergevin et al. 2002). Nicotine may thus potentiate opioid-induced antinociception via nAChR activation.

1 Introduction

The acute rewarding aspects of drug use and conditioned learning associated with craving and relapse seem to be mediated by the mesolimbic dopaminergic pathway, whereas adaptations in the mesocortical and corticofugal glutamatergic pathways play a role in the loss of inhibitory control and continued drug seeking behaviors (Feltenstein and See 2008). Other brain regions and systems, such as the insula, thalamus and cerebellum and stress-related brain systems, also have modulatory functions in the development and maintenance of drug dependence (Baler and Volkow 2006; Zorrilla et al. 2014; Korpi et al. 2015). Since the mechanisms, brain regions and systems involved in polysubstance drug dependence are greatly diverse and complex, it is no wonder that effective pharmacological treatments are challenging to develop. The ideal pharmacotherapeutic treatment should be able to treat withdrawal symptoms, induce abstinence, reduce substance use and prevent relapse.

2 Review of the Literature

2. REVIEW OF THE LITERATURE

2.1 Neuronal nicotinic acetylcholine receptors

2.1.1 Structure and classification

The nAChRs belong to the cys-loop receptor superfamily of ligand-gated receptors, which includes GABAA, , and 5-HT3 () receptors and the muscle-type nicotinic receptors (Miller and Smart 2010). They are often referred as the cys-loop receptors since they share a similar topology with a disulphide-bridged cys-loop in the extracellular domain. Mammalian nAChRs are composed of five subunits arranged around a pore filled with water (Fig. 1) (Dani 2015; Fasoli and Gotti 2015). Each subunit is composed of a long extracellular N-terminal domain, where the ligand binding site is located, followed by four hydrophobic transmembrane regions (M1–M4). Between transmembrane regions M3 and M4 there is a large intracellular loop, where the phosphorylation sites for proto-oncogene -protein kinase Src family kinases are (Charpantier et al. 2005). The M4 region ends with a short extracellular C- terminus.

Figure 1. The structure of α7, α3β4 and α4β2 nAChRs. The nAChRs are pentameric ion channel receptors with long extracellular N-terminal domains, followed by four transmembrane regions (M1–M4). The M4 region ends with a short extracellular C-terminus. The ionic pore of nAChRs is mainly lined by the M2 transmembrane segment. The heteromeric receptors have two orthosteric ligand binding sites at the extracellular N-terminal domain, whereas the homomeric receptors have five binding sites.

The ionic pore of nAChRs is mainly lined by the M2 transmembrane segment with some contribution from the M1 segment (Hucho et al. 1986; Karlin 2002). The ionic pore contains residues important for the ion selectivity, permeability, and channel gating of nAChRs. The other transmembrane segments (M1, M3 and M4) separate the M2 segment from the hydrophobic cell membrane (Papke 2014). The intracellular loop between M3 and M4 transmembrane segments, which holds the phosphorylation sites for Src family kinases, is the most variable element in nAChRs and has a profound

3 Review of the Literature influence upon receptor assembly, targeting and ion channel properties (Charpantier et al. 2005; Kracun et al. 2008; Pollock et al. 2009).

The extracellular N-terminal domain contains the orthosteric agonist-binding site in nAChRs (Brejc et al. 2001; Bartos et al. 2009). The orthosteric binding site is composed of many amino acid residues grouped into loops A, B, and C (the principal component) and D, E, and F (the complementary component) (Fig. 2) (Changeux and Taly 2008). The agonists bind to a hydrophobic pocket formed by loops A, B, D, and F, after which loop C closes this binding pocket. When antagonists bind, loop C stays in the open conformation.

Figure 2. Structure of the acetylcholine (ACh)-binding site. Schematic representation of the ACh-binding site illustrating the loops A-C (the principal component) and loops D-F (the complementary component). Upon agonist binding, loop C closes the binding pocket formed by loops A, B, D and F.

The mammalian nAChR subunits are divided into alpha (α2–α7, α9, and α10) (the α8 nAChRs are only found in avian species) and beta (β2–β4) subunits based on the presence of adjacent groups in the extracellular domain of only the α subunits (Gotti et al. 2009). These subunits form either homo- or heteromeric pentameric receptors; the great diversity of different combinations also makes the functionality diverse. Two main classes of nAChR subtypes have been identified: α-bungarotoxin (α-Bgtx) -sensitive and -insensitive receptors. The α-Bgtx-sensitive receptors are homomeric or heteromeric receptors made up of α7, α9 and/or α10 subunits, and they bind α-Bgtx with high affinity. The α-Bgtx- insensitive receptors are heteromeric receptors formed by combinations of α (α2–α6) and β (β2–β4) subunits, which do not bind α-Bgtx. The α-Bgtx-insensitive receptors have a higher affinity for nicotine and other nicotinic agonists than the α-Bgtx-sensitive receptors (Gotti and Clementi 2004).

The homomeric α7 and α9 nAChRs have five identical binding sites to which the same subunit contributes both the principal and complementary component (Fig. 1) (Zoli et al. 2015). The heteromeric nAChRs have two agonist binding sites at the interface between two adjacent, asymmetric subunits. The principal component of the binding site in heteromeric nAChRs is formed by the α2, α3, α4, α6, α7, or α9 subunits and the complementary site is formed by the β2, β4, α7, α9, or α10 subunits (Figs. 1 and 2). The α5 subunit is not a true α subunit since it only forms functional receptors when co-expressed with a principal and complementary subunit. The α10 subunit cannot act as a principal subunit at the agonist binding

4 Review of the Literature site, and only functions when it is associated with the α9 subunit (Sgard et al. 2002). The β3 subunit is also an accessory subunit since it needs to be co-expressed with a principal and complementary subunit. The presence of accessory subunits has an effect on the pharmacological and functional properties of nAChRs.

In addition to different subunit combinations, subunit stoichiometries also have an effect on nAChR function. The α4β2 and α3β4 nAChR subtypes are example of receptors that can exist in two different stoichiometric arrangements: (α4)2(β2)3 or (α4)3(β2)2, and (α3)2(β4)3 or (α3)3(β4)2 (Moroni et al. 2006; Krashia et al. 2010). The different stoichiometric combinations can have variable agonist sensitivities and Ca2+ permeabilities (Tapia et al. 2007). The α4β2 subtype exhibits biphasic agonist concentration– response curves since the (α4)2(β2)3 combination has a higher sensitivity for agonists, such as nicotine and ACh, than the (α4)3(β2)2 stoichiometry (Zwart and Vijverberg 1998; Nelson et al. 2003; Marks et al.

2010). Chronic nicotine exposure favors increased assembly of the high-sensitivity (α4)2(β2)3 stoichiometry (Moroni et al. 2006; Fasoli et al. 2016; Fasoli et al. 2016). The two stoichiometries of the α3β4 subtype exhibit similar agonist sensitivities, but only the subtype with two α subunits is susceptible to enhancement by low Zn2+ concentrations (Krashia et al. 2010). The two stoichiometries also exhibit substantially different channel conductance and kinetics.

2.1.2 Distribution and localization

The nAChRs are widely and unevenly distributed in the brain and, in most cases, they have presynaptic or preterminal localization, but some are also located post-synaptically in somatodendritic synapses (Fig. 3) (Jones et al. 1999). The nAChRs operate in the brain by multiple mechanisms, yet the most well-studied process is the modulation of release by presynaptic nAChRs (Gray et al. 1996; Role and Berg 1996; Albuquerque et al. 1997; Wonnacott 1997; Dani and Bertrand 2007). When presynaptic nAChRs are activated, the level of intracellular calcium rises, which enhances neurotransmitter release (Vijayaraghavan et al. 1992; Vernino et al. 1992; Rathouz and Berg 1994; Engelman and MacDermott 2004). Axonal and preterminal nAChRs modulate neuron excitability and neurotransmitter release indirectly by activating voltage-operated Ca2+ channels (VOCCs) and initiating action potentials (Lena et al. 1993; Albuquerque et al. 2000). Somatodendritic nAChRs modulate plasticity and information flow by initiating or modulating synaptic inputs to the cell body (Pidoplichko et al. 2013).

5 Review of the Literature

Figure 3. Different localization of nAChR subtypes at synaptic sites. A) Postsynaptic nAChRs bind ACh released from the presynaptic terminal. This type of nicotinic synaptic transmission is fast and direct. B) The presynaptic nAChRs can influence the release of synaptic vesicles. Presynaptic nAChRs initiate a direct and indirect increase of calcium in the presynaptic terminal, which enhances the release of the neurotransmitter. C) The preterminal or axonal nAChRs are located in a position along the axon where they can influence the excitability of the axon.

The α4* nAChRs bind radiolabeled nicotine with the highest affinity and the α4 subunit is predominantly expressed with the β2 subunit in the vertebrate brain (Flores et al. 1992; Clementi et al. 2000). The α4β2 nAChRs are implicated in nicotine self-administration, reward and dependence, and in Alzheimer's disease and epilepsy (Picciotto et al. 1998; Tapper et al. 2004; Steinlein and Bertrand 2010; Jurgensen and Ferreira 2010). This subtype is widely expressed in the mammalian brain and in specific subregions such as the cerebral cortex, , superior colliculus, nucleus geniculatus lateralis and cerebellum (Zoli et al. 2002; Turner and Kellar 2005; Gotti et al. 2005). In addition to the β2 subunit, the α4 subunit can also assemble with other subunits, such as β4 or α5 subunits (Kuryatov et al. 2008; Hamouda et al. 2009). The α4β2 subtype is expressed in both dopaminergic and non-dopaminergic cells in the striatum, whereas the α4α5β2 subtype is localized in dopaminergic terminals (Zoli et al. 2002).

The α7 nAChRs are highly expressed in the brain (e.g., in the cortex, hippocampus and subcortical limbic regions) and have presynaptic, postsynaptic or somatic localization (Picciotto et al. 2001; Jones and Wonnacott 2004). The localization of α7 nAChRs can be readily studied with the α7-specific ligand, α- Bgtx. The channel kinetics of α7 nAChRs are rapid and they are highly permeable to calcium. In addition to the homomeric α7 nAChRs, emerging evidence demonstrates the existence of heteromeric α7 nAChRs, in which α7 subunits are co-assembled with β2 subunits to form a novel type of α7β2 nAChR (Wu et al. 2016). The α7β2 nAChRs have been found in rodents as well as in human basal forebrain neurons and cerebral cortical neurons (Moretti et al. 2014; Thomsen et al. 2015). Compared to the homomeric α7 nAChRs, α7β2 nAChRs have slower whole-cell current amplitudes and decay kinetics (Liu et al. 2009; Liu et al. 2012; Zwart et al. 2014).

The α3* nAChRs are expressed in multiple brain areas such as the pineal gland, medial habenula (MHb), dorsal nucleus of the vagus nerve, anterior thalamus, dopaminergic ventral midbrain, locus coeruleus, and retinal ganglionic neurons (Zoli et al. 1998; Lena et al. 1999; Whiteaker et al. 2002). α3β2* nAChRs are found in the visual pathway of the retina, superior colliculus and nucleus geniculatus lateralis (Gotti

6 Review of the Literature et al. 2005), whereas α3β4* nAChRs are localized in the pineal gland, cerebellum, retina, hippocampus and the habenulo-interpeduncular pathway (Luo et al. 1998; Grady et al. 2001; Hernandez et al. 2004; Turner and Kellar 2005).

Distribution of the α6 subunit in the central nervous system (CNS) is limited. α6* nAChRs are highly expressed in regions such as the substantia nigra, VTA, locus coeruleus, retina, interpeduncular nucleus and MHb, where it often co-localizes with the β3 subunit (Le Novere et al. 1996; Champtiaux et al. 2003). The α6 subunit co-assembles with the β2 subunit in the striatum and retina (Champtiaux et al. 2002; Gotti et al. 2007). The two major α6* subtypes in rodent striatum and retina are the α6α4β2β3 and α6β2β3 nAChRs, which have different binding affinities and sensitivities to α- MII and methyllycaconitine (MLA) (Zoli et al. 2002; Champtiaux et al. 2003).

α2* nAChRs are expressed in small amounts in different brain regions (e.g., in the interpeduncular nucleus, putamen, globus pallidus, motor and somatosensory cortex and thalamus) (Wada et al. 1989). Expression of the α5 subunit is also restricted, with the highest expression in the substantia nigra, VTA and MHb (Picciotto et al. 2001).

2.1.3 Functionality of nAChRs

The subunit composition of each nAChR subtype determines the functional characteristics of the channel. Depending on the subunit composition, each subtype has unique channel kinetics, ion conductance and ion selectivity (Albuquerque et al. 2009). Channel gating is a reversible process that leads to channel opening or closure after a ligand has been bound to the receptor. The gate is located at the ion channel pore where the M2 segment provides the amino acid residues necessary for channel opening and closure. In the closed conformation, the hydrophobic residues located in the middle of the channel approach each other to narrow the channel. Ion flow in a closed state is prevented by the hydrophobic environment, which is energetically unfavorable for ion permeation (Unwin and Fujiyoshi 2012). The channel opens upon agonist binding by concerted tilting of the M2 helices, the M2–M3 loop, and the M3 segment, which increases the diameter of the pore near the middle of the membrane (Taly et al. 2009). The agonist binding region is linked to the channel through a series of interacting residues which transmit the conformational change of the receptor upon ligand binding (Lee and Sine 2005; Lee et al. 2009). nAChRs are permeable to small mono- and divalent cations that are small enough to fit through the channel in the open conformation (Dani and Eisenman 1987; Albuquerque et al. 2009). Sodium and potassium ions contribute to the majority of the ion current, but some nAChR subtypes, such as the homomeric α7 nAChRs, are especially permeable to calcium (Fucile et al. 2003; Fucile 2004). Binding of an agonist stabilizes the open conformation of the nAChR, so that small cations flow through the channel for several milliseconds before the channel closes by going back to a resting state or a desensitized state that is unresponsive to agonists. The cation flow causes depolarization of the cell.

The Ca2+ influx through nAChRs modulates several Ca2+-dependent cellular processes, such as neurotransmitter release, synaptic plasticity and cell motility. The subunit composition of nAChRs influences their Ca2+ permeability. Heteromeric nAChRs have a fractional Ca2+ current of 2–5%, whereas with homomeric α7 nAChRs the fractional Ca2+ current ranges from 6% to 12% in vitro (Fucile et al. 2003;

7 Review of the Literature

Fucile 2004). The incorporation of accessory subunits may change the Ca2+ permeability of nAChRs. For instance, the Ca2+ permeability is increased if the α5 subunit is coassembled with α3 subunit (Gerzanich et al. 1998). The flow of Ca2+ through nAChRs causes an increase in intracellular calcium levels. In addition to this direct flow of ions, VOCCs are also activated by depolarization, which augments the primary Ca2+ signals (Fig. 4) (Dajas-Bailador et al. 2002). Furthermore, Ca2+ is released from the intracellular stores which creates long-lasting Ca2+ signals (Brain et al. 2001). The Ca2+ release from ER is regulated by -

1,4,5- trisphosphate receptors (IP3R), ryanodine receptors (RyR) and sarco-/endoplasmic reticulum calcium ATPase (SERCA) (Dajas-Bailador et al. 2002; Stutzmann and Mattson 2011).

Figure 4. Calcium homeostasis in a neuronal cell. Sources of calcium influx include the glutamate-type receptors (N- methyl-D-aspartate (NMDA) and α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptors), neuronal nicotinic acetylcholine receptors (nAChRs), voltage-operated calcium channels (VOCCs) and conventional transient receptor potential (TRPC) channels. Calcium efflux is mediated by plasma membrane calcium ATPase (PMCA) and sodium-calcium exchanger (NCX). Intracellular calcium levels are also regulated by mitochondria, calcium binding proteins, and the endoplasmic reticulum (ER), where the calcium homeostasis is regulated by inositol-1,4,5- trisphosphate receptors (IP3R), ryanodine receptors (RyR) and sarco-/endoplasmic reticulum calcium ATPase (SERCA).

The nAChRs have three basic conformational states: rest, open, and desensitized (Fig. 5) (Katz and Thesleff 1957; Changeux et al. 1984). In the resting state, the channel is closed until the conformation − upon agonist binding − changes to the open state, allowing ions to flow through the channel. Upon prolonged exposure to an agonist, the receptor conformation changes to a desensitized state, where the channel is closed. The desensitized receptor has a higher affinity for agonists than the resting or open conformations, and thus slow application of low agonist concentrations can cause desensitization

8 Review of the Literature without activation. The kinetics of transitions between different conformational states depends, among other things, on receptor subtype. Therefore, the same agonist may cause different conformational responses depending on the nAChR subtype. For instance, agonist concentrations needed for α4α5β2 nAChR desensitization are, on average, 8-fold higher than what is needed for desensitization of the α4β2 subtype (Wageman et al. 2014). The kinetics of conformational changes are also ligand-dependent. The onset and recovery from desensitization, for example, depend on both duration of ligand exposure and ligand concentration. Long exposures to low concentrations of agonists cause “deeper”, longer-lasting levels of desensitization (Lester and Dani 1994; Dani and Heinemann 1996).

Figure 5. The basic conformational states of nAChRs: rest, open, intermediate and desensitized states. In the resting state, the channel is closed and affinity for antagonists is high. In the open state, the channel is open and affinity is low for agonists and there is little if any affinity for antagonists. In the desensitized state, the channel is closed and affinity for agonists and antagonists is high. Agonists, such as ACh or nicotine, stabilize the open conformation which transiently lets through small cations for several milliseconds before closing upon switch to the resting state or the desensitized state that is unresponsive to agonists. Prolonged exposure to agonists, such as nicotine, produces significant desensitization of nAChRs to the unresponsive closed state. The intermediate state is a conformation between ligand binding and channel opening and between open state and full desensitization. Modified from Changeux et al., (1984).

2.1.4 Regulation of nAChR numbers

Chronic nicotine exposure paradoxically increases the number of nAChRs, whereas long-lasting repeated agonist stimulation usually leads to a reduction in the number of receptors (Creese and Sibley 1981; Wonnacott 1990). nAChRs are upregulated in the brain of tobacco smokers, which has been detected by [3H]nicotine binding in postmortem of smokers and by single-photon emission computed tomography (SPECT) and positron emission tomography (PET) imaging (Benwell et al. 1988; Breese et al. 1997; Perry et al. 1999; Mamede et al. 2007; Mukhin et al. 2008; Wullner et al. 2008; Brody et al. 2013; Jasinska et al. 2014). Furthermore, chronic nicotine has been shown to cause upregulation of nAChRs in numerous in vitro and in vivo experiments (Marks et al. 1983; Schwartz and Kellar 1983; Marks et al.

9 Review of the Literature

1985; Bencherif et al. 1995; Walsh et al. 2008). Nicotine-induced upregulation is brain region-specific; upregulation is found in the brain stem, cerebellum, prefrontal cortex, and corpus callosum, but not in the thalamus (Pauly et al. 1991; Marks et al. 1992; Nguyen et al. 2003; Mukhin et al. 2008; Brody et al. 2013).

In addition to region specificity, upregulation is also nAChR subtype-specific. α4β2 nAChRs have been shown to be upregulated both by low (≤100 nM) and high (>1 μM) nicotine concentrations (Peng et al.

1994; Srinivasan et al. 2011); the high-sensitivity (α4)2(β2)3 stoichiometry is preferentially upregulated by nicotine (Fasoli et al. 2016). The concentration activating 50% of maximum (EC50) for the nicotine- induced upregulation of α4β2 nAChRs varies between tens to hundreds of nM, depending on the assay type (Peng et al. 1994). The steady state plasma concentration of nicotine in the blood of smokers is approximately 150−200 nM (Benowitz 1990; Schneider et al. 2001), and the calculated typical concentrations of free brain nicotine after one are 265 nM after overnight abstinence and 465 nM during afternoon smoking (Rose et al. 2010). Therefore, it is likely, that the α4β2 nAChRs are upregulated by smoking, which is also supported by postmortem human studies (Benwell et al. 1988; Breese et al. 1997; Perry et al. 1999). α3β4* nAChRs on the other hand are not believed to be upregulated at nicotine concentrations achieved by smoking, as they require much higher (≥10 μM) nicotine concentrations for upregulation (Peng et al. 1997; Mazzo et al. 2013). α3β2 nAChRs are more sensitive to nicotine-induced upregulation than α3β4 nAChRs (Walsh et al. 2008). Similar to α4β2 nAChRs, α3β4 nAChRs also exist in two stoichiometries and the (α3)2(β4)3 stoichiometry is preferentially upregulated by chronic nicotine exposure (Mazzo et al. 2013). α7 nAChRs are upregulated only by high concentrations of nicotine, not achieved by moderate smoking, and the magnitude of upregulation is smaller than for α6* and α4* nAChRs (Peng et al. 1997). The regulation of α6* nAChRs depends on the subunit composition: α4α6β2* nAChRs are not upregulated by nicotine, whereas α6[non-α4]β2* are upregulated by high nicotine concentrations (Tumkosit et al. 2006; Perez et al. 2008; Walsh et al. 2008). α6[non-β3]* nAChRs in the striatum are downregulated by nicotine, whereas those containing β3 are unaffected (Perry et al. 2007).

The nicotine-induced upregulation of nAChRs is concentration-dependent in humans (Breese et al. 1997; Perry et al. 1999), in vivo (Rowell and Li 1997) and in vitro (Gopalakrishnan et al. 1997; Govind et al. 2012). Heavy smokers (more than 14 a day) have 25-330% higher nAChR levels compared to non- smokers. The upregulation of nAChRs is reversible: smokers, who have been abstinent more than 3 weeks, have nAChR levels similar to non-smokers (Breese et al. 1997; Mamede et al. 2007). In in vivo animal models nAChR levels return to normal after 8−14 days of withdrawal, depending on assay type (Marks et al. 1985; Collins et al. 1990; Fasoli et al. 2016). Maximal upregulation levels are reached in 1- 14 days in vitro (Peng et al. 1994; Walsh et al. 2008; Srinivasan et al. 2011) and in 10-14 days in vivo (Marks et al. 1983; Nashmi et al. 2007).

Ion flow through nAChRs or receptor activation are not necessary for nAChR upregulation since both NCAs, such as mecamylamine (Meca), and competitive antagonists, such as dihydro-β-erythroidine, upregulate nAChRs (Peng et al. 1994; Kishi and Steinbach 2006). Meca also has an additive effect on upregulation, when administered with nicotine (Peng et al. 1994). Furthermore, upregulation without activation of nAChRs has been shown by using loss-of-function-mutated receptors which allow nicotinic ligands to bind without opening the channel (Kuryatov et al. 2005). Nevertheless, nicotinic agonists are more potent pharmacological chaperones than antagonists, thus inducing upregulation more readily

10 Review of the Literature since the assembly of activated or desensitized conformations is more efficient than the closed conformation. Although nAChR activation is not a requirement for upregulation (Peng et al. 1994), the binding of a nicotinic ligand is necessary. If ligand binding to nAChRs is impaired using mutated nAChRs, the upregulation of α4β2 nAChRs is diminished or completely abolished (Kishi and Steinbach 2006).

The nicotine-induced upregulation of nAChRs is independent of transcriptional events since the messenger ribonucleic acid (mRNA) levels are unchanged in response to nicotine (Marks et al. 1983; Marks et al. 1992; Peng et al. 1994; Bencherif et al. 1995). Several posttranslational mechanisms have been proposed to contribute to nicotine-induced upregulation, such as nAChR trafficking to the cell surface, reduced receptor turnover, nAChR subunit maturation and assembly in the ER, inhibition of subunit degradation in the ER, changes in subunit stoichiometry and nAChR conformational changes (Fig. 6) (Peng et al. 1994; Harkness and Millar 2002; Nashmi et al. 2003; Nelson et al. 2003; Darsow et al. 2005; Sallette et al. 2005; Kuryatov et al. 2005; Ficklin et al. 2005; Vallejo et al. 2005; Moroni et al. 2006; Rezvani et al. 2007).

Figure 6. The suggested mechanisms for nAChR upregulation. The major models of nicotine-induced nAChR upregulation are: 1) increased nAChR trafficking to the cell surface, 2) reduced receptor turnover, 3) increased nAChR subunit maturation and assembly in the ER/Golgi, 4) inhibition of subunit degradation in the ER, 5) changes in subunit stoichiometry and 6) nAChR conformational changes.

One of the mechanisms suggested to contribute to the nicotine-induced upregulation is increased receptor trafficking to the cell surface (Harkness and Millar 2002). The nAChRs are transported from the ER/Golgi to the plasma membrane through the secretory pathway, which is tightly controlled so that only fully assembled receptors eventually reach the cell surface (Wang et al. 2002). In order to be transported from the ER, the nAChRs must be correctly folded into pentamers and recruited at the ER-exit sites. The correct assembly of the five subunits buries a motif responsible for the retention of unassembled subunits, thus, only assembled pentamers are allowed to be delivered to the membrane. When the

11 Review of the Literature secretory pathway from the Golgi apparatus is blocked with brefeldin A, nicotine-induced upregulation of surface α4β2 nAChRs is blocked, whereas upregulation of the total binding sites is not inhibited (Darsow et al. 2005). This indicates that the upregulation of intracellular nAChRs is independent of the transport mechanism through the secretory pathway, but this pathway is necessary for the upregulation of surface nAChRs.

The reduced turnover of cell-surface receptors has been suggested to contribute to the nicotine-induced upregulation of nAChRs since nicotine was shown to slow down the turnover of surface α4β2 nAChRs (Peng et al. 1994). The nAChRs treated with high concentrations of nicotine (5 μM) remained on the plasma membrane longer than untreated nAChRs. Subsequent studies produced mixed results: some groups were unable to replicate these results (Vallejo et al. 2005; Darsow et al. 2005; Sallette et al. 2005), whereas others succeeded (Kuryatov et al. 2005). More studies are still needed to provide evidence regarding whether nAChR stability and turnover is altered by chronic nicotine exposure.

Increased receptor assembly and subunit maturation in the ER is one suggested mechanism for nAChR upregulation, and many studies have shown that receptor assembly and maturation are increased during α4β2 nAChR upregulation (Harkness and Millar 2002; Nashmi et al. 2003; Sallette et al. 2005; Kuryatov et al. 2005). The increase in the assembly process leads to growing receptor numbers on the cell surface. Nicotine may promote the assembly of subunits by acting as a molecular chaperone in the ER and may also increase the half-life of surface receptors (Kuryatov et al. 2005).

Degradation and trafficking of nAChRs is regulated partly by the ubiquitin–proteosome system (Yi and Ehlers 2007). Blocking the ER-associated proteosome-mediated degradation of nAChRs can increase the receptor numbers expressed on the cell surface. In the case of α7 nAChRs, nicotine has been shown to inhibit the activity of the proteasome directly, without activating nAChRs, which leads to increased receptor numbers (Rezvani et al. 2007). Overexpression of ubiquilin-1, a ubiquitin-like protein with the capacity to interact with both the proteosome and ubiquitin ligases, abolishes the nicotine-induced upregulation of surface α3* nAChRs by promoting ER-associated proteosome-mediated degradation (Ficklin et al. 2005).

α4β2 nAChRs exist in two different stoichiometries and nicotine selectively upregulates the high- sensitivity stoichiometry (α4)2(β2)3 (Nelson et al. 2003; Kuryatov et al. 2005; Tapia et al. 2007; Son et al. 2009; Srinivasan et al. 2011). The affinity of nicotine for the high-sensitivity α4β2 nAChRs is about 100- fold higher than for the low-sensitivity α4β2 nAChRs; thus, nicotine acts as a chaperone mainly at the high-sensitivity subtype to which it binds best (Nelson et al. 2003; Kuryatov et al. 2005). When oocytes are injected with different subunit ratios of mRNAs coding for α4 and β2 subunits, α4β2 nAChRs with high- and low-sensitivity are formed, yet only the high-sensitivity type is notably upregulated by nicotine (Lopez-Hernandez et al. 2004). Nicotine-induced stabilization of the high-sensitivity α4β2 nAChRs happens at the level of the ER and its exit sites, where the ER export of the high-sensitivity subtype is enhanced (Srinivasan et al. 2011) and the number of high-sensitivity α4β2 nAChRs on the cell surface is increased (Moroni et al. 2006).

The change in the nAChR conformation is one of the suggested mechanisms of nicotine-induced upregulation. According to this theory, there is a nicotine-induced increase in the number of high-affinity binding sites, with no increase in the number of receptors or change in the assembly, trafficking, or cell-

12 Review of the Literature surface turnover of α4β2 nAChRs (Vallejo et al. 2005). Nicotine regulates the transition between two conformational states: the resting and the upregulated state. In the resting state, the affinity for agonists is low/normal and nicotine binding activates and desensitizes the receptors as usual. The suggested mechanism is that chronic nicotine exposure induces a transition to the upregulated state, where the affinity for the agonists is high and an increase in the agonist binding sites is observed. The conformational change also converts the functional state of the receptor. Nicotine is able to induce functional upregulation of α4β2 nAChRs (Buisson et al. 2000; Buisson and Bertrand 2001), caused by a conformational change in which the desensitization rate, single-channel conductance and ligand concentration-dependence of activation are changed. On the other hand, it has been demonstrated that the increase in α4β2* nAChR binding sites in mouse brain results from increases in assembled nAChR subunit proteins (Marks et al. 2011).

Since there is a lot of data about other upregulation mechanisms, the conformational change of nAChRs induced by chronic nicotine is unlikely to be the sole regulatory mechanism of upregulation. Instead, upregulation seems to be a multiple-step process occurring at different rates and through different mechanisms. The kinetic studies of upregulation have shown that there are two components of upregulation: an initial fast component that saturates after approximately 4h and a second slower phase which results in higher receptor numbers (Govind et al. 2012). It is hypothesized that the initial component is due to nicotine-induced conformational changes in nAChRs, and the second component results from increased receptor assembly and decreased subunit degradation (Darsow et al. 2005; Ficklin et al. 2005; Vallejo et al. 2005; Govind et al. 2012).

2.2 Nicotinic acetylcholine receptor ligands

2.2.1 Nicotinic agonists and antagonists nAChRs exhibit multiple ligand-binding sites: the orthosteric site, the allosteric sites and the ion channel (Cecchini and Changeux 2015). Nicotinic agonists are ligands that bind to the orthosteric binding site of the receptor located in the extracellular domain at the interface between subunits leading to the activation of the receptor. Partial agonists also bind to the orthosteric binding site, but they only have partial efficacy at the receptor in relation to full agonists. For instance, and are partial α4β2 nAChR agonists that produce a more moderate and sustained increase in DA levels in the reward pathway compared to nicotine, a full agonist (Peng et al. 2013). At the same time, partial agonists prevent the binding of other ligands to the orthosteric site, thus blocking their effect. Competitive antagonists, on the other hand, bind to the orthosteric site, have no efficacy, prevent the binding of other orthosteric ligands and prevent the allosteric transition to the open channel state (Dwoskin and Crooks 2001; Wyllie and Chen 2007). Classical competitive nAChRs antagonist are, for example, dihydro-β-erythroidine, MLA, and α-Bgtx. NCAs either physically block the ion channel or act via allosteric mechanisms (Arias et al. 2006). Channel blockers are NCAs, that bind to the transmembrane domain of nAChRs and prevent ion flux by sterically occluding the channel pore. Many agents act as nAChR channel blockers (Tassonyi et al. 2002).

13 Review of the Literature

2.2.2 Allosteric modulation

Allosteric modulators are compounds that are able to modulate nAChR function by binding to sites distinct from the orthosteric binding site (Chatzidaki and Millar 2015). Positive allosteric modulators (PAMs) potentiate the effects of agonist activation with no . There are several proposed mechanisms of action: PAMs may reduce the energy barrier between closed and open conformations or increase the energy barrier between open and desensitized states, leading to increased agonist efficacy (Bertrand and Gopalakrishnan 2007; Williams et al. 2011; Cecchini and Changeux 2015). Some PAMs may potentiate activation of nAChRs at low agonist concentrations, whereas at high agonist concentrations their effect is weak. PAMs can be divided into type I and type II modulators based on their functional properties (Bertrand and Gopalakrishnan 2007). Type I modulators enhance agonist-induced nAChR activation without affecting desensitization kinetics, while type II modulators enhance agonist-induced nAChR activation by stabilizing the open-channel conformation and slowing down desensitization. Allosteric agonists are ligands that can induce nAChR activation in the absence of an orthosteric agonist (Chatzidaki and Millar 2015). Allosteric agonists may structurally resemble type II PAMs, with only a minor change in the ligand structure (Gill et al. 2012; Gill-Thind et al. 2015).

Negative allosteric modulators (NAMs) inhibit agonist-induced activation without binding to the orthosteric binding site and are thus NCAs of nAChRs (Arias 2010; Chatzidaki and Millar 2015). NAMs inhibit agonist-induced activation either by stabilizing a non-conducting conformational state of the nAChR or by increasing the nAChR desensitization rate. The binding sites of NAMs differ from those of NCAs, although these binding sites are similarly located within the ion channel or at the extracellular- transmembrane interface. For instance, the ethidium binding site is localized at the extracellular portion of the receptor and the quinacrine binding site is localized within the transmembrane domain (Arias 1998; Pratt et al. 2000).

Silent allosteric modulators (SAMs) are compounds that do not have a positive or negative modulatory effect on responses evoked by an orthosteric ligand, but they can block the effect of other allosteric modulators (Chatzidaki and Millar 2015; Gill-Thind et al. 2015). SAMs bind to the same allosteric site as other allosteric modulators or to an overlapping allosteric site. The chemical structure of SAMs is similar to PAMs and NAMs, with only a small change in the structure (e.g., a methyl substitution of a single aromatic ring).

2.3 The connection between opioid ligands, nAChRs and nicotine

2.3.1 Mechanism of nicotine dependence

Nicotine is considered to be the main addictive chemical in tobacco smoke (Benowitz 2009; Le Foll and Goldberg 2009). Following a cigarette puff, nicotine enters the bloodstream from the lungs and crosses the blood–brain barrier, reaching the brain within 10–20 seconds (Le Houezec 2003; Tutka et al. 2005). Tobacco is described as a chronic disorder with compulsive drug-seeking and drug-taking behavior (McLellan et al. 2000). Although most smokers want to quit smoking, only a small percentage succeeds. Chronic tobacco use induces adaptive changes in the CNS, such as desensitization and

14 Review of the Literature upregulation of nAChRs, which could contribute to the development of drug dependence. In addition to nicotine, tobacco smoke contains several hundred other chemical substances, some of which may enhance the addictive and reinforcing effects of nicotine (Fowler et al. 1996a; Fowler et al. 1996b).

The basis of nicotine addiction is a combination of positive reinforcement and avoidance of the negative consequences (i.e., withdrawal symptoms), in addition to conditioning. Tobacco smoking and nicotine produce feelings of pleasure and reward and reduce stress and (De Biasi and Dani 2011). Some of the effects of smoking can be positive, such as improved concentration, better reaction time, and better performance at certain tasks (Levin et al. 2006). The withdrawal symptoms, on the other hand, are negative: irritability, , restlessness, anxiety, difficulty concentrating, and craving (McLaughlin et al. 2015). Nicotine withdrawal often causes anhedonia, inability to experience pleasure from activities usually found enjoyable (Cook et al. 2015). The withdrawal symptoms, mood disorders, anhedonia and tobacco craving are thought to result from a relative deficiency in DA release.

Activation of nAChRs by nicotine releases several in the brain, most importantly DA, in a similar manner as other drugs of abuse (Imperato et al. 1986; Di Chiara and Imperato 1988). Nicotine induces DA release in multiple brain areas (e.g., the mesolimbic area, corpus striatum, and frontal cortex). The mesolimbic pathway, connecting the VTA to the NAc, is particularly important for the rewarding effects of nicotine (Dani and De Biasi 2001; Nestler 2005). Acute nicotine administration increases brain reward function by releasing DA, and if this release is blocked by, for instance, lesioning DA neurons in vivo, nicotine self-administration is reduced (Corrigall et al. 1992). In addition to DA release, other neurotransmitters, such as ACh, serotonin, GABA, , glutamate and endorphins, also mediate the effects of nicotine. Nicotine-induced release of neurotransmitters is both direct and indirect, the latter being the primary mode of action (Wonnacott 1997). Nicotine enhances glutamatergic inputs and inhibits the GABAergic inputs to the DA neurons in the VTA, which leads to a net increase in excitation of the DA neurons and augmented DA release (Mansvelder and McGehee 2002). Tobacco smoke contains components that inhibit brain (MAO-A) and B (MAO-B) activity, which can potentiate nicotine's addictive effects by increasing the levels of monoamine neurotransmitters, such as DA and norepinephrine (Yu and Boulton 1987; Berlin and Anthenelli 2001; Lewis et al. 2007). Nicotine self-administration studies in rats indicate that MAO inhibition (particularly MAO-A inhibition) increases the rewarding effect of low doses of nicotine, possibly via dopaminergic and mechanisms (Villegier et al. 2007; Villegier et al. 2011; Smith et al. 2016).

Chronic nicotine administration leads to adaptive mechanisms, such as desensitization and upregulation of nAChRs (Wonnacott 1990; Wang and Sun 2005). Desensitization has been suggested to be one of the mechanisms contributing to nicotine dependence and tolerance (Picciotto et al. 2008). Typical daily cigarette smoking leads to nearly complete occupancy of the α4β2 nAChRs, indicating that nAChR saturation and desensitization is maintained throughout the day (Brody 2006). During longer periods of nicotine abstinence (e.g., during sleep or smoking cessation), nicotine levels drop and some of the desensitized nAChRs recover to a responsive state, which is suggested to be the cause of nicotine withdrawal and craving symptoms (Dani and Heinemann 1996). It is speculated that smoking maintains α4β2 nAChRs in a desensitized state so that these negative symptoms are avoided.

In addition to maintaining plasma nicotine levels high enough to prevent the withdrawal symptoms, smokers may also continue smoking because of the conditioned reinforcers associated with smoking

15 Review of the Literature

(Chiamulera 2005; Le Foll and Goldberg 2005a; Conklin 2006; Stoker and Markou 2015). Conditioning is the association between smoking cues and anticipated drug effects which results in the urge to smoke. The cues can be, for example, environmental situations, social interactions or the taste or feel of smoke, which becomes repeatedly associated with the pleasurable effects of smoking. Conditioned cues can also be unpleasant experiences associated with tobacco abstinence, such as irritability provoked by not smoking (Baker et al. 2004). Smokers may even come to perceive irritability from any source, such as stress, as a cue for smoking after repeated experiences (Perkins and Grobe 1992; Childs and de Wit 2010). The high rate of relapse observed in smokers wanting to quit is most likely related to conditioning factors, since smoking cues are long-lasting and resistant to interventions. Exposure to smoking-related cues in nicotine-deprived smokers activate both brain reward and attention circuits (Due et al. 2002), whereas cues associated with nicotine withdrawal decrease brain reward function (Kenny and Markou 2005). The effect of environmental stimuli on the reinforcing effects of nicotine can be studied in vivo by, for instance, intravenous nicotine self-administration and conditioned place preference (CPP) procedures (Goldberg and Henningfield 1988; Tzschentke 1998; Caggiula et al. 2002; Le Foll and Goldberg 2005b).

2.3.2 Opioid receptors and their signaling

The classical opioid receptors expressed in the CNS are the μ-, δ- and κ-receptors (Waldhoer et al. 2004; Trescot et al. 2008). Although the is genetically related to classical opioid receptors, it is not classed as one since it does not bind the same ligands and has rather a modulatory role in μ-opioid receptor-mediated actions (Toll et al. 2016). Opioid receptors are expressed primarily in the cortex, brain stem and limbic system in the brain (Le Merrer et al. 2009). The classical opioid receptors belong to the class A γ-subgroup of the G protein-coupled receptor (GPCR) superfamily (Fredriksson et al. 2003; Katritch et al. 2013). The GPCRs all have a similar structure with seven transmembrane domains. GPCRs associate with heterotrimeric G-proteins composed of three different subunits: α, β, and ɣ. Upon opioid receptor activation, guanosine-5'-triphosphate (GTP) is converted to guanosine diphosphate (GDP) and the G protein dissociates into active Gα and Gβɣ subunits (Stein 2016). The Gα subunit inhibits adenylyl cyclase and reduces the levels of adenosine 3’, 5’-cyclic monophosphate (cAMP), and the Gβɣ subunit interacts with different ion channels expressed on the cell membrane. The opioid receptors modulate presynaptic Ca2+ channels by suppressing the influx of Ca2+, lowering the excitability of neurons (Tedford and Zamponi 2006). Opioid receptor activation also prevents neuronal excitation and the propagation of action potentials by opening G protein-coupled inwardly-rectifying K+ channels (Luscher and Slesinger 2010; Nockemann et al. 2013). Furthermore, opioid agonists inhibit Na+ channels, VOCCs, transient receptor potential vanilloid-1 and acid-sensing ion channels (Gold and Levine 1996; Endres- Becker et al. 2007; Cai et al. 2014).

Opioid receptor activation commonly prevents the elevation of free intracellular calcium concentration 2+ 2+ ([Ca ]i) by inhibiting VOCCs, yet opioid receptor activation can also increase [Ca ]i (Samways and 2+ Henderson 2006). The opioid receptor-mediated increase in [Ca ]i usually requires simultaneous 2+ activation of Gq-coupled receptors, which leads to Ca release from intracellular stores via the inositol phosphate pathway (Connor and Henderson 1996; Werry et al. 2003). Nevertheless, opioid agonists have 2+ also been reported to be able to increase [Ca ]i without concomitant Gq-coupled receptor activation (Okajima et al. 1993; Allouche et al. 1996; Spencer et al. 1997; Thorlin et al. 1998). Opioids have also been 2+ 2+ shown to increase [Ca ]i by stimulating Ca entry across the plasma membrane in some cell types,

16 Review of the Literature possibly via L-type Ca2+ channels (Jin et al. 1992; Bao et al. 2003). Furthermore, opioids may increase Ca2+ influx into cells by increasing Ca2+ flux through ligand-gated cation channels, such as the P2X receptors (Chizhmakov et al. 2005).

2.3.3 Opioid dependence and pharmacological treatment options

Opioid addiction results from repeated, long-lasting exposure to opioids which leads to changes in the mesolimbic dopaminergic system, increased tolerance and to receptor desensitization and downregulation. In particular, DA is implicated in the establishment of reward in opioid addiction and also in the rewarding effects of other drugs of abuse (Nutt et al. 2015). Opioids increase the firing rate of DA neurons and thereby increase the DA levels in the NAc (Di Chiara and Imperato 1988; Volkow and Morales 2015). Another neurotransmitter particularly involved in opioid addiction is norepinephrine, which is implicated in motivating drug-seeking behaviors and establishing drug–environment pairings necessary for CPP (Weinshenker and Schroeder 2007). Other neurotransmitters, such as GABA, serotonin, opioid peptides and glutamate, have been implicated indirectly in the acute reinforcing properties and may cooperate with the DA system or work via independent pathways of reinforcement (Le Merrer et al. 2009; Kranz et al. 2010).

The pharmacological treatment options for opioid addiction can generally be divided into abstinence- oriented treatments and opioid maintenance (Lobmaier et al. 2010). The abstinence-oriented treatments involve a preliminary phase where the withdrawal symptoms are eliminated or reduced by pharmacological treatments, after which, abstinence is further supported (Gowing and Ali 2006). The most common treatments used in opioid detoxification are tapered methadone, other opioid agonists, , , and buprenorphine (Gish et al. 2010; Meader 2010; Ducharme et al. 2012; Amato et al. 2013). Opioid maintenance is an option for treatment if the abstinence-oriented approaches have failed. The purpose is not to achieve a drug-free state, but to replace the illicit opioid with a safer option. Methadone is the most widely used opioid in maintenance therapy, but buprenorphine is also increasingly used (Bell 2014). The naloxone is used in the treatment of and for opioid dependence combined with buprenorphine, and naltrexone is in use as an extended- release injection for opioid dependence (Gerra et al. 2006; van Dorp et al. 2007; Syed and Keating 2013).

2.3.4 Methadone

Methadone is a μ-opioid receptor agonist with low affinity for δ- and κ-opioid receptors (Kristensen et al. 1995). Similar to other opioid agonists, methadone is in clinical use for treating moderate to severe pain (Bieter and Hirsch 1948; Fredheim et al. 2008), but it has also been used since the 1960s for the maintenance replacement treatment of opioid addiction (Kramer 1970; Sim 1973; Bell 2014). Smoking rates are unusually high among patients with approximately 77–97% of them being smokers (Nahvi et al. 2006; Richter et al. 2007; Elkader et al. 2009; Pajusco et al. 2012; Chisolm et al. 2013). As the methadone dose increases, the satisfaction experienced from smoking rises, which leads to increased smoking (Chait and Griffiths 1984). Heavy smokers also use higher doses of methadone (Frosch et al. 2000). Methadone users have more severe nicotine dependence than regular smokers (Clarke et al. 2001), and they are less successful at smoking cessation (Stein et al. 2006; Okoli et al. 2010).

17 Review of the Literature

Methadone is commonly used in maintenance treatment for heroin addiction, and heroin use is also linked to increased smoking rates (Mello et al. 1980). Heroin addicts smoke more cigarettes when heroin self-administration is allowed than during drug-free periods. Quitting smoking appears to be more difficult than quitting opioid use for addicts (Story and Stark 1991).

At the receptor level, methadone has been shown to be an agonist of human α7 nAChRs (Pakkanen et al. 2005). Methadone inhibits [3H]MLA and [3H]epibatidine ([3H]EPI) from binding to nAChRs. Both optical 2+ isomers of methadone increase [Ca ]i in SH-SY5Y and SH-EP1-hα7 cell lines and evoke nAChR-mediated inward currents in patch-clamp studies in SH-SY5Y cells. Furthermore, nAChR antagonists are able to block these methadone-induced responses. Methadone exposure for 3 days increases [3H]EPI binding in the SH-SY5Y cell line. In addition to interaction with the α7 nAChRs, both of methadone also have non-competitive antagonistic activity at α3β4 nAChRs in rats (Xiao et al. 2001).

2.3.5 Morphine

Morphine is a classical μ-receptor agonist used as an (Wolff et al. 1940; Pasternak and Pan 2013). Morphine‘s effect on nAChRs has not been studied on the receptor level, yet in vivo studies have revealed the interplay between morphine and nicotine dependence. Both nicotinic and opioid antagonists reduce CPP to both morphine and nicotine in mice (Zarrindast et al. 2003). Furthermore, the reinstatement of morphine-induced CPP is diminished by pretreatment with specific α4β2 and α7 nAChR- subtype antagonists, dihydroxy-β-erythroidine and MLA, which suggests that these nAChR subtypes may contribute to the reinstatement of morphine-induced CPP (Feng et al. 2011). Nicotine is able to reduce naloxone-induced withdrawal symptoms in morphine-tolerant mice (Zarrindast and Farzin 1996). The effect of morphine on locomotor activity and reinforcement is further enhanced by chronic nicotine administration (Vihavainen et al. 2006; Vihavainen et al. 2008). The α3β4* nAChRs are involved in the mediation of on morphine, since mice with increased expression of α5, α3 and β4 nAChR subunits exhibit enhanced somatic signs of morphine withdrawal, and the blockade of these nAChRs attenuates morphine withdrawal symptoms (Muldoon et al. 2014).

2.3.6 Other opioids

Buprenorphine is a of the μ-opioid receptor, antagonist of the κ-opioid receptor and also a partial agonist of the (Cowan et al. 1977; Leander 1987; Kamei et al. 1995; Bloms- Funke et al. 2000). Buprenorphine does not have the same intrinsic activity as full μ-opioid agonists, such as heroin or methadone, which makes it safer, compared to full agonists, with respect to respiratory depression and overdose (Kimber et al. 2015). Buprenorphine has a long duration of action and a relatively low addiction potential (Jasinski et al. 1978). Nevertheless, buprenorphine is increasingly misused (Lofwall and Walsh 2014) and it is the most common intravenously abused opioid in Finland (Simojoki and Alho 2013). Clinically, buprenorphine is used as an analgesic and for the treatment of opioid addiction (Johnson et al. 2005; Vadivelu and Anwar 2010; Bell 2014). Smoking rates are increased during opioid substitution therapy with buprenorphine in a similar manner as in methadone maintenance therapy (Mello et al. 1985; Mutschler et al. 2002; Pajusco et al. 2012), and buprenorphine detoxification

18 Review of the Literature treatment reduces smoking (Patrick et al. 2014). Smoking cessation treatment with has shown to be ineffective among buprenorphine-maintained patients (Mooney et al. 2008).

Opioid antagonists naloxone and naltrexone are clinically used for the treatment of opioid overdose, for opioid addiction combined with buprenorphine and for dependence (Fudala et al. 1998; Bart 2012; Zindel and Kranzler 2014; Noble et al. 2015). The effect of naloxone and naltrexone in smoking cessation is controversial (David et al. 2013). In vitro, naloxone and naltrexone inhibit nAChRs (Madsen and Albuquerque 1985; Almeida et al. 2000; Almeida et al. 2004) and naltrexone blocks nicotine-induced upregulation of α4β2 nAChRs (Almeida et al. 2000).

Dextromethorphan is structurally related to opioid agonists and commonly used as an antitussive (Brown et al. 2004; Taylor et al. 2016). is a NCA at α3β4, α4β2 and α7 nAChRs (Hernandez et al. 2000; Damaj et al. 2005) and has been shown to reduce nicotine self-administration in rats (Glick et al. 2001; Briggs et al. 2016). Another antitussive agent with an opioid structure, codeine, is a PAM of α4β2 and α7 nAChRs (Storch et al. 1995; Iorga et al. 2006). Tramadol, a weak μ-receptor agonist, used as an analgesic, has been shown to inhibit the function of α7 nAChRs (Lewis and Han 1997; Shiraishi et al. 2002). Furthermore, tramadol use increases the severity of nicotine dependence in tramadol addicts dose-dependently, and the increase in nicotine dependence seems to have the same effect on tramadol intake (Shalaby et al. 2015).

19 Aims of the Study

3. AIMS OF THE STUDY

The purpose of this study was to find out whether opioid ligands have any effect on nAChRs at the receptor level. This question was approached by studying the receptor binding and functional effect of opioid ligands, both after acute drug treatment and after prolonged incubation. More specifically, the aims were:

x To determine, whether opioid agonists, such as morphine and methadone, bind to nAChRs expressed in SH-SY5Y, SH-EP1-hα4β2 and SH-EP1-hα7 cells (I, II).

x To investigate the acute functional effects of methadone and morphine on nAChRs expressed in SH-SY5Y, SH-EP1-hα4β2 and SH-EP1-hα7 cells (I, II). x To investigate whether buprenorphine, codeine, oxycodone, tramadol, naloxone and naltrexone bind to nAChRs expressed in SH-SY5Y, SH-EP1-hα4β2 and SH-EP1-hα7 cells and whether they have any effect on nAChR function (III, unpublished results). x To characterize the effects of prolonged methadone and morphine treatments on nAChR numbers and function (IV).

20 Materials and Methods

4. MATERIALS AND METHODS

4.1 Drugs and reagents

The (–)-nicotine tartrate, bovine serum albumin, Meca hydrochloride, MLA citrate salt hydrate, naltrexone hydrochloride, Bradford reagent, poly-d-, poly(ethyleneimine) solution, cytisine and carbamylcholine chloride were purchased from Sigma-Aldrich (St. Louis, MO, USA). The bovine serum albumin standards were from Thermo Fisher Scientific Inc. (Rockford, IL, USA). The rubidium-86 isotope (86Rb+) and [3H]EPI radioligand were purchased from PerkinElmer (Waltham, MA, USA). The nicotine ditartrate dehydrate was from Acros Organics (Geel, Belgium). Morphine hydrochloride was purchased from University Pharmacy (Helsinki, Finland). The (±)-methadone hydrochloride was purchased from Star (Tampere, Finland) and from University Pharmacy (Helsinki, Finland). Cell culture plasticware was purchased from Nunc (Roskilde, Denmark) and cell culture media and media supplements and Fluo-3AM were from Invitrogen (Carlsbad, CA, USA). Potassium phosphate buffer was composed of 50 mM K2HPO4,

50 mM KH2PO4, 1 mM ethylenenediaminetetra-acetic acid (EDTA) and the pH was adjusted to 7.4.

Phosphate-buffered saline (PBS) was composed of 150 mM NaCl, 8 mM K2HPO4, 2 mM KH2PO4, and the 86 + pH was adjusted to 7.4. Rb efflux buffer was composed of 130 mM NaCl, 5.4 mM KCl, 2 mM CaCl2, 5 mM glucose and 50 mM HEPES and the pH was adjusted to 7.4. Tyrode’s salt solution (TSS) was composed of 137 mM NaCl, 2.7 mM KCl, 1 mM MgCl2, 1.8 mM CaCl2, 0.2 mM NaH2CO3 and 5.5 mM glucose, pH adjusted to 7.4.

4.2 Cell culture

The cell lines used were SH-SY5Y, SH-EP1-hα4β2 and SH-EP1-hα7 cells. The SH-SY5Y cell line is a neuroblastoma cell line of human origin that is thrice cloned from the SK-N-SH line (Ross et al. 1983; Gould et al. 1992; Lukas et al. 1993; Peng et al. 1997; Groot Kormelink and Luyten 1997). SH-SY5Y cells natively express α3, α5, α7 and β2, β3, β4 subunit cDNAs, and these subunits form functional pentameric receptors of different combinations, such as α3β4 and α7 nAChRs. In addition to nAChRs, the SH-SY5Y cells also express functional muscarinic and μ- and δ-opioid receptors (Adem et al. 1987; Kazmi and Mishra 1987; Kukkonen et al. 1992).

The SH-EP1 cell line has no native expression of nAChRs, but expresses the μ- and δ-opioid receptors (Baumhaker et al. 1993). SH-EP1-hα4β2 and SH-EP1-hα7 cells were transfected with human cDNAs coding for α4 and β2 subunits or α7 subunit, respectively, and thus express these human nAChR subtypes exclusively (Peng et al. 1999; Zhao et al. 2003; Eaton et al. 2003). Both of these cell lines were kindly provided by Dr. Ronald J. Lukas (Barrow Neurological Institute, St Joseph's Hospital and Medical Center, Phoenix, AZ, USA). The SH-EP1-hα4β2 cell line was constructed as described by Eaton et al. (2003). In brief, cDNAs encoding human α4 and β2 subunits we used to generate the pcDNA3.1/zeo-hα4 and pcDNA3.1/hygro-hβ2 constructs. Native nAChR-null SH-EP1 cells were simultaneously transfected with both α4 and β2 constructs using electroporation. The culture medium was supplemented with selection antibiotics zeocin and hygromycin B in order to sort out the cells expressing dual drug resistance. Ring cloning was used to isolate single, resistant cell colonies, which were then expanded. Functional screening was done with the 86Rb+ efflux assay, and a clone exhibiting high expression of α4β2 nAChRs

21 Materials and Methods was further subcloned by dilution and the ring-cloning method. The SH-EP1-hα7 cells were constructed as described in Zhao et al. (2003). Briefly, the SH-EP1 cells were transfected using electroporation with cDNA encoding the human α7 subunit. Hygromycin was used as the selection antibiotic, and positive transfectants were isolated by ring cloning. The expression of α7 nAChRs was confirmed by radiolabeled α-Bgtx binding.

All cell cultures were maintained in 5% CO2/humidified air at 37 °C, the cells were split twice weekly and passage numbers 10–30 were used in experiments. SH-SY5Y cells were grown in a mixed (1:1) Dulbecco’s modified Eagle:Ham’s F12 (DMEM/Ham F12) medium supplemented with fetal bovine serum (12.7%), penicillin and streptomycin (1.7%) and non-essential amino acids (0.8%). SH-EP1-hα7 cells were grown in DMEM (with high glucose, l- and sodium pyruvate) supplemented with horse serum (8.5%), fetal bovine serum (4.3%), penicillin and streptomycin (0.9%), amphotericin B (0.7%) and hygromycin (0.7%), and the same mixture supplemented with zeocin (0.3%) was used for the SH-EP1-hα4β2 cells.

4.3 [3H]Epibatidine binding

Epibatidine is an that was extracted from the skin of Epipedobates tricolor, an endemic Ecuadorian frog (Daly et al. 1978). Epibatidine is a potent analgesic and its analgesic action is not blocked by the opioid antagonist naloxone, whereas the nAChR antagonist Meca is able to antagonize its analgesic effect (Qian et al. 1993; Li et al. 1993; Badio and Daly 1994). Both isomers of epibatidine have an equal affinity for nAChRs, whereas their affinities for muscarinic, opioid, serotonin, or GABA receptors are weak or nonexistent (Badio and Daly 1994). Although epibatidine is known to be 200 times more potent than morphine in terms of its analgesic effect, its toxicity makes it unsuitable for clinical use. Doses not much higher than what are needed for antinociception cause , respiratory paralysis, seizures and death in vivo (Sullivan et al. 1994; Bonhaus et al. 1995).

Epibatidine‘s high affinity for nAChRs makes it an excellent ligand for studying nAChRs. [3H]EPI binds with high affinity to α2, α3, α4, β2 and β4 subunit-containing nAChRs expressed in X. laevis oocytes or HEK 293 cells (Parker et al. 1998; Xiao and Kellar 2004) and to α6β2* and α6β4* nAChRs expressed in X. laevis oocytes (Kuryatov et al. 2000). In mouse brain, [3H]EPI binding is biphasic by nature meaning that there are binding sites with both high affinity (equilibrium binding constant (Kd) ≈ 0.02 nM) and low affinity (Kd ≈ 5 nM)(Gerzanich et al. 1995; Marks et al. 2006). Deletion of α7, β2, and β4 nAChR subunits eliminates virtually all [3H]EPI binding sites in mouse brain (Marks et al. 2006). The lower affinity binding sites are α- Bgtx-sensitive, α7-containing receptors, and α-Bgtx-resistant receptors, which contain either β2 or β4 subunits. The high affinity binding sites can be divided into cytisine-sensitive sites, which require the β2 subunit, and to cytisine-resistant sites, which require the presence of either a β2 or β4 subunit.

[3H]EPI binding can be used to assess the site of action for ligands interacting with nAChRs (Lukas et al., 2002). Additionally, radio-labeled agonist binding can be used to measure the change in receptor numbers (e.g., upregulation). Ligands that act as competitive antagonists or agonists of nAChRs inhibit the binding of [3H]EPI to these receptors. NCAs have either no effect on [3H]EPI binding or inhibit it only at very high concentrations. In order to measure the affinities of different nAChR ligands by [3H]EPI binding, the number of binding sites must be assessed by saturation binding. The amount of specific

22 Materials and Methods binding is determined by measuring both total and nonspecific binding (the binding of [3H]EPI to something other than its designated receptors) and subtracting the nonspecific binding from the total.

[3H]EPI binding was used for saturation binding studies and for competition binding. For the acute nicotine and opioid studies, the cells were grown in 175 cm2 flasks. The medium was aspirated at confluency, followed by three washes with ice-cold PBS. Cells were mechanically harvested into potassium phosphate buffer and homogenized using an ultrasonic homogenizer (Ringo Ultrasonics Bio 70, Romanshorn, Switzerland) (75% amplitude, 2×15 s) on ice. Homogenates were next centrifuged at 13,000g for 20 min at 4 °C. The supernatant was discarded and the pellet was resuspended in potassium phosphate buffer and frozen at −70 °C. The frozen cell homogenates were thawed and re-homogenized by sonication, and protein concentrations were measured by the Bradford method. [3H]EPI binding was done in 96-well filter plates (Millipore MultiScreen HTS FC) in a final volume of 250 μl with samples containing 5–50 μg of protein. For saturation binding studies, the cell homogenates were incubated with 25–5000 pM [3H]EPI for 2 h at room temperature. In competition binding assays, the cell membranes were incubated with 150 pM [3H]EPI and serial dilutions of nicotinic and/or opioid ligands. Nicotine (1 mM) was used to determine the amount of nonspecific binding. Optiphase HiSafe 3 scintillant (Perkin Elmer, Turku, Finland) was added to the wells and radioactivity was quantified by scintillation counting (1450 MicroBeta TriLux Liquid Scintillation Counter & Luminometer, PerkinElmer, Waltham, MA, USA).

For prolonged nicotine and opioid studies, cells grown in 175 cm2 flasks were treated (final concentration) with either nicotine (1 and 10 μM), methadone (1 and 10 μM) or morphine (0.1 and 1 μM). The studied drugs were added to cell culture medium and flasks were maintained at 37 °C in 5% CO2/humidified air for 3 days. For the control group, cells were incubated with normal cell culture medium for 3 days. On the day of assay, cultures were rinsed for 3×10 min with warm medium; then incubated for 3 h at 37°C before a final wash with PBS to remove all traces of drugs from the cultures. After that, the cell homogenates were prepared as described above. Cell membranes were incubated with 25–6400 pM [3H]EPI; otherwise the protocol was the same as described above.

4.4 Calcium fluorometry

Cells were grown in 75 cm2 flasks, from which the cells were seeded into 96-well BD Falcon microplates (in a 100 μl volume) coated with poly-d-lysine the day before the experiment. Culture medium was removed from 96-well plates and the cells were washed twice with TSS. The loading medium was composed of Fluo-3 AM (10 μM) and probenecid (2.5 mM) in TSS. Probenecid was prepared fresh every day as a stock concentration of 250 mM solubilized 1:1 in 1.0 M NaOH and TSS. Cells were incubated with the loading medium (50 μl/well) at room temperature for 1 h in the dark. After incubation, cells were washed twice with TSS (200 μl/well). After washings, 100 μl of TSS with or without antagonists was added, and the plate was transferred to a FlexStation microplate reader (Molecular Devices, Sunnyvale, CA, USA). If pretreatments were used, the cells were preincubated with drugs for 10 min in the dark. The test drug was then injected (50 μl/well), and the fluorescence was monitored for 80 s at 485/525 nM wavelength. The magnitudes of the drug-induced signals were calculated by subtracting the background signal (measured for 10 s before injection) from the drug-induced signal. The cells were also stimulated 2+ by injecting vehicle (TSS) as a control in each experiment. The changes in [Ca ]i were expressed as a percentage of nicotine-evoked increase in fluorescence.

23 Materials and Methods

4.5 86Rb+ efflux nAChR function can be studied with the 86Rb+ efflux assay, which is a specific and highly sensitive ion flux assay, where the 86Rb+ acts as a tracer for potassium movement across the cell membrane (Lukas and Cullen 1988; Lukas et al., 2002). When nAChRs are stimulated, the 86Rb+ ions flow out of the cell down a concentration gradient and the amount of outflow ions can be measured using Cerenkov counting. The 86Rb+ ions are loaded by the Na+/K+-ATPase into cells by incubating them with cell medium containing a fixed amount of 86Rb+. The 86Rb+ has a short half-life (18.66 days), thus, the amount of 86Rb+ in the loading medium must be adjusted to an adequate level before every experiment. The “flip-plate” method, where cells are seeded to 24-well plates and liquid changes are handled by flipping the plates, offers gentle sample handling, high throughput, good temporal accuracy and reproducibility. In this method, the rinse buffer and ligands being studied are also pipetted into 24-well plates and the liquid changes are handled by aligning the plates on top of each other so that liquid from all 24 original wells flows to the other 24 wells almost simultaneously. Liquid aspiration is handled by a multichannel aspiration manifold, so that several wells can be handled at the same time.

For the 86Rb+ efflux assays, the cells were grown in 75 cm2 flasks, which were maintained at 37 °C in 5%

CO2/humidified air. For the prolonged drug treatments, the cell culture medium was supplemented with drugs being studied (nicotine, methadone or morphine) and flasks were maintained in the incubator for 3 days. Cells were detached from flasks with trypsin-EDTA and seeded in a 500 μl volume to 24-well plates. Plates were placed in a 37 °C incubator for at least four hours to let the cells adhere. Next, the culture medium was replaced with medium (250 μl per well) supplemented with 250,000–300,000 cpm of 86Rb+ and the drugs being studied for the prolonged drug treatments. A day after seeding the cells, the 86Rb+ efflux was measured with the “flip-plate” method (Lukas et al., 2002). Each well was rinsed once with 1.2 ml of 86Rb+ efflux buffer for 1 min to remove extracellular 86Rb+, except for prolonged studies where each well was rinsed twice with buffer for altogether 10 min, in order to remove any traces of drugs. Cells were exposed to efflux buffer containing the ligands studied for 5 min. Six of the wells on each plate were reserved for controls: three for maximum 86Rb+ efflux (1 mM carbamylcholine) and three for non-specific 86Rb+ efflux (buffer only). Radioactivity was counted with Cerenkov counting after inserting the cross-talk-minimizing inserts into wells. Specific efflux in each well was obtained by subtracting the nonspecific efflux from the total efflux. Typical values for specific maximum 86Rb+ efflux (depending on cell type and density and the concentration of 86Rb+ in the loading medium) were 25,000−50,000 cpm over the non-specific backgrounds of 6,000–15,000 cpm. The amount of remaining intracellular 86Rb+ was counted to ensure that the sum of 86Rb+ released into the efflux plate and 86Rb+ remaining in the cell plate were the same for each well. To determine the amount of remaining intracellular 86Rb+, 0.1% sodium dodecyl sulfate/0.1 M NaOH (1.5 ml per well) was added to lyse the cells and the plates were counted with Cerenkov counting. The 86Rb+ efflux assays could not be conducted with the SH-EP1-hα7 cell line because α7 nAChRs are rapidly inactivating channels that are not open long enough to give significant ion flux signal (Lukas et al., 2002).

4.6 Statistical analysis

All data were analyzed using GraphPad Prism (version 5.02 in study I, version 6.02 in studies II and IV);

GraphPad Software, Inc., La Jolla, CA, USA). Briefly, the Kd and maximum specific binding (Bmax) values

24 Materials and Methods from saturation binding experiments in studies I and II were determined by non-linear regression by fitting the data points to a one-site or two-site ligand binding model. The equilibrium dissociation constant (Ki) values in the competition binding assay were calculated using one-site and two-site curve- fitting models, and the Ki values were obtained from best fit (p-value < 0.05, extra sum-of-squares F test). The calcium responses were calculated as the percentage of the increase in fluorescence produced by a maximally effective concentration of nicotine (10 μM for SH-EP1-hα4β2 and SH-EP1-hα7 cells and 30 μM for SH-SY5Y cells). The data were analyzed statistically using one-way analysis of variance (ANOVA) and Dunnett's multiple-comparison post hoc test. Specific 86Rb+ efflux signal was defined as total 86Rb+ efflux (1 mM carbamylcholine) minus non-specific 86Rb+ efflux (buffer) and drug-induced 86Rb+ efflux was normalized to specific efflux after subtraction of nonspecific efflux. The specific 86Rb+ efflux was fit to the variable slope model or the biphasic concentration–response model after comparing these two different fits (p-value < 0.05, extra sum-of-squares F test). EC50 values were determined using the biphasic concentration–response model, and concentration inhibiting 50% of maximum (IC50) values were determined using the variable slope model.

In study IV, the Bmax and Kd values for the saturation binding assays were calculated using one-site specific binding with Hill slope model. Changes in the Bmax values were analyzed using one-way ANOVA and Dunnett's multiple comparisons test. Values of p < 0.05 were taken to be statistically significant. In the 86 + Rb efflux assay, the EC50 and logEC50 values were determined using the biphasic concentration- response model after comparing it to the variable slope model. The biphasic concentration-response model was a better fit (p-value < 0.05, extra sum-of-squares F test).

25 Results

5. RESULTS

5.1 Methadone

5.1.1 Effects of acute methadone treatment (II)

3 Methadone displaced [ H]EPI in SH-SY5Y and SH-EP1-hα7 cells, but not in SH-EP1-hα4β2 cells. Ki and logKi values are presented in Table 1. In SH-SY5Y cells, methadone displaced [3H]EPI with high and low affinity, 2+ and the fraction of high-affinity binding sites was 49 ± 20%. Methadone significantly increased [Ca ]i in SH-SY5Y and SH-EP1-hα7-cells, but not in SH-EP1-hα4β2 cells. In SH-EP1-hα4β2 cells, pre-incubation with 2+ methadone (10 and 50 μM) inhibited the nicotine (10 μM) -induced increase in [Ca ]i by 72 ± 6% and 85 ± 14%, respectively. In SH-SY5Y cells, pre-incubation with methadone (10 and 50 μM) inhibited the 2+ nicotine (30 μM) -induced increase in [Ca ]i by 90 ± 11% and 90 ± 7%, respectively. In SH-EP1-hα7 cells, 2+ pre-incubation with methadone (10 and 50 μM) inhibited the nicotine (10 μM) -induced increase in [Ca ]i by 65 ± 12% and 74 ± 17%, respectively. In SH-EP1-hα4β2 cells, pre-incubation with Meca (50 μM) 2+ inhibited the methadone (100 μM) -induced increase in [Ca ]i by 43 ± 9%. MLA, naloxone and naltrexone had no effect on the methadone-induced response. In SH-SY5Y cells, pre-incubation with Meca (100 μM), MLA (50 μM), naloxone (100 μM) and naltrexone (100 μM) inhibited the methadone (100 μM) -induced 2+ response in [Ca ]i by 36 ± 4%, 53 ± 10%, 40 ± 10% and 37 ± 8%, respectively. In SH-EP1-hα7 cells, pre- incubation with Meca (50 μM), MLA (100 μM), naloxone (50 μM) and naltrexone (100 μM) inhibited the 2+ methadone (100 μM) -induced increase in [Ca ]i by 30 ± 7%, 30 ± 7%, 34 ± 9% and 32 ± 8%, respectively.

Methadone had no effect on the 86Rb+ efflux in SH-EP1-hα4β2 and SH-SY5Y cells. However, methadone did inhibit nicotine-induced 86Rb+ efflux in SH-EP1-hα4β2 and SH-SY5Y cells in a concentration-dependent manner. To determine whether this antagonism is competitive or non-competitive, nicotine concentration–response assays were also performed in the presence of several different methadone concentrations, all of which inhibited the nicotine-induced 86Rb+ efflux non-competitively in both cell lines studied. IC50 and logIC50 values are presented in Table 2.

26 Results

3 Table 1. Equilibrium dissociation constant (Ki and logKi) values for opioid ligands in [ H]EPI competition binding in the SH-EP1-hα4β2, SH-SY5Y and SH-EP1-hα7 cell lines.

SH-EP1-hα4β2 SH-SY5Y SH-EP1-hα7

KiHi,KiLow Compound Ki (μM) logKi ±S D logKiHi ± SD, logKiLow ± SD Ki (μM) logKi ± SD (μM) Methadone >3000 NA 19.4, 1008 -4.71±0.28, -3.00±0.50 6.3 -5.20±0.06 Morphine 13.2 -4.88±0.03 0.2, 126 -6.80±0.42, -3.90±0.12 43.7 -4.36±0.06 Buprenorphine >3000 NA 1234 -2.098±0.45 14.6 -4.84±0.14 Codeine 366.9 -3.44±0.05 165.5 -3.78±0.08 - - Oxycodone 296.3 -3.53±0.03 105.5 -3.98±0.06 - - Tramadol >3000 NA >3000 NA - - Naloxone >3000 NA 0.3, 44.5 -6.47±0.59, -4.35±0.13 10.4 -4.98±0.07 Naltrexone >3000 NA 3.9, 131.8 -5.41±0.36, -3.88±0.18 5.3 -5.27±0.07

KiHi/KiLow, the equilibrium dissociation constant value for the high-/low-affinity binding sites; logKiHi /logKiLow, the log equilibrium dissociation constant values for the high-/low-affinity binding sites; SD, standard deviation; NA, not applicable; -, not measured.

86 + Table 2. IC50 and logIC50 values for opioid ligands in Rb efflux assay in the SH-EP1-hα4β2 and SH-SY5Y cell lines.

SH-EP1-hα4β2 SH-SY5Y Nicotine 10 μM Nicotine 1 μM Nicotine 30 μM Nicotine 10 μM Compound IC50 IC50 IC50 IC50 logIC50 ± SD logIC50 ± SD logIC50 ± SD logIC50 ± SD (μM) (μM) (μM) (μM) Methadone 7.4 -5.13±0.03 5.6 -5.25±0.08 0.8 -6.08±0.03 1.0 -6.02±0.05 Morphine NA NA NA NA 342 -3.47±0.06 189 -3.72±0.06 Buprenorphine 68.3 -4.17±0.10 198 -3.70±0.15 17.9 -4.75±0.03 16.4 -4.79±0.04 Codeine 1721 -2.76±0.42 NA NA 278 -3.56±0.08 154 -3.81±0.10 Oxycodone 247 -3.61±0.05 189 -3.72±0.19 164 -3.79±0.06 105 -3.98±0.08 Tramadol 64.7 -4.19±0.03 58.7 -4.23±0.06 9.4 -5.03±0.04 17.1 -4.77±0.04 Naloxone 104.2 -3.98±0.03 21.9 -4.66±0.07 73.9 -4.13±0.03 50.4 -4.30±0.05 Naltrexone 1866 -2.73±0.09 564 -3.25± 0.09 99.0 -4.00±0.05 62.9 -4.20±0.05

IC50, the half-maximal inhibitory concentration; logIC50, the log half-maximal inhibitory concentration; SD, standard deviation.

5.1.2 Effects of prolonged methadone treatment (IV)

Prolonged exposure to methadone (1 and 10 μM) increased the [3H]EPI binding sites in SH-SY5Y (Fig. 7B) and SH-EP1-hα7 (Fig. 7C) cells, but decreased [3H]EPI binding sites in SH-EP1-hα4β2 cells (Fig. 7A).

27 Results

Figure 7. [3H]EPI saturation binding profiles in SH-EP1-hα4β2 (A), SH-SY5Y (B) and SH-EP1-hα7 (C) cell membranes. The cells were exposed for 72 h to either normal cell culture medium or to cell culture medium containing 1 or 10 μM methadone. The [3H]EPI saturation binding curve for no pre-treatment is presented in each picture as a control. The assay mixtures contained 25-6400 pM [3H]EPI and 5-50 μg cell protein. The data were fit to a one-site competition binding model, after comparing it with the two-sites model (the better fit was determined using the extra sum-of- squares F test, p-value < 0.05). The values are the mean ± standard deviation (SD) from five to six assays, with each point assayed in triplicate at least.

In the 86Rb+ efflux assay, nicotine showed biphasic full agonist activity in the SH-EP1-hα4β2 cell line with

EC50 values of 0.1 μM and 1.3 μM (Fig. 8A). The EC50 values for nicotine were 0.08 μM and 5.7 μM when pre-treated with 1 μM methadone for 4 days, and 0.07 μM and 1.6 μM with 10 μM methadone pre- treatment. Nicotine also showed biphasic full agonist activity in the SH-SY5Y cell line, with EC50 values of 2.5 μM and 11.9 μM (Fig. 8B). When SH-SY5Y cells were pre-treated with 1 μM methadone for 4 days, the EC50 values for nicotine were 3.2 μM and 10.5 μM, and with 10 μM methadone the EC50 values were 1.4 and 7.5 μM.

Figure 8. Dose–response profiles of nAChR function for nicotine using an 86Rb+ efflux assay in SH-EP1-hα4β2 (A) and SH-SY5Y (B) cells. The specific 86Rb+ efflux was determined in the presence of nicotine and the cells were pretreated with 1 or 10 μM methadone for 96 h. The specific efflux for each drug concentration was assessed using 1 mM carbamylcholine (total efflux) and efflux buffer (non-specific efflux) controls. The cells were exposed to nicotine for 5 min. The data were fit to the biphasic concentration–response equation. The values are the mean ± SD from six separate assays.

28 Results

5.2 Morphine

5.2.1 Effects of acute morphine treatment (I)

3 Morphine displaced [ H]EPI from all cell lines studied. The Ki and logKi values are presented in Table 1. 2+ Morphine was also able to increase [Ca ]i in all cell lines, and in SH-EP1-hα7 cells the effect was comparable to that of nicotine. The effect of morphine on the nicotine-induced response in SH-EP1- hα4β2 cells was statistically significant only at 100 μM, inhibiting 74% of the nicotine-induced response. In SH-SY5Y cells, morphine (10 μM and 50 μM) preincubation increased the nicotine-induced response by 70% and 49%, respectively. In SH-EP1-hα7 cells, morphine (50 μM) preincubation increased the nicotine-induced response by 61%. In the 86Rb+ efflux assay, concentration–response profiles showed partial agonist efficacy for morphine in SH-EP1-hα4β2 cells with an EC50 value of 53.3 μM. Meca concentration-dependently attenuated the morphine-induced 86Rb+ efflux response. Opioid antagonists naloxone and naltrexone modified the morphine-induced 86Rb+ efflux response by shifting the curve to the right. Morphine enhanced the 86Rb+ ion efflux induced by low nicotine concentrations. In SH-SY5Y cells, morphine showed no efficacy in the 86Rb+ efflux assay. Morphine had inhibitory efficacy on nicotine- induced 86Rb+ ion flux. In order to clarify the mechanism of nAChR antagonism or agonism, the 86Rb+ efflux nicotine concentration–response studies were also performed in the presence of multiple morphine concentrations. In SH-EP1-hα4β2 cells, morphine increased the nicotine-stimulated response at low nicotine concentrations and decreased the nicotine-stimulated response at high nicotine concentrations. In SH-SY5Y cells, morphine had virtually no effect on the nicotine-stimulated response.

5.2.2 Effects of prolonged morphine treatment (IV)

Prolonged exposure to morphine (0.1 and 1 μM) increased [3H]EPI binding sites in SH-SY5Y (Fig 9B) and SH-EP1-hα7 (Fig 9C) cells, but decreased [3H]EPI binding sites in SH-EP1-hα4β2 cells (Fig 9A).

Figure 9. [3H]EPI saturation binding profiles in SH-EP1-hα4β2 (A), SH-SY5Y (B) and SH-EP1-hα7 (C) cell membranes. The cells were exposed for 72 h to either normal cell culture medium or to cell culture medium containing 0.1 or 1 μM morphine. The [3H]EPI saturation binding curve without any pre-treatment is presented in each picture as a control. The assay mixtures contained 25-6400 pM [3H]EPI and 5-50 μg cell protein. The data were fit to a one-site competition binding model, after comparing it with the two-sites model (the better fit was determined using the extra sum-of- squares F test, p-value < 0.05). The values are the mean ± SD from five to six assays, with each point assayed in triplicate at least.

29 Results

86 + In the Rb efflux assay, nicotine had biphasic full agonist activity with EC50 values of 0.1 μM and 1.3 μM in SH-EP1-hα4β2 cells (Fig. 10A). The EC50 values for nicotine were 1.3 μM and 21.4 μM when pre-treated with 0.1 μM morphine for 4 days, and 0.2 μM and 7.8 μM with 1 μM morphine pre-treatment. Nicotine also showed biphasic full agonist activity in the SH-SY5Y cell line, with EC50 values of 2.5 μM and 11.9 μM

(Fig 10B). After pre-treatment with 0.1 μM morphine for 4 days the EC50 values for nicotine were 2.2 μM and 10.5 μM, and with 1 μM morphine the EC50 values were 2.8 and 11.8 μM.

Figure 10. Dose–response profiles of nAChR function for nicotine using an 86Rb+ efflux assay in SH-EP1-hα4β2 (A) and SH-SY5Y (B) cells. The specific 86Rb+ efflux was determined in the presence of nicotine and the cells were pretreated with 0.1 or 1 μM morphine for 96 h. The specific efflux for each drug concentration was assessed using 1 mM carbamylcholine (total efflux) and efflux buffer (non-specific efflux) controls. The cells were exposed to nicotine for 5 min. The data were fit to the biphasic dose–response equation. The values are the mean ± SD from six separate assays.

5.3 Buprenorphine (unpublished)

5.3.1 Effects of acute buprenorphine treatment

Buprenorphine displaced [3H]EPI in SH-EP1-hα7 cells, but not in SH-EP1-hα4β2 or SH-SY5Y cells (Fig. 11).

The Ki and logKi values are presented in Table 1. Preincubation with buprenorphine inhibited the nicotine- 2+ induced increase in [Ca ]i in SH-SY5Y, SH-EP1-hα4β2 and SH-EP1-hα7 cells (Fig. 12). In SH-SY5Y cells, buprenorphine (10 μM, 50 μM and 100 μM) preincubation decreased the nicotine-induced response by 68%, 89% and 99%, respectively. The effect of buprenorphine on the nicotine-induced response in SH- EP1-hα4β2 and SH-EP1-hα7 cells was statistically significant only at 50 μM and 100 μM, inhibiting 75% and 82% of the nicotine-induced response in SH-EP1-hα4β2 cells and 59% and 70% in SH-EP1-hα7 cells.

30 Results

Figure 11. [3H]EPI competition binding profiles in SH-EP1-hα4β2 (A), SH-SY5Y (B) and SH-EP1-hα7 (C) cell membranes for nicotine and buprenorphine. The assay mixtures contained 150 pM [3H]EPI and 5–50 μg cell protein and nicotine or buprenorphine at increasing concentrations. The data were fit to a one-site competition binding model. The values are the mean ± SD from three to five assays, with each point assayed at least in triplicate. Ki and logKi values are provided in Table 1.

2+ Figure 12. Effects of buprenorphine on nicotine-induced increases in [Ca ]i in SH-SY5Y, SH-EP1-hα4β2 and SH-EP1- hα7 cells. The Fluo-3 AM-loaded cells were incubated for 10 min with vehicle (TSS) or 10 μM, 50 μM or 100 μM buprenorphine and then stimulated with 10 μM or 30 μM nicotine. The responses are expressed as the percentage of the nicotine response. The values are the mean ± standard error of mean (SEM) from three to six separate assays. In each experiment, there were at least six replicates for each condition. Significantly different from nicotine response **p < 0.01, and ***p < 0.001 using one-way ANOVA.

2+ Buprenorphine (100 μM) induced increase in [Ca ]i in all cell lines studied (data not shown). Preincubation with nAChR partial agonist cytisine or with opioid antagonist naltrexone inhibited the 2+ buprenorphine-induced increase in [Ca ]i in SH-EP1-hα4β2, SH-SY5Y, and SH-EP1-hα7 cells (Fig. 13). 2+ Opioid antagonist naloxone was able to inhibit the buprenorphine-induced increase in [Ca ]i in SH-EP1- hα4β2 and SH-SY5Y cells.

31 Results

2+ Figure 13. Effects of cytisine and opioid antagonists on buprenorphine-evoked increases in [Ca ]i in SH-EP1-hα4β2 (A), SH-SY5Y (B) and SH-EP1-hα7 (C) cells. The Fluo-3 AM-loaded cells were incubated for 10 min with vehicle (TSS), cytisine (Cyt), naloxone (Nalo) or naltrexone (Nalt) and then stimulated with 100 μM buprenorphine. The responses are expressed as the percentage of the buprenorphine response. The values are the mean ± SEM from three to six separate assays. In each experiment, there were at least six replicates for each condition. Significantly different from buprenorphine response *p < 0.05, **p < 0.01, using one-way ANOVA.

In the 86Rb+ efflux assay, buprenorphine showed no efficacy in either SH-EP1-hα4β2 or SH-SY5Y cells (data not shown). Buprenorphine had inhibitory efficacy on nicotine-induced 86Rb+ ion flux in both cell lines (Fig. 14A, C). To determine the mechanism of nAChR antagonism, the 86Rb+ efflux nicotine dose–response studies were also performed in the presence of multiple buprenorphine concentrations (Fig. 14B, D). In both cell lines, buprenorphine decreased the nicotine-stimulated response non-competitively. The IC50 and logIC50 values are presented in Table 2.

32 Results

Figure 14. 86Rb+ efflux dose–response profiles of nAChR function for buprenorphine in SH-EP1-hα4β2 (A) and SH- SY5Y (C) cells and the mechanisms of antagonism of nAChR function in SH-EP1-hα4β2 (B) and SH-SY5Y (D) cells. The specific 86Rb+ efflux was determined in the presence of various concentrations of nicotine and increasing concentrations of buprenorphine in SH-EP1-hα4β2 (A) and SH-SY5Y (C) cells. To determine the mechanism of antagonism the nicotine- stimulated specific 86Rb+ efflux was determined in the absence or presence of various concentrations of buprenorphine in SH-EP1-hα4β2 (B) and SH-SY5Y (D) cells. The cells were exposed to the ligands for 5 min. The data were fit to the variable slope model. The values are the mean ± SD from three to six separate assays. The IC50 and logIC50 values are provided in Table 2.

5.4 Codeine (unpublished)

5.4.1 Effects of acute codeine treatment

3 Codeine only displaced [ H]EPI at high concentrations in SH-EP1-hα4β2 and SH-SY5Y cells (Fig. 15). The Ki 86 + and logKi values are presented in Table 1. In the Rb efflux assay, codeine showed no efficacy in either SH-EP1-hα4β2 or SH-SY5Y cells (data not shown). Codeine had inhibitory efficacy on nicotine-induced 86Rb+ ion flux in SH-SY5Y cells (Fig. 16C) and this inhibition was non-competitive by nature (Fig. 16D). The

IC50 and logIC50 values are presented in Table 2. In SH-EP1-hα4β2 cells, codeine had a potentiative effect on nicotine-induced 86Rb+ efflux at low nicotine concentrations (Fig. 16A, B).

33 Results

Figure 15. [3H]EPI competition binding profiles in SH-EP1-hα4β2 (A) and SH-SY5Y (B) cell membranes for nicotine and codeine. The assay mixtures contained 150 pM [3H]EPI, 5–50 μg cell protein and nicotine or codeine at increasing concentrations. The data were fit to a one-site competition binding model. The values are the mean ± SD from three to five assays, with each point assayed at least in triplicate. Ki and logKi values are provided in Table 1.

Figure 16. 86Rb+ efflux dose–response profiles of nAChR function for codeine in SH-EP1-hα4β2 (A) and SH-SY5Y (C) cells and the mechanisms of antagonism or agonism of nAChR function in SH-EP1-hα4β2 (B) and SH-SY5Y (D) cells. The specific 86Rb+ efflux was determined in the presence of various concentrations of nicotine and increasing concentrations of codeine in SH-EP1-hα4β2 (A) and SH-SY5Y (C) cells. To determine the mechanism of antagonism or agonism, the nicotine-stimulated specific 86Rb+ efflux was determined in the absence or presence of various concentrations of codeine in SH-EP1-hα4β2 (B) and SH-SY5Y (D) cells. The cells were exposed to the ligands for 5 min.

The data were fit to the variable slope model. The values are the mean ± SD from three to six separate assays. The IC50 and logIC50 values are provided in Table 2.

34 Results

5.5 Oxycodone (unpublished)

5.5.1 Effects of acute oxycodone treatment

Oxycodone only displaced [3H]EPI at high concentrations in SH-EP1-hα4β2 and SH-SY5Y cells (Fig. 17). The 86 + Ki and logKi values are presented in Table 1. In the Rb efflux assay, oxycodone showed no agonist efficacy in either SH-EP1-hα4β2 or SH-SY5Y cells (data not shown). Oxycodone had inhibitory efficacy on nicotine-induced 86Rb+ ion flux in both cell lines (Fig. 18A, C) and this inhibition was non-competitive by nature in the SH-SY5Y cells and mixed competitive/noncompetitive antagonism in the SH-EP1-hα4β2 cells, since oxycodone increased the EC50 value for nicotine, yet the surmountability of the blockage was not evident (Fig. 18B, D). The IC50 and logIC50 values are presented in Table 2.

Figure 17. [3H]EPI competition binding profiles in SH-EP1-hα4β2 (A) and SH-SY5Y (B) cell membranes for nicotine and oxycodone. The assay mixtures contained 150 pM [3H]EPI, 5–50 μg cell protein and nicotine or oxycodone at increasing concentrations. The data were fit to a one-site competition binding model. The values are the mean ± SD from three to five assays, with each point assayed at least in triplicate. Ki and logKi values are provided in Table 1.

35 Results

Figure 18. 86Rb+ efflux dose–response profiles of nAChR function for oxycodone in SH-EP1-hα4β2 (A) and SH-SY5Y (C) cells and the mechanisms of antagonism of nAChR function in SH-EP1-hα4β2 (B) and SH-SY5Y (D) cells. The specific 86Rb+ efflux was determined in the presence of various concentrations of nicotine and increasing concentrations of oxycodone in SH-EP1-hα4β2 (A) and SH-SY5Y (C) cells. To determine the mechanism of antagonism, the nicotine- stimulated specific 86Rb+ efflux was determined in the absence or presence of various concentrations of oxycodone in SH-EP1-hα4β2 (B) and SH-SY5Y (D) cells. The cells were exposed to the ligands for 5 min. The data were fit to the variable slope model. The values are the mean ± SD from three to six separate assays with each point assayed at least in triplicate. The IC50 and logIC50 values are provided in Table 2.

5.6 Tramadol (unpublished)

5.6.1 Effects of acute tramadol treatment

Tramadol did not displace [3H]EPI in either SH-EP1-hα4β2 or SH-SY5Y cells (Fig. 19). In the 86Rb+ efflux assay, tramadol showed no efficacy in either SH-EP1-hα4β2 or SH-SY5Y cells (data not shown). Tramadol had inhibitory efficacy on nicotine-induced 86Rb+ ion flux in both cell lines (Fig. 20A, C) and this inhibition was non-competitive by nature (Fig. 20B, D). The IC50 and logIC50 values are presented in Table 2.

36 Results

Figure 19. [3H]EPI competition binding profiles in SH-EP1-hα4β2 (A) and SH-SY5Y (B) cell membranes for nicotine and tramadol. The assay mixtures contained 150 pM [3H]EPI, 5–50 μg cell protein and nicotine or tramadol at increasing concentrations. The data were fit to a one-site competition binding model. The values are the mean ± SD from three to five assays, with each point assayed at least in triplicate.

Figure 20. 86Rb+ efflux dose–response profiles of nAChR function for tramadol in SH-EP1-hα4β2 (A) and SH-SY5Y (C) cells and the mechanisms of antagonism of nAChR function in SH-EP1-hα4β2 (B) and SH-SY5Y (D) cells. The specific 86Rb+ efflux was determined in the presence of various concentrations of nicotine and increasing concentrations of tramadol in SH-EP1-hα4β2 (A) and SH-SY5Y (C) cells. To determine the mechanism of antagonism, the nicotine- stimulated specific 86Rb+ efflux was determined in the absence or presence of various concentrations of tramadol in SH- EP1-hα4β2 (B) and SH-SY5Y (D) cells. The cells were exposed to the ligands for 5 min. The data were fit to the variable slope model. The values are the mean ± SD from three to six separate assays with each point assayed at least in triplicate. The IC50 and logIC50 values are provided in Table 2.

37 Results

5.7 Naloxone and naltrexone (unpublished)

5.7.1 Effects of acute naloxone and naltrexone treatments

Naloxone and naltrexone displaced [3H]EPI in SH-SY5Y and SH-EP1-hα7 cells, but not in SH-EP1-hα4β2 cells (Fig. 21). The Ki and logKi values are presented in Table 1. Preincubation with naloxone and 2+ naltrexone inhibited the nicotine-induced increase in [Ca ]i in SH-SY5Y and SH-EP1-hα7 cells, but not in SH-EP1-hα4β2 cells (Fig. 22).

Figure 21. [3H]EPI competition binding profiles in SH-EP1-hα4β2 (A), SH-SY5Y (B) and SH-EP1-hα7 (C) cell membranes for nicotine, naltrexone and naloxone. The assay mixtures contained 150 pM [3H]EPI, 5–50 μg cell protein and nicotine, naltrexone or naloxone at increasing concentrations. The data were fit to a one-site competition binding model. The values are the mean ± SD from three to five assays, with each point assayed at least in triplicate. Ki and logKi values are provided in Table 1.

2+ Figure 22. Effects of naloxone and naltrexone on nicotine-induced increases in [Ca ]i in SH-EP1-hα4β2, SH-SY5Y and SH-EP1-hα7 cells. The Fluo-3 AM-loaded cells were incubated for 10 min with vehicle (TSS) or 100 μM naloxone or naltrexone and then stimulated with 10 μM or 30 μM nicotine. The responses are expressed as the percentage of the nicotine response. The values are the mean ± SEM from three to six separate assays. In each experiment, there were at least six replicates for each condition. Significantly different from nicotine response: *p < 0.05, **p < 0.01, and ***p < 0.001 using one-way ANOVA.

In the 86Rb+ efflux assay, naloxone and naltrexone showed no agonist efficacy in either SH-EP1-hα4β2 or SH-SY5Y cells (data not shown). Naloxone and naltrexone had inhibitory efficacy on nicotine-induced 86Rb+ ion flux in both cell lines, and this inhibition was non-competitive by nature in the SH-SY5Y cells and

38 Results mixed competitive/noncompetitive antagonism in the SH-EP1-hα4β2 cells, since both naloxone and naltrexone increased the EC50 value for nicotine, yet the surmountability of blockage was not evident

(Fig. 23 and 24). The IC50 and logIC50 values are presented in Table 2.

Figure 23. 86Rb+ efflux dose–response profiles of nAChR function for naloxone (A) and naltrexone (C) and the mechanisms of antagonism of nAChR function (B, D) in SH-EP1-hα4β2 cells. The specific 86Rb+ efflux was determined in the presence of various concentrations of nicotine and increasing concentrations of naloxone or naltrexone in SH- EP1-hα4β2 cells. To determine the mechanism of antagonism, the nicotine-stimulated specific 86Rb+ efflux was determined in the absence or presence of various concentrations of naloxone or naltrexone in SH-EP1-hα4β2 cells. The cells were exposed to the ligands for 5 min. The data were fit to the variable slope model. The values are the mean ±

SD from three to six separate assays with each point assayed at least in triplicate. The IC50 and logIC50 values are provided in Table 2.

39 Results

Figure 24. 86Rb+ efflux dose–response profiles of nAChR function for naloxone (A) and naltrexone (C) and the mechanisms of antagonism of nAChR function (B, D) in SH-SY5Y cells. The specific 86Rb+ efflux was determined in the presence of various concentrations of nicotine and increasing concentrations of naloxone or naltrexone in SH-SY5Y cells. To determine the mechanism of antagonism, the nicotine-stimulated specific 86Rb+ efflux was determined in the absence or presence of various concentrations of naloxone or naltrexone in SH-SY5Y cells. The cells were exposed to the ligands for 5 min. The data were fit to the variable slope model. The values are the mean ± SD from three to six separate assays with each point assayed at least in triplicate. The IC50 and logIC50 values are provided in Table 2.

40 Discussion

6. DISCUSSION

6.1 Effects of opioids on α4β2 nAChRs

The most abundant nAChR subtype in the human brain, the α4β2 nAChRs, have an important role in cognitive functions such as memory, attention and learning and in mood and motor function (Grupe et al. 2015). The α4β2 nAChRs are implicated in nicotine dependence as well as various neurological diseases, such as Parkinson's disease, Alzheimer's disease, attention deficit hyperactivity disorder (ADHD), schizophrenia, depression and epilepsy (Picciotto et al. 1998; Tapper et al. 2004; Steinlein and Bertrand 2010; Jurgensen and Ferreira 2010; Quik and Wonnacott 2011; Sarter et al. 2012; Yu et al. 2014; Potter et al. 2014). In addition to substance use patients, smoking rates are especially high among people with psychiatric disorders (Besson and Forget 2016; Sharma et al. 2016). Smoking has been proposed to act as self-medication among psychiatric patients and both smoking and mental illness may share a common cause.

In our studies, only morphine was able to displace [3H]EPI from α4β2 nAChRs expressed in SH-EP1-hα4β2 cells. Morphine was also the only opioid that showed partial agonist efficacy in the 86Rb+ efflux assay, and this response was attenuated completely and non-competitively by Meca, confirming that the effect is nAChR mediated. The opioid antagonists naloxone and naltrexone inhibited the morphine-induced ion flux competitively. Morphine elevated the 86Rb+ ion flux induced by low nicotine concentrations and attenuated the ion flux with high nicotine doses. Furthermore, the nicotine- 2+ induced elevation of [Ca ]i was reduced by high morphine concentrations, whereas low morphine concentrations had a small additive effect with nicotine. These findings suggest that morphine acts as a partial agonist at α4β2 nAChRs.

All of the other opioid ligands studied, methadone, buprenorphine, codeine, oxycodone, tramadol, naloxone and naltrexone, either did not displace [3H]EPI from α4β2 nAChRs or the concentrations needed for displacement were very high. In the 86Rb+ efflux assay, only methadone inhibited nicotine-induced ion flow with adequate concentration. None of these opioid ligands showed agonistic efficacy in the 86Rb+ efflux assay. Codeine had a potentiative effect on nicotine-induced 86Rb+ efflux at low nicotine concentrations. Interestingly, the EC50 values for nicotine in the presence of naloxone, naltrexone or oxycodone were higher than in the absence of these ligands, but surmountability of blockage was not clearly evident, suggesting a mixture of competitive and noncompetitive blockage. Methadone and 2+ buprenorphine inhibited the nicotine-induced rise in [Ca ]i, whereas naloxone and naltrexone did not 2+ inhibit the nicotine-induced increase in [Ca ]i. These results suggest that methadone is an NCA at α4β2 nAChRs, that codeine has a positive modulatory effect on α4β2 nAChRs, that naloxone, naltrexone and oxycodone are mixed competitive/noncompetitive antagonists of α4β2 nAChRs, and that tramadol and buprenorphine are weak NCAs at α4β2 nAChRs. Our results support the previous findings of codeine’s positive modulatory effect on α4β2 nAChRs (Storch et al. 1995; Iorga et al. 2006).

Both prolonged methadone and morphine treatment down-regulated α4β2 nAChRs and changed nicotine’s effect on α4β2 nAChRs, whereas sustained nicotine caused substantial upregulation. The α4β2 nAChRs are upregulated by chronic nicotine as well as by other nicotinic agonists and antagonists in vitro and in vivo (Peng et al. 1994; Gopalakrishnan et al. 1997; Whiteaker et al. 1998; Hussmann et al. 2011;

41 Discussion

Marks et al. 2015). Studies with non-competitive α4β2 nAChR antagonists and partial agonists have produced mixed results; some have been shown to cause upregulation, whereas some either have no effect on receptor numbers or have a down-regulating effect (Peng et al. 1994; Whiteaker et al. 1998; Xiao et al. 2006; Hussmann et al. 2012; Marks et al. 2015). Whereas smoking upregulates nAChRs, reduced α4β2 nAChR numbers are associated with many neuropsychiatric disorders, such as Parkinson's disease and Alzheimer’s disease (Meyer et al. 2009; Kendziorra et al. 2011; Hurst et al. 2013). Furthermore, a low baseline level of α4β2* nAChRs has been suggested to be one of the predisposing factors for nicotine dependence since increased nicotine self-administration is associated with lower levels of midbrain α4β2 nAChRs (Le Foll et al. 2009). Our findings that both methadone and morphine downregulate α4β2 nAChRs and that prolonged exposure to these opioid agonists changes nicotine’s effect on α4β2 nAChRs may thus have clinical importance. The high smoking rates and difficulties in smoking cessation among patients on opioid maintenance therapy may partially be a consequence of this receptor-level interaction. If α4β2 nAChRs are down-regulated in the brains of opioid-maintained patients, they may smoke more in order to compensate the reduced effect of nicotine. The plasma concentration of methadone in maintenance patients can reach 1 μM (de Vos et al. 1995; Dyer et al. 1999) and the plasma concentration of morphine in cancer patients varies between 0.02–0.60 μM (Sakurada et al. 2010). The effects seen in our upregulation studies are likely to also be valid in the clinical setting, since the opioid concentrations used fall in the concentration ranges of 1–10 μM for methadone and 0.1–1 μM for morphine.

6.2 Effects of opioids on α3* nAChRs

The 15q25 gene cluster, which contains the CHRNA5/CHRNA3/CHRNB4 genes coding for the α5, α3 and β4 nAChRs, is linked to the risk of heavy smoking, nicotine dependence and smoking-related diseases (Bierut 2009). The role of α5, α3 and β4 nAChRs in nicotine withdrawal and aversion has also been confirmed in in vivo studies (Salas et al. 2004; Salas et al. 2009; Jackson et al. 2010; Frahm et al. 2011). The majority of α3β4* nAChRs reside in the MHb and interpeduncular nucleus, which are involved in nicotine withdrawal and intake (Quick et al. 1999; Whiteaker et al. 2002; Salas et al. 2009; Fowler and Kenny 2012).

The SH-SY5Y cells express the α3, α5, α7 and β2, β3, β4 subunits and these subunits form functional pentameric receptors of different combinations, such as the α3β4 or α7 nAChRs (Ross et al. 1983; Gould et al. 1992; Lukas et al. 1993; Peng et al. 1997; Groot Kormelink and Luyten 1997). [3H]EPI binds to all nAChR subtypes expressed in SH-SY5Y cells, however, the β2 subunit-containing receptors have a higher affinity for [3H]EPI than the β4 or α7 subunit-containing receptors (Gerzanich et al. 1995; Wang et al. 1996). In the [3H]EPI binding assay, methadone, morphine, naloxone and naltrexone displaced [3H]EPI, whereas buprenorphine, codeine, oxycodone and tramadol either did not inhibit [3H]EPI binding or the concentrations needed for inhibition were very high. Methadone, morphine, naloxone and naltrexone also displaced [3H]EPI from α7 nAChRs in the SH-EP1-hα7 cell line, so the [3H]EPI displacement identified in the SH-SY5Y cells could result from either α7 or α3* nAChRs.

In the 86Rb+ efflux assay, methadone, tramadol and buprenorphine inhibited the nicotine-induced ion flux with reasonably low IC50 values in the SH-SY5Y cells, whereas other opioids were less effective. The rank order for ion flux inhibition potency by IC50 value was methadone < tramadol < buprenorphine <

42 Discussion naloxone < naltrexone < oxycodone < codeine < morphine. Buprenorphine and methadone also inhibited 2+ 2+ the nicotine-induced increase in [Ca ]i, however the effect of tramadol, oxycodone or codeine on [Ca ]i 2+ was not studied. Methadone itself increased [Ca ]i in a concentration-dependent manner, as has been shown before (Pakkanen et al. 2005). In the same study, preincubation with an α3* nAChR antagonist, 2+ dihydro-β-erythroidine, had no effect on methadone-evoked increases in [Ca ]i in SH-SY5Y cells, whereas an α7 nAChR antagonist, MLA, partially antagonized the effects of methadone. These results suggest that 2+ the methadone-induced increase in [Ca ]i is at least partly mediated through the α7 nAChRs and that the α3* nAChRs expressed in SH-SY5Y cells do not participate in the Ca2+ influx evoked by methadone. The α7 nAChRs have especially high permeability for Ca2+ (Quik et al. 1997), and therefore the methadone- induced Ca2+ flow is also most likely mediated through α7 nAChRs. In SH-SY5Y cells, opioids require ongoing muscarinic receptor activation in order to mobilize Ca2+ from intracellular stores (Connor and 2+ Henderson 1996), and the activation of α7 nAChRs may provide the initial increase in [Ca ]i allowing the opioid-like action of methadone. The 86Rb+ efflux assay measures the function of α3* nAChRs in SH-SY5Y cells; because methadone inhibited the nicotine-induced ion flow without shifting the EC50 value for nicotine, and the inhibition curve is insurmountable, it seems that methadone is an NCA of α3* nAChRs. 86 + In addition to methadone, tramadol also inhibited the nicotine-induced Rb efflux at low IC50 values without displacing [3H]EPI in the SH-SY5Y cells, suggesting that tramadol is also an NCA of α3* nAChRs. Tramadol has previously been shown to inhibit the function of α7 and the other subclasses of nAChRs expressed in adrenal chromaffin cells, such as the α3β4 nAChRs (Shiraishi et al. 2002), and our results are in line with this study regarding the α3* nAChR inhibition.

Both prolonged methadone and morphine treatment, as well as nicotine, upregulated nAChRs in SH-SY5Y cells, as measured by [3H]EPI binding; however, this upregulation had almost no effect on nicotine- induced ion flux in the 86Rb+ efflux assay. This may be explained by the fact that the 86Rb+ efflux assay measures the function of receptors on the cell surface only, whereas the majority of the increase in [3H]EPI binding sites takes place in the intracellular compartment of SH-SY5Y cells (Peng et al. 1997). The intracellularly accumulated α3* nAChRs in SH-SY5Y cells have been shown to been previously exposed on the cell surface by an antigenic modulation technique, which suggests that chronic nicotine may induce internalization of α3* nAChRs without decomposition of these receptors. Chronic treatment with nicotine produces a 500–600% increase in α3* nAChRs in SH-SY5Y cells, but only a 30% increase in α7 nAChRs (Peng et al. 1997); consequently, the receptors upregulated by methadone and morphine treatments are most likely α3* nAChRs. Furthermore, since only the function of the α3* nAChRs can be measured by 86Rb+ efflux, the effect of possible α7 nAChR upregulation cannot be measured using this method. The upregulated nAChRs on the cell surface may also be nonfunctional and therefore no effect is seen although receptor numbers measured by [3H]EPI binding are elevated. Nevertheless, these results are in line with an earlier study where chronic nicotine treatment did not increase the affinity for nicotine despite upregulation (Peng et al. 1997). Prolonged methadone treatment has previously been shown to upregulate the nAChRs in SH-SY5Y cells (Pakkanen et al. 2005), whereas there are no studies on morphine’s effect on nAChR numbers.

6.3 Effects of opioids on α7 nAChRs

α7 nAChRs are particularly expressed in brain regions linked with cognitive function, such as the hippocampus, cortex and subcortical limbic regions as well as in the thalamus and (Marks

43 Discussion and Collins 1982; Clarke et al. 1985; Rubboli et al. 1994; Gotti et al. 2006). α7 nAChRs are implicated in the pathophysiology and pathogenesis of many diseases and disorders such as schizophrenia, autism, Alzheimer's disease and Down syndrome (Young and Geyer 2013; Deutsch et al. 2015; Shen and Wu 2015). The expression of α7 nAChRs is reduced in postmortem brains of schizophrenic patients (Freedman et al. 1995; Court et al. 1999). Most schizophrenics are heavy smokers, and smoking increases the expression of CHRNA7 mRNA and protein levels, raising them up to the levels of healthy controls (Dalack et al. 1998; de Leon and Diaz 2005; Mexal et al. 2010). α7 nAChR protein expression and binding are elevated in the cerebellum of autistic patients (Lee et al. 2002; Martin-Ruiz et al. 2004). Losses in α7 nAChR binding sites have also been reported in several brain regions of Alzheimer's disease patients (Burghaus et al. 2000).

Methadone, morphine, buprenorphine, naloxone and naltrexone displaced [3H]EPI from the α7 nAChRs expressed in SH-EP1-hα7 cells, with methadone and naltrexone having the lowest Ki values. Also, all of 2+ these opioid ligands studied, except for morphine, inhibited the nicotine-induced increase in [Ca ]i. These results suggest that buprenorphine, naloxone and naltrexone are weak α7 antagonists. Naltrexone has previously been shown to be a non-competitive α7 nAChR antagonist and to block the nicotine- induced upregulation of α7 nAChRs (Almeida et al. 2000; Almeida et al. 2004). Morphine (50 μM) 2+ preincubation had an additive effect on the nicotine-induced increase in [Ca ]i and morphine itself also 2+ increased [Ca ]i in a similar manner as nicotine. This could result from a synergetic effect of morphine and nicotine acting through their own specific receptors or due to interaction at nAChRs via a positive allosteric mechanism. Morphine‘s effect on nAChRs has not been studied before at the receptor level; however, in vivo pretreatment with specific α7 nAChR antagonist, MLA, diminished the reinstatement of morphine-induced CPP, which suggests that α7 nAChRs may contribute to the reinstatement of morphine-induced CPP (Feng et al. 2011).

An earlier in vitro study suggested that methadone is an agonist of human α7 nAChRs (Pakkanen et al. 2005), and our results are in line with this study. Methadone inhibits [3H]EPI binding in SH-EP1-hα7 cells as well as in the SH-SY5Y cell line, which expresses the α3, α5, α7, β2 and β4 nAChR subunits. Unfortunately, the 86Rb+ efflux assay is not suitable for studying α7 nAChR function due to the very rapid desensitization kinetics (Lukas et al., 2002). Functional studies, however, can be conducted with calcium fluorometry since the α7 nAChRs have a high permeability to Ca2+ (Castro and Albuquerque 1995). 2+ Methadone increased the [Ca ]i levels in SH-EP1-hα7 cells in a similar manner as nicotine, which supports the theory of methadone as an α7 nAChR agonist. When SH-EP1-hα7 cells were pre-incubated with 2+ methadone, the nicotine-induced increase in [Ca ]i was reduced, presumably as a result of the desensitization of α7 nAChRs. Nicotinic ligands MLA and Meca, as well as opioid antagonists naloxone 2+ and naltrexone, inhibited the methadone-induced increase in [Ca ]i levels in SH-EP1-hα7 cells. This inhibition could result from inhibition of both the α7 nAChRs and opioid receptors expressed in SH-EP1- hα7 cells, since naloxone and naltrexone appear to be weak antagonists at the α7 nAChRs.

Both prolonged methadone and morphine treatment upregulated α7 nAChRs measured by [3H]EPI binding. α7 nAChRs are upregulated by chronic nicotine exposure to a lesser extent than α4β2 nAChRs, and the nicotine concentration required for the upregulation might be higher than what is attained in smokers’ blood (Fenster et al. 1997; Ke et al. 1998; Alkondon et al. 2000). Both chronic and antagonist treatments upregulate α7 nAChRs and may enhance receptor function (Molinari et al. 1998; Ridley et al. 2001; Nuutinen et al. 2006).

44 Discussion

6.4 General discussion

Nicotine appears to interact with several opioid ligands, and this interplay seems to be at least in part mediated through their pharmacological activity at nAChRs. The main findings of this receptor level interaction are summarized in Table 3. The high prevalence of smoking with opioid use may also be a consequence of interactions at the level of DA, since nicotine has the ability to increase extracellular DA levels in mesolimbic brain regions, likewise for opioids and other drugs of abuse. Nicotine exposure alters the reinforcing and rewarding effects of opioids and vice versa; the use of other drugs may enhance the pleasurable effect of smoking (Henningfield and Griffiths 1981; Chait and Griffiths 1984; Amos et al. 2004). In addition to opioids, nicotine is also commonly used with other substances of abuse, such as , alcohol and (Burling and Ziff 1988; Roll et al. 1997; Viveros et al. 2006). Nicotine has, for instance, been shown to alleviate opioid and alcohol withdrawal (Elkader et al. 2009; Perez et al. 2015), to enhance the effects of stimulants (Wiseman and McMillan 1998), and to enhance opioid tolerance and antinociception (Zarrindast et al. 1997; Schmidt et al. 2001; Zarrindast et al. 2003).

Table 3. The main findings of the receptor-level effects of opioids studied in SH-EP1-hα4β2, SH-SY5Y and SH-EP1- hα7 cell lines.

Compound SH-EP1-hα4β2 SH-SY5Y SH-EP1-hα7 Methadone NCA NCA agonist Morphine partial agonist weak NCA PAM? Buprenorphine weak NCA weak NCA weak antagonist Codeine PAM weak NCA - Oxycodone mixed CA/NCA weak NCA - Tramadol weak NCA NCA - Naloxone mixed CA/NCA weak NCA weak antagonist Naltrexone mixed CA/NCA weak NCA weak antagonist NCA, non-competitive antagonist; CA, competitive antagonist; PAM, positive allosteric modulator; -, not studied;

Smoking increases health risks not only in the general population, but to an even greater magnitude among patients on opioid maintenance therapy. Opioid users are at a greater risk of death due to respiratory diseases, cancer, and cardiovascular diseases, and the relative risk of death due to tobacco- related diseases is 1.8 times higher among opioid abusers than in the general population (Hurt et al. 1996; Maxwell et al. 2005). Smoking cessation is challenging for opioid maintenance patients, yet the majority of them are motivated to quit (Shoptaw et al. 2002; Nahvi et al. 2006). Smoking cessation would be very beneficial for opioid-maintained patients, since success in smoking cessation is also associated with greater abstinence from opioid use (Lemon et al. 2003). Unfortunately, the current pharmacotherapeutic options for smoking cessation, bupropion, varenicline and nicotine replacement therapy, are ineffective among opioid-substituted patients (Okoli et al. 2010; Zirakzadeh et al. 2013; Nahvi et al. 2014). The nicotinic–opioid interaction is a built-in property of varenicline, since its chemical structure is derived from morphine and cytisine (Coe et al. 2005; Rollema et al. 2007). Varenicline is a α4β2 and α6β2* nAChR partial agonist and a full agonist at α3β4* and α7 nAChRs (Coe et al. 2005; Grady et al. 2010), upregulating α4β2, α3β4*, α7 nAChRs and down-regulating α6β2* nAChRs (Turner et al. 2011; Marks et al. 2015). Both methadone and varenicline seem to regulate nAChR numbers

45 Discussion differentially, which might explain why combined use of these substances results in low smoking abstinence rates among methadone users. If α4β2 nAChRs are down-regulated in the brains of methadone-maintained patients, the effect of varenicline might be submaximal, and lead to lower smoking cessation rates. Patients on high methadone doses are more severely addicted to nicotine, and their withdrawal symptoms and nicotine cravings are higher compared to the general population (Clarke et al. 2001; Elkader et al. 2009). In our studies, chronic methadone down-regulated the α4β2 nAChRs and reduced nicotine’s potency and efficacy on α4β2 nAChRs; thus, it can be speculated that methadone- maintained patients smoke more to overcome this receptor-level interaction of nicotine and methadone. Even though a direct comparison of in vitro and human data might not be relevant, our findings support the hypothesis of nicotine–opioid interactions at the level of nAChRs, which should be taken into account in the planning of smoking cessation treatments and pharmacotherapy for opioid addicts. Polysubstance abuse represents a unique challenge for treatment schemes, requiring a more precise understanding of the behavioral and biochemical background of the nicotine–opioid interplay.

46 Summary and Conclusions

7. SUMMARY AND CONCLUSIONS

Although nicotine and opioids have both their own molecular mechanisms of action and designated receptors, it seems that these substances interact at the level of nAChRs. The aims of these studies were to investigate in vitro the effect of various opioid agonists and antagonists on nAChR function and whether these opioid ligands bind to the orthosteric binding site of nAChRs. Additionally, the effect of prolonged opioid treatment on nAChR numbers and function was studied with methadone and morphine. The major findings are as follows:

1. Morphine acts as a partial agonist and codeine as a PAM at α4β2 nAChRs. Methadone is a NCA at α4β2 nAChRs, whereas tramadol and buprenorphine are weak NCAs. Naloxone, naltrexone and oxycodone are mixed competitive/noncompetitive antagonists of α4β2 nAChRs.

2. All of the opioid ligands studied are NCAs of α3* nAChRs: methadone and tramadol have

the lowest IC50 values, and morphine and codeine are the weakest inhibitors.

3. Methadone is an agonist of human α7 nAChRs, whereas buprenorphine, naloxone and naltrexone are weak α7 antagonists. Morphine has a positive synergistic effect with nicotine on α7 nAChR function.

4. α3*, α4β2 and α7 nAChRs are differentially regulated by prolonged exposure to methadone and morphine. Methadone and morphine upregulate α3* and α7 nAChRs, whereas α4β2 nAChRs are down-regulated. The down-regulation of α4β2 nAChRs changes nicotine’s effect on these receptors.

47 Acknowledgements

8. ACKNOWLEDGEMENTS

The work described in this thesis has been carried out at the Division of Pharmacology and Pharmacotherapy, Faculty of Pharmacy, University of Helsinki during the years 2009-2016. I wish to express my gratitude to the following people:

First, I would like to express my sincere gratitude to my supervisors Docent Outi Salminen and Professor and Head of the Division of Pharmacology and Pharmacotherapy, Raimo Tuominen for their continuous support of my Ph.D. research and studies. I appreciate that you always found time for advising me despite your demanding and busy schedules. I thank you for your patience, motivation, and immense knowledge. Without your guidance and encouragement, this work would not have been completed.

I am deeply grateful to the reviewers of my thesis: Professor Jyrki Kukkonen and Docent Petri Hyytiä, for their thorough investigation of my thesis and valuable comments and suggestions which improved my thesis substantially. I would like to thank Professor Neil Millar for agreeing to act as my opponent of this dissertation. I would like to thank the committee of my research plan defense: Professor Jyrki Kukkonen, Docent Sanna Janhunen and Docent Ilkka Reenilä for their insightful comments and questions which encouraged me to broaden my research from various perspectives. I am also thankful to Professor Jyrki Kukkonen for allowing me to use his lab facilities in the Faculty of Veterinary Medicine.

My sincere thanks also goes to Dr. Ronald J. Lukas and Dr. Paul Whiteaker who provided me an opportunity to visit their research facility and gave me access to work at their laboratory in the Barrow Neurological Institute in Phoenix. I would also like to thank them for the cell lines they provided for my experiments. I am deeply grateful to J. Brek Eaton for all of the technical instructions and methods he taught me, as well as for his kindness, friendship and generosity. Without his support, it would not have been possible to conduct this research. I would also like to thank Syndia Marxer and all of the other members of Lukas-Whiteaker Lab for the hospitality they showed me.

I am grateful to Kati Rautio, Marjo Vaha and Anna Niemi for their excellent technical assistance. I would like to thank Mari Havia and Juri Meijer for conducting some of the experiments presented in this thesis as a part of their Pro Gradu thesis projects under my supervision. I thank all my colleagues at the Division of Pharmacology and Pharmacotherapy for the pleasant working environment and all your support. In particular, I am grateful for the friendship of Iida Peltonen, Susanne Bäck, Virpi Talman, Marjo Piltonen and Nadia Schendzielorz.

I wish to thank the personnel of the Arabianranta pharmacy for the warm and supportive atmosphere. The practical, customer-oriented work with your delightful company has kept me sane throughout these years.

I am very grateful to my friends for their support, friendship and all the fun we have had and all the fun yet to come. I would especially like to thank Johanna Vesterinen, Anna Ikonen, and Anne Lehtinen. I am very lucky to have friends like you in my life.

48 Acknowledgements

My sincere thanks goes to my father Jukka and late mother Tuulikki and my brother Ismo for all their support throughout my life. I know my mother would have wanted to be there for me the day I defend my thesis and would have been so proud of me.

Finally, I owe my deepest and dearest gratitude to my beloved family: my children, Aapo and Pinja, and husband Tommi. Thank you for reminding me of what truly matters in life.

Espoo, April 2017

Reeta

49 References

9. REFERENCES

Adem A, Mattsson ME, Nordberg A, Pahlman S (1987) Muscarinic receptors in human SH-SY5Y neuroblastoma cell line: regulation by phorbol ester and retinoic acid-induced differentiation. Brain Res 430: 235-242.

Albuquerque EX, Alkondon M, Pereira EF, Castro NG, Schrattenholz A, Barbosa CT, Bonfante-Cabarcas R, Aracava Y, Eisenberg HM, Maelicke A (1997) Properties of neuronal nicotinic acetylcholine receptors: pharmacological characterization and modulation of synaptic function. J Pharmacol Exp Ther 280: 1117- 1136.

Albuquerque EX, Pereira EF, Alkondon M, Rogers SW (2009) Mammalian nicotinic acetylcholine receptors: from structure to function. Physiol Rev 89: 73-120.

Albuquerque EX, Pereira EF, Mike A, Eisenberg HM, Maelicke A, Alkondon M (2000) Neuronal nicotinic receptors in synaptic functions in humans and rats: physiological and clinical relevance. Behav Brain Res 113: 131-141.

Alkondon M, Pereira EF, Almeida LE, Randall WR, Albuquerque EX (2000) Nicotine at concentrations found in cigarette smokers activates and desensitizes nicotinic acetylcholine receptors in CA1 interneurons of rat hippocampus. Neuropharmacology 39: 2726-2739.

Allouche S, Polastron J, Jauzac P (1996) The delta-opioid receptor regulates activity of ryanodine receptors in the human neuroblastoma cell line SK-N-BE. J Neurochem 67: 2461-2470.

Almeida LE, Pereira EF, Alkondon M, Fawcett WP, Randall WR, Albuquerque EX (2000) The opioid antagonist naltrexone inhibits activity and alters expression of alpha7 and alpha4beta2 nicotinic receptors in hippocampal neurons: implications for smoking cessation programs. Neuropharmacology 39: 2740-2755.

Almeida LE, Pereira EF, Camara AL, Maelicke A, Albuquerque EX (2004) Sensitivity of neuronal nicotinic acetylcholine receptors to the antagonists naltrexone and naloxone: receptor blockade and up- regulation. Bioorg Med Chem Lett 14: 1879-1887.

Amato L, Davoli M, Minozzi S, Ferroni E, Ali R, Ferri M (2013) Methadone at tapered doses for the management of . Cochrane Database Syst Rev (2):CD003409. doi: CD003409.

Amos A, Wiltshire S, Bostock Y, Haw S, McNeill A (2004) 'You can't go without a fag...you need it for your hash'--a qualitative exploration of smoking, cannabis and young people. Addiction 99: 77-81.

Arias HR (2010) Positive and negative modulation of nicotinic receptors. Adv Protein Chem Struct Biol 80: 153-203.

Arias HR (1998) Binding sites for exogenous and endogenous non-competitive inhibitors of the nicotinic acetylcholine receptor. Biochim Biophys Acta 1376: 173-220.

Arias HR, Bhumireddy P, Bouzat C (2006) Molecular mechanisms and binding site locations for noncompetitive antagonists of nicotinic acetylcholine receptors. Int J Biochem Cell Biol 38: 1254-1276.

Badio B, Daly JW (1994) Epibatidine, a potent analgetic and nicotinic agonist. Mol Pharmacol 45: 563- 569.

Baker TB, Brandon TH, Chassin L (2004) Motivational influences on cigarette smoking. Annu Rev Psychol 55: 463-491.

Baler RD, Volkow ND (2006) Drug addiction: the neurobiology of disrupted self-control. Trends Mol Med 12: 559-566.

50 References

Bao L, Jin SX, Zhang C, Wang LH, Xu ZZ, Zhang FX, Wang LC, Ning FS, Cai HJ, Guan JS, Xiao HS, Xu ZQ, He C, Hokfelt T, Zhou Z, Zhang X (2003) Activation of delta opioid receptors induces receptor insertion and neuropeptide secretion. Neuron 37: 121-133.

Bart G (2012) Maintenance medication for opiate addiction: the foundation of recovery. J Addict Dis 31: 207-225.

Bartos M, Corradi J, Bouzat C (2009) Structural basis of activation of cys-loop receptors: the extracellular- transmembrane interface as a coupling region. Mol Neurobiol 40: 236-252.

Baumhaker Y, Wollman Y, Goldstein MN, Sarne Y (1993) Evidence for mu-, delta-, and kappa-opioid receptors in a human neuroblastoma cell line. Life Sci 52: PL205-10.

Bell J (2014) Pharmacological maintenance treatments of opiate addiction. Br J Clin Pharmacol 77: 253- 263.

Bencherif M, Fowler K, Lukas RJ, Lippiello PM (1995) Mechanisms of up-regulation of neuronal nicotinic acetylcholine receptors in clonal cell lines and primary cultures of fetal rat brain. J Pharmacol Exp Ther 275: 987-994.

Benowitz NL (2009) Pharmacology of nicotine: addiction, smoking-induced disease, and therapeutics. Annu Rev Pharmacol Toxicol 49: 57-71.

Benowitz NL (1990) Pharmacokinetic considerations in understanding nicotine dependence. Ciba Found Symp 152: 186-200; discussion 200-9.

Benwell ME, Balfour DJ, Anderson JM (1988) Evidence that tobacco smoking increases the density of (-)- [3H]nicotine binding sites in human brain. J Neurochem 50: 1243-1247.

Bergevin A, Girardot D, Bourque MJ, Trudeau LE (2002) Presynaptic mu-opioid receptors regulate a late step of the secretory process in rat ventral tegmental area GABAergic neurons. Neuropharmacology 42: 1065-1078.

Berlin I, Anthenelli RM (2001) Monoamine oxidases and tobacco smoking. Int J Neuropsychopharmacol 4: 33-42.

Bertrand D, Gopalakrishnan M (2007) Allosteric modulation of nicotinic acetylcholine receptors. Biochem Pharmacol 74: 1155-1163.

Besson M, Forget B (2016) Cognitive Dysfunction, Affective States, and Vulnerability to Nicotine Addiction: A Multifactorial Perspective. Front Psychiatry 7: 160.

Bierut LJ (2009) Nicotine dependence and genetic variation in the nicotinic receptors. Drug Alcohol Depend 104 Suppl 1: S64-9.

Bieter RN, Hirsch SA (1948) Methadone in internal medicine. Ann N Y Acad Sci 51: 137-144.

Bloms-Funke P, Gillen C, Schuettler AJ, Wnendt S (2000) Agonistic effects of the opioid buprenorphine on the nociceptin/OFQ receptor. Peptides 21: 1141-1146.

Bonhaus DW, Bley KR, Broka CA, Fontana DJ, Leung E, Lewis R, Shieh A, Wong EH (1995) Characterization of the electrophysiological, biochemical and behavioral actions of epibatidine. J Pharmacol Exp Ther 272: 1199-1203.

Brain KL, Trout SJ, Jackson VM, Dass N, Cunnane TC (2001) Nicotine induces calcium spikes in single nerve terminal varicosities: a role for intracellular calcium stores. Neuroscience 106: 395-403.

Breese CR, Marks MJ, Logel J, Adams CE, Sullivan B, Collins AC, Leonard S (1997) Effect of smoking history on [3H]nicotine binding in human postmortem brain. J Pharmacol Exp Ther 282: 7-13.

51 References

Brejc K, van Dijk WJ, Klaassen RV, Schuurmans M, van Der Oost J, Smit AB, Sixma TK (2001) Crystal structure of an ACh-binding protein reveals the ligand-binding domain of nicotinic receptors. Nature 411: 269-276.

Briggs SA, Hall BJ, Wells C, Slade S, Jaskowski P, Morrison M, Rezvani AH, Rose JE, Levin ED (2016) Dextromethorphan interactions with and serotonergic treatments to reduce nicotine self- administration in rats. Pharmacol Biochem Behav 142: 1-7.

Brody AL (2006) Functional brain imaging of tobacco use and dependence. J Psychiatr Res 40: 404-418.

Brody AL, Mukhin AG, La Charite J, Ta K, Farahi J, Sugar CA, Mamoun MS, Vellios E, Archie M, Kozman M, Phuong J, Arlorio F, Mandelkern MA (2013) Up-regulation of nicotinic acetylcholine receptors in cigarette smokers. Int J Neuropsychopharmacol 16: 957-966.

Brown C, Fezoui M, Selig WM, Schwartz CE, Ellis JL (2004) Antitussive activity of sigma-1 receptor agonists in the guinea-pig. Br J Pharmacol 141: 233-240.

Buisson B, Bertrand D (2001) Chronic exposure to nicotine upregulates the human (alpha)4((beta)2 nicotinic acetylcholine receptor function. J Neurosci 21: 1819-1829.

Buisson B, Vallejo YF, Green WN, Bertrand D (2000) The unusual nature of epibatidine responses at the alpha4beta2 nicotinic acetylcholine receptor. Neuropharmacology 39: 2561-2569.

Burghaus L, Schutz U, Krempel U, de Vos RA, Jansen Steur EN, Wevers A, Lindstrom J, Schroder H (2000) Quantitative assessment of nicotinic acetylcholine receptor proteins in the cerebral cortex of Alzheimer patients. Brain Res Mol Brain Res 76: 385-388.

Burling TA, Ziff DC (1988) Tobacco smoking: a comparison between alcohol and drug abuse inpatients. Addict Behav 13: 185-190.

Caggiula AR, Donny EC, Chaudhri N, Perkins KA, Evans-Martin FF, Sved AF (2002) Importance of nonpharmacological factors in nicotine self-administration. Physiol Behav 77: 683-687.

Cai Q, Qiu CY, Qiu F, Liu TT, Qu ZW, Liu YM, Hu WP (2014) Morphine inhibits acid-sensing ion channel currents in rat dorsal root ganglion neurons. Brain Res 1554: 12-20.

Castro NG, Albuquerque EX (1995) alpha-Bungarotoxin-sensitive hippocampal nicotinic receptor channel has a high calcium permeability. Biophys J 68: 516-524.

Cecchini M, Changeux JP (2015) The nicotinic acetylcholine receptor and its prokaryotic homologues: Structure, conformational transitions & allosteric modulation. Neuropharmacology 96: 137-149.

Chait LD, Griffiths RR (1984) Effects of methadone on human cigarette smoking and subjective ratings. J Pharmacol Exp Ther 229: 636-640.

Champtiaux N, Gotti C, Cordero-Erausquin M, David DJ, Przybylski C, Lena C, Clementi F, Moretti M, Rossi FM, Le Novere N, McIntosh JM, Gardier AM, Changeux JP (2003) Subunit composition of functional nicotinic receptors in dopaminergic neurons investigated with knock-out mice. J Neurosci 23: 7820-7829.

Champtiaux N, Han ZY, Bessis A, Rossi FM, Zoli M, Marubio L, McIntosh JM, Changeux JP (2002) Distribution and pharmacology of alpha 6-containing nicotinic acetylcholine receptors analyzed with mutant mice. J Neurosci 22: 1208-1217.

Changeux JP, Devillers-Thiery A, Chemouilli P (1984) Acetylcholine receptor: an allosteric protein. Science 225: 1335-1345.

Changeux JP, Taly A (2008) Nicotinic receptors, allosteric proteins and medicine. Trends Mol Med 14: 93- 102.

52 References

Charpantier E, Wiesner A, Huh KH, Ogier R, Hoda JC, Allaman G, Raggenbass M, Feuerbach D, Bertrand D, Fuhrer C (2005) Alpha7 neuronal nicotinic acetylcholine receptors are negatively regulated by tyrosine phosphorylation and Src-family kinases. J Neurosci 25: 9836-9849.

Chatzidaki A, Millar NS (2015) Allosteric modulation of nicotinic acetylcholine receptors. Biochem Pharmacol 97: 408-417.

Chiamulera C (2005) Cue reactivity in nicotine and tobacco dependence: a "multiple-action" model of nicotine as a primary reinforcement and as an enhancer of the effects of smoking-associated stimuli. Brain Res Brain Res Rev 48: 74-97.

Childs E, de Wit H (2010) Effects of acute psychosocial stress on cigarette craving and smoking. Nicotine Tob Res 12: 449-453.

Chisolm MS, Fitzsimons H, Leoutsakos JM, Acquavita SP, Heil SH, Wilson-Murphy M, Tuten M, Kaltenbach K, Martin PR, Winklbaur B, Jansson LM, Jones HE (2013) A comparison of cigarette smoking profiles in opioid-dependent pregnant patients receiving methadone or buprenorphine. Nicotine Tob Res 15: 1297- 1304.

Chizhmakov I, Yudin Y, Mamenko N, Prudnikov I, Tamarova Z, Krishtal O (2005) Opioids inhibit purinergic nociceptors in the sensory neurons and fibres of rat via a G protein-dependent mechanism. Neuropharmacology 48: 639-647.

Clarke JG, Stein MD, McGarry KA, Gogineni A (2001) Interest in smoking cessation among injection drug users. Am J Addict 10: 159-166.

Clarke PB, Schwartz RD, Paul SM, Pert CB, Pert A (1985) Nicotinic binding in rat brain: autoradiographic comparison of [3H]acetylcholine, [3H]nicotine, and [125I]-alpha-bungarotoxin. J Neurosci 5: 1307-1315.

Clementi F, Fornasari D, Gotti C (2000) Neuronal nicotinic receptors, important new players in brain function. Eur J Pharmacol 393: 3-10.

Coe JW, Brooks PR, Vetelino MG, Wirtz MC, Arnold EP, Huang J, Sands SB, Davis TI, Lebel LA, Fox CB, Shrikhande A, Heym JH, Schaeffer E, Rollema H, Lu Y, Mansbach RS, Chambers LK, Rovetti CC, Schulz DW, Tingley FD,3rd, O'Neill BT (2005) Varenicline: an alpha4beta2 nicotinic receptor partial agonist for smoking cessation. J Med Chem 48: 3474-3477.

Collins AC, Romm E, Wehner JM (1990) Dissociation of the apparent relationship between nicotine tolerance and up-regulation of nicotinic receptors. Brain Res Bull 25: 373-379.

Conklin CA (2006) Environments as cues to smoke: implications for human extinction-based research and treatment. Exp Clin Psychopharmacol 14: 12-19.

Connor M, Henderson G (1996) delta- and mu-opioid receptor mobilization of intracellular calcium in SH- SY5Y human neuroblastoma cells. Br J Pharmacol 117: 333-340.

Cook JW, Piper ME, Leventhal AM, Schlam TR, Fiore MC, Baker TB (2015) Anhedonia as a component of the tobacco withdrawal syndrome. J Abnorm Psychol 124: 215-225.

Corrigall WA, Franklin KB, Coen KM, Clarke PB (1992) The mesolimbic dopaminergic system is implicated in the reinforcing effects of nicotine. Psychopharmacology (Berl) 107: 285-289.

Court J, Spurden D, Lloyd S, McKeith I, Ballard C, Cairns N, Kerwin R, Perry R, Perry E (1999) Neuronal nicotinic receptors in dementia with Lewy bodies and schizophrenia: alpha-bungarotoxin and nicotine binding in the thalamus. J Neurochem 73: 1590-1597.

Cowan A, Lewis JW, Macfarlane IR (1977) Agonist and antagonist properties of buprenorphine, a new antinociceptive agent. Br J Pharmacol 60: 537-545.

53 References

Creese I, Sibley DR (1981) Receptor adaptations to centrally acting drugs. Annu Rev Pharmacol Toxicol 21: 357-391.

Dajas-Bailador FA, Mogg AJ, Wonnacott S (2002) Intracellular Ca2+ signals evoked by stimulation of nicotinic acetylcholine receptors in SH-SY5Y cells: contribution of voltage-operated Ca2+ channels and Ca2+ stores. J Neurochem 81: 606-614.

Dalack GW, Healy DJ, Meador-Woodruff JH (1998) Nicotine dependence in schizophrenia: clinical phenomena and laboratory findings. Am J Psychiatry 155: 1490-1501.

Daly JW, Brown GB, Mensah-Dwumah M, Myers CW (1978) Classification of skin from neotropical poison-dart frogs (Dendrobatidae). Toxicon 16: 163-188.

Damaj MI, Flood P, Ho KK, May EL, Martin BR (2005) Effect of dextrometorphan and on nicotine and neuronal nicotinic receptors: in vitro and in vivo selectivity. J Pharmacol Exp Ther 312: 780- 785.

Dani JA (2015) Neuronal Nicotinic Acetylcholine Receptor Structure and Function and Response to Nicotine. Int Rev Neurobiol 124: 3-19.

Dani JA, Bertrand D (2007) Nicotinic acetylcholine receptors and nicotinic cholinergic mechanisms of the central nervous system. Annu Rev Pharmacol Toxicol 47: 699-729.

Dani JA, De Biasi M (2001) Cellular mechanisms of nicotine addiction. Pharmacol Biochem Behav 70: 439- 446.

Dani JA, Eisenman G (1987) Monovalent and divalent cation permeation in acetylcholine receptor channels. Ion transport related to structure. J Gen Physiol 89: 959-983.

Dani JA, Heinemann S (1996) Molecular and cellular aspects of nicotine abuse. Neuron 16: 905-908.

Darsow T, Booker TK, Pina-Crespo JC, Heinemann SF (2005) Exocytic trafficking is required for nicotine- induced up-regulation of alpha 4 beta 2 nicotinic acetylcholine receptors. J Biol Chem 280: 18311-18320.

Davenport KE, Houdi AA, Van Loon GR (1990) Nicotine protects against mu-opioid receptor antagonism by beta-funaltrexamine: evidence for nicotine-induced release of endogenous opioids in brain. Neurosci Lett 113: 40-46.

David SP, Lancaster T, Stead LF, Evins AE, Prochaska JJ (2013) Opioid antagonists for smoking cessation. Cochrane Database Syst Rev (6):CD003086. doi: CD003086.

De Biasi M, Dani JA (2011) Reward, addiction, withdrawal to nicotine. Annu Rev Neurosci 34: 105-130. de Leon J, Diaz FJ (2005) A meta-analysis of worldwide studies demonstrates an association between schizophrenia and tobacco smoking behaviors. Schizophr Res 76: 135-157. de Vos JW, Geerlings PJ, van den Brink W, Ufkes JG, van Wilgenburg H (1995) of methadone and its primary metabolite in 20 opiate addicts. Eur J Clin Pharmacol 48: 361-366.

Deutsch SI, Burket JA, Urbano MR, Benson AD (2015) The alpha7 nicotinic acetylcholine receptor: A mediator of pathogenesis and therapeutic target in autism spectrum disorders and Down syndrome. Biochem Pharmacol 97: 363-377.

Dhatt RK, Gudehithlu KP, Wemlinger TA, Tejwani GA, Neff NH, Hadjiconstantinou M (1995) Preproenkephalin mRNA and methionine- content are increased in mouse striatum after treatment with nicotine. J Neurochem 64: 1878-1883.

Di Chiara G, Imperato A (1988) Drugs abused by humans preferentially increase synaptic dopamine concentrations in the mesolimbic system of freely moving rats. Proc Natl Acad Sci U S A 85: 5274-5278.

54 References

Ducharme S, Fraser R, Gill K (2012) Update on the clinical use of buprenorphine: in opioid-related disorders. Can Fam Physician 58: 37-41.

Due DL, Huettel SA, Hall WG, Rubin DC (2002) Activation in mesolimbic and visuospatial neural circuits elicited by smoking cues: evidence from functional magnetic resonance imaging. Am J Psychiatry 159: 954-960.

Dwoskin LP, Crooks PA (2001) Competitive neuronal nicotinic receptor antagonists: a new direction for drug discovery. J Pharmacol Exp Ther 298: 395-402.

Dyer KR, Foster DJ, White JM, Somogyi AA, Menelaou A, Bochner F (1999) Steady-state pharmacokinetics and in methadone maintenance patients: comparison of those who do and do not experience withdrawal and concentration-effect relationships. Clin Pharmacol Ther 65: 685-694.

Eaton JB, Peng JH, Schroeder KM, George AA, Fryer JD, Krishnan C, Buhlman L, Kuo YP, Steinlein O, Lukas RJ (2003) Characterization of human alpha 4 beta 2-nicotinic acetylcholine receptors stably and heterologously expressed in native nicotinic receptor-null SH-EP1 human epithelial cells. Mol Pharmacol 64: 1283-1294.

Elkader AK, Brands B, Selby P, Sproule BA (2009) Methadone-nicotine interactions in methadone maintenance treatment patients. J Clin Psychopharmacol 29: 231-238.

Endres-Becker J, Heppenstall PA, Mousa SA, Labuz D, Oksche A, Schafer M, Stein C, Zollner C (2007) Mu- opioid receptor activation modulates transient receptor potential vanilloid 1 (TRPV1) currents in sensory neurons in a model of inflammatory pain. Mol Pharmacol 71: 12-18.

Engelman HS, MacDermott AB (2004) Presynaptic ionotropic receptors and control of transmitter release. Nat Rev Neurosci 5: 135-145.

Fasoli F, Gotti C (2015) Structure of neuronal nicotinic receptors. Curr Top Behav Neurosci 23: 1-17.

Fasoli F, Moretti M, Zoli M, Pistillo F, Crespi A, Clementi F, Mc Clure-Begley T, Marks MJ, Gotti C (2016) In vivo chronic nicotine exposure differentially and reversibly affects upregulation and stoichiometry of alpha4beta2 nicotinic receptors in cortex and thalamus. Neuropharmacology 108: 324-331.

Feltenstein MW, See RE (2008) The neurocircuitry of addiction: an overview. Br J Pharmacol 154: 261- 274.

Feng B, Xing JH, Jia D, Liu SB, Guo HJ, Li XQ, He XS, Zhao MG (2011) Blocking alpha4beta2 and alpha7 nicotinic acetylcholine receptors inhibits the reinstatement of morphine-induced CPP by drug priming in mice. Behav Brain Res 220: 100-105.

Fenster CP, Rains MF, Noerager B, Quick MW, Lester RA (1997) Influence of subunit composition on desensitization of neuronal acetylcholine receptors at low concentrations of nicotine. J Neurosci 17: 5747-5759.

Ficklin MB, Zhao S, Feng G (2005) Ubiquilin-1 regulates nicotine-induced up-regulation of neuronal nicotinic acetylcholine receptors. J Biol Chem 280: 34088-34095.

Flores CM, Rogers SW, Pabreza LA, Wolfe BB, Kellar KJ (1992) A subtype of nicotinic cholinergic receptor in rat brain is composed of alpha 4 and beta 2 subunits and is up-regulated by chronic nicotine treatment. Mol Pharmacol 41: 31-37.

Fowler CD, Kenny PJ (2012) Habenular signaling in nicotine reinforcement. Neuropsychopharmacology 37: 306-307.

Fowler JS, Volkow ND, Wang GJ, Pappas N, Logan J, MacGregor R, Alexoff D, Shea C, Schlyer D, Wolf AP, Warner D, Zezulkova I, Cilento R (1996a) Inhibition of in the brains of smokers. Nature 379: 733-736.

55 References

Fowler JS, Volkow ND, Wang GJ, Pappas N, Logan J, Shea C, Alexoff D, MacGregor RR, Schlyer DJ, Zezulkova I, Wolf AP (1996b) Brain monoamine oxidase A inhibition in cigarette smokers. Proc Natl Acad Sci U S A 93: 14065-14069.

Frahm S, Slimak MA, Ferrarese L, Santos-Torres J, Antolin-Fontes B, Auer S, Filkin S, Pons S, Fontaine JF, Tsetlin V, Maskos U, Ibanez-Tallon I (2011) Aversion to nicotine is regulated by the balanced activity of beta4 and alpha5 nicotinic receptor subunits in the medial habenula. Neuron 70: 522-535.

Fredheim OM, Moksnes K, Borchgrevink PC, Kaasa S, Dale O (2008) Clinical pharmacology of methadone for pain. Acta Anaesthesiol Scand 52: 879-889.

Fredriksson R, Lagerstrom MC, Lundin LG, Schioth HB (2003) The G-protein-coupled receptors in the human genome form five main families. Phylogenetic analysis, paralogon groups, and fingerprints. Mol Pharmacol 63: 1256-1272.

Freedman R, Hall M, Adler LE, Leonard S (1995) Evidence in postmortem brain tissue for decreased numbers of hippocampal nicotinic receptors in schizophrenia. Biol Psychiatry 38: 22-33.

Frosch DL, Shoptaw S, Nahom D, Jarvik ME (2000) Associations between tobacco smoking and illicit drug use among methadone-maintained opiate-dependent individuals. Exp Clin Psychopharmacol 8: 97-103.

Fucile S (2004) Ca2+ permeability of nicotinic acetylcholine receptors. Cell Calcium 35: 1-8.

Fucile S, Renzi M, Lax P, Eusebi F (2003) Fractional Ca(2+) current through human neuronal alpha7 nicotinic acetylcholine receptors. Cell Calcium 34: 205-209.

Fudala PJ, Yu E, Macfadden W, Boardman C, Chiang CN (1998) Effects of buprenorphine and naloxone in morphine-stabilized opioid addicts. Drug Alcohol Depend 50: 1-8.

Gerra G, Fantoma A, Zaimovic A (2006) Naltrexone and buprenorphine combination in the treatment of opioid dependence. J Psychopharmacol 20: 806-814.

Gerzanich V, Peng X, Wang F, Wells G, Anand R, Fletcher S, Lindstrom J (1995) Comparative pharmacology of epibatidine: a potent agonist for neuronal nicotinic acetylcholine receptors. Mol Pharmacol 48: 774- 782.

Gerzanich V, Wang F, Kuryatov A, Lindstrom J (1998) alpha 5 Subunit alters desensitization, pharmacology, Ca++ permeability and Ca++ modulation of human neuronal alpha 3 nicotinic receptors. J Pharmacol Exp Ther 286: 311-320.

Gill-Thind JK, Dhankher P, D'Oyley JM, Sheppard TD, Millar NS (2015) Structurally similar allosteric modulators of alpha7 nicotinic acetylcholine receptors exhibit five distinct pharmacological effects. J Biol Chem 290: 3552-3562.

Gill JK, Dhankher P, Sheppard TD, Sher E, Millar NS (2012) A series of alpha7 nicotinic acetylcholine receptor allosteric modulators with close chemical similarity but diverse pharmacological properties. Mol Pharmacol 81: 710-718.

Gish EC, Miller JL, Honey BL, Johnson PN (2010) Lofexidine, an {alpha}2-receptor agonist for opioid detoxification. Ann Pharmacother 44: 343-351.

Glick SD, Maisonneuve IM, Dickinson HA, Kitchen BA (2001) Comparative effects of dextromethorphan and dextrorphan on morphine, , and nicotine self-administration in rats. Eur J Pharmacol 422: 87-90.

Gold MS, Levine JD (1996) DAMGO inhibits E2-induced potentiation of a TTX-resistant Na+ current in rat sensory neurons in vitro. Neurosci Lett 212: 83-86.

Goldberg SR, Henningfield JE (1988) Reinforcing effects of nicotine in humans and experimental animals responding under intermittent schedules of i.v. drug injection. Pharmacol Biochem Behav 30: 227-234.

56 References

Goodwin RD, Sheffer CE, Chartrand H, Bhaskaran J, Hart CL, Sareen J, Bolton J (2014) Drug use, abuse, and dependence and the persistence of nicotine dependence. Nicotine Tob Res 16: 1606-1612.

Gopalakrishnan M, Molinari EJ, Sullivan JP (1997) Regulation of human alpha4beta2 neuronal nicotinic acetylcholine receptors by cholinergic channel ligands and second messenger pathways. Mol Pharmacol 52: 524-534.

Gotti C, Clementi F (2004) Neuronal nicotinic receptors: from structure to pathology. Prog Neurobiol 74: 363-396.

Gotti C, Clementi F, Fornari A, Gaimarri A, Guiducci S, Manfredi I, Moretti M, Pedrazzi P, Pucci L, Zoli M (2009) Structural and functional diversity of native brain neuronal nicotinic receptors. Biochem Pharmacol 78: 703-711.

Gotti C, Moretti M, Gaimarri A, Zanardi A, Clementi F, Zoli M (2007) Heterogeneity and complexity of native brain nicotinic receptors. Biochem Pharmacol 74: 1102-1111.

Gotti C, Moretti M, Zanardi A, Gaimarri A, Champtiaux N, Changeux JP, Whiteaker P, Marks MJ, Clementi F, Zoli M (2005) Heterogeneity and selective targeting of neuronal nicotinic acetylcholine receptor (nAChR) subtypes expressed on retinal afferents of the superior colliculus and lateral geniculate nucleus: identification of a new native nAChR subtype alpha3beta2(alpha5 or beta3) enriched in retinocollicular afferents. Mol Pharmacol 68: 1162-1171.

Gotti C, Zoli M, Clementi F (2006) Brain nicotinic acetylcholine receptors: native subtypes and their relevance. Trends Pharmacol Sci 27: 482-491.

Gould J, Reeve HL, Vaughan PF, Peers C (1992) Nicotinic acetylcholine receptors in human neuroblastoma (SH-SY5Y) cells. Neurosci Lett 145: 201-204.

Govind AP, Walsh H, Green WN (2012) Nicotine-induced upregulation of native neuronal nicotinic receptors is caused by multiple mechanisms. J Neurosci 32: 2227-2238.

Gowing LR, Ali RL (2006) The place of detoxification in treatment of opioid dependence. Curr Opin Psychiatry 19: 266-270.

Grady SR, Drenan RM, Breining SR, Yohannes D, Wageman CR, Fedorov NB, McKinney S, Whiteaker P, Bencherif M, Lester HA, Marks MJ (2010) Structural differences determine the relative selectivity of nicotinic compounds for native alpha 4 beta 2*-, alpha 6 beta 2*-, alpha 3 beta 4*- and alpha 7-nicotine acetylcholine receptors. Neuropharmacology 58: 1054-1066.

Grady SR, Meinerz NM, Cao J, Reynolds AM, Picciotto MR, Changeux JP, McIntosh JM, Marks MJ, Collins AC (2001) Nicotinic agonists stimulate acetylcholine release from mouse interpeduncular nucleus: a function mediated by a different nAChR than dopamine release from striatum. J Neurochem 76: 258-268.

Gray R, Rajan AS, Radcliffe KA, Yakehiro M, Dani JA (1996) Hippocampal synaptic transmission enhanced by low concentrations of nicotine. Nature 383: 713-716.

Groot Kormelink PJ, Luyten WH (1997) Cloning and sequence of full-length cDNAs encoding the human neuronal nicotinic acetylcholine receptor (nAChR) subunits beta3 and beta4 and expression of seven nAChR subunits in the human neuroblastoma cell line SH-SY5Y and/or IMR-32. FEBS Lett 400: 309-314.

Grupe M, Grunnet M, Bastlund JF, Jensen AA (2015) Targeting alpha4beta2 nicotinic acetylcholine receptors in central nervous system disorders: perspectives on positive allosteric modulation as a therapeutic approach. Basic Clin Pharmacol Toxicol 116: 187-200.

Hamouda AK, Jin X, Sanghvi M, Srivastava S, Pandhare A, Duddempudi PK, Steinbach JH, Blanton MP (2009) Photoaffinity labeling the agonist binding domain of alpha4beta4 and alpha4beta2 neuronal nicotinic acetylcholine receptors with [(125)I]epibatidine and 5[(125)I]A-85380. Biochim Biophys Acta 1788: 1987-1995.

57 References

Harkness PC, Millar NS (2002) Changes in conformation and subcellular distribution of alpha4beta2 nicotinic acetylcholine receptors revealed by chronic nicotine treatment and expression of subunit chimeras. J Neurosci 22: 10172-10181.

Henningfield JE, Griffiths RR (1981) Cigarette smoking and subjective response: effects of d- . Clin Pharmacol Ther 30: 497-505.

Hernandez SC, Bertolino M, Xiao Y, Pringle KE, Caruso FS, Kellar KJ (2000) Dextromethorphan and its metabolite dextrorphan block alpha3beta4 neuronal nicotinic receptors. J Pharmacol Exp Ther 293: 962- 967.

Hernandez SC, Vicini S, Xiao Y, Davila-Garcia MI, Yasuda RP, Wolfe BB, Kellar KJ (2004) The nicotinic receptor in the rat pineal gland is an alpha3beta4 subtype. Mol Pharmacol 66: 978-987.

Higgins ST, Budney AJ, Hughes JR, Bickel WK, Lynn M, Mortensen A (1994) Influence of use on cigarette smoking. JAMA 272: 1724.

Hucho F, Oberthur W, Lottspeich F (1986) The ion channel of the nicotinic acetylcholine receptor is formed by the homologous helices M II of the receptor subunits. FEBS Lett 205: 137-142.

Hurst R, Rollema H, Bertrand D (2013) Nicotinic acetylcholine receptors: from basic science to therapeutics. Pharmacol Ther 137: 22-54.

Hurt RD, Offord KP, Croghan IT, Gomez-Dahl L, Kottke TE, Morse RM, Melton LJ,3rd (1996) Mortality following inpatient treatment. Role of tobacco use in a community-based cohort. JAMA 275: 1097-1103.

Hussmann GP, Turner JR, Lomazzo E, Venkatesh R, Cousins V, Xiao Y, Yasuda RP, Wolfe BB, Perry DC, Rezvani AH, Levin ED, Blendy JA, Kellar KJ (2012) Chronic sazetidine-A at behaviorally active doses does not increase nicotinic cholinergic receptors in rodent brain. J Pharmacol Exp Ther 343: 441-450.

Hussmann GP, Yasuda RP, Xiao Y, Wolfe BB, Kellar KJ (2011) Endogenously expressed muscarinic receptors in HEK293 cells augment up-regulation of stably expressed alpha4beta2 nicotinic receptors. J Biol Chem 286: 39726-39737.

Imperato A, Mulas A, Di Chiara G (1986) Nicotine preferentially stimulates dopamine release in the limbic system of freely moving rats. Eur J Pharmacol 132: 337-338.

Iorga B, Herlem D, Barre E, Guillou C (2006) Acetylcholine nicotinic receptors: finding the putative binding site of allosteric modulators using the "blind docking" approach. J Mol Model 12: 366-372.

Islami F, Torre LA, Jemal A (2015) Global trends of lung cancer mortality and smoking prevalence. Transl Lung Cancer Res 4: 327-338.

Jackson KJ, Marks MJ, Vann RE, Chen X, Gamage TF, Warner JA, Damaj MI (2010) Role of alpha5 nicotinic acetylcholine receptors in pharmacological and behavioral effects of nicotine in mice. J Pharmacol Exp Ther 334: 137-146.

Jasinska AJ, Zorick T, Brody AL, Stein EA (2014) Dual role of nicotine in addiction and cognition: a review of neuroimaging studies in humans. Neuropharmacology 84: 111-122.

Jasinski DR, Pevnick JS, Griffith JD (1978) Human pharmacology and abuse potential of the analgesic buprenorphine: a potential agent for treating addiction. Arch Gen Psychiatry 35: 501-516.

Jin W, Lee NM, Loh HH, Thayer SA (1992) Dual excitatory and inhibitory effects of opioids on intracellular calcium in neuroblastoma x glioma hybrid NG108-15 cells. Mol Pharmacol 42: 1083-1089.

Johnson RE, Fudala PJ, Payne R (2005) Buprenorphine: considerations for . J Pain Symptom Manage 29: 297-326.

58 References

Jones IW, Wonnacott S (2004) Precise localization of alpha7 nicotinic acetylcholine receptors on glutamatergic axon terminals in the rat ventral tegmental area. J Neurosci 24: 11244-11252.

Jones S, Sudweeks S, Yakel JL (1999) Nicotinic receptors in the brain: correlating physiology with function. Trends Neurosci 22: 555-561.

Jurgensen S, Ferreira ST (2010) Nicotinic receptors, amyloid-beta, and synaptic failure in Alzheimer's disease. J Mol Neurosci 40: 221-229.

Jääskeläinen M, Virtanen S (2015) Tobacco statistics 2015. Helsinki, Finland: Official Statistics of Finland, National Institute for Health and Welfare.

Kamei J, Saitoh A, Suzuki T, Misawa M, Nagase H, Kasuya Y (1995) Buprenorphine exerts its antinociceptive activity via mu 1-opioid receptors. Life Sci 56: PL285-90.

Karlin A (2002) Emerging structure of the nicotinic acetylcholine receptors. Nat Rev Neurosci 3: 102-114.

Katritch V, Cherezov V, Stevens RC (2013) Structure-function of the G protein-coupled receptor superfamily. Annu Rev Pharmacol Toxicol 53: 531-556.

Katz B, Thesleff S (1957) A study of the desensitization produced by acetylcholine at the motor end-plate. J Physiol 138: 63-80.

Kazmi SM, Mishra RK (1987) Comparative pharmacological properties and functional coupling of mu and delta opioid receptor sites in human neuroblastoma SH-SY5Y cells. Mol Pharmacol 32: 109-118.

Ke L, Eisenhour CM, Bencherif M, Lukas RJ (1998) Effects of chronic nicotine treatment on expression of diverse nicotinic acetylcholine receptor subtypes. I. Dose- and time-dependent effects of nicotine treatment. J Pharmacol Exp Ther 286: 825-840.

Kendziorra K, Wolf H, Meyer PM, Barthel H, Hesse S, Becker GA, Luthardt J, Schildan A, Patt M, Sorger D, Seese A, Gertz HJ, Sabri O (2011) Decreased cerebral alpha4beta2* nicotinic acetylcholine receptor availability in patients with mild cognitive impairment and Alzheimer's disease assessed with positron emission tomography. Eur J Nucl Med Mol Imaging 38: 515-525.

Kenny PJ, Markou A (2005) Conditioned nicotine withdrawal profoundly decreases the activity of brain reward systems. J Neurosci 25: 6208-6212.

Kimber J, Larney S, Hickman M, Randall D, Degenhardt L (2015) Mortality risk of opioid substitution therapy with methadone versus buprenorphine: a retrospective cohort study. Lancet Psychiatry 2: 901- 908.

Kishi M, Steinbach JH (2006) Role of the agonist binding site in up-regulation of neuronal nicotinic alpha4beta2 receptors. Mol Pharmacol 70: 2037-2044.

Korpi ER, den Hollander B, Farooq U, Vashchinkina E, Rajkumar R, Nutt DJ, Hyytia P, Dawe GS (2015) Mechanisms of Action and Persistent Neuroplasticity by Drugs of Abuse. Pharmacol Rev 67: 872-1004.

Kosten TR, George TP (2002) The neurobiology of opioid dependence: implications for treatment. Sci Pract Perspect 1: 13-20.

Kracun S, Harkness PC, Gibb AJ, Millar NS (2008) Influence of the M3-M4 intracellular domain upon nicotinic acetylcholine receptor assembly, targeting and function. Br J Pharmacol 153: 1474-1484.

Kramer JC (1970) Methadone maintenance for opiate dependence. Calif Med 113: 6-11.

Kranz GS, Kasper S, Lanzenberger R (2010) Reward and the serotonergic system. Neuroscience 166: 1023- 1035.

59 References

Krashia P, Moroni M, Broadbent S, Hofmann G, Kracun S, Beato M, Groot-Kormelink PJ, Sivilotti LG (2010) Human alpha3beta4 neuronal nicotinic receptors show different stoichiometry if they are expressed in Xenopus oocytes or mammalian HEK293 cells. PLoS One 5: e13611.

Kristensen K, Christensen CB, Christrup LL (1995) The mu1, mu2, delta, kappa opioid receptor binding profiles of methadone stereoisomers and morphine. Life Sci 56: PL45-50.

Kukkonen J, Ojala P, Nasman J, Hamalainen H, Heikkila J, Akerman KE (1992) Muscarinic receptor subtypes in human neuroblastoma cell lines SH-SY5Y and IMR-32 as determined by receptor binding, Ca++ mobilization and northern blotting. J Pharmacol Exp Ther 263: 1487-1493.

Kuryatov A, Luo J, Cooper J, Lindstrom J (2005) Nicotine acts as a pharmacological chaperone to up- regulate human alpha4beta2 acetylcholine receptors. Mol Pharmacol 68: 1839-1851.

Kuryatov A, Olale F, Cooper J, Choi C, Lindstrom J (2000) Human alpha6 AChR subtypes: subunit composition, assembly, and pharmacological responses. Neuropharmacology 39: 2570-2590.

Kuryatov A, Onksen J, Lindstrom J (2008) Roles of accessory subunits in alpha4beta2(*) nicotinic receptors. Mol Pharmacol 74: 132-143.

Le Foll B, Chefer SI, Kimes AS, Shumway D, Stein EA, Mukhin AG, Goldberg SR (2009) Baseline expression of alpha4beta2* nicotinic acetylcholine receptors predicts motivation to self-administer nicotine. Biol Psychiatry 65: 714-716.

Le Foll B, Goldberg SR (2009) Effects of nicotine in experimental animals and humans: an update on addictive properties. Handb Exp Pharmacol (192):335-67. doi: 335-367.

Le Foll B, Goldberg SR (2005a) Control of the reinforcing effects of nicotine by associated environmental stimuli in animals and humans. Trends Pharmacol Sci 26: 287-293.

Le Foll B, Goldberg SR (2005b) Nicotine induces conditioned place preferences over a large range of doses in rats. Psychopharmacology (Berl) 178: 481-492.

Le Merrer J, Becker JA, Befort K, Kieffer BL (2009) Reward processing by the opioid system in the brain. Physiol Rev 89: 1379-1412.

Le Novere N, Zoli M, Changeux JP (1996) Neuronal nicotinic receptor alpha 6 subunit mRNA is selectively concentrated in catecholaminergic nuclei of the rat brain. Eur J Neurosci 8: 2428-2439.

Leander JD (1987) Buprenorphine has potent kappa opioid activity. Neuropharmacology 26: 1445-1447.

Lee M, Martin-Ruiz C, Graham A, Court J, Jaros E, Perry R, Iversen P, Bauman M, Perry E (2002) Nicotinic receptor abnormalities in the cerebellar cortex in autism. Brain 125: 1483-1495.

Lee WY, Free CR, Sine SM (2009) Binding to gating transduction in nicotinic receptors: Cys-loop energetically couples to pre-M1 and M2-M3 regions. J Neurosci 29: 3189-3199.

Lee WY, Sine SM (2005) Principal pathway coupling agonist binding to channel gating in nicotinic receptors. Nature 438: 243-247.

Lemon SC, Friedmann PD, Stein MD (2003) The impact of smoking cessation on drug abuse treatment outcome. Addict Behav 28: 1323-1331.

Lena C, Changeux JP, Mulle C (1993) Evidence for "preterminal" nicotinic receptors on GABAergic axons in the rat interpeduncular nucleus. J Neurosci 13: 2680-2688.

Lena C, de Kerchove D'Exaerde A, Cordero-Erausquin M, Le Novere N, del Mar Arroyo-Jimenez M, Changeux JP (1999) Diversity and distribution of nicotinic acetylcholine receptors in the locus ceruleus neurons. Proc Natl Acad Sci U S A 96: 12126-12131.

60 References

Lester RA, Dani JA (1994) Time-dependent changes in central nicotinic acetylcholine channel kinetics in excised patches. Neuropharmacology 33: 27-34.

Levin ED, McClernon FJ, Rezvani AH (2006) Nicotinic effects on cognitive function: behavioral characterization, pharmacological specification, and anatomic localization. Psychopharmacology (Berl) 184: 523-539.

Lewis A, Miller JH, Lea RA (2007) Monoamine oxidase and tobacco dependence. Neurotoxicology 28: 182- 195.

Lewis KS, Han NH (1997) Tramadol: a new centrally acting analgesic. Am J Health Syst Pharm 54: 643-652.

Li T, Qian C, Eckman J, Huang D, Shen T (1993) The analgesic effect of epibatidine and isomers. Bioorg Med Chem Lett 3: 2759-2764.

Liu Q, Huang Y, Shen J, Steffensen S, Wu J (2012) Functional alpha7beta2 nicotinic acetylcholine receptors expressed in hippocampal interneurons exhibit high sensitivity to pathological level of amyloid beta peptides. BMC Neurosci 13: 155-2202-13-155.

Liu Q, Huang Y, Xue F, Simard A, DeChon J, Li G, Zhang J, Lucero L, Wang M, Sierks M, Hu G, Chang Y, Lukas RJ, Wu J (2009) A novel nicotinic acetylcholine receptor subtype in basal forebrain cholinergic neurons with high sensitivity to amyloid peptides. J Neurosci 29: 918-929.

Lobmaier P, Gossop M, Waal H, Bramness J (2010) The pharmacological treatment of opioid addiction-- a clinical perspective. Eur J Clin Pharmacol 66: 537-545.

Lofwall MR, Walsh SL (2014) A review of buprenorphine diversion and misuse: the current evidence base and experiences from around the world. J Addict Med 8: 315-326.

Lopez-Hernandez GY, Sanchez-Padilla J, Ortiz-Acevedo A, Lizardi-Ortiz J, Salas-Vincenty J, Rojas LV, Lasalde-Dominicci JA (2004) Nicotine-induced up-regulation and desensitization of alpha4beta2 neuronal nicotinic receptors depend on subunit ratio. J Biol Chem 279: 38007-38015.

Lukas RJ, Fryer JD, Eaton JB, Gentry CL: Some methods for studies of nicotinic acetylcholine receptor pharmacology. Kirjassa: Nicotinic Receptors and the Nervous Syste, s. 3-27, 1. painos. Toim. Levin ED, CRC Press, Boca Raton 2002

Lukas RJ, Cullen MJ (1988) An isotopic rubidium ion efflux assay for the functional characterization of nicotinic acetylcholine receptors on clonal cell lines. Anal Biochem 175: 212-218.

Lukas RJ, Norman SA, Lucero L (1993) Characterization of Nicotinic Acetylcholine Receptors Expressed by Cells of the SH-SY5Y Human Neuroblastoma Clonal Line. Mol Cell Neurosci 4: 1-12.

Luo S, Kulak JM, Cartier GE, Jacobsen RB, Yoshikami D, Olivera BM, McIntosh JM (1998) alpha-conotoxin AuIB selectively blocks alpha3 beta4 nicotinic acetylcholine receptors and nicotine-evoked norepinephrine release. J Neurosci 18: 8571-8579.

Luscher C, Slesinger PA (2010) Emerging roles for G protein-gated inwardly rectifying potassium (GIRK) channels in health and disease. Nat Rev Neurosci 11: 301-315.

Madsen BW, Albuquerque EX (1985) The narcotic antagonist naltrexone has a biphasic effect on the nicotinic acetylcholine receptor. FEBS Lett 182: 20-24.

Mamede M, Ishizu K, Ueda M, Mukai T, Iida Y, Kawashima H, Fukuyama H, Togashi K, Saji H (2007) Temporal change in human nicotinic acetylcholine receptor after smoking cessation: 5IA SPECT study. J Nucl Med 48: 1829-1835.

Mansvelder HD, McGehee DS (2002) Cellular and synaptic mechanisms of nicotine addiction. J Neurobiol 53: 606-617.

61 References

Marks MJ, Burch JB, Collins AC (1983) Effects of chronic nicotine infusion on tolerance development and nicotinic receptors. J Pharmacol Exp Ther 226: 817-825.

Marks MJ, Collins AC (1982) Characterization of nicotine binding in mouse brain and comparison with the binding of alpha-bungarotoxin and quinuclidinyl benzilate. Mol Pharmacol 22: 554-564.

Marks MJ, McClure-Begley TD, Whiteaker P, Salminen O, Brown RW, Cooper J, Collins AC, Lindstrom JM (2011) Increased nicotinic acetylcholine receptor protein underlies chronic nicotine-induced up- regulation of nicotinic agonist binding sites in mouse brain. J Pharmacol Exp Ther 337: 187-200.

Marks MJ, Meinerz NM, Brown RW, Collins AC (2010) 86Rb+ efflux mediated by alpha4beta2*-nicotinic acetylcholine receptors with high and low-sensitivity to stimulation by acetylcholine display similar agonist-induced desensitization. Biochem Pharmacol 80: 1238-1251.

Marks MJ, O'Neill HC, Wynalda-Camozzi KM, Ortiz NC, Simmons EE, Short CA, Butt CM, McIntosh JM, Grady SR (2015) Chronic treatment with varenicline changes expression of four nAChR binding sites in mice. Neuropharmacology 99: 142-155.

Marks MJ, Pauly JR, Gross SD, Deneris ES, Hermans-Borgmeyer I, Heinemann SF, Collins AC (1992) Nicotine binding and nicotinic receptor subunit RNA after chronic nicotine treatment. J Neurosci 12: 2765-2784.

Marks MJ, Stitzel JA, Collins AC (1985) Time course study of the effects of chronic nicotine infusion on drug response and brain receptors. J Pharmacol Exp Ther 235: 619-628.

Marks MJ, Whiteaker P, Collins AC (2006) Deletion of the alpha7, beta2, or beta4 nicotinic receptor subunit genes identifies highly expressed subtypes with relatively low affinity for [3H]epibatidine. Mol Pharmacol 70: 947-959.

Martin-Ruiz CM, Lee M, Perry RH, Baumann M, Court JA, Perry EK (2004) Molecular analysis of nicotinic receptor expression in autism. Brain Res Mol Brain Res 123: 81-90.

Maxwell JC, Pullum TW, Tannert K (2005) Deaths of clients in methadone treatment in Texas: 1994-2002. Drug Alcohol Depend 78: 73-81.

Mazzo F, Pistillo F, Grazioso G, Clementi F, Borgese N, Gotti C, Colombo SF (2013) Nicotine-modulated subunit stoichiometry affects stability and trafficking of alpha3beta4 nicotinic receptor. J Neurosci 33: 12316-12328.

McClure EA, Acquavita SP, Dunn KE, Stoller KB, Stitzer ML (2014) Characterizing smoking, cessation services, and quit interest across outpatient treatment modalities. J Subst Abuse Treat 46: 194-201.

McLaughlin I, Dani JA, De Biasi M (2015) Nicotine withdrawal. Curr Top Behav Neurosci 24: 99-123.

McLellan AT, Lewis DC, O'Brien CP, Kleber HD (2000) Drug dependence, a chronic medical illness: implications for treatment, insurance, and outcomes evaluation. JAMA 284: 1689-1695.

Meader N (2010) A comparison of methadone, buprenorphine and alpha(2) adrenergic agonists for opioid detoxification: a mixed treatment comparison meta-analysis. Drug Alcohol Depend 108: 110-114.

Mello NK, Lukas SE, Mendelson JH (1985) Buprenorphine effects on cigarette smoking. Psychopharmacology (Berl) 86: 417-425.

Mello NK, Mendelson JH, Sellers ML, Kuehnle JC (1980) Effects of heroin self-administration on cigarette smoking. Psychopharmacology (Berl) 67: 45-52.

Mexal S, Berger R, Logel J, Ross RG, Freedman R, Leonard S (2010) Differential regulation of alpha7 nicotinic receptor gene (CHRNA7) expression in schizophrenic smokers. J Mol Neurosci 40: 185-195.

62 References

Meyer PM, Strecker K, Kendziorra K, Becker G, Hesse S, Woelpl D, Hensel A, Patt M, Sorger D, Wegner F, Lobsien D, Barthel H, Brust P, Gertz HJ, Sabri O, Schwarz J (2009) Reduced alpha4beta2*-nicotinic acetylcholine receptor binding and its relationship to mild cognitive and depressive symptoms in Parkinson disease. Arch Gen Psychiatry 66: 866-877.

Miller PS, Smart TG (2010) Binding, activation and modulation of Cys-loop receptors. Trends Pharmacol Sci 31: 161-174.

Molinari EJ, Delbono O, Messi ML, Renganathan M, Arneric SP, Sullivan JP, Gopalakrishnan M (1998) Up- regulation of human alpha7 nicotinic receptors by chronic treatment with activator and antagonist ligands. Eur J Pharmacol 347: 131-139.

Mooney ME, Poling J, Gonzalez G, Gonsai K, Kosten T, Sofuoglu M (2008) Preliminary study of buprenorphine and bupropion for opioid-dependent smokers. Am J Addict 17: 287-292.

Moretti M, Zoli M, George AA, Lukas RJ, Pistillo F, Maskos U, Whiteaker P, Gotti C (2014) The novel alpha7beta2-nicotinic acetylcholine receptor subtype is expressed in mouse and human basal forebrain: biochemical and pharmacological characterization. Mol Pharmacol 86: 306-317.

Moroni M, Zwart R, Sher E, Cassels BK, Bermudez I (2006) Alpha4beta2 Nicotinic Receptors with High and Low Acetylcholine Sensitivity: Pharmacology, Stoichiometry, and Sensitivity to Long-Term Exposure to Nicotine. Mol Pharmacol 70: 755-768.

Mukhin AG, Kimes AS, Chefer SI, Matochik JA, Contoreggi CS, Horti AG, Vaupel DB, Pavlova O, Stein EA (2008) Greater nicotinic acetylcholine receptor density in smokers than in nonsmokers: a PET study with 2-18F-FA-85380. J Nucl Med 49: 1628-1635.

Muldoon PP, Jackson KJ, Perez E, Harenza JL, Molas S, Rais B, Anwar H, Zaveri NT, Maldonado R, Maskos U, McIntosh JM, Dierssen M, Miles MF, Chen X, De Biasi M, Damaj MI (2014) The alpha3beta4* nicotinic ACh receptor subtype mediates physical dependence to morphine: mouse and human studies. Br J Pharmacol 171: 3845-3857.

Mutschler NH, Stephen BJ, Teoh SK, Mendelson JH, Mello NK (2002) An inpatient study of the effects of buprenorphine on cigarette smoking in men concurrently dependent on cocaine and opioids. Nicotine Tob Res 4: 223-228.

Nahvi S, Ning Y, Segal KS, Richter KP, Arnsten JH (2014) Varenicline efficacy and safety among methadone maintained smokers: a randomized placebo-controlled trial. Addiction 109: 1554-1563.

Nahvi S, Richter K, Li X, Modali L, Arnsten J (2006) Cigarette smoking and interest in quitting in methadone maintenance patients. Addict Behav 31: 2127-2134.

Nashmi R, Dickinson ME, McKinney S, Jareb M, Labarca C, Fraser SE, Lester HA (2003) Assembly of alpha4beta2 nicotinic acetylcholine receptors assessed with functional fluorescently labeled subunits: effects of localization, trafficking, and nicotine-induced upregulation in clonal mammalian cells and in cultured midbrain neurons. J Neurosci 23: 11554-11567.

Nashmi R, Xiao C, Deshpande P, McKinney S, Grady SR, Whiteaker P, Huang Q, McClure-Begley T, Lindstrom JM, Labarca C, Collins AC, Marks MJ, Lester HA (2007) Chronic nicotine cell specifically upregulates functional alpha 4* nicotinic receptors: basis for both tolerance in midbrain and enhanced long-term potentiation in perforant path. J Neurosci 27: 8202-8218.

Nelson ME, Kuryatov A, Choi CH, Zhou Y, Lindstrom J (2003) Alternate stoichiometries of alpha4beta2 nicotinic acetylcholine receptors. Mol Pharmacol 63: 332-341.

Nestler EJ (2005) Is there a common molecular pathway for addiction? Nat Neurosci 8: 1445-1449.

Nguyen HN, Rasmussen BA, Perry DC (2003) Subtype-selective up-regulation by chronic nicotine of high- affinity nicotinic receptors in rat brain demonstrated by receptor autoradiography. J Pharmacol Exp Ther 307: 1090-1097.

63 References

Noble F, Lenoir M, Marie N (2015) The opioid receptors as targets for drug abuse medication. Br J Pharmacol 172: 3964-3979.

Nockemann D, Rouault M, Labuz D, Hublitz P, McKnelly K, Reis FC, Stein C, Heppenstall PA (2013) The K(+) channel GIRK2 is both necessary and sufficient for peripheral opioid-mediated analgesia. EMBO Mol Med 5: 1263-1277.

Nutt DJ, Lingford-Hughes A, Erritzoe D, Stokes PR (2015) The dopamine theory of addiction: 40 years of highs and lows. Nat Rev Neurosci 16: 305-312.

Nuutinen S, Ekokoski E, Lahdensuo E, Tuominen RK (2006) Nicotine-induced upregulation of human neuronal nicotinic alpha7-receptors is potentiated by modulation of cAMP and PKC in SH-EP1-halpha7 cells. Eur J Pharmacol 544: 21-30.

Okajima F, Tomura H, Kondo Y (1993) Enkephalin activates the phospholipase C/Ca2+ system through cross-talk between opioid receptors and P2-purinergic or bradykinin receptors in NG 108-15 cells. A permissive role for pertussis toxin-sensitive G-proteins. Biochem J 290 ( Pt 1): 241-247.

Okoli CT, Khara M, Procyshyn RM, Johnson JL, Barr AM, Greaves L (2010) Smoking cessation interventions among individuals in methadone maintenance: a brief review. J Subst Abuse Treat 38: 191-199.

Pajusco B, Chiamulera C, Quaglio G, Moro L, Casari R, Amen G, Faccini M, Lugoboni F (2012) Tobacco addiction and smoking status in heroin addicts under methadone vs. buprenorphine therapy. Int J Environ Res Public Health 9: 932-942.

Pakkanen JS, Nousiainen H, Yli-Kauhaluoma J, Kylanlahti I, Moykkynen T, Korpi ER, Peng JH, Lukas RJ, Ahtee L, Tuominen RK (2005) Methadone increases intracellular calcium in SH-SY5Y and SH-EP1-halpha7 cells by activating neuronal nicotinic acetylcholine receptors. J Neurochem 94: 1329-1341.

Papke RL (2014) Merging old and new perspectives on nicotinic acetylcholine receptors. Biochem Pharmacol 89: 1-11.

Parker MJ, Beck A, Luetje CW (1998) Neuronal nicotinic receptor beta2 and beta4 subunits confer large differences in agonist binding affinity. Mol Pharmacol 54: 1132-1139.

Pasternak GW, Pan YX (2013) Mu opioids and their receptors: evolution of a concept. Pharmacol Rev 65: 1257-1317.

Patrick ME, Dunn KE, Badger GJ, Heil SH, Higgins ST, Sigmon SC (2014) Spontaneous reductions in smoking during double-blind buprenorphine detoxification. Addict Behav 39: 1353-1356.

Pauly JR, Marks MJ, Gross SD, Collins AC (1991) An autoradiographic analysis of cholinergic receptors in mouse brain after chronic nicotine treatment. J Pharmacol Exp Ther 258: 1127-1136.

Peng C, Stokes C, Mineur YS, Picciotto MR, Tian C, Eibl C, Tomassoli I, Guendisch D, Papke RL (2013) Differential modulation of brain nicotinic acetylcholine receptor function by cytisine, varenicline, and two novel bispidine compounds: emergent properties of a hybrid molecule. J Pharmacol Exp Ther 347: 424- 437.

Peng JH, Lucero L, Fryer J, Herl J, Leonard SS, Lukas RJ (1999) Inducible, heterologous expression of human alpha7-nicotinic acetylcholine receptors in a native nicotinic receptor-null human clonal line. Brain Res 825: 172-179.

Peng X, Gerzanich V, Anand R, Wang F, Lindstrom J (1997) Chronic nicotine treatment up-regulates alpha3 and alpha7 acetylcholine receptor subtypes expressed by the human neuroblastoma cell line SH-SY5Y. Mol Pharmacol 51: 776-784.

Peng X, Gerzanich V, Anand R, Whiting PJ, Lindstrom J (1994) Nicotine-induced increase in neuronal nicotinic receptors results from a decrease in the rate of receptor turnover. Mol Pharmacol 46: 523-530.

64 References

Perez E, Quijano-Carde N, De Biasi M (2015) Nicotinic Mechanisms Modulate Withdrawal and Modify Time Course and Symptoms Severity of Simultaneous Withdrawal from Alcohol and Nicotine. Neuropsychopharmacology 40: 2327-2336.

Perez XA, Bordia T, McIntosh JM, Grady SR, Quik M (2008) Long-term nicotine treatment differentially regulates striatal alpha6alpha4beta2* and alpha6(nonalpha4)beta2* nAChR expression and function. Mol Pharmacol 74: 844-853.

Perkins KA, Grobe JE (1992) Increased desire to smoke during acute stress. Br J Addict 87: 1037-1040.

Perry DC, Davila-Garcia MI, Stockmeier CA, Kellar KJ (1999) Increased nicotinic receptors in brains from smokers: membrane binding and autoradiography studies. J Pharmacol Exp Ther 289: 1545-1552.

Perry DC, Mao D, Gold AB, McIntosh JM, Pezzullo JC, Kellar KJ (2007) Chronic nicotine differentially regulates alpha6- and beta3-containing nicotinic cholinergic receptors in rat brain. J Pharmacol Exp Ther 322: 306-315.

Picciotto MR, Addy NA, Mineur YS, Brunzell DH (2008) It is not "either/or": activation and desensitization of nicotinic acetylcholine receptors both contribute to behaviors related to nicotine addiction and mood. Prog Neurobiol 84: 329-342.

Picciotto MR, Caldarone BJ, Brunzell DH, Zachariou V, Stevens TR, King SL (2001) Neuronal nicotinic acetylcholine receptor subunit knockout mice: physiological and behavioral phenotypes and possible clinical implications. Pharmacol Ther 92: 89-108.

Picciotto MR, Zoli M, Rimondini R, Lena C, Marubio LM, Pich EM, Fuxe K, Changeux JP (1998) Acetylcholine receptors containing the beta2 subunit are involved in the reinforcing properties of nicotine. Nature 391: 173-177.

Pidoplichko VI, DeBiasi M, Williams JT, Dani JA (1997) Nicotine activates and desensitizes midbrain dopamine neurons. Nature 390: 401-404.

Pidoplichko VI, Prager EM, Aroniadou-Anderjaska V, Braga MF (2013) alpha7-Containing nicotinic acetylcholine receptors on interneurons of the basolateral amygdala and their role in the regulation of the network excitability. J Neurophysiol 110: 2358-2369.

Pollock VV, Pastoor T, Katnik C, Cuevas J, Wecker L (2009) Cyclic AMP-dependent protein kinase A and protein kinase C phosphorylate alpha4beta2 nicotinic receptor subunits at distinct stages of receptor formation and maturation. Neuroscience 158: 1311-1325.

Potter AS, Schaubhut G, Shipman M (2014) Targeting the nicotinic cholinergic system to treat attention- deficit/hyperactivity disorder: rationale and progress to date. CNS Drugs 28: 1103-1113.

Pratt MB, Pedersen SE, Cohen JB (2000) Identification of the sites of incorporation of [3H]ethidium diazide within the Torpedo nicotinic acetylcholine receptor ion channel. Biochemistry 39: 11452-11462.

Qian C, Li T, Shen TY, Libertine-Garahan L, Eckman J, Biftu T, Ip S (1993) Epibatidine is a nicotinic analgesic. Eur J Pharmacol 250: R13-4.

Quick MW, Ceballos RM, Kasten M, McIntosh JM, Lester RA (1999) Alpha3beta4 subunit-containing nicotinic receptors dominate function in rat medial habenula neurons. Neuropharmacology 38: 769-783.

Quik M, Philie J, Choremis J (1997) Modulation of alpha7 nicotinic receptor-mediated calcium influx by nicotinic agonists. Mol Pharmacol 51: 499-506.

Quik M, Wonnacott S (2011) alpha6beta2* and alpha4beta2* nicotinic acetylcholine receptors as drug targets for Parkinson's disease. Pharmacol Rev 63: 938-966.

65 References

Rao TS, Correa LD, Adams P, Santori EM, Sacaan AI (2003) Pharmacological characterization of dopamine, norepinephrine and serotonin release in the rat prefrontal cortex by neuronal nicotinic acetylcholine receptor agonists. Brain Res 990: 203-208.

Rathouz MM, Berg DK (1994) Synaptic-type acetylcholine receptors raise intracellular calcium levels in neurons by two mechanisms. J Neurosci 14: 6935-6945.

Rezvani K, Teng Y, Shim D, De Biasi M (2007) Nicotine regulates multiple synaptic proteins by inhibiting proteasomal activity. J Neurosci 27: 10508-10519.

Rice ME, Cragg SJ (2004) Nicotine amplifies reward-related dopamine signals in striatum. Nat Neurosci 7: 583-584.

Richter KP, Hamilton AK, Hall S, Catley D, Cox LS, Grobe J (2007) Patterns of smoking and methadone dose in drug treatment patients. Exp Clin Psychopharmacol 15: 144-153.

Ridley DL, Rogers A, Wonnacott S (2001) Differential effects of chronic drug treatment on alpha3* and alpha7 nicotinic receptor binding sites, in hippocampal neurones and SH-SY5Y cells. Br J Pharmacol 133: 1286-1295.

Role LW, Berg DK (1996) Nicotinic receptors in the development and modulation of CNS synapses. Neuron 16: 1077-1085.

Roll JM, Higgins ST, Tidey J (1997) Cocaine use can increase cigarette smoking: evidence from laboratory and naturalistic settings. Exp Clin Psychopharmacol 5: 263-268.

Rollema H, Coe JW, Chambers LK, Hurst RS, Stahl SM, Williams KE (2007) Rationale, pharmacology and clinical efficacy of partial agonists of alpha4beta2 nACh receptors for smoking cessation. Trends Pharmacol Sci 28: 316-325.

Rose JE, Mukhin AG, Lokitz SJ, Turkington TG, Herskovic J, Behm FM, Garg S, Garg PK (2010) Kinetics of brain nicotine accumulation in dependent and nondependent smokers assessed with PET and cigarettes containing 11C-nicotine. Proc Natl Acad Sci U S A 107: 5190-5195.

Ross RA, Spengler BA, Biedler JL (1983) Coordinate morphological and biochemical interconversion of human neuroblastoma cells. J Natl Cancer Inst 71: 741-747.

Ross S, Peselow E (2009) The neurobiology of addictive disorders. Clin Neuropharmacol 32: 269-276.

Rowell PP, Li M (1997) Dose-response relationship for nicotine-induced up-regulation of rat brain nicotinic receptors. J Neurochem 68: 1982-1989.

Rubboli F, Court JA, Sala C, Morris C, Chini B, Perry E, Clementi F (1994) Distribution of nicotinic receptors in the human hippocampus and thalamus. Eur J Neurosci 6: 1596-1604.

Sakurada T, Takada S, Eguchi H, Izumi K, Satoh N, Ueda S (2010) Relationship between plasma concentrations of morphine and its metabolites and pain in cancer patients. Pharm World Sci 32: 737- 743.

Salas R, Pieri F, De Biasi M (2004) Decreased signs of nicotine withdrawal in mice null for the beta4 nicotinic acetylcholine receptor subunit. J Neurosci 24: 10035-10039.

Salas R, Sturm R, Boulter J, De Biasi M (2009) Nicotinic receptors in the habenulo-interpeduncular system are necessary for nicotine withdrawal in mice. J Neurosci 29: 3014-3018.

Sallette J, Pons S, Devillers-Thiery A, Soudant M, Prado de Carvalho L, Changeux JP, Corringer PJ (2005) Nicotine upregulates its own receptors through enhanced intracellular maturation. Neuron 46: 595-607.

66 References

Samways DS, Henderson G (2006) Opioid elevation of intracellular free calcium: possible mechanisms and physiological relevance. Cell Signal 18: 151-161.

Sarter M, Lustig C, Taylor SF (2012) Cholinergic contributions to the cognitive symptoms of schizophrenia and the viability of cholinergic treatments. Neuropharmacology 62: 1544-1553.

Schmidt BL, Tambeli CH, Gear RW, Levine JD (2001) Nicotine withdrawal hyperalgesia and opioid- mediated analgesia depend on nicotine receptors in nucleus accumbens. Neuroscience 106: 129-136.

Schneider NG, Olmstead RE, Franzon MA, Lunell E (2001) The nicotine inhaler: clinical pharmacokinetics and comparison with other nicotine treatments. Clin Pharmacokinet 40: 661-684.

Schwartz RD, Kellar KJ (1983) Nicotinic cholinergic receptor binding sites in the brain: regulation in vivo. Science 220: 214-216.

Sgard F, Charpantier E, Bertrand S, Walker N, Caput D, Graham D, Bertrand D, Besnard F (2002) A novel human nicotinic receptor subunit, alpha10, that confers functionality to the alpha9-subunit. Mol Pharmacol 61: 150-159.

Shalaby AS, El-Hady Sweilum OA, Ads MK (2015) Does Tramadol Increase the Severity of Nicotine Dependence? A Study in an Egyptian Sample. J Psychoactive Drugs 47: 197-202.

Sharma R, Gartner CE, Hall WD (2016) The challenge of reducing smoking in people with serious mental illness. Lancet Respir Med 4: 835-844.

Shen J, Wu J (2015) Nicotinic Cholinergic Mechanisms in Alzheimer's Disease. Int Rev Neurobiol 124: 275- 292.

Shiraishi M, Minami K, Uezono Y, Yanagihara N, Shigematsu A, Shibuya I (2002) Inhibitory effects of tramadol on nicotinic acetylcholine receptors in adrenal chromaffin cells and in Xenopus oocytes expressing alpha 7 receptors. Br J Pharmacol 136: 207-216.

Shoptaw S, Rotheram-Fuller E, Yang X, Frosch D, Nahom D, Jarvik ME, Rawson RA, Ling W (2002) Smoking cessation in methadone maintenance. Addiction 97: 1317-28; discussion 1325.

Sim SK (1973) Methadone. Can Med Assoc J 109: 615-619.

Simojoki K, Alho H (2013) A Five-Year Follow-up of Buprenorphine Abuse Potential. J Alcohol Drug Depend 1: 1-6.

Smith TT, Rupprecht LE, Cwalina SN, Onimus MJ, Murphy SE, Donny EC, Sved AF (2016) Effects of Monoamine Oxidase Inhibition on the Reinforcing Properties of Low-Dose Nicotine. Neuropsychopharmacology 41: 2335-2343.

Son CD, Moss FJ, Cohen BN, Lester HA (2009) Nicotine normalizes intracellular subunit stoichiometry of nicotinic receptors carrying mutations linked to autosomal dominant nocturnal frontal lobe epilepsy. Mol Pharmacol 75: 1137-1148.

Spencer RJ, Jin W, Thayer SA, Chakrabarti S, Law PY, Loh HH (1997) Mobilization of Ca2+ from intracellular stores in transfected neuro2a cells by activation of multiple opioid receptor subtypes. Biochem Pharmacol 54: 809-818.

Spiga R, Schmitz J, Day J,2nd (1998) Effects of nicotine on methadone self-administration in humans. Drug Alcohol Depend 50: 157-165.

Srinivasan R, Pantoja R, Moss FJ, Mackey ED, Son CD, Miwa J, Lester HA (2011) Nicotine up-regulates alpha4beta2 nicotinic receptors and ER exit sites via stoichiometry-dependent chaperoning. J Gen Physiol 137: 59-79.

67 References

Stein C (2016) Opioid Receptors. Annu Rev Med 67: 433-451.

Stein MD, Weinstock MC, Herman DS, Anderson BJ, Anthony JL, Niaura R (2006) A smoking cessation intervention for the methadone-maintained. Addiction 101: 599-607.

Steinlein OK, Bertrand D (2010) Nicotinic receptor channelopathies and epilepsy. Pflugers Arch 460: 495- 503.

Stoker AK, Markou A (2015) Neurobiological Bases of Cue- and Nicotine-induced Reinstatement of Nicotine Seeking: Implications for the Development of Smoking Cessation . Curr Top Behav Neurosci 24: 125-154.

Storch A, Schrattenholz A, Cooper JC, Abdel Ghani EM, Gutbrod O, Weber KH, Reinhardt S, Lobron C, Hermsen B, Soskic V (1995) Physostigmine, galanthamine and codeine act as 'noncompetitive nicotinic receptor agonists' on clonal rat cells. Eur J Pharmacol 290: 207-219.

Story J, Stark MJ (1991) Treating cigarette smoking in methadone maintenance clients. J Psychoactive Drugs 23: 203-215.

Stutzmann GE, Mattson MP (2011) Endoplasmic reticulum Ca(2+) handling in excitable cells in health and disease. Pharmacol Rev 63: 700-727.

Sullivan JP, Decker MW, Brioni JD, Donnelly-Roberts D, Anderson DJ, Bannon AW, Kang CH, Adams P, Piattoni-Kaplan M, Buckley MJ (1994) (+/-)-Epibatidine elicits a diversity of in vitro and in vivo effects mediated by nicotinic acetylcholine receptors. J Pharmacol Exp Ther 271: 624-631.

Syed YY, Keating GM (2013) Extended-release intramuscular naltrexone (VIVITROL(R)): a review of its use in the prevention of relapse to opioid dependence in detoxified patients. CNS Drugs 27: 851-861.

Taly A, Corringer PJ, Guedin D, Lestage P, Changeux JP (2009) Nicotinic receptors: allosteric transitions and therapeutic targets in the nervous system. Nat Rev Drug Discov 8: 733-750.

Tapia L, Kuryatov A, Lindstrom J (2007) Ca2+ permeability of the (alpha4)3(beta2)2 stoichiometry greatly exceeds that of (alpha4)2(beta2)3 human acetylcholine receptors. Mol Pharmacol 71: 769-776.

Tapper AR, McKinney SL, Nashmi R, Schwarz J, Deshpande P, Labarca C, Whiteaker P, Marks MJ, Collins AC, Lester HA (2004) Nicotine activation of alpha4* receptors: sufficient for reward, tolerance, and sensitization. Science 306: 1029-1032.

Tassonyi E, Charpantier E, Muller D, Dumont L, Bertrand D (2002) The role of nicotinic acetylcholine receptors in the mechanisms of . Brain Res Bull 57: 133-150.

Taylor CP, Traynelis SF, Siffert J, Pope LE, Matsumoto RR (2016) Pharmacology of dextromethorphan: Relevance to dextromethorphan/ (Nuedexta(R)) clinical use. Pharmacol Ther 164: 170-182.

Tedford HW, Zamponi GW (2006) Direct G protein modulation of Cav2 calcium channels. Pharmacol Rev 58: 837-862.

Thomsen MS, Zwart R, Ursu D, Jensen MM, Pinborg LH, Gilmour G, Wu J, Sher E, Mikkelsen JD (2015) alpha7 and beta2 Nicotinic Acetylcholine Receptor Subunits Form Heteromeric Receptor Complexes that Are Expressed in the Human Cortex and Display Distinct Pharmacological Properties. PLoS One 10: e0130572.

Thorlin T, Eriksson PS, Persson PA, Aberg ND, Hansson E, Ronnback L (1998) Delta-opioid receptors on astroglial cells in primary culture: mobilization of intracellular free calcium via a pertussis sensitive G protein. Neuropharmacology 37: 299-311.

Toll L, Bruchas MR, Calo' G, Cox BM, Zaveri NT (2016) Nociceptin/Orphanin FQ Receptor Structure, Signaling, Ligands, Functions, and Interactions with Opioid Systems. Pharmacol Rev 68: 419-457.

68 References

Trescot AM, Datta S, Lee M, Hansen H (2008) Opioid pharmacology. Pain Physician 11: S133-53.

Tumkosit P, Kuryatov A, Luo J, Lindstrom J (2006) Beta3 subunits promote expression and nicotine- induced up-regulation of human nicotinic alpha6* nicotinic acetylcholine receptors expressed in transfected cell lines. Mol Pharmacol 70: 1358-1368.

Turner JR, Castellano LM, Blendy JA (2011) Parallel -like effects and upregulation of neuronal nicotinic acetylcholine receptors following chronic nicotine and varenicline. Nicotine Tob Res 13: 41-46.

Turner JR, Kellar KJ (2005) Nicotinic cholinergic receptors in the rat cerebellum: multiple heteromeric subtypes. J Neurosci 25: 9258-9265.

Tzschentke TM (1998) Measuring reward with the conditioned place preference paradigm: a comprehensive review of drug effects, recent progress and new issues. Prog Neurobiol 56: 613-672.

Unwin N, Fujiyoshi Y (2012) Gating movement of acetylcholine receptor caught by plunge-freezing. J Mol Biol 422: 617-634.

Vadivelu N, Anwar M (2010) Buprenorphine in postoperative pain management. Anesthesiol Clin 28: 601- 609.

Vallejo YF, Buisson B, Bertrand D, Green WN (2005) Chronic nicotine exposure upregulates nicotinic receptors by a novel mechanism. J Neurosci 25: 5563-5572. van Dorp E, Yassen A, Dahan A (2007) Naloxone treatment in opioid addiction: the risks and benefits. Expert Opin Drug Saf 6: 125-132.

Vernino S, Amador M, Luetje CW, Patrick J, Dani JA (1992) Calcium modulation and high calcium permeability of neuronal nicotinic acetylcholine receptors. Neuron 8: 127-134.

Vihavainen T, Mijatovic J, Piepponen TP, Tuominen RK, Ahtee L (2006) Effect of morphine on locomotor activity and striatal monoamine in nicotine-withdrawn mice. Behav Brain Res 173: 85-93.

Vihavainen T, Piltonen M, Tuominen RK, Korpi ER, Ahtee L (2008) Morphine-nicotine interaction in conditioned place preference in mice after chronic nicotine exposure. Eur J Pharmacol 587: 169-174.

Vijayaraghavan S, Pugh PC, Zhang ZW, Rathouz MM, Berg DK (1992) Nicotinic receptors that bind alpha- bungarotoxin on neurons raise intracellular free Ca2+. Neuron 8: 353-362.

Villegier AS, Belluzzi JD, Leslie FM (2011) Serotonergic mechanism underlying tranylcypromine enhancement of nicotine self-administration. Synapse 65: 479-489.

Villegier AS, Lotfipour S, McQuown SC, Belluzzi JD, Leslie FM (2007) Tranylcypromine enhancement of nicotine self-administration. Neuropharmacology 52: 1415-1425.

Viveros MP, Marco EM, File SE (2006) Nicotine and : parallels, contrasts and interactions. Neurosci Biobehav Rev 30: 1161-1181.

Volkow ND, Morales M (2015) The Brain on Drugs: From Reward to Addiction. Cell 162: 712-725.

Wada E, Wada K, Boulter J, Deneris E, Heinemann S, Patrick J, Swanson LW (1989) Distribution of alpha 2, alpha 3, alpha 4, and beta 2 neuronal nicotinic receptor subunit mRNAs in the central nervous system: a hybridization histochemical study in the rat. J Comp Neurol 284: 314-335.

Wageman CR, Marks MJ, Grady SR (2014) Effectiveness of nicotinic agonists as desensitizers at presynaptic alpha4beta2- and alpha4alpha5beta2-nicotinic acetylcholine receptors. Nicotine Tob Res 16: 297-305.

Waldhoer M, Bartlett SE, Whistler JL (2004) Opioid receptors. Annu Rev Biochem 73: 953-990.

69 References

Walsh H, Govind AP, Mastro R, Hoda JC, Bertrand D, Vallejo Y, Green WN (2008) Up-regulation of nicotinic receptors by nicotine varies with receptor subtype. J Biol Chem 283: 6022-6032.

Wang F, Gerzanich V, Wells GB, Anand R, Peng X, Keyser K, Lindstrom J (1996) Assembly of human neuronal nicotinic receptor alpha5 subunits with alpha3, beta2, and beta4 subunits. J Biol Chem 271: 17656-17665.

Wang H, Sun X (2005) Desensitized nicotinic receptors in brain. Brain Res Brain Res Rev 48: 420-437.

Wang JM, Zhang L, Yao Y, Viroonchatapan N, Rothe E, Wang ZZ (2002) A transmembrane motif governs the surface trafficking of nicotinic acetylcholine receptors. Nat Neurosci 5: 963-970.

Weinshenker D, Schroeder JP (2007) There and back again: a tale of norepinephrine and drug addiction. Neuropsychopharmacology 32: 1433-1451.

Werry TD, Wilkinson GF, Willars GB (2003) Mechanisms of cross-talk between G-protein-coupled receptors resulting in enhanced release of intracellular Ca2+. Biochem J 374: 281-296.

Whiteaker P, Peterson CG, Xu W, McIntosh JM, Paylor R, Beaudet AL, Collins AC, Marks MJ (2002) Involvement of the alpha3 subunit in central nicotinic binding populations. J Neurosci 22: 2522-2529.

Whiteaker P, Sharples CG, Wonnacott S (1998) Agonist-induced up-regulation of alpha4beta2 nicotinic acetylcholine receptors in M10 cells: pharmacological and spatial definition. Mol Pharmacol 53: 950-962.

Williams DK, Wang J, Papke RL (2011) Positive allosteric modulators as an approach to nicotinic acetylcholine receptor-targeted therapeutics: advantages and limitations. Biochem Pharmacol 82: 915- 930.

Wiseman EJ, McMillan DE (1998) Rationale for cigarette smoking and for mentholation preference in cocaine- and nicotine-dependent outpatients. Compr Psychiatry 39: 358-363.

Wolff HG, Hardy JD, Goodell H (1940) Studies on Pain. Measurement of the Effect of Morphine, Codeine, and Other on the Pain Threshold and an Analysis of their Relation to the Pain Experience. J Clin Invest 19: 659-680.

Wonnacott S (1997) Presynaptic nicotinic ACh receptors. Trends Neurosci 20: 92-98.

Wonnacott S (1990) The paradox of nicotinic acetylcholine receptor upregulation by nicotine. Trends Pharmacol Sci 11: 216-219.

Wu J, Liu Q, Tang P, Mikkelsen JD, Shen J, Whiteaker P, Yakel JL (2016) Heteromeric alpha7beta2 Nicotinic Acetylcholine Receptors in the Brain. Trends Pharmacol Sci 37: 562-574.

Wullner U, Gundisch D, Herzog H, Minnerop M, Joe A, Warnecke M, Jessen F, Schutz C, Reinhardt M, Eschner W, Klockgether T, Schmaljohann J (2008) Smoking upregulates alpha4beta2* nicotinic acetylcholine receptors in the human brain. Neurosci Lett 430: 34-37.

Wyllie DJ, Chen PE (2007) Taking the time to study competitive antagonism. Br J Pharmacol 150: 541- 551.

Xiao Y, Fan H, Musachio JL, Wei ZL, Chellappan SK, Kozikowski AP, Kellar KJ (2006) Sazetidine-A, a novel ligand that desensitizes alpha4beta2 nicotinic acetylcholine receptors without activating them. Mol Pharmacol 70: 1454-1460.

Xiao Y, Kellar KJ (2004) The comparative pharmacology and up-regulation of rat neuronal nicotinic receptor subtype binding sites stably expressed in transfected mammalian cells. J Pharmacol Exp Ther 310: 98-107.

Xiao Y, Smith RD, Caruso FS, Kellar KJ (2001) Blockade of rat alpha3beta4 nicotinic receptor function by methadone, its metabolites, and structural analogs. J Pharmacol Exp Ther 299: 366-371.

70 References

Yi JJ, Ehlers MD (2007) Emerging roles for ubiquitin and protein degradation in neuronal function. Pharmacol Rev 59: 14-39.

Young JW, Geyer MA (2013) Evaluating the role of the alpha-7 nicotinic acetylcholine receptor in the pathophysiology and treatment of schizophrenia. Biochem Pharmacol 86: 1122-1132.

Yu LF, Zhang HK, Caldarone BJ, Eaton JB, Lukas RJ, Kozikowski AP (2014) Recent developments in novel targeting alpha4beta2-nicotinic acetylcholine receptors. J Med Chem 57: 8204-8223.

Yu PH, Boulton AA (1987) Irreversible inhibition of monoamine oxidase by some components of cigarette smoke. Life Sci 41: 675-682.

Zale EL, Dorfman ML, Hooten WM, Warner DO, Zvolensky MJ, Ditre JW (2015) Tobacco Smoking, Nicotine Dependence, and Patterns of Prescription Opioid Misuse: Results From a Nationally Representative Sample. Nicotine Tob Res 17: 1096-1103.

Zarrindast MR, Faraji N, Rostami P, Sahraei H, Ghoshouni H (2003) Cross-tolerance between morphine- and nicotine-induced conditioned place preference in mice. Pharmacol Biochem Behav 74: 363-369.

Zarrindast MR, Farzin D (1996) Nicotine attenuates naloxone-induced jumping behaviour in morphine- dependent mice. Eur J Pharmacol 298: 1-6.

Zarrindast MR, Pazouki M, Nassiri-Rad S (1997) Involvement of cholinergic and opioid receptor mechanisms in nicotine-induced antinociception. Pharmacol Toxicol 81: 209-213.

Zhao L, Kuo YP, George AA, Peng JH, Purandare MS, Schroeder KM, Lukas RJ, Wu J (2003) Functional properties of homomeric, human alpha 7-nicotinic acetylcholine receptors heterologously expressed in the SH-EP1 human epithelial cell line. J Pharmacol Exp Ther 305: 1132-1141.

Zindel LR, Kranzler HR (2014) Pharmacotherapy of alcohol use disorders: seventy-five years of progress. J Stud Alcohol Drugs Suppl 75 Suppl 17: 79-88.

Zirakzadeh A, Shuman C, Stauter E, Hays JT, Ebbert JO (2013) Cigarette smoking in methadone maintained patients: an up-to-date review. Curr Drug Abuse Rev 6: 77-84.

Zoli M, Lena C, Picciotto MR, Changeux JP (1998) Identification of four classes of brain nicotinic receptors using beta2 mutant mice. J Neurosci 18: 4461-4472.

Zoli M, Moretti M, Zanardi A, McIntosh JM, Clementi F, Gotti C (2002) Identification of the nicotinic receptor subtypes expressed on dopaminergic terminals in the rat striatum. J Neurosci 22: 8785-8789.

Zoli M, Pistillo F, Gotti C (2015) Diversity of native nicotinic receptor subtypes in mammalian brain. Neuropharmacology 96: 302-311.

Zorrilla EP, Logrip ML, Koob GF (2014) Corticotropin releasing factor: a key role in the neurobiology of addiction. Front Neuroendocrinol 35: 234-244.

Zwart R, Strotton M, Ching J, Astles PC, Sher E (2014) Unique pharmacology of heteromeric alpha7beta2 nicotinic acetylcholine receptors expressed in Xenopus laevis oocytes. Eur J Pharmacol 726: 77-86.

Zwart R, Vijverberg HP (1998) Four pharmacologically distinct subtypes of alpha4beta2 nicotinic acetylcholine receptor expressed in Xenopus laevis oocytes. Mol Pharmacol 54: 1124-1131.

71 REETA TALKA Effects of Opioids on Neuronal Nicotinic Acetylcholine Receptors Recent Publications in this Series

10/2017 Veera Pohjolainen Health-Related Quality of Life and Cost-Utility in Bulimia Nervosa and Anorexia Nervosa in Women 11/2017 Lotta von Ossowski Interaction of GluA1 AMPA Receptor with Synapse-Associated Protein 97 12/2017 Emma Andersson dissertationes scholae doctoralis ad sanitatem investigandam Characterization of Mature T-Cell Leukemias by Next-Generation Sequencing and Drug universitatis helsinkiensis 30/2017 Sensitivity Testing 13/2017 Solja Nyberg Job Strain as a Risk Factor for Obesity, Physical Inactivity and Type 2 Diabetes – a Multi-cohort Study 14/2017 Eero Smeds REETA TALKA Cortical Processes Related to Motor Stability and Proprioception in Human Adults and Newborns Effects of Opioids on Neuronal Nicotinic 15/2017 Paavo Pietarinen Effects of Genotype and Phenotype in Personalized Drug Therapy Acetylcholine Receptors 16/2017 Irene Ylivinkka Netrins in Glioma Biology: Regulators of Tumor Cell Proliferation, Motility and Stemness 17/2017 Elisa Lázaro Ibáñez Extracellular Vesicles: Prospects in Prostate Cancer Biomarker Discovery and Drug Delivery 18/2017 Anu Kaskinen Measurement of Lung Liquid and Outcome after Congenital Cardiac Surgery 19/2017 Taru Hilander Molecular Consequences of Transfer-RNA Charging Defects 20/2017 Laura Teirilä Activation of the Inflammatory Response by Fungal Components 21/2017 Laura Sokka Burnout in the Brain at Work 22/2018 Martti Rechardt Metabolic and Inflammatory Factors in Upper Extremity Soft-Tissue Disorders 23/2017 Jaana Hautala Improving the Palatability of Minitablets for Feline Medication 24/2017 Satu Lehti Extracellular Lipid Particles in Atherosclerosis and Aortic Stenosis 25/2017 Asko Wegelius Influence of Birth Weight on the Risk and Clinical Presentation of Schizophrenia 26/2017 Siva P.R. Maddirala Venkata Public Health and Patient Care Aspects in Pharmacy Education and Pharmacists’ Role in National Health Care Programs in India 27/2017 Kristyna Spillerova The Role of the Angiosome Concept in the Treatment of below the knee Critical Limb Ischemia 28/2017 Anna-Riia Holmström Learning from Medication Errors in Healthcare — How to Make Medication Error Reporting Systems Work? 29/2017 Aaro Haapaniemi DIVISION OF PHARMACOLOGY AND PHARMACOTHERAPY Laryngeal Cancer Recurrence, Prognostic Factors and Management FACULTY OF PHARMACY DOCTORAL PROGRAMME IN DRUG RESEARCH UNIVERSITY OF HELSINKI 30/2017

Helsinki 2017 ISSN 2342-3161 ISBN 978-951-51-3126-3